Skip to main content
NIHPA Author Manuscripts logoLink to NIHPA Author Manuscripts
. Author manuscript; available in PMC: 2021 Jul 29.
Published in final edited form as: ACS Macro Lett. 2021 Jan 20;10(2):243–257. doi: 10.1021/acsmacrolett.0c00855

Geared Toward Applications: A Perspective on Functional Sequence-Controlled Polymers

Cangjie Yang 1, Kevin B Wu 1, Yu Deng 1, Jingsong Yuan 1, Jia Niu 1,*
PMCID: PMC8320758  NIHMSID: NIHMS1672536  PMID: 34336395

Abstract

Sequence-controlled polymers are an emerging class of synthetic polymers with a regulated sequence of monomers. In the past decade, tremendous progress has been made in the synthesis of polymers with the sophisticated sequence control approaching the level manifested in biopolymers. In contrast, the exploration of novel functions that can be achieved by controlling synthetic polymer sequences represents an emerging focus in polymer science. This Viewpoint will survey recent advances in the functional applications of sequence-controlled polymers and provide a perspective on the challenges and outlook for pursuing future applications of this fascinating class of macromolecules.

Graphical Abstract

graphic file with name nihms-1672536-f0001.jpg

1. Introduction

The ability to synthesize and decode sequence-defined biopolymers such as DNA, RNA, and proteins is the molecular essence of life that enables the evolution of highly sophisticated structures and functions. For example, using transfer RNA (tRNA) as cofactors, the ribosome can decode the genetic information stored in messenger RNA (mRNA) and synthesize polypeptides from amino acids using mRNAs as templates in remarkable efficiency and fidelity. Programmed by their primary sequences, these polypeptides fold into secondary, tertiary, and quaternary structures and can perform a wide array of functions ranging from molecular recognition to catalysis. Aspiring to a myriad of complex functions possessed by biopolymers, generations of polymer chemists took on the formidable challenge to emulate biopolymers and achieve the precise sequence control in synthetic polymers. In his Nobel Lecture, Hermann Staudinger suggested that the studies on synthetic polymers with enhanced structure and sequence control could help to give insight into the structure-function relationship of biopolymers.1 Several landmark discoveries in polymer science, such as the solid-phase synthesis2 and living polymerization,3 were associated with chemists’ efforts to precisely control the sequences and chain structures of macromolecules. In this Viewpoint, we refer to the polymers in which the arrangement of the monomeric building blocks follows a defined order along the polymer chain as “sequence-controlled polymers”.4 It is noteworthy that such a definition includes not only monodispersed and polydispersed synthetic polymers with regulated monomer sequences, but also analogs of biopolymers that are monodispersed and consist of extensive non-natural structures and functionalities, such as xenobiotic nucleic acids (XNAs, nucleic acids consisting of non-natural sugar backbones or modified nucleobases56) and peptidomimetics. As a result, the sequence-controlled polymers discussed in this Viewpoint are divided into two main categories: sequence-defined polymers (Ð = 1) and polydispersed sequence-controlled polymers (Ð > 1) (Figure 1). We believe that this broad definition is consistent with the spirit of Staudinger’s macromolecular hypothesis,78 as well as the seminal review by Lutz, Ouchi, Liu, and Sawamoto.9 Moreover, it reflects the highly interdisciplinary nature of this field of research and its promising applications in chemistry, biology, and materials science.

Figure 1.

Figure 1.

Sequence-controlled polymers discussed in this Viewpoint.

Over the past decade, the development of sequence-controlled polymers was proclaimed as the next “Holy Grail” in polymer science, a topic of paramount importance.10 New chemical and biological polymerization methods, including iterative coupling, living/controlled chain-growth polymerization, step-growth polymerization, templated polymerization, and enzymatic polymerization, allowed for the preparation of sequence-controlled polymers with diverse sequences and chain structures. Because these advances in the synthetic strategies of sequence-controlled polymers have been comprehensively discussed in many excellent recent research and review articles,4, 922 this Viewpoint will avoid repeating the discussion on synthesis. Rather, the discussion will be centered on two fundamental questions about function: (1) what novel properties do these sequence-controlled polymers have compared to classic synthetic polymers, and (2) how can these properties address some of the pressing challenges in our society? Hence, this Viewpoint aims to provide a perspective on functional sequence-controlled polymers by highlighting recent examples of the emerging applications of sequence-controlled polymers, such as data storage, biomaterials and biocatalysis, self-assembled nanostructures, and functional soft materials.

2. Functional Applications of Sequence-Controlled Polymers

2.1. Data storage

The amount of data generated and transmitted has experienced an explosive increase since the beginning of the digital age, and it is anticipated that this trend will continue into the foreseeable future. As a result, there is a critical need for new methods and materials that enable efficient and secure data storage. Evidently, DNA as a sequence-defined biopolymer is a promising solution for storing a large amount of information both in living organisms and in laboratories.23 The defined sequence of the four nucleotides A, G, C, and T gives DNA an information storage capability of 5.5 petabits/mm3, much higher than the traditional silicon-based media. The principle of storing information within the monomer sequence of polymers is not limited to DNA, and sequence-defined polymers should also enable the storage of large amounts of digital information at the molecular level. Compared to DNA, the chemical stability of synthetic polymers offers unique potential advantages for data storage. The ability to incorporate diverse functional groups gives sequence-defined polymers another edge on DNA in applications involving non-biological environments.1617, 2428 For example, Lutz et al. encoded the 0-bit and 1-bit in binary sequence-defined polyphosphates with degrees of polymerization up to 104.2930 By incorporating an alkoxyamine linkage into each phosphoramidite unit, the information-containing polyphosphates were amenable for tandem mass spectrometry (MS/MS) sequencing thanks to the easily cleavable alkoxyamine bonds.31 The authors further developed sequence-defined poly(alkoxyamine amides)32 (Figure 2a) and polyurethanes3334 (Figure 2b) that contain weak C–ON bonds and C–O carbamate bonds between monomer units, respectively, to allow for higher bond cleavage efficiency by MS/MS. Similarly, Gao and coworkers prepared positively charged sequence-defined polymers using alternating Menschutkin reactions and copper(I)-catalyzed alkyne-azide cycloaddition (CuAAC) reactions in the solution phase.35 The on-source cleavage of the quaternary ammonium group in matrix-assisted laser desorption/ionization time-of-flight mass spectrometry (MALDI-TOF MS) sequencing allowed for the generation of distinct fragments. Besides the on-source fragmentation within the mass spectrometer, controlled end-to-end chemical degradation has also been used to decode the sequence-defined polymers. For instance, Anslyn and coworkers synthesized sequence-encoded oligourethanes which can be sequenced without the need of MS/MS.36 These oligomers can undergo self-immolation to release 2-oxazolidinones via base-promoted intramolecular cyclization from the terminal alcohol. This process could be monitored by liquid chromatography-mass spectrometry (LC-MS) to provide the sequence information. While these works have demonstrated the potential of sequence-defined polymers in data storage applications, major limitations that require further improvements include: (1) high cost and laborious synthesis limit the amount of information that can be “written” into sequence-defined polymers; (2) high-throughput sequencing of sequence-defined polymers to “read” the stored information remains challenging; (3) unlike DNA, few methods have been developed to date for the amplification of sequence-defined polymers and the information stored in them.

Figure 2.

Figure 2.

Sequence-defined polymers for data storage. (a) Oligo(alkoxyamine amide)s consisting of a binary sequence. Reproduced with permission from Ref. 32. Copyright 2015, Springer Nature. (b) The synthesis and MS/MS sequencing of oligourethanes consisting of a binary sequence of 101010. Reproduced with permission from Ref. 34. Copyright 2016, Elsevier Inc.

Data security is another important aspect of data storage. To this end, erasable sequence-defined polymers were developed to undergo structural change in response to external stimuli such as light, temperature, and chemicals. For example, the poly(alkoxyamine amides)s reported by Lutz et al. are temperature-labile and able to degrade at elevated temperature (120 °C) in the presence of a large excess of 2,2,6,6-tetramethylpiperidine-1-oxyl (TEMPO), resulting in the erasing of the encoded information.32 They also decorated the aforementioned phosphoramidite monomers with light-labile o-nitrobenzyl/o-nitroveratryl ether and light-inert p-nitrobenzyl ether side chains and assembled them into photo-editable binary encoded polymers.37 These polymers can be read out using MS/MS but became unreadable after UV exposure. Synthetic sequence-defined polymers also offer the possibility to be used for anticounterfeiting and cryptography. For example, a small amount of binary encoded oligourethanes synthesized by Lutz et al. could be blended into industrial plastics such as polystyrene and poly(methacrylate).3334 The information stored in these oligomers can be extracted from the mixed matrix by selective dissolution in methanol or tetrahydrofuran followed by MS/MS decoding. To enhance the information-carrying capacity, the authors further developed a 2D coding approach by writing information in a set of binary coded oligo(alkoxyamine amide)s. The selective identification by LC-MS and MS/MS could be used to decode the information stored in these 2D barcodes.38

2.2. Bioactivities

2.2.1. Recognition of biomolecules

Glycan recognition by receptors plays important roles in a variety of biological activities such as cell-cell interaction, signal transduction, and host–pathogen recognition.39 To elucidate the relationship between the structure of glycan ligands and their receptor-binding properties, glycomimetic sequence-defined polymers have been developed as synthetic analogs of natural polysaccharides. For example, Hartmann et al. developed the sequence-defined glycopolymers carrying different sugar ligands including mannose (Man), galactose (Gal) and glucose (Glc) at precise positions on the peptide backbone (Figure 3a).4041 They found variations in both the number and sequence of the sugar ligands could alter the binding affinity. Specifically, the presence of non-binding galactose and glucose moieties could suppress receptor clustering and promote steric shielding of the heterotrivalent glycopolymer, and thus enhanced their affinity to Concanavalin A (Con A). Polymer-receptor binding has also been studied in the context of polydispersed sequence-controlled glycopolymers. For instance, Haddleton et al. prepared a library of multiblock glycopolymers with block sequences consisting of mannose, glucose, fucose side chains via atom transfer radical polymerization (ATRP) (Figure 3b).42 The dissociation constants (KD) and half maximal inhibitory concentration (IC50) of these glycopolymers to dendritic cell-specific intercellular adhesion molecule-3-grabbing non-integrin (DC-SIGN) exhibited a general correlation to the mannose content. Collectively, these examples demonstrated that sequence control in both sequence-defined polymers and polydispersed sequence-controlled polymers can significantly influence the complex, multivalent nature of the interactions with proteins.

Figure 3.

Figure 3.

