Sei sulla pagina 1di 10

Thin Solid Films 583 (2015) 6069

Contents lists available at ScienceDirect

Thin Solid Films


journal homepage: www.elsevier.com/locate/tsf

Layer-by-layer assembled highly absorbing hundred-layer lms


containing a phthalocyanine dye: Fabrication and photosensibilization
by thermal treatment
Alena S. Sergeeva a,b,, Elena K. Volkova a, Daniil N. Bratashov a, Mikhail I. Shishkin a, Vsevolod S. Atkin a,
Aleksey V. Markin a, Aleksandr A. Skaptsov a, Dmitry V. Volodkin b, Dmitry A. Gorin a
a
b

Saratov State University, Astrakhanskaya 83, 410012 Saratov, Russia


Fraunhofer Institute for Cell Therapy and Immunology, Branch Bioanalytics and Bioprocesses (Fraunhofer IZI-BB), Muehlenberg 13, 14476 Potsdam, Germany

a r t i c l e

i n f o

Article history:
Received 14 November 2013
Received in revised form 15 March 2015
Accepted 20 March 2015
Available online 28 March 2015
Keywords:
Layer-by-layer assembly
Copper phthalocyanine
Polyelectrolytes
Multilayer thin lms
Absorption spectra
Surface morphology
Thermogravimetry analysis

a b s t r a c t
Highly absorbing hundred-layer lms based on poly(diallyldimethylammonium chloride) (PDADMAC) of
various molecular weights and on sulfonated copper phthalocyanine (CuPcTs) were prepared using layer-bylayer assembly. The multilayer lms grew linearly up to 54 bilayers, indicating that the same amount of CuPcTs
was adsorbed at each deposition step. This amount, however, was dependent on the molecular weight of
PDADMAC in the range 100500 kDa: the higher the molecular weight, the more CuPcTs molecules were
adsorbed. This can be explained by the larger surface charge number density specic to longer polymer chains.
Domains of pure PDADMAC and of the PDADMAC/CuPcTs complex were formed in the lms during the assembly.
Uniform distribution of CuPcTs over the lms could be achieved by thermal treatment, leading to an phase
transition in phthalocyanine at 300 C. Annealing caused changes in the lm absorbance spectra, resulting in a
30-nm red shift of the peak maxima and in a strong (up to 62%) decrease in optical density. Thermogravimetric
analysis revealed thermodegradation of PDADMAC during annealing above 270 C, giving rise to micrometersized cracks within the lms, as evidenced by scanning electron microscopy.
2015 Elsevier B.V. All rights reserved.

1. Introduction
Nano- and microstructured thin lms containing organic materials
(such as phthalocyanine dyes) serve as components of highly integrated
devices in different elds of science and technology, including optoelectronics [17]. A wide variety of methods can be used to fabricate such
coatings. Layer-by-layer (LbL) assembly is one of the simplest techniques to design multilayer lms with controlled thickness and chemical composition [8]. This approach makes it possible to use substrates
of various shapes and dimensions (up to the micro- and nano-scale)
for the fabrication of planar and curved nanomaterials with required
properties [916]. The LbL technique has been employed to pattern
surfaces with various particles, including microcapsules, micelles,
and nanoparticles [1722]. The nature of the polyelectrolytes used in
LbL assembly is important, owing to the strong inuence of interpolyelectrolyte complexation [23,24]. The thickness of deposited layers
Corresponding author at: 410012, Saratov State University, Astrakhanskaya St. 83,
Saratov, Russia. Tel.: +7 845 2511181, +7 927 1399975.
E-mail addresses: alenasergeeva@mail.ru (A.S. Sergeeva), ekvolkova87@rambler.ru
(E.K. Volkova), dn2010@gmail.com (D.N. Bratashov), shishkin1mikhail@gmail.com
(M.I. Shishkin), ceba91@list.ru (V.S. Atkin), markinav@mail.ru (A.V. Markin),
skaptsov@yandex.ru (A.A. Skaptsov), dmitry.volodkin@izi-bb.fraunhofer.de
(D.V. Volodkin), gorinda@mail.ru (D.A. Gorin).

http://dx.doi.org/10.1016/j.tsf.2015.03.050
0040-6090/ 2015 Elsevier B.V. All rights reserved.

depends strongly on polyelectrolyte molecular weight (MW), surface


charge density, and polymer transport to the lm surface [2528]. The
formation of domains or phase separation can take place during layer
deposition [27,29,30] or as a result of some post-treatment [3133],
leading to lm microstructurization. In the case of phthalocyanine
polymer lms, thermal treatment can cause both phase separation
and formation of a homogeneous phase [29].
LbL lms containing conductive polymers, carbon nanostructures
(single- or multiwall tubes, fullerenes), and dyes can serve as electrodes
or photoactive layers in multilayer photovoltaic converters [7,3440].
Indeed, the hybrid photoelement structure can be fabricated only
by all-solution processes [7,3440]. The main advantages of such
multilayer organic elements are that they are cheap, easy to process,
and ecologically safe; however, the photovoltaic characteristics are not
satisfactory. Therefore, the next promising direction in optoelectronics
(apart from the creation of thin-lm hybrid solar cells) is the development of unconventionally constructed photoelements [4146]. Modern
approaches to modifying photovoltaic converters are based on the
use of textured templates [41], nanowire arrays [42], plasmonic nanoparticles [45], spherically shaped elements [43], and photonic crystal
structures [44]. Alternative constructions with the abovementioned
changes in design can provide enhanced photovoltaic converter
efciency. The performance improvement can be achieved both by

A.S. Sergeeva et al. / Thin Solid Films 583 (2015) 6069

optimizing absorbance spectra and by enlarging the space-charge


region (increasing the photogenerated charge separation area). Surface
patterning reduces the reection and increases the number of adsorbed
photons [4144], leading to enhanced efciency.
A nonconventional design for a thin-lm photovoltaic converter was
suggested recently by Regan et al. [46]. The proposed construction is
a screening-engineered eld-effect photoelement based on a singledoped inorganic semiconductor [46]. The multilayer architecture is
similar to the metalinsulatorsemiconductor structure and consists
of a bottom contact on a solid substrate, a single-doped semiconductor
layer, an ultrathin top contact, and two transparent layers: a gate
contact and a gate insulator. This unconventional construction shows
acceptable characteristics, which can be additionally improved by applying an external voltage to the gate contact and blocking the recombination current [46].
Regan et al.'s [46] work inspired us and became a starting point
for this research. Si and Cu2O lms can be replaced with an organic
layer (such as a sulfonated phthalocyanine complex acting as a singledoped semiconductor), so that the described construction [46] could
be fabricated using an all-solution process. One challenge is to obtain
the LbL photoactive organic layers with a proper thickness of about
30200 nm [13] in general for hybrid solar cells based on fullerene
phthalocyanine evaporated lms. Thus, to prepare sufciently thick
phthalocyanine lms by LbL assembly, one should use an acceptable
cationic polyelectrolyte [e.g., poly(diallyldimethylammonium chloride)
(PDADMAC), poly(ethyleneimine) (PEI), poly(allyl amine hydrochloride)
(PAH), or polyaniline] coupled to an anionic phthalocyanine to consecutively deposit the necessary number of layers. For a single LbL bilayer, a
thickness of 0.81.2 nm has been reported for MPcTs/PDADMAC and
MPcTs/PAH structures [4749] (here, MPcTs is sulfonated phthalocyanine of a certain metal). Thus, it is important to fabricate hundred-layer
structures to achieve a proper thickness for a photoactive semiconductor
lm. Note that CuPcTs/PDADMAC multilayers possess electrochromic as
well as photovoltaic properties [49].
Fabrication of thick micrometer-sized polyelectrolyte multilayers
with large numbers of bilayers (tens or more) has been demonstrated
previously for different polymers [8,50]. Such thick lms usually serve
as reservoirs for bioactive molecules and are utilized in various biological applications to load, store, and release biomolecules to biological
cells [5154], also upon external stimulation [55,56]. However, the literature describing the LbL fabrication of hundred-layer lms, especially
thick coatings containing phthalocyanines, is very limited [47,49].
LbL assembly was applied to fabricate highly absorbing hundredlayer organic semiconductor coatings from PDADMAC as a cationic
component and from sulfonated copper phthalocyanine (CuPcTs) as
an anionic single-doped organic semiconductor. PDADMACs of different
MWs were used to examine the effect of the polymer MW on the thickness of deposited lms [25,50,54,5762].
The choice of phthalocyanine as an anion is driven by its photovoltaic properties and its strong optical absorbance between 550 and
750 nm [47,6373]. In fact, phthalocyanine lms can be in the amorphous -phase or in the crystalline -phase, depending on the fabrication approach. As both - and -phthalocyanines demonstrate
photogeneration, both forms are used in photovoltaic devices. In both
cases, the planar phthalocyanine macrocyclic molecules are arranged
in one-dimensional stacks; however, the relative arrangement of these
stacks and the stacking overlap of adjacent molecules differ between
- and -phases. The main distinction between the two phases is the
difference in the stacking axis of the phthalocyanine molecules and
the normal to the surface plane of the dye molecule; this can be distinguished by using a strong diffractive peak near 6.9 in X-ray diffraction
spectra of phthalocyanine lms [69,70]. The overlap between the
neighboring molecular stacks in the -phase is smaller (0.38 nm), as
compared with that for the -phase (0.48 nm). The closer spacing of
the molecular stacks and the orientation of molecules lying at (facedown) on the substrate (in contrast to the standing-up (on-edge)

