Rationally Designed Redox-Sensitive Protein ... - ACS Publications

Oct 4, 2016 - were obtained from New England Biolabs (Ipswich, MA). The Pierce ... with the Tris-HCl buffer supplemented with 300 mM NaCl and 5 mM imi...
0 downloads 0 Views 3MB Size
Subscriber access provided by Columbia Univ Libraries

Article

Rationally Designed Redox-Sensitive Protein Hydrogels with Tunable Mechanical Properties Ming-Liang Zhou, Zhigang Qian, Liang Chen, David L Kaplan, and Xiaoxia Xia Biomacromolecules, Just Accepted Manuscript • DOI: 10.1021/acs.biomac.6b00973 • Publication Date (Web): 04 Oct 2016 Downloaded from http://pubs.acs.org on October 10, 2016

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 free 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 accessible to all readers and 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.

Biomacromolecules 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 30

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

1

Biomacromolecules

Submitted to Biomacromolecules as an Article

2 3 4

Rationally Designed Redox-Sensitive Protein Hydrogels with Tunable

5

Mechanical Properties

6

Ming-Liang Zhou,† Zhi-Gang Qian, † Liang Chen, † David L. Kaplan,‡ and Xiao-Xia Xia*,†

7 8 9



State Key Laboratory of Microbial Metabolism, Joint International Research Laboratory of

10

Metabolic & Developmental Sciences, and School of Life Sciences and Biotechnology, Shanghai

11

Jiao Tong University, 800 Dongchuan Road, Shanghai 200240, P. R. China

12 13



Department of Biomedical Engineering, Tufts University, 4 Colby Street, Medford,

Massachusetts 02155, United States

14 15 16 17 18

AUTHOR INFORMATION

19

Corresponding Author

20

*E-mail: [email protected].

21

Notes

22

The authors declare no competing financial interest.

1 ACS Paragon Plus Environment

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

23

ABSTRACT

24

Protein hydrogels are an important class of materials for applications in biotechnology and

25

medicine. The fine tuning of their sequence, molecular weight, and stereochemistry offers unique

26

opportunities to engineer biofunctionality, biocompatibility, and biodegradability into these

27

materials. Here we report a new family of redox-sensitive protein hydrogels with controllable

28

mechanical properties, composed of recombinant silk-elastin-like protein polymers (SELPs). The

29

SELPs were designed and synthesized SELPs with different ratios of silk-to-elastin blocks that

30

incorporated periodic cysteine residues. The cysteine-containing SELPs were thermally

31

responsive in solution and rapidly formed hydrogels at body temperature under physiologically

32

relevant, mild oxidative conditions. Upon addition of a low concentration of hydrogen peroxide

33

at 0.05% (w/v), gelation occurred within minutes for the SELPs with a protein concentration of

34

approximately 4% (w/v). The gelation time and mechanical properties of the hydrogels were

35

dependent on the ratio of silk to elastin. These polymer designs also significantly affected redox-

36

sensitive release of a highly polar model drug from the hydrogels in vitro. Furthermore, oxidative

37

gelation was performed at other physiologically relevant temperatures, and this resulted in

38

hydrogels with tunable mechanical properties, thus providing a secondary level of control over

39

hydrogel stiffness. These newly developed injectable SELP hydrogels with redox-sensitive

40

features and tunable mechanical properties may be potentially useful as biomaterials with broad

41

applications in controlled drug delivery and tissue engineering.

42 43

Keywords: silk-elastin-like protein polymers, redox-sensitive hydrogel, controllable drug

44

delivery, genetic engineering, tissue engineering

2 ACS Paragon Plus Environment

Page 2 of 30

Page 3 of 30

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

45

INTRODUCTION

46

Protein hydrogels are an important class of biomaterials for potential applications in

47

biotechnology and medicine. They are typically fabricated via physical or chemical cross-linking

48

of soluble protein polymers that form insoluble three dimensional networks with an ability to

49

retain an appreciable amount of water and mimic aspects of the microenvironment of native

50

tissue extracellular matrices.1-4 Recent trends in the design of protein hydrogels have shifted

51

from static to responsive systems, to address broader biomedical needs such as controllable drug

52

release and tissue engineering.5 Two types of dynamic, responsive protein hydrogels are of

53

particular interest. In the first scenario, protein polymers are designed to undergo dynamic

54

gelation upon triggers of physiological relevance that initiate a sol-gel phase transition.6-8 For

55

example, an injectable solution of recombinant elastin-like polypeptide was developed that

56

underwent rapid in situ gelation following intramuscular and intratumoral administration in

57

mice.9 In the second scenario, responsive protein hydrogels are designed and fabricated that can

58

significantly change their volume, shape, pore size, mechanical properties, optical transparency

59

and other properties in response to stimuli such as temperature, pH, and certain biological signals.

60

10,11

In particular, disulfide cross-linked protein hydrogels have attracted much attention because

61

they are redox-sensitive, as disulfides are rapidly reduced to thiols under the reductive

62

environment inside cells, thus allowing the quantitative release of the payload incorporated

63

within the materials. Moreover, such redox-sensitive protein hydrogels as three-dimensional cell-

64

culture scaffolds can be degraded under cytocompatible mild reductive conditions without

65

affecting the vitality of the embedded cells.12,13

66

Genetically engineered protein polymers offer a unique opportunity for the design of

67

responsive hydrogels with tunable mechanical properties, dynamic responses, and utility.14-16

3 ACS Paragon Plus Environment

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

68

First, compared to synthetic polymers, protein polymers are generally biocompatible and fully

69

degradable in vivo, which is critical for many biomedical needs.15,16 Second, nature has evolved a

70

variety of highly functional materials that are composed of unique peptide motifs or domains

71

with well-defined structures, such as α-helices, β-sheets, β-turns, and coiled coils.17-19 These

72

peptide motifs or domains can be employed either individually or in combinations for the design

73

of novel responsive hydrogels.20-22 Indeed, the conserved carboxyl-terminal domain (CTD) of

74

spider dragline silk protein, with a structure of five α-helix bundle fold, was recently fabricated

75

into protein hydrogels with dual thermosensitive behavior.23 With advances in genetic and

76

protein engineering, protein hydrogels can be modularly redesigned at the polypeptide level to

77

introduce functions by exploiting the constituent modules of fibrous proteins and even of

78

globular, structural proteins. Third, protein polymers are relatively easy to modify at the genetic

79

level for functionalization through insertion of diverse amino acids in distinct regions of the

80

polymer backbone. 24 Therefore, protein polymers are ideal for designing new stimuli-responsive

