directing the self-assembly of mesostructured hybrid materials:...

15
Directing the Self-Assembly of Mesostructured Hybrid Materials: Effect of Polymer Concentration and Solvent Type a Torsten Pietsch, Ezzeldin Metwalli, Stephan Volker Roth, Ronald Gebhardt, Nabil Gindy, Peter Mu ¨ller-Buschbaum,* Amir Fahmi* Introduction Well-defined surface patterns are required in numerous applications ranging from simple coatings to biotechnol- ogy, sensors, optics, and electronics. In principle there are two main approaches to fabricate these patterns—top- down and bottom-up. The top-down approach relies on structuring larger entities using a complex sequence of process steps including lithography, etching, and deposi- tion techniques. This concept has been well proven to enable the fabrication of surface patterns with complex geometries and a high degree of perfection. [1] However, the onward request for smaller dimensions to meet the requirements of future applications may not be accom- plished by means of conventional techniques. Thus, the bottom-up approach is highly beneficial to create nano- structures with dimensions in the range of 1–100 nm by assembling basic building blocks into complex architec- tures. Full Paper T. Pietsch, N. Gindy, A. Fahmi School of Mechanical, Materials, and Manufacturing Engineering; The University of Nottingham; University Park, Nottingham, NG7 2RD, United Kingdom Fax: (þ44) 1159514140; E-mail: [email protected]; [email protected]; [email protected] E. Metwalli, P. Mu ¨ller-Buschbaum Technische Universita ¨t Mu ¨nchen, Physik-Department, LS E13, James-Franck-Str.1, 85747 Garching, Germany E-mail: [email protected]; [email protected] S. V. Roth HASYLAB at DESY, Notkestrasse 85, 22603 Hamburg, Germany E-mail: [email protected] R. Gebhardt ESRF, BP 220, F-38043 Grenoble Cedex 09, France E-mail: [email protected] a : Supporting information for this article is available at the bottom of the article’s abstract page, which can be accessed from the journal’s homepage at http://www.mcp-journal.de, or from the author. Herein we explore the possibility to control the fabrication of non-equilibrium nano-patterns of spin-coated organic-inorganic hybrid materials based on diblock copolymers and metal nanoparticles in thin films. It is demonstrated that the type of solvent and the initial solution concentration, among other factors, can serve as tools to direct the morphology of spin-coated thin films. The driv- ing forces leading to the pattern formation are reviewed with respect to these parameters—type of solvents and polymer concentration. As a result well-defined surface patterns of functional hybrid materials are obtained. Moreover, the same tools used to direct the pattern for- mation can be applied to gain control over the particle size and size distribution. 864 Macromol. Chem. Phys. 2009, 210, 864–878 ß 2009 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim DOI: 10.1002/macp.200900099

Upload: others

Post on 16-Jul-2020

5 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Directing the Self-Assembly of Mesostructured Hybrid Materials: …bib-pubdb1.desy.de/record/92824/files/Pietsch_et_al-2009... · 2016-03-16 · Directing the Self-Assembly of Mesostructured

Full Paper

864

Directing the Self-Assembly of MesostructuredHybrid Materials: Effect of PolymerConcentration and Solvent Typea

Torsten Pietsch, Ezzeldin Metwalli, Stephan Volker Roth, Ronald Gebhardt,Nabil Gindy, Peter Muller-Buschbaum,* Amir Fahmi*

Herein we explore the possibility to control the fabrication of non-equilibrium nano-patternsof spin-coated organic-inorganic hybrid materials based on diblock copolymers and metalnanoparticles in thin films. It is demonstrated that the type of solvent and the initial solutionconcentration, among other factors, can serve as tools todirect the morphology of spin-coated thin films. The driv-ing forces leading to the pattern formation are reviewedwith respect to these parameters—type of solvents andpolymer concentration. As a result well-defined surfacepatterns of functional hybrid materials are obtained.Moreover, the same tools used to direct the pattern for-mation can be applied to gain control over the particle sizeand size distribution.

T. Pietsch, N. Gindy, A. FahmiSchool of Mechanical, Materials, and Manufacturing Engineering;The University of Nottingham; University Park, Nottingham, NG72RD, United KingdomFax: (þ44) 1159514140;E-mail: [email protected]; [email protected];[email protected]. Metwalli, P. Muller-BuschbaumTechnische Universitat Munchen, Physik-Department, LS E13,James-Franck-Str.1, 85747 Garching, GermanyE-mail: [email protected]; [email protected]. V. RothHASYLAB at DESY, Notkestrasse 85, 22603 Hamburg, GermanyE-mail: [email protected]. GebhardtESRF, BP 220, F-38043 Grenoble Cedex 09, FranceE-mail: [email protected]

a : Supporting information for this article is available at the bottom ofthe article’s abstract page, which can be accessed from the journal’shomepage at http://www.mcp-journal.de, or from the author.

Macromol. Chem. Phys. 2009, 210, 864–878

� 2009 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

Introduction

Well-defined surface patterns are required in numerous

applications ranging from simple coatings to biotechnol-

ogy, sensors, optics, and electronics. In principle there are

two main approaches to fabricate these patterns—top-

down and bottom-up. The top-down approach relies on

structuring larger entities using a complex sequence of

process steps including lithography, etching, and deposi-

tion techniques. This concept has been well proven to

enable the fabrication of surface patterns with complex

geometries and a high degree of perfection.[1] However, the

onward request for smaller dimensions to meet the

requirements of future applications may not be accom-

plished by means of conventional techniques. Thus, the

bottom-up approach is highly beneficial to create nano-

structures with dimensions in the range of 1–100 nm by

assembling basic building blocks into complex architec-

tures.

DOI: 10.1002/macp.200900099

Page 2: Directing the Self-Assembly of Mesostructured Hybrid Materials: …bib-pubdb1.desy.de/record/92824/files/Pietsch_et_al-2009... · 2016-03-16 · Directing the Self-Assembly of Mesostructured

Directing the Self-Assembly of Mesostructured Hybrid Materials . . .

Self-assembly is a bottom-up approach that generates

regular structures byminimizing the free energy balance of

a system.[2] The advantage of self-assembly lies in its

simplicity as a fast and cost-effectiveway to fabricatewell-

defined nanostructures. A system that has been reviewed

extensively during the past decade is A-B diblock copoly-

mers, itsability toself-assemble stemsfromthemicrophase

separation,which isa resultof thechemical incompatibility

of its building blocks (A and B). Depending on the relative

block length of the blocks (A and B) mainly five different

morphologies can be obtained in the bulk material.[3] A

much larger diversity of morphologies was discovered for

ABA and ABC triblock copolymers.[4] When the polymer is

confined to a thin film, the number of possible morphol-

ogies increases due to the interaction with the underlying

substrate.[5] However, targeting a certain thin film mor-

phology remains rather demanding. Recently, great efforts

have been made for example to control the orientation of

cylinders in a block copolymer thin film perpendicular to

thefilm’s surface; anorientation thatdoesnot represent the

most stable configuration.[6] Such a high degree of control is

essential, since the fabrication of nearly defect-free films

with a preferential orientation (anisotropy) of the micro-

phase separated domains is required in numerous applica-

tions such as coatings, optics, and electronics among many

others.

Herein we focus on controlling the fabrication of non-

equilibrium nanostructures; a possibility that until now

received little attention. Such non-equilibrium patterns in

thin films are obtained during solvent evaporation when a

copolymer solution is spin-coated onto a solid substrate.

Theadvantageover simplemicrophase separated thinfilms

is that amuch larger diversityofmorphologies is accessible.

An attempt to explore the formation of nano-patterns in

spin-coated copolymer filmswas presented by Zhao et al.[7]

They found that both the solvent and the nature of the

substrate significantly affect the microstructure in thin

film. Mechanisms to control the morphology were also

explored by Muller-Buschbaum et al.[8] who studied the

interplay between (micro-)phase-separation and dewet-

ting of a thin (co-)polymer film. Despite some recent efforts

attempting to evaluate the role of individual factors, there

are only few studies covering a wider range of parameters.

Hence, controlling the microstructure of self-assembled

thin films is still a challenging aim, especially for non-

equilibrium morphologies. Additionally, the particularly

interesting case of using hybrid materials to generate

functionalnanostructureshas received little attention. Self-

assembled block copolymer thin films are interesting from

a scientific point of view but possess only limited

functionality. As one of the first, Brinker et al.[9] developed

a simple concept to self-assemble functionalmaterials into

nanoscale hierarchical architectures by means of sol-gel

chemistry. This approach, known as evaporation-induced

Macromol. Chem. Phys. 2009, 210, 864–878

� 2009 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

self-assembly (EISA) is usually applied to fabricate meso-

structured thin films of silica/surfactant hybrid materials.

