Multilevel Manipulation of Supramolecular Structures of Giant

May 17, 2018 - ... Advertisers · Institutional Sales; Live Chat. Partners. Atypon · CHORUS · COPE · COUNTER · CrossRef · CrossCheck Depositor · Orcid ...
2 downloads 0 Views 3MB Size
Letter Cite This: ACS Macro Lett. 2018, 7, 635−640

pubs.acs.org/macroletters

Multilevel Manipulation of Supramolecular Structures of Giant Molecules via Macromolecular Composition and Sequence Wei Zhang,† Shuailin Zhang,† Qingyun Guo,† Xinlin Lu,† Yuchu Liu,† Jialin Mao,‡ Chrys Wesdemiotis,†,‡ Tao Li,§,∥ Yiwen Li,*,⊥ and Stephen Z. D. Cheng*,† †

Department of Polymer Science, College of Polymer Science and Polymer Engineering, The University of Akron, Akron, Ohio 44325-3909, United States ‡ Department of Chemistry, The University of Akron, Akron, Ohio 44325-3601, United States § X-ray Science Division, Advanced Photon Source, Argonne National Laboratory, Argonne, Illinois 60439, United States ∥ Department of Chemistry and Biochemistry, Northern Illinois University, DeKalb, Illinois 60115, United States ⊥ College of Polymer Science and Engineering, State Key Laboratory of Polymer Materials Engineering, Sichuan University, Chengdu, 610065, China S Supporting Information *

ABSTRACT: We have successfully synthesized a series of monodispersed chain-like giant molecules with precisely controlled macromolecular composition and sequence based on polyhedral oligomeric silsesquioxane (POSS) nanoparticles using an orthogonal “click” strategy. Their nonspherical supramolecular structures, such as lamellae, double gyroids, and hexagonal packed cylinders, are mainly determined by the composition (namely, the number of incorporated amphiphilic nanoparticles). In addition, by precisely alternating the sequence of arranged nanoparticles in the giant molecules with identical chemical compositions, the domain sizes of their supramolecular structures could be fine-tuned. This is attributed to the macromolecular conformational differences caused by collective hydrogen bonding interactions in each set of sequence isomeric giant molecules. This work has demonstrated multilevel manipulation of supramolecular structures of giant molecules: coarse tuning by composition and fine-tuning by sequence.

N

having the exact molecular mass at the value of Mn, assuming the molecular mass distribution obeys a Gaussian distribution. Note that the polydispersity parameter could also lead to the phase structure shift in the composition window of block copolymer self-assemblies.8 Furthermore, those polymerization methodologies for monomer sequence control have also been advanced to afford random, block, gradient copolymers, and others.9−13 Yet, to precisely control the sequences at the single monomer level remains to be difficult.2 Very recently, a few pioneering works have been devoted to address this issue by using iteratively performed, efficient, and orthogonal organic transformations.2,13−23 For instance, Johnson, Lutz, and Meier et al. have used “click” and epoxide opening reactions, phosphoramidite couplings, and Passerini three-component reactions to construct sequence-defined macromolecules. 16,20,23−25 Although several synthetic routes have been documented, it remains to be very interesting of further knowing how

atural polymers are usually composed with small building blocks via precise chemical connections in terms of molecular mass, composition, sequence, topology, and stereochemistry. The functions and properties of natural polymers could be vividly different when one of those macromolecular parameters is changed. They have inspired polymer chemists to develop synthetic macromolecular materials with controllable chemical parameters and functions in a similar precise manner.1,2 For well controlling macromolecular weight or composition, several polymerization methodologies have been developed to achieve narrow and controlled molecular mass distributions, such as living anionic polymerization (LAP), atom-transfer radical-polymerization (ATRP),3,4 ring-opening metathesis polymerization (ROMP),5 and others.6,7 Many investigations have demonstrated that the molecular masses and chemical compositions are key parameters to affect the structures and properties of polymeric materials. Nevertheless, in most cases, precisely controlled, monodispersed macromolecular systems have not yet been achieved. For example, a polystyrene from well-controlled polymerization having a number-average molecular mass (Mn) of about 4000 Da with a PDI = 1.01 or 1.02 only contains about a 10% or 7% fraction © XXXX American Chemical Society

