Synthesis of Modular Brush Polymer-Protein Hybrids using

Here, we demonstrate the synthesis of modular protein-bottlebrush hybrids using .... shows protein ladder (lane 1), positive control alkyne-PEG (20 kD...
1 downloads 0 Views 2MB Size
Subscriber access provided by The University of British Columbia Library

Article

Synthesis of Modular Brush Polymer-Protein Hybrids using Diazotransfer and Copper Click Chemistry Luis A. Navarro, Daniel L. French, and Stefan Zauscher Bioconjugate Chem., Just Accepted Manuscript • DOI: 10.1021/acs.bioconjchem.8b00309 • Publication Date (Web): 12 Jul 2018 Downloaded from http://pubs.acs.org on July 16, 2018

Just Accepted “Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted online prior to technical editing, formatting for publication and author proofing. The American Chemical Society provides “Just Accepted” as a service to the research community to expedite the dissemination of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts appear in full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been fully peer reviewed, but should not be considered the official version of record. They are citable by the Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered to authors. Therefore, the “Just Accepted” Web site may not include all articles that will be published in the journal. After a manuscript is technically edited and formatted, it will be removed from the “Just Accepted” Web site and published as an ASAP article. Note that technical editing may introduce minor changes to the manuscript text and/or graphics which could affect content, and all legal disclaimers and ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors or consequences arising from the use of information contained in these “Just Accepted” manuscripts.

is published by the American Chemical Society. 1155 Sixteenth Street N.W., Washington, DC 20036 Published by American Chemical Society. Copyright © American Chemical Society. However, no copyright claim is made to original U.S. Government works, or works produced by employees of any Commonwealth realm Crown government in the course of their duties.

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

Bioconjugate Chemistry

Synthesis of Modular Brush Polymer-Protein Hybrids using Diazotransfer and Copper Click Chemistry Luis A. Navarro,* Daniel L. French, Stefan Zauscher* Department of Mechanical Engineering and Materials Science, Duke University, 101 Science Drive, Durham, North Carolina 27708, United States KEYWORDS. Elastin-like polypeptide. Poly(oligoethylene glycol methacrylate). Proteoglycan mimics. Bottlebrush. Grafting to. Alkyne terminated polymers. Azide modification.

ABSTRACT. Proteoglycans are important brush-like biomacromolecules which serve a variety of functions in the human body. While protein-bottlebrush hybrids are promising proteoglycan mimics, many challenges still exist to robustly produce such polymers. In this paper, we report the modular synthesis of protein-brush hybrids containing elastin-like polypeptides (ELP) as model proteins by copper-catalyzed azide-alkyne cycloaddition. We exploit the recently discovered imidazole-1-sulfonyl azide (ISA) in a diazotransfer reaction to introduce an Nterminal azide onto an ELP. Next, we use a click reaction to couple the azido-ELP to an alkyneterminated amine-rich polymer followed by a second diazotransfer step to produce an azide-rich backbone that serves as a scaffold. Finally, we used a second click reaction to graft alkyneterminated poly(oligoethylene glycol methacrylate) (POEGMA) bristles to the azide-rich

ACS Paragon Plus Environment

1

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

Page 2 of 39

backbone to produce the final protein-bottlebrush hybrid. We demonstrate the effectiveness of this synthetic path at each step through careful characterization with 1H-NMR, FTIR, GPC, and diagnostic test reactions on SDS-PAGE. Final reaction products could be consistently obtained for a variety of different molecular weight backbones with final total grafting efficiencies around 70 %. The high-yielding reactions employed in this highly modular approach allow for the synthesis of protein-bottlebrush hybrids with different proteins and brush polymers. Additionally, the mild reaction conditions used have the potential to avoid damage to proteins during synthesis.

Introduction Proteoglycans are a class of biomacromolecules with important functional roles in a variety of human tissues. They serve key roles in maintaining osmotic pressure in cartilage,1 protecting cartilage from cell overgrowth and degradation,2-6 protecting the respiratory tract,7 lubricating eyes,8 and protecting the epithelium.9 The functions of these brush-like macromolecules critically depend on the oligosaccharide chains that are densely grafted to a polypeptide backbone. For example, stripping a single saccharide unit from proteoglycan 4’s side chains dramatically reduces its boundary lubrication ability.10 Despite the interest in designing new proteoglycans, their synthesis typically requires mammalian cells to achieve a large degree of post-translational modification, making it challenging to generate authentic proteoglycans in large quantities.11-13 In spite of this, researchers have been able to take major structural inspiration from various proteoglycans to produce functional biomimetic macromolecules. For example, Banquy et al. made a purely synthetic proteoglycan mimic that served as a boundary lubricant on mica,14 while Lawrence et al. showed that a purely biological mimic reduces friction

ACS Paragon Plus Environment

2

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

Bioconjugate Chemistry

between mated cartilage surfaces.15 This growing body of work is indicative of a rising demand for ways to produce heavily hydrated brush-like structures that mimic proteoglycans.6, 16, 17 Although bottlebrush polymers have attracted a large amount of attention due to their ability to mimic proteoglycans, they additionally offer some unique advantages over linear polymers. Where linear polymers have a persistence length which is fixed by the choice of monomer, bottlebrushes of the same chemical composition can exhibit different persistence lengths by varying the lengths or grafting density of the bristles.18-22 Furthermore, bottlebrushes can be used to make extremely soft structural materials.23 In conjunction with existing proteoglycan work, there is thus great interest in making hybrids with structural proteins and synthetic brush polymers. Recently, click chemistry has revolutionized the field of bioconjugation.24 Specifically, copper-catalyzed azide-alkyne cycloaddition has gained popularity in the bioconjugate space due to its high conjugation efficiency, fast reaction times, compatibility with aqueous media, and high functional group tolerance.25 These click reactions are especially effective for the preparation of bottlebrushes where they enable high grafting densities.26-28 Issues in these reactions generally arise from the incorporation of functional azide or alkyne groups into synthetic polymer backbones. Most notably, alkynes are vulnerable to chain transfer,29 polymerization through the triple bond,30 and Glaser coupling.31 Azides suffer from side reactions especially at elevated temperatures that allow either cyclization of azide with monomer alkenes32 or decomposition with evolution of nitrogen.30 One popular approach for alkyne incorporation is to either use protected alkynes33-35 or introduce alkyne functional groups into polymers after polymerization. For example, Gao used carbodiimide chemistry to couple alkynes to poly(2-hydroxyethyl methacrylate) (PHEMA).36 Azides are typically included in monomers32,

