• No results found

University of Groningen Photoresponsive supramolecular soft materials in aqueous media Chen, Shaoyu

N/A
N/A
Protected

Academic year: 2021

Share "University of Groningen Photoresponsive supramolecular soft materials in aqueous media Chen, Shaoyu"

Copied!
27
0
0

Bezig met laden.... (Bekijk nu de volledige tekst)

Hele tekst

(1)

Photoresponsive supramolecular soft materials in aqueous media

Chen, Shaoyu

DOI:

10.33612/diss.107818650

IMPORTANT NOTE: You are advised to consult the publisher's version (publisher's PDF) if you wish to cite from it. Please check the document version below.

Document Version

Publisher's PDF, also known as Version of record

Publication date: 2019

Link to publication in University of Groningen/UMCG research database

Citation for published version (APA):

Chen, S. (2019). Photoresponsive supramolecular soft materials in aqueous media. Rijksuniversiteit Groningen. https://doi.org/10.33612/diss.107818650

Copyright

Other than for strictly personal use, it is not permitted to download or to forward/distribute the text or part of it without the consent of the author(s) and/or copyright holder(s), unless the work is under an open content license (like Creative Commons).

Take-down policy

If you believe that this document breaches copyright please contact us providing details, and we will remove access to the work immediately and investigate your claim.

Downloaded from the University of Groningen/UMCG research database (Pure): http://www.rug.nl/research/portal. For technical reasons the number of authors shown on this cover page is limited to 10 maximum.

(2)
(3)
(4)

Chapter 7

Biocompatible Macroscopic Strings

of Motor Amphiphiles for Cell

Growth

Manuscript in preparation:

Shaoyu Chen, Liangliang Yang, Franco King-Chi Leung,* Takashi Kajitani, Marc C.

A. Stuart, Takanori Fukushima, Chaoxia Wang, Patrick van Rijn* and Ben L.

(5)

Abstract: Macroscopic unidirectionally aligned and stimuli-responsive hierarchical

supramolecular structures have much potential in adaptive materials, such as biomedical components and soft actuating systems. However, to control the macroscopic morphological transformation in isotropic soft materials, i.e., organogels and hydrogels, remains highly challenging. Towards the design of artificial muscles and future soft robotic systems with excellent biocompatibility, we present in this chapter a macroscopic soft material based on the hierarchical supramolecular assembled structure of motor amphiphiles (MA). Detailed investigations on the photoisomerization by UV-vis absorption and NMR spectroscopies, analysis of the microscopic assembled structure by electronic microscopy, degree of alignment by X-ray diffraction, the actuation process, and biocompatibility are presented. Macroscopic motor amphiphile strings prepared from a calcium chloride solution provide a high degree of alignment and fast response actuations in water. In preliminary experiments, the MA strings show no significant cytotoxicity to the cell cultures, which cell growth observed on the MA strings surface, indicating the potential in developing future biocompatible photoresponsive soft robotic systems.

(6)

7.1 Introduction

Hierarchical supramolecular polymeric structures are commonly found in living systems to serve vital roles in key biological functions, e.g., cell division and movement.1–5 Taking inspiration from nature’s precisely controlled supramolecular

polymerization processes, numerous synthetic supramolecular polymers have been developed with tunable organic functionalization and biomimetic structural transformations.6–15 By implementation of stimuli-responsive motifs, i.e., molecular

machines, into the hierarchical supramolecular polymers, various artificial responsive functions in soft materials can be achieved ranging from the molecular level to macroscopic length-scales.10,13 Stimuli-responsive amphiphiles, at the

microscopic level, have shown the capability for assembling into one-dimensional supramolecular polymers to allow advanced functions, such as supramolecular structural transformation16–24 and control of cell growth.23,25,26 One-dimensional

supramolecular polymers can be assembled hierarchically, at the macroscopic level, into a randomly entangled three-dimensional network, with an excellent stimuli-responsiveness, allowing macroscopic gel-sol physical state transformations27–34 and

isotropic gel contractions.35,36 Alternatively, three-dimensional unidirectionally

aligned hierarchical supramolecular structures can be obtained from one-dimensional supramolecular polymers by various macroscopic alignment methodologies, e.g., shear-flow, shear force.37,38 These high anisotropic soft

materials show important opportunities towards applications in anisotropic actuators,39–42 electronic and optoelectronic materials,43–45 and biomedical

materials.46–48 We have demonstrated the first photoresponsive unidirectionally

aligned hierarchical supramolecular structures of motor amphiphiles (MA) in aqueous media, allowing a photocontrolled macroscopic muscle-like actuation in water and in air.39,41,42 This hierarchical supramolecular approach provides an

alternative methodology to macroscopic actuators based on stimuli-responsive crystals, polymeric gels, and polymeric liquid crystals. The large anisotropic macroscopic structural transformations of MA strings have shown interesting macroscopic soft robotic functions in cargo transport processes and weight lifting behaviors,39,41,42 which remains as highly challenging tasks in isotropic motions of

randomly entangled three-dimensional supramolecular networks. Although various robotic functions have been successfully demonstrated, the biocompatibility of the macroscopic MA strings remains unexplored.

On the basis of the recent developments of the cell alignments in hierarchical supramolecular structures of amphiphiles, reported by Stupp’s group,46–48

biocompatible amphiphiles have shown challenging applications in regenerative biomedical materials in arrhythmia, spinal cord injury and tissue engineered blood vessels.49,50 Low amounts of amphiphiles allow to trap cells into macroscopic gels

(7)

with a high anisotropically aligned structure, providing a mimic for circumferential alignment as seen in native arteries.49 In this chapter, we demonstrate a

biocompatible hierarchical supramolecular soft materials of MA and its potential in soft robotic materials. The MA nanofibers are assembled by a shear-flow method in a calcium chloride solution to afford unidirectionally aligned macroscopic strings. The macroscopic strings of MA are actuated upon photoirradiation, while the MA strings show no significant cytotoxicity to the cell cultures, allowing for cell growth onto the surface of the MA strings.

Figure 7.1 Schematic illustrations of (a) the reversible photoisomerization and

thermal helix inversion of molecular motor amphiphiles (MA) and (b) cell growth on the surface of macroscopic MA strings and their photoactuation process.

7.2 Results and Discussion

7.2.1 Molecular design and synthesis

Our earlier molecular motor amphiphiles,39,41 for forming macroscopic strings, were

designed with a dodecyl chain attached to the upper half of a second-generation molecular motor core and two carboxyl end-groups connected with two alkyl-linkers to the lower half of the motor core, which need the addition of 2 equiv of sodium hydroxide to allow their dissolution in aqueous media. Such alkaline condition (pH ~11) maybe unsuitable for employing them in biological systems. In this regard, the carboxyl end-groups in our earlier amphiphiles were replaced by phosphate and

(8)

sulfate end-groups, which were anticipated to enable the dissolution of the corresponding molecular motor amphiphiles (MAP1 and MAS1) in mild aqueous

media (pH ~7), enabling potential applications for cell growth (Figure 7.1). The general synthesis is summarized in Scheme 7.1.

Scheme 2.1 Synthesis of motor amphiphiles MAP1 and MAS1.