The self-assembly process is dependent on a large number

of parameters such as chemical composition and the

concentration of both the surfactant and the inorganic

precursor, the wetting behavior of the sol on the substrate,

as well as the humidity.[10] Control the formation of

mesostructures in thin films using the EISA approach

becomes increasingly difficult as the film thickness

decreases.[11] However, a number of reports demonstrate

the successful preparation of mesoporous structures in

ultra-thin films.[11,12]

The present work explores the possibility to fabricate

and control well-defined nanostructures in ultra-thin films

of functional hybrid materials based on amphiphilic

diblock copolymers and metallic nanoparticles. Addition-

ally the block copolymer not only stabilizes the nanopar-

ticles but allows to control the self-assembly. A brief

overview over some of the mechanism driving the self-

assembly in thin films is given with respect to the control

parameters. It is demonstrated that themorphology of self-

assembled thin films can be directed by a narrow selection

of tools including the nature of the solvent and the initial

polymer concentration. Simultaneously, these parameters

are used to control the particle size and size distribution

resulting in a flexible approach to design functional nano-

patterns.

Recently it was established that microphase separated

block copolymers can be used to template the preparation

of inorganicmaterials.[13] This approach already provides a

flexible way to generate well-defined and periodic

nanostructures of inorganic materials. However, multiple

step processes are required that increase both the complex-

ity and time consumption of this technology, while

the number of possible morphologies is still limited.

Wepresent, herein, a simpler approachwhich is to generate

inorganic nanoparticles in situ and simultaneously

direct their spatial arrangement through the self-assembly

of the diblock copolymer. Wiesner and coworkers[14]

applied this concept to generate inorganic nano-objects

in a block copolymer matrix by means of sol-gel synthesis.

An early way to generate inorganic nanostructures in

thin film was presented by Moller and coworkers,[15] who

used metal loaded block copolymer micelles to create

hexagonally ordered arrays of metallic nanoparticles. The

micelle geometry provides an excellent nano-reactor to

synthesize and stabilize inorganic nanoparticles with

controlled size and size distribution. Recently a number

of research groups presented fairly similar routes to

generate metallic nanostructures by depositing metallic

elements from copolymeric micelles.[16] For example, our

recent work demonstrated the fabrication of nanoporous

thin films by controlled deposition of metal loaded

micelles.[17]

www.mcp-journal.de 865

Page 3: Directing the Self-Assembly of Mesostructured Hybrid Materials: …bib-pubdb1.desy.de/record/92824/files/Pietsch_et_al-2009... · 2016-03-16 · Directing the Self-Assembly of Mesostructured

T. Pietsch et al.

866

The scope of this article is to control the fabrication of

nanostructured hybrid materials in thin film. The nano-

patterns are generated by spin-coating solutions of the

hybrid material onto solid substrates. It is demonstrated

that the type of solvent and the polymer concentration can

serve as tools to direct the pattern formation. The same

parameters are used to control the particle size and size

distribution. Simultaneously, mechanisms are explored

that can be used to guide the self-assembly into specific

surface patterns.

Experimental Part

Block Copolymer Solutions

Polystyrene-block-poly(4-vinyl pyridine) (PS327-b-P4VP28)

[Mn(PS)¼ 34000 g �mol�1; Mn(P4VP)¼ 2 900 g �mol�1] was pur-

chased from Polymer Source Inc. and used as received. The

polydispersity index was Mw=Mn ¼1.05. Solutions of the block

copolymer were prepared at various concentrations between 0.01

and 100.0mg �mL�1 using different solvents and solventmixtures.

The solvents were purchased from Fisher Scientific and used

without further purification. Metallic nanoparticles were gener-

ated in situ via complexation and reduction of an inorganic

precursor within the copolymer scaffold.

Dynamic Light Scattering

Dynamic light scattering (M802 DLS, Viscotek) was performed to

determine the radius of gyration and the hydrodynamic radius of

the copolymer-hybrid material in solution.

Thin Film Preparation

Oxidized siliconwafers (h111i) were used as a substrate; theywere

treated in a piranha solution (H2SO4/H2O2/H2O¼2:1:1) at 80 8C for

20 min to remove organic contaminations from the surface. A

second cleaning step in RCA-1 solution (H2O2/NH4OH/H2O¼ 1:1:1)

wasperformedat 60 8C for another 20min. After rinsing in distilled

water the substrateswereblownwithdryair. Theblock-copolymer

solutionswere spin-coated on the substrates at 3 000 rpm. The film

thickness was measured using atomic force microscopy (AFM) by

partially removing the film with a razor blade (SI-1).

Atomic Force Microscopy

The morphological characterization was performed by Tapping

ModeTM AFM in air at ambient conditions using a Dimension IVa

Nanoscope (Digital Instruments). We used standard silicon

cantilevers with a resonance frequency of about 330 kHz, a spring

constant of 45 N �m�1, and a tip radius of less than 10 nm

(PointprobeSPMCantilevers,Nanoworld). Theoperating frequency

was chosen to be far on the repulsive side of the resonance

frequency to increase scanning performance and stability.

Macromol. Chem. Phys. 2009, 210, 864–878

� 2009 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

Transmission Electron Microscopy (TEM and HR-TEM)

Thin films were produced by spin-coating the polymer solution on

carbon coated copper grids (400mesh/AGAR Scientific) attached to

a solid substrate support. The copper gridwas then peeled off from

the substrate and analyzed in a TECNAI Biotwin (FEI Ltd.)

transmission electron microscope at 100 k � eV�1. The instrument

was operated at low beam intensities to prevent electron damage

of the polymer samples. HR-TEM measurements were performed

using a JEOL JEM-2110F operated at 200 kV; the instrument was

equipped with an Orius SC1000 digital camera.

X-Ray Diffraction

Powder diffraction (XRD) measurements were performed on a

Bruker D8 Advance. The instrument was operated with CuKa1

(1.54056 A) and a Vario monochromator.

Sub-Microbeam Grazing Incidence Small Angle

X-Ray Scattering

The sub-microbeamgrazing incidence small angle X-ray scattering

(sub-mGISAXS) experimentswere performed at the ID13 beamline

(ESRF, Grenoble).[18] Awavelength of l¼0.0997 nmwas chosen. To

emphasize on small lateral structures the detector (MARCCD,

2048�2048 pixels array) was placed at 0.84m from the sample. A

fixed beamstop was used to shield the high intensity of the direct

beam and therefore to protect the detector. With a movable and

semi-transparent beamstop the specular peak was shielded. Thus

the non-specular as well as the specular intensity was recorded a

function of the exit angle af and the out-of plane angle c. A fixed

incident angle ofai¼0.638was selected to allow for a goodangular

separation between the Yoneda peak and the specular peak.

Successively increasing counting times were used in combination

with a lateral shifting of the samplewith respect to the x-ray beam

to test for radiation damage and to ensure absence of radiation

damage in the measurements. Due to the used sub-micrometer

beamsize it was possible to avoid over illumination of small

samples.

X-Ray Photoelectron Spectroscopy

The samples were analyzed in a Kratos AXIS ULTRA instrument

with amonochromated Al KaX-ray source (1 486.6 eV) operated at

15 mA emission current and 10 kV anode potential. For non-

conductingsamplesachargeneutralizerfilamentabovethesample

surface gives a flux of low energy electrons providing uniform

charge neutralization. The ULTRA was used in FAT (fixed analyzer

transmission)mode,withpass energy of 80 eV forwide scans.High

resolution scans with pass energy of 20 eV were collected at an

electronemissionangleof908withrespect to the sampleplane.The

sampling depth at this take-off angle was estimated to be

approximately 10 nm.

DOI: 10.1002/macp.200900099

Page 4: Directing the Self-Assembly of Mesostructured Hybrid Materials: …bib-pubdb1.desy.de/record/92824/files/Pietsch_et_al-2009... · 2016-03-16 · Directing the Self-Assembly of Mesostructured

Directing the Self-Assembly of Mesostructured Hybrid Materials . . .

Results and Discussion

The morphology of self-assembled block copolymer thin

films depends on a large number of factors such as the

chemical nature of the copolymer, its volume fraction, the

character of the solvent and the substrate material among

many others. Controlling these factors provides a powerful

tool to direct the self-assembly process with the ultimate

goal of targeting specific surface patterns. In the present

article we study the effects of the type of solvent and the

polymer concentration structure formation in spin-coated

thin films. These are two important factors that can be used

to direct the morphology of self-assembled functional

hybrid materials in thin film. The influence of substrate,

molecularweight of the copolymer, and its volume fraction

cannot be neglected. Therefore, these factors were kept

constant throughout the experiments presented here; we

used only one diblock copolymer (PS327-b-P4VP28) on Si/

SiO2 substrates. A detailed study of the influence of the

molecular weight of the block copolymer, its volume

fraction, and the substrate material will appear in

future publications.