ACS Paragon Plus Environment

3

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

37, 38

Page 4 of 39

or incorporated by nucleophilic substitution using sodium azide.31, 35 Azides can also be

incorporated by diazotransfer, which converts primary amines to azides. However, to date, diazotransfer has been a far less popular choice due to the extremely hazardous and harsh in situ preparation of the typically employed tosyl azide in dichloromethane.39,

40

Recently,

diazotransfer has become much more popular due to the discovery of imidazole-1-sulfonyl azide salts (ISA), which are the first shelf-stable diazotransfer reagents usable in purely aqueous media.40, 41 ISA can be used to functionalize amines in polypeptide N-termini and lysine side chains.42, 43 This development opens new opportunities for the synthesis of protein-brush hybrids. Here, we demonstrate the synthesis of modular protein-bottlebrush hybrids using copperassisted click chemistry. The critical advance in this paper is the way in which we exploit recent advances in diazotransfer to generate these hybrids. Specifically, we use diazotransfer to prepare an azide-containing protein, and use that azide group to conjugate it to an alkyne-terminated, amine-rich polymer backbone (Scheme 1). We then use diazotransfer to convert the amines on the backbone to azides, and then graft alkyne-terminated polymer bristles. Notably, the use of ISA allows these transformations to be carried out in aqueous media, which is essential for handling a wide variety of proteins. Furthermore, the use of click reactions allows for the modular assembly of the brush with high conversion. To demonstrate the modularity and reproducibility of this approach, we carried out the synthesis with three different polymer backbone lengths (see Table 1). To discern the effect of the protein block on the physical properties of the final products, we compare our protein-brush hybrids to bottlebrush polymers lacking the terminal protein.

ACS Paragon Plus Environment

4

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

Bioconjugate Chemistry

≡-

NH2 NH2 NH2

NH2

-N3 ELP

NH2 NH2 NH2

NH2

ISA N3 N3

N3 N3

≡POEGMA

ELP-(PAMA-g-POEGMA) Brush

Scheme 1. Modular synthesis of ELP-brush polymer hybrid by sequential copper click and diazotransfer reactions. HEMA and fluorescein-O-methacrylate comonomers present in the amine-rich polymer have been omitted for clarity.

ACS Paragon Plus Environment

5

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

Page 6 of 39

Table 1. Molecular weight statistics of polymers employed in this study. GPCb Light Scatteringa Sample Mw (kDa) Rg (nm) Rh (nm) Mn (kDa) PDI x c x c x d 40.06 28.9 19.43 1.555d Alkyne-POEGMA P(azido-AMA-co-HEMA) 30 38.46 12.6 7.5 30.70 1.264 P(azido-AMA-co-HEMA) 92 83.48 19.3 12.3 92.57 1.588 P(azido-AMA-co-HEMA)101 127.7 20.0 11.5 101.01x 1.389 e d NMR GPC Brush Samplef Bristle Xn Rho (%) Mn (kDa) Mn (kDa) PDI TIPS-Brush( 30 kDa) 42.8 68.3 543.9 237.2 1.19 TIPS-Brush( 92 kDa) 21.8 70.1 596.5 690.2 1.52 TIPS-Brush(101 kDa) 26.2 63.6 924.1 803.9 1.31 ELP -Brush( 30 kDa) 19.6 96.5 381.9 1013x.x 1.24 ELP -Brush( 92 kDa) 26.7 75.4 778.1 608.6 1.35 ELP -Brush(101 kDa) 26.7 71.7 1063x.x 709.1 1.31 a Light scattering in 1 M acetic acid, 0.3 M NaH2PO4 (pH 3.3) at 25 oC prior to diazotransfer. b GPC (MALS) with DMF (10 mM LiBr) mobile phase after diazotransfer. c Light scattering in PBS at 25 oC. d GPC (MALS) with THF. e 1H NMR in pyridine-d5 at 500 MHz with statistics determined by integrals for triazole (8.4 ppm), PHEMA methylene (4.1 ppm), and POEGMA methoxy (3.4 ppm). For reference, bristle Xn determined by THF GPC of precursors was 64.7. f Brush samples are P((AMA-g-POEGMA)-co-HEMA) where the number in parentheses corresponds to the molar mass of the P(AMA-co-HEMA) backbone polymer.

Results and Discussion Conjugate Design We chose to use poly(2-aminoethyl methacrylate-co-2-hydroxyethyl methacrylate) (P(AMAco-HEMA)) as a polymer backbone. The amine-containing monomers serve as handles to introduce azide groups for later grafting of alkyne-terminated polymer bristles. PHEMA was copolymerized into the backbone to serve as a spacer between grafting points and to determine grafting efficiency in the final product. Poly(oligoethyleneglycol methacrylate) (POEGMA) was chosen as a bristle polymer due to its hydrophilicity, relevance as a biocompatibilizer in biotechnology, compatibility with radical polymerization, and ability to be handled and

