Porsch 2011

Download as pdf or txt
Download as pdf or txt
You are on page 1of 10

View Article Online / Journal Homepage / Table of Contents for this issue

Polymer Dynamic Article Links < C


Chemistry
Cite this: Polym. Chem., 2011, 2, 1114
www.rsc.org/polymers PAPER
Thermo-responsive cellulose-based architectures: tailoring LCST using
poly(ethylene glycol) methacrylates†‡
Christian Porsch, Susanne Hansson, Niklas Nordgren and Eva Malmstr€om*
Received 21st December 2010, Accepted 19th January 2011
DOI: 10.1039/c0py00417k
Published on 16 February 2011. Downloaded on 25/10/2014 02:33:23.

There is a growing interest in designing advanced macromolecular architectures applicable for instance
in drug delivery systems. Employing cellulose in these systems is particularly favorable due to attractive
properties such as biocompatibility and low price. Additionally, thermo-responsive polymers of
poly(ethylene glycol) methacrylates are promising in this field owing to their biocompatibility and non-
toxicity. In the present study, amphiphilic thermo-responsive homo- and copolymers of oligo(ethylene
glycol) methyl ether methacrylate (OEGMA300) and di(ethylene glycol) methyl ether methacrylate
(DEGMA) were synthesized via ARGET ATRP. Both linear copolymers of DEGMA/OEGMA300 as
well as comb architectures with copolymers of DEGMA/OEGMA300 grafted from hydroxypropyl
cellulose were produced. The lower critical solution temperature of the linear copolymers was readily
tailored by altering the monomer feed ratio. The grafting of the thermo-responsive polymers from
hydroxypropyl cellulose resulted in a consistent decrease of the lower critical solution temperature
compared to the linear analogues; however, interestingly the ability to tune the transition temperature
remained. Moreover, the amphiphilic comb architectures formed polymeric micelles with low critical
micelle concentrations. Consequently, these advanced architectures combine the favorable properties
of hydroxypropyl cellulose with the interesting thermo-responsive and stealth properties of
poly(ethylene glycol) methacrylates, and may, therefore, find potential applications in biomedicine.

Introduction microcrystalline cellulose (MCC),7 microfibrillated cellulose


(MFC),8 and hydroxypropyl cellulose (HPC).9 HPC is a linear
Cellulose is the most abundant polysaccharide, and an important cellulose derivative consisting of a glucose-unit backbone, which
sustainable raw material.1 During the last decades, cellulose and has been propoxylated (see Fig. 1). Owing to the decreased
its derivatives have attracted significant interest in many fields, as polarity, HPC is fairly soluble, which in combination with the
a result of increased environmental concerns. Additionally, superb inherent properties of cellulose mentioned above makes it
cellulose exhibits many attractive properties, such as renew- of great interest in biomedical applications. Recently,
ability, biocompatibility, and good availability in nature. Malmstr€ om et al.10,11 reported how HPC can be grafted with
However, due to the hydrophilic nature of cellulose, modification PCL-block-PAA arms, obtaining nanocontainers suitable for
is often required to tailor the final properties for specific appli- guest molecule encapsulation.
cations. This can be accomplished by grafting polymers onto the
cellulose surface, thus enhancing existing properties or adding
new desired ones. Our group has extensively been grafting
various polymers from a number of different cellulose substrates,
for example: filter paper,2–5 cellulose microspheres,6

KTH Royal Institute of Technology, School of Chemical Engineering,


Dept. of Fibre and Polymer Technology, SE-100 44 Stockholm, Sweden.
E-mail: [email protected]; Fax: +46 8 790 8283; Tel: +46 8 790 7225
† This work was supported by AB Willhelm Beckers Jubileumsfond, the
Swedish Research Council, and the Swedish Foundation for Strategic
Research (via Biomime, the Swedish Center for Biomimetic Fiber
Engineering).
‡ Electronic supplementary information (ESI) available: Calculations of
the molar substitution of propoxy groups and the total degree of
substitution of the initiators on HPC, NMR spectra, calculation of
copolymer compositions, kinetic plots, CMC and DLS measurements. Fig. 1 Repeating unit of hydroxypropyl cellulose (HPC) and the
See DOI: 10.1039/c0py00417k monomers used in the present study.

1114 | Polym. Chem., 2011, 2, 1114–1123 This journal is ª The Royal Society of Chemistry 2011
View Article Online

The ability to tailor material properties by employing monomer. However, none of the available PEGMAs shows an
advanced macromolecular architectures has gained increased LCST close to the physiological temperature, which has been
attention in biomedicine.12 Earlier, interesting architectures like obstructing their possibility to compete with PNIPAM in
graft copolymers suffered from poor control during polymeri- biomedical applications. In 2006, Lutz et al.29,32 elegantly showed
zation of the grafts, which obstructed their use in many appli- that the LCST of PEGMAs could be finely adjusted by random
cations. However, the progress of controlled radical copolymerization of di(ethylene glycol) methyl ether methacry-
polymerization techniques has enabled synthesis of products late (DEGMA) and oligo(ethylene glycol) methyl ether metha-
with well-defined architectures, pre-determined molecular crylate (OEGMA475) via ATRP; thus, widening the applicability
weights, and low polydispersity indices.13 The most common of these polymers. Later, Schubert et al.31 showed that the
controlled radical polymerization technique, atom transfer tailoring of the LCST also can be achieved by using RAFT
radical polymerization (ATRP), was introduced in 1995 inde- polymerization. These progresses in tailoring the responsive
pendently by the groups of Matyjaszewski14 and Sawamoto,15 behavior of oligo(ethylene glycol) methacrylate-based polymers,
and has proved to be a versatile method for preparing cellulose- combined with their biocompatibility and desirable stealth
based graft copolymers.16–19 In ATRP, a catalyst complex properties,27 have resulted in extensive research during the last
ensures a low concentration of radicals, thus suppressing irre- years.28,33
Published on 16 February 2011. Downloaded on 25/10/2014 02:33:23.