81

hydrogels and have drawn increased attention. To date, a variety of proteins have been studied for the development of responsive protein

82 83 84

hydrogels, including elastin,9,25,26 collagen,27 fibrinogen,28 silkworm silk fibroin,29 and spider silk. 23,30

Elastin and recombinant elastin-like polypeptides are particularly attractive since the

85

pioneering studies of Urry et al. provided a foundation that linked peptide sequence chemistry,

86

the coacervation of elastin-like proteins in solution, and stimuli-responses.31 While the stimuli-

87

responsive properties of elastin-like polypeptides were mostly explored when they were in

88

solution, quite a few examples have been shown where these proteins can be cross-linked for

89

temperature sensitive, soft, elastomeric matrices.6,9,25,26,32-36 However, these hydrogel

90

biomaterials are usually weak in mechanical strength9,26,32,34,35 and exhibit sensitivity to

4 ACS Paragon Plus Environment

Page 4 of 30

Page 5 of 30

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

91

restrained environmental conditions such as temperature,6,9,25,26,32-36 pH,26 salt,25 and elastolytic

92

enzymes,11 which thus limits their uses for drug delivery and tissue engineering. To overcome

93

these obstacles, we and others have recently begun to explore the feasibility of designing silk-

94

elastin-like polymers (SELPs) and SELP hydrogels with an attempt to combine diverse

95

responsive properties of elastin and high tensile strength of silk.37-39, 41-44 This is triggered by the

96

assumption that the silk blocks in SELPs might be able to crystallize into β-sheets via hydrogen

97

bonding, thus enabling robust materials formation.40 Our initial exploration demonstrated that the

98

ratio of silk-to-elastin blocks in a SELP’s repeating unit played an essential role in tuning the

99

fundamental self-assembly characteristics of these remarkable polymers in solution.41,42 Going a

100

step further, we recently developed a robust high-throughput synthesis and characterization

101

method to screen for SELPs that were responsive against diverse environmental stimuli,

102

including temperature, pH, ionic strength, redox, phosphorylation, and electric field.43 The new

103

family of SELPs with tyrosine residues in the polymer backbone was more recently cross-linked

104

with horseradish peroxidase and hydrogen peroxide to form robust hydrogels with temperature

105

responsive properties.44 However, SELP hydrogels with responsiveness other than temperature

106

remain to be fully explored, to fill the high need for such materials in biomedical-related use.

107

Herein, we report the design and fabrication of a new family of SELP hydrogels through

108

rapid gelation of the protein polymers under physiologically relevant, mild oxidative conditions.

109

First, we synthesized SELPs genetically engineered with cysteine residues in the elastin block

110

and varying the ratio of silk-to-elastin blocks in each repeating unit. The proteins were then

111

characterized for thermally-induced phase transition behavior in solution. Next, the protein

112

polymers were treated with hydrogen peroxide at concentrations as low as 0.05% (w/v) and body

113

temperature to rapidly form disulfide crosslinked hydrogels. The utility of these hydrogels for

5 ACS Paragon Plus Environment

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

114

dithiothreitol-responsive release of a model polar drug was examined in vitro. To further

115

modulate mechanical properties of the hydrogels, oxidative gelation was performed at other

116

physiologically relevant temperatures, thus providing a secondary level of control over hydrogel

117

stiffness.

118 119

MATERIALS AND METHODS

120

Materials.

121

Chemically competent cells of E. coli DH5α and BL21(DE3), TIANprep Mini Plasmid Kit and

122

TIANgel Midi Purification Kit were purchased from TIANGEN Biotech (Beijing, China).

123

Restriction enzymes, alkaline phosphatase and T4 DNA ligase were obtained from New England

124

Biolabs (Ipswich, MA). The Pierce™ BCA Protein Assay Kit was purchased from Thermo

125

Fisher Scientific Inc. (Rockford, IL). The membrane dialysis tubing with molecular weight cut

126

off (MWCO) at 3.5 kDa was obtained from Spectrumlabs (Phoenix, AZ). Ampicillin, β-

127

mercaptoethanol, dithiothreitol (DTT), hydrogen peroxide 30% (w/v) solution, imidazole and

128

isopropyl-β-ᴅ-thiogalactopyranoside (IPTG) were purchased from Sangon Biotech (Shanghai,

129

China). Ni-NTA agarose (Catalog # 30230) and rhodamine B (Catalog # 83689) were obtained

130

from Qiagen (Hilden, Germany) and Sigma (St. Louis, MO), respectively. Coomassie Brilliant

131

Blue R-250 (Catalog #161-0400) was obtained from Bio-Rad (Hercules, CA). Tryptone and

132

yeast extract were obtained from Oxoid (Basingstoke, Hampshire, UK). All other chemicals were

133

of the highest purity available from commercial suppliers.

134

Construction of Expression Plasmids.

135

The tailor-made vector pET-19b3, was employed to construct plasmids for recombinant

136

expression of the SELPs under the IPTG-inducible T7 promoter.41 A DNA sequence encoding the

6 ACS Paragon Plus Environment

Page 6 of 30

Page 7 of 30

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

137

silk-elastin-like sequence SE8C [(GVGVP)4(GCGVP)(GVGVP)3(GAGAGS)] was purchased as

138

a synthetic gene that was cloned into the EcoRV site of plasmid pUC57 from Genewiz (Suzhou,

139

China). The pUC57 derivative was digested with the restriction enzyme BanII, and the liberated

140

monomer DNA isolated by agarose gel electrophoresis and purified using the TIANgel

141

Purification Kit. The purified monomer DNA was then self-ligated by T4 DNA ligase at 16 °C

142

for 12 h. The mixture containing the resulting DNA multimers was added with the BanII- and

143

alkaline phosphatase-treated pET-19b3, and incubated at 16 °C for an additional 12 h. The

144

ligation mixture was then used to transform chemically competent cells of E. coli DH5α. The

145

transformants were selected on lysogeny broth (LB) agar plates (per liter: 10 g tryptone, 5 g

146

yeast extract, 10 g NaCl, and 15 g agar) supplemented with 50 µg mL-1 of ampicillin. The

147

transformants were then grown in the LB liquid medium for extraction of the recombinant

148

plasmids using the TIANprep Mini Plasmid Kit. The expression plasmids carrying the repetitive

149

SE8C genes of varying lengths were identified by double digest with NcoI and BamHI and

150

further confirmed by dideoxy sequencing with primers derived from the T7 promoter and T7

151

terminator sequences. Plasmid pSE8C-12 was thus obtained for the recombinant expression of

152

12 repeats of SE8C. Plasmids pS2E8C-12 and pS4E8C-11 were constructed as described

153

previously,43

154

[(GVGVP)4(GCGVP)(GVGVP)3(GAGAGS)2]

155

[(GVGVP)4(GCGVP)(GVGVP)3(GAGAGS)4], respectively.