ACS Paragon Plus Environment

6

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

Bioconjugate Chemistry

characterized in both organic and aqueous media. We used triisopropylsilyl (TIPS) protecting groups on the backbone alkyne to prevent potential side reactions during polymerization. We chose an elastin-like polypeptide (ELP) to serve as a model protein for conjugation due to its lack of secondary structure, lack of extraneous amines, ease of purification, and repetitive nature which allowed for simpler characterization. Without loss of generality, the ELP could be replaced with another protein. ELP Synthesis Here, we chose to use ELP (sequence (GVGVP)30 with a C-terminal collagen affinity tag (WYRGRL),15 for further detail in Table S1) because it is an intrinsically disordered model protein for its ease of purification, lack of susceptibility to denaturation in organic solvents, lack of lysine residues, and repetitive nature producing relatively simple NMR spectra. We introduced azide functionality into the ELP by diazotransfer with ISA. Because the entire protein lacks lysines, the only amine available for diazotransfer is that on the N-terminus. This allows us to controllably add no more than one brush polymer to each protein. After diazotransfer, a combination of 1H-NMR, IR, MALDI-TOF, ESI-TOF, and diagnostic SDS-PAGE were used to confirm the purity of the protein and introduction of an azide. Diagnostic gels show reaction with Cy3-alkyne after diazotransfer (Figure S31), thus confirming the introduction of at least one azide group into the protein. Cyanine labeling further reveals the presence of ELP-V30 (i.e., ELP lacking the collagen binding tag) impurity in the ELP sample. ELP-V30 impurities are nearly invisible through routine Coomassie staining because the only available positive charge is on the N-terminus, while the collagen binding tag contains several positively charged groups. Cyanine labeling also allowed for accurate quantification of ELP-V30 impurities.

ACS Paragon Plus Environment

7

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

Page 8 of 39

) y3

(-C A A M M G G r 3 de EG 3-N OE OE d P P P y La ≡- C ≡- ≡-

75k

25k

Figure 1. Cyanine3 (Cy3) scan of a diagnostic SDS-PAGE of alkyne-POEGMA bristles. The gel shows protein ladder (lane 1), positive control alkyne-PEG (20 kDa) reacted with Cy3-azide (lane 2), unreacted Cy3-azide dye (lane 3), alkyne-POEGMA (19 kDa) without dye (lane 4), and alkyne-POEGMA reacted with Cy3-azide (lane 5). Bristle Synthesis We used AGET ATRP to synthesize alkyne-terminated POEGMA bristles (see Supporting Information). Their chemical composition and purity were verified by NMR and IR. GPC traces show a bimodal molecular weight distribution with a second peak containing less polymer. For the bristle synthesis, we used unprotected alkyne initiators to obtain enough material to produce brush polymers with a large excess of bristles. However, the use of unprotected alkynes also led to undesirable side reactions, most notably Glaser coupling, which negates the ability of terminal alkynes to react. We know this because the same reaction conditions performed on a small scale with TIPS-protected alkyne initiators showed unimodal distributions (data not shown). A

ACS Paragon Plus Environment

8

Page 9 of 39

diagnostic gel reveals a successful reaction between Cy3-N3 and alkyne-POEGMA bristles,

A)

La dd PE er G PE - ≡ G EL -N3 P≡- N3 PA M A

demonstrating the presence of active alkyne groups after polymerization (Figure 1).

B)

75k

Cy3

75k

Fluorescein

25k

75k 25k

75k 25k

Lane:

ELP-PAMA

Cy3-X N3 ≡ ≡ N 3 N3 ≡ N 3 ≡ N 3 ≡ MW (kDa) 20 10 17 30 30 30 92 92 101101

Coomassie

Fluorescein

Cy3

AgF - + - + + + Cy3-N3 - + + + + + + MW (kDa) 101 101 30 92 101 30 92 101

Coomassie

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

Bioconjugate Chemistry

1

2

3

4

5

6

7

8

9

25k

75k 25k

75k 25k

Lane: 1

10

2

3

4

5

6

7

8

9 10 11

Figure 2. Diagnostic SDS-PAGE to test alkyne activity in different molecular weight P(AMAco-HEMA) by reaction with Cy3 dyes and visualization by Cy3 (top), fluorescein (middle), and Coomassie (bottom). For reference, gels show protein ladder (lane 1), positive control alkynePEG (20 kDa, lane 2), azido-PEG (10 kDa, gel B, lane 3), and azido-ELP (gel B, lane 4). (A) Gel A shows polymers before (AgF –) and after (AgF +) AgF deprotection by reaction with (Cy3 +) or without (Cy3 –) Cy3-azide (color overlay available in Figure S65). (B) Gel B shows deprotected P(AMA-co-HEMA) before (lane 5) and after (lanes 6-11) conjugation with azidoELP and reacted with Cy3-azide (N3) or Cy3-alkyne (≡).

Backbone Design AGET ATRP proved to be effective for producing the high molecular weight backbone polymers (see Supporting Information). We note that switching from PMDETA to the stronger, pyridyl-based ligand, TPMA, allowed the polymerization to reach high conversion while following first order rate kinetics. Backbone polymer chemical composition and purity were determined by 1H NMR and IR, with

1

H NMR additionally revealing comonomer

ACS Paragon Plus Environment

9

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

Page 10 of 39

incorporation ratios that closely matched the monomer feed ratios in every batch, independent of conversion or molecular weight. This observation is further supported by research by QuiñonezAngulo et al. who showed that the kinetics of P(DEAEAM-co-OEGMA) random copolymers proceed independent of the ratio of OEGMA to amine-based monomer.44 Therefore, we suspect that, despite the charged monomers, the polymers produced with this method are truly random copolymers. This is important for the production of brush polymers with roughly uniform bristle spacing. In some cases, alkyne headgroup activity in smaller polymers can be determined directly by 1

H NMR.29, 33 However, such direct approaches fail for larger polymers where the mass of the

polymer dwarfs signals from headgroups. Verification of the activity of single alkynes or azides is thus typically performed directly by mass spectrometry42, 43, 45, 46 and/or indirectly by reaction with an azide followed by mass spectrometry,33, 45 GPC,47 or fluorescent detection.48 However, polydisperse polyelectrolytes yield exceedingly complex mass spectra, and typically have major issues in chromatographic separation due to complications arising from interactions between the stationary phase, polymers, and their counterions. Furthermore, neutralization of PAMA-based polymers causes degradation through intramolecular and intermolecular reactions with the amine,49 potentially complicating analysis dramatically. We overcame these obstacles by using a diagnostic click reaction in which we labeled the reactive headgroups with cyanine dyes (Figure 2). Fortunately, the polycations employed in this work migrate well in SDS-PAGE, enabling us to use this technique to separate the reaction products. The inclusion of a different fluorophore (0.5 mol% fluorescein O-methacrylate) in the backbone copolymers enabled the fluorescent detection of polymer chains independent of their reaction with cyanine dye, allowing us to verify that both signals are produced by the same species. Additionally, we discovered that these

