Hierarchical Self-Assembly of a Porphyrin into Chiral Macroscopic

Nov 22, 2017 - Molecular structures of amphiphilic histidine derivative (LHC18 or DHC18) and TCPP and illustration on self-assembly of amphiphilic his...
0 downloads 9 Views 2MB Size
Subscriber access provided by READING UNIV

Article

Hierarchical Self-Assembly of A Porphyrin into Chiral Macroscopic Flowers with Superhydrophobic and Enantioselective Property Hejin Jiang, Li Zhang, Jie Chen, and Minghua Liu ACS Nano, Just Accepted Manuscript • DOI: 10.1021/acsnano.7b06484 • Publication Date (Web): 22 Nov 2017 Downloaded from http://pubs.acs.org on November 23, 2017

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.

ACS Nano 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

ACS Nano

Hierarchical Self-Assembly of A Porphyrin into Chiral Macroscopic Flowers with Superhydrophobic and Enantioselective Property Hejin Jiang, †,Ⅱ Li Zhang,*, † Jie Chen † and Minghua Liu* ,†,‡,§,Ⅱ †

Beijing National Laboratory for Molecular Science (BNLMS), CAS Key Laboratory of Colloid,

Interface and Chemical Thermodynamics, Institute of Chemistry, Chinese Academy of Sciences, Beijing, 100190, China. ‡

National Center for Nanoscience and Technology, Beijing, 100190, China.

§

Collaborative Innovation Center of Chemical Science and Engineering, Tianjin, 300072, China. University of Chinese Academy of Sciences, Beijing, 100049, China.

Email address: [email protected]; [email protected]

ABSTRACT: Supramolecular self-assembly provides an efficient way to fabricate simple units into various hierarchical nano/microstructures, which could mimic the bio-self-assembly and develop functional materials.

Since chiral molecules and chiral nanostructures are widely

adopted by biological systems, an introduction of the chiral factor into the self-assembly process will provide better understanding of the biological systems. Here, using a chiral amphiphilic

ACS Paragon Plus Environment

1

ACS Nano

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 2 of 30

histidine to assist the self-assembly of a porphyrin with four carboxylic acids, we obtained hierarchical chiral nano to micro structures. We have found that through the hydrogen bonds/electrostatic interactions between the porphyrin and histidine derivatives, the π−π stacking between the porphyrins, and hydrophobic interactions between the amphiphilic histidine, the two components could self-assemble into chiral nanohelices and microflowers. The supramolecular chirality of these structures was confirmed by SEM images as well as the CD spectra, which was found to follow the molecular chirality of the histidine derivative. More interestingly, the microflower structures formed a superhydrophobic and chiral surface, which exhibited macroscopic enantioselective recognition of some L- and D-amino acids via contact angle measurements.

KEYWORDS: self-assembly, porphyrin, chiral, microflowers, enantioselectivity

Chirality, as a ubiquitous phenomenon in nature, expressed in biological system at various hierarchical levels like molecular amino acid, sugar, supramolecular DNA, protein and micro tubular and cell structures.1, 2 These structures are strongly related to their functions such as the duplicate of the DNA and the enzymatic reaction of the proteins. Self-assembly provides an efficient way to obtain hierarchical nano to micro structures starting from simple molecular units.3, 4 Since most of biological systems exhibit chiral sense, the chiral self-assembly based on or assisted by the chiral components could provide better understanding and mimic the biological self-assembly.5, 6 For instance, some peptides,7-9 amino acid derivatives,10-12 sugar derivatives,13 nanoparticles,14-16 and even some achiral molecules17, 18 have been designed and self-assembled into various chiral nanostructures19 or hybrid materials.20-22 These provide the well understanding

ACS Paragon Plus Environment

2

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

ACS Nano

between molecular and supramolecular as well as nanoscale chirality.23,

24

However, it still

remains an important issue if the controllability of the molecular chirality can be extended to even a larger scale and how chirality is transferred during such process. Some recent works show such possibility.15, 25 For example, Mckee and co-workers obtained the chiral, macroscopic vaterite toroidal suprastructure of calcium carbonate, which could show counterclockwise or clockwise spiraling morphology via the inductions of different enantiomeric amino acids.25 Kotov et al. have assembled mesoscale helices with near-unity enantiomeric excess using chiral semiconductors.15 On the other hand, flower-shaped structures generally possess large surface area, easy accessibility to reaction sites, confinement effect and hierarchical nanostructures extending to a larger scale, which could possibly exhibited enhanced enzyme activity26 and superhydrophobicity.27 Although some flower-shape structures have been obtained in selfassemblies,28 chiral flower-like structures have rarely been reported in pure organic systems. Herein, chiral flowers-shaped structures were fabricated based on the self-assembly of porphyrins and their surface properties were investigated.

Porphyrin is a well-known building block that has heterocyclic macrocycle ring and a large πconjugated system, which has widely used in the catalysis, photovoltaic devices and chemical sensors.29, 30 Their planar aromatic macrocycle is optimal for π-π stacking and it is a versatile platform for peripheral decoration with groups that can offer other interaction motifs, such as metal-ligand coordination, hydrogen bonding and electrostatic interaction. By virtue of these non-covalent interactions, porphyrin is extensively used as a brick to fabricate various structures such as hollow hexagonal nanoprisms, nanodiscs, nanorods, nanowires, nanospheres, vesicles, nanoarrays and showed some special or enhanced functions.31-45 Meanwhile, porphyrin plays an important role in organisms such as hemoglobin, vitamin B12, cytochrome and chlorophylls in

ACS Paragon Plus Environment

3

ACS Nano

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 4 of 30

human, plants and microorganism. Because of their significance in many biological processes, it is expected to assemble the porphyrin in a chiral environment or into chiral nanostructures. So far, only rare examples of chiral porphyrin structures have been reported, in which porphyrin derivatives are directly linked to the chiral units by covalent bonds.46-49 Here, by means of supramolecular interactions, we construct the chiral self-assembly of an achiral porphyrin assisted by an chiral amphiphilic histidine.

