Download as pdf or txt
Download as pdf or txt
You are on page 1of 13

View Online / Journal Homepage

Journal of Dynamic Article Links < C


Materials Chemistry C
Cite this: DOI: 10.1039/c2tc00026a
www.rsc.org/materialsC FEATURE ARTICLE
Development of in situ studies of spin coated polymer films
Daniel T. W. Toolan and Jonathan R. Howse*
Received 22nd August 2012, Accepted 22nd August 2012
DOI: 10.1039/c2tc00026a
Published on 07 August 2012 on http://pubs.rsc.org | doi:10.1039/C2TC00026A
Downloaded by UNIVERSITY OF SOUTH AUSTRALIA on 08 October 2012

High quality, uniform thin polymer films are routinely produced by the technique of spin coating.
Applications for such polymer films beyond photoresist layer fabrication include photovoltaics and
light-emitting diodes, where device performance is dependent upon an appropriate phase separated
morphology. Developing an understanding of the factors involved in the development of such
morphologies is therefore an important goal. The spin coating of polymer blends is a high speed, non-
equilibrium process and as such, it is difficult to monitor in situ, with most studies inferring structure
development from the resultant final morphology. Over the past 20 years various in situ experimental
techniques have been developed, providing insight into the details of the spin coating process itself and
opening a route to understanding and controlling morphological development. The majority of studies
have been based upon interferometry and light scattering, ‘‘based in reciprocal space’’ and have been
able to verify theoretical models of spin coating and the associated phase separation of polymer blends,
with direct real-space, in situ studies now a possibility.

Introduction Under ideal conditions, spin-coating a solution of a single


polymer from a single solvent results in the formation of a
Spin coating is a widely used technique for making uniform, smooth, unstructured film. However, spin-coating a blend of two
high-quality thin polymer films. Interest in developing and immiscible polymers from a common solvent results in a phase-
understanding the spin-coating process has been driven by the separated morphology. The morphology produced is the result
potential applications which arise from the rich variety of of a liquid–liquid phase transition, which occurs due to solvent
accessible morphologies. This attention has recently intensified evaporation and a subsequent liquid to solid transition, through
due to the attraction of utilising polymeric materials in electronic crystallisation or vitrification as illustrated in Fig. 1.
devices, such as in photovoltaics, where device efficiency is These processes and the resultant morphology have been
greatly dependent on the resulting nano-scale phase separation.1 shown to be greatly dependant upon the nature of the individual
polymers (molecular-weight,2 solubility,3 surface tension and
viscosity), the solution composition (solid content,4 solvent
Department of Chemical and Biological Engineering, The University of
Sheffield, Sheffield, S1 3JD, UK. E-mail: j.r.howse@sheffield.ac.uk
properties3 and blend ratio5) and the substrate (chemistry,3

Daniel Toolan (left) studied chemistry and completed his Master’s thesis
under the supervision of Prof. Tony Ryan at the University of Sheffield in
2011. He is currently in the second year of his PhD working on organic
photovoltaics and the development of in situ measurement techniques
utilising both microscopy and scattering and recently attended
HERCULES 2012, the Higher European Research Course for Users of
Large Experimental Systems.

Jonathan Howse (right) obtained his PhD in 2000 (Physical Chemistry,


Sheffield). Following a 2 year research position at the Berlin Neutron
Scattering Centre and the TU-Berlin he returned to Sheffield to conduct
research on a variety of ‘‘soft-nanotechnology’’ projects. He was
appointed lecturer in 2007 and has since led pioneering work on strobo-
Daniel T: W: Toolan and Jonathan R: Howse scopic interference microscopy for looking at spin-coating (discussed
here), polymer vesicle formation, and colloidal nanoswimmers.

This journal is ª The Royal Society of Chemistry 2012 J. Mater. Chem. C


View Online

concentration profile which develops in the film.10 An overview


of the spin coating process is shown in Fig. 2, for both pure
solvent and a ternary mixture of two incompatible polymers
dissolved in a common solvent.
There have been numerous attempts to deduce the mechanisms
controlling morphology development in spin coated blends. Due
to the fact that many studies have been restricted to analysis of
the structure of the final film and the fact that phase separation is
a complex non-equilibrium process it has been very difficult to
deduce details about the process from the end-point alone.11
Walheim suggested that an important mechanism for lateral
phase separation in thin films was the creation of a transiently
layered film by a process of wetting, followed by the break-up of
Published on 07 August 2012 on http://pubs.rsc.org | doi:10.1039/C2TC00026A
Downloaded by UNIVERSITY OF SOUTH AUSTRALIA on 08 October 2012

the layers because of an interfacial instability.12 Buxton and


Fig. 1 Ternary phase diagram for a 2 wt% polymer blend in a common Clarke conducted a numerical study that indicated that phase
solvent, showing the mixture going from the single phase region to the separation initially occurred at the air–film interface, where the
two phase region with example morphologies obtained from different solvent concentration is at its lowest. As the solvent diffuses
points of the two phase region. through the film, and evaporates from the surface, phase sepa-
ration becomes energetically favourable progressively
roughness, size and shape) as well as various parameters of the throughout the film. However, phase separation is now initiated
spin-coating process such as spin speed,6 acceleration and and directed by regions closer to the air surface which have
surrounding atmosphere.7 already undergone phase separation. This, in effect, produces an
There have been many attempts to model the spin coating ordering front which propagates down through the film and
process, the first of which was proposed by Emslie, Boner and leaves an ordered lateral morphology in its wake.13
Peck; commonly referred to as the EBP model.8 This consists of a In situ studies have been a key tool in developing further
viscous fluid on a rotating plane and is based on the assumptions understanding of phase-separation in spin coating, however, the
that the fluid shows Newtonian character, the rotating plane is process is still not fully understood. A summary of the key
infinite in extent, the rotating plane is horizontal and that the analytical developments over the past 20 years is given in Table 1.
liquid layer is radially symmetric so that shear resistance is only The most widely used techniques for in situ studies of spin
applicable in the horizontal planes. In their analysis EBP gave coating are based on interferometry. In this process a narrow
the thinning rate as: bandwidth light source (usually a laser (Dl < 1 nm)) is incident
on the substrate and the fluctuations in specular reflectivity are
dh recorded. Through analysis of the variations of specular reflec-
¼ 2Kh3 (1)
dt tivity over time it is possible to extract drying rate curves using
where h is the film thickness, t is time and K is a system constant Bragg’s law (constructive interference occurs when 2ndcos qi ¼
defined as: ml, where n ¼ refractive index of the film, d is the film thickness,
qi is the angle of incidence relative to the normal, l is the wave-
ru2 length and m is an integer value).14,15 This approach is greatly
K¼ (2)
3h simplified by many polymers having similar refractive indices to
where r is the fluid density, u is the rotation rate in radians per organic solvent (i.e. polystyrene (PS) ¼ 1.592, poly(methyl
second and h is the viscosity in poise. In the early stage of spin methacrylate) (PMMA) ¼ 1.491, toluene ¼ 1.496 at 20  C).
coating known as hydrodynamic thinning, the rate of thinning is
sufficiently described by eqn (1). However, later on in the spin- Specular reflectivity interferometry
ning contributions from solvent evaporation can no longer be
ignored. Meyerhofer was the first to estimate the effect of solvent The first in situ studies of spin coated polymer films were con-
evaporation on the thinning rate, by simply adding an evapo- ducted by Graves et al. between 1990 and 1991.16 The motivation
ration term to eqn (1):9 behind the study was to understand the effect of substrate surface
topography on spin coating, as previous manufacturing steps
dh
¼ 2Kh3  e (3) often left an uneven surface topography on the substrate. Their
dt strategy was to make systematic comparisons between film
where e is the evaporation rate [ml s1 cm2] and is dependant profiles over topography, taken during the process, against
upon the rate of rotation and the rate of diffusion through a model predictions of the wet film profile.
vapour boundary layer above the spinning substrate. Eqn (3) was Their experimental set-up is shown in Fig. 3. A laser was
not solved analytically, instead Meyerhofer assumed spin coating directed onto the spinning substrate and the resultant interfer-
was a two stage process; where the first stage was dominated by ence patterns where recorded using a conventional 35 mm
flow and the second stage was dominated by evaporation. This camera. The laser and 35 mm camera where synchronously
assumption allowed prediction of the final coating thickness in pulsed so as to image the same point once per revolution. The
terms of several key solution properties. Over the years this spin coating experiment produced a series of photographs, each
model has been improved through taking into account the showing a contour map of the thin film at a preselected point on

