You are on page 1of 10

Macromolecular

Chemistry and Physics Full Paper

Oxidant-Dependent REDOX Responsiveness


of Polysulfides

Paolo Carampin, Enrique Lallana, Jureerat Laliturai, Sabrina C. Carroccio,


Concetto Puglisi, Nicola Tirelli*

Poly(propylene sulfide) (PPS) is studied as an oxidation-responsive macromolecular building


block, in the perspective of anti-inflammatory therapies. Here, we show that the nature of
the oxidant has profound effects on the outcome of the oxidation process. PPS was exposed
to hydrogen peroxide (H2O2) and hypochlorite (OCl−), which are oxidizing species commonly
encountered during inflammatory processes. It was found
that the oxidation with H2O2 converted thioethers into sulfox-
ides, producing water-soluble macromolecular products with
extremely low toxicity (L929 fibroblasts); on the contrary, the
reaction with NaOCl produced sulfones in addition to sulfox-
ides, and this was accompanied by depolymerization, which
appears to considerably affect the toxicity of the oxidation
products.

1. Introduction radical (O2•−) from oxygen and nicotinamide adenine


dinucleotide phosphate (NADPH) in a reaction catalyzed
An ubiquitous feature of inflamed tissues is the presence by NADPH oxidases, which are membrane-bound enzy-
of a number of phagocyte-generated oxidants, which are matic systems typically dormant in resting phagocytes,
generally produced from molecular oxygen through a cas- but activated by a variety of external and internal stimuli,
cade of one-electron reduction steps.[1,2] Most commonly, for example, bacteria and soluble factors, such as certain
this cascade starts with the generation of superoxide anion inflammatory polypeptides (e.g., C5a).[3] The released O2•−
can either react with nitric oxide (NO) to form highly reac-
tive peroxynitrite (ONOO−)[4] or alternatively can be rapidly
P. Carampin, J. Laliturai[a]
converted into hydrogen peroxide (H2O2) under the influ-
School of Pharmacy and Pharmaceutical Sciences, ence of superoxide dismutase. In the presence of Fe2+, H2O2
University of Manchester, Oxford Road, Manchester, can also produce the more damaging hydroxyl radical (•OH)
M13 9PT, United Kingdom through Fenton chemistry.[5] Moreover, REDOX cycling of
E. Lallana, N. Tirelli Fe2+ and Fe3+ through a combination of Haber–Weiss and
School of Materials, University of Manchester, Fenton chemistry can also result in the direct formation of
Oxford Road, Manchester, M13 9PT, United Kingdom •
OH from O2•−.[6] The heme peroxidases such as myeloper-
E-mail: nicola.tirelli@manchester.ac.uk oxidase and eosinophil peroxidase, secreted by inflam-
S. C. Carroccio, C. Puglisi matory cells, can also catalyze the formation of the very
Istituto di Chimica e Tecnologia dei Polimeri,
potent and damaging oxidants hypochlorous acid (HOCl)
UOS Catania, Consiglio Nazionale delle Ricerche,
and hypobromous acid (HOBr) from H2O2 in the presence of
Via P. Gaifami 18, 95126 Catania, Italy
N. Tirelli
chloride (Cl−) and bromide (Br−), respectively. On the other
School of Biomedicine, University of Manchester, hand, oxidants generated as a result of the activation of
Oxford Road, Manchester, M13 9PT, United Kingdom inflammatory cells can lead to further activation processes,
[a] The Government Pharmaceutical Organization, 75/1 Rama either directly or through the mediation of oxidized bio-
6 Road, Ratchathewi, Bangkok 10400, Thailand molecules.[7] For example, the exposure of Immunoglobin

Macromol. Chem. Phys. 2012, 213, 2052−2061


2052 © 2012 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim wileyonlinelibrary.com DOI: 10.1002/macp.201200264
Macromolecular
Oxidant-Dependent REDOX Responsiveness of Polysulfides . . . Chemistry and Physics
www.mcp-journal.de

G (IgG) to •OH or HOCl in the inflamed rheumatoid joint nonactivated macrophages, and can extensively circulate
has been claimed to generate an aggregated IgG form that in rats,[31] although this behavior may strongly depend on
increases neutrophil activation,[8] while complement acti- the nature of the chemical groups present on the surface
vation is reportedly increased by the action of oxygen radi- of the nanoparticle.[32]
cals.[9] In addition, in many forms of tissue injury, including Comparatively little has been done to understand
mechanical disruption, the extent of free-radical reactions the chemistry of the oxidation process, and specifically
can be increased by inactivation of cellular antioxidants, as whether the action of different oxidants may result not
well as by the release of transition metal ions from their only in different oxidation kinetics but also in different
sites of sequestration within cells.[10] reaction products. Few literature reports have focused
The above summary illustrates the role of oxidant spe- on the oxidation behavior of organic polysulfides,[33–38]
cies in inflammatory processes, which is the rationale for showing that they can be oxidized in vitro to polysul-
the use of REDOX-responsive drug delivery systems sensi- foxides,[33,37] although an earlier report from Marvel and
tive to inflammation-triggered oxidation reactions.[11] In Weil hinted to polysulfones too.[39] It is conceivable that
this regard, we have focused our attention on a class of under mild oxidation conditions (e.g., H2O2 as an oxidizing
organic polymers, poly(1,2-alkylene sulfides), where the agent) the reaction could generate lesser oxidized species,
conversion of sulphur (II) to higher oxidation states (sul- that is, sulfoxides, while using harsher conditions (e.g.,
foxides or sulfones) results in a large change in polarity oxidation with organic peroxyacids) polymeric sulfones
and, consequently, in an increase of their water solu- can be obtained.[34,36,38,40] However, there are conflicting
bility/swellability.[12–15] The resulting morphological reor- evidences: for example, Orgeretravanat and Galin[37]
ganization can trigger the release of encapsulated active reported that the oxidation of PPS with meta-chloroper-
principles in an oxidative (inflamed) environment. This oxybenzoic acid produced poly(propylene sulfoxide) and
scheme can be applied to intra- or extracellular oxidation, not sulfones; surprisingly, this was accompanied by a
depending on a number of additional factors, for instance severe depolymerization of the substrate, which would be
the recognition of the carrier and its cellular uptake.[16] more typical of a sulfone-containing structure.[37] Addi-
Polysulfide preparative chemistry is overwhelm- tionally, the effects of other biological oxidants, such as
ingly based on the living anionic polymerization of ClO−, or of heterogeneous conditions still remain to be
episulfides,[17] which allows the preparation of narrow addressed.
dispersity macromolecules with an excellent control The knowledge of the oxidized products is also impor-
over terminal groups,[18] possibly featuring a blocky tant: sulfoxide-containing polymers were initially
structure.[19–21] Episulfide polymerization can be car- studied by the group of Ringsdorf in the 1970s with
ried out under mild conditions, and can be performed in very promising results in terms of extremely low in vivo
protic media,[22] provided that disulfide impurities are toxicity.[41,42] On the other hand, low-molecular-weight
removed.[23,24] This versatility has allowed to produce a depolymerization products and/or sulfone-containing
variety of oxidation-sensitive colloidal carriers based on substrates may be considerably more toxic.
poly(propylene sulfide) (PPS): vesicles[13] and micelles,[25] In the present work, PPS has been synthesized from a
which are based on the self-assembly in water of dithiolate initiator to form linear bifunctional polymer
amphiphilic block copolymer structures, and nanoparti- chains. The polymer was then oxidized under different
cles,[14,26] which are produced through in situ emulsion physicochemical environments using H2O2 and NaOCl,
polymerization techniques. Upon exposure to H2O2, the which are typical markers of inflammatory reactions
gradual oxidation of PPS transforms vesicles into higher but are also characterized by different solubilities in an
curvature structures, such as worm-like micelles, spher- organic environment.
ical micelles, unimolecular micelles, eventually yielding
soluble polymers.[27,28] Cross-linked PPS nanoparticles
respond to H2O2 through a bulk oxidation mechanism, 2. Experimental Section
with the oxidant diffusion into the nanoparticles occur-
ring more rapidly than their oxidation; this determines 2.1. Materials
a gradual and homogeneous swelling of the polysulfide
All chemicals were used as received unless otherwise stated. Pro-
network in water, allowing the corresponding release of
pylene sulfide (PS), acetyl chloride, ethyl bromoacetate, 2,2′-(ethy-
encapsulated drugs such as doxorubicin.[29,30] PPS nano-
lenedioxy)diethanethiol, Live/Dead Cell Double Staining Kit,
particles also released hydrophobic payloads such as Nile NaOCl aqueous solution (available chlorine 10%–15%), tributy-
red and Reichardt’s dye after oxidation with NaOCl.[15] lphosphine (TBP), fungizone, dichloromethane, triethylamine
We have further demonstrated that PEGylated (TEA), and Dulbecco’s Modified Eagle’s Medium (DMEM) with
PPS nanoparticles exhibit a substantially “stealth” high glucose (4500 mg L−1) were purchased from Sigma–Aldrich
behavior (very slow endocytic uptake) in the presence of (Gillingham, UK). Ethyl acetate (AcOEt), tetrahydrofuran (THF),

