Different Origins of Strain-Induced Chirality Inversion of Co2+-

Publication Date (Web): March 7, 2018. Copyright © 2018 American Chemical Society. *E-mail: [email protected]. Cite this:Chem. Mater. 30, 6, 2074-208...
0 downloads 0 Views 3MB Size
Article pubs.acs.org/cm

Cite This: Chem. Mater. 2018, 30, 2074−2083

Different Origins of Strain-Induced Chirality Inversion of Co2+Triggered Supramolecular Peptide Polymers Hyesong Park,†,§ Ka Young Kim,†,§ Sung Ho Jung,† Yeonweon Choi,† Hisako Sato,‡ and Jong Hwa Jung*,† †

Department of Chemistry and Research Institute of Natural Sciences, Gyeongsang National University, Jinju 660-701, Republic of Korea ‡ Graduate School of Science and Engineering, Department of Chemistry, Ehime University, Matsuyama 790-8577, Japan S Supporting Information *

ABSTRACT: We report a distinctly different dynamic helix inversion pathway of self-assembled terpyridine-based ligands composed of different numbers of peptide moieties with Co2+ and its amplification of strain-induced chirality from an achiral terpyridine moiety. The helical chirality of the metal centers, coordinated by terpyridine ligands, is controlled by strain-induced chirality with complex ligandto-Co2+ ratios. We also show that the distinct helical inversion mechanism is significantly dependent on the number of peptides attached to ligands. The helical inversion pathway of the self-assembled ligand (R-1 and S-1) complexes composed of one alanine analogue (R- or S-2-amino-1-propyl moiety) and one long saturated alkyl chain relies on two steps of chirality with different complex geometries, first from strain-induced chirality originating from an octahedral structure to octahedral structure with different helical direction and then on to helical chirality in a square-pyramidal structure. In contrast, the helix inversion of the selfassembled R-2 and S-2 complexes containing an alanine analogue and two glycine moieties with Co2+ was followed by one step to form two distinct coexisting complex geometries having the same helical direction. In particular, the circular dichroism (CD) intensities of the self-assembled R-1 and R-2 complexes with Co2+ were 900−1500-fold amplified compared to those of free R-1 and R-2. The Gibbs free energies of the self-assembled complexes with different geometries were also calculated by temperaturedependent CD observation; the square-pyramidal structure of the self-assembled R-1 complex with Co2+ was more stable than the self-assembled R-2 complex with Co2+. Furthermore, the self-assembled R-1 and S-1 complexes with 1.0 equiv of Co2+ could classify amino acids by their chirality. dialkylpolysilanes27 as well as the temperature-induced helicity inversions of switchable polyacetylenes,28−32 polydiarylsilanes,33 and polyisocyanates,34−36 among others.37−44 In particular, a variety of specific noncovalent interactions have been used to build such assemblies.45−49 Among them, metal-bound self-assemblies are the most extensively used strategy for constructing supramolecular architectures with chiral geometries.50−60 In most previous examples, these systems have relied on the presence of chiral ligands to bias the handedness of the dynamic metal complexes. Yashima and co-workers50 demonstrated remote-controlled dynamic chirality of metal centers coordinated by 2,2-bispyridine ligands bearing dynamic helical oligopeptides. In addition, Miyake et al. 21,41,51−53 demonstrated that a peptide-based ligand composed of a chiral hexacoordinated metal center and two achiral pentapeptide chains with a Co2+ ion dynamically converted a left-handed M cis-α structure to a right-handed P cis-α one upon addition of an NO3− anion with a change in coordination geometry by one step. We also reported that the

1. INTRODUCTION Control of helicity in dynamic supramolecular assemblies is an attractive challenge with implications for our understanding of functional biological helicity as well as for the possible development of novel chiral materials that could be useful in the production and separation of enantiomers. Furthermore, the supramolecular assemblies of which helicity can be controlled have a great potential to be extended to the development of useful chiral materials such as sensors,1−3 catalysts,4−6 switches,7,8 and memory devices9 with specific functionalities that involve chiral recognition and sensing, asymmetric catalysis, and chiral optoelectronic properties.10,11 The well-known right-handed double-helical structure representative of B-form DNA, for instance, can undergo inversion to generate the left-handed Z-form DNA helix by exposure to high-salt environments, chemical modification, or complexation with cations and other chemicals.12−20 While the diversity of synthetic systems adopting helical conformations is extensive,21−24 investigations showing controlled transformation between different helical senses are surprisingly rare. Existing synthetic systems known to undergo helical switching as a result of changes in their environment include the solventdependent inversion of helical metal complexes25,26 and © 2018 American Chemical Society