Preparation of Organic-Inorganic HybridNanostructures

Functional hybrid materials are prepared as shown in

Figure 1. Due to the interesting properties of Au-nanopar-

ticles, tetrachloro-auric acid (HAuCl4) was used as a gold

precursor.[19] However, the same concept can be applied to

generate a variety of metallic and semi-conducting

nanoparticles.[20] Starting from the neat PS-b-P4VP diblock

copolymer in step I, a new hybrid material is designed

successively. Initially, the block copolymer is dissolved in a

selective or non-selective solvent. In step II an inorganic

precursor (HAuCl4) is coordinated to the block copolymer

in stoichiometric amounts with respect to the number of

pyridine units. The inorganic precursor protonates

the P4VP units and due to strong acid-base interaction

Figure 1. Schematic representation of the preparation process offunctional hybrid materials based on a PS-b-P4VP diblock copo-lymers.

Macromol. Chem. Phys. 2009, 210, 864–878

� 2009 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

the chloro-aurate counterion forms a complex with the

pyridine groups.[21] Metallic Au-nanoparticles are formed

uponreductionof theprecursor (Au3þ) tometallicgold (Au0)

due to the addition of reducing agent (LiBH4) to the

copolymer solution (Step III). The nanoparticles are

generated in situ within the block copolymer matrix,

which act as nano-reactors and allow controlling the

particle size and size distribution. Additionally, the

nanoparticles are located solely in the P4VP block. Thus,

self-assembly of the PS-b-P4VP block copolymer provides

means to direct the spatial arrangement of the nanopar-

ticles. Spin-coating this solution on solid substrates results

in the formation of compact films of spherical aggregates of

the hybridmaterial. We found that themorphology of self-

assembled thin films can be controlled much easier, if the

precursor loaded material PS-b-P[4VP(HAuCl4)] (Step II) is

used to deposit the films prior to the reduction step.

Subsequently, different methods such as wet-chemical

processes,[22] UV irradiation,[23] or O2-plasma etching[24]

can be applied to generate Au-nanoparticles.

The size and size distribution of metallic nanoparticles

was determined by transmission electron microscopy

(TEM) measurements. Figure 2a shows a typical TEM

micrograph; the Au-nanoparticles appear as dark spots due

to their higher electron density. The inset is a HR-TEM

micrograph of an individual Au-nanoparticle; the presence

of fringes suggests that these particles are single crystals.

Further, the existence of crystalline gold nanoparticles was

confirmed by XRD. A typical diffraction pattern is shown in

Figure 2b; the peak positions corresponding to the fcc

structure of the gold lattice (space group Fm3m) are

observed.

Oxygen plasma is used to selectively etch the polymer

matrix, leaving behindmetallic Au-nanoparticles. Since the

particles are generated within the P4VP block of the

copolymer, their spatial distribution represents the self-

assembled morphology of the copolymeric matrix. Quanti-

tative analysis of the chemical composition is performed

using X-ray photoelectron spectroscopy (XPS). Figure 2c

shows the XPS spectra of the copolymer thin films

containing Au-nanoparticles after O2-plasma treatment.

Compared to the non-etched samples, the carbon signal

decreases significantly, indicating complete removal of the

organic matrix. In contrast to that a strong gold (Au4f)

signal is observed before and after the etching process,

signifying the existence of Au nanoparticles.

Effect of Solvent on the Pattern Formation

The typeof solventused to spin-coat thinfilmsof thehybrid

material PS-b-P[4VP(HAuCl4)] onto a solid substrate has

been found to influence the polymer structure formation

significantly. The reason is that different solubilities of the

www.mcp-journal.de 867

Page 5: Directing the Self-Assembly of Mesostructured Hybrid Materials: …bib-pubdb1.desy.de/record/92824/files/Pietsch_et_al-2009... · 2016-03-16 · Directing the Self-Assembly of Mesostructured

T. Pietsch et al.

Figure 2. (a) TEM micrograph of PS-b-P[4VP(Au)], the inset showsa HR-TEM micrograph of an individual Au-nanoparticle togetherwith the corresponding FFT image; the scale bar is 3 nm, (b) XRDspectrum of the hybrid material confirming the existence ofcrystalline Au-nanoparticles, (c) XPS spectrum of a thin film afterO2-plasma treatment.

868

twoblocks of the block copolymer can lead to the formation

of aggregates such as spherical micelles. The quality of a

solvent for a given polymer depends on the solubility

parameters of both blocks in that solvent. The polymer-

Macromol. Chem. Phys. 2009, 210, 864–878

� 2009 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

solvent interaction can be expressed in terms of Hildeb-

randt solubility parameters.[25]

x ¼ xb þVS

RTdT;S � dT;P� �2

(1)

Where xb is the entropic contribution, VS is the molar

volume of the solvent, R is the gas constant, and T is the

absolute temperature. The total or Hildebrandt solubility

parameter (dT) of the solvent (S) and the polymer (P) are

given by

dT ¼ d2d þ d2p þ d2h

� �1=2(2)

where the factors dd, dp, and dh denote the dispersion-, the

polar-, and the hydrogen bonding component or Hansen

solubility parameters.

Based on its solubility parameters, at a given tempera-

ture, a solvent canbeeither selectiveornon-selective for the

PS-b-P4VP diblock copolymer (Table 1). In a common- or

non-selective solvent [(dT,S� dT, PS)2� (dT,S� dT, P4VP)

2] both

blocks of the diblock copolymer are soluble. Thus the

copolymer chains remain practically free. In contrast to

that, a selective solvent dissolves only one of the blocks

(x<0.5), while the other one is not soluble (x> 0.5)[27] e.g.,

[(dT,S� dT, PS)2� (dT,S� dT, P4VP)

2]. As a result micelles

are formed above the critical micelle concentration (cmc).

The copolymer self-assembles in such a way that the

interface between the less-soluble block and the solvent

is minimized. Different micelle morphologies can be

observed, depending on the stretching of core-forming

chains and the interfacial tension between the core and the

shell. Besides simple spherical micelles for example,

cylindrical aggregates and vesicles were found.[28]

Whenapolymerfilmis spin-coated fromasolutionbased

on different types of solvents, regular surface patternsmay

evolve. Avarietyofmorphologies canbe obtainedas shown

in Figure 3. These results clearly demonstrate that using

different types of solvents dramatically affects the thinfilm

morphology. The films are obtained by spin-coating the

same hybrid system of PS-b-P[4VP(HAuCl4)] onto a native

oxidized silicon wafer using both selective- and non-

selective solvents.

Dissolving PS-b-P[4VP(HAuCl4)] in selective solvents for

the PS block, such as toluene (Figure 3a) and chlorobenzene

(Figure 3b), spherical micelles are formed prior to the

deposition process. Subsequently, solvent evaporation

drives the self-assembly into hexagonally ordered patterns

(Figure 3a),which results from interactions among the shell

(PS block) of the micelles and the substrate. As demon-

strated in the TEM micrograph (inset in Figure 3a), the

nanoparticles are located in the core of themicelles. Hence,

DOI: 10.1002/macp.200900099

Page 6: Directing the Self-Assembly of Mesostructured Hybrid Materials: …bib-pubdb1.desy.de/record/92824/files/Pietsch_et_al-2009... · 2016-03-16 · Directing the Self-Assembly of Mesostructured

Directing the Self-Assembly of Mesostructured Hybrid Materials . . .

Table 1. Properties of solvents and polymers.[7,26]

Solvent dT dd dp dh Solvent

characterMPa1/2 MPa1/2 MPa1/2 MPa1/2

Toluene 18.2 18.0 1.4 2.0 Selective (PS)

Chloroform 19.0 17.8 3.1 5.7 Non-selective

THF 19.4 16.8 5.7 8.0 Non-selective

Chlorobenzene 19.6 19.0 6.3 3.3 Selective (PS)

DMF 24.8 17.4 13.7 11.3 Non-selective

Ethanol 26.5 15.8 8.8 19.4 Selective (P4VP)

Polystyrene 19.0 18.1 1.0 4.1 –

Poly(vinyl pyridine) 21.8 19.0 8.8 5.9 –

Figure 3. AFM images demonstrating the influence of the solventon the morphology of self-assembled block copolymer hybrid thinfilms PS-b-P[4VP(HAuCl4)]. The insets show the correspondingFFT image, in (a) and (c) the insets also display TEM micrographsof the structures. The films were spin-coated from solutions in(a) toluene, (b) chlorobenzene, (c) chloroform, (d) THF, and(e) DMF. In (f) the corresponding values of the most prominentin-plane length are shown. The corresponding film thicknessesare (a) 5.1 nm, (b) 5.4 nm, (c) 8.7 nm, (d) 9.3 nm, and (e) 4.6 nm.