61

molecule arrangement) facilitate the electron transport that is of


importance for photovoltaic converter efciency. The usual way to
obtain -phase phthalocyanine lms is to employ high-temperature
evaporation techniques and special conditions (i.e., heating of substrates), resulting in high crystallinity and good electrical properties
of the structures. Alternatively, a high-temperature postannealing is
applied to amorphous phthalocyanine lms (e.g., those prepared by
wet chemical deposition techniques) to cause an phase transition
accompanied by improvement of the molecular structure as well as
spectral and electrical characteristics. The inuence of annealing on
evaporated pure phthalocyanine lms has been studied [6670]. For
30150-nm-thick coatings prepared by vacuum evaporation, a phase
transition was achieved if the temperatures from 240 to 250 C
[68,69] to 300 C [66] were used. The minimum known phase transition
temperature is 200 C for 80-nm-thick vacuum-evaporated ZnPc lms
[70]. The phase transition point for phthalocyaninepolyelectrolyte
multilayer lms may differ from that for pure phthalocyanine owing
to the presence of the polyelectrolyte matrix, which can lead to phase
separation [27,2933] and affect phthalocyanine thermostability [47].
In this work, we examine the high-temperature behavior of
hundred-layer PEI/CuPcTs/(PDADMAC/CuPcTs)n lms, taking into
account lm composition and morphology.

2. Experimental details
The general denition of the LbL method, used to prepare multilayer
thin-lm structures on solid substrates, was rst given by Decher and
Hong [8]. A detailed description of this process in the context of this
work is given below.

2.1. Materials
Poly(ethyleneimine) (MW 6001000 kDa), poly(diallyldimethylammonium chloride) (MW b 100 kDa, 35 wt.%), poly(diallyldimethylammonium chloride) (MWs 100200 kDa, 200350 kDa, and 400
500 kDa; 20 wt.%), and tetrasulfonated copper phthalocyanine
(CuPcTs) were purchased from Sigma-Aldrich and used without further
purication. Aqueous polymer solutions (2 mg mL 1) and homogeneous dispersions of CuPcTs (0.25 mg mL1) were prepared in deionized water. Glass slides coated with uorine-doped tin oxide (FTO)
(20 30 3.2 mm3; surface resistance, 8 square1; transmittance,
8081.5% in the visible range; Sigma-Aldrich) were used as substrates
to assemble the lms.

2.2. Deposition of multilayer thin lms


Thin lms of polyelectrolytes and sulfonated copper phthalocyanine
were prepared by LbL assembly [8] on ethanol-precleaned glass slides
coated with conductive FTO [74,75]. Films were deposited onto the
slides by vertical immersion of the glass substrate into PDADMAC and
CuPcTs solutions at a controlled rate (5 mm min1) by using a dipping
mechanism (part of the KSV Nima LangmuirBlodgett Trough). The
rst layer was made by adsorption of the cationic polyelectrolyte
(2 mg mL1 water solution), with subsequent three rinsing steps in
deionized water. The second layer was prepared by adsorption of
anionic copper phthalocyanine (0.25 mg mL1 water solution) followed
by three rinsing steps. The fabricated multilayer structure was air dried
during a variable period after each bilayer deposition. Either long-time
drying (1 min for every layer and 20 min after every 3 bilayers) or
short-time drying (1 min for every layer) was used during LbL deposition. The deposition procedure was repeated n times to assemble (PDADMAC/CuPcTs)n lms. In this way, glass/FTO/PEI/CuPcTs/
(PDADMAC/CuPcTs)n structures (n, 1554) were obtained (Table 1).

62

A.S. Sergeeva et al. / Thin Solid Films 583 (2015) 6069

Table 1
Structure of the fabricated LbL multilayers and data about an average MW of the used
PDADMAC polyelectrolytes.
Multilayer structure of thin lms formed on the
FTO-coated substrates

Polyelectrolyte
MW, kDa

Number of
layers

PEI/CuPc/(PDDA/CuPc)15
PEI/CuPc/(PDDA/CuPc)15
PEI/CuPc/(PDDA/CuPc)15
PEI/CuPc/(PDDA/CuPc)54

PDDA, b100
PDDA, 100200
PDDA, 200350
PDDA, 400500

32
32
32
110

2.3. Thermal treatment


The hundred-layer lms were annealed in air at 200300 C. The
oven with the samples was heated from 25 C to the required temperature at a rate of ~20 C min1; this temperature was kept constant for
3 h. Various annealing temperatures were used to examine the inuence of heating temperature on the PDADMAC/CuPcTs multilayers and
to achieve an phase transition [6670].

2.4. Research methods


The surface morphology of the fabricated multilayer lms was studied by atomic force microscopy (AFM), scanning electron microscopy
(SEM), and optical microscopy (OM). AFM images of the coatings
were obtained in the tapping mode with a NTEGRA Spectra microscope
(NT-MDT, Russia) and with a CombiScope scanning probe microscope
(AIST-NT, Russia). The AFM images were processed using Gwyddion
open source software. OM images were obtained in the bright-eld
transmission mode with an Olympus IX71 microscope (part of the
NTEGRA Spectra Setup). For imaging and laser irradiation, a 100 /
0.90 M Plan FL N objective was used. An Altami MET 5 optical microscope was also used to obtain optical images of multilayer structures
(in the reection and transmission modes). All optical images were captured from the video signal of the microscope camera. SEM analysis was
performed with a MIRA II LMU scanning electron microscope (TESCAN,
Czech Republic) operated at an accelerating voltage of 730 kV.
A Lambda 950 UVvis spectrophotometer (PerkinElmer, USA) was
used to obtain UVvis absorbance spectra of the fabricated multilayer
lms.
For Raman spectroscopy analysis, a NTEGRA Spectra Setup (NT-MDT,
Russia) was used. A heliumneon laser (wavelength, 632.8 nm; maximum power, 5 mW) was used for sample irradiation.
Thermogravimetric analysis (TGA) was performed on a TGA Q500
V20.10 Build 36 thermogravimetric analyzer (TA Instruments, USA)
under argon gas ow. TGA measurements were performed between
24 and 1000 C, and the temperature increase rate was adjusted for
each sample according to the intrinsic properties of a given substance.
For PDADMAC, the temperature increase rate was 1 C min1 in the
range 25250 C and 10 C min1 in the range 250998 C; for CuPcTs,
it was 10 C min1 for the temperature ranges 25270 C and
6001000 C and was 1 C min1 for the range 270600 C.