Received: January 5, 2018 Revised: March 7, 2018 Published: March 7, 2018 2074

DOI: 10.1021/acs.chemmater.8b00057 Chem. Mater. 2018, 30, 2074−2083

Article

Chemistry of Materials helicity of terpyridine-based ligands containing one alanine analogue moiety and one cis-double-bond-bridged long alkyl chain complex with Co2+ could be controlled upon addition of Co2+.61 More interestingly, microscopic observation in the gel state showed that both the left-handed M and the right-handed P of terpyridine-based ligand complexes with Co2+ coexisted in the range of 0.2−0.8 equiv of Co2+. Such mono- and oligopeptide groups have been used as chiral cores to transfer helicity to an achiral moiety at the molecular level and/or self-assembled supramolecules.62,63 Thus, the helicity of the self-assembled supramolecular complexes was mainly controlled by the stereoenantiomer of the peptide and/or complex geometry such as octahedral and square-pyramidal structures. Here, we report a distinctly different dynamic helical inversion mechanism of self-assembled terpyridine-based ligands composed of different numbers of peptide moieties in the presence of Co2+ and its dramatic enhancement of straininduced chirality from an achiral terpyridine moiety. When the chirality-switchable metal complex was combined with an alanine analogue and/or glycine moieties, the dynamic inversion of the strain-induced chirality of the self-assembled R-1 and S-1 complexes with Co2+ occurred in two steps: through octahedral, to octahedral with a different helical direction, to square-pyramidal structures.64,65 In contrast, the helix inversion of the self-assembled R-2 and S-2 complexes containing an alanine analogue and two glycine moieties with Co2+ occurred in one step with two distinct coexisting complex geometries having the same helical direction. To the best of our knowledge, these observations constitute the first observation of dynamic chirality inversion by a distinctly different pathway, which strongly depended upon the number of peptides. The dramatically enhanced helicity of the self-assembled complexes also originated from the strain-induced chirality of the achiral terpyridine moieties of the complex formation with Co2+ and not from the stereoenantiomeric alanine analogue.

2.2. Helicity Inversion of the Self-Assembled R-1 and S-1 Complexes. To determine the chiroptical response, we observed CD spectra of the self-assembled R-1 and S-1 complexes with Co(NO3)2 in THF. The CD spectrum of the self-assembled R-1 without Co2+ exhibited a weak positive signal at 285 nm, and the signal intensity gradually increased up to 0.5 equiv of Co2+ with taking the form of one symmetrical band with no shoulder, indicating one complex structure with right-handed (P) helicity (Figure 1A).21,41,51−53,61 The CD

2. RESULTS AND DISCUSSION 2.1. UV−Vis Study of Self-Assembled Complexes with Co2+. The absorption spectra of R-1 in tetrahydrofuran (THF) are shown in Figure S1A. Without Co2+, compound R-1 exhibited two absorption bands at 275 nm (ε275 = 3.02 × 104 M−1 cm−1) with a shoulder at 310 nm (ε310 = 4.62 × 103 M−1 cm−1), which was ascribed to the π−π* and n−π* transitions. The addition of Co2+ to the solution of R-1 in THF significantly changed the absorption spectra (Figure S1 in Supporting Information), increasing the intensity of bands at 310 and 320 nm. In particular, the absorption band at 320 nm was newly generated upon addition of Co2+, and this originated from metal−ligand charge transfer (MLCT) between Co2+ and the terpyridine moieties of R-1.66−69 The absorption band at 310 nm showed two-step slope changes at approximately 0.5 and 1.0 equiv of Co2+ (Figure S1B) that were ascribed to the high concentration of R-1. These results indicate that the selfassembled R-1 formed 1:2 (Co2+/R-1) and 1:1 complexes with Co2+ in THF. We also observed the d−d transition−absorption band of the self-assembled R-1 complexes with 0.5 and 1.0 equiv of Co2+. As shown in Figure S1C and E, the absorption band of the self-assembled R-1 in the presence of a Co2+ equivalent of 0.083 mM Lphenylalanine was added, whereas it was not changed upon addition of D-phenylalanine (Figures 6C and S19B). The helicity inversion of the self-assembled R-1 complex with 1.0 equiv of Co2+ upon addition of D-phenylalanine was also evidenced by the microscopic analysis. The SEM images clearly exhibited the morphological change from left-handed (M) helicity to right-handed (P) helicity upon addition of 0.166 mM D-phenylalanine (Figure S20). We further evaluated the sensing behavior of the self-assembled R-1 and S-1 complexes with 1.0 equiv of Co2+ on several classes of amino acids including valine, serine, leucine, histidine, alanine, and arginine. All of the tested amino acids exhibited classifiability by their chirality with the self-assembled R-1 and S-1 complexes. The negative CD signal of the self-assembled R-1 with 1 equiv of Co2+ was inverted upon addition of D-form amino acids with similar intensities; however, it was not changed with L-form amino acids (Figure S21). With the self-assembled S-1 complex with 1 equiv of Co2+, L-form amino acids were sensitive, showing exactly opposite tendencies. To investigate the influence of the amino acids on the helicity of the self-assembled complexes, we measured IR spectra of the self-assembled complexes of R-1 and S-1 with 1.0 equiv of Co2+ in the absence and the presence of amino acids. As shown in Figure S22A, CO bands of amide group in Dphenylalanine (1696 cm−1) and that in the self-assembled R-1 with 1.0 equiv of Co2+ (1645 cm−1) were shifted to a lower wavenumber and to a higher wavenumber, respectively, when D-phenylalanine was added to the self-assembled R-1 complex, indicating that the hydrogen-bonding interaction was formed between the amide group in R-1 and that in D-phenylalanine. In contrast, any shift was not shown upon addition of Lphenylalanine, implying the absence of the intermolecular interaction between the self-assembled R-1 complex and L-