Macromol. Chem. Phys. 2009, 210, 864–878

� 2009 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

their spatial distribution represents the hexagonal order of

the micellar films.

In contrast, in non-selective solvents such as chloroform

(Figure 3c), tetrahydrofuran (THF) (Figure 3d), and N,N-

dimethylformamide (DMF) (Figure 3e) the copolymer does

not form aggregates at the initial solution concentration.

Therefore, the film morphology is purely a result of the

solvent evaporation. Moreover, in non-selective solvents

both blocks of the diblock copolymer contribute to the

structure formation. The number of interactions increases

because the copolymer chains are practically free in

solution, permitting interactions among both blocks of

the copolymer, the inorganic component, the solvent, and

the substrate. Hence a larger diversity of morphologies is

accessible if the hybrid material is spin-coated from non-

selective solvents, where no aggregates are formed. In this

case, the inorganic nanoparticles are randomly distributed

within the P4VP phase (inset Figure 3c). Competing effects

such as the microphase separation of the PS-b-P4VP

copolymer, preferential wetting on the substrate, and thin

film instabilities due to solvent evaporation can induce

order in the spatial arrangement of nanoparticles. The film

spin-coated from DMF (Figure 3e) shows a compact worm-

like structure. The solvent (DMF) is a good solvent for both

blocks of the copolymer and additionally wets the Si/SiO2

substrate well. Hence, during spin-coating the solvent

(DMF) mediates the wetting of both blocks of the

copolymer. As the solvent evaporates and the polymer

concentration increases, stable localized patterns are

formed. This is a consequence of the different solvent

evaporation rates fromthePSand theP4VP(HAuCl4)phases

in the film.[29] Despite its non-selective character, DMF is a

better solvent for the P4VP block than the PS block (see

Table1). Therefore, theevaporation rate fromthePSphase is

higher, leadingtosurfaceundulations in thefilmandearlier

solidification of the PS block. As the solvent evaporates

completely the wetting behavior is dominated by the

www.mcp-journal.de 869

Page 7: Directing the Self-Assembly of Mesostructured Hybrid Materials: …bib-pubdb1.desy.de/record/92824/files/Pietsch_et_al-2009... · 2016-03-16 · Directing the Self-Assembly of Mesostructured

T. Pietsch et al.

870

interactions of both blocks of the copolymer with the

substrate. While the non-polar PS block dewetts, the polar

P[4VP(HAuCl4)] block preferentially wets the Si/SiO2

substrate. However, due to the low mobility of polymer

chains in the last stage of solvent evaporation, the

formation of dewetting polymer droplets is suppressed

and the film is trapped in a non-equilibrium state. Hence,

dewetting polymer droplets are rarely observed after spin-

coating of block copolymer solutions.

A characteristic length scale can be observed in the

patterns shown in Figure 3; this is a result of the surface

undulations caused by the different solvent evaporation

rates frombothblocks of the copolymer and itsmicrophase.

Eachmorphology can be described by a characteristic value

e.g., the most prominent in-plane length (L), which is

accessible by radial averaging over the corresponding FFT

image (insets Figure 3a–e) and sub-m GISAXS measure-

ments (example shown in Figure 7). As shown in Figure 3f,

the values ofL vary between 30 nm for chlorobenzene and

400 nm for chloroform. These characteristic length scales

are dependent on the evaporation rate of the solvent, the

different solubility of both blocks of the copolymer, and the

wettability of the polymer solution on the substrate.

Therefore, the type of solvent also provides a means to

control the lateral dimensions of self-assembled thin films

and allows tuning the morphology.

In order toutilize these tools, thedriving forces leading to

the formationofnanostructuredsurfacepatternshavetobe

understood. Describing the evolution of nanostructures in

thin liquid films under solvent evaporation is complicated

due to the complex hydrodynamic behavior, the inherent

unstable nature of the process, as well as heat- and mass

transfer within the film.[10] Besides the EISA approach,

various mechanisms have been proposed to describe the

generation of regular patterns in evaporating films; these

include dewetting and surface tension driven effects such

as the Marangoni instability.[30] These phenomena are,

among other factors, dependent on the surface tension and

the film thickness, which are a function of solvent and

polymer concentration. For instance, the self-assembly of

diblock copolymer in anevaporating thinfilm is drivenbya

concentration gradient, which is the result of continuous

solvent evaporation from the film surface.[10] This evapora-

tion process is characterized by the formation of a

copolymer-rich surface layer, while the copolymer con-

centration near the substrate surface represents the initial

solution concentration. During the drying process, the film

composition is time-dependent and varies over the

thickness of the film. Hence concentration gradients are

formedwithin the film; the thickness of the fluid layer, and

the evaporation rate of the solvents are themost important

factors that determine the steepness of the concentration

gradient and therefore significantly effects the pattern

formation. A reasonable approximation is that evaporation

Macromol. Chem. Phys. 2009, 210, 864–878

� 2009 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

is a constant throughout spinning, as long as the rotation

rate isheld constant,which results in theequationobtained

by Meyerhofer:[31]

@h

@t¼ �2Kh3 � ewithK ¼ rv2

3hand e ¼ Ce

ffiffiffiffiv

p(3)

Where h is the film thickness, t the time, and e the

evaporation rate of the solvent, which is described by

the scaling constant Ce and accounts for the vapor pressure

and diffusion through the boundary layer over the

substrate. The factor K constitutes the fluid density (r),

viscosity (h), and the angular velocity (v) during spin-

coating. This approach is strictly valid only when K is

constant. However, for spin coating of sol-gel or other

complex solutions this may not hold true during all stages

of spinning. Both viscosity and density are expected to

increase as evaporation progresses. The exact estimation

of the polymer concentration due to the absence of a more

elaborate theoretical approach is difficult.

Integration over the Meyerhofer equation (3) is a strong

approximation to estimate the polymer concentration and

film thickness during the spin-coating process as a function

of time,because it assumesK tobeconstantwhich is strictly

valid only for single solvent systems. Thus the result of this

integration shown in Figure 4 has to be understood as a

pictorial way to illustrate the time-dependent evolution of

the concentration. Due to solvent evaporation from the top

of the film, its thickness is decreasing, while both the

polymer concentration and viscosity of the liquid film

are increasing over several orders of magnitude. When the

copolymer concentration reaches c> cmc, aggregates begin

to form. Depending on the actual solvent composition

regular assemblies of copolymer chains are formed. This

mechanism is similar to the organization of surfactant

molecules in the EISA approach.[9,10] Due to rapid solvent

evaporation, the high molecular weight, and high concen-

tration of copolymer in the liquid film, the system quickly

reaches the overlap concentration (c�) and may be

characterized as a highly viscous solution (c> c�) with a

lowmobilityofpolymerchains.Therefore, the timescale for

rearrangements of copolymer chains increases dramati-

cally and the micelle formation may be delayed or even

totally prevented in the late stages of spinning. As the

remaining solvent evaporates the film morphology is

trapped in a non-equilibrium state.

In order to target specific morphologies the type of

solvent has to be chosen not only from the point of its

solvation power for a specific copolymer but also with

respect to its evaporation rate, viscosity, and surface

tension. For example, very volatile solvents, such as THF

and chloroform, create a strong concentration gradient

upon evaporation. Hence, the nano-patterns obtained are

far from the thermodynamic equilibrium. In contrast, less

DOI: 10.1002/macp.200900099

Page 8: Directing the Self-Assembly of Mesostructured Hybrid Materials: …bib-pubdb1.desy.de/record/92824/files/Pietsch_et_al-2009... · 2016-03-16 · Directing the Self-Assembly of Mesostructured

Directing the Self-Assembly of Mesostructured Hybrid Materials . . .

Figure 4. Pictorial representation of the development of the filmthickness and the polymer concentration during the spin-coatingprocess of PS-b-P[4VP(Au)] from a DMF solution. As the solventevaporates, the polymer concentration increases; when the cmcis reached, aggregates are formed. As the concentration reachesthe overlap concentration (c�), the aggregates (micelles) interactleading to the formation of a nanoscopically ordered highlyviscous solution. When the solvent evaporated completely, thisstructure is trapped in a glassy copolymer film. The final stagemay be extended over a long period of time, especially for low-volatile solvents such as DMF. The inset shows a zoom into theconcentration regime close to the point of complete solventevaporation.

volatile solvents, which evaporate slower, allow relaxation

of the copolymer chains to reach structures closer to the

equilibrium state. Depending on the solvent evaporation

rate, thefilmmorphology is trapped inakinetic equilibrium

state without reaching thermodynamic equilibrium con-

ditions. The latter can be achieved by post-deposition

treatments such as thermal- or solvent annealing over long

periods.[32]

Regarding the special case of organic-inorganic hybrid

materials, we have found that the character of the solvent

not only affects the morphology in thin film but also the

particle formation. For example, the particle size and size

distribution significantly depend on the chosen solvent.