156

Protein Expression, Purification and Identification.

157

The recombinant plasmids were transformed into the common expression host, E. coli BL21

158

(DE3), and plated on the LB agar plates with 50 µg mL-1 of ampicillin. A single colony was

159

inoculated in a 15 mL tube containing 4 mL of LB medium and cultured overnight at 37 °C and

which

allowed

expression

of and

12 11

7 ACS Paragon Plus Environment

repeats repeats

of

S2E8C

of

S4E8C

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

160

220 rpm in a rotary shaker. Unless otherwise indicated, 50 µg mL-1 of ampicillin was routinely

161

added into the culture media for the selection of plasmid-carrying recombinant cells.

162

Subsequently, 1 mL of the overnight culture was transferred into a 250 mL Erlenmeyer flask

163

containing 100 mL of the fresh LB medium. Cell growth was monitored by measuring the

164

absorbance at 600 nm (OD600) using an Eppendorf BioPhotometer plus spectrophotometer

165

(Hamburg, Germany). After cultivation for 4 h (OD600 ~3-4), the 100 mL seed culture was

166

transferred into a 2 L baffled flask containing 800 mL of the Terrific broth (per liter: 12 g

167

tryptone, 24 g yeast extract, 5 g glycerol, 2.31 g KH2PO4, 12.54 g K2HPO4). The cultures were

168

incubated at 37 °C and 220 rpm for ~6 h, shifted to 16 °C, and induced overnight with IPTG at

169

the final concentration of 1 mM. Cells were harvested by centrifugation at 9000g for 15 min at

170

10 °C. The cell pellets were resuspended in the 20 mM Tris-HCl buffer (pH 8.0) and then lysed

171

using a high pressure homogenizer (AH-1500; ATS Engineering Limited, Vancouver, Canada).

172

The homogenate was centrifuged at 9000g for 10 min at 10 °C. The resulting supernatant was

173

loaded onto a Ni-NTA agarose column that had been equilibrated with the Tris-HCl buffer

174

supplemented with 300 mM NaCl and 5 mM imidazole. The column was washed and eluted with

175

the Tris-HCl buffer containing 300 mM NaCl and imidazole at 50 mM and 250 mM, respectively.

176

To minimize disulfide bond formation, the eluted proteins of interest were added with 10 mM

177

DTT and then dialyzed against deionized water at 4 °C for 2 days, with water changes every 4-8

178

h. Following dialysis, the protein solutions were concentrated to ~50 mg mL-1 at 4 °C using

179

Amicon Ultra-0.5 mL 10K centrifugal filter devices (Millipore, Billerica, MA). The purity of the

180

purified proteins was analyzed by 10% sodium dodecyl sulfate polyacrylamide gel

181

electrophoresis (SDS-PAGE) using a gel loading buffer with 1% β-mercaptoethanol (a reducing

182

agent). The gels were stained with Coomassie Brilliant Blue R-250, and the stained gels were

8 ACS Paragon Plus Environment

Page 8 of 30

Page 9 of 30

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

183

scanned by the model GS-800 Calibrated Imaging Densitometer (Bio-Rad, Hercules, CA).

184

Protein concentrations were quantified using the BCA Protein Assay Kit with bovine serum

185

albumin as the standard. To verify molecular weights of the purified proteins, mass spectrometry

186

was performed on a SolariX-70FT-MS Bruker spectrometer (Bruker Daltonics Inc., Billerica,

187

MA). All the protein samples were freshly prepared and temporarily maintained at 4 °C in a

188

refrigerator before use.

189

Hydrogel Formation.

190

Hydrogen peroxide induced hydrogel formation was performed for the recombinant proteins in

191

glass tubes. Briefly, 90 microliters of each protein solution at 4.5% (w/v) were gently mixed with

192

10 µL of 0.5% (w/v) H2O2, and promptly incubated at 37 °C in a water bath for 10 min. The

193

proteins without H2O2 treatment were taken as controls and also incubated at 37 °C for 10 min.

194

Tubes were then taken out and inversed to evaluate the formation of a self-supporting hydrogel.

195

Photographs were quickly taken using a Canon EOS 700D camera (Canon, Tokyo, Japan).

196

Characterization of Phase Transition.

197

The phase transition behavior of the proteins was characterized by monitoring the absorbance of

198

protein solutions at 350 nm (OD350) as a function of temperature on a Shimadzu UV-2600 UV-

199

Vis spectrophotometer (Shimadzu Corp., Kyoto, Japan) equipped with a constant-temperature

200

water circulator (Model SDC-6; Scientz, Ningbo, China). Each protein solution at either 1 mg

201

mL-1 or 40 mg mL-1 was loaded in a Suprasil quartz micro cuvette (10 mm lightpath; Hellma,

202

Müllheim, Germany) and tested from 15 °C with a heating rate of 1 °C min-1. The inverse

203

transition temperature (Tt) was defined as the solution temperature corresponding to 50% of the

204

maximum value of OD350.

9 ACS Paragon Plus Environment

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 10 of 30

205

Rheological Monitoring of Gelation.

206

Rheology monitoring of gelation was performed using a stress-controlled AR-G2 rheometer (TA

207

Instruments, New Castle, DE) with a 40-mm parallel-plate configuration. A 360 µL aliquot of

208

each protein solution (4.5%, w/v) was gently mixed with 40 µL of 0.5% (w/v) H2O2, and the

209

mixture was immediately transferred onto the Peltier plate that had been precooled at 10 °C. The

210

top plate was then lowered to set the gap distance at 300 µm, and hydrogenated silicone oil was

211

added to the outer edge of the samples to minimize water evaporation. Time sweeps were carried

212

out at 37 °C with a frequency of 1 Hz and a strain of 1%, which was within the linear

213

viscoelastic range. Frequency sweep tests were performed after 30 min gelation at the indicated

214

temperatures with a constant strain of 1% and logarithmic ramping from 0.1 to 100 rad s-1.

215

Dynamic Light Scattering (DLS).

216

DLS was carried out on a Zetasizer Nano S system (Malvern Instruments, Worcestershire, UK)

217

equipped with a temperature controller. A mixture of each protein at 1 mg mL-1 and 0.05% (w/v)

218

H2O2 was introduced into quartz cuvettes and the samples stabilized at the desired temperatures

219

(10, 25, 45 or 65 °C) for 10 min prior to measurement. Number distribution data were collected

220

from three biological replicates and analyzed using the Zetasizer version 7.04 software (Malvern

221

Instruments).

