Вы находитесь на странице: 1из 128

1

LANGMUIR-BLODGETT (LB) AND


SELF-ASSEMBLED (SA) POLYMERIC FILMS
Osvaldo N. Oliveira Jr
1
., Maria Raposo
2
and Anantharaman Dhanabalan
3
1
Instituto de Fsica de So Carlos, USP, CP 369, 13569-970 So Carlos, SP, Brasil.
2
Departamento de Fsica, Faculdade de Cincias e Tecnologia,
Universidade Nova de Lisboa, 2825 Monte Caparica, Portugal.
3
Laboratory for Macromolecular and Organic Chemistry, Eindhoven University of Technology,
P.O. Box. 513, 5600 MB Eindhoven, The Netherlands.

Telephone: +55 16 2715365; Fax: +55 16 2713616; E-mail: chu@if.sc.usp.br

Keywords : Langmuir-Blodgett (LB), self-assembly (SA), monolayer, layer-by-layer, ultra-thin films, conducting
polymers, photosensitive polymers, physical adsorption, electrical properties, optical properties, molecular control,
polyaniline, polypyrrole, polythiophene, poly(p-phenylene vinylene), polydiacetylene, azobenzene polymers,
organic LEDs, gas sensors, atomic force microscopy.


Table of contents
I. Introduction
II. Langmuir-Blodgett (LB) of functional polymers
2.1. LB manipulation of materials a general overview
2.2. LB manipulation of polypyrroles
1. Electrochemical polymerization in LB multilayers of amphiphilic pyrrole monomers
2. Polymerization of the pyrrole monomer at the air-water interface
3. Parent/substituted polypyrrole doped with amphiphilic dopants
4. Metal-fatty acid salt LB film templates
5. Amphiphilic polypyrroles
6. Composite LB approach
2.3. LB manipulation of polythiophenes
1. Low molecular weight polymers
2. Mixed LB film approach
3. Functionalized polythiophenes
4. Optical, Structural and Electrical Characteristics
2.4. LB manipulation of polyanilines
1. Surface pressure-area isotherms of parent/substituted polyanilines
2. Surface potential-area isotherms of parent/substituted polyanilines
3. Brewster angle microscopy (BAM) of polyanilines
4. Polymerization at the air-water interface
5. The mixed monolayer approach
6. Fabrication and optical and structural characteristics
7. Electrical and electrochemical properties
8. Post-deposition treatment
9. Devices using polyaniline LB films
2.5. LB manipulation of azobenzene polymers
1. Azobenzene-containing monomeric amphiphiles
2. Azobenzene polymers
3. Storage characteristics of azobenzene polymers
2.6. LB manipulation of polydiacetylene
1. Polymerization of diacetylenes as applied to LB films
2. Langmuir monolayers of diacetylene amphiphiles
3. Mixed monolayers of diacetylene/polydiacetylenes
2
4. Preparation and characteristics of LB films of polydiacetylenes
5. Possible applications of polydiacetylene LB films
III. Self-assembled (SA) polymer films
3.1. The Self-assembly technique
3.2. Physical adsorption of polymers
1. Kinetics of adsorption
2. Adsorption Isotherms
3. Mechanisms for polymer adsorption
3.3. Characterization of SA films
3.4. Possible applications of SA polymeric films
IV. Concluding remarks
V. References

I. Introduction
The growing interest in organic materials requires research that can be broadly divided into four main
categories, as shown schematically in Scheme - 1: a) material design; b) synthesis; c) processing and
characterization; d) testing for applications. It is obviously an interdisplinary effort from scientists belonging to
various branches including chemistry, physics, materials science, electrical engineering and so on.
Design Synthesis
Testing for
application
Processing

Scheme 1. Organic Materials Science Cycle

This is not necessarily a one-way route, since often one has to go back to previous steps for
optimization/improvement. The present review is mainly focused on the third category, i.e. materials processing.
Having the designed material synthesized, in most cases an important step prior to testing a proposed application is
the processing of the materials into a desired form, preferably in the form of ultrathin films that often possess a large
surface area. For organic materials, in particular, the most commonly used techniques are simple drop casting, spin
casting, dip coating, Langmuir-Blodgett (LB) technique and self-assembly. A common feature in all these
techniques is that the material of interest is first dissolved in a suitable solvent, being subsequently either
physisorbed or chemisorbed onto a surface. In addition to these methods through which films are fabricated at room
temperature and atmospheric pressure, there are also techniques such as chemical vapor deposition and molecular
beam epitaxy in which the material is either vaporized (CVD) or ionized (MBE) for obtaining thin films. The latter
methods will not be discussed at any length here.
3
In comparison to other techniques, the LB and self-assembly methods are looked upon as viable means of
producing ultrathin films with controllable film thickness and molecular packing. While the former involves the
formation of a monolayer film at the air-water interface and subsequent transfer as a multilayer film onto a solid
substrate (a, Fig. 1), the latter technique is essentially based on the ionic interaction between alternate layers (b, Fig.
1). This chapter will concentrate on these techniques.
Substrate
LB Films
Substrate
+
+
+
+
+
+
-
-
-
-
-
-
+
+
+
+
+
+
-
-
-
-
-
-
+
+
+
+
+
+
-
-
-
-
-
-
+
+
+
+
+
+
-
-
-
-
-
-
SA Films
(a)
(b)

Figure 1. Schematic representation of (a) Langmuir-Blodgett (LB) films and (b) Self-Assembly (SA) films


II. Langmuir-Blodgett (LB) of functional polymers
The possibility offered by the LB technique for producing ultrathin films with controllable thickness and
molecular architecture prompted researchers across diversified fields to extend this technique to non-classical semi-
amphiphilic and non-amphiphilic molecules, such as polymers and other macromolecules. During the late 1980s,
there was an intense search for suitable approaches for producing thin films of functional polymers that exhibit fine-
tunable optical, electrical and electronic optical properties. Such studies were particularly aimed at understanding
the polymer properties in a more reproducible way and at exploiting these polymers in a variety of applications.
Also, most polymers, in general, possess amorphous structure with isotropic properties with no or little steric
regulation. The LB technique was recognized as a viable route for producing polymer thin films wit h highly
anisotropic properties.


2.1. LB Manipulation of materials a general overview
4
Detailed accounts on the LB processing of different materials can be found in reviews and books in ref. [1-
23]. The Langmuir-Blodgett (LB) technique has been tradit ionally applied to amphiphilic molecules that possess a
hydrophilic head group attached with a hydrophobic alkyl chain, e.g. stearic acid. In a common LB experiment, a
known amount of amphiphilic material dissolved in a water-immiscible, volatile organic solvent such as chloroform
is placed on the water surface. After evaporation of the solvent, the monolayer material is compressed with movable
barriers to form a monolayer, referred to as Langmuir monolayer. Monolayer formation is usually monitored with
surface pressure ()-area (A) isotherms, as shown in Fig. 2 for stearic acid. Also shown is the schematic diagram of
a Langmuir trough that is usually made up of hydrophobic materials like Teflon to contain the subphase water and
movable barrier(s) that span over the water surface. It also contains measuring devices including a Wilhelmy plate
connected to an electrobalance for measuring surface pressure, a Kelvin probe for measuring surface potential and a
dipper system for transfer of the monolayer onto a solid substrate.

Area per molecule (
2
)
S
u
r
f
a
c
e

p
r
e
s
s
u
r
e

(
m
N
/
m
)

Figure 2. Typical Surface pressure-Area (-A) isotherm of a Langmuir film of stearic acid at the air-water interface.
Also shown is the schematic diagram of a Langmuir trough.


5
The -A isotherm usually comprises regions with distinct compressibility, representing the various levels of
molecular packing at the air-water interface. At very large areas per molecule, the molecules interact poorly with
each other (referred to as 2D gas phase) until they are compressed to a liftoff area at which the head groups start to
interact, while the hydrophobic chains are still randomly oriented (2D liquid phase). Further compression leads to a
closely packed monolayer where both head and tail groups are packed in an ordered manner (3D solid phase).
Attempts to further compress the monolayer result in collapse, which is reflected in a sudden increase of
compressibility. Stringent conditions must be adopted for obtaining reproducible monolayers, including use of pure
materials, spreading solvents and ultrapure water, accurate weighing of monolayer material and a clean and
vibration-free environment. Under optimized conditions, the monolayer may be amenable to transfer onto
hydrophilic/hydrophobic solid supports either by the vertical dipping method or horizontal touching method (also
named Langmuir-Schaeffer method). First hand information on the quality of transferred films may be obtained
from the transfer ratio (TR), which is the area moved by the barrier during monolayer transfer divided by the area of
substrate coated with the monolayer. For an ideal transfer process, TR should be one. Transfer can be of Y-type
(transfer during both insertion and withdrawal of the substrate), Z-type (transfer only during withdrawal) or X-type
(transfer only during insertion) (Fig. 3).
Hydrophilic
substrate
Y - type film
Z - type film
X - type film
Hydrophilic
substrate
Hydrophobic
substrate

Figure 3. Different types of LB films

6
In comparison to model amphiphilic molecules, the LB manipulation of polymeric materials is far from simple.
Different molecular engineering approaches have been employed for LB processing of polymeric materials, as
depicted in Scheme-2.
Monomer
Langmuir film
of monomer
LB film of
monomer
Polymer
Langmuir film
of polymer
LB film of
polymer
[a]
[b]
[c]
[A]
[B]
[C]
[I]
[II]
[III]

Scheme2. Different molecular engineering approaches employed in the LB processing of functional polymeric
materials

Route I-II-III, i.e., LB processing of pre-formed polymers, is the approach commonly employed for conventional
polymers and also for functional polymers such as polyanilines, polypyrroles, polyalkylthiophenes. Step-I is usually
achieved via chemical or electrochemical polymerization of the corresponding monomer. Such pre-formed polymers
are directly spread on the water surface along with the spreading solvent as a monolayer (step-II) whose packing is
determined by the lateral cohesive interactions of polymer chains, which in turn depends on pendent groups in the
polymer chain, with the water surface. When spread at the air-water interface, polymers normally assume a random
2D/3D conformation (Fig. 4) which calls for an entirely different perspective from the one based on monolayers
from model amphiphilic molecules. For the polymers consist of chains with different lengths (i.e. they are
polydisperse) and different end groups and most often defects along the polymer main, where these parameters
depend mainly on the synthetic procedure to obtain the polymers.
7
in solution
expanded
2D monolayer
condensed
2D monolayer
collapsed
3D monolayer

Figure-4. Spreading behavior of polymers at the air-water interface

Also, with some pendent groups, segments of the polymer chains can form a loop-like structure (Fig. 5) depending
upon the lateral compression which makes the analysis of the monolayer results in general, and the area calculations
in particular, less straightforward.

low
moderate
high

Figure 5. Organization of polymers at different stages of compression at the air-water interface

The packing arrangement of the polymer chains can also change during monolayer transfer onto a substrate as
multilayer LB films. Route A-B-C, which involves the spreading of monomeric material as a monolayer on the
water surface and subsequent polymerization of monomers, is the method developed by Tredgold [18] for obtaining
maleic anhydride vinyl copolymers. This approach has later been applied in the LB manipulation of various
functional polymers such as polyalkylanilines, poly(N-alkylpyrrole)s, polydiacetylenes. In the latter example, the
polymerization is achieved by UV-irradiation rather than through chemical polymerization of monomers. In the
8
approach (a-b-c), polymerization is performed after transferring the monolayer of monomers onto a support, either
by electrochemical means or by UV-irradiation. Typical examples of polymers processed in this manner include
polydiacetylenes and polypyrroles.
In contrast to simple monomeric monolayer materials, the polymers in general possess limited solubility in
the spreading solvents used in LB film fabrication. This solubility problem is alleviated by using highly polar water-
miscible solvents such as N-methylpyrrolidone, N,N-dimethylformamide, N,N-dimethylacetamide,
dimethylsulfoxide or mixtures of these solvent(s) with chloroform. Polymer derivatization with groups such as long
alkyl chains also helps to solubilize these otherwise insoluble polymers. Despite these approaches, most of the
functional polymers do not dissolve into the spreading solvent readily and completely and one has to ultrasonicate
the spreading solution and filter the undissolved material to get a clear spreading solution to use in LB experiments.
Precipitation of the polymer from the spreading solution upon storing is also commonly observed, and therefore the
use of freshly prepared spreading solution is recommended. Such difficulties in obtaining true solutions in which
the monolayer material is supposedly molecularly dissolved can lead to irreproducible monolayer results.
In the following sections, we focus on the LB manipulation of three well -studied conducting polymers,
namely polypyrroles, polythiophenes and polyanilines and two photosensitive polymers, namely polydiacetylenes
and azobenzene polymers, with particular emphasis on film preparation and on the optical, structural and electrical
characteristics of these LB films.

2.2. LB Manipulation of polypyrroles
The approaches used in the LB manipulation of polypyrrole are listed below, accompanied by a short
discussion on each approach:

1. Electrochemical polymerization in LB multilayers of amphiphilic pyrrole monomers [24-26]
Electrochemical polymerization has been accomplished with the pyrrole derivatives shown in Fig. 6 [24,
26] and appears to be the first attempt to fabricate polypyrrole -based LB films. Among these three amphiphilic
pyrrole derivatives (1-3), the one with simple alkyl substitution on the N atom of pyrrole (1) did not form a
monolayer on pure water even after mixing with builder materials such as octadecane, though an expanded
monolayer could be obtained on acidic subphases.
9
N
C
18
H
37
N
CH
3
C
18
H
37
HOOC
N
H
C H
3
CO
2
C
18
H
37
(1) (3) (2)

Figure 6. Structures of different amphiphilic pyrrole monomers used in ref. [24].


However, the polar group and long alkyl chain substituted pyrroles (2 and 3) did form stable monolayers upon
mixing with octadecane, which underlines the necessity to consider an adequate balance between hydrophilic and
hydrophobic groups in designing new molecules to be used with the LB technique. These monomeric mixed LB
films were transferred onto conducting substrates as Y-type films and subsequently electropolymerized by partially
dipping the LB film in an electrolyte that contained oxidant (Fig. 7). The course of polymerization was monitored
through visible-near-IR spectrum, with the appearance of an absorption band at about 500 nm corresponding to the
-* transition and a broad band
LiCl
4
-CH
3
CN
ITO
Pyrrole
monomer
Pt wire
Substrate
LiClO
4
-CH
3
CN

Figure7. Electrochemical polymerization of the LB film of amphiphilic pyrrole monomers [26]

at about 1500 nm, and by attenuated total reflection-infrared spectrum, through the disappearance of the band at 780
cm
-1
assigned to the C-H out-of-plane bending vibration. The structure of these electropolymerized films was
investigated with X-ray diffraction, from which bilayer distances of 56.7 and 58.1 were observed for monomeric
and polymeric LB films, respectively. In the cross sectional transmission electron microscopic (TEM) image
obtained after staining the polymeric LB film with RuO
4
, a striped pattern of dark (due to polypyrrole) and light
10
region (due to alkyl chain region) was observed and the spacing between these strips was close to the value obtained
from XRD measurements [25]. The bilayer distance varied with the deposition surface pressure and the substrate
onto which the monolayer was transferred. These films also exhibited high anisotropic conductivity with the in-
plane conductivity being 10 orders of magnitude higher than the conductivity across the sample bulk [26]. Since the
degree of electropolymerization could not be determined, it is possible that the LB films obtained through this self-
assisted electropolymerization route comprised a mixture of monomers, low molecular weight oligomers and high
molecular weight polymer.

2. Polymerization of the pyrrole monomer at the air-water interface [27-41]
This approach was originally developed by Rubner et al [27], where the surface active pyrrole monomers
shown in Fig. 8 were deposit ed along with the pyrrole monomer onto an aqueous subphase containing oxidizing
agents such as ferric chloride.
N
H
(CH
2
)
17
CH
3
(CH
2
)
16
CH
3
N
H
O
(4)
(5)

Figure-8

Pure 3-octadecylpyrrole (4) formed a stable and condensed monolayer both at the air-water interface and at the air-
ferric chloride solution interface [27], but it did not polymerize on the surface of ferric chloride solution. Such a
failure was attributed to the structure of polypyrrole, which was believed to consist of a linear chain with alternating
pyrrole rings pointing in opposite directions in a way that the neighboring pendent alkyl chains are inverted in
relation to each other (a, Fig. 9). Polymerization could still be achieved when the amphiphilic pyrrole monomer was
co-deposited with the pyrrole monomer which can act as a bridging molecule (b, Fig. 9). The formation of
copolymers of 3-octadecylpyrrole and pyrrole and homopolymer of the parent polypyrrole was inferred using
several techniques.
11
(CH
2
)
14
CH
3
(CH
2
)
14
CH
3
(CH
2
)
14
CH
3
(CH
2
)
14
CH
3
(CH
2
)
14
CH
3
(CH
2
)
14
CH
3
(CH
2
)
14
CH
3
N N
N
N
N
N
N
* * n
H
H
H
H
H
H
H
Air
Water
(CH
2
)
14
CH
3
(CH
2
)
14
CH
3
(CH
2
)
14
CH
3
(CH
2
)
14
CH
3
N N
N
N
N
N
N
* * n
H
H
H
H
H
H
H
Air
Water
( a )
( b )

Figure-9

The partial solubility of the film in most organic solvents indicated that the film comprised a mixture of copolymer
and polypyrrole, for while the copolymer is soluble, polypyrrole is insoluble in most organic solvents. Near edge X-
ray absorption fine structure (NEXAFS) showed increased disorder in the copolymer LB film in comparison to the
monomeric film [28, 29]. Spectroscopy in the visible-near IR region displayed a broad absorption extending deep
into the near IR region, which was attributed to free charge carriers such as bipolarons [30] that also caused a
featureless broad absorption in the 1800-4000 cm
-1
region of a Fourier transform infrared spectrum. In the FTIR
spectrum, characteristic polypyrrole absorption bands were observed in the 1600-400 cm
-1
region apart from
methylene group stretching and bending vibrations. Such LB films possessed an average thickness per layer of about
30 and in-plane conductivity between 10
-1
and 10
-2
S/cm [30].
In other experiments, instead of 3-octadecylpyrrole, 3-octadecanoyl pyrrole (5) was used. The latter was
expected not to undergo polymerization due to deactivation of the pyrrole ring by the carbonyl group, but formed
stable monolayers at the air-water interface thereby leading to homopolymerization of the pyrrole monomer alone
[30, 32]. While the pyrrole-3-alkylpyrrole copolymer monolayer could only be transferred with the horizontal lifting
method, the monolayer comprising polypyrrole and 3-octadecanoylpyrrole could easily be transferred by the
12
conventional vertical dipping method [32]. This approach was extended to mixed LB films comprising pyrrole, 3-
hexadecylpyrrole and -ferrocene derivatized 3-alkyl pyrrole [31, 34] and also to produce multicomponent LB films
consisting of polypyrrole, 3-octadecanoyl pyrrole and poly(3-hexylthiophene) [35]. In the latter case, the pyrrole
monomer was polymerized at the air-ferric chloride subphase in the presence of poly(3-hexylthiophene) and 3-
octadecanoyl pyrrole and subsequently transferred (Z-type) as multicomponent LB films in which the polypyrrole
chains were found to preferentially orient parallel to the film surface [35]. The thickness and the conductivity of
copolymer LB films varied with the mole ratio of pyrrole:amphiphilic pyrrole, with highest conductivity for the film
obtained by spreading 5000:1 of pyrrole and amphiphilic pyrrole. The reason behind the need to use such a large
excess of pyrrole monomer was not completely understood. An anisotropy in conductivity of about three orders of
magnitude was observed in these films.
Duran et al [36] observed that the amphiphilic pyrrole on the water surface containing oxidant can be
polymerized when the monolayer was kept in the compressed state. For polymerization, monolayers of 3-
alkylpyrrole on an aqueous subphase containing oxidant (without pyrrole monomer) were initially compressed to a
particular surface pressure and held at that pressure while the change in mean molecular area (due to
polymerization) was monitored with time [36,38]. Unlike amphiphilic aniline, to be discussed later, the rate of
polymerization did not depend on the applied surface pressure. Polymerization has also been attempted in a two-
compartment LB trough [36], where one compartment contained pure water onto which the compressed pyrrole
monolayer was formed (so-called pre-orientation of the monomer). The monolayer was then moved using confining
barriers to another compartment containing the subphase with oxidizing agent. Polymerization at the air-water
interface was confirmed by nuclear magnetic resonance and gel permeation chromatography analysis of monolayer
material. Scanning tunneling microscopy features of monomer and polymer monolayers transferred onto substrates
have also been reported [37]. The monolayer polymer film exhibited a high degree of positional and orientational
order, and the multilayer polymer film was disordered due to the curvature of polymer chains. Such information has
been used to speculate on the possible organization of monomer/polymer at the air-water interface [37]. Similar
polymerization of amphiphilic pyrroles has been reported by other research groups [38, 39]. Polypyrrole films
alternated with lipid modified glucose oxidase on platinum electrodes were useful as glucose sensors employing
amperometric measurements [38], and films alternated with polyimide exhibited high thermal stability [39].
Preparation and electrochemistry of the alkylpyrrole LB films have been recently reported [41].

3. Parent/substituted polypyrrole doped with amphiphilic dopants [42,43]
LB films were obtained from parent polypyrrole/substituted polypyrrole which were amphiphilized (and
hence more processible) by chemical/electrochemical doping with amphiphilic dopants [42,43]. For instance, 3,4-
13
dibutylpyrrole, electrochemically doped with perfluorooctadecanoic acid, was soluble in organic solvents and could
be spread at the air-water interface using a mixture of trifluoroethanol and benzene as spreading solvent [42]. A
stable monolayer with a mean molecular area per pyrrole repeat unit of 32
2
was formed, which could be
transferred at 20 mN/m with the horizontal lifting method. The thickness per layer obtained from the TEM cross
section analysis was about 16 and the anisotropy in conductivity was of the order of three. In another report [43],
Langmuir monolayers were produced from the soluble part of parent polypyrrole that was obtained through
chemical polymerization in the presence of dodecylbenzene sulfonic acid [44,45]. The parent polypyrrole was
dissolved in a mixture of m-cresol and chloroform, and the undissolved high molecular weight/cross-linked polymer
was filtered. Surface pressure (a, Fig. 10) and surface potential isotherms (b, Fig. 11) were obtained at different
subphase temperatures.
2 4 6 8
0
100
200
300
400
500
2 4
1 7
1 2
7
o
C
S
u
r
f
a
c
e

p
o
t
e
n
t
i
a
l

/

m
V
Area per monomer /
2
2 4 6
0
10
20
30
40
50
24
17
12
7
O
C
S
u
r
f
a
c
e

p
r
e
s
s
u
r
e

/

m
N

m

-
1
Area per monomer / 2

Figure-10

The surface pressure increased monotonically with no detectable phase changes until collapse and the limiting mean
molecular was about 5
2
per repeat unit. The compression speed affected only slightly the isotherm, which was
shifted to lower mean molecular areas with increasing subphase temperature. The onset for surface potential, at
which the potential increases from 0 to approximately 200 mV, occurred at larger areas when the subphase
temperature decreased from 24 to 7 C. These results may be understood as the formation of less compressible
aggregates of polymer molecules at the air-water interface at lower subphase temperatures. These monolayers could
be transferred using the conventional vertical dipping method as Z-type LB films. The UV-vis and FTIR analysis of
the transferred multilayer LB film pointed to a partial dedoping of the polymer while at the air-water interface. A
direct comparison of the mean molecular area reported in ref. 42 (about 32
2
) and that observed in ref. 43 (about 5

2
) is not possible since a substituted polypyrrole was used in the former case in contrast to parent polypyrrole in
14
the latter case. Also, the dopants used were different - while the dopant itself formed a monolayer in the case of the
former, the dopant is water soluble in the latter case.

4. Metal-fatty acid salt LB film templates [46-55]
This approach involves the sequential exposure of LB films of metal-fatty acid salt to hydrogen chloride
gas and then to pyrrole vapor for obtaining thin doped polypyrrole films within the fatty acid LB multilayers (Fig.
11).
O O
-
O O
-
O O
-
O
O
-
O
O
-
O
O
-
Fe
3+
Fe
3+
O H
O O
O H
O H
O
OH O OH O
OH
O
FeCl
3
FeCl
3
HCl
gas
O H O O
O H
O H O
OH
O OH
O
OH O
pyrrole
vapour

Figure-11

Rubner et. al [46] were the first to demonstrate the applicability of such a simple course of solid state chemical
reactions in LB templates, using preformed ferric stearate LB films. The chemical reactions could be followed by
infrared spectroscopy, with the disappearance of the carbonyl stretching vibration of the carboxylate group at 1523
cm
-1
, and the appearance of strong absorption at 1700 cm
-1
due to the free carboxylic acid during HCl treatment. In
addition, absorption peaks appeared in the low frequency region of 1600-800 cm
-1
corresponding to polypyrrole,
upon pyrrole vapor treatment. The change in the orientation of alkyl chains of fatty salt/acid during the solid state
reaction was inferred through transmission/reflection Fourier transform infrared spectroscopy. Major orientational
changes were only observed during the acid treatment of ferric stearate films, with no significant changes being
noticed in the subsequent pyrrole vapor treatment [48]. Also inferred was the random orientation of polypyrrole
formed in the polar planes of the LB matrix, which is in contrast to the preferred in-plane orientation of polypyrrole
in the film obtained using the second approach [33]. Such random orientation of polymer chains has also been
inferred through UV-vis spectral measurements with polarized light. As one could anticipate, the diffraction peaks
normally observed in XRD patterns for precursor ferric stearate LB films disappear upon HCl treatment. The XRD
spectra obtained after pyrrole vapor treatment exhibited new diffraction peaks, which may indicate the emergence of
new layered structures within the film. This has not been fully understood as yet. The in-plane conductivity obtained
15
with such films lied in the range from 1 to 20 Scm
-1
[47, 48, 50, 55]. A high degree of anisotropy in conductivity,
viz. eight orders of magnitude, has been found for these films [47]. This approach was adopted later by a number of
groups with slight modifications, which include LB films with different fatty acid salts such as copper stearate [50],
silver arachidate [51], ferric palmitate [55] and different reagents such as perchloric acid instead of HCl [51]. It was
also extended for producing thin films of copolymer of pyrrole and N-methylpyrrole [49]. In the process of
optimizing the experimental parameters using this approach, one important variable identified was the concentration
of ferric chloride in the subphase [47]. Ferric chloride is known to undergo hydrolysis, thereby producing ferric
hydroxide gels at high concentrations. This specific problem may be alleviated by using silver arachidate LB films
prepared by exposing an arachidic acid film to a solution containing silver ions [51]. These polypyrrole films have
also been visualized through scanning electron microscopy which essentially indicated the presence of densely
packed spherical granules, a characteristic feature of polypyrrole thin films [47] and characterized by X-ray
photoelectron spectroscopy (XPS) and static secondary ion mass spectroscopy (SIMS) [52, 53]. From the latter two
techniques, evidence was obtained of the presence of polypyrrole within the LB films, namely observation of N(1s)
peak in XPS and peaks corresponding to C-N
-
and (C
2
H
2
N)
-
in SIMS [52, 53]. The thickness per layer increased
with the exposure to HCl gas and also during the exposure to pyrrole vapor, from about 2.5 nm for pristine film to
about 5.8 nm for HCl exposed film and to nearly 6.0 nm for subsequently pyrrole vapor exposed films [55].
These thin polypyrrole films were employed as active material in gas sensor devices [52, 54, 55], where use
was made of changes in the film resistance when exposed to organic vapors [52, 55]. Exposure to alcohol vapors
resulted in a significant change, but only a small change in resistance was noticed upon exposure to vapors of
acetone, ethyl acetate and toluene [52]. An ammonia sensor was prepared with a thin polypyrrole film deposited on
a surface acoustic wave (SAW) device, in which sensing was effected by monitoring the change in the acoustic
wave velocity, measured by the phase shift technique at the output of the SAW device [54]. The film exhibited
excellent selectivity, in the presence of other interfering gases, with a possible detection limit of 18 ppm of ammonia
in air.

5. Amphiphilic polypyrroles [56, 57]
Substituted polypyrroles derived from the corresponding substituted pyrrole were directly spread at the air-
water interface and transferred onto substrates as multilayer LB films [38, 57]. The structure of these polymers is
shown in Fig. 12. The preformed poly-3-decylpyrrole (6) monolayer could be transferred homogeneously in a layer-
by-layer manner, but there was no evidence for the presence of periodic structure within the film in the X-ray
diffraction measurements [38], which was attributed to the coil structure of the
16
N
NH
H
n
O
N
C
10
H
21
n
H
(6)
(7)

Figure-12

polymer. However, evidence was obtained from linear dichroism analysis of the polarized ATR spectral data for the
preferred orientation of polymer chains in LB films of amphiphilic polymer (7) derived from the chemical
polymerization of 5-acetamido-4,5,6,7-tetrahydro-2H-benzo[c]pyrrole monomer. The monolayer characteristics
were studied by surface pressure and surface potential measurements, with the onset of surface pressure occurring at
about 13
2
and the monolayer collapse at approximately 7
2
. Hysteresis was present in surface pressure and
surface potential curves and the maximum surface potential was about 400 mV.

6. Composite LB film approach [58]
The composite LB film approach has been extensively used in the LB manipulation of polyalkylthiophenes,
consisting in co-spreading the polymer with builder materials such as stearic acid/cadmium stearate (for details, see
the next section). Note that there is no chemical interaction between the polymer and the builder materials but the
presence of the latter imparts enhanced stability to the monolayer and also facilitates uniform monolayer transfer.
This composite LB film approach has been adopted in the LB manipulation of soluble parent pyrrole [58]. The area
per molecule of the mixed monolayer (calculated based on the amount of the builder material) increases with
increasing weight percentage of polypyrrole, thus indicating the presence of polymer along with the builder
material. The pressure-area isotherms of polypyrrole and cadmium stearate were featureless, unlike mixed
monolayers of cadmium stearate and polyalkylthiophenes, where a plateau was attributed to the squeezing out of the
polymer from the cadmium stearate monolayer. The maximum surface potential was in between that observed for
pure cadmium stearate (120 mV) and pure polypyrrole monolayer (400 mV). The composite monolayer could be
transferred as Y-type LB films and uniform transfer of polymer during each transfer cycle was inferred through the
linear increase of absorption at 415 nm (due to polypyrrole) with increasing number of layers. As observed in the
pure polypyrrole, a partial dedoping was identified for mixed LB films as well.
17
From the reports described so far, it is clear that the LB manipulation of polypyrrole is far from simple, due
to obvious difficulties in handling the rather non-processible parent polypyrrole. The LB fabrication is usually
accomplished by employing innovative molecular engineering approaches, the most common ones being the use of
metal and fatty acid salt LB templates and the polymerization at the air-water interface, though there are a few
reports on the use of pre-polymerized material itself. LB manipulation of soluble and amphiphilic pyrrole based
copolymers with spacer groups between the pyrrole mers (refer to Fig. 9) would be worth investigating and such
study may provide a better understanding of the material properties in the form of a monolayer and as a multilayer
thin film onto a substrate. Recently, an attempt has been made along this direction. The monolayer characteristics of
the AB type copolymer in which N-hexylpyrrole and thiophene are alternated along the polymer chain have been
studied by surface pressure, surface potential and Brewster angle microscopy [59]. In comparison to earlier reports,
this copolymer is rather highly processible (soluble in most organic solvents) and hence could be easily spread at the
air-water interface. Under optimized conditions, the monolayer was found to be sufficiently stable to be transferred,
without the need to mix with builder materials. However, the observed lower limiting mean molecular area pointed
to the fact that the monolayer is not a true monomolecular layer, instead a multilayer stack, similar to that observed
with most polymeric materials [59].

2.3. LB manipulation of polythiophenes
Parent polythiophene is insoluble in most organic solvents and hence unsuitable to be used in the LB
technique. The most widely studied polythiophene derivatives, poly(3-alkylthiophenes) (8, Fig. 13), on the other
hand, are soluble but still do not form a stable monolayer at the air-water interface on their own. They tend to form
aggregated islands of polymer during monolayer compression, resulting in a lower area per repeat unit (less than 10

2
[60, 61]) in contrast to what was found with the pyrrole counterpart.
S
R
n
R = alkyl
(8)

Figure-13

Attempts to transfer such poorly stable non-monomolecular layer usually led to poor quality LB films [60]. An
exception is represented by a recent report by Bjornholm et al [62] who studied monolayer characteristics of tailor-
made regioregular amphiphilic polythiophenes (9 and 10, Fig. 14).
18
S
S *
*
O
O
O
C
12
H
25
n
(9)
S
S
C
12
H
25
R
R = -COOC4H9; -COOCH(CH3)2;
-COON+(CH3)4;
O
N R =
(10)

Figure-14

The LB processing of polyalkylthiophenes was, in general, therefore often achieved through either one or the
combination of two of the following methods:

1. Low molecular weight oligomers
After a few scattered reports on the LB processing of oligothiophenes [63-65], the first systematic LB study
on oligo- and polythiophenes was performed by Nakahara et al in 1988 [66]. In their work, well-defined oligomers
consisting of different numbers of thiophene mers, namely, ter- (3), quarter- (4), quinque- (5), sexi- (6) septi- (7)
thiophenes (11) were employed (Fig. 15).
S
n
n = 3, 4, 5, 6 and 7
(11)
S S S
CH
2
OCOC
17
H
35
(12)

Figure-15

A stable and transferable monolayer was obtained with -terthiophene esters with alkyl chains (12), and the higher
homologues without any functionalities formed stable mixed monolayers with builder materials such as arachidic
acid. Evidence for the near vertical orientation of the thiophene mers in these monolayer assemblies was obtained
from polarized UV-vis absorption spectral data [66]. Except for this and a few other recent reports on oligomers
which will be commented upon later [67-72], most monolayer and LB studies on polythiophene derivatives were
carried out with high molecular weight polymer, i.e. Mn >10,000 g/mol with a polydispersity higher than 2.
19

2. Mixed LB film approach
Poly(3-alkylthiophenes) do not form a stable monolayer on their own but they do form stable mi xed
monolayers when cospread with builder materials such as fatty acid/fatty acid salts [28, 31, 35, 61, 73-81]. Other
builder materials have also been used, including poly(isobutylmethacrylate) [60], an amphiphilic polydiacetylene,
namely pentacosa-10,12-diynoic acid [82], an unsaturated fatty acid, namely oleic acid [83], a macromolecule,
namely, tetra-tert-butylphthalocyanine [84], an amphiphilic pyrrole derivative, 3-octadecanoylpyrrole [85] and a
mixture of arachidic acid and C
60
[86]. Different parameters studied include the influence of the molecular weight
[77] and the regioregularity of the polymer [78, 79], influence of the length of the alkyl chains [73, 75] and the
mixing ratio of the polymer and the builder materials [74]. These mixed monolayers were mostly studied by surface
pressure-area isotherms and in a few cases with surface potential-area isotherms [77, 83]; no microscopic analysis
on these mixed monolayers is available yet.
It is customary to calculate the area per molecule in surface pressure-area isotherms of mixed monolayers
based on the number of builder molecules in the mixture so that one can ascertain of the influence of the added
polymer. Obviously, one can also directly get the average area per molecule corresponding to both polymer and
builder material employing an average molecular weight. In the latter case, the molecular weight of the repeat unit in
the polymer chain is normally used for the calculation instead of a molecular weight derived from size exclusion
chromatography (SEC) analysis. For mixed monolayers, the surface pressure onset usually occurs at a relatively
large area per molecule and the surface pressure rises monotonically till reaching a characteristic plateau transition
region, at approximately 25 mN/m, in which there is no or little increase of surface pressure. After this plateau
region, the surface pressure rises rather sharply until monolayer collapse, which is seen as the sudden change of
slope in the isotherm. Interestingly, the extrapolated area from the post-plateau steep portion of the isotherm curve
matched well with the area per molecule calculated for the pure builder material (for example, about 20
2
for
cadmium stearate). A typical surface pressure-area isotherm for a mixed monolayer of polyalkylthiophene (hereafter
abbreviated as PAT) and cadmium stearate (5:1 ratio) is shown in Fig. 16.
20
20 30 40 50 60
0
10
20
30
40
50
60
S
u
r
f
a
c
e

p
r
e
s
s
u
r
e

(
m
N
/
m
)
Area per molecule (
2
)

Figure-16

The appearance of the plateau region in the -A isotherm of the mixed monolayer was attributed to the squeezing
out of the polymer from the matrix of the builder material. It is not known as whether the rejected polymer is
stacked on top or underneath the condensed monolayer. Interestingly, no plateau region was seen in isotherms of
mixed monolayers containing the polymer derivatized with a long alkyl chain (for example, poly(3-
octadecylthiophene)), probably due to the strong interaction between the long alkyl chains in the polymer and the
hydrocarbon tails of the builder material. Information on the level of mixing was inferred through the plot of
average area per molecule and mole percentage of the polymer in the mixture, which essentially indicated a phase
separation between the two materials. The extent of phase separation seemed to decrease with increasing length of
the alkyl chain attached to the polymer. Rikukawa [78,79] recently investigated monolayer characteristics of
regioregular head-tail (HT)-PAT mixed with cadmium stearate and observed a better mixing even with short alkyl
chain derivatized polythiophenes, in comparison to the regiorandom PAT counterpart. Such improved molecular
organization was also reflected in the transferred mixed LB films, as evidenced from XRD, UV-vis and conductivity
data discussed below. In the study of the influence of molecular weight of PAT on the organization of mixed
monolayers, Pawlicka et al [77] observed a difference in the collapse behavior, a minimum after the collapse for low
molecular weight polymer (15,300 g/mol) and only a change of slope in the surface pressure isotherm for a high
molecular weight polymer (2,50,000 g/mol). Also, the area per monomer was smaller for the high molecular weight
polymer. However, there was no significant difference in the critical area for the onset of surface potential and the
maximum surface potential (about 350 mV). The analysis of these results is complicated since the measured
collapse pressure was close or higher than the maximum surface pressure (73 mN/m), possibly due to the
displacement of the Wilhelmy plate by the rigid polymer monolayer. Experimental artifacts in the surface pressure
measurement due to the displacement of the Wilhelmy plate by a rigid biopolymer monolayer have been reported by
21
Constantino et al. [87] recently. The mixed LB film approach has also been applied for the fabrication of LB films
containing poly(bromooctylthiophene-co-vinylhexylthiophene) and deuterated stearic acid [88].
The discussion above was mainly related to mixed monolayers in which there is no significant chemical
interaction with the polymer, and the builder molecules only assist the physical ordering of the polymer molecules.
If oleic acid is employed as builder material, on the other hand, then there is significant change in monolayer
properties and in the absorption spectra of the mixed monolayers, which depend upon whether the spreading
solution was exposed to light or not. This is explained by the formation of a charge transfer complex between oleic
acid and PAT [83]. When a topochemically polymerizable amphiphilic diacetylene (see section of polydiacetylene
LB films for further details) was used as builder material, the molecular packing in the monolayer as well as in the
LB films depends on the mixing ratio between the number of PAT repeating units and the number of diacetylene
molecules [82]. An unhindered topochemical polymerization of the diacetylene amphiphile was achieved under UV-
light illumination/thermal treatment for all compositions of the mixed monolayer, which could be considered as
evidence for the formation of a phase separated system. However, the square root of the second harmonic signal
(derived from PAT) was found to increase non-linearly with the area per molecule when the mixing ratio exceeded
one. Based on these observations, Tsumura et al. [82] proposed a double layer structure in which some of the
diacetylene molecules are superimposed on PAT when the ratio was less than one, and aggregation of squeezed out
polymer as separate domains took place when the ratio was increased above one. These mixed monolayers were
amenable to transfer by the conventional vertical dipping method as Y-type films, with the transfer ratio close to
unity up to a large number of layers.

