Sei sulla pagina 1di 5

Communication

Tuning Surface Wettability of


Poly(3-sulfopropyl methacrylate) Brushes by
Cationic Surfactant-Driven Interactions

Markus Dobbelin, Garciela Arias, Iraida Loinaz, Irantzu Llarena,


David Mecerreyes,* Sergio Moya*

In this communication, polyanionic poly(potassium 3-sulfopropyl methacrylate) (PSPM)


brushes were switched from hydrophilic to hydrophobic by exchange of the counter cations.
First, poly(potassium 3-sulfopropyl methacrylate) brushes were grown by means of atom
transfer radical polymerization (ATRP) from thiol monolayers of initiating v-mercaptoundecyl
bromoisobutyrate and mixed monolayers of thiol initiator and 1-undecanothiol (blank thiol)
attached to gold surfaces. The kinetics of the polymerization reaction were followed by means
of the quartz microbalance technique with dissipation (QCM-D). The collapse of PSPM brushes
in the presence of cationic surfactants like quaternary ammonium salts (tetraethylammonium
bromide, hexadecyltrimethylammonium bromide) and imidazolium salts (1-dodecyl-3-
methylimidazolium bromide, 1H, 1H, 2H, 2H-
perfluoro-1-decyl-3-methylimidazolium bromide)
was shown by QCM-D. Water contact angle
measurements proved that the wettability of
the surface could be tuned reversibly from hydro-
philic values (<30 8) to hydrophobic ones (>85 8).

Introduction films have indeed been used to protect surfaces, as barriers


for selective permeability, for the anchorage of active
Polymer thin films provide a wide platform for tuning molecules, as anti-bacterial layers or to adjust wetting
the surface properties of materials, both through their properties.[14] As a particular case of polymer thin films,
chemical and topological characteristics. Polymer thin polyelectrolyte brushes are monolayers of polyelectrolyte
molecules, where the polymer chains are tethered on
M. Dobbelin, I. Loinaz, D. Mecerreyes one side to a surface while the other end of the chain is
New Materials Department, CIDETEC, Centre for Electrochemical free. The dense arrangement of the polymer chains in a
Technologies and CIC NANOGUNE Consolider, Parque Tecnolo- brush provides the system with very interesting physico-
gico de San Sebastian, Paseo Miramon 196, Donostia-San chemical properties. For example, the conformational state
Sebastian 20009, Spain
of the charged tethered chains is highly dependent on
E-mail: dmecerreyes@cidetec.es
the ionic strength and, in some cases, on the pH.[5] The
I. Llarena, S. Moya
CIC BIOMAGUNE, Parque Tecnologico de San Sebastian, Paseo
repulsive coulombic interactions between their intrinsic
Miramon 182 C, Donostia-San Sebastian 20009, Spain charges leads to an extended (stretched) conformation of
E-mail: smoya@cicbiomagune.es the solvated polymer chains. Upon increasing the ionic
G. Arias strength, the repulsive interactions are screened and the
CIQA, Blv. Enrique Reyna, Saltillo, Coahuila, Mexico brush changes conformation into a collapsed state. When

Macromol. Rapid Commun. 2008, 29, 871875


2008 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim DOI: 10.1002/marc.200800071 871
M. Dobbelin, G. Arias, I. Loinaz, I. Llarena, D. Mecerreyes, S. Moya