Compound 1 and 5 were prepared by our reported procedure.39 The key step in the

synthesis of MAP1 and MAS1 is the formation of the central overcrowded olefin bond

by diazo-thioketone coupling. The precursor 3 was obtained by a Williamson ether formation of thioxanthone 1 with alkyl bromide 2 in the presence of K2CO3 in DMF,

followed by conversion into the thioketone 4 with Lawesson’s reagent in toluene. Hydrazone 5 was in-situ oxidized with diacetoiodobenzene in DMF into the corresponding diazo compound, and subsequent addition of the freshly prepared thioketone 4 provided the corresponding episulfide 6. Desulfurization with triphenylphosphine in toluene gave the overcrowded alkene 7. Hydrolysis of 7 in the presence of NaOH in water and THF gave overcrowded alkene 8. The motor amphiphiles were functionalized via two different synthetic pathways. MAS1 was

(9)

obtained by a sulfate ester formation of 8 with sulfur trioxide pyridine complex, while the MAP1 was provided by phosphonylation of 8 with diphenyl phosphite and

subsequent hydrolysis. The detailed procedures of synthesis and the structures of all new compounds, unambiguously determined by 1H, 13C NMR and high-resolution

ESI mass spectrometry, are provided in the section of Experimental Data.

7.2.2 Photoisomerization of motor amphiphiles

The photochemical and thermal isomerization steps of MAP1 and MAS1 were

examined by 1H NMR and UV-vis spectroscopy. Essentially identical 1H NMR

signal shifts are observed in CD2Cl2 solutions of MAP1 and MAS1 (Figure 7.2), and

upon extending irradiation time, photostationary states (PSS) with an unstable/stable isomer ratio of 85:15 are formed in the CD2Cl2 solutions of MAP1 and MAS1. In the

UV-vis absorption spectra of CH3CN solutions of MAP1 and MAS1, an increase in

the absorption around 310 nm with a concomitant decrease of the absorption band from 330 nm to 370 nm is observed upon irradiating with 365 nm light, which is essentially identical to that observed in MAC10,39,41 indicating the isomerization from

stable configuration to unstable configuration (Figure 7.3). Additionally, an isosbestic point at 327 nm over the course of irradiation indicates that a selective photoisomerization process occurs (Figure 7.3). The transformation from unstable isomer to stable isomer can be induced by heating. The thermal helix inversion processes of MAP1 and MAS1 in CH3CN solutions were studied by means of Eyring

analysis (Figure 7.4). The activation parameters and half-life of MAP1 and MAS1 are

presented in Table 7.1. For example, a Gibbs free energy of activation (Δ‡Go) of

MAP1 was 102.49 kJ mol–1, which corresponded to a half-life (t1/2) of 40.1 h at 25 °C.

Figure 7.2 Selected parts of 1H NMR spectra (CD

2Cl2, 25 °C, 500 MHz) of (a) MAP1

and (b) MAS1 in stable state (black) and a photostationary state mixture (red)

containing 85% unstable isomers after irradiation. For the proton assignment, see Figure 7.1.

(10)

Figure 7.3 UV-vis absorption spectra of (a) MAP1 and (b) MAS1 in CH3CN solutions (6.5x10–5 M) before 365 nm light irradiation (black), upon irradiated for 1 min to 3 min (yellow) and after irradiation (red).

Figure 7.4 Kinetic studies of the thermal helix inversion step of (a) MAP1 and (b) MAS1

in CH3CN solutions (55 °C, 60 °C, 65 °C, 70 °C and 75 °C) by UV-vis absorption spectral changes at 342 nm.

Table 7.1. Activation parameters and half-lives.

Samples t1/2 at 293.15 K (h) Δ (kJ/mol) Δ (kJ/mol) Δ (J/K/mol) MAP1 ~40.1 102.50.9 48.40.7 -181.52.1 MAS1 ~77.3 103.83.8 59.92.8 -147.28.4

7.2.3 Supramolecular assembly of motor amphiphiles

A freshly prepared aqueous solution of MAP1 was heated at 80 °C for 10 min and

cooled down to room temperature to afford a colorless transparent solution, indicating that the MAP1 was soluble up to a concentration of 180 mM. Notably, MAP1 also showed excellent solubility in tris-buffer (pH 7.4) by the identical

(11)

preparation as mentioned. A Nile red fluorescence assay (NRFA), which probes the internal hydrophobicity of the assemblies,51 revealed a decrease in blue shift when

diluting beyond 0.01 mM. The critical aggregation concentration (CAC) of MAP1 is

determined as 0.76 μM (Figure 7.5a). The MAP1 assemblies were imaged using

cryogenic transmission electronic microscopy (cryo-TEM) to capture their solution-state morphologies. MAP1 assembles into fibers from hundreds of nanometers to

micrometers in length at 3.9 mM concentration (Figure 7.6a). An aqueous solution of MAS1 was prepared using the identical procedure to that for the preparation of a

solution of MAP1. The maximum solubility of MAS1 in water and tris-buffer (pH 7.4)

were identical (180 mM). A higher CAC is observed for MAS1 (1.51 M, Figure 7.5b)

than the one observed for MAP1 (Figure 7.5a). MAS1 assembles into vesicles of

20~30 nm in diameter at a concentration of 3.9 mM (Figure 7.6b).

Figure 7.5 Nile Red fluorescence assay for the determination of critical aggregation

concentration of (a) MAP1 and (b) MAS1 (concentration: 5.0 x 10–7 to 2.0 mM).

Figure 7.6 Cryo-TEM images of aqueous solutions of (a) MAP1 (3.9 mM, above CAC)

(12)

A macroscopic string of MAP1 was prepared according to our reported

procedure.39,41,42 Typically, an aqueous solution of MA

P1 (50 mM) was manually

drawn into an aqueous solution of CaCl2 (150 mM) from a pipette, and a noodle-like

string with an arbitrary length was formed. Scanning electronic microscopy image (SEM) of the string, prepared from the solution of CaCl2, shows arrays of

unidirectionally aligned nanofiber bundles (Figure 7.7a), which is essentially identical to that of the MAC10 string prepared from a CaCl2 solution. No macroscopic

string was observed using the shear-flow method with an aqueous solution of MAS1

(50 mM), i.e., a direct dissolution of MAS1 was observed. The freshly prepared string

of MAP1 from the CaCl2 solution shows uniform birefringence in the direction of the

string long axis in polarized optical microscopy (POM) images (Figure 7.7b), which is essentially identical to POM images of that of the MAC10 string. To provide the

structural parameters and orientational order, i.e., degree of alignment, of the MAP1

nanofibers in the macroscopic strings, we carried out through-view small-angle X-ray scattering (SAXS) measurements. In the 2D SAXS image of the MAP1 string,

prepared from the CaCl2 solution on a sapphire substrate at 25 °C (Figure 7.7c), a

pair of spot-like scatterings is observed in a smaller-angle region (q = 0.1–0.45 nm– 1) (Figure 7.7c, inset), which is due to scattering from the unidirectionally aligned

nanofiber bundles. The diffraction arcs with d-spacing of 6.15 nm (Figure 7.7d), are attributed to the diffraction from (001) plane of a lamellar structure, which is constructed by the unidirectionally aligned nanofibers of MAP1 with ionic interaction

between Ca2+ and phosphites of MA

P1 as interfibrillar interaction. The layer spacing

of the lamellar structure (c = 6.15 nm) of the MAP1 prepared from CaCl2 solution is

longer than that observed in a MAC10 string prepared from the CaCl2 solution (c =

5.48 nm), possibly due to a loose packing between MAP1 nanofibers and Ca2+. The

angular dependency of the peak intensity of the diffraction from the (001) plane converted from the through-view 2D SAXS image of the MAP1 string prepared from

the CaCl2 solution (Figure 7.7c) shows intensity maxima at 0° and 180°. The peak

intensity of the diffraction from the (001) plane was quantified by full-width half-maximum (fwhm) to obtain an ~100° azimuthal angle, with a smaller azimuthal angle representing a larger degree of unidirectional alignment. Given that MAC10

string prepared from the CaCl2 solution showed an ~65° azimuthal angle, the results

indicated that a lower degree of unidirectional alignment was present in the MAP1

string. With the good degree of alignment in the MAP1 string, a fast response

macroscopic photoactuation is expected, indeed, a MAP1 string prepared in a cuvette

containing CaCl2 solution bent towards light source from an initial position of 0° to

a flexion angle (90°), upon 365 nm irradiation for 60 s, with an actuation speed of 1.5 °/s (Figure 7.8).