Macromol. Chem. Phys. 2012, 213, 2052−2061


www.MaterialsViews.com © 2012 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim 2053
Macromolecular
Chemistry and Physics P. Carampin et al.

www.mcp-journal.de

and 30% (w/w) H2O2 in water were purchased from Fisher (delayed extraction, time lag) were optimized for each sample to
(Leicester, UK). 0.5 M sodium methoxide (NaOMe), methanolic obtain the highest molar mass values. The laser irradiance was
solution, and 1,8-diazabicyclo[5.4.0]undec-7-ene (DBU) were pur- maintained slightly above threshold. Samples for MALDI-TOF
chased from Fluka (Gillingham, UK). Phosphate-buffered saline analysis (PPS and PPS oxidized with NaOCl) were prepared by
(PBS) Dulbecco A tablets were purchased from Oxoid (Hampshire, mixing 10 μL of the matrix solution [0.1 M 2-(4-hydroxyphenylazo)
UK). CellTiter 96 AQueous One Solution Cell Proliferation Assay benzoic acid (HABA) in THF] and the appropriate volume of 5 mg
was purchased from Promega (Southampton, UK). Foetal bovine mL−1 polymer solution in THF to obtain a 1:1 or 1:3 (v/v) (sample/
serum (FBS), penicillin-streptomycin and L-glutamine 200 × 10−3M matrix) mixture. Finally, 1 μL of each sample/matrix mixture was
(×100) were purchased from Invitrogen (Paisley, UK). spotted on the sample holder and slowly dried to allow matrix
For all operations, THF was degassed by bubbling argon under crystallization. Note: different matrices, that is HABA, dithranol,
an inert atmosphere for 1 h before use. and trans-3-indoleacrylic acid (IAA), and solvents, that is, THF
and/or THF/CHCl3, were unsuccessfully tried for the analysis of
PPS oxidized with H2O2, but the high polarity of the resulting
polymers apparently prevented their detachment.
2.2. Physicochemical Characterization
1H NMR spectra were recorded on a Bruker 500 MHz spectro-