versible termination reactions. However, due to ATRP’s sensi- In the present work, ARGET ATRP is employed as
tivity to air and need for high amounts of catalyst, a straightforward technique for preparation of thermo-respon-
Matyjaszewski and Pintauer20 developed an improved technique: sive copolymers of OEGMA300 and DEGMA with tunable
activators regenerated by electron transfer (ARGET) ATRP. In LCSTs. Furthermore, comb copolymers of HPC and the thermo-
contrast to conventional ATRP, a reducing agent is employed in responsive copolymers are synthesized, employing surface-initi-
ARGET ATRP, which significantly reduces the amount of ated ARGET ATRP. These interesting architectures are an
catalyst required to obtain a controlled polymerization. The additional step towards the understanding of the fascinating
technique also allows the polymerization to be conducted in the solution properties, and potential applicabilities, of advanced
presence of small amounts of air. Consequently, less effort can be poly(ethylene glycol) methacrylate structures.
devoted to deoxygenation, which simplifies the reaction proce-
dure. Experimental section
Cellulose-based, amphiphilic graft copolymers are investigated
as components in polymeric micelles, which can be used as Materials
carriers in drug delivery systems.16,21,22 Above the critical micelle Hydroxypropyl cellulose (HPC, Mw ¼ 80 000 g mol1, Mn ¼
concentration (CMC) these structures may associate into a core– 10 000 g mol1 according to manufacturer (Aldrich); Mw ¼
shell morphology, where the hydrophobic core enables solubili- 74 000 g mol1, Mn ¼ 34 000 g mol1 was obtained using SEC
zation of a hydrophobic drug, and the hydrophilic shell ensures with THF as a mobile phase; molar substitution of propoxy
micelle solubility in the water solution. Additional benefits of groups (MS) was determined to be 2.9 using 1H-NMR (Fig. S1,
polymeric micelles as drug delivery systems are for example long ESI‡)). 4-(Dimethylamino)pyridine (DMAP, 99%), ethyl
circulation times, improved biodistribution, and enhanced 2-bromoisobutyrate (EBiB, 98%), 2-bromisobutyryl bromide
possibility to cross physiological barriers. (BiB, 98%), N,N,N0 ,N00 ,N00 -pentamethyldiethylenetriamine
Thermo-responsive polymers have the ability to go through (PMDETA, 99%), 2,20 -bipyridyl (Bipy, >99%), ascorbic acid
a phase transition in aqueous solution, switching from a hydro- (AsAc, >99%) and a-bromoisobutyric acid were purchased from
philic to hydrophobic character upon heating. This phase tran- Aldrich. Copper(II) bromide (Cu(II)Br2, 99%) and N, N0 -dicy-
sition, referred to as the lower critical solution temperature clohexylcarbodiimide (DCC, 99%) were purchased from Acros
(LCST), is present due to inter- and intramolecular hydrogen Organics. The monomers oligo(ethylene glycol) methyl ether
bonds between water molecules and the polymer chains. methacrylate (OEGMA300, Mw 300 g mol1, Aldrich) and
Recently, Hu et al. demonstrated that HPC exhibits thermo- di(ethylene glycol) methyl ether methacrylate (DEGMA, Mw
responsive properties with a phase transition temperature of 188 g mol1, Aldrich) were prior to use activated by passing it
41  C.23 Further, the most widely studied thermo-responsive through a column of neutral aluminium oxide. Phosphate buff-
polymer is poly(N-isopropylacrylamide) (PNIPAM), mainly due ered saline (PBS) was prepared according to the recipe in the ESI
to the fact that its LCST (32  C) is close to the physiological S2‡.
temperature.24 However, recently published results by Vihola
et al.25 suggest that acrylamide-based polymers show cytotoxicity
Instrumentation
at the physiological temperature, which is believed to be related
to the presence of unreacted monomeric residues. A promising Nuclear magnetic resonance. 1H- and 13C-NMR spectra were
group of thermo-responsive polymers is thus poly(ethylene recorded on a Bruker Avance AM 400 NMR instrument using
glycol) methacrylates (PEGMAs), which have shown to be non- CDCl3 and MeOD as solvents. The residual solvent peak was
toxic25,26 and biocompatible.27–30 Furthermore, the LCST of used as an internal standard.
these polymers can be tailored by changing the length of the
ethylene glycol segment in the side chain.31 Longer ethylene Fourier transform infrared spectroscopy. The absorption
glycol segments result in a more hydrophilic nature of the spectra were collected by employing a Perkin-Elmer spectrum
polymer, thus increasing the LCST. As a result, the LCST can be 2000 FT-IR, equipped with a MKII Golden Gate, single reflec-
tuned in the range of 26–90  C by the appropriate choice of tion ATR system from Specec Ltd. The ATR crystal was a MKII

This journal is ª The Royal Society of Chemistry 2011 Polym. Chem., 2011, 2, 1114–1123 | 1115
View Article Online

heated diamond 45 ATR top plate. Sixteen scans were recorded terminated by removal of the flask from the oil bath, exposure to
for each spectrum. air, and dilution with THF. Thereafter, the reaction mixture was
allowed to pass through a short column of neutral aluminium
Size exclusion chromatography. SEC using THF (1.0 ml min1) oxide to remove the copper complex. Subsequently, the solution
as the mobile phase was performed at 35  C using a Viscotek was concentrated through rotary evaporation and precipitated in
TDA model 301 equipped with two GMHHR-M columns with diethyl ether cooled with dry ice. The solvent was decanted off
TSK gel (mixed bed, Mw resolving range: 300–100 000 Da) from and the sticky product was dried under vacuum. 1H-NMR
Tosoh Biosep, a VE 5200 GPC autosampler, a VE 1121 GPC (CDCl3): d 0.91 (br), 1.06 (br), 1.61 (s), 1.84 (br), 1.94 (br), 3.42
solvent pump, and a VE 5710 GPC degasser (all from Viscotek (s), 3.59 (br), 3.66–3.69 (br), 4.13 (br) ppm. Typical yields of
Corp.). A calibration method was created using broad and approximately 25% were achieved.
narrow linear polystyrene standards. Corrections for the flow
rate fluctuations were made using toluene as an internal stan-
Synthesis of 2-bromoisobutyryl anhydride
dard. Viscotek OmniSEC version 4.0 software was used to
process the data. The procedure used was adopted from Malmstr€ om et al.10 2-
Bromoisobutyric acid (10.0 g, 59.9 mmol) was dissolved in
Published on 16 February 2011. Downloaded on 25/10/2014 02:33:23.