222

Atomic Force Microscopy (AFM).

223

A solution of each protein was mixed with hydrogen peroxide to reach the final concentrations of

224

1 mg mL-1 and 0.05% (w/v), respectively. The mixtures were casted on mica surfaces at

225

indicated temperatures and allowed to dry for ∼2 days. AFM was performed in tapping mode

226

using a multimode Nanoscope IIIa atomic force microscope (Bruker, Germany). The silicon tip

227

probe had a spring constant of ~3 N m-1, and the scan rate was 1.97 Hz. The AFM images were

10 ACS Paragon Plus Environment

Page 11 of 30

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

228

collected with a scanning window of 10 µm and analyzed using the Nanoscope analysis v5.30

229

software (Bruker).

230

Scanning Electron Microscopy (SEM).

231

The recombinant proteins were allowed to form hydrogels at 37 °C upon H2O2 treatment as

232

described above. The hydrogels were then lyophilized with a FreeZone Plus 6 Liter cascade

233

console freeze dry system (Labconco, Kansas City, MO), and the lyophilized hydrogels were

234

sputter coated with gold for SEM observation. Images of the microstructure of the hydrogels

235

were taken using a Hitachi S-3400N scanning electron microscope (Tokyo, Japan).

236

Fourier Transform Infrared Spectroscopy (FTIR)

237

FTIR analysis of the hydrogels was carried out in transmission mode using a Nicolet 6700

238

spectrometer (Thermo Fisher Scientific Inc., Madison, WI) equipped with a deuterated triglycine

239

sulfate detector. For each measurement, the wave numbers ranged from 400 to 4000 cm-1 with a

240

resolution of 4 cm-1. The infrared spectra covering the Amide I region (1600-1700 cm-1) were

241

analyzed using the OMNIC software (Thermo Fisher Scientific). The background spectra were

242

taken under the same conditions and subtracted from each sample scan.

243

Drug Release Assay.

244

The fluorescent dye, rhodamine B, was used as a model of a highly polar drug to examine in

245

vitro drug release from the protein hydrogels. Briefly, 90 µL of each protein at 4.5% (w/v) was

246

first mixed with 5 µL of an aqueous rhodamine B solution (1 mg mL-1) and then with 5 µL of 1%

247

(w/v) H2O2 on a 96-well cell culture plate (Nest Biotechnology Co., Ltd, Wuxi, China). The

248

mixtures were incubated at 37 °C for 30 min to allow the formation of hydrogels. Rhodamine B

249

release was initiated by dropping on the hydrogel surface 200 µL of phosphate buffered saline

250

(PBS) either with or without 10 mM DTT. A 10 µL aliquot of the PBS solution was sampled

11 ACS Paragon Plus Environment

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 30

251

from the wells at the indicated time points, and an equal volume of pre-warmed fresh buffer was

252

added back to each well. Rhodamine B concentrations in the PBS solutions were quantified

253

using a standard calibration curve experimentally obtained. The rhodamine B fluorescence (553

254

nm excitation, 627 nm emission) was measured on a fluorescence microplate reader (SpectraMax

255

M5; Molecular Devices Corp., Sunnyvale, CA). Data are the average of three biological

256

replicates with standard deviation.

257 258

RESULTS AND DISCUSSION

259

Design and Biosynthesis of Cysteine-Containing SELPs.

260

Our earlier study demonstrated that a cysteine-containing SELP with silk-to-elastin block ratio at

261

1:4 displayed redox-sensitive properties in solution.43 Inspired by this observation, we

262

hypothesized that the cysteine residues in the elastin block could be potentially oxidized to form

263

intramolecular and intermolecular disulfide crosslinks leading to a supramolecular network, and

264

the covalent bonds in the resulting hydrogels might also be reduced under a mild reductive

265

condition. To test this hypothesis, a new family of hydrogels was designed based on three SELPs

266

that were genetically engineered with varying ratios of silk-to-elastin blocks at 1:8, 1:4, and 1:2,

267

and

268

(GVGVP)4(GXGVP)(GVGVP)3. Sequence features in this design included the soft elastin

269

domain providing elasticity and stimuli-sensitive properties, and the hard silk domain,

270

GAGAGS, providing tunable mechanical stiffness.

a

cysteine

residue

at

the

X

position

of

the

central

elastin

block

271

We next constructed expression plasmids encoding the desirable cysteine-containing SELPs,

272

which consisted of 12 repeats of SE8C [(GVGVP)4(GCGVP)(GVGVP)3(GAGAGS)], 12 repeats

273

of

S2E8C

[(GVGVP)4(GCGVP)(GVGVP)3(GAGAGS)2],

12 ACS Paragon Plus Environment

and

11

repeats

of

S4E8C

Page 13 of 30

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

274

[(GVGVP)4(GCGVP)(GVGVP)3(GAGAGS)4], respectively (Figure 1A). The three SELPs,

275

hereafter termed as SE8C, S2E8C and S4E8C, had a comparable number of the monomer repeats

276

and theoretical molecular weights (Figure 1B). These recombinant proteins were expressed with

277

an N-terminal tag (MGHHHHHHHHHHSSGHIDDDDKHMGAGAGS) in the expression host E.

278

coli BL21(DE3) and purified using immobilized-metal-affinity chromatography. The yield of the

279

purified SELPs was in the range of 80-100 mg L-1 of bacterial culture in shake flasks. All the

280

purified protein polymers had a purity greater than 95% confirmed by SDS-PAGE analysis

281

(Figure 1C). These proteins were further verified by mass spectrometry, displaying identified

282

molecular weights at 47.48 kDa, 52.16 kDa, and 56.89 kDa, respectively, which were all within

283

0.27% difference of the expected theoretical values (Figure S1).

284

285

Figure 1. (A) Constructs of recombinant SELPs that contain cysteine residues and varying ratios

286

of silk-to-elastin blocks in each monomer repeat. (B) The number of monomer repeats, the

287

number of amino acids, theoretical and mass spectrometry-identified molecular weight (Mw) of

288

the three SELPs. (C) Coomassie-stained 10% SDS-PAGE gel analysis of the purified SELPs.

13 ACS Paragon Plus Environment

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

289

Page 14 of 30

Thermally Responsive Behavior of the SELPs in Solution.