The TEM micrographs presented in Figure 5 demonstrate

that the type of solvent significantly affects the size and

size distribution of Au-nanoparticles obtained from

PS327-b-P[4VP(HAuCl4)28]. In case of selective solvents for

the PS block, such as toluene (Figure 5a) and chlorobenzene

(Figure 5b), sphericalmicelles are formed,whose core (P4VP

block) is loadedwithgoldnanoparticles.Dueto theconfined

environment of themicelle core a very narrow particle size

distribution can be achieved having amean particle size of

3.5 nm. In the opposite case of using anon-selective solvent

Macromol. Chem. Phys. 2009, 210, 864–878

� 2009 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

such as chloroform (Figure 5c) and THF (Figure 5d) a much

wider particle size distribution is obtained with a

maximumat around 4.5 nm. Significantly smaller particles

(2 nm) are formed in DMF (Figure 5e); however, the particle

size distribution is wider than in the case of selective

solvent, where micelles are formed. The reason for smaller

particles inDMF lies in the stretching of the P[4VP(HAuCl4)]

block due to the swelling with solvent. Additionally the

solvent (DMF) may also contribute to the stabilization of

gold ions through the interaction with the lone electron

pairs. A comparison of the particle sizes and size distribu-

tions ispresented inFigure5f. Ifno reductionagent isadded,

the gold precursor (Au3þ) is reduced due to the exposure to

electron beam during the TEMmeasurements. In this case,

the particle growth is limited due to the lowmobility of Au

clusters in the glassy polymer matrix. In contrast, if the

reduction agent (LiBH4) is added to the solution prior to film

deposition, the gold precursor (Au3þ) is reduced simulta-

neously to Au0 followed by particle growth; thus the

average metal particle diameter increases with increase in

the time of the chemical treatment. As a result large

particles and relatively broad size distributions are

obtained in common solvents such as THF, DMF, and

chloroformwhile the confined environment of themicelles

in selective solvents (toluene, chlorobenzene) limits both

the particle size and size distribution.

Based on these observations it is assumed that the

particle size is correlated to the aggregation number of

copolymeric aggregates and the radius of gyration (Rg) of

the particle-host P4VP block. Therefore, factors such as

relative block chain length and solvent quality are of

significant importance to control the particle size. Table 2

gives a comparison between the hydrodynamic radius of

the PS327-b-P4VP28 copolymer measured by DLS and the

particle size (DP) obtained upon metallization and reduc-

tion. However, a direct relationship had not been derived

due to the contribution of the PS block to Rh as well as

micellization in case of selective solvents.

It can be difficult to find a solvent for thehybridmaterial,

which also fulfils the requirements to generate the desired

morphology in self-assembled thin films. An alternative

way to tune the solution properties is based onmixing two

or more different solvents in order to solubilize the

copolymer, control thegenerationof aggregates in solution,

and finally direct the pattern formation upon solvent

evaporation. The aggregation behavior in such solvent

mixtures can also be described in terms of solubility

parameters, which are estimated by:

dd ¼Xi

fidd;i; dp ¼Xi

fidp;i and dh ¼Xi

fidh;i (4)

Where dd,i, dp,i, and dh,i are the Hansen solubility

parameters of the ith solvent in the mixture and fi is its

www.mcp-journal.de 871

Page 9: Directing the Self-Assembly of Mesostructured Hybrid Materials: …bib-pubdb1.desy.de/record/92824/files/Pietsch_et_al-2009... · 2016-03-16 · Directing the Self-Assembly of Mesostructured

T. Pietsch et al.

Figure 5. Particle size as a function of the solvent: (a) toluene, (b) chlorobenzene, (c) chloroform, (d)THF, (e) DMF. The inset in each image shows the corresponding particle size distribution asobtained from TEM measurements. In (f) the average particle size for each system is shown; theerror bars represent the standard deviation of the respective particle size distribution.

Table 2. Hydrodynamic radius and particle size of PS327-b-P[4VP(Au)]28.

PS-b-P[4VP(Au)] in Rh DP

nm nm

Toluenea) 18.7 3.5

Chlorobenzenea) 6.5 3.6

THF 4.2 4.3

Chloroform 5.1 4.5

DMF 4.2 2.3

a)Due to c> cmc reverse micelles are formed.

872Macromol. Chem. Phys. 2009, 210, 864–878

� 2009 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

volume fraction. Based on that,

a non-selective solvent can be

designed by adjusting the

volume ratio of different sol-

vents in a blend to yield

solubility parameters between

those of the blocks of the

diblock copolymer. Considering

the case of hybrid materials, the

effect of the inorganic precursor

or the nanoparticles on the

solubility has to be taken into

account. For example due to the

polar character of the gold

precursor, micelles may form

in solutions of the PS-b-

P[4VP(HAuCl4)] hybrid material

even in a non-selective solvent

for the neat copolymer.[33]

These effects can be compen-

sated by re-adjusting the

volume ratio of solvents in the

mixture. Moreover, the compo-

sition of the solvent blend

provides a powerful tool that

can be used to direct the struc-

ture formation in thin films.

An example is presented in

Figure 6; well-defined nanos-

tructures are obtained by spin-

coating the copolymer hybrid

material PS-b-P[4VP(HAuCl4)]

from mixtures of DMF and

toluene. The sub-m GISAXS

experiments confirm the quite

well-developed lateral order in

these films. As an example,

Figure 7 shows a typical sub-m

GISAXS pattern together with a

horizontal slice of the 2D inten-

sity called out-of-plane scan. The existence of a strong first

order Bragg peak is a fingerprint of the lateral order.

The morphology can be tuned simply by changing the

volume fraction of the solvent mixture. Spherical micelles

are obtained in pure toluene (6a), which is a selective

solvent for the PS block of the diblock copolymer. Above

DMF contents of 16% DLS showed no micellization

because the solvent mixture dissolves both blocks of the

copolymer. Thus anon-selective solvent forhybridmaterial

is designed by blending a selective solvent for the PS block

(toluene) with a non-selective solvent (DMF). Spin-coating

of PS327-b-P[4VP(HAuCl4)]28 from DMF-toluene mixtures

ontonativeoxidized siliconwafers leads to the formationof

nano-patterns, morphologies of which differ from the

DOI: 10.1002/macp.200900099

Page 10: Directing the Self-Assembly of Mesostructured Hybrid Materials: …bib-pubdb1.desy.de/record/92824/files/Pietsch_et_al-2009... · 2016-03-16 · Directing the Self-Assembly of Mesostructured

Directing the Self-Assembly of Mesostructured Hybrid Materials . . .

Figure 6. AFM images of thin films of PS-b-P[4VP(HAuCl4)], spin-coated from DMF-toluenemixtures with DMF contents of (a) 0%, (b) 50%, (c) 100%. The inset in each figure represents thecorresponding FFT image. In (d) a plot of the most prominent in-plane length as a function ofthe DMF content is shown.

Figure 7. (a) Example of typical 2D sub-m GISAXS intensity distri-bution from sample PS327-b-P[4VP(HAuCl4)]28 spin-coated frommixtures of DMF and toluene having a ratio of 1:1 showing themain features: specular peak (shadowed with semi-transparentbeamstop) and Yoneda peak. (b) Selected out-of plane cut fromthe 2D intensity with peak positions marked with the arrows. Thesolid line is a fit to the data and the dashed line shows theresolution limit.

Macromol. Chem. Phys. 2009, 210, 864–878

� 2009 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

structureobtained fromnon-mixed

puresolvents.Anexample is shown

in Figure 6b; the worm-like struc-

ture was obtained by spin-coating

PS327-b-P[4VP(HAuCl4)]28 from

mixtures of DMF and toluene hav-

inga ratio of 1:1 (Figure 7 shows the

corresponding sub-m GISAXS data).

A more detailed series of morphol-

ogiesat intermediateDMFcontents

is shown in the supporting infor-

mation (SI-1). At high DMF/toluene

ratios, the patterns are comparable

to the structure obtained from

DMF. Therefore the characteristic

length scale of the nano-patterns

can be adjusted in a wide range by

changing the solvent composition

of the polymer solution (Figure 6d).

Describing the pattern forma-

tion in thin films of hybrid materi-

als spin-coated from solvent mix-

tures is intricate compared to using

a single solvent. Again, a concen-

tration gradient over the film thick-

ness drives the self-assembly of the

copolymer hybrid material but the

process becomes more complex

whensolventmixturesareinvolved.