Dynamic light scattering. The polymer solutions were analyzed CH2Cl2 (DCM; 50 ml). N,N0 -Dicyclohexylcarbodiimide (DCC;
with a Malvern Zetasizer NanoZS using polymer concentrations 6.80 g, 32.9 mmol) was dissolved in DCM (25 ml) and added to
of 100 and 500 mg l1 in deionized water. At changed tempera- the solution. A milky white mixture was formed and was stirred
ture, the samples were allowed to equilibrate for 2 minutes prior at ambient temperature overnight. The precipitate was filtered
to the measurement. off, the filtrate was concentrated through rotary evaporation,
and precipitated in n-heptane cooled with dry ice. The product
Fluorescence spectroscopy. Fluorescence Spectroscopy was was filtered, washed with cold n-heptane and finally dried under
performed on a Varian Cary Eclipse collecting emission spectra vacuum. The purity of the white powder product was confirmed
using an excitation wavelength of 332 nm. Measurements were by NMR and achieved in a yield of 55%. 1H-NMR (CDCl3):
performed in PBS buffer solutions at temperatures below LCST d 2.03 ppm. 13C-NMR (CDCl3): d 30.12, 55.10, 165.84 ppm.
of the polymers (HPC-g-PDEGMA 15  C, HPC-g-POEGMA
and HPC-g-P(OEGMA-co-DEGMA) 25  C). Pyrene was loaded
Synthesis of hydroxypropyl cellulose macroinitiators (HPC-Is)
into the polymer micelles by physical entrapment. Different
polymer concentrations were prepared by dissolving the correct Two separate routes were employed to synthesize the HPC
amount of polymer in PBS buffer during gentle stirring. 25 ml of macroinitiator (Scheme 1, top). In one route 2-bromoisobutyryl
the pyrene solution (0.8 mM) was added to a vial and the acetone bromide was utilized to achieve the pendant a-bromoesters, and
was evaporated for 3 h in rt. The polymer solution was then in the other 2-bromoisobutyryl anhydride was employed.
added and allowed to equilibrate overnight prior to the
measurements.
Synthesis of HPC-I using 2-bromoisobutyryl anhydride (HPC-
I1.4)
Surface tension measurements. Surface tension measurements
were performed on a Thermo Cahn (Radian series 300) Micro- The synthesis was conducted as described earlier in our group,10
balance, using the Wilhelmy plate method. Prior to every and a light brown polymer was obtained in 85% yield. 1H-NMR
measurement the platinum plate was rinsed with EtOH and (CDCl3): d 1.16 (br), 1.28 (br), 1.94 (s), 2.70–4.20 (br), 3.60 (value
MilliQ water, and subsequently flamed. Measurements were of the sharpest peak inside broad peak), 4.35 (br), 5.05 (br) ppm.
13
performed in PBS buffer solutions at temperatures below LCST C-NMR (CDCl3): d 16.39, 17.49, 30.75, 34.80, 56.14, 71.39,
of the polymers (HPC-g-PDEGMA 15  C, HPC-g-POEGMA 73.44, 75.05, 99.99, 170.75–171.00 ppm.
and HPC-g-P(OEGMA-co-DEGMA) 25  C).
Synthesis of HPC-I using 2-bromoisobutyryl bromide (HPC0.6)
General procedure for polymerization of linear polymers
The synthesis procedure was adopted from Xu et al.18 Hydroxy-
A typical polymerization of the linear polymers was conducted as propyl cellulose (HPC; 0.5 g) was completely dissolved in
described below, and the comonomer composition was altered to anhydrous DCM (25 ml), and the flask subsequently put on
achieve different copolymers. The targeted monomer composi- a water/ice-bath (0  C). 2-Bromoisobutyryl bromide (BiB; 0.81
tion (20 mmol) was added to a 25 ml round bottom flask mmol) was dissolved in 5 ml anhydrous DCM, and cooled in the
equipped with a magnetic stirrer. Anisole (50 wt%), PMDETA water/ice-bath. Thereafter, the solution was slowly added to the
(8.7 mg, 50 mmol) and AsAc (8.7 mg, 50 mmol) were added, and HPC solution over a time period of ten minutes, the water/ice-
allowed to dissolve while stirring. The reaction mixture was put bath was removed, and the reaction was allowed to proceed for
on a water/ice bath (0  C), EBiB (7.4 ml, 50 mmol) was added, and 150 min at room temperature. The reaction was ended by
5 minutes of purging with argon was conducted. Subsequently, precipitation in 300 ml diethyl ether cooled with dry ice. The
Cu(II)Br2 (1.1 mg, 5 mmol) was added followed by additional polymer was filtered off and washed with cold diethyl ether. The
15 minutes of purging with argon. The polymerization was per- pure product was dried under vacuum overnight, and a yield of
formed at 40  C and NMR aliquots were withdrawn under argon 40% was achieved. 1H-NMR (MeOD): d 1.16 (br), 1.93 (s), 3.00–
flow to monitor the reaction (Fig. S12, ESI‡). The reaction was 4.10 (br), 4.50 (br), 4.99 (s) ppm. 13C-NMR (MeOD): d 16.18,

1116 | Polym. Chem., 2011, 2, 1114–1123 This journal is ª The Royal Society of Chemistry 2011
Published on 16 February 2011. Downloaded on 25/10/2014 02:33:23. View Article Online

Scheme 1 Schematic illustration of the comb polymers synthesized in this study.