The amphiphilic histidine, abbreviated as LHC18 or DHC18, has a chiral center and easily interacts with carboxylic acid groups through the interaction between the imidazole and carboxylic acid groups.50 Herein, tetrakis(4-carboxyphenyl)porphyrin (TCPP) is chosen to coassemble with LHC18 and DHC18. By adjusting the molar ratio of the two components and the mixed solvents, we obtained chiral nano/micro structures, as verified via the SEM and further by CD spectra. Interestingly, the chiral structures can be extended to the microscale level and the handedness of the structures was controlled by the molecular chirality of the amphiphiles, as shown in Figure 1. Moreover, the chiral flower showed superhydrophobic surface due to the hierarchical nano/micro structures. Interestingly, the chiral microflowers further exhibited enantioselectivity to aspartic acids, one of the proteinogenic amino acids. So far, although various porphyrin nano and micro structures have been reported, the chiral architectures and corresponding functions were rarely explored. The present paper provided an example of chiral microflower with superhydrophobic and enantioselective property.

ACS Paragon Plus Environment

4

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

ACS Nano

Figure 1. Molecular structures of amphiphilic histidine derivative (LHC18 or DHC18) and tetrakis(4-carboxyphenyl)porphyrin (TCPP), and illustration on self-assembly of amphiphilic histidine/TCPP into chiral microflowers, in which the chiral sense originates from preference stacking of nanosheets and is determined by molecular chirality of histidine derivatives. RESULTS Hierarchical self-Assembly of TCPP with LHC18 (DHC18) into chiral microflowers. The self-assembly was performed in the mixed solvents of DMF and water. Typically, 4 mM LHC18 (or DHC18) and 1mM TCPP were dissolved in DMF by heating and then water was added dropwisely into the DMF solution. Upon cooling, the transparent solution turned into an opaque dispersion, indicating that TCPP formed aggregates by the aid of histidine derivatives. Figure 2 shows the nanostructures formed by the self-assembly of LHC18/TCPP under various molar ratio of LHC18 to TCPP. LHC18 could self-assemble into entangled nanobelt structures with the width of about 300 nm in DMF/ H2O mixed solvent, as shown in Figure S1a. On the other hand, individual TCPP could form straight nanofibers in DMF/H2O solution (Figure S1b). Upon mixing TCPP with LHC18 or DHC18, significant changes in the morphologies are observed due to the interactions between carboxylic acid group of TCPP and the histidine moiety of LHC18. The formed nanostructures can be regulated by the molar ratio of LHC18 to TCPP. Non-uniform

ACS Paragon Plus Environment

5

ACS Nano

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

Page 6 of 30

amorphous structures can be obtained under the ratio of LHC18 to TCPP at 1:1. With increasing of molar ratio to 2:1, obvious chiral twists with right-handed sense are observed. The pitch of twist is about 470 nm and the width is around 200 nm. When the molar ratio reaches to 3:1, bundles of nanotwists formed by single nanotwist entangled together are observed. It is interesting to note that the flower-like structures appear when the molar ratio increasing to 4:1. These flower-like structures consisted of bent or curved nanolayers. This result demonstrates that molar ratio of the two components has a great effect on the morphology of co-assemblies. It should be noted that the LHC18-TCPP heteroaggregates can be formed when stoichiometry of the LHC18/TCPP is at 2:1, 3:1 and 4:1, according to a Job’s plot analysis (Figure S2). However, the flower-like structures were only observed at ratio of 4:1, then we selected the ratio of 4:1 as an optimal condition to fabricate the flower structures. When DHC18 was used to co-assemble with TCPP, similar nano/micro structures were obtained. However, their handedness of the chiral nanostructures was just in a mirror image with that of LHC18/TCPP assemblies. This suggests that the supramolecular chirality of assemblies follows molecular chirality of amphiphilic histidine (Figure S3).

ACS Paragon Plus Environment

6

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

ACS Nano

Figure 2. SEM images of LHC18/TCPP assemblies at different molar ratio of LHC18/TCPP (a) 1:1; (b) 2:1; (c) 3:1; (d) 4:1 in DMF/H2O (7:3) mixed solvents; and at different volume ratio of DMF/H2O mixed solvents (e) 3:7; (f) 6:4; (g) 7:3; (h) 8:2. The effect of solvent on the co-assemblies are also investigated. Taken LHC18 to TCPP at 4: 1 as an example, the co-assemblies formed three kinds of structures with the variation of volume ratio of the miscible solvents. As shown in Figure 2, when the ratio of the solvents (DMF/H2O) is less than 5:5, some nanobricks present. While the ratio reaches to 5:5 or 6:4, spherical structures with relative smooth surface are obtained. However, the size of microsphere is not uniform, and the diameter of microsphere ranges from 460 nm to 1.5 µm. Specifically, flowerlike structures appear when the solvent ratio increases to 7:3 and 8:2, and then we fixed the solvent at 7:3 for the following experiment. In the case of other mixed solvents, such as THF/H2O, DMSO/H2O and DMF/chloroform, we could not observe the flower like assemblies (Figure S4).