3. Functionalized polythiophenes
The LB manipulation of several tailored, functionalized polythiophenes with no need to mix with builder
materials was reported in [62, 89-92]. The structure of substituted PATs (13-16) is depicted in Fig. 17. It is recalled
that pure polyalkylthiophenes do not form a stable and transferable monolayer on their own, owing to the lack of
balance between the hydrophilic and hydrophobic ends, particularly because of the weak hydrophilicity associated
with the thiophene ring.
22
S
OC
4
H
9
C
4
H
9
O
n S
CH
2
COOC
17
H
35
n
S
CH
2
COOCH
2
CH
2
(CF
2
)
2
CF
3
n
S
(CH
2
)
3
C
6
F
13
n
(13) (14)
(15) (16)

Figure-17

In designing polythiophene derivatives to form a stable monolayer on their own, the main focus was to enhance the
hydrophilicity of the polar end. This has been achieved by attaching alkoxy substituents instead of alkyl substituents.
Callender et al [89] reported the formation of stable and transferable monolayers from poly(dibutoxythiophene),
PDBT (13). At a subphase temperature of 11C, PDBT formed a stable monolayer that could be transferred
uniformly by the vertical dipping method. Low and high molecular weight poly(heptadecyl 3-thiopheneacetates)
(14) were only transferred onto hydrophobic substrates as Z-type LB films [90]. The layered structure of the LB
film, consisting of layers of the polythiophene backbone separated by alkyl chains, was inferred through XRD and
TEM measurements [90]. Stable monolayers produced from polythiophene substituted with partially fluorinated,
long alkyl chain esters (15) could be transferred as Z-type films [91]. AFM measurements pointed to high quality
LB films, but no evidence for the layered structure was observed through XRD measurements. Also, the superlattice
LB films obtained by alternately depositing the polymer and mercury stearate exhibited only the diffraction peaks
corresponding to the latter, and again there was no evidence for the layered structure of the polymer [91]. The
formation of stable monolayers from polythiophenes derivatized with a partially fluorinated long alkyl chain,
poly(1,1,1,2,2,3,3,4,4,5,5,6,6-tridecafluorononylthiophene) (16) was reported by Robitaille et. al [92], indicating that
the highly hydrophobic fluorine atoms influence the formation of stable monolayers of PAT. Detailed in-situ
synchrotron X-ray diffraction and reflection measurements made with Langmuir films of 10 indicated a preferred
orientation of -stacks of polythiophene along the subphase water [62]. Liu et al. [70] found that the terthiophene
substituted with a long alkyl chain at the -position forms a stable and transferable monolayer. There are few reports
on the fabrication of LB films of functionalized monomeric thiophenes [93-96] but they will not be discussed at
length here. Noe et al [93] reported the fabrication of mixed LB films containing 3,4-didecyloxy -2,5-di(4-
nitrophenylazomethine)thiophene and stearic acid and obtained refractive index and thickness values of the mixed
23
LB films through surface plasmon measurements. The preparation of highly organized LB films from 3-
thienylpentadecanoic acid was reported by Schmelzer et al [94-96].

4. Optical, Structural and Electrical Characteristics
The UV-visible absorption characteristics of mixed PAT:builder material LB films resembled those of cast
films made from pure PAT, with an onset wavelength for the -* transition of the conjugated backbone of the
polymer at ca. 650 nm (2.0 eV), indicating that the polymer did not have its chemical composition changed during
LB manipulation. However, the wavelength of maximum absorbance (
max
) for LB films was red-shifted in
comparison to the spectrum of the polymer in a good solvent. For poly(3-octylthiophene),
max
was shifted from 440
nm in chloroform to 500 nm in the mixed LB film [61].
max
of the LB film varied between 440 and 550 nm
depending on the length of the alkyl chains attached to the polymer and for a given polymer, the absorbance at
max

increased linearly with the number of layers transferred, indicating the transfer of near equal amount of polymer
during each transfer step. Such uniform transfer of the polymer was also inferred through the plot of reciprocal
capacitance (1/C) versus number of layers, which again exhibited a linear relationship. For a given number of layers,
the maximum absorbance increased with increasing polymer content in the mixture and also with the deposition
surface pressure. Specific to the latter, the absorbance was about 1.4 times higher for films deposited in the post-
plateau surface pressure regime, in comparison to those deposited in the pre-plateau region. This observation was
corroborated by thickness measurements in which a higher thickness per layer was observed for films fabricated at
higher surface pressures. In the study of thermochromic effects in the mixed LB film of PAT with arachidic acid,
Ahlskog et al [80] observed a partially reversible blue shift of the absorption maximum of PAT with increasing
temperatures and the temperature at which such shift occurred was lower for the LB films than in spin-cast films.
The X-ray diffraction pattern for these mixed LB films [75] exhibited a set of intense diffraction peaks with a
corresponding bilayer distance similar to that observed for LB films of the pure builder materials, though the
diffraction peaks for the mixed LB films were noticeably broader. Interestingly, the optical absorbance and the
thickness depended on the deposition surface pressure, which could imply a bilayer structure in the LB film, but no
such dependence appeared in XRD patterns of LB fil ms produced at different surface pressures. The most probable
model for the structure of these mixed LB films is therefore one in which the polymer domains are randomly but
uniformly dispersed within a highly ordered matrix of the builder material, as depicted in Fig. 18.
24

Figure-18

Such model was also supported by analysis of near-edge X-ray absorption fine structure (NEXAFS) results, in
which a very strong polarization dependence of (C-H)* peak (alkyl chains) was observed in the carbon K-edge
spectra, while there was no dependence of polarization in the sulfur L edge spectra (thiophene) [31]. A layered
packing of polymer could nevertheless be observed from XRD measurements on mixed LB films from regioregular
HT-PAT and cadmium stearate. The observed bilayer distance of about 17
2
matched well with the expected
interlayer distance between two polymer chains. This result points to the importance of the polymer structure in its
organization in a layered architecture.
As-deposited mixed LB films are insulating, but they can be rendered conductive upon doping with
oxidizing agents such as nitrosyl hexafluorophosphate (NOPF
6
). This can be estimated from the decrease in the -*
absorption with a concomitant appearance of peaks in the long wavelength region (at about 800 and 2200 nm)
indicative of the creation of bipolaron defect states along the polymer chain. The in-plane d.c. conductivity of doped
films varied with the amount of polymer in the mixture, within the range from 10
-2
to 10 Scm
-1
. In comparison to
these values for the regiorandom polymer, an enhancement of conductivity to 50-100 Scm
-1
was observed for mixed
LB films of the regioregular HT-PAT [78]. The dependence of the conductivity of doped and undoped PAT-
containing mixed LB films on the temperature, applied electric field and the applied magnetic field has also been
investigated [80, 85]. The conductivity decreased with increasing temperatures [80]. The electric field dependence
25
for the conductivity in the doped films could be fitted with the theory of charging-energy-limited tunneling between
highly conducting regions [85]. For undoped LB films, the magnitude of the positive magnetoresistance increased
with decreasing temperatures [85]. The doped LB films exhibited a high and bias-dependent Hall mobility, of 10
3

cm
2
/Vs below 40 K [85].
Electroluminescence (EL) characteristics of mixed LB films of poly-/oligo- thiophenes and arachidic acid
in a LED structure have been studied by Pal et al [67, 68]. The LED device comprised a bottom ITO electrode, a
middle layer of mixed LB film and a top electrode made either from aluminum or calcium. Evidence for aggregation
of quinquethiophene was derived from the optical absorption and photoluminescence (PL) spectra. A greenish
electroluminescence was observed for devices with quinquethiophene mixed LB films, with quantum efficiencies of
the order of 10
-3
%, which is one order of magnitude lower than for the device containing PAT mixed LB films. For
both PAT and quinquethiophene mixed LB films, the electroluminescence and the photoluminescence spectra
exhibited an identical profile, which confirms that the device emitting layer is indeed the PAT/quinquethiophene LB
layer. Interestingly, even a 5-layer LB film was found to exhibit the same luminance as thicker LB films. The lower
turn-on current with the calcium electrode in comparison to aluminum electrodes has been related to the lower work
function of calcium, which permits the required electron injection at a lower diode current [67]. From the transient
electroluminescence measurements made with quinquethiophene mixed LB films with different thicknesses and
applied fields, the observed response time of the EL emission (the time lag between the voltage pulse and the first
appearance of the EL emission) has been related to the charge carrier transit time which enables one to calculate the
drift mobility through the film [68]. The results varied with the thickness; for an anisotropic mobility was observed
in thicker films, while the interface between the LB film and the electrode played an important role in the case of
thinner films [68]. The role of the interface has also been studied by introducing insulating LB layers of
polymethylmethacrylate (PMMA) between the electrode and the emitting LB layer. Such study indicated that the
time delay observed in the transient EL measurements made with LEDs without insulating LB layers is mainly due
to accumulation of charge carriers at the interface and not due to the carrier transient time [97]. The response time
for the electric field redistribution due to injected and trapped charged carriers in LED devices containing mixed
PAT-AA LB films as active layer, as inferred through charge collection measurements with the time -of-flight
technique, was of the order of 5-200 s [98]. This value agrees closely with the delay time observed in time-
resolved EL measurements made with similar LB films [98].
The fabrication of a LED device with a PAT mixed LB film sandwiched between EB polyaniline (EB
PANi) as active layer (ITO/EB-PANi/PAT-AA/EB-PANi/Al) has also been reported [99, 100]. The high-frequency
limit of a LED, in general, depends on the field across the device and also on the thickness of the active layer, apart
from the carrier mobility. In ITO/EB-PANi/PAT-AA/EB-PANi/Al LED device, it has been shown that the high-
26
frequency operating limit could be moderately increased by reducing the thickness of the active layer (by depositing
a lesser number of layers) [99]. Even an LB film with only two layers of emitting material was sufficient for light
emission [100]. In another systematic study, the possibility of fine-tuning the quantum efficiency was investigated
[101], either by the use of hole- and electron- transporting LB layers together with the active emitting LB layer or by
using LB films of a blend of the hole- and electron transporting materials with the emitting material. The quantum
efficiency of the LED device varied with the structure employed, where a multilayer containing a separate layer of
hole- or electron- transporting material seemed to provide a better control of the charge carrier injection than the
blend structure. Also observed was a decrease in quantum efficiency when an EB-PANi layer was interposed
between PAT-AA mixed LB films and Al, or when poly(vinylcarbazole) (PVK) was placed between ITO and PAT-
AA mixed LB films [101]. In heterostructured LED devices, a single LB mixed monolayer of quinquethiophene as
active emitting layer displayed a quantum efficiency comparable to that of thicker films, if an additional electron
transporting LB layer of 2-(4-biphenyl)-5-(4-tert-butylphenyl)-1,3,4-oxadiazole (PBD) was incorporated [69].
Bolognesi et al [102] observed a strong thermochromic effect in photoluminescence (PL) of spin cast films of
poly(3-decylmethoxythiophene), PDMT, while the PL from Z-type LB films of PDMT exhibited negligible
temperature dependence [102]. They also observed a blue shift of EL with respect to PL for the LB film of PDMT,
which was attributed to partial polymer degradation [102].
Elshocht et al [103] found that the second harmonic generation from LB films transferred using the
horizontal lifting method of PAT with the chiral side chain (poly{3-[2-((S)-2-methylbutoxy)ethyl]thiophene}) (17,
Fig. 19) depended on the chirality of the polymer and cannot be simply correlated to the electric dipoles.
S
O
CH
3
n
(17)

Figure-19

The third order non-linearity of these LB films measured around the excitonic resonance using the four wave mixing
method was also obtained and the measured
3
was used to infer the effective conjugation length of the polymer
[104].

Figure-19

27
2.4. LB manipulation of polyanilines
Earlier developments in the LB manipulation of polyanilines can be found in refs [105, 106] and hence
only a short description of such research on monolayers and LB films of polyaniline is provided here, which will be
followed by discussion on more recent developments. In comparison to polypyrrole and polythiophene, a
significantly larger number of monolayer studies have been carried out with parent polyaniline (in partial oxidation
state, namely emeraldine base EB form). The poor processibility associated with parent polyaniline has been
overcome by the use of either highly polar solvents such as N-methylpyrrolidone [107-111] or functionalized acids
such as camphor sulfonic acid (CSA) or dedecylbenzene sulfonic acid (DBSA) [112-116]. LB manipulation of
polyaniline has also been accomplished by (a) the polymerization of monomeric long chain aniline derivatives (18,
Fig. 20) at the air-water interface and subsequent transfer as multilayer LB films [117-121]; (b) use of substituted
polyanilines [122-130] and (c) use of oligo anilines instead of high molecular weight polyanilines [131-133].
NH
2
R
NH
2
R
NH
2
R
(18)
R = -(CH
2
)
15
CH
3

Figure-20

The structure of the parent polyaniline and substituted polyanilines (19-24) employed in LB studies are
provided in Fig. 21. Also shown in the figure are chemical structures of processing aids employed in solubilizing
polyaniline in common organic solvents such as chloroform. It should be noted that the substitution of polyaniline
with smaller groups such as methoxy and ethoxy (20, 21) (not necessary with longer alkyl chains) has rendered
polyaniline soluble in most common organic solvents.
28
CH
3
C H
3
CH
2
SO
3
H
O
SO
3
H C
12
H
25
CH
3
OH
DBSA
CSA
m-cresol
Processing aids:
N
H
N
H
N N
y x
(x = 1-y)
N
H
N
H
N N
y x
R R R R
(x = 1-y)
R = -OCH
3
(20) -OC
2
H
5
(21), -OC
10
H
21
(22), -OC
18
H
37
(23)
(19)

Figure-21

A recent advance in the LB manipulation of polyanilines, namely the enhancement of monolayer stability with the
acidic subphases (pH ~2.0), is proven effective for obtaining reproducible monolayer characteristics [113-116, 128,
129]. Similar to polythiophenes, the mixed LB film approach, apart from leading to improved monolayer stability, is
helpful in the transfer of the poly/oligo anilines/substituted aniline monolayers onto substrates [131-138].

1. Surface Pressure-Area Isotherms of Parent/Substituted Polyanilines
In contrast to simple amphiphilic molecules such as fatty acids, the mean molecular area calculation for
parent polyaniline monolayer is complicated due to the following reasons: (a) Difficulties in determining the exact
molecular weight and its distribution; (b) The concentration of the polymer in the spreading solution may not be
known to any high accuracy, since in most cases the solution must be filtered for removing undissolved particles,
which possibly arise from high molecular weight or some cross-linked polymer. For the sake of comparison and
simplicity, it is customary to use the weight of the polymer repeat unit (similar to other conducting polymers) in the
29
mean molecular area calculations, though there are few reports which considered the actual molecular weight of the
polymer as obtained from gel permeation chromatography (GPC) [125, 139]. In the specific case of polyanilines,
researchers have used different repeat units (one to four aniline units), and thereby obtained different areas per
molecule. Readers should be aware of that when analyzing data from different research groups.
Isotherms of parent polyaniline show no clear phase transition, unlike those of model fatty acids, which
usually exhibit distinct phase transitions such as 2D gas-liquid, 2D liquid-solid and 2D solid-bilayer/3D solid. As
Fig. 22 shows, there are only changes in compressibility at different parts of the isotherm. Collapse is usually
denoted by a change
2 4 6 8 10 12
0
10
20
30
S
u
r
f
a
c
e

p
r
e
s
s
u
r
e

(
m
N
/
m
)
Mean molecular area (
2
)

Figure-22
in the slope with which the pressure rises in the condensed part of the isotherm. The collapse pressures for a
polyaniline monolayer are, in general, lower (in the order of 20-40 mN/m depending upon the experimental
conditions) than for typical amphiphilic molecules. Such differences to typical film-forming materials are not
unexpected considering its non-amphiphilic and polymeric nature. Furthermore, the mean molecular area, calculated
on the basis of the repeat unit, indicates that polyaniline does not form a true monomolecular layer, but the way
polymer chains are packed in the monolayer structure is not yet understood. There are discrepancies in the literature
as to whether any preferential direction in the plane should exist, which can only be resolved with further detailed
molecular modeling studies.
With regard to stability, polyaniline monolayers are less stable than typical amphiphilic materials, but
stabilization may be achieved after a short period of time under specific experimental conditions, which then allows
transfer onto a substrate. Monolayer stability is improved when doped (emeraldine salt) instead of undoped
(emeraldine base) polyaniline is employed or if the monolayer is spread on acidic subphases (pH ~2). As a typical
example, the doping of polyaniline with HCl is shown in Scheme-3
30
N
H
N
H
N
H
N
H
* *
x
. + . +
Cl
- Cl
-
N
H
N
H
N N
* *
y x
(x = 1-y)
+
HCl

Scheme3. Acid doping of polyaniline


When the spreading solution containing doped polyaniline (e.g. with CSA or DBSA) is placed on a neutral water
surface (pH~6.0), there will be an instantaneous dedoping. Hence, the functionalized acids in these cases only
enhance the solubility of polyaniline in the spreading solution. However, upon acidic subphases, polyaniline
becomes doped as soon as it is placed on the subphase, regardless of its previous doping state. Reports have
indicated that the type of acid used in pre-doping polyaniline or in the subphase is important for the monolayer
properties. Upon spreading the monolayer, the counter ions in the doped polyaniline solution are readily exchanged
with those in the subphase. Interestingly, it has been noticed that for poly(o-ethoxyaniline) the doping state affects
the monolayer characteristics [140], even when the same acid is used for both solubilization and in the subphase.
Apparently there is a kind of memory effect due to which the polymers are packed in a different way. Such
memory effect was also observed when the spreading solution contained m-cresol and CSA [141]. In a systematic
study on the influence of functionalized acids (like CSA, DBSA, trifluoroacetic acid (TFA)) to pre-dope polyaniline
[114], even the order of addition of either the functionalized acid or m-cresol to polyaniline seems to influence its
doping level in the spreading solution.
With m-cresol in the spreading solution, one expects both enhanced solubility and the so-called secondary
doping effect which has been suggested to explain the enhanced conductivity of polyaniline upon exposure to m-
cresol vapor [142]. But, as far as the LB experiments are concerned, one would expect an immediate dissolution of
m-cresol into the bulk water subphase, following spreading. In fact, FITR results of transferred LB films support this
hypothesis, since no traces of m-cresol were found in the LB films. However, the presence of m-cresol in the
spreading solution had a clear influence on the monolayer characteristics [141]. A significant increase in area per
31
molecule and critical area (refer to next section on surface potential for details) was noticed when m-cresol was
introduced in the spreading solution, in addition to CSA (Fig. 23 and b). This memory effect is consistent with
secondary doping in which the conformation of polyaniline chains changes from the compact-coil to expanded coil
structure. The monolayer experiments clearly revealed that such expanded structure persists even after the sudden
and complete removal of CSA and m-cresol from the polymer [141].
18 21 24 27 30
0.00
0.05
0.10
0.15
0.20
(b)
(a)
S
u
r
f
a
c
e

P
o
t
e
n
t
i
a
l

(
V
)
Mean molecular area (
2
)
0 5 10 15 20 25 30 35
0
10
20
30
40
(b)
(a)
S
u
r
f
a
c
e

P
r
e
s
s
u
r
e

(
m
N
/
m
)
Mean Molecular Area (
2
)

Figure-23

The influence of experimental parameters such as subphase temperature, subphase pH and compression speed on the
nature of the isotherms has been reported in several instances. Upon changing the subphase pH by introducing acids
containing different counter ions such as Cl
-
, S0
4
-
, CSA
-
(camphor sulfonate) and in the presence of neutral salts, the
area increases in the order above [115], which is understood to reflect the size of the counter ions.
An increase in area per molecule, accompanied by a decrease in compressibility, was observed in ref.
[110], but in another study the area was not affected by changing the subphase temperature [109]. These differences
in behavior, which includes the measurement of different mean molecular areas for a given temperature, might
originate from the use of different spreading solvents, since in ref. [110] only NMP was used as the spreading
solvent while a mixture of NMP and chloroform was employed in ref. [109]. In a systematic study, monolayers of
polyaniline processed by different methods with varying levels of doping were investigated as a function of
subphase temperature, pH and ionic strength of the subphase and spreading volume [143]. Surface pressure-area
isotherms of polyaniline processed with different processing aids on the surface of pure water at a subphase
temperature of 20 C are shown in Fig. 24.
32
0 5 10 15 20
0
10
20
30
40
50
60
PANi/NMP
PANi+AcOH/NMP
PANi+CSA/CHCl
3
PANi+CSA+
m-cresol/CHCl
3
S
u
r
f
a
c
e

p
r
e
s
s
u
r
e

(
m
N
/
m
)
Area per monomer (
2
)

Figure-24

Mello et al [144] investigated the influence of subphase pH and type of doping acid on monolayer
properties of poly(o-ethoxyaniline) (21, Fig. 21). A decrease in subphase pH resulted in a decrease in the mean
molecular area for poly(o-ethoxyaniline) [128]. Sworakowski et al [145] observed an increase of both area per
molecule and collapse pressure of poly(o-methoxyaniline) (20, Fig. 21) monolayers when the compression speed
was increased.
Readers may find a detailed account on the behavior of polyaniline monolayers, as observed by different
research groups, in ref. [106]. Discrepancies even for parent polyaniline obtained under similar conditions may be
justifiable since polyaniline does not form a true monomolecular film at the air-water interface. Therefore the chains
are folded and entangled to an extent that depends on the polyaniline structure which in turn depends on the
synthetic procedures employed.

2. Surface Potential-Area Isotherms of Parent/Substituted Polyanilines
Surface potential isotherms of polyaniline monolayers have been investigated and related to measured
surface potentials of polyaniline LB films deposited on metal substrates [146]. Unlike the case of simple fatty
acid/salt monolayers, a direct quantitative analysis of surface potentials for polymeric monolayers cannot be made.
The surface potential isotherm can still be employed to infer about the coming together of polymeric domains at
large mean molecular areas, long before the onset in surface pressure, since a non-zero surface potential is only
detected if large islands were formed from the start. For instance, for the polyaniline monolayer spread on acidic
subphase, an almost zero surface potential is observed at very large areas, indicating the poor interaction between
33
polymer chains/domains at this stage. However, at the critical area, a sharp rise in surface potential could be
observed until reaching a maximum value that varies between 350-400 mV. The lack of major changes in V after
the maximum is reached indicates the absence of any major structural rearrangement in the monolayer. In a study on
the influence of protonation of both polyaniline and poly(o-ethoxyaniline) on the measured surface potentials, Mello
et al. [147] observed an increase in surface potential of polyaniline with the change of acid concentration in the
subphase and such behavior could be explained in terms of the double layer potential. Interestingly, despite the
complex nature of the polymeric monolayer, the Gouy-Chapman (GC) theory, which is normally employed with
simple fatty acid molecules, could be used to estimate the changes in the double layer potential. For poly(o-
ethoxyaniline) monolayers, on the other hand, changes in the double layer potential did not suffice to explain the
surface potential dependency with the acid concentration, especially when the subphase contained trifluoroacetic
acid [147] The interaction of the hydrophobic counter ions (trifluoroacetate ions) with the polymer is probably the
main cause for the failure, since the intrinsic dipole moments of the polymer chains were affected.
For pure polyaniline monolayer, both the critical area and the maximum surface potential (V
max
) were
higher when the spreading solution contained m-cresol, in addition to CSA [141]. A near zero surface potential was
observed only at a very large molecular area when the spreading solution contained m-cresol. The V
max
values
observed in the presence and in the absence of m-cresol in the spreading solution were about 130 and 170 mV,
respectively (Fig. 23b). As mentioned before, such changes were correlated to the secondary doping effect caused
by m-cresol, even though m-cresol seems to dissolve in the aqueous subphase.

3. Brewster Angle Microscopy (BAM) of polyanilines
There are a few reports on Brewster angle microscopy (BAM) of polyaniline monolayers. BAM results on
poly(o-ethoxyaniline) monolayers on pure water and on aqueous acidic solutions were used to optimize the
deposition pressure range [128]. On pure water, the smaller domains formed immediately after spreading come
together and form large domains, even at areas per molecule where the surface pressure was still zero. Such an
observation, i.e., interaction of domains of the polymer at large area per molecule, is similar to that observed in
surface potential-area isotherms discussed above. Upon compression, the molecular density increases until forming
a rather uniform film with less defects. This film breaks and tends to form aggregates near the collapse pressure and
larger polymer aggregates visible to the naked eye upon monolayer collapse. Decompression does not bring about
the 3D-2D transformation, indicating the irreversibility of the compression-decompression cycle. On the other hand,
quite different BAM images were obtained for poly(o-ethoxyaniline) monolayers spread on acidic subphases,
especially at the initial and final stages of compression [128]. BAM analysis of the composite monolayers of
polyaniline and cadmium stearate [148] has also been reported. In the compressed state (at ca. 25 mN/m), there were
34
two distinct types of domain which could be of cadmium stearate and polyaniline. The phase separation is consistent
with results from -A isotherms and X-ray diffraction of mixed LB films, which indicate that polyaniline and
cadmium stearate are present in separate domains. The formation of similar phase separated domains, whose
morphological features seemed to depend on monolayer spreading and also on the speed with which the monolayer
is compressed, was inferred for a mixed monolayer consisting of poly(o-pentadecylaniline) and a ferroelectric liquid
crystal termed as 10PPB2, using the tapping mode of atomic force microscopy (AFM) of a transferred monolayer
onto a mica substrate [120].

4. Polymerization at the air-water interface
Taking advantage of the control of intermolecular distance offered by the LB technique, Duran et al. has
carried out extensive studies on the polymerization of substituted aniline monomers (18, Fig. 20) at the air-water
interface in order to obtain anisotropic LB films of high molecular weight polyanilines [117-121]. The
polymerization of pentadecyl anilines was performed on the surface of acidic subphases after holding the
monomeric monolayer at a fixed, high surface pressure and the course of the reaction was usually monitored by the
decrease in mean molecular area with time [117,118]. The substitution pattern (whether o-, m- or p-) of alkylaniline
seemed to have profound influence on the rate and degree of polymerization [121]. The polymerized monolayer
could be uniformly transferred as evidenced from scanning tunneling microscopy (STM) [119]. Spontaneous
oligomerization of an amphiphilic aniline derivative, namely o-octadecyloxyaniline organized in a monolayer by
oxygen contained in air and/or water, was reported by Sagisaka et al [149]. While the rate at the beginning of such
oligomerization was estimated to be 0.003 min
-1
, conversion reached 80% after 3 h under constant molecular area
(455
2
mol
-1
). The maximum chain length of the oligomer was 5 monomer units. The monolayer oligomerization
occurred even on the strong basic aqueous subphase (pH 12) with re latively high conversion (50-80%) [149].
Ultrathin polymeric LB films could also be produced from parent aniline and 4-hexadecylaniline monomeric
monolayers at 20C using polymerization mediated by the enzyme horseradish peroxidase (HRP) and the
subsequent transfer onto solid substrates [150].

6. The mixed monolayer approach
The composite monolayer approach in which the polymeric material is cospread with builder materials has
been adopted with a two-fold objective, namely to improve monolayer stability and transferability [134-138] and
also to fabricate functional composite LB films comprising polyaniline and functional guest materials like glucose
oxidase (a biosensor) [151], calixarene (a gas sensor) [152] and crown-ether (an ion sensor) [153]. Note that the
builder/functional materials are not linked to the polyaniline covalently but just immobilized, and they are phase
35
separated. There is also a report on sulfonated polyaniline complexed with stearylamine to form a polyion complex
(24, Fig. 21) [154,155]. In contrast to the non-covalent mixtures mentioned earlier, in this case the components are
held together chemically and they are miscible at the molecular level. As discussed in sections related to mixed
monolayers of either polythiophenes or polypyrroles, information on phase separation can be obtained by plotting
the change in area per molecule with the relative contents of the components in the mixture. While a simple,
additive behavior with the area varying linearly with the relative contents indicates a phase separated mixture, any
deviation from linearity can be related to the molecular interaction between the components. For the non-covalently
mixed monolayers - e.g. polyaniline and stearic acid, polyaniline-cadmium stearate and 16-mer polyaniline-
cadmium stearate - a linear relationship has been observed (Fig. 25) while the polyion complex monolayer
(sulfonated polyaniline and stearylamine) exhibited a negative deviation.
0 30 60 90 120
0
5
10
15
20
A
r
e
a

p
e
r

m
o
l
e
c
u
l
e

(

2
)
Weight % of 16-mer

Figure-25

V-A isotherms of composite monolayers containing polyaniline/16mer polyaniline and cadmium stearate
have also been reported [131, 138]. Analogously to either pure cadmium stearate or pure polyaniline monolayers on
acidic subphases, a near zero surface potential was measured at very large areas per molecule, provided that the
monolayers contained a low fraction of the polymer/oligomer. However, if the relative contents of the
polymer/oligomer was increased, an initial non-zero surface potential would be observed, indicating the formation
of large domains at large areas. Also, the onset of surface potential at the critical area would be increasingly less
steep as the fraction of the polymer/oligomer in the mixture increased. Despite the dispersion in the maximum
surface potential values, pointing to the sensitivity of the measurement, a decreasing trend was observed with
increasing polymer contents in the mixture. The only noticeable difference in V-A isotherms of mixed monolayers
of high molecular weight polyaniline and of the oligomeric (16-mer) counterpart was the increase in the maximum
36
surface potential with the amount of 16-mer in the mixture [131, 138], which perhaps underlines the dependence of
monolayer organization on the polymer chain length. Similar to that observed with pure polyaniline monolayer, both
the critical area and the maximum surface potential of the mixed monolayer of polyaniline and cadmium stearate
were higher when the spreading solution contained m-cresol, in addition to CSA [141].
Hua et al [156] reported a hydrogen-bond driven mi xed monolayer formation of stearic acid and aniline
monomer. When the monolayer was compressed to a surface pressure of 50 mN/m, the hydrogen bonds seemed to
be rearranged in such way that the molecule changed from a horizontal into a vertical orientation, which led to a
packing with higher thickness. Such mixed monolayers could be transferred onto conducting substrates and
polymerized.

6. Fabrication and Optical and Structural Characteristics
There are varying reports on the transfer of polyaniline monolayers. The most common trend seems to be
as follows: pure polyaniline monolayers are transferred mostly as Z-type (transfer only during upstrokes), and a
perfect Y-type deposition is observed while transferring mixed monolayers containing polyaniline and builder
materials. As mentioned before, using acidic subphases improved transferability significantly, apart from improving
the monolayer stability. Paul et al [157] reported the successful transfer of both undoped and protonated poly(o-
methoxyaniline) as multilayer Langmuir-Blodgett films. In both cases, scanning electron micrographs showed a
distinct linear array of polymer grains [157]. Z-type deposited LB films of polyaniline, as a function of number of
layers, deposition pressure and dipping speed, were investigated with scanning force microscopy [158]. A single
layer of polyaniline, obtained with a dipping speed of 1 mm/min. and deposition pressure of about 10 mN/m,
exhibited a platelet-like structure. Transfer at lower surface pressures resulted in incomplete layers, while much
higher surface pressures resulted in clumps of polyaniline bundles in the finished films. While clumped bundles are
seen in chemical cast and evaporated films, flat platelet features with much greater smoothness and structural
homogeneity were observed for LB films [158]. A fibrillar-like structure, with the fibril width ranging from about
60 to 160 nm, was observed for Z-type deposited polyaniline [114].
Preliminary information on the macroscopic film uniformity of transferred LB films can be obtained from
surface potential measurements on films transferred onto conducting substrates [131,138]. For mixed LB films with
cadmium stearate, the surface potential increased with the amount of polymer and was always positive, regardless of
whether the LB film contained an odd or even number of layers, unlike the LB films of pure cadmium stearate that
displayed positive and negative potentials for odd and even numbers of layers, respectively. This difference in
behavior has been explained by stabilization of packing of cadmium stearate molecules/domains by the surrounding
polyaniline matrix [146].
37
To date, no direct evidence has been obtained from X-ray diffraction measurements for the multilayer
structure of polyaniline LB films. Ram et al [129] observed Kiessing fringes in the 2 = 1-2 region in the XRD
pattern of Langmuir-Schaefer films of poly(o-methoxyaniline) [129]. The clear diffraction peaks for the mixed LB
films of polyaniline and builder materials such as cadmium stearate arise from domains of builder materials. The
stacking order of the builder material seems to depend on the polyaniline content in the mixed films, the order of
which decreases upon increasing the polymer content [132, 138]. UV-vis results indicated that even though
polyaniline was in the doped state in the spreading solution (with the use of functionalized acids), films transferred
from pure water surface were in the emeraldine base state. The films were partially doped with the employment of
acidic subphase, as evidenced from the absorption spectrum. The UV-visible spectral changes for the polyaniline LB
films obtained with m-cresol in the spreading solution depended on the doping state of the monolayer. Such effects
probably originate from changes in molecular conformation since m-cresol is not retained in the transferred film.
Despite the changes in the absorption spectra and in the molecular characteristics as discussed above, the processing
with m-cresol brought no change to the electrical conductivity nor to the stacking order of cadmium stearate
domains in the transferred LB films, as evidenced from XRD studies [141]. Polyaniline LB films obtained with m-
cresol in the spreading solution have also been investigated by Raman spectroscopy [159]. The average thickness
per layer reported for the polyaniline LB films is in the range of 20-25
2
[114, 129]. For the mixed LB film, the
thickness obtained from spectroscopic ellipsometry depended on the film composition, being mostly determined by
the cadmium stearate domains up to 50 wt. % of the polymer [132].