J. Mater. Chem. C This journal is ª The Royal Society of Chemistry 2012


View Online

Fig. 2 Overview of the spin coating process for (a) solvent and (b) a ternary mixture consisting of two incompatible polymers dissolved in a common
solvent.
Published on 07 August 2012 on http://pubs.rsc.org | doi:10.1039/C2TC00026A
Downloaded by UNIVERSITY OF SOUTH AUSTRALIA on 08 October 2012

Table 1 Summary of analytical developments of in situ studies of spin-coating

Year Set-up Source Detector 1D/2D System studied Real/reciprocal space Reference

1990 Specular reflectivity Laser Photographic film 2D Photo-resists Real space Graves, Manske
and Perrunge16
1993 Specular reflectivity Laser Photo-diode 1D Silica spin-on Reciprocal space Horowitz14
glasses by a
sol–gel method
1996 Specular reflectivity Laser Photo-diode 1D Pure solvents Reciprocal space D. P. Birnie15
1997 Specular reflectivity Laser Photo-diode 1D Solvent mixtures Reciprocal space D. P. Birnie18
2005 Specular reflectivity Laser Photo-diode and 2D Blend F8BT/PFB Reciprocal space P. C, Jukes, S. Y. Heriot,
and off-specular progressive scan J. S. Sharp and
scattering CCD camera R. A. L. Jones19
2010 Colour video calibrated Laser and Photo-diode and 2D n-Propyl Real space D. P. Birnie26
with laser interferometry white light colour video camera
2011 Direct stroboscopic LED EMCCD 2D PS/PI from Real space Ebbens and Howse28
imaging (z10 nm) (electron-multiplying o-xylene
CCD) QE ¼ 95%

the wafer at a certain time. The space between interference azimuthal symmetry17 was found to be in good agreement with
fringes showed the differences in height between two points, and the obtained experimental profiles of thin films. Predictions for
not the absolute thickness. dry profiles, assumed constant throughout the film, where found
The study indicated that an applicable model proposed by to be inaccurate for small feature sizes on substrate topography.
Stillwagon and Larson for spin coating over topography with The ability of the technique to measure film profiles was hindered
by the coherence of the laser producing speckle, internal reflec-
tions within the microscope, and the exposure time of the
photographic film.
Horowitz conducted an interferometric technique similar to
that of Graves. However, rather than photographic film,
Horowitz’s experiment utilised two photodetectors and an
analogue to digital converter to record the intensity.14 The ratio
between the reflected and incident intensity gave a value for
reflectivity. An example plot of modulated reflectance against
time (dubbed ‘‘optospinogram’’) is shown in Fig. 4a for spin-
coating silica spin-on glass a by sol–gel process in open air and in
a saturated solvent atmosphere.
Changes in optical thickness as a function of time where
obtained from the optospinogram by counting a quarter wave
every time reflectance evolved between two successive extremes.
Plots of optical thickness against time for the sol–gel in air and a
saturated solvent atmosphere showed film thinning occurring at
Fig. 3 (a) Schematic of experiment. As the slit on the wheel attached to
a faster rate for the sample in air compared to the saturated
the spindle passes between the light emitting diode and the photodiode, a
TTL signal is sent to the synchronization control, which triggers the solvent atmosphere. This study indicated that the role of
excimer laser. A 20 ns pulse of light from the excimer pumped dye laser is evaporation could not be disregarded when modelling the spin-
passed to the illuminator of the microscope. The interferometric images coating process and disproved the early assumption of a separate
are captured by a 35 mm camera; (b) interferogram of 50 mm line 17 s into solvent evaporation stage after convection has ceased as
spin; (c) interferogram of 500 mm line 1.5 s into spin. Taken from ref. 16. proposed by Meyerhofer.9

This journal is ª The Royal Society of Chemistry 2012 J. Mater. Chem. C


View Online

Fig. 4 (a) Double-beam experimental set-up with a He–Ne laser followed by Att. (attenuator), M. mirror, BS. beamsplitter, L 1–3. condenser lenses,
ch.1incident intensity at photodiode 1 (reference), transmitted intensity at photodiode 2 (sample), (b) ‘‘optospinogram’’ for spin coating of a sol–gel
solution in a saturated solvent atmosphere. Taken from ref. 14.
Published on 07 August 2012 on http://pubs.rsc.org | doi:10.1039/C2TC00026A
Downloaded by UNIVERSITY OF SOUTH AUSTRALIA on 08 October 2012