ACS Paragon Plus Environment

7

ACS Nano

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 8 of 30

Further, the concentration of LHC18 and TCPP (the molar ratio is kept at 4:1) is changed to explore its effect on the co-assemblies. The concentration of LHC18 varied from 1 mM to 10 mM, and correspondingly, the concentration of TCPP increased from 0.25 mM to 2.5 mM. Generally, the flower-like microstructures are found under all the concentrations (Figure 3 and Figure S5), indicating the molar ratio of LHC18 to TCPP plays a key role rather than the concentration in the forming flower-like structures. However, when the concentration of the twocomponents increases, nanotwists and flower exist simultaneously. As shown in Figure 3, when the concentration is relatively low, the flower-like structures composed of some thin and bent nanobelts, which are curved in the same direction under a certain chiral molecule and stacked into flower-like structures. With the concentration increasing, although microflowers consisted of nanobelts appear, some twisted nanofibers with obvious right-handed chiral sense are also obtained. These twisted nanofibers entangled together and also turned into flower. Transmission electron microscopy (TEM) confirmed that the flower-like structures were formed (Figure S6). Furthermore, DHC18/TCPP co-assemblies in different concentrations and solvents were obtained, which was similar to the LHC18/TCPP co-assemblies in shape but with opposite chirality (Figure S7).

ACS Paragon Plus Environment

8

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

ACS Nano

Figure 3. SEM images of LHC18/TCPP assemblies at concentration of LHC18 at (a) 1mM; (b) 3mM; (c) 5mM; (d) 7mM. (molar ratio of LHC18 to TCPP is 4:1 and solvent ratio is (DMF/H2O=7:3)). LHC18 is a chiral molecule, which can effectively transfer its chiral information to some achiral components containing carboxylic group to form chiral supramolecular assemblies through hydrogen bonding.51 In this system we also got the chiral structures by achiral TCPP via interacting with the chiral LHC18. Unlike those common chiral nanostructures such as nanotwist, helix, and helical tubes, we fabricated the nanobelt or nanosheet with a preferential curved direction, which stacked further into the micro flowers. As shown in Figure 4a, b, the chiral sense of co-assemblies of LHC18 and TCPP grew along a clockwise direction of the nanobelt from the inside to the outside, which could be defined as right-handed.24 On the other hand, DHC18/TCPP flowers grew along counterclockwise direction (left-handed), as shown in Figure 4d and Figure 4e. These chiral flowers are quite stable and can be preserved for at least one month, as shown in Figure S8, which guarantee the realization of corresponding property and functions. We also investigated the co-assemblies containing LHC18 and some metal porphyrins, such as ZnTCPP and FeTCPP, as shown in Figure S9. It was found that no macroscopic chiral flowers architectures can be obtained. In order to confirm that the chiral flowers were induced by chiral amphiphilic histidine derivatives, the racemic molecules including equal molar of LHC18 and DHC18 were utilized to interact with TCPP. As shown in Figure 4g, h, although both flower-like and nanorod structures are observed, the flower as well as the nanorod do not show any chiral preference. Thus, it can be concluded that the chirality of the flower-like structures was determined by chiral LHC18 and

ACS Paragon Plus Environment

9

ACS Nano

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

DHC18, respectively. This result was also confirmed by CD spectra, which will be shown in following Figure 5b.

Figure 4. SEM images of (a, b) LHC18 with TCPP microflowers with nanosheets arranged in a clockwise manner; (d, e) DHC18 and TCPP flower-like structures composed of nanosheets arranged in a counterclockwise manner; (g, h) racemate of LHC18/ DHC18 and TCPP with no obvious chiral sense; (c, f, i) carton pictures of LHC18/TCPP, DHC18/TCPP and racemate/ TCPP. Spectroscopic characterization of the chiral microflowers. The as-formed nanostructure was monitored by UV-vis spectra, as shown in Figure 5a. The UV-vis spectrum of LHC18 with TCPP co-assemblies in DMF/H2O mixed solution shows a Soret band (B-band) at 418 nm. At the same time, from 500 nm to 700 nm four obvious Q-band could be observed. The UV-vis spectra of cast film of TCPP and LHC18/TCPP were also characterized. The B-band of

ACS Paragon Plus Environment

10

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

ACS Nano

LH18/TCPP cast film shifts from 428 nm to 443 nm and the Q-band also has a red shift relative to TCPP alone, which suggests TCPP formed J-like aggregates by the aid of LHC18. To further confirm the interaction between LHC18 and TCPP, the fluorescence decay of LHC18/TCPP complex together with TCPP itself was measured, as show in Figure S10. Both of the decay curves of the cast films can be fitted using a double exponential function, and lifetimes can be obtained. For the pristine TCPP, the lifetime is 3.20 ns, while the lifetime increases to 6.86 ns at LHC18/TCPP complex, indicating the π−π stacking between TCPP molecules may be slightly inhibited after interacting with LHC18. The Fourier transform infrared spectra (FT-IR) of LHC18/TCPP provide more information for the interaction between LHC18 and TCPP. As shown in Figure 5c, LHC18 exhibits vibration band at 3351, 1760, 1637 and 1576 cm-1, which is ascribed to N-H stretching band, ester carbonyl group, amide I and amide II band, respectively. TCPP shows a typical hydrogenbonded C=O group in 1702 cm-1. After interacting with LHC18, C=O of TCPP shows blue shift from 1702 cm-1 to 1691 cm-1, indicating more stronger hydrogen bonding between histidine with carboxylic acid of LHC18. In addition, N-H vibration band shift from 3351 to 3348 cm-1 also support the hydrogen bonding between LHC18 and TCPP. More important, the chiroptical properties of the hierarchically nanostructure were investigated by circular dichroism (CD) spectra. It could be found that an obvious negative Cotton effect is observed at around 420 nm for LHC18/TCPP system (Figure 5b), which is in accordance with the UV-vis absorbance of Soret band (B-band) of TCPP. The mirror signal for CD spectrum is obtained when the DHC18 is used, indicating that the chiral sense of LHC18 is effective to be transferred to TCPP chromophores. The hydrogen bonding between LHC18 and TCPP was supposed to cause this chirality transfer.