290

To test whether the cysteine-containing SELPs exhibited thermal transition behavior,

291

changes in optical density were monitored at 350 nm (OD350) of the protein solutions upon

292

heating from 15 to 80 °C. When SE8C, S2E8C and S4E8C were tested at a low concentration of

293

1 mg mL-1, they all exhibited an inverse phase transition typical of elastin, with Tt at

294

approximately 42 °C, 45 °C and 62 °C, respectively (Figure 2A). This indicated a link between

295

the ratio of silk-to-elastin blocks and Tt of the resulting SELPs. A likely explanation was that

296

coacervation of the elastin blocks was hampered in the presence of a higher proportion of the silk

297

blocks in the SELPs. We also examined the thermally triggered phase transition behavior for the

298

SELPs at 40 mg mL-1, a protein concentration relevant for oxidative gelation. In this scenario,

299

the sharpness of the phase transition weakened and the Tt of the three SELPs was also

300

significantly decreased (Figure 2B). SE8C had a Tt of ~30 °C, which is below body temperature

301

(37°C), whereas S2E8C and S4E8C had a Tt of 40 and 45°C, respectively. Collectively, the

302

results indicated that the three SELPs exhibited inverse phase transition and their Tt could be

303

finely modulated by adjusting molecular design and the concentrations of the protein polymers.

304 14 ACS Paragon Plus Environment

Page 15 of 30

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

305

Figure 2. Thermally responsive behavior of the protein solutions at 1 mg mL-1 (A) and 40 mg

306

mL-1 (B) as a function of temperature. Turbidity profiles were obtained by monitoring optical

307

density at 350 nm as the aqueous solutions were heated at a rate of 1 °C min-1.

308 309 310

Hydrogel Formation Under Oxidative Condition.

311

To test whether the cysteine-containing SELPs underwent oxidative gelation via disulfide bond

312

formation, solutions of the recombinant proteins were treated with an externally added oxidant.

313

H2O2 was chosen for gelation studies as it is a common, mild oxidant that is not toxic at

314

reasonable concentrations.9,44 Hydrogel formation of the three SELPs was judged initially at

315

body temperature, 37 °C, with the addition of a low concentration of H2O2 at 0.05% (w/v), which

316

was minimally required according to our preliminary experiments. All the protein polymers

317

formed self-supporting hydrogels at a protein concentration of 4.05% (w/v) within 10 min in the

318

presence of H2O2, while these proteins did not gel without H2O2 treatment (Figure 3A). In

319

addition, SE8C formed an opaque hydrogel, while S2E8C and S4E8C formed a semi-transparent

320

and transparent hydrogels, respectively. SE8C at 4.05% (w/v) should have undergone a complete

321

phase transition into coacervates at 37 °C, while the degree of coacervation was significantly

322

lower for S2E8C and S4E8C under the same conditions (Figure 2B). It appeared that

323

coacervation of the protein polymers contributed to optical transparency of the H2O2-triggered

324

hydrogels. The phenomenon coincided with that observed earlier, in which a transparent elastin-

325

like polypeptide hydrogel that was formed on ice became turbid upon incubation at a temperature

326

above the solution Tt of the polymer.9

15 ACS Paragon Plus Environment

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 16 of 30

327 328

Figure 3. (A) Formation of self-supporting SELP hydrogels. Vials containing 4.05% (w/v)

329

protein with or without 0.05% (w/v) H2O2 (as controls) were incubated at 37 °C for 10 min, and

330

then inverted for image collection. (B) Oscillatory rheological profiles for the SELP hydrogels. A

331

mixture of 4.05% (w/v) protein solution and 0.05% (w/v) H2O2 was loaded into the rheometer,

332

and time sweeps were carried out at 37 °C with a frequency of 1 Hz and a strain of 1% (linear

333

regime). Elastic modulus (G’) and loss modulus (G”) are shown as a function of time.

334 335

To study gelation kinetics and quantify the hydrogel mechanical behavior, the storage (G′)

336

and loss (G″) moduli were recorded as a function of temperature using oscillatory rheology

337

(Figure 3B). For all the SELP hydrogels, G′ values were larger than their respective G″ values

338

from the very beginning, indicating immediate formation of a network hydrogel structure. For

339

accurate estimation of the time needed to form a robust hydrogel, gelation time was defined as

340

the time when G′ reached a plateau value in the rheological studies. The gelation times for the

341

SE8C, S2E8C and S4E8C were approximately 300 s, 500 s and 900 s, respectively. Notably, the

342

H2O2-triggered SE8C hydrogel at 37 °C exhibited appreciably high elastic modulus, with plateau

343

G′ value at ~870 Pa, which was significantly higher than those of S2E8C (~470 Pa) and S4E8C

344

(~160 Pa), respectively. Taken together, the results indicated that the ratio of silk-to-elastin 16 ACS Paragon Plus Environment

Page 17 of 30

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

blocks, one of the key design parameters for the SELPs, was critical in modulating the gelation

346

time and mechanical properties of the resulting hydrogels. Previously, a family of cysteine-

347

containing elastin-like polypeptides (ELPs) with different degrees of hydrophobicity was

348

designed that displayed tunable, thermally responsive behavior, and these ELPs at 2.5 wt% could

349

be rapidly fabricated into disulfide cross-linked hydrogels with storage moduli in the range of

350

~20-200 Pa. Again, this earlier study stressed the important role of molecular design in

351

determining the polymer properties in solution and hydrogel states.

352 353

Redox-Sensitivity of the SELP Hydrogels.

354

To explore the utility of these hydrogels in biomedical needs such as controllable drug delivery,

355

we tested whether the hydrogels permitted the redox-sensitive release of small molecule drugs.

356

Rhodamine B, which is highly hydrophilic, was selected to simulate polar drugs because it

357

exhibits strong fluorescent signals, which can be accurately monitored by fluorescence

358

spectrometry.45 A mild reductive reagent, dithiothreitol (DTT), was used as the trigger to reduce

359

the disulfide bonds in the hydrogels. The kinetics of release from SE8C, S2E8C and S4E8C

360

hydrogels are shown in Figure 4, which tracks the cumulative rhodamine B released (%) as a

361

function of time at 37 °C. For all the hydrogels, the release of rhodamine B was enhanced upon

362

the introduction of 10 mM DTT, implying redox-sensitive behavior. For example, the SE8C

363

hydrogel released 44.4% of the loaded dye at 72 h with DTT in the release buffer, while this

364

hydrogel released only 28.9% in the buffer without DTT. In addition, we observed a significant

365

difference in the release rate among the three hydrogels. The SE8C hydrogel exhibited a

366

significantly lower release rate than those of S2E8C and S4E8C hydrogels, which might be

367

related with their microstructures. To verify this, we performed scanning electron microscopy

368

(SEM) on the hydrogel samples following lyophilization, a procedure generally known to have a 17 ACS Paragon Plus Environment

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 30

369

minimal impact on the structure of hydrogels.23,30 The SE8C hydrogel exhibited a more compact

370

structure with pore sizes at 20-35 μm, which coincided with its higher mechanical strength as

371

observed from the oscillatory rheology analysis (Figure 5). In contrast, S2E8C and S4E8C

372

showed a relatively loose and weak structure with larger pore sizes (40-80 µm).