the collapse takes place, the solvent is largely excluded and Quartz Crystal Microbalance with Dissipation
the brush acquires such conformations that the chain
A quartz crystal microbalance with dissipation allows changes in
density is at a maximum.[69] Recently, it has been shown
both frequency and dissipation of an excited quartz crystal to be
that polycation brushes can be used to change the followed as a function of time. A QCM crystal consists of a thin
wettability of surfaces by simply choosing the proper layer of quartz between two electrodes; the crystal is excited by
anions. It has been reported that brushes of poly[2- applying an RF voltage across the electrodes close to the resonance
(methacryloyloxy)-ethyl-trimethylammonium chloride] frequency. The frequency of oscillation is related to the mass of the
(PMETAC) showed significant changes in their wetting crystal, and additions of mass to the surface of the crystal will
character when exposed to different anions. Perchlorate decrease the frequency. The dissipation of the crystal is measured
(ClO4 ) coordinated PMETAC brushes showed an increased
by recording the response of the freely oscillating crystal that
water contact angle from an initial 35 8 in the uncollapsed has been vibrated at its resonant frequency. The values of the
dissipation are related to the viscoelasticity of the film. The
state up to 79 8 in the collapsed state.[10] In contrast, when
dissipation is defined as DD Edissipated/2Estored. The equipment
PMETAC brushes were coordinated with SO 4 anions, their
has a chamber with an 80 mL volume, which is closed at one side
surface turned to highly hydrophilic.[11] This interesting
by the quartz crystal. The cell can be filled and the fluid exchanged
anion sensitive behavior is completely reversible and it has using the standard Q-Sense flow system. The quartz crystals,
been demonstrated with different anions, such as trifluor- purchased from Q-Sense, had a main resonance frequency of
oborate and amidosulfonates, as well as for other poly- 5 MHz.[1620]
cationic brushes, such as poly[1-ethyl 3-(2-methacryloyloxy
ethyl) imidazolium chloride].[12] Polycation brushes can be Typical Procedure for Synthesis of PSPM Brushes
used to fabricate smart surfaces, whose contact angles Atom transfer radical polymerization of poly(potassium
can be easily varied by exposure of the brush to specific ion 3-sulfopropyl methacrylate) (PSPM) brushes from monolayers of
solutions. initiating thiols was conducted following well established
To the best of our knowledge, unlike polycation brushes, methods described in the literature.[1,21,22] The samples for the
polyanion brushes have not been studied with regard water contact angle measurements were grown on gold coated
to their collapse with specific anions and their capacity pre-annealed glass substrates from arrandee (www.arrandee.com)
to change the wetting properties of surfaces. In this by immersing them in the polymerization solution for several
hours. Samples for the QCM-D were synthesized from quartz
communication, we report the tuning of surface wett-
crystals from Q-Sense. In a typical reaction, 1.24 g of SPMA, 0.75 mL
ability by exposing anionic poly(potassium 3-sulfopropyl
of distilled water and 1.25 mL of DMF were introduced into a
methacrylate) brushes to different cationic surfactants, Schlenk flask and degassed with nitrogen. Aqueous Cu-bpy stock
such as quaternary ammonium and quaternary imidazo- solution was prepared by adding 2 mL of degassed water to
lium molecules. 39.6 mg of CuCl, 53.8 mg of CuCl2 and 312 mg of bpy. 19.5 mg
of 2-EBiB and 0.5 mL of the Cu-bpy stock solution were added to
the SPMA solution for the polymerization.

Experimental Part Typical Procedure for Cation Exchange

Chemicals For the cation exchange reactions, 10  103 M solutions of the


ammonium and imidazolium salts were prepared. The substrates
Potassium 3-sulfopropylmethacrylate (SPMA), 2,20 -bipyridine (bpy), with PSPM brushes were immersed in the polymerization
CuICl, CuIICl2, ethyl 2-bromoisobutyrate (2-EBiB), mercaptounde- solutions for 60 min with stirring and then washed with water.
cane (blank thiol), tetraethylammonium bromide (TEAB), tetra- For the QCM-D measurements, the chamber was filled with the
butylammonium bromide (TBAB), tetrahexylammonium bromide salt solutions at suitable time periods.
(THAB), hexadecyltrimethylammonium bromide (HDTMAB) and
all other reagents were purchased from Aldrich. v-Mercaptoundecyl
bromoisobutyrate (thiol initiator) and 1-dodecyl-3-methylimida-
zolium bromide were synthesized as described elsewhere.[13,14] Results and Discussion
1H,1H,2H,2H-Perfluoro-1-decyl-3- methylimidazolium iodide was
synthesized according to the literature procedure.[15] The synthetic pathway to PSPM brushes and the sub-
sequent cation exchange procedure by a cationic surfac-
tant is shown in Figure 1. In the first step, PSPM brushes
were grown by ATRP from gold surfaces similarly to
Measurements procedures recently reported (Figure 1(a)).[21] The scheme
ATRP and cation exchange reactions were followed using a quartz in Figure 1(b) shows the brush monolayer with K as
crystal microbalance with dissipation (QCM-D) from Q-Sense, the counter ion. In a second step, the K counter ions in the
Gothenburg, Sweden. Water contact angles were measured on a polyanionic brushes were replaced by voluminous cations,
CAM-200 from KSV Instruments Ltd. such as tetraalkylammonium and alkylimidazolium ones

Macromol. Rapid Commun. 2008, 29, 871875


872 2008 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim DOI: 10.1002/marc.200800071
Tuning Surface Wettability of Poly(3-sulfopropyl methacrylate) Brushes . . .