ACS Paragon Plus Environment

10

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

Bioconjugate Chemistry

polycations could be intensely stained with Coomassie brilliant blue, a dye that is routinely used to stain proteins through interactions with positively charged groups. Coomassie staining effectively adds an orthogonal, redundant check for visualizing these migrating polymers. These diagnostic click reactions on SDS-PAGE were paired with positive controls (PEG-azide or PEGalkyne) to verify that the test reactions were working. Similarly, negative controls (polymer without dye) were included to ensure cyanine signals came only from reactive dyes and not partial fluorescein excitation. This approach allowed us to simultaneously screen for functional azides and/or alkyne groups in the desired polymers as well as unreacted starting material impurities. Diagnostic gels of TIPS-protected P(AMA-co-HEMA) with Cy3-N3 reveal that a small amount of TIPS protecting group was removed during polymerization, which is known to happen to some degree in the presence of bases.33 However, after deprotection with AgF, backbone polymers show much more reaction with cyanine dyes (Figure 2A). Despite challenges in quantitatively determining the degree of deprotection, this result indicates that AgF successfully deprotects these alkynes. The position of the cyanine signal matches the position of fluorescein and Coomassie signals in the same gel, and properly correlates to less migration for higher molecular weight polymers. Alkyne-terminated backbones were then conjugated to azido-ELP with a large excess of protein to ensure minimal residual unfunctionalized polymer. SDS-PAGE revealed that no high molecular weight polycations reacted with Cy3-N3 following conjugation (Figure 2B), implying that nearly all available alkyne head groups were consumed. The gel also revealed that a significant amount of unreacted protein impurity is present in the final products. At this step, the majority of ELP was removed by precipitation, while the remaining protein was later removed

ACS Paragon Plus Environment

11

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

Page 12 of 39

after the second diazotransfer reaction. We also noted that a small amount of conjugated polycation unexpectedly reacted with Cy3-alkyne in lanes 7, 9, and 11 (Figure 2B). Upon closer inspection, the ratio of fluorescein to cyanine signal intensity of these bands increases with migration, whereas the alkyne-P(AMA-co-HEMA) precursor has a constant ratio of fluorescein to cyanine intensity (Figure 2A). Specifically, it seems as though the conjugate contains a mixture of Cy3-active, lower mobility species and a Cy3-inactive, higher mobility species. We therefore hypothesize that the proteins employed potentially contained multiple azides, allowing some proteins to react with both polymer and dye. Furthermore, the presence of the Cy3inactive, high mobility fluorescein signal indicates the potential degradation of the polymers’ alkyne head groups. Backbone Diazotransfer The presence of a strong peak at 2100 cm-1 in polymer IR spectra confirms the successful incorporation of azide groups into the amine-rich polymers. No unreacted amines are visible in 1

H-NMR (DMSO-d6), although trace acetate ion and/or acetic acid is visible (>87 % conversion).

Unfortunately, low water solubility made SDS-PAGE diagnostic analysis unreliable. Brush Polymer Synthesis We chose to use copper-assisted click reactions for polymer brush assembly to achieve high grafting efficiency26,

36

and to avoid reactive intermediates present in many nucleophilic

substitution-based approaches.27 Furthermore, it is known that the triazole products of brush polymers in “grafting to” approaches serve as copper ligands which autocatalyze the coupling reaction.50 Copper-based click reactions are robust against a wide variety of functional groups which may be found in proteins or synthetic polymers, making it suitable for this modular synthesis.

ACS Paragon Plus Environment

12

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

Bioconjugate Chemistry

To detect the presence of remaining bristles after purification, the brush polymers were passed through GPC with a THF mobile phase (Figure 3). Chromatograms clearly show longer retention times for brush polymers with longer backbone lengths. This seemingly contradictory trend has been previously reported for other brush polymers.51 Furthermore, brush polymer samples with short peptide units showed consistently shorter retention times than their peptidefree counterparts. The observed one minute reduction in retention time roughly corresponds to a threefold difference in the molecular weight for a linear POEGMA based polymer on this column (data not shown). Therefore, we argue that this large difference in retention time can only come from a conformational change in the brush polymer to accommodate the extremely THF-insoluble protein with the hydrophilic POEGMA side chains. 1

H-NMR of the brush polymers in pyridine proved to be the most effective way to determine

brush polymer molecular weight, grafting density, and bristle length (see Table 1). Specifically, pyridine substantially shifts protons near alcohol groups downfield relative to their positions in more common NMR solvents, which allowed us to effectively separate peaks for our PHEMArich backbone from those of the POEGMA bristles. In this medium, separate peaks were observed for the POEGMA terminal methoxy group (3.3 ppm), triazole head group (8.4 ppm), and backbone PHEMA CH2 (4.1 ppm), which allowed for determination of the effective grafted bristle lengths and grafting efficiency. NMR spectra clearly showed incorporated bristle lengths that are consistently shorter than those determined by GPC of the precursors. Therefore, it is plausible this grafting process proceeds with a strong preference for the shortest available chains. Furthermore, the apparent grafting efficiencies consistently reached 70 % despite the bulky nature of the POEGMA bristles. This figure includes the reaction efficiency of the diazotransfer and click reactions. However, overlap between then PHEMA CH2 and POEGMA CH2 peaks in

ACS Paragon Plus Environment

13

Bioconjugate Chemistry

NMR causes this figure to be a slight underestimate. We note that light scattering proved to be quite limited for molecular weight determination of the brush polymers, consistent with previous reports on other bottlebrushes (see also Supporting Information).51 1.5

1

Light Scattering Chromatogram

B)

Headgroup Backbone TIPS ELP 30 kDa 92 kDa x x 101 kDa POEGMA

0.5

Backbone

1

Headgroup TIPS ELP

30 kDa 92 kDa x 101 kDa POEGMA

x

0.5

0

0 10

Refractive Index Chromatogram

1.5

Relative Scale

A) Relative Scale

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

Page 14 of 39

12

14

16

18

10

12

14

16

18

Time (min)

Time (min)

Figure 3. THF GPC chromatograms of P((AMA-g-POEGMA)-co-HEMA) brush polymers with 30 kDa (black), 92 kDa (red), and 101 kDa (blue) backbone lengths conjugated to ELP (dash) or not (solid). Unreacted POEGMA bristle (black dot) is shown for reference. Chromatograms are shown for detection by A) light scattering and B) refractive index.