7. Electrical and Electrochemical Properties
Undoped polyaniline LB films were insulating as expected, whereas the reported conductivity of doped
polyaniline LB films varied from 10
-3
to 1 S/cm. Most doping s tudies were performed by treating the polyaniline LB
film with mineral acids such as HCl, but doping with iodine and X-ray irradiation (see next subsection) has also
been reported. The conductivity measured in-plane and normal to the surface of the doped polyaniline LB film
differed by several orders of magnitude (10
3
10
7
), indicating the formation of highly anisotropic films. Similar
conductivity values are observed for mixed LB films containing polyaniline and builder materials which seemed to
suggest that the mixing with insulating materials did not lower the macroscopic conductivity of the polyaniline
[136]. However, Dimitriev et al [160] observed a suppression of conductivity of polyaniline in the presence of the
amphiphilic C
15
-TCNQ (pentadecyltetracyanoquinodimethane) in a mixed LB film. Attempts to measure the
conductivity of a single monolayer have failed, since no ohmic behavior was observed during the measurement. The
thickness of the monolayer (about 24 ) is so small in comparison to the thickness of the electrode (40 nm), that led
to a breakdown in the film. The conductivity of a 40-layer film was ca. 0.1 S/cm, which was attributed to the
38
increase in charge defects - polarons/bipolarons - with the increase in number of layers [129]. The temperature
dependence of conduction in polyaniline LB films has been related to one-dimensional variable range hopping
[157]. The charge transport in LB films of polyaniline has been studied as a function of film thickness, temperature
and the electric field, and the results could be understood on the basis of theoretical models for variable-range
hopping in a parabolic quasi-gap [161]. In the comparison among thin films made by spin-cast, evaporation and LB
techniques, Agbor et al. [162] observed differences in in-plane conductivity, which were attributed to differences in
the layer morphology and chemical structure. Paloheimo et al [163] reported lower conductivity for doped
oligoaniline LB films than for high molecular weight polyaniline LB films.
Cyclic voltammograms (CV) have been used to investigate the electrochemical characteristics of
polyaniline LB films [108, 110, 113, 164]. The redox features are similar to those of thick polyaniline films, despite
the fact that LB films are very thin and compact. Furthermore, the currents at the anodic and cathodic limits are
significantly higher and the redox peaks are relatively broader for polyaniline LB films than for electrochemically
deposited or chemically cast films [108, 110, 113]. Broadening of redox peaks is predicted for an increasing number
of layers in ref. [108], but no appreciable change in the nature of CV upon increasing the number of layers was
found by Dhanabalan et al. [110]. Nevertheless, in both cases, a linear increase of peak current with sweep rate has
been observed, which could be related to surface confined species. The substrate surface onto which the polyaniline
LB film is deposited affects the nature of the redox peaks [110]. Films transferred onto platinum and gold substrates
displayed broader redox peaks, but films on ITO exhibited sharp redox peaks, analogously to electrochemically
deposited films [110]. For polyaniline LB films deposited at different surface pressures, the redox charge capacity,
q
redox
, of the polyaniline LB film varied with the deposition pressure. q
redox
was practically constant for deposition
pressures between 13 and 15 mN/m (in the condensed part of the isotherm), but increased for films transferred at
lower surface pressures (e.g. 8 mN/m), being minimum for films transferred near the collapse point (30 mN/m).
These changes were attributed to differences in the packing density of the LB films transferred at different surface
pressures [110]. The electroactivity of LB films of poly(o-alkoxyaniline) was compared with that from drop coated
and electrochemically deposited films [165], and no significant difference in the redox characteristics could be
noticed between the films produced by the different methods. The final potential, film thickness, sweep rate and
electrolyte solution influenced the electroactivity of poly(o-alkoxyanilines) LB films [166]. Also noticed was that
the LB film took a shorter period of time for stabilization - as far as the CV response is concerned - in comparison to
thicker cast films [166]. The redox features of composite LB films containing polyaniline and cadmium arachidate
are also similar to those of pure polyaniline LB films [136], but LB films of polyaniline with different chemical
structural repeat units (obtained through a controlled chemical synthesis) have displayed different electrochemical
responses [137]. The electrochemistry of polyaniline LB films has also been studied using cyclic voltammetry
39
coupled with a quartz crystal microbalance (QCM) [167]. Cyclic voltammograms were obtained for films
transferred at various surface pressures and also for films containing different numbers of layers and the results were
compared with those of electrochemically deposited polyaniline films. Such comparison was used to infer the
packing behavior of polymer chains in the LB film structure. The total redox charges calculated from the
voltammogram increased linearly with increasing number of layers, as did the mass associated with anion inclusion
during the anodic scan. The multilayer LB films are electroactive but the kinetics of counter ion transport in these
films is slower than that for electrochemically deposited films [167]. Goldenberg et al. [164] observed clear
electrochromic transitions for a single monolayer polyaniline LB film on ITO, but only a poor electrochromic
switching behavior was observed for the multilayer LB films of polyaniline [167]. These differences have been
attributed to the close packed structure of polymer chains in the LB film which seemed to restrict the conformational
changes and the ion transport through the film [167]. Thin multilayer LB films of 21 have also exhibited reversible
electrochromism similar to polyaniline LB films [165]. The response of polyaniline in terms of alteration in
conformation and/or electrical characteristics to changes in the surroundings was investigated in mixed monolayers
containing polyaniline and a crown ether. For example, the conformational change of guest crown ethers, which are
embedded in the polyaniline matrix after incorporation of host ions, was inferred by the change in mean molecular
area. In addition, changes in the crown-ether also appear to cause conformational changes in polyaniline, which then
affect the thin film electrical characteristics [151].

8. Post-Deposition Treatment
Annealing of LB films has reportedly brought significant changes in structural, optical and electrical
characteristics of both EB-polyaniline and poly(alkoxyaniline) LB films [109, 168]. The decrease in thickness (24-
18.1 ) of each monolayer of 20 has been observed for annealing at 150. The conductivity of LB films varied from
0.4 to 10
-8
S cm
-1
as a function of annealing temperature. A continuous shift of UV absorption band from 840 to 590
nm with annealing temperature has been related to the simultaneous change of charge defects in the LB films, as the
differential scanning calorimetry (DSC) studies show a complete evaporation of dopant ions occurring at 180 C in
the powder form of 20. Complete removal of dopant ions for LB films have been observed for annealing at 150.
The Cole -Cole plot obtained at 80 for the LB film of 20 shows marked deviation from a semicircle, which indicates
a complex transport mechanism in the LB films [168].
A systematic study on the X-ray doping has been made [169, 170], in which composite LB films from
polyaniline and cadmium stearate have been irradiated with various exposure times. Upon X-ray irradiation, the
absorption maximum shifted from 600 nm to 800 nm with the LB film color changing from blue to green, similarly
to t hat observed during the acid doping of polyaniline [169] (Fig. 26).
40
400 600 800 1000
0,0
0,2
0,4
0,6
10
9
8
7
6
5
4
3
2
1
A
b
s
o
r
b
a
n
c
e
Wavelength (nm)

Figure. 26

The exposure time period needed to observe the shift of absorption increased linearly with film thickness, indicating
a surface controlled process [170] (Fig. 27).
0 1000 2000 3000
0
50
100
150
200
250
D
o
p
i
n
g

i
r
r
a
d
i
a
t
i
o
n

t
i
m
e

(
m
i
n
)
LB film thickness ()

Figure-27

The environment in which the film was exposed to X-ray also influenced the doping. While no change was observed
when the film was exposed under vacuum or under dry N
2
, O
2
and argon atmospheres, the increase in humidity
resulted in changes within a short exposure time [170]. The changes in the FTIR spectra upon irradiation are also
similar to those observed upon acid doping of polyaniline. When compared with acid doping, two major differences
were observed for the LB films exposed to X-rays. First, the packing order of the cadmium stearate domains in the
41
composite LB films - as observed by X-ray diffraction - is not affected by X-ray irradiation. In addition, no
significant increase in the DC conductivity was noted after the X-ray exposure whereas similar LB films have their
conductivity increased by a few orders of magnitude upon acid doping. These differences may be explained by
considering that the inter-domain contribution to the conductivity is increased by the acid doping because the
insulating cadmium stearate domains are destroyed, which does not occur with the X-ray irradiation [169, 170].
Mixed LB films of polyaniline-cadmium arachidate were used to obtain polyaniline-cadmium sulfide thin films,
through post-deposition treatment of precursor mixed LB film with H
2
S gas and subsequent selective removal of
arachidic acid [171].

9. Devices using polyaniline LB films
The fabrication of molecular photodiodes which exhibit both rectifying and photoswitching characteristics
has been reported by Kim et al [172], using hetero-LB films of electron acceptor/sensitizer/donor arrays stacked on
ITO glass coated with composite LB films of I
2
-doped polyaniline-stearic acid as top electrode. The usefulness of
polyaniline LB films as gas sensors was also reported, where the surface plasmon resonance curves obtained for
polyaniline LB films were influenced by NO
2
and H
2
S [173]. The effects were partially reversible, with lower
detection limits of ~50 vapor ppm at room temperatures [173]. Conductometric gas sensors, sensitive to NH
3
or HCl
in the ppm range, were based on mixed LB films from polyaniline and phosphorylated calix[4]resorcinolarene
derivative (CA) [152]. The enzyme glucose oxidase (GOX), entrapped between layers of a polyaniline LB film on
ITO, was tested for retention of its catalytic activity and its ability to perform like a reagentless glucose biosensor.
From the anodic peak current on 0.7 V with respect to Ag/AgCl measured by cyclic voltammetry, a calibration
curve for a PANI-LB-GOX based glucose biosensor was obtained [151]. Composite LB films of poly(o-
methoxyaniline) and stearic acid (3:1 molar ratio) deposited onto a QCM plate were employed for chemical sensing
by monitoring the change in resonance frequency upon adsorption/reaction with sensing material [130]. This sensor
is capable of selectively sensing fatty acid vapors such as formic acid, acetic acid, propionic acid and butyric acid
with no or little response to ketone, alcohol, amine and hydrocarbon (hexane). A sharp drop in sensit ivity was
observed at about 65 C owing to the melting of stearic acid which causes the film morphology to change. The
sensitivity increased with the number of layers, but then there was loss of selectivity [130]. As indicated in the
previous section, insulating polyaniline (EB) LB films have been used at the electrode interfaces in
polyalkylthiophene-based light emitting diodes to control device operation [68, 101].

2.5. LB manipulation of azobenzene polymers
42
Azobenzene derivatives exist in two kinds of molecular configuration with two corresponding energy
states, namely the trans and cis forms. These molecules are spectroscopically characterized by a low energy n-*
band in the visible region and a high energy -* band in the UV region. The trans form of azobenzene is
thermodynamically more stable than the cis and hence the photostationary state, at room temperature and ambient
light, contains a higher population of trans isomers. However, upon irradiation with UV-light, the equilibrium is
shifted to a larger number of cis isomers. These cis- isomers are stable in the dark and at low temperatures but
slowly reverse to trans isomers according to a first order kinetics. Irradiation with visible light or heating quickly
shifts the equilibrium of the photostationary state to that possessing more trans isomers [174]. The transformations
of either trans to cis or cis to trans occur in an unhindered manner in solution and can conveniently be followed by
absorption spectral changes. The cis isomer occupies a larger volume than the trans isomer, for while the elongated
rod-like trans form is flat, the phenyl rings in the bent banana-like cis form are not coplanar [175] (Fig. 28), and for
this reason photoisomerization of azobenzene is usually prohibited in the solid state.

N
N
N N
9.0
5.5






= 360 nm
= 470 nm
trans - cis -

Figure-28

The concept of free volume has been put forward to explain the absence of photoisomerization of azobenzene
chromophores, since a free volume (about 0.12 nm
3
) should be provided around the chromophore for a facile
reversible cis-trans photoisomerization in the LB film [176].
Experiments performed in the femtosecond scale revealed that azobenzene photoisomerization is an
ultrafast process, possibly occurring via inversion at one of the nitrogen atoms [177]. For details on the
photoinduced molecular orientation phenomena in polymers, liquid crystals and LB films of light absorbing
chromophores, the readers are referred to a recent review by Blinov [178]. Several questions such as how fast trans-
43
cis-trans transitions and slow reorientation accompany each other in photoinduced optical anisotropy are yet to be
answered.
In a solid matrix such as an LB film, the azobenzene chromophores are usually aggregated which can be
inferred from the shift in
max
of the -* transition of the azobenzene to shorter/higher wavelengths (blue / red
shift), in comparison to the spectrum in solution where the molecules are considered to be molecularly dissolved.
Such a shift is either due to the parallel or antiparallel orientation of transition moments along the long axes of the
azobenzene group (H/J-aggregation). The photo-induced trans-cis-trans isomerization of the azobenzene group is
usually suppressed in a rigid matrix such as LB films [179], owing to the spatial inhibition resulting in the lack of
free volume necessary for the geometrical changes accompanying photoisomerization. Different molecular
engineering approaches have been put forward to overcome such a difficulty, with direct implications for the
development of optical devices based on azobenzene compounds. Results obtained with monomeric azobenzenes
and azobenzene polymers are summarized separately below.

1. Azobenzene-containing Monomeric Amphiphiles
Systematic monolayer and LB experiments with azobenzene containing amphiphiles have essentially
started with researchers from Japan in the early 1980s. Nakahara et al. [180] reported the control of orientation of
azobenzene chromophores at the air-water interface, whether parallel or perpendicular or tilted, by varying the
number and position of long alkyl substituents attached to the azobenzene group (25-29, Fig. 29). This was inferred
through surface pressure-area isotherms of monolayers and polarization dependent infrared spectroscopy of the LB
films transferred by the horizontal lifting method.
N
N
NH
HN
O
O
N
N
HN
NH
O
O
(28)
( 29)
N
N
NH
O
O
N
N
HN
NH
O
N
N
NH
HN
O
O
(25) (26) (27)
28
2
47
2
49
2
58
2 74
2

Figure-29

44
An unhindered cis-trans photoisomerization of azobenzene chromophores in guest-host LB films [179,
181, 182] containing -cyclodextrin (host) and simple non-amphiphilic azobenzene derivatives (guest) appears to be
one of the first successful approaches leading to a facile cis -trans photoisomerization of azobenzene chromophore in
an LB matrix (Fig. 30). Also shown in the figure are the structures of azobenzene compounds used in such study.
O
O OH
OH
CH
2
NHC
12
H
25
7
N X
N Y
X = NMe
2
/ H ; Y = COOH / SO
3
Na / H

N
X
N
Y
NH
NH
NH
NH

Figure-30

The use of amphotropic spacer polymers as host material (30, Fig. 31) to incorporate different low molecular weight
functional guest materials including azobenzene derivatives (31, Fig. 31) for obtaining functional polymeric LB
films has been reported by Laschewsky et. al [183]. This mixed LB film methodology offered a way to organize the
functional guest materials in an ordered fashion since these guest materials do not self-organize at the air-water
interface on their own.
45
O O (CH
2
)
6
OOC CH
CH
2
CH
CH
2
COO(CH
2
)
2
OH
CH
2
CH
3
C
2
H
5
H
*
1
3.5
N O
N CN (CH
2
)
6
COO C H
2
CH
3
(30)
(31)

Figure-31

Liu et. al [184] reported an excellent reversible photoisomerization of an amphiphilic azobenzene derivative (32,
Fig. 32) in the LB matrix, if the LB films were pretreated with UV irradiation which seems to result in film
expansion.
N C
8
H
17
N O (CH
2
)
5
COOH
(32)

Figure-32

Reversible photoisomerization in LB films of a specially designed azobenzene amphiphile bearing two alkyl chains
(33, Fig. 33) has also been demonstrated [185]. It is a typical example where the required free volume has been
achieved by introducing bulky hydrophobic parts, in contrast to the strategies of having larger head groups [185].
N N
N
CH
3
(CH
2
)
16
CH
2
CH
3
(CH
2
)
16
CH
2
(33)

Figure-33

The aggregation of azobenzene-containing long chain fatty acids, similar to 32 on the surface of pure water and on
the water containing barium ions, has been studied through in-situ UV-vis absorption spectral measurements. The
results were compared with the aggregation in the transferred LB films. The extent of aggregation depended on
46
whether the LB films were obtained from monolayers spread on pure water/water containing barium ions [186]. The
level of aggregation of the azobenzene amphiphiles (34-36, Fig. 34) in a mixed monolayer and its dependence on the
structure of the azobenzene amphiphiles have
N
N OC
10
H
20
COOH
N
N COOH
C
10
H
20
O
N
N
C
10
H
20
O
NO
2
(34)
(35)
(36)

Figure-34

been studied through surface pressure-area isotherms of the monolayer as well as through UV-vis spectral changes
of the LB films [187-189]. Apart from in-situ UV-vis measurements, the photo-induced cis-trans isomerization of
azobenzene amphiphiles has been monitored through surface potential and Brewster reflectivity changes [190, 191].
In a systematic study performed with azobenzene amphiphiles with a pyridinium head group (37), the trans isomer
(35
2
) was found to occupy a smaller area than the corresponding cis isomer (76
2
). In a mixed monolayer of 37
with dimyristoylphosphatidic acid (DMPA) which contained negative head groups, a molecular level mixture was
realized (Fig. 35) which guaranteed an unhindered reversible cis -trans photoisomerization. In such a mixed system,
the change in surface pressure/area owing to cis-trans photoisomerization is expected to be small. Nevertheless,
significant changes in surface potential (about 100 mV) and Brewster reflectivity (by a factor of 4) were observed
during cis-trans photoisomerization [190, 191].
CH
3
(CH
2
)
7
N
N O(CH
2
)
20
N +
Br
-
(37)

47
N
(CH
2
)
20
O
N
N
C
8
H
17
+
D
M
P
A
N
( CH
2
)
20
O
N
N
C
8
H
17
+
D
M
P
A
N
(CH
2
)
20
O
N
N
C
8
H
17
+
D
M
P
A
N
(CH
2
)
20
O
N
N
C
8
H
17
+
D
M
P
A
N
( CH
2
)
20
O
N
N
C
8
H
17
+
WATER
AIR

Figure-35

Langmuir monolayers of cis and trans azocrown ethers in which the azobenzene is part of the crown ether
(38 and 39, Fig. 36) have been studied by surface pressure and surface potential isotherms [192]. As one expects, the
monolayer characteristics such as the limiting area, compressibility and surface potential depended on the size of the
alkyl side chains as well as on the geometry around the azobenzene chromophore. While the trans to cis isomer
conversion could be effected by the UV-light irradiation, the cis to trans could be achieved with suitable alkali metal
ions in the subphase. The driving force for the latter seems to be the selective complexation of metal ions with the
trans isomer [192].
N N
O
R
O
R
O
N N
O
R
O
R
O O
(38)
(39)

Figure-36

Similarly, non-linear optical characteristics of the LB films of specially designed monomeric azobenzocrown ether
amphiphiles (40 and 41, Fig. 37) have been reported by Yao et. al. [193]. The molecular hyper polarizability ()
values deduced from second harmonic generation measurements were about 10
-28
-10
-29
esu.
48
O
O
O
O
N
O
N
OH
C
16
H
33
O
O
O
O
O
N
N H
2
N C
16
H
33
(40)
(41)

Figure-37

Phase separated mixed monolayers were formed with various mole percentages of cadmium salts of an azo acid
(6Az10COOH) and arachidic acid. Evidence of phase separation was obtained from surface pressure-area isotherms
and X-ray diffraction (XRD) of mixed LB films, with the azobenzene molecules being H-aggregated within such
domains. Interestingly, the mixed LB film seemed to consist of three phases of molecular organizates: domains
containing pure cadmium arachidate (CdAr) and pure cadmium salt of 6Az10COOH and a third type in which both
CdAr and Cd-6Az10COO
-
salts straddle each other [194]. A similar H-type aggregation of stearoyl ester of disperse
red-13 (DR13St) (42, Fig. 38) in mixed LB films of DR13St and cadmium stearate has been shown to hamper
photoisomerization of azochromophores and hence their optical storage capability [195].
O
N
Cl
O
2
N N N
C
2
H
5
C
18
H
37
O
(42)

Figure-38

Mixed Langmuir monolayers of azobenzene amphiphiles and phospholipids were studied by in-situ
Maxwell displacement current (MDC) measurements, which can be related to molecular dipole moments [196, 197],
similarly to surface potential measurements. The close packing of azobenzene amphiphiles in the Langmuir film
seemed to be affected by spacer phospholipids, resulting in a facile cis-trans photoisomerization of the azobenzene
chromophore due to the availability of free volume. Since the cis-trans isomerization of azobenzene is
49
accompanied by a significant change in the vertical component of the dipole moment, transient MDC pulses that are
molecular area dependent [198] are observed during photoisomerization. The charge flow and MDC pulses
generated during cis-trans photoisomerization have also been investigated as a function of monolayer compression
cycles [199]. The influence of the nature of alkyl chains of phospholipids on the photo switching of azobenzene
amphiphiles has also been studied. The aggregation of azobenzene chromophores was reduced both in Langmuir
monolayers and LB films, by the introduction of phospholipids possessing either unsaturated long alkyl chains
(DOLPC) or saturated long alkyl chains (d-DPPC). Regardless of the nature of the alkyl chains, the surface pressure
at a constant molecular area changed during the cis-trans photoisomerization process. On the one hand, MDC was
generated from DOLPC-azobenzene mixed monolayers over the entire range of molecular areas, whereas the
generation of MDC from d-DPPC-azobenzene mixed monolayers was limited to certain molecular areas [200-202].
Morphological changes upon UV-irradiation of LB films of an azobenzene amphiphile have been
investigated by atomic force microscopy [203]. Though no signature for the photo-induced trans-cis isomerization
was noticed in UV-spectral measurements, owing to the tight packing of azobenzene amphiphiles in the LB film
structure, significant changes in the crystalline packing were observed when these films were exposed to UV-
irradiation. The photoinduced isomerization (both trans-cis and cis-trans) of azobenzene amphiphiles enhances J-
aggregation of cyanine dye in a mixed LB film structure [204]. No significant morphological change was observed
while alternately irradiating LB films of pure azobenzene amphiphiles, but a cone like structure was found to
develop during irradiation of mixed LB films consisting of azobenzene amphiphile and cyanine dyes. These
structures have been related to the self-organization of cyanine dyes into a three-dimensional structure. The cyanine
dyes on their own do not photoisomerize and apparently the energy necessary for the self-organization has been
provided by the coexisting azobenzene amphiphiles which undergo photoisomerization upon light absorption [204].
Azobenzene amphiphiles tend to aggregate when they spread at the air-water interface and such
aggregation remains in the transferred LB films. It may be circumvented using the approaches outlined above.
Another alternative is the attachment of azobenzene chromophores as side chains to a polymer main chain or the
electrostatic complexation of azobenzene amphiphiles with ionic polymers. In these cases, chromophore aggregation
is expected to be significantly reduced. In the following paragraphs, we discuss monolayer properties of azobenzene
polymers both with reference to their monolayer properties and the LB film characteristics.

2. Azobenzene Polymers
Azobenzene polymers studied for their monolayer properties include polyamides [205], polylysine [206],
polyglutamate [207] and poly(vinylalcohol)s [208]. For instance, the photochemical response (change of area) under
light stimuli is faster, with a completely reversible isomerization, in monolayers of azobenzene polymer (43, Fig.
50
39) in comparison to simple azobenzene amphiphiles [209], owing to reduced chromophore aggregation. In contrast,
in the case of monomeric azobenzene amphiphile (44, Fig. 39), a lower photoresponse was observed when the head
group was either ionized or complexed with divalent ions such as cadmium [209].
N
O
N
C
6
H
13
(CH
2
)n
O
O
CH
CH
2
CH
2
CH
OH
500
x
1-x
N
O
N
C
6
H
13
(CH
2
)n
O
O H
(43)
(44)

Figure-39

The photoisomerization of azobenzene amphiphiles in an LB matrix could however be achieved with the polyion
complex approach in which the head group of the azobenzene amphiphile formed a complex with ionic sites of the
polymer dissolved in the subphase [210-212]. The structure of some ionic polymers employed is depicted in Fig. 40.
CH
CH
2
NH
2
CH
2
n
CH CH
2
NH
3
+
n
Cl
N
CH CH
2
CH
3
+
n
I
N
CH
2
CH
2
C H
3 CH
3
+
n
Cl
CH
CH
2
N C H
3
CH
3
CH
3
+
Cl
CH
2
n
PAA PVA
PVC1
PDAA
PVB

Figure-40

51
The complexation was found to be pH dependent and the higher cross-sectional area per molecule could be
controlled by choosing the polymer that eventually led to a favorable environment (free volume) for facile reversible
cis-trans isomerization. Using this approach, the photoresponsive switching in conductivity of LB films of specially
designed azobenzene amphiphiles has been studied by Tachibana et. al [213-215]. In this case, the azobenzene
group acts as a switching unit connected to the working unit (TCNQ group) (Fig. 41). When such LB films were
irradiated alternately with UV and visible lights, a change in conductivity of ca. 20% was observed.
N N (CH
2
)
7
O (CH
2
)
12
C H
3
N
+
(TCNQ)
2
-
switching unit
working unit
transmission
unit

Figure-41

In LB films from polyion complexes of an azobenzene amphiphile bearing two azobenzene units with negative head
group (45, Fig. 42) and cationic polymers, the fraction of cis isomer produced at the photostationary state upon UV-
irradiation increased with increasing area per molecule at which the LB films were made [216].
N C
8
H
17
N O(CH
2
)
2
OCOCH
2
N C
8
H
17
N O(CH
2
)
2
OCOCH SO
3
Na
(45)

Figure-42

Monolayer and LB films of poly(vinyl alcohol)s with covalently attached azobenzene side chains via methylene
spacers were studied by Seki et. al. [217, 218]. Monolayers of both cis- and trans- isomers were obtained and an
efficient and reversible photoisomerization of azobenzene chromophore could be exercised in the transferred LB
films. These azobenzene polymer LB films were useful as command surfaces for photochemical alignment of a
nematic liquid crystal [219]. This alignment was studied as a function of molecular structure, number of deposited
layers, number density of azobenzene units, deposition method and method of light irradiation [220]. A single
azobenzene monolayer was capable of inducing liquid crystal alignment, and the performance of these command
surfaces depended on the length of the spacer group connecting the azobenzene to the polymer main chain.
Monolayer characteristics of a similar polyvinylalcohol attached to an azobenzene side chain have been reported by
Takahashi et al. [221].
52
Brewster angle microscopy was used to investigate morphological changes caused by photoisomerization
of an azobenzene polymer monolayer [222]. The trans isomer exhibited rigid iceberg-like domains at large areas per
molecule and a homogeneous pattern for a condensed monolayer, while a fluid-like, homogeneous monolayer was
observed for the cis isomer. Furthermore, reflected light was less intense for the cis isomer owing to the reduced
monolayer thickness and lower number density. Upon UV-light irradiation, the domain structure characteristics of
trans isomers disappeared and a homogeneous monolayer typical of cis isomers was observed [222].
The formation of a stable and transferable monolayer from a methacrylic homopolymer of disperse red-13
and disperse red-1 dyes (HPDR1(46) and HPDR13 (47), Fig. 43), without long alkyl substituents, was reported
[223, 224]. Interestingly, neither the disperse red dyes nor their methacylate derivatives form stable monolayers
owing to complete/partial solubility in the subphase water. The monolayer could be transferred as Z-type LB films
using the vertical dipping method up to 30 layers, after which the transfer ratio decreased. However, upon using the
mixed LB film approach, in which HPDR13 was
*
CH
3
C
H
2
*
O
O
CH
2
CH
2
N
C
2
H
5
N
N
Cl
NO
2
n
*
CH
3
C
H
2
*
O
O
CH
2
CH
2
N
C
2
H
5
N
N
NO
2
n
(46)
(47)

Figure-43

codeposited with cadmium stearate, Y-type LB films with up to a few hundred layers could be obtained [225]. The
mixed monolayers containing various weight percentage of the polymer and cadmium stearate were studied through
surface pressure and surface potential isotherms and BAM. The maximum surface potential lied between the values
for pure HPDR13 (160 mV) and pure cadmium stearate (120 mV) monolayers. BAM analysis indicated the
formation of a large continuous domain structure, prior to monolayer compression. In both pure and mixed LB
films, HPDR13 seemed to be J-aggregated, as inferred from the red-shift of the absorption maxima in comparison to
53
that of chloroform solution. The non-linear optical properties of the mixed LB films containing HPDR13 and
cadmium stearate were studied by the Z-scan technique with Fourier analysis. The latter technique enabled one to
conclude that the non-linear effects are mainly of thermal origin. The non-linear coefficient (n
2
) obtained was about
6.0 X 10
-5
cm
2
/kW and the thermal diffusion coefficient (D) was about 2.5 X 10
-3
cm
2
/s [226].
Langmuir monolayers of cis (UV-irradiated) and trans (dark) forms of poly (L-lysine)s with 43% and
100% substitution of azobenzene side chains (48 and 49, Fig. 44) were studied by surface potential-area
measurements, in addition to surface pressure-area isotherms [227]. The trans isomer displayed higher surface
potentials than the cis isomer,
NH CH CO NH
(CH
2
)
4
NH
2
CH CO
(CH
2
)
4
NH
CO
N
N
57% 43%
N
H
CH CO
(CH
2
)
4
NH
n
N
N
SO
2
(48)
(49)

Figure-44

in both polymers, similar to an earl ier report [190], and these results were interpreted using the Demchak and Fort
model [228] and Oliveiras inverted cone model [229] to infer the angle of inclination of the dipole moment of the
azobenzene chromophores. The required thickness values were obtained by in-situ ellipsometric measurements.
From such analysis, the chromophore was found to lie perpendicular to the interface in the compressed monolayer in
both polymers. These results should be considered as preliminary for the possible changes in dipole moment owing
to distinct conformations of the polymers due to chain substitution were not taken into account. The formation of
expanded monolayers with collapse pressures ranging from 40-50 mN/m from three methacrylate copolymers with
azobenzene side chain substitution of 25, 50 and 75 mole percentages (50, Fig. 45) was reported by Ren et al. [230].
54
N OC
6
H
12
O N COOCH
2
CH*(CH
3
)C
2
H
5
O
CH
2
CH
2
OCH
2
CH
2
OH
O
x
y
x = 0.25, 0.5, 0.75
(50)

Figure-45

Z-type LB films obtained from these monolayers were characterized by polarized UV-vis and IR spectroscopies
which provided a clue for the preferential orientation of azochromophores along the film normal. Irrespective of the
azobenzene substitution, azobenzene chromophores undergo reversible trans-cis-trans photoisomerization. The
copolymer with lower azobenzene substitution has been used for photoinduced alignment of liquid crystals on top of
the LB films [230]. A clear transition in monolayer behavior was observed when the percentage of dye incorporation
in methacrylate copolymers was varied (51, Fig. 46). Copolymers with low dye contents presented an expanded
monolayer, but those with higher dye content formed a
*
C
H
2
O
O
CH
2
C H
2
N
C
2
H
5
N
N
Cl
NO
2
CH
3
CH
3
C
H
2
*
O
O
CH
2
CH
2
OH
y x
x + y = 1.0
x = 0.02, 0.11, 0.20,
0.29, 0.51
(51)

Figure-46

condensed monolayer. Again, monolayers of pure copolymers could not be transferred onto solid substrates, but
mixed monolayers with cadmium stearate did form Y-type LB films [231]. The same applied to monolayers from
disperse red-19 isophorone polyurethane (52, DR19-IPPU, Fig. 47) mixed with cadmium stearate [232].
55
N
OCH
2
CH
2
CH
2
CH
2
O
O
N
H
C
H
2
N
H
O
n
(52)

Figure-47

Annealing and UV-irradiation on LB films from a thermotropic copolyacrylate with alkoxyazobenzene and
hydroxyethyl side chains have caused an order-disorder transition, as investigated by angular dependent absorption
spectroscopy [233]. The photoinduced orientation of azochromophores was restricted in the as -deposited films, but
it occurred in a reversible way once the intermolecular interactions (aggregation of the azobenzene chromophores)
were weakened either by annealing/UV-irradiation. The photoinduced dichroism in these LB films was stable and
could be erased by irradiation/heating [233].
The photoisomerization characteristics of azobenzene polymer LB films, deposited onto a thin gold film,
have been investigated by surface plasmon resonance (SPR) [234, 235]. This technique employs the collective
resonant oscillation of free electrons at a metal-dielectric (here LB film) interface. It is usually employed in the ATR
mode and provides information about changes of thickness and refractive index of dielectrics on a metal film. A
reversible change in the resonance angle of the guided reflective mode, which is related to the change in refractive
index of the dielectric, was noticed in the SPR spectra of azobenzene polymer LB films when the film was
alternately irradiated with UV and blue lights. This optical switching process could be repeated several times.
Sawodny et. al. [236] have used azobenzene LB films deposited on a metal-coated glass substrate for making read
and write patterns with surface plasmon microscopy.
The fine thickness control offered by the LB technique has been exploited for the fabrication of quarter-
wave and rugate optical interference filters with low scattering behavior, using LB films from polysiloxane with
azobenzene side chains [237]. The structural changes in LB films of polyglutamates with azobenzene side chains
caused by photoisomerization/annealing were studied by small angle X-ray scattering and UV-vis spectroscopy
[238]. Upon photoisomerization/annealing, a transition was observed from a deformed hairy-rod arranged bilayer
structure to one with symmetrically distributed side chains around the main polymer chain. The extent of
chromophore aggregation and its dependence on the length of the spacer methylene group also varied [238].
Substantial morphological changes have been observed through AFM in a single-layer LB film of an
anionic azobenzene amphiphile complexed with a polycation [239]. It was suggested that the 2D LB film structure
was substantially modified to accommodate the increase in cross-sectional area of the azobenzene when the trans-
56
cis photoisomerization occurred. This observation is not obvious, as one does not expect significant changes in the
2D LB structure during photoisomerization, based on the free volume concept according to which the
photoisomerization is expected to proceed within the layer. The occurrence of reversible photoisomerization in LB
films has been confirmed with UV-vis and IR spectral changes. The film was smooth with surface undulation of less
than 1 nm before UV-light illumination, but the number of hills with height of 5 nm and the base of 100 nm was
seen after the light illumination and these hills were found to disappear upon illuminating with visible light. Similar
changes were observed even for 2- and 3-layer LB films, though these films contained defects (holes/small hills) to
start with. Such hill formation in the film upon UV-light illumination has been understood as the curvature to
release the stress developed within the film when the increase of cross-sectional area due to trans-cis
photoisomerization exceeds that available within the layer of the LB film [239] (Scheme-4).
-
+ + + + + +
- - - - -
-
+ + + + + +
- - - - -
-
-
-
-
-
-
-
-
-
-
-
-
UV VIS

Scheme4. Model proposed for the structural change of a single layer LB film of an azobenzene containing polyion
complex upon photoisomerization [239].