Each step of phthalocyanine deposition changed the absorbance peak


in the lm's spectra.
Optical spectra (absorption and transmission) were measured
for the prepared lms after every bilayer deposition in the whole
process of multilayer structure fabrication. The absorbance spectra
of the PEI/CuPcTs/(PDADMAC/CuPcTs)54 structures (PDADMAC MW,
400500 kDa) fabricated on FTO-coated glass substrates are shown in
the inset in Fig. 1. The graphs demonstrate spectra of the deposited coating with the subtraction of the native spectrum of FTO-coated glass. In
good agreement with typical spectra of phthalocyanine [47,6668,
7073], the absorbance spectra had peaks in the range 300400 nm
(Soret band) and in the 550750 nm region (Q bands) (Fig. 1), with a
maximum at 330 nm and a dominant peak near 620 nm (with a prominent shoulder at 670 nm), respectively. Absorbance increased for all
peaks (inset in Fig. 1), and this increase was accompanied by a decrease
in transmission (data not shown). The dependence of the intensity
(absorbance peak at 620 nm) on the number of deposited bilayers
was linear throughout the fabrication of the hundred-layer structures
(Fig. 1). This means that the same amount of phthalocyanine molecules
was adsorbed onto the substrate during every PDADMAC/CuPcTs
bilayer deposition. This straight trend line is given by the equation
A = 0.048n (R2 = 0.99) (Fig. 1), where A is the absorbance at the
peak near 620 nm and n is the number of deposited PDADMAC/CuPcTs
bilayers.
As shown by many research groups [25,27,50,54,5761], the relative
molecular weight and the surface charge density () of polyelectrolytes
inuence the adsorption process. However, the effect of the polymer
MW on LbL assembly with CuPcTs has not been studied before. Thus,
three other relative MWs (b 100 kDa, 100200 kDa, and 200350 kDa)
were used in addition to the 400500 kDa PDADMAC to study the effect
of the polymer MW on lm deposition. The optical spectra of the multilayer lms fabricated using PDADMAC of three MWs were similar
in terms of the peak maximum position. Linear growth of intensity at
absorbance maxima with an increasing number of PDADMAC/CuPcTs
bilayers was observed for the 100500 kDa MW (Fig. 2(a)). However,
the slope magnitudes for the three MWs differed perceptibly. The
higher the relative MW was, the greater was the tilt angle, with the
maximal value being observed for the 400500 kDa MW (Fig. 2(b)).
These data demonstrate the inuence of the polymer MW on CuPcTs
adsorption: a higher concentration of CuPcTs has been found in lms
based on higher-MW polymers. This agrees with the previous results
[25,27] showing that lms are deposited better when polyelectrolytes
of higher MWs and with higher values are used. Indeed, the thickness

3. Results and discussion


3.1. Film fabrication
The actual work describes the formation of the hundred-layer lms
based on a water-soluble copper phthalocyanine dye and polyelectrolytes. The central metal atom and substituents on the benzene rings of
the phthalocyanine molecule inuenced UVvis absorbance; CuPcTs
had strong absorption bands around 650 nm (Q bands) [6367,
7073]. The characteristic maxima of CuPcTs were in the 300800 nm
range of the absorbance spectra, which was convenient for the analysis
of the semitransparent thin PEI/CuPcTs/(PDADMAC/CuPcTs)n lms.

Fig. 1. Dependence of the absorbance (at maximum near 620 nm) on the number of deposited bilayers n for PEI/CuPc/(PDADMAC/CuPc)54 lms; the inset demonstrates the corresponding absorbance spectra. MW of PDADMAC is 400500 kDa.

A.S. Sergeeva et al. / Thin Solid Films 583 (2015) 6069

63

Fig. 2. (a) Dependence of the absorbance (at maximum near 620 nm) on the number of deposited bilayers n for PEI/CuPc/(PDADMAC/CuPc)15 lms and (b) dependence of the slope
angle on MW of the PDADMAC for the straight lines in (a). Circles containing various number of PDADMAC molecules with the different chain length illustrate that at equal concentration of polymer solutions (2 mgmL1) the PDADMAC content is not equal; the amount
of molecules is different in solutions of PDADMAC with various MWs. MWs of PDADMAC
are b100 100200="" 200350="" kda="" and="" 400500="" span=""N.

of multilayers can be affected by various factors such as substrate properties, deposition conditions, and ionic strength of polyelectrolytes, and
it is always strongly determined by and MW of the polyelectrolytes
used [25]. The dependence of the deposited layer thickness on the
surface charge density is nonmonotonic for different adsorption
regimes [25]. Initially, it is inversely proportional to surface charge number density (1); in the second regime, it decreases according to 1/3
because the thickness of the adsorbed layer is determined by the energy
gain from electrostatic attraction to the charged surface and the connement entropy loss from chain compression [25]. Within the third
interval, increases as 1/3, and, nally, the thickness of the adsorbed
layer saturates at high surface charge densities when the surface counterions dominate the screening of the surface potential. Moreover, an
addition of salt lowers the amount of polymer adsorbed, because the
salt ions also take part in the screening of the surface charge [25]. At
the same time, surface overcharging exhibits a nonmonotonic saltconcentration dependence, rst increasing until a critical salt concentration and then decreasing with further rising salt concentration [25].
Thus, most high-molecular-weight polymers (including PDADMAC)
adsorb better owing to the larger surface charge number density,
which provides stronger surfacesubstance interaction because of the
larger amount of ions per chain [25,27]. This is why the 400500 kDa
PDADMAC adsorbed best, as compared with the lower-MW polyelectrolytes, and, as seen from the absorbance spectra (Figs. 1 and 2), the
CuPcTs concentration in the fabricated lms was the highest. Note

Fig. 3. (a) Absorbance spectra of the two identical PEI/CuPc/(PDADMAC/CuPc)54 lms fabricated using the various drying times during LbL assembly (the long-time drying means
1 min air drying after each layer deposition and 20 min after every 3 bilayers; the shorttime drying means 1 min air drying after each layer adsorption). (b) Absorbance spectra
of PEI/CuPc/(PDADMAC/CuPc)54 lms before and after annealing at different temperatures (200300 C) during 3 h. MW of PDADMAC is 400500 kDa.

that the amount of PDADMAC molecules was different for solutions


with the different MWs, despite the weight concentrations of the polymer solutions being equal (Fig. 2(b)). However, all PDADMAC solutions
had an excess concentration (2 mg mL 1), so adsorption did not
depend on the polymer concentration even if the amount of molecules
differed between solutions of the different MW polymers. The drying
time during the lm formation inuenced the deposition process also.
This fact can be proven by the analysis of absorbance spectra of two similarly structured (PEI/CuPcTs/(PDADMAC/CuPcTs)49 coatings fabricated
under different drying conditions (Fig. 3(a)). Long-time drying (1 min
for every layer, 20 min for every 3 bilayers) was applied in the fabrication of the rst lm, and the total time of coating formation was
9.5 days (including stand-by periods during the night). Short-time
drying (1 min for every layer) was used for the second lm, and the
total fabrication time was 4 days. As a result, the total absorbance of
the lm subjected to long-time drying was almost two times higher
than that for the lm subjected to short-time drying (Fig. 3(a)). The
difference in absorbance means that longer drying during LbL deposition provided better adsorption of the phthalocyanine molecules on
the surface (Fig. 3(a)).