Biomolecular recognition by sequence-controlled polymers. (a) Glycopolymers with heteromutivalent and homomultivalent structures consisting of mannose (Man), glucose (Glc) and galactose (Gal) side chains at defined positions for lectin binding. Reproduced with permission from Ref. 40. Copyright 2014, American Chemical Society. (b) Multiblock glycopolymer capable of inhibiting the DC-SIGN-gp120 recognition. Reproduced with permission from Ref. 42. Copyright 2013, WILEY-VCH. (c) SOMAmers bind their targets with exquisite shape complementarity and utilize hydrophobic modifications at the binding interface. Reproduced with permission from Ref. 47. Copyright 2014, Elsevier Inc. (d) Click-PD strategy for two-color screening of lectin-specific base-modified aptamers and FACS (fluorescence-activated cell sorting) plot with two-color screening for affinity and specificity in parallel. Reproduced with permission from Ref. 51. Copyright 2019, American Chemical Society.

XNA-based affinity reagents, often referred to as “XNA aptamers”, are a class of sequence-defined polymers extensively studied for binding to biomolecular targets. Two important advantages of XNA aptamers compared to their DNA/RNA counterparts are the ability to incorporate customized functional groups to optimize binding affinity according to its applications and the resistance to nuclease-mediated degradation in vivo. Romesberg et al. identified XNA aptamers with 2’-fluoro and 2’-OMe substitutions of the sugar backbone binding to human neutrophil elastase (HNE), a serine protease associated with inflammatory diseases.43 Building upon their work on polymerase engineering, Holliger et al. prepared a 2′-deoxy-2′-fluoroarabino nucleic acid (FANA) library and selected FANA aptamers targeting HIV-1 reverse transcriptase (RT) and integrase (IN), respectively, via systematic evolution of ligands by exponential enrichment (SELEX), with affinities comparable to DNA and RNA aptamers.4445 While modifications on the sugar backbone could improve XNA’s resistances to nucleases, side chain modifications on the nucleobases were explored to expand the scope of functionalities in aptamers. Eaton et al. developed a 5-modified dUTP derivatives that incorporate a series of hydrophobic groups through an amide linkage and incorporate these 5-modified dUTP derivatives in an XNA library for in vitro selection.46 Comparably, the incorporation of hydrophobic/aromatic functional side chains at the 5-position of uracil generated slow dissociating aptamers to a wider range of protein targets and enhanced the success rate of SELEX. These so-called SOMAmers (Slow Off-rate Modified Aptamers) utilize their hydrophobic modifications at the binding interface to complement the target protein surfaces that are significantly more hydrophobic, increasing their binding potential to a broader scope of target proteins (Figure 3c).47 Using the T4 DNA ligase-mediated DNA-templated polymerization strategy, Liu et al. and Hili et al. independently generated highly functionalized nucleic acid polymer (HFNAP) libraries and identified HFNAP aptamers targeting therapeutically relevant proteins such as PCSK9, interleukin-6, and thrombin via SELEX. These HFNAP aptamers feature low nanomolar affinity and high selectivity, highlighting the importance of expanding the chemical space of side chain modifications in XNA aptamers.4850 Soh, Niu, and coworkers combined a CuAAC click chemistry-based DNA modification strategy with particle display (PD) screening platform to select carbohydrate-modified aptamers for the recognition of carbohydrate-binding proteins (Figure 3d).51 Specifically, emulsion polymerase chain reaction (PCR) was performed to produce monoclonal aptamer particles in which the thymine base was replaced by an alkyne-modified uridine. Azide-modified mannose was subsequently conjugated to the aptamer particles via CuAAC. The carbohydrate-modified aptamer particles were screened for binding to fluorescently labeled lectin targets in a two-color fluorescence-activated cell sorting (FACS). The selected aptamers exhibited low nanomolar affinity and exceptional specificity to ConA, with minimal binding to other structurally similar lectins. In another notable example, Krauss et al. developed DNA-supported clusters of carbohydrates through selection with modified aptamers (SELMA) to mimic the epitope of 2G12, an antibody that protects against HIV infection.5255 This technology involves using DNA as cluster scaffolding materials and sequence-specifically introducing glycan modifications such as dense high mannose glycans (Man4) via CuAAC coupling. Glycoclusters discovered in SELMA displayed low nanomolar affinity to 2G12. These glycoclusters have shown promising implications for the design of future drugs and vaccines.

Sequence-defined XNA oligomers have also been applied to binding to DNA/RNA in vivo and inhibiting gene expression. Known as antisense oligonucleotides, they function through forming sequence-specific complexes with the target DNA or RNA and disrupting their replication or transcription. A variety of XNA backbones have been studied for antisense applications, including locked nucleic acid, peptide nucleic acid, morpholino nucleic acid, phosphorothioate nucleic acid, etc.5658 Compared to DNA- and RNA-based reagents, XNA-based antisense oligonucleotides possess improved stability, pharmacokinetics, and antisense activity in vivo. To date, several XNA-based antisense oligonucleotide drugs, such as Vitravene (fomivirsen) and Kynamro (mipomersen sodium), Inotersen (Tegsedi), Eteplirsen (Exondys 51), golodirsen (Vyondys 53) have already received FDA approval, with several others undergoing clinical trials.5960

Peptidomimetics such as peptoids, β-peptides, and γ-peptides are another class of sequence-defined biomimetic polymers used for biomolecular recognition. Compared to peptides, advantages of peptidomimetics include high proteolytic stability, low mammalian cytotoxicity, and the high flexibility of incorporating functional groups. Zuckermann and coworkers first synthesized peptoid libraries and screened them for their binding abilities to protein targets.61 Since then, a plethora of peptoid ligands to therapeutically relevant proteins has been discovered. For example, Kodadek et al. identified a peptoid oligomer that can bind to biomarkers related to Parkinson’s disease,62 HIV,63 Alzheimer’s disease,64 and other diseases.65 In addition, Seo et al. developed functional glycopeptoids through N-alkylaminooxy glycosylation. Serving as glycopeptide mimics or carbohydrate-presenting materials, these glycopeptoids created opportunities for revealing new insights into carbohydrate-protein interactions.66 Schepartz et al. systematically studied β-peptides as protein inhibitors. They reported a β3-peptide 14-helix structure that can inhibit the human oncogene product double minute 2 (hDM2), an important cancer therapy target, with nanomolar affinity.67 Later on, they developed β-peptides that contain diether and hydrocarbon bridges to improve cell permeability and circumvent the need for an Arg8 sequence for cellular delivery.68 Besides inhibiting the p53-hDM2 interactions, β3-peptide has also been developed to serve as HIV fusion inhibitors. They further developed a β3-decapeptide with a non-natural trifluoromethylbenzene side chain that inhibited the HIV gp41-mediated fusion with a CC50/EC50 ratio of 8, a major improvement compared to previous works.6970 Cai et al. developed an impressive class of sulfono- γ -AApeptides that can reproducibly fold into well-defined helical structures.7173 They demonstrated that these sulfono- γ -AApeptide helices were excellent mimicries of α-helices in several therapeutically relevant protein-protein interactions and could serve as inhibitors or agonists to these interactions, such as BCL9-β-catenin, p53-MDM2/MDMX, and glucagon peptide-GLP-1R.7477

2.2.2. Antimicrobial and antifouling applications

Antimicrobial peptides (AMPs) have attracted significant research and development activities given their wide-spectrum efficacy against bacterial pathogens and the difficulty for pathogens to develop resistance. However, limitations of AMPs such as poor scalability, high cost, and protease degradability have impeded their applications. To address these issues, a wide range of synthetic mimics of AMPs composed of cationic sequence-controlled polymers have been developed.7882 These antimicrobial polymers often carry cationic and hydrophobic functional groups. Their antimicrobial activities arise from the electrostatic interactions between the cationic side chains and the negatively charged phospholipids in the bacterial cell membranes, and the membrane disruption by the hydrophobic functionalities of the bound polymers.8384 As a result, the antimicrobial activities of these polymers are strongly influenced by the sequence, charge, hydrophilic/hydrophobic moiety ratio, molecular weight, and self-assembled morphologies. Baran and coworkers prepared 27 peptoids with antimicrobial structures and found a number of these peptoids had potent antimicrobial activities against a wide range of Gram-positive and Gram-negative strains.85 Beyond peptoids, other synthetic sequence-defined polymers have also been discovered to possess antimicrobial activities. For instance, Alabi and coworkers synthesized a unique class of sequence-defined oligothiotheramides from allyl acrylamide monomers using liquid-phase iterative synthesis and fluorous tag-facilitated purification.7880 The precise sequence control in these oligothiotheramides enabled the precise tuning of the hydrophobicity and the antimicrobial activity. Based on the mechanism of action of the antimicrobial polymers discussed above, having a defined sequence and monodispersity may not be a prerequisite for antimicrobial activities. Indeed, Boyer et al. have shown that polydispersed multiblock copolymers could also be mimics of antimicrobial peptides. They prepared a library of 32 multiblock copolymers consisting of cationic, hydrophilic, and hydrophobic blocks and screened the library against a wide range of bacterial pathogens.81 The results confirmed that by tuning the distribution of monomers within blocks a balance between the desired antimicrobial activity and undesired hemolytic activity could be achieved.

Biofouling is a process in which living organisms and organic matter produced by the deposit on surfaces.86 Biofouling is highly undesirable in many cases such as adhesion of marine organisms on the ship hull and biofilm formation on the surface of biomedical devices. Antifouling sequence-controlled polymers have been developed as coating materials to render the modified surfaces resistant to fouling by proteins and bacteria. Segalman et al. conjugated a series of sequence-defined polypeptoids composed of two monomers, a hydrophilic N-(2-methoxyethyl)glycine and a hydrophobic N-(heptafluorobutyl)glycine, to the surface to prevent the biofouling by the algae Ulva linza.87 They found that the position of fluorinated monomer in the polypeptoid sequence imposed a major influence on the antifouling property of the coating against Ulva linza. Messersmith et al. prepared a library of zwitterionic peptoid brushes to evaluate their antifouling properties.88 Their results showed that at a sufficiently low ionic strength, the polarity of the charged residues near the termini of the polypeptoid chains determined the electrostatic adsorption of the proteins to the surface.