This journal is ª The Royal Society of Chemistry 2011 Polym. Chem., 2011, 2, 1114–1123 | 1117
View Article Online

18.47, 29.83, 55.46, 66.04, 66.39, 75.01, 76.61, 78.91, 83.00, polymerizations with polydispersity indices between 1.1 and 1.6
102.33, 172.14 ppm. (Table 1). Lower molecular weights were received from SEC than
expected theoretical values calculated from 1H-NMR. This is
most likely an effect of the brush-like structure of POEGMA and
General procedure for polymerization from the macroinitiator PDEGMA, where SEC underestimates their molecular weight
(HPC-I) since calibration is performed with linear polystyrene standards.
HPC-I1.4 or HPC-I0.6 (50 mg, 0.08 mmol and 0.04 mmol of Br, In addition, further attempts to determine the molecular weights
respectively) was added to a round bottom flask, and completely and molecular weight distributions of the polymers were con-
dissolved in anisole or methanol (80 wt%), respectively. Bipy ducted using MALDI-TOF (see Fig. S3, ESI‡). However,
(15.6 mg, 0.10 mmol) and AsAc (17.3 mg, 0.10 mmol) were although several different matrices were employed, all attempts
added, and allowed to dissolve during gentle stirring. The flask failed.
was put in a water/ice bath, and the desired monomer mixture of
OEGMA300 and DEGMA (34 mmol in total) was added. The
solution was purged with argon for 5 minutes, Cu(II)Br2 (2.2 mg, Thermo-responsive behavior of linear polymers
10 mmol) was added, and additionally 15 minutes of purging was
Published on 16 February 2011. Downloaded on 25/10/2014 02:33:23.

performed. The reactions were carried out in a 40  C tempered oil The solution properties of the linear polymers in water were
bath until the viscosity of the reaction mixture had increased studied with dynamic light scattering (DLS), using polymer
significantly. The reaction was terminated by withdrawal from concentrations of 100 mg l1. As expected, all synthesized poly-
the oil bath, air exposure, and the solution was diluted with THF. mers showed water solubility at room temperature. Below LCST,
The product was purified through 48 h of dialysis (MWCO 6– all polymers were present as associated structures with a dia-
8000) with THF that was changed after 2, 6 and 36 h. The meter of 4–12 nm. However, at LCST they displayed a clear
product was concentrated through rotary evaporation, dried phase transition upon heating, going from soluble to insoluble in
under vacuum and typical yields of approximately 32% were the aqueous solution. The phase transition at LCST occurs when
obtained. 1H-NMR (CDCl3): d 0.91 (br), 1.06 (br), 1.45 (s), 1.66 hydrogen bonds between water and the ethylene glycol segments
(s), 1.83 (br), 1.92 (br), 3.42 (s), 3.59 (s), 3.65–3.71 (br), 4.13 (br) in the polymer side chains are broken. The more favorable
ppm. 13C-NMR (CDCl3): d 17.08, 18.85, 25.59, 30.29, 44.78, polymer–polymer interaction then results in a structural collapse
53.44, 59.01, 63.78, 67.95, 70.20, 71.91, 176.16–177.04 ppm. and subsequently, a more hydrophobic polymer chain. Conse-
quently, the collapsed chains start to aggregate, thus going from
4–12 nm to, in most cases, several hundred nanometres. As
Results and discussion expected, further increase of the temperature above LCST
resulted in additional increase in size due to further aggregation
ARGET ATRP of linear homo- and copolymers
(see Fig. S3, ESI‡).
In the first part of this work, ARGET ATRP was employed to The obtained LCSTs, reported in Table 1, correlate well with
achieve controlled polymerizations of homo- and copolymers of earlier established results for these polymers synthesized by other
OEGMA300 and DEGMA. It was found that ARGET ATRP controlled polymerization techniques.31,34 As expected, the
provides a simple and attractive polymerization of polar, func- PDEGMA homopolymer showed the lowest value of LCST at
tional vinyl monomers as OEGMA300 and DEGMA. The 27  C. Moreover, as shown recently by Lutz et al.,34 the transi-
controlled polymerizations of the monomers into linear polymers tion temperatures for P(OEGMA-co-DEGMA) copolymers
were easily conducted by mixing the desired monomer compo- increase with increasing molar fraction of OEGMA300 in the
sition, anisole, PMDETA, AsAc, and EBiB, followed by 5 copolymer (Fig. 2; -). Consequently, ARGET ATRP proved to
minutes degassing. The polymerization was initiated by the enable a tunable responsive behavior of PEGMAs comparable to
addition of CuBr2, 15 min degassing and temperature set to other controlled polymerization techniques. However, note-
40  C. Copolymers with different compositions of OEGMA300 worthy is that the obtained LCST for the POEGMA homo-
and DEGMA were prepared by aiming at different initial polymer (73  C) is about 8–10  C higher than earlier reported
monomer feeds, and polymers of OEGMA300, DEGMA and results.34,35 This is believed to be an effect of the lower molecular
OEGMA300-co-DEGMA were readily synthesized. The achieved weight of this specific polymer sample. Results demonstrated by
products were confirmed with 1H-NMR spectroscopy, and their Lutz et al.29 and Chu et al.36 indicate that shorter ethylene glycol-
characteristics are reported in Table 1. The copolymer compo- based polymers may result in increased transition temperatures
sitions were calculated by comparing the overall integration of in comparison to longer ones.
the methoxy protons with the overall integration of the ethylene As stated above, the achieved results confirm that ARGET
glycol protons (see Fig. S2, ESI‡). A Good agreement was found ATRP is a viable technique for producing tunable responsive
between feed amounts and final composition of OEGMA and PEGMAs, in similarity with other controlled radical polymeri-
DEGMA in the copolymers. This confirms that synthesizing zation techniques reported by others. However, as discussed
P(OEGMA-co-DEGMA) copolymers employing ARGET earlier, ARGET ATRP provides advantages compared to ordi-
ATRP under reported conditions enables the production of pre- nary ATRP, such as more simple reaction conditions, and
determined polymer compositions. improved environmental conditions due to the reduced amount
Results from size exclusion chromatography (SEC) in THF, of catalyst used. To the authors’ best knowledge, this is the first
calibrated with linear polystyrene standards with toluene as the time tunable LCSTs of P(OEGMA-co-DEGMA) have been
internal standard, suggest relatively good control of the reported using ARGET ATRP.

