Glucose-Responsive Polymeric Micelles via ... - ACS Publications

Jan 4, 2019 - ... Chemistry, University of Hamburg, Bundesstrasse 45, D-20146 Hamburg, Germany ... II).1,2 It requires a life-long treatment to keep g...
1 downloads 0 Views 2MB Size
Subscriber access provided by Iowa State University | Library

Article

Glucose-responsive polymeric micelles via boronic acid-diol complexation for insulin delivery at neutral pH Heba Gaballa, and Patrick Théato Biomacromolecules, Just Accepted Manuscript • DOI: 10.1021/acs.biomac.8b01508 • Publication Date (Web): 04 Jan 2019 Downloaded from http://pubs.acs.org on January 4, 2019

Just Accepted “Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted online prior to technical editing, formatting for publication and author proofing. The American Chemical Society provides “Just Accepted” as a service to the research community to expedite the dissemination of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts appear in full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been fully peer reviewed, but should not be considered the official version of record. They are citable by the Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered to authors. Therefore, the “Just Accepted” Web site may not include all articles that will be published in the journal. After a manuscript is technically edited and formatted, it will be removed from the “Just Accepted” Web site and published as an ASAP article. Note that technical editing may introduce minor changes to the manuscript text and/or graphics which could affect content, and all legal disclaimers and ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors or consequences arising from the use of information contained in these “Just Accepted” manuscripts.

is published by the American Chemical Society. 1155 Sixteenth Street N.W., Washington, DC 20036 Published by American Chemical Society. Copyright © American Chemical Society. However, no copyright claim is made to original U.S. Government works, or works produced by employees of any Commonwealth realm Crown government in the course of their duties.

Page 1 of 36 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Biomacromolecules

1

Glucose-responsive polymeric micelles via boronic

2

acid-diol complexation for insulin delivery at neutral

3

pH

4

Heba Gaballa 1 and Patrick Theato 1,2,3*

5

1

6

45, D-20146 Hamburg, Germany.

7

2

8

(KIT), Engesser Strasse. 18, D-76131 Karlsruhe, Germany.

9

3

Institute for Technical and Macromolecular Chemistry, University of Hamburg, Bundesstrasse

Institute for Chemical Technology and Polymer Chemistry, Karlsruhe Institute of Technology

Soft Matter Synthesis Laboratory, Institute for Biological Interfaces III, Karlsruhe Institute of

10

Technology (KIT), Herrmann-von-Helmholtz-Platz 1, D-76344 Eggenstein-Leopoldshafen,

11

Germany.

12

13 14

Dedicated to Prof. Dr. Rudolf Zentel on the occasion of his 65th birthday.

15

16

1 Environment ACS Paragon Plus

Biomacromolecules 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

17

ABSTRACT

18

In the present study, glucose-responsive polymeric complex micelles based on the complexation

19

of phenylboronic acid (PBA) based block copolymer and diol-functionalized polymers are

20

reported. The phenylboronic acid and diol-based block copolymers were successfully synthesized

21

in only two reaction steps using RAFT polymerization and post-polymerization modification of

22

the

23

pentafluorophenylacrylate)] poly[(AMP-b-(AMP-co-PFPA)] reactive block copolymer. The self-

24

assembly of the complex micelles was investigated under neutral conditions using DLS and TEM.

25

The interaction of the PBA block copolymer and diol-containing polymer via boronate ester bond

26

was investigated by 1H-NMR and UV-Vis spectroscopy. The complex micelles enhanced the

27

glucose responsiveness under physiological conditions compared to simple PBA micelles.

28

Furthermore, the successful glucose-triggered release of FITC-insulin from the polymeric micelles

29

was investigated.

30

INTRODUCTION

31

Diabetes mellitus (hyperglycemia) is a metabolic disorder that is characterized by high blood

32

glucose levels (Type I and Type II).1,2 It requires a life-long treatment to keep glucose

33

concentration at a normal level by monitoring glucose concentration and administrating insulin by

34

multiple daily injections. Inconvenient and painful diabetic treatments inspired scientists to exploit

35

alternative strategies for the treatment of diabetes. Consequently, glucose-responsive polymers

36

have raised great interest in the past decades because of their promising applications in self-

37

regulated insulin delivery.3–6 The common strategies that have been reported in this field are based

38

on enzymatic reactions between glucose oxidase (GODx) and glucose or carbohydrate-binding

obtained

poly[(N-acryloylmorpholine-block-(N-acryloylmorpholine-co-

2 Environment ACS Paragon Plus

Page 2 of 36

Page 3 of 36 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Biomacromolecules

39

lectin proteins, such as concanavalin A (Con A) as a complimentary binder to glucose.7–12

40

However, these protein-based systems lacked stability for long-term use and showed cytotoxicity,

41

which limited their use for clinical trials.13 To address the limitations of protein-based glucose

42

oxidase and Con (A), enzyme-free boronic acid and derivatives have been developed as synthetic

43

alternatives to these materials. For example, phenylboronic acid C6H5B(OH)2 (PBA) and

44

derivatives have proven their capability to bind with cis-1,2-diols and cis-1,3-diols of common

45

sugars, including glucose, through reversible ester formation offering a promising class of glucose-

46

responsive

47

(uncharged/hydrophobic) form and a tetrahedral (charged/hydrophilic) form in aqueous solutions

48

with a reported pKa ≈8-9.19–28 It is noteworthy that PBA-based polymers have shown a notable

49

glucose responsiveness only under high pH conditions (pH9) because of their relatively high pKa,

50

which limits their in vivo studies and drug delivery applications.29–35 Kataoka et al.36 synthesized

51

earlier chemically crosslinked hydrogels based on poly(N-isopropylacrylamide) (PNIPAm) and 3-

52

acrylamidophenylboronic acid (AAPBA), and the gels were used for a glucose-regulated release

53

of insulin. However, the addition of glucose increased the fraction of charged borate anions only

54

at pH 9. For the past several years, ongoing research on boronic acid polymers adapted many

55

strategies to enable the glucose-responsive behavior of these polymers under physiological

56

conditions.37 For example, Wulff-type styrenic monomers have demonstrated a higher affinity than

57

PBA to bind with monosaccharides at a neutral pH.38–40 However these benzoboroxle monomers

58

are either complicated to prepare or they revealed poor yields after polymerization, both limiting

59

their practical applications.41 Consequently, another strategy was adapted to lower the pKa of

60

phenylboronic acid-based polymers which consisted of introducing electron-withdrawing groups,

61

such as amino or carbonyl groups either into the polymer backbone or in the vicinity of the

materials.14–18

PBA

exists

in

an

equilibrium

3 Environment ACS Paragon Plus

between

a

triangular

Biomacromolecules 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

62

phenylboronic acid moiety to decrease the apparent pKa via the coordination between nitrogen or

63

oxygen and boron.42–47 Another reported method is the complexation of PBA-containing polymer

64

and glycopolymer resulting in the formation of boronate ester bonds, which are highly responsive

65

to physiological pH when compared to phenylboronic acid and thereby allow pH and glucose-

66

responsive drug release under physiological conditions.48–52 Shi and coworkers53,54demonstrated

67

an interesting strategy to prepare polymeric micelles based on the complexation of phenylboronic

68

acid and glycopolymers. Interestingly, the results revealed that the release of insulin could be

69

enhanced in the presence of glucose under physiological conditions due to the interaction of

70

boronic acid moieties and glycopolymer-forming boronate ester crosslinking bonds. Recently, we

71

reported on the successful synthesis of boronic acid-based polymers via RAFT polymerization and

72

post-polymerization modification techniques and studied their pH- and sugar-responsive

73

behaviors.55 However, the prepared PBA block copolymer showed a higher affinity to bind with

74

fructose than glucose under basic conditions because of their relatively high pKa (~ 9). More

75

recently, we published56 a paper on the self-assembly of a block copolymer containing

76

phenylboronic acid and glycine under physiological pH where its relative pKa was significantly