2.2.3. Drug delivery

Biocompatible polymers are well known as essential carriers to deliver drugs or genes for therapeutical applications. One of the advantages of using polymeric carriers is that the intracellular trafficking of molecular cargos could be regulated by macromolecular architectures including composition, topology, and self-assembled morphologies. Tew et al. prepared three synthetic polymers to mimic the protein transduction domains (PTDs) via ring-opening metathesis polymerization (ROMP).8990 These homopolymers, gradient copolymers, and block copolymers had the same length and overall composition of cationic (guanidine) and hydrophobic (phenyl) groups, but had different degrees of segregation of hydrophobic side chains. While all of them could be internalized in Jurkat-T cells and HEK293T cells, the gradient copolymer with an intermediate degree of functional group segregation had the optimal balance of internalization activity, solubility, and toxicity. Perrier et al. investigated the influence of monomer sequence on the cellular uptake of polymers, which are multiblock polymers consisting of guanidinium-functionalized acrylamide and two other less hydrophilic monomers. These studies demonstrate that the sequence of blocks has an influence on both the uptake efficiency and the mechanism of internalization.91 Wagner et al. synthesized a library of oligo(ethane amino) amides with precise modification patterns and topology to study the structure-activity relationship between the polymer sequence and nucleic acid delivery activity.9297 This work indicated that the optimization of the location of hydrophobic and crosslinkable functional groups can improve the siRNA polyplex stability and transfection efficiency. Two candidates of a 38-oligomer library have promising properties of siRNA-binding and gene-silencing activities and achieved excellent efficacy in inhibiting tumor growth in vivo when formulated with therapeutic EG5 siRNA. In another example, Zuckermann et al. developed “lipitoids” composed of a cationic peptoid oligomer and a hydrophobic lipid moiety for siRNA delivery.9899 These lipitoids can self-assemble into spherical nanoparticles with siRNA. It was found a particular repeating trimer sequence of cationic-hydrophobic-hydrophobic demonstrated high transfection efficiency and low cytotoxicity.100 However, no correlation was found between the transfection efficiency and the structures of the DNA-peptoids/lipitoids complexes, suggesting that further investigation is needed to probe the hidden structure-function relationships.101

2.2.4. Catalysis

Similar to DNA and RNA that can form ribozymes and DNAzymes, single-stranded XNAs can be selected to fold into particular motifs and serve as molecular catalysts. Furthermore, modifications to the sugar backbone and nucleobases provided the XNA catalysts (XNAzymes) superior stability and/or enhanced catalytic activities. Holliger et al. presented an impressive example of XNAzyme with RNA endonuclease and ligase activities.102 They subjected chimeric DNA–XNA libraries of four types of XNA (arabinonucleic acid (ANA), FANA, hexitol nucleic acid (HNA), cyclohexyl nucleic acid (CeNA)), all derived from DNAzyme motifs, into in vitro selections for RNA cleavage and ligation. While the activities varied, the selected XNAzymes all exhibited RNA cleavage/ligation activities. They also used the same strategy to identify the first successful XNA-cleaving FANAzyme with a rate constant of 0.02 min−1. Chaput et al. developed an RNA-cleaving FANAzyme with higher catalytic efficiency (kcat = 0.2 ± 0.01 min−1) and Michaelis-Menten kinetics (Figure 4a).103 The discoveries of these XNAzymes proved that catalytic activities can emerge from synthetic XNA polymers, a general principle that can also be potentially expanded to other types of sequence-defined synthetic polymers.

Figure 4.

Figure 4.

Catalytic XNAs and peptidomimetics. (a) Cleavage of a chimeric DNA/RNA substrate contains a single unpaired ribo-G residue by FANAzyme NGS12–7. Reproduced with permission from Ref. 103. Copyright 2018, Springer Nature. (b) The foldamer catalyst for macrocycle formation. Reproduced with permission from Ref. 105. Copyright 2019, AAAS.

Molecular catalysts based on peptidomimetics have also been developed. Kirshenbaum et al. synthesized a series of TEMPO conjugated peptoid helices to achieve the catalytic function of enantioselective alcohol-ketone transformation.104 They discovered that the enantioselectivity of the catalytic peptoids depends on the handedness of the helical scaffold, the position of the catalytic center along the peptoid backbone, and the folding of the peptoid scaffold. In another example, Gellman et al. reported a peptidomimetic foldamer containing both α- and β-amino acid residues to be used as catalysts on the ring-closing in macrocycles synthesis. When folding into a three-dimensional helix structure, this foldamer presents a primary amine and a secondary amine in defined positions to promote aldol reactions that form rings containing 14 to 22 atoms (Figure 4b).105

2.2.5. Protein mimics

Empowering synthetic polymers with protein-like structures and functions has been a longstanding goal for polymer chemists. Recent advances in peptidomimetic polymers have shed light on this important research field. Zuckermann et al. designed a peptoid helix bundle structure equipped with thiol and imidazole handles that can bind to zinc cations and bridge two helix bundles in a tertiary structure. Such a structure resembles the zinc-binding motif of zinc-finger domains found in DNA-binding proteins.106 Schepartz et al. studied the development of β-peptide bundles that mimic protein functions including carbohydrate recognition, metal binding, and catalytic activity.107 For example, they incorporated β-amino acid analogs with a boronic acid side chain into the β-peptide bundle to achieve high affinity to sorbitol, highlighting the potential to develop artificial β-peptide lectins in the future.108 Similarly, incorporating β3 -homocysteine (β3 C) residues in β-peptide bundles enabled selective Cd2+ ion binding activity over other metal ions tested (Hg2+, Pb2+, and Zn2+).109 They also designed a β-peptide bundle catalyst aiming to operate as esterases. The β-peptide bundle sequence was optimized to increase the affinity to the substrate, resulting in catalytic efficiency of kcat/KM = 98 M−1 min−1. Structural studies of the β-peptide bundle esterase reveal that bundle formation plays a critical role in catalysis, paving the road for future research on polymer-based protein mimics.110

2.3. Self-Assembled Nanostructures

Through judicious design, sequence-controlled polymers can self-assemble into nanostructures that have promising prospective applications in fields of lithography, drug delivery, sensors, and optoelectronics. Studies on these sequence-controlled polymers will also reveal important insights into the relationship between monomer sequences and the morphologies of polymer self-assembly.

2.3.1. Self-assembly of sequence-controlled synthetic copolymers

Johnson et al. prepared monodisperse diblock 32-mer copoly(triazoles) with different stereochemical sequences via iterative exponential growth (IEG). The group found that different monomer stereochemistry and sequences led to dramatic changes in polymer self-assembly (Figure 5a).111 Small-angle X-ray scattering (SAXS) and transmission electron microscopy (TEM) analyses showed that L/L diblock 32-mer, D/L diblock 32-mer, and alt/L diblock 32-mer yielded three distinct morphologies: double-gyroid, lamellar, and hexagonal cylinder, respectively. The incorporation of a small-molecule additive (L)-tartaric acid can swell the hydrophilic block and further induce the morphology transformation from gyroid to lamellar for L/L 32mer, and hexagonal cylinder to lamellar for alt/L 32mer. In another example of the self-assembly of sequence-defined polymers, Cheng et al. prepared a small library of polymers composed of two types of giant molecules polyhedral oligomeric silsesquioxane (POSS): hydrophilic DPOSS and hydrophobic BPOSS.112 The collective hydrogen bonding and hydrophobic interactions induced the nanophase separation. Tailoring the location of DPOSS and BPOSS in the polymer backbones results in different well-defined supramolecular lattices including lamellar, double gyroid, hexagonal cylinder, Frank–Kasper (F-K) A15, and body-centered cubic. Regulating sequence in polydispersed sequence-controlled polymers has also been used to tune their assembled structure in the bulk interface of two phases. For instance, Coates et al.113114 utilized polyethylene/isotactic polypropylene (PE/iPP) multiblock copolymers with precise control over block length as the compatibilizer to weld common grades of commercial PE and iPP together. With comparably short block length of PE and iPP, tetrablock copolymers PP36PE20PP34PE24 exhibited better adhesive strength than diblock PP24PE31 due to the formation of entangled loops in the interfaces. Simmons et al. employed a molecular-dynamics simulation-based genetic algorithm to predict the performance of sequence-specific copolymers as a compatibilizer. They showed that sequence-specific copolymers offered the potential to yield a larger reduction in interfacial energy than either block or random copolymers.115 In another example, Cai et al. investigated the “sequence-controlled polymerization-induced self-assembly” of block polyions in aqueous media.116 Three types of block polyions with different sequences of ionic monomers, i.e., block, gradient, and alternating zwitterionic sequences, were prepared via photoswitchable reversible addition-fragmentation chain-transfer (RAFT) copolymerization using polyethylene glycol chain transfer agents. While block zwitterionic copolymerization resulted in precipitation, and polyions with alternating zwitterionic copolymerization were water-soluble, various gradient zwitterionic sequences led to the formation of different nanostructures such as spheres, lamellae, and vesicles. These studies confirmed that polymer sequences can significantly affect inter- and intramolecular interactions, and effectively mediate the self-assembled nanostructures in bulk and in solution.

Figure 5.

Figure 5.

Self-assembled nanostructures formed by sequence-controlled polymers. (a) Unimolecular stereoisomeric diblock 32-mers synthesized by IEG. Their 1D intensity SAXS and WAXS plots indicate distinct phases. Reproduced with permission from Ref. 111. Copyright 2018, American Chemical Society. (b) 3D structure of tetrahedral assembled from AuNP-labelled XNA strands confirmed by TEM at different tilting angles. Reproduced with permission from Ref. 117. Copyright 2016, Wiley-VCH. (c) Antibody-mimetic peptoid nanosheets for molecular recognition. Reproduced with permission from Ref. 120. Copyright 2013, American Chemical Society.

2.3.2. XNA nanostructures

It is well established that natural nucleic acids such as DNA are highly programmable and can fold into complex 3D structures. Such programmability and complex nanostructure assembly have recently been extended to XNAs. For example, Holliger and coworkers described XNA-based 70 kDa Turberfield tetrahedron nanostructures from four different classes of XNAs: 2’F-RNA, FANA, HNA, and CeNA (Figure 5b).117 Impressively, despite their conformational heterogeneity, all XNAs successfully formed the same tetrahedra. A 600 kDa all-FANA octahedron was also assembled. Further analysis of the FANA octahedron structure revealed differences in shape and structural orientation compared to that of natural DNA, which may be due to the conformational differences in sugar puckering between FANA and DNA. Notably, the full XNA nanostructures were found to be significantly more stable than those made of natural DNA in the presence of nucleases.

2.3.3. Peptoid nanostructures

The facile synthesis and functionalization make peptoids a great material for building self-assembled nanostructures. Zuckermann et al. discovered a class of polypeptoids that can fold into water-soluble and precisely oriented nanosheets.118 The self-assembly of these polypeptoids formed a highly ordered, planar bilayer structure that could extend to hundreds of units long. This process was driven by the phase separation of the hydrophobic and charged monomers in the same peptoid sequence.119 These peptoid nanosheets had a chemically defined interior and exterior that were only 3 nm thick, and could be custom modified with various functional groups on the exterior surface. The authors further demonstrated the utility of these nanosheets by fabricating an antibody-mimetic two-dimensional substrate. Based on the structural component of natural antibodies, they decorated the exterior surface of the nanosheets with antibody-mimicking peptide loops. The high-surface area of the nanosheets served as a scaffold for a variety of functional peptide loops that could be used for binding to target biomolecules (Figure 5c).120121 Cryogenic transmission electron microscopy imaging of the crystal lattice of the diblock copolypeptoid nanosheets revealed further insights into how the aromatic side chains in the polypeptoid sequence influenced the crystal structure of the nanosheet crystals at the atomic level.122 Similarly, Zhang et al. found that progressively placing positive charge away from the hydrophobic block led to the increase of the micellar radius and aggregation number in the micelles assembled by amphiphilic ionic block copolypeptoids.123 Zuckermann et al. demonstrated that the microphase separation behavior of sequence-defined peptoid diblock copolymers in the solid state differed from the traditional block copolymers.124 In particular, only lamellar and disordered morphologies were observed over the entire window of the composition and temperature variation. Furthermore, the maximum order-disorder transition temperature was reached when the volume fraction of the hydrophilic block was 0.24 in these peptoid diblock copolymers, compared to 0.5 for the same parameter in conventional block copolymers.