(13)

Figure 7.7 (a) SEM and (b) POM images of a macroscopic aligned string composed

of MAP1 prepared from solutions of CaCl2 (150 mM) under crossed polarizers. The POM images of the string was tilted at 45°, 135°, 225°, and 315° relative to the transmission axis of the analyzer. Scale bar applied for all panels. (c) 2D SAXS image of a macroscopic aligned string composed of MAP1 (inset: enlarged 2D image for q

= 0.1–0.45 nm–1 at 25 °C. (d) 1D SAXS patterns of a macroscopic aligned string composed of MAP1 of 2D SAXS images in (c), showing the diffraction pattern in the

direction perpendicular to long axis of the string.

Figure 7.8 Photoisomerization step of MAP1 and photoactuation of a MAP1 string

prepared from a CaCl2 solution, after irradiation with 365 nm light source for 60 s. Scale bar, 5.0 mm.

(14)

7.2.4 Biocompatibility and cell growth onto the macroscopic strings

To investigate the biocompatibility of the motor amphiphiles (MA), including

MAC10 and MAP1, human bone marrow-derived mesenchymal stem cells (hBM-MSCs) were cultured in solutions with different concentrations of MAs and

analyzed using an XTT assay.52 Freshly prepared aqueous solutions of MA C10 (50

mM) or MAP1 (50 mM) were diluted with growth medium to obtain a range of

concentrations from 0.5 μM (below CAC) to 2200 μM (above CAC). A solution mixture of hBM-MSCs (cell density: 2500 cells per well) and growth medium (100 μL, containing MAs at a range of concentration: 0.5 μM, 5 μM and 2200 μM) was placed in 96-well plates. The mixture was incubated at 37 oC at 5% CO

2 for 24 h,

and subsequently the growth media were replaced by the fresh media (100 μL). Freshly prepared solutions (50 μL) of XTT reagent and activation solution (v/v = 50/1) were added to the well plates and incubated at 37 oC in a humidified

atmosphere at 5% CO2 for 3 h. During the incubation time, an orange color is formed,

the intensity of which can be measured by a microplate reader. The intensity of the dye is proportional to the number of metabolically active cells, i.e., the greater the number of metabolically active cells in the well, the greater the activity of mitochondrial enzymes, and the higher the concentration of the dye formed.52

Therefore, the dye absorbance is proportional to the number of cells in each well. The absorbances are plotted as a function of the concentration of motor amphiphiles (Figure 7.9). Compared to the absorbance of the sample without MAs, no significant differences of absorbances in the samples containing MAC10 or MAP1 at a

concentration lower than 5 μM were found, indicating that both MAC10 and MAP1

showed no obvious cytotoxicity to the hBM-MSCs.

Figure 7.9 Cytotoxicity measurements of MAC10 or MAP1 solutions. The human bone

marrow derived mesenchymal stem cells (hBM-MSCs) were cultured in the solutions with different concentrations of MAs and analyzed using an XTT assay.

(15)

To provide a deep insight in the biocompatibility of MAC10 and MAP1, preliminary

studies towards the growth conditions of hBM-MSCs onto the material surface of macroscopic strings of MAC10 and MAP1 were performed by confocal fluorescence

microscopy. Freshly prepared macroscopic strings of MAC10 and MAP1 were placed

in 6-well plates and washed with growth media (0.5 mL) for 3 times. Subsequently, the hBM-MSCs (cell density: 2500 cells per well) were seeded into the medium with

MA strings for cell growth. The samples were incubated at 37 oC and 5% CO 2 for

72 h. The hBM-MSCs were fixed with 3.7% paraformaldehyde in phosphate-buffered saline (PBS) for 20 min at room temperature, and subsequently washed 3 times with PBS. Subsequently, the cell membrane was permeabilized with a 0.5% TritonX-100 solution for 3 min. The obtained hBM-MSCs were stained with 4′,6-diamidino-2-phenylindole (DAPI) and tetramethylrhodamine isothiocyanate(TRITC)–phalloidin for the cell nuclei and F-actin, respectively, and observed using a confocal laser scanning microscopy (CLSM).53 The fluorescence

images are showed in Figure 7.10.

Figure 7.10 Fluorescence images of hBM-MSCs growth in the present of

macroscopic strings of MAC10 and MAP1 for 72 h. (a) Schematic illustration of

fluorescence images taken by changing the focus positions from the bottom of the strings, i.e., the surface of the well plate, to the middle and to the surface of the strings. The images of a MAC10 string by changing the focus positions from (b)

bottom, to (c) middle and (d) surface. The images of a MAP1 string by changing the

focus positions from (e) bottom, to (f) middle and (g) surface. hBM-MSCs are stained using phalloidin for F-actin (red) and DAPI for the nucleus (blue), Scale bar: 100 μm. It is generally accepted, with few exceptions, that a small, round cell shape is typically indicative of a cell entering apoptosis,54,55 whereas a well-spread and

aligned cell shape is most often quantified as being in a viable state.53,56,57 As depicted

in Figure 7.10, cells spread with well-defined actin stress fibers onto the surfaces of well plates in the present of the MA strings (Figure 7.10b,e), indicating that both

(16)

macroscopic strings of MAC10 and MAP1 show no cytotoxicity. Furthermore,

well-defined actin stress fibers and nuclei of hBM-MSCs are observed by focusing on the surface of MA strings (Figure 7.10d,g), indicating that the hBM-MSCs also adhere and grow onto the surface of the macroscopic strings of MAC10 and MAP1.

The initial results demonstrated that both MAC10 and MAP1 show no significant

cytotoxicity to cell cultures, allowing for cell growth on the MA strings surface. This might be a first step towards the development of biocompatible soft robotic systems operating in living organisms.

7.3 Conclusions

Motor amphiphiles with different charged end groups at the lower half of the motor moiety were synthesized and probed for their self-assembling properties. Nanofibers of MAP1 and vesicles of MAS1 were observed by cryo-TEM. As shown by NRFA, a

lower CAC was observed in MAP1 solution than that of observed in MAS1. By

applying a shear-flow method, macroscopic strings of MAP1 could be prepared from

a calcium chloride solution, while direct dissolution of MAS1 was observed. A high

degree of alignment and fast photoactuations were found for the MAP1 string. In

preliminary experiments, both MAP1 and MAC10 showed no significant cytotoxicity

in the XTT assays, allowing for cell growth on the surface of MAP1 and MAC10

strings. The current approach demonstrates the potential of developing multiple responsive macroscopic actuators with potential for future biocompatible soft robotic systems.