Figure 8a demonstrates the evolu-

tion of the film composition during

solvent evaporation of block copo-

lymer dissolved in a binary solvent

blendofDMFandtoluene.Dueto its

higher vapor pressure toluene eva-

porates faster than DMF. Hence the

composition of the liquid film changes during the deposi-

tion process depending on the evaporation rate of both the

solvents. Asmost of the toluene evaporated, the remaining

solvent (DMF) dominates the structure formation provided

that the mobility of polymer chains is sufficiently high.

Hence adjusting the composition of theDMF-toluene blend

allows tuning the surface patterns. In Figure 8b the

evolution of the film composition is shown in a schematic

phase diagram for spin-coating PS327-b-P[4VP(HAuCl4)]28fromDMF-toluenemixtures. The startingpointAmarks the

initial composition of the block copolymer solution.

Depending on these initial conditions (A) the composition

evolves along different trajectories in the phase diagram

(Figure 8b). As the concentration of copolymer increases,

the cmc (B) is reached, leading to the formation of micellar

aggregates. However, with increase in copolymer concen-

tration, also the viscosity of the liquid film increases, thus

reducing the mobility of polymer chains and preventing

www.mcp-journal.de 873

Page 11: Directing the Self-Assembly of Mesostructured Hybrid Materials: …bib-pubdb1.desy.de/record/92824/files/Pietsch_et_al-2009... · 2016-03-16 · Directing the Self-Assembly of Mesostructured

T. Pietsch et al.

Figure 8. (a) Development of the composition of a 130 nm thickfilm of PS-b-P[4VP(Au)] during spin-coating from a DMF-toluenemixture. (b) Ternary phase diagram showing the evolution of thefilm composition during spin-coating of PS-b-P[4VP(Au)] fromDMF-toluene mixtures with different mixing ratios. Importantstages of the process are marked: (A) initial composition,(B) copolymer concentration reaches the cmc, and (C) the solventevaporation is completed, some solvent remains in the film.Different phase boundaries are shown schematically asrepresented by dashed lines.

874

reorganization of copolymer chains. At this stage interac-

tions among the residual solvent, both blocks of the hybrid

material, and the substrate surface gains significant

importance.[34] Instabilities in the liquid film due to

thermally exited perturbations and non-uniform solvent

evaporation may lead to local breakup of the film, a

consequence of which regular surface patterns are formed

having a wavelength which is dependent on the film

thickness and the surface tension of the film. In the case of

PS-b-P[4VP(HAuCl4)] spin-coated from DMF-toluene mix-

tures (Figure 6b) onto native oxidized silicon, the PS block

Macromol. Chem. Phys. 2009, 210, 864–878

� 2009 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

dewets thehydrophilic Si/SiO2 surface,while both the P4VP

block and the residual solvent (DMF) wet the substrate.

Sincebothpolymerblocksarewell-soluble inDMF, swelling

of the PS block with solvent mediates the wetting of the PS

block with the hydrophilic substrate. As solvent evapora-

tion is completed (C), the filmmorphology is trapped in this

non-equilibrium state; furthermodification requires a high

mobility of polymer chains, e.g., thermal annealing above

Tg or swellingwith solvent vapor.[35] It is important to note,

that the hybrid material PS327-b-P[4VP(HAuCl4)]28 in a

binary solvent blend displays a rich phase behavior. The

simple model given in Figure 8b does not yield a sufficient

description of the phase diagram. Yet even this simple

model signifies that depending on the initial composition

(A), the material undergoes various phase transitions (B)

until the final composition (C) is reached, where some

solvent remains entrapped in the film. The morphology of

spin-coated thinfilms reflects the entire depositionprocess,

summarizing the phase transitions occurring during

solvent evaporation. Therefore, the initial solution compo-

sition is an important factor that can be adjusted to control

morphology of spin-coated thin films.

Effect of Polymer Concentration

The morphology of the hybrid material in thin film also

depends on the initial block copolymer concentration.

There are two main factors relating the polymer concen-

tration with the pattern formation in spin-coated thin

films. These factors are thefilm thickness obtained after the

deposition process and the viscosity of the copolymer

solution during solvent evaporation. Both parameters

changeover several orders ofmagnitude. Thefilmthickness

increases with the initial concentration simply because

morematerial is appliedonto the substrate.Additionally, as

indicated by equation (3), the solution viscosity signifi-

cantly affects the thinning of a film during spin-coating.

Therefore, besides controlling the solvent quality and

evaporation rate, also the initial polymer concentration

(c0) can also be adjusted in order to control themorphology

of the hybrid material in thin films.

The effect of the polymer concentration on themorphol-

ogy in thin films of PS-b-P[4VP(HAuCl4)] is demonstrated in

Figure 9a–d. The filmswere deposited on Si/SiO2 substrates

from toluene solutions of varying initial solution concen-

trations between 0.05 and 10 mg �mL�1. At very low

polymer concentrations such as 0.05mg �mL�1 (Figure 9a) a

disordered, micellar morphology is formed. These micelles

are randomly distributed within the film. However, drying

mediated phenomena and dewetting of both the corona

forming chains (PS block) and the solvent (toluene) can lead

to the formation of regular surface patterns such as rings

shown in Figure 9a. As the initial polymer concentration

DOI: 10.1002/macp.200900099

Page 12: Directing the Self-Assembly of Mesostructured Hybrid Materials: …bib-pubdb1.desy.de/record/92824/files/Pietsch_et_al-2009... · 2016-03-16 · Directing the Self-Assembly of Mesostructured

Directing the Self-Assembly of Mesostructured Hybrid Materials . . .

Figure 9. AFM height images of micellar thin films of PS-b-P[4VP(Au)] spin-coated from toluene atdifferent concentrations of (a) 0.05, (b) 0.5, (c) 1.0 and (d) 3 mg �mL�1. The corresponding sectionsacross the images are shown below. The inset in (b) is a 10� 10 mm image of the same sampledemonstrating the formation of regular surface patterns. The inset in (d) shows a magnifiedimage (0.5�0.5 mm2) indicating the formation of a densely packed micellar film.

increases the film coverage is increased and hexagonally

ordered regions of compact micellar layers are formed

(Figure 9b). Yet the film is not continuous; consequently

networks of densely packed micelles are formed (inset

Figure 9b). At concentrations of 1.0 mg �mL�1 a continuous

monolayer of hexagonally ordered micelles are obtained

(Figure 9c). Well-defined, hexagonally ordered domains are

formed over short length scales (�1 mm). However, defects

can be cured through fine tuning of the film thickness

followed by thermal- or solvent annealing, leading to

perfectly ordered thin films. As expected thick multilayer

films having a hexagonal close-packing of spherical

micelles are formed at high initial solution concentrations

(Figure 9d). The film surface appears smooth and almost

featureless; however as indicated by the inset in Figure 9d

Figure 10. Effect of polymer concentration on micellar films. (a) Schematic representation of filmmorphology as a function of polymer concentration. The plots show the characteristic distance(b) and the micellar bump height (c) of micellar films of PS-b-P[4VP(Au)].

the top-most layer consists of

denselypackedsphericalmicelles.

The development of micellar

films with increase in solution

concentrations is shown sche-

matically in Figure 10; three

different regions can be defined

with respect to the initial solu-

tion concentration. At low poly-

mer concentrations, well-sepa-

rated micelle bumps are formed

(region A); both the solvent

(toluene) and the PS shell of the

micelles dewet the hydrophilic

Si/SiO2 substrate. Micellar

monolayer can be fabricated,

when a certain concentration

threshold is reached (region B).

Entanglements among the shell

formingPS chainsof themicelles

lead to relatively smooth films.

Therefore, the micellar bump

Macromol. Chem. Phys. 2009, 210, 864–878

� 2009 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

height (hM) is reduced signifi-

cantly. Additionally, the average

spacing between the micellar

bumps decreases due to the

formation of domains with hex-

agonal-closest packing of sphe-

rical micelles. At much higher

concentration, compact multi-

layers are formed; both the

micellar spacing and the micelle

bump height remain constant

(region C).

The characteristic distance (L)

and micelle bump height as a

function of the initial solution

concentration are shown in

Figure 10b and c; the regions A,

B, and C are highlighted in both plots. The transition

between the A and C marks the threshold (�0.15 wt.%¼1.3mg �mL�1) for the formationof amonolayer (regionB) of

closely packed spherical micelles of PS-b-P[4VP(HAuCl4)].

Below this threshold discontinuous films are formed; there

is a strong correlation between the initial solution

concentration and the micellar bump height as well as

the average spacing between surface micelles. In the

opposite case, high concentrations result in the formation

of multilayer. Hence, adjusting the initial solution con-

centration of thehybridmaterial providesmeans to control

both the dimensions and the morphology in micellar thin

films.

