ArticlePDF Available

Tribological Properties of Poly(methyl methacrylate) Brushes Prepared by Surface-Initiated Atom Transfer Radical Polymerization

Authors:

Abstract and Figures

Surface-initiated living radical polymerization of methyl methacrylate (MMA) using copper/bipyridyl complex was carried out from silicon wafer immobilized with a 2-bromoisobutylate moiety, resulting in the assembly of polymer chains tethered by one end to a surface, a so-called “polymer brush”. The thickness of the PMMA brush, which was estimated by ellipsometry and atomic force microscopy, increased linearly with the molecular weight of the chains, and was controlled by the free initiator concentration. The graft density of the PMMA brushes was estimated to be as high as 0.56 chains/nm2. The frictional properties of the high density PMMA brush were characterized by sliding the stainless ball probe on its surface across a width of 20 mm at a rate of 90 mm/min under loading of 0.49 N in air at room temperature. The PMMA brush was found to have a lower frictional coefficient and much better wear resistance than the corresponding spin-coated PMMA film because of the anchoring of the chain ends in the brush. In addition, the friction coefficient of the polymer brush significantly decreased in response to soaking in acetone and toluene, and increased in response to immersion in hexane. The tribological properties depended on the solvent quality.
Content may be subject to copyright.
Tribological Properties of Poly(methyl methacrylate)
Brushes Prepared by Surface-Initiated Atom
Transfer Radical Polymerization
Hiroki SAKATA, Motoyasu KOBAYASHI, Hideyuki OTSUKA, and Atsushi TAKAHARA
y
Institute for Materials Chemistry and Engineering, Kyushu University,
6-10-1 Hakozaki Higashi-ku, Fukuoka 812-8581, Japan
(Received May 16, 2005; Accepted July 4, 2005; Published October 15, 2005)
ABSTRACT: Surface-initiated living radical polymerization of methyl methacrylate (MMA) using copper/bipyr-
idyl complex was carried out from silicon wafer immobilized with a 2-bromoisobutylate moiety, resulting in the assem-
bly of polymer chains tethered by one end to a surface, a so-called ‘‘polymer brush’’. The thickness of the PMMA brush,
which was estimated by ellipsometry and atomic force microscopy, increased linearly with the molecular weight of the
chains, and was controlled by the free initiator concentration. The graft density of the PMMA brushes was estimated to
be as high as 0.56 chains/nm2. The frictional properties of the high density PMMA brush were characterized by sliding
the stainless ball probe on its surface across a width of 20 mm at a rate of 90mm/min under loading of 0.49 N in air at
room temperature. The PMMA brush was found to have a lower frictional coefficient and much better wear resistance
than the corresponding spin-coated PMMA film because of the anchoring of the chain ends in the brush. In addition, the
friction coefficient of the polymer brush significantly decreased in response to soaking in acetone and toluene, and
increased in response to immersion in hexane. The tribological properties depended on the solvent quality.
[DOI 10.1295/polymj.37.767]
KEY WORDS Tribology / Friction Coefficient / ATRP / Surface-Initiated Polymerization /
Poly(methyl methacrylate) (PMMA) / Polymer Brush / Wear Resistance /
Wear-resistant materials with low friction are
required for various modern technologies ranging
from a hard disk drives to roller bearings in automo-
bile. In particular nonlubricanted moving devices will
be increasingly needed for nano-machine technologies
in future. To improve the physicochemical properties
of solid surfaces, we have been investigating the tribo-
logical properties of ultrathin films prepared by chem-
ical vapor adsorption of organosilane compounds with
long alkyl chains, which are immobilized on a silicon
substrate by means of silyloxy bonds, forming a high-
density monolayer structure.
1
This immobilized
monolayer has a low friction coefficient and great
wear resistance, probably because of the low interfa-
cial energy of monolayer surface and strong adhesion
between monolayer and substrate. Consequently, the
following study focused on a surface-grafted polymer,
referred to as a ‘‘polymer brush’’.
Over the last decade, high-density and well-defined
polymer brushes have been readily synthesized, since
controlled/‘‘living’’ polymerization techniques
2–4
have been applied to the surface-initiated polymeriza-
tion.
5–7
One of the advantages of living polymeriza-
tion is quantitative initiation, which means that poly-
mers are propagated efficiently from the initiators on
a substrate. Practically, only 1 of 10 initiators bound
to the surface is expected to initiate a polymerization
because of the steric hindrance of the growing poly-
mer, however, it is sufficient to afford a high-density
brush. Atom transfer radical polymerization (ATRP)
has been widely employed for the formation of poly-
mer brushes because ATRP is compatible with various
functionalized monomers,
8–12
and the living character
of the ATRP process yields polymers with a low poly-
dispersity. Many researchers are investigating the sur-
face properties of tailored polymer brushes,
13
includ-
ing adhesion,
14
chromatography,
15
biomaterials,
16
wettability,
17
and biocompatibility.
18
Although the molecular mechanism of tribology on
the brush is still difficult to understand experimentally
due to the complexity of friction, lubrication, and
wear phenomena, some research groups have tried
to analyze the frictional properties of the polymer
brush. Klein et al. have found a reduction in the fric-
tional forces between solid surfaces bearing polymer
brushes using a newly developed surface force bal-
ance,
19,20
and they have also reported that brushes of
a charged polymer (polyelectrolyte) can act as effi-
cient lubricants between mica surfaces in an aqueous
medium.
21,22
Similarly, Osada and his coworkers have
reported that the well-defined polyelectrolyte brushes,
prepared by controlled radical polymerization using
TEMPO, reduce the surface friction of hydrogels
in water.
23
They have also found that the friction
y
To whom correspondence should be addressed (Tel: +81-92-642-2721, Fax: +81-92-642-2715, E-mail: takahara@cstf.kyushu-u.ac.jp).
767
Polymer Journal, Vol. 37, No. 10, pp. 767–775 (2005)
depends on the brush length; i.e., gels with longer
polymer brushes show higher friction. The lubrication
of a polymer brush is interesting from both a scientific
and a technological perspective. Tsujii et al. have
measured topographic images and force-distance pro-
files of high-density PMMA brushes by scanning force
microscopy (SFM) using a micro silica sphere attached
to a cantilever head.
24,25
The highly anisotropic struc-
ture of the swollen brush in toluene revealed extremely
strong resistance against compression. These results
imply the potential of the polymer brush to serve as
a low-friction and a wear-resistant film, although its
macro tribology and mechanical properties have not
yet been well studied, despite their very importance
with regard to practical applications. Therefore we
started a project to synthesize tethered polymer brush-
es on flat silicon substrate by surface-initiated ATRP
of MMA, investigating the sliding friction with a stain-
less probe and the wear resistance of the brush surface
under a load of 108Pa pressure. The various solvent
effects on the tethered brush during sliding friction
tests are also described in this paper.
EXPERIMENTAL
Materials
Anisole was stirred with sodium tips at 383 K for
6 h, followed by distillation from sodium under
reduced pressure. Copper bromide (CuBr, Wako Pure
Chemicals Industries Ltd.) was purified by washing
with acetic acid and dried under vacuum.