77

lowered due to the coordination between boronic acid moiety and amine and carbonyl groups in

78

the glycine segment. Polymeric micelles formed from block copolymers showed a notable glucose

79

responsiveness but only at a higher glucose concentration (~ 50 mg mL-1). Nevertheless,

80

considering the ease and success of the synthetic strategy via post-polymerization modification in

81

the preparation of well-defined boronic acid block copolymers, we considered it would be worthy

82

to utilize and optimize this strategy. In the present work, well-defined block copolymers containing

83

phenylboronic acid and diols are prepared in a facile one-pot RAFT polymerization of

84

pentafluorophenyl ester containing monomer (PFPA) to prepare a reactive precursor block

4 Environment ACS Paragon Plus

Page 4 of 36

Page 5 of 36 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Biomacromolecules

85

copolymer, which was then post-modified. Poly(acryloylmorpholine) (PAMP) is a promising

86

hydrophilic substitute to poly(ethylene glycol) (PEG) in a wide range of biomedical applications

87

because it is a water soluble, non-toxic and biocompatible polymer and can be synthesized using

88

conventional polymerization techniques.57,58Glucosamine and 3-amino-1, 2-propanediol as diol-

89

containing amines are utilized for the modification because they are available at low cost and allow

90

the preparation of water-soluble diol-containing polymers. Therefore, we expect that the

91

interaction of PBA-containing polymer and diol-functionalized block copolymers leads to the

92

formation of complex micelles through boronate ester bonds between pendent boronic acids and

93

diol moieties at neutral pH. The sizes of micelles in terms of polymer feed ratios are studied. Then,

94

the addition of different concentrations of glucose and the corresponding change in micelle

95

structure due to replacing complexation of boronic acid and diols with glucose is studied. Lastly,

96

the self-regulated insulin delivery in response to 2 g L-1 glucose (hyperglycemia) is studied under

97

physiological conditions (pH 7.4, 37oC).  

98

1. EXPERIMENTAL SECTION

99

Materials

100

Trioxane (99%), D-(-)-glucose (Glu, 99%), D-(+)-glucosamine hydrochloride (≥99%), (±)-3-

101

amino-1,2-propanediol (97%), 1,4-dioxane (99.9%) and N,N-dimethylformamide ( DMF, 99.9%)

102

were used as received from Sigma-Aldrich, unless otherwise stated. Tetrahydrofuran (THF,

103

99,9%) and n-hexane (99.9%) were purchased and used as received from VWR.

104

Azobisisobutyronitrile (AIBN, Sigma-Aldrich) was recrystallized twice from methanol and stored

105

at 4oC. 3-Aminophenylboronic acid hydrochloride (98%) was purchased from Alfa Aesar. 4-

106

Cyano-4-[(dodecylsulfanylthiocarbonyl)sulfanyl] pentanoic acid (CDTPA) was prepared

5 Environment ACS Paragon Plus

Biomacromolecules 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 6 of 36

107

following an earlier published report.59 Pentafluorophenyl acrylate was synthesized by adapting

108

the published procedure.60 N-Acryloylmorpholine (AMP, 98%, TCI) was passed over a short

109

column of basic aluminum oxide prior to use in order to remove the inhibitor. Insulin (27.5 IU mg-

110

1

111

Hydrochloric acid (HCl, 1N) was purchased from ROTH. Fluorescein 5-Isothiocyanate (FITC,

112

97%) was obtained from TCI. Triethylamine (Et3N, 99%) was purchased from Grüssing.

113

) was purchased from Serva. Sodium hydroxide (NaOH, 97%) was obtained from Geyer.

Synthesis of poly[(AMP-b-(AMP-co-PFPA)]-PF1

114

AMP (4.19 g, 14.16 mmol), CTA (59.5 mg, 014 mmol), and AIBN (1.21 mg, 0.007 mmol) were

115

placed in a Schlenk tube in a ratio of [AMP]: [CTA]: [I] = [100]: [1]: [0.05]. 1,4-dioxane (6 mL)

116

and 1,3,5-trioxane (214.2 mg) as an internal standard were added. The flask was sealed with a

117

rubber septum and the reaction mixture was subjected to 3 freeze-pump-thaw cycles and then was

118

placed into a 70oC preheated oil bath under stirring for 30 min. AMP conversion was monitored

119

by 1H-NMR spectroscopy (chloroform-d, 300 MHz). Meanwhile, PFPA (2.35 g, 8.30 mmol) was

120

purged with argon in a separate flask for 30 min. When the AMP polymerization had been stirred

121

for 30 min (22% monomer conversion calculated by 1H-NMR spectroscopy), the PFPA was

122

cannulated into the reaction vessel. After 60 min (86% AMP conversion calculated by 1H-NMR

123

spectroscopy and 80% PFPA conversion calculated by

124

removed from the heat and opened to the atmosphere, and placed into liquid nitrogen. The solution

125

was added drop-wise to cold hexane to precipitate a yellow polymer. The polymer was re-

126

precipitated into colder hexane from THF, centrifuged, and vacuum-dried, yielding a yellow

127

powder (3.5 g, yield= 83.5%).

128 129

19

F-NMR spectroscopy) the tube was

Post-modification of poly [(AMP-b-(AMP-co-PFPA)] with 3-amino-phenylboronic acid (P1), glucosamine (P2), and 3-amino-1,2-propanediol (P3) 6 Environment ACS Paragon Plus

Page 7 of 36 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Biomacromolecules

130

In a typical modification reaction, poly[(AMP-b-(AMP-co-PFPA)] 100 mg (0.0078 mmol of

131

PFPA) was dissolved in 3 mL of DMF in a small vial with a stirring bar, then 1.2 equiv. with

132

respect to PFPA of the desired amine was added into the solution to which 2 equiv. of triethylamine

133

was added. The reaction mixture was stirred at 40oC for 24 hrs. Upon completion, the solution was

134

purified by dialysis against water (MWCO 6000 Da) and then lyophilized to give the desired block

135

copolymers. Finally, the polymer samples were analyzed by 1H,

136

spectroscopy.

137

19

F NMR, and FT-IR

Preparations of block copolymer micelles.

138

First, 20 mg of block copolymer P1 was dissolved in basic water (pH 10) and then acidic water

139

(pH 2) was slowly added into the polymer solution under vigorous stirring until the appearance of

140

opalescence. This micelle solution was dialyzed against water (pH 7) with MWCO 6000 for 24

141

hrs. Finally, the micelle solution was diluted to a concentration of 1 mg mL-1 by PBS buffer (pH

142

7.4). For the preparation of complex micelles, P2 and P3 were dissolved in PBS buffer (pH 7.4)

143

with a concentration of 1.5 mg mL-1. Then a given volume of P1 basic solution was quickly added

144

into P2 and P3 under vigorous stirring. The mixed solutions became opalescent indicating the

145

formation of complex micelles. The micelle solutions were dialyzed against water and finally, the

146

solutions were diluted with PBS buffer (pH 7.4) to the desired concentrations.

147

Boronic acid-Alizarin red S (ARS) assay

148

ARS was used as a catechol model dye to study the response of the complex micelles to glucose

149

under physiological conditions. The complexation of ARS and polymeric micelles was performed

150

in PBS buffer solution (pH 7.4) at room temperature. Briefly, ARS solution with a concentration

151

of (1.0 x 10-4 M) was prepared. Polymeric micelles of P1, P1-P2, P1-P3 (1 mg mL-1, pH 7.4) were

7 Environment ACS Paragon Plus

Biomacromolecules 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

152

prepared as reported previously. The polymeric solutions were added slowly under stirring into

153

the ARS solution with a final concentration of (0.5 x 10-4 M). To study the glucose responsiveness,

154

glucose (2 mg mL-1) was added into these solutions under stirring. The solutions were then kept

155

for 12 hr to allow the glucose and the micelles to equilibrate. The UV-vis and fluorescence spectra

156

were recorded for ARS and the polymeric micelles in the absence and presence of glucose.