373 374

Figure 4. Cumulative release of rhodamine B from the hydrogels in an in vitro assay. A mixture

375

of 50 µg mL-1 rhodamine B, 0.05% (w/v) H2O2, and protein at 4.05% (w/v) were allowed to form

376

hydrogels at 37 °C on a 96-well plate, and rhodamine B release was initiated by dropping on the

377

hydrogel surface PBS buffer either with or without 10 mM DTT.

378

379 380

Figure 5. Scanning electron microscopy (SEM) of the lyophilized hydrogels fabricated with 4.05% 18 ACS Paragon Plus Environment

Page 19 of 30

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

381

Biomacromolecules

(w/v) protein and 0.05% (w/v) H2O2 at 37 °C.

382 383

Oxidative Gelation at Diverse Temperatures Resulted in Hydrogels with Tunable

384

Mechanical Properties.

385

Having demonstrated oxidative gelation and redox-sensitivity of the resulting hydrogels at body

386

temperature, we next investigated gelation of the polymers at other physiologically relevant

387

temperatures. This was inspired by the fact that all the three SELPs were thermally responsive in

388

solution, and their coacervation states might affect disulfide crosslinking and thus the mechanical

389

properties of the resulting hydrogels. Therefore, we performed gelation of 4.05% (w/v) SELPs

390

with 0.05% (w/v) H2O2 at 25 °C, 37 °C, and 45 °C on the rheometer. After gelation equilibrium

391

for 30 min to obtain a stable network hydrogel, the rheological data were obtained by linear

392

oscillatory frequency sweep (Figure 6). As expected, the SE8C hydrogels showed superior

393

material stiffness (G′), followed by the S2E8C and S4E8C hydrogels, if the three proteins were

394

gelled at the same temperature. In addition, for each type of protein polymer, an increase in

395

gelation temperature resulted in hydrogels with elevated material stiffness. Notably, when 4.05%

396

(w/v) SE8C was gelled at 37 °C corresponding to full coacervation of the protein in solution, the

397

hydrogel stiffness was enhanced, compared with gelation at 25 °C under which the polymer was

398

partially coacervated. In addition, a further increase in gelation temperature from 37 to 45 °C

399

was beneficial for enhancing hydrogel stiffness, even though 4.05% (w/v) SE8C was fully

400

coacervated at these two temperatures. Taken together, the results demonstrated that gelation

401

temperature played a significant yet complicated role in modulating mechanical properties of the

402

hydrogels fabricated from the thermally responsive SELPs.

19 ACS Paragon Plus Environment

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 20 of 30

403 404

Figure 6. Storage moduli (G′) as a function of frequency for the SELP hydrogels formed from

405

4.05% (w/v) protein with 0.05% H2O2 (w/v) at 25 °C, 37 °C, and 45 °C, respectively.

406 407

To examine whether the silk blocks in the SELPs were directly involved in hydrogel

408

formation, we performed Fourier transform infrared spectroscopy (FTIR) analysis for the

409

lyophilized hydrogels (Figure 7). It is generally recognized that the region from 1600 to 1640

410

cm-1 of FTIR spectra is related to the intermolecular and intramolecular beta-sheet bands,

411

whereas the region between 1640 and 1660 cm-1 is associated with the presence of random coils

412

and alpha-helices.4,46 Surprisingly, the SE8C and S2E8C hydrogels fabricated at all the three

413

temperatures showed signals in the region between 1600 and 1640 cm-1, which is indicative of

414

the formation of β-sheet structures, whereas S4E8C with higher ratio of silk-blocks did not show

415

obvious β-sheet structures. This result indicated that silk crystallization also contributed to

416

formation of the SE8C and S2E8C hydrogels except disulfide crosslinking, which might explain,

417

at least partially, why the SE8C and S2E8C hydrogels showed higher material stiffness than

418

those of S4E8C (Figure 6).

20 ACS Paragon Plus Environment

Page 21 of 30

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

419 420

Figure 7. FTIR absorbance spectra of the SELP hydrogels. The hydrogels were fabricated from

421

4.05% (w/v) protein with 0.05% H2O2 (w/v) at 25 °C, 37 °C, or 45 °C for 30 min, and freeze

422

dried before FTIR analysis. The shift toward lower wavenumbers is indicative of the formation

423

of β-sheets.

424 425

Microscopic Structures of the SELPs.

21 ACS Paragon Plus Environment

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 22 of 30

426 427

Figure 8. Dynamic light scattering (DLS) size distribution profiles. A mixture of each protein at

428

1 mg mL-1 and 0.05% (w/v) H2O2 was incubated at the indicated temperatures before DLS

429

analysis.

430 431

To further explore the molecular and structural events that gave rise to the macroscopic

432

properties described above, we performed dynamic light scattering (DLS) analysis for solutions

433

of the three SELPs with H2O2 oxidation at temperatures ranging from 10 to 65 °C (Figure 8). At

434

low temperatures of 10 and 25 °C (below Tt), SE8C existed mostly as nanostructures with small

435

hydrodynamic diameter sizes of 12.98 ± 0.74 nm and 15.79 ± 1.36 nm, respectively, which was

436

suggestive of the presence of free chains in solution. Upon increasing the temperature to 45 °C

437

(above Tt), the size of SE8C particles was approximately 622 nm, indicating coacervation of this 22 ACS Paragon Plus Environment

Page 23 of 30

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

438

protein polymer. Interestingly, self-assembly of S2E8C into large particles with diameter sizes of

439

~150 nm was observed even at 10 and 25 °C, although this polymer mainly existed as free chains

440

at these low temperatures below Tt of the polymer. However, most of S2E8C was self-assembled

441

into larger nanostructures with diameters of 218.87 ± 17.59 nm as the temperature increased to

442

45 °C (around Tt). More interestingly, most of S4E8C existed as self-assembled particles with

443

diameters of 97.38 ± 10.55 nm at 10 °C, and the size distribution of this polymer exhibited as

444

similar profiles at 25 and 45 °C (below Tt). This might be due to the existence of more silk

445

blocks leading to the preassembly of the proteins into micellar-like particles at a low temperature,

446

which partially masked the effect of elevated temperature on coacervation of the elastin blocks in

447

the SELPs.