1118 | Polym. Chem., 2011, 2, 1114–1123 This journal is ª The Royal Society of Chemistry 2011
View Article Online

Table 1 Characteristics of homo- and copolymers of OEGMA300 and DEGMA polymerized via ARGET ATRP

fOEGMA (mol%)

Polymer Feed Finalb Conv.b (%) Mn SECc/g mol1 Mn NMRb/g mol1 PDIc Rhd/nm LCSTe/ C

PDEGMA 0 0 45 30 400 50 800 1.2 7 27


P(OEGMA15-co-DEGMA85)a 13 15 46 25 800 36 800 1.2 6 33
P(OEGMA19-co-DEGMA81)a 19 19 51 33 000 41 500 1.3 6 36
P(OEGMA27-co-DEGMA73)a 24 27 63 53 400 53 000 1.6 12 38
P(OEGMA37-co-DEGMA63)a 36 37 75 58 100 66 700 1.3 11 42
P(OEGMA54-co-DEGMA46)a 50 54 18 21 200 17 600 1.1 6 55
P(OEGMA73-co-DEGMA27)a 70 73 41 30 000 43 700 1.3 9 59
POEGMA 100 100 13 14 100 15 600 1.1 4 73
a
P(OEGMAx-co-DEGMAy) where x ¼ mol% of OEGMA and y ¼ mol% of DEGMA in the final polymer. b Determined by 1H-NMR. c Assessed by
THF SEC (polystyrene calibration). d Assessed by DLS measurements (conc. 100 mg l1) at a temperature 1  C below LCST of the polymer. A
characteristic size distribution from DLS is shown in Fig. S13, ESI‡. e Assessed by DLS measurements (conc. 100 mg l1).
Published on 16 February 2011. Downloaded on 25/10/2014 02:33:23.

Synthesis of HPC-macroinitiators Consequently, DMF facilitates the esterification better than


DCM. The lower amount of initiator in HPC-I0.6 was further
Two different macroinitiators with different initiator densities
confirmed using ICP-SMS analysis, which determined the
were synthesized employing routes adopted from Malmstr€ om
bromine content in HPC-I1.4 to be 20.1 wt% and in HPC-I0.6 to
et al.10 and Xu et al.,18 respectively (Scheme 1, top). The
be 11.4 wt%. This was used to calculate the total degree of
successful formation of the initiating moieties was confirmed by
substitution (DS) of the macroinitiators (see the ESI S1‡), and
FT-IR and 1H-NMR spectroscopies, as well as ICP-SMS anal-
1.41 and 0.59 of the 3 available hydroxyl groups of each glucose
ysis.37 In the 1H-NMR spectra of HPC-I1.4 (Fig. S8, ESI‡) and
unit were converted for HPC-I1.4 and HPC-I0.6, respectively.
HPC-I0.6 (Fig. S9, ESI‡), the appearance of a singlet at 1.94 ppm
This states that two macroinitiators with different initiator
originates from the new methyl protons originating from the
densities were achieved, thus enabling synthesis of comb
initiating moiety, thus corroborating successful esterification.
copolymers with different grafting densities.
FT-IR proved to be a very powerful tool to confirm successful
reactions, and spectra of native HPC (a), HPC-I0.6, and HPC-I1.4
are presented in Fig. 3. The reduction of the peak at 3411 cm1, ARGET ATRP from the macroinitiators
corresponding to the available hydroxyl groups of HPC, shows
By employing ARGET ATRP, thermo-responsive polymers of
that both synthetic routes resulted in successful functionaliza-
OEGMA300, DEGMA and OEGMA300-co-DEGMA were
tion. Additionally, the appearance of a peak related to the
grafted from HPC, forming comb polymer architectures. The
carbonyl stretching of the initiator at 1729 cm1 further confirms
HPC-Is were, prior to the reaction, allowed to dissolve
a successful reaction. The spectrum for HPC-I0.6 shows a weaker
completely in the solvent to prevent uneven and uncontrolled
carbonyl signal than HPC-I1.4, as well as more unreacted
grafting. Grafting from the HPC-I1.4 macroinitiator was, as for
hydroxyl groups, thus indicating that the route employing 2-
the linear polymers, conducted in anisole. However, due to the
bromoisobutyric acid is not as effective, and provides a macro-
lower amount of ATRP-initiators on HPC-I0.6, the more polar
initiator with a lower density of initiating moieties. This result is
character made it insoluble in anisole. Instead, methanol was
assumed to be an effect of the solvent, rather than the lower
used as solvent in these polymerizations. All reactions were
reactivity of the bromo acid compared to the anhydride. DMF is
performed in highly diluted systems. Due to the high number of
a more efficient solvent than DCM, thus suggesting that the
initiating sites, dilution is important to keep the local concen-
hydroxyl groups are more available for substitution.
tration of radicals low.10 By doing so, undesired termination and
crosslinking reactions are avoided. In addition, due to the

Fig. 2 LCST as a function of OEGMA content in P(OEGMA-co-


DEGMA) polymers (-) and HPC-g-P(OEGMA-co-DEGMA) poly- Fig. 3 FT-IR of native HPC (a), and the two macroinitiators HPC-I0.6
mers (O). (b) and HPC-I1.4 (c).

This journal is ª The Royal Society of Chemistry 2011 Polym. Chem., 2011, 2, 1114–1123 | 1119
View Article Online

diluting effect of the monomer, the reactions were stopped at low


conversions (typically 20–25%), since higher monomer conver-
sions increase the probability of radical coupling reactions.
Characteristics of the synthesized comb copolymers are
reported in Table 2. The polymerizations performed from the
HPC-I0.6 macroinitiator in methanol proceeded in a repeatable
manner, thus forming HPC-based comb copolymers in 4 hours.
However, the polymerizations using the HPC-I1.4 macroinitiator
were less reproducible and showed tendencies of undesired
intermolecular coupling and gelation, even at much diluted
systems. This corroborates that the higher concentration of
radicals, due to the significantly higher density of initiators of
HPC-I1.4, increases the probability of radical coupling reactions.
Fig. 4 FT-IR of HPC-I0.6 (a) and HPC0.6-g-PDEGMA (b).
As an effect, HPC-I1.4 was successfully grafted only with
PDEGMA, which most likely is due to the shorter side chains of
Published on 16 February 2011. Downloaded on 25/10/2014 02:33:23.