meter (Bruker UK Limited, UK). NMR samples were prepared in 2.3. Preparative Operations
CDCl3 at a concentration of approximately 1 wt%. Chemical shifts
2.3.1. Synthesis of Protected Bifunctional Initiator:
are reported in ppm (δ units) upfield using internal tetramethy-
2,2′-(ethylenedioxy)diethanethioacetate
lsilane (0.00 ppm) as a reference.
FTIR spectra were recorded in ATR mode on a Tensor 27 Bruker 3.8 g of 2,2′-(ethylenedioxy)diethanethiol (20.8 mmol, corre-
spectrometer (Bruker UK Limited, UK) equipped with a 3000 sponding to 41.6 mmol of thiols) was dissolved in 300 mL of THF
Series TM High Stability Temperature Controller with RS232 under an inert atmosphere and cooled to 0 °C. 17.4 mL of TEA
Control (Specac, UK). A drop of a solution of the substrate (in THF (124.8 mmol, 3:1 molar ratio to thiols) was added under stirring.
or water) was placed on a Golden Gate Heated Diamond ATR top 5.9 mL of acetyl chloride (83.2 mmol, 2:1 molar ratio to thiols) was
plate and allowed to completely dry at 50 °C. dissolved in 30 mL of THF and added dropwise. The reaction mix-
Molecular weight and molecular weight distribution of ture was allowed to reach room temperature and stirred for 4 h.
polymers were determined using a Polymer Laboratories The insoluble TEA hydrochloride was removed via filtration, and
PL-GPC50 integrated GPC (Polymer Laboratories, UK) comprising the supernatant was concentrated under reduced pressure. The
a PLgel 5 μm Guard and two PolyPore 5 μm columns operating at resulting residue was dissolved in 100 mL of dichloromethane
30 °C. THF was used as an eluent at a flow rate of 1.0 mL min−1. and extracted with brine (3 × 25 mL). The organic layer was dried
A series of near-mono-dispersed linear polystyrene standards over Na2SO4 and concentrated under reduced pressure to give a
(Fluka; Gillingham, UK) was used for calibration with a refractive crude product that was purified by column chromatography on
index detector for the analysis of the polymers. silica gel 230-400 mesh (hexane/AcOEt 3:1), finally recovering a
SLS measurements were carried out on a BI-APD Brookhaven colorless oil. Yield: 3.12 g = 56% wt.
instrument (Brookhaven, USA) equipped with a 35 mW He–Ne FTIR (film on ATR crystal), δ (cm−1): 2926 (νas CH2), 1687
laser emitting vertically polarized light at 632.8 nm, a BI-200SM (ν CO), 1131, 1099 (νas C—O—C), 952 (νs C—O—C).
goniometer, and a BI-9000 digital correlator scanning over the 1
H MNR (CDCl3), δ (ppm): 2.33 (s, 6, C2—C2—C2—S—
angular range 15–155°. Completely oxidized PPS with molecular COCH3), 3.05–3.10 (t, 4,—C2—C2—S—COCH3), 3.55–3.60
weight around 3000 g mol−1 (2 equivalents of H2O2 for 20 h) was (m, 8,—SCH2CH2—OCH2CH2OCH2CH2S—).
dissolved in MilliQ water prefiltered through a 0.2 μm pore size
syringe filter to remove dust particles (concentration in the range
2.3.2. Synthesis of Poly(propylene sulfide)
4.0–9.0 mg mL−1). Each polymer solution was filtered directly
into the light scattering cell for 30 min through an on-line Polymerization reactions were carried out in a FirstMate parallel
filtration system composed of a Stepdos 03S liquid pump (KNF- reactor (Argonaut Technologies), which was purged with argon
Flodos, Switzerland) connected to a 0.2 μm pore size filter. SLS for 5 min prior to the introduction of reagents. The reaction ves-
measurements were performed at 25 °C mounting the sample sels were maintained under a positive argon pressure during
cell in a temperature-controlled bath filled with cis–trans polymerization. All stock solutions employed were prepared
decahydronaphthalene (decalin). Under these experimental from solvents degassed by argon bubbling for 1 h. Following
conditions, a Zimm plot was constructed using BI-ZMP software the polymerization/end-capping procedure, PPS with molecular
from which the polymer Mw and second virial coefficient (A2) weights around 3000, 6000, and 9000 g mol−1 (PPS3000, PPS6000,
were calculated. Refractive index increments (dn/dc) were and PPS9000, respectively) were prepared. In a typical procedure,
determined at 25 °C using an ABBE refractometer (Belingham 2 mL of a 2,2′-(ethylenedioxy)diethanethioacetate stock solu-
and Stanley Limited, UK). tion (133 mg mL−1 in THF, 1.0 mmol), 1.62 mL of TBP stock solu-
MALDI-TOF mass spectra were recorded in linear and tion (1.25 mg mL−1 in THF, 10-fold excess of TBP per thioacetate
reflectron mode using a Voyager-DETM STR mass spectrometer group), and degassed THF (5 mL) were introduced into the reactor
working in positive ion mode. The instrument was equipped vessel. 2.10 equivalents of NaOMe from an aliquot of 0.5 M
with a nitrogen laser emitting at 337 nm. The accelerating NaOMe methanolic solution (4.2 mL, 2.1 mmol) were added next,
voltage was set at 20 kV, and the grid voltage and the delay time and the final mixture was stirred for 5 min at room temperature.

Macromol. Chem. Phys. 2012, 213, 2052−2061


2054 © 2012 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim www.MaterialsViews.com
Macromolecular
Oxidant-Dependent REDOX Responsiveness of Polysulfides . . . Chemistry and Physics
www.mcp-journal.de

In order to obtain PPS with molecular weights around 3000, 2.3.5. Oxidation With NaOCl in THF (8.0-15.9 wt% Water
6000, and 9000 g mol−1, an appropriate amount of PS (3.2, 6.3, or Content)
12.6 mL, corresponding to 40, 80, and 120 equivalents, compared
to thioacetate groups) was added to the solution. The polymeriza- An aliquot of 1.88 M NaOCl in water (373, 560 or 746 μL, corre-
tion mixture was stirred for 45 min at room temperature. After- sponding to 0.70, 1.05, and 1.40 mmol and then to 0.5, 0.75, and
ward, 1.10 equivalents of acetic acid and 1.15 equivalents of DBU 1.0 equivalents relative to thioether groups, respectively) was
were sequentially added to neutralize the excess of NaOMe and added to 5 mL of a THF solution containing 100 mg of PPS3000
adjust the pH. Finally, 3.6 equivalents of end-capping reagent (1.40 mmol of thioether groups). In typical experiments, the mix-
ethyl bromoacetate (0.4 mL, 3.6 mmol) were added, and the reac- ture was allowed to react for 20 h at 37 ºC, but experiments con-
tion mixture was stirred for 2 h at room temperature. The solvent ducted with 1 equivalent of NaOCl were also reacted for times
was then evaporated under reduced pressure, and the obtained ranging between 2 min and 96 h. After the addition of 2 mL of
crude product was dissolved in 30 mL of dichloromethane and MilliQ water (precipitation observed), the mixture was dialyzed
extracted with water (3 × 6 mL). The organic layer was dried over against MilliQ water (six changes during 48 h) through 1000 Da
Na2SO4 and concentrated under reduced pressure to give a vis- MWCO dialysis membranes, and freeze-dried; the water-soluble
cous oil that was washed with methanol (3 × 5 mL). The polymer and water-insoluble fractions were separated as described above.
was separated by decantation after centrifugation and dried
under high vacuum for 24 h at 40 °C. Yield: 65-70% wt. 2.3.6. Oxidation With NaOCl in Dichloromethane/Water
End-capping yield: ≥95% (calculated by 1H NMR from the
integration of —SCH2CH2—OCH2CH2OCH2CH2S—(3.51–3.68 ppm) An aliquot of 1.88 M NaOCl in water (3.7 mL, 7.0 mmol,
and H3CH2—O(O)C—(4.05–4.26 ppm)). corresponding to 5 equivalents relative to thioethers) was
FTIR (film on ATR crystal), δ (cm−1): 2959 (νas CH3), 2917 (νas added to a solution of PPS3000 (100 mg, 0.03 mmol) in a
CH2), 2865 (νs CH3 and νs CH2), 1730 (ν CO ester), 1448 (δs CH2), dichloromethane/water mixture (5.0 and 1.3 mL, respec-
1372, 1263, 1221, 1172, 1103 (νas C—O—C), 943, 852 (νs C—O—C) tively). The mixture was stirred for 0.5, 24, 48, 72, or 96 h
cm−1. at 37 °C. Afterward, the dichloromethane was evaporated
1H NMR (CDCl ), δ (ppm): 1.26–1.30 (t, 6H, CH CH —OOC—), 1.35–
3 3 2 and the resulting aqueous phase was freeze-dried. Finally,
1.45 (d, CH3 in PPS chain), 2.55–2.75 (m, PPS chain: 1 diastereotopic both the water-soluble and water-insoluble fractions of PPS
CH2), 2.71–2.78 (t, 4H,—SCH2CH2OCH2CH2OCH2CH2S—), 2.85-
oxidation products were separated as described above.
3.05 (m, PPS chain: CH and 1 diastereotopic CH2), 3.16-3.25
(d, 1H,—S—CH2—COO—: 1 diastereotopic CH2), 3.25-3.37
(d, 1H,—S—CH2—COO—: 1 diastereotopic CH2), 3.51–3.68 2.4. Cell Culture
(m, 8H,—SCH2CH2—OCH2CH2OCH2CH2S—), 4.05–4.26
(q, 4H, CH3CH2—O(O)C—). The mouse fibroblast L929 cell line was obtained from the Euro-
pean Collection of Animal Cell Cultures (ECACC, Rockville, UK). Cells
were routinely maintained in culture medium composed of DMEM
2.3.3. Oxidation Experiments supplemented with 10% (v/v) FBS, 2 × 10−3M L-glutamine, 1% (v/v)
Fungizone, and 0.5% (v/v) penicillin–streptomycin. Cells in sup-
The responsiveness of PPS was studied under different physico- plemented DMEM medium were grown as a confluent monolayer
chemical scenarios: H2O2 in THF, NaOCl in THF with low or high culture on 75 cm2 polystyrene tissue culture flasks (Falcon, Oxford,
water content, and in dichoromethane/water 1:1 mixture. Unless UK). Cell cultures were maintained at 37 °C in a humidified atmos-
otherwise stated, PPS3000 was used as a substrate. Experiments phere of 5% CO2. The culture medium was changed three times a
with PPS6000 and PPS9000 were carried out under identical experi- week. Adherent cells approaching 90% confluence were harvested
mental conditions as described for PPS3000. with trypsin and subcultured. Passages were always below 6.