2.4. Macroscopic Material Properties

The ability to regulate the sequence along the polymer chain created an important opportunity to fine tune the macroscopic properties of sequence-controlled polymers including degradability,125 thermal properties,126128 mechanical properties,129 optoelectronic properties,130133 etc. Meyer et al. investigated the impact of sequence on the degradation of poly(lactic-co-glycolic acid) (PLGA).125 They observed that, under neutral phosphate buffer, random copolymers showed exponential decay of 70% weight loss in four weeks, while the weight loss of the alternating copolymer was considerably smaller over the same period. After a rapid initial weight loss, the degradation of the alternating copolymers exhibited a near zero-order behavior. The authors attributed the slower degradation rates of the alternating copolymers to their unique sequence: the alternating copolymer has only one connectivity pattern, i.e., lactic-glycolic, while the random copolymers have two more connectivity patterns, i.e., lactic-lactic and glycolic-glycolic (Figure 6a). Since it was expected that the degradation rate followed the order of glycolic-glycolic > lactic-glycolic > lactic-lactic, the random copolymer degraded faster at the beginning and gradually slowed down. In contrast, the degradation rate of alternating copolymer remained steady over the course of the reaction.

Figure 6.

Figure 6.

Macroscopic material properties of sequence-controlled polymers. (a) Sequence-controlled PLGA copolymers demonstrated distinct hydrolytic degradation rates compared to random copolymers; Reproduced with permission from Ref. 125. Copyright 2011, American Chemical Society. (b) The sequence of polyurethane oligomers has an influence on network topology and mechanical properties upon crosslinking using dithiol linkers. Reproduced with permission from Ref. 129. Copyright 2020, American Chemical Society. (c) Diastereo- and enantioselective ROP of a diastereomeric mixture of rac/meso-8DLMe could generate diblock stereocopolymers it-sb-st-P3HB consisting of it- and st-stereoblocks of P3HB. Reproduced with permission from Ref. 136. Copyright 2019, AAAS.

Alabi et al. reported a clever strategy for the scalable production of sequence-defined polyurethane networks (Figure 6b).129 By utilizing highly efficient reductive amination and amine-anhydride condensation reactions, large scale synthesis of sequence-defined polyurethane macromers (SD-PUM) was readily achieved. Specifically, SD-PUM with three different sequences AAmmmA, mAmAmA, and AAAmmm (where A and m represent the units with allyl group and methyl group on urethane, respectively) were prepared and crosslinked by dithiol crosslinkers. It was revealed that polyurethane networks with AAAmmm sequence always showed lower Tg than mAmAmA- and AAmmmA networks. They also exhibited ~40% and ~30% lower rubbery modulus than that of mAmAmA- and AAmmmA networks. The authors explained that the relative amounts of intramolecular loops and dangling chain ends may account for different thermal and mechanical properties.

The stereosequence also plays an important role in influencing polymer properties. A typical example is that atactic poly(propylene succinate) synthesized by alternating copolymerization of racemic propylene epoxide and succinic anhydride showed Tg of −39 °C,126 while the isotactic poly(propylene succinate) from enantiopure propylene epoxide had a much lower Tg of −4 °C. This shows the huge influence of stereosequence on polyester thermal properties.134 Recently, Chen et al. reported an ingenious method to make stereosequenced polyhydroxyalkanoates from diastereomeric monomer mixtures.135 By using highly stereoselective catalysts, rac (R,R and S,S) or meso (R,S) cyclic diolide (8DLMe, Me denotes the methyl group on the 8DL ring) were polymerized both enantioselectively and diastereoselectively to form isotactic and syndiotactic poly(3-hydroxybutyric acid) (P3HB), respectively (Figure 6c).136 In addition, the polymerization of rac-8DLMe was faster than that of meso-8DLMe, thus the synthesis of a tapered stereodiblock polymer it-P3HB-sb-st-P3HB was possible. By changing the rac/meso ratio, it-P3HB-sb-st-P3HB with similar molecular weights (16.8–17.3 kg/mol) showed different Tm values. Changing the methyl group on monomer into an ethyl group allowed for the synthesis of more complex tapered copolymers with improved mechanical properties. These results clearly demonstrated the significance of stereosequence in determining the polymer properties.

3. Conclusion and Outlook

Since Staudinger’s seminal paper Über Polymerisation7 a century ago, polymer chemists have made great strides in controlling the molecular weight, dispersity, sequence, microstructure, and chain-end groups of synthetic polymers. Evidenced by this Viewpoint, the study on the sequence-controlled polymers is a burgeoning and highly interdisciplinary field featuring the convergence of many previously independent research topics such as polymer chemistry, biochemistry, glycoscience, nucleic acid chemistry, protein engineering, immunology, and information technology. The integration of ideas and expertise from these diverse disciplines helped to create sequence-controlled polymers that not only exhibited improved functionalities in the existing applications of synthetic polymers, but also stimulated novel applications inaccessible by traditional polymers, such as data storage, molecular recognition, programmable self-assembly, vaccine, and catalysis. Particular applications often require specific molecular features and degree of sequence control. For example, the application of data storage requires high polymer molecular weight, defined building block sequence, and monodispersity to ensure high data storage capacity and fidelity. Biological applications such as molecular recognition, antimicrobial activity, and drug delivery need the sequence-controlled polymers to be water soluble or amphiphilic. Cationic functionalities are often incorporated in the sequence-controlled polymers used for the antimicrobial application and RNA delivery due to the need for the electrostatic interaction with negatively charged molecules. As advancements continue to emerge, there are great opportunities for the development of new sequence-controlled polymers to provide improved or completely new functions in the future.

New functionalities of sequence-controlled polymers are synergistically connected with the development of new synthetic approaches. For instance, methods that can significantly increase the molecular weight and the scale of synthesis of sequence-defined polymers are essential for their information-carrying ability to approach that of DNA in data storage applications. Toward these goals, the IEG strategy developed by Johnson et al.111, 137 and the single unit monomer insertion (SUMI) approach developed by Xu and Boyer138140 have demonstrated promising potentials (Figure 7a), but further efforts are still needed: while the density of variable monomer units incorporated by IEG requires improvement, SUMI is still limited to short oligomers. Furthermore, compared to DNA, another limitation of existing sequence-defined polymers is their inability to undergo a replication and amplification process analogous to PCR of DNA, making it challenging for the decoding, amplification, and transmission process of the information carried in these polymers. New methods that can allow for replication and amplification of synthetic polymers, exemplified by recent advances in XNA technologies6 and DNA-templated polymerization,141 will offer important advantages in data storage over existing sequence-defined polymers.

Figure 7.

Figure 7.

Strategies to access sequence-controlled polymers with more complex structure and functions. (a) Schematic illustration of the single unit monomer insertion (SUMI) approach to access sequence-defined oligomers. Reproduced with permission from Ref. 139. Copyright 2017 Wiley-VCH. (b) Covalent folding of linear synthetic polymer chains could be achieved through strategic positioning of crosslinkable side chains in the polymer structure. Reproduced with permission from Ref. 142. Copyright 2011, Springer Nature. (c) Three methods for the sequence identification of functional sequence-controlled polymers: bioabstraction, combinatorial screening, and rational design. Reproduced with permission from Ref. 13. Copyright 2019, Royal Society of Chemistry.

One of the ultimate goals of the studies on sequence-controlled polymers is to generate polymers that can fold into protein-like globular structures and execute sophisticated chemical and biochemical functions such as receptor recognition and catalysis. Toward this goal, two distinct, but converging strategies have been adopted: (1) controlling the folding and the positioning of functional groups of single-chain polymers by regulating their sequence and composition, and (2) engineering existing biopolymers or recapitulating the structures of biopolymers by their synthetic mimics. An excellent example of the former approach was reported by Lutz et al., in which they added discrete amounts of functionalized maleimides at various time points during the ATRP polymerization of styrene to precisely introduce crosslinkable functional groups through the incorporated maleimides in the polymer chain, allowing for controlled polymer folding (Figure 7b).142 The latter approach has been explored by many researchers, yielding promising results thus far. The XNA nanostructures developed by Holliger143 and Chaput,59 and the peptoid and β-peptide bundles reported by Zuckermann106 and Schepartz,107 respectively, are notable examples. One particularly noteworthy direction that holds a great promise to generate protein-like synthetic polymers is to engineer the ribosome as a polymer synthesis machine. This formidable task may require many iterations of mutagenesis and structural optimization to the natural ribosome-tRNA machinery, but the seminal work by Fahnestock and Rich,144 and recent reports by Suga,145147 Schepartz,148150 and Jewett151 have shown important advances in engineering the ribosome-tRNA machinery to incorporate non-natural β - and γ -amino acids in polypeptides as well as α -hydroxyacids in polyesters.

In the process to generate functional sequence-controlled polymers, knowing what sequence to make is equally if not more important than developing the methods to make it.  Three distinct strategies: “bioabstraction”, rational de novo design, and selection/screening from combinatory libraries were originally summarized by Brummelhuis, Wilke, and Börner for identifying functional peptides,152 and were further elaborated by Austin and Rosales in a recent review for identifying sequence-controlled polymers (Figure 7c).13 “Bioabstraction” implies the direct translation of the biopolymer sequence into the sequence of synthetic polymers, which has been widely adopted to generate biomimetic polymers such as the peptidomimetic affinity reagents discussed in Section 2.2.1 and the antimicrobial polymers discussed in Section 2.2.2. Capitalizing on the recent advances in de novo protein design153 and artificial intelligence,154 the de novo design of sequence-controlled polymers has few published examples to date but has emerged as a highly active area of research widely anticipated to evolve rapidly in the near future. A significant obstacle to this strategy is that unlike the substantial databases of proteins, existing knowledge on the sequence-structure-functional relationship of synthetic polymers is few and far between, making it challenging to develop training sets for the computational algorithms. Therefore, further developments of the de novo design strategy for discovering functional sequence-controlled polymers will likely require inputs from new experimental approaches, as well as adapt from existing computational platforms for biopolymer structural prediction.155 Perhaps one of the most promising experimental approaches to discover functional sequence-controlled synthetic polymers is the selection/screening from combinatorial libraries. Reported approaches in this direction range from the screening of one-bead-one-compound (OBOC) library156 to the selection of DNA-encoded polymers.141 In all of these methods, with the exception of some XNA and peptidomimetics that can be sequenced directly, a key step is to tag each member of the combinatory library with a unique molecular barcode such that the sequence information of the hits from the selection/screening can be easily tracked and decoded. To this end, biopolymers that can be conveniently sequenced by existing technologies, such as peptides and DNA, are among the popular candidates for molecular barcodes. Methods that expand the scope of the molecular barcodes to include other information-carrying synthetic polymers (e.g., the sequence-defined polymers for data storage discussed in Section 2.1), and polymer sequencing methods (e.g., mass spectrometry-based and nanopore-based sequencing techniques157), will help to simplify the combinatorial library synthesis of novel sequence-controlled polymers and improve the efficiency of the selection/screening processes.