(Figure 1(c)). Both processes were


followed using the quartz crystal
microbalance with dissipation.
Figure 2 shows changes in
frequency and dissipation on a
quartz crystal during the poly-
merization in the chamber. The
quartz crystal was first coated
with a monolayer of initiating
thiols from a 10  103 M stock
solution of v-mercaptoundecyl
bromoisobutyrate. For the poly-
merization, the chamber was
firstly filled with a mixture of
2 parts solvent and 1 part mono-
mer to remove all oxygen present
in the chamber. This stage was
followed by the addition of the
same deoxygenated mixture of
solvents and monomer, but with
catalyst (CuCl, CuCl2), ligand (bpy)
and initiator (2-EBiB). After the
reaction, the chamber was washed
with ethanol and water. Fre-
quency changes are related to
the mass growth on the surfaces
and can be used to assess when
Figure 1. Schematic of: (a) a mixed initiator and blank thiol monolayer; (b) surface initiating the reaction comes to a conclu-
polymerization of PSPM; (c) cation exchange reaction.
sion. Obtained dissipation values
were typical. QCM-D allows com-
parison of the yield of the reac-
tion for different reaction conditions, such as catalyst,
co-catalyzer ratio, density of initiators, solvent character-
istics or temperature.[11,13,2327]
Once the PSPM brushes had been grown, the cation
exchange process and the collapse of the brush was trigged
by exposing it to solutions of different cationic surfactants,
such as tetraethylammonium bromide, tetrabutylammonium
bromide,tetrahexylammoniumbromide,hexadecyltrimethyl-
ammonium bromide, 1-dodecyl-3-methylimidazolium
bromide and 1H,1H,2H,2H-perfluoro-1-decyl-3-methyl-
imidazolium iodide. As an illustrative example shown in
Figure 3, changes in frequency and dissipation for PSPM
brushes exposed to tetraethylammonium bromide can be
seen. For the cation exchange, the chamber of the QCM-D
was filled with 10  103 M tetraethylammonium bromide
(aq). After the addition of the surfactant solution, the
frequency started to increase slowly while the dissipation
went down. Washing with water did not recover the
original frequency and dissipation values. The original
frequency and dissipation values were recovered after
washing the brush with 1 M NaCl (aq) followed by water
Figure 2. Changes in: (a) frequency; (b) dissipation of a quartz (not shown). The observed changes, both in frequency and
crystal during the polymerization of PSPM brushes. dissipation, were consistent with those expected for the

Macromol. Rapid Commun. 2008, 29, 871875


2008 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim www.mrc-journal.de 873
M. Dobbelin, G. Arias, I. Loinaz, I. Llarena, D. Mecerreyes, S. Moya