The photomechanical behavior of monolayers of azobenzene-containing poly(vinyl) alcohol has also been
investigated [240-242]. In recent studies, monolayers of azobenzene functionalized polyvinylalcohol were studied
by surface potential and surface pressure isotherms, BAM and in-situ UV-vis absorption spectroscopy [241-242].
The surface potential increased at a relatively large surface area where the surface pressure was still zero and
reached a maximum value of about 200 mV in the compressed state for the trans isomer. For the cis isomer,
however, the initial potential of ca. 400 mV decreased to ca. 200 mV upon film compression, which is in accordance
with the model proposed by Seki et. al [241] where the cis azobenzene is in contact with the water surface while the
trans azobenzene stays in the air. It is in contrast, however, to an earlier report where the surface potential increased
monotonically during monolayer compression [227]. Such discrepancies may be attributed to the difference in the
polymer to which the azobenzene chromophore is attached. The expansion and contraction process which
57
accompanies photoisomerization of azobenzene chromophores could be observed even at zero surface pressure by
monitoring changes in a single domain through BAM. In combination with the macroscopic photomechanical
responses (the change in area at constant surface pressure), BAM results indicated that the expansion and
contraction mechanical processes are unsymmetrical in terms of light energy needed and the molecular motions
involved [242] (Scheme-5). The photomechanical response of the azobenzene polymer monolayer has also been
inferred through dancing of photo-inactive substances such as camphor grains dispersed onto a compressed
monolayer under light irradiation [209].
N
N
N
N
N
N
N
N
Air
Water
N
N
N
N
N
N
Air
Water
expansion
contraction
trans
cis
365 nm
436 nm
(UV) (vis)

Scheme5. Schematic representation of the reversible photoinduced area change of an azobenzene containing polyion
complex at the air-water interface [242].

A reversed Duckweed polymer consisting of hexanediamine-epichlorohydrin hydrophilic network and hydrophobic
chains containing p-nitroazobenzene groups (Fig. 48) forms a monolayer with an average available area per
azobenzene group of about 47
2
, which is higher than the cross sectional area of an azobenzene group (about 30

2
) [243]. Z-type LB films could be deposited fromthese monolayers [243].
58
N N N
CH
3
(CH
2
)
6
CH
2
-
=

Figure-48

Using novel acrylic polymers containing azobenzene with long alkyl side chains, Crosswell et al. [244] produced LB
films with effective electro-optic coefficients ranging from 1 to 8.7 pm V
-1
, and where an enhanced value of 13 pm
V
-1
could be obtained when the azobenzene polymer layer was combined in an alternate layer film structure.
Recently, Weener et al. [245-246] reported the synthesis and photoswitching properties of Langmuir films of
azobenzene derivatized amphiphilic poly(propylene imine) dendrimers of 1-5 generations [245-246]. The fifth
generation poly(propylene imine) dendrimer functionalized with azobenzene containing alkyl chains is shown in
Fig. 49a. Due to strong H-type azobenzene aggregation which seemed to result in a tight packing of chromophores,
no cis-trans switching for monolayer of this macromolecule was observed upon irradiation. However, a
unimolecular inverted dendritic micelle, with a random shell structure containing equal amounts of alkyl chains with
and without azobenzene chromophores (Fig. 49b), exhibited a reversible cis-trans switching behavior, in Langmuir
monolayers as well as in LB films [246].
O
O
N N
O
O
O
N N
O
O
O
O
O
N
N
O
O
O
N
N
O
O
O
O
O N
N O
O
O N
N O
O
O
O
O
N
N
O
O
O
N N
O
O
O
O
O
N
N
O
O
O
N
N
O
O
O
O
O
N
N
O
O
O
N
N
O
O
O
O
O
N
N
O
O
O
N
N
O
O
O
O
O
N
N
O
O
O
N
N
O
O
O
O
O
N
N
O
O
O
N
N
O
O
O
O
O
N
N
O
O
O
N
N
O
O
O
O
O
N
N
O
O
O
N
N
O
O
O
O
O
N
N
O
O
O
N
N
O
O
O
O
O
N N
O
O
O
N N
O
O
O
O
O
N N
O
O
O
N
N
O
O
O
O
O
N
N
O
O
O
N
N
O
O
O
O
O
N N
O
O
O N
N O
O
O
N
N
N
N
N
N
N
N
N
N
N
N
N
N N N
N
N
N
N
N
N
N
N
N
N
N
N
N
N
N N
N
N
N
N
N
N
N N
N
N
N
N
N
N
N
N
N
N
N
N
N
N
N
N
N
N
N
N
N
N
HN
HN
HN
HN
HNHNHNHN
HN
HN
HN
HN
HN
HN
HN
HN
HN
HN
HN
HN
HN
HN
HN
HN
HN
HN
HN
HN
HNHNHNHNNHNHNH
NH
NH
NH
NH
NH
NH
NH
NH
NH
NH
NH
NH
NH
NH
NH
NH
NH
NH
NH
NH
NH
NH
NH
NH
NH
NH
NH NH NH
NHNHNH
NH
NH
NH
NH
NH
NH
NH
NH
NH
NH
NH
NH
NH
NH
NH
NH
NH
NH
NH
NH
NH
NH
NH
NH
NH
NH
NH NH NH HN HN HN
HN
HN
HN
HN
HN
HN
HN
HN
HN
HN
HN
HN
HN
HN
HN
HN
HN
HN
HN
HN
HN
HN NH HN HN
HN
HN
HN
HN
N
N
N
N
N
N
N
N
N
N
N
N
N
N
N
N
N
N
N
N
N
N
N N
N
N
N
N
N
N
N N
N
N
N
N
N
N
N
N
N
N
N
N
N
N
N N N
N
N
N
N
N
N
N
N
N
N
N
N
N
O N
N O
O
O
N
N
O
O
O
N
N
O
O
O
N
N
O
O
O
N
N
O
O
O
N
N
O
O
O
N
N
O
O
O
N
N
O
O
O
N
N
O
O
O
N
N
O
O
O
N
N
O
O
O
N
N
O
O
O
N
N
O
O
O
N
N
O
O
O
N
N
O
O
O
N
N
O
O O
N N
O
O
O
N N
O
O
O
N N
O
O
O
N N
O
O
O N
N O
O O N
N O
O
O
N
N
O
O
O
N N
O
O
O
N
N
O
O
O
N
N
O
O
O
N N
O
O
O
N N
O
O
O
N
N
O
O
O
N
N
O
O
O
N
N
O
O
O
N
N
O
O
O
N
N
O
O
O
N
N
O
O
O
N
N
O
O
O
N
N
O
O
O
N N
O
O
O
N N
O
O
O
N N
O
O
O
N
N
O
O
O
N
N
O
O
O
N
N
O
O
O N
N O
O
O N
N O
O
O
N N
O
O
O
N N
O
O
O
N N
O
O
O
N N
O
O
O
N
N
O
O
O
N
N
O
O
O
N
N
O
O
O
N
N
O
O
O
N
N
O
O
O
N
N
O
O
O
N
N
O
O
O
N
N
O
O
O
N
N
O
O
O
N
N
O
O
O
N
N
O
O
O
N
N
O
O
O
N
N
O
O
O
N
N
O
O
O
N
N
O
O
O N
N O
O

59

3. Storage Characteristics of azobenzene polymers
Fujishima et. al. [247] proposed a high-density information storage system with trans-cis
photoisomerization comprising the selective electrochemical reduction of the cis-isomer and the exclusive
reoxidation of the reduced form to trans isomer of LB films from azobenzene amphiphiles (Scheme -6). The highest
isomerization rate constant was measured
N N
trans-AZO
N N
cis-AZO
N
H
N
H
hydro-AZO
h
+ 2H
+
+ 2e
-
- 2H
+
- 2e
-

Scheme6. Photoisomerization and electrochemical oxidation-reduction of azobenzene [247].

at +0.85 V vs Ag/AgCl which represented an increase of about 3 orders of magnitude in comparison to the thermal
isomerization process, though the activation for both these processes are nearly equal [248]. The rate constant at a
given positive potential also depended on the molecular structure of the azobenzene amphiphiles, especially on the
length of the spacer methylene groups between the azobenzene group and the head -COOH group and on the length
of the tail alkyl group [248].
By alternately irradiating azo compounds with appropriate light, one can induce trans-cis-trans
isomerization in a reversible manner. In principle, systems with stable trans and cis states can be used in optical
switching (ON-OFF), as in azobenzenes without active substituents. Azobenzenes attached with push-pull
substituents exhibit a fast trans-cis-trans photo/thermal isomerization that can be useful in optical storage and for
creation of surface relief gratings. If circularly polarized light is used, reorientation of molecules upon absorption of
light is prevented. However, molecular reorientation results from a series of trans-cis-trans photoisomerization
processes when linearly polarized light is employed. For after such trans-cis-trans isomerization process, the
chromophore can adopt any orientation including its original one. In the next process, chromophores whose dipole
moment is perpendicular to the polarization direction are not affected, which will lead, after several cycles, to a net
population of chromophores oriented in this particular direction. This imbalance is seen as light-induced
birefringence in the film (WRITE). Furthermore, when the source light is switched-off, some molecules relax, but
60
still a considerable number of molecules remain oriented, leading to a stable birefringence pattern (STORE). The
appearance of birefringence is usually inferred from the change in transmittance of a weak probe beam (READ) that
passes through two crossed polarizers before and after the sample. The birefringence can be completely erased (i.e.
randomization of molecular orientation) by heating or overwriting the test spot with circularly polarized light
(ERASE).
Optical storage of azobenzene polymers has mainly been exploited in spin cast films [249-257]. This has
been extended to LB films recently, in order to take advantage of the thickness control and surface uniformity. For
in thick films the intensity of the writing beam may vary substantially across the film, which can lead to uneven
writing patterns [258]. There have been experiments indicating the optically induced birefringence in LB films of
small azo dyes (53 and 54, Fig. 50) [259-262].
N N
N N C
18
H
37
H
+

S N
O
O
NH
2
N N C
4
H
9
H
+

(53)
(54)

Figure-50

Reversible optical birefringence and dichroism have been demonstrated in LB films of a reversed Duckweed
polymer containing a hexanediamine-epichlorohydrin hydrophilic network and hydrophobic chains with p-
nitroazobenzene groups (refer to Fig. 45) [243]. Though no birefringence was observed prior to the laser light
irradiation, owing to the isotropic orientation of the azobenzene chromophores, significant birefringence was
induced upon irradiation as observed with a weak read beam. The angular dependence of the absorbance at a fixed
wavelength clearly pointed to dichroism in the LB film [243]. Photoinduced molecular orientation in the LB film of
a polyimide with azobenzene pendent unit has been studied by Yokoyama et al. [263,264] and the induced
orientational anisotropy displayed an angular independence to the polarization of the incident UV light.
Photoinduced optical birefringence in mixed LB films containing different azobenzene derivatized polymethacrylate
and polyurethanes and cadmium stearate has been studied by Oliveira and co-workers [231, 232, 265]. The result of
61
a typical optical storage experiment carried out with a mixed LB film containing an azobenzene polymer and
cadmium stearate is shown in Fig. 51.
0 20 40 60 80
0.0
0.2
0.4
0.6
0.8
1.0
C


B
A
T
r
a
n
s
m
i
t
t
e
d

S
i
g
n
a
l
Time (seconds)

Figure-51

When the writing beam is off, no transmission of the probe that passes through the film and the crossed polarizers is
observed. However, when the writing beam was switched-on at point A, the transmission of the probe beam
increased until reaching saturation and the transmission remained at this saturation value as long as the writing beam
was on, indicating the creation of birefringence in that film area. When the writing beam was switched off at point-
B, the transmission decreased sharply reaching about 35% of initial saturation value (point-C) and remained at that
level for several days, revealing that a significant number of molecules still remained oriented. Such induced
birefringence could nevertheless be erased either by optical means (overwriting the test spot with circularly
polarized light) or by thermal means (heating the sample). In practice, it has been observed that several such cycles
can be performed without significant changes in the maximum birefringence and material degradation. For a 100-
layer mixed LB film of HPDR13 and cadmium stearate, a laser power of 2 mW is sufficient to induce maximum
birefringence. The amplitude of induced birefringence increases with increasing amounts of HPDR13 in mixtures
with cadmium stearate, but the time to achieve 50% of maximum birefringence and the remaining birefringence are
not affected significantly with the change of composition. Interestingly, the maximum birefringence decreases with
increasing number of layers, probably due to the decrease in the packing order in the thicker LB films. A similar
observation has been made for mixed LB films from DR-19-IPPU and cadmium stearate. For the mixed LB films of
methacrylate copolymer with different dye contents and cadmium stearate, the maximum induced birefringence
increased with the weight percentage of the dye in the copolymer, but the time to achieve 50% of saturated
62
birefringence and the remaining birefringence decreased with increasing dye content (Fig. 52). The latter behavior
was explained in terms of the cooperative interaction among neighboring azo groups and also thermal effects.
0 10 20 30 40 50 60
-2
0
2
4
6
8
10
12
14
C
B
A
(b)
(c)
(d)
(a)
(e)
T
r
a
n
s
m
i
t
t
e
d

S
i
g
n
a
l

(
a
r
b
.

u
n
i
t
s
)
Time (s)
Figure-52

When thin films of azo polymers are exposed to the interference pattern of laser beams, a large-scale molecular
migration is observed which leads to regularly spaced surface relief gratings on the film surface. The mechanism of
mass transport upon exposure to the interference pattern is still the subject of intensive research, but it is generally
believed that the gratings are formed via photoinduced trans-cis-trans isomerization of azochromophores. The
molecular dipoles interact with the electric field gradient of the laser light, resulting in a worm-like physical
movement of the molecule. A typical surface relief grating pattern created on the mixed LB film of HPDR13 and
cadmium stearate, as observed through atomic force microscopy, is shown in Fig. 53. No such regular grating on the
film surface was observed prior to exposure to the interference pattern. The grating spacing was about 2.6 m and
the peak-valley height was in the range of 50-60 nm.
63

Figure-53

These values are similar to those observed with spin cast films of azo polymers. It has been observed that when a
low power laser (10 mW/cm
2
) instead of the laser with a power of 180 mW/cm
2
is used to create a surface relief
grating, only index gratings are observed by diffraction of the He-Ne laser [265]. UV-vis and FTIR spectra of the
exposed areas were obtained and compared with that of the unexposed areas. The AFM phase image analysis
indicated that the non-photoactive cadmium stearate molecules have also moved along with the photoactive
HPDR13 molecules [265].

2.6. LB manipulation of polydiacetylenes
Before going into details of the LB manipulation of polydiacetylenes, a short account on the topochemical
polymerization of diacetylenes into polydiacetyelenes is provided. Diacetylenes (DAs) containing two conjugated
triple bonds are the most extensively investigated monolayer materials for their post polymerization characteristics,
either in the form of monolayers at the air-water interface or as multilayer LB films. They can be polymerized by
exposure to UV or -ray or by thermal means (Scheme-7), where the polymerization reaction is believed to proceed
via the formation of carbene (two nonbonded electrons on a carbon atom) intermediates, which rapidly transform
into polydiacetylene (PDA). It makes the quenching of carbene intermediates by oxygen less significant.
64
R C C
C C R'
C C C
C
R
R'
[
]
UV / /


Scheme7. Topochemical polymerization of diacetylene into polydiacetylene

This polymerization occurs in single crystals of the diacetylene monomer and is sensitive to the size and
shape of the substituents attached to the diacetylene group, thus representing a classical example of
topochemical/lattice controlled polymerization reaction. Polymerization can be monitored by the appearance of
absorption bands corresponding to the conjugated polymer backbone in the visible region of the spectrum. No
polymerization is expected when diacetylene is in the melt state/dissolved in a good solvent where the molecules
have a higher degree of freedom for their movement. Even in the solid state, the diacetylene groups must be suitably
oriented for mutual interaction and only a limited tolerance of distortion of the unit cell is allowed during
polymerization (Scheme-8). For instance, the polymerization reaction is possible only when (i) the difference
between the translational period in the monomer (d
1
) and in the polymer (d
2
) approaches zero. For the diacetylenes
which are known to be reactive, d
1
lies between 4.7-5.2 and d
2
is almost constant (4.93 ). (ii) S
1
, the distance
between the reacting diacetylene groups of the adjacent molecules (see scheme-8), should be smaller than 4 with a
lower limit of 3.4 . (iii) the angle between the diacetylene rod and the translational vector () should be close to
45 [266]. Polydiacetylene mostly retains the crystal structure of the monomer from which it was formed. Despite its
structure, the conductivity of the polydiacetylene is not so high as one could have anticipated, owing to the difficulty
in introducing dopants into such tightly packed and well organized polymeric matrix. Nevertheless, PDAs are
looked upon as promising materials for optical sensors as well as non-linear optical materia ls. Chromic transitions of
PDAs (blue to red) are used in various sensing applications including biosensors.
R'
R
R
R' R'
R
S
1
d
1

R
R'
R
R' R'
R d
2

Scheme8. Topochemical requirements for diacetylene polymerization.


65
1. Polymerization of diacetylenes as applied to LB films
Polydiacetylene LB films can be obtained: (a) by polymerization of the diacetylene monolayer and
subsequent transfer onto the substrate, (b) by post-polymerization of the diacetylene LB film and (c) by spreading
soluble polydiacetylene as a monolayer that is transferred onto a solid substrate. The resulting polydiacetylene LB
film is envisaged to possess a one-dimensional conjugated backbone. The last approach is very limited because
simple amphiphilic polydiacetylenes are insoluble in most organic solvents. The use of mixtures of solvents such as
phenol/chloroform to spread quiniline-substituted polydiacetylene has been reported recently [267]. Soluble
polydiacetylenes can be obtained by attaching substituents such as urethanes to the diacetylene group. As far as LB
films are concerned, the polymerization of the monomer proceeds within the same layer but possibly in different
directions, the extent of which is determined by the size and shape of the domains in the polydiacetylene LB film.
Due to the rigidity of the resulting polydiacetylene, the polymerization of diacetylene monomers in the LB film is
known to lead to cracks in the LB film, which is one stumbling block for using polydiacetylenes for device
applications.
The topochemical polymerization of diacetylene and the fabrication of LB films represent an interesting
combination in the sense that while the latter is usually employed to pack the molecules in a desired fashion, the
former proceeds only when the molecules are packed in a specific arrangement. It would mean that the topochemical
polymerization can be used as a tool to infer the degree of order in the LB film. It should be stressed that
polydiacetylene LB films are not single crystals, but usually consist of an array of two-dimensional domains whose
size depends on the experimental conditions of film preparation. These domain structures are formed as soon as the
material is spread and compressed at the air-water interface and remain during film transfer and polymerization.
Since diacetylene polymerizes into polydiacetylene upon exposure to UV-light, one has to control ambient light
during preparation of monolayer/LB films from diacetylene amphiphiles. Use of red light in the preparation room is
recommended. In practice, diacetylene monolayers are polymerized by exposing the trough area to a UV light
source, just above the water surface. It is believed that during polymerization of monomeric diacetylene
amphiphiles, e.g. cadmium salt of a diacetylenic acid, in the form of LB films, different polymer phases are formed,
namely blue or blue green. In the formation of the blue phase, the 2-dimensional unit cell of the monomeric LB
film shrunk due to the covalent bond formation among the molecules, in contrast to no change in molecular packing
during the formation of bluish-green PDA. Electron diffraction analysis revealed differences in the repeating unit
distances of the main chains for the two forms, implying that they had distinct resonance backbone structures. For
instance, the bluish-green form comprised butatrienic-type main chains, whereas acetylenic chains were observed
for the blue form [268, 269].

66
2. Langmuir monolayers of diacetylene amphiphiles
Most monolayer studies related to diacetylene amphiphiles were carried with molecules having the
following general structure [270]:

CH
3
-(CH
2
)
n
-CC-CC-(CH
2
)
m
-X

with the variation of the number of methylene groups in the hydrophobic tail (n), the spacer methylene groups
between the polar head group and diacetylene group (m) and the polar groups such as COOH, CH
2
OH, etc. The
structural modifications of diacetylene amphiphiles, which include the nature of head groups and the position of
diacetylene in the hydrophobic chain, do influence the monolayer properties as well as the polymerization
characteristics. Surface pressure-area isotherms of diacetylene amphiphiles with either acid or alcohol head group, in
general, can be classified as of condensed type with no or a little phase transition prior to the monolayer collapse.
Langmuir monolayers of 4-(10,12-pentacosa-diynoicamidomethyl)-pyridine, a diacetylenic amphiphile with
pyridine head group (55, Fig. 54), exhibited a clear liquid expanded to liquid condensed phase transition in the
surface pressure-area isotherm.
(CH
2
)
8
CH
3
(CH
2
)
11
O
N
H
CH
2
N C
(55)

Figure-54

This phase transition depended on the subphase temperature and also on the absence/presence of complexing ions in
the subphase [271]. Monolayers of amino acid-derivatized diacetylene lipid, 10,12-pentacosadiynoic acid
succinimidyl ester on water were studied using EPI-fluorescence microscopy. The surface pressure-area isotherm
indicated the formation of a stable monolayer with hydrophilic lipids where an attenuation of liquid expanded (LE)-
liquid condensed (LC) phase transition was noticed upon decreasing the subphase pH, as evidenced by EPI-
fluorescence microscopy [272]. Apart from these studies where the role of the head group is investigated, there are
few reports on the influence of substitution in the tail group, as indicated below.
Para substituted 13-phenyl-10,12-tridecynoic acid (PTAs, 56, Fig. 55) are capable of forming stable
Langmuir monolayers that may be transferred as Y-type LB films and be photopolymerized with UV-irradiation,
whereas parent PTA did not form monolayers [273].
67
(CH
2
)
8
COOH X
(56)

Figure-55

Upon using m-alca-1,3-diynylbenzoic acids (57, Fig. 56) with various chain lengths, it has been found that only
those with 8 or more methylene chains formed condensed monolayers that could be transferred and
photopolymerized [274].
N O
2
N
CH
3
H
CH
2
OH CH
2
n
(57)

Figure-56

In addition to the molecular structure of diacetylenes, there are a number of parameters which seemed to
influence the monolayer properties as well as the polymerization characteristics within the monolayer/LB film
structure. The experimental parameters extensively studied t o achieve control over the polymerization of diacetylene
include the experimental conditions used during monolayer formation such as addition of divalent metal ions into
the subphase, choice of the spreading solvent, subphase temperature, application of electric field to the monolayer,
nature of the builder materials (in the case of mixed monolayers) and the experimental conditions for monolayer
transfer, such as deposition pressure, mode of deposition (whether, Z or Y or X-type), and nature of the substrate
(refer to next section for film transfer).
From a systematic monolayer study performed with diacetylenic acid purified using the zone refining
technique, it has been concluded that trace amounts of impurities can influence the monolayer characteristics, both
in terms of the area per molecule and the collapse pressure [275]. According to Tieke et al [276], the solvent used to
spread the monolayer material may affect the nature of domains formed at the initial stages. The change in surface
morphology of a diacetylene monolayer as a function of the concentration of the spreading solution and the area of
the spreading surface has also been investigated through in situ fluorescence microscopy measurements. 2D spiral
structures with a diameter as large as 1 mm were obtained by spreading a solution of relatively high concentration,
and a monolayer which consisted of uniformly oriented large single crystal domains of over 20 mm X 5 mm was
68
obtained when the area of the water surface was slowly increased by moving a barrier after spreading a highly
concentrated solution onto a small bounded water surface. The surface structure of the monolayer during
photopolymerization depended on the atmosphere in which the monolayer was formed. For instance, the surface
morphology was unchanged in the argon atmosphere and a large deformation occurred in the unit cell in air [277-
279].
Owing to the poor stability of monolayers from diacetylene amphiphilic acids, most studies were
performed with the corresponding cadmium salt, similarly to simple fatty acids. In a study with amphiphilic
diacetylene acids with various hydrophobic chain lengths (n= 8, 9, 11 and 13) and at different subphase pHs, it has
been concluded that the cadmium salt formation starts at pH ~ 5.0 and is complete at pH ~ 6.5. This was inferred
through the change in collapse pressure and infrared characterization of transferred LB films (a strong band at 1550
cm
-1
for cadmium salt) [280], analogously to that observed with simple fatty acid monolayers. Therefore, there is no
obvious interference of the diacetylene group in the hydrophobic tail with the ionization of the head COOH group.
Srikhirin et al [281] reported the stabilization and polymerization of Langmuir monolayers of two diacetylene
monomers, 14-{4' -[(methoxy)methoxy] biphenyl}-10,12-tetradecadiynoic acid) and (14-{4'-[(nitro)methoxy]
biphenyl}-10,12-tetradecadiynoic acid) (58 and 59, Fig. 57) by addition of Ba
2+
and Cd
2+
salts into the subphase.
OCH
2 O
2
N
(CH
2
)m COOH
OCH
2 MeO
(CH
2
)m COOH
(58)
(59)

Figure-57

It is noted that these monomers neither form a stable monolayer nor undergo polymerization on the surface of pure
water [281]. However, stable monolayers of 61 that can be transferred and polymerized are formed if spread on
water containing metal salts. On the other hand, 62 did not form a stable monolayer even when metal ions were
added to the subphase [281]. The carboxylate-counterion interactions in the model diacetylenic acid have been
studied by in-situ external IR reflection absorption spectroscopy (ERAIR). In this case, photopolymerization of a
long-chain diacetylene monocarboxylic acid, 10,12-pentacosadiynoic acid (DA), on aqueous subphases containing
divalent metal ions such as Ba
2+
(pH 7.7), Cd
2+
(pH 6.8), and Pb
2+
(pH 6.0), was monitored [282]. The changes in
69
the IR band corresponding to the carboxylate antisymmetric stretching vibration during film compression indicated
discrete changes in the coordination / association states of the carboxylate groups. On the basis of ab initio
calculations, it was proposed that during film compression the carboxylate group in the DA monolayer on both Pb
2+

/ Ba
2+
subphases changes its coordination mode from a bridging state to a bidentate one, while on Cd
2+
subphase it
keeps a bidentate coordination state. Interestingly, IR spectral changes in the
as
(COO-) and
s
(COO-) regions
during photopolymerization were similar to that observed during film compression without irradiation, which
indicated that polymerization induces more densely packed states of the carboxylate groups in the monolayers [282].
The amphiphilic diacetylene compound, N,N' -bis(10,12-pentacosadiynyloxycarbonyl-methyl)-dithiooxamide
formed a stable monolayer on a subphase containing Cu (II) ions, but not on pure water. At low surface pressures,
loosely packed islands were formed, whereas a more homogeneous monolayer was inferred through electron
microscopy at high pressures [283]. In monolayers of diacetylenic acid (m=12 and n=8), the area per molecule
during UV-irradiation at a constant pressure depends on both the surface pressure, on the feedback parameters of the
trough control mechanism and also on the vibration characteristics of the trough environment. Indeed, induced
vibration seems to reorganize the monolayer suitable/not suitable to photopolymerization [284]. A decrease in
reactivity for the photochromic transition of the polydiacetylene monolayer (blue to red) with the decrease in
molecular area was observed for the cadmium salt of 10,12-pentacosadiynoic acid, indicating the sensitivity of the
photopolymerization to the molecular arrangement and density [285]. The color transition was reversible at the air-
water interface, but not in the corresponding LB film, which means that the color transitions of the PDA polymer are
due to the structural disorder of side chains attached to the polymer backbone [285]. A similar chromic transition
dependence on the side chain interaction has been observed during in situ reflection spectroscopic investigations of
monolayers of various amphiphilic diacetylenes [286].

3. Mixed monolayers of diacetylene/polydiacetylenes
Analogously to the other polymer materials discussed in the previous sections, the mixed LB film approach
has also been used for processing polydiacetylenes, where the main aim is to understand the topochemical
polymerization characteristics of diacetylene amphiphiles in an environment of builder materials. The stability and
transferability of a non-amphiphilic diacetylene, 1,4-bis(3-quinolyl)buta-1,3-diyne (DQ), were improved by
cospreading with builder materials such as cadmium arachidate [267]. Pure DQ monolayers exhibited a liquid
expanded region with a limiting mean molecular area of 7.5
2
and a collapse pressure of about 40 mN/m. In mixed
monolayers, DQ molecules were squeezed out from the matrix of the builder material. Attempts to photopolymerize
DQ to PDQ in pure and mixed LB films failed, even though DQ is known to photopolymerize rapidly in the bulk
solid, indicating that the packing arrangement of DQ molecules is not suitable for photopolymerization [267]. The
70
possibility of preparing PDQ LB films directly from PDQ monolayers has recently been attempted [287]. From the
analysis of the collapse pressures and the excess free energies of the isotherms obtained for the pure and mixed
monolayers of 10,12-pentacosadiyne-1-N-phenylamide -D-mannopyranoside (MPDA) and 10,12-
pentacosadiynoic acid (PDA), it has been found that MPDA/PDA monolayers have good miscibility [288]. Wang et
al [289] observed a first order plateau phase transition in mixed monolayers of 10,12-pentacosadiynoic acid
(PDA)/(N-(11-O--D-mannopyranosyl-3,6,9-trioxa)undecyl-10,12-pentacosadiynamide (PDTM) monolayers, but
not in any of the pure compounds. An overshot at the beginning of the phase transition plateau has been understood
as the hydrogen bonding between the headgroups of PDA and PDTM. Interestingly, the absorption spectra of the
mixed monolayer with high ratio of PDTM revealed a reduction in the effective conjugation length of the
polydiacetylene backbone which prevented the appearance of the blue film phase [289]. They have also studied a
dual chemical and photo induced polymerization of a ternary mixture of octadecyltrimethoxysilane (C
18
TMS), PDA
and PDTM at various subphase pHs using surface pressure-area isotherms and UV-vis spectral measurements [290].
At pH 1, C
18
TMS formed a long linear polymer network where PDA and PDTM were present as small domains,
which could be polymerized weakly upon UV exposure. At pH 2.6, the weak condensation of C
18
TMS formed short
linear polymers, which caused phase separation between C
18
TMS and PDA/PDTM and the observed strong
photopolymerization of the PDA/PDTM rich domain was similar to the PDA/PDTM binary mixed monolayer. At
pH 4.0, the partially hydrolyzed C
18
TMS was miscible with PDA/PDTM which apparently slowed the
polymerization of PDA/PDT upon UV irradiation, revealing the geometry and kinetic controls of the PDA/PDTM
polymerization in 2D by using the C
18
TMS polymer network [290]. The nature of mixing in Langmuir films of two
amphiphilic diacetylenic acids, tricosa-2,4-diynoic acid (TCDA) and 10,12-pentacosadiynoic acid (PDA) and -
glucoside, dioctadecylglyceryl ether- -glucosides (DGG), has been studied by surface pressure-area isotherms and
BAM measurements [291]. From the free energy calculation made from surface pressure-area isotherm data, it has
been found that a good miscibility was achieved in the mixed monolayer containing TCDA and DGG, but the
components repelled each other in the PDA/DGG mixed monolayer, as it was also evidenced from BAM images
[291]. A complex phase with a 1:1 composition was found in mixed monolayers formed by codepositing two
diacetylenic acids (or a diacetylenic acid mixed with a diacetylenic alcohol), in addition to the phase corresponding
to the individual components [292]. The influence of subphase temperature on the pressure-induced crystallization
of diacetylenic lipids with two alkyl chains (called as Bronco, 60, Fig. 58) was studied by fluorescence and electron
microscopies [293, 294].
71
( CH
2
)
8
) CH
3
(CH
2
)
12
COO(CH
2
)
2
N
CH
3
CH
3
(CH
2
)
8
) CH
3
(CH
2
)
12
COO(CH
2
)
2
+
Br
(60)

Figure-58

The plateau region related to the liquid-solid phase transition in surface pressure-area isotherms disappeared upon
lowering the subphase temperature. The morphological features of the monolayer, as inferred through fluorescence
microscopic studies, were also different, since elongated domains were seen at higher subphase temperatures, but
the domains appeared as two-dimensional plates at low subphase temperatures, and these shapes did not change
upon polymerization [293, 294]. The application of an electric field on the molecules on the water surface prior to
and during deposition also permits control of size and shape of domains [295].