Acknowledgments

The authors acknowledge the support by the Arnold and Mabel Beckman Foundation through a Beckman Young Investigator award to J.N. and the NIH (1DP2HG011027-01 to J.N.). We thank Gavin Giardino for proofreading the manuscript.

Footnotes

Conflict of Interest

The authors declare no conflict of interest.

References

  • 1.Staudinger H, Macromolecular chemistry. Nobel Lecture 1953, 397–419. [Google Scholar]
  • 2.Merryfield R, Solid phase peptide synthesis. J. Am. Chem. Soc 1963, 85, 2149. [Google Scholar]
  • 3.Szwarc M; Levy M; Milkovich R, Polymerization initiated by electron transfer to monomer. A new method of formation of block polymers1. J. Am. Chem. Soc 1956, 78 (11), 2656–2657. [Google Scholar]
  • 4.Lutz JF, Defining the Field of Sequence-Controlled Polymers. Macro. Rapid. Commun 2017. [DOI] [PubMed] [Google Scholar]
  • 5.Chaput JC; Herdewijn P, What is XNA? Angew. Chem. Int. Ed 2019, 58 (34), 11570–11572. [DOI] [PubMed] [Google Scholar]
  • 6.Pinheiro VB; Holliger P, Towards XNA nanotechnology: new materials from synthetic genetic polymers. Trends Biotechnol 2014, 32 (6), 321–328. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 7.Staudinger H; Ochiai E, Über hochpolymere Verbindungen. Zeitschrift für Physikalische Chemie 1932, 158 (1), 35–55. [Google Scholar]
  • 8.Mülhaupt R, Hermann Staudinger and the origin of macromolecular chemistry. Angew. Chem. Int. Ed 2004, 43 (9), 1054–1063. [DOI] [PubMed] [Google Scholar]
  • 9.Lutz J-F; Ouchi M; Liu DR; Sawamoto M, Sequence-controlled polymers. Science 2013, 341 (6146), 1238149. [DOI] [PubMed] [Google Scholar]
  • 10.Lutz J-F, Sequence-controlled polymerizations: the next Holy Grail in polymer science? Polym. Chem 2010, 1 (1), 55–62. [Google Scholar]
  • 11.Badi N; Lutz J-F, Sequence control in polymer synthesis. Chem. Soc. Rev 2009, 38 (12), 3383–3390. [DOI] [PubMed] [Google Scholar]
  • 12.Lutz J-F; Lehn J-M; Meijer E; Matyjaszewski K, From precision polymers to complex materials and systems. Nat. Rev. Mater 2016, 1 (5), 16024. [Google Scholar]
  • 13.Austin MJ; Rosales AM, Tunable biomaterials from synthetic, sequence-controlled polymers. Biomater. Sci 2019, 7 (2), 490–505. [DOI] [PubMed] [Google Scholar]
  • 14.Nanjan P; Porel M, Sequence-defined non-natural polymers: synthesis and applications. Polym. Chem 2019, 10 (40), 5406–5424. [Google Scholar]
  • 15.Meier MA; Barner-Kowollik C, A New Class of Materials: Sequence-Defined Macromolecules and Their Emerging Applications. Adv. Mater 2019, 31 (26), 1806027. [DOI] [PubMed] [Google Scholar]
  • 16.Colquhoun H; Lutz J-F, Information-containing macromolecules. Nat. Chem 2014, 6 (6), 455–456. [DOI] [PubMed] [Google Scholar]
  • 17.Lutz J-F, 100th Anniversary of Macromolecular Science Viewpoint: Toward Artificial Life-Supporting Macromolecules. ACS Macro Lett 2020, 9 (2), 185–189. [DOI] [PubMed] [Google Scholar]
  • 18.ten Brummelhuis N, Controlling monomer-sequence using supramolecular templates. Polym. Chem 2015, 6 (5), 654–667. [Google Scholar]
  • 19.Solleder SC; Schneider RV; Wetzel KS; Boukis AC; Meier MA, Recent progress in the design of monodisperse, sequence-defined macromolecules. Macro. Rapid. Commun 2017, 38 (9), 1600711. [DOI] [PubMed] [Google Scholar]
  • 20.Ouchi M; Sawamoto M, Sequence-controlled polymers via reversible-deactivation radical polymerization. Polym. J 2018, 50 (1), 83–94. [Google Scholar]
  • 21.Szymański JK; Abul-Haija YM; Cronin L, Exploring strategies to bias sequence in natural and synthetic oligomers and polymers. Acc. Chem. Res 2018, 51 (3), 649–658. [DOI] [PubMed] [Google Scholar]
  • 22.Yang C; Flynn JP; Niu J, Facile Synthesis of Sequence-Regulated Synthetic Polymers Using Orthogonal SuFEx and CuAAC Click Reactions. Angew. Chem. Int. Ed 2018, 57 (49), 16194–16199. [DOI] [PubMed] [Google Scholar]
  • 23.Church GM; Gao Y; Kosuri S, Next-generation digital information storage in DNA. Science 2012, 337 (6102), 1628–1628. [DOI] [PubMed] [Google Scholar]
  • 24.Lutz J.-F. o., Coding macromolecules: inputting information in polymers using monomer-based alphabets. Macromolecules 2015, 48 (14), 4759–4767. [Google Scholar]
  • 25.Martens S; Landuyt A; Espeel P; Devreese B; Dawyndt P; Du Prez F, Multifunctional sequence-defined macromolecules for chemical data storage. Nat. Commun 2018, 9 (1), 1–8. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 26.Huang Z; Shi Q; Guo J; Meng F; Zhang Y; Lu Y; Qian Z; Li X; Zhou N; Zhang Z, Binary tree-inspired digital dendrimer. Nat. Commun 2019, 10 (1), 1–7. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 27.Boukis AC; Reiter K; Frölich M; Hofheinz D; Meier MA, Multicomponent reactions provide key molecules for secret communication. Nat. Commun 2018, 9 (1), 1–10. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 28.Zydziak N; Konrad W; Feist F; Afonin S; Weidner S; Barner-Kowollik C, Coding and decoding libraries of sequence-defined functional copolymers synthesized via photoligation. Nat. Commun 2016, 7 (1), 1–13. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 29.Al Ouahabi A; Charles L; Lutz J.-F. o., Synthesis of non-natural sequence-encoded polymers using phosphoramidite chemistry. J. Am. Chem. Soc 2015, 137 (16), 5629–5635. [DOI] [PubMed] [Google Scholar]
  • 30.Al Ouahabi A; Kotera M; Charles L; Lutz J.-F. o., Synthesis of monodisperse sequence-coded polymers with chain lengths above DP100. ACS Macro Lett 2015, 4 (10), 1077–1080. [DOI] [PubMed] [Google Scholar]
  • 31.Cavallo G; Al Ouahabi A; Oswald L; Charles L; Lutz J-F, Orthogonal synthesis of “easy-to-read” information-containing polymers using phosphoramidite and radical coupling steps. J. Am. Chem. Soc 2016, 138 (30), 9417–9420. [DOI] [PubMed] [Google Scholar]
  • 32.Roy RK; 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. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 33.Karamessini D; Petit BE; Bouquey M; Charles L; Lutz JF, Identification-tagging of methacrylate-based intraocular implants using sequence defined polyurethane barcodes. Adv. Func. Mater 2017, 27 (3), 1604595. [Google Scholar]
  • 34.Gunay US; Petit BE; Karamessini D; Al Ouahabi A; Amalian J-A; Chendo C; Bouquey M; Gigmes D; Charles L; Lutz J-F, Chemoselective synthesis of uniform sequence-coded polyurethanes and their use as molecular tags. Chem 2016, 1 (1), 114–126. [Google Scholar]
  • 35.Zhao B; Gao Z; Zheng Y; Gao C, Scalable Synthesis of Positively Charged Sequence-Defined Functional Polymers. J. Am. Chem. Soc 2019, 141 (11), 4541–4546. [DOI] [PubMed] [Google Scholar]
  • 36.Dahlhauser SD; Escamilla PR; VandeWalle AN; York JT; Rapagnani RM; Shei JS; Glass SA; Coronado JN; Moor SR; Saunders DP, Sequencing of Sequence-Defined Oligourethanes via Controlled Self-Immolation. J. Am. Chem. Soc 2020, 142 (6), 2744–2749. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 37.König NF; Al Ouahabi A; Oswald L; Szweda R; Charles L; Lutz J-F, Photo-editable macromolecular information. Nat. Commun 2019, 10 (1), 1–9. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 38.Laure C; Karamessini D; Milenkovic O; Charles L; Lutz JF, Coding in 2D: using intentional dispersity to enhance the information capacity of sequence-coded polymer barcodes. Angew. Chem. Int. Ed 2016, 55 (36), 10722–10725. [DOI] [PubMed] [Google Scholar]
  • 39.Kiessling LL; Grim JC, Glycopolymer probes of signal transduction. Chem. Soc. Rev 2013, 42 (10), 4476–4491. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 40.Ponader D; Maffre P; Aretz J; Pussak D; Ninnemann NM; Schmidt S; Seeberger PH; Rademacher C; Nienhaus GU; Hartmann L, Carbohydrate-Lectin recognition of sequence-defined heteromultivalent glycooligomers. J. Am. Chem. Soc 2014, 136, 2008–2016. [DOI] [PubMed] [Google Scholar]
  • 41.Ponader D; Wojcik F; Beceren-Braun F; Dernedde J; Hartmann L, Sequence-defined glycopolymer segments presenting mannose: Synthesis and lectin binding affinity. Biomacromolecules 2012, 13, 1845–1852. [DOI] [PubMed] [Google Scholar]
  • 42.Zhang Q; Collins J; Anastasaki A; Wallis R; Mitchell DA; Becer CR; Haddleton DM, Sequence-controlled multi-block glycopolymers to inhibit DC-SIGN-gp120 binding. Angew. Chem. Int. Ed 2013, 52 (16), 4435–4439. [DOI] [PubMed] [Google Scholar]
  • 43.Thirunavukarasu D; Chen T; Liu Z; Hongdilokkul N; Romesberg FE, Selection of 2′-fluoro-modified aptamers with optimized properties. J. Am. Chem. Soc 2017, 139 (8), 2892–2895. [DOI] [PubMed] [Google Scholar]