26
Methyl
methacrylate (MMA) purchased from Wako Pure
Chemicals was distilled under reduced pressure over
CaH2before use. 4,40-Di-n-heptyl-2,20-bipyridine
(Hbpy) was prepared by the dilithiation of 4,40-dimeth-
yl-2,20-bipyridine followed by coupling with 1-bromo-
hexane according to the method of Matyjaszewski
et al.
27
Ethyl 2-bromoisobutylate (EB), purchased
from Tokyo Chemical Inc., was distilled before use.
Water for contact angle measurements and frictional
tests was purified with the NanoPure Water system
(Millipore, Inc.). All other reagents were purchased
from commercial sources and used as received.
Synthesis of 60-dimethoxymethyliylyhexyl 2-bromoiso-
butylate (DMSB)
Bromoisobutyl bromide (51.5 mmol) was added
dropwise to a stirred solution of 6-undecenyl-1-ol
(51.2 mmol) and triethylamine (71.9 mmol) in dry di-
chloromethane (85 mL) under a nitrogen atmosphere
with cooling in an ice bath, and the solution was then
stirred overnight at room temperature (Scheme 1).
7
After the filtration with suction to remove the hydro-
chloric salt, the reaction solution was washed with
0.5 N HCl and water, and dried over MgSO4. The
product was distilled under reduced pressure (321–
322 K/0.1 mmHg) to give 60-hexenyl 2-bromoisobu-
tylate as a colorless liquid (31.9 mmol, 76%). The ob-
tained 60-undecenyl 2-bromoisobutylate (16.2 mmol)
and chloroplatinic acid (0.019 mmol, 2-propanol solu-
tion) were placed in a three-necked flask, dimethoxy-
methylsilane (32.6 mmol) was slowly added to the
mixture over a period of 30 min. During the reaction
mixture was stirred at 323 K for 12 h, the conversion
was checked by NMR. After unreacted dimethoxy-
methylsilane was removed under a vacuum, the solu-
tion was passed through a short column of anhydrous
Na2SO4to remove the catalyst, and distilled to give
12.8 mmol (79%) of 60-dimethoxymethylslilylhexyl
2-bromoisobutylate (DMSB) as a colorless oily resi-
due. Further purification was not carried out before
use. 400 MHz 1H NMR (CDCl3): 0.1 (SiCH3), 0.6
(SiCH2), 1.4 (–CH2–), 1.7 (OCH2CH2–), 1.9 (CH3),
3.5 (SiOCH3), 4.2 (COOCH2). 100 MHz 13 C NMR
(CDCl3): 5:7(SiCH3), 13.1 (SiCH2), 22.7 (SiCH2-
CH2), 25.5 (SiCH2CH2CH2), 28.3 (OCH2CH2), 30.9
(CH3), 32.8 (OCH2CH2CH2), 50.2 (SiOCH3), 56.1
(CBr), 66.2 (OCH2), 171.8 (C=O).
OH Br Br Br
C
O
C
CH3
CH3
Et3N
C
O
C
CH3
CH3
O
HSiMe(OMe)2
H2PtCl6
MMA
CuBr / Hbpy
Si wafer
CVA
MMA
CuBr / Hbpy
EB
O
O
Si(CH2)6OOC
Me
CBr
CH3
CH3
EtOOC C Br
CH3
CH3
EtOOC C CH2
CH3
CH3
C
CH3
Br
COOCH3
free polymer
O
O
Si(CH2)6OOC
Me
CCH
2
CH3
CH3
C
CH3
Br
COOCH3
n
C
O
CBr
CH3
CH3
(CH2)6O(CH3O)2Si
CH3
DMSB
PMMA brush
n
+
Scheme 1.
H. SAKATA et al.
768 Polym. J., Vol. 37, No. 10, 2005
Preparation of Initiator-Immobilized Silicon Substrate
The silicon(111) wafers (40 mm 8mm) were
immersed in a mixture of conc. H2SO4and 30% H2O2
aqueous solution (70/30, v/v) at 373 K for 1 h to
remove the organic contaminant from their surface.
Successively, substrates were further cleaned by expo-
sure to vacuum ultraviolet-ray (VUV, ¼172 nm)
for 5 min under reduced pressure at 15 mmHg to result
in the hydrophilic surface. These silicon wafers and a
glass vessel filled with 10% toluene solution of DMSB
were packed in a Teflon container purged with N2gas,
and were allowed to stand in an autoclave at 373 K for
5 h. During the heating at this temperature, DMSB
vapor adsorbed on the surface of the wafers to make
an organosilane monolayer, which is known as the
chemical vapor adsorption (CVA) method.
28,29
After
these wafers were rinsed with toluene and ethanol,
they were dried in vacuo at 373 K for 10 min and
was stored in a dark place.
Surface-Initiated ATRP
Typical polymer brush growth was achieved by
placing the DMSB-immobilized substrates in a glass
tube equipped with a stop cock under argon gas and
adding a degassed anisole solution of CuBr (0.020
mmol), Hbpy (0.040 mmol), MMA (50.0 mmol), and
ethyl 2-bromoisobutylate (EB, 0.010 mmol) as a free
initiator. The total volume of the polymerization solu-
tion was approximately 10 mL. The polymerization
solution was degassed by repeating the freeze-and-
thaw process, and the glass tube was sealed off under
the vacuum condition. The polymerization was then
allowed to proceed for a set reaction time (2–24 h)
at 363 K, and was terminated by cooling the solution
to 273 K and the addition of a small amount of meth-
anol under ambient pressure. The conversion of MMA
was estimated by 1H NMR spectra of the polymeriza-
tion solution, comparing the relative intensities of the
signals due to the unreacted monomer and the pro-
duced polymer. The silicon substrates were washed
with toluene using a Soxhlet apparatus for 12 h to re-
move the free polymer absorbed on their surfaces, and
were dried under the reduced pressure at 373 K for 1 h.
The polymer solution was passed through the alumina
column using THF to remove catalyst, and was poured
into the methanol to precipitate the free polymer. Us-
ing the obtained free polymer, spin coat films on sili-
con wafers were also prepared from toluene solution.
The thickness of the spin coat film was adjusted to that
of the corresponding polymer brush by changing the
spinning rate and concentration of toluene solution.
Measurements
The number-average molecular weights (Mn) and
molecular weight distribution (MWD) of the free
polymer were determined by size exclusion chroma-
tography (SEC) recorded on a Tosoh GPC-8010 sys-
tem using polystyrene standards calibration, which
runs through two directly connected polystyrene gel
columns (Shodex GPC KF-804L, 1.0 mL/min) using
THF as an eluent at 313 K. The NMR spectra were
measured in CDCl3with a Jeol EX-400 (1H 400
MHz) system. IR spectra were obtained with a Per-
kin-Elmer Spectra-One KY type (Perkin-Elmer) sys-
tem coupled with a Mercury Chromium Tell detector.