Birnie utilised laser interferometry to study solvent thinning flow and evaporation constants was achieved by performing a
behaviour on spinning silicon wafers.15 In contrast to the work of simple linear regression fit. Evaporation rates varied as a func-
Horowitz, relative reflectivity was measured rather than absolute tion of the square-root of spin speed as predicted by Meyerhofer.
reflectance, to allow interference fringe counting. As the bright/ When evaporation constants where derived for several solvents,
dark interference fringe spacing only measured thinning rate, a strong correlation was found between evaporation rates and
thickness was determined once thinning had proceeded to a their room temperature vapour pressures.
completely dry wafer by counting back from the reference point, Birnie followed on from this study by studying the flow and
the last intensity maximum corresponded to the time when the evaporation rate of mixed solvent solutions.18 The time evolution
wafer first became dry. A plot of log thickness against log time of fluid thickness for a series of two component solutions was
for methanol, ethanol, butanol and butyl acetate emphasized the studied as in the previous study. Methanol-butanol mixtures
similarity found for solutions thinning in this geometry and is where studied, due to their relative differences in volatility. In
shown in Fig. 5. addition a sol–gel solution was tested using the same methodo-
The results indicated that the process is initially dominated by logy, as a measure to test the effectiveness of this experimental
mass flow and later in the process solvent volatility dominates, as technique.
indicated by the each sequence reaching the dry state. By plotting As in the previous study, Meyerhofer plots for pure butanol
thinning rate (dh/dt) as a function of two times the cube of film and pure methanol were linear. However, butanol–methanol
thickness (2h3) (Fig. 6) (commonly referred to as a Meyerhofer mixtures exhibited more complex behaviour. Fig. 7 shows the
plot), Birnie was able to determine both fluid viscosity and the Meyerhofer plot of a 1 : 1 mixture of butanol : methanol. Two
evaporation rate for a variety of solvents. Analysis of the evap- distinct regions where observed, instead of one as with pure
oration data substantiated an earlier conclusion of the spin-speed solvents. Early in the spinning process the slope and intercept
dependence of evaporation rate, and confirmed that evaporation yield viscosities and evaporation rates that are intermediates
rate is limited by vapour diffusion though the laminar boundary between that of pure butanol and pure methanol. The second
layer above the disk.9 region present, later on in the spinning process, corresponded to
The flow behaviour was found to be largely consistent with the an evaporation rate and viscosity of that expected of pure
EBP model, however, Meyerhofer’s modification was required to butanol. Birnie assumed that the second stage represents a
account for evaporating fluids. Simultaneous extraction of both transition where viscosity and evaporation behaviour are both

Fig. 6 Meyerhofer plot for pure butanol spinning at 2000 rpm. The line
Fig. 5 Thickness development for several volatile fluids spinning at 2000 is a linear regression to the observed data. Taken and modified from
rpm. Taken from ref. 18. ref. 15.

J. Mater. Chem. C This journal is ª The Royal Society of Chemistry 2012


View Online
Published on 07 August 2012 on http://pubs.rsc.org | doi:10.1039/C2TC00026A
Downloaded by UNIVERSITY OF SOUTH AUSTRALIA on 08 October 2012

Fig. 7 Meyerhofer plot for a solution of equal volumes methanol and


butanol spinning at 2000 rpm. Solid line is a regression to early data
points, while the dashed line shows the expected response for pure Fig. 8 Experimental set-up for in situ light scattering consisting of a
butanol. Near the end of spinning the fluid behavior changes gradually by HeNe laser (633 nm), spin coater rotation stage, photo-diode detectors
evaporation of methanol, leaving behind a butanol-rich liquid. Taken screen and a CCD camera. Taken from ref. 11.
and modified from ref. 18.

light was detected by a photodiode, off-specular diffracted light


changing toward those of pure butanol. For solvent mixtures was collected on a screen and recorded by a CCD camera.
which where more unbalanced (1 : 9 or 9 : 1 – buta- Blends of polyfluorene derivatives where studied; poly(9,90 -
nol : methanol) the second transition region was more difficult to dioctylfluorene-co-benzothiadiazole) (F8BT) and poly(9,90 -
resolve. Which was attributed to evaporation and viscosity dioctylfluorene-co-ibs-N,N0 -(4-butylphenyl)-bis-N,N0 -phenyl-
behaviour being dominated by the higher component solvent. 1,4-phenylenediamine) (PFB). The F8BT/PFB is a blend of two
The sol–gel precursor solution showed behaviour consistent conjugated polymers (PFB is a hole acceptor and F8BT is an
with a single thinning region (with exception to the latter data electron acceptor), which is of interest for applications in blend-
points). However, the later data points suggested that the solvent based photodiodes.
depletion effect, as in butanol–methanol is occurring in sol–gel During spin coating of these systems the speculary reflectivity
coatings. Therefore the majority of the thinning behaviour could data exhibited a series of peaks and troughs corresponding to
be easily modelled, until the final moments of the process. constructive and destructive interference conditions of the
Comparison of the sol–gel precursor evaporation and viscosity reflected light as the film thickness decreased. Final film thickness
values to that of pure methanol and butanol, suggested that there was determined via AFM, which allowed deduction of the time
was no substantial polymerisation of the sol–gel, during the dependence of the film thickness. A linear change in refractive
majority of the spin coating process. However, the evaporation index with time from the mixed polymer–solvent state to the final
constant was below what might was expected for a solution with solid polymer film was assumed, in order to correct for the
a significant fraction of methanol, which indicated that evapo- changing refractive index. Fig. 9 shows specular reflectivity data
ration may be slowed by a gel-like surface skin on the spinning and a corresponding thickness-time curve extracted from these
solution. data.
Birnie hypothesised that if a volatile solvent is necessary to
stabilize the precursor components, then instability might occur
before the coating has set completely, leaving behind inferior
coatings. Therefore improved coating solutions may require
selection of an alternate solvent system with associated modifi-
cation to other solution chemical parameters. For greater
control, it was suggested to use solvents which are less volatile.18

Specular reflectivity combined with off-specular scattering


Heriot et al. integrated a light scattering set-up with an integrated
spin coater measuring both specular and off-specular scattering
as shown in Fig. 8.19 This allowed both the evolution of film
thickness and the development of the lateral phase separated
structure to be monitored as a function of time. As in previous
studies, specular reflectivity interferometry was used to monitor
changes in film thickness, while off-specular scattering allowed Fig. 9 Specular reflectivity for a 44/56 F8BT/PFB blend spin-cast at
observation of the onset of phase-separation and then evolution 2000 rpm from a 2% solution in xylene. The inset shows the thickness-
of length scales in the phase-separated blend. Specularly reflected time curve extracted from these data. Taken from ref. 19.

This journal is ª The Royal Society of Chemistry 2012 J. Mater. Chem. C


View Online

The off-specular scattering provided information on the of the domains roughly followed a t1/2 dependence on the
development of lateral structure in the film. Radially averaged spinning time, t. The result was surprising as it directly contra-
scattering data are shown as a function of time in Fig. 10. The dicted the Walheim model, where domains nucleate from the
scattering patterns obtained where radially symmetric and initially homogeneous mixture below a critical solvent concen-
remain through-out the phase-separation process. The plot tration and the grow until their size reaches the thickness of the
shows radially averaged data taken from around the first frame film.12
that exhibits scattering at 9200  30 ms to the point where the The proposed explanation for these results was that an initial
scattering pattern ceases to evolve at approximately 10 000 ms, lateral structure may form, (not from bulk-like phase separation
believed to be of a consequence of polymer solution becoming within the film), as a result of an instability in transient wetting
concentrated enough to form a glass. layers at the surface and substrate. The spinning fluid may
The point at which the off-specular scattering was first detec- initially form a layered structure, driven by the difference in
ted corresponded to the cloudpoint of a bulk phase separation surface energy between the two polymer solutions. As the solvent
experiment, i.e. the onset of lateral phase separation in the evaporates, surface and interfacial tensions that stabilise the
Published on 07 August 2012 on http://pubs.rsc.org | doi:10.1039/C2TC00026A
Downloaded by UNIVERSITY OF SOUTH AUSTRALIA on 08 October 2012