ASSOCIATED CONTENT

* Supporting Information S

The Supporting Information is available free of charge on the ACS Publications website at DOI: 10.1021/acs.chemmater.8b00057. Experimental section, CD spectra, UV−vis spectra, ESI mass spectra, and SEM photograph data (PDF) 2080

DOI: 10.1021/acs.chemmater.8b00057 Chem. Mater. 2018, 30, 2074−2083

Article

Chemistry of Materials



(12) Dumat, B.; Larsen, A. F.; Wilhelmsson, L. M. Studying Z-DNA and B- to Z-DNA transitions using a cytosine analogue FRET-pair. Nucleic Acids Res. 2016, 44, No. e101. (13) Chen, F. Y.-H.; Park, S.; Otomo, H.; Sakashita, S.; Sugiyama, H. Investigation of B-Z transitions with DNA oligonucleotides containing 8-methylguanine. Artif. DNA PNA XNA 2014, 5, No. e28226. (14) Bhanjadeo, M. M.; Nayak, A. K.; Subudhi, U. Cerium chloride stimulated controlled conversion of B-to-Z DNA in self-assembled nanostructures. Biochem. Biophys. Res. Commun. 2017, 482, 916−921. (15) Choi, J. K.; D’Urso, A.; Balaz, M. Chiroptical properties of anionic and cationic porphyrins and metalloporphyrins in complex with left-handed Z-DNA and right-handed B-DNA. J. Inorg. Biochem. 2013, 127, 1−6. (16) Bae, S.; Kim, D.; Kim, K. K.; Kim, Y.-G.; Hohng, S. Intrinsic ZDNA Is Stabilized by the Conformational Selection Mechanism of ZDNA-Binding Proteins. J. Am. Chem. Soc. 2011, 133, 668−671. (17) Gregoliński, J.; Ślepokura, K.; Lisowski, J. Lanthanide Complexes of the Chiral Hexaaza Macrocycle and Its meso-Type Isomer: Solvent-Controlled Helicity Inversion. Inorg. Chem. 2007, 46, 7923−7934. (18) Gregoliński, J.; Starynowicz, P.; Hua, K. T.; Lunkley, J. L.; Muller, G.; Lisowski, J. Helical Lanthanide(III) Complexes with Chiral Nonaaza Macrocycle. J. Am. Chem. Soc. 2008, 130, 17761−17773. (19) Gregoliński, J.; Lisowski, J. Helicity Inversion in Lanthanide(III) Complexes with Chiral Nonaaza Macrocyclic Ligands. Angew. Chem., Int. Ed. 2006, 45, 6122−6126. (20) Leonzio, M.; Melchior, A.; Faura, G.; Tolazzi, M.; Zinna, F.; Di Bari, L.; Piccinelli, F. Strongly Circularly Polarized Emission from Water-Soluble Eu(III)- and Tb(III)-Based Complexes: A Structural and Spectroscopic Study. Inorg. Chem. 2017, 56, 4413−4421. (21) Miyake, H.; Yoshida, K.; Sugimoto, H.; Tsukube, H. Dynamic Helicity Inversion by Achiral Anion Stimulus in Synthetic Labile Cobalt(II) Complex. J. Am. Chem. Soc. 2004, 126, 6524−6525. (22) D’Alonzo, D.; Froeyen, M.; Schepers, G.; Di Fabio, G.; Van Aerschot, A.; Herdewijn, P.; Palumbo, G.; Guaragna, A. 1′,5′Anhydro-L-ribo-hexitol Adenine Nucleic Acids (α-L-HNA-A): Synthesis and Chiral Selection Properties in the Mirror Image World. J. Org. Chem. 2015, 80, 5014−5022. (23) Yu, Z.