  • 44.Alves Ferreira-Bravo I; Cozens C; Holliger P; DeStefano JJ, Selection of 2′-deoxy-2′-fluoroarabinonucleotide (FANA) aptamers that bind HIV-1 reverse transcriptase with picomolar affinity. Nucleic Acids Res 2015, 43 (20), 9587–9599. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 45.Rose KM; Alves Ferreira-Bravo I; Li M; Craigie R; Ditzler MA; Holliger P; DeStefano JJ, Selection of 2′-Deoxy-2′-Fluoroarabino Nucleic Acid (FANA) Aptamers That Bind HIV-1 Integrase with Picomolar Affinity. ACS Chem. Biol 2019, 14 (10), 2166–2175. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 46.Vaught JD; Bock C; Carter J; Fitzwater T; Otis M; Schneider D; Rolando J; Waugh S; Wilcox SK; Eaton BE, Expanding the chemistry of DNA for in vitro selection. J. Am. Chem. Soc 2010, 132 (12), 4141–4151. [DOI] [PubMed] [Google Scholar]
  • 47.Rohloff JC; Gelinas AD; Jarvis TC; Ochsner UA; Schneider DJ; Gold L; Janjic N, Nucleic acid ligands with protein-like side chains: modified aptamers and their use as diagnostic and therapeutic agents. Mol. Ther. Nucleic Acids 2014, 3, e201. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 48.Chen Z; Lichtor PA; Berliner AP; Chen JC; Liu DR, Evolution of sequence-defined highly functionalized nucleic acid polymers. Nat. Chem 2018, 10 (4), 420–427. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 49.Lichtor PA; Chen Z; Elowe NH; Chen JC; Liu DR, Side chain determinants of biopolymer function during selection and replication. Nat. Chem. Biol 2019, 15 (4), 419–426. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 50.Kong D; Yeung W; Hili R, In vitro selection of diversely functionalized aptamers. J. Am. Chem. Soc 2017, 139 (40), 13977–13980. [DOI] [PubMed] [Google Scholar]
  • 51.Gordon CK; Wu D; Pusuluri A; Feagin TA; Csordas AT; Eisenstein MS; Hawker CJ; Niu J; Soh HT, Click-particle display for base-modified aptamer discovery. ACS Chem. Biol 2019, 14 (12), 2652–2662. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 52.MacPherson IS; Temme JS; Habeshian S; Felczak K; Pankiewicz K; Hedstrom L; Krauss IJ, Multivalent glycocluster design through directed evolution. Angew. Chem. Int. Ed 2011, 50 (47), 11238–11242. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 53.Temme JS; MacPherson IS; DeCourcey JF; Krauss IJ, High temperature SELMA: Evolution of DNA-supported oligomannose clusters which are tightly recognized by HIV bnAb 2G12. J. Am. Chem. Soc 2014, 136 (5), 1726–1729. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 54.Horiya S; Bailey JK; Temme JS; Guillen Schlippe YV; Krauss IJ, Directed evolution of multivalent glycopeptides tightly recognized by HIV antibody 2G12. J. Am. Chem. Soc 2014, 136 (14), 5407–5415. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 55.Nguyen DN; Xu B; Stanfield RL; Bailey JK; Horiya S; Temme JS; Leon DR; LaBranche CC; Montefiori DC; Costello CE, Oligomannose glycopeptide conjugates elicit antibodies targeting the glycan core rather than its extremities. ACS Cent. Sci 2019, 5 (2), 237–249. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 56.Shen X; Corey DR, Chemistry, mechanism and clinical status of antisense oligonucleotides and duplex RNAs. Nucleic Acids Res 2018, 46 (4), 1584–1600. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 57.Milner N; Mir KU; Southern EM, Selecting effective antisense reagents on combinatorial oligonucleotide arrays. Nat. Biotechnol 1997, 15 (6), 537–541. [DOI] [PubMed] [Google Scholar]
  • 58.Sohail M; Southern E, Hybridization of antisense reagents to RNA. Curr. Opin. Mol. Ther 2000, 2 (3), 264–271. [PubMed] [Google Scholar]
  • 59.Rinaldi C; Wood MJ, Antisense oligonucleotides: the next frontier for treatment of neurological disorders. Nat. Rev. Neurol 2018, 14 (1), 9. [DOI] [PubMed] [Google Scholar]
  • 60.Quemener AM; Bachelot L; Forestier A; Donnou-Fournet E; Gilot D; Galibert MD, The powerful world of antisense oligonucleotides: From bench to bedside. Wiley Interdiscip. Rev. RNA 2020, e1594. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 61.Zuckermann RN; Martin EJ; Spellmeyer DC; Stauber GB; Shoemaker KR; Kerr JM; Figliozzi GM; Goff DA; Siani MA; Simon RJ, Discovery of nanomolar ligands for 7-transmembrane G-protein-coupled receptors from a diverse N-(substituted) glycine peptoid library. J. Med. Chem 1994, 37 (17), 2678–2685. [DOI] [PubMed] [Google Scholar]
  • 62.Yazdani U; Zaman S; Hynan LS; Brown LS; Dewey RB; Karp D; German DC, Blood biomarker for Parkinson disease: peptoids. NPJ Parkinsons Dis 2016, 2 (1), 1–7. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 63.Gearhart TL; Montelaro RC; Schurdak ME; Pilcher CD; Rinaldo CR; Kodadek T; Park Y; Islam K; Yurko R; Marques ET Jr, Selection of a potential diagnostic biomarker for HIV infection from a random library of non-biological synthetic peptoid oligomers. J. Immunol. Methods 2016, 435, 85–89. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 64.Reddy MM; Wilson R; Wilson J; Connell S; Gocke A; Hynan L; German D; Kodadek T, Identification of candidate IgG biomarkers for Alzheimer’s disease via combinatorial library screening. Cell 2011, 144 (1), 132–142. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 65.Raveendra BL; Wu H; Baccala R; Reddy MM; Schilke J; Bennett JL; Theofilopoulos AN; Kodadek T, Discovery of peptoid ligands for anti-aquaporin 4 antibodies. Chem. Biol 2013, 20 (3), 351–359. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 66.Seo J; Michaelian N; Owens SC; Dashner ST; Wong AJ; Barron AE; Carrasco MR, Chemoselective and microwave-assisted synthesis of glycopeptoids. Org. Lett 2009, 11 (22), 5210–5213. [DOI] [PubMed] [Google Scholar]
  • 67.Kritzer JA; Lear JD; Hodsdon ME; Schepartz A, Helical β-peptide inhibitors of the p53-hDM2 interaction. J. Am. Chem. Soc 2004, 126 (31), 9468–9469. [DOI] [PubMed] [Google Scholar]
  • 68.Bautista AD; Appelbaum JS; Craig CJ; Michel J; Schepartz A, Bridged β3-peptide inhibitors of p53-hDM2 complexation: Correlation between affinity and cell permeability. J. Am. Chem. Soc 2010, 132 (9), 2904–2906. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 69.Stephens OM; Kim S; Welch BD; Hodsdon ME; Kay MS; Schepartz A, Inhibiting HIV fusion with a β-peptide foldamer. J. Am. Chem. Soc 2005, 127 (38), 13126–13127. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 70.Bautista AD; Stephens OM; Wang L; Domaoal RA; Anderson KS; Schepartz A, Identification of a β3-peptide HIV fusion inhibitor with improved potency in live cells. Bioorganic Med. Chem. Lett 2009, 19 (14), 3736–3738. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 71.Sang P; Shi Y; Huang B; Xue S; Odom T; Cai J, Sulfono-γ-AApeptides as Helical Mimetics: Crystal Structures and Applications. Acc. Chem. Res 2020, 9916–9920. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 72.She F; Teng P; Peguero-Tejada A; Wang M; Ma N; Odom T; Zhou M; Gjonaj E; Wojtas L; van der Vaart A, De Novo Left-Handed Synthetic Peptidomimetic Foldamers. Angew. Chem. Int. Ed 2018, 130 (31), 10064–10068. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 73.Teng P; Gray GM; Zheng M; Singh S; Li X; Wojtas L; van Der Vaart A; Cai J, Orthogonal Halogen-Bonding-Driven 3D Supramolecular Assembly of Right-Handed Synthetic Helical Peptides. Angew. Chem. Int. Ed 2019, 131 (23), 7860–7864. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 74.Sang P; Zhang M; Shi Y; Li C; Abdulkadir S; Li Q; Ji H; Cai J, Inhibition of β-catenin/B cell lymphoma 9 protein− protein interaction using α-helix–mimicking sulfono-γ-AApeptide inhibitors. Proc. Natl. Acad. Sci 2019, 116 (22), 10757–10762. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 75.Sang P; Shi Y; Higbee P; Wang M; Abdulkadir S; Lu J; Daughdrill G; Chen J; Cai J, Rational Design and Synthesis of Right-Handed d-Sulfono-γ-AApeptide Helical Foldamers as Potent Inhibitors of Protein–Protein Interactions. J. Org. Chem 2020, 85 (16), 10552–10560. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 76.Sang P; Shi Y; Lu J; Chen L; Yang L; Borcherds W; Abdulkadir S; Li Q; Daughdrill G; Chen J, α-Helix-mimicking sulfono-γ-AApeptide inhibitors for p53–MDM2/MDMX protein–protein interactions. J. Med. Chem 2020, 63 (3), 975–986. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 77.Sang P; Zhou Z; Shi Y; Lee C; Amso Z; Huang D; Odom T; Nguyen-Tran VT; Shen W; Cai J, The activity of sulfono-γ-AApeptide helical foldamers that mimic GLP-1. Sci. Adv 2020, 6 (20), eaaz4988. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 78.Porel M; Thornlow DN; Phan NN; Alabi CA, Sequence-defined bioactive macrocycles via an acid-catalysed cascade reaction. Nat. Chem 2016, 8 (6), 590. [DOI] [PubMed] [Google Scholar]
  • 79.Porel M; Thornlow DN; Artim CM; Alabi CA, Sequence-defined backbone modifications regulate antibacterial activity of oligoTEAs. ACS Chem. Biol 2017, 12 (3), 715–723. [DOI] [PubMed] [Google Scholar]
  • 80.Artim CM; Phan NN; Alabi CA, Effect of composition on antibacterial activity of sequence-defined cationic oligothioetheramides. ACS Infect. Dis 2018, 4 (8), 1257–1263. [DOI] [PubMed] [Google Scholar]