4. Preparation and characteristics of LB films of polydiacetylenes
Different research groups reported distinct types of deposition in multilayer LB films for divalent metal
salts of diynoic acid, which could be Y-type, Z-type and mixed Z-Y type in which a switch-over from Z- to Y-type
occurred after few deposition cycles. The reasons behind such dispersion are still poorly understood, even though
parameters such as the aging of the monolayer, the concentration of spreading solution and the nature of the
substrate seemed to influence the type of deposition [296]. XRD analysis of the Y-type, Z-type and Z-Y-type
transferred LB films indicated a significant difference in the packing of molecules within the LB film, though the Z-
type deposited films are known to rearrange to more stable Y-type films. The polymerization characteristics of these
films also differed, for while a more rapid transformation from blue to red occurred in Y-type deposited films in
comparison to Z-type films, the Z-Y deposited film exhibited a superimposed behavior. As the photochromic
transition is sensitive to the molecular packing, these observations may indicate that the Y-type films arising from
overturning in the Z-type build-up process are structurally quite different from the Y-type films obtained from the
Y-type build-up process [296]. In another study, the monolayer of cadmium salt of 10,12-tricosadiynoic acid was
transferred both by the vertical dipping method (Y-type) and by the horizontal lifting method (X-type). The bilayer
spacings obtained from X-ray diffraction are similar, indicating the overturning of molecules either during the build-
up process or soon after deposition [297]. These films exhibited different photopolymerization characteristics, with
the Y-film being transformed into the blue and then to the red form, and with only the blue form being achieved in
X-type LB films. The latter behavior is similar to that of AB-type multilayer LB films comprising cadmium salts of
a diacetylene acid and arachidic acid [298]. Analogously to the LB films of the divalent metal salt 10,12-
72
pentacosadiynoate, Z-type LB films of tetravalent titanium diynoate undergo photopolymerization upon UV
irradiation. No clear evidence for the layer structure in such an LB film was derived from XRD measurements,
indicating the presence of domains consisting of diacetylene molecules with favorable packing arrangement for
polymerization [299].
The nature of the substrate can also be of importance, as suggested by Shirai et al. [300]. With surface-
enhanced resonance Raman scattering (SERRS) spectroscopy, it has been shown that the photopolymerization of a
one-layer LB film of cadmium salt of PDA on an evaporated silver island film with a mass thickness of 60
proceeds in a way that is similar to that of the multilayer LB films on a smooth silver substrate with a mass thickness
of 1000 . That is to say, there is a transition from the blue to the red phase. The rate of photopolymerization on the
SERS-active substrate, however, is higher than that observed for an LB film (d = 1000 ) on the smooth SERS-
inactive substrate. This has been attributed to an accelerated propagation reaction (an addition of DA to reactive
polydiacetylene oligomers) resulting from an enhanced band-to-band transition of the oligomers caused by a strong
coupling of the transition to the localized plasmon resonance of the silver substrates [300]. Yamamoto et al [301]
reported the generation of defects in PDA LB films on Si substrates, even in the first monolayer, through XPS
measurements. The defects are due to molecular elimination generated selectively on the thickly oxidized area when
the second layer is deposited.
LB films produced from both diacetylenic acid and alcohol also possess similar crystal packing, according
to diffraction measurements, and undergo a similar phase transition during photopolymerization. A noticeable
difference, however, was the appearance of cracks in the acid LB films due to a greater volume change during the
phase transition [270]. Another structural variation, i.e. the position of the diacetylene group in the hydrophobic
chain, has been studied in relation to polymerization of diacetylene in the LB film structure [302]. For diacetylene
amphiphiles in which m=0 and n=17 and X=COOH, i.e., the polar COOH group is directly attached to diacetylene
without any methylene spacer, the LB film of divalent metal salts of the acid did not undergo polymerization. In
contrast, LB films of pure acid and monovalent salts could be polymerized, but these polymers were unstable owing
to decarboxylation reactions [302].
The structure and photopolymerization behavior of LB films consisting of alternating layers of two
different diacetylene amphiphiles, one with acid head group (A) and another with pyridine head group (B), has also
been investigated. A regular layer structure with a bilayer spacing of 55.2 , which increases to 57.9 during
photopolymerization, was observed using small angle X-ray diffraction (SAXR). Different extents of tilting of the
aliphatic tails of A (34) and B (45) were estimated from FTIR measurements. Also inferred was the all -trans
conformation for the aliphatic alkyl chains of both monomeric amp hiphiles. The symmetric stretching of the
methylene groups is in the substrate plane and there is a partial transformation into a more irregular packing with
73
more gauche conformations during photopolymerization [303]. Intermolecular rather than intramolecular
photopolymerization occurred in LB films of amphiphilic -cyclodextrin attached with diacetylene side chains
(Diace--CD) (61, Fig. 59), which led to the immobilization of Diace--CD onto the substrate [304].
O
CH
2
H
H
H
H
H
O
OMe
OMe
NH
O
(CH
2
)
8
(CH
2
)
9
CH
3
7
(61)

Figure-59

The preparation of ordered and polymerizable LB films from an amphiphilic diacetylene attached with an alkyl
substituted carbazole group (62, Fig. 60) was reported by Pedriali et al [305].
N
CH
2
(CH
2
)
8
COOH
(CH
2
)
11
CH
3 CH
3
(CH
2
)
11
(62)

Figure-60

In these Y-type anisotropic LB films (dichroic ratio of about 2), no significant change of bilayer distance (63 ) was
observed during photopolymerization in contrast to the increase of bilayer distance for LB films made from model
diacetylene amphiphiles [305]. The fabrication and surface characteristics of ordered polydiacetylene thin films
based on 6-(2-methyl-4-nitroanilino)-2,4-hexadiyn-1-ol derivatives (DAMNA) have been studied by Wolfe et al
[306]. The fabrication of LB films from a perfluorodiacetylene, C
24
F
42
, was reported by Terasawa et al. [307], who
obtained highly hydrophobic surfaces with a contact angle of 162 for films deposited on KRS-5 plates. A
comparison of the reflection-absorption and transmission IR spectra showed that the difluoromethylene chains in the
74
LB film were oriented perpendicular to the substrate surface. The LB film on the quartz plate could be
photopolymerized yielding a highly electroconductive film without doping or further treatment [307]. The
deposition pressure also influences photopolymerization of the diacetylene amphiphiles. For instance, LB films of
diacetylenic acid with m=14 and n=8, transferred at 45 mN/m, could be polymerized, but no polymerization was
observed for films transferred at 25 mN/m [308]. Grando et al [309] investigated the influence of molecular
weight (for soluble polydiacetylenes) on monolayer characteristics. Monolayer and LB films of low (UV-
polymerized) and high (-ray polymerized) molecular weight soluble poly(3-butoxycarbonylmethylurethanes)
(P3BCMU) were investigated. The phase transition during monolayer compression leads to a larger increase of
order in the polymer chains of -P3BCMU which was further supported by visible and Raman spectra of multilayer
LB films. Though the number of repeating units in both polymers was much larger than the value corresponding to
the optimum effective conjugation length, a better arrangement of the conjugated chains was found in films of
higher molecular weight polymers [309]. Another soluble asymmetrically substituted polydiacetylene studied was
poly(BPOD) (63, Fig. 61) whose monolayer could be transferred as Z-type films. Spectral measurements indicated a
preferential orientation of the main chain along the dipping direction with a mean fluctuation of 30 between the
director axis and the chain axis. Polarization-dependent second-harmonic generation (SHG) in multilayers indicates
a high degree of orientational anisotropy of the backbone, with the increase of SHG intensities with the number of
layers [310, 311].
N
N
(CH
2
)
4
O
O
NHCH
2
COO(CH
2
)
5
CH
3
(63)
n

Figure-61

The Langmuir film of N,N' -bis(10,12-pentacosadiynyloxycarbonyl-methyl)-dithiooxamide on the surface of water
containing Cu
2+
ions could be transferred onto solid substrates as Y-type LB films, whose layer structure was shown
to be irregular in SAXR measurements, while XPS revealed the presence of Cu(II) i ons in the films. A partial
photopolymerization accompanied by partial decomposition was observed while exposing the LB film to UV-light
[283].
75
The fabrication of sensitized mixed LB films in which diacetylene amphiphiles were codeposited with
surface active acridinium and anthraquinone dyes, which are known to sensitize the photopolymerization of
diacetylenes to visible light, was reported by Bubeck et al [312]. The polymer yield obtained by using these dye
sensitizers was similar to that obtained with UV-polymerization. The sensitization process has been explained in
terms of the electron transfer between the sensitizer and the diacetylenes [312]. In the case of LB films produced
from a diacetylenic lipids with two alkyl chains (Bronco, 63, refer to Fig. 60), the crystal structures with varying
alkyl chain tilts, as evidenced from electron diffraction, were found to depend on the subphase temperature and that
led to distinct polymerization characteristics of the LB film [293, 294]. The influence of incorporated subphase
water (during film preparation) on the packing of diacetylene amphiphiles and on polymerization has also been
reported. For instance, exposure of the as-deposited films to high vacuum resulted in an increase of the bilayer
distance and decrease in reactivity for polymerization [280]. Pre-annealing LB films of diacetylenic acids affects
photopolymerization [313,314]. The annealed film exhibited increased intensity of X-ray diffraction peaks and also
polymerized rapidly to a blue-green form (
max
= 700 nm) in comparison to photopolymerization of as-deposited
which was transformed to the blue form (
max
= 640 nm) upon UV-irradiation. Such changes have been attributed to
the change of packing order of molecules upon annealing [314]. The cubic susceptibilities (
(3)
) measured for LB
films of polydiacetylene depended on whether the precursor diacetylene LB film was annealed or not prior to
photopolymerization. The blue polymer obtained by photopolymerization of as-deposited LB films possessed higher
(
(3)
= 1.2 X 10
-10
esu) value than the green polymer obtained upon photopolymerizing annealed diacetylene LB
films (
(3)
= 0.97 X 10
-10
esu). This shows that even though an ordered structure was achieved by annealing, as
inferred through absorption spectral measurements, -electrons seemed to be less delocalized in the green polymer
[315].
Polymerization in diacetylene LB films has been monitored by a combination of analytical techniques. For
instance, the rate and extent of polymerization of DA in LB films of cadmium salt of 10,12-pentacosadiynoic acid
(DA) [316] were monitored by the change in relative intensity of the infrared bands due to CC stretching
vibrations, which were correlated with orientation changes of the alkyl groups of DA. Furthermore , resonance
Raman scattering (RRS) helped elucidate the changes in conjugation length of the PDA backbone in the CC
stretching vibration region. Polymerization seemed to proceed in two steps. In the first step (transition from
colorless to blue), the PDA backbone keeps a fully extended conformation without any interruption of its
conjugation, and the long axis of the alkyl backbone changed from a tilted state to a less tilted one. The axis was
kept perpendicular to the long axis within the substrate surface plane, which explains the increase in interlayer
spacing of LB films of long-chain diacetylene monocarboxylic acids during photopolymerization. In the second
76
step, blue to red, the regular plane of the all-trans alkyl group is converted to an irregular one with gauche
conformations and consequent interruption of the fully extended backbone structure, thus reducing the average
conjugation length [316].
LB films of cadmium salts of diacetylenic acid usually exhibit sharp reflection up to high orders in small
angle X-ray scattering (SAXS) patterns, due to the periodicity of the LB film and to the large difference in electron
density between the hydrophobic portion including the diacetylene group and the cadmium salt. The bilayer
distances obtained for different hydrophobic chain lengths indicated a head-head/tail-tail (Y-type) packing
arrangement of molecules within the LB multilayers. Transmission electron microscopy (TEM), on the other hand,
indicated that the films consisted of domains of uniquely oriented molecules whose diameters are of the order of
several hundred microns. It should be stressed that the individual domains in adjacent layers are not in register with
each other, i.e. there is no columnar like structure along the Z-axis. The similarity between TEM images of such
oriented areas and those observed with polyethylene single crystals points to the zig-zag orientation of the
methylene groups, at least for the two methylene groups close to the diacetylene group. This allows the acetylene
bonds of neighboring molecules to be parallel to each other, in contrast to the all trans conformation of methylene
groups usually observed for simple fatty acid salt films [276, 293, 294]. Seki et al [317] observed a lowering of
ionization threshold energy from 6.7 eV to 5.1 eV during photopolymerization of LB films of tricosa-10,12-diynoic
acid through UV photoelectron spectroscopic (UPS) measurements and such changes have been ascribed to the
formation of delocalized -electrons along the polymer chain, similarly to what was observed in evaporated films.
However, X-ray absorption near-edge structure (XANES) features showed that polymerization did not proceed
regularly in the topmost layer of the LB film, unlike the case of evaporated films [317].
In an interesting study, the photopolymerization behavior of one-layer Langmuir-Blodgett (LB) films of
10,12-pentacosadiyonic acid (PDA) deposited onto a monolayer of poly(vinyl alcohol) containing an azobenzene
side chain (6Az10-PVA) was investigated by UV-visible absorption spectroscopy. The conjugation state of the
resulting polydiacetylene depended on the initial isomerization form of the 6Az10-PVA monolayer beneath the
diacetylene LB film. The trans- and cis- surfaces led to a blue and red colored polydiacetylene film, respectively
[318, 319]. An increase in surface roughness during photopolymerization was observed through atomic force
microscopy while preserving the LB layer structure, as inferred through diffraction measurements [320]. The
thermochromic phase transition behavior from blue to bluish-green LB films of cadmium salt of PDA was
investigated by electron diffraction, ER-FTIR, SAXS, and XPS spectral measurements. At elevated temperatures,
both PDA LB films were converted into the red form and this chromatic phase transition was also accompanied by
the characteristic crystal phase transition according to electron diffraction measurements. Furthermore, the
77
difference in the chromatic phase of PDA LB film could be explained by the difference in the exciton energy level
being strongly dependent on both the crystal symmetry and on the resonance backbone [321].
The LB technique has been exploited in the fabrication of stable non-centrosymmetric Z-type films with
stable polar surfaces. In general, Z-type films are prone to molecular overturning to the thermodynamically more
favorable Y-type. The fabrication of stable Z-type LB films from 10,12-nonocosadiynoyl-S-lysine and 10,12-
nonocosadiynoyl-S-ornithine, both containing diacetylene as well as amide functionalities in the hydrophobic chain,
was reported by Popovitz-Biro et al [322], where the Z-type structure appears to be established by
photopolymerization and H-bonding between adjacent molecules. The formation of stable Z-type LB films was
inferred from the repeat dis tance obtained through low-angle X-ray diffraction, the ellipsometric thickness and
reflection-absorption Fourier transform infra red (RA-FTIR) measurements. These amphiphilic diacetylenes could
be photopolymerized both at the air-water interface and in the LB multilayers, with the extent of polymerization
depending on the position of the H-bonding forming amide groups in the hydrophobic chain [322]. As the
photopolymerization of diacetylene amphiphiles in the LB films proceeds within the same layer, all the molecules
within that layer are expected to be involved. This property has been made use of in preparing LB films with stable
polar surfaces which essentially involved the three following steps: (1) Hydrophobizing the hydrophilic glass
substrate with an odd number of diacetylenic fatty acid multilayers; (2) Subsequent transfer of an even number of
amphiphilic diacetylenic fatty acids with polar groups, COOH, NH
2
and OH, and at this stage the substrate is inside
the water; (3) Coating the LB film formed by a protective monolayer of stearic acid while lifting the film to the air.
After stabilizing the LB film by UV-polymerization, the top coating of stearic acid can be removed by treatment
with a basic solution, and one ends up with the LB films possessing a stable polar surface [323]. Attempts to prepare
LB films with polar surfaces, by simple control of the deposition cycle, usually lead to non-uniform surfaces due to
the repacking of molecules in the top most layer [146].
The impact of polymerization and preparation conditions of LB films of PDA on the surface properties of
the layers was investigated through direct force and contact angle measurements [324]. Humidity-dependent
variations both in the water contact angles and in the directly measured interfacial energies of these layers were used
to assess film stability under various environmental conditions. Direct force measurements demonstrated that
polymerization prevents molecular reorientation and consequent changes in the interfacial properties of the polymer
films. Also, the polymerized diacetylene layers were stable to repeated subjection to large compressive loads for
several hours. A major limitation, however, is the difficulty of forming homogeneous, defect-free films over large
areas [324]. The mechanism of the thermochromic blue to red color transition of Langmuir-Schaefer films of 10,12-
pentacosadiynoic acid functionalized with a sialic acid lipid was investigated using atomic force microscopy, visible
absorption and Fourier transform IR spectroscopy [325, 326]. The blue films seemed to consist of micro -sized
78
domains with stripe-like morphology on top of a nearly complete layer. The thermochromic transition temperature
was between 70 and 90. At the molecular level, the film order increased at the thermochromic transition and
remained the same for temperatures well above the transition (e.g., 130). No evidence of the previously suggested
entanglement or disordering of the alkyl side chains could be observed. Apparently, the pendent side chains
rearrange from a partially disordered configuration characteristic of the blue film, to a well -ordered close-packed
hexagonal arrangement in the red form [325, 326]. AFM revealed that indium-tin oxide (ITO) coated glass
substrates had their surface corrugation decreased when coated with polymerizable diacetylenic lecithin LB films
[327]. The orientation of a nematic mono-domain based on these substrates was changed from homeotropic to tilted
upon UV irradiation. Unique star-like features were found on the monolayer LB film, indicating that the
conformation of the diacetylene chains changed during photopolymerization, thus causing the liquid crystal to
undergo a transition from homeotropic to a tilted alignment [327]. Smoothening of the surface of LB films of PDA
deposited on glass substrates upon UV-polymerization was inferred through X-ray reflectivity measurements [328].
Using electron paramagnetic resonance (EPR), Takamura et al. [329] observed spin magnetism in LB films of
manganese salt of polydiacetylenic acid. The monomer PDA LB film seemed to possess 1D spin chain interaction,
but the polymerized PDA LB films exhibited a drastic extinction of the EPR signal due to the onset of
antiferromagnetic coupling owing to the decrease in the distance between Mn ions.

5. Possible applications of polydiacetylene LB films
The usefulness of LB films of diacetylene/polydiacetylene as synthetic membranes has been demonstrated
with diacetylene/polydiacetylene coated polypropylene support in which a reduction by a factor of 30 was observed
in the gas flow compared to the bare polypropylene support [330]. Gas sensors from LB films of diacetylene
derivatives of amides and pyridine were produced based on the reversible adsorption of SO
2
and H
2
S, as deduced
from ellipsometric measurements [331]. The amount of adsorbed gas was proportional to the relative amount of
receptor-sites and number of layers in the film, which led to a model of layer-by-layer adsorption of gases by the LB
film [331]. The fabrication of composite gas-separating membranes by deposition of multibilayer diacetylene LB
films onto a poly(dimethylsiloxane)-polycarbonate membrane was also reported [332], where the LB modification
seemed to increase both membrane permeability to O
2
and He gases and cause dispersion of local permeability over
the membrane surface. In particular, coating with PDA LB films increased the selectivity of O
2
/N
2
and He/N
2

separation by 1.5 and 1.2 times, respectively [332]. The concept of a constructive addition of nonlinear-nonlinear
optical materials, in which the NLO response of an embedded quantum-dot semiconductor (CdS) can add
constructively to that of a NLO polymer (PDA) having complementary wavelength characteristics, was
demonstrated by Schwerzel et al. [333]. The nonlinear refractive index (n
2
) of a nanocomposite PDA LB film
79
containing thiophenol-capped CdS nanocrystals, as measured by Z-scan and degenerate 4-wave mixing (DFWM)
techniques, is 11 X 10
-8
cm
2
/MW, while that of an undoped PDA film is only 3 X 10
-8
cm
2
/MW [333]. Cheong et al
[334] estimated the second-order nonlinear coefficient (d33) to be 1.52 pm/V at 1064, after correcting for
absorption, for Z-type deposited poly(BPOD) LB films. The change in optical transmission at different wavelengths
which depends on the state of polymerization offers the possibility of using these LB films as a radiation dosimeter
in the dose range of interest to the radiation processing industry [335]. Photopolymerization of diacetylene
ammonium lipid LB monolayers deposited on plastic optical fibers (POF) [336] decreased the optical intensity of
the light from a light emitting diode traversing the film, which levels off after 90 s. The decrease was attributed to
changes in the refractive index and color phases as polydiacetylene is formed [336]. LB films of the cell membrane
receptor (monosialioganglioside, GM1) mixed with diacetylenic lipids, which undergo blue-red transition upon
exposure to cholera toxin, have been used as colorimetric biosensors [337]. Both monolayer stability and
transferability of GM1 were enhanced upon mixing with diacetylenic lipids and the colorimetric response of the LB
film could be optimized through modification of the ratio of receptor GM1 to diacetylene [337]. Cheng et al [338]
reported the utilization of a protein conformational change to activate a colorimetric solid-state sensor, using a lipid
polydiacetylene LB thin film immobilized with hexokinase. The conformational change caused by binding glucose
at the active site has been found to induce colorimetric changes in polydiacetylene [338].

III. Self-assembled polymer films
3.1. The Self-assembly technique
The search for new materials for device applications has been a strong motivation in the development of
fabrication methods. The need to build multilayers arises because the adsorption of a single layer hardly reaches the
required device thickness. For a long time the Langmuir-Blodgett (LB) technique was the only one to provide
thickness and architecture control while building organic multilayers. The self-assembly (SA) method appeared as
an alternative, initially based on the chemical adsorption of organic molecules with specific functional groups (see,
for example, [16, 339-342]). In order to obtain multilayers it was necessary to chemically adsorb a layer onto an
already deposited layer, which required the synthesis of molecules with adequate affinity. This is not a
straightforward task, which then imposes a severe limitation in the use of this self-assembly method. Decher's
approach [343-345] circumvented this problem, by producing films out of the electrostatic attraction between
oppositely charged layers, i.e. by adsorbing alternately anionic and cationic polyelectrolytes on solid supports [343-
345]. The formation of these multilayers is based on the electrostatic interactions and on the tendency of pairs of
80
oppositely charged polyelectrolytes to form complexes [346]. It must be mentioned that a similar technique was first
proposed by Iler in 1966 for colloidal particles [347]. An advantage of the layer-by-layer technique is that the
adsorption processes are independent of the substrate size and topology, with film thickness increasing linearly with
the number of adsorbed bilayers [343-345]. Finally, the fabrication of layer-by-layer films is quite environmentally
safe since is employs aqueous solutions. Another advantage lies in the possibility of building up heterostructures
with different materials with a specified functionality. The potential of this method may be illustrated by the wide
range of materials that have been employed, such as polyelectrolytes [343-346], conducting polymers [348-358],
polymers containing dyes [359-362], biological materials [363-366], ceramics [367-368] and colloids. A summary
of several materials used is shown in Tables I and II.
In the layer-by-layer or self-assembly (SA) method, the substrate is immersed in a polycationic or
polyanionic solution for a given period of time (usually a few minutes). The substrate is then washed in a solution of
approximately the same pH of the polymeric solutions, in order to remove non-adsorbed molecules, and usually
dried as well. The substrate+film system is now charged which allows adsorption of an oppositely charged layer
upon immersion in the other polymeric solution. Following adsorption of this second layer, the substrate is again
washed and dried. Multilayer structures can then be built by repeating the steps above as long as required.
Apparently, there are no limitations for the number of bilayers that can be deposited; films with hundreds of layers
have been reported [355]. In figure 62 a schematic diagram of the layer-by-layer process is shown.
+
+
+
+ +
+
+
+
+
+
+ +
+
+
+
+
+
+
+
+
+
+
+
-
-
-
-
-
-
-
-
-
-
-
- -
-
-
-
-
-
-
+
-
-
-
-
-
-
-
-
-
-
-
-
-
-
-
-
-
-
+
+ +
+
+
+
+
+
+
+
+
+
+
+
+
+
+
+
+
+
+
+
+
+
+
+
+
+
+
+
+
+
-
-
-
-
-
-
-
-
-
-
-
- -
-
-
-
-
-
-
+
-
-
-
-
-
-
-
-
-
-
-
-
-
-
-
-
-
-
+
+ +
+
+
+
+
+
+
+
+
+
+
+
+
+
+
+
+
+
+
+
+
+
+
+
+
+
+
+
+
+
-
-
-
-
-
-
-
-
-
-
-
-
-
-
-
-
- -
-
-
-
-
-
-
-
-
-
-
-
-
-
-
-
-
+
-
-
-
-
-
-
-
-
-
-
-
-
-
-
-
-
-
-
+
+ +
+
+
+
+
+
+
+
+
+
+
+
+
+
+
+
+
+
+
+
+
+
+
+
+
+
+
+
+
+
-
-
-
-
-
-
-
-
-
-
-
-
-
-
-
-
-
- -
-
-
-
-
-
- -
- -
-
-

A
B C D E

Figure- 62

Several substrates such as hydrophilic and hydrophobic glass, quartz, mica, silicon and gold have been used
in the preparation of layer-by-layer films. In addition, efforts are being made by MacCarthy and collaborators to
modify the surface of organic polymeric substrates. Poly(4-methyl-1-pentene)(PMP) was oxidized [385] and a
carboxylic acid-functionalized PMP surface (PMP-CO2
-
) was obtained. Poly(ethylene terephthalate) (PET) was
hydrolyzed (PET-CO2
-
) and amidated (PET-NH3
+
) [386]. A functionalized alcohol was introduced to treat
poly(chlorotrifluorethylene) (PCTFE) [387] and poly(tetrafluorethylene) (PTFE) was modified with allylamine
81
plasma polymerization [388]. POMA/PVS layer-by-layer films were prepared onto tetrafluorethylene propylene
[355] without surface treatment. Poly(4-methyl-1-pentene) [385] was used in the study of gas permeability.
Colloidal particles as negatively charged polystyrene latex particles were also used to prepare layer-by-layer films
[372, 373]. The catalase enzyme crystal was used as substrate for adsorption of layer-by-layer films in order to
obtain enzyme encapsulation [374]. Figure 63 shows the process of enzyme encapsulating by layer-by-layer
technique film production. The enzyme release is obtained by rupturing the polymer capsule by exposure to
solutions which decomposes the enzyme [374].

Figure - 63

The most important feature of layer-by-layer films is the linear increase in thickness or in the adsorbed
amount per unit area with the number of bilayers. Figure 64 shows the linear buildup of a layer-by-layer film [392].
However, the thickness or the adsorbed amount usually does not increase linearly from zero [343], as one might
have expected. Indeed, it is now well established that substrate effects are important for the first few layers, leading
to a non-linear build up in the initial stages of adsorption indicating the different types of interaction between the
substrate and polyelectrolytes. Such interactions can be weakly or strongly attractive [355,393], and deviation from
the linear behavior may also be due to the time of adsorption allowed for each layer. Also, when layer-by-layer films
grow linearly, as for the POMA/PVS system [355], the adsorbed amount or thickness per bilayer is substrate
independent. We have analyzed this feature very carefully for the POMA/PVS system[355].
82

Figure - 64

Fig. 65 shows the buildup of POMA/PVS films on several substrates [355]. Glass I and glass II refer to
glass substrates that have been submitted to different preparation procedures prior to adsorption. It can been seen
that deviations from the linear behavior occurred for the first few layers, especially for the Teflon FEP substrate.
The latter presented the least stable SA films owing to lack of adhesion. In addition, it is possible that the 3 min. of
immersion employed was far below the necessary for Teflon. In a subsequent investigation for POMA/PVS on
glass, the amount of adsorbed POMA was found to increase exponentially in the first few layers and only increase
linearly after the sixth or seventh bilayer as is shown in fig. 64 [392]. This was attributed to weak attractive
interactions between the substrate and the polymer, which could be demonstrated by the shift in the UV-vis.
absorption peak as the adsorption of the first POMA layer proceeded. The interaction with the substrate caused
POMA to be dedoped thus shifting the absorption curve to lower wavelengths due to a mechanism involving H-
bonding as it will be discussed later on. The increase in the amount of adsorbed POMA with the number of layers is
explained by the increase in adsorption sites when some polymer is already adsorbed. Such an observation was
confirmed in atomic force microscopy (AFM) measurements that showed an increase in film roughness as additional
layers were adsorbed. Fig. 66 shows that the mean roughness increases for the first few bilayers, until reaching a
constant value, in much the same way as the amount of POMA adsorbed per layer [355, 393-395].
83
0
0.1
0.2
0.3
0.4
0 5 10 15 20
Glass I
Glass II
ITO
Teflon FEP
M
a
x
i
m
u
m

A
b
s
o
r
b
a
n
c
e
Number of Bilayers

Figure - 65
0
1
2
3
4
5
6
7
0
1
2
3
4
5
6
7
8
0 2 4 6 8 10 12
A
d
s
o
r
b
e
d

A
m
o
u
n
t

(
m
g
/
m
2
)

M
e
a
n

R
o
u
g
h
n
e
s
s

(
n
m
)
Number of Bilayers

Figure - 66

An opposite behavior was observed in layer-by-layer films from poly(o-methoxyaniline) alternated with
poly(thiophene acetic acid) (POMA/PTAA). Interestingly, the amount of material adsorbed is very large in the first
bilayers but then it becomes constant, leading to a linear increase from the fourth layer (2 bilayers) on [393]. The
reason for this lies in the strong attractive interaction between the glass substrate and PTAA molecules. Similar
results were obtained by Ferreira et al [348] in a similar polymer system. Owing to the importance of substrate
effects for the first few bilayers, results in the literature now tend to fall in these categories of non-linear buildup,
84
with either a superlinear or a sublinear increase in the amount of material adsorbed with the number of layers, for
the first few bilayers. A linear increase from zero has become scarce but was already observed by Decher et al [343]
in films of PSS/poly-4-vinylbenzyl-(N,N-diethyl-N-methyl-) ammonium iodide.
The buildup of multilayers has so far been discussed with no mention to the time allowed for adsorption,
whether it was sufficient for the formation of a complete layer or not. In fact, a linear build up may be observed even
when the time allowed is not sufficient for saturation of the adsorption process for a given layer. This occurs
provided that the period of adsorption is maintained constant for all layers, as illustrated in Figures 65 [392] and 66
[355]. Similar results were obtained for self-assembled films of ammonium octamolybdate ((NH4)4[Mo8O26]) and
poly(allylaminehydrochloride)(PAH) [373]. The amount of adsorbed material was estimated using a quartz crystal
microbalance whose frequency shift was plotted against the immersion time. The time period of adsorption of
((NH4)4[Mo8O26] is varied for the three series of experiments as indicated in the figure, with PAH being adsorbed
for 15 minutes in all experiments [373]. A similar behavior was found for POMA/PVS films where saturation in the
adsorption process takes approximately 2 hours, but a linear build up (after few bilayers) of the multilayer structure
may be obtained using a constant immersion time of only three minutes for each PVS layer.
The layer-by-layer method was conceived for the fabrication of organized structures, which is more likely
to be achieved in the perpendicular direction to the substrate, owing to the layered structure [343-345]. Nevertheless,
some ordering in the plane may also be obtained if adequate materials are employed, as demonstrated by Lvov [360]
for polymers with side chain nonlinear optical dyes. They obtained orientation of chromophore side-chain groups of
the polymer poly[1-[4-(3-carboxyl-4-hydroxyphenylazo)benzene sulfonamide]-1,2-ethenediyl, Na salt] (PAZO)
alternated with poly(dimethyldiallylammonium chloride) (PDDA), which was confirmed by second harmonic
generation (SHG) measurements. Maximum orientation was obtained for 4 PDDA/PAZO bilayers. Another example
of molecular organization was obtained by He et al [361] who produced layer-by-layer films of purple membrane
(PM), isolated from Halobacterial halobium (R1M1), and poly(dimethyldiallylammonium chloride) (PDAC). Using
second harmo nic generation, the latter authors measured the second-order susceptibility of PDAC/PM layer-by-layer
films (12 bilayers) and obtained a larger value compared with electrophoretically sedimented films [361].
Structural investigations of layer-by-layer films have also been carried out. Lsche et al [396] showed that
the internal structure of PSS/PAH layer-by-layer films consists of stratified structures with the individual layers
interdigitating one another intimately. In systems containing polyanilines such interdigitation may be even more
pronounced, as demonstrated by AFM studies for POMA/PVS where the surface roughness for a thick film was
twice the thickness per bilayer [393, 394, 397].
Recently Ladam et al [398] from results of streaming potential measurements (SPM) and scanning angle
reflectometry (SAR) showed that the PAH/PSS layer-by-layer could be schematically represented, as is shown if
85
figure 67, by three zones: (I) a precursor zone corresponding to the layers near the substrate which present different
behavior compared with the film bulk; (II) The core zone or bulk zone which presents a zwitterionic nature; and (III)
the outer zone where the excess charge is located. It must be pointed out that the results which confirm this model
were obtained with constant adsorption times per layer.

Figure - 67

3.2 Physical adsorption of polymers
The fabrication of layer-by-layer films is based on polyelectrolyte adsorption from solid-liquid interfaces
and consequently the formation of each layer is a particular case of polyelectrolyte adsorption onto a solid substrate.
To treat these phenomena theoretically, one must take into account studies over the last four decades, as the first
systematic study appeared in 1965 [399]. It is now clear that adsorption is the final result from competition among
several interactions that include polymer/substrate, polymer/polymer, polymer/solvent and solvent/substrate
interactions. Furthermore, polymer molecules may adopt a completely different conformation when adsorbed as
compared to that in solution. Consequently, as far as experimental conditions are concerned, the adsorption process
depends on the concentration, pH and ionic strength of the polymeric solutions, polyelectrolyte charge density, type
of substrate and immersion time. Establishing meaningful adsorption theories is made difficulty by the complexity
of the molecules involved and the dependency on so many parameters. Several statistical theories have been
developed (see for example [400-402], but they are not easily applicable to adsorption studies. One may also
analyze separate aspects of adsorption in a phenomenological way in order to gain insights into the physical
86
mechanisms involved. Adsorption mechanisms that occur in each layer, such as nucleation and growth controlled by
diffusion, could be inferred by investigating adsorption and desorption kinetics. Upon obtaining adsorption
isotherms, i.e. plots of the adsorbed amount versus the equilibrium solution concentration, one may estimate
adsorption energies using adequate analytical models, especially if experiments are performed at various
temperatures. This could be done, however, by using extremely simplified models of adsorption while interpreting
the data.

1. Kinetics of adsorption
For each pair of polyelectrolyte solution and substrate one can obtain a time function of the amount
adsorbed (or thickness) of a given layer. Full coverage of the substrate for the first layer, for instance, may take
distinct immersion periods depending on the substrate, on the material to be adsorbed and on the solution
concentration, pH and ionic strength. From the point of view of the physical mechanisms, establishing the kinetics
of adsorption is obviously illustrative of the types of interaction that prevail in the adsorption processes. Ferreira and
Rubner [350] were the first to investigate systematically the kinetics of adsorption in layer-by-layer films, where a
number of experiments were conducted for the poly(thiophene-3-acetic acid) (PTAA) and PAH system with various
concentrations and types of substrate. Figure 68 shows their data for adsorption of PTAA on 5 PTAA/PAH bilayers
deposited on a positively charged substrate for various PTAA solution concentrations. Light absorption was
measured at 420 nm [350], corresponding to the absorption peak for PTAA. The adsorbed amount increases rapidly
with time, thus indicating that PTAA has a high affinity for the film-covered substrate, with the adsorbed amount
reaching 72%, 78% and 80% of the final adsorbed amount value, respectively for concentrations of 10
-4
M, 10
-3
M
and 10
-2
M, within the first minute of adsorption. A diffusion-controlled adsorption process was suggested to explain
the initial stages of adsorption, but the effective diffusion coefficients varied widely with the concentration. This
made Ferreira and Rubner [350] conclude that the Langmuir-Schaefer relationship is not valid for PTAA in the time
regime accessible by UV-vis experiments.
87

Figure - 68

The kinetics of adsorption may also depend on the experimental conditions. In fact, it was observed that if
the adsorption process is interrupted several times (e.g. for measuring absorption), the adsorbed amount increases
substantially. This is illustrated in figure 69 where the amount of adsorbed POMA onto hydrophilized glass is shown
as a function of the substrate immersion time. In curve I the substrate was immersed a few times in the solution
while in curve II a larger number of immersions were performed. Similar results were obtained with PPV kinetics
curves [403]. One can therefore conclude that in certain polymeric solutions the kinetics curves can be drastically
influenced by the number of immersions of the substrate. For polyanilines, such a dependency was associated with
hydrogen bonding between POMA molecules, which is consistent with the findings of Stockton and Rubner [354]
who stated that hydrogen bonding is the driving force for adsorption of emeraldine base polyaniline. Raposo and
Oliveira [394] showed that both electrostatic and hydrogen bonding are present in POMA emeraldine salt. Hence,
true kinetics curves must be obtained by immersing the substrate only once in the polyelectrolyte solution. From
the kinetics results of figure 69 it was also observed that the adsorbed amount follows the empirical equation:

,
_

,
_

,
_

,
_

,
_


n
2
2
1
1
t
exp 1 k
t
exp 1 k A (1)
88
where A is the absorbance at the peak which is proportional to the POMA adsorbed amount per unit area, k1 and k2
are constants, 1 and 2 are the characteristic times and n is a constant. Therefore, POMA adsorption may be
attributed to two types of process: an initial, fast first order kinetics with characteristic times of a few seconds,
followed by a much slower process that is fitted with a Johnson-Mehl-Avrami function with much higher
characteristic times (hundreds of seconds). The latter function is generally believed to occur in processes of
diffusion; when 1.5<n<2.5 the process occurs with a diffusion-controlled growth of domains and decreasing
nucleation rate. For n=1.5 the process is due to diffusion controlled growth of domains and zero nucleation rate.
Then, the first process would correspond to nucleation while in the second process there is only growth of domains.
This hypothesis was corroborated by atomic force microscopy (AFM) measurements [355], which showed that
within a few seconds of immersion, nuclei appeared all over the sample, and for longer immersion times larger
domains were observed. Also, Gabrielli et al [404] showed that the adsorption kinetics of the anionic surfactant and
a non-ionic polymer follows the Avrami law with n=1.5. The authors justified this value by a two dimension growth
(on the surface) limited by diffusion. An interesting result was obtained from analysis of AFM micrographs results
of POMA/PVS [395] where a growth exponent of 1.5 was estimated for the buildup of POMA/PVS multilayers.
0
0.01
0.02
0.03
0.04
0.05
0.06
0 100 200 300 400 500 600
M
a
x
i
m
u
m

A
b
s
o
r
b
a
n
c
e
Immersion Time [s]
I
II
III

Figure - 69

As for the kinetics of a POMA layer onto an already formed POMA/PVS film, it was originally thought
that adsorption would become faster as the number of bilayers increased. However, this conclusion was based on
experiments in which the substrate was immersed several times in the polymeric solution for the data points to be
obtained. When the true kinetics was measured, i.e. with substrates immersed only once in the polymeric solution
for that given layer, it was noticed that the kinetics of adsorption of the nth layer also followed Eq. (1) [355]. The
89
only major difference was that the characteristic time for the second process (domain growth) was much longer
because a larger quantity of material was adsorbed in the nth layer, as will be discussed later.
A two-stage adsorption process was also proposed by Tsukruk et al [405] for layer-by-layer films of poly-
(styrenesulfonate) (PSS) and poly(allylamine) (PAA). They based this proposal on surface morphology changes.
During the first stage (1-2 minutes) of film formation the polymer chains are inhomogeneously adsorbed on selected
sites of the substrate (in this case an oppositely charged self-assembled monolayer), which had a high local charge
concentration as in scratches, holes, edges and foreign microparticles. At the second stage, for longer deposition
times (10-30 minutes) the polymer islands are gradually spread out and the roughness decreases. This feature was
interpreted as caused by the high charge concentration in macromolecular chains [405]. In the case of POMA/PVS
samples the sample roughness increased with the number of bilayers, tending to a constant value [394]. Even though
Ferreira and Rubner [350] did not mention the possibility of a two-stage kinetics process, a close inspection in their
data (Figure 68) points to a small plateau starting at approximately 2 minutes of adsorption, which could also be
interpreted as the appearance of a second adsorption process [393].
There are reports in the literature, however, in which only one stage is observed for adsorption. For
example, Ichinose et al [373] studied the kinetics of adsorption of ammonium octamolybdate ((NH4)4[Mo8O26])
onto films of ((NH4)4[Mo8O26])/PAH by plotting the quartz crystal microbalance frequency shift as a function of
the immersion time. In contrast to the examples above no saturation occurred for the observed period of time (30
minutes). The adsorbed amount for a very short adsorption time (say 5 s) is proportional to the polyelectrolyte
concentration. For POMA this initial adsorbed amount is practically equal to the equilibrium value, thus indicating
that the kinetics curve for higher polyelectrolyte concentrations is controlled by the first order kinetics equation,
which is consistent with the simple Langmuir model to be described in the next section.