157

Preparation of Fluorescein-labeled insulin (FITC)

158

Based on a previous report, the preparation of FITC-insulin was followed.61 Briefly, insulin (15

159

mg, 0.0025 mmol) was added into 2 mL PBS buffer (pH 7.4) into a flask. Then fluorescein 5-

160

isothiocyanate (FITC) (2.1 mg, 0.005 mmol) was dissolved in dried DMSO (0.4 mL) and added

161

dropwise to the solution containing insulin. The mixture was stirred at 4oC under argon and

162

protected from light for 12 hrs. The solution was dialyzed against water to remove the unreacted

163

FITC. FITC-insulin was precipitated through adjustment of the pH to 4 - 5 with HCl (0.1 M) and

164

stored at 4oC.

165

Loading and release of FITC-insulin.

166

FITC-insulin (3mg) was dissolved in PBS buffer (pH 7.4) and mixed with P2 and P3 prepared

167

solutions. A basic solution of polymer P1 was prepared and then added dropwise under stirring to

168

the mixture of P2-FITC-Ins and P3-FITC-Ins until the solution changed from transparent to

169

yellowish translucent. The micelle solution was dialyzed against water with a dialysis membrane

170

(MWCO: 3500 Da) for 24 h and replaced with fresh water every 6 hrs to remove the unloaded

171

FITC-insulin. At last, FITC-loaded micelle solutions were adjusted to a concentration of 1 mg mL-

172

1

173

mL of FITC-Ins loaded micelles were sealed in a dialysis membrane with MWCO 3500 Da and

. The release profile was studied in PBS buffer solution (pH 7.4) by dialysis method. Briefly, 10

8 Environment ACS Paragon Plus

Page 8 of 36

Page 9 of 36 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Biomacromolecules

174

then dialyzed against PBS buffer (pH 7.4) with glucose concentration 2 g L-1 under gentle stirring

175

at 37oC. At each predetermined interval, 1 mL of micelle solution was drawn from the dialysis bag

176

for fluorescence measurement (Ex= 450 nm, Em= 581 nm) and replaced by 1 mL fresh solution.

177

The cumulative release was calculated according to the following formula:

178

Cumulative release % =

I0 - It x 100 % I0

179

Where I0 denotes the fluorescence intensity of FITC-insulin loaded micelles at t = 0, while It

180

denotes the fluorescence intensity at different sampling times during the release.

181

Characterizations

182

1

183

in CDCl3 and D2O/NaOD.

184

spectrometer in CDCl3 and D2O/NaOD. Chemical shifts are reported relative to CDCl3 at 7.27

185

ppm or D2O at 4.79 ppm. The molecular weight and corresponding molecular weight distribution

186

(Mw/Mn) of poly[(AMP-b-(AMP-co-PFPA)] was determined by gel permeation chromatography

187

(GPC) using polystyrene standards at room temperature using THF as eluent with a flow rate of

188

1.0 mL min-1. The instrument was equipped with an intelligent AI12 pump, an RI 101 detector,

189

two pre-columns MZ-Gel SDplus 50 × 8 mm with 50 Å and 100 Å, respectively, and a column

190

MZ-Gel SDplus 300 × 8 mm linear 5 µm. FT-IR spectra before and after post-polymerization

191

modification were recorded using the ATR (Smart iTR) unit on a Thermo Scientific Nicolet IS10

192

FT-IR spectrometer. The self-assembly behavior of block copolymer aqueous solutions and the

193

hydrodynamic diameters (Rh) and size distributions were determined by dynamic light scattering

194

(DLS). The DLS instrument consisted of a Malvern Zetasizer Nano-ZS90 apparatus equipped with

H-NMR spectra of block copolymers were recorded on a Bruker Fourier 300 NMR spectrometer 19

F-NMR spectra were obtained on a Varian Gemini 2000BB

9 Environment ACS Paragon Plus

Biomacromolecules 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

195

a He–Ne laser operated at 632 nm and the measurements were made at a scattering angle of 90°.

196

Transmission electron microscopy (TEM) measurement was performed using a Hitachi H600

197

electron microscope instrument, operated at an acceleration voltage of 100 kV. TEM samples

198

were prepared by dropping micelle solution (10 µL) onto a carbon-coated copper grid for 2 min

199

following removal of excess by a filter paper. Absorption spectra were measured on a JASCO V-

200

630 UV/Vis-photo-spectrometer, utilizing a Starna Silica (quartz) cuvette with 10 mm path

201

lengths, two faces polished. Fluorescence spectra were measured using a Jobin-Yvon-Horiba

202

Fluoromax-4 spectrofluorometer utilising Starna Silica (quartz) cuvette with 10 mm path lengths,

203

four faces polished, and the excitation wavelength was 450 nm.

204 205 206 207 208 209 210 211 212

RESULTS AND DISCUSSION

10 Environment ACS Paragon Plus

Page 10 of 36

Page 11 of 36 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Biomacromolecules

213 214

Scheme 1. Synthesis of poly [(AMP)22-b-(AMP60-co-PFPA45)] -PF1, poly[(AMP)-b-(AMP-co-

215

PBA)]-P1, poly[(AMP)-b-(AMP-co-GA)]-P2 and poly[(AMP-b-poly(AMP-co-PRD)]-P3.

216 217 218 219

Synthesis of poly [(AMP)22-b-(AMP60-co-PFPA45)]

11 Environment ACS Paragon Plus

Biomacromolecules 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 12 of 36

220

The synthetic route for the preparation of a precursor block copolymer containing reactive PFPA

221

was already reported previously by our group.62,63 However, in the present study, we modified the

222

synthesis procedure yielding a block copolymer containing PPFPA in one step to overcome any

223

unwanted side reaction of PFPA during the polymerization (Scheme 1). By this newly developed

224

methodology, poly[(AMP)-b-(AMP-co-PFPA)] was synthesized by RAFT polymerization in 1,4-

225

dioxane at 70oC by a sequential addition of PFPA. AMP monomer conversion was monitored by

226

1

227

added to the polymerization mixture for 60 more minutes. The polymerization was then quenched

228

by cooling down in liquid nitrogen and exposure to air. The calculated overall AMP conversion

229

was 86%, as calculated from 1H-NMR (DPtarget = 96, DPfirst block = 22, DPsecond block = 60), and PFPA

230

conversion had reached 80%, as calculated from

231

monomer conversion was calculated by comparing the proton signals of AMP at δ = 5.6 ppm, 6.2

232

ppm and 6.4 ppm with the proton signal of 1,3,5-trioxane at δ = 5.1 ppm. Due to overlapping

233

signals in 1H-NMR, PFPA monomer conversion was calculated by comparing the integration of

234

PFPA monomer and PPFPA polymer signals at o-(δ = −153 ppm), p-(δ = −156ppm) and m-(δ =

235

−161.8 ppm) in 19F-NMR. GPC results showed a unimodal molecular weight distribution with Mn

236

= 2.6x104 g mol-1 and a dispersity of Đ = 1.3 (See ESI S1). The yielded block copolymer was

237

purified by precipitation in n-hexane and then characterized by 1H and

238

(Figure 1) in chloroform. The characteristic peaks for both blocks could be assigned; however,

239

they were partly overlapped. Nevertheless, the chemical shifts for (-CH3) protons of the Z-end

240

group in the CTA are shown at (0.88 ppm), indicating that the CTA is still intact after

241

polymerization. In addition, the characteristics peaks of PPFPA are shown at (-153 ppm), (-156

H-NMR spectroscopy, and after 30 min 22% of AMP monomer was consumed; PFPA was then

19

F-NMR (DPtarget = 56, DP = 45). The AMP

12 Environment ACS Paragon Plus

19

F-NMR spectroscopy

Page 13 of 36 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Biomacromolecules

242

ppm) and (-161 ppm) in the 19F spectrum of the purified polymer, indicating the successful chain

243

extension with PFPA.