  • 81.Judzewitsch PR; Nguyen TK; Shanmugam S; Wong EH; Boyer C, Towards Sequence-Controlled Antimicrobial Polymers: Effect of Polymer Block Order on Antimicrobial Activity. Angew. Chem. Int. Ed 2018, 130 (17), 4649–4654. [DOI] [PubMed] [Google Scholar]
  • 82.Kuroki A; Sangwan P; Qu Y; Peltier R; Sanchez-Cano C; Moat J; Dowson CG; Williams EG; Locock KE; Hartlieb M, Sequence control as a powerful tool for improving the selectivity of antimicrobial polymers. ACS Appl. Mater. Inter 2017, 9 (46), 40117–40126. [DOI] [PubMed] [Google Scholar]
  • 83.Zasloff M, Antimicrobial peptides of multicellular organisms. nature 2002, 415 (6870), 389–395. [DOI] [PubMed] [Google Scholar]
  • 84.Timofeeva L; Kleshcheva N, Antimicrobial polymers: mechanism of action, factors of activity, and applications. Appl. Microbiol. Biotechnol 2011, 89 (3), 475–492. [DOI] [PubMed] [Google Scholar]
  • 85.Czyzewski AM; Jenssen H; Fjell CD; Waldbrook M; Chongsiriwatana NP; Yuen E; Hancock RE; Barron AE, In vivo, in vitro, and in silico characterization of peptoids as antimicrobial agents. PloS one 2016, 11 (2), e0135961. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 86.Dürr S; Thomason JC, Biofouling John Wiley & Sons: 2009. [Google Scholar]
  • 87.van Zoelen W; Buss HG; Ellebracht NC; Lynd NA; Fischer DA; Finlay J; Hill S; Callow ME; Callow JA; Kramer EJ; Zuckerman RN; Segalman RA, Sequence of hydrophobic and hydrophilic residues in amphiphilic polymer coatings affects surface structure and marine antifouling/fouling release properties. ACS Macro Lett 2014, 3 (4), 364–368. [DOI] [PubMed] [Google Scholar]
  • 88.Lau KHA; Sileika TS; Park SH; Sousa AM; Burch P; Szleifer I; Messersmith PB, Molecular Design of Antifouling Polymer Brushes Using Sequence-Specific Peptoids. Adv. Mater. Interfaces 2015, 2 (1), 1400225. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 89.Sgolastra F; Minter LM; Osborne BA; Tew GN, Importance of sequence specific hydrophobicity in synthetic protein transduction domain mimics. Biomacromolecules 2014, 15 (3), 812–820. [DOI] [PubMed] [Google Scholar]
  • 90.Backlund CM; Sgolastra F; Otter R; Minter LM; Takeuchi T; Futaki S; Tew GN, Increased hydrophobic block length of PTDMs promotes protein internalization. Polym. Chem 2016, 7 (48), 7514–7521. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 91.Martin L; Peltier R; Kuroki A. s.; Town JS; Perrier S. b., Investigating Cell Uptake of Guanidinium-Rich RAFT Polymers: Impact of Comonomer and Monomer Distribution. Biomacromolecules 2018, 19 (8), 3190–3200. [DOI] [PubMed] [Google Scholar]
  • 92.Schaffert D; Troiber C; Salcher EE; Fröhlich T; Martin I; Badgujar N; Dohmen C; Edinger D; Kläger R; Maiwald G, Solid-phase synthesis of sequence-defined T-, i-, and U-shape polymers for pDNA and siRNA delivery. Angew. Chem. Int. Ed 2011, 50 (38), 8986–8989. [DOI] [PubMed] [Google Scholar]
  • 93.Fröhlich T; Edinger D; Kläger R; Troiber C; Salcher E; Badgujar N; Martin I; Schaffert D; Cengizeroglu A; Hadwiger P, Structure–activity relationships of siRNA carriers based on sequence-defined oligo (ethane amino) amides. J. Control. Release 2012, 160 (3), 532–541. [DOI] [PubMed] [Google Scholar]
  • 94.Klein PM; Kern S; Lee D-J; Schmaus J; Höhn M; Gorges J; Kazmaier U; Wagner E, Folate receptor-directed orthogonal click-functionalization of siRNA lipopolyplexes for tumor cell killing in vivo. Biomaterials 2018, 178, 630–642. [DOI] [PubMed] [Google Scholar]
  • 95.Maier K; Martin I; Wagner E, Sequence defined disulfide-linked shuttle for strongly enhanced intracellular protein delivery. Mol. Pharm 2012, 9 (12), 3560–3568. [DOI] [PubMed] [Google Scholar]
  • 96.Lächelt U; Wittmann V; Müller K; Edinger D; Kos P; Höhn M; Wagner E, Synthetic polyglutamylation of dual-functional MTX ligands for enhanced combined cytotoxicity of poly (I: C) nanoplexes. Mol. Pharm 2014, 11 (8), 2631–2639. [DOI] [PubMed] [Google Scholar]
  • 97.Zhang P; He D; Klein PM; Liu X; Röder R; Döblinger M; Wagner E, Enhanced Intracellular Protein Transduction by Sequence Defined Tetra-Oleoyl Oligoaminoamides Targeted for Cancer Therapy. Adv. Func. Mater 2015, 25 (42), 6627–6636. [Google Scholar]
  • 98.Konca YU; Kirshenbaum K; Zuckermann RN, Nanometer-scale siRNA carriers incorporating peptidomimetic oligomers: physical characterization and biological activity. Int. J. Nanomedicine 2014, 9, 2271. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 99.Huang C-Y; Uno T; Murphy JE; Lee S; Hamer JD; Escobedo JA; Cohen FE; Radhakrishnan R; Dwarki V; Zuckermann RN, Lipitoids—novel cationic lipids for cellular delivery of plasmid DNA in vitro. Chem. Biol 1998, 5 (6), 345–354. [DOI] [PubMed] [Google Scholar]
  • 100.Utku Y; Dehan E; Ouerfelli O; Piano F; Zuckermann RN; Pagano M; Kirshenbaum K, A peptidomimetic siRNA transfection reagent for highly effective gene silencing. Mol. Biosyst 2006, 2 (6–7), 312–317. [DOI] [PubMed] [Google Scholar]
  • 101.Lobo BA; Vetro JA; Suich DM; Zuckermann RN; Middaugh CR, Structure/function analysis of peptoid/lipitoid: DNA complexes. J. Pharm. Sci 2003, 92 (9), 1905–1918. [DOI] [PubMed] [Google Scholar]
  • 102.Taylor AI; Pinheiro VB; Smola MJ; Morgunov AS; Peak-Chew S; Cozens C; Weeks KM; Herdewijn P; Holliger P, Catalysts from synthetic genetic polymers. Nature 2015, 518 (7539), 427–430. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 103.Wang Y; Ngor AK; Nikoomanzar A; Chaput JC, Evolution of a general RNA-cleaving FANA enzyme. Nat. Commun 2018, 9 (1), 1–10. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 104.Maayan G; Ward MD; Kirshenbaum K, Folded biomimetic oligomers for enantioselective catalysis. Proc. Natl. Acad. Sci 2009, 106 (33), 13679–13684. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 105.Girvin ZC; Andrews MK; Liu X; Gellman SH, Foldamer-templated catalysis of macrocycle formation. Science 2019, 366 (6472), 1528–1531. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 106.Lee B-C; Chu TK; Dill KA; Zuckermann RN, Biomimetic nanostructures: Creating a high-affinity zinc-binding site in a folded nonbiological polymer. J. Am. Chem. Soc 2008, 130 (27), 8847–8855. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 107.Wang PS; Schepartz A, β-Peptide bundles: Design. Build. Analyze. Biosynthesize. Chem. Commun 2016, 52 (47), 7420–7432. [DOI] [PubMed] [Google Scholar]
  • 108.Melicher MS; Walker AS; Shen J; Miller SJ; Schepartz A, Improved carbohydrate recognition in water with an electrostatically enhanced β-peptide bundle. Org. Lett 2015, 17 (19), 4718–4721. [DOI] [PubMed] [Google Scholar]
  • 109.Miller JP; Melicher MS; Schepartz A, Positive allostery in metal ion binding by a cooperatively folded β-peptide bundle. J. Am. Chem. Soc 2014, 136 (42), 14726–14729. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 110.Wang PS; Nguyen JB; Schepartz A, Design and high-resolution structure of a β3-peptide bundle catalyst. J. Am. Chem. Soc 2014, 136 (19), 6810–6813. [DOI] [PubMed] [Google Scholar]
  • 111.Golder MR; Jiang Y; Teichen PE; Nguyen HV-T; Wang W; Milos N; Freedman SA; Willard AP; Johnson JA, Stereochemical sequence dictates unimolecular diblock copolymer assembly. J. Am. Chem. Soc 2018, 140 (5), 1596–1599. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 112.Zhang W; Lu X; Mao J; Hsu CH; Mu G; Huang M; Guo Q; Liu H; Wesdemiotis C; Li T, Sequence-Mandated, Distinct Assembly of Giant Molecules. Angew. Chem. Int. Ed 2017, 129 (47), 15210–15215. [DOI] [PubMed] [Google Scholar]
  • 113.Eagan JM; Xu J; Di Girolamo R; Thurber CM; Macosko CW; LaPointe AM; Bates FS; Coates GW, Combining polyethylene and polypropylene: Enhanced performance with PE/iPP multiblock polymers. Science 2017, 355 (6327), 814–816. [DOI] [PubMed] [Google Scholar]
  • 114.Xu J; Eagan JM; Kim S-S; Pan S; Lee B; Klimovica K; Jin K; Lin T-W; Howard MJ; Ellison CJ, Compatibilization of Isotactic Polypropylene (i PP) and High-Density Polyethylene (HDPE) with i PP–PE Multiblock Copolymers. Macromolecules 2018, 51 (21), 8585–8596. [Google Scholar]
  • 115.Meenakshisundaram V; Hung J-H; Patra TK; Simmons DS, Designing sequence-specific copolymer compatibilizers using a molecular-dynamics-simulation-based genetic algorithm. Macromolecules 2017, 50 (3), 1155–1166. [Google Scholar]
  • 116.Wang L; Ding Y; Liu Q; Zhao Q; Dai X; Lu X; Cai Y, Sequence-controlled polymerization-induced self-assembly. ACS Macro Lett 2019, 8 (5), 623–628. [DOI] [PubMed] [Google Scholar]
  • 117.Taylor AI; Beuron F; Peak-Chew SY; Morris EP; Herdewijn P; Holliger P, Nanostructures from synthetic genetic polymers. ChemBioChem 2016, 17 (12), 1107. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 118.Nam KT; Shelby SA; Choi PH; Marciel AB; Chen R; Tan L; Chu TK; Mesch RA; Lee B-C; Connolly MD, Free-floating ultrathin two-dimensional crystals from sequence-specific peptoid polymers. Nat. Mater 2010, 9 (5), 454–460. [DOI] [PubMed] [Google Scholar]