The incident angle of the p-polarized infrared beam to
the silicon wafer was 73.7(Brewstar angle). The
thickness of the polymer brush and the spin coat film
on the silicon substrate were determined by an imag-
ing ellipsometer (Nippon Laser & Electronics Lab.)
equipped with a YAG laser (532.8 nm). The polarizer
angle was fixed at 50, and a refractive index of 1.49
was used for the calculations of the film thickness.
Atomic force microscopic (AFM) observation was
done with SPA 400 with an SPI 3800N controller
(Seiko Instruments Industry Co., Ltd.) in air at room
temperature, using a Si3N4integrated tip on a com-
mercial triangle 100 mmcantilever (Olympus Co.,
Ltd.) with a spring constant of 0.09 N/m. XPS meas-
urements were carried out on a PHI ESCA 5800 (PHI
Electronics Co., Ltd.) at 105Pa using a monochro-
matic Al-KX-ray source. The contact angles against
water were recorded with a drop shape analysis sys-
tem DSA10 Mk2 (KRU
¨SS Inc.) equipped with a video
camera. The frictional coefficient of the polymer
brushes and the cast films were recorded on a Tribos-
tation Type32 (Shinto Scientific Co., Ltd.) by sliding a
stainless ball (10 mm) on the substrates over a width
of 20 mm at a sliding velocity of 90 mm/min under
loading of 0.20–0.98 N at 298 K (Figure 1). The fric-
tion force on the ball probe was transmitted to a stress
gauge attached to a probe end, and was recorded auto-
matically. Friction tests in various solvents were also
carried out using the polymer brush substrates, which
were immersed in the corresponding solvents for
24 h, in advance. The morphologies of the wear trace
of the thin films were observed with an S-4300SE
20 ~ 100 g
Stainless ball
PMMA brush
Silicon wafer
Load
Sliding platform
Friction Probe
Figure 1. Schematic description of the friction tester setup.
Tribological Properties of PMMA Brushes
Polym. J., Vol. 37, No. 10, 2005 769
field-emission SEM (Hitachi Co., Ltd.) equipped with
an X-ray microanalysis system (Genesis 7000, EDAX
Co., Ltd.) in order to examine the elements on the
wear surface.
RESULTS AND DISCUSSION
Surface-Initiated ATRP of MMA
The CVA method is often used to prepare the
high-density monolayer films of organosilane com-
pounds.
30,31
Takai et al. have carried out AFM analy-
sis of the monolayer film prepared by CVA, and they
found that the monolayer surface is very smooth, with
no aggregates and low number of defects.
32
We pre-
pared here the silicon substrate immobilized with rad-
ical initiator, DMSB, by the CVA method. The typical
water contact angle of silicon wafer irradiated by
VUV is lower than 5, but water contact angle on
the substrate increased to 87after the immobilization
of DMSB by CVA. XPS spectra of initiator-immobi-
lized silicon wafer showed a carbon signal (C1s)at
286 eV associated with the organic portion of the
attachable initiator along with the bromide (Br3d) sig-
nal at 71 eV. The C1s signals attributed to C=O and
C–O bonds were also observed in the narrow scan
mode. These results are indicative of formation of a
DMSB thin layer on the silicon wafer.
Surface-initiated radical polymerizations of MMA
were carried out in the presence of EB as a free initia-
tor coupled with CuBr and Hbpy (mole ratio EB/
CuBr/Hbpy/MMA = 1/2/4/1000). Figure 2 shows
the plots of Mnand the Mw=Mnindex of free polymer
produced in the solution as a function of monomer
conversion, where the Mnand MWD values were
determined by polystyrene-calibrated SEC. The poly-
dispersities of the free polymers were relatively low,
and the Mnvalues were proportional to the monomer
conversion, with the slope being very close to the
theoretical value. These findings indicate that the con-
trolled polymerization proceeded with a restriction of
transfer and termination reactions. The Mnof surface-
grafted PMMA on a silicon wafer cannot be directly
determined yet, however, a polymer brush should
have the same molecular weight as the value of the
corresponding free polymer.
24
As shown in Figure 3,
the thickness of the polymer brushes increased linear-
ly with molecular weight. The thickness of the ob-
tained polymer brush was smaller than the theoretical
value of the all-trans conformation given by 0:254N
and was larger than that of the random coil conforma-
tion calculated by 2ðNb2=6Þ1=2, where Nand bare the
degree of polymerization and the statistical segment
length of 0.68 nm, respectively.
33
According to the
proportional relationship between the thickness Ld
(nm) and Mn, the graft density was estimated to
be ca. 0.56 (chains/nm2) by following equation,
¼dLdNA1021=Mn
where dand NAare the assumed density of bulk
PMMA at 293 K and Avogadro’s number, respective-
ly. This graft density is comparatively high, taking
into account the volume fraction of the polymer chain.
Hence, the tethered PMMA chains would have a rela-
tively extended conformation along the direction nor-
mal to the substrate surface. The AFM observation
revealed that a homogeneous polymer layer was
formed on the substrate. The root mean square
(RMS) of the surface roughness was found to be ap-
proximately 1.0 nm in a 10 10 mm2scanning area
at any location. The water contact angle of PMMA
brush was 78, which was very close to the value
(80) for spin-coated PMMA film. The thickness of
the brush did not change in response to repeated rins-
ing with toluene using a Soxhlet apparatus; therefore,
the polymer chains were not physically adsorbed but
were chemically anchored on the substrate. Formation
of the PMMA brush was also confirmed with XPS and
0
10
20
30
40
50
1.0
1.5
2.0
2.5
0102030405060
Mn x 10-3
Mw / Mn
Conversion/ %
Mn (theo)
Figure 2. Mnand MWD of free PMMA obtained with EB/
CuBr/Hbpy in anisole at 363 K: ( )Mn;( ) MWD. The dashed
line indicates the theoretical value of Mngiven by the ratio of
[MMA]/[EB] and conversion.
0
10
20
30
40
0 1020304050
Thickness of PMMA brush / nm
Mn x 10-3
Figure 3. Relationship between brush thickness and Mnof
free polymer.
H. SAKATA et al.
770 Polym. J., Vol. 37, No. 10, 2005
Brewstar FT-IR spectra, although the results are not
shown here.