7.4 Acknowledgements

This work was supported financially by the China Scholarship Council (no. 201706790063 to S.Y.C.), the Croucher Foundation (Croucher Startup Allowance to F.K.C.L), the Netherlands Organization for Scientific Research (NWO-CW), the European Research Council (ERC; advanced grant no. 694345 to B.L.F.), the Ministry of Education, Culture and Science (Gravitation program no. 024.001.035), and a Grant-in-Aid for Scientific Research on Innovative Areas “π-Figuration” (no. 26102008 and no. 15K21721) of The Ministry of Education, Culture, Sports, Science and Technology (MEXT), Japan. The synchrotron XRD experiments were performed at the BL45XU in the 8 with the approval of the RIKEN SPring-8 Center (proposal no. 20160027).

(17)

7.5 Experimental Data

7.5.1 Materials and Methods

For the general materials and methods, please refer to section A, Materials and Methods, in the Appendix.

For the Cryo-TEM measurement, Nile Red Fluorescence Assay, please see chapter 6.

Synchrotron Radiation X-ray Diffraction Experiments.39,41,42 Through view

X-ray diffraction (XRD) image of the strings of molecular amphiphiles (MA) were obtained using the BL45XU beamline at SPring-8 (Hyogo, Japan) equipped with an R-AXIS IV++ (Rigaku) imaging plate area detector or with a Pilatus3X 2M (Dectris) detector. The scattering vector, q = 4πsinθ/λ, and the position of incident X-ray beam on the detector were calibrated using several orders of layer reflections from silver behenate (d = 58.380 Å), where 2θ and λ refer to the scattering angle and wavelength of the X-ray beam (1.0 Å), respectively. The sample-to-detector distances for through-view XRD measurements were 2.02 m, respectively. The obtained diffraction patterns and images were integrated along the Debye-Scherrer ring to afford one-dimensional (1D) intensity data using the FIT2D software. The lattice parameters were refined using the CellCalc ver. 2.10 software.

Scanning Electron Microscopy and Polarized Optical Microscopy Analysis.39,41,42 Polarized optical microscopy (POM) was performed on a Nikon

model Eclipse LV100POL optical polarizing microscope. Scanning electron microscopy (SEM) was performed on a Hitachi S-5500 Field Emission SEM (FE-SEM). Preparation of a MA string on a sapphire substrate: When an aqueous solution of MA (50 mM) was manually drawn into an aqueous solution of calcium chloride (CaCl2) (150 mM) from a pipette, a noodle-like string with an arbitrary length was

formed. After removal of the solution of CaCl2, the string was washed with deionized

water (three times), and the resulting string was used directly for POM and XRD experiments. A motor string for SEM was dried in air for 48 h before measurement.

Standardized Preparation and Actuation of Motor Amphiphile Strings in Water.39,41,42 Actuation experiments have been performed in water with standardized

parameters as follows: diameter of the strings (300 ± 30 μm) and length of the strings in water (8.0 ± 2.0 mm) (Figure 7.8). The photoactuation speed was determined from the bending process of a string with a saturated flexion angle of 90° within a particular time. All photoactuation speeds were averaged from three set of actuation experiments and recorded with a Nikon Coolpix A900 Digital Camera. On the basis of the UV-Vis absorption and 1H NMR studies in solution shown in Figures 7.2 and

(18)

with a ratio of 85:15 (unstable-MA:stable-MA). Therefore, we expect a similar extent of switching in the MA string when the saturated flexion angle of the MA string is reached.

Cell Culture.53-54 Human bone marrow-derived mesenchymal stem cells (hBM-MSCs) obtained from Lonza were used for the cell experiments. The growth medium

consisted of Alpha modifed Eagle medium (Gibco), 10% (v/v) fetal bovine serum (Gibco), 0.1% (v/v) ascorbic acid 2-phosphate (Sigma) and 2% penicillin/streptomycin (Gibco). Cells were incubated at 37 oC, 5% CO

2. The cells

were harvested at ~ 80 – 90% confluence from T75 culture flasks by trypsin for 3–5 min at 37 °C for further subcultures.

XTT assay.52 The cells were cultivated in a flat 96-well plate. To each well was

added 100 µL of growth medium (containing MA at a range of concentration: 0.5 µM, 5 µM and 2200 µM) and 50 µL of freshly prepared reaction solution. The reaction solutions consisted of XTT reagent solutions and activation solution (v/v = 1/50), which were defrosted in a 37 oC bath immediately prior to use. All the samples

were stored in an incubator at 37 oC and 5% CO

2 for 24 h. The well plates, gently

shaken to evenly distribute the dye, were subjected to absorbance measurement against a background control as a blank with a spectrophotometer (BMG LABTECH, Offenburg, Germany) at the wavelength of 450 – 500 nm.

Cell growth on the surface of MA string.53,56 Freshly prepared macroscopic strings

of MAC10 and MAP1 were placed in 6-well plates and washed with growth media

(0.5 mL) for 3 time. Subsequently, the hBM-MSCs with a density of 2500 cells per well were seeded onto the macroscopic strings for cell growth. All samples were stored in an incubator at 37 oC and 5% CO

2 for 72 h. The hBM-MSCs were fixed

with 3.7% paraformaldehyde (Sigma-Aldrich) in phosphate-buffered saline (PBS) for 20 min at room temperature, and subsequently washed three times with PBS. Next, the cell membrane was permeabilized with 0.5% TritonX-100 (Sigma-Aldrich) solution for 3 min. In addition, 4′,6-diamidino-2-phenylindole (DAPI) and tetramethylrhodamine isothiocyanate(TRITC)–phalloidin were used to stain the cell nuclei and F-actin, respectively. Cells were observed using a LEICA TCS SP2 confocal laser scanning microscopy (CLSM) equipped with a 40 × NA 0.80 water immersion objective.

7.5.2 Synthesis and Characterization Compound 3:

(19)

A DMF suspension (20 mL) of compound 1 (180 mg, 0.74 mmol) was added K2CO3 (407 mg, 2.95 mmol) and alkyl

bromide 2 (516 mg, 1.85 mmol) at 25 °C, and the mixture was stirred at 85 °C for 18 h. The reaction mixture was allowed to cool to 25 °C and then poured into water (20 mL). The residue

was washed successively with water (50 mL) and ethyl acetate (20 mL), and dried under reduced pressure. The residue was subjected to column chromatography on SiO2 (ethyl acetate/ pentane; v/v = 1/5, Rf = 0.5) to allow isolation of compound 3

(268 mg, 0.42 mmol, 57% yield) as a pale yellow powder.

1H NMR (400 MHz, CDCl 3) δ (ppm) 8.18 (d, J = 8.0 Hz, 2H), 7.35 (t, J = 8.0 Hz, 2H), 7.06 (d, J = 8.0 Hz, 2H), 4.12 (t, J = 6.4 Hz, 4H), 4.00 (t, J = 6.8 Hz, 4H), 2.00 (s, 6H), 1.95 – 1.81 (m, 4H), 1.64 – 1.46 (m, 8H), 1.44 – 1.21 (m, 20H). 13C NMR (100 MHz, CDCl 3) δ (ppm) 181.4, 172.3, 155.5, 131.1, 129.5, 127.0, 122.4, 114.2, 70.6, 65.7, 30.7, 30.6, 30.4, 30.2, 29.8, 27.2, 27.1, 22.1.