  • 119.Kudirka R; Tran H; Sanii B; Nam KT; Choi PH; Venkateswaran N; Chen R; Whitelam S; Zuckermann RN, Folding of a single-chain, information-rich polypeptoid sequence into a highly ordered nanosheet. Pept. Sci 2011, 96 (5), 586–595. [DOI] [PubMed] [Google Scholar]
  • 120.Olivier GK; Cho A; Sanii B; Connolly MD; Tran H; Zuckermann RN, Antibody-mimetic peptoid nanosheets for molecular recognition. ACS nano 2013, 7 (10), 9276–9286. [DOI] [PubMed] [Google Scholar]
  • 121.Kim JH; Kim SC; Kline MA; Grzincic EM; Tresca BW; Cardiel J; Karbaschi M; Dehigaspitiya DC; Chen Y; Udumula V, Discovery of Stable and Selective Antibody Mimetics from Combinatorial Libraries of Polyvalent, Loop-Functionalized Peptoid Nanosheets. ACS nano 2019, 14 (1), 185–195. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 122.Xuan S; Jiang X; Spencer RK; Li NK; Prendergast D; Balsara NP; Zuckermann RN, Atomic-level engineering and imaging of polypeptoid crystal lattices. Proc. Natl. Acad. Sci 2019, 116 (45), 22491–22499. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 123.Sternhagen GL; Gupta S; Zhang Y; John V; Schneider GJ; Zhang D, Solution self-assemblies of sequence-defined ionic peptoid block copolymers. J. Am. Chem. Soc 2018, 140 (11), 4100–4109. [DOI] [PubMed] [Google Scholar]
  • 124.Sun J; Teran AA; Liao X; Balsara NP; Zuckermann RN, Nanoscale phase separation in sequence-defined peptoid diblock copolymers. J. Am. Chem. Soc 2013, 135 (38), 14119–14124. [DOI] [PubMed] [Google Scholar]
  • 125.Li J; Stayshich RM; Meyer TY, Exploiting sequence to control the hydrolysis behavior of biodegradable PLGA copolymers. J. Am. Chem. Soc 2011, 133 (18), 6910–3. [DOI] [PubMed] [Google Scholar]
  • 126.Darensbourg DJ; Poland RR; Escobedo C, Kinetic Studies of the Alternating Copolymerization of Cyclic Acid Anhydrides and Epoxides, and the Terpolymerization of Cyclic Acid Anhydrides, Epoxides, and CO2 Catalyzed by (salen)CrIIICl. Macromolecules 2012, 45 (5), 2242–2248. [Google Scholar]
  • 127.Stayshich RM; Meyer TY, New insights into poly (lactic-co-glycolic acid) microstructure: using repeating sequence copolymers to decipher complex NMR and thermal behavior. J. Am. Chem. Soc 2010, 132 (31), 10920–10934. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 128.McGraw ML; Clarke RW; Chen EY-X, Compounded Sequence Control in Polymerization of One-Pot Mixtures of Highly Reactive Acrylates by Differentiating Lewis Pairs. J. Am. Chem. Soc 2020, 142 (13), 5969–5973. [DOI] [PubMed] [Google Scholar]
  • 129.Hoff EA; De Hoe GX; Mulvaney CM; Hillmyer MA; Alabi CA, Thiol-Ene Networks from Sequence-Defined Polyurethane Macromers. J Am Chem Soc 2020, 142 (14), 6729–6736. [DOI] [PubMed] [Google Scholar]
  • 130.Meng H; Zheng J; Lovinger AJ; Wang B-C; Van Patten PG; Bao Z, Oligofluorene− thiophene derivatives as high-performance semiconductors for organic thin film transistors. Chem. Mater 2003, 15 (9), 1778–1787. [Google Scholar]
  • 131.Yu H; Li S; Schwieter KE; Liu Y; Sun B; Moore JS; Schroeder CM, Charge Transport in Sequence-Defined Conjugated Oligomers. J. Am. Chem. Soc 2020, 142 (10), 4852–4861. [DOI] [PubMed] [Google Scholar]
  • 132.Tsai C-H; Fortney A; Qiu Y; Gil RR; Yaron D; Kowalewski T; Noonan KJ, Conjugated Polymers with Repeated Sequences of Group 16 Heterocycles Synthesized through Catalyst-Transfer Polycondensation. J. Am. Chem. Soc 2016, 138 (21), 6798–6804. [DOI] [PubMed] [Google Scholar]
  • 133.Norris BN; Zhang S; Campbell CM; Auletta JT; Calvo-Marzal P; Hutchison GR; Meyer TY, Sequence matters: modulating electronic and optical properties of conjugated oligomers via tailored sequence. Macromolecules 2013, 46 (4), 1384–1392. [Google Scholar]
  • 134.Longo JM; DiCiccio AM; Coates GW, Poly(propylene succinate): a new polymer stereocomplex. J. Am. Chem. Soc 2014, 136 (45), 15897–900. [DOI] [PubMed] [Google Scholar]
  • 135.Tang X; Chen EYX, Chemical Synthesis of Perfectly Isotactic and High Melting Bacterial Poly(3-hydroxybutyrate) from Bio-Sourced Racemic Cyclic Diolide. Nat. Commun 2018, 9 (1), 2345. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 136.Tang XY; Westlie AH; Watson EM; Chen EYX, Stereosequenced crystalline polyhydroxyalkanoates from diastereomeric monomer mixtures. Science 2019, 366 (6466), 754–+. [DOI] [PubMed] [Google Scholar]
  • 137.Barnes JC; Ehrlich DJ; Gao AX; Leibfarth FA; Jiang Y; Zhou E; Jamison TF; Johnson JA, Iterative exponential growth of stereo-and sequence-controlled polymers. Nat. Chem 2015, 7 (10), 810–815. [DOI] [PubMed] [Google Scholar]
  • 138.Xu J; Shanmugam S; Fu C; Aguey-Zinsou K-F; Boyer C, Selective photoactivation: From a single unit monomer insertion reaction to controlled polymer architectures. J. Am. Chem. Soc 2016, 138 (9), 3094–3106. [DOI] [PubMed] [Google Scholar]
  • 139.Xu J; Fu C; Shanmugam S; Hawker CJ; Moad G; Boyer C, Synthesis of Discrete Oligomers by Sequential PET-RAFT Single-Unit Monomer Insertion. Angew. Chem. Int. Ed 2017, 56 (29), 8376–8383. [DOI] [PubMed] [Google Scholar]
  • 140.Xu J, Single unit monomer insertion: a versatile platform for molecular engineering through radical addition reactions and polymerization. Macromolecules 2019, 52 (23), 9068–9093. [Google Scholar]
  • 141.Niu J; Hili R; Liu DR, Enzyme-free translation of DNA into sequence-defined synthetic polymers structurally unrelated to nucleic acids. Nat. Chem 2013, 5 (4), 282. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 142.Schmidt BV; Fechler N; Falkenhagen J; Lutz J-F, Controlled folding of synthetic polymer chains through the formation of positionable covalent bridges. Nat. Chem 2011, 3 (3), 234–238. [DOI] [PubMed] [Google Scholar]
  • 143.Pinheiro VB; Taylor AI; Cozens C; Abramov M; Renders M; Zhang S; Chaput JC; Wengel J; Peak-Chew S-Y; McLaughlin SH, Synthetic genetic polymers capable of heredity and evolution. Science 2012, 336 (6079), 341–344. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 144.Fahnestock S; Rich A, Ribosome-catalyzed polyester formation. Science 1971, 173 (3994), 340–343. [DOI] [PubMed] [Google Scholar]
  • 145.Katoh T; Suga H, Ribosomal incorporation of consecutive β-amino acids. J. Am. Chem. Soc 2018, 140 (38), 12159–12167. [DOI] [PubMed] [Google Scholar]
  • 146.Huang Y; Wiedmann MM; Suga H, RNA display methods for the discovery of bioactive macrocycles. Chem. Rev 2018, 119 (17), 10360–10391. [DOI] [PubMed] [Google Scholar]
  • 147.Ohta A; Murakami H; Higashimura E; Suga H, Synthesis of polyester by means of genetic code reprogramming. Chem. Biol 2007, 14 (12), 1315–1322. [DOI] [PubMed] [Google Scholar]
  • 148.Melo Czekster C; Robertson WE; Walker AS; Söll D; Schepartz A, In Vivo Biosynthesis of a β-Amino Acid-Containing Protein. J. Am. Chem. Soc 2016, 138 (16), 5194–5197. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 149.Ad O; Hoffman KS; Cairns AG; Featherston AL; Miller SJ; Söll D; Schepartz A, Translation of Diverse Aramid- and 1,3-Dicarbonyl-peptides by Wild Type Ribosomes in Vitro. ACS Cent. Sci 2019, 5 (7), 1289–1294. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 150.Tharp JM; Ad O; Amikura K; Ward FR; Garcia EM; Cate JH; Schepartz A; Söll D, Initiation of Protein Synthesis with Non-Canonical Amino Acids In Vivo. Angew. Chem. Int. Ed 2020, 132 (8), 3146–3150. [DOI] [PubMed] [Google Scholar]
  • 151.Lee J; Schwarz KJ; Moore JS; Jewett MC, Ribosome-mediated polymerization of long chain carbon and cyclic amino acids into peptides in vitro. Nat. Commun 2020, 11 (1), 1–8. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 152.ten Brummelhuis N; Wilke P; Börner HG, Identification of functional peptide sequences to lead the design of precision polymers. Macro. Rapid. Commun 2017, 38 (24), 1700632. [DOI] [PubMed] [Google Scholar]
  • 153.Huang P-S; Boyken SE; Baker D, The coming of age of de novo protein design. Nature 2016, 537 (7620), 320–327. [DOI] [PubMed] [Google Scholar]
  • 154.Paladino A; Marchetti F; Rinaldi S; Colombo G, Protein design: from computer models to artificial intelligence. Wiley Interdiscip. Rev. Comput 2017, 7 (5), e1318. [Google Scholar]
  • 155.Senior AW; Evans R; Jumper J; Kirkpatrick J; Sifre L; Green T; Qin C; Žídek A; Nelson AW; Bridgland A, Improved protein structure prediction using potentials from deep learning. Nature 2020, 577 (7792), 706–710. [DOI] [PubMed] [Google Scholar]
  • 156.Doran TM; Dickson P; Ndungu JM; Ge P; Suponitsky-Kroyter I; An H; Kodadek T, Synthesis and screening of bead-displayed combinatorial libraries. In Methods in Enzymology, Elsevier: 2019; Vol. 622, pp 91–127. [DOI] [PubMed] [Google Scholar]
  • 157.Mutlu H; Lutz JF, Reading polymers: sequencing of natural and synthetic macromolecules. Angew. Chem. Int. Ed 2014, 53 (48), 13010–13019. [DOI] [PubMed] [Google Scholar]

RESOURCES