The structure formation is different in the case of non-

selective solvents, where the hybridmaterial does not form

www.mcp-journal.de 875

Page 13: Directing the Self-Assembly of Mesostructured Hybrid Materials: …bib-pubdb1.desy.de/record/92824/files/Pietsch_et_al-2009... · 2016-03-16 · Directing the Self-Assembly of Mesostructured

T. Pietsch et al.

876

aggregates. As demonstrated for mixtures of DMF and

toluene, a non-selective solvent can be designed by

blending two different solvents. The concentration depen-

dence of themorphology of thin films deposited from these

solvent mixtures is shown in Figure 11. At low concentra-

tions (0.05mg �mL�1) a disorderedmorphology is obtained

(Figure 11a); dewetting droplets of PS-b-P[4VP(HAuCl4)] are

randomlydistributed over the substrate. In contrast to that,

nano-patterns with a characteristic spacing of L¼ 41 nm

and hexagonal short-range order (Figure 11b) can be

generated at higher polymer concentrations (0.1mg �mL�1).

. Since the surface energy of PS is lower than those of P4VP,

the PS block enriches at the polymer-air interface,while the

precursor loaded P4VP block P[4VP(HAuCl4)] preferentially

wets the hydrophilic Si/SiO2 substrate. Hence the surface

patterns consist of a coreof P[4VP(HAuCl4)] andadewetting

shell of PS.

As the solution concentration increases, ribbon-like

morphologies are formed having a spacing of 48 nm

(Figure11candd).Eventually, these ribbonsmergeto forma

continuous worm-like morphology with L¼ 50 nm

(Figure 11e). The corresponding FFT image (inset) shows a

sharp ring indicating a regular but isotropic structure. As

expected a continuous and mainly featureless thin film

(Figure 11f) is generated when a certain concentration

threshold (�1 mg �mL�1) is reached. Moreover, at higher

Figure 11. AFM height images demonstrating the development of the sDMF-toluene solution on native oxidized silicon. The initial solution c(h) 5.0 mg �mL�1, leading to a average heights of the surface featucorresponding FFT images are shown in the insets.

Macromol. Chem. Phys. 2009, 210, 864–878

� 2009 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

polymerconcentrations (Figure11gandh)aggregatesof the

hybrid material are formed during the deposition process;

the resultingmorphology represents network-structures of

spherical and rod-like micelles.

It is well known, that the film thickness plays an

important role in microphase separated block copolymer

films. For example, for a symmetrical diblock copolymer

having a lamellar morphology with domain orientations

parallel to the substrate, the total film thickness (h) has to

obey the relation h¼ (nþ 1/2)L0 to form a continuous film;

where n is a positive integer and L0 is the bulk lamellar

period of the copolymer.[36] If the film thickness does not

match this natural film thickness of the copolymer, islands

or holes are formed to compensate for the excess or

deficiency material. Spin-coated thin films are usually

trapped in non-equilibrium morphologies rather than an

ordered microphase separated structure; therefore this

quantization of the film thickness does not hold true.

However, the formation of non-equilibrium patterns in

spin-coated thin films also depends on the film thickness.

For instance the pattern formation at low film thicknesses

(<100nm) isdominatedbymolecular interactionsbetween

the fluid (copolymer and solvent) layer and the sub-

strate.[37] In thick films however, thermo-capillary effects

play an important role. For example the characteristic

wavelength (L) of liquid film instabilities such as spinodal

urface morphology of PS-b-P[4VP(HAUCl4)] films, spin-coated from aoncentration was (a) 0.05, (b) 0.1, (c) 0.3, (d) 0.5, (e) 0.7, (f) 1.0, (g) 3.0,res of 7.3, 8.6, 7.9, 7.6, 6.5, 6.2, 12.4 and 16.1 nm respectively. The

DOI: 10.1002/macp.200900099

Page 14: Directing the Self-Assembly of Mesostructured Hybrid Materials: …bib-pubdb1.desy.de/record/92824/files/Pietsch_et_al-2009... · 2016-03-16 · Directing the Self-Assembly of Mesostructured

Directing the Self-Assembly of Mesostructured Hybrid Materials . . .

dewetting[38] and thermo-capillary effects[39] is a function

of the film thickness, the viscosity, and the surface tension

of the film; which are strongly related to the polymer

concentration.

The high evaporation rate during the spin-coating

process relates to a steep concentration gradient; resulting

in highMarangoni numbers [69,70] which in return lead to

relative short wavelengths of the corresponding surface

patterns. Unfortunately, there is no easy way of assessing

the Ma of polymer blend solutions during spin-coating

directly. In theory, the Ma instability leads to regular

surface patterns, when the typical film thickness is much

smaller than the lateral length scale. In our experiments,

this requirement is not fulfilled for the patterns presented

inFigure11,wherecharacteristic lengthscalesofL� 50nm

are observed in much thinner films. At a polymer

concentration of 5 mg �mL�1 (Figure 11e) a compact film

is formed; theaveragefilmthickness isonly16nm.At lower

concentration even thinner films are obtained; in these

films the influence of the Ma effect on the structure

formation can be neglected. Therefore, the morphology of

ultra-thin spin-coated thin films of PS-b-P[4VP(HAuCl4)]

cannot be described by the Ma effect only; additionally

(spinodal-) dewetting of both polymer blocks (PS and P4VP)

as well as polymer-solvent and solvent-substrate interac-

tions have to be taken into account. However, theMa effect

will gain significant importance in thicker films deposited

and spin-coated at higher polymer concentrations.

Altogether, controlling themorphology of block copolymer

hybrid materials in thin film is an important issue. The

morphology of thin films of hybrid materials based on a

diblock copolymer and inorganic nanoparticles is largely

unpredictable. Yet, herein we demonstrated that both the

type of solvent and the initial polymer concentration can

serve as tools to control themorphology of spin-coated thin

films.

Conclusion

In this article we presented a way to control the formation

of non-equilibrium nano-patterns in spin-coated thin

films of functional hybrid materials. The hybrid material

based on the diblock copolymer (PS-b-P4VP) and metallic

nanoparticles was designed through metallization of the

pyridine units’ in situ using a wet chemical process.

Subsequently, solutions were spin-coated on a solid

substrate to obtain nanostructured thin films. It was found

that the typeof solvents and thepolymer concentration can

be used as parameters to direct the morphology of spin-

coated thinfilms. Inparticular, control over themorphology

of hybrid materials, particle size and size distribution is

achieved. We demonstrated that the type of solvent not

only affects the pattern formation but also the particle size

Macromol. Chem. Phys. 2009, 210, 864–878

� 2009 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

and size distribution. Moreover, the initial concentration

canalso serveasa tool todirect thepattern formationdue to

its effect on thermo-capillary forces. As a result not only the

morphology and the particle size can be controlled but also

the characteristic length scale and the height of the surface

patterns may be adjusted. Other structural determinant

factors suchas themolecularweight, thevolume fractionof

the copolymer, and the influence of annealing are currently

under investigation.

Acknowledgements: This work was funded by the EPSRC throughthe Nottingham IMRC. Peter Muller-Buschbaum acknowledgessupport by BMBF project 05 KS7WO1. We acknowledge theEuropean Synchrotron Radiation Facility for provision of synchro-tron radiation facilities and we thank M. Burghammer andC. Riekel for assistance in using beamline ID13. We also thankM. Al-Hussein, R. Dormann, R. Gehrke, V. Korstgens, and U. Vainiofor their help during the GISAXS experiments.

Received: March 9, 2009; Accepted: March 12, 2009; DOI: 10.1002/macp.200900099

Keywords: block copolymer; hybrid materials; nanoparticle;nanotechnology; self-assembly; thin film

[1] [1a] M. Geissler, Y. N. Xia, Adv. Mater. 2004, 16, 1249; [1b] H.Ahmed, Proc. IEEE 1991, 79, 1140; [1c] S. Matsui, Proc. Ieee1997, 85, 629.

[2] T. P. Russell, Science 2002, 297, 964.[3] [3a] I. W. Hamley, Nanotechnology 2003, 14, R39; [3b] T. P.

Lodge, Macromol. Chem. Phys. 2003, 204, 265.[4] G. Krausch, S. Ludwigs, A. Boker, A. Voronov, N. Rehse, R.

Magerle, Nat. Mater. 2003, 2, 744.[5] [5a] H. C. Kim, T. P. Russell, J. Polym. Sci., Polym. Phys. 2001, 39,

663; [5b] G. Brown, A. Chakrabarti, J. Chem. Phys. 1994, 101,3310.