ACS Paragon Plus Environment

11

ACS Nano

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

Figure 5. (a) UV-visible spectra of TCPP cast film (black line), LHC18/TCPP cast film (red line) and LHC18/TCPP solution (blue line); (b) CD spectra of LHC18/TCPP (black line), DHC18/TCPP (red line) and racemate of LHC18/ DHC18 with TCPP (blue line) ; (c) FT-IR spectra of LHC18 ( black line), TCPP (red line) and LHC18/TCPP (blue line); (d) XRD patterns of TCPP (black line) and LHC18/TCPP (red line). The internal structures of the microflowers structures composed of LHC18 and TCPP were investigated by XRD analysis. As shown in Figure 5d, the XRD pattern of the TCPP cast film shows narrow diffraction peak at 2θ = 6.3 º(1.42 nm) in the low angle range, which is ascribed to the diffraction from the (100) plane. This (100) plane also gives higher order diffraction at 0.71 nm (200), 0.47 nm (300) in the wide angle region of the XRD pattern. On the basis of the geometry optimization and energy-minimized molecular structures of TCPP, the width of TCPP is about 1.42 nm. Thus, the strong peak observed at 1.42 nm can be assigned to the width of TCPP. TCPP was reported to form two-dimensional (2D) square-grid polymeric arrays and these layers stack tightly one on top of the other in an offset manner along the normal direction with the offset about 0.5 nm.52 Thus, a weak Bragg peak with a spacing of 0.45 nm might be ascribed to the offset of center of successive layers. Further, a refraction peak at 2θ = 25.6° corresponds

ACS Paragon Plus Environment

12

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

ACS Nano

to 0.34 nm, which can be assigned to the stacking distance between neighboring porphyrin rings along the direction perpendicular to the porphyrin rings. After interacting with LHC18, refraction bands at (100), (200), (300) are still observed, indicating that basic packing of TCPP is preserved in the complex. But the peak at 0.34 nm is relative weak comparing to that of TCPP, suggesting that the π−π stacking between porphyrin is weakened to some extent because of the interaction with amphiphilic histidine. In particular, a weak Bragg peak at 21.5° is observed, which corresponds to a spacing of 0.42 nm and agrees well with the lateral packing of long alkyl chains in its crystal structure.53 Then the XRD pattern suggests that the synergism of hydrophobic interaction between long alkyl chains and π−π stacking between porphyrin rings cause the hierarchical assembly into microflowers. Superhydrophobic and enantioselectivity of the chiral microflowers. It has been reported that the micro-and nanoscale hierarchical structures could induce super-hydrophobic structures with large contact angle.54-56 Inspired by this knowledge, we tested the wettability of surface containing microflower structures formed by LHC18 (DHC18)/TCPP complex. We made a cast film, which was composed of flower-like co-assemblies, by casting the solution on silicon slice. After evaporated the mixed solvents naturally, we obtained the cast films. Actually, the surface of the cast film was nearly superhydrophobic to the ultra-pure water (Milli-Q water) and the contact angle was about 140° (Figure S11). Due to the co-assemblies formed chiral flowers on the surface of silicon slice, the pre-modified surface may show chiral recognition to enatiomers.57 In this report, aspartic acid (Asp), glutamic acid (Glu), tyrosine (Tyr), phenylalanine (Phe), arginine (Arg) and alanine (Ala) were chosen for chiral recognition. Firstly, amino acids were

ACS Paragon Plus Environment

13

ACS Nano

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 14 of 30

dissolved into Milli-Q water by heating and the concentration of aqueous solution were 0.1 mM. The homogeneous distribution chiral films were made by casting the co-assemblies solution on the silicon slice. Then, we tested the contact angles (CA) of the made cast film under different amino acid solutions. Each chiral amino acid aqueous were tested five times using one chiral cast film, as shown in Figure 6. The average contact angle of L-Asp on LHC18/TCPP surface was about 159.2±3.8°, while contact angle for D-Asp was nearly 142.2±2.8°. It seems that the chiral film could recognize enantiomeric Aspartic acid. In contrast, DHC18/TCPP cast films were also investigated. As we expected, the average contact angle for L-Asp was 152.5±3.4°, while for the D-Asp was 161.9±4.1°, which confirmed the chiral films exhibiting enantiomeric recognition on Aspartic acid. Table S1 presents the contact angle of different cast films to enantiomer of Aspartic acid. It was found that the surface containing flower like structures also exhibited chiral recognition of enantiomeric Tyr and Phe (Figure S12, S13, Table S2, S3). However, the chiral surface showed almost the same contact angle to enantiomeric Glu, Arg and Ala (Figure S14-16 and Table S4-6). This indicated that the chiral superstructures had no recognition preference to Glu, Arg and Ala, meaning that the chiral interfaces of the microflowers are sensitive only to some amino acids although we did not know exactly the reason.

ACS Paragon Plus Environment

14

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

ACS Nano

Figure 6. Contact angle of (a) L-Asp on LHC18/TCPP cast film; (b) D-Asp on LHC18/TCPP cast film; (C) L-Asp on DHC18/TCPP cast film; (d) D-Asp on DHC18/TCPP cast film. DISCUSSION Based on the above FT-IR and XRD results, the self-assembly process as well as the formation of the flower-like structures is proposed, as illustrated in Scheme 1. Since TCPP molecule has four carboxylic acids and histidine derivative has an imidazole group, it is easy to self-assemble in DMF/H2O mixed solvents driven by the double hydrogen bonding and electrostatic interactions.51 It has been confirmed by FT-IR spectra that the hydrophobic interaction between LHC18 or DHC18 and hydrogen bonding between urea moiety of amphiphilic histidine link LHC18 or DHC18 together. By the aid of hydrogen bonding/electrostatic interactions between imidazole and carboxylic acid, LHC18 or DHC18 works as a “glue” to attach TCPP into a two-dimensional layered structures, as illustrated in Scheme 1. Similar 2D porphyrin layer structures formed through the coordination and surfactant

ACS Paragon Plus Environment

15

ACS Nano

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

assistance have been reported.58 Multilayered structures containing amphiphilic histidine stabilized TCPP layers start to aggregate into nanosheets or nanobelts. Because of the chirality of histidine, the nanosheet could bend at a certain direction. This chiral preference of bending accumulated and finally expressed as chiral microflowers. As shown in Scheme 1, LHC18 could induce the nanosheets bending at a clockwise direction and form right-handed flowers, while DHC18 led to an opposite bending direction.