system. At the cloudpoint there was no drop in specular reflec- configuration are continuously changing, resulting in the devel-
tivity intensity, showing that phase separation could not be opment of interfacial instabilities (the origin of which was not
detected using specular reflectivity alone, and indicating that clear). These instabilities lead to dewetting and breakup of the
multiple scattering was not likely to be a significant layered arrangement and the development of the lateral domain
complication. structure with an exact morphology dependent on the initial
The scattering data showed two prominent features. The first composition of the blend. The lateral structure would then be
is that the scattering exhibited pulsations in intensity with a detected as soon as the layers begin to breakup.
periodicity that mirrored that of the specular reflectivity. Which Evidence in support of this proposed mechanism came from
was due to constructive–destructive interference effects within reports that spin-coating the same polymer from isodurene, a
the film which affect both specular and off-specular scattering. highly viscous and volatile solvent, forms a bilayer with the lower
By accounting for the subsequent periodic modulation of overall surface energy F8BT on top.20 In this instance the structural
scattering intensity, the scattering curves initially showed a evolution was frozen at the layered stage before lateral domains
strong similarity with that would arise from the scattering in bulk can develop. More direct evidence of this layering during the spin
phase separation systems. A scattering peak, characterised by a coating process was found through detailed analysis of the time
well-defined intensity maximum, grows smoothly out of the dependence of specular reflectivity in blends of polystyrene/
background, in a way strongly reminiscent of spinodal polyisoprene.21
decomposition. The study by Heriot and Jukes showed that it was possible to
However, as phase separation proceeds, the behaviour devi- measure the kinetics of film formation and phase separation in a
ated from that of conventional bulk separation. Fig. 10 inset spin-cast polymer blend using in situ reflectivity and light scat-
shows that the q value of maximum scattering as a function of tering techniques. The results of these experiment where
time, qmax moved to a higher value as phase separation pro- explained by two possible mechanisms: one in which morphology
ceeded, implying that the overall length scale decreased as phase arose by the unstable growth of either a composition fluctuation
separation proceeded. Further analysis of Fig. 10 showed that and another in which the morphology results from an interfacial
the average size of the phase separated domains, as deduced from instability initiated by the loss of solvent. A single length scale is
qmax, is initialy 5.7 mm which rapidly shrinks to a final size of initially selected, but contrary to the normal situation in bulk
3.6 mm. This result was in stark contrast to behaviour seen in spinodal decomposition following a temperature quench, this
the bulk, where qmax decreases as a result of coarsening. The size length scales subsequently appears to shrink rather than grow.
Heriot and Jones then investigated a PS:PMMA polymer
system using the above mentioned technique of measuring
specular reflectivity and off-specular scattering to study the
development of structure directly.11 Specular reflectivity data was
obtained for both pure PMMA and a PS:PMMA (1 : 1) blend.
Off-specular scattering was plotted as a intensity graph for the
PS:PMMA blend and is shown in Fig. 11. The plot shared the
same features as the F8BT/PBT data, in that there was a smooth
appearance of a broad scattering peak as in bulk phase-separa-
tion. As previously seen unlike bulk phase-separation, there was
a superimposed rise in scattering intensity, a periodic increase
and decrease in overall intensity (with a period matching specular
reflectivity) and the scattering vector qmax moved to a higher
value as phase separation proceeded.
Fig. 10 In situ light scattering taken during the spin-coating of a 44/56
The appearance of scattering with a well-defined peak scat-
F8BT/PFB blend from a 2% solution in xylene at 2000 rpm. The data tering vector is indicative of the development of lateral structure
shown are for a portion of the spinning event from the ‘‘cloudpoint’’ to within or at the surface of the film, in Fig. 11 the length scale
the point where the length scale stops evolving. The inset shows the time characterising the lateral structure is around 30 mm. Two
dependence of the peak wave vector qmax. Taken from ref. 19. mechanisms where proposed to account for the production of

J. Mater. Chem. C This journal is ª The Royal Society of Chemistry 2012


View Online
Published on 07 August 2012 on http://pubs.rsc.org | doi:10.1039/C2TC00026A
Downloaded by UNIVERSITY OF SOUTH AUSTRALIA on 08 October 2012

Fig. 12 Fringe visibility for PMMA and PS:PMMA with corresponding


Fig. 11 Contour plot showing the radially averaged light-scattering models. Taken from ref. 11.
intensity out of the specular direction for the blend film. Taken from
ref. 11.

that, during the initial phase of the instability, the r.m.s rough-
such a peak. The first is the bulk mechanism of spinodal ness increases to a value of 120 nm, at a time when the total film
decomposition, in which thermal composition fluctuations are thickness was 3.44 mm. The end of this phase was marked by an
amplified by a process that selects the fastest growing length abrupt shift in the position of the peak in off-specular scattering
scale. The second was that the initial phase-separation within the intensity; the shift corresponded to a decrease in the charac-
film is perpendicular to the plane of the film, resulting in a lateral teristic length scale from a value of 53 to 25 mm.
interface between layers of different composition.22 The interface They where then able to model reflectivity as a film thins using
then becomes unstable leading to a wavelength-selective ampli- a classical expression for the reflectivity of a dielectric film on a
fication of a capillary wave in a process analogous to spinodal de- substrate, for both pure PMMA and a PS:PMMA blend. Solvent
wetting.23 concentration was assumed to be uniform throughout the film,
The periodic modulation in intensity observed in these and interfacial roughness was included. The model fitted the
experiments was attributed due to the thin-film geometry. As the obtained data for pure PMMA well, except that it shows lower
film continues to thin, interference effects cause a periodic vari- visibility than the experimental data, due to the model not taking
ation with film thickness of the electric-field intensity at the into account of concentration gradients in the film. The model
position of the roughening interfaces from which scattering is for the PS:PMMA blend, did not fit the experimental data as
arises. well. However, the model did capture the most striking features
A more detailed analysis was performed on the specular of the data (i.e. the periodic modulation of fringe modulation).
reflectivity data, and offered support in favour of the second Fig. 13 summarises the conclusions from this study. Phase
suggestion for the mechanism producing the well defined scat- separation initially takes place by the formation of wetting layers
tering peak observed in Fig. 11. Fig. 12 shows the fringe visibility, at the surface and substrate (i)–(iii). When the thickness
calculated as (Rmax  Rmin)/(Rmax + Rmin). For the pure PMMA decreases to a critical value, the interface becomes unstable;
film there was little variation throughout the process, only a slow thermally excited capillary waves are amplified with a mechanism
decrease in fringe visibility, followed by a sharper increase which to select a length scale, resulting in the observed off-specular
was attributed to a coarsening of the film. For the PS:PMMA scattering (iv).
film at early stages of the spinning process, the blend showed a From these results Heriot and Jones speculated that instability
systematic modulation of fringe visibility with time that was not at the polymer–polymer interface arose because of a solvent-
observed with the pure polymer. This was interpreted to indicate concentration gradient through the film. Due to the solvent at the
that the film split into two layers, one rich in PS and the other film surface evaporating at a faster rate than the solvent in the
rich in PMMA. Later in the spinning process, at a time corre- bulk can diffuse through the film, resulting in a significantly
sponding to the onset of off-specular scattering, the fringe visi- lower solvent concentration at the surface compared to that of
bility decreased whereas the average reflectivity remained the substrate.23 The interfacial tension at a polymer–polymer
constant. which was accounted for by assuming the average interface is a strong function of the solvent concentration,25 so if
roughness of the free surface increases, leading to a reduction in an interface is in the plane of the film, it is very probable that it
the reflectance of that interface. would be subject to a Marangoni-like instability.
They went on to semi-quantitatively model the effect by As the film thinned further the instability is expected to grow,
assuming that the effect of small-length-scale roughness on the independent of its origin, leading to an increase in the off-spec-
reflectance of the top surface can be accounted for by a Debye– ular scattered intensity and an increase in the roughness of the
Waller-like factor as derived for X-rays.24 Analysis indicated top surface of the film. At some point, the amplitude of the