Received: April 13, 2018 Accepted: May 15, 2018

635

DOI: 10.1021/acsmacrolett.8b00275 ACS Macro Lett. 2018, 7, 635−640

Letter

ACS Macro Letters

Figure 1. (a) Orthogonal “click” strategy for preparing the precise chain-like giant molecules; (b) cartoons of the resulting giant molecules; (c) chemical structures of the key building blocks: DIBO-V-CHO, DIBO-B-CHO, and the “click adaptor”.

monomer sequences in those synthetic macromolecules affect their supramolecular structures as well as their properties.26 Our group tried to use “giant molecules”, which are macromolecules with nanobuilding blocks, such as polyhedral oligomeric silsesquioxane (POSS) cages,21,26−31 as the model system to study the influence of molecular composition and sequence. Our first effort was to prepare a series of amphiphilic chainlike giant molecules (DPOSS)(BPOSS)n (n = 1−5) based on two kinds of POSS macromonomers (i.e., DPOSS denotes 14 hydroxyl groups functionalized hydrophilic POSS cage, BPOSS refers to 7 isobutyl groups functionalized hydrophobic POSS cage). It was found that versatile self-assembled phase structures, including conventional lamellae (LAM), double gyroids (DG), hexagonal packed cylinder (HEX), and bodycentered cubic (BCC) structures, and unconventional FrankKasper and dodecagonal quasicrystal phase structures, can be formed by tuning the molecular composition and sequence.26 More interestingly, sequence-mandated different spherical phase structures was observed for the first time in the isomers of (DPOSS)(BPOSS)4 and (DPOSS)(BPOSS)5. However, the number of sequence isomers is limited for short chains (n < 3) to study the sequence effect in the nonspherical phase structure (i.e., LAM, DG, and HEX) formations. Moreover, the sequence change so far has only been focused on a single hydrophilic DPOSS cage associated with different numbers of hydrophobic BPOSS cages. What will happen if we introduce two DPOSS cages associated with BPOSS cages with different but symmetrical sequences? To fully address this issue, we have specifically designed and prepared another series of amphiphilic giant molecules with a broader variety of sequence macromolecular isomers (Figure 1). POSS cages with different surface chemistry were connected in predesigned positions. The resulting giant molecules possess

molecular masses of up to larger than 10k Da yet are still monodispersed. The collective interactions of two DPOSS cages in the sequence could greatly alter single molecular conformation and thus, generate large enough size differences in nonspherical self-assembled structures. The synthetic strategy employs sequentially performed strain-promoted azide−alkyne cycloadditions (SPAAC) and oxime ligations, as shown in Figure 1a. The detailed synthetic routes and chemical structures are available in Supporting Information (SI). The building blocks such as DIBO-V-CHO, DIBO-B-CHO, and the linear “click adaptor” (Figure 1c) were presynthesized as previously reported.21,32 A small molecule with two azido sites was employed as the starting material for controlled macromolecular growth, then two kinds of POSS cages (e.g., BPOSS and VPOSS cages, the latter one is the precursor for further transferring into hydrophilic DPOSS) could be incorporated in one reaction cycle. After four reaction cycles, eight POSS cages with a total molecular weight of over 10k Da could be precisely attached in a linear configuration. Some typical characterizations are performed (as shown in Figure 2) to confirm their precisely defined macromolecular structure and purity. For example, the representative peaks from VPOSS (δ 6.5−5.5 ppm) and BPOSS (δ 1.8−2.0 ppm) increase correspondingly in the 1H NMR spectra (Figure 2a) as we attach the additional NPs. The GPC curves (Figure 2b) also show consistently shifts toward low retention time when the number of POSS cages increases. Furthermore, the matrixassisted laser desorption/ionization (MALDI-TOF) mass spectra display single component m/z peaks that match the calculated molecular mass of the giant molecules (Figure 2c). Moreover, to introduce amphiphilicity for further self-assembly studies, thiol−ene click reactions are used to convert VPOSSs into DPOSSs.33,34 After the reaction, the vinyl groups (δ 6.5− 5.5 ppm) are completely consumed and the signals from the 636