64

A.S. Sergeeva et al. / Thin Solid Films 583 (2015) 6069

3.2. Film thermal treatment


As stated before, it would be better for photovoltaic applications to
use phthalocyanine in the -phase, because high crystallinity gives
good photoconductive properties. The -phase can be obtained from
an -lattice by high-temperature annealing, which improves the molecular structure of lms. LbL-assembled lms with a hundred layers are
amorphous; hence, annealing is necessary for an phase transition
of CuPcTs contained in the fabricated lms. Thermal treatment was performed at different temperatures to determine if any domain formation
or phase transition took place in the multilayer lms.
The annealing temperature regime for the fabricated hundredlayer lms was chosen for the following reason: it was reported that annealing of 80-nm-thick ZnPc lms at 200 C for 3 h leads to structure
modications and to a transition to the photoactive -phase [70].
This is the minimum known temperature for the transition
of thin (30150 nm) ZnPc lms [6670]. In our study, therefore,
200 C was chosen as a starting annealing point for PEI/CuPcTs/
(PDADMAC/CuPcTs)n lms. However, this temperature was insufcient
for an phase transition to occur, probably owing to the presence
of the PDADMAC matrix, which can affect the properties of phthalocyanine [47]. Further, the temperature was increased step by step until
phthalocyanine transformed into the -phase (at 300 C). The processes
taking place during the annealing of the hundred-layer PEI/CuPcTs/
(PDADMAC/CuPcTs)n lms were investigated and are described below.
The phase transition causes some changes in both crystallinity
and optical spectra of coatings, which helps to determine the phasetransition point [6670,76]. According to the obtained results, thermal
treatment caused a red shift of the absorbance peaks from 633 and
695 nm to 650 and 752 nm, respectively. Moreover, an intensity
decrease was observed for these peaks, as shown earlier [70,76].
Fig. 3(b) shows the absorbance spectra of the fabricated PEI/CuPcTs/
(PDADMAC/CuPcTs)n lms before and after annealing. The treatment
time was 3 h, and the temperature range was 200300 C. The background spectrum used for the analysis of these UVvis spectra was a
spectrum of pure phthalocyanine, because PDADMAC does not make a
strong contribution to absorbance within the analyzed wavelength
range [47,77,78]. Annealing at 200 C did not modify the molecular
structure of the lms, because there were no changes in the absorbance
spectra (Fig. 3(b)). This means that the presence of PDADMAC in the
multilayers affected the thermal sensitivity of CuPcTs and that as a
result of the PDADMAC/CuPcTs interactions, the phase-transition temperature increased [47]. A possible reason for this effect might be a
phase separation during the LbL assembly or a kind of domain formation
in the lms (Fig. 5). A 50 C increase in the annealing temperature
induced no signicant changes in the spectra (~ 12% absorbance decrease and 12-nm red shift for the peak near 615 nm), demonstrating
that the lm was highly stable. Under annealing at 270 C, the molecular
structure underwent a change and the absorbance decreased by 47%
(Q band maximum) without a wavelength shift (Fig. 3(b)). Raising
the annealing temperature to 300 C induced a stronger modication
of the crystal lattice, which was accompanied by an approximately
30-nm red shift of the Q-band peaks to longer wavelengths (from
615 nm to 645 nm and from 670 nm to 700 nm), with a decrease in
absorbance (to 62%) and with intensity redistribution between the two
peaks of the Q band (Fig. 3(b)). As explained above, such changes in
UVvis spectra of phthalocyanine indicate that an phase transition has occurred.
3.3. Raman spectra characterization
The annealing-mediated changes in the lm molecular structure
could be seen in the Raman scattering spectra (RS) owing to the inuence of thermal treatment on the chemical bonds within the lms.
Fig. 4 presents both calculated spectrum of CuPcTs by itself and spectra
of the PEI/CuPcTs/(PDADMAC/CuPcTs)n lms experimentally measured

Fig. 4. Raman scattering spectra of CuPcTs and PEI/CuPc/(PDADMAC/CuPc)54 lms


annealed at different temperatures during 3 h; spectra obtained using on-resonance
(633 nm) excitation. MW of PDADMAC is 400500 kDa.

by using resonance excitation (wavelength, 633 nm). Resonance


excitation was chosen because it excites the electron states of phthalocyanine, giving well-dened peaks over the frequency range examined
(5001800 cm 1). The strongest peaks were observed for chemical
bonds with a central copper atom, corresponding to Raman shifts of
756, 860, 978, 1158, 1200, 1293, 1341, 1459, and 1547 cm1.
All measured spectra had several indicative peaks between 500 and
1800 cm 1 corresponding to certain chemical bonds of CuPcTs, and
they were in the agreement with the results reported previously
[7072]. As seen from the obtained data, most of the band positions
observed in the Raman spectra were similar for the analyzed samples
(both as-prepared and annealed) in the examined frequency range
(5001800 cm1) (Fig. 4). Consequently, there was no strong difference
between the Raman spectra of the - and -phthalocyanines in the
5001800 cm1 interval. The shifts of several cm1 and the different
intensities were not sufcient to identify the CuPcTs crystalline structure
by Raman spectroscopy [72]. However, the preservation of the characteristic peaks of CuPcTs in Raman spectra could be evidence of the thermal
stability of the used phthalocyanine (resistance of a molecule to decomposition) in the examined temperature range [79,80].
Composite polymer-containing multilayers giving well-dened RS
can be used for molecular analysis by Raman spectroscopy [81,82].
Raman-sensitive platforms have been fabricated with CaCO3 microspheres [83]. Calcium carbonate microspheres can serve as templates
to prepare protein-loaded particles [8487], opening a way to employ
these complexes for theranostic applications.
3.4. OM, AFM, and SEM characterization
The morphology of the fabricated hundred-layer lms was examined by OM, AFM, and SEM to investigate the effect of thermal treatment. The optical images obtained for the prepared FTO/PEI/CuPcTs/
(PDADMAC/CuPcTs)n multilayers show a morphology typical of organic
LbL thin lms (Fig. 5(ab)). The surface of the multilayers is rough
and has no macroscopic defects. Both the reection-mode and the
transmission-mode images (Fig. 5(a,b)) show colored areas. In both
images, it is possible to distinguish small dark dots and a two-colored
network consisting of elongated islands (some islands are indicated
by the arrows in Fig. 5(a,b)). Various colors from pink to light blue
(Fig. 5(a)) are observed owing to the difference in density between
the corresponding areas: a denser structure provides a darker pattern
(black arrows in Fig. 5(a)), whereas softer places give light islands

A.S. Sergeeva et al. / Thin Solid Films 583 (2015) 6069

Fig. 5. Optical images of PEI/CuPc/(PDADMAC/CuPc)54 lms before (ab) and after (cf)
annealing at 300 C. MW of PDADMAC is 400500 kDa. Pictures were obtained using reection (a,c,e) and transmission (b,d,f) mode. Scale bar is 10 m.

(white arrows in Fig. 5(a)). Such a patterned surface can be explained


by the formation of domains within the lm. The dark grains in Fig. 5
might be large aggregates of CuPcTs molecules, whereas the pinkblue
patterns consist of pure PDADMAC islands (lighter) and CuPcTs
PDADMAC areas (darker). In fact, phthalocyanine molecules tend to

65

aggregate in a kind of phase separation, e.g., if they are surrounded by


a polymer matrix [29]. Under some conditions, phthalocyanine aggregation takes place in various complexes, including phthalocyanineaggregated Pluronic nanoparticles, used in photodynamic therapy
[88], and phthalocyaninefullerene complexes, used to fabricate bulk
heterojunctions for photovoltaic applications [6365].
Thermal treatment at 300 C led to a change in lm morphology
(Fig. 5(c,d)). The lm structure was light blue, with nanosized dark
dots (Fig. 5(c,d)) similar to the dots observed before annealing
(Fig. 5(a,b)). The colored network disappeared, probably because the
polymer matrix had melted, resulting in a uniform component redistribution within the lms. Moreover, some defects were formed in the
lms (Fig. 5(e,f)). This can be explained by the different thermal behaviors of pure CuPcTs aggregates, pure PDADMAC, and CuPcTsPDADMAC
complexes with various component ratios. Whereas thermal treatment
at 300 C led to an phase transition of the CuPcTs molecules, the
surrounding phases of CuPcTsPDADMAC and pure PDADMAC started
to melt (see the TGA data). Thus, thermal treatment gave rise to a liquid
polymeric phase with uniformly distributed nanosized solid CuPcTs aggregates (Fig. 5(c,d)). At the same time, the islands of pure PDADMAC
(white arrows in Fig. 5(a)) degraded at above 270 C (see the TGA
data) and microdefects were formed in the corresponding places
(Fig. 5(e,f)). The microcracks and pores in the annealed lms were better visible in the SEM images.
It was difcult to study the multilayer lms by AFM and SEM owing
to the granular nature of the conductive FTO coating that covered the
glass substrate (Figs. 6 and 7). The average size of the FTO granules
was 100500 nm, in agreement with the data described previously [74,
75]. It was possible to observe the microstructure of the hundred-layer
PEI/CuPcTs/(PDADMAC/CuPcTs)54 lms by AFM (Fig. 6). Thermal treatment induced the following changes in lm morphology (Fig. 6(bd)):
the prepared amorphous PEI/CuPcTs/(PDADMAC/CuPcTs)54 lms
were characterized by a surface nanopatterned with some spherical
nanograins ~ 30 nm in size (Fig. 6(b)). The spherical grains probably
were solid CuPcTs aggregates surrounded by the PDADMAC matrix
(dark dots in Fig. 5, bright dots in Fig. 6(b,c)). Such a surface pattern
is characteristic of phthalocyanine coatings, which typically consist
of small spherical particles [9,66].
Annealing the PEI/CuPcTs/(PDADMAC/CuPcTs)54 lms at 250 C
caused changes in coating morphology (Fig. 6(b,c)). The average

Fig. 6. AFM-images: (a) the granular conductive FTO-coated glass substrate; (b) the air dried PEI/CuPc/(PDADMAC/CuPc)54 lm on the FTO-coated substrate; (cd) the glass/FTO/PEI/
CuPc/(PDADMAC/CuPc)54 lm after thermal treatment at 250 C (c) and 300 C (d). Images (a) and (d) show a real surface relief; images (b) and (c) were treated with a high-pass lter
for the better morphology visualization. Scale bar is 200 nm.