Figure 4. Representative tapping mode AFM images in air of (left) TIPS-P((AMA-gPOEGMA)-co-HEMA) and (right) ELP-P((AMA-g-POEGMA)-co-HEMA) spin coated onto

ACS Paragon Plus Environment

14

Page 15 of 39

freshly cleaved mica. Apparent bottlebrush polymer contour lengths were 87 +/- 30 nm (40 brushes) with protein tag and 102 +/- 33 nm (17 brushes) without protein tag (t-test p = 0.123) when averaged across several AFM images. The brush diameters for both bottlebrush polymers were approximately 17 nm. To confirm the final bottlebrush architecture, we imaged brush polymer on mica surfaces by tapping mode atomic force microscopy (AFM) in air (Figure 4). AFM images show bottlebrushlike objects of comparable lengths (87.2 +/- 30.4 nm with ELP, and 102.2 +/- 33.4 nm without ELP) and diameters (~17 nm for both). The apparent dimensions are consistent with the estimated contour lengths of the backbone (179 nm) and bristles (16 nm), which set upper bounds for the dimensions of the polymers at maximum extension.

Cy3

Cy3-≡ Biotin-Cy3

Coomassie

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

Bioconjugate Chemistry

-

Strep - + + -

Strep-N3 Strep-PAMA - + - + - + - + -

75k 25k

75k 25k

Lane:

1

2

3

4

5

6

7

8

9

10

Figure 5. Diagnostic SDS-PAGE to test azide activity in different streptavidin derivatives. The gel shows native streptavidin (lanes 2-4), azido-streptavidin (lanes 5-7), and streptavidinP(AMA-co-HEMA) hybrid (lanes 8-10) reacted with Cy3-alkyne (lanes 4, 7, and 10), biotin-Cy3 (lanes 3, 6, and 9), or with no additive (lanes 2, 5, and 8). A ladder is included for reference (lane 1). Scans for Cy3 and Coomassie are shown. For color overlay and fluorescein scan, see Figure S76.

Synthesis of a Biohybrid with a Folded Protein Certain applications require biohybrid constructs in which the protein secondary/tertiary structure must be preserved. The biggest obstacle inherent to extending our approach to folded

ACS Paragon Plus Environment

15

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

Page 16 of 39

proteins is its potential denaturation due to interactions with the polycationic AMA-rich backbone, as the use of organic solvents can be avoided with different choices of filler comonomers and bristle polymers (for an extended discussion see Supporting Information). To demonstrate that AMA-based copolymers are not intrinsically destructive to folded protein structures, we show the synthesis of a functional streptavidin-P(AMA-co-HEMA) conjugate by diazotransfer and Huisgen cycloaddition (Scheme 1). The gel in Figure 5 shows the introduction of azides into streptavidin following diazotransfer (lane 7) and the consumption of some of those azides following conjugation to alkyne-P(AMA-co-HEMA) (lane 10). Importantly, streptavidinP(AMA-co-HEMA) retains its biotin-binding ability, as visualized by reaction with Cy3-labeled biotin (lanes 3, 6, and 9). This suggests that our AMA-based copolymers can be compatible with functional proteins.

Conclusions We have demonstrated that modern diazotransfer reagents (specifically, ISA) can be exploited in new ways to produce complex, branched protein-polymer hybrids, such as the proteinbottlebrushes shown here. We employ mild reaction conditions that are favorable for the preparation of hybrids because the protein only needs to be present during two diazotransfer reactions and two click reactions, both of which can be performed at room temperature in buffered aqueous media. Furthermore, our “grafting to” approach allows synthetic polymers to be separately produced in bulk, handled with organic solvents, and characterized for their molecular weight distributions. This is valuable in situations where the protein of interest is precious, as the molecular weight distribution of the final products can be estimated a priori. We showed that our amine-rich, synthetic polymer backbones are amenable to characterization by SDS-PAGE using fluorescent detection and staining procedures that are routine in protein

ACS Paragon Plus Environment

16

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

Bioconjugate Chemistry

analysis. Diazotransfer to these polymers reliably yielded azide-rich backbones suitable for the preparation of brush polymers. Furthermore, our click-based grafting of POEGMA to backbone azides consistently produced high grafting densities despite our use of sterically encumbering bristles. Finally, our synthetic strategy can be compatible with folded proteins, as shown by the synthesis of a functional streptavidin-P(AMA-co-HEMA) hybrid. The synthetic approach taken in this paper paves the way for the design of a general class of highly branched biohybrids, such as proteoglycan mimics. While we showed successful coupled to an ELP N-terminus, our approach extends to any protein N-terminal amine, protein lysine residue, DNA nucleobase excepting thymidine,52 amine-bearing polysaccharide, or existing azide modification. Furthermore, our procedures can be easily adapted to prepare hybrids with different bristle chemistries, including hydrophilic synthetic polymers, proteins, and glycopolymers. Lastly, by using alkyne-P(AMA-co-HEMA) in place of the bristles, more highly branched architectures should be accessible.