DOI: 10.1021/acsmacrolett.8b00275 ACS Macro Lett. 2018, 7, 635−640

Letter

ACS Macro Letters

different from those in our previous studies, where esterification reactions were repeatedly performed to connect the POSS cages, and only up to six POSSs could be achieved in a single giant molecule through six reaction cycles with tedious purifications. In this study, the employment of efficient “click” reactions make the procedure easier. The introduction of two growing sites reduces by half the reaction cycles needed for a given number of POSS cages in the chain and also indicates more kinds of sequence isomers existed in low BPOSS/DPOSS ratio, offering new opportunities for us to study the sequence effect of those giant molecules in the nonspherical phase regions. As previously studied, hydrophobic BPOSS cages and hydrophilic DPOSS cages in one giant molecule also tend to phase separate with each other due to the strong immiscibility.26,35 Those giant molecules could also self-assemble into various supramolecular structures after 1 h thermal annealing at 150 °C, similar to the condition we used previously.26 The annealing temperature is chosen to be above the melting point of the BPOSS (Figure S8) so that they could have large mobility to form the thermodynamic stable structures. When the BPOSS number increases, the supramolecular structures of BnD2Bn (n = 1, 2, 3) evolve from lamellae (LAM) to double gyroids (DG), and further to hexagonal packed cylinder (HEX). In detail, for BD2B, the volume fraction of BPOSS (vfB) is about 0.50 and it forms the LAM phase with a d-spacing of 7.82 nm, which is confirmed by a small-angle X-ray scattering (SAXS) pattern (Figure 3a) and a bright-field (BF) transmission electron microscopy (TEM) image (Figure 3d). When increasing the number of BPOSS to four, B2D2B2 with vfB ≈ 0.66 shows a characteristic q ratio of √6:√8 in a SAXS pattern (Figure 3b, d1 = 7.44 nm), indicating a DG phase (a BF TEM image is shown in Figure 3e). Further increasing the number of BPOSS to vfB ≈ 0.75, B3D2B3 self-organizes into a HEX phase, evident by q ratio of 1:√3:√4:√7 in a SAXS pattern (Figure 3c, d1 = 8.05 nm) and hexagonal packing morphology in a BF TEM image (Figure 3f). This phase evolution is similar to traditional block copolymers, by increasing the volume fractions of hydrophobic part, which leads to an increase of interfacial curvature, and resulting in different phase structures.26,36,37

Figure 2. (a) 1H NMR, (b) GPC, and (c) MALDI-TOF MS spectra of the chain-like giant molecules and some related intermediates.

DPOSS appear clearly (δ 3.8−3.6 ppm; Figure 2a). Other kinds of sequence isomers with distinct DPOSS locations could be achieved in a similar way by introducing the VPOSSs at different reaction cycles. In this way, we could prepare a series of giant molecules with precise amphiphilic blocks modularly. Note that the chains grow from both sides at the same time due to the indistinguishable two click sites, leading to more POSS cages included in those resulting giant molecules. The summary of all samples is outlined in Figure 1b. This series of samples is

Figure 3. (a−c) SAXS patterns and (d−f) TEM images of BD2B, B2D2B2, and B3D2B3, respectively. The blue and red balls represent BPOSS and DPOSS, respectively, in the inset cartoon. 637

DOI: 10.1021/acsmacrolett.8b00275 ACS Macro Lett. 2018, 7, 635−640

Letter

ACS Macro Letters

Figure 4. SAXS patterns of the three sequence isomers of B3D2B3: (a) B2DB2DB2, (b) BDB4DB, and (c) DB6D; (d) Illustration of HEX structure formation; (e−h) Cartoon illustration of the NP sequence induced geometrical difference of those four isomeric giant molecules.