66

A.S. Sergeeva et al. / Thin Solid Films 583 (2015) 6069

Fig. 7. SEM-images: (a) the granular conductive FTO-coated glass substrate; (b) the air
dried PEI/CuPc/(PDADMAC/CuPc)54 lm on the FTO-coated substrate; (c) the glass/FTO/
PEI/CuPc/(PDADMAC/CuPc)54 lm annealed at 300 C during 3 h.

size of the nanograins (aggregated CuPcTs molecules surrounded by


PDADMAC) was reduced from 30 to 20 nm (Fig. 6(b,c)). This can be
explained by the lm behavior under the applied temperature. First,
all water molecules evaporated from the lm. Moreover, as found by
TGA, the PDADMAC matrix did not degrade at 250 C, but the polymer
structure became more compact. As a result, solid aggregates of heatstable phthalocyanine molecules, which are dispersed in the polymer
matrix, were seen better (Fig. 6(b,c)).
As noted above, annealing of the hundred-layer structure at 300 C
led to changes in molecular structure (Fig. 5) and to an phase
transition of CuPcTs (Fig. 3(b)). The obtained AFM images show that
increasing the annealing temperature from 250 to 300 C modied
the lm morphology further. After thermal treatment at 250 C, the
PEI/CuPcTs/(PDADMAC/CuPcTs)54 lms became more compact and patterned with 20-nm nanograins. Conversely, after annealing at 300 C,
the 20-nm patterns disappeared (Fig. 6(d)) and the surface of the coated FTO granules became similar in appearance to that of the uncoated
ones (Figs. 6(a), 7(a)). This effect can be explained by the different
thermal sensitivities of the lm constituents (described in the rst
paragraph of this section). Briey, at 300 C, CuPcTs is stable, whereas
the PDADMAC/CuPcTs complex starts to melt and pure PDADMAC
degrades at above 270 C [89,90]. As a result, the melted polymer matrix
with dispersed CuPcTs aggregates spreads over the FTO surface to create
a uniform layer repeating the relief of the FTO granular surface.
The SEM images gave no information about the nanomorphology
of the freshly prepared lms, because the lm multilayer structure
was transparent for SEM observation. The blurred borders of the FTO
granules were the only visible result of layer deposition (Fig. 7(a,b)).
However, thermal treatment at 270300 C changed the structure of

Fig. 8. TG-curves measured for (a) PDADMAC of MW 400500 kDa (the temperature increase rate is 1 C/min for the range 25250 C and 10 C/min for the range 250998 C);
(b) CuPcTs (the temperature increase rate is 10 C/min for the ranges 25270 C and
6001000 C, and 1 C/min for the range 270600 C).

the hundred-layer coatings. The lms developed holes and microcracks,


which could be seen by SEM (Fig. 7(c)) and OM (Fig. 5). The formation
of such micropores was complex and was caused only by the melting
and decomposition of the pure polymer (white arrows in Fig. 5(a)) at
above 270 C (TG measurements, Fig. 8(a)). Indeed, annealing of CuPcTs
at 300 C could not destroy the lms or give rise to microcracks, because
this material is heat resistant and decomposes at above 450 C (TG
measurements, Fig. 8(b)). The presence of characteristic peaks in the
absorbance (insets in Figs. 1 and 3(b)) and Raman (Fig. 4) spectra
after annealing also conrms that CuPcTs was thermally stable. Most
polymers, including PDADMAC, are not heat resistant, and the organic
matrix decomposes completely at 300500 C, depending on the
precursors employed [8993]. Therefore, the only possible reason for
such microdefects to form in the PEI/CuPcTs/(PDADMAC/CuPcTs)n
lms (Figs. 5 and 7) was the degradation of the PDADMAC at above
270 C. As a result, the lms were broken in certain places (islands
of the pure polymer, white arrows in Fig. 5) and the FTO granules
could be seen through the emerged differently shaped and sized pores
(Fig. 7(c)).

3.5. Characterization by TGA


TGA was conducted to conrm whether the pores emerged because
of the nature of the PDADMAC polymer. TGA measurements were made
between 24 and 1000 C to determine the individual decomposition
temperatures of the 400500 kDa PDADMAC (Fig. 8(a)) and CuPcTs
(Fig. 8(b)). During annealing, the PDADMAC went through a sequence
of exothermic processes with continuous weight loss. It is possible to
detect three curve segments in Fig. 8(b) within the following temperature limits: 24260 C, 260470 C, and 4701000 C. These three

A.S. Sergeeva et al. / Thin Solid Films 583 (2015) 6069

segments are characterized by different tilt angles describing the specific behavior of the PDADMAC in the corresponding temperature range.
The rst segment (24270 C) demonstrates that the sample weight
was reduced to approximately 11% of its initial value. This slow and
uniform weight reduction can be due to the release of the inner water
molecules, making the polymer matrix more compact without degrading
it. These data agree with the obtained AFM images of the hundred-layer
lms, showing that annealing at 250 C made the lms textured and
compact (Fig. 6(b,c)).
The second curve segment (270470 C) shows a rapid weight loss
of ~ 77% (Fig. 8(a)). This strong constant decrease is responsible for
the decomposition of the PDADMAC matrix and for the oxidation of
organic carbonaceous residues [8993]. The PDADMAC polymer becomes soft and then liquid and its constituents are burned away [89].
These processes brought the sample to the nal stable weight of the
third curve segment (4701000 C) (Fig. 8(a)). The observed thermal
degradation of PDADMAC agrees with the previous results [89,90] and
explains the behavior of the multilayer PEI/CuPcTs/(PDADMAC/CuPcTs)n
lms under annealing.
The TG curve obtained for sulfonated phthalocyanine consists of
two main sections within the temperature ranges 20455 C and
4551000 C. The rst segment (20455 C) demonstrates that the
dye was resistant to temperatures of up to ~455 C (Fig. 8(b)). Between
20 and 455 C, the sample lost ~ 9% of its initial weight. Moreover, 8%
of this weight loss at 20200 C was associated with water evaporation
(similarly to what is observed in the TG curve of PDADMAC) (Fig. 8(b)).
Subsequently, all water molecules were released and the weight of
CuPcTs became almost constant at 200455 C. This effect could be
due to the heat resistance of phthalocyanines.
The second segment of the curve (4551000 C) is characterized by
constant weight loss during the whole period. Between 455 and 535 C,
the weight of the sample underwent a large (~15%) decrease (Fig. 8(b)).
Weight decreased by 40%, with several visible steps being observed
between 540 and 1000 C. Note that 1000 C is not enough to bring
the sample to the nal stable weight and even higher temperatures
are needed to decompose it. As a result, the total weight loss by CuPcTs
was 54% of its initial value after annealing at 241000 C (Fig. 8(b)).
In summary, the organic PDADMAC matrix decomposed completely
between 270 and 470 C (Fig. 8(a)), whereas CuPcTs was stable in this
range (Fig. 8(b)). This proves our assumption that the pores in the
annealed hundred-layer lms emerged in the areas of pure PDADMAC
(Fig. 5).
4. Conclusions
LbL assembly has been employed to fabricate highly absorbing
hundred-layer PEI/CuPcTs/(PDADMAC/CuPcTs)n lms (n, 1554) on
FTO-coated slides. Four PDADMAC MWs (b100 kDa, 100200 kDa,
200350 kDa, and 400500 kDa) have been used to study the effect
of the polymer MW on the adsorption of CuPcTs molecules. Film deposition has been monitored by absorbance spectra measurements. The
absorbance of the lms depended linearly on the number of deposited
PDADMAC/CuPcTs bilayers, which means that the same amount of
CuPcTs molecules was adsorbed at each deposition step. The absorbance of the lms was enhanced by increasing the MW of the
PDADMAC, with maximal absorbance observed for the highest-MW
polymer (400500 kDa). This might be because longer PDADMAC
molecules, having higher surface charge number density, adsorbed
more phthalocyanine molecules. Moreover, it has been possible to
enhance absorbance almost twofold by using long-time drying during
the LbL assembly. The thin hundred-layer lms were annealed at high
temperatures, leading to an phase transition in CuPcTs at
300 C. Thermal treatment has resulted in the formation of domains
within the lms that were composed of pure CuPcTs aggregates, pure
PDADMAC, and complex CuPcTs/PDADMAC mixtures, as found by
morphological and TG analyses.