2.3.4. Oxidation With H2O2 in THF (7.0 wt% Water 2.4.1. Cytotoxicity
Content)
The oxidation of MTS ([3-(4,5-dimethylthiazol-2-yl)-5-(3-
An aliquot of 30 wt% H2O2 in water (285 μL, 2.80 mmol) was carboxymethoxyphenyl)-2-(4-sulfophenyl)-2H-tetrazolium, inner
added to 3 mL of a THF solution containing 100 mg of PPS3000 salt]) to the corresponding colored formazan (CellTiter 96 One
(1.40 mmol of thioether groups, corresponding to a 2:1 peroxide/ Solution Cell Proliferation Assay) and the fluorescence of cal-
thioether ratio). The mixture was stirred for 1, 5, 10, 14, 17, or cein acetoxymethyl ester (calcein-AM) and propidium iodide (as
20 h at room temperature; after addition of 2 mL of deionized markers of live and dead cells, respectively; Live/Dead Cell Double
water, it was dialyzed against MilliQ water (six changes during Staining Kit) was used to determine the cytotoxicity of PPS oxi-
48 h) through 1000 Da MWCO (molecular weight cut-off) dial- dation products on L929 mouse fibroblasts. Cells in culture media
ysis membranes (SpectraPor 7 regenerated cellulose tubings, were seeded and incubated in a 96 well plate at 37 °C in a humid-
Spectrum Laboratories, Inc., USA), and finally freeze-dried. The ified atmosphere of 5% CO2 at a density of 8000 cells/well. After
obtained solid was suspended in 10 mL of water, isolating the 24 h, the culture medium was replaced with a solution of fresh
water-soluble fraction via centrifugation at 2000 rpm for 5 min medium containing varying concentrations of the oxidation
and then freeze-drying it. The water-insoluble fraction was dried products of linear PPS or PPS nanoparticles, which was previously
under vacuum for 24 h at 40 °C. filtered through a sterile 0.22 μm filter. After 24 h, the incubation

Macromol. Chem. Phys. 2012, 213, 2052−2061


www.MaterialsViews.com © 2012 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim 2055
Macromolecular
Chemistry and Physics P. Carampin et al.

www.mcp-journal.de

medium was removed, the cells were rinsed


twice with PBS at pH 7.4, and their viability
was analyzed. For the MTS assay, a mixture
containing MTS and culture medium without
FBS and phenol red was added into each well.
After 3 h of incubation, the absorbance was
read on a microplate reader (μQuant, Biotek,
UK) at 490 nm. The relative cell viability
(%) was calculated by the formula [A]test/
[A]control∗100, where [A]test is the absorbance Figure 1. Preparation of PPS by in situ deprotection of the dithioacetate precursor, ani-
of the test sample, and [A]control is the absorb- onic ring-opening episulfide polymerization, and end-capping of the resulting bifunc-
ance of the control cells incubated solely with tional PPS.
culture medium. For the live/dead assay, cells
were simultaneously incubated with calcein-AM and propidium this precursor in the presence of a reducing agent (TBP) allows
iodide in PBS (pH 7.4) for 10 min at 37 °C, washed three times to minimize the presence of disulfides, which behave as
with PBS (pH 7.4), and immediately analyzed for green (live cells) chain transfer agents in episulfide polymerization. The addi-
and red (dead cells) fluorescence using a fluorescence microscope tion of PS produced a bis(thiol)-terminated polysulfide that
(Leica DMI 6000 B, Leica Microsystem). The relative cell viability
was finally end-capped with ethyl bromoacetate following
(%) was calculated as the ratio between the number of live cells
a recently optimized protocol.[18,22] Using this procedure, PPS
(green) and the total number of cells (live and death: green and
with a targeted degree of polymerization of 40, 80, and 120
red, respectively) multiplied by 100.
(PPS3000, PPS6000, and PPS9000, respectively) were prepared, with
a 65%–70% overall yield and polydispersity values around 1.1
3. Results and Discussion (GPC and MALDI-TOF; Table 1), confirming the living and con-
trolled character of the episulfide polymerization.
3.1. Synthesis of PPS
3.2. Oxidation Behavior of PPS
PPS was synthesized from a bifunctional initiator [2,2′-(ethy-
lenedioxy)diethanethiol], which was generated in situ from Since PPS is hydrophobic, its oxidation may proceed in
its acetylated precursor (Figure 1); the in situ methanolysis of homogeneous or heterogeneous conditions depending on

Table 1. Composition and molecular weight of PPS3000 oxidized with H2O2 in THF.