Note that DB2D is a “sequence isomer” to BD2B with vfB ≈ 0.50. It also shows a LAM phase with a d-spacing of 7.88 nm from the SAXS result (Figure S4), which is almost identical to BD2B (7.82 nm). At vfB ≈ 0.66, B2D2B2 has another two kinds of “sequence isomers” (BDB2DB and DB4D). These isomers also form DG structures as clearly indicated by the characteristic q ratio of √6:√8 in the SAXS results (Figures S5 and S6). However, the dimensions of their formed phase structures are different (d1 = 6.65 and 7.37 nm). A similar result is also observed for B3D2B3 with vfB ≈ 0.75, who has another three “sequence isomers” (B2DB2DB2, BDB4DB, and DB6D) that all form HEX phases with different dimensions (Figure 4a−c, d1 = 6.66, 6.95, and 8.09 nm). These results reveal that, in this series of giant molecules, the supramolecular structures are determined by the chemical compositions. However, within their sequence isomers, the structure dimensions are different although their phase structures are identical. It would be interesting to understand how the sequence change affects the phase structure formation of giant molecules. During the self-assembly process studied here, two DPOSS cages with strong collective hydrogen bonding are able to aggregate with each other into one phase, while the hydrophobic BPOSSs will stay in another phase. When the symmetric sequence changes, so does the arrangement of molecular conformation. This may significantly alter the interfaces area between these two phases. In the condensed state, the densities of both phases are constant. For each giant molecule, when its cross-section area between hydrophilic and hydrophobic parts increases, the phase thickness (d-spacing) should be decreased. Their scaling relationships are as follows. In the LAM structure, the cross-section area can be calculated by eq 1 as

A0 =

2MB dρB NAvfB



1 d

where d is the layer spacing (d = d1 in Figure 3a that is corresponding to the d-spacing of the first diffraction peak in SAXS), MB is the molecular mass of the BPOSS part, ρB is the density of BPOSS, and NA is Avogadro constant. In the DG phase, the cross-section area can be calculated by eq 2 as38,39 A0 =

xMB aρB NAvfB



1 a

(2)

where a is the cubic unit cell length (a = √6d1 in Figure 3b) and x is the specific surface area to bulk volume ratio of the DG structure as a function of its volume fraction. In HEX, the cross-section area can be calculated by eq 3 as A0 =

4MB 1 − vfB RvfBρB NA



1 R

(3)

where R is the intercolumn distance of hexagonal packed cylinders (R = 2d1/√3 in Figure 3c). It is evident that the dimensions of the formed supramolecular structures (d, a, or R for LAM, DG, or HEX, respectively) are all inverse proportional to the cross-section area of the two phases (A0). For BD2B and DB2D, their crosssection areas are almost identical since each DPOSS cage is only adjacent to one BPOSS cage. Therefore, the dimensional differences in their LAM structures are not noticeable. However, the situations become different in the longer chainlike giant molecules with symmetric sequences, such as in the cases of B2D2B2, BDB2DB, and DB4D. Note that there are two DPOSS cages in each of the giant molecule. When they are located in the center (i.e., B2D2B2) or at the end (i.e., DB4D), each DPOSS cage is adjacent to one BPOSS in the sequence, while for BDB2DB, each DPOSS is adjacent to two BPOSS cages in the sequence. Therefore, the B/D interfacial area (A0) of BDB2DB must be larger, and thus, due to the inverse proportional relationship, it shows smaller dimensions (d1 =