Friction Behavior in Dried State
Dynamic friction tests were carried out by sliding a
stainless ball on the substrates at a rate of 90 mm/min
in air under the normal load ranging from 20–100 g at
room temperature. In the case of a non modified sili-
con wafer under a normal load of 50 g (0.49 N), the
theoretical contact area between a stainless probe and
substrate can be calculated to be 2:43 109(m2)by
Hertz’s theory,
34
and the average pressure on the con-
tact area was estimated to be 201 MPa. Although the
actual contact area on the PMMA brush substrate
might be larger than the theoretical value, the average
pressure supposed to be more than 102MPa. We
attempted to demonstrate here whether the polymer
brush could resist such a high pressure and friction
for industrial application. As shown in Figure 4(a),
the dynamic friction coefficient of the polymer brush
was found to be 0.20–0.22 under a normal load of
20 g, and 0.24–0.25 under a normal load of 50 and
100 g. However, the magnitude of the friction coeffi-
cient of the polymer brush was independent of the
brush thickness and sliding velocity,
35
and was almost
constant at all normal loads. Large error margins were
observed in short brush around 5–20 nm, while the
margins of the error bars decreased in size with the
thickness of the brush layer, and the friction coeffi-
cient converged to a constant value. These results
indicate that stable sliding was achieved on the sub-
strates, and wear resistance was improved due to the
tethered polymer chain end by covalent bonds on
the substrate. On the other hand, the dynamic friction
coefficient of the cast film increased with film thick-
ness. In addition, a larger friction force was detected
than that of brush; for example, the friction coeffi-
cients of cast film and brush under loading of 100 g
were 0.31 and 0.24, respectively. The error margins
of the cast films were also larger than those of brush
at any film thickness. The thicknesses of these cast
films and the molecular weights of spin-coated poly-
mer were almost the same as those of the correspond-
ing polymer brushes; therefore, the grafting structure
of polymer brush should contribute to lowering the
friction, as described later.
To evaluate the wear resistance of polymer brush, a
continuous friction test for 600 s were performed by
sliding a stainless ball on substrates covered with
20 nm of polymer brush or cast film under a load of
0.49 N at a rate of 90 mm/min (Figure 5). In the early
stage of friction test, the friction coefficient of the
brush surface increased to 0.5 from 0.25, but there
was no further increase during continuous friction,
as shown in Figure 5(a). On the other hand, the
friction coefficient of the cast film was gradually
0.10
0.15
0.20
0.25
0.30
0 5 10 15 20 25 30 35
Brush [100g]
Brush [50g]
Brush [20g]
Friction coefficient
Thickness of PMMA brush / nm
0.10
0.15
0.20
0.25
0.30
0 5 10 15 20 25 30 35
Cast [100g]
Cast [50g]
Cast [20g]
Friction coefficient
Thickness of cast film / nm
(a)
(b)
Polymer brush
Cast film
Figure 4. Friction coefficient under dry condition of PMMA
brush (a) and cast film (b).
0.0
0.50
1.0
1.5
2.0
0 100 200 300 400 500 600
Friction coefficient
Sliding time /s
0.0
0.50
1.0
1.5
2.0
0 100 200 300 400 500 600
Friction coefficient
Sliding time /s
(a) Polymer brush
(b) Cast film
Figure 5. Friction time dependence of friction coefficient on
PMMA brush (a) and cast film (b) under a load of 0.49 N at a slid-
ing velocity of 90 mm/min in air: Mnof PMMA ¼26000; Thick-
ness of brush and cast film = 20 nm.
Tribological Properties of PMMA Brushes
Polym. J., Vol. 37, No. 10, 2005 771
increasing with friction time [Figure 5(b)], and the
surface polymer gradually peeled away to form many
debris (‘‘wear elements’’), some of which were ad-
sorbed on the surface of the probe (‘‘transfer parti-
cles’’) or left on the wear track, interfering with
smooth sliding of the probe. A sudden increase of fric-
tion coefficient to 1.5 was observed after 400 s, prob-
ably because the polymer layer on the wear track had
completely peeled off.
The trace scratched by the sliding probe can be seen
in the SEM image of wear tracks after the 600-s fric-
tion test (Figure 6). No topographic difference in the
width and morphology of the scratched tracks could
be found between the polymer brush and cast film.
An accumulation of wear elements produced by
scratching can also be observed at both ends of the
sliding trace in both pictures. However, the elemental
analysis showed different results. Signals due to car-
bon Kand oxygen Kcould be clearly observed at
the point of wear track on polymer brush by energy-
dispersive X-ray (EDX) spectra, which indicated that
the PMMA component remained. The observed peak
at 1.8 keV attributed to silicon Kmust have originat-
ed from the silicon substrate. As the matter of course,
all peaks attributed with carbon, oxygen, and silicon
atoms were observed at the outside of wear tracks.
On the contrary, peaks due to neither carbon nor oxy-
gen were observed in the EDX spectrum from the
wear track of the spin-cast film. Accordingly, a sliding
probe readily scratched the polymer layer on cast film
because the polymer molecules on cast film were not
anchored on the substrates. In the case of polymer
brush, polymer chains were immobilized on the sub-
strates and were highly extended in a perpendicular
direction for steric reasons related to the high graft
density.
36
We suppose that such a bound and aniso-
tropic structure restricted the mobility of the end-
grafted chain to prevent stretching and scattering by
the sliding probe. Of course, some of the brushes were
worn away under the strong pressure, as shown in
SEM image, although the rest of the tethered compo-
nents remained during the friction test to result in the
low magnitude of friction coefficient.
Effects of Solvent Quality on Friction Behavior
Since the early stages of polymer brush synthesis,
many researchers have suggested that the thickness
of polymer brush changes in response to soaking in
various solvents; the polymer brush can be stretched
in good solvents, and is compressed in poor solvents
or when in a dry state on the substrate. Furthermore,
the surface morphology and roughness can also be
arranged using solvation of polymer brushes. For
example, Zhao et al. have reported observing that dif-
ferent surface morphologies of polystyrene-block-
PMMA brush on the substrates layer by AFM after
treatment with cyclohexane and dichloromethane.
37,38
The solvent quality affects whether the polymer brush
will be stretched or compressed. We next attempted to
investigate the influence of solvent quality on the fric-
tion of polymer brushes. Friction tests of cast films
were not performed because the cast films can be
dissolved in organic solvents and removed from the
substrates.
The friction coefficients of the PMMA brush with
various thickness were measured by sliding a stainless
probe under a load of 50 g at room temperature in
acetone as a good solvent, and in n-hexane as a poor
0.0 1.0 2.0
Intensity / cps
/keV
C
O
Si
0.0 1.0 2.0
Intensity / cps
/ keV
C
Κ
α
Κ
α
Κ
α
Κ
α
Κ
α
Κ
α
O
Si
100µm
250 µm
250 µm
50 µm
Cast film
Brush film
(a)
(b)
Figure 6. EDX spectra and SEM images of wear track on the PMMA brush (a) and the cast film (b) after the sliding friction test in
Figure 5.
H. SAKATA et al.
772 Polym. J., Vol. 37, No. 10, 2005
solvent [Figure 7(a)]. Compared with the friction
coefficients of polymer brush in the dry state, the val-
ue decreased in both organic solvents due to the fluid
lubrication effect. The friction coefficient in acetone
was lower than that in hexane, and it decreased with
increases in the brush thickness. The friction coeffi-
cient of the PMMA brush with a 30-nm thickness
was 0.05, which is a half magnitude of the non modi-
fied substrate, suggesting that the polymer in good
solvents performs as a good lubricant. In n-hexane,
however, the friction coefficient of brush with any
thickness was almost as the same as that of non modi-
fied silicon substrate. This finding suggests that the
polymer brushes in poor solvents did not perform well
as lubricants, in other words, only fluid lubrication oc-
curred. Similar results were observed in toluene and
cyclohexane solution, as shown in Figure 7(b). The
dynamic friction coefficient in toluene approached to
a value of 0.06 on a 25-nm thickness brush, while
the friction coefficient for sliding friction in a cyclo-
hexane system was higher than that in toluene, main-
taining a constant value around 0.14, regardless of the
brush thickness. These results suggest that the interac-
tion between the brush surface and the stainless ball
(friction probe) was moderated because acetone and
toluene were good solvents for PMMA. On the other
hand, the PMMA brush surface would be unwilling
to be in contact with poor solvent such as hexane and
cylcohexane, and prefers to interact with the stainless
probe, thus giving a higher friction coefficient.