PDEGMA compared to both POEGMA and P(OEGMA-co- the phase transition temperature compared to their linear
DEGMA). Although only one copolymer was synthesized from analogues. DLS measurements were performed in aqueous
HPC-I1.4, the more substituted HPC backbone is interesting in solutions, and all comb polymers were water soluble. For the
order to understand how the architectural difference affects the comb copolymers, polymer concentrations of 500 mg l1, instead
characteristics of the polymer. of 100 mg l1 as for the linear polymers, were required for DLS
Successful grafting from the macroinitiators was confirmed measurements. As a consequence, to enable accurate compari-
with 1H-NMR and FT-IR spectroscopies. In the 1H-NMR sons, DLS measurements of linear PDEGMA, P(OEGMA54-co-
spectrum of HPC0.6-g-PDEGMA (see the ESI S11‡), peaks from DEGMA46) and POEGMA in Fig. 5 were performed at a 500 mg
PDEGMA can easily be observed, confirming successful grafting l1 concentration.
of the polymers onto the cellulose backbone. However, due to the In the fully soluble state, i.e. below the LCST of the grafts, the
very small amount of HPC backbone compared to polymer comb polymers formed self-assemblies with sizes in the range 20–
grafts, as well as a concealing effect, no peaks originating from 40 nm. This correlates well with earlier reported results for
HPC could be detected. This has also been reported earlier.10,38 similar structures,21,22 and shows that the comb polymers form
Grafting was further confirmed by FT-IR analysis (Fig. 4). The considerably larger micellar assemblies than the linear polymers.
shift of the peak corresponding to the carbonyl stretching from Moreover, at the LCST the comb polymers show the same
1732 cm1 to 1725 cm1, and its somewhat broadened appearance distinct and instant increase in size as the linear polymers,
in the HPC0.6-g-PDEGMA spectrum, compared to the spectrum resulting in a collapse and the formation of micellar aggregates
of the HPC-I0.6 macroinitiator, is most certainly due to the (Fig. 5). However, the observed LCSTs were all lower than for
presence of new ester moieties in the new material. Furthermore, the linear analogues (Table 2). This decrease in LCST has
the clear change of appearance in the region around 2900 cm1 is previously been observed for similar systems,17 and is most likely
a consequence of the dominance of methylenes in the ethylene due to the increased hydrophobic character of the comb polymer
glycol segments of HPC0.6-g-PDEGMA. structures. Additionally, the restricted mobility of the polymer
chains when grafted to the cellulose backbone is believed to
further increase this effect. However, it is noteworthy to consider
Thermo-responsive behavior of comb polymers
the slightly lower OEGMA300 content in HPC0.6-g-
The solution properties of the comb copolymers were determined P(OEGMA52-co-DEGMA48) compared to P(OEGMA54-co-
by DLS, fluorescence spectroscopy and surface tension DEGMA46), which surely contributes to the lowered LCST in
measurements, and are reported in Table 2. DLS was employed this case. In addition, the significantly higher molecular weights
to obtain information about the polymers’ response to changes in of the comb polymers may contribute to the lowered LCSTs. The
temperature, and to understand how the new architectures affect fact that the lowering of the LCST for HPC0.6-g-PDEGMA is

Table 2 Characteristics of HPC-based comb copolymers synthesized via surface-initiated ARGET ATRP

CMC/mg ml1

Polymer fOEGMAa (mol%) Conv.a (%) Rhb/nm LCSTc/ C Fluorescenced Surface tensione

HPC1.4-g-PDEGMA 0 20 40 22 0.12 11.3


HPC0.6-g-PDEGMA 0 21 20 23 0.65 10.5
HPC0.6-g-P(OEGMA52-co-DEGMA48) 52 28 30 45 0.55 11.9
HPC0.6-g-POEGMA 100 35 30 62
a
Final OEGMA content determined by 1H-NMR. b Assessed by DLS measurements (conc. 500 mg l1) at a temperature 1  C below LCST of the
polymer. A characteristic size distribution from DLS is shown in Fig. S14, ESI‡. Since DLS measurements are performed on carbohydrates the size
distributions are somewhat broad. Therefore, the reported values are rounded off to tenth nanometres. c Assessed by DLS measurements (conc. 500
mg l1). d Determined by the fluorescent probe technique. e Determined by surface tension measurements.

1120 | Polym. Chem., 2011, 2, 1114–1123 This journal is ª The Royal Society of Chemistry 2011
View Article Online

above the CMC since the amphiphilic molecules will then end up
in micelles.
Results from fluorescent spectroscopy and surface tension
measurements show that HPC1.4-g-PDEGMA, HPC0.6-g-
PDEGMA and HPC0.6-g-P(OEGMA52-co-DEGMA48) form
micelles above the CMC, as reported in Table 2. Fig. 6 shows the
change in fluorescence characteristics of pyrene as a function of
HPC1.4-g-PDEGMA concentration. The observed decrease in
intensity ratio (I1/I3) originates from the entrapment of pyrene in
the hydrophobic core of the formed polymeric micelles. The
obtained CMCs from the fluorescent probe technique are in the
range 0.12–0.65 mg ml1. This is in good agreement with reported
values for similar architectures,21 and indicates that the synthe-
sized materials self-assemble into micelles at low concentrations,
Fig. 5 Size as a function of temperature for HPC-based comb polymers which is essential for possible use as drug carriers. The difference
Published on 16 February 2011. Downloaded on 25/10/2014 02:33:23.