2. Adsorption Isotherms
The concentration of the polymeric solution is an important parameter to be controlled in the fabrication of
layer-by-layer films. By varying the concentration, one may draw the so-called adsorption isotherms which must be
obtained in the stationary regime of adsorption, i.e., when the kinetics of adsorption curves are time independent.
The determination of adsorption isotherms is important for two main reasons. First, the amount of adsorbed material
depends on the polymer concentration in the solution. In addition, when models are applied to the adsorption
isotherms results, one may get information on the interaction forces involved in the stability of the adsorbed layers.
Analytical models for the liquid-solid adsorption can be employed, the simplest of which is the well -known
Langmuir model, with the following equilibrium equation:
90

( )

des ads
k 1 x k
dt
d
(2)
where is the surface coverage (the amount at a given concentration and the maximum amount ratio), x is the molar
fraction of polyelectrolyte monomers in solution which is assumed to be proportional to the polyelectrolyte solution
concentration. It must be emphasized that the adsorption models assume equilibrium, which is not strictly valid for
the experiments. However, the experimentally measured adsorbed amount per unit area in a metastable-equilibrium
state can still be treated theoretically according to Pan and Liss [406] provided that the ideal equilibrium constant is
taken to be higher than the real equilibrium constant:

( )
( ) kx 1
kx
x
+
(3)
where k is the equilibrium adsorption constant which is associated with the adsorption energy.
The Langmuir model was used to fit the adsorption of POMA onto glass (Fig 70) for concentrations lower
than 0.25 g/L and the thermodynamic parameters were estimated [358]. The adsorption energy was ~-35 kJ/mol, the
enthalpy of adsorption was 202 kJ/mol, while the entropy of adsorption was 183 J/Kmol. From these values one
could conclude that POMA adsorption onto glass is controlled by entropy, that is, the increase in entropy in the
solvent makes the polymer adsorption favorable, in spite of the positive enthalpy. It must be stressed that a positive
adsorption enthalpy does not mean that the interactions between the substrate and the solute are repulsive. It simply
means that energy must be supplied for adsorption to occur. By analyzing substrate effects it was concluded that
such energy is spent in breaking hydrogen bonds between substrate sites and water molecules that are immediately
adsorbed upon contact of the substrate with the aqueous solution [394]. One of the consequences of these substrate
effects is that the polymer adsorbed in the initial stages is less doped than in solution [355]. POMA at pH=3 is 50%
positively charged and has an absorbance peak at about 720 nm. When POMA loses some of its charge the peak
shifts toward lower wavelengths. This occurs during the nucleation phase, but the peak returns to higher
wavelengths after 200 s of immersion time, when the polymer being adsorbed probably does not interact with the
substrate directly.
91
0
0.5
1
1.5
2
2.5
3
3.5
0 0.05 0.1 0.15 0.2 0.25
A
d
s
o
r
b
e
d

A
m
o
u
n
t

(
m
g
/
m
2
)
Concentration (g/L)

Figure 70

It must be pointed out that the Langmuir simple model was developed for adsorption of non-interacting
hard spheres of negligible size, which are premises that are not fulfilled by polymer systems. In addition, a round
curve such as that of Fig. 70 could be attributed to heterodispersity of the polymer [407], where the area per volume
ratio should also influence the isotherm [400, 407]. For the POMA experiments, however, this ratio was rather low
and therefore the isotherms were still expected to be round. Since experimental conditions such as the extent of
POMA aggregation in solution could affect the shape and temperature dependence, some authors take the view
[408] that the energy values extracted from the fittings with Langmuir isotherms are meaningless. Owing to the lack
of theoretical models that can be applied to complex systems such as heterodisperse polymers, a number of other
authors still employ the Langmuir isotherm fittings in order to get estimates of the energies involved [409]. In
particular, such analysis allowed us to identify the substrate influence on the adsorption process [358]. For an
opposite trend in the temperature dependence was observed for the adsorption isotherms when an nth POMA layer
was deposited on an already formed POMA/PVS film, with smaller adsorbed amounts at higher temperatures and
low concentrations. In contrast to the first POMA layer, a negative enthalpy was obtained which demonstrates that
substrate effects are no longer important and that adsorption on sites containing some polymer is favored, consistent
with the increase in the amount of adsorbed material with the number of bilayers deposited [392]. Nevertheless,
before independent calorimetric measurements of enthalpy are made, experimenters should be aware of the strong
reservations expressed by Fleer [407] and by Senger et al [410] in the analysis, even if phenomenological, of their
adsorption data. In fact the Langmuir model cannot account for experimental adsorption data over a long time scale
[references within 410] and the first order exponential Langmuir adsorption kinetics cannot explain the two
92
processes observed for POMA. In conclusion, the Langmuir model must be used carefully while estimating
adsorption energy values.
The adsorption isotherms can be easily used to identify possible interactions other than ionic ones in layer-
by-layer films, since if only ionic interactions exist the adsorbed amount per layer must be independent of the
solution concentration and should be the same for each and every layer deposited. However, a clear dependence on
concentration was observed for POMA self-assembled films that were obtained with long periods of immersion (ca.
2 hours). For the true kinetics, where substrates with bilayers already deposited were immersed only once in the
POMA solution, 2 hours of immersion led to an adsorbed amount close to equilibrium [394]. Fig. 71 shows
adsorption isotherms at 25C of POMA onto self-assembled POMA/PVS films with various numbers of bilayers.
The already deposited layers were obtained with a POMA concentration of 0.13 0.01 g/L. Each point in the
isotherm for the various solution concentrations employed for obtaining the top POMA layer was the average from
four distinct self-assembled films. The analysis of adsorption processes in heterodisperse polymers such as POMA is
complicated since theoretical models cannot deal with complex systems like this, as mentioned above. We had to
resort therefore to simple models, only to find that despite the crude limitations in its derivation the Langmuir
isotherm (Eq. (1)) can account for the data obtained for POMA, reflecting an adsorption based on dispersion forces,
hydrogen or hydrophobic bonding [407]. The solid lines in Fig. 71 represent the fittings; the equilibrium adsorption
constant and the corresponding free energy of adsorption were also calculated taking into account that the Langmuir
model gives an average free energy that is independent of the solution concentration (~-34 kJ/mole). This free
energy is practically independent of the number of bilayers. It should be noted that for bare glass substrates at 25C
a Langmuir isotherm can explain only the experiments at low concentrations [358].
93

Figure - 71

The POMA adsorbed amount increases with the number of bilayers until reaching a plateau at 25 mg/m
2
.
This increase can be interpreted as follows. After the adsorption of a first layer, which is essentially controlled by
nuclei formation followed by a diffusion-controlled growth [355], subsequent layers tend to adsorb preferentially on
the sites where some polymer was adsorbed. Hence, surface roughness is expected to increase which was confirmed
in AFM studies, with the mean roughness varying from 1.4 nm for the first POMA layer to 5.6 nm for the 11th
POMA layer. After several bilayers the roughness becomes constant, which coincides with the plateau being
reached in the adsorption isotherm [393].
The Langmuir isotherm does not take into account interactions between adsorbed particles and the Gibbs
energy of adsorption is independent of the surface coverage. The simplest model that takes into account interactions
with the adsorbate is the Frumkin model [412], which assumes that the interactions are proportional to the surface
coverage, and is represented by:

,
_

RT RT
G
exp x
1
F
0
ads
(4)
where

is the surface coverage, x is the solute monomer molar fraction,


0
ads
G is an energy component that is
independent of the amount already adsorbed, R is the gas constant, and T is the absolute temperature.
F
is the
interaction parameter that is proportional to the surface coverage, which can be positive when the adsorbed particles
94
are repelled or negative if they attract each other. When
F


= 0 this isotherm is reduced to the Langmuir isotherm.


F
is positive for a POMA layer adsorbed on a bare hydrophilized glass substrate (0 bilayers), which indicates
repulsive interaction. When the number of POMA/PVS bilayers increases,
F
becomes negative. When the
activation energy is plotted against the number of bilayers already deposited on the substrate, the energy increases
much in the same way as the amount of material adsorbed. This implies that the activation energy does depend on
film roughness, as the latter also follows the behavior of the adsorbed amount when the numb er of layers is varied
[392]. The equilibrium constant depends on the adsorbed amount already adsorbed, leading to a concentration-
dependent energy which is more realistic.
From the analytical models available in the literature, Filippovas model [413] is perhaps the one which
reflects more closely the experimental conditions for adsorption of a polyelectrolyte on planar surfaces since it
assumes a local quasi-equilibrium with the ion concentration at the outer boundary of the double layer. In this
model, Fillippova employs the rate equation relating adsorption and desorption processes:
( )
( ) ( ) [ ] ( ) ( )
( ) [ ] n i 1 , t 1
2
z
exp K
t z t
2
z
exp t 1 t , 0 c K
dt
t d
2
i i
i
des
i
0 i
2
i i
i
ad
i
i

'


,
_

1
]
1

,
_


(5)
where
i
is the surface coverage for the specie I (polyelectrolyte),
ad
i
K and
des
i
K are the rate constants for
adsorption (ad) and desorption (des) reaction, respectively, for the i-th species, ( ) t , 0 c
i
is the surface concentration
of polyelectrolytes,

is the total surface coverage for the mixture of ions,


0
(t) is the electrical surface potential
which depends on the adsorption,
i
z is the charge number (valence) and ( ) RT / H
i i
where ( )
i
H is
the activation energy of adsorption for the I-th species, characterizing interactions between polyelectrolyte/interface,
polyelectrolyte/polyelectrolyte and polyelectrolyte/solvent. Considering non-symmetrical polyelectrolytes, i.e. with
two types of ion, with electrical charges
z1 > >
and
z2
, and assuming the amount of polyelectrolyte adsorbed to be much
larger than the amount of adsorbed ions (
o1 > >
), which is consistent with Hsieh et al [388] and Caruso et al [390], who
showed that only trace amounts of small counterions are detected in the layers, Filippova obtained an adsorption
isotherm in the form:

( ) ( )
1
]
1

,
_


1
1 2 1
1 1 1
2 1
4
exp

C r (6)
95
where

1
is the ratio between the polyelectrolyte adsorbed amount and the total adsorbed amount, C
1
is the
dimensionless polyelectrolyte concentration in the bulk, r
1

is a constant,
1
is a parameter that takes into account
interactions of the polyelectrolyte in the double layer and adsorbed layers, with ( ) RT H
1 1

1 H 1 ( )R T
where
( )
1
H is the activation energy of adsorption between polyelectrolyte/interface, polyelectrolyte/polyelectrolyte
and polyelectrolyte/solvent. Equation (6) can be written in the form:

( )
1
]
1

,
_

1
1
1
2
1
1
2 1
2
exp

r
C (7)

The experimental data for POMA adsorbed onto POMA/PVS films were also successfully explained using
this model. The activation energy increases in the same way as the amount of material adsorbed which implies that
this energy does depend on film roughness, as the latter also follows the behavior of the adsorbed amount when the
number of layers is varied. The activation energies are smaller than those obtained by Filippova for the cationic
poly(vinylamine) hydrochloride polyelectrolyte which could be explained by the significant smaller number of
electrical charges in POMA molecules, about half of the other polyelectrolyte. Moreover, the adsorption kinetics of
poly(vinylamine) hydrochloride onto silicon, obtained by Filippova, displays a similar behavior to that of POMA.
However, a detailed analysis of adsorption kinetics curves, probably including other polyelectrolytes, is necessary to
generalize the Filippova model.
The irreversibility of the polyelectrolyte adsorption is one of the weaknesses of analytical models since
these models assume an adsorption/desorption equilibrium. The simplest model which takes into account the
irreversible adsorption is the random sequential adsorption (RSA) model, which is entirely based on statistics and
geometric deposition [414-416]. Adsorption of polyelectrolytes has also been extensively studied with models using
conformational statistics, involving exact enumeration, Monte Carlo methods, mean-field lattice models and train-
loop-tail models [401 and references therein]. Such models usually require knowledge of experimental parameters
that is not available. For instance, even though the enthalpy of adsorption can be measured directly using
calorimetric methods, the entropy and the free energy of adsorption cannot be measured easily. Additional
difficulties are associated with the large variety of experimental conditions that affect the adsorption processes. Each
trio of solute-solvent-substrate represents a new case, and in layer-by-layer films it is necessary to investigate the
processes for at least two materials that form the alternating layers. Finally, one can conclude that the analytical
96
models are advantageous for providing estimates of energies. Filippova [417], for instance, showed that the heat of
adsorption measured by microcalorimetry can be explained by her model for a wide range of adsorbate
concentrations.

3. Mechanisms for polymer adsorption
Electrostatic interactions are expected to predominate in the original paradigm of the formation of layer-by-
layer films, which was confirmed in the study by Lowach and Helm [372], using the surface force apparatus, for the
sulfonated polystyrene (PSS) and polyallylamine (PAH) system. They noted that only one third to one half of the
PAH positive charges are neutralized by PSS molecules upon adsorption of the latter. Such results were interpreted
with a Cohen-Stuart model [372,418] which states that polyelectrolyte molecules adsorb only until they are
electrostatically repelled from the surface and not until the binding places are saturated, which would correspond to
a "true equilibrium". The Cohen-Stuart model also explains why it is so difficult to remove adsorbed
polyelectrolytes from a surface under adverse conditions such as during the washing process. For instance, no
desorption is observed when POMA/PVS layer-by-layer films are placed in aqueous solutions for several days at
room temperature. Evidence of the predominance of electrostatic interactions for the stability of multilayers was
obtained by Hoogeveen et al. [346], who showed that when the polymer with a very low charge (>5%) is adsorbed
onto a surface, practically all polymer charges are bonded to the substrate and there is no adsorption of the next
layer owing to the small number of remaining sites for further adsorption. When the polymer charge is between 5-
20% the next layer may be formed but the conformation of the first layer will somewhat adjust itself to the new
layer. In some cases, during such adjustment bonds between the first layer and the substrate are broken, with both
layers being desorbed. When the polymer charge is even higher (>20%) the bonds between polymer-substrate and
polymer layers are strong, and stable multilayers can be formed [346].For amphoteric polymers, Kaniyama and
Israelachvili [419] suggested an adsorption mechanism where the number of discrete ionic bonds that can be formed
between the charged groups on the polymer and on the surface plays an imp ortant role. According to their rationale,
adsorption would be determined by the details of the charge distribution and not simply by the average charge of the
polymer and the surface, and therefore adsorption is assumed to be dependent on short -range electrostatic
(coulombic) interactions rather than on a long-range or continuum-type interaction.
For certain polymers hydrogen bonding also contribute for adsorption. An example is the case of
polyanilines where the hydrogen bonding was the driving force for adsorption of polyanilines in layer-by-layer films
as pointed out by Stockton and Rubner [354]. Also, Benjamin et al [420] produced layer-by-layer films via
hydrogen bonding between the amine group of poly(ethylleneimine) and the hydroxy group of the co(PHydroxyV-
PV) conjugated polymer. We have confirmed the importance of H-bonding by investigating the POMA/PVS system
97
at pH=3.0 where POMA is protonated (i.e. charged) and also at higher pHs where POMA is neutral. Even at
pH=3.0, thermally stimulated desorption data showed three types of interaction in the POMA/PVS multilayers,
namely van der Waals forces, hydrogen bonds and ionic interactions [393]. The thermally stimulated POMA
desorbed amount per unit area is shown as a function of temperature in figure 72. The importance of H-bonding
was demonstrated by FTIR spectroscopy of the films, and also by the pH dependency of the amount of adsorbed
POMA. For the first as well as for the nth layer of POMA, the maximum amount of adsorbed material occurred for a
pH = 6, at which POMA is completely dedoped, and therefore ionic attraction is vanishing. In fact, polyaniline
molecules can form H-bonds among themselves, in addition to the PVS molecules, which explains why the amount
of adsorbed POMA increases with the numb er of layers already deposited [394] and also the non-self-limiting
adsorption of POMA and PANi (polyaniline) under certain experimental conditions [403]. H-bonding promotes
aggregation, with the probability of adsorption increasing in sites where some polymer is already adsorbed. This
also causes the roughness to increase with the number of layers, as observed by AFM measurements, in contrast to
the results of Lowack and Helm [372] where defects and roughness of surface polyelectrolyte multilayer films were
found to be unimportant for the adsorption processes. As mentioned before, in their films the electrostatic attraction
was the predominant factor.

Figure 72

Several other works investigated details of the adsorption mechanisms. For example, Caruso et al [390]
studied the electrostatic binding of two anionic fluorescence probes to layer-by-layer films deposited onto
98
polystyrene (PS) latex particles. Hoogeveen et al [346] showed that there is a minimum charge density to grow a
stable layer-by-layer film. Upon investigating the influence of ionene polyelectrolyte linear charge density, Arys et
al [421] concluded that the formation of regular layer-by-layer films is limited by a critical charge density, from 0.9
to 1.1, below which the preparation of such films from solution is not possible. Also investigated were the effects
from varying the ionic strength of the polyelectrolyte solution.
Polymer adsorption is largely influenced by the molecular weight [422, 423], which brings additional
complexity because most polymers are polydisperse. Evidence has been gathered from the literature on polymer
adsorption that small molecules are the first to be adsorbed and at equilibrium the small molecules are substituted by
higher weight molecules [422]. In principle, this could also apply to adsorption in layer-by-layer films since most
polyelectrolytes (conductive polymers) are polydisperse. However the experimental conditions used in the studies of
adsorption and in the preparation of layer-by-layer films are quite different. In the first case the polymer molecules
are injected into the solvent where the substrate is already inside. In the case of layer-by-layer films the substrate is
immersed in the polyelectrolyte solution. For the POMA/PVS system, it was shown that within a short period of
adsorption (5s) the adsorbed amount per unit area is proportional to the polyelectrolyte solution concentration [394].
This result is easily explained by the adsorption of POMA molecules which are near the substrate caused by the
electrostatic interaction between positive charges of POMA molecules and the substrate which is negatively charged
due to the last PVS layer. Consequently, small as well as large molecules are adsorbed. Atomic force microscopic
studies of polypeptide poly[benzyl-L-glutamate] adsorption from polydisperse solutions suggest that neither long
nor short macromolecules replace each other on the substrate surface [424].
For adsorption of charged particles the polydispersity in particle size also influences the kinetics of
adsorption. The smaller particles adsorb preferentially due to their higher diffusion coefficients; already adsorbed
molecules that are electrically charged may prevent adsorption of larger molecules due to repulsion while the
smaller molecules can still be adsorbed in the interstices since repulsion is weaker. Semmler et al [425] showed,
both experimentally and using the random sequential model, that polydispersivity affects deposition of highly
charged polystyrene latex spheres with amidine surface head groups, with the small particles being adsorbed
preferentially. Polydispersity was also important for the changes with time of the size distribution of deposited
particles [425] as shown in fig. 73. No systematic study has been developed for polyelectrolytes, but it is reasonable
to assume a similar behavior in the formation of layer-by-layer films.
99

Figure - 73

The ionic strength of the polyelectrolyte solution has a strong influence in the adsorbed amount and in the surface
roughness of the layer. Indeed, an increase in solution ionic strength results in screening of the repulsion between
the electrical charges of the polyelectrolyte molecule [400]. This contributes for a coiling of the molecules and the
adsorbed amount or film thickness tends to increase. Lvov and Decher [426] showed that the growth per bilayer of
PVS/PAA and PSS/PAA layer-by-layer films is proportional to the square root of the solution ionic strength.
Recently, Steitz et al [427] confirmed this behavior in PSS/PAH films (fig. 74).
100

Figure - 74

3.3 Characterization of SA films
The experimental techniques which have been used in the characterization of layer-by-layer films are the
same for other organic thin films and a detailed description of these techniques can be found in [16, 428, 429]. The
most extensively used is UV-visible -near-infrared spectroscopy since most candidate materials for the SA method
absorbs light in this wavelength region, and this spectroscopy provides a direct means of estimating the amount of
material adsorbed and also the multilayer buildup. The multilayer thickness can be measured by X-rays and neutron
reflectivity, ellipsometry, profilometry and atomic force measurements. Lobo el al [397] presented a method to
measure thickness of soft films, in which a furrow is made on the film with the atomic force microscope tip and the
topography of the furrow is measured by scanning the sample. Figure 75 shows the POMA/PVS film thickness
measured by this method and by profilometry [397]. Photoluminescence and electroluminescence have also been
measured for the SA films made from luminescent polymers such as PPV [356,357]. The order of SA films has been
probed by neutron reflectivity measurements [371] and small angle X-ray reflectivity (SAXR) [344]. Although the
original model of layer-by-layer films consists of strictly stratified layers, interpenetration or overlapping of layers
has been shown to occur [388, 390, 430]. This was confirmed by the results obtained with POMA/PVS films, where
the amount of adsorbed material increased linearly with the number of layers even when a given layer was not
101
completely adsorbed. The AFM results show that the adsorption of these films consists of adsorption of small
nucleus which could lead to interpenetration.
0
100
200
300
400
0 20 40 60 80 100 120
AFM
Profilometer
T
h
i
c
k
n
e
s
s

(
n
m
)
Number of Bilayers

Figure - 75

The quartz crystal microbalance has been used for obtaining the adsorbed amount which is determined
from the frequency shift of the quartz resonator and using the Sauerbrey equation [349, 373]. Other techniques
employed in the investigation of SA film properties include atomic force microscopy (AFM) [355, 361, 375, 397],
X-ray photoelectron spectroscopy (XPS) [385], contact angle [375], cyclic voltammetry [431], scanning angle
reflectometry [432, 398], total internal reflectance fluorescence-TIRF [423], Near-Brewster reflectivity [423],
surface plasmon resonance [381], steady state fluorescence measurements [390], surface Force apparatus [372],
second harmonic generation [361, 362], electro -optic coefficient [433], single particle light scattering (SPLS)
measurements [390], streaming potential measurements [398] and thermally stimulated desorption [392].

3.4 Possible applications of SA polymeric films
As for the possible applications of SA films, the first extensive efforts have been made by Rubner's group
in the fabrication of light emitting diodes (LEDs) employing PPV initially [356, 357] and then ruthenium II
complexes (Ru(bpy)3
2+
polyester) [434]. Fou el al [351] showed that the type of polyanion used to assemble the film
affects dramatically the behavior and performance of the device. Figures 76 and 77 show the light-voltage and
current-voltage curves of PMA/PPV and SPS/PPV layer-films adsorbed onto ITO and covered with an aluminum
electrode. These films were thermally converted at 210C for 11h. The differences observed in the behavior of
102
PMA/PPV and SPS/PPV devices were attributed to a doping effect caused by the SPS sulfonic group. Also, these
authors showed that the device performance depends on the layer that is in contact with the aluminum electrode. The
device luminance and efficiency also depend on the type and sequence of PPV bilayers in layer-by-layer films [357].
This is demonstrated in figure 78 where light-current curves for different PPV layer-by-layer films were plotted.
Polymer light emitting diodes were also prepared with poly[2-(3-thienyl)ethanol hydroxycarbonyl-methyl methane]
(H-PURET) and europium (Eu(III)) layer-by-layer films [435].

Figure - 76

Figure - 77
103

Figure - 78

Similarly to other thin solid films, layer-by-layer films may be applied to sensors, integrated optics, friction
reducing coating and surface orientation layers. Raposo [436] showed that POMA/PVS layer-by-layer films
prepared from pH=3 are moisture sensitive and present a conductivity of about 5 orders of magnitude higher than
POMA casting films obtained from solutions with the same pH. Such an increase in conductivity was attributed to
protonation caused by the PVS sulfonic group as shown in figure 79. Dai et al [383] also reported that sulfonated
fullerene derivatives and dendrimers are protonic acid dopants of polyanilines in the emeraldine base form [383].
S O O
O
-
Na
+
n
-
Cl
+
H
N
N
N
H H
H
3
CO
H
3
CO

Figure - 79

Clark et al [374] discovered a process of patterning ionic layer-by-layer films with micron-sized features, exploiting
the ability of reverting the polyelectrolyte deposition on the self-assembled monolayer surfaces at very high salt
concentration, thus creating a negative to the original positive structure. This may lead to fabrication of more
complex structures, and then applied to optical or electrical devices [374].
The layer-by-layer method has been used to prepare films with non-linear optical proprieties. One of these
properties is the second order generation but results [362] of adsorption of ionene-type polycation demonstrate that
104
the layer-by-layer film does not possess a centrosymmetric structure as assumed by Lvov el al [360] who observed
second harmonic generation in PDDA/PAZO films. Ribeiro et al [433] prepared layer-by-layer films from
poly(allylamine hydrochloride) (PAH)methacrylic copolymer functionalized with the azo chromophore 4-[N-ethyl-
N-(2-hydroxiethyl)]-amino-2-chloro-4-nitroazobenzene (MMA-DR13) and observed that these films do not present
an initial orientation. Nevertheless, chromophore orientation could be induced by corona poling or optical poling,
leading to electrooptic coefficient values of 3 pm/V. The layer-by-layer films could be used as modified tribological
surfaces for boundary lubrication [437, 438]. Another important application for layer-by-layer films is the integrated
coating with gradient of refractive index along the normal to the surface plane which could be obtained by preparing
layer-by-layer films dendrimers [438].
Levsalmi and McCarthy [385] have determined the gas permeability of membranes of PAH/PSS films
onto poly(4-methyl-1-pentene) for hydrogen and oxygen. The PAH/PSS layer-by-layer films are 18000 times better
barriers to nitrogen than poly(4-methyl-1-pentene). The incorporation of metal nanosized particles onto a matrix is
also interesting in connection with layer-by-layer films that could be used in specific applications such as in
fabricating recording media, ferrofluids, magnetic refrigeration and color imaging [379 and references therein].
Dante et al [379] studied the nucleation and growth of iron oxy -hydroxide of polycrystalline particles of akaganite.
In order to obtain a glucose sensor Sun et al [382] produced layer-by-layer films of glucose oxidase and poly(4-
vinylpyridine) complex of osmium. Gas separation can also be obtained with layer-by-layer film, as demonstrated
by Ackern et al [377]. They showed that a 20-bilayer film of PAH/PSS reduces the argon flow to 7% of the initial
value while a 60-bilayer film reduces it to 0.1%. Similar results were obtained for gas flow rates of oxygen, nitrogen
and argon and the flow rate of carbon dioxide was higher by a factor of 2.4 [377]. Krasemann and Tieke [439]
showed that layer-by-layer films favor separation of mono and divalent ions by Donnan exclusion of the divalent
ions. This rejection increases with increasing number of bilayers, as shown in figure 80 where the permeation rate
(PR) was plotted against the number of bilayers [439]. These authors also showed that the ion permeation depends
on the concentration of excess charges in the individual polyelectrolyte layers. A high charge density favors a dense
less permeable membrane exhibiting improved rejection of ions. Also, Harris and Bruening [440] concluded that the
permeability of PAH/PSS films changes with the number of bilayers, ambient pH and film preparation conditions.
The permeability also depends on the ions to be separated and on the chemical structure of the polyelectrolyte.
Layer-by-layer films were also used in pervaporation [377] of polar and non-polar liquid mixtures. Shimazaki et al
[441] reported that it is possible to prepare layer-by-layer films of nonionic polymers, which have electron-accepting
groups and electron-donating groups, respectively at the ends of the side-chain molecules. Recently Caruso el al
[391] reported the encapsulation of the enzyme catalase by coating the enzymes crystal templates with the layer-by-
layer technique.
105

Figure 80
106
IV. Concluding Remarks
It was during mid-sixties, nearly four decades after its invention, and following the reports on energy
transfer experiments in multilayers by Hans Kuhn and his co-workers, that the LB technique was exploited for
producing ultrathin films with controllable molecular packing and film thickness. Since then, the applicability of the
LB technique has been extended to a variety of functional materials including polymers and other macromolecules.
Especially after the pioneering work of Tredgold on maleic anhydride polymer LB films, a number of different
functional polymers were subjected to LB manipulation. With polymeric materials, the inherent stability problem
associated with simple organic materials is expected to diminish. Over the years, significant understanding on the
interfacial properties of different monolayer material both at the air-water interface and on the solid substrate has
been gained. This is due to the intimate scientific collaboration among researchers from various areas of science, to
the advance of optical, electrical, microscopic and surface characterization techniques and also the development of
molecular modelling studies. For most of the proposed solid state device applications based on organic materials, an
important requirement is the control over molecular packing and film thickness. LB films from functional polymers
can readily be anisotropic, even though one may not achieve great control over the molecular packing. The advent of
the electrostatically -based self-assembly (SA) (or layer-by-layer) deposition method brought considerable progress
to the search of organic thin films with molecular control. In addition to being capable of producing multilayer
polymer structures similarly to LB films, the SA technique has the advantage of the simplicity and low cost of
production, which makes it considerably more attracting for technological applications.

Acknowledgements
The authors are grateful to Fapesp and CNPq (Brazil) and to FTC-Subprograma Cincia e
Tecnologia do 2o. Quadro Comunitrio de Apoio (Portugal) for the financial support.