The wear resistance in solvents was also measured
under a normal load of 0.49 N and a sliding velocity of
90 mm/min for 10 min in toluene and cyclohexane.
As we expected, better wear resistance of the polymer
brush was observed in toluene compared with that in
cyclohexane. Figure 8 shows the SEM image and
EDX spectra of the worn surface after the wear-resist-
ance test. The wear track in toluene seems to be slight
compared with the scratched surface under the non
solvent condition in Figure 6. Almost no accumula-
tion of wear elements was observed at the sides and
ends of the sliding trace. Strong peaks due to carbon
and oxygen Kwere observed in EDX spectra from
the worn track surface, indicating that PMMA brush
components still covered the substrate surface. We
believe that the polymer brush, together with toluene,
worked as a lubricant, and reduced the interaction
between the stainless probe and the brush. The molec-
ular motion and characteristic structure of polymer
brush might impact the wear resistance, especially in
a solution. As mentioned above, polymer brush chains
in an good solvents typically form an extended struc-
ture along the perpendicular direction and show high
repulsion against the compression, while the polymer
chains tend to be compressed or collapse in poor sol-
vents.
39
However, brush structures in the presence and
absence of a solvent would be squashed by a probe
under as high pressure as 102MPa. Therefore, we sup-
pose that the sliding friction and wear resistance
0.0
0.050
0.10
0.15
0 5 10 15 20 25 30 35
Thickness of PMMA brush / nm
In cyclohexane
In toluene
Friction coefficient
0.0
0.050
0.10
0.15
0 5 10 15 20 25 30 35
Friction coefficient
Thickness of PMMA brush / nm
In hexane
In acetone
(a)
(b)
Figure 7. Friction coefficient of the PMMA brush in hexane
and acetone (a), in cyclohexane and toluene (b) under a load of
0.49 N at room temperature.
0.0 1.0 2.0 / keV
C
Κ
α
Κ
α
Κ
α
O
Si
Intensity / cps
In toluene
100 µm
100 µm
Figure 8. EDX spectra and SEM photograph of wear track on PMMA brush after sliding friction test in toluene under a load of 0.49 N
at a rate of 90 mm/min for 600 s: Mnof PMMA ¼28000; Thickness of brush = 20 nm.
Tribological Properties of PMMA Brushes
Polym. J., Vol. 37, No. 10, 2005 773
depend more on the interaction of the brush and probe
than the extended structure of polymer brush.
CONCLUSIONS
The high-density PMMA brushes were obtained by
surface-initiated atom transfer radical polymerization
using a silicon wafer covered with a flat monolayer
of 2-bromoisobutylate derivatives prepared by the
CVA method. According to the proportional relation-
ship between the thickness and Mnof the polymer
brush, the graft density was estimated to be 0.56
chains/nm2, which is the so-called high-density brush.
The friction coefficient of the polymer brush under a
normal load of 0.49 N in air at room temperature
was a lower than that of spin-coated film having the
same Mnand thickness as the corresponding polymer
brush. The polymer brush revealed better wear resist-
ance than the spin cast film because of the end-grafted
structure of the polymer brush. The tribological prop-
erties of brush in solution were dramatically changed
by the solvent quality. The friction coefficient of
PMMA brush in a good solvent such as toluene,
decreased further compared with those in air and in
a poor solvent such as hexane. In good solvent, poly-
mer brush performed as a lubricant to reduced interac-
tion between the probe and brush surface; as such, the
wear resistance was also significantly improved. From
these results, it can be concluded that polymer brush
prepared by surface-initiated polymerization exhibits
excellent tribological properties compared with those
prepared by spin-coating.
Acknowledgment. This work was partially sup-
ported by a Grant-in-Aid for the 21st century COE
Program ‘‘Functional Innovation of Molecular Infor-
matics’’ from the Ministry of Education, Culture,
Sports, Science and Technology of Japan. The FE-
SEM observation was performed using S-4300SE
(Hitachi Co., Ltd.) at the Collabo-station II, Kyushu
University.
REFERENCES
1. H. Ishida, T. Koga, M. Morita, H. Otsuka, and A. Takahara,
Tribology Lett.,19, 523 (2005).
2. K. Matyjaszewski and J. Xia, Chem. Rev.,101, 2921 (2001).
3. C. J. Hawker, A. W. Bosman, and E. Harth, Chem. Rev.,
101, 3661 (2001).
4. M. Kamigaito, T. Ando, and M. Sawamoto, Chem. Rev.,
101, 3689 (2001).
5. M. Ejaz, S. Yamamoto, K. Ohno, Y. Tsujii, and T. Fukuda,
Macromolecules,31, 5934 (1998).
6. B. Zhao and W. J. Brittain, J. Am. Chem. Soc.,121, 3557
(1999).
7. M. Husseman, E. E. Malmstrom, M. McNamara, M. Mate,
D. Mecerreyes, D. G. Benoit, J. L. Hedrick, P. Mansky,
E. Huang, T. P. Russell, and C. J. Hawker, Macromolecules,
32, 1424 (1999).
8. T. von Werne and T. E. Patten, J. Am. Chem. Soc.,121,
7409 (1999).
9. M. Ejaz, K. Ohno, Y. Tsujii, and T. Fukuda, Macromole-
cules,33, 2870 (2000).
10. C. Perruchot, M. A. Khan, A. Kamitsi, S. P. Armes,
T. von Werne, and T. E. Patten, Langmuir,17, 4479 (2001).
11. W. Huang, J. B. Kim, M. L. Bruening, and G. L. Baker,
Macromolecules,35, 1175 (2002).
12. M. Ejaz, S. Yamamoto, Y. Tsujii, and T. Fukuda, Macromo-
lecules,35, 1412 (2002).
13. K. Kato, M. Uchida, E.-T. Kang, Y. Uyama, and Y. Ikada,
Prog. Polym. Sci.,28, 209 (2003).
14. J. J. Lin, J. A. Silas, H. Bermudez, V. T. Milam, F. S. Bates,
and D. A. Hammer, Langmuir,20, 5493 (2004).
15. X. Huang and M. J. Wirth, Anal. Chem.,69, 4577 (1997).
16. T. Matsuda, M. Kaneko, and S. Ge, Biomaterials,24, 4507
(2003).
17. A. M. Granville and W. J. Brittain, Macromol. Rapid
Commun.,25, 1298 (2004).