-P.; Ma, C.-H.; Wang, Q.; Liu, N.; Yin, J.; Wu, Z.-Q. Polyallene-block-polythiophene-block-polyallene Copolymers: One-Pot Synthesis, Helical Assembly, and Multiresponsiveness. Macromolecules 2016, 49, 1180−1190. (24) Mathieu, L.; Legrand, B.; Deng, C.; Vezenkov, L.; Wenger, E.; Didierjean, C.; Amblard, M.; Averlant-Petit, M.-C.; Masurier, N.; Lisowski, V.; Martinez, J.; Maillard, L. T. Helical Oligomers of Thiazole-Based γ-Amino Acids: Synthesis and Structural Studies. Angew. Chem., Int. Ed. 2013, 52, 6006−6010. (25) Merten, C.; McDonald, R.; Xu, Y. Strong Solvent-Dependent Preference of Δ and Λ Stereoisomers of a Tris(diamine)nickel(II) Complex Revealed by Vibrational Circular Dichroism Spectroscopy. Inorg. Chem. 2014, 53, 3177−3182. (26) Nagata, Y.; Yamada, T.; Adachi, T.; Akai, Y.; Yamamoto, T.; Suginome, M. Solvent-Dependent Switch of Helical Main-Chain Chirality in Sergeants-and-Soldiers-Type Poly(quinoxaline-2,3-diyl)s: Effect of the Position and Structures of the “Sergeant” Chiral Units on the Screw-Sense Induction. J. Am. Chem. Soc. 2013, 135, 10104− 10113. (27) Suzuki, N.; Fujiki, M.; Kimpinde-Kalunga, R.; Koe, J. R. Chiroptical Inversion in Helical Si−Si Bond Polymer Aggregates. J. Am. Chem. Soc. 2013, 135, 13073−13079. (28) Miyagawa, T.; Furuko, A.; Maeda, K.; Katagiri, H.; Furusho, Y.; Yashima, E. Dual Memory of Enantiomeric Helices in a Polyacetylene Induced by a Single Enantiomer. J. Am. Chem. Soc. 2005, 127, 5018− 5019. (29) Maeda, K.; Miyagawa, T.; Furuko, A.; Onouchi, H.; Yashima, E. Dual Memory of Enantiomeric Helices in Poly(phenylacetylene)s Induced by a Single Enantiomer through Helix Inversion and Dual Storage of the Enantiomeric Helicity Memories. Macromolecules 2015, 48, 4281−4293.

AUTHOR INFORMATION

Corresponding Author

*E-mail: [email protected]. ORCID

Hisako Sato: 0000-0002-3927-1794 Jong Hwa Jung: 0000-0002-8936-2272 Author Contributions §

H.P. and K.Y.K. contributed equally.

Notes

The authors declare no competing financial interest.



ACKNOWLEDGMENTS This work was supported by the NRF (2015R1A2A2A05001400, 2018R1A2B2003637, and 2017R1A4A1014595) supported by the Ministry of Education, Science and Technology, Korea. In addition, this work was partially supported by a grant from the Next-Generation BioGreen 21 Program (SSAC, Grant no. PJ013186052018), Rural Development Administration, Korea.