series of surfactants as measured with the contact angle


technique. Initially the brush was hydrophilic with a water
contact angle of 23 8.
Increasing the length of aliphatic chains in the cationic
surfactant increased the contact angle. The tetraethylam-
monium coordinated brush showed a water contact angle
of 38 8 (not shown). For longer aliphatic chains, the water
contact angles reached values of 40 8 for hexadecyltri-
methylammonium bromide and 56 8 for 1-dodecyl-3-
methylimidazolium bromide. Hydrophobicity increased
even more when we made the brush collapse with a
fluorinated surfactant (1H,1H,2H,2H-perfluoro-1-decyl-
3-methylimidazolium), to around 69 8. The increase in
hydrophobicity was nevertheless not as steep as with
polycation brushes using fluorinated salts.[10] This is
probably due to the characteristics of the sulfonate groups
Figure 3. Changes in: (a) frequency; (b) dissipation of a quartz
crystal during the exposure of a PSPM brush to 10  103 M
whose charges are more diffuse and less easy to screen
tetraethylammonium bromide (TEAB) solution. than the quaternary ammonium groups of PMETAC
brushes. It was then thought that using a brush with
more accessibility for the surfactant molecules to the
brushes during the collapse. The reduction in frequency interior of the brush could be a good strategy to obtain a
can be related to a loss of water from the system, while the brush with contact angles over 80 8. It has been shown by
decrease in dissipation can be understood as the result of Ruhe et al. that the grafting density strongly influences the
conformational changes in the polymer chains during degree of binding of surfactants to a brush and the degree
collapse; the brush passes from being a soft interface with of binding of alkyl ammonium salts to polymethacrylic
high dissipation when the chains are extended to a solid acids increases with decreasing grafting.[28] Therefore, we
material when the chains collapse. In order to remove the synthesized a less dense poly(potassium 3-sulfopropyl
surfactant, the surfactant sulfonate interaction must be methacrylate) brush by diluting the amount of initiator on
weakened using NaCl in a high concentration. the surface. Thus, the initiator molecules were assembled
In some cases, washing with ethanol was necessary on the substrates from a 10  103 M thiol solution bearing
after the treatment with NaCl to fully remove the 5% of initiator, while the remaining percentage of the
surfactant and return it to its original
state. The slow speed of the collapse
process contrasts with that reported
for inorganic salts.[11] We assume that
this is due to the accessibility of the
surfactant in the brush. For the series
of surfactants we employed, collapse
was always observed, but the kinetics
of the process varied significantly
from one molecule to the next. Experi-
ments are ongoing to properly under-
stand the nature of this process.
The collapse, as we have shown,
implies a reduction in the water
content of the brush. The loss of water
in the system and the pairing of the
charges of the pendant groups and
the organic cations, respectively, will
affect surface properties and the
wetting characteristics of the surface.
Figure 4. Changes in wettability corresponding to a gold surface modified with poly(sulfo-
In Figure 4(a) we can observe the propyl methacrylate) brushes with different counter ions. Line (a) is a gold surface initially
changes in the wetting character of modified with 100% v-mercaptoundecyl bromoisobutyrate and (b) is a mixture of 5%
the PSPM brush when treated with a v-mercaptoundecyl bromoisobutyrate/95% blank thiol (non-initiating).

Macromol. Rapid Commun. 2008, 29, 871875


874 2008 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim DOI: 10.1002/marc.200800071
Tuning Surface Wettability of Poly(3-sulfopropyl methacrylate) Brushes . . .