This journal is ª The Royal Society of Chemistry 2012 J. Mater. Chem. C


View Online

Fig. 13 A schematic model describing the film formation during the spin-coating process, and the final film morphology. Taken from ref. 11.
Published on 07 August 2012 on http://pubs.rsc.org | doi:10.1039/C2TC00026A
Downloaded by UNIVERSITY OF SOUTH AUSTRALIA on 08 October 2012

instability becomes such that the liquid–liquid interface meets the off-specular scattering data and optical micrographs of films,
surface. At this point there is a rapid movement of the contact taken at the centre and near the edge of the sample for PS/
lines to yield the laterally phase separated structure shown in step PMMA blends spun at different evaporation rates.
(v) of Fig. 13. This stage is marked by a decrease in surface Fig. 15b and c show the typical spin-cast film structures for a
roughness and an abrupt change in the peak in the scattering film spun in an ambient atmosphere. At the centre of rotation (b)
pattern. Simultaneously, or subsequently, as the total polymer there are isotropic islands of PMMA in a cellular pattern,
concentration is increasing and the equilibrium phase boundaries whereas off centre (c) the structure is radially orientated stria-
change, phase separation may be initiated in one or both phase, tions. The isotropic island and striations have a similar length
leading to a hierarchical, secondary phase separation step (vi). scale (50 mm), which is an indication that the hydrodynamics
Mokarian-Tabari et al. performed a quantitative evaluation of break the symmetry of the instability when the shear field is large,
the effect of evaporation rate during spin-coating of polymer without affecting the physics that controls the selection of a
blend films. Unlike Bernie’s experiments where the effect of length scale. Fig. 15a is a intensity plot showing the radially
different solvents with different evaporation rates was examined, averaged off-specular light scattering intensity for the blend film
toluene was spun-cast on silicon, inside a cell capable of spun cast in an ambient environment, which was the greatest
producing a controlled solvent atmosphere, allowing control evaporation rate (e ¼ 3.15 mm s1). The q(t) profile provides
over vapour pressure and consequently evaporation rate.7 A information on structure evolution during film formation. The
schematic of this experimental set-up is shown in Fig. 14a. profile shows three different regions. First, a smooth and broad
Experimental curves for the evolution of film thickness for a scattering peak shifts towards the largest scattering vector with
variety of values of the nominal applied vapour pressure are time, indicative of the onset of instability in the film. This was
shown in Fig. 14b. There where two clearly observable stages of then followed by the breakup of that structure and the formation
film formation; firstly, early on the spin-coating process hydro- of a pattern 12 mm. Finally a fixed pattern with a length scales
dynamic thinning in which the solvent does not strongly effect of 50 mm forms.
the thinning rate, secondly, an evaporation dominated stage in Fig. 15d–f, show data at a decreased evaporation rate (e ¼
which the rate of thinning could be systematically changed by the 1.03 mm s1), which has a different scattering pattern and final
solvent vapour pressure. The evaporation rate was extracted structure to that observed at a higher evaporation rate. The q(t)
from the thickness-time profiles using Meyerhofer’s model. As profile (Fig. 15d) shows no dominant q, nor scattering ring and
expected increasing the nominal vapour pressure lead to a takes longer to develop a fixed structure (6.5 s instead of 3.5 s).
decrease in the evaporation rate. Fig. 15e and f show the absence of a strong cellular pattern in the
The structure evolution for films spun from mixtures of PS and centre and that striations start to disappear. Exploiting slower
PMMA at controlled evaporation rates was then studied using evaporation rates, showed the above trend continued and at one
specular reflectivity and off-specular scattering. Fig. 15 shows the extreme e ¼ 0.42 mm s1 there was no lateral phase separation. As
shown in Fig. 15g, the q(t) profile does not reveal any dominant
q, indicative of a lack of any specific length scale in the film.
Evaporation in this film was very slow as it took 13 s for the
toluene to fully evaporate. The conclusions drawn from these
results were that due to a low evaporation rate, the concentration
gradient in the film is not large enough to trigger the Marangoni
instability. Therefore the interface oscillates between rough and
smooth but the magnitude of the instability is never large enough
to breakup the film, resulting in stratification. Selective washing
Fig. 14 (a) Schematic diagram showing the integrated light scattering
set-up with an integrated spin coater and environmental cell; (b) thick- was performed on the film formed at the slowest evaporation
ness–time profiles for toluene at different evaporation rates. The evapo- rate, which indicated a PMMA layer at the substrate with a PS
ration rate, e, is controlled by changing Pt, the nominal toluene vapour layer on top. The ion scattering techniques forward recoil spec-
pressure inside the cell. The points and lines are data and fits, respectively. trometry (FReS) and nuclear reaction analysis (NRA) where
Taken from ref. 7. then used to quantitatively analyse the layers. The results of

J. Mater. Chem. C This journal is ª The Royal Society of Chemistry 2012


View Online
Published on 07 August 2012 on http://pubs.rsc.org | doi:10.1039/C2TC00026A
Downloaded by UNIVERSITY OF SOUTH AUSTRALIA on 08 October 2012