Experimental Procedure Materials Acetic acid (glacial, 99.5 %, Mallinckrodt), acetone (99.5 %, Sigma-Aldrich), acetonitrile (99.99 %, EM Science), N-acetyl-L-cysteine (99 %, Sigma-Aldrich), alumina (basic, 150 mesh, Sigma-Aldrich), 2-aminoethyl hydrochloride (99 %, Sigma-Aldrich), ammonium chloride (99.5 %, Sigma-Aldrich), ammonium sulfate (99 %, MP Biomedicals), L-ascorbic acid (99 %, Sigma-Aldrich), azide-PEG3-biotin (444.6 Da, 89.5 %, Sigma-Aldrich), biotin (99 %, Sigma-Aldrich), biotin-labeled bovine serum albumin (biotin-BSA, 80 % protein, 8-16 biotin per protein, Sigma-Aldrich), α -bromoisobutyryl bromide (98 %, Sigma-Aldrich), 3-bromo-1propanol (97 %, Sigma-Aldrich), chloroform-d (CDCl3, 99.8 % D, Sigma-Aldrich), copper (II) bromide (CuBr2, anhydrous, 99 %, Acros), copper (II) chloride (CuCl2, 99.999 %, Sigma-

ACS Paragon Plus Environment

17

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

Page 18 of 39

Aldrich), copper (II) sulfate pentahydrate (CuSO4 ・ 5H2O, 98.9 %, Fisher Chemicals), cyanaine3-alkyne (Cy3-alkyne, Lumiprobe), cyanine3-azide (Cy3-azide, 10 mM in DMSO, Lumiprobe), deuterium oxide (D2O, 99.9 % D, Sigma-Aldrich), dichloromethane (DCM, 99.5 %, Alfa Aesar), diethyl ether (anhydrous, 99.7 %, Sigma-Aldrich), N,N-dimethylformamide (DMF, HPLC grade, 99.9 %, Honeywell Research Chemicals), dimethylsulfoxide (DMSO, 99.99 %, Mallinckrodt/Covidien), dimethylsulfoxide-d6 (DMSO-d6, 99.9 % D, Cambridge Isotope Labs), ethanol (EtOH, absolute, KOPTEC), ethyl acetate (99 %, Fisher Chemicals), ethyl magnesium bromide (3.0 M in ether, Sigma-Aldrich), ethylenediaminetetraacetic acid disodium salt (Na2EDTA, 99 %, Sigma-Aldrich), fluorescein O-methacrylate (97 %, Sigma-Aldrich), glycerol (99.5 %, EM Science), heptane (99 % Aldrich), hexane (98.5 %, BDH, VWR Analytical), hydrochloric acid (37 % solution, Mallinkrodt), hydroquinone (99.5 %, Sigma-Aldrich), (4-(2hydroxyethyl)-1-piperazineethanesulfonic acid) (HEPES, 99.5 %, Sigma Life Science), (4-(2hydroxyethyl)-1-piperazineethanesulfonic acid) sodium salt (HEPES, 99.5 %, Sigma), imidazole (Mallinckrodt), isopropyl alcohol (99.5 %, BDH, VWR Analytical), Laemli buffer (2x, BioRad), lithium bromide (LiBr, anhydrous, 99.995 %, Beantown Chemical), magnesium sulfate (MgSO4, anhydrous, 99 %, Mallinckrodt), methacryloyl chloride (97 %, Sigma-Aldrich), methanol

(MeOH,

99.8

%,

EMD

Millipore

Corporation),

N,N,N’,N’’,N’’-

pentamethyldiethylenetriamine (PMDETA, 99 % Sigma-Aldrich), polyethyleneglycol methyl ether alkyne (PEG-alkyne, Laysan Bio, Inc., MW 20,000, Lot # 134-165), polyethyleneglycol methyl ether azide (PEG-azide, Laysan Bio, Inc., MW 10,000, Lot # 134-77), potassium chloride (99.995 %, Alfa Aesar), potassium permanganate (99 %, Sigma-Aldrich), potassium phosphate monobasic (99.8 %, J.T. Baker), propargyl alcohol (99 %, Sigma-Aldrich), pyridine-d5 (99.5 % D, Cambridge Isotope Labs), silica gel (100-200 mesh, Fisher Chemicals), silver (I) fluoride

ACS Paragon Plus Environment

18

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

Bioconjugate Chemistry

(AgF, 99.9 %, Sigma-Aldrich), sodium azide (99 %, EMD Millipore), sodium bicarbonate (99.5 %, Sigma-Aldrich), sodium chloride (99 %, Macron Chemicals), sodium hydroxide (98.2 %, Fisher Chemicals), sodium L-ascorbate (99 %, Sigma-Aldrich), sodium nitrite (NaNO2, 97 %, Sigma-Aldrich), sodium phosphate dibasic (99 %, Sigma-Aldrich), streptavidin (from Streptomyces avidinii, 13 U/mg, 65-100 % protein, Sigma-Aldrich), sulfuric acid (95 %, EM Science), sulfuryl chloride (SO2Cl2, 98.5 %, Fischer Scientific), tert-butyl alcohol (tBuOH, 99 %, Alfa Aesar), tetrahydrofuran (THF, anhydrous, 99.9 %, Sigma-Aldrich), triethylamine (NEt3, 99 %, Sigma-Aldrich), trifluoroacetic acid (TFA, 99 %, Sigma-Aldrich), triisopropylsilyl chloride (TIPS-Cl, 97 %, Sigma-Aldrich), tripropargylamine (98 %, Sigma-Aldrich), tris(2pyridylmethyl)amine

(TPMA,

98

%,

Sigma-Aldrich),

and

tris(3-

hydroxypropyltriazolylmethyl)amine (THPTA, 95 %, Sigma-Aldrich) were purchased and used as received. Potassium carbonate (anhydrous, 99 %, Fluka) was heated in an oven for several hours and allowed to cool in a desiccator before use to ensure accurate stoichiometry. 2-Hydroxyethyl methacrylate (HEMA, 99 %, Sigma-Aldrich) and poly(ethylene glycol) methyl ether methacrylate (OEGMA, Mn = 300, Sigma-Aldrich) were passed through basic alumina to remove inhibitor and used within three hours prior to any polymerization. While technical grade 2-aminoethyl methacrylate hydrochloride (AMA-HCl) is commercially available, its low 90 % purity prohibits its use in this synthesis. AMA-HCl was thus prepared according to literature procedures.49 Diagnostic click reactions for gels in Figures 1 and 2A were conducted using THPTA purchased from Sigma Aldrich. THPTA used for the gel shown in Figure 2B, conjugation reactions, and grafting reactions was prepared on a large scale synthetically. Initiators, polymers, proteins, ISA, buffers, and staining solutions were prepared with full synthesis procedures available in Supporting Information.