Scheme 1. Schematic illustration of the interaction modes between histidine derivatives and TCPP. By the aid of hydrophobic interaction and hydrogen bonding between LHC18 ( DHC18) molecules, TCPP can form two-dimensional sheet. And the nanosheets could bend at a certain direction in the presence of chiral amphiphilic histidine. Since these microflowers contain the hierarchical nano and microstructures and the alkyl chains will occupy the surface, they show superhydrophobic property as revealed by many similar system.54 When these surfaces interact with the acidic amino acids, some of the amino acids could penetrate into the flower and react with the imidazolium, which changed the surface properties and corresponding contact angle. Due to the existence of chiral hierarchical

ACS Paragon Plus Environment

16

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

ACS Nano

architectures, as illustrated above, the enantioselectivity of this surface to some amino acids appeared via contact angle measurements. CONCLUSIONS

In summary, we have successfully fabricated the flower-like structures by using an achiral porphyrin and chiral amphiphilic histidine. Chiral sense was obviously observed, in which curved nanosheets arranged in a clockwise manner or counterclockwise manner depending on the absolute configuration of histidine. Both the SEM and CD spectral measurements confirmed that the molecular chirality of amphiphilic histidine (LHC18 or DHC18) could be transferred to achiral TCPP as well as the microflower structures. Due to the hierarchical nano/micro structures, the flower structures showed a superhydrophobic property. Combining with the chiral nature of the microflowers, it exhibited further enantioselectivity to some amino acids via contact angle measurement. This study demonstrated an example for large scale chirality transfer and control from a molecular level to a complex microflower, which could further feature the superhydrophobic and enantioselective sensing surface.

EXPERIMENTAL Materials. The amphiphilic histidine derivatives (LHC18 and DHC18) containing L/Dhistidine and octadecyl isocyanate were synthesized according to the method reported previously by our group.50 Tetrakis(4-carboxyphenyl)porphyrin (TCPP) was purchased from J&K. All chemical solvents were purchased from Beijing Chemicals. Milli-Q water (18.2 MΩ·cm) was used in all cases.

ACS Paragon Plus Environment

17

ACS Nano

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

Hierarchical co-assemblies formed in DMF/H2O mixed solvent. A typical procedure for co-assemblies in mixed solvent of N, N-dimethylformamide (DMF) and water is as follows: both LHC18 and TCPP firstly dissolved in 700 µL DMF until absolutely turned into transparent purple solution in a sealed vial, then added 300 µL water dropwisely into the as-prepared hot solution. Upon cooling to room temperature, the transparent solution turned into opaque suspension. Instruments. UV−vis spectra were recorded in quartz cuvettes (light path 0.1 mm) on a Hitachi U-3900 Spectrophotometer. Fourier transform-infrared (FT-IR) studies were recorded by using a Bruker TENSOR-27 spectrophotometer in the range of 400-4000 cm-1 with a wavenumber resolution of 4 cm-1 at room temperature. CD spectra were recorded in quartz cuvettes (light path 0.1 mm) on a JASCO J-810 spectrophotometer. For the measurement of the CD spectra, the quartz cuvettes were placed perpendicular to the light path of CD spectrometer and rotated within the quartz cuvettes plane to rule out the possibility of the birefringence phenomena and eliminate the contamination of Linear Dichroism (LD).59, 60 The fluorescence lifetime measurements were recorded on Edinburg FLS-980 fluorescence spectrometer using time-correlated single photon counting (TCSPC). SEM images were recorded on a Hitachi S4800 FE-SEM instrument with an accelerating voltage of 10 kV. Before SEM measurement, the samples on silicon wafers were coated with a thin layer of Pt to increase the contrast. XRD analysis was performed on a Rigaku D/Max-2500 X-ray diffractometer (Japan) with Cu Kα radiation (λ = 1.5406 Å), which was operated at a voltage of 40 kV and a current of 200 mA. Samples were cast on silicon substrates and vacuum-dried for XRD measurements.

ACS Paragon Plus Environment

18

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

ACS Nano

Contact Angle Measurement. The suspension liquid about 900 µL was slowly casted on an 1.5 × 1.5 cm silicon slice, which had an hydrophilic interface, and then the sample was evaporated naturally. The flower-likes structures are occupied the whole substrate, as shown in Figure S17. Water static contact angles were measured on an OCA20 machine (Data Physics, Germany) at ambient temperature. A water drop (2 µL) was suspended with a metal ring and controlled to squeeze the surface at a constant speed (0.005 mm·s−1) and then allowed to relax. The digital photographs were taken using a Nikon (Tokyo, Japan) D90 single Lens reflex camera. ASSOCIATED CONTENT Supporting Information. SEM and TEM images of LHC18 (DHC18)/TCPP assemblies, and contact angle results. This material is available free of charge via the Internet at http://pubs.acs.org. AUTHOR INFORMATION Corresponding Author * [email protected]; [email protected] ORCID Minghua Liu: 0000-0002-6603-1251 Funding Sources Authors Li Zhang and Minghua Liu received funding from the National Natural Science Foundation of China Grant (21473219 and 91427302), Basic Research Development Program (2013CB834504), and Strategic Priority Research Program” of the Chinese Academy of Sciences (XDB12020200).

ACS Paragon Plus Environment

19

ACS Nano

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

ACKNOWLEDGMENT This work was supported by the National Natural Science Foundation of China (Nos. 21473219 and 91427302), the Basic Research Development Program (2013CB834504), and Strategic Priority Research Program” of the Chinese Academy of Sciences (XDB12020200). REFERENCES 1.