Fig. 15 (a) The off-specular scattering profile shows the structural evolution when there is no environmental toluene (partial) pressure and the solvent
evaporation is fastest. There are three different stages: firstly smooth layering occurs, then the structure breaks up with features on many length scales
and finally (t > 3.5 s) a fixed pattern with a length scale of 50 mm forms. Optical micrographs taken at the centre (b) and at the edge (c) of the sample show
a laterally phase-separated structure (50 mm) with radial striations towards the edge of the sample. (d) Due to a lower evaporation rate in this sample, the
Marangoni-like instability is weaker and does not completely break through to the surface, so much less off-specular scattering is observed. The film
evolution is slower, with its structure frozen in after 6.5 s. In the optical micrographs (e and f) there is no cellular pattern and the striations are not
dominant. (g) The slowest evaporation rate used and therefore the concentration gradient which drives the Marangoni-like instability is most sup-
pressed. The magnitude of the roughness is never sufficient to break through to the surface. The micrographs showing the structures from the centre (h)
and the edge (i) of the sample reveal neither cellular patterns nor striations. Ion scattering reveals a layered structure in these samples. Taken from ref. 7.

which showed clear stratification of the film, with a clear segre- However, for fast evaporation in which diffusion cannot re-
gation of PS to the surface and towards the rear of the film, a equilibrate the concentration through the film, large solvent
PMMA-rich layer. concentration gradients exist near the surface of the film. Such
The conclusions from this study where that at high evapora- gradients destabilise the polymer flow causing the breakup of the
tion rates (e ¼ 1.8 mm s1), the film was unstable during evapo- (transient) wetting layer in a Marangoni-type instability, leading
ration, resulting in lateral phase separation. For evaporation to lateral phase separation.
rates between 1.5 and 1.8 mm s1 striations in the film may
develop. Decreasing the evaporation rate below e ¼ 1.6 mm s1
Video observation studies
suppresses the instability, allowing vertical phase separation to
proceed. For these low rates of evaporation, the solvent Studies based upon light scattering are based in reciprocal-space,
remained in the film long enough to allow the polymer surface where data is in the form of the Fourier transform of an image.
energies to dominate and control the final morphology. Therefore information on the morphological development is

Fig. 16 Frames captured from a video sequence taken of n-propanol spinning on a 50 mm diameter silicon wafer rotating at 2000 rpm. The three images
occur sequentially separated by 66.7 ms (every other frame from this sequence). (a), (b) and (c) correspond to times that are 3.87, 3.93, and 4.00 s into the
spinning run, respectively. Taken and modified from ref. 26.

This journal is ª The Royal Society of Chemistry 2012 J. Mater. Chem. C


View Online

The thinning was first tracked qualitatively by following the


progression of the yellowish zone that starts in the centre of
Fig. 16a. This coloured zone spreads centrifugally resulting in the
appearance of a colourful, growing ring as shown by the
expanding yellowish ring in Fig. 16b and c. The yellow colour
was matched to a corresponding film thickness from laser
interferometry and they assumed that the ensuing yellow rings
where indicative of the fluid reaching the same thickness at radial
positions further from the wafer centre. The process was con-
ducted for every observable colour displayed at the film centre, in
order to dynamically map the film thickness across the entire
Fig. 17 Thickness profile for fluid for four different frames 3.87, 3.93,
4.00, and 4.07 s into the spinning run (from top to bottom). Taken from wafer throughout the bulk of the latter stage of spin-coating.
ref. 26. Fig. 17 shows the resultant fluid thickness profiles as a function
Published on 07 August 2012 on http://pubs.rsc.org | doi:10.1039/C2TC00026A
Downloaded by UNIVERSITY OF SOUTH AUSTRALIA on 08 October 2012

of radius for 4 different time intervals, at 66.7 ms second gaps. As


expected from the visible colour variations seen in Fig. 16, there
inferred from the changes in the observed length-scale of the is a substantial increase in the fluid thickness with increasing
scattering sample. radius. With each time step there is an increment of solvent
In 2009 Birnie used colour video based observations in evaporation and a similar increment of outward radial flow, that
conjunction with thickness calibration with laser interferometry contribute to thinning over time.
to map the thickness profile evolution with time for fluid flowing
radially outward on a rotating silicon wafer during spin
Real-space studies
coating.26 This built upon previous work by Birnie, where colour
video image where used to determine film thickness for trans- In 2011 Ebbens et al. used a direct imaging approach that
parent coatings on silicon post processing.27 demonstrated a method to perform micrometer-scale lateral and
During the latter stages of the spin coating process the fluid nanometer vertical precision metrology during the spin coating
reaches thickness values such that interference effects cause the of a phase separating polymer blend, providing lateral in plane
film to exhibit distinct colours which correlate directly to thick- information in real-space with height information inferred from
ness. This effect arises as fluid thickness is dependent on radius. interference information.28
The study was performed in order to afford a clearer insight into The method employed synchronized stroboscopic LED illu-
the dynamics governing the spin coating process. mination with a high sensitivity electron-multiplying charge-
Laser interferometry was performed at the centre of rotation coupled device camera (EMCCD) connected to an episcopic
to determine fluid thickness as in previous experiments.14,15 optical microscope to optically observe structure evolution of
Matching fluid thickness at the centre of rotation at a specific polymer blends during spin coating. The sample was irradiated
point in time with the colour intensity observed at the centre of with a filtered high intensity near monochromatic green LED (Dl
rotation on the corresponding video frame provided an instan-  10 nm), which allowed interference reconstruction of three-
taneous calibration for determining the fluid thickness for other dimensional topographies of the spin-coated film as it dries and
locations, due to constant lighting conditions across the phase separates with nanometre precision.28 The experimental
substrate. Laser interferometry was performed at normal inci- set-up, shown in Fig. 18, is somewhat similar to Graves’, in that a
dence and video observation was performed away from the light source was synchronously triggered with a charge coupled
normal incidence. The video camera captures a colour image by device and a series of stroboscopically frozen images where
detecting the varying intensities of the reflected incident red, recorded. Images were recorded once every full revolution,
green and blue (RGB) light at a particular position on the wafer. allowing for a movie of the evolution of phase separation at the
By linking the evolution of colour intensity at the centre, to the same location to be recorded. An illumination source with a
corresponding fluid thickness measured by interferometry, Birnie narrow bandwidth was used so that interferometric height
was able possible to extend the intensity analysis to nearby reconstruction could be performed, as such contrast in the
positions on the wafer and by comparison determine the fluid stroboscopic images reflects instantaneous height differences at
thickness profile at a particular time. the sample surface and is as a consequence of constructive/
N-propyl alcohol was used to test this experimental study, as it destructive interference.
had a sufficiently low volatility to allow 60 distinct RGB intensity High molecular-weight polystyrene (PS) and polyisoprene (PI)
variations with which to generate colour matching calibration blends in o-xylene where studied in order to generate better
points. understanding of the processes involved in the fundamental
A typical fluid height vs. time plot was generated as previously phase separating behaviour of polymer blends. The high
described. Then the fluid thickness at the centre of rotation was molecular-weights of the polymers and relatively low vapour
matched to the colour of the fluid at an instant in time, thus a pressure of o-xylene ensured that length scales where optically
thickness value was assigned to a particular colour/hue. Then visible and that the phase separation and drying where relatively
from the interferometrically determined thickness vs. time slow.
dependence they effectively assigned each (R*, B*) value a Fig. 19 shows extracts from a movie which directly observes, in
particular thickness, where R* and B* are the scaled fractional real-space, the evolution of microstructure during polymer-blend
red and blue colour balance values of a given pixel. spin-coating, with resolution and clarity comparable to that