Reaction time Water-insoluble fractiona) Conversion to sulfoxide Mw d) Mn d) Mw / M


n d)
[h] [wt%] [mol%] [g mol−1] [g mol−1]
1H NMRb) FTIRc)
0 100 0 0 3420 (3228)e) 3120 (3046)e) 1.09 (1.06)e)
1 100 4.1 5.8 3190 2890 1.11
5 100 7.0 6.0 3010 2700 1.12
10 100 10.4 6.3 3280 2970 1.12
14 100 55.5 40.1 1490 430 3.43
17 45 52.9f) 68.6f) 1320f) 430f) 3.06f)
77.6g) 83.0g) = = =
20 0 93.5 93.5 3800h) = =
a)The THF solution containing the oxidized polymer was transferred into a water environment and then dialyzed against water (1000
Da MWCO) and freeze-dried. The water-soluble fraction was isolated by suspending the obtained solid in 10 mL of water followed by
filtration (water-insoluble fraction) and freeze-drying of supernatant; b)Calculated from the integration of signals corresponding to the
methyl group next to sulfoxides (1.38–1.58 ppm) and the total content of the methyl groups (1.29–1.58 ppm); c)Calculated from the
height of the band at 1030 cm−1 (sulfoxide stretching). The intensity of the signals was normalized against the νas CH3 band at 2959 cm−1.
To provide absolute values, the conversion of the sample at 20 h was assumed equal to that obtained from 1H NMR (93.5%); d)Calculated
from GPC measurements in THF, unless otherwise stated; e)Calculated from MALDI-TOF measurements;[43] f)THF-soluble fraction; g)THF-
insoluble (water-dispersible) fraction. Although insoluble in THF, this material aggregated in water and most polar solvents, hindering
SLS analysis; h)Calculated from SLS measurements in water. Mw = 3780 ± 52 and 3826 ± 46 g mol−1 were obtained from the extrapola-
tions to zero angle and zero concentration, respectively. A2 = 1.14 ± 0.24 was calculated from the Zimm plot, a positive value that sug-
gests lack of aggregation in water. dn/dc = 0.14 ± 0.005 mg mL−1.

Macromol. Chem. Phys. 2012, 213, 2052−2061


2056 © 2012 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim www.MaterialsViews.com
Macromolecular
Oxidant-Dependent REDOX Responsiveness of Polysulfides . . . Chemistry and Physics
www.mcp-journal.de

oxidized species) and methyl resonances


(1.29–1.38 ppm for sulfides, 1.38–1.58
ppm for oxidized species) (Figure 2A to
C). It is noteworthy that the development
of a νSO band can be related to the
increasing presence of sulfinic acids or
of sulfoxides, which respectively imply
or not depolymerization reactions. How-
ever, no significant change was recorded
below 1000 cm−1, which would exclude
the presence of sulfinic acids (νS—O). At
the same time, no signal possibly related
to sulfone (or sulfonic) groups could be
recorded in FTIR spectra.
Both NMR and FTIR revealed the
formation of oxidized sulphur species
already after 1 h (Table 1 and Figure 2A
to C), with an initially slow kinetics and
a significant acceleration after 10 h
(Figure 3), to yield an almost quantita-
tive conversion after 20 h (Figure 2C,
disappearance of the initial methyl
Figure 2. Top left (A): FTIR spectra of oxidized PPS3000 in THF treated with H2O2 (2:1 molar
ratio with sulfides). Top right (B): 900-1400 cm−1 window of the same spectra, showing
resonance).
the dramatic increase in absorbance of the SO stretching vibration. Arrows indicate After 14 h of reaction all samples were
variations in diagnostic bands. The intensity of all spectra was normalized against the opaque, with significant flocculation
νas CH3 band at 2959 cm−1. Bottom left (C):1H NMR spectra of PPS3000 in THF treated at later stages. Correspondingly, GPC in
with H2O2 (2:1 molar ratio with sulfides) as a function of the reaction time. The broader THF showed a markedly higher reten-
appearance of the resonances associated to the oxidized polymers is an indication of the
aggregation of the high dipole moment sulfoxides also in CDCl3. Bottom right (D): GPC
tion with increasing oxidation, which
traces in THF for PPS3000 treated with H2O2 (2:1 molar ratio with sulfides) as a function could be ascribed to coil contraction due
of time. The shift to longer retention times is particularly evident for the broad peaks to poorer solvation or depolymerization
recorded after 14 and 17 h of oxidation. This phenomenon is likely related to the poor (Figure 2D). MALDI-TOF analysis was not
expansion of the sulfoxide-containing polymer coils in THF. possible, due to poor desorption of the
polymers from the matrix, but the fully
the polarity of the oxidant. We have studied PPS oxidation oxidized product (20 h) presented a Mw = 3800 g mol−1 in
using two model oxidants, H2O2 and NaOCl: on one hand, water (SLS, Table 1; consider the increase in mass due to
they are among the most common reactive oxygen species
in inflammation processes[6]; on the other hand, due to
their different polarity and charge, they allow to study the
reactions, respectively, under homogeneous and heteroge-
neous conditions.

3.2.1. Oxidation With H2O2 (Homogeneous Oxidation)


THF was chosen as one of the few organic solvents where
both PPS and H2O2 are soluble and it is also easy to remove
via evaporation. The experiments were carried out at room
temperature after addition of H2O2 ([H2O2] ≈ 3 wt%, 2:1
molar ratio with sulfides; the medium contained about 7
wt% of water). The progress of the reaction was monitored
at different time intervals by FTIR (Figure 2A and B) and
1
H NMR (Figure 2C), respectively using the SO stretching
vibration band at 1030 cm−1 and a vibration at 1170 cm−1 Figure 3. Conversion of sulfide to sulfoxide (expressed as mol%
(possibly related to the sulfide group), and methylene and of the total sulphur atoms) in PPS3000 oxidized in THF with H2O2
methine (2.54–2.96 ppm for sulfides, 2.58–3.49 ppm for (2:1 molar ratio with sulfides).