HRMS (ESI): calcd. for C37H52O7S [M+Na] 663.3326, found 663.3302. Compound 6:

A toluene solution (10 mL) of compound 3 (267 mg, 0.42 mmol) was added Lawesson’s reagent (680 mg, 1.7 mmol) and the mixture was stirred at 110 °C for 1 h. After solvent removal in vacuum, the residue was subjected to column chromatography on SiO2 (ethyl acetate/ pentane; v/v = 1/5, Rf = 0.5) to allow isolation

of the thioketone compound 4 (252 mg, 0.38 mmol, 91% yield) as a green solid. Meanwhile, a DMF solution (3 mL) of

compound 5 (164 mg, 0.384 mmol) at –50 oC, a DMF solution (2 mL) of

(diacetoxyiodo)benzene (123.7 mg, 0.384 mmol) was added dropwise. After complete addition, the obtained pale pink solution was stirred at –50 °C for 60 s followed by addition of the thioketone compound 4 (240.2 mg, 0.366 mmol) in THF (2 mL). The reaction mixture was stirred at 25 oC for 2 h. Subsequently, the solvent

was removed using rotary evaporation and the residue was subjected to column chromatography on SiO2 (ethyl acetate/pentane; v/v = 1/9, Rf = 0.5) to allow isolation

of episulfide compound 6 (191 mg, 0.181 mmol, 50% yield) as a colorless oil.

1H NMR (400 MHz, CDCl 3) δ (ppm) 8.61 (d, J = 8.0 Hz, 1H), 7.67 (d, J = 8.0 Hz, 1H), 7.50 (d, J = 8.0 Hz, 1H), 7.37 – 7.27 (m, 2H), 7.20 (t, J = 8.0 Hz, 1H), 6.86 (d, J = 8.0 Hz, 1H), 6.70 (s, 1H), 6.31 (dd, J = 20.0, 8.0 Hz, 2H), 6.12 (t, J = 8.0 Hz, 1H), 4.23 – 4.12 (m, 1H), 4.04 (m, 5H), 3.95 – 3.77 (m, 4H), 2.65 – 2.52 (m, 2H), 2.11 (m, 1H), 2.04 (s, 6H), 1.99 – 1.69 (m, 6H), 1.68 – 1.23 (m, 46H), 1.18 (d, J = 4.0 Hz, 3H), 0.89 (t, J = 8.0 Hz, 3H).

(20)

13C NMR (100 MHz, CDCl 3) δ (ppm) 172.2, 168.3, 155.3, 154.5, 153.6, 135.8, 133.5, 132.4, 132.1, 131.5, 130.6, 127.8, 127.0, 125.8, 125.4, 125.2, 125.1, 124.9, 124.3, 123.5, 122.7, 110.9, 110.7, 107.0, 70.6, 70.4, 70.3, 66.3, 65.6, 62.7, 41.6, 36.5, 32.9, 30.7, 30.7, 30.7, 30.6, 30.6, 30.5, 30.5, 30.4, 30.4, 30.4, 30.3, 30.3, 30.2, 30.0, 29.6, 29.6, 27.0, 27.0, 27.0, 26.9, 26.9, 23.7, 22.0, 15.1.

HRMS (ESI): calcd. for C63H88O7S3 [M+H] 1053.5765, found 1053.5748. Compound 7:

To a toluene solution (10 mL) of episulfide compound 6 (191 mg, 0.181 mmol), triphenyl phosphine (95 mg, 0.36 mmol) was added and the mixture stirred at 135 °C for 16 h. The solvent was removed in vacuum and the residue was subjected to column chromatography on SiO2 (ethyl acetate/pentane; v/v =

1/9, Rf = 0.4) to allow isolation of compound 7 (158 mg, 0.155 mmol, 85% yield) as a colorless oil.

1H NMR (400 MHz, CDCl 3) δ (ppm) 7.42 (dd, J = 16.0, 8.0 Hz, 2H), 7.23 (d, J = 8.0 Hz, 1H), 7.14 (d, J = 8.0 Hz, 1H), 7.01 (t, J = 8.0 Hz, 1H), 6.91 (s, 1H), 6.81 (t, J = 8.0 Hz, 2H), 6.32 (t, J = 8.0 Hz, 1H), 6.24 (d, J = 8.0 Hz, 1H), 6.04 (d, J = 4.0 Hz, 1H), 4.19 – 4.07 (m, 3H), 4.04 (m, 6H), 3.98 – 3.81 (m, 2H), 3.72 (dd, J = 12.0, 8.0 Hz, 1H), 3.06 (dd, J = 12.0, 4.0 Hz, 1H), 2.01 (s, 6H), 1.90 (m, 4H), 1.78 (m, 2H), 1.67 – 1.17 (m, 46H), 0.86 (t, J = 4.0 Hz, 3H), 0.74 (d, J = 4.0 Hz, 3H). 13C NMR (100 MHz, CDCl 3) δ (ppm) 172.2, 156.5, 155.6, 153.8, 139.7, 137.3, 136.7, 133.1, 132.8, 132.6, 128.7, 127.4, 127.2, 127.0, 126.4, 126.0, 125.4, 124.8, 123.8, 122.6, 120.5, 110.3, 110.2, 106.0, 70.0, 69.9, 69.7, 65.6, 37.6, 32.9, 31.5, 30.7, 30.6, 30.6, 30.6, 30.5, 30.5, 30.5, 30.5, 30.4, 30.4, 30.3, 30.3, 30.3, 30.2, 30.2, 30.1, 29.6, 27.1, 27.1, 27.0, 26.9, 23.7, 22.0, 19.3, 15.1.

HRMS (ESI): calcd. for C63H88O7S2 [M+H] 1021.6044, found 1021.6017. Compound 8:

To a THF/methanol solution (v/v = 1/1, 6 mL) of compound 7 (158 mg, 0.155 mmol), an aqueous NaOH solution (4 M, 0.34 mL, 1.4 mmol) was added and the mixture stirred at 90 oC for 1

h. The reaction mixture was allowed to cool to 25 °C. The solvent was removed in vacuum and then the residue was washed successively with water (50 mL) and ethyl acetate (20 mL). The organic layer was separated and the water layer was extracted

with ethyl acetate (20 mL). The combine organic layers were washed with brine and dried over Na2SO4. The solvent was removed in vacuum and the residue was

(21)

1/19, Rf = 0.5) to allow isolation of compound 8 (142 mg, 0.151 mmol, 98% yield) as a colorless oil. 1H NMR (400 MHz, CDCl 3) δ (ppm) 7.42 (dd, J = 12.0, 8.0 Hz, 2H), 7.23 (d, J = 8.0 Hz, 1H), 7.14 (d, J = 8.0 Hz, 1H), 7.02 (t, J = 8.0 Hz, 1H), 6.91 (s, 1H), 6.81 (t, J = 8.0 Hz, 2H), 6.32 (t, J = 8.0 Hz, 1H), 6.24 (d, J = 8.0 Hz, 1H), 6.04 (d, J = 8.0 Hz, 1H), 4.12 (m, 3H), 4.03 (m, 2H), 3.98 – 3.89 (m, 1H), 3.89 – 3.80 (m, 1H), 3.72 (dd, J = 12.0, 8.0 Hz, 1H), 3.60 (t, J = 8.0 Hz, 4H), 3.06 (dd, J = 12.0, 4.0 Hz, 1H), 1.91 (m, 4H), 1.78 (m, 2H), 1.68 – 1.17 (m, 48H), 0.86 (t, J = 8.0 Hz, 3H), 0.74 (d, J = 8.0 Hz, 3H). 13C NMR (100 MHz, CDCl 3) δ (ppm) 156.5, 155.6, 153.8, 139.7, 137.3, 136.7, 133.1, 132.8, 132.6, 128.7, 127.4, 127.4, 127.1, 127.0, 126.4, 126.1, 125.4, 124.8, 123.8, 122.6, 120.5, 110.3, 110.3, 106.0, 70.0,69.9, 69.7, 64.0, 37.6, 33.8, 32.9, 31.5, 30.7, 30.6, 30.6, 30.6, 30.6, 30.6, 30.6, 30.5, 30.5, 30.4, 30.4, 30.3, 30.3, 30.1, 30.1, 27.1, 27.1, 27.0, 26.8, 23.7, 19.3, 15.1.