REFERENCES

(1) Fu, C.; Liu, C.; Li, Y.; Guo, Y.; Luo, F.; Wang, P.; Guo, L.; Qiu, B.; Lin, Z. Homogeneous Electrochemical Biosensor for Melamine Based on DNA Triplex Structure and Exonuclease III- Assisted Recycling Amplification. Anal. Chem. 2016, 88, 10176−10182. (2) Suk, J.-m.; Naidu, V. R.; Liu, X.; Lah, M. S.; Jeong, K.-S. A Foldamer-Based Chiroptical Molecular Switch That Displays Complete Inversion of the Helical Sense upon Anion Binding. J. Am. Chem. Soc. 2011, 133, 13938−13941. (3) 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 Alanine-Appended BenzeneTricarboxamide Gelators. ACS Appl. Mater. Interfaces 2016, 8, 14102−14108. (4) Zheng, Y.; Tan, Y.; Harms, K.; Marsch, M.; Riedel, R.; Zhang, L.; Meggers, E. Octahedral Ruthenium Complex with Exclusive MetalCentered Chirality for Highly Effective Asymmetric Catalysis. J. Am. Chem. Soc. 2017, 139, 4322−4325. (5) Taura, D.; Hioki, S.; Tanabe, J.; Ousaka, N.; Yashima, E. Cobalt(II)-Salen-Linked Complementary Double-Stranded Helical Catalysts for Asymmetric Nitro-Aldol Reaction. ACS Catal. 2016, 6, 4685−4689. (6) Aoki, T.; Kaneko, T.; Maruyama, N.; Sumi, A.; Takahashi, M.; Sato, T.; Teraguchi, M. Helix-Sense-Selective Polymerization of Phenylacetylene Having Two Hydroxy Groups Using a Chiral Catalytic System. J. Am. Chem. Soc. 2003, 125, 6346−6347. (7) Pijper, D.; Jongejan, M. G. M.; Meetsma, A.; Feringa, B. L. Control of Dynamic Helicity at the Macro- and Supramolecular Level. J. Am. Chem. Soc. 2008, 130, 4541−4552. (8) Maeda, K.; Mochizuki, H.; Watanabe, M.; Yashima, E. Switching of Macromolecular Helicity of Optically Active Poly(phenylacetylene)s Bearing Cyclodextrin Pendants Induced by Various External Stimuli. J. Am. Chem. Soc. 2006, 128, 7639−7650. (9) Maeda, K.; Miyagawa, T.; Furuko, A.; Onouchi, H.; Yashima, E. Dual Memory of Enantiomeric Helices in Poly(phenylacetylene)s Induced by a Single Enantiomer through Helix Inversion and Dual Storage of the Enantiomeric Helicity Memories. Macromolecules 2015, 48, 4281−4293. (10) Mukai, A.; Fukai, T.; Matsumoto, M.; Ishikawa, J.; Hoshino, Y.; Yazawa, K.; Harada, K.; Mikami, Y. Transvalencin Z, A New Antimicrobial Compound with Salicylic Acid Residue from Nocardia transvalensis IFM 10065. J. Antibiot. 2006, 59, 366−369. (11) Suzuki, J.; Ishida, T.; Shibuya, I.; Toda, K. Development of a New Acaricide, Etoxazole. J. Pestic. Sci. 2001, 26, 215−223. 2081