(1) 638

DOI: 10.1021/acsmacrolett.8b00275 ACS Macro Lett. 2018, 7, 635−640

ACS Macro Letters



6.65 nm) than the former two (d1= 7.44 and 7.37 nm, respectively) when they form the same type of supramolecular structures. In the cases of B3D2B3 and DB6D molecules, the number of adjacent BPOSS cages for each DPOSS is one, while that is two for B2DB2DB2 and BDB4DB. Therefore, both of the B3D2B3 and DB6D molecules possess qualitatively larger A0 values and, thus, smaller d-spacings (d1= 6.66 and 6.95 nm, respectively) compared with the former two (d1 = 8.05 and 8.09 nm, respectively). A more visual illustration is given in Figure 4d−h for the samples that form the HEX phase. For constructing this phase, we can view that each giant molecule exhibits a fan-like topology with a fan-angle (α). The fan-like shapes assemble into 360° to form disks, which could further stack into columns (Figure 4d). Such a hierarchical assembly process of precisely defined ABn-type molecules like amphiphilic giant molecules and janus dendrimers has been previously demonstrated by our group and many others.21,37,40,41 At a fixed composition, larger interfacial area (A0) also means a larger fan angle, and less molecules are needed to reach 360°. From the measured density of about 1.25−1.26 g/cm3 for these molecules, we could also calculate fan angles (see SI for detail), which are 59°, 85°, 80°, and 59° for B3D2B3, B2DB2DB2, BDB4DB, and DB6D, respectively. Now we could understand how nanoparticle sequence in giant molecules influence the supramolecular structure formation in general based on the obtained results and pervious data. That might be relied on the phase regions of different supramolecular structures formed by each giant molecule with defined volume fraction (vfB), which is changed discretely due to the large size of the POSS-containing monomer in our system. For those unconventional phases with very narrow existing windows in the phase diagrams of the linear macromolecules, like A15, sigma and quasicrystal phases, the sequence change of giant molecule might be able to finally induce their phase structure change across the boundary line. However, for the phases with wide existing windows in the phase diagrams, like in the LAM and HEX phases, it is highly possible that the sequence change of giant molecular could only affect the domain sizes of the formed phase structures but not alter the type of structures. Note that the DG structures usually also perform relatively narrow existing window in the phase diagram, so either phase transformation or dimension change variation could possibly be observed by tuning the sequences. The importance of sequence should thus not be overlooked. In summary, using a series of amphiphilic chain-like giant molecules based on symmetrically arranged POSS cages building blocks, we have found that their self-assembled supramolecular structures could be determined by the number of incorporated POSS cages (composition) and their connecting sequences. Particularly in those nonspherical phases, the sequence of the symmetrically connected POSS cages could affect the cross-section areas of the hydrophobic/ hydrophilic POSS domains, finally resulting in adjustment of domain sizes of the supramolecular structures. Those results reveal that the supramolecular structure of synthetic macromolecules could be tuned in multiple levels by precise control over the chemical composition and monomer sequence.

Letter

ASSOCIATED CONTENT

S Supporting Information *

The Supporting Information is available free of charge on the ACS Publications website at DOI: 10.1021/acsmacrolett.8b00275. Detailed characterization methods, calculations, synthetic schemes and procedures, additional synthetic and X-ray characterizations (PDF).



AUTHOR INFORMATION

Corresponding Authors

*E-mail [email protected];. *E-mail [email protected]. ORCID

Wei Zhang: 0000-0002-9321-6411 Yuchu Liu: 0000-0001-9780-8724 Jialin Mao: 0000-0002-5011-4378 Chrys Wesdemiotis: 0000-0002-7916-4782 Yiwen Li: 0000-0002-6874-0350 Stephen Z. D. Cheng: 0000-0003-1448-0546 Notes

The authors declare no competing financial interest.



ACKNOWLEDGMENTS This work was supported by NSF (DMR-1408872 to S.Z.D.C, and CHE-1308307 to C.W.) and NSFC (51603133 to Y. L). This research used resources of the Advanced Photon Source, a U.S. Department of Energy (DOE) Office of Science User Facility operated for the DOE Office of Science by Argonne National Laboratory under Contract No. DE-AC0206CH11357.