Addadi, L.; Weiner, S. Biomineralization-Crystals, Asymmetry and Life. Nature 2001,

411, 753-755. 2.

Hegstrom, R. A.; Kondepudi, D. K. The Handedness of the Universe Sci. Am. 1990, 262,

108−115. 3.

Lehn, J. M. Perspectives in Supramolecular Chemistry - From Molecular Recognition

towards Molecular Information-Processing and Self-Organization. Angew. Chem., Int. Ed. 1990, 29, 1304-1319. 4.

Aida, T.; Meijer, E. W.; Stupp, S. I. Functional Supramolecular Polymers. Science 2012,

335, 813-817. 5.

Wang, Y.; Xu, J.; Wang, Y.; Chen, H. Emerging Chirality in Nanoscience. Chem. Soc.

Rev. 2013, 42, 2930-2962. 6.

Liu, M.; Zhang, L.; Wang, T. Supramolecular Chirality in Self-Assembled Systems.

Chem. Rev. 2015, 115, 7304-7397. 7.

Cui, H.; Webber, M. J.; Stupp, S. I. Self-Assembly of Peptide Amphiphiles: From

Molecules to Nanostructures to Biomaterials. Biopolymers, 2010, 94, 1-18.

ACS Paragon Plus Environment

20

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

ACS Nano

8.

Wang, M.; Zhou, P.; Wang, J.; Zhao, Y.; Ma, H.; Lu, J. R.; Xu, H. Left or Right: How

Does Amino Acid Chirality Affect the Handedness of Nanostructures Self-Assembled from Short Amphiphilic Peptides? J. Am. Chem. Soc. 2017, 139, 4185-4194. 9.

Wang, Y.; Qi, W.; Huang, R.; Yang, X.; Wang, M.; Su, R.; He, Z. Rational Design of

Chiral Nanostructures from Self-Assembly of a Ferrocene-Modified Dipeptide. J. Am. Chem. Soc. 2015, 137, 7869-7880. 10.

Wang, X.; Duan, P.; Liu, M. Universal Chiral Twist via Metal Ion Induction in the

Organogel of Terephthalic Acid Substituted Amphiphilic L-Glutamide. Chem. Commun. 2012, 48, 7501-7503. 11.

Cao, H.; Zhu, X.; Liu, M. Self-Assembly of Racemic Alanine Derivatives: Unexpected

Chiral Twist and Enhanced Capacity for the Discrimination of Chiral Species. Angew. Chem., Int. Ed. 2013, 52, 4122-4126. 12.

Liu, Y.; Wang, T.; Huan, Y.; Li, Z.; He. G.; Liu, M. Self-Assembled Supramolecular

Nanotube Yarn. Adv. Mater. 2013, 25, 5875-5879. 13.

Cui, J.; Zheng, Y.; Shen, Z.; Wan, X. Alkoxy Tail Length Dependence of Gelation

Ability and Supramolecular Chirality of Sugar-Appended Organogelators. Langmuir 2010, 26, 15508-15515. 14.

Zhou, Y.; Marson, R. L.; van Anders, G.; Zhu, J.; Ma, G.; Ercius, P.; Sun, K.; Yeom, B.;

Glotzer, S. C.; Kotov, N. A. Biomimetic Hierarchical Assembly of Helical Supraparticles from Chiral Nanoparticles. ACS Nano 2016, 10, 3248-3256.

ACS Paragon Plus Environment

21

ACS Nano

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

15.

Page 22 of 30

Feng, W.; Kim, J.-Y.; Wang, X.; Calcaterra, H. A.; Qu, Z.; Meshi, L.; Kotov, N. A.

Assembly of Mesoscale Helices with Near-Unity Enantiomeric Excess and Light-Matter Interactions for Chiral Semiconductors, Sci. Adv. 2017, 3, e1601159. 16.

Ma, W.; Xu, L.; de Moura, A. F.; Wu, X.; Kuang, H.; Xu, C.; Kotov, N. A. Chiral

Inorganic Nanostructures. Chem. Rev. 2017, 117, 8041-8093. 17.

Song, B.; Liu, B.; Jin, Y.; He, X.; Tang, D.; Wu, G.; Yin, S. Controlled Self-Assembly of

Helical Nano-Ribbons Formed by Achiral Amphiphiles. Nanoscale 2015, 7, 930-935. 18.

Zhang, Y.; Wang, S.; Liu, Y.; Jin, Y.; Xia. Y.; Song, B. Bilayers Directly Scrolling up to

Form Nanotubes via Self-Assembly of an Achiral Small Molecule. Nanoscale 2017, 9, 14911495. 19.

Jin, Q.; Zhang, L.; Liu, M. Solvent-Polarity-Tuned Morphology and Inversion of

Supramolecular Chirality in a Self-Assembled Pyridylpyrazole-Linked Glutamide Derivative: Nanofibers, Nanotwists, Nanotubes, and Microtubes. Chem. -Eur. J. 2013, 19, 9234-9241. 20.

Huo, H.; Wang, S.; Lin, S.; Li, Y.; Li, B.; Yang, Y. Chiral Zirconia Nanotubes Prepared

Through a Sol-Gel Transcription Approach. J. Mater. Chem. A 2014, 2, 333-338. 21.

Lidor-Shalev, O.; Pliatsikas, N.; Carmiel, Y.; Patsalas, P.; Mastai, Y. Chiral Metal-Oxide

Nanofilms by Cellulose Template Using Atomic Layer Deposition Process. ACS Nano 2017, 11, 4753-4759. 22.

Levi, G.; Scolnik, Y.; Mastai, Y. Imprinting Chirality in Silica Nanotubes by N-Stearoyl-

serine Template. ACS Appl. Mater. Interface 2016, 8, 23356-23361.

ACS Paragon Plus Environment

22

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

ACS Nano

23.