DOI: 10.1021/acs.chemmater.8b00057 Chem. Mater. 2018, 30, 2074−2083

Article

Chemistry of Materials (30) Sakurai, S.; Okoshi, K.; Kumaki, J.; Yashima, E. TwoDimensional Surface Chirality Control by Solvent-Induced Helicity Inversion of a Helical Polyacetylene on Graphite. J. Am. Chem. Soc. 2006, 128, 5650−5651. (31) Zhou, Y.; Zhang, C.; Qiu, Y.; Liu, L.; Yang, T.; Dong, H.; Satoh, T.; Okamoto, Y. Temperature-Triggered Switchable Helix-Helix Inversion of Poly(phenylacetylene) Bearing L-Valine Ethyl Ester Pendants and Its Chiral Recognition Ability. Molecules 2016, 21, 1583. (32) Shimomura, K.; Ikai, T.; Kanoh, S.; Yashima, E.; Maeda, K. Switchable enantioseparation based on macromolecular memory of a helical polyacetylene in the solid state. Nat. Chem. 2014, 6, 429−434. (33) Tang, K.; Green, M. M.; Cheon, K. S.; Selinger, J. V.; Garetz, B. A. Chiral Conflict. The Effect of Temperature on the Helical Sense of a Polymer Controlled by the Competition between Structurally Different Enantiomers: From Dilute Solution to the Lyotropic Liquid Crystal State. J. Am. Chem. Soc. 2003, 125, 7313−7323. (34) Maxein, G.; Zentel, R. Photochemical Inversion of the Helical Twist Sense in Chiral Polyisocyanates. Macromolecules 1995, 28, 8438−8440. (35) Roth, J.; O’Leary, D. J.; Wade, C. G.; Miller, D.; Armstrong, K. B.; Thoburn, J. D. Conformational Analysis of Alkylated Biuret and Triuret: Evidence for Helicity and Helical Inversion in Oligoisocyanates. Org. Lett. 2000, 2, 3063−3066. (36) Pijper, D.; Feringa, B. L. Molecular transmission: controlling the twist sense of a helical polymer with a single light-driven molecular motor. Angew. Chem., Int. Ed. 2007, 46, 3693−3696. (37) Nagata, Y.; Takeda, R.; Suginome, M. Pressure-dependent helix inversion of poly(quinoxaline-2,3-diyl)s containing chiral side chains in non-aqueous solvents. Chem. Commun. 2015, 51, 11182−11185. (38) Nagata, Y.; Nishikawa, T.; Suginome, M. Poly(quinoxaline-2,3diyl)s Bearing (S)-3-Octyloxymethyl Side Chains as an Efficient Amplifier of Alkane Solvent Effect Leading to Switch of Main-Chain Helical Chirality. J. Am. Chem. Soc. 2014, 136, 15901−15904. (39) Nagata, Y.; Kuroda, T.; Takagi, K.; Suginome, M. Ether solventinduced chirality inversion of helical poly(quinoxaline-2,3-diyl)s containing L-lactic acid derived side chains. Chem. Sci. 2014, 5, 4953−4956. (40) Yoshinaga, Y.; Yamamoto, T.; Suginome, M. ChiralitySwitchable 2,2′-Bipyridine Ligands Attached to Helical Poly(quinoxaline-2,3-diyl)s for Copper-Catalyzed Asymmetric Cyclopropanation of Alkenes. ACS Macro Lett. 2017, 6, 705−710. (41) Gregoliński, J.; Hikita, M.; Sakamoto, T.; Sugimoto, H.; Tsukube, H.; Miyake, H. Redox-Triggered Helicity Inversion in Chiral Cobalt Complexes in Combination with H+ and NO3− Stimuli. Inorg. Chem. 2016, 55, 633−643. (42) Zhao, D.; van Leeuwen, T.; Cheng, J.; Feringa, B. L. Dynamic control of chirality and self-assembly of double-stranded helicates with light. Nat. Chem. 2017, 9, 250−256. (43) Kumar, J.; Nakashima, T.; Kawai, T. Inversion of Supramolecular Chirality in Bichromophoric Perylene Bisimides: Influence of Temperature and Ultrasound. Langmuir 2014, 30, 6030−6037. (44) Cheng, Z.; Li, K.; Wang, F.; Wu, X.; Xiao, J.; Zhang, H.; Cao, H.; Yang, H. A helix inversion from the temperature-dependent intramolecular chiral conflict. Liq. Cryst. 2011, 38, 633−638. (45) Wang, C. X.; Utech, S.; Gopez, J. D.; Mabesoone, M. F. J.; Hawker, C. J.; Klinger, D. Non-Covalent Microgel Particles Containing Functional Payloads: Coacervation of PEG-Based Triblocks via Microfluidics. ACS Appl. Mater. Interfaces 2016, 8, 16914− 16921. (46) Beyler, M.; Heitz, V.; Sauvage, J.-P.; Ventura, B.; Flamigni, L.; Rissanen, K. Three-Component Noncovalent Assembly Consisting of a Central Tetrakis-4-pyridyl Porphyrin and Two Lateral Gable-Like Bis-Zn Porphyrins. Inorg. Chem. 2009, 48, 8263−8270. (47) Yang, H.-B.; Ghosh, K.; Northrop, B. H.; Zheng, Y.-R.; Lyndon, M. M.; Muddiman, D. C.; Stang, P. J. A Highly Efficient Approach to the Self-Assembly of Hexagonal Cavity-Cored Tris[2]pseudorotaxanes from Several Components via Multiple Noncovalent Interactions. J. Am. Chem. Soc. 2007, 129, 14187−14189.