REFERENCES

(1) Matyjaszewski, K. Architecturally Complex Polymers with Controlled Heterogeneity. Science 2011, 333, 1104−1105. (2) Lutz, J.-F.; Ouchi, M.; Liu, D. R.; Sawamoto, M. SequenceControlled Polymers. Science 2013, 341, 1238149. (3) Matyjaszewski, K.; Xia, J. Atom Transfer Radical Polymerization. Chem. Rev. 2001, 101, 2921−2990. (4) Matyjaszewski, K. Atom Transfer Radical Polymerization (ATRP): Current Status and Future Perspectives. Macromolecules 2012, 45, 4015−4039. (5) Bielawski, C. W.; Grubbs, R. H. Living ring-opening metathesis polymerization. Prog. Polym. Sci. 2007, 32, 1−29. (6) Hawker, C. J.; Bosman, A. W.; Harth, E. New Polymer Synthesis by Nitroxide Mediated Living Radical Polymerizations. Chem. Rev. 2001, 101, 3661−3688. (7) Boyer, C.; Bulmus, V.; Davis, T. P.; Ladmiral, V.; Liu, J.; Perrier, S. Bioapplications of RAFT Polymerization. Chem. Rev. 2009, 109, 5402−5436. (8) Widin, J. M.; Schmitt, A. K.; Schmitt, A. L.; Im, K.; Mahanthappa, M. K. Unexpected Consequences of Block Polydispersity on the SelfAssembly of ABA Triblock Copolymers. J. Am. Chem. Soc. 2012, 134, 3834−3844. (9) Lutz, J.-F.; Lehn, J.-M.; Meijer, E. W.; Matyjaszewski, K. From precision polymers to complex materials and systems. Nat. Rev. Mater. 2016, 1, 16024. (10) Hibi, Y.; Ouchi, M.; Sawamoto, M. A strategy for sequence control in vinyl polymers via iterative controlled radical cyclization. Nat. Commun. 2016, 7, 11064. (11) Engelis, N. G.; Anastasaki, A.; Nurumbetov, G.; Truong, N. P.; Nikolaou, V.; Shegiwal, A.; Whittaker, M. R.; Davis, T. P.; Haddleton, D. M. Sequence-controlled methacrylic multiblock copolymers via 639