244 245

Figure 1. 1H- and 19F-NMR of poly[(AMP)-b-(AMP-co-PFPA)].

246 247 248

Poly[(AMP)-b-(AMP-co-PBA)]-P1,

poly[(AMP)-b-(AMP-co-GA)]-P2

poly[(AMP-b-(AMP-co-PRD)]-P3

13 Environment ACS Paragon Plus

and

Biomacromolecules 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

249

Synthesis of polymers featuring boronic acid groups is often challenging due to difficult synthesis

250

of both monomers and (co)polymers, their poor solubility in solvents, possible crosslinking during

251

polymerization and their complicated purification and characterization procedures.64 We

252

developed a straightforward route to prepare PBA-based polymers by post-polymerization

253

modification of PFPA-reactive pendant groups. Here lies the diversity and versatility of our

254

strategy as we prepared block copolymers containing PBA and diol functionalities avoiding any

255

complicated synthetic steps. All the reactions were conducted in DMF at 40oC for 24 hrs with 1.2

256

equiv. of amines. Figure 2 illustrates the successful installation of functional amines with

257

quantitative yield characterized by 1H-NMR spectroscopy. The reaction was conducted with 3-

258

aminophenyl boronic acid, glucosamine, and 3-amino-1,2-propanediol, yielding P1, P2, and P3,

259

respectively. In Figure 3, the peaks in the FT-IR spectra corresponding to PFPA ester at 1780 and

260

1540 cm-1 completely disappeared after modification in P1, P2 and P3, confirming the successful

261

substitution.

14 Environment ACS Paragon Plus

Page 14 of 36

Page 15 of 36 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Biomacromolecules

262 263

Figure 2. 1H-NMR of modified polymers: P1 (black), P2 (blue) and P3 (fuchsia).

15 Environment ACS Paragon Plus

Biomacromolecules

D

C

Transmitance (%)

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 16 of 36

B

A

4000

3500

3000

2500

2000

1500

1000

Wavenumber (cm -1) 264 265

Figure 3. FT-IR spectra (A): Block copolymer PAMP-b-(AMP-co-PFPA) (PF1) before

266

modification, (B) polymer modified with 3-aminophenyl boronic acid (P1), (C) polymer modified

267

with glucosamine (P2), and (D) polymer modified with 3-amino-1,2-propanediol (P3).

268 269

Self-assembly and the formation of complex micelles

270

Boronic acids are known to form reversible covalent bonds with cis 1,2- and cis 1,3 diol containing

271

molecules in aqueous media. In this study, the amphiphilic block copolymer P1 and the water-

272

soluble block copolymers P2 and P3 micellar solutions were prepared. We hypothesized that PBA

273

and diol-containing block copolymers could form a cross-linking via boronic ester formation with

274

the boronic acid-diol segments (Scheme 2).

275

concentration of 1 mg mL-1 at pH 7.4 below its pKa, where the PBA is in its uncharged/hydrophobic

Self-assembly of P1 was investigated at a

16 Environment ACS Paragon Plus

Page 17 of 36

276

form. DLS analysis revealed particle sizes with an average size (Dh) of ~ 171 nm, and TEM

277

imaging showed nearly spherical particles in the range of 100 – 150 nm (Figure 4) with a proposed

278

structure of hydrophobic PBA in the core and hydrophilic AMP as the outer shell. P2,P3

P1 HO

OH B OH

OH Diol

O

B

OH O

O

B

OH O

Glucose

Diol

Complexation with glucose

Complexation with diol Self-assembly

Disassembly

O N

279

O

280

Scheme 2. Schematic illustration of the formation of complex micelles of poly[(AMP)-b-(AMP-

281

co-PBA)]-P1 and poly[(AMP)-b-(AMP-co-GA)]-P2 or poly[(AMP-b-poly(AMP-co-PRD)]-P3.

(b)

(a) Number (%)

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Biomacromolecules

10

100

1000

Size (d.nm)

282 283

Figure 4. (a) DLS data showing the particle size distribution and (b) TEM image of P1 polymeric

284

micelle solution (1 mg mL-1).

285 286

Moreover, crosslinked complexes of P1-P2 and P1-P3 were characterized by 1H-NMR

287

spectroscopy in D2O (see Figure 5). The aromatic protons of PBA between 7.2-7.5 ppm of P1

17 Environment ACS Paragon Plus

Biomacromolecules 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 18 of 36

288

clearly diminished upon complexation with P2 and P3, demonstrating the formation of ester

289

bonding between PBA and diol-containing polymers. What is noteworthy is the appearance of

290

characteristic peaks of glucosamine (α,β C1H1-) at 5.3 ppm and 1, 2-propanediol (-CH2-CH-CH2)

291

at 3.4 ppm.

(a)

(b)

(c)

292 293

Figure 5. 1H NMR spectra of (a) P1 and (b) complex micelle P1-P2 and (c) complex micelle P1-

294

P3. ∗ denotes the residual solvent peak from D2O-d.

295 296

To prepare stable complex micelles, different feed ratios of P1, P2, and P3 were prepared in order

297

to study the appropriate conditions for the complexation (see Table 1). It should be noted that a

298

too high content of glucosamine in polymer P2 compared to P1 was unfavorable because unstable

299

micelles were formed with an average size ca. 50 nm (CM1_G) and could be disintegrated with

300

time. This might be due to the increase in the total hydrophilic content by diffusing more

18 Environment ACS Paragon Plus

Page 19 of 36 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Biomacromolecules

301

glucosamine chains into the micelle core and subsequently, the micelles would fall apart. However,

302

when the feed ratio was 1:1 (w/w) (CM5_G), the average size increased to ca. 200 nm and the

303

formed micelles were more stable. Nevertheless, it was observed that the size of the complex

304

micelles increased when increasing the ratio of 1, 2-propanediol content P3 with respect to P1. For

305

example, the complex feed ratio of 1:1 (w/w) (CM5_P) showed aggregation with an average size

306

of ca. 900 nm which might be attributed to the increase in the crosslinking points forming the

307

micelle core through PBA-diol complexation.

308

Table 1. Complex micelles with different compositions, all micelle solutions had a final

309

concentration of 0.5 mg mL-1.

Code

Composition (w/w)

PBA/diol(mol/mol)

Dh (nm)

CM1_G

1.0:0.2

1:6.9

50±10

CM2_G

1.0:0.4

1:3.4

95±20

CM3_G

1.0:0.6

1:2.3

80±20

CM4_G

1.0:0.8

1:1.7

130±20

CM5_G

1.0:1.0

1:1.3

200±30

CM1_P

1.0:0.2

1:2.4

50±20

CM2_P

1.0:0.4

1:1.2

80±20

CM3_P

1.0:0.6

1:0.8

400±50

CM4_P

1.0:0.8

1:0.6

1000±20

CM5_P

1.0:1.0

1:0.4

900±20

310 311

Glucose responsiveness of the complex polymeric micelles

19 Environment ACS Paragon Plus

Biomacromolecules 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

312

The most efficient conditions in micelle complex formation yielding stable micelles and less

313

polydispersity was selected for further investigations. The complex micelles formed upon the

314

complexation of P1 and P2 or P1 and P3 were monitored by DLS in terms of the hydrodynamic

315

diameter (Dh) (Figure 6 a,b). The results revealed particles with average size ca. 230 nm upon

316

complexation between P1 and P2 suggesting that the incorporated glucosamine-containing

317

polymer would form not only a cross-linked core between PBA and glucosamine units but also a

318

swollen core increasing the size of the complex micelles. Contrarily, the complex micelles of P1

319

and P3 showed a particle size of ca. 80 nm, which could be possibly due to the significant increase

320

in the hydrophilicity of the cross-linked micelles and the decrease in the hydrophobic core. The

321

observed size changes in DLS were in good agreement with the TEM images, demonstrating

322

spherical particles with a significant increase in the size of P1-P2 micelles and a decrease in the