67

Acknowledgments
This work was supported by an individual grant of the Foundation for
Assistance to Small Innovative Enterprises in Science and Technology
under the program U.M.N.I.K. (project no. 17259) and by grant of the
Russian Scientic Foundation (project no. 14-12-00275) (work related
to phase separation and the effect of thermal treatment on phase separation in the nanostructured lms). We thank Mikhail Trusov (AIST-NT)
for supplying the equipment used in this study and for helping with
AFM microscopy. We are also grateful to Ilya Gorbachev (Scientic Research Institute of Nanostructures and Biosystems, Saratov State University, Saratov, Russia) for his help with TG measurements and to Roman
Sergeev for his help with OM imaging of samples after thermal
treatment.
References
[1] M.G. Walter, A.B. Rudine, C.C. Wamser, Porphyrins and phthalocyanines in solar
photovoltaic cells, J. Porphyrins Phthalocyanines 14 (2010) 759.
[2] A. Varotto, C.-Y. Nam, I. Radivojevic, J.P.C. Tom, J.A.S. Cavaleiro, C.T. Black, C.M.
Drain, Phthalocyanine blends improve bulk heterojunction solar cells, J. Am.
Chem. Soc. 132 (2010) 2552.
[3] A.P. Yuen, A. Hor, J. Preston, R. Loutfy, Photovoltaic properties of M-phthalocyanine/
fullerene organic solar cells, Sol. Energy 86 (2012) 1683.
[4] S. Rajaputra, S. Vallurupalli, V.P. Singh, Copper phthalocyanine based Schottky diode
solar cells, J. Mater. Sci. Mater. Electron. 18 (2007) 1147.
[5] K. Yoshida, Fabrication and characterization of phthalocyanine/C.60. Solar cells with
inverted structure, Adv. Chem. Eng. Sci. 02 (2012) 461.
[6] L. Li, Q. Tang, H. Li, W. Hu, X. Yang, Z. Shuai, Y. Liu, D. Zhu, Organic thin-lm transistors of phthalocyanines, Pure Appl. Chem. 80 (2008) 2231.
[7] N.M. Bamsey, A.P. Yuen, A.-M. Hor, R. Klenkler, J.S. Preston, R.O. Loutfy, Integration of
an M-phthalocyanine layer into solution-processed organic photovoltaic cells for
improved spectral coverage, Sol. Energy Mater. Sol. Cells 95 (2011) 1970.
[8] G. Decher, J.D. Hong, Buildup of ultrathin multilayer lms by a self-assembly process:
II. Consecutive adsorption of anionic and cationic bipolar amphiphiles and polyelectrolytes on charged surfaces, Ber. Bunsenges. Phys. Chem. 95 (1991) 1430.
[9] J.-S. Lee, J. Cho, C. Lee, I. Kim, J. Park, Y.-M. Kim, H. Shin, J. Lee, F. Caruso, Layer-bylayer assembled charge-trap memory devices with adjustable electronic properties,
Nat. Nanotechnol. 2 (2007) 790.
[10] L.G. Paterno, M.A.G. Soler, F.J. Fonseca, J.P. Sinnecker, E.H.C.P. Sinnecker, E.C.D. Lima,
M.A. Novak, P.C. Morais, Layer-by-layer assembly of bifunctional nanolms: surfacefunctionalized maghemite hosted in polyaniline, J. Phys. Chem. C 113 (2009)
5087.
[11] Y.H. Kim, J. Park, P.J. Yoo, P.T. Hammond, Selective assembly of colloidal particles on
a nanostructured template coated with polyelectrolyte multilayers, Adv. Mater. 19
(2007) 4426.
[12] K. Ariga, J.P. Hill, Q. Ji, Layer-by-layer assembly as a versatile bottom-up
nanofabrication technique for exploratory research and realistic application, Phys.
Chem. Chem. Phys. 9 (2007) 2319.
[13] Y. Li, X. Wang, J. Sun, Layer-by-layer assembly for rapid fabrication of thick polymeric
lms, Chem. Soc. Rev. 41 (2012) 5998.
[14] N. Madaboosi, K. Uhlig, S. Schmidt, M.S. Jger, H. Mhwald, C. Duschl, D.V. Volodkin,
Microuidics meets soft layer-by-layer lms: selective cell growth in 3D polymer
architectures, Lab Chip 12 (2012) 1434.
[15] C.S. Peyratout, L. Dhne, Tailor-made polyelectrolyte microcapsules: from multilayers to smart containers, Angew. Chem. Int. Ed. 43 (2004) 3762.
[16] M.A. Pechenkin, H. Mhwald, D.V. Volodkin, pH- and salt-mediated response of
layer-by-layer assembled PSS/PAH microcapsules: fusion and polymer exchange,
Soft Matter 8 (2012) 8659.
[17] X.L. Chen, S.A. Jenekhe, Supramolecular self-assembly of three-dimensional nanostructures and microstructures: microcapsules from electroactive and photoactive
rodcoilrod triblock copolymers, Macromolecules 33 (2000) 4610.
[18] M.V. Kiryukhin, S.R. Gorelik, S.M. Man, G.S. Subramanian, M.N. Antipina, H.Y. Low,
G.B. Sukhorukov, Individually addressable patterned multilayer microchambers for
site-specic release-on-demand, Macromol. Rapid Commun. 34 (2013) 87.
[19] P. Yusan, I. Tuncel, V. Btn, A.L. Demirel, I. Erel-Goktepe, pH-responsive layer-bylayer lms of zwitterionic block copolymer micelles, Polym. Chem. 5 (2014) 3777.
[20] A.S. Sergeeva, D.A. Gorin, D.V. Volodkin, Polyelectrolyte microcapsule arrays:
preparation and biomedical applications, BioNanoScience 4 (2014) 1.
[21] D. Kohler, N. Madaboosi, M. Delcea, S. Schmidt, B.G. de Geest, D.V. Volodkin, H.
Mhwald, A.G. Skirtach, Patchiness of embedded particles and lm stiffness control
through concentration of gold nanoparticles, Adv. Mater. 24 (2012) 1095.
[22] A.G. Skirtach, D.V. Volodkin, H. Mhwald, Bio-interfaces-interaction of PLL/HA thick
lms with nanoparticles and microcapsules, ChemPhysChem 11 (2010) 822.
[23] R. von Klitzing, Internal structure of polyelectrolyte multilayer assemblies, Phys.
Chem. Chem. Phys. 8 (2006) 5012.
[24] D.V. Volodkin, R.v. Klitzing, Competing mechanisms in polyelectrolyte multilayer
formation and swelling: polycationpolyanion pairing vs. polyelectrolyteion
pairing, Curr. Opin. Colloid Interface Sci. 19 (2014) 25.
[25] A.V. Dobrynin, M. Rubinstein, Theory of polyelectrolytes in solutions and at surfaces,
Prog. Polym. Sci. 30 (2005) 1049.