Duan, Y.; Liu, X.; Han, L.; Asahina, S.; Xu, D.; Cao,Y.; Yao, Y.; Che, S. Optically

Active Chiral CuO "Nanoflowers". J. Am. Chem. Soc. 2014, 136, 7193-7196. 24.

Zhang, L.; Wang, T.; Shen, Z.; Liu, M. Chiral Nanoarchitectonics: Towards the Design,

Self-Assembly, and Function of Nanoscale Chiral Twists and Helices. Adv. Mater. 2016, 28, 1044-1059. 25.

Jiang, W.; Pacella,M. S.; Athanasiadou, D.; Nelea, V.; Vali, H.; Hazen, R. M.; Gray, J. J.;

McKee, M. D. Chiral Acidic Amino Acids Induce Chiral Hierarchical Structure in Calcium Carbonate. Nat. Commun. 2017, 8, 15066. 26.

Ge, J.; Lei, J.; Zare, R. N. Protein-Inorganic Hybrid Nanoflower. Nat. Nanotech. 2012, 7,

428-432. 27.

Nakanishi, T.; Michinobu, T.; Yoshida, K.; Shirahata, N.; Ariga, K.; Mohwald, H.;

Kurth, D. G. Nanocarbon Superhydrophobic Surfaces Created from Fullerene-Based Hierarchical Supramolecular Assemblies. Adv. Mater. 2008, 20, 443–446. 28.

Salimimarand, M.; Duc La, D.; Kobaisi, Al M.; Bhosale, S. V. Flower-Like

Superstructures of AIE-Active Tetraphenylethylene through Solvophobic Controlled SelfAssembly. Sci. Rep. 2017, 7, 42898. 29.

Xie, Y.; Tang, Y.; Wu, W.; Wang, Y.; Liu, J.; Li, X.; Tian, H.; Zhu, W. Porphyrin

Cosensitization for a Photovoltaic Efficiency of 11.5%: A Record for Non-Ruthenium Solar Cells Based on Iodine Electrolyte. J. Am. Chem. Soc. 2015, 137, 14055-14058. 30.

Ding, Y.; Zhu, W.; Xie, Y. Development of Ion Chemosensors Based on Porphyrin

Analogues. Chem. Rev. 2017, 117, 2203-2256.

ACS Paragon Plus Environment

23

ACS Nano

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

31.

Page 24 of 30

Rong, Y.; Chen, P.; Liu, M. Self-Assembly of Water-Soluble TPPS in Organic Solvents:

From Nanofibers to Mirror Imaged Chiral Nanorods. Chem. Commun. 2013, 49, 10498-10500. 32.

Wang, L.; Chen, Y.; Jiang, J. Controlling the Growth of Porphyrin Based Nanostructures

for Tuning Third-Order NLO Properties. Nanoscale 2014, 6, 1871-1878. 33.

Sun, R.; Wang, L.; Tian, J.; Zhang, X.; Jiang, J. Self-Assembled Nanostructures of

Optically Active Phthalocyanine Derivatives. Effect of Central Metal Ion on the Morphology, Dimension, and Handedness. Nanoscale 2012, 4, 6990-6996. 34.

Zhao, Q.; Wang, Y.; Xu, Y.; Yan, Y.; Huang, J. Out-of-Plane Coordinated Porphyrin

Nanotubes with Enhanced Singlet Oxygen Generation Efficiency. Sci. Rep. 2016, 6, 31339. 35.

Bai, F.; Sun, Z.; Wu, H.; Haddad, R. E.; Coker, E. N.; Huang, J.; Rodriguez, M. A.;

Fan, H. Porous One-Dimensional Nanostructures through Confined Cooperative Self-Assembly. Nano Lett. 2011, 11, 5196-5200. 36.

Hu, J.; Guo, Y.; Liang, H.; Wan, L.; Jiang, L. Three-Dimensional Self-Organization of

Supramolecular Self-Assembled Porphyrin Hollow Hexagonal Nanoprisms. J. Am. Chem. Soc. 2005, 127, 17090-17095. 37.

Guo, P.; Chen, P.; Liu, M.; Porphyrin Assemblies via a Surfactant-Assisted Method:

From Nanospheres to Nanofibers with Tunable Length. Langmuir 2012, 28, 15482-15490. 38.

Zhao, L.; Qu, R.; Li, A.; Ma, R.; Shi, L. Cooperative Self-Assembly of Porphyrins with

Polymers Possessing Bioactive Functions. Chem. Commun. 2016, 52, 13543-13555.

ACS Paragon Plus Environment

24

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

ACS Nano

39.

Mandal, S.; Nayak, S. K.; Mallampalli, S.; Patra, A. Surfactant-Assisted Porphyrin Based

Hierarchical Nano/Micro Assemblies and Their Efficient Photocatalytic Behavior. ACS Appl. Mater. Interface 2014, 6, 130-136. 40.

Qiu, Y.; Chen, P.; Liu, M. Evolution of Various Porphyrin Nanostructures via an

Oil/Aqueous Medium: Controlled Self-Assembly, Further Organization, and Supramolecular Chirality. J. Am. Chem. Soc. 2010, 132, 9644-9652. 41.

Huang, C.; Li, Y.; Yang, J.; Cheng, N.; Liu, H.; Li, Y. Construction of Multidimensional

Nanostructures by Self-Assembly of A Porphyrin Analogue. Chem. Commun. 2010, 46, 31613163. 42.

Tian, Y.; Busani, T. G.; Uyeda, H.; Martin, K. E.; van Swol, F.; Medforth, C. J.;

Montano, G. A.; Shelnutt, J. A. Hierarchical Cooperative Binary Ionic Porphyrin Nanocomposites. Chem. Commun. 2012, 48, 4863-4865. 43.