448 449

Figure 9. Representative AFM images of the cross-linked nanostructures for the SELPs under

450

oxidative condition. A mixture of each protein at 1 mg mL-1 and 0.05% (w/v) H2O2 was

451

deposited on mica surfaces and allowed to dry at the indicated temperatures before analysis.

452 23 ACS Paragon Plus Environment

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

453

To verify the microscopic structures, atomic force microscopy (AFM) analysis was

454

performed for the three SELPs under oxidative condition (Figure 9). As expected, the formation

455

of globules was observed, suggestive of micellar-like particles for S2E8C and S4E8C at 25 °C,

456

respectively, whereas SE8C did not form any obvious particles at this temperature. This result

457

stressed that the assembly capability of the cysteine-containing SELPs was dependent on the

458

ratio of silk-to-elastin blocks, which coincided well with that observed for the tyrosine-

459

containing SELPs in our earlier study.41 On the other hand, we also observed the formation of

460

large spherical particles for SE8C and S2E8C at 45 °C, which were most probably formed

461

through cross-linking of the protein coacervates. For S4E8C, no obvious large particles were

462

observed, which might be due to the fact that 45 °C was not high enough to trigger coacervation

463

of this protein.

Page 24 of 30

464

Taken together, oxidative formation of the above SELPs hydrogels at diverse temperatures

465

was complicated and affected by the dual cross-linking mechanisms of disulfide formation and

466

silk crystallization, different from the previously reported SELP hydrogels that were formed

467

from soluble protein polymers at high concentrations via silk physical cross-linking.3,10,37,38 In

468

another interesting study of Fernández-Colino et al., a dual physical gelation mechanism was

469

proposed for gelation of a silk-elastin-like corecombinamer from a 15 wt % aqueous solution of

470

the (EIS)×2 corecombinamer.4 In the first stage, a rapid, thermally driven gelation of the polymer

471

solution occurred upon an increase in temperature due to self-assembly of the elastin blocks as a

472

result of their characteristic inverse transition temperature. In the second stage, folding of the

473

elastin blocks favored the interaction between the silk blocks, which triggered the emergence and

474

maturation of irreversible β-sheet structures with silk annealing at a time scale of days to months.

475

With regard to our newly developed SELP hydrogels, an in-depth investigation of the dual cross-

24 ACS Paragon Plus Environment

Page 25 of 30

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

476

linking mechanisms at the molecular level may be necessary in future studies to further

477

understand the synergies between these interactions to gain further fine control of the assembly

478

and resulting material properties.

479 480

CONCLUSIONS

481

We have developed a new family of redox-sensitive protein hydrogels via fast oxidative gelation

482

with low concentrations of the mild oxidant, H2O2. This type of material was completely

483

composed of protein polymers that vary in the ratio of silk-to-elastin blocks and contain multiple

484

periodic cysteine residues that provide a means of chemical cross-linking through disulfide bond

485

formation. Such protein designs made it possible for the concerted action of the dynamic, elastin

486

blocks and the silk blocks. By doing so, proteins under physiologically relevant temperatures

487

displayed diverse coacervation states with varying microscopic nanostructures, which can be

488

further crosslinked into hydrogels with tunable mechanical properties and redox-responsive

489

behavior. Tunable control over the material properties is thus possible at polymer design and

490

hydrogel fabrication levels. Such cysteine-containing SELP hydrogels with redox responsiveness

491

and tunable mechanical properties are anticipated to find broader applications in drug delivery,

492

tissue engineering and regenerative medicine.

493 494

ASSOCIATED CONTENT

495

Supporting Information

496

The amino acid sequences of the three SELPs, quantification of free thiol groups (Table S1), and

497

mass spectroscopy analysis of the purified recombinant proteins (Figure S1). This material is

498

available free of charge via the Internet at http://pubs.acs.org.

25 ACS Paragon Plus Environment

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 26 of 30

499

ACKNOWLEDGEMENTS

500

Financial support was provided by the National Natural Science Foundation of China (31470216,

501

21406138, 21674061), and the Shanghai Pujiang Program (14PJ1405200 to Z.-G.Q). Support

502

from the National Institutes of Health (NIH P41 EB002520) is also greatly appreciated. The

503

authors appreciate Instrumental Analysis Center of Shanghai Jiao Tong University for allowing

504

us to use the AFM and SEM equipments.

505 506

REFERENCES

507

(1) Langer, R.; Tirrell, D. A. Nature 2004, 428, 487-492.

508

(2) Silva, R.; Fabry, B.; Boccaccini, A. R. Biomaterials 2014, 35, 6727-6738.

509

(3) Gustafson, J. A.; Ghandehari, H. Adv. Drug Deliv. Rev. 2010, 62, 1509-1523.

510

(4) Fernández-Colino, A.; Arias, F. J.; Alonso, M.; Rodríguez-Cabello, J. C. Biomacromolecules

511

2014, 15, 3781-3793.

512

(5) Stuart, M. A.; Huck, W. T.; Genzer, J.; Müller, M.; Ober, C.; Stamm, M.; Sukhorukov, G. B.;

513

Szleifer, I.; Tsukruk, V. V.; Urban, M.; Winnik, F.; Zauscher, S.; Luzinov, I.; Minko, S. Nat.

514

Mater. 2010, 9, 101-113.

515

(6) Glassman, M. J.; Olsen, B. D. Biomacromolecules 2015, 16, 3762-3773.

516

(7) Rombouts, W. H.; de Kort, D. W.; Pham, T. T.; van Mierlo, C. P.; Werten, M. W.; de Wolf, F.

517

A.; van der Gucht, J. Biomacromolecules 2015, 16, 2506-2513.

518

(8) Zhang, Y. N.; Avery, R. K.; Vallmajo-Martin, Q.; Assmann, A.; Vegh, A.; Memic, A.; Olsen,

519

B. D.; Annabi, N.; Khademhosseini, A. Adv. Funct. Mater. 2015, 25, 4814-4826.

520

(9) Asai, D.; Xu, D.; Liu, W.; Garcia Quiroz, F.; Callahan, D. J.; Zalutsky, M. R.; Craig, S. L.;

521

Chilkoti, A. Biomaterials 2012, 33, 5451-5458.

26 ACS Paragon Plus Environment

Page 27 of 30

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

522

(10) Price, R.; Poursaid, A.; Cappello, J.; Ghandehari, H. J. Control. Release. 2014, 195, 92-98.

523

(11) Bandiera, A.; Markulin, A.; Corich, L.; Vita, F.; Borelli, V. Biomacromolecules 2014, 15,

524

416-422.