68

A.S. Sergeeva et al. / Thin Solid Films 583 (2015) 6069

[26] S. Schwarz, S. Bratskaya, W. Jaeger, B.-R. Paulke, Effect of charge density, molecular
weight, and hydrophobicity on polycations adsorption and occulation of polystyrene latices and silica, J. Appl. Polym. Sci. 101 (2006) 3422.
[27] P. Somasundaran, Encyclopedia of Surface and Colloid Science, 2nd ed. Taylor &
Francis, New York, 2006.
[28] N. Madaboosi, K. Uhlig, M.S. Jger, H. Mhwald, C. Duschl, D.V. Volodkin,
Microuidics as a tool to understand the build-up mechanism of exponential-like
growing lms, Macromol. Rapid Commun. 33 (2012) 1775.
[29] K.M. Kadish, K.M. Smith, R. Guilard, Phthalocyanines, Properties and Materials,
Academic Press, Amsterdam, 2003.
[30] International Electron Devices Meeting, 2006 International Electron Devices Meeting,
San Francisco, CA, December 1113, 2006, Institute of Electrical and Electronics
Engineers, Piscataway, N.J., 2006
[31] X.L. Chen, J.Q. Sun, Fabrication of macroporous lms with closed honeycomb-like
pores from exponentially growing layer-by-layer assembled polyelectrolyte multilayers, Chem. Asian. J. 9 (2014) 2063.
[32] W. Lin, Y. Guan, Y. Zhang, J. Xu, X.X. Zhu, Salt-induced erosion of hydrogen-bonded
layer-by-layer assembled lms, Soft Matter 5 (2009) 860.
[33] J.D. Mendelsohn, C.J. Barrett, V.V. Chan, A.J. Pal, A.M. Mayes, M.F. Rubner, Fabrication
of microporous thin lms from polyelectrolyte multilayers, Langmuir 16 (2000)
5017.
[34] J.Y. Kim, K. Lee, N.E. Coates, D. Moses, T.-Q. Nguyen, M. Dante, A.J. Heeger, Efcient
tandem polymer solar cells fabricated by all-solution processing, Science 317
(2007) 222.
[35] H. Tokuhisa, P. Hammond, Solid-state photovoltaic thin lms using TiO2, organic
dyes, and layer-by-layer polyelectrolyte nanocomposites, Adv. Funct. Mater. 13
(2003) 831.
[36] J.K. Mwaura, M.R. Pinto, D. Witker, N. Ananthakrishnan, K.S. Schanze, J.R. Reynolds,
Photovoltaic cells based on sequentially adsorbed multilayers of conjugated poly(pphenylene ethynylene)s and a water-soluble fullerene derivative, Langmuir 21
(2005) 10119.
[37] K.Y.K. Man, H.L. Wong, W.K. Chan, A.B. Djurii, E. Beach, S. Rozeveld, Use of a
ruthenium-containing conjugated polymer as a photosensitizer in photovoltaic
devices fabricated by a layer-by-layer deposition process, Langmuir 22 (2006) 3368.
[38] W.K. Chan, K. Man, K.W. Cheng, H.L. Wong, A. Djurisic, Fabrication of organic photovoltaic devices by the layer-by-layer polyelectrolyte deposition method, IEEE 2
(2005) 221.
[39] J. You, L. Dou, K. Yoshimura, T. Kato, K. Ohya, T. Moriarty, K. Emery, C.-C. Chen, J. Gao,
G. Li, Y. Yang, A polymer tandem solar cell with 10.6% power conversion efciency,
Nat. Commun. 4 (2013) 1.
[40] G. Li, R. Zhu, Y. Yang, Polymer solar cells, Nat. Photonics 6 (2012) 153.
[41] J. Zhu, C.-M. Hsu, Z. Yu, S. Fan, Y. Cui, Nanodome solar cells with efcient light
management and self-cleaning, Nano Lett. 10 (2010) 1979.
[42] P. Uthirakumar, J.H. Kang, S. Senthilarasu, C.-H. Hong, The different types of ZnO
materials on the performance of dye-sensitized solar cells, Physica E 43 (2011)
1746.
[43] T. Minemoto, H. Takakura, Y. Hamakawa, Chemical bath deposition for the fabrication of antireective coating of spherical silicon solar cells, Sol. Energy Mater. Sol.
Cells 90 (2006) 3576.
[44] D. Zhou, R. Biswas, Photonic crystal enhanced light-trapping in thin lm solar cells,
J. Appl. Phys. 103 (2008) 93102.
[45] H.A. Atwater, A. Polman, Plasmonics for improved photovoltaic devices, Nat. Mater.
9 (2010) 205.
[46] W. Regan, S. Byrnes, W. Gannett, O. Ergen, O. Vazquez-Mena, F. Wang, A. Zettl,
Screening-engineered eld-effect solar cells, Nano Lett. 12 (2012) 4300.
[47] V. Zucolotto, M. Ferreira, M.R. Cordeiro, C.J.L. Constantino, D.T. Balogh, A.R. Zanatta,
W.C. Moreira, O.N. Oliveira, Unusual interactions binding iron tetrasulfonated
phthalocyanine and poly(allylamine hydrochloride) in layer-by-layer lms, J. Phys.
Chem. B 107 (2003) 3733.
[48] D. Li, A. Bishop, Y. Gim, X.B. Shi, M.R. Fitzsimmons, Q.X. Jia, Conduction properties of
metal/organic monolayer/semiconductor heterostructures, Appl. Phys. Lett. 73
(1998) 2645.
[49] K. Shinbo, K. Kato, F. Kaneko, K. Onishi, R.C. Advincula, X. Fan, Fabrication and
electrochromic properties of layer-by-layer self-assembled ultrathin lms containing
water-soluble phthalocyanine, Mol. Cryst. Liq. Cryst. 407 (2003) 97.
[50] C. Porcel, P. Lavalle, G. Decher, B. Senger, J.-C. Voegel, P. Schaaf, Inuence of the polyelectrolyte molecular weight on exponentially growing multilayer lms in the linear
regime, Langmuir 23 (2007) 1898.
[51] D. Volodkin, A. Skirtach, H. Mhwald, LbL lms as reservoirs for bioactive molecules,
Adv. Polym. Sci. 240 (2011) 135.
[52] S. Schmidt, N. Madaboosi, K. Uhlig, D. Khler, A. Skirtach, C. Duschl, H. Mhwald,
D.V. Volodkin, Control of cell adhesion by mechanical reinforcement of soft polyelectrolyte lms with nanoparticles, Langmuir 28 (2012) 7249.
[53] Z. Tang, Y. Wang, P. Podsiadlo, N.A. Kotov, Biomedical applications of layer-by-layer
assembly: from biomimetics to tissue engineering, Adv. Mater. 18 (2006) 3203.
[54] T. Boudou, T. Crouzier, R. Auzly-Velty, K. Glinel, C. Picart, Internal composition
versus the mechanical properties of polyelectrolyte multilayer lms: the inuence
of chemical cross-linking, Langmuir 25 (2009) 13809.
[55] D. Volodkin, A. Skirtach, H. Mhwald, Bioapplications of light-sensitive polymer
lms and capsules assembled using the layer-by-layer technique, Polym. Int. 61
(2012) 673.
[56] D. Volodkin, A. Skirtach, N. Madaboosi, J. Blacklock, R. von Klitzing, A. Lankenau, C.
Duschl, H. Mhwald, IR-light triggered drug delivery from micron-sized polymer
biocoatings, J. Control. Release 148 (2010) e70.
[57] T. Boudou, T. Crouzier, K. Ren, G. Blin, C. Picart, Multiple functionalities of polyelectrolyte multilayer lms: new biomedical applications, Adv. Mater. 22 (2010) 441.