J. Mater. Chem. C This journal is ª The Royal Society of Chemistry 2012


View Online

was likely to be Marangoni type instabilities, due to solvent


evaporation. However, in the case of the PS/PI blend, Marangoni
type instabilities had little influence on the final structure. This
technique provides a new method to explore the extent to which
phase separation and Marangoni phenomena inter-relate to form
the resultant final morphology in a range of polymer systems.
Analysis of the intensity at each image pixel as function of time
allowed the corresponding drying rate of the film to be deter-
mined and mapped across the entire field of view, through using
methodology analogous to that of spatially averaged specular
laser methods previously mentioned. This required the assump-
tion of flat uniform film morphology at a sufficiently early stage
of the film formation, to be used as a starting point where all
Published on 07 August 2012 on http://pubs.rsc.org | doi:10.1039/C2TC00026A
Downloaded by UNIVERSITY OF SOUTH AUSTRALIA on 08 October 2012

oscillations of the specular reflectivity intensity would be in-


Fig. 18 Schematic diagram of a stroboscopic interference microscope. phase.
The drying curves extracted from the analysis revealed a
consistent difference in the drying rate between optically ‘‘bright’’
obtained from conventional optical microscopy of a static and ‘‘dark’’ domains as seen in the dry film under mono-
sample. As the height of the film is continuously decreasing due chromatic light. Resulting in the predicted height difference
to hydrodynamic flow/solvent evaporation, periodic interference between these locations, such that optically ‘‘dark’’ regions dried
contrast inversions occur according to Bragg’s law (as seen in the at a slower rate, resulting in a higher film height than the ‘‘bright’’
final two frames of Fig. 19). Early on in the spinning process (the regions. Applying the same analysis to the homopolymer samples
first frame, t ¼ 8 s) the image shows no contrast, as the surface is revealed that the pure PS homo-polymer dried completely within
flat and uniform. After approximately 12 seconds onwards, weak 12 s of spin coating, whereas, the pure PI homo-polymer took
large-scale fluctuations roughly 100 mm in length become 22 s to dry. The blend dry film end point lied at the midpoint of
observable, which then become increasingly dominated by these two values. The drying rate over the final 6 s was faster for
shorter scale fluctuations of approximately 20 mm, which appear PS than PI, and ellipsometry indicated that the final PS film
to have a fixed position until the final dry film at 17.2 s reflecting thickness was less than that for PI. Suggesting the faster drying
the expected stages of the spin-coating process. The radially ‘‘bright’’ domains where PS rich and the slower drying ‘‘dark’’
integrated Fourier transformed data shows the evolution of the domains where PI rich. This assignment was confirmed via
two lengths cales during the spin coating process. The longer selective washing, and confirmed that such drying rate curve
length scale decreased in length as the film dried, converging with analysis can allow material identification.
the length scale corresponding to the obvious phase separated Through application to every pixel value the method was
domain structure. The images and FFT analysis in Fig. 19 extended to reconstruct by interpolation the absolute quantita-
indicated the final phase-separated structure appeared very early tive sample height across the whole field of view at all stages
and evolved very little once the liquid–liquid transition had during the drying process, as shown in Fig. 20.
occurred, despite the fact solvent evaporation occurred for a The validity of the assumption of uniform film morphology
further 5 s. It was believed that the larger length scale structure with a low roughness as the start point for the analysis was tested

Fig. 19 (a) a series of unprocessed stroboscopic optical reflectance images recorded during spin coating of a 1 : 1 blend of PS:PI (2 wt% at 1500 rpm)
from o-xylene are shown. The images were illuminated with narrow-bandwidth green light at the indicated times during the spin-coating process (b)
corresponding background-corrected 2D Fourier transforms are shown underneath each frame, with the radial integral of the Fourier transform for a
section of the movie above. Taken from ref. 28.

This journal is ª The Royal Society of Chemistry 2012 J. Mater. Chem. C


View Online
Published on 07 August 2012 on http://pubs.rsc.org | doi:10.1039/C2TC00026A
Downloaded by UNIVERSITY OF SOUTH AUSTRALIA on 08 October 2012

Fig. 20 2D relative height reconstructions during spin coating a PS:PI blend at 1500 rpm obtained from pixel by pixel drying curve arrays generated
from fringe counting the stroboscopic optical images. (Bottom) 3D surface morphology reconstructions. Taken from ref. 28.

by comparing the final reconstructed morphology with AFM z-direction within 10% accuracy. Fig. 21 schematically illustrates
metrology carried out at the same region of the final sample how the bis-continuous phase-separated morphology appeared
surface. Cross-section analysis revealed a quantitative agreement well before the end point of the drying process and that no
between the reconstructed and ‘‘real’’ AFM data in the significant lateral translation occurred after this point. Apparent
features in the final surface topography where initially visible
early on in the spinning process where the film thickness was
greater than 2 mm (10 greater than final thickness). FFT
analysis of the raw optical images (Fig. 19), indicated features at
this length scale where first visible after 11 s after spin coating
begins (6 s before the sample dried). The early surface structure
observed supported the theoretical prediction that that phase
separation originated at the surface first and that the original
structure forms a motif to direct the subsequent phase separation
of solvent richer underlying regions.13
The morphological reconstructions confirmed that phase
separation in the PI/PS system followed a straightforward spi-
nodal decomposition process, with little influence by Marangoni
instabilities. However, fine details of the phase-separated
morphology develop over fractions of seconds of the drying
process and appear fixed once formed.

Conclusions
In situ studies of spin coating have been vital in developing
knowledge of the many different parameters that determine the
final film thickness and the rich morphologies generated. The
early studies of Horowitz, and later Birnie, confirmed the validity
of the EBP model in the early stages of the process. The work by
Birnie also validated the Meyerhofer’s extension of the EBP
model to include evaporation rate and showed that a two stage
process adequately described the spin coating of simple solvents
and binary solvent mixtures. Work conducted by Heriot,
Fig. 21 (Left) Schematic cross section during the phase separation Jones and Jukes combined the techniques of reflectivity
process. (Right) Quantitative reconstructed surface topography at 0.25 s measurement and light scattering, in order to investigate more
intervals (start time 13 s). Estimated polymer content labeled on right- complex polymer blend systems. which cast insight into the
hand axis. Taken from ref. 28. processes involved in spin coating and phase separation. The