(48) Rousseaux, S. A. L.; Gong, J. Q.; Haver, R.; Odell, B.; Claridge, T. D. W.; Herz, L. M.; Anderson, H. L. Self-Assembly of Russian Doll Concentric Porphyrin Nanorings. J. Am. Chem. Soc. 2015, 137, 12713−12718. (49) Li, F.; Yager, K. G.; Dawson, N. M.; Jiang, Y.-B.; Malloy, K. J.; Qin, Y. Stable and Controllable Polymer/Fullerene Composite Nanofibers through Cooperative Noncovalent Interactions for Organic Photovoltaics. Chem. Mater. 2014, 26, 3747−3756. (50) Ousaka, N.; Takeyama, Y.; Iida, H.; Yashima, E. Chiral information harvesting in dendritic metallopeptides. Nat. Chem. 2011, 3, 856−861. (51) Miyake, H.; Sugimoto, H.; Tamiaki, H.; Tsukube, H. Dynamic helicity inversion in an octahedral cobalt(II) complex system via solvato-diastereomerism. Chem. Commun. 2005, 0, 4291−4293. (52) Miyake, H. Supramolecular Chirality in Dynamic Coordination Chemistry. Symmetry 2014, 6, 880−895. (53) Miyake, H.; Kamon, H.; Miyahara, I.; Sugimoto, H.; Tsukube, H. Time-Programmed Peptide Helix Inversion of a Synthetic Metal Complex Triggered by an Achiral NO3‑ Anion. J. Am. Chem. Soc. 2008, 130, 792−793. (54) Akine, S.; Sairenji, S.; Taniguchi, T.; Nabeshima, T. Stepwise Helicity Inversions by Multisequential Metal Exchange. J. Am. Chem. Soc. 2013, 135, 12948−12951. (55) 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. (56) El-ghayoury, A.; Hofmeier, H.; Schenning, A. P. H. J.; Schubert, U. S. Self-assembled chiral terpyridine ruthenium complexes. Tetrahedron Lett. 2004, 45, 261−264. (57) Wang, P.; Moorefield, C. N.; Panzer, M.; Newkome, G. R. Helical and polymeric nanostructures assembled from benzene tri- and tetracarboxylic acids associated with terpyridine copper(II) complexes. Chem. Commun. 2005, 0, 465−467. (58) Yeung, C.-T.; Yeung, H.-L.; Tsang, C.-S.; Wong, W.-Y.; Kwong, H.-L. Supramolecular double helical Cu(I) complexes for asymmetric cyclopropanation. Chem. Commun. 2007, 0, 5203−5205. (59) Leung, S. Y.-L.; Yam, V. W.-W. Hierarchical helices of helices directed by Pt···Pt and π−π stacking interactions: reciprocal association of multiple helices of dinuclear alkynylplatinum(II) complex with luminescence enhancement behavior. Chem. Sci. 2013, 4, 4228−4234. (60) Fu, H. L.-K.; Leung, S. Y.-L.; Yam, V. W.-W. A rational molecular design of triazine-containing alkynylplatinum(II) terpyridine complexes and the formation of helical ribbons via Pt···Pt, π−π stacking and hydrophobic−hydrophobic interactions. Chem. Commun. 2017, 53, 11349−11352. (61) Park, S. H.; Jung, S. H.; Ahn, J.; Lee, J. H.; Kwon, K.-Y.; Jeon, J.; Kim, H.; Jaworski, J.; Jung, J. H. Reversibly tunable helix inversion in supramolecular gels trigged by Co2+. Chem. Commun. 2014, 50, 13495−13498. (62) Wechsel, R.; Maury, J.; Fremaux, J.; France, S. P.; Guichard, G.; Clayden, J. Inducing achiral aliphatic oligoureas to fold into helical conformations. Chem. Commun. 2014, 50, 15006−15009. (63) Solà, J.; Helliwell, M.; Clayden, J. N- versus C-Terminal Control over the Screw-Sense Preference of the Configurationally Achiral, Conformationally Helical Peptide Motif Aib8GlyAib8. J. Am. Chem. Soc. 2010, 132, 4548−4549. (64) Zeller, E.; Beruda, H.; Kolb, A.; Bissinger, P.; Riede, J.; Schmidbaur, H. Change of coordination from tetrahedral gold− ammonium to square-pyramidal gold−arsonium cations. Nature 1991, 352, 141−143. (65) Chisca, D.; Croitor, L.; Coropceanu, E. B.; Petuhov, O.; Baca, S. G.; Krämer, K.; Liu, S.-X.; Decurtins, S.; Rivera-Jacquez, H. J.; Masunov, A. E.; Fonari, M. S. From pink to blue and back to pink again: changing the Co(II) ligation in a two-dimensional coordination network upon desolvation. CrystEngComm 2016, 18, 384−389. (66) Indelli, M. T.; Bignozzi, C. A.; Scandola, F.; Collin, J.-P. Design of Long-Lived Ru(II) Terpyridine MLCT States. Tricyano Terpyridine Complexes. Inorg. Chem. 1998, 37, 6084−6089. 2082