V. References
1. I. Langmuir, J. Am. Chem. Soc. 39, 1848 (1917).
2. K.B. Blodgett, J. Am. Chem. Soc. 57, 1007 (1935).
3. E.K. Rideal, Surface Chemistry, 2
nd
Ed., Cambridge University Press, London, (1930).
107
4. N.K. Adams, The Physics and Chemistry of Surfaces, 3
rd
Ed., Chapter II, Oxford University Press, London,
(1941).
5. W.D. Harkins, Physical Chemistry of Surface Films, Reinhold, New York, (1952).
6. J.T. Davies and E.K. Rideal, Interfacial Phenomena, Academic Press, New York, (1961).
7. D.W. Stephens, Gas/Liquid and Liquid/Liquid Interfaces: A Bibliography, Joseph Crosfield, England, (1962).
8. G.L. Gaines Jr., Insoluble Monolayers at Liquid-Gas Interfaces, Interscience, New York, (1966).
9. A.W. Adamson, The Physical Chemistry of Surfaces, 2
nd
Ed., Interscience, New York, (1967).
10. H. Kuhn, D. Mbius and H. Bucher, Physical Methods of Chemistry, Eds., A. Weissberger, B. Rossiter, Vol.1,
Wiley, New York, (1972).
11. V.K. Srivastava, Physics of Thin Films, Eds., G. Hass, M.H. Francombe, R.W. Hoffman, Vol. 7, 311-397,
Academic Press, New York, (1973).
12. R. Aveyard and R. Haidon, An Introduction to the Principles of Surface Chemistry, Cambrigde University Press,
(1973).
13. M. Pomerantz, Phase Transitions in Surface Films, Plenum Press, New York, (1980).
14. G. Roberts, Ed. Langmuir-Blodgett Films, Plenum Press, New York, (1990).
15. C.M. Knobler, Adv. Chem. Phys. 77, 397 (1990)
16. A. Ulman, An Introduction to Ultrathin Organic Films, from Langmuir-Blodgett to Self-Assembly, Academic
Press, New York, (1991).
17. A. Ulman, Ed. A Guide to Ultrathin Organic Films, Academic Press, Boston, (1991).
18. R.H. Tredgold, Order in Thin Organic Films, Cambridge University Press, Cambridge, (1994)
19. A. Ulman, Organic Thin Films and Surfaces: Directions for the Nineties, Academic Press, London, (1995).
20. A. Ulman, Ed. Characterization of Organic Thin Films, Manning, Greenwich, (1995).
21. M.C. Petty, Langmuir-Blodgett Films - An Introduction, Cambridge University Press, Cambridge, (1996).
22. C.M. Knobler and D.K. Schwartz, Curr. Opinion in Colloid and Interface Sci. 4, 46 (1999).
23. Also refer the proceedings of the LB conferences in Thin Solid Films, 99 (1983), 132-134 (1985), 159-160
(1988), 178-180 (1989), 210-211 (1992), 243/244 (1994), 284-285 (1996), 327-329 (1998).
24. T. Iyoda, M. Ando, T. Kaneko, A. Ohtani, T. Shimadzu and K. Honda, Tetrahedron Lett., 27, 5633 (1986).
25. T. Iyoda, M. Ando, T. Kaneko, A. Ohtani, T. Shimadzu and K. Honda, Langmuir, 3, 1169 (1987).
26. T. Shimadzu, T. Iyoda, M. Ando, A. Ohtani, T. Kaneko and K. Honda, Thin Solid Films, 160, 67 (1988).
27. K. Hong and M.F. Rubner, Thin Solid Films, 160, 187 (1988).
28. X.Q. Yang, J. Chen, P.D. Hale, T. Inagaki, T.A. Skotheim, D.A. Fischer, Y. Okamoto, L. Samuelsen, S.
Tripathy, K. Hong, I. Watanabe, M.F. Rubner and M.L. den Boer, Langmuir, 5, 1288 (1989).
108
29. A.K.M. Rahman, L. Samuelson, D. Minehan, S. Clough, S. Tripathy, T. Inagaki, X.Q. Yang, T.A. Skotheim and
Y. Okamoto, Synth. Met., 28, C237 (1989).
30. K. Hong and M.F. Rubner, Thin Solid Films, 179, 215 (1989).
31. T.A. Skotheim, X.Q. Yang, J. Chen, T. Inagaki, M. Den Boer, S. Tripathy, L. Samuelsen, M.F. Rubner, K.
Hong, I. Watanabe and Y. Okamoto, Thin Solid Films, 178, 233-242 (1989).
32. K. Hong, R.B. Rosner and M.F. Rubner, Chem. Mater. 2, 82 (1990).
33. J. Cheung, R.B. Rosner, I. Watanabe and M.F. Rubner, Mol. Cryst. Liq. Cryst., 190, 133 (1990).
34. T.A. Skotheim, H.S. Lee, P.D. Hale, H.I. Karan, Y. Okamoto, L. Samuelson and S. Tripathy, Synth. Met., 41-43,
1433 (1991).
35. M. Rikukawa and M.F. Rubner, J. Mater. Sci., Pure Appl. Chem., A31, 793 (1994).
36. R.S. Duran and H.C. Zhou, Polymer, 33, 4019 (1992).
37. W.M. Sigmund, T.S. Bailey, M. Hara, H. Sasabe, W. Knoll and R.S. Duran, Langmuir, 11, 3153 (1995).
38. C.A. Nicolae, M.P. Fontana, M. Lazzeri and G. Ruggeri, Thin Solid Films, 284-285, 170 (1996).
39. M. Rikukawa, M. Nakagawa, N. Nishizawa, K. Sanui and N. Ogata, Synth. Met., 85, 1377 (1997).
40. M.P. Srinivasan and F.J. Jing, Thin Solid Films, 327-329, 127 (1998).
41. Y.-H. Kim and Y. -T. Kim, Langmuir, 15, 1876 (1999).
42. M. Ando, Y. Watanabe, T. Iyoda, K. Honda and T. Shimidzu, Thin Solid Films, 179, 225 (1989).
43. S.V. Melo, A. Dhanabalan and O.N. Oliveira Jr., Synth. Met., 102, 1433 (1999).
44. J.Y. Lee, D.Y. Kim and C.Y. Kim, Synth. Met., 74, 103 (1995).
45. J.Y. Lee and H. Lee, Korea Polymer Journal, 5, 207 (1997).
46. R.B. Rosner and M.F. Rubner, J. Chem. Soc., Chem. Commu., 1449 (1991).
47. D. Sarkar, A. Paul and T.N. Misra, Thin Solid Films, 227, 105 (1993).
48. R.B. Rosner and M.F. Rubner, Chem. Mater., 6, 581 (1994).
49. A. Paul, D. Sarkar and T.N. Misra, Solid State Commun., 89, 363 (1994).
50. A. Paul, D. Sarkar and T.N. Misra, J. Phys. D: Appl. Phys., 28, 899 (1995).
51. C.W. Yuan, C.R. Wu, J.J. Bai, W.Y. Yang and Y. Wei, Langmuir, 11, 5 (1995).
52. E. Milella, F. Musio and M.B. Alba, Thin Solid Films, 284-285, 908 (1996).
53. E. Milella, F. De Riccardis, C. Gerardi, C. Massaro, Synth. Met., 87, 19 (1997).
54. E. Milella and M. Penza, Thin Solid Films, 327-329, 694 (1998).
55. R. Casalini, L.M. Goldenberg, C. Pearson, B.K. Tanner and M.C. Petty, J. Phys. D: Appl. Phys., 31, 1504
(1998).
109
56. C.A. Nicolae, M.P. Fontana, R. Capelletti, R. Paradiso, A. Bonfiglio, M.T. Parodi, F. Ciardelli and G. Ruggeri,
Mol. Cryst. Liq. Cryst., 266, 277 (1995).
57. W.M. Sigmund, W.A. Goedel, R. Souto- Maior, A. C. Tenorio and C.P. de Melo, Langmuir, 15, 3273 (1999).
58. A. Dhanabalan, S.V. Mello and O.N. Oliveira Jr., Macromolecules, 31, 1827 (1998).
59. S.V. Mello, Patrycja Dynarowicz-Latka, A. Dhanabalan, R. F. Bianchi, R. Onmori,
R.A.J. Janssen

and O.N. Oliveira Jr., Abstract submitted to Ninth International conference on Organized Molecular
Films (LB9) to be held in Germany (2000).
60. M. Sato, S. Okada, H. Matsuda, H. Nakanishi and M. Kato, Thin Solid Films, 179, 429 (1989).
61. P. Yli-Lahti, E. Punkka, H. Stubb, P. Kuivalainen and J. Laakso, Thin Solid Films, 179, 221 (1989).
62. T. Bjornholm, D.R. Greve, N. Reitzel, T. Hassenkam, K. Kjaer, P.B. Howes, N.B. Larsen, J. Bogelund, M.
Jayaraman, P.C. Ewbank and R.D. McCullough, J. Am. Chem. Soc. 120, 7643 (1998).
63. U. Schoeler, K.H. Tews and H. Kuhn, J. Chem. Phys., 61, 5009 (1974).
64. N. Yamamoto, T. Ohnishi, M. Hatakeyama and H. Tsubomura, Thin Solid Films, 68, 191 (1980).
65. S. Tasaka, H.E. Tatz, R.s. Hutton. J. Orenstein, G.H. Fredrickson and T.T. Wang, Synth. Met., 16, 17 (1986).
66. H. Nakahara, J. Nakayama, M. Hoshino and K. Fukuda, Thin Solid Films, 160, 87 (1988).
67. A.J. Pal, J. Palheimo and H. Stubb, Appl. Phys. Lett., 67, 3909 (1995).
68. A.J. Pal, R. Osterbacka, K.-M. Kallman and H. Stubb, Appl. Phys. Lett., 71, 228 (1997).
69. A.J. Pal, T. Ostergard, J. Paloheimo and H. Stubb, Phys. Rev. B., 55, 1306 (1997).
70. Y.Q. Liu, Y. Xu and D.B. Zhu, Synth. Met., 84, 197 (1997).
71. L.M. Goldenberg, A.Donat-Bouillud, M. Leclerc and M.C. Petty, J. Electroanal. Chem., 443, 266 (1998).
72. L.M. Goldenberg, M. Leclerc, A.Donat-Bouillud, C. Pearson and M.C. Petty, Thin Solid Films, 327-329, 715
(1998).
73. I. Watanabe, K. Hong and M.F. Rubner, J. Chem. Soc. Chem. Commun., 123, (1989).
74. I. Watanabe, K. Hong and M.F. Rubner, Thin Solid Films, 179, 199-206, 1989.
75. I. Watanabe, K. Hong, M.F. Rubner, Langmuir, 6, 1164 (1990).
76. I. Watanabe and T. Shimidzu, Thin Solid Films, 205, 270 (1991).
77. A. Pawlicka, R.M. Faria, M. Yonashiro, S.V. Canevarolo Jr. and O.N. Oliveira Jr., Thin Solid Films, 244, 723
(1994).
78. M. Rikukawa, M. Nakagawa, H. Abe, K. Ishida, K. Sanui and N. Ogata, Thin Solid Films, 273, 240 (1996).
79. M. Rikukawa, M. Nakagawa, K. Ishida, H. Abe, K. Sanui and N. Ogata, Thin Solid Films, 284-285, 636 (1996).
80. M. Ahlskog, J. Paloheimo, E. Punkka and H. Stubb, Synth. Met., 55-57, 3830 (1993).
81. A.J. Pal, T. Ostergard, J. Paloheimo and H. Stubb, Appl. Phys. Lett., 69, 1137 (1996).
110
82. A. Tsumura, T. Kurata, S. Suzuki, H. Nobutoki, H. Koezuka and T. Moriwaki, Thin Solid Films, 178, 393
(1989).
83. J.P.K. Peltonen and J.B. Rosenholm, Thin Solid Films, 179, 543 (1989).
84. M. Rikukawa and M.F. Rubner, Thin Solid Films, 210-211, 274 (1992)
85. E. Punkka, M.F. Rubner, J.D. Hettinger, J.S. Brooks and S.T. Hannahs, Synth. Met., 55-57, 4997 (1993).
86. Y. Liu, Y. Xu and D. Zhu, Synth. Met., 90, 143 (1997).
87. C.J.L. Constantino, A. Dhanabalan and O.N. Oliveira Jr., Rev. Sci. Instrum., 70, 3674 (1999).
88. M. Rikukawa, Y. Wang and M.F. Rubner, Thin Solid Films, 239, 144 (1994).
89. C.L. Callender, C.A. Carere, G. Daoust and M. Leclerc, Thin Solid Films, 204, 451 (1991).
90. S. Sagisaka, M. Ando, T. Iyoda and T. Shimidzu, Thin Solid Films, 230, 65 (1993).
91. M. Bardosova, B. Stiller, R.H. Tredgold, M. Wooley, P. Hodge and L. Brehmer, Thin Solid Films, 284-285, 450
(1996).
92. L. Robitaille, J.-Y. Bergeron, G. DAprano, M. Leclerc and C.L. Callender, Thin Solid Films, 244, 728 (1994).
93. L.J. Noe, M. Tomoaia -Cotisel, M. Casstevens and P.N. Prasad, Thin Solid Films, 208, 274 (1992).
94. M. Schmelzer, S. Roth, P. Bauerle and R. Li, Thin Solid Films, 229, 255 (1993).
95. M. Schmelzer, M. Burghard, P. Bauerle and S. Roth, Synth. Met., 61, 97 (1993).
96. M. Schmelzer, M. Burghard, P. Bauerle and S. Roth, Thin Solid Films, 243, 620 (1994).
97. T. Ostergard, A.J. Pal and H. Stubb, J. Appl. Phys., 83, 2338 (1998).
98. R. Osterbacka, G. Juska, K. Arlauskas, A.J. Pal, K.-M. Kallman and H. Stubb, J. Appl. Phys., 84, 3359 (1998).
99. A.J. Pal, R. Osterbacka, K.-M. Kallman and H. Stubb, Appl. Phys. Lett., 70, 2022 (1997).
100. R. Osterbacka, A.J. Pal, K. -M. Kallman and H. Stubb, J. Appl. Phys., 83, 1748 (1998).
101. T. Oestergard, J. Paloheimo, A.J. Pal and H. Stubb, Synth. Met., 88, 171 (1997).
102. A. Bolognesi, C. Botta, G. Bajo, R. Osterbacka, T. Ostergard and H. Stubb, Synth. Met., 98, 123 (1998).
103. S.V. Elshocht, T. Verbiest, M. Kauranen, A. Persoons, B.M.W. Langeveld-Voss and E.W. Meijer, J. Chem.
Phys., 107, 8201 (1997).
104. S. Kishino, Y. Ueno, K. Ochiiai, M. Rikukawa, K. Sanui, T. Kobayashi, H. Kunugita and K. Ema, Phys. Rev.
B., 58, R13430 (1998).
105. L.H.C. Mattoso, O.N. Oliveira Jr. and M. Ferreira, CRC Polymeric Materials, 1432 (1996).
106. A. Dhanabalan, L.H.C. Mattoso and O.N. Oliveira Jr., Curr. Trends in Poly. Sci., (in press).
107. N.E. Agbor, M.C. Petty, A.P. Monkman, and M. Harris, Synth. Met., 55-57, 3789 (1993).
108. M.K. Ram, N.S. Sundaresan and B.D. Malhotra, J. Phys. Chem., 97, 11580 (1993).
109. M.K. Ram and B.D. Malhotra, Polymer, 37, 4809 (1996).
111
110. A. Dhanabalan, R.B. Dabke, N. Prasanth Kumar, S.S. Talwar, S. Major, R. Lal and A.Q. Contractor,
Langmuir, 13, 4395 (1997).
111. M.K. Ram, R. Gowri and B.D. Malhotra, J. Appl. Poly. Sci., 63, 141 (1997).
112. E. Punkka, K. Laakso, H. Stubb, K. Levon and W.-Y. Zheng, Thin Solid Films, 243, 515 (1994).
113. A. Riul Jr., L.H.C. Mattoso, S.V. Mello, G.D. Telles and O.N. Oliveira Jr., Synth. Met., 71, 2067 (1995).
114. A. Riul Jr., L.H.C. Mattoso, G.D. Telles, P.S.P. Herrmann, L.A. Colnago, N.A. Parizotto, V. Baranauskas,
R.M. Faria and O.N. Oliveira Jr., Thin Solid Films, 284-285, 177 (1996).
115. P. Granholm, J. Paloheimo and H. Stubb, , Phys. Rev., 55, 13658 (1997).
116. S.V. Mello, A. Riu l Jr., L.H.C. Mattoso, R.M. Faria and O.N. Oliveira Jr., Synth. Met., 84, 773 (1997).
117. H. Zhou, R. Stern, C. Batich, and R.S. Duran, , Makromol. Chem. Rapid Commun., 11, 4099 (1990).
118. R.S. Duran and H.C. Zhou, Polymer, 33, 4019 (1992).
119. P. Quint, W. Knoll, M. Hara, H. Sasabe and R.S. Duran, Macromolecules, 28, 4029 (1995).
120. T.E. Herod and R.S. Duran, Langmuir, 14, 6956 (1998).
121. L.J. Kloeppner and R.S. Duran, Langmuir, 14, 6734 (1998).
122. M. Ando, Y. Watanabe, T. Iyoda, K. Honda and T. Shimidzu, Thin Solid Films, 179, 225 (1989).
123. A.T. Royappa and M.F. Rubner, Langmuir, 8, 3168 (1992).
124. J.H. Cheung, E. Punkka, M. Rikukawa, R.B. Rosner, A.T. Royappa and M.F. Rubner, Thin Solid Films, 210-
211, 246 (1992).
125. D. Goncalves, R.M. Faria, O.N. Oliveira Jr. and J. Sworakowski, Synth. Met., 55-57, 3819 (1993).
126. L.H.C. Mattoso, S.V. Mello, A. Riul Jr., O.N. Oliveira Jr. and R.M. Faria, , Thin Solid Films, 244, 714 (1994).
127. L. Robitaile, J.-Y. Bergeron, G. DAprano, M. Leclerc and C.L. Callender, Thin Solid Films, 244, 728 (1994).
128. S. Paddeu, M.K. Ram and C. Nicolini, J. Phys. Chem. B, 101, 4759 (1997).
129. M.K. Ram, S. Carrara, S. Paddeu, E. Maccioni and C. Nicolini, Langmuir, 13, 2760 (1997).
130. Z. Chen, S. Ng, S. F.Y. Li, L. Zhong, L. Xu, H.S.O. Chan, Synth. Met., 87, 201 (1997).
131. A. Dhanabalan, A. Riul Jr., L.H.C. Mattoso and O.N. Oliveira Jr., Langmuir, 13, 4882 (1997).
132. A. Riul Jr., A. Dhanabalan, L.H.C. Mattoso, L.M. de Souza, E.A. Ticianelli and O.N. Oliveira Jr., Thin Solid
Films, 327-329, 576 (1998).
133. A. Riul Jr., A. Dhanabalan, M.A. Cotta, P.S.P. Herrmann, L.H.C. Mattoso, A.G. MacDiarmid and O.N.
Oliveira Jr., Synth. Met., 101, 830 (1999).
134. J.H. Cheung, , Ph.D., thesis, Massachusetts Institute of Technology, Cambridge, U.S.A (1993).
135. J.H. Cheung and M.F. Rubner, Thin Solid Films, 244, 990 (1994);
112
136. A. Dhanabalan, R.B. Dabke, S.N. Datta, N. Prasanth Kumar, S.S. Major, S.S. Talwar and A.Q. Contractor,
Thin Solid Films, 295, 255 (1997).
137. T. Suwa, M. Kakimoto, Y. Imai, T. Araki and K. Iriyama, , Mol. Cryst. Liq. Cryst., 255, 45 (1994).
138. A. Dhanabalan, A. Riul Jr. and O.N. Oliveira Jr., Supramolecular Sci., 5, 75 (1998).
139. R.M. Faria, L.H.C. Mattoso, M. Ferreira, O.N. Oliveira Jr., D. Goncalves and L.O.S. Bulhoes, Thin Solid
Films, 221, 5 (1992).
140. S.V. Mello, L.H.C. Mattoso, R.M. Faria and O.N. Oliveira Jr., Synth. Met., 71, 2039 (1995).
141. A. Dhanabalan, A. Riul Jr., D. Goncalves and O.N. Oliveira Jr., Thin Solid Films, 327-329, 60 (1998).
142. A.G. MacDiarmid and A.J. Epstein, Synth. Met., 65, 103 (1994).
143. A. Dhanabalan, A. Riul Jr., C.J.L. Constantino and O.N. Oliveira Jr., Synth. Met., 101, 690 (1999).
144. S.V. Mello, L.H.C. Mattoso, O.N. Oliveira, Jr. and R.M. Faria, Thin Solid Films, 284-285, 187 (1996).
145. J. Sworakowski, D. Goncalves, O.N. Oliveira Jr. and R.M. Faria, Chem. Listy, 90, 52 (1995).
146. C.J.L. Constantino, A. Dhanabalan, A. Riul Jr. and O.N. Oliveira Jr., Synth. Met., 101, 688 (1999).
147. S.V. Mello, A. Riul Jr., L.H.C. Mattoso, R.M. Faria and O.N. Oliveira Jr., Synth, Met., 84, 773 (1997).
148. A. Riul Jr., H. Haas, A. Dhanabalan, M.A. Cotta and O.N. Oliveira Jr., , Proc. of 4
th
congress of Brasilian
Polymer Soc. held in Salvador, Brazil (1997).
149. S. Sagisaka, S. Yoshida, M. Ando, T. Iyoda, T. Shimidzu, Thin Solid Films, 271, 138 (1995).
150. F.F. Bruno, J.A. Akkara, L.A. Samuelson, D.L. Kaplan, B.K. Mandal, K.A. Marx and S.K. Tripathy, Adv. Sci.
Technol., 10, 99 (1995).
151. K. Ramanathan, M.K. Ram, B.D. Malhotra and A.S.N. Murthy, , Materials Sci. Engg., C3, 159 (1995).
152. N.V. Lavrik, D. De Rossi, Z.I. Kazantseva, A.V. Nabok, B.A. Nesterenko, S.A. Piletsky, V.I. Kalchenko, A.N.
Shivaniuk and L.N. Markovskiy, Nanotechnology, 7, 315 (1996).
153. R.B. Dabke, G.D. Singh, A. Dhanabalan, R. Lal and A.Q. Contractor, Anal. Chem., 69, 724 (1997).
154. A. T. Royappa and M.F. Rubner, Langmuir, 8, 3168 (1992).
155. J.H. Cheung, E. Punkka, M. Rikukawa, R.B. Rosner, A.T. Royappa and M.F. Rubner, Thin Solid Films,
210/211, 246 (1992).
156. B. Hua, Z. Xie, Y. Chen and Y. Zhang, Dianhuaxue, 1, 432 (1995) (CA Number: 126: 157882).
157. A. Paul, T.N. Misra and D. Talukdar, Solid State Commun., 99, 633 (1996).
158. T.L. Porter, D. Thompson and M. Bradley, Thin Solid Films, 288, 268 (1996).
159. J.E.P. Da Silva, S. I. C. Detorresi, M.L.A. Temperini, D. Goncalves and O.N. Oliveira Jr., Synth. Met., 101,
691 (1999).
160. O.P. Dimitriev and N.V. Lavrik, Synth. Met., 98, 173 (1998).
113
161. O.P. Dimitriev and N.V. Lavrik, Synth. Met., 98, 173 (1998).
162. N.E. Agbor, M.C. Petty, and A.P. Monkman, Sens. Actuators, B28, 173 (1995)].
163. J. Paloheimo, A.J. Pal, H. Stubb, H, P. Granholm and H. Isotalo, Mater. Res. Soc. Symp. Proc., 413, 517
(1996).
164. L.M. Goldenberg, M.C. Petty and A.P. Monkman, J. Electrochem. Soc., 141, 1573 (1994).
165. S.V. Mello, L.H.C. Mattoso, J.R. Santos, Jr., D. Goncalves, R.M. Faria, and O.N. Oliveira Jr., Electrochimica
Acta, 40, 1851 (1995).
166. D. Goncalves, L.O.S. Bulhoes, S.V. Mello, L.H.C. Mattoso, R.M. Faria and O.N. Oliveira Jr., Thin Solid
Films, 243, 544 (1994).
167. R.B. Dabke, A. Dhanabalan, S. Major, S.S. Talwar, R. Lal and A.Q. Contractor, Thin Solid Films, 335, 203
(1998).
168. M.K. Ram, S. Carrara, S. Paddeu and C. Nicolini, Thin Solid Films, 302, 89 (1997).
169. A. Dhanabalan, J.A. Malmonge, A. Riul Jr., R.M. Faria and O.N. Oliveira Jr., Thin Solid Films, 327-329, 808
(1998).
170. J.A. Malmonge, A. Dhanabalan, A. Riul Jr., R.M. Faria and O.N. Oliveira Jr., Synth. Met., 101, 801 (1999).
171. A. Dhanabalan, S.S. Talwar, A.Q. Contractor, N.P. Kumar, S.N. Narang, S.S. Major, K.P. Muthe and J.C.
Vyas, J. Mater. Sci. Lett., 18, 603 (1999).
172. J.I. Kim, B.K. Oh, S.Y. Oh, J.-W. Choi, H. -W. Rhee and W.H. Lee, Mol. Cryst. Liq. Cryst. Sci. Technol., Sect.
A, 316, 397 (1998).
173. N.E. Agbor, J.P. Cresswell, M.C. Petty and A.P. Monkman, Sens. Actuators, B, B41, 137 (1997).
174. E. Fischer, J. Am. Chem. Soc., 82, 3249 (1960).
175. H.H. Jaffe, S.J. Yeh and W.R. Gardner, J. Mol. Spectrosc., 2, 120 (1958).
176. G. Kumar and D. Neckers, Chem. Rev. 89, 1915 (1989).
177. T. Nagele, R. Hoche, W. Zinth and J. Wachtveitl, Chem. Phys. Lett., 272, 489 (1997).
178. L.M. Blinov, J. Non-linear Opt. Phys. Mater., 5, 165 (1996).
179. A. Yabe, Y. Kawabata, H. Niino, M. Matsumoto, A. Ouchi, H. Takahashi, S. Tamura, W. Tagaki, H. Nakahara
and K. Fukuda, Thin Solid Films, 160, 33 (1988).
180. H. Nakahara and K. Fukuda, J. Colloid Interface Sci., 93, 530 (1983).
181. M. Tanaka, Y. Ishizuka, M. Matsumoto, T. Nakamura, A. Yabe, H. Nakanishi, Y. Kawabata, H. Takahashi, S.
Tamura, W. Tagaki, H. Nakahara and K. Fukuda, Chem. Lett., 1307 (1987).
182. A. Yabe, Y. Kawabata, H. Niino, M. Tanaka, A. Ouchi, H. Takahashi, S. Tamura, W. Tagaki, H. Nakahara and
K. Fukuda, Chem. Lett., 1 (1988).
114
183. A. Laschewsky, W. Paulus, H. Ringsdorf, A. Schuster, G. Frick and A. Mathy, Thin Solid Films, 210/211, 191
(1992).
184. Z. Liu, B.H. Loo, R. Baba and A. Fujishima, Chemistry Lett., 1023 (1990).
185. J. Anzai, N. Sugaya and T. Osa, J. Chem. Soc. Perkin Trans., 2, 1897 (1994).
186. T. Kawai, J. Umemua and T. Takenaka, Langmuir, 5, 1378 (1989).
187. X. Xu, M. Era, T. Tsutsui and S. Saito, Thin Solid Films, 173, L135 (1989).
188. X. Xu, M. Era, T. Tsutsui and S. Saito, Thin Solid Films, 178, 541 (1989).
189. X. Xu, Y. Munakata, M. Era, T. Tsutsui and S. Saito, Thin Solid Films, 179, 65 (1989).
190. J. Maack, R.C. Ahuja, D. Mobius, H. Tachibana and M. Matsumoto, Thin Solid Films, 242, 122 (1994).
191. J. Maack, R.C. Ahuja and H. Tachibana, J. Phys. Chem., 99, 9210 (1995).
192. I. Zawisza, R. Bilewicz, E. Luboch and J.F. Biernat, Thin Solid Films, 348, 173 (1999).
193. Z.Q. Yao, P. Liu, R.Z. Yan, L.Y. Liu, X.H. Liu and W.C. Wang, Thin Solid Films, 210/211, 208 (1992).
194. V.R. Shembekar, A. Dhanabalan, S.S. Talwar and A.Q. Contractor, Thin Solid Films, 342, 270 (1999).
195. D.S. Santos Jr., C.R. Mendonca, D.T. Balogh, A. Dhanabalan, A. Cavalli, L. Misoguti, J.A. Giacometti, S.C.
Zilio and O.N. Oliveira Jr., Chem. Phys. Lett., 317, 1 (2000).
196. K.-S. Lee and M. Iwamoto, J. Coll. Interface. Sci. 177, 414 (1996).
197. M. Iwamoto, T. Kubota and Z. Ou-Yang, J. Chem. Phys. 104, 736 (1996).
198. X. Xu and M. Iwamoto, Ferroelectrics, 196, 269 (1997).
199. K.H. Park and M. Iwamoto, J. Colloid Interface Sci. 193, 71 (1997).
200. X. Xu and M. Iwamoto, Jpn. J. Appl. Phys. Part 1, 36, 7348 (1997).
201. X.B. Xu, Y. Majima and M. Iwamoto, Thin Solid Films 331, 239 (1998).
202. K.H. Park and M. Iwamoto, Mol. Cryst. Liq. Cryst. Sci. A 316, 145 (1998).
203. M. Velez, S. Mukhopadhyay, I. Muzikante, G. Matisova and S. Vieira, Langmuir, 13, 870 (1997).
204. M. Matsumoto, H. Tachibana, F. Sato and S. Terrettaz, J. Phys. Chem. B., 101, 702 (1997).
205. H.S. Blair and C.B. McArdle, Polymer, 25, 1347 (1984).
206. B.R. Malcolm and O. Pieroni, Biopolymers, 29, 1121 (1990).
207. H. Menzel, Macromol. Chem. Phys., 195, 3747 (1994).
208. T. Seki and T. Tamaki, Chem. Lett., 1739 (1993).
209. T. Seki, K. Ichimura, R. Fukuda and T. Tamaki, Thin Solid Films, 284-285, 365 (1996).
210. M. Shimomura and T. Kunitake, Thin Solid Films, 132, 243 (1985).
211. K. Nishiyama and M. Fujihira, Chem. Lett., 1257 (1988).
212. K. Nishiyama, M. Kurihara and M. Fujihira, Thin Solid Films, 179, 477 (1989).
115
213. H. Tachibana, A. Goto, T. Nakamura, M. Matsumoto, E. Manda, H. Niino, A. Yabe and Y. Kawabata, Thin
Solid Films, 179, 207 (1989).
214. H. Tachibana, R. Azumi, T. Nakamura, M. Matsumoto and Y. Kawabata, Chem. Lett., 173 (1992).
215. H. Tachibana, T. Nakamura, M. Matsumoto, H. Komizu, E. Manda, H. Niino, A. Yabe and Y. Kawabata, J.
Am. Chem. Soc., 111, 3080 (1989).
216. H. Tachibana, R. Azumi, M. Tanaka, M. Matsumoto, S. Sako, H. Sakai, M. Abe, Y. Kondo, N. Yoshino, Thin
Solid Films, 284-285, 73 (1996).
217. T. Seki and K. Ichimura, Polym. Commun., 30, 108 (1989).
218. T. Seki and K. Ichimura, Thin Solid Films, 179, 77 (1989).
219. T. Seki, T. Tamaki, Y. Suzuki, Y. Kawanishi and K. Ichimura, Macromolecules, 22, 3505 (1989).
220. T. Seki, M. Sakuragi, Y. Kawanishi, Y. Suzuki, T. Tamaki, R. Fukuda and K. Ichimura, Langmuir, 9, 211
(1993).
221. T. Takahashi, Y.M. Chen, A.K. Rahaman, J. Kumar and S.K. Tripathy, Thin Solid Films, 210/211, 202 (1992).
222. T. Seki, H. Sekizawa and K. Ichimura, Polymer, 38, 725 (1997).
223. A. Dhanabalan, D.T. Balogh, A. Riul Jr., J.A. Giacometti and O.N. Oliveira Jr., Thin Solid Films, 323, 257
(1998)
224. A. Dhanabalan, D.T. Balogh, C.J.L. Constantino, A. Riul Jr., O.N. Oliveira Jr. and J.A. Giacometti, Mat. Res.
Soc. Symp. Proc., 488, 927 (1998).
225. A. Dhanabalan, D.T. Balogh, C.R. Mendonca, A. Riul Jr., C.J.L. Constantino, J.A. Giacometti, S.C. Zilio and
O.N. Oliveira Jr., Langmuir, 14, 3614 (1998).
226. C.R. Mendonca, L. Misoguti, A. Dhanabalan, D.T. Balogh, A. Riul Jr., C.J.L. Constantino, J.A. Giacometti,
O.N. Oliveira Jr., and S.C. Zilio, Proc. Symp. Laser and Appln., 204 (1997).
227. A. Ahluwalia, R. Piolanti, D. De Rossi and A. Fissi, Langmuir, 13, 5909 (1997).
228. R.J. Demchak and T. Fort Jr., J. Coll. Interface Sci., 46, 191 (1974).
229.O.N. Oliveira Jr., D.M. Taylor and H. Morgan, Thin Solid Films, 210-211, 76 (1992).
230. Y. Ren, Y. Tian, R. Sun, S. Xi, Y. Zhao and X. Huang, Langmuir, 13, 5120 (1997).
231. A. Dhanabalan, C. R. Mendona, D. Balogh, L. Misoguti, C.J.L. Constantino, J.A. Giacometti, S.C. Zilio and
O.N. Oliveira Jr., Macromolecules, 32, 5277 (1999).
232. A. Dhanabalan, D. S. Dos Santos, Jr., C. R. Mendona, L. Misoguti, D. T. Balogh, J. A. Giacometti, S. C. Zilio,
and O. N. Oliveira, Jr., Langmuir, 15, 4560 (1999).
233. Th. Geue, A. Ziegler and J. Stumpe, Macromolecules, 30, 5729 (1997).
234. M. Buchael, Z. Sekkat, S. Paul, B. Weichart, H. Menzel, W. Knoll, Langmuir, 11, 4460 (1995).
116
235. J. Gu, B. Liang, L. Liu, Y. Tian, Y. Chen, B. Lu and Z. Lu, Thin Solid Films, 327-329, 427 (1998).
236. M. Sawodny, A. Schmidt, M. Stamm, W. Knoll, C. Urban and H. Ringsdorf, Polym. Adv. Technol., 2, 127
(1991).
237. N. Carr, M.J. Goodwin, K.J. Harrison and K.L. Lewis, Thin Solid Films, 230, 59 (1993).
238. H. Menzel, B. Weichart, A. Schmidt, S. Paul, W. Knoll, J. Stumpe and T. Fischer, Langmuir, 10, 1926 (1994).
239. M. Matsumoto, D. Miyazaki, M. Tanaka, R. Azumi, E. Manda, Y. Kondo, N. Yoshino and H. Tachibana, J.
Am. Chem. Soc., 120, 1479 (1998).
240. T. Seki, H. Sekizawa, R. Fukuda, T. Tamaki, M. Yokoi, K. Ichimura, Polym. J., 28, 613 (1996).
241. T. Seki, H. Sekizawa and K. Ichimura, Polym. Commnu., 38, 725 (1997).
242. T. Seki, H. Sekizawa, S. Morino and K. Ichimura, J. Phys. Chem. B., 102, 5313 (1998).
243. H. Li, L. Zhang, X. Zhang, J. Shen, Y. Yang and H. Fei, Polymer Bulletin, 40, 735 (1998).
244. J.P. Cresswell, M.C. Petty, I. Ferguson, M. Hutchings, S. Allen, T.G. Ryan, C.H. Wang and B.S. Wherrett,
Adv. Mater. Opt. Electron., 6, 33 (1996)]
245. A.P.H.J. Schenning, C. Elissen-Roman, J. Weener, M.W.P.L. Baars, S.J. van der Gaast and E.W. Meijer, J.
Am. Chem. Soc., 120, 8199 (1998).
246. J. Weener and E.W. Meijer, Adv. Mater., (in press).
247. Z. -F. Liu, K. Hashimoto, A. Fujishima, Nature, 347, 658 (1990).
248. T. Enomoto, H. Hagiwara, D. A. Tryk, Z.-F. Liu, K. Hashimoto and A. Fujishima, J. Phys. Chem., B., 101,
7422 (1997).
249. A. Natansohn, P. Rochon, J. Gosselin and S. Xie, Macromolecules, 25, 2268 (1992)
250. S. Xie, A. Natansohn and P. Rochon, Chem. Mater., 5, 403 (1993).
251. A. Natansohn, P. Rochon, C. Barrett, A. Hay, Chem. Mater., 7, 1612 (1995).
252. M. Ho, A. Natansohn, C. Barrett and P. Rochon, Can. J. Chem., 73, 1773 (1995)
253. M. Ho, A. Natansohn and P. Rochon, Macromolecules, 29, 44 (1996).
254. T.S. Lee, D.Y. Kim, X.L. Jiang, L. Li, J. Kumar and S. Tripathy, Macromol.Chem. Phys., 198, 2279 (1997)
255. S.K. Tripathy, D. Kim, L. Li and J. Kumar, Chemtech, 28, 34 (1998).
256. A.-C. Etil, C. Fiorini, F. Charra and J.-M. Nunzi, Phys. Rev. A, 56, 3888 (1997).
257. C. Fiorini, J.-M. Nunzi, F. Charra, F. Kajzar, M. Lequan, R.-M. Lequan and K. Chane-Ching, Chem. Phys.
Lett., 271, 335 (1997).
258. P. Rochon, D. Bissonnette, A. Natansohn and S. Xie, Appl. Opt., 32, 7277 (1993).
259. M.I. Barnik, S.P. Palto, V.A. Khavrichev, N.M. Shtykov, S. G. Yudin, Thin Solid Films, 179, 493 (1989).
117
260. S.P. Palto, L.M. Blinov, S.G. Yudin, G. Grewer, M. Schoenhoff, M. Loeschem, Chem. Phys. Lett., 202, 308
(1993).
261. M. Schoenhoff, L.F. Chi, H. Fuchs, M. Loesche, Langmuir, 11, 163 (1995).
262. M.I. Barnik, V.M. Kozenkov, N.M. Shtykov, S.P. Palto and S.G. Yudin, J. Mol. Electron., 5, 54 (1988).
263. S. Yokoyama, M. Kakimoto, Y. Imai, Langmuir, 10, 4594 (1994).
264. S. Yokoyama, T. Yamada, K. Kajikawa, M. Kakimoto, Y. Imai, H. Takezoe and A. Fukuda, Langmuir, 10,
4599 (1994).
265. C.R. Mendonca, A. Dhanabalan, D.T. Balogh, L. Misoguti, D.S. dos Santos Jr., M.A. Pereira -da-Silva, J.A.
Giacometti, S.C. Zilio and O.N. Oliveira Jr., Macromolecules, 32, 1493 (1999).
266. B. Tieke, Adv. Polym. Sci., 71, 79 (1985).
267. A. Dhanabalan, S. Major and S. S. Talwar, Thin Solid Films, 327-329, 378 (1998).
268. K. Kuriyama, H. Kikuchi and T. Kajiyama, Chem. Lett. 1071 (1995).
269. K. Kuriyama, H. Kikuchi and T. Kajiyama, Langmuir, 12, 6468 (1996).
270. M. Sarkar and J.B. Lando, Thin Solid Films, 99, 119 (1983).
271. P.J. Werkman and A.J. Schouten, Thin Solid Films, 284-285, 24 (1996).
272. J.A. Fu, Q. Cheng and R. Stevens, Berkeley Sci., 2, 99 (1998).
273. M.D.P. Carreon, G. Burillo, L. Fomina, T. Ogawa, Polym. J. (Tokyo), 30, 95 (1998).
274. P.C.M. Del, G. Burillo, V. Agabekov and T. Ogawa, Polym. J. (Tokyo), 29, 103 (1997).
275. F. Grunfeld and C.W. Pitt, Thin Solid Films, 99, 249 (1983).
276. B. Tieke, G. Lieser and K. Weiss, Thin Solid Films, 99, 95 (1983).
277. S. Ya mada and Y. Shimoyama, Jpn. J. Appl. Phys., Part 1, 35, 4480 (1996)
278. S. Yamada and Y. Shimoyama, Jpn. J. Appl. Phys., Part 1, 36, 5242 (1997)
279. S. Yamada and Y. Shimoyama, Kobunshi Ronbunshu, 55, 157 (1998).
280. G. Lieser, B. Tieke and G. Wegner, Thin Solid Films, 68, 77 (1980).
281. T. Srikhirin, P.-M. Cham, J.A. Mann Jr., J.B. Lando, J. Polym. Sci., Part A: Polym. Chem., 37, 1771 (1999).
282. C. Ohe, H. Ando, N. Sato, Y. Urai, M. Yamamoto and K. Itoh, J. Phys. Chem. B, 103, 435 (1999).
283. P.J. Werkman, A. Schasfoort, R.H. Wieringa and A.J. Schouten, Thin Solid Films, 323, 243 (1998).
284. G. Veale, D.R.J. Milverston and M.N. Wybourne, Thin Solid Films, 136, 141 (1986).
285. H. Tamura, N. Mino and K. Ogawa, Thin Solid Films, 179, 33 (1989).
286. Y. Tomoika, N. Tanaka and S. Imazeki, Thin Solid Films, 179, 27 (1989).
287. D.T. Balogh, A. Dhanabalan, S.S. Talwar and O.N. Oliveira Jr., Abstract submitted to Ninth International
conference on Organized Molecular Films (LB9) to be held in Germany (2000).
118
288. S. Wang, Y. Li, L. Shao, J. Ramirez, P.G. Wang, R.M. Leblanc, Langmuir, 13, 1677 (1997).
289. S. Wang, J. Ramirez, Y. Chen, P.G. Wang and R.M. Leblanc, Langmuir, 15, 5623 (1999).
290. S. Wang, S. Vidon and R.M. Leblanc, J. Colloid Interface Sci., 207, 303 (1998).
291. Z. Ma, J. Li and L. Jiang, Langmuir, 15, 489 (1999).
292. B. Tieke and G. Lieser, J. Collod Interface Sci., 83, 230 (1981).
293. H.D. Gobel and H. Mhwald, Thin Solid Films, 159, 63 (1988).
294. H.D. Gobel, K. Kjaer, J. Als -Nielsen and H. Mhwald, Thin Solid Films, 179, 41 (1989).
295. F. Grunfeld and C.W. Pitt, Thin Solid Films, 99, 249 (1983).
296. A. Dhanabalan, S. S. Talwar and S. Major, Thin Solid Films, 279, 221 (1996).
297. K. Fukuda, Y. Shibasaki and H. Nakahara, Thin Solid Films, 160, 43 (1988).
298. K. Fukuda, Y. Shibasaki and H. Nakahara, Thin Solid Films, 160, 43 (1988).
299. A. Dhanabalan, S. S. Talwar and S. Major, Mater. Sci. Eng., C3, 235 (1995).
300. E. Shirai, Y. Urai and K. Itoh, J. Phys. Chem. B, 102, 3765 (1998).
301. S. Yamamoto, T. Ogawa, K. Mochiji and Y. Tomioka, Jpn. J. Appl. Phys., Part 1, 35, 5453 (1996).
302. F. Kajzar and J. Messier, Thin Solid Films, 99, 109 (1983).
303. P.J. Werkman, R.H. Wieringa and A.J. Schouten, Langmuir, 13, 6755 (1997).
304. H. Nino, H. Miyasaka, A. Ouchi, Y. Kawabata and A. Yabe, Thin Solid Films, 179, 53 (1989).
305. R. Pedriali, C. Cuniberti, D. Comoretto, G. Dellepiane, C. DellErba, M. Novi, A. Bolognesi, G. Bajo and W.
Porzio, Thin Solid Films, 284-285, 36 (1996).
306. D.B. Wolfe, M.S. Paley and D.O. Frazier, Polym. Prepr. 38, 197 (1997).
307. N. Terasawa, Y. Hayakawa, H. Fukaya, E. Hayashi, K. Kato, S. Fujii, H. Sawada, K. Li and J. Kyokane,
Polymer, 39, 5889 (1998).
308. F. Grunfeld and C.W. Pitt, Thin Solid Films, 99, 249 (1983).
309. D. Grando, S. Sottini and G. Gabrielli, Thin Solid Films, 327-329, 336 (1998).
310. H.C. Wang, D.W. Cheong, J. Kumar, C. Sung and S.K. Tripathy, Mater. Res. Soc. Symp. Proc., 440, 221
(1997).
311. D. Cheong, W. Kim, L.A. Samuelson, J. Kumar and S.K. Tripathy, Macromolecules, 29, 1416 (1996).
312. C. Bubeck, K, Weiss and B. Tieke, Thin Solid Films, 99, 103 (1983).
313. B. Tieke and K. Weiss, J. Colloid Interface Sci., 101, 129 (1984)
314. K. Fukuda, Y. Shibasaki and H. Nakahara, Thin Solid Films, 160, 43 (1988).
315. S. Hara, T. Arise, R. Yamamoto, M. Uchimura, M. Sasabe, H. Matsuda, H. Nakanishi, Thin Solid Films, 284-
285, 600 (1996).
119
316. A. Saito, Y. Urai and K. Itoh, Langmuir, 12, 3938 (1996).
317. K. Seki, I. Morisada, H. Tanaka, K. Edamatsu, M. Yoshiki, Y. Takata, T. Yokoyama, T. Ohta, S. Asada, H.
Inokuchi, H. Nakahara and K. Fukuda, Thin Solid Films, 179, 15 (1989).
318. T. Seki, K. Tanaka and K. Ichimura, Mol. Cryst. Liq. Cryst. Sci. Technol., Sect. A, 298, 511 (1997)
319. T. Seki, K. Tanaka and K. Ichimura, Adv. Mater., 9, 561 (1997).
320. T. Seki, K. Tanaka and K. Ichimura, Polym. J. (Tokyo), 30, 646 (1998).
321. K. Kuriyama, H. Kikuchi and T. Kajiyama, Langmuir, 14, 1130 (1998).
322. R. Popovitz-Biro, D.J. Hung, E. Shavit, M. Lahav and L. Leiserowitz, Thin Solid Films, 178, 203 (1989).
323. L.R. McLean, A.A. Durrani, M.A. Whittam, D.S. Johnston and D. Chapman, Thin Solid Films, 99, 127
(1983).
324. S.R. Sheth and D.E. Leckband, Langmuir, 13, 5652 (1997).
325. A. Lio, A. Reichert, J.O. Nagy, M. Salmeron and D.H. Charych, J. Vac. Sci. Technol., B, 14, 1481 (1996).
326. A. Lio, A. Reichert, D.J. Ahn, J.O. Nagy, M. Salmeron and D.H. Charych, Langmuir, 13, 6524 (1997).
327. D.-F. Gu, C. Rosenblatt and Z. Li, Liq. Cryst., 19, 489 (1995).
328. B. Cull, Y. Shi, S. Kumar and M. Schadt, Phys. Rev. E., 53, 3777 (1996).
329. T. Takamura, K. Matsushita and Y. Shimoyama, Hyomen Kagaku, 16, 486 (1995).
330. O. Albrecht, A. Laschewsky and H. Ringsdorf, J. Membrane Sci. 22, 187, (1985).
331. V.N. Kruchinin, L.L. Sveshnikova, S.M. Repinskii, L.G. Fedenok, E.V. Plashchenyuk and M.S. Shvartsberg,
Zh. Fiz. Khim., 71, 1470 (1997).
332. A.V. Maximychev, V.D. Matyukhin, N.D. Stepina and L.G. Yanusova, Thin Solid Films, 284-285, 866 (1996).
333. R.E. Schwerzel, K.B. Spahr, J.P. Kurmer, V.E. Wood, and J.A. Jenkins, J. Phys. Chem. A, 102, 5622 (1998).
334. D.-W. Cheong, W.-H. Kim, L.A. Samuelson, J. Kumar and S.K. Tripathy, Macromolecules, 29, 1416 (1996).
335. N.M. Ali, C.E. Tucker and F.A. Smith, Thin Solid Films, 289, 267 (1996).
336. N. Higashi, T. Kato and M. Niwa, Polym. J. (Tokyo), 30, 935 (1998).
337. Q. Cheng and R.C. Stevens, Chem. Phys. Lipids, 87, 41 (1997).
338. Q. Cheng and R.C. Stevens, Adv. Mater., 9, 481 (1997).
339. N. Tillman, A. Ulman and T. L. Penner, Langmuir 5, 101 (1989)
340. J. Richer, L. Stolberg and J. Lipkowski, Langmuir 2, 630 (1986)
341. R. Maoz and J. Sagiv, Langmuir 3, 1045, (1987)
342. S.D. Evans, E. Urankar, A. Ulman and N. Ferris, J. Am. Chem. Soc. 113, 4121 (1991)
343. G. Decher, J.D. Hong and J. Schmitt, Thin Solid Films 210/211, 831 (1992)
344. Y. Lvov, H. Haas, G. Decher and H. Mhwald, J. Phys. Chem. 97, 12835 (1993)
120
345. Y. Lvov, G. Decher, H. Mhwald, Langmuir 9, 481 (1993)
346. N.G. Hoogeveen, M. A. Cohen-Stuart, G. F. Fleer and M.R. Bhmer, Langmuir 12, 3675 (1996)
347. R.K. Iler, J. Colloid Interface. Sci. 21 , 569 (1966)
348. M. Ferreira, J.H. Cheung and M.F. Rubner, Thin Solid Films 244, 806 (1994)
349. J. H. Cheung, A.F. Fou and M. F. Rubner, Thin Solid Films 244, 985 (1994)
350. M. Ferreira and M.F. Rubner, Macromolecules 28, 7107 (1995)
351. M.C. Fou and M.F. Rubner, Macromolecules 28, 7115 (1995)
352. L.H.C. Mattoso, V. Zucolotto, L.G. Patterno, R. van Griethuijsen, M. Ferreira, S.P. Campana and O. N.
Oliveira Jr, Synth. Met. 71, 2037 (1995)
353. L.H.C. Mattoso, L.G. Patterno, S.P Campana and O.N. Oliveira Jr., Synth. Met. 84, 123 (1997)
354. W.B. Stockton and M.F. Rubner, Macromolecules 30, 2717 (1997)
355. M. Raposo, R.S. Pontes, L.H.C. Mattoso and O.N.Oliveira Jr., Macromolecules 30, 6095 (1997)
356. A.C. Fou, O. Onitsuka, M. Ferreira, M.F. Rubner and B. R. Hsieh, J. Appl. Phys. 79(10), 7501 (1996)
357. O. Onitsuka, A.C. Fou, M. Ferreira, B. R. Hsieh and M.F. Rubner, J. Appl. Phys. 80 (7), 1996, 4067
358. M. Raposo, L.H.C. Mattoso and O.N. Oliveira Jr. , Thin Solid Films 327/329, 739 (1998)
359. K. Ariga, Y. Lvov and T. Kunitake, J. Am. Chem. Soc. 119, 2224, (1997)
360. Y. Lvov, S. Yamada and T. Kunitake, Thin Solid Films 300, 107 (1997)
361. J. A. He, L. Samuelson, L. Li, J. Kumar and S. K. Tripathy, Langmuir 14, 1674 (1998)
362. A. Laschewsky, E. Wischerhoff, M. Kauranen and A. Persoons, Macromolecules 30, 8304 (1997)
363. Y. Lvov, K. Ariga, I. Ichinose and T. Kunitake, J. Am. Chem. Soc. 117, 6117 (1995)
364. Y. Lvov, K. Ariga, I. Ichinose and T. Kunitake, Thin Solid Films 284, 797 (1996)
365. M. Onda, Y. Lvov, K. Ariga and T. Kunitake, Jpn. J. Appl. Phys. 36, L1608 (1997)
366. C.E. Borato, P.S.P. Herrmann, L.A. Colnago, O.N. Oliveira Jr., L.H.C. Mattoso, Braz. J. Chem. Eng. 14, 367
(1997)
367. Y. Lvov, K. Ariga, I. Ichinose and T. Kunitake, Langmuir 12, 3038 (1996)
368. Y. Lvov, K. Ariga, M. Onda, I. Ichinose and T. Kunitake, Langmuir 13, 6195 (1997)
369. G. Decher, Y. Lvov and J. Schmitt, Thin Solid Films 244, 772 (1994)
370. J.H. Cheung, W.B.Stockton and M. F. Rubner, Macromolecules 30, 2712 (1997)
371. G. J. Kellogg, A. M. Mayes, W. B. Stockton, M. Ferreira, M. F. Rubner, S.K. Satija, Langmuir 12, 5109 (1996)
372. K. Lowack and C. A. Helm, Macromolecules 3, 823 (1998)
373. I. Ichinose, H. Tagawa, S. Mizuki, Y. Lvov and T. Kunitake, Langmuir 14, 187 (1998)
374. S. L. Clark, M. F. Montague, and P. T. Hammond, Macromolecules 30, 7237 (1997)
121
375. D. Yoo, S. S. Shiratori and M. F. Rubner, Macromolecules 31, 4309 (1998)
376. K. Araki, M.J. Wagner and M.S. Wrighton, Langmuir 12, 5393 (1996)
377. F. van Ackern, L. Krasemann and B. Tieke, Thin Solid Films 327-329, 762 (1998)
378. H. Fukumoto and Y. Yonezawa, Thin Solid Films, 327-329, 748 (1998)
379. S. Dante, Z. Hou, S. Risbud and P. Stroeve, Langmuir 15, 2176 (1999)
380. K. Yang, S. Balasubramanian, X. Wang, J. Kumar and S. Tripathy, Appl. Phys. Lett. 73(23), 3325 (1998)
381. R. Advincula, E. Aust, W. Meyer and W. Knoll, Langmuir 12, 3536 (1996)
382. Y. Sun, J. Sun, X. Zhang, C. Sun, Y. Wang, J. Shen, Thin Solid Films 327-329, 730(1999)
383. L. Dai, J. Lu, B. Mattews and A.W.H. Mau, J. Phys. Chem. B 102, 4049 (1998)
384. F. Caruso, D.G. Kurth, D. Volkmer, M.J. Koop and A. Mller, Langmuir 14, 3462 (1998)
385. J. -M. Levsalmi and T.J. McCarthy, Macromolecules 30, 78 (1997)
386. W. Chen and T. J. McCarthy, Macromolecules, 30, 78 (1997)
387. A.J. Dias and T.J. Mc Carthy, Macromolecules 20, 2068 (1987)
388. M.C. Hsieh, R.J. Farris and T.J. MacCarthy, Macromolecules 30, 8453 (1997)
389. E. donath, D. Walther, V.N. Shilov, E. Knippel, A. Budde, K. Lowack, C.A. Helm and H. Mhwald, Langmuir
13, 5294 (1997)
390. F. Caruso, H. Lichtenfeld, E. Donath and H. Mhwald, Macromolecules 32, 2317 (1999)
391. F. Caruso, D. Trau, H. Mhwald and Renneberg, Langmuir 16, 1485 ( 2000)
392. M. Raposo and O.N. Oliveira Jr., Langmuir, 16(6), 2839 (2000)
393. M. Raposo and O.N.Oliveira Jr., Braz. J. Phys. 28 (4), 2-14 (1998)
394. M. Raposo and O.N. Oliveira Jr., to be submitted
395. R.F.M. Lobo, M. A. Pereira-da-Silva, M. Raposo, R.M. Faria and O.N. Oliveira Jr., to be submitted
396. M. Lsche, J. Schmitt, G. Decher, W. G. Bouwman and K. Kjaer, Macromolecules 32, 8893 (1998)
397. R.F.M. Lobo, M. Pereira-da-Silva, M. Raposo, R.M. Faria and O.N. Oliveira Jr., Nanotechnology 10, 389
(1999)
398. G. Ladam, P. Schaad, J. C. Voegel, P. Schaad, J.C. Voegel, P. Schaaf, G. Decher and F. Cuisinier, Langmuir,
16, 1249, 2000
399. R.H. Ottewill, C.H. Rochester and A.L. Smith, Adsorption from solution, 1983
400. F. th. Hesselink, in Adsorption from solution at the solid/liquid interface, edited by G.D. Parfitt and C.H.
Rochester, Academic Press, 1983, p.377
122
401. G.J. Fleer, M.A. Cohen Stuart, J.M.H.M. Scheutjens, T. Cosgrove and B. Vincent, Polymers at Interfaces,
(Chapman&Hall, 1993)
402. J. Liklema, Fundamentals of Interface and Colloid Science Vol.II Solid-Liquid Interfaces(Academic Press,
1995)
403. R.S. Pontes, M. Raposo, C.S. Camilo, A. Dhanabalan, M. Ferreira and O.N. Oliveira Jr., Physica Status Solidi
(a) 173, 41-50 (1999)
404. G. Gabrielli, F. Cantale, G.G.T. Guarini, Colloids and Surfaces A: Physics Chemical and Engineering Aspects
119, 163 (1996)
405. V. V. Tsukruk, V. N. Bliznynk, D. Visser, A. L. Campbell, T. J. Bunning and W. W. Adams, Macromolecules
30, 6615 (1997)
406. G. Pan and P.S. Liss, J. Colloid Interface Science 201, 71 (1998)
407. G.J. Fleer and J. Lyklema, in "Adsorption from solution at the solid/liquid interface", edited by G.D. Parfitt and
C.H. Rochester, (Academic Press, 1983), p. 153
408. G.J. Fleer, Personal Communication
409. M.A. Awan, V.L. Dimonie, L.K. Filippov and M.S. El-Aasser, Langmuir 13, 130 (1997)
410. B. Senger, J.-C. Voegel and P. Schaaf, Colloids and Surfaces A: Physicochemical and Engineering Aspects
165, 255 (2000)
411. G.D. Parfitt and C.H. Rochester, Adsorption from solution at the solid/liquid interface, Academic Press, p. 4,
1983
412. W. Schinickler, Interfacial Electrochemistry, Oxford University Press, New York, 1996
413. N. L. Filippova, Langmuir 14, 1162 (1998)
414. G. Tarjus, P. Schaaf, J. Talbot, J. Chem. Phys. 93(11), 8352 (1990)
415. P.R. Van Tassel, P. Viot, G. Tarjus, P. Schaaf and J. Talbot, J. Chem. Phys. 101(8), 7064 (1994)
416. P.R. Van Tassel, P. Viot, G. Tarjus, J.J. Ramsden and J. Talbot, J. Chem. Phys. 112(3), 1483 (2000)
417. N. L. Filippova, J. Colloid Interface Science 197, 170 (1998)
418. M.A. Cohen-Stuart, C.W. Hoogendam and A. de Keizer, J.Phys.: Condens. Matter. 9, 7767 (1997)
419. Y. Kaniyama and J. Israelachvili, Macromolecules 25, 5081 (1992)
420. I. Benjamim, H. Hong, Y. Avny, D. Davidov and R. Neumann, J. Mater. Chem. 8(4), 919 (1998)
421. X. Arys, A.M. Jonas, B. Laguitton, A. Laschewsky, R. Legras and E. Wischerhoff, Thin Solid Films 327/329,
734 (1998)
422. J.C. Dijt, PhD Thesis, Wageningen Univ. (1993).
123
423. M. Santore and Z. Fu, Macromolecules 30, 8516 (1997)
424. V. Kitaev and E. Kumacheva, Langmuir 14, 5568 (1998)
425. M. Semmler, J. Ricka, M. Borkovec, Colloids and Surfaces A: Physicochemical and Engineering Aspects 165,
79 (2000)
426. Y. M. Lvov and G. Decher, Crystallography Reports 39(4), 628 (1994)
427. R. Seitz, V. Leiner, R. Siebrecht and R.V. Klitzing, Colloids and Surfaces A: Physicochemical and Engineering
Aspects 163, 63 (2000)
428. V.V. Tsukruk, Prog. Polym. Sci 22, 247 (1997)
429. A.K. Bajpai, Prog. Polym. Sci 22, 523 (1997)
430. G. Decher, Science 277,1232 (1997)
431. M.K. Ram, M. Salerno, M. Adami, P. Faraci and C. Nicolini, Langmuir, 15, 1252 (1999)
432. E.S. Pagac, D. C. Prieve and R.D. Tilton, Langmuir 14, 2333 (1998)
433. P.A. Ribeiro, M. Raposo, D.S. dos Santos Jr., D.T. Balogh, J.A. Giacometti and O.N. Oliveira Jr., Materials
Research Society 1999 Fall Meeting Abstract Book, Boston, p.510, 1999
434. A. Wu, J. Lee and M.F. Rubner, Mat. Res. Soc. Symp. Proc. Vol-488, (Materials Research Society 1998), p.63.
435. J. Kim, K.G. Chittibabu, M. J. Cazeca, W. Kim, J. Kumar and S.K. Tripathy, Mat. Res. Soc. Symp. Proc. Vol-
488, (Materials Research Society 1998), p. 527
436. M. Raposo, PhD Thesis, Instituto de Fsica de So carlos, Universidade de So Paulo, Brazil, 1999
437. B. Bhushan, J. Israelachvili and U. Landman, Nature 374, 607 (1995)
438. V.V. Tsukruk, F. Rinderspacher and V.N. Bliznyuk, Langmuir 13(8), 2171 (1997)
439. L. Krasemann and B. Tieke, Langmuir 16, 287 (2000)
440. J.J. Harris and M.L. Bruening, Langmuir, 16, 2006 (2000)
441. Y. Shimazaki, M. Mitsuishi, S. Ito and M. Yamamoto, Langmuir 14, 2768 (1998)
124
Figure captions