J. Mater. Chem. C This journal is ª The Royal Society of Chemistry 2012


View Online

results indicated that phase separation in thin films did not occur 11 S. Y. Heriot and R. A. L. Jones, An interfacial instability in a
via the same mechanisms as in the bulk case. The later work of transient wetting layer leads to lateral phase separation in thin spin-
cast polymer-blend films, Nat. Mater., 2005, 4(10), 782–786.
Birnie then Ebbens and Howse, utilised novel experimental 12 S. Walheim, M. Boltau, U. Steiner and G. Krausch, Phase Separation
approaches to investigate further intricacies of the processes in Thin Films of Strongly Incompatible Polymer Blends. 1999, pp. 75–
involved. Technological development of EMCCD’s and high 99.
intensity pulsed LED’s (which don’t suffer from speckle) now 13 G. A. Buxton and N. Clarke, Ordering polymer blend morphologies
via solvent evaporation, Europhys. Lett., 2007, 78(5), 1.
make such studies possible, which provide insight into both the 14 F. Horowitz, E. Yeatman, E. Dawnay and A. Fardad, Real-time
details of the spin coating process and opening a route to optical monitoring of spin-coating, J. Phys. III, 1993, 3(11), 2059–
understanding and controlling morphological development. 2063.
15 D. Birnie, Combined flow and evaporation of fluid on a spinning disk,
Phys. Fluids, 1997, 9(4), 870.
16 (a) L. Manske, Dynamic measurements of film thickness over local
topography in spin coating, Appl. Phys. Lett., 1990, 56(23), 2348;
References (b) L. M. Peurrung and D. B. Graves, Film thickness profiles over
Published on 07 August 2012 on http://pubs.rsc.org | doi:10.1039/C2TC00026A

topography in spin coating, J. Electrochem. Soc., 1991, 138(7),


Downloaded by UNIVERSITY OF SOUTH AUSTRALIA on 08 October 2012

1 C. J. Brabec, M. Heeney, I. McCulloch and J. Nelson, Influence of 2115–2124.


blend microstructure on bulk heterojunction organic photovoltaic 17 L. E. Stillwagon and R. G. Larson, Leveling of Thin Films over Uneven
performance, Chem. Soc. Rev., 2011, 40(3), 1185–1199. Substrates During Spin Coating, AIP, 1990, vol. 2, pp. 1937–1944.
2 X. Li, R. Xing, Y. Zhang, Y. Han and L. An, Molecular weight effects 18 D. P. Birnie, Combined flow and evaporation during spin coating of
on the phase morphology of PS/P4VP blend films on homogeneous complex solutions, J. Non-Cryst. Solids, 1997, 218, 174–178.
SAM and heterogeneous SAM/Au substrates, Polymer, 2004, 45(5), 19 P. C. Jukes, S. Y. Heriot, J. S. Sharp and R. A. L. Jones, Time-
1637–1646. resolved light scattering studies of phase separation in thin film
3 S. Walheim, M. Boltau, J. Mlynek, G. Krausch and U. Steiner, semiconducting polymer blends during spin-coating,
Structure formation via polymer demixing in spin-cast films, Macromolecules, 2005, 38(6), 2030–2032.
Macromolecules, 1997, 30(17), 4995–5003. 20 N. Corcoran, et al., Increased efficiency in vertically segregated thin-
4 A. D. F. Dunbar, P. Mokarian-Tabari, A. J. Parnell, S. J. Martin, film conjugated polymer blends for light-emitting diodes, Appl. Phys.
M. W. A. Skoda and R. A. L. Jones, A solution concentration Lett., 2003, 82(2), 299–301.
dependent transition from self-stratification to lateral phase 21 A. Budkowski, et al., Substrate-determined shape of free surface
separation in spin-cast PS:d-PMMA thin films, Eur. Phys. J. E: profiles in spin-cast polymer blend films, Macromolecules, 2003,
Soft Matter Biol. Phys., 2010, 31 (Copyright (C) 2011 American 36(11), 4060–4067.
Chemical Society (ACS). All Rights Reserved), 369–375. 22 J. W. Cahn, Phase separation by spinodal decomposition in isotropic
5 K. Dalnoki-Veress, J. A. Forrest, J. R. Stevens and J. R. Dutcher, systems, J. Chem. Phys., 1965, 42(1), 93–99.
Phase separation morphology of thin films of polystyrene/ 23 P. G. de Gennes, Instabilities during the evaporation of a film: non-
polyisoprene blends, J. Polym. Sci., Part B: Polym. Phys., 1996, glassy polymer plus volatile solvent, Eur. Phys. J. E: Soft Matter
34(17), 3017–3024. Biol. Phys., 2001, 6(5), 421–424.
6 L. Fang, M. Wei, C. Barry and J. Mead, Effect of spin speed and 24 L. Nevot and P. Croce, Characterization of surfaces by grazing x-ray
solution concentration on the directed assembly of polymer blends, reflection – application to study of polishing of some silicate-glasses,
Macromolecules, 2010, 43(23), 9747–9753. Rev. Phys. Appl., 1980, 15(3), 761–779.
7 P. Mokarian-Tabari, M. Geoghegan, J. R. Howse, S. Y. Heriot, 25 D. Broseta, et al., A theoretical and experimental-study of interfacial-
R. L. Thompson and R. A. L. Jones, Quantitative evaluation of tension of immiscible polymer blends in solution, J. Chem. Phys.,
evaporation rate during spin-coating of polymer blend films: control 1987, 87(12), 7248–7256.
of film structure through defined-atmosphere solvent-casting, Eur. 26 D. P. Birnie, D. E. Haas and C. M. Hernandez, Laser interferometric
Phys. J. E: Soft Matter Biol. Phys., 2010, 33 (Copyright (C) 2011 calibration for real-time video color interpretation of thin fluid layers
U.S. National Library of Medicine), 283–289. during spin coating, Opt. Lasers Eng., 2010, 48(5), 533–537.
8 A. G. Emslie, F. T. Bonner and L. G. Peck, Flow of a viscous liquid 27 D. P. Birnie, Optical video interpretation of interference colors from
on a rotating disk, J. Appl. Phys., 1958, 29(5), 858–862. thin transparent films on silicon, Mater. Lett., 2004, 58(22–23), 2795–
9 D. Meyerhofer, Characteristics of Resist Films Produced by Spinning, 2800.
AIP, 1978, vol. 49, pp. 3993–3997. 28 S. Ebbens, R. Hodgkinson, A. J. Parnell, A. Dunbar, S. J. Martin,
10 (a) C. J. Lawrence, The mechanics of spin coating of polymer-films, P. D. Topham, N. Clarke and J. R. Howse, In situ imaging and
Phys. Fluids, 1988, 31(10), 2786–2795; (b) D. E. Bornside, height reconstruction of phase separation processes in polymer
C. W. Macosko and L. E. Scriven, Spin coating – one-dimensional blends during spin coating, ACS Nano, 2011, 5 (Copyright (C) 2011
model, J. Appl. Phys., 1989, 66(11), 5185–5193. American Chemical Society (ACS). All Rights Reserved), 5124–5131.

This journal is ª The Royal Society of Chemistry 2012 J. Mater. Chem. C

You might also like