Macromol. Chem. Phys. 2012, 213, 2052−2061


www.MaterialsViews.com © 2012 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim 2057
Macromolecular
Chemistry and Physics P. Carampin et al.

www.mcp-journal.de

in a hydrophobic environment (THF,


water content ranging between 8.0 and
15.9 wt%), where it is likely present
in a dispersed state with a high local
concentration, (2) in a biphasic system
(dichloromethane/water 1:1), where the
oxidation should predominantly occur
interfacially.
(1) NaOCl dispersed in a hydrophobic
environment. No phase separation was
recorded during oxidation, while a sig-
nificant fraction of the polymer could
be precipitated at the end of reaction
by addition of water; after dialysis,
the amount of water-insoluble mate-
rial decreased with increasing amount
Figure 4 . Cell viability of L929 fibroblasts after exposure for 24 h to variable concentra-
tions of water-soluble oxidized PPS3000 with H2O2 (2:1 molar ratio with sulfides) for 20 h. of NaOCl (55% of the original weight
for 0.5 equivalents of NaOCl; 39 wt%
for 1 equivalent of NaOCl/20 h); on the
the introduction of oxygen atoms), while no appreciable other hand, no polymeric substance could be revealed by
mass loss was recorded upon dialysis (1000 Da MWCO) SLS analysis of the water phase, nor any material could
and the presence of possibly terminal groups in the form be recovered from it after freeze-drying, suggesting that
of sulfinic or sulfonic acids was already excluded on the the remaining mass of oxidized PPS3000 was converted
basis of the FTIR results. We are therefore inclined to to volatile products. In the water-insoluble phase, oxi-
dized groups were recorded (both FTIR and 1
exclude the occurrence of any depolymerization reaction. H NMR)
The autoacceleration can therefore be explained on the already after treatment with 0.5 equivalents of NaOCl per
basis of the phase separation of oxidized PPS segments thioether (Figure 6 ) and as soon as 2–3 min after the addi-
into polar and thus water-rich domains, where the higher tion of the oxidant (Figure 1SI, Supporting Information).
local concentration of H2O2 would increase the rate of In the FTIR spectra, it was possible to detect the devel-
oxidation. opment of a νSO band at 1030 cm−1 and of two bands
The cytotoxicity of the water-soluble/dispersible PPS (1215 and 1370 cm−1) likely associated to the symmetric
oxidation products was evaluated on L929 murine fibrob-
lasts by investigating the cell viability through a live/
dead and a metabolic (MTS) assay. The two tests pro-
vided a reasonably good agreement showing no cytotoxic
effects for fully oxidized PPS3000 up to a concentration of
at least 10 mg mL−1 (Figure 4). Further, while PPS3000 and
PPS6000 presented IC50 = 33–35 mg mL−1, the larger PPS9000
did not cause any significant decrease in cell viability up
to a concentration of 50 mg mL−1 (5 wt%) (Figure 5). Prod-
ucts of partial oxidation showed somehow higher tox-
icity, for example, 80% oxidized PPS3000 (17 h oxidation)
exhibited an IC50 = 15 mg mL−1; this effect is likely due
to amphiphilicity, which may confer some membrane-
disrupting activity. However, even for the partially
oxidized products, the IC50 values are far above (>1 wt%)
than the concentrations used for carriers in drug delivery.

3.2.2. Oxidation With NaOCl (Heterogeneous Oxidation)


NaOCl is substantially insoluble in any organic environ- Figure 5. Cell viability (MTS assay) of L929 murine fibroblast after
exposure for 24 h to various concentrations of water-soluble oxi-
ment, and may therefore be useful to study the effects of dized PPS3000, PPS6000, and PPS9000 with H2O2 (2:1 molar ratio with
a heterogeneous oxidation process; we have specifically sulfides) for 20 h and oxidized PPS3000 with H2O2 (2:1 molar ratio
used two different conditions of dispersion: (1) dispersed with sulfides) for 17 h.

Macromol. Chem. Phys. 2012, 213, 2052−2061


2058 © 2012 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim www.MaterialsViews.com
Macromolecular
Oxidant-Dependent REDOX Responsiveness of Polysulfides . . . Chemistry and Physics
www.mcp-journal.de

ascribed to structures (SA to SD) com-


patible with alkylene sulfone depolym-
erization because they feature terminal
double bonds and sulfinic acids. This is
not a surprising phenomenon because
aliphatic polysulfones [poly(alkenyl sul-
fones)] are well known to depolymerize
under mild conditions: the weak C—S
bond can be easily cleaved thermally[44]
or via irradiation (e.g., X-rays[45,46] or
electron beams[47]), releasing ultimately
SO2 and olefins. It is unclear whether
Figure 6. Left: FTIR spectra of PPS3000 after 20 h of reaction in low water content THF the observed reduction in molecular
mixtures as a function of NaOCl concentration. Arrows indicate variations in dia- weight can be ascribed to depolymeri-
gnostic bands. The intensity of all spectra was normalized against the νas CH3 band at
zation occurring during oxidation, or to
2959 cm−1. A plot of the time evolution of the band at 1030 cm−1 is provided in the
Supporting Information, Figure 1SI. Right: H NMR spectra of oxidized PSS3000 in CDCl3 a fragmentation during MALDI-TOF
1

corresponding to the same reaction conditions. analysis. GPC analysis in THF of the
water-insoluble phase showed the same
aggregation-related problems as in
and asymmetric stretching of an SO2 group, respectively. Figure 2D. The absence of S—O stretching vibrations from
We are inclined to exclude the presence of sulfinic or sul- FTIR spectra of oxidized material apparently excluded the
fonic acids, either in protonated or deprotonated form, presence of sulfinic acids, suggesting the second option.
due to the absence of any new band below 1000 cm−1 (cor- On the other hand, the very significant weight loss and
responding to νS—O) following oxidation. The intensity of absence of polymers in the water phase would favor the
the carbonyl band increased in comparison to that of νas first hypothesis.
CH3 of PPS chain methyl groups at 2959 cm−1, which was It can be therefore concluded that the oxidation with
used to normalize the FTIR spectra. This higher weight of ClO− in THF introduced significant amounts of sulfones
terminal groups could indicate a depolymerization ran- in the PPS chain, and this led to an overall labilization
domly occurring in the PPS chain, coher-
ently with the recorded weight loss.
The 1H NMR spectra broadly con-
firmed the presence of two different oxi-
dized groups: a peak at 1.33–1.58 ppm
could be associated to methyl groups
close to sulfoxides, while one at
1.61–1.66 ppm, due to the larger down-
field shift, could be assigned to more
polar sulfones (Figure 6, right). On the
contrary, the resonances of methylene
and methyl groups were only marginally
affected (peak broadening).
MALDI-TOF analysis showed a sharp
decrease in molecular weight with
increasing amounts of NaOCl: already
at 0.5 equivalents of ClO− (per thioether),
a majority of the material appeared to
have a molar mass lower than 1500 g Figure 7. MALDI-TOF spectra of oxidized PPS3000 as a function of NaOCl concentration
mol−1 (Figure 7). The degradation prod- after 20 h of reaction in THF/water mixtures. The peak labels in the left inset corre-
ucts still exhibited a PS-based repeating spond to compounds SA-SD (showed in the right inset) and the further oxidized products
structure (74 Da spacing) with the with 16 Da spacing peaks derived thereof (sulfoxides and/or sulfones). It is noteworthy
appearance of oxygen-containing species that structures SA-SD are non- or poorly oxidized and they do not appear to contain the
central initiator-derived segment, while the (hydrolyzed) terminal groups appear to be
(presence of peaks at +16 and +32 Da). present; this suggests that even mildly oxidized larger fragments are adsorbed irrevers-
Two sets of peaks can be recognized ibly, as previously observed for the H2O2-oxidized polymers and that only the shortest
(left inset in Figure 7), which can be fragments, that is, those between a sulfone and chain terminus, can be detected.