and their linear analogues: PDEGMA (-), HPC1.4-g-PDEGMA (+), in grafting densities of HPC1.4-g-PDEGMA and HPC0.6-g-
HPC0.6-g-PDEGMA (,), P(OEGMA54-co-DEGMA46) (:), HPC0.6-g- PDEGMA resulted in an increase of CMC from 0.12 mg ml1 to
P(OEGMA52-co-DEGMA48) (O), POEGMA (C), and HPC0.6-g- 0.65 mg ml1, thus demonstrating that the micelle formation is
POEGMA (B). Insets: (a) polymer sample below LCST in aqueous affected by the polymer architecture. The longer grafts in
solution and (b) polymer sample above LCST in aqueous solution. HPC0.6-g-PDEGMA seem to facilitate micelle formation more
than the shorter and more dense grafts of HPC1.4-g-PDEGMA.
less pronounced, in relation to its linear analogue, than observed A recent study44 from our group suggests that the architecture, as
in the case of both HPC0.6-g-P(OEGMA52-co-DEGMA48) and well as the molecular weight, clearly affects the critical micelle
HPC0.6-g-POEGMA, further indicates the relevance of the concentration of advanced amphiphilic architectures.
molecular weight. In Fig. 7, surface tension is shown as a function of HPC1.4-g-
As observed when comparing HPC0.6-g-PDEGMA with the PDEGMA concentration. An increased concentration of the
more dense HPC1.4-g-PDEGMA, the increase in grafting density comb polymer lowers the surface tension of the solution,
has a rather small influence on the transition temperature, corroborating surface active properties. The stabilization of the
lowering the LCST by only 1  C. However, this indicates that the surface tension at increased concentration verifies the formation
degree of substitution influences the solution properties for these of micelles, i.e. CMC is reached. The CMCs determined by
kind of structures. The effects of substitution on the aggregation surface tension measurements, ranging from 10.5–11.9 mg ml1,
behavior for ungrafted cellulose structures have recently been are consistently higher compared to the values obtained using
demonstrated by Bodvik et al.39 In addition, a striking result is fluorescence spectroscopy. This difference between the methods
that although the LCST is generally lowered for the HPC-grafted has been observed earlier, for both low42 and high41 molecular
polymers, the ability to tune the transition temperature with weight surfactants. However, important to bear in mind is that
OEGMA content remains (see Fig. 2; O). This suggests that the the polymers discussed in this study exhibit very elaborate
attractive behavior of PEGMAs favourably can be utilized in architectures, resulting in a more complex association behavior
advanced macromolecular designs in the same way as for the compared to ideal low molecular weight surfactant systems.
linear polymers. Furthermore, factors such as polydispersity and physical
restraints of the architecture most certainly affect their assembly.
In addition, the pyrene–copolymer interaction may affect the
Self-assembly of comb polymers
formation of micelles, thereby influencing the obtained values of
One potential application for these advanced architectures is as CMC. This effect has earlier been observed for low molecular
carriers in drug delivery systems. In the present study, micelli-
zation of the HPC-based copolymers was studied using the
fluorescent probe technique and surface tension measurements.
Pyrene used as fluorescent probe exhibits different fluorescence
characteristics depending on the polarity of the solubilizing
medium. This can favorably be employed to determine the crit-
ical micelle concentration (CMC) of a polymer in solution.40
Furthermore, since earlier studies41,42 indicate discrepancies
between CMCs obtained with different methods, surface tension
measurements were used as a complementary technique to
determine CMC of the polymers. Surface active amphiphilic
polymers decrease the surface tension of an aqueous solution
with increasing polymer concentration. At the CMC the air–
water interface is saturated with amphiphilic molecules and
micelles start to form in the bulk liquid.43 Consequently, the Fig. 6 Change in the fluorescence characteristics of pyrene as a function
surface tension is almost constant for polymer concentrations of HPC1.4-g-PDEGMA concentration.

This journal is ª The Royal Society of Chemistry 2011 Polym. Chem., 2011, 2, 1114–1123 | 1121
View Article Online

may be of interest for biomedical applications, which also will be


the topic of future studies to be conducted in our laboratory.

References
1 D. Klemm, B. Heublein, H.-P. Fink and A. Bohn, Angew. Chem., Int.
Ed., 2005, 44, 3358–3393.
2 A. Carlmark and E. Malmstr€ om, Biomacromolecules, 2003, 4, 1740–
1745.

3 S. Hansson, E. Ostmark, A. Carlmark and E. Malmstr€ om, ACS Appl.
Mater. Interfaces, 2009, 1, 2651–2659.
4 J. Lindqvist, D. Nystr€ €
om, E. Ostmark, P. Antoni, A. Carlmark,
M. Johansson, A. Hult and E. Malmstr€ om, Biomacromolecules,
2008, 9, 2139–2145.
5 R. Westlund, A. Carlmark, A. Hult, E. Malmstr€ om and I. M. Saez,
Fig. 7 Change in surface tension as a function of concentration of Soft Matter, 2007, 3, 866–871.
HPC1.4-g-PDEGMA solutions. 6 N. Nordgren, H. L€ onnberg, A. Hult, E. Malmstr€ om and
M. W. Rutland, ACS Appl. Mater. Interfaces, 2009, 1, 2098–2103.
Published on 16 February 2011. Downloaded on 25/10/2014 02:33:23.

7 J. Lindqvist and E. Malmstr€ om, J. Appl. Polym. Sci., 2006, 100, 4155–
weight surfactant systems when using pyrene as a probe in 4162.
fluorescent measurements.42 8 H. L€ onnberg, L. Fogelstr€ om, L. Berglund, E. Malmstr€ om and
In the case of HPC0.6-g-POEGMA no CMC was observed A. Hult, Eur. Polym. J., 2008, 44, 2991–2997.

9 E. Ostmark, J. Lindqvist, D. Nystr€ om and E. Malmstr€ om,
neither by fluorescence spectroscopy nor surface tension
Biomacromolecules, 2007, 8, 3815–3822.
measurements (in the investigated concentration range). €
10 E. Ostmark, S. Harrisson, K. L. Wooley and E. E. Malmstr€ om,
However, the hydrodynamic radius obtained from DLS (30 nm) Biomacromolecules, 2007, 8, 1138–1148.