Lu, G.; Chen, Y.; Zhang, Y.; Bao, M.; Bian, Y.; Li, X.; Jiang, J. Morphology Controlled

Self-Assembled Nanostructures of Sandwich Mixed (Phthalocyaninato)(Porphyrinato) Europium Triple-Deckers. Effect of Hydrogen Bonding on Tuning the Intermolecular Interaction. J. Am. Chem. Soc. 2008, 130, 11623-11630. 44.

Gao, Y.; Zhang, X.; Ma, C.; Li, X.; Jiang, J. Morphology-Controlled Self-Assembled

Nanostructures of 5,15-Di[4-(5-acetylsulfanylpentyloxy)phenyl]porphyrin Derivatives. Effect of Metal-Ligand Coordination Bonding on Tuning the Intermolecular Interaction. J. Am. Chem. Soc. 2008, 130, 17044-17052.

ACS Paragon Plus Environment

25

ACS Nano

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

45.

Page 26 of 30

Wang, L.; Chen, Y.; Bian, Y.; Jiang, J. Porphyrin Nanocrystal Synthesized via Chemical

Reaction Route: pH-Sensitive Reversible Transformation between Nanocrystals and Bulk Single Crystal. J. Phys. Chem. C 2013, 117, 17352-17359. 46.

Rananaware, A.; La, D. D.; Al Kobaisi, M.; Bhosale, R. S.; Bhosale, S. V.; Bhosale, S. V.

Controlled Chiral Supramolecular Assemblies of Water Soluble Achiral Porphyrins Induced By Chiral Counterions. Chem. Commun. 2016, 52, 10253-10256. 47.

Lu, H.; Kobayashi, N. Optically Active Porphyrin and Phthalocyanine Systems. Chem.

Rev. 2016, 116, 6184-6261. 48.

Charalambidis, G.; Georgilis, E.; Panda, M. K.; Anson, C. E.; Powell, A. K.; Doyle, S.;

Moss, D.; Jochum, T.; Horton, P. N.; Coles, S. J.; et al. A Switchable Self-Assembling and Disassembling Chiral System Based on A Porphyrin-Substituted Phenylalanine-Phenylalanine Motif. Nat. Commun. 2016, 7, 12657. 49.

Vela, S.; Augusto Berrocal, J.; Atienza, C.; Meijer, E. W.; Martín, N. Mesoscopic Helical

Architectures via Self-Assembly of Porphyrin-Based Discotic Systems. Chem. Commun. 2017, 53, 4084-4087. 50.

Chen, J.; Wang, T.; Liu, M. Chaperone Gelator for the Chiral Self-Assembly of All

Proteinogenic Amino Acids and Their Enantiomers. Chem. Commun. 2016, 52, 6123-6126. 51.

Chen, C.; Chen, J.; Wang, T.; Liu, M. Fabrication of Helical Nanoribbon Polydiacetylene

via Supramolecular Gelation: Circularly Polarized Luminescence and Novel Diagnostic Chiroptical Signals for Sensing. ACS Appl. Mater. Interface 2016, 8, 30608-30615.

ACS Paragon Plus Environment

26

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

ACS Nano

52.

George, S.; Lipstman, S.; Muniappan, S.; Goldberg, I. Porphyrin Network Solids:

Examples of Supramolecular Isomerism Noncentrosymmetric Architectures and Competing Solvation. CrystEngComm, 2006, 8, 417–424. 53.

Bao, Y.; Bi, L.; Wu, L.; Mal, S. S.; Kortz, U. Preparation and Characterization of

Langmuir-Blodgett Films of Wheel-Shaped Cu-20 Tungstophosphate and DODA by Two Different Strategies. Langmuir 2009, 25, 13000-13006. 54.

Feng, L.; Li, S.; Li, Y.; Li, H.; Zhang, L.; Zhai, J.; Song, Y.; Liu, B.; Jiang, L.; Zhu, D.

Super-Hydrophobic Surfaces: From Natural to Artificial. Adv. Mater. 2002, 14, 1857-1860. 55.

Sun, T.; Feng, L.; Gao, X.; Jiang, L. Bioinspired Surfaces with Special Wettability. Acc.

Chem. Res. 2005, 38, 644-652. 56.

Wang, S.; Liu, K.; Yao, X.; Jiang, L. Bioinspired Surfaces with Superwettability: New

Insight on Theory, Design, and Applications. Chem. Rev. 2015, 115, 8230-8293. 57.

Jung, S. H.; Kim, K. Y.; Ahn, A.; Choi, M. Y.; Jaworski, J.; Jung, J. H. Determining

Chiral Configuration of Diamines via Contact Angle Measurements on Enantioselective AlanineAppended Benzene-Tricarboxamide Gelators. ACS Appl. Mater. Interface 2016, 8, 14102-14108. 58.

Zhao, M.; Wang, Y.; Ma, Q.; Huang, Y.; Zhang, X.; Ping, J.; Zhang, Z.; Lu, Q.; Yu, Y.;

Xu, H.; et al. Ultrathin 2D Metal–Organic Framework Nanosheets. Adv. Mater. 2015, 27, 7372– 7378.

ACS Paragon Plus Environment

27

ACS Nano

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

59.

Page 28 of 30

Yuan, J.; Liu, M. Chiral Molecular Assemblies From a Novel Achiral Amphiphilic 2-

(Heptadecyl) Naphtha 2,3 Imidazole through Interfacial Coordination. J. Am. Chem. Soc. 2003, 125, 5051-5056. 60.

Tsuda, A.; Alam, M. A.; Harada, T.; Yamaguchi, T.; Ishii, N.; Aida, T. Spectroscopic

Visualization of Vortex Flows Using Dye-Containing Nanofibers. Angew. Chem., Int. Ed. 2007, 46, 8198-8202.

ACS Paragon Plus Environment

28

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

ACS Nano

For Table of Contents Only

ACS Paragon Plus Environment

29

ACS Nano

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

Table of content 78x39mm (300 x 300 DPI)

ACS Paragon Plus Environment

Page 30 of 30