18. R. Iwata, P. Suk-In, V. P. Hoven, A. Takahara, K. Akiyoshi,
and Y. Iwasaki, Biomacromolecules,5, 2308 (2004).
19. J. Klein, E. Kumacheva, D. Mahalu, D. Perahia, and L.
Fetters, Nature,370, 634 (1994).
20. R. Tadmor, J. Janik, and J. Klein, Phys. Rev. Lett.,91,
115503 (2003).
21. U. Raviv, S. Giasson, N. Kamph, J.-F. Gohy, R. Je
´ro
ˆme, and
J. Klein, Nature,425, 163 (2003).
22. J. Klein, U. Raviv, S. Perkin, N. Kampf, L. Chai, and S.
Giasson, J. Phys.: Condens. Matter,16, S5437 (2004).
23. Y. Ohsedo, R. Takashina, J. P. Gong, and Y. Osada, Lang-
muir,20, 6549 (2004).
24. S. Yamamoto, M. Ejaz, Y. Tsujii, M. Matsumoto, and T.
Fukuda, Macromolecules,33, 5602 (2000).
25. S. Yamamoto, M. Ejaz, Y. Tsujii, and T. Fukuda, Macromo-
lecules,33, 5608 (2000).
26. T. Grimaud and K. Matyjaszewski, Macromolecules,30,
2216 (1997).
27. K. Matyjaszewski, T. E. Patten, and J. Xia, J. Am. Chem.
Soc.,119, 674 (1997).
28. T. Koga, H. Otsuka, and A. Takahara, Chem. Lett.,31, 1196
(2002).
29. A. Takahara, H. Sakata, M. Morita, T. Koga, and H. Otsuka,
Compos. Interfaces,10, 489 (2003).
30. H. Tada and H. Nagayama, Langmuir,10, 1472 (1994).
31. K. Hayashi, N. Saito, H. Sugimura, O. Takai, and N.
Nagagiri, Langmuir,18, 7469 (2002).
32. A. Hozumi, Y. Yokogawa, Y. Kameyama, H. Sugimura, K.
Hayashi, H. Shinohara, and O. Takai, J. Vac. Sci. Technol.,
A,19, 1812 (2001).
33. K. Tanaka, K. Kojio, R. Kimura, A. Takahara, and T.
Kajiyama, Polym. J.,35, 44 (2003).
34. If a circle with radius a(m) is regarded as a contact area
between a stainless probe and substrate under a normal load
P(0.49 N), Hertz’s theory afford a following relationship
using Young’s modulus of stainless and silicon wafer, EA
H. SAKATA et al.
774 Polym. J., Vol. 37, No. 10, 2005
(1:96 1011 Pa), EB(1:30 1011 Pa), and Poisson’s ratio vA
(0.30), vB(0.28);
2=E¼ð1vA
2Þ=EAþð1vB
2Þ=EB
a¼ð3=42=EPRAÞ3=2
where RAis curvature radius (5:00 103m) of stainless
ball.
35. Unpublished data. Even if the stainless probe slided on the
surface of polymer brush with 20–30 nm thickness at rate
of 90 and 900 mm/min, the friction coefficients were almost
same within error margin.
36. S. Yamamoto, T. Tsujii, and T. Fukuda, Macromolecules,
35, 6077 (2002).
37. B. Zhao, W. J. Brittain, W. Zhou, and S. Z. G. Cheng,
J. Am. Chem. Soc.,122, 2307 (2000).
38. S. G. Boyes, W. J. Brittain, X. Weng, and S. Z. G. Cheng,
Macromolecules,35, 4960 (2002).
39. A. Roters, M. Schimmel, J. Ruhe, and D. Johannsmann,
Langmuir,14, 3999 (1998).
Tribological Properties of PMMA Brushes
Polym. J., Vol. 37, No. 10, 2005 775
... The pioneers of polymer brushes for tribology, Klein et al., also realised the importance of a good solvent to facilitate swelling and therefore better sliding performance [288]. The immersion of polymer brushes in good solvent allows brush swelling, known to help lubricate as illustrated in Figure 3.6 [223,[289][290][291][292][293][294]. Polymer brushes can also be based on zwitterionic monomers which are usually hydrated by water solutions, however these will not be covered in this review due to their poor characteristics in non-aqueous solvents such as motor oil. ...
... This is due to the repulsion of brush-brush interaction within the system which gives rise to entropic stabilisation. Also key to the low friction force is the grafting density of the polymer brush, as has been shown by multiple researchers comparing spin-coated or adsorbed PMMA samples and the corresponding tribological results [272,291]. Yamamoto et al. also concludes that a higher density of PMMA results in a higher repulsive ability in addition to more resistance to compression [227]. ...
... The relative success of the bare silicon nitride contact may be due to the hydrated silicon layer offering a lubricating effect [357]. Sakata et al. also suggested that in a poor solvent, PMMA brushes may preferentially interact with stainless steel probes resulting in a relatively high friction force [291]. No clear relationship between the thickness and Ff was found in water lubricated PMMA brushes. ...
Thesis
This thesis presents a novel lubrication solution for silicon nitride hybrid bearings developed through the use of polymer brush technology, specifically brushes created using surface initiated atom transfer radical polymerisation. This work also details a novel testing regime utilising custom colloidal probes to replicate, for the first time, this hybrid bearing under atomic force microscopy in both dry and lubricated conditions. Due to their promising tribological properties polymer brushes have the potential to be a lubrication solution for the hybrid bearing system where current lubrication solutions are not tailored to the surfaces and contain harmful components such as sulphur and phosphorus. Polymer brush systems have generated considerable interest in the academic community as a possible new greener lubrication solution. To further understand the mechanism by which an effective polymer brush can be employed in a tribological contact this study was initiated. As the first known study to investigate the effect of the polymer brushes on the silicon nitride-steel contact, previous literature findings have been reapplied to a novel material for a novel application. Grafting from the silicon nitride surface ensures that less additive competition will occur. Poly(methyl methacrylate) (PMMA) brushes were chosen for the reduction of steric hindrance within the polymer chain therefore allowing a higher density brush and better load carrying capacity in a tribological sense. These brushes act synergistically with a poly-alpha-olefin, a high quality base oil lubricant present in the type of engine where these hybrid bearing operate. The synergy here refers to the swelling effect in which the anchored macromolecule and base oil work as one to repel the asperity contacts, reducing friction whilst the brush system protects itself. The formation of polymer brushes on a silicon nitride surface utilises atom transfer radical polymerisation (ATRP) and activators regenerated by electron transfer (ARGET) coupled with a surface initiation step. Initiating from the surface allows a strong covalent bond to the contact surface ensuring stability when the final brush is subjected to physical interactions, the main advantage being that by adding monomer molecules individually in situ the steric interaction of the chain-chain iterations in the growing brush is reduced so denser films can be formed, especially with small molecules such as MMA. By applying recent developments such as ARGET synthesis of the polymer brush is made much easier, as this technique allows reactions to occur with limited amounts of oxygen present as well as reducing the quantity of the copper complex needed for the reaction. By investigating the chemical and mechanical properties of the polymer brush with techniques such as ellipsometry, atomic force microscopy (AFM), and x-ray photoelectron spectroscopy (XPS), it is possible to suggest explanations for the tribological properties of the polymer brush system. Detailed XPS analysis shows that bromine is still present at the surface, a key indicator that the functional sites of the polymer process are still available to be bonded for increasing the chain length while also indicating that fewer termination reactions had occurred. With a lack of silicon visible in the XPS sample spectra it is clear that the polymer has achieved good surface coverage and should therefore exhibit better tribological characteristics. By using novel, custom made stainless steel colloidal probes, it has been possible for the first time to replicate the hybrid contact on the nanoscale, which allows high quality testing by accurately replicating the materials in contact and thus an effective evaluation of the lubrication solution. The importance of the polymer thickness, measured by ellipsometry, and the liquid in which they are solvated, is clearly elucidated by testing in multiple fluids, when highly synergetic fluids like the poly-alpha-olefin result in a significant reduction in friction whereas poor solvents like water can even be detrimental when compared to the bare surfaces in contact. In the worst case scenario under the highest load using the novel probes the lubricated polymer brush reduced the friction force successfully from 3.3 nN to 1.3 nN when compared to the bare nitride surface. Preliminary work has been completed in respect to the transition from the nanoscale to the macroscale, and the polymerisation reaction has been scaled from 1 cm2 silicon nitride wafers up to a 10 cm diameter silicon nitride discs. One of the reasons why the polymerisation can be scaled in such a way is due to the ARGET technique which allows polymerisations to occur in the presence of limited amounts of air. A tribological study of these PMMA modified disc surfaces using a pin-on-disc setup shows favourable results and on average a reduction of friction of 15% when comparing PMMA modified surfaces with unmodified ones in an oil lubricated environment.