ACS Paragon Plus Environment

19

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

Page 20 of 39

For biosynthesis, AcuI (New England Biolabs, Ipswich, MA), agar (Ultrapure, Affymetrix), agarose (molecular biology grade, Bio-Rad), aldolase (Sigma-Aldrich), BamHI (as BamHI-HF, New England Biolabs), BglI (New England Biolabs), BL21 (DE3) competent E. coli (New England Biolabs), BseRI (New England Biolabs), calf intestinal phosphatase (CIP, New England Biolabs), Cutsmart Buffer (New England Biolabs), isopropyl β-D-1-thiogalactopyranoside (IPTG, 99 %, Gold Biotechnology), kanamycin (USP grade, Gold Biotechnology), NEB10β competent E. coli (New England Biolabs), polyethyleneimine (PEI, 50 % aqueous solution, 50100 kDa, MP Biomedicals), QIAprep Spin Miniprep Kit (Qiagen, Valencia, CA), QIAquick Gel Extraction Kit (Qiagen), Quick Ligase (New England Biolabs), Quick Ligase Buffer (New England Biolabs), SDS-Tris-Glycine Buffer (Bio-Rad), SybrSafe (Invitrogen), Terrific Broth II (MO Bio), tris-acetate-EDTA buffer (TAE, Bio-Rad), trypsin (Trypsin-ultra, mass spectrometry grade, New England Biolabs), Trypsin-ultra reaction buffer (2x concentrated, 50 mM tris HCl, 20 mM CaCl2, pH 8, New England Biolabs), and XbaI (New England Biolabs) were purchased and used as received. Forward and reverse oligonucleotides (Ultramer) encoding for a collagen-binding domain (WYRGRL) and flexible linker (G4S)3 were ordered from Integrated DNA Technologies. The plasmid, JMD2, was donated from the Chilkoti Laboratory (Duke University, Durham, NC) JMD2 was adapted from pET-24 via previously described procedures.53 A JMD2 plasmid containing a gene for (VPGVG)30 (abbreviated ELP-V30) was supplied by Prof. Ashutosh Chilkoti’s laboratory (Duke University), and was produced through the same procedures.53 Full protein and DNA sequence information is available in Table S1. All water used was Milli-Q grade and obtained using a NANOpure Diamond water purification system (Barnstead, 18.2 MΩ·cm) and filtered through a 0.22 µm Millipak 40

ACS Paragon Plus Environment

20

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

Bioconjugate Chemistry

Gamma Gold (Millipore) filter. Regenerated cellulose dialysis membranes (Spectra/Por, molecular weight cut-off 8-10 kDa and 25 kDa) were purchased from VWR and used as received. Regenerated cellulose dialysis membranes (Snakeskin, molecular weight cutoff 3.5 kDa) were purchased from Thermo Scientific and used as received. Instrumentation: Chemical Characterization: NMR spectra were obtained using Bruker 500 MHz and Bruker 400 MHz spectrometers at 25 oC with VnmrJ version 3.2. NMR data was analyzed using MestReNova (v8.0.1-10878). Fourier transform infrared-attenuated total reflection (FTIR-ATR) (500-4000 cm-1) spectra were obtained using an ABB FTLA2000 spectrometer and processed using GRAMS/AI (Thermo Galactic, v7.00). Direct injection mass spectra were obtained using an Agilent Technologies 1100 series LC/MSD Trap with electrospray ionization. Mass spectra were analyzed using DataAnalysis for LC/MSD Trap (v3.3). Mass spectroscopy samples were loaded by direct injection using a Harvard Apparatus 20 syringe pump. LCMS was obtained with an Agilent 1100 series HPLC with capillary pump (G1376A), degasser (G1379A), microALS (G1389A), ALS Therm (G1330A), and column compartment (G1316A). The HPLC fed into a Luna C18(2) column (Phenomenex, 3 µm, 100 Å, 50x1 mm) with detection of UV absorbance by a diode array detector (DAD, G1315B). Mass spectrometry data for LCMS was recorded and processed with the same hardware and software as directly injected samples. Melting points were obtained manually using a heated oil bath. Liquid densities were obtained gravimetrically using a 1 mL volumetric flask. Biological Characterization: Matrix-assisted laser desorption/ionization with time-of-flight detection (MALDI-TOF) was performed using a Bruker AutoflexSpeed MALDI Mass Spectrometer (Billerica, MA) calibrated with an aldolase standard. MALDI samples were