Macromol. Chem. Phys. 2012, 213, 2052−2061


www.MaterialsViews.com © 2012 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim 2059
Macromolecular
Chemistry and Physics P. Carampin et al.

www.mcp-journal.de

components that could be isolated are


still insoluble in water. For a qualitative
comparison, nanoparticles composed of
PPS of similar molecular weight (com-
pared to PPS3000) were prepared and
cross-linked via Michael-type addi-
tion, similarly to previous reports.[31]
The nanoparticles dispersed in water
at the concentration of 4 mg mL−1 and
oxidized with 1 equivalent of NaOCl
per thioether for 20 h at ambient tem-
perature showed a significantly higher
Figure 8. Left: FTIR spectra of PPS3000 oxidized with 5 equivalents of NaOCl in dichlo- toxicity than the oxidation products of
romethane/water 1:1 at different reaction times. No significant alteration of the spec- PPS with H2O2; their IC50 value was in
trum can be recorded at reaction times > 30 min. The presence of sulfone groups cannot the range of 1 mg mL−1 (Figure 3SI, Sup-
be unequivocally confirmed. Arrows indicate variations in diagnostic bands. The inten- porting Information). It must be noted
sity of all spectra was normalized against the νas CH3 band at 2959 cm−1. Right:1H NMR
that these values are, however, not dis-
spectra of PPS3000 before and after a 30 min oxidation with 5 equivalents of NaOCl for 20
h in dichloromethane/water 1:1. In analogy to what seen in Figure 6, we have tentatively tant from those of polymers often con-
ascribed the peak appearing at 1.66 ppm to methyl groups in beta to sulfones. sidered to be rather biocompatible, such
as chitosan.[48]

of the macromolecular structure, eventually causing


depolymerization. 4. Conclusions
Note also that (a) compounds of similar structure of
SA but featuring a sulfonic acid terminal group could The reaction of PPS with hydrogen peroxide (H2O2) in a
be also present within SA + 16 Da mass peaks; (b) the homogeneous THF/water environment appeared to have
sulfoxide in SB is placed on the terminal sulphur atom, an autoaccelerating character, suggesting that the local
under the hypothesis the vicinal oxidation may promote concentration of H2O2 is a major rate-determining factor
the sulfone cleavage; however, the same peaks may be for this oxidation. This process caused the conversion of
attributed to SC structures cationized with potassium. sulfides into sulfoxides, without significant production of
(2) NaOCl in a biphasic system. FTIR and 1H NMR spectra sulfones or depolymerization of the macromolecule; the
(Figure 8 left and right, respectively) showed signs of oxi- resulting polysulfoxides were soluble in water and charac-
dation only with a stoichiometric excess of NaOCl. As for terized by very low cytotoxicity.
the reaction in THF, all the recovered oxidation products Hypochlorite (ClO−) was used under two heterogeneous
were soluble in organic solvents. Using a large stoichio- conditions, likely differing in the local concentration of
metric excess of NaOCl (5 equivalents), sulfoxides were the oxidant; despite some differences, in both cases, the
rapidly introduced in the polymer structure (FTIR peak at reaction proceeded much more rapidly than with H2O2,
1030 cm−1, NMR resonance centered at 1.49 ppm). Smaller and provided macromolecules with clear tendency to
amounts of other oxidized species could be detected in depolymerization due to the presence of sulfones in addi-
the 1H NMR spectrum by the presence of a resonance cen- tion to sulfoxides.
tered at 1.66 ppm; although it was impossible to unequiv- The different response of polysulfides to different
ocally identify them in the FTIR spectra, we are inclined oxidants may allow different biomedical applications:
to recognize these groups as sulfones, in analogy to the for example, ClO− released during neutrophile oxida-
reaction in THF. Finally, MALDI-TOF results were virtually tive bursts may lead to an instantaneous response, for
indistinguishable from those of the product obtained in example, the release of a payload, accompanied by the
THF (Figure 3SI, Supporting Information), suggesting a rapid degradation of polysulfides carriers to low molec-
similar lability of the oxidized macromolecular structure. ular weight, water-soluble products; the latter may offer
In summary, despite requiring significantly higher the disadvantage of higher toxicity and the advantage
amounts of hypochlorite and introducing more sul- of an easier elimination through kidneys. On the other
foxide than sulfone groups, the results of the oxidation hand, the action of H2O2 may cause a sustained release of
in a biphasic system were qualitatively similar to those a payload, with a longer permanence of possibly less toxic
recorded in THF. oxidation products.
The toxicity of the material oxidized by NaOCl could We can conclude that the responsiveness of polysulfides
not be easily evaluated because the only macromolecular is sensitive to the details of the oxidative environment