323

size of P1-P3 micelles (Figure 6 c,d). To evaluate the glucose responsiveness of the complex

324

micelles, the particle size change was monitored by DLS upon the addition of glucose. When 2

325

mg mL-1 of glucose was added, the particle size increased to ca. 530 nm for P1-P2 and to ca. 190

326

nm for P1-P3 which could be attributed to the swelling of the hydrophobic core due to the

327

formation of boronate ester bonds between boronic acid and glucose replacing the complexation

328

between PBA and diol units. The increase in glucose concentration to 5 mg mL-1 further increased

329

the size of P1-P2 micelles to ca. 730 nm, revealing the increase in swelling degree caused by the

330

increase in the amount of boronate ester, however P1-P3 complex showed a size of ≈ 6 nm,

331

confirming the disintegration of the micelles into unimers suggesting that 5 mg mL-1 of glucose

332

was sufficient for replacing the complexation between P1-P3 by glucose inducing the cross-linked

333

core to collapse. However, the complete disintegration of P1-P2 micelles was only observed in the

334

presence of glucose with a concentration of 10 mg mL-1, leading to unimers with an average size

20 Environment ACS Paragon Plus

Page 20 of 36

Page 21 of 36 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Biomacromolecules

335

of ~ 10 nm, indicating the full disintegration of the polymeric micelles. The results revealed the

336

increased degree of complexation between P1-P2 more than P1-P3 and thus higher glucose

337

concentration was needed to allow the micelles to be fully disintegrated.

21 Environment ACS Paragon Plus

Biomacromolecules

P1+P2+ 0 mg/mL Glu P1+P2+ 2 mg/mL Glu P1+P2+ 5 mg/mL Glu P1+P2+10 mg/mL Glu

Intensity (%)

40 30 20 10

1

10

100

Dh (nm)

1000

10000

P1+P3+ 0 mg/mL Glu P1+P3+ 2 mg/mL Glu P1+P3+ 5 mg/mL Glu

40

Intensity (%)

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 22 of 36

30 20 10

1

10

100

Dh (nm)

1000

10000

338 339 340

Figure 6. DLS data in the absence and presence of glucose a) complex micelles of P1-P2, b)

341

complex micelles of P1-P3, and TEM images of c) P1-P2 and d) P1-P3.

342 343

It is also known that boronic acid can form a complex with Alizarin Red S (ARS), as shown in

344

Scheme 3. This complex could be perturbed by the addition of glucose, setting up a second

22 Environment ACS Paragon Plus

Page 23 of 36 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Biomacromolecules

345

complex between the boronic acid and the glucose-forming complex (4). Here, the complexation

346

of the phenylboronic acid-based polymer P1 and the diol-containing polymers P2 and P3 were

347

investigated by probing the spectral changes upon reaction with ARS. In a buffer solution (pH

348

7.4), ARS existed in its neutral form (1), exhibiting a maximum absorption band at ~520 nm and

349

another peak at ~ 334 nm. Upon the addition of ARS to P1, a blue shift could be observed with a

350

maximum absorption band at ~ 489 nm, confirming the binding interaction between P1 and ARS.

351

When glucose was added (2 mg mL-1), the maximum absorption band did not change from 489

352

nm, which might be attributed to the lower pH of the solution (pH 7.4) which is less than the

353

reported pKa of PBA-based polymers. Interestingly, there was a slight shift of the maximum

354

absorption band to ~ 494 nm (Figure 7A) when the pH of the solution increased to pH 9 which is,

355

indeed, enhanced by the binding affinity of glucose with PBA, leaving ARS in form (1). It has

356

been reported that the complexation of PBA with diols could significantly lower the pKa of PBA

357

based polymers through the formation of boronate ester bonding, which would enhance the

358

glucose-responsiveness under neutral conditions.14 To further investigate the influence of

359

complexation of P1 with P2 and P3, we studied their UV-Vis spectra with ARS at pH 7.4. As

360

shown in Figure 7B, the absorption band of ARS with the P1-P2 complex was recorded at ~ 486

361

nm indicating the formation of an ester bond between ARS and boronic acid, resulting in the

362

absorption shift. After the addition of glucose, a maximum absorption showed at ~ 492 nm with a

363

similar shift compared to P1 at pH 9. The results demonstrated that the addition of glucose altered

364

the interaction between boronic acid and diol units in P2 and formed a second binding between

365

the boronic acid and glucose through the reversible boronate ester formation. The spectrum of

366

ARS with the P1-P3 complex showed an absorption band at ~ 498 nm with a ~13 nm shift

367

compared with P1-P2 (Figure 7C) and a significant shift to ~ 524 nm upon the addition of glucose.

23 Environment ACS Paragon Plus

Biomacromolecules 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

368

The results revealed the higher binding affinity of P1 with P3 than ARS, not allowing PBA to

369

freely bind with ARS. Nevertheless, glucose could competitively replace the complexation of P1-

370

P3 than ARS leading to a significant shift in the spectrum. Another reason could be that the P1-

371

P2 complex micelles were entirely intact and more stable even upon the addition of 2 mg mL-1

372

glucose, while the P1-P3 complex was completely replaced upon the addition of glucose shifting

373

the absorption band of ARS. This enabled us to observe the glucose responsiveness of the complex

374

micelles under neutral conditions with an enhanced binding affinity of glucose with P1-P3

375

complex rather with P1-P2 complex. To further probe the responsiveness of complex micelles to

376

glucose, we performed fluorescence spectroscopy studies on the same solutions of ARS and

377

polymers. Interestingly, the fluorescence spectroscopic results are in a good agreement with the

378

UV-Vis results. The results in Figure S2 showed that the fluorescence intensity of ARS-P1 only

379

decreased by 22% upon the addition of glucose at pH 9 where the boronate ester can form between

380

glucose and PBA as in form 4, however in the case of P1-P2 and P1-P3 the intensity decreased by

381

12% and 51%, respectively under the physiological pH, suggesting lowering the apparent pKa of

382

the crosslinked micelles. Moreover, the results revealed the significant response of P1-P3 complex

383

to glucose which is observed by the significant decrease in the fluorescence intensity to overlap

384

with the ARS spectrum with weak fluorescent properties.

24 Environment ACS Paragon Plus

Page 24 of 36

Page 25 of 36 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Biomacromolecules

O

O

OH OH

O

1

3

O S O O Na

Glucose HO

Non-fluorescence

OH B OH

OH

OH B 4 O O

O OH B O

O

O S O Na O

2

Fluorescence

diol

385 386

Scheme 3. Schematic illustration of ARS and boronic acid complex and the interaction upon the

387

addition of glucose.

25 Environment ACS Paragon Plus

Biomacromolecules

388 (A)

(B) 0.6

0.6

ARS ARS_P1 ARS_P1_Glu_pH 7.4 ARS_P1_Glu_pH 9

0.4

520 nm

0.3

400

520 nm

0.3

0.2

486 nm

494 nm

0.1

0.0 300

0.4

489 nm

0.2

ARS ARS_P1_P2 ARS_P1_P2_Glu

0.5

Abs

0.5

Abs

500

492 nm

0.1

600

0.0 300

700

350

400

450

500

550

600

650

700

Wavelength (nm)

Wavelength (nm)

(C) 0.6

ARS ARS_P1_P3 ARS_P1_P3_Glu

0.5

0.4

Abs

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 26 of 36

520 nm

0.3

0.2

498 nm 0.1

524 nm 0.0 300

350

400

450

500

550

600

650

700

Wavelength (nm)

389 390

Figure 7. UV-vis absorption spectra of ARS upon complexation with polymeric micelle solutions

391

of (A) P1, (B) P1-P2, and (C) P1-P3. The measurements were conducted in the absence of glucose

392

and upon addition of (2 mg mL-1) glucose.