thiols was non-initiating. The initial poly(potassium [1] M. Husseman, E. E. Malmstrom, M. McNamara, M. Mate,
3-sulfopropyl methacrylate) brush showed, in this case, D. Mecerreyes, D. G. Benoit, J. L. Hedrick, P. Mansky, E. Huang,
T. P. Russell, C. J. Hawker, Macromolecules 1999, 32, 1424.
a contact angle of 36 8, which was 13 8 more than the brush
[2] J. C. Tiller, C. J. Liao, K. Lewis, A. M. Klibanov, PNAS 2001, 98,
from the 100% initiator modified surface (Figure 4(b)). 5981.
When we performed the treatment with the surfactants, in [3] J. Lin, S. Y. Qiu, K. Lewis, A. M. Klibanov, Biotechnology
all cases higher water contact angles were observed, and in Progress 2002, 18, 1082.
the case of the fluorinated surfactant the contact angle [4] U. Raviv, S. Glasson, N. Kampf, J. Gohy, R. Jerome, J. Klein,
reached a value of 89 8. Nature 2003, 425, 163.
[5] K. Sakai, E. G. Smith, G. B. Webber, M. Baker, E. J. Wanless,
V. Butun, S. P. Armes, S. Biggs, Langmuir 2006, 22, 8435.
[6] E. B. Zhulina, O. V. Borisov, V. A. Pryamitsyn, T. M. Birshtein,
Conclusion Macromolecules 1991, 24, 140.
[7] R. S. Ross, P. Pincus, Macromolecules 1992, 25, 2177.
We have shown that the surface wettability of polyanionic [8] R. C. Advincula, W. J. Brittain, R. C. Caster, J. Ruhe, Polymer
PSPM brushes can be tuned by cation exchange with ionic Brushes, Wiley-VCH. Weinheim: 1998, p. I/1 ff.
[9] R. Konradi, J. Ruhe, Macromolecules 2004, 37, 6954.
liquid-like surfactants and ammonium surfactants. The
[10] O. Azzaroni, A. A. Brown, W. T. S. Huck, Adv. Mater. 2007, 19,
growth and collapse of the brush was followed with 151.
QCM-D. A characteristic increase in frequency and [11] S. Moya, O. Azzaroni, T. Farhan, V. L. Osborne, W. T. S. Huck,
decrease in dissipation due to a reduction in water content Angew. Chem. Int. Ed. 2005, 44, 4578.
and conformational changes of the brushes were observed [12] B. Yu, F. Zhou, H. Hu, C. W. Wang, W. M. Liu, Electrochim. Acta
2007, 53, 487.
when coordinated with different cations. Interestingly,
[13] D. M. Jones, A. A. Brown, W. T. S. Huck, Langmuir 2002, 18,
higher water contact angles were obtained when applying 1265.
the cation exchange to the less dense brushes. This [14] T. Inoue, B. Dong, L. Q. Zheng, J. Colloid Interface Sci. 2007, 307,
situation was explained by better accessibility of the 578.
surfactant in the brush. By choosing the proper counter [15] R. P. Singh, S. Manandhar, J. M. Shreeve, Tetrahedron Lett.
2002, 43, 9497.
cations, the water contact angle of poly(3-sulfopropyl
[16] S. E. Moya, A. A. Brown, O. Azzaroni, W. T. S. Huck, Macromol.
methacrylate) brushes could be increased reversibly Rapid Commun. 2005, 26, 1117.
from 36 8 up to 89 8. This observation highlights the use [17] J. He, Y. Wu, J. Wu, X. Mao, L. Fu, T. Qian, J. Fang, Ch., Xiong,
of brush density in the design of complex surfaces with J. Xie, H. Ma, Macromolecules 2007, 40, 3090.
tailored properties. [18] F. Hook, M. Rodahl, B. Kasemo, P. Brzezinski, Proc. Natl. Acad.
Sci. USA 1998, 95, 12271.
[19] F. Hook, B. Kasemo, T. Nylander, C. Fant, K. Scott, H. Elwing,
Anal. Chem. 2001, 73, 5796.
[20] G. Zhang, Macromolecules 2004, 37, 6553.
Acknowledgements: S. E. Moya is a Ramon y Cajal fellow and
[21] M. Ramstedt, N. Cheng, O. Azzaroni, D. Mossialos, H. J.
would like to thank this program as well as project (Ma
Mathieu, W. T. S. Huck, Langmuir 2007, 23, 3314.
T2007-60458) from the Spanish Ministry of Science and Education.
[22] K. Matyjaszewski, J. H. Xia, Chem. Rev. 2001, 101, 2921.
The authors wish to thank the Gobierno Vasco and the Diputacion
[23] M. T. Muller, X. P. Yan, S. W. Lee, S. S. Perry, N. D. Spencer,
de Gipuzkoa for financial support through the NANOTRON Etortek
Macromolecules 2005, 38, 5706.
project. They also acknowledge Marie Curie RTN THREADMILL
[24] M. T. Muller, X. P. Yan, S. W. Lee, S. S. Perry, N. D. Spencer,
(MRTN-CT-2006-036040) for financial support.
Macromolecules 2005, 38, 3861.
[25] F. T. Limpoco, R. C. Advincula, S. S. Perry, Langmuir 2007, 23,
12196.
Received: February 1, 2008; Revised: February 27, 2008; Accepted:
February 28, 2008; DOI: 10.1002/marc.200800071 [26] N. Ishida, S. Biggs, Langmuir 2007, 23, 11083.
[27] S. E. Moya, O. Azzaroni, T. Kelby, E. Donath, W. T. S. Huck,
Keywords: atom transfer radical polymerization (ATRP); mono- J. Phys. Chem. 2007, 111, 7034.
layers; polyelectrolytes; surfaces; surfactants [28] R. Konradi, J. Ruhe, Macromolecules 2005, 38, 6140.

Macromol. Rapid Commun. 2008, 29, 871875


2008 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim www.mrc-journal.de 875

Potrebbero piacerti anche