1. Schematic representation of (a) Langmuir-Blodgett (LB) films and (b) Self-Assembly (SA) films
2. Typical Surface pressure-Area (-A) isotherm of a Langmuir film of stearic acid at the air-water interface. Also
shown is the schematic diagram of a Langmuir trough.
3. Different types of LB films
4. Spreading behavior of polymers at the air-water interface
5. Organization of polymers at different stages of compression at the air-water interface
6. Structures of different amphiphilic pyrrole monomers used in ref. [24].
7. Electrochemical polymerization of the LB film of amphiphilic pyrrole monomers [26]
8. Structures of surface active pyrrole monomers used in ref. [27].
9. Possible organization of (a) polypyrrole and (b) pyrrole -alkylpyrrole copolymer, at the air-water interface.
10. Surface pressure -Area (-A) (a) and surface potential-area (V-A) (b) isotherms, of polypyrrole at air-water
interface at different subphase temperatures [43].
11. Schematic representation of solid state reaction in metal-fatty salt LB templates for producing thin polypyrrole
films [48].
12. Structures of amphiphilic polypyrroles used in refs [56, 57].
13. Structure of semi-amphiphilic polyalkylthiophene
14. Functionalized regio-regular amphiphilic polythiophenes used in ref. [62].
15. Structures of oligothiophenes used in ref. [66].
16. Typical surface pressure-area (-A) isotherm of a mixed monolayer of polyalkylthiophene and cadmium stearate
(5:1 ratio).
17. Structures of functionalized polythiophenes used in refs [89-92].
18. Schematic model proposed for a multilayer LB film containing polyalkylthiophene and cadmium stearate [75].
19. Structure of the chiral side chain substituted polythiophene used in ref. [103].

20. Structures of aniline monomers used for electrochemical polymerization at the air-water interface [117-120].
21. Structures of parent and substituted polyanilines used for LB processing. Also shown are the structures of
processing aids used to functionalize polyanilines.
22. Typical surface pressure-area (-A) isotherm of a polyaniline Langmuir film at the air-water interface.
23. Surface pressure - Area (-A) (A) and surface potential-area (V-A) (B) isotherms, of polyaniline at the air-
water interface, in the presence (a) and in the absence (b) of m-cresol in the spreading solution [141].
24. Surface pressure - Area (-A) isotherms of polyaniline processed with different processing aids at the air-water
interface [143].
25. Plot of weight percentage of 16-mer polyaniline in the mixed monolayer versus the limiting mean molecular area
calculated based on the weighed average molecular weights of 16-mer polyaniline and stearic acid [131].
26. Absorption spectral changes of a composite LB film containing polyaniline and cadmium stearate, upon X-ray
irradiation, 1) as deposited, 2) after 20 min. of irradiation, 3) 40 min., 4) 70 min., 5) 100 min., 6) 130 min., 7) 160
min., 8) 220 min., 9) 280 min., 10) 340 min [170].
27. Plot of the time required for the complete doping upon X-ray irradiation versus the number of layers of
composite LB films of polyaniline and cadmium stearate [170].
28. Reversible photoisomerization of azobenzene.
29. Structures of amphiphilic azobenzenes used in ref. [180].
30. Structure of host (-cyclodextrin) and guests (substituted azobenzenes), along with the proposed structure of the
host-guest complex [181].
31. Structure of host amphotropic spacer polymer and guest azobenzene derivative [183].
32. Structure of the amphiphilic azobenzene used in ref. [184].
33. Structure of the amphiphilic azobenzene with two alkyl chains used in ref. [185].
34. Structure of amphiphilic azobenzenes used in ref. [188]
35. Structure of amphiphilic azobenzene and schematic illustration of molecular packing arrangement of a mixed
monolayer of DMPA and 37 [190].
36. Structure of azocrown ethers used in ref. [192].
37. Structure of amphiphilic azobenzocrown ethers used in ref. [193].
38. Structure of disperse red 13 dye used in ref. [195].
39. Structures of polymeric and monomeric azobenzene amphiphiles used in ref. [209].
40. Structures of cationic and anionic poly mers used to form polyion complexes with azobenzene amphiphiles.
41. Structure of tailor-made azobenzene amphiphile used to achieve photoinduced switching of conductivity [215].
42. Structure of azobenzene amphiphiles with two azobenzene units used in ref. [216].
43. Structure of disperse red dye derivatized methacrylate homopolymers from ref. [224].
44. Structures of azobenzene polymers from ref. [227].
125
45. Structure of methacrylate copolymers with different amounts of azobenzene substitution [230].
46. St ructure of disperse red dye derivatized methacrylate copolymers [231].
47. Structure of disperse red-19 isophorone polyurethane from ref. [232].
48. Structure of azobenzene derivatized reversed Duckweed polymer used in ref. [243].
49. Structure of fifth generation amphiphilic poly(propylene imine) dendrimers: (a) fully azobenzene functionalized
and (b) partially azobenzene functionalized.
50. Structure of monomeric azo dyes used in optical storage studies.
51. Writing-erasing sequences on a mixed LB film containing an azobenzene polymer and cadmium stearate
52. Maximum induced birefringence (a) and time to achieve 50 % of saturated birefringence as a function of the
weight percentage of dye in the methacrylate copolymer [231].
53. Atomic force microscopy 3-D topography image of surface relief gratings on a 100-layer mixed LB film of
HPDR13 and cadmium stearate [265].
54. Structure of amphiphilic diacetylene with pyridine head group used in ref. [271].
55. Structure of PTSs from ref. [273].
56. Structure of amphiphilic diacetylenic acids from ref. [274].
57. Structure of amphiphilic diacetylenic acids from ref. [281].
58. Structure of Bronco used in ref. [293].
59. Structure of an amphiphilic -cyclodextrin with diacetylene side chain from ref. [304].
60. Structure of amphiphilic diacetylene functionalized with carbazole group from ref. [305]
61. Structure of poly(BPOD) from ref. [311].
62. Schematic diagram of layer-by-layer films buildup: A- Charged substrate; B-The substrate is immersed in the
cationic polyelectrolyte solution; C-Substrate covered with the cationic polymer; D-The substrate is rinsed in a HCl
aqueous solution and then immersed in the anionic polyelectrolyte.; E-Substrate covered with one bilayer.
63. Process to encapsulate enzymes using biocrystals as templates for the deposition of polymer multilayers,
subsequent enzyme solubilization and release, and the formation of hollow polymer capsules. (1, 2) Polyelectrolyte
layers are deposited in a stepwise fashion onto the crystals by making use of the surface charge reversal that occurs
upon adsorption of each layer. Each polyelectrolyte layer deposited bears an opposite charge to that already
adsorbed. Excess, unadsorbed polyelectrolyte is removed by repeated centrifugation/wash redispersion cycles before
the next layer is deposited. (3) Solubilization of the enzyme inside the polymer capsule by exposure to solutions of
pH > 6 or acidic solution (pH < 4) results in a morphology change of the polymer capsule. (4) Release of the
enzyme by rupturing the polymer capsule, achieved by exposure to solutions of pH > 11. (5) Exposure of the
encapsulated enzyme to an oxidizing solution results in decomposition of the enzyme which then is expelled from
the interior through the polymer walls, leaving behind hollow polymer capsules that originally encapsulate the
enzyme [391].
64. POMA adsorbed amount per unit area versus number of bilayers of POMA/PVS self-assembled films on
hydrophilized glass, changing the immersion time in POMA solution [392].
65. Maximum absorbance versus number of bilayers of POMA/PVS films adsorbed on: Teflon FEP, ITO, glass I
and glass II [355].
66. The adsorbed amount and mean roughness of POMA/PVS SA films versus the number of bilayers [392].
67. Schematic representation of a multilayer. Multilayer films can be subdivided into three regions: I, the region
close to the substrate. This region is typically composed of only a few layers that are different from region II owing
to the influence of the substrate. II, bulk multilayer film. Here all anionic layers possess equal thickness, and all
cationic layers possess equal thickness. In most cases polyanion/polycation stoichi-ometry is observed to be 1:1 or at
least close to that value. III, the region close to the film surface. This region is typically composed of only a few
layers. It could be described as a transition zone between the charge-compensated region II and the charged surface.
Please note that the transitions between regions I and II and between II and III are gradual and not as sharp as
schematically depicted here [398].
68. Kinetics of adsorption of PTAA onto 5 bilayers of PTAA/PAH for various concentrations. The absorbance was
measured at 420 nm [350].
69. Kinetics of adsorption of poly(o-methoxyaniline) onto a glass substrate: I- Few points were taken using only one
substrate. II- Points were taken at every 5 s using only one substrate. III- Several substrates, immersed only once for
distinct periods of time [355].
70. Adsorption isotherms of poly(o-methoxyaniline) onto glass at 25C.
71. Adsorption isotherms of POMA onto self-assembled films with distinct numbers of bilayers at 25C. The pH of
the POMA solutions was 3.0 0.2. The solid lines represent the fittings with Langmuir isotherms [392].
72. Thermally stimulated POMA desorbed amount per unit area as a function of temperature. The dashed line
corresponds to the fitting of the experimental data with a polynomial function. The time derivative of the fitted data
is also shown as a solid line [392].
73. Experimental surface size distribution Psurf (dark gray columns) and corresponding simulations with the
effective hard-sphere model (dashed columns) for different adsorption times of diffusional deposition. The
simulations represent an average over five individual runs, the experimental data are from 8002500 particles. The
126
adsorption times t are indicated. The longest time shows only the simulation result for 1 year of purely diffusional
deposition, which is inaccessible in our experiment [425].
74. Thickness of a single PSS:PAH bilayer as a function of the root mean square of the NaCl content of the
polyelectrolyte solution. The error bars correspond to the symbol size. The salt concentration also includes the free
counterions of the polyelectrolyte (3*10
-3
monomol/l) [427].
75. Film thickness measured with AFM and the profilometer versus the number of POMA/PVS film onto glass
[397].
76. Light-voltage and current-voltage plots of ITO/(PMA/PPV)n/Al devices thermally converted at 210 C for 11 h
[356].
77. Light-voltage and current-voltage plots of ITO/(SPS/PPV)n/Al devices thermally converted at 210 C for 11 h
[356].
78. Light-current plot of ITO/(PMA/PPV)20/Al and ITO/(SPS/PPV)5/(PMA/PPV)15/Al devices. Open symbols
represent films that have a PPV/Al interface whereas closed symbols represent films that have a PMA/Al interface
[357].
79. Protonation of POMA by PVS molecules in POMA/PVS layer-by-layer films.
80. Effect of the thickness of the separating membrane on permeation rates PR of NaCl and MgCl2. Separating
membrane: PAH/PSS [439].


Table I - Polycations used in the buildup of layer-by-layer films

POLYCATIONS ACRONYM REFERENCES
, , , -tetrakis(1-N-methylpyridyl)
porphine tetrakis(p-toluenesulfonate)
TMPyP 359
Calf thymus type YIII-S histone ---- 363
Chicken egg white lysozyme Lys 363, 364
Cytochrome c from horse heart Cyt c. 363, 364
Horse heart myoglobin Mb 363, 364
Horse hemoglobin Hb 364
Methylviologen MV 359
-{meso-tetra(4-pyridyl)-porphyrin}zinc-
tetrakis{bis(bipyridine)(chloro)ruthenium}
376
Peroxidase Per 364
Poly(allylaminehydrochloride) PAH 344, 350, 363, 369,
372, 373, 375, 377
Polyaniline PAN 370, 375
poly(ethylenimine) PEI 363-365, 359,.360,
367, 373
poly(diallyldimethylammonium) PDDA 360, 363, 364, 367,
373, 378-380
Poly(L- lysine hydrobromide) PL 372
Poly(2-vinylpyridine) PVP 372
poly(dimethylamino) ethylmethacrylate PMA 346
poly(dimethyldiallylammonium chloride) PDAC 361, 374
Poly[(N-methyl-pyridinium-2-
yl)acetylene]
PMPA 350
Poly(N,N- dimethyl-N, N-dioctadecyl-
1,6-hexanediamine)
Ionene-1 381
Poly(p-phenylenevinylene) precursor PPV precursor 350, 356, 357
poly(o- methoxyaniline) POMA 352, 355, 358
partially iodomethane qyarternized poly(4-
vinylpyridine) complex of osmium
Pos 382
partially quaternized polyvinylimidazole PVI+ 346

127

128

Table II- Polyanions used in the buildup of layer-by-layer films.

POLYANIONS ACRONYM REFERENCES
3,3-Disulfopropyl-5,5-dichloro-
thiacyanine triethylammonium
378
3,3-Disulfopropyl-5,5-dichloro-9-ethyl-
thiacarbocyanine triethylammonium
378
1,4-diketo-3,6-diphenylpyrrolo-[3,4-c]-
pyrrole-4,4-disulfonic acid
DPPS 377
10,12-docosadiyne-1,22-disulfatedisodium
salt
DCDS 377
Acid Red 26 AR26 359
Acid Red 27 AR27 359
Ammonium octamolybdate (NH4)4[Mo8O2
6]
373
Bacteriorhodopsin BR 361
Bovine liver catalase Cat 364
Concanavalin A Canavalia Ensiformis Con A 364
Congo Red LR 359
Diaphorase DIA 364
Ferritin FER 365
Glucoamylase Aspergillus niger GA 363, 364
Glucose oxidase GOD 363-365, 382
Hydrogensulfated fullerenol --- 383
Indigo Carmine IC 359
Meso-tetra(4-phenyl)porphyrin sulfonate 376
Molybdenum(VI) polyoxometalate cluster Mo57 384
Poly(1-(4-(3-carboxy-4-
hydroxyphenylazo)benzenesulfonamido)1,
2-ethanediyl)sodium salt
PAZO 381
Poly(acrylic acid) PAA 346, 375
Poly(anilinepropanesulfonic acid) PAPSA 363
Poly(ethenesulfonic acid) PVS 353, 355, 358
Polyimide precursor --- 350
Poly(methacrylic acid) PMA 356, 357, 346
Poly(thiophene-3-acetic acid) PTAA 375, 393
Poly(styrene sulfonate) SPS (or PSS) 346, 349,350,356,
357, 359, 360,
363-365, 367, 369,
372-375, 377, 379
poly(1-(4-(3-carboxi-4-
hydroxyphenylazo)-
benzenesulphonamido)-1,2-ethanediyl)
PAZO 345

Sulfonated dendrimer --- 383
Sulfonated polyaniline SPAn 350

Вам также может понравиться