393 394

The glucose-responsive release of FITC-insulin

395

The normal glucose concentration level is (100 mg/dL, 1 g L-1), however, the typical diabetic

396

glucose level is (200 mg/dL, 2 g L-1) in hyperglycemia.54 Constant administration of insulin can

397

cause pain, inconvenience, infections, and hypoglycemia which is induced when blood glucose

398

decreases below normal levels. Thus, self-regulated insulin delivery systems are beneficial to 26 Environment ACS Paragon Plus

Page 27 of 36 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Biomacromolecules

399

overcome the inconvenience caused by injections as well as a better-controlled release and

400

enhanced bioavailability. Therefore, glucose-responsive polymeric micelles have a special

401

superiority as self-regulated insulin delivery systems, including a high encapsulation efficiency

402

and a better-controlled release of insulin in response to a change in glucose concentration.

403

Therefore, we investigated the FITC-insulin release profile of the complex micelles in response to

404

normal glucose level (1 g L-1) and a hyperglycemic level (2 g L-1) of glucose at pH 7.4 and 37oC.

405

As shown in Figure 8, in the presence of glucose with a concentration of 1 g L-1 the release profile

406

was very slow from P1-P3 complex micelles and only about 30% insulin was released in 24 h,

407

however when the glucose concentration increased to 2 g L-1, the amount of insulin released

408

significantly increased to about 90 % indicating the stability of complex micelles under normal

409

glucose concentration and better insulin release enhanced under diabetic glucose concentration.

410

As for P1-P2 complex micelles, the insulin release profile was relatively fast and not stable with

411

about 91% and 100% at glucose concentrations of 1 and 2 g L-1, respectively. This might be due

412

to the complete disintegration of the crosslinked core of P1-P2 in the presence of glucose and as a

413

consequence the fast release of insulin occurred. The results demonstrated that the release of

414

insulin from P1-P3 complex micelles was dependent on glucose concentration with better stability

415

under normal glucose concentration while the instability of P1-P2 complex micelles showed

416

similar insulin release at normal and hyperglycemic level.

27 Environment ACS Paragon Plus

100

100

80

80

Released FTIC_Insulin (%)

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Released FTIC_Insulin (%)

Biomacromolecules

60

40

20

P1_P2_Glu 2 g/L P1_P2_Glu 1 g/L

0 0

5

10

15

20

Page 28 of 36

60

40

20

P1_P3_Glu 2 g/L P1_P3_Glu 1 g/L

0

25

0

5

10

15

20

25

Release Time (h)

Release Time (h)

417 418

Figure 8. Glucose-responsive release of FITC-insulin from complex micelles (A) P1-P2, (B) P1-

419

P3. All experiments were carried out under 1 and 2 g L-1 glucose at, pH 7.4, and 37oC.

420 421

CONCLUSIONS

422

In this study, well-defined phenylboronic acid and diol-based block copolymers were successfully

423

prepared from a precursor PPFPA block copolymer. The self-assembly and complex micelles of

424

PBA and diol polymers via boronate ester bonding were investigated. It is interesting to note, that

425

the complexation of the propanediol-based block copolymer with PBA (P1-P3) displayed a notable

426

glucose responsiveness compared to the glucosamine-containing block copolymer (P1-P2) in

427

terms of self-assembly, morphology, and binding with ARS dye under physiological pH. The

428

cumulative release of FITC-insulin from P1-P3 complex micelles was successfully triggered under

429

physiological conditions with 2 g L-1 glucose (hyperglycemia) with better stability with 1 g L-1

430

(normoglycemia). Thus, we believe that resultant glucose-responsive complex micelles with better

431

sensitivity to glucose under physiological conditions may be potentially used as a promising

432

candidate for a self-regulated insulin delivery system.

433

28 Environment ACS Paragon Plus

Page 29 of 36 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Biomacromolecules

434

ASSOCIATED CONTENT

435

Supporting Information

436

Additional data on polymer characterizations (GPC analysis of poly[(N-acryloylmorpholine-

437

block-(N-acryloylmorpholine-co-pentafluorophenylacrylate)]

438

PFPA45)]) as a function of elution time and molecular weight) and Emission spectra (λex = 470

439

nm) of the ARS– polymers solutions of (A) P1, (B) P1-P2, and (C) P1-P3. The measurements were

440

conducted in the absence of glucose and upon addition of (2 mg mL-1) glucose.

(Poly[(AMP)22-b-(AMP60-co-

441 442

AUTHOR INFORMATION

443

Corresponding Author

444

* E-mail: [email protected].

445

ORCID

446

Patrick Theato: 0000-0002-4562-9254

447

Heba Gaballa: 0000-0002-1444-947X

448

Notes

449

The authors declare no competing financial interest.

450

Author Contributions

451

All authors have given approval to the final version of the manuscript.

29 Environment ACS Paragon Plus

Biomacromolecules 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

452

ACKNOWLEDGMENT

453

We thank Hamburg University for financial support during the study.

454 455

REFERENCES

456

(1) Daneman, D. Type 1 diabetes. Lancet 2006, 367, 847–858.

457 458

(2) Stumvoll, M.; Goldstein, B. J.; Haeften, T. W. Van. Type 2 diabetes: Principles of pathogenesis and therapy. Lancet 2005, 365, 1333–1346.

459 460

(3) Rege, N. K.; Phillips, N. F. B.; Weiss, M. A. Curr. Development of Glucose-Responsive “Smart” Insulin Systems. Opin. Endocrinol. Diabetes Obes. 2017, 24, 267-278.

461 462

(4) Webber, M. J.; Anderson, D. G. Smart approaches to glucose-responsive drug delivery. Journal of Drug Targeting 2015, 23, 651–655.

463 464 465

(5) Shiino, D.; Murata, Y.; Kataoka, K.; Koyama, Y.; Yokoyama, M.; Okano, T.; Sakurai, Y. Preparation and characterization of a glucose-responsive insulin-releasing polymer device. Biomaterials 1994, 15, 121–128.

466 467 468

(6) Bakh, N. A.; Cortinas, A. B.; Weiss, M. A.; Langer, R. S.; Anderson, D. G.; Gu, Z.; Dutta, S.; Strano, M. S. Glucose-responsive insulin by molecular and physical design. Nat. Chem. 2017, 9, 937–943.

469 470

(7) Zhao, L.; Wang, L.; Zhang, Y.; Xiao, S.; Bi, F.; Zhao, J.; Gai, G.; Ding, J. Glucose oxidasebased glucose-sensitive drug delivery for diabetes treatment. Polymers 2017, 9, 255.

471 472

(8) Sung, W. J.; Bae, Y. H. A glucose oxidase electrode based on polypyrrole with polyanion/PEG/enzyme conjugate dopant. Biosens. Bioelectron 2003, 18, 1231–1239.

473 474

(9) Coulibaly, F. S.; Youan, B. B. C. Concanavalin A–Polysaccharides binding affinity analysis using a quartz crystal microbalance. Biosens. Bioelectron. 2014, 59, 404–411.

475 476

(10) Zhao, L.; Xiao, C.; Wang, L.; Gai, G.; Ding, J. Glucose-sensitive polymer nanoparticles for self-regulated drug delivery. Chem. Commun. 2016, 52, 7633–7652.

477 478 479

(11) Wang, J.; Ye, Y.; Yu, J.; Kahkoska, A. R.; Zhang, X.; Wang, C.; Sun, W.; Corder, R. D.; Chen, Z.; Khan, S. A.; Buse, J. B.; Gu, Z. Core–Shell Microneedle Gel for Self-Regulated Insulin Delivery. ACS Nano 2018, 12, 2466–2473.

30 Environment ACS Paragon Plus

Page 30 of 36

Page 31 of 36 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Biomacromolecules

480 481 482

(12) Tai, W.; Mo, R.; Di, J.; Subramanian, V.; Gu, X.; Buse, J. B.; Gu, Z. Bio-Inspired Synthetic Nanovesicles for Glucose-Responsive Release of Insulin. Biomacromolecules 2014, 15, 3495– 3502.