HRMS (ESI): calcd. for C59H84O5S2 [M+H] 937.5833, found 937.5812. Motor amphiphile (MAP1):

To a pyridine solution (3 mL) of compound 8 (110 mg, 0.117 mmol), diphenyl phosphite (192 mg, 0.819 mmol) was added and the mixture stirred at 25 oC for 16 h. Subsequently, a solution

of triethylamine/water (v/v = 1/1, 2 mL) was added to the reaction mixture and stirring continued at 25 °C for 1 h. The solvent was removed in vacuum and then the residue was washed successively with water (50 mL) and dichloromethane (20 mL). The organic layer was separated and the water layer was

extracted with dichloromethane (20 mL). The combine organic layers were washed with brine and dried over Na2SO4. The solvent was removed in vacuum and the

residue was subjected to column chromatography on SiO2 (methanol (with 10%

triethylamine)/dichloromethane; v/v = 1/9, Rf = 0.5) to allow isolation of MAP1 (48

mg, 0.0379 mmol, 32% yield) as a pale yellow oil.

1H NMR (400 MHz, CDCl 3) δ (ppm) 7.59 (s, 1H), 7.42 (t, J = 12.0 Hz, 2H), 7.25 (t, J = 8.0 Hz, 1H), 7.15 (d, J = 8.0 Hz, 1H), 7.03 (t, J = 8.0 Hz, 1H), 6.92 (s, 1H), 6.82 (t, J = 8.0 Hz, 2H), 6.33 (t, J = 8.0 Hz, 1H), 6.25 (d, J = 8.0 Hz, 1H), 6.05 (d, J = 8.0 Hz, 2H), 4.20 – 4.09 (m, 3H), 4.08 – 3.98 (m, 2H), 3.98 – 3.90 (m, 1H), 3.86 (m, 5H), 3.73 (dd, J = 12.0, 8.0 Hz, 1H), 3.14 – 2.92 (m, 13H), 1.91 (m, 4H), 1.79 (m, 2H), 1.68 – 1.19 (m, 66H), 0.87 (t, J = 6.7 Hz, 3H), 0.74 (d, J = 6.7 Hz, 3H). 13C NMR (150 MHz, CDCl 3) δ (ppm) 156.5, 155.5, 153.7, 139.6, 137.2, 136.6, 133.0, 132.8, 132.5, 128.6, 127.3, 127.1, 126.9, 126.3, 126.0, 125.4, 125.4, 124.7, 123.8,

(22)

122.5, 120.4, 110.3, 110.2, 105.9, 70.0, 69.9, 69.6, 65.0 65.0, 46.3, 37.6, 32.8, 31.8, 31.7, 31.4, 30.6, 30.6, 30.6, 30.6, 30.5, 30.4, 30.4, 30.4, 30.3, 30.3, 30.3, 30.3, 30.1, 30.1, 27.1, 27.1, 27.0, 26.8, 23.6, 19.3, 15.1, 9.4.

HRMS (ESI): calcd. for C59H86O9P2S2 [M-H] 1063.5105, found 1063.5129. Motor amphiphile (MAS1):

A mixture of compound 8 (36 mg, 0.038 mmol), sulfur trioxide pyridine complex (37 mg, 0.23 mmol) and triethylamine solution (0.1 mL) in dichloromethane (6 mL) was stirred at 25

oC for 18 h. The solvent was removed in vacuum and the

residue was subjected to column chromatography on SiO2

(methanol (with 10% triethylamine)/dichloromethane; v/v = 1/9, Rf = 0.5) to allow isolation of MAS1 (23 mg, 0.018 mmol,

47% yield) as a pale yellow oil.

1H NMR (400 MHz, CDCl 3) δ (ppm) 9.58 (s, 2H), 7.42 (t, J = 8.0 Hz, 2H), 7.26 (d, J = 8.0 Hz, 1H), 7.15 (d, J = 8.0 Hz, 1H), 7.03 (t, J = 8.0 Hz, 1H), 6.92 (s, 1H), 6.82 (d, J = 8.0 Hz, 2H), 6.33 (t, J = 8.0 Hz, 1H), 6.26 (d, J = 8.0 Hz, 1H), 6.04 (d, J = 8.0 Hz, 1H), 4.25 – 4.08 (m, 3H), 4.02 (m, 6H), 3.98 – 3.80 (m, 2H), 3.73 (m, 1H), 3.11 (m, 13H), 1.92 (m, 4H), 1.78 (m, 2H), 1.66 (m, 4H), 1.61 – 1.19 (m, 60H), 0.87 (t, J = 8.0 Hz, 3H), 0.74 (d, J = 8.0 Hz, 3H). 13C NMR (150 MHz, CDCl 3) δ (ppm) 156.5, 155.5, 153.7, 139.7, 137.3, 136.7, 133.0, 132.8, 132.5, 128.7, 127.3, 127.2, 127.0, 126.3, 126.0, 125.4, 125.3, 124.7, 123.8, 122.5, 120.4, 110.3, 110.3, 106.0, 70.0, 69.9, 69.7, 69.0 55.7, 47.5, 37.6, 32.8, 31.5, 30.6, 30.6, 30.6, 30.5, 30.5, 30.5, 30.4, 30.4, 30.4, 30.3, 30.3, 30.3, 30.2, 30.1, 30.1, 27.1, 27.0, 26.9, 26.8, 23.6, 19.3, 15.0, 9.6.

HRMS (ESI): calcd. for C59H84O11S4 [M-H] 1095.4812, found 1095.4813.

7.6 References

(1) Ebasm, S.; Endo, M. Calcium Ion and Muscle Contraction. Prog. Biophys. Mol. Biol.

1968, 18, 123–183.

(2) Rayment, I.; Holden, H. M.; Whittaker, M.; Yohn, C. B.; Lorenz, M.; Holmes, K. C.; Milligan, R. A. Structure of the Actin-Myosin Complex and Its Implications for Muscle Contraction. Science 1993, 261, 58–65.

(3) Alberts, B. The Cell as a Collection of Protein Machines : Preparing the Next Generation of Molecular Biologists. Cell 1998, 92, 291–294.

(4) Vale, R. D.; Milligan, R. A. The Way Things Move : Looking Under the Hood of Molecular Motor Proteins. Science 2000, 288, 88–95.

(5) Zehr, E. P.; Duysens, J. Regulation of Arm and Leg Movement during Human Locomotion. Neurosci. 2004, 10, 347–361.

(23)

Angew. Chem. Int. Ed. 2007, 46, 2366–2393.

(7) de Greef, T. F. A.; Meijer, E. W. Supramolecular Polymers. Nature 2008, 453, 171– 173.

(8) Palmer, L. C.; Stupp, S. I. Molecular Self-Assembly into One-Dimensional Nanostructures. Acc. Chem. Res. 2008, 41, 1674–1684.

(9) Wojtecki, R. J.; Meador, M. A.; Rowan, S. J. Using the Dynamic Bond to Access Macroscopically Responsive Structurally Dynamic Polymers. Nat. Mater. 2011, 10, 14–27.