[58] B. Schoeler, G. Kumaraswamy, F. Caruso, Investigation of the inuence of polyelectrolyte charge density on the growth of multilayer thin lms prepared by the
layer-by-layer technique, Macromolecules 35 (2002) 889.
[59] Z. Sui, D. Salloum, J.B. Schlenoff, Effect of molecular weight on the construction
of polyelectrolyte multilayers: stripping versus sticking, Langmuir 19 (2003)
2491.
[60] B. Sun, C.M. Jewell, N.J. Fredin, D.M. Lynn, Assembly of multilayered lms using welldened, end-labeled poly(acrylic acid): inuence of molecular weight on exponential
growth in a synthetic weak polyelectrolyte system, Langmuir 23 (2007) 8452.
[61] L. Shen, P. Chaudouet, J. Ji, C. Picart, pH-amplied multilayer lms based on
hyaluronan: inuence of HA molecular weight and concentration on lm growth
and stability, Biomacromolecules 12 (2011) 1322.
[62] R. Lingstrm, L. Wgberg, Polyelectrolyte multilayers on wood bers: inuence of
molecular weight on layer properties and mechanical properties of papers from
treated bers, J. Colloid Interface Sci. 328 (2008) 233.
[63] H. Wei, J. Li, Z. Xu, Y. Cai, J. Tang, Q.Y. Li, Thermal annealing-induced vertical phase
separation of copper phthalocyanine: fullerene bulk heterojunction in organic
photovoltaic cells, Appl. Phys. Lett. 97 (2010) 83302-1.
[64] C. Schnemann, D. Wynands, L. Wilde, M.P. Hein, S. Pftzner, C. Elschner, K.-J.
Eichhorn, K. Leo, M. Riede, Phase separation analysis of bulk heterojunctions in
small-molecule organic solar cells using zinc-phthalocyanine and C{60}, Phys. Rev.
B 85 (2012) 245314.
[65] J. Jiang, . Bekarolu, Functional Phthalocyanine Molecular Materials, Springer,
Berlin, 2010.
[66] S. Senthilarasu, S. Velumani, R. Sathyamoorthy, G. Canizal, P. Sebastian, J. Chavez, R.
Perez, A. Subbarayan, J. Ascencio, Characterization of zinc phthalocyanine (ZnPc) for
photovoltaic applications, Appl. Phys. A Mater. Sci. Process. 77 (2003) 383.
[67] S. Senthilarasu, S. Baek, S. Chavhan, J. Lee, S. Lee, Effect of temperature on stacking
orientations of zinc phthalocyanine thin lms, J. Nanosci. Nanotechnol. 8 (2008)
5414.
[68] M. Kozlik, S. Paulke, M. Gruenewald, R. Forker, T. Fritz, Determination of the optical
constants of - and -zinc (II)-phthalocyanine lms, Org. Electron. 13 (2012) 3291.
[69] Y.-L. Lee, W.-C. Tsai, C.-H. Chang, Y.-M. Yang, Effects of heat annealing on the lm
characteristics and gas sensing properties of substituted and un-substituted copper
phthalocyanine lms, Appl. Surf. Sci. 172 (2001) 191.
[70] L. Gaffo, M.R. Cordeiro, A.R. Freitas, W.C. Moreira, E.M. Girotto, V. Zucolotto, The
effects of temperature on the molecular orientation of zinc phthalocyanine lms,
J. Mater. Sci. 45 (2010) 1366.
[71] D.N. Bratashov, A. Masic, A.M. Yashchenok, M.F. Bedard, O.A. Inozemtseva, D.A.
Gorin, T. Basova, T.K. Sievers, G.B. Sukhorukov, M. Winterhalter, H. Mhwald, A.G.
Skirtach, Raman imaging and photodegradation study of phthalocyanine containing
microcapsules and coated particles, J. Raman Spectrosc. 42 (2011) 1901.
[72] C. Defeyt, P. Vandenabeele, B. Gilbert, J. Van Pevenage, R. Cloots, D. Strivay, Contribution to the identication of a-, b- and -copper phthalocyanine blue pigments in
modern artists' paints by X-ray powder diffraction, attenuated total reectance
micro-Fourier transform infrared spectroscopy and micro-Raman spectroscopy,
J. Raman Spectrosc. 43 (2012) 1772.
[73] B. Schechtman, W. Spicer, Near infrared to vacuum ultraviolet absorption spectra
and the optical constants of phthalocyanine and porphyrin lms, J. Mol. Spectrosc.
33 (1970) 28.
[74] N. Noor, I.P. Parkin, Enhanced transparent-conducting uorine-doped tin oxide
lms formed by Aerosol-Assisted Chemical Vapour Deposition, J. Mater. Chem. C 1
(2013) 984.
[75] K. Subba Ramaiah, V. Sundara Raja, Structural and electrical properties of uorine
doped tin oxide lms prepared by spray-pyrolysis technique, Appl. Surf. Sci. 253
(2006) 1451.
[76] B. Clifford I-I. Grifths, P. Michael S. Walker, Direct alpha to X phase conversation of
metal containing phthalocyanine, USA patent, 1973.
[77] G. Yang, Z. Wu, P. Zhang, Study on the tribological behaviors of polyelectrolyte
multilayers containing copper hydroxide nanoparticles, Tribol. Lett. 25 (2006) 55.
[78] T. Zeng, R. Claus, Y. Liu, F. Zhang, W. Du, K.L. Cooper, Piezoelectric ultrathin polymer
lms synthesized by electrostatic self-assembly processing, Smart Mater. Struct. 9
(2000) 801.
[79] S.A. Znoiko, V.E. Maizlish, G.P. Shaposhnikov, N.S. Lebedeva, E.A. Mal'kova, Stability
of benzotriazolyl-substituted phthalocyanines with respect to thermal oxidative
decomposition, Russ. J. Phys. Chem. 87 (2013) 352.
[80] G. Lbbert, Phthalocyanines, Ullmann's Encyclopedia of Industrial Chemistry, BASF
Aktiengesellschaft, Ludwigshafen, Federal Republic of Germany, 2000.
[81] M. Harz, M. Kiehntopf, S. Stckel, P. Rsch, T. Deufel, J. Popp, Analysis of single blood
cells for CSF diagnostics via a combination of uorescence staining and microRaman spectroscopy, Analyst 133 (2008) 1416.
[82] C. Krafft, B. Dietzek, J. Popp, Raman and CARS microspectroscopy of cells and tissues,
Analyst 134 (2009) 1046.
[83] I.Y. Stetciura, A.V. Markin, A.N. Ponomarev, A.V. Yakimansky, T.S. Demina, C.
Grandls, D.V. Volodkin, D.A. Gorin, New surface-enhanced Raman scattering
platforms: composite calcium carbonate microspheres coated with astralen and silver
nanoparticles, Langmuir 29 (2013) 4140.
[84] D. Volodkin, Colloids of pure proteins by hard templating, Colloid Polym. Sci. 292
(2014) 1249.
[85] A.I. Petrov, D.V. Volodkin, G.B. Sukhorukov, Protein-calcium carbonate coprecipitation:
a tool for protein encapsulation, Biotechnol. Prog. 21 (2005) 918.
[86] S. Schmidt, D. Volodkin, Microparticulate biomolecules by mild CaCO3 templating,
J. Mater. Chem. B 1 (2013) 1210.
[87] S. Schmidt, M. Behra, K. Uhlig, N. Madaboosi, L. Hartmann, C. Duschl, D. Volodkin,
Mesoporous protein particles through colloidal CaCO3 templates, Adv. Funct.
Mater. 23 (2013) 116.

A.S. Sergeeva et al. / Thin Solid Films 583 (2015) 6069


[88] C.-K. Lim, J. Shin, Y.-D. Lee, J. Kim, K.S. Oh, S.H. Yuk, S.Y. Jeong, I.C. Kwon, S. Kim,
Phthalocyanine-aggregated polymeric nanoparticles as tumor-homing nearinfrared absorbers for photothermal therapy of cancer, Theranostics 2 (2012) 871.
[89] S. Francis, L. Varshney, S. Sabharwal, Thermal degradation behavior of radiation
synthesized polydiallyldimethylammonium chloride, Eur. Polym. J. 43 (2007) 2525.
[90] B. Sadeghi, A. Pourahmad, Synthesis of silver/poly (diallyldimethylammonium
chloride) hybride nanocomposite, Adv. Powder Technol. 22 (2011) 669.

69

[91] K. Pielichowski, J. Njuguna, Thermal Degradation of Polymeric Materials, Rapra


Technology, Shawbury, 2005.
[92] P.J. DiNenno, SFPE Handbook of Fire Protection Engineering, 2nd ed. National Fire
Protection Association; Society of Fire Protection Engineers, Quincy, Mass, Boston,
Mass, 1995.
[93] P.M. Visakh, Y. Arao, Thermal Degradation of Polymer Blends, Composites, and
Nanocomposites, Springer Verlag, Switzerland, 2014.

Potrebbero piacerti anche