DOI: 10.1021/acs.chemmater.8b00057 Chem. Mater. 2018, 30, 2074−2083

Article

Chemistry of Materials (67) Fortage, J.; Dupeyre, G.; Tuyèras, F.; Marvaud, V.; Ochsenbein, P.; Ciofini, I.; Hromadová, M.; Pospísil, L.; Arrigo, A.; Trovato, E.; Puntoriero, F.; Lainé, P. P.; Campagna, S. Molecular Dyads of Ruthenium(II)− or Osmium(II)−Bis(terpyridine) Chromophores and Expanded Pyridinium Acceptors: Equilibration between MLCT and Charge-Separated Excited States. Inorg. Chem. 2013, 52, 11944− 11955. (68) Chung, S.-K.; Tseng, Y.-R.; Chen, C.-Y.; Sun, S.-S. A Selective Colorimetric Hg2+ Probe Featuring a Styryl Dithiaazacrown Containing Platinum(II) Terpyridine Complex through Modulation of the Relative Strength of ICT and MLCT Transitions. Inorg. Chem. 2011, 50, 2711−2713. (69) Lebon, E.; Sylvain, R.; Piau, R. E.; Lanthony, C.; Pilmé, J.; Sutra, P.; Boggio-Pasqua, M.; Heully, J.-L.; Alary, F.; Juris, A.; Igau, A. Phosphoryl Group as a Strong σ-Donor Anionic Phosphine-Type Ligand: A Combined Experimental and Theoretical Study on LongLived Room Temperature Luminescence of the [Ru(tpy)(bpy)(Ph2PO)]+ Complex. Inorg. Chem. 2014, 53, 1946−1948. (70) Inada, Y.; Hotta, N.; Kuwabara, H.; Funahashi, S. Spectrophotometric Analysis of 5-Coordinate Cobalt(II) Species for Ligand Substitution of Hexakis(acetonitrile)cobalt(II) with Bulky 1,1,3,3-Tetramethylurea in Noncoordinating Nitromethane. Anal. Sci. 2001, 17, 187−191. (71) Pal, R. R.; Higuchi, M.; Kurth, D. G. Optically Active MetalloSupramolecular Polymers Derived from Chiral Bis-terpyridines. Org. Lett. 2009, 11, 3562−3565. (72) Dawson, D. M.; Ke, Z.; Mack, F. M.; Doyle, R. A.; Bignami, G. P. M.; Smellie, I. A.; Bühl, M.; Ashbrook, S. E. Calculation and experimental measurement of paramagnetic NMR parameters of phenolic oximate Cu(II) complexes. Chem. Commun. 2017, 53, 10512−10515. (73) Smulders, M. M. J.; Nieuwenhuizen, M. M. L.; de Greef, T. F. A.; van der Schoot, P.; Schenning, A. P. H. J.; Meijer, E. W. How to Distinguish Isodesmic from Cooperative Supramolecular Polymerisation. Chem. - Eur. J. 2010, 16, 362−367. (74) Saito, N.; Kobayashi, H.; Yamaguchi, M. Inverse” thermoresponse: heat-induced double-helix formation of an ethynylhelicene oligomer with tri(ethylene glycol) termini. Chem. Sci. 2016, 7, 3574− 3580. (75) Markvoort, A. J.; ten Eikelder, H. M. M.; Hilbers, P. A. J.; de Greef, T. F. A.; Meijer, E. W. Temperature-Triggered Switchable Helix-Helix Inversion of Poly(phenylacetylene) Bearing l-Valine Ethyl Ester Pendants and Its Chiral Recognition Ability. Nat. Commun. 2011, 2, 509. (76) Lee, Y.-M.; Kim, H.-E.; Park, C.-J.; Lee, A.-R.; Ahn, H.-C.; Cho, S.-J.; Choi, K.-H.; Choi, B.-S.; Lee, J.-H. NMR Study on the B−Z Junction Formation of DNA Duplexes Induced by Z-DNA Binding Domain of Human ADAR1. J. Am. Chem. Soc. 2012, 134, 5276−5283.

2083

DOI: 10.1021/acs.chemmater.8b00057 Chem. Mater. 2018, 30, 2074−2083