(10) Krieg, E.; Rybtchinski, B. Noncovalent Water-Based Materials: Robust yet Adaptive.

Chem. Eur. J. 2011, 17, 9016–9026.

(11) Aida, T.; Meijer, E. W.; Stupp, S. I. Functional Supramolecular Polymers. Science

2012, 335, 813–817.

(12) Yan, X.; Wang, F.; Zheng, B.; Huang, F. Stimuli-Responsive Supramolecular Polymeric Materials. Chem. Soc. Rev. 2012, 41, 6042–6065.

(13) Li, S. L.; Xiao, T.; Lin, C.; Wang, L. Advanced Supramolecular Polymers Constructed by Orthogonal Self-Assembly. Chem. Soc. Rev. 2012, 41, 5950–5968. (14) Dong, R.; Zhou, Y.; Huang, X.; Zhu, X.; Lu, Y.; Shen, J. Functional Supramolecular

Polymers for Biomedical Applications. Adv. Mater. 2015, 27, 498–526.

(15) Mulder, A.; Huskens, J.; Reinhoudt, D. N. Multivalency in Supramolecular Chemistry and Nanofabrication. Org. Biomol. Chem. 2004, 2, 3409–3424.

(16) Frisch, H.; Nie, Y.; Raunser, S.; Besenius, P. pH-Regulated Selectivity in Supramolecular Polymerizations : Switching between Co- and Homopolymers.

Chem. Eur. J. 2015, 21, 3304–3309.

(17) Ardona, H. A. M.; Tovar, J. D. Energy Transfer within Responsive Pi-Conjugated Coassembled Peptide-Based Nanostructures in Aqueous Environments. Chem. Sci.

2015, 6, 1474–1484.

(18) Wang, Y.; Huang, Z.; Kim, Y.; He, Y.; Lee, M. Guest-Driven Inflation of Self-Assembled Nano Fi Bers through Hollow Channel Formation. J. Am. Chem. Soc.

2014, 136, 16152–16155.

(19) Dehsorkhi, A.; Castelletto, V.; Hamley, I. W.; Adamcik, J.; Mezzenga, R. The Effect of pH on the Self-Assembly of a Collagen Derived Peptide Amphiphile. Soft Matter

2013, 9, 6033–6036.

(20) Shin, S.; Lim, S.; Kim, Y.; Kim, T.; Choi, T.; Lee, M. Supramolecular Switching between Flat Sheets and Helical Tubules Triggered by Coordination Interaction. J.

Am. Chem. Soc. 2013, 135, 2156–2159.

(21) Zhang, J.; Hao, R.; Huang, L.; Yao, J.; Chen, X.; Shao, Z. Self-Assembly of a Peptide Amphiphile Based on Hydrolysed Bombyx Mori Silk Fibroin. Chem. Commun. 2011,

47, 10296–10298.

(22) Huang, Z.; Kang, S. K.; Lee, M. Induction of Supramolecular Chirality in Self-Assembled Nanofibers Triggered by Environmental Change. J. Mater. Chem. 2011,

21, 15327–15331.

(23) Huang, Z.; Lee, H.; Lee, E.; Kang, S.; Nam, J.; Lee, M. Responsive Nematic Gels from the Self-Assembly of Aqueous Nanofibres. Nat. Commun. 2011, 2, 455–459. (24) Kuad, P.; Miyawaki, A.; Takashima, Y.; Yamaguchi, H.; Harada, A. External

Stimulus-Responsive Supramolecular Structures Formed by a Stilbene Cyclodextrin Dimer. J. Am. Chem. Soc. 2007, 129, 12630–12631.

(25) Lin, B. F.; Megley, K. A.; Viswanathan, N.; Krogstad, D. V; Drews, L. B.; Kade, M. J.; Tirrell, M. V. pH-Responsive Branched Peptide Amphiphile Hydrogel Designed for Applications in Regenerative Medicine with Potential as Injectable Tissue Scaffolds. J. Mater. Chem. 2012, 22, 19447–19454.

(24)

Amphiphilic Structure Direct Supramolecular Behavior in Biofunctional Assemblies.

J. Am. Chem. Soc. 2014, 136, 14746–14752.

(27) Wu, A.; Sun, P.; Sun, N.; Zheng, L. Responsive Self-Assembly of Supramolecular Hydrogel Based on Zwitterionic Liquid Asymmetric Gemini Guest. Chem. Eur. J.

2018, 24, 10452–10459.

(28) Choi, Y. J.; Kim, J. T.; Yoon, W. J.; Kang, D. G.; Park, M.; Kim, D. Y.; Lee, M.; Ahn, S.; Jeong, K. Azobenzene Molecular Machine: Light-Induced Wringing Gel Fabricated from Asymmetric Macrogelator. ACS Macro Lett. 2018, 7, 576–581. (29) Goujon, A.; Mariani, G.; Lang, T.; Moulin, E.; Rawiso, M.; Buhler, E.; Giuseppone,

N. Controlled Sol-Gel Transitions by Actuating Molecular Machine Based Supramolecular Polymers. J. Am. Chem. Soc. 2017, 139, 4923–4928.

(30) van Herpt, J. T.; Areephong, J.; Stuart, M. C. A.; Browne, W. R.; Feringa, B. L. Light-Controlled Formation of Vesicles and Supramolecular Organogels by a Cholesterol-Bearing Amphiphilic Molecular Switch. Chem. Eur. J. 2014, 20, 1737–1742. (31) Wang, C.; Chen, Q.; Sun, F.; Zhang, D.; Zhang, G.; Huang, Y.; Zhao, R.; Zhu, D.

Multistimuli Responsive Organogels Based on a New Gelator Featuring Tetrathiafulvalene and Azobenzene Groups : Reversible Tuning of the Gel-Sol Transition by Redox Reactions and Light Irradiation. J. Am. Chem. Soc. 2010, 132, 3092–3096.

(32) Kim, J. H.; Seo, M.; Kim, Y. J.; Kim, S. Y. Rapid and Reversible Gel−Sol Transition of Self-Assembled Gels Induced by Photoisomerization of Dendritic Azobenzenes.

Langmuir 2009, 25, 1761–1766.

(33) Zhou, Y.; Xu, M.; Yi, T.; Xiao, S.; Zhou, Z.; Li, F.; Huang, C. Morphology-Tunable and Photoresponsive Properties in a Self-Assembled Two-Component Gel System.

Langmuir 2007, 23, 202–208.

(34) Koumura, N.; Kudo, M.; Tamaoki, N. Photocontrolled Gel-to-Sol-to-Gel Phase Transitioning of Meta-Substituted Azobenzene Bisurethanes through the Breaking and Reforming of Hydrogen Bonds. Langmuir 2004, 20, 9897–9900.

(35) Borre, E.; Stumbe, J. F.; Bellemin-Laponnaz, S.; Mauro, M. Light-Powered Self-Healable Metallosupramolecular Soft Actuators. Angew. Chem. Int. Ed. 2016, 55, 1313–1317.

(36) Borre, E.; Stumbe, J. F.; Bellemin-Laponnaz, S.; Mauro, M. Control of the Light-Response in Supramolecular Metallopolymeric Gels by Tuning the Coordination.

Chem. Commun. 2017, 53, 8344–8347.

(37) Li, Q. Liquid Crystals Beyond Displays: Chemistry, Physics, and Applications; John Wiley & Sons, 2012.