483 484

(13) Matsumoto, A.; Ishii, T.; Nishida, J.; Matsumoto, H.; Kataoka, K.; Miyahara, Y. A synthetic approach toward a self-regulated insulin delivery system. Angew. Chemie 2012, 124, 2166–2170.

485 486

(14) Springsteen, G.; Wang, B. A detailed examination of boronic acid–diol complexation. Tetrahedron 2002, 58, 5291–5300.

487 488

(15) Ma, R.; Shi, L. Phenylboronic acid-based glucose-responsive polymeric nanoparticles: synthesis and applications in drug delivery. Polym. Chem. 2014, 5, 1503.

489 490 491 492

(16) Bull, S. D.; Davidson, M. G.; Van Den Elsen, J. M. H.; Fossey, J. S.; Jenkins, A. T. A.; Jiang, Y. B.; Kubo, Y.; Marken, F.; Sakurai, K.; Zhao, J.; James, T. D. Exploiting the Reversible Covalent Bonding of Boronic Acids: Recognition, Sensing, and Assembly. Acc. Chem. Res. 2013, 46, 312– 326.

493 494

(17) Wu, X.; Li, Z.; Chen, X.-X.; Fossey, J. S.; James, T. D.; Jiang, Y.-B. Selective sensing of saccharides using simple boronic acids and their aggregates. Chem. Soc. Rev. 2013, 42, 8032.

495 496

(18) Wang, H. C.; Lee, A. R. Recent developments in blood glucose sensors. Journal of Food and Drug Analysis, 2015, 23, 191–200.

497 498 499

(19) Kitano, S.; Koyama, Y.; Kataoka, K.; Okano, T.; Sakurai, Y. J. A novel drug delivery system utilizing a glucose responsive polymer complex between poly (vinyl alcohol) and poly (N-vinyl2-pyrrolidone) with a phenylboronic acid moiety. Control. Release 1992, 19, 161–170.

500 501 502

(20) Matsumoto, A.; Yoshida, R.; Kataoka, K. Glucose-Responsive Polymer Gel Bearing Phenylborate Derivative as a Glucose-Sensing Moiety Operating at the Physiological pH. Biomacromolecules 2004, 5, 1038–1045.

503 504

(21)Vancoillie, G.; Hoogenboom, R. Responsive Boronic Acid-Decorated (Co)polymers: From Glucose Sensors to Autonomous Drug Delivery. Sensors 2016, 16, 1736.

505 506 507

(22) Jiang, G.; Jiang, T.; Chen, H.; Li, L.; Liu, Y.; Zhou, H.; Feng, Y.; Zhou, J. Preparation of multi-responsive micelles for controlled release of insulin. Colloid Polym. Sci. 2014, 293, 209– 215.

508 509 510

(23) Zhao, L.; Ding, J.; Xiao, C.; He, P.; Tang, Z.; Pang, X.; Zhuang, X.; Chen, X. Glucosesensitive polypeptide micelles for self-regulated insulin release at physiological pH. J. Mater. Chem. 2012, 22, 12319 -12328.

31 Environment ACS Paragon Plus

Biomacromolecules 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

511 512 513

(24) Matsumoto, A.; Yamamoto, K.; Yoshida, R.; Kataoka, K.; Aoyagi, T.; Miyahara, Y. Glucosesensitive polypeptide micelles for self-regulated insulin release at physiological pH. Chem. Commun. 2010, 46, 2203.

514 515 516

(25) Xia, F.; Ge, H.; Hou, Y.; Sun, T.; Chen, L.; Zhang, G.; Jiang, L. Multiresponsive Surfaces Change Between Superhydrophilicity and Superhydrophobicity. Adv. Mater. 2007, 19, 2520– 2524.

517 518 519

(26) Vancoillie, G.; Hoogenboom, R.; Sundararaman, A.; Jäkle, F.; Jäkle, F.; Fukuda, K.; Kumaki, D.; Tokito, S.; Nishiyama, N.; Kataoka, K.; Smith, D.; Domschke, A.; Lowe, C. R. Synthesis and polymerization of boronic acid containing monomers. Polym. Chem. 2016, 7, 5484–5495.

520 521

(27) James, T. D.; Samankumara Sandanayake, K. R. A.; Shinkai, S. Saccharide Sensing with Molecular Receptors Based on Boronic Acid. Angew. Chem. Ed Eng 1996, 35, 1910–1922.

522 523 524 525

(28) Zou, J.; Zhang, S.; Shrestha, R.; Seetho, K.; Donley, C. L.; Wooley, K. L. pH-Triggered reversible morphological inversion of orthogonally-addressable poly(3-acrylamidophenylboronic acid)-block-poly(acrylamidoethylamine) micelles and their shell crosslinked nanoparticles. Polym. Chem. 2012, 3, 3146–3156.

526 527

(29) Matsumoto, A.; Ishii, T.; Nishida, J.; Matsumoto, H.; Kataoka, K.; Miyahara, Y. A synthetic approach toward a self-regulated insulin delivery system. Angew. Chem. Ed Eng 2012, 2124–2128.

528 529

(30) Krismastuti, F. S. H.; Brooks, W. L. a.; Sweetman, M. J.; Sumerlin, B. S.; Voelcker, N. H. A photonic glucose biosensor for chronic wound prognostics. J. Mater. Chem. B 2014, 2, 3972.

530 531

(31) Cambre, J. N.; Roy, D.; Sumerlin, B. S. Tuning the sugar-response of boronic acid block copolymers. J. Polym. Sci. Part A Polym. Chem. 2012, 50, 3373–3382.

532 533 534

(32) Roy, D.; Cambre, J. N.; Sumerlin, B. S. Triply-responsive boronic acid block copolymers: solution self-assembly induced by changes in temperature, pH, or sugar concentration. Chem. Commun. (Camb). 2009, 7345, 2106–2108.

535 536

(33) Yuan, W.; Shen, T.; Wang, J.; Zou, H. Formation–dissociation of glucose, pH and redox triply responsive micelles and controlled release of insulin. Polym. Chem. 2014, 5, 3968–3971.

537 538 539

(34) Brooks, W. L. A.; Vancoillie, G.; Kabb, C. P.; Hoogenboom, R.; Sumerlin, B. S. Triple responsive block copolymers combining pH-responsive, thermoresponsive, and glucoseresponsive behaviors. J. Polym. Sci. Part A Polym. Chem. 2017, 55, 2309–2317.

540 541

(35) Hoare, T.; Pelton, R. Charge-Switching, Amphoteric Glucose-Responsive Microgels with Physiological Swelling Activity. Biomacromolecules 2008, 9, 733–740.

32 Environment ACS Paragon Plus

Page 32 of 36

Page 33 of 36 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Biomacromolecules

542 543 544

(36) Kataoka, K.; Miyazaki, H.; Bunya, M.; Okano, T.; Sakurai, Y. Totally synthetic polymer gels responding to external glucose concentration: Their preparation and application to on-off regulation of insulin release. J. Am. Chem. Soc. 1998, 120, 12694-12695.

545 546 547

(37) Chai, Z.; Ma, L.; Wang, Y.; Ren, X. Phenylboronic acid as a glucose-responsive trigger to tune the insulin release of glycopolymer nanoparticles. J. Biomater. Sci. Polym. Ed. 2016, 27, 599– 610.

548 549 550

(38) Kim, H.; Kang, Y. J.; Kang, S.; Kim, K. T. Monosaccharide-Responsive Release of Insulin from Polymersomes of Polyboroxole Block Copolymers at Neutral pH. J. Am. Chem. Soc. 2012, 134, 4030–4033.

551 552

(39) Liang, L.; Liu, Z. A self-assembled molecular team of boronic acids at the gold surface for specific capture of cis-diol biomolecules at neutral pH. Chem. Commun. 2011, 47, 2255.

553 554 555

(40) Kotsuchibashi, Y.; Agustin, R. V. C.; Lu, J. Y.; Hall, D. G.; Narain, R. Temperature, pH, and Glucose Responsive Gels via Simple Mixing of Boroxole- and Glyco-Based Polymers. ACS Macro Lett. 2013, 2, 260–264.