Macromol. Chem. Phys. 2012, 213, 2052−2061


2060 © 2012 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim www.MaterialsViews.com
Macromolecular
Oxidant-Dependent REDOX Responsiveness of Polysulfides . . . Chemistry and Physics
www.mcp-journal.de

and this can be used to fine tune the biomedical applica- [18] G. Kilcher, L. Wang, C. Duckham, N. Tirelli, Macromolecules
tions of this class of materials. 2007, 40, 5141.
[19] A. Napoli, N. Tirelli, G. Kilcher, J. A. Hubbell, Macromolecules
2001, 34, 8913.
[20] P. Hu, N. Tirelli, React. Funct. Polym. 2011, 71, 303.
Supporting Information [21] L. Wang, P. Hu, N. Tirelli, Polymer 2009, 50, 2863.
[22] A. Rehor, N. Tirelli, J. A. Hubbell, Macromolecules 2002, 35,
Supporting Information is available from the Wiley Online Library 8688.
or from the author. [23] G. Kilcher, L. Wang, N. Tirelli, J. Polym. Sci.,Polym. Chem.
2008, 46, 2233.
[24] L. Wang, G. Kilcher, N. Tirelli, Macromol. Chem. Phys. 2009,
Acknowledgements: Financial support from EPSRC (grant
210, 447.
No. EP/C543564/1 and Advanced Research Fellowship for
[25] L. Wang, G. Kilcher, N. Tirelli, Macromol. Biosci. 2007, 7, 987.
NT) is gratefully acknowledged. J.L. thanks The Government
[26] G. Kilcher, C. Duckham, N. Tirelli, Langmuir 2007, 23, 12309.
Pharmaceutical Organization of Thailand for a research grant
[27] A. Napoli, M. J. Boerakker, N. Tirelli, R. J. M. Nolte,
and a PhD studentship.
N. Sommerdijk, J. A. Hubbell, Langmuir 2004, 20, 3487.
[28] A. Napoli, H. Bermudez, J. A. Hubbell, Langmuir 2005, 21,
Received: May 17, 2012; Revised: July 10, 2012; Published online: 9149.
August 28, 2012; DOI: 10.1002/macp.201200264 [29] A. Rehor, N. E. Botterhuis, J. A. Hubbell, N. Sommerdijk,
N. Tirelli, J. Mater. Chem. 2005, 15, 4006.
Keywords: biomaterials; depolymerization; polysulfides; redox [30] V. V. Khutoryanskiy, N. Tirelli, Pure Appl. Chem. 2007, 80,
polymers; sulfones 1703.
[31] A. Rehor, H. Schmoekel, N. Tirelli, J. A. Hubbell, Biomaterials
2008, 29, 1958.
[32] S. T. Reddy, A. J. van der Vlies, E. Simeoni, V. Angeli,
[1] A. Mantovani, P. Allavena, A. Sica, F. Balkwill, Nature 2008, G. J. Randolph, C. P. O’Neill, L. K. Lee, M. A. Swartz,
454, 436. J. A. Hubbell, Nat. Biotechnol. 2007, 25, 1159.
[2] C. Cordon-Cardo, C. Prives, J. Exp. Med. 1999, 190, 1367. [33] T. Oyama, Y. Chujo, Macromolecules 1999, 32, 7732.
[3] B. M. Babior, Blood 1999, 93, 1464. [34] K. Sato, M. Hyodo, M. Aoki, X. Q. Zheng, R. Noyori, Tetrahe-
[4] G. L. Squadrito, W. A. Pryor, Free Radical Bio. Med. 1998, 25, dron 2001, 57, 2469.
392. [35] G. Barbieri, M. Cinquini, S. Colonna, F. Montanar, J. Chem.
[5] C. J. Morris, J. R. Earl, C. W. Trenam, D. R. Blake, Int. J. Soc. C 1968, 659.
Biochem. Cell B. 1995, 27, 109. [36] M. Madesclaire, Tetrahedron 1986, 42, 5459.
[6] I. Fridovich, Arch. Biochem. Biophys. 1986, 247, 1. [37] C. Orgeretravanat, J. C. Galin, Makromol. Chem. 1988, 189,
[7] B. Halliwell, J. M. C. Gutteridge, Methods Enzymol. 1990, 186, 2017.
1. [38] J. C. Lee, M. H. Litt, C. E. Rogers, J. Polym. Sci. Polym. Chem.
[8] J. Lunec, D. R. Blake, S. J. McCleary, S. Brailsford, P. A. Bacon, 1998, 36, 793.
J. Clin. Invest. 1985, 76, 2084. [39] C. S. Marvel, E. D. Weil, J. Am. Chem. Soc. 1954, 76, 61.
[9] K. T. Oldham, K. S. Guice, G. O. Till, P. A. Ward, Surgery 1988, [40] D. C. Dittmer, G. C. Levy, J. Org. Chem. 1965, 30, 636.
104, 272. [41] H. G. Batz, V. Hofmann, H. Ringsdorf, Makromol. Chem. 1973,
[10] B. Halliwell, J. M. C. Gutteridge, D. Blake, Philos. Trans. R. Soc. 169, 323.
A 1985, 311, 659. [42] V. Hofmann, H. Ringsdorf, G. Muacevic, Makromol. Chem.
[11] C. Fang, B. Shi, Y. Y. Pei, M. H. Hong, J. Wu, H. Z. Chen, 1975, 176, 1929.
Eur. J. Pharm. Sci. 2006, 27, 27. [43] E. Lallana, T. Ferreri, S. Carroccio, A. M. Puga, N. Tirelli, Rapid
[12] A. Mercanzini, S. T. Reddy, D. Velluto, P. Colin, A. Maillard, Commun. Mass Spectrom. 2012, 26, DOI: 10.1002/rcm.6337.
J.-C. Bensadoun, J. A. Hubbell, P. Renaud, J. ControlledRelease [44] M. H. Yang, Polym. Degrad. Stabil. 2000, 68, 451.
2010, 145, 196. [45] M. J. Bowden, L. F. Thompson, J. P. Ballantyne, J. Vac. Sci.
[13] A. Napoli, M. Valentini, N. Tirelli, M. Muller, J. A. Hubbell, Technol. 1975, 12, 1294.
Nat. Mater. 2004, 3, 183. [46] F. Iacona, G. Marletta, Nucl. Instrum. Methods Phys. Res. Sect.
[14] A. Rehor, J. A. Hubbell, N. Tirelli, Langmuir 2005, 21, 411. B-Beam Interact. Mater. Atoms 2000, 166, 676.
[15] B. L. Allen, J. D. Johnson, J. P. Walker, Acs Nano 2011, 5, 5263. [47] A. Watanabe, T. Sakakibara, S. Ito, H. Ono, Y. Yoshida,
[16] N. Tirelli, Curr. Opin. Colloid Interface 2006, 11, 210. S. Tagawa, M. Matsuda, Macromolecules 1992, 25, 692.
[17] C. D. Vo, G. Kilcher, N. Tirelli, Macromol. Rapid Commun. [48] S. Ungphaiboon, D. Attia, G. G. d’Ayala, P. Sansongsak,
2009, 30, 299. F. Cellesi, N. Tirelli, Soft Matter 2010, 6, 4070.

Macromol. Chem. Phys. 2012, 213, 2052−2061


www.MaterialsViews.com © 2012 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim 2061

You might also like