(38) Broer, D. J.; Crawford, G. P.; Zumer, S. Cross-Linked Liquid Crystalline Systems:

From Rigid Polymer Networks to Elastomers; CRC Press, 2011.

(39) Chen, J.; Leung, F. K. C.; Stuart, M. C. A.; Kajitani, T.; Fukushima, T.; van der Giessen, E.; Feringa, B. L. Artificial Muscle-like Function from Hierarchical Supramolecular Assembly of Photoresponsive Molecular Motors. Nat. Chem. 2018,

10, 132–138.

(40) Chin, S. M.; Synatschke, C. V.; Liu, S.; Nap, R. J.; Sather, N. A.; Wang, Q.; Álvarez, Z.; Edelbrock, A. N.; Fyrner, T.; Palmer, L. C.; Szleifer, I.; Olvera de la Cruz, M.; Stupp, S. I. Covalent-Supramolecular Hybrid Polymers as Muscle-Inspired Anisotropic Actuators. Nat. Commun. 2018, 9, 2395.

(41) Leung, F. K.; van den Enk, T.; Kajitani, T.; Chen, J.; Stuart, M. C. A.; Kuipers, J.; Fukushima, T.; Feringa, B. L. Supramolecular Packing and Macroscopic Alignment Controls Actuation Speed in Macroscopic Strings of Molecular Motor Amphiphiles.

J. Am. Chem. Soc. 2018, 140, 17724–17733.

(25)

Dual-Controlled Macroscopic Motions in a Supramolecular Hierarchical Assembly of Motor Amphiphiles. Angew. Chem. Int. Ed. 2019, 58 (32),10985-10989.

(43) Wall, B. D.; Diegelmann, S. R.; Zhang, S.; Dawidczyk, T. J.; Wilson, W. L.; Katz, H. E.; Mao, H. Q.; Tovar, J. D. Aligned Macroscopic Domains of Optoelectronic Nanostructures Prepared via Shear-Flow Assembly of Peptide Hydrogels. Adv. Mater.

2011, 23, 5009–5014.

(44) Besar, K.; Ardoña, H. A. M.; Tovar, J. D.; Katz, H. E. Demonstration of Hole Transport and Voltage Equilibration in Self-Assembled π-Conjugated Peptide Nanostructures Using Field-Effect Transistor Architectures. ACS Nano 2015, 9, 12401–12409.

(45) Sanders, A. M.; Magnanelli, T. J.; Bragg, A. E.; Tovar, J. D. Photoinduced Electron Transfer within Supramolecular Donor-Acceptor Peptide Nanostructures under Aqueous Conditions. J. Am. Chem. Soc. 2016, 138, 3362–3370.

(46) Matson, J. B.; Stupp, S. I. Self-Assembling Peptide Scaffolds for Regenerative Medicine. Chem. Commun. 2012, 48, 26–33.

(47) Stupp, S. I.; Zha, R. H.; Palmer, L. C.; Cui, H.; Bitton, R. Self-Assembly of Biomolecular Soft Matter. Faraday Discuss. 2013, 166, 9–30.

(48) Webber, M. J.; Berns, E. J.; Stupp, S. I. Supramolecular Nanofibers of Peptide Amphiphiles for Medicine. Isr. J. Chem. 2013, 53, 530–554.

(49) Mcclendon, M. T.; Stupp, S. I. Biomaterials Tubular Hydrogels of Circumferentially Aligned Nano Fi Bers to Encapsulate and Orient Vascular Cells. Biomaterials 2012,

33, 5713–5722.

(50) Zhang, S.; Greenfield, M. A.; Mata, A.; Palmer, L. C.; Bitton, R.; Mantei, J. R.; Aparicio, C.; Olvera de la Cruz, M.; Stupp, S. I. A Self-Assembly Pathway to Aligned Monodomain Gels. Nat. Mater. 2010, 9, 594–601

(51) Tantakitti, F.; Boekhoven, J.; Wang, X.; Kazantsev, R. V.; Yu, T.; Li, J.; Zhuang, E.; Zandi, R.; Ortony, J. H.; Newcomb, C. J.; Palmer, L. C.; Shekhawat, G. S.; Olvera de la Cruz, M.; Schatz, G. C.; Stupp, S. I. Energy Landscapes and Functions of Supramolecular Systems. Nat. Mater. 2016, 15, 469−476.

(52) Zhou, Q.; Kühn, P. T.; Huisman, T.; Nieboer, E.; van Zwol, C.; van Kooten, T. G.; van Rijn, P. Directional Nanotopographic Gradients: A High-Throughput Screening Platform for Cell Contact Guidance. Sci. Rep. 2015, 5, 1–12.

(53) Zhou, Q.; Castañeda Ocampo, O.; Guimarães, C. F.; Kühn, P. T.; van Kooten, T. G.; van Rijn, P. Screening Platform for Cell Contact Guidance Based on Inorganic Biomaterial Micro/Nanotopographical Gradients. ACS Appl. Mater. Interfaces 2017,

9 (37), 31433–31445.

(54) Chen C. S.; Mrksich M.; Huang S.; Whitesides G. M.; Ingber D. E. Geometric Control of Cell Life and Death. Science 1997, 276 (5317), 1425–1428.

(55) Zheng, Y.; Chen, J.; Craven, M.; Choi, N. W.; Totorica, S.; Diaz-Santana, A.; Kermani, P.; Hempstead, B.; Fischbach-Teschl, C.; Lopez, J. A.; Stroock, A. D. In Vitro Microvessels for the Study of Angiogenesis and Thrombosis. Proc. Natl. Acad.

Sci. 2012, 109 (24), 9342–9347.

(56) Zhou, Q.; Zhao, Z.; Zhou, Z.; Zhang, G.; Chiechi, R. C.; van Rijn, P. Directing Mesenchymal Stem Cells with Gold Nanowire Arrays. Adv. Mater. Interfaces 2018,

5 (14), 1–8..

(57) Zhou, Q.; Ge, L.; Guimarães, C. F.; Kühn, P. T.; Yang, L.; van Rijn, P. Development of a Novel Orthogonal Double Gradient for High-Throughput Screening of Mesenchymal Stem Cells–Materials Interaction. Adv. Mater. Interfaces 2018, 5 (18), 4–11.

(26)
(27)

Referenties

GERELATEERDE DOCUMENTEN

The results showed clearly that by manipulation of molecular geometrical transformations with light and heat stimuli, the macroscopic foaming ratio also can be

Ten slotte is in Hoofdstuk 7 het ontwerp en de synthese van biocompatibele fotoresponsieve amfifielen op basis van een tweede generatie moleculaire motoren met

To Anouk: Dear Anouk, thank you for helping me to correct the summary and the Dutch translations. When I began to write my thesis, I am very grateful that you told me: “If you

Photoresponsive azobenzene amphiphiles with negligible variation of equilibrium surface tension can present significant differences in dynamic surface tension (Langmuir

Besides providing good predictability for clinical trials, these models provide valuable tools to study the mechanisms underlying complex processes like cancer initiation,

Hier komt de tekst voor de rug; hoe dikker de rug, hoe groter de tekst Pulling the strings on anti-cancer immunity K elly K er sten

BSVMM cannot be covered by existing convergence results, and in order to understand its convergence, it is necessary to exploit the special structure of the problem; thirdly, BSVMM

Evaluates the ⟨floating point expression⟩ and converts the result to a string according to the ⟨format specification⟩. The ⟨style⟩ can be. • e for scientific notation, with