DOI: 10.1021/acsmacrolett.8b00275 ACS Macro Lett. 2018, 7, 635−640

Letter

ACS Macro Letters sulfur-free RAFT emulsion polymerization. Nat. Chem. 2017, 9, 171− 178. (12) Gutekunst, W. R.; Hawker, C. J. A General Approach to Sequence-Controlled Polymers Using Macrocyclic Ring Opening Metathesis Polymerization. J. Am. Chem. Soc. 2015, 137, 8038−8041. (13) Anastasaki, A.; Oschmann, B.; Willenbacher, J.; Melker, A.; Van Son, M. H. C.; Truong, N. P.; Schulze, M. W.; Discekici, E. H.; McGrath, A. J.; Davis, T. P.; Bates, C. M.; Hawker, C. J. One-Pot Synthesis of ABCDE Multiblock Copolymers with Hydrophobic, Hydrophilic, and Semi-Fluorinated Segments. Angew. Chem., Int. Ed. 2017, 56, 14483−14487. (14) Xi, W.; Pattanayak, S.; Wang, C.; Fairbanks, B.; Gong, T.; Wagner, J.; Kloxin, C. J.; Bowman, C. N. Clickable Nucleic Acids: Sequence-Controlled Periodic Copolymer/Oligomer Synthesis by Orthogonal Thiol-X Reactions. Angew. Chem., Int. Ed. 2015, 54, 14462−14467. (15) Roy, R. K.; Meszynska, A.; Laure, C.; Charles, L.; Verchin, C.; Lutz, J.-F. Design and synthesis of digitally encoded polymers that can be decoded and erased. Nat. Commun. 2015, 6, 7237. (16) Solleder, S. C.; Meier, M. A. R. Sequence Control in Polymer Chemistry through the Passerini Three-Component Reaction. Angew. Chem., Int. Ed. 2014, 53, 711−714. (17) Porel, M.; Alabi, C. A. Sequence-Defined Polymers via Orthogonal Allyl Acrylamide Building Blocks. J. Am. Chem. Soc. 2014, 136, 13162−13165. (18) Barnes, J. C.; Ehrlich, D. J. C.; Gao, A. X.; Leibfarth, F. A.; Jiang, Y.; Zhou, E.; Jamison, T. F.; Johnson, J. A. Iterative exponential growth of stereo- and sequence-controlled polymers. Nat. Chem. 2015, 7, 810−815. (19) Lutz, J.-F. Coding Macromolecules: Inputting Information in Polymers Using Monomer-Based Alphabets. Macromolecules 2015, 48, 4759−4767. (20) Ouahabi, A. A.; Kotera, M.; Charles, L.; Lutz, J.-F. Synthesis of Monodisperse Sequence-Coded Polymers with Chain Lengths above DP100. ACS Macro Lett. 2015, 4, 1077−1080. (21) Zhang, W.; Huang, M.; Su, H.; Zhang, S.; Yue, K.; Dong, X.-H.; Li, X.; Liu, H.; Zhang, S.; Wesdemiotis, C.; Lotz, B.; Zhang, W.-B.; Li, Y.; Cheng, S. Z. D. Toward Controlled Hierarchical Heterogeneities in Giant Molecules with Precisely Arranged Nano Building Blocks. ACS Cent. Sci. 2016, 2, 48−54. (22) Ji, Y.; Zhang, L.; Gu, X.; Zhang, W.; Zhou, N.; Zhang, Z.; Zhu, X. Sequence-Controlled Polymers with Furan-Protected Maleimide as a Latent Monomer. Angew. Chem., Int. Ed. 2017, 56, 2328−2333. (23) Wu, Y.-H.; Zhang, J.; Du, F.-S.; Li, Z.-C. Dual Sequence Control of Uniform Macromolecules through Consecutive Single Addition by Selective Passerini Reaction. ACS Macro Lett. 2017, 6, 1398−1403. (24) Laure, C.; Karamessini, D.; Milenkovic, O.; Charles, L.; Lutz, J.F. Coding in 2D: Using Intentional Dispersity to Enhance the Information Capacity of Sequence-Coded Polymer Barcodes. Angew. Chem., Int. Ed. 2016, 55, 10722−10725. (25) Golder, M. R.; Jiang, Y.; Teichen, P. E.; Nguyen, H. V. T.; Wang, W.; Milos, N.; Freedman, S. A.; Willard, A. P.; Johnson, J. A. Stereochemical Sequence Dictates Unimolecular Diblock Copolymer Assembly. J. Am. Chem. Soc. 2018, 140, 1596−1599. (26) Zhang, W.; Lu, X.; Mao, J.; Hsu, C.-H.; Mu, G.; Huang, M.; Guo, Q.; Liu, H.; Wesdemiotis, C.; Li, T.; Zhang, W.-B.; Li, Y.; Cheng, S. Z. D. Sequence-Mandated, Distinct Assembly of Giant Molecules. Angew. Chem., Int. Ed. 2017, 56, 15014−15019. (27) Zhang, W.-B.; Yu, X.; Wang, C.-L.; Sun, H.-J.; Hsieh, I. F.; Li, Y.; Dong, X.-H.; Yue, K.; Van Horn, R.; Cheng, S. Z. D. Molecular Nanoparticles Are Unique Elements for Macromolecular Science: From “Nanoatoms” to Giant Molecules. Macromolecules 2014, 47, 1221−1239. (28) Yue, K.; Huang, M.; Marson, R. L.; He, J.; Huang, J.; Zhou, Z.; Wang, J.; Liu, C.; Yan, X.; Wu, K.; Guo, Z.; Liu, H.; Zhang, W.; Ni, P.; Wesdemiotis, C.; Zhang, W.-B.; Glotzer, S. C.; Cheng, S. Z. D. Geometry induced sequence of nanoscale Frank−Kasper and quasicrystal mesophases in giant surfactants. Proc. Natl. Acad. Sci. U. S. A. 2016, 113, 14195−14200.