... 87,92, There have also been many reviews concentrating on various aspects of PBs from their synthesis and promising applications, 24,116 such as friction control 32,56,61,117,118 and biopassivity, 33 to theory and computational modelling. 47,48,94,119,120 Experimentally, the tribological performance of PBs has been studied against various control parameters such as grafting density, 37,68,71,74,78,84,86,89,90,121 the degree of polymerization, 8,10,37,65,70,73 or architecture of the grafted chains 11,33,56,87 as well as for different solvents, 6,9,55,64,66,72,[75][76][77][80][81][82][83] which for charged brushes could also include different counterions or salt concentrations. 9,83,[122][123][124] The measurement of the interaction forces between PBs was originally performed using the surface forces apparatus (SFA), 59,64,[67][68][69]125 which is a microscale technique. ...
Article
Full-text available
When polymer chains are grafted to solid surfaces at sufficiently high density, they form brushes that can modify the surface properties. In particular, polymer brushes are increasingly being used to reduce friction in water-lubricated systems close to the very low levels found in natural systems, such as synovial joints. New types of polymer brush are continually being developed to improve with lower friction and adhesion, as well as higher load-bearing capacities. To complement experimental studies, molecular simulations are increasingly being used to help to understand how polymer brushes reduce friction. In this paper, we review how molecular simulations of polymer brush friction have progressed from very simple coarse-grained models toward more detailed models that can capture the effects of brush topology and chemistry as well as electrostatic interactions for polyelectrolyte brushes. We pay particular attention to studies that have attempted to match experimental friction data of polymer brush bilayers to results obtained using molecular simulations. We also critically look at the remaining challenges and key limitations to overcome and propose future modifications that could potentially improve agreement with experimental studies, thus enabling molecular simulations to be used predictively to modify the brush structure for optimal friction reduction.
... In particular, polymer brushes with a normalised graft density higher than 10% is known as a 'concentrated polymer brush (CPB)' [9]. CPBs in good solvents have been reported to exhibit ultralow friction in both micro [10] and macro [11][12][13][14] scale sliding; therefore, their application to mechanical components such as bearings and oil seals is expected. The challenge of applying CPBs for actual industrial applications is their durability against friction. ...
Article
Full-text available
Concentrated polymer brushes (CPBs) are promising soft-material coatings for improving tribological properties under severe sliding conditions, even in the macroscopic scale. Therefore, they are expected to be applied to mechanical sliding components. However, the durability of CPBs has remained challenging for industrial applications. Previous studies revealed that applying a groove texture to the CPB substrate is effective in improving the durability of CPBs. In order to achieve further improvement of durability of CPBs, we attempted to apply nano-periodic structures, whereas the groove texture applied in previous studies has widths and depths in micrometres. In this study, the effect of the nano-periodic structure in addition to the groove texture applied to the CPB substrate on the durability of CPB is investigated. The results demonstrate a significant improvement in the durability of CPBs by up to 90% compared with non-textured CPB when an appropriate nano-periodic structure is applied (i.e. a nano-periodic structure oriented parallel to the groove texture). Graphic Abstract
Article
The effects of a cross-linking layer on the wear resistance of polymer brush were investigated by using molecular dynamics-based sliding simulations. We found that a cross-linking layer improved wear resistance. The cross-linking layer suppressed the interpenetration of polymer chains on the counter surface and thus lowered the frictional force and wear. The degrees of interpenetration decreased as the cross-linking layer closed to the tip of the chain. A cross-linking layer in the tip of the polymer chains was thus found to improve wear resistance most effectively.
Article
Surface-tethered polymers have been shown to be an efficient lubrication strategy for boundary and mixed lubrication by providing a solvated film between solid surfaces. We have assessed the performance of various graft copolymers as friction modifier additives in oil and revealed important structure–property relationships for this application. The polymers consisted of an oil-soluble, grafted poly(lauryl acrylate) segment and a polar, linear poly(4-acryloylmorpholine) anchor group. Reversible addition–fragmentation chain transfer polymerization was used to access various architectures with control of the grafting density and position of the anchor group. Macrotribological studies displayed promising results with ≈50% reduction in friction coefficient at low polymer treatment rates. QCM-D experiments, neutron reflectometry, small-angle neutron scattering, and atomic force microscopy were used to gather detailed information on these polymers’ surface adsorption characteristics, film structure, and solution behavior.
Article
Full-text available
High-density, end-anchored macromolecules that form so-called polymer brushes are popular components of bio-inspired surface coatings. In a bio-mimetic approach, they have been utilized to reduce friction, repel contamination and control wetting, in particular in the development of biomedical materials. For reliable application of these coatings, it is critical that the performance of these coatings does not degrade in time. Yet, it is well-known that polymer brushes can deteriorate and degraft when exposed to water(-vapor) and this strongly limits the durability of these coatings. In this article, we provide an overview of the current status of research on the stability of polymer brushes. Moreover, we review different synthetic strategies, some of which are bio-inspired by itself, to enhance the long-term stability of these brushes. Based on this overview, we identify open question and issues to be resolved for brushes to be applied as durable bio-inspired surface coatings.