525

(12) Mishina, M.; Minamihata, K.; Moriyama, K.; Nagamune, T. Biomacromolecules 2016, 17,

526

1978-1984.

527

(13) Singh, S.; Topuz, F.; Hahn, K.; Albrecht, K.; Groll, J. Angew. Chem. Int. Ed. Engl. 2013, 52,

528

3000-3003.

529

(14) Ehrick, J. D.; Deo, S. K.; Browning, T. W.; Bachas, L. G.; Madou, M. J.; Daunert, S. Nat.

530

Mater. 2005, 4, 298-302.

531

(15) Li, H.; Kong, N.; Laver, B.; Liu, J. Small 2016, 12, 973-987.

532

(16) Gomes, S.; Leonor, I. B.; Mano, J. F.; Reis, R. L.; Kaplan, D. L. Prog. Polym. Sci. 2012, 37,

533

1-17.

534

(17) Wegst, U. G.; Bai, H.; Saiz, E.; Tomsia, A. P.; Ritchie, R. O. Nat. Mater. 2015, 14, 23-36.

535

(18) Miserez, A.; Weaverc, J. C.; Chaudhurid, O. J. Mater. Chem. B 2015, 3, 13-24.

536

(19) Xia, X. X.; Qian, Z. G.; Ki, C. S.; Park, Y. H.; Kaplan, D. L.; Lee, S. Y. Proc. Natl. Acad.

537

Sci. U.S.A. 2010, 107, 14059-14063.

538

(20) Banwell, E. F.; Abelardo, E. S.; Adams, D. J.; Birchall, M. A.; Corrigan, A.; Donald, A. M.;

539

Kirkland, M.; Serpell, L. C.; Butler, M. F.; Woolfson, D. N. Nat. Mater. 2009, 8, 596-600.

540

(21) Kopeček, J.; Yang, J. Angew. Chem. Int. Ed. Engl. 2012, 51, 7396-7417.

541

(22) Schacht, K.; Scheibel, T. Curr. Opin. Biotechnol. 2014, 29, 62-69.

542

(23) Qian, Z. G.; Zhou, M. L.; Song, W. W.; Xia, X. X. Biomacromolecules 2015, 16, 3704-3711.

543

(24) DiMarco, R. L.; Heilshorn, S. C. Adv. Mater. 2012, 24, 3923-3940.

544

(25) Annabi, N.; Mithieux, S. M.; Weiss, A. S.; Dehghani, F. Biomaterials 2009, 30, 1-7.

27 ACS Paragon Plus Environment

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 28 of 30

545

(26) Lim, D. W.; Nettles, D. L.; Setton, L. A.; Chilkoti, A. Biomacromolecules 2007, 8, 1463-

546

1470.

547

(27) Rubert Pérez, C. M.; Rank, L. A.; Chmielewski, J. Chem. Commun. 2014, 50, 8174-81766.

548

(28) Elvin, C. M.; Brownlee, A. G.; Huson, M. G.; Tebb, T. A.; Kim, M.; Lyons, R. E.; Vuocolo,

549

T.; Liyou, N. E.; Hughes, T. C.; Ramshaw, J. A.; Werkmeister, J. A. Biomaterials 2009, 30, 2059-

550

2065.

551

(29) Pritchard, E. M.; Kaplan, D. L. Expert Opin. Drug Deliv. 2011, 8, 797-811.

552

(30) Schacht, K.; Scheibel, T. Biomacromolecules 2011, 12, 2488-2495.

553

(31) Urry, D. W.; Urry, K. D.; Szaflarski, W.; Nowicki, M. Adv. Drug Deliv. Rev. 2010, 62, 1404-

554

1455.

555

(32) Betre, H.; Setton, L. A.; Meyer, D. E.; Chilkoti, A. Biomacromolecules 2002, 3, 910-916.

556

(33) Trabbic-Carlson, K.; Setton, L. A.; Chilkoti, A. Biomacromolecules 2003, 4, 572-580.

557

(34) McHale, M. K.; Setton, L. A.; Chilkoti, A. Tissue Eng. 2005, 11, 1768-1779.

558

(35) Xu, D.; Asai, D.; Chilkoti, A.; Craig, S. L. Biomacromolecules 2012, 13, 2315-2321.

559

(36) Desai, M. S.; Wang, E.; Joyner, K.; Chung, T. W.; Jin, H. E.; Lee, S. W. Biomacromolecules

560

2016, 17, 2409-2416.

561

(37) Poursaid, A.; Price, R.; Tiede, A.; Olson, E.; Huo, E.; McGill, L.; Ghandehari, H.; Cappello,

562

J. Biomaterials 2015, 57, 142-152.

563

(38) Price, R.; Poursaid, A.; Cappello, J.; Ghandehari, H. J. Control. Release 2015, 213, 96-102.

564

(39) Gustafson, J. A.; Price, R. A.; Frandsen, J.; Henak, C. R.; Cappello, J.; Ghandehari, H.

565

Biomacromolecules 2013, 14, 618-625.

566

(40) Matsumoto, A.; Chen, J.; Collette, A. L.; Kim, U. J.; Altman, G. H.; Cebe, P.; Kaplan, D. L.

567

J. Phys. Chem. B 2006, 110, 21630-21638.

28 ACS Paragon Plus Environment

Page 29 of 30

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

568

(41) Xia, X. X.; Xu, Q.; Hu, X.; Qin, G.; Kaplan, D. L. Biomacromolecules 2011, 12, 3844-3850.

569

(42) Xia, X. X.; Wang, M.; Lin, Y.; Xu, Q.; Kaplan, D. L. Biomacromolecules 2014, 15, 908-914.

570

(43) Wang, Q.; Xia, X.; Huang, W.; Lin, Y.; Xu, Q.; Kaplan, D. L. Adv. Funct. Mater. 2014, 24,

571

4303-4310.

572

(44) Huang, W.; Tarakanova, A.; Dinjaski, N.; Wang, Q.; Xia, X.; Chen, Y.; Wong, J. Y.; Buehler,

573

M. J.; Kaplan, D. L. Adv. Funct. Mater. 2016, 26, 4113-4123.

574

(45) Lammel, A. S.; Hu, X.; Park, S. H.; Kaplan, D. L.; Scheibel, T. R. Biomaterials 2010, 31,

575

4583-4591.

576

(46) Hu, X.; Wang, X.; Rnjak, J.; Weiss, A. S.; Kaplan, D. L. Biomaterials 2010, 31, 8121-8131.

29 ACS Paragon Plus Environment

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

577 578

579

580

Table of Contents Graphic.

30 ACS Paragon Plus Environment

Page 30 of 30