[6] [6a] T. T-. Albrecht, J. DeRouchey, T. P. Russell, H. M. Jaeger,Macromolecules 2000, 33, 3250; [6b] A. Boker, A. Knoll, H. Elbs,V. Abetz, A. H. E. Muller, G. Krausch,Macromolecules 2002, 35,1319.

[7] J. Zhao, S. Tian, Q. Wang, X. Liu, S. Jiang, X. Ji, L. An, B. Jiang,Eur. Phys. J. E Soft Matter 2005, 16, 49.

[8] [8a] P. Muller-Buschbaum, E. Bauer, O. Wunnicke, M. Stamm,J. Phys. :Condens. Matter 2005, 17, S363; [8b] P. Muller-Busch-baum, J. S. Gutmann, M. Stamm, J. Macromol. Sci. Phys. 1999,B38, 577.

[9] C. J. Brinker, Y. F. Lu, A. Sellinger, H. Y. Fan, Adv. Mater. 1999,11, 579.

[10] D. Grosso, F. Cagnol, G. J. D. A. Soler-Illia, E. L. Crepaldi, H.Amenitsch, A. B. Bruneau, A. Bourgeois, C. Sanchez, Adv.Funct. Mater. 2004, 14, 309.

[11] T. Brezesinski, M. Groenewolt, A. Gibaud, N. Pinna, M. Anto-nietti, B. M. Smarsly, Adv. Mater. 2006, 18, 2260.

[12] [12a] Y. J. Cheng, J. S. Gutmann, J. Am. Chem. Soc. 2006, 128,4658; [12b] Y. J. Cheng, S. Y. Zhou, J. S. Gutmann, Macromol.

www.mcp-journal.de 877

Page 15: Directing the Self-Assembly of Mesostructured Hybrid Materials: …bib-pubdb1.desy.de/record/92824/files/Pietsch_et_al-2009... · 2016-03-16 · Directing the Self-Assembly of Mesostructured

T. Pietsch et al.

878

Rapid Commun. 2007, 28, 1392; [12c] Y. J. Cheng, P. Muller-Buschbaum, J. S. Gutmann, Small 2007, 3, 1379.

[13] [13a] T. T. Albrecht, R. Steiner, J. DeRouchey, C. M. Stafford, E.Huang, M. Bal, M. Tuominen, C. J. Hawker, T. P. Russell, Adv.Mater. 2000, 12, 787; [13b] J. Y. Cheng, A. M. Mayes, C. A. Ross,Nat. Mater. 2004, 3, 823; [13c] M. Li, C. K. Ober, Mater. Today2006, 9, 30.

[14] [14a] R. Ulrich, A. Du Chesne, M. Templin, U. Wiesner, Adv.Mater. 1999, 11, 141; [14b] S. C. Warren, F. J. Disalvo, U.Wiesner, Nat. Mater. 2007, 6, 156.

[15] [15a] R. Glass, M. Moller, J. P. Spatz, Nanotechnology 2003, 14,1153; [15b] J. P. Spatz, M. Moller, M. Noeske, R. J. Behm, M.Pietralla, Macromolecules 1997, 30, 3874.

[16] [16a] L. Meli, Y. Li, K. T. Lim, K. P. Johnston, P. F. Green,Macromolecules 2007, 40, 6713; [16b] C.-A. Dai, Y.-L. Wu, Y.-H. Lee, C.-J. Chang, W.-F. Su, J. Cryst. Growth 2006, 288, 128;[16c] Z. Liu, R. Levicky, Nanotechnology 2004, 15, 1483.

[17] A. Fahmi, T. Pietsch, N. Gindy, Macromol. Rapid Commun.2007, 28, 2300.

[18] [18a] C. Riekel, Rep. Prog. Phys. 2000, 63, 233; [18b] P.Muller-Buschbaum, S. V. Roth, M. Burghammer, A. Diethert,P. Panagiotou, C. Riekel, Europhys. Lett. 2003, 61, 639.

[19] [19a] M.-C. Daniel, D. Astruc, Chem. Rev. 2004, 104, 293; [19b]B. Kim, S. L. Tripp, A. Wei, J. Am. Chem. Soc. 2001, 123, 7955.

[20] [20a] F. S. Diana, S. H. Lee, P. M. Petroff, E. J. Kramer,Nano Lett.2003, 3, 891; [20b] S. H. Kim, M. J. Misner, L. Yang, O. Gang,B. M. Ocko, T. P. Russell, Macromolecules 2006, 39, 8473.

[21] J. P. Spatz, S. Sheiko, M. Moller, Macromolecules 1996, 29,3220.

[22] S. Forster, M. Antonietti, Adv. Mater. 1998, 10, 195.[23] D. H. Kim, Z. C. Sun, T. P. Russell, W. Knoll, J. S. Gutmann, Adv.

Funct. Mater. 2005, 15, 1160.[24] J. P. Spatz, S. Mossmer, C. Hartmann, M. Moller, T. Herzog, M.

Krieger, H.-G. Boyen, P. Ziemann, B. Kabius, Langmuir 2000,16, 407.

[25] [25a] L. Cui, Y. Xuan, X. Li, Y. Ding, B. Y. Li, Y. C. Han, Langmuir2005, 21, 11696; [25b] J. Brandrup, E. H. Immergut, E. A.Grulke, ‘‘Polymer Handbook’’, 4th edition, Wiley, New York

Macromol. Chem. Phys. 2009, 210, 864–878

� 2009 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

1999, p. 2288; [25c] A. J. Marzocca, Eur. Polym. J. 2007, 43,2682; [25d] T. Lindvig, M. L. Michelsen, G. M. Kontogeorgis,Fluid Phase Equilib. 2002, 203, 247.

[26] [26a] C.M. Hansen, Prog. Org. Coat. 2004, 51, 77; [26b] D. L. Ho,C. J. Glinka, J. Polym. Sci., Polym. Phys. 2004, 42, 4337.

[27] A. Guner, G. Kibarer, Eur. Polym. J. 2001, 37, 619.[28] [28a] D. E. Discher, A. Eisenberg, Science 2002, 297, 967; [28b]

C. J. Clarke, L. F. Zhang, J. Y. Zhu, K. Yu, R. B. Lennox, A.Eisenberg, Macromol. Symp. 1997, 118, 647; [28c] T. Pietsch,N. Gindy, A. Fahmi, Polymer 2008, 49, 914.

[29] S. Walheim, E. Schaffer, J. Mlynek, U. Steiner, Science 1999,283, 520.

[30] [30a] S. G. Yiantsios, B. G. Higgins, Phys. Fluids 2006, 18,082103; [30b] R. J. Deissler, A. Oron, Phys. Rev. Lett. 1992,68, 2948.

[31] D. Meyerhofer, J. Appl. Phys. 1978, 49, 3993.[32] [32a] T. H. Kim, J. Hwang, W. S. Hwang, J. Huh, H. Kim, S. H.

Kim, J. M. Hong, E. L. Thomas, C. Park, Adv. Mater. 2008, 20,522; [32b] J. Bang, S. H. Kim, E. Drockenmuller, M. J. Misner,T. P. Russell, C. J. Hawker, J. Am. Chem. Soc. 2006, 128, 7622;[32c] Y. Xuan, J. Peng, L. Cui, H. F. Wang, B. Y. Li, Y. C. Han,Macromolecules 2004, 37, 7301; [32d] P. Muller-Buschbaum,J. S. Gutmann, C. L. Haas, B. Mahltig, M. Stamm, W. Petry,Macromolecules 2001, 34, 7463.

[33] N. J. Jain, V. K. Aswal, P. S. Goyal, P. Bahadur, Colloids Surf. APhysicochem. Eng. Asp. 2000, 173, 85.

[34] K. Kargupta, R. Konnur, A. Sharma, Langmuir 2001, 17, 1294.[35] J. Zhao, S. Jiang, X. Ji, L. An, B. Jiang, Polymer 2005, 46,

6513.[36] B. H. Sohn, B. W. Seo, S. I. Yoo, J. Mater. Chem. 2002, 12, 1730.[37] B. Yu, P. C. Sun, T. C. Chen, Q. H. Jin, D. T. Ding, B. H. Li, A. C. Shi,

Phys. Rev. Lett. 2006, 96, 138306.[38] [38a] A. Sharma, Eur. Phys. J E 2003, 12, 397; [38b] P. Muller-

Buschbaum, J. S. Gutmann, C. L. Haas, O. Wunnicke, M.Stamm, W. Petry, Macromolecules 2002, 35, 2017.

[39] [39a] H. Kaya, B. Jerome, P. Colinet, Europhys. Lett. 2006, 74,861; [39b] G. Reiter, A. Sharma, A. Casoli, M.-O. David, R.Khanna, P. Auroy, Langmuir 1999, 15, 2551.

DOI: 10.1002/macp.200900099