(29) Liu, H.; Luo, J.; Shan, W.; Guo, D.; Wang, J.; Hsu, C.-H.; Huang, M.; Zhang, W.; Lotz, B.; Zhang, W.-B.; Liu, T.; Yue, K.; Cheng, S. Z. D. Manipulation of Self-Assembled Nanostructure Dimensions in Molecular Janus Particles. ACS Nano 2016, 10, 6585− 6596. (30) Wu, K.; Huang, M.; Yue, K.; Liu, C.; Lin, Z.; Liu, H.; Zhang, W.; Hsu, C.-H.; Shi, A.-C.; Zhang, W.-B.; Cheng, S. Z. D. Asymmetric Giant “Bolaform-like” Surfactants: Precise Synthesis, Phase Diagram, and Crystallization-Induced Phase Separation. Macromolecules 2014, 47, 4622−4633. (31) Kuo, S.-W. Building Blocks Precisely from Polyhedral Oligomeric Silsesquioxane Nanoparticles. ACS Cent. Sci. 2016, 2, 62−64. (32) Chu, Y.; Zhang, W.; Lu, X.; Mu, G.; Zhang, B.; Li, Y.; Cheng, S. Z. D.; Liu, T. Rational controlled morphological transitions in the selfassembled multi-headed giant surfactants in solution. Chem. Commun. 2016, 52, 8687−8690. (33) Zhang, W.; Chu, Y.; Mu, G.; Eghtesadi, S. A.; Liu, Y.; Zhou, Z.; Lu, X.; Kashfipour, M. A.; Lillard, R. S.; Yue, K.; Liu, T.; Cheng, S. Z. D. Rationally Controlling the Self-Assembly Behavior of Triarmed POSS−Organic Hybrid Macromolecules: From Giant Surfactants to Macroions. Macromolecules 2017, 50, 5042−5050. (34) Li, Y.; Dong, X.-H.; Zou, Y.; Wang, Z.; Yue, K.; Huang, M.; Liu, H.; Feng, X.; Lin, Z.; Zhang, W.; Zhang, W.-B.; Cheng, S. Z. D. Polyhedral oligomeric silsesquioxane meets “click” chemistry: Rational design and facile preparation of functional hybrid materials. Polymer 2017, 125, 303−329. (35) Huang, M.; Hsu, C.-H.; Wang, J.; Mei, S.; Dong, X.; Li, Y.; Li, M.; Liu, H.; Zhang, W.; Aida, T.; Zhang, W.-B.; Yue, K.; Cheng, S. Z. D. Selective assemblies of giant tetrahedra via precisely controlled positional interactions. Science 2015, 348, 424−428. (36) Kim, H.-C.; Park, S.-M.; Hinsberg, W. D. Block Copolymer Based Nanostructures: Materials, Processes, and Applications to Electronics. Chem. Rev. 2010, 110, 146−177. (37) Feng, X.; Zhang, R.; Li, Y.; Hong, Y.-l.; Guo, D.; Lang, K.; Wu, K.-Y.; Huang, M.; Mao, J.; Wesdemiotis, C.; Nishiyama, Y.; Zhang, W.; Zhang, W.; Miyoshi, T.; Li, T.; Cheng, S. Z. D. Hierarchical SelfOrganization of ABn Dendron-like Molecules into a Supramolecular Lattice Sequence. ACS Cent. Sci. 2017, 3, 860−867. (38) Scherer, M. R. J. Double-Gyroid-Structured functional materials: synthesis and applications; Springer International Publishing, Switzerland, 2013. (39) Wang, X.-M.; Shao, Y.; Xu, J.; Jin, X.; Shen, R.-H.; Jin, P.-F.; Shen, D.-W.; Wang, J.; Li, W.; He, J.; Ni, P.; Zhang, W.-B. Precision Synthesis and Distinct Assembly of Double-Chain Giant Surfactant Regioisomers. Macromolecules 2017, 50, 3943−3953. (40) Percec, V.; Cho, W.-D.; Ungar, G.; Yeardley, D. J. P. Synthesis and Structural Analysis of Two Constitutional Isomeric Libraries of AB2-Based Monodendrons and Supramolecular Dendrimers. J. Am. Chem. Soc. 2001, 123, 1302−1315. (41) Percec, V.; Won, B. C.; Peterca, M.; Heiney, P. A. Expanding the Structural Diversity of Self-Assembling Dendrons and Supramolecular Dendrimers via Complex Building Blocks. J. Am. Chem. Soc. 2007, 129, 11265−11278.

640

DOI: 10.1021/acsmacrolett.8b00275 ACS Macro Lett. 2018, 7, 635−640