Article
Full-text available
We show that a surface-grafted polymer brush, 1-n-butyl-3-vinyl imidazolium bromide-based poly(ionic liquids), is able to reduce the interfacial friction by up to 66% and 42% in dodecane and water, respectively. AFM-based force spectroscopy reveals that the polymer brush adopts distinctively different interfacial conformations: swollen in water but collapsed in dodecane. Minimal surface adhesion was observed with both polymer conformations, which can be attributed to steric repulsion as the result of a swollen conformation in water or surface solvation when the hydrophobic fraction of the polymer was exposed to the dodecane. The work brings additional insight on the polymer lubrication mechanism, which expands the possible design of the polymer architecture for interfacial lubrication and modification.
Article
Concentrated polymer brushes (CPBs) are swollen polymeric materials promising for tribological applications owing to their ultralow friction. However, their insufficient durability has limited their practical use. In this study, to improve the durability, poly(methyl methacrylate) CPBs fabricated on microgrooved substrates were experimentally investigated in two types of tests (i.e., “sliding tests” and “nanoindentation tests”). Sliding tests revealed that parallel grooves with adequate dimensions significantly improve their durability in lubricated sliding. Nanoindentation tests revealed that the CPB on a flat substrate exhibits a layered structure comprising two layers (i.e., a “diluted surface layer” and a “concentrated bulk layer”). An additional third layer (i.e., a “reinforced tough layer”) provided by the parallel grooves appears to improve the durability.
Article
Full-text available
Well-defined poly(methyl methacrylate) (PMMA) brush layers were prepared onto silicon wafers by surface-initiated atom transfer radical polymerization using 2-(4-chlorosulfonylphenyl)-ethyltrichlorosilane as an initiator. Based on molecular weight and layer thickness measurements, it was deduced that the apparent graft density was 0.6–0.8 chains nm-2 depending on the polymerization time and that the conformation of tethered chains was highly extended. Surface relaxation behavior of the PMMA brush layer and the spin-coated PMMA film was examined by lateral force microscopy. The αa- and β-relaxation processes were discernibly observed at both surfaces. Although surface molecular motion of the brush layer and the spin-coated film was markedly different from the bulk one, both were hardly distinguishable in terms of relaxation phenomena.
Article
Metal-catalyzed living radical polymerization was discussed. This process is used for the precision synthesis of various polymers with controlled architectures. The advantage of this process is that it is versatile toward a variety of monomers and is feasible in a wide range of recation conditions.
Article
Well-ordered amino-terminated self-assembled monolayers (SAMs) were reproducibly prepared on Si substrates covered with native oxide in a vapor of 12.5 vol % solution of N-(6- aminohexyl)-3-aminopropyltrimethoxysilane (H2N(CH2)6NHCH2CH2CH2Si(OCH3)3,AHAPS) diluted with absolute toluene. Although aggregated AHAPS molecules were excessively adsorbed on the deposited AHAPS-SAM films, they were removed by sonication in ethanol, toluene, NaOH, and HNO3 aqueous solutions, conducted in that order. The thickness of the AHAPS-SAM as estimated by ellipsometry was 1.3±0.1 nm. The AHAPS-SAM surfaces observed by atomic force microscopy appeared very smooth with a root mean square roughness of about 0.15 nm in a several micron square area. This resulted in low hysteresis between the advancing and receding water-contact angles, which were determined to be 62±3° and 57±2°, respectively. Micropatterning of the SAM was also demonstrated on the basis of photolithography using an excimer lamp radiating vacuum ultraviolet light of 172 nm in wavelength. A microstructure composed of 5 μm×25 μm rectangular features was successfully fabricated on an AHAPS-SAM surface. © 2001 American Vacuum Society.
Article
Organosilane monolayers are novel ultrathin films used to control the physicochemical properties, such as friction and wear, of solid surfaces. In this study, the authors prepared alkylsilane and fluoroalkylsilane monolayers with a series of chain lengths by a chemical vapor adsorption method. The monolayers’ tribological properties were investigated by lateral force microscope (LFM) and friction tester. LFM nanoscale measurements of tribological properties showed that alkylsilane monolayer gave lower lateral force than the Si substrate surface. The lateral force decreased as the length of the alkyl chain increased. On the macroscale, friction test revealed that the organosilane monolayers gave lower dynamic friction coefficients than the Si substrate surface in air at room temperature. The longer the alkyl chain, the greater the wear resistance of the organosilane monolayers. Friction experiments using tetradecane as a lubricant showed better tribological properties than were obtained in air. Furthermore, microscopically line-patterned two-component organosilane monolayers were prepared and their macroscopic friction behavior was investigated. Even though the height difference between the two-components was less than 1 nm, friction force anisotropy between the parallel and perpendicular directions against the line pattern was observed.
Article
Three-component micropatterned organosilane monolayer was successfully fabricated on Si-wafer substrate by a stepwise photolithography process. Scanning force microscopic observations revealed that three kinds of organosilane monolayers with different physicochemical properties were area-selectively immobilized on a Si wafer.
Article
The use of lubricants to reduce friction and wear between rubbing surfaces has been documented since antiquity. Recent approaches have focused on boundary lubrication by surfactant-like species coating the surfaces, whereby the friction between them is replaced by the weaker forces required for shear of adhesive contacts between the surfactant layers. An alternative approach is to tether polymer chains to the surfaces by one end which, when swollen by a solvent, then act as molecular 'brushes' that may facilitate sliding. The normal forces between sliding brush-bearing surfaces have been previously investigated, but the lateral forces, which are the most important from the point of view of lubrication, are harder to measure. Here we report the measurement of lateral forces in such a system. We find a striking reduction in the effective friction coefficient mu(sub b) between the surfaces to below our detection limit (mu(sub b) less than 0.001), for contact pressures of around 1 MPa and sliding velocities from zero to 450 nm s(sup -1). We believe that this effect is due to the long-range repulsion, of entropic origin, between the brushes, which acts to keep the surfaces apart while maintaining a relatively fluid layer at the interface between them.
Article
Multi-component micropatterned organosilane monolayers are fabricated on a Si-wafer substrate by stepwise site-specific vacuum ultraviolet (VUV)-ray photodecomposition and chemisorption. The introduction of different organosilane components is confirmed by X-ray photoelectron spectroscopy (XPS). Atomic force microscopic and lateral force microscopic observations reveal that the line-widths of the micropatterned surface correspond to those of the photomask. Contact angle measurement reveals that the micropatterning of the surface functional groups influences the magnitudes of surface free energy. A line pattern with high wetting contrast shows anisotropic water condensing behavior. Also, the patterned surface is used for the site-specific polymerization and site-specific adsorption of microparticles.
Article
The preparation of a wide variety of unique polymer brush structures can be accomplished by “living” free radical polymerization of vinyl monomers from surface-tethered alkoxyamines or from tethered α-halo esters in the presence of (PPh3)2NiBr2. The use of a “living” free radical process permits the molecular weight and polydispersity of the covalently attached polymer chains to be accurately controlled while also allowing the formation of block copolymers by the sequential growth of monomers from the surface. These block and random copolymer brushes have been used to control surface properties.