556 557

(41) Li, H.; Liu, Y.; Liu, J.; Liu, Z. A Wulff-type boronate for boronate affinity capture of cis-diol compounds at medium acidic pH condition. Chem. Commun. 2011, 47, 8169 -8171.

558 559 560 561

(42) Pablos, J. L.; Vallejos, S.; Ibeas, S.; Muñoz, A.; Serna, F.; García, F. C.; García, J. M. Acrylic Polymers with Pendant Phenylboronic Acid Moieties as “Turn-Off” and “Turn-On” Fluorescence Solid Sensors for Detection of Dopamine, Glucose, and Fructose in Water. ACS Macro Lett. 2015, 4, 979–983.

562 563 564

(43) Okasaka, Y.; Kitano, H. Direct spectroscopic observation of binding of sugars to polymers having phenylboronic acids substituted with an ortho-phenylazo group. Colloids Surfaces B Biointerfaces 2010, 79, 434–439.

565 566 567

(44) Wang, B.; Ma, R.; Liu, G.; Li, Y.; Liu, X.; An, Y.; Shi, L. Glucose-Responsive Micelles from Self-Assembly of Poly(ethylene glycol)-b-Poly(acrylic acid-co-acrylamidophenylboronic acid) and the Controlled Release of Insulin. Langmuir 2009, 25, 12522–12528.

568 569 570

(45) Wang, B.; Ma, R.; Liu, G.; Liu, X.; Gao, Y.; Shen, J.; An, Y.; Shi, L. Effect of Coordination on the Glucose-Responsiveness of PEG-b-(PAA-co-PAAPBA) Micelles. Macromol. Rapid Commun. 2010, 31, 1628–1634.

571 572 573

(46) Farooqi, Z. H.; Wu, W.; Zhou, S.; Siddiq, M. Engineering of Phenylboronic Acid Based Glucose-Sensitive Microgels with 4-Vinylpyridine for Working at Physiological pH and Temperature. Macromol. Chem. Phys. 2011, 212, 1510–1514.

33 Environment ACS Paragon Plus

Biomacromolecules 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

574 575 576

(47) Prosperi-Porta, G.; Kedzior, S.; Muirhead, B.; Sheardown, H. Phenylboronic-Acid-Based Polymeric Micelles for Mucoadhesive Anterior Segment Ocular Drug Delivery. Biomacromolecules 2016, 17, 1449–1457.

577 578 579

(48) Lee, J.; Kim, J.; Lee, Y. M.; Park, D.; Im, S.; Song, E. H.; Park, H.; Kim, W. J. Self-assembled nanocomplex between polymerized phenylboronic acid and doxorubicin for efficient tumortargeted chemotherapy. Acta Pharmacol. Sin. 2017, 38, 848–858.

580 581 582

(49) Wang, C.; Wang, J.; Chen, X.; Zheng, X.; Xie, Z.; Chen, L.; Chen, X. Phenylboronic AcidCross-Linked Nanoparticles with Improved Stability as Dual Acid-Responsive Drug Carriers. Macromol. Biosci. 2017, 17, 1600227.

583 584 585

(50) Ren, J.; Zhang, Y.; Zhang, J.; Gao, H.; Liu, G.; Ma, R.; An, Y.; Kong, D.; Shi, L. pH/Sugar Dual Responsive Core-Cross-Linked PIC Micelles for Enhanced Intracellular Protein Delivery. Biomacromolecules 2013, 14, 3434–3443.

586 587

(51) Yao, Y.; Wang, X.; Tan, T.; Yang, J. A facile strategy for polymers to achieve glucoseresponsive behavior at neutral pH. Soft Matter 2011, 7, 7948–7951.

588 589

(52) Wang, Y.; Zhang, X.; Mu, J.; Li, C. Synthesis and pH/sugar/salt-sensitivity study of boronate crosslinked glycopolymer nanoparticles. New J. Chem. 2013, 37, 796–803.

590 591 592

(53) Ma, R.; Yang, H.; Li, Z.; Liu, G.; Sun, X.; Liu, X.; An, Y.; Shi, L. Phenylboronic Acid-Based Complex Micelles with Enhanced Glucose-Responsiveness at Physiological pH by Complexation with Glycopolymer. Biomacromolecules 2012, 13, 3409–3417.

593 594 595

(54) Yang, H.; Sun, X.; Liu, G.; Ma, R.; Li, Z.; An, Y.; Shi, L. Glucose-responsive complex micelles for self-regulated release of insulin under physiological conditions. Soft Matter 2013, 9, 8589–8599.

596 597 598

(55) Gaballa, H.; Lin, S.; Shang, J.; Meier, S.; Theato, P. A synthetic approach toward a pH and sugar-responsive diblock copolymer via post-polymerization modification. Polym. Chem. 2018, 9, 3355–3358.

599 600

(56) Gaballa, H.; Shang, J.; Meier, S.; Theato, P. The glucose-responsive behavior of a block copolymer featuring boronic acid and glycine. J. Polym. Sci. Part A Polym. Chem. 2018, 1–10.

601 602 603

(57) Torchilin, V. P.; Trubetskoy, V. S.; Whiteman, K. R.; Caliceti, P.; Ferruti, P.; Veronese, F. M. New Synthetic Amphiphilic Polymers for Steric Protection of Liposomes in Vivo. J. Pharm. Sci. 1995, 84, 1049–1053.

604 605 606

(58) Boursier, T.; Georges, S.; Mosquet, M.; Rinaldi, D.; D’Agosto, F. Synthesis of poly(Nacryloylmorpholine) macromonomers using RAFT and their copolymerization with methacrylic acid for the design of graft copolymer additives for concrete. Polym. Chem. 2016, 7, 917–925.

34 Environment ACS Paragon Plus

Page 34 of 36

Page 35 of 36 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Biomacromolecules

607 608

(59) Lin, S.; Das, A.; Theato, P. CO2-Responsive graft copolymers: synthesis and characterization. Polym. Chem. 2017, 8, 1206–1216.

609 610 611

(60)Eberhardt, M.; Mruk, R.; Zentel, R.; Théato, P. Synthesis of pentafluorophenyl(meth)acrylate polymers: New precursor polymers for the synthesis of multifunctional materials.Eur. Polym. J. 2005, 41, 1569–1575.

612 613

(61) Yao, Y.; Wang, X.; Tan, T.; Yang, J. A facile strategy for polymers to achieve glucoseresponsive behavior at neutral pH. Soft Matter 2011, 7, 7948–7951.

614 615

(62) He, L.; Shang, J.; Theato, P. Preparation of dual stimuli-responsive block copolymers based on different activated esters with distinct reactivities. Eur. Polym. J. 2015.

616 617

(63) Doncom, K. E. B.; Hansell, C. F.; Theato, P.; O’Reilly, R. K. pH-switchable polymer nanostructures for controlled release. Polym. Chem. 2012, 3, 3007.

618 619

(64) Vancoillie, G.; Hoogenboom, R. Synthesis and polymerization of boronic acid containing monomers. Polym. Chem. 2016, 20, 1602.

620 621 622 623 624 625 626 627 628 629

35 Environment ACS Paragon Plus

Biomacromolecules

630

Table of Contents graphic

631

Glucose-responsive polymeric micelles via boronic acid-diol complexation for insulin

632

delivery at neutral pH

633

Heba Gaballa and Patrick Theato

100

80

Released Insulin (%)

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 36 of 36

60

40

20

P1-P2 P1-P3

0 0

5

10

15

20

25

Release Time (h)

634 635

Complex glucose responsive polymeric micelles containing phenylboronic acid and diols were

636

prepared for self-regulated insulin delivery application.

637

KEYWORDS

638

Diabetes – Glucose responsive polymers – Phenylboronic acid polymers – Complex micelles

639

36 Environment ACS Paragon Plus