11 E. Ostmark, D. Nystr€om and E. Malmstr€ om, Macromolecules, 2008,
suggests that HPC0.6-g-POEGMA forms self-assembled struc-
41, 4405–4415.
tures in solution. In addition, the surface tension of the solution 12 S. M. Grayson and W. T. Godbey, J. Drug Targeting, 2008, 16, 329–
was lowered, indicating that HPC0.6-g-POEGMA exhibits 356.
surface active properties. 13 M. Tizzotti, A. Charlot, E. Fleury, M. Stenzel and J. Bernard,
Macromol. Rapid Commun., 2010, 31, 1751–1772.
14 J.-S. Wang and K. Matyjaszewski, J. Am. Chem. Soc., 1995, 117,
5614–5615.
15 M. Kato, M. Kamigaito, M. Sawamoto and T. Higashimura,
Conclusions Macromolecules, 1995, 28, 1721–1723.
16 X. Li, M. Yin, G. Zhang and F. Zhang, Chin. J. Chem. Eng., 2009, 17,
Linear homo- and copolymers of oligo(ethylene glycol) methyl 145–149.
17 Y. Li, R. Liu, W. Liu, H. Kang, M. Wu and Y. Huang, J. Polym. Sci.,
ether methacrylate (OEGMA300) and di(ethylene glycol) methyl Part A: Polym. Chem., 2008, 46, 6907–6915.
ether methacrylate (DEGMA) were successfully synthesized via 18 F. J. Xu, Y. Ping, J. Ma, G. P. Tang, W. T. Yang, J. Li, E. T. Kang
ARGET ATRP. By altering the initial monomer feed ratio, and K. G. Neoh, Bioconjugate Chem., 2009, 20, 1449–1458.
copolymers with tunable LCSTs were produced, thus enabling 19 F. J. Xu, Y. Zhu, F. S. Liu, J. Nie, J. Ma and W. T. Yang,
Bioconjugate Chem., 2010, 21, 456–464.
polymers with LCSTs in between 27 and 73  C. This widens the 20 T. Pintauer and K. Matyjaszewski, Chem. Soc. Rev., 2008, 37, 1087–
use of these benign and non-toxic materials in different fields. 1097.
Furthermore, this group of thermo-responsive polymers was 21 J.-J. Tan, Y.-X. Li, R.-G. Liu, H.-L. Kang, D.-Q. Wang, L. Ma, W.-
grafted from hydroxypropyl cellulose, resulting in complex Y. Liu, M. Wu and Y. Huang, Carbohydr. Polym., 2010, 81, 213–218.
22 Q. Yan, J. Yuan, F. Zhang, X. Sui, X. Xie, Y. Yin, S. Wang and
macromolecular comb architectures. The comb architecture Y. Wei, Biomacromolecules, 2009, 10, 2033–2042.
preserved the thermo-responsive behavior, however, the LCST 23 J. Gao, G. Haidar, X. Lu and Z. Hu, Macromolecules, 2001, 34, 2242–
was consistently lowered compared to the linear analogues. This 2247.
24 H. G. Schild, Prog. Polym. Sci., 1992, 17, 163–249.
is believed to be an effect of the distorted hydrophilic/hydro- 25 H. Vihola, A. Laukkanen, L. Valtola, H. Tenhu and J. Hirvonen,
phobic balance of the structures and the restrained mobility of Biomaterials, 2005, 26, 3055–3064.
the polymer grafts. Additionally, the significantly higher mole- 26 J.-F. Lutz, J. Andrieu, S. Uezguen, C. Rudolph and S. Agarwal,
cular weight is believed to contribute to this effect. Interestingly, Macromolecules, 2007, 40, 8540–8543.
27 R. B. Greenwald, Y. H. Choe, J. McGuire and C. D. Conover, Adv.
although grafting of PEGMAs onto hydroxypropyl cellulose Drug Delivery Rev., 2003, 55, 217–250.
resulted in lower LCSTs, the ability to tune the transition 28 J.-F. Lutz, J. Polym. Sci., Part A: Polym. Chem., 2008, 46, 3459–3470.
temperature with the OEGMA content of these architectures 29 J.-F. Lutz, O. Akdemir and A. Hoth, J. Am. Chem. Soc., 2006, 128,
13046–13047.
remained. Furthermore, the amphiphilic comb polymers’ ability 30 L. Tao, G. Mantovani, F. Lecolley and D. M. Haddleton, J. Am.
to form polymeric micelles was examined by dynamic light Chem. Soc., 2004, 126, 13220–13221.
scattering, fluorescence spectroscopy and surface tension 31 C. R. Becer, S. Hahn, M. W. M. Fijten, H. M. L. Thijs,
measurements. The results show that, above their CMCs, R. Hoogenboom and U. S. Schubert, J. Polym. Sci., Part A: Polym.
Chem., 2008, 46, 7138–7147.
HPC1.4-g-PDEGMA, HPC0.6-g-PDEGMA and HPC0.6-g- 32 J.-F. Lutz and A. Hoth, Macromolecules, 2006, 39, 893–896.
P(OEGMA52-co-DEGMA48) associates into micellar structures 33 Z.-B. Hu, T. Cai and C.-L. Chi, Soft Matter, 2010, 6, 2115–2132.
in the size range of 20–40 nm. Conclusively, the obtained results 34 J.-F. Lutz, A. Hoth and K. Schade, Des. Monomers Polym., 2009, 12,
343–353.
imply that these architectures, which combine the beneficial
35 S. J. Holder, G. G. Durand, C.-T. Yeoh, E. Illi, N. J. Hardy and
properties of the biocompatible HPC with the thermo-responsive T. H. Richardson, J. Polym. Sci., Part A: Polym. Chem., 2008, 46,
and stealth properties of poly(ethylene glycol) methacrylates, 7739–7756.

1122 | Polym. Chem., 2011, 2, 1114–1123 This journal is ª The Royal Society of Chemistry 2011
View Article Online

36 T. Liu, Z. Zhou, C. Wu, V. M. Nace and B. Chu, Macromolecules, 40 K. Kalyanasundaram, F. Grieser and J. K. Thomas, Chem. Phys.
1997, 30, 7624–7626. Lett., 1977, 51, 501–505.
37 Conducted at ALS Scandinavia AB; Aurorum 10, 97 775 Lule a, 41 E. Amado, C. Augsten, K. Maeder, A. Blume and J. Kressler,
Sweden. Macromolecules, 2006, 39, 9486–9496.
38 D. Shen, H. Yu and Y. Huang, J. Polym. Sci., Part A: Polym. Chem., 42 E. D. Goddard, N. J. Turro and P. L. Kuo, Langmuir, 1985, 1, 352–355.
2005, 43, 4099–4108. 43 B. J€
onsson, B. Lindman, K. Holmberg and B. Kronberg, Surfactants
39 R. Bodvik, A. Dedinaite, L. Karlson, M. Bergstr€ om, P. B€averb€ack, and Polymers in Aqueous Solution, John Wiley & Sons, 2003.
J. S. Pedersen, K. Edwards, G. Karlsson, I. Varga and 44 P. Lundberg, M. V. Walter, M. I. Monta~ nez, D. Hult, A. Hult,
P. M. Claesson, Colloids Surf., A, 2010, 354, 162–171. A. Nystr€om and M. Malkoch, Polym. Chem., 2011, 2, 394–402.
Published on 16 February 2011. Downloaded on 25/10/2014 02:33:23.

This journal is ª The Royal Society of Chemistry 2011 Polym. Chem., 2011, 2, 1114–1123 | 1123

You might also like