ACS Paragon Plus Environment

21

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

Page 22 of 39

prepared by mixing 5 µL analyte (10 µM in H2O) with 5 µL sinapinic acid (saturated solution in 69.9 % water, 30 % acetonitrile, 0.1 % trifluoroacetic acid), and spotting the mixture onto MALDI sample plates. ESI-TOF was taken using an Agilent 6224 TOF LC/MS system with a dual ESI source and a 1200 HPLC, using Agilent MassHunter software. When applicable, C18 ZipTip pipette tips (Millipore-Sigma) were used to desalt samples. DNA sequencing was performed by submitting plasmid samples to Eton Bioscience (Durham, NC branch), and sequencing data was analyzed using SeqBuilder (v14.1, DNASTAR). DNA sequencing chromatograms were analyzed by FinchTV (v1.4) to ensure data quality. Ion Exchange Chromatography. Protein purification was carried out with an AKTApurifier system (GE Healthcare) equipped with modules P-960, PV-908, PV-907, NV-907, and M-925. Samples were separated by cation exchange columns (three HI-Trap SPFF columns in series) with 300-600 µL injections of roughly 40 mg/mL protein. Separation was monitored using an inline UV-detector (UPC-900) and analyzed by Unicorn (v5.1). Gel Permeation Chromatography: Gel permeation chromatography (GPC) with THF eluent (HPLC grade, 100 ppm BHT) was performed on 2 Agilent PLgel 105 Å, 7.5 x 300 mm, 5 μm columns (part number PL1110-6550) at room temperature at a flow rate of 1.0 mL/min. The flow rate was set using Agilent 1260 Infinity Isocratic pump. THF GPC samples were filtered through a PTFE membrane with 0.2 µm pore prior to injection. UV absorbance was measured with an in line Agilent 1260 Infinity UV detector (Agilent Technologies) in the range of 190-800 nm with a 2.0 nm step and 4.0 nm slit width. Molecular weights were calculated using inline Wyatt Optilab T-rEX refractive index detector (RI, Wyatt Technologies Corp.) and Wyatt miniDAWN TREOS multi-angle light scattering detector (MALS, Wyatt Technologies Corp.). Refractive index increments (dn/dc) values for POEGMA polymers were determined inline, assuming 100 %

ACS Paragon Plus Environment

22

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

Bioconjugate Chemistry

mass recovery based on known injection mass. These dn/dc values were corroborated with data from an Abbemat 500 refractometer (Anton Paar). Dn/dc values for brush polymers were assumed to match bristle dn/dc’s for purposes of molecular weight determination. GPC data was analyzed using Astra (v6.1.5.22, Wyatt Technologies Inc.). POEGMA molecular weights and polydispersities were determined using MALS and RI data fit to a Zimm model. GPC with dimethylformamide (DMF, HPLC grade, 10 mM LiBr) eluent was performed with Styragel HR 4, Styragel HR 4E, and Styragel HR 3 columns (Mn ranges: 5−600, 0.05− 100 and 0.5−30 kDa) in series at 25 °C with a flow rate of 0.25 mL/min. The flow rate was set using a Waters 2695 separations module. DMF GPC samples were filtered through a polypropylene membrane with 0.2 µm pore prior to injection. Molecular weights were calculated using an inline Wyatt Optilab Rex refractive index detector (RI, Wyatt Technologies Corp.) and Wyatt miniDawn multi angle light scattering detector (MALS, Wyatt Technologies Corp.). The light scattering was manually calibrated using six polystyrene standards (Fluka) ranging from 2 to 700 kDa. Refractive index increments (dn/dc) values for P(AMA-co-HEMA) polymers in DMF were determined inline, assuming 100 % mass recovery based on known injection mass. GPC data was analyzed using Astra (v6.1.5.22, Wyatt Technologies Inc.). P(AMA-co-HEMA) molecular weights and polydispersities were determined using MALS and RI data fit to a Zimm model. Light Scattering: Static light scattering (SLS) and dynamic light scattering (DLS) data was obtained with a CGS-3 compact goniometer system (ALV) and LSE-5004 correlator (ALV). SLS data was fit to a Zimm formalism using ALV Static and Dynamic FIT and PLOT software (v4.48). DLS data was processed using AfterALV (v1.0), autocorrelation functions were fit at each acquired angle, a line was fit for hydrodynamic radius versus scattering angle, and the reported Rh corresponds to the extrapolated hydrodynamic radius at zero angle. SLS samples

ACS Paragon Plus Environment

23

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

Page 24 of 39

containing POEGMA were filtered through a cellulose acetate membrane with 0.2 µm pore prior to measurement. DLS and SLS samples containing P(AMA-co-HEMA) were filtered through an Anotop 10 0.1 µm pore syringe filter followed by a Anotop 10 0.02 µm pore filter immediately prior to measurement. Gel Electrophoresis: SDS-PAGE gels were run using a Mini-PROTEAN Tetra cell, 4-20 % TGX Stain Free polyacrylamide gel, and PowerPac Basic power supply purchased from BioRad. The chambers were filled with 1x tris/glycine/SDS buffer diluted from a 10x concentrated solution purchased from Bio-Rad. Precision Plus Protein Dual Color (Bio-Rad) were loaded onto gels as a protein ladder. Gels were imaged using a Typhoon 9410 for Cy3 and fluorescein signals (532 nm laser with 580 nm emission filter) and imaged using a Universal Hood II (Bio-Rad) for Coomassie stained gels. Gels to be stained with Coomassie were imaged for Cy3 prior to staining. Coomassie staining was achieved by rocking the gel in a 0.05 % solution of Coomassie Brilliant Blue R-250 (Bio-Rad) in 10 % acetic acid and 50 % methanol for 10-15 minutes, and then destaining in deionized water for a day. Gel data was analyzed using ImageJ (v2.0.0-rc-59) with Fiji to determine absorbed intensities relative to background. DNA gels were taken by loading samples into a 1.5 % agarose gel (with 0.01 % SybrSafe dye) and separated electrophoretically (130 mV, 30 minutes). The DNA ladder used was 1 kbp DNA ladder (N3232S, New England Biolabs). The gel was imaged using a Universal Hood II (BioRad) and assessed qualitatively for plasmid gene insertion. Atomic Force Microscopy. Samples were spin coated onto fresh-cleaved mica from a 100 µg/mL aqueous solution (filtered through a 0.2 µm regenerated cellulose filter, 20 µL of solution onto a 1.5 cm diameter disc, spun at 3,000 rpm for 60 seconds). Samples were imaged using TESPW AFM probes (Veeco, 20-80 N/m, 230-410 kHz) for tapping mode in air using a

ACS Paragon Plus Environment

24

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

Bioconjugate Chemistry

Bruker MultiMode AFM with Nanoscope V controller and Research Nanoscope (Bruker, v9.0, build R1Sr4.98954) software. Images were processed with Nanoscope Analysis (Bruker, v1.50, build R2.103555) to plane fit, flatten, and export images, then analyzed with ImageJ (v2.0.0-rc59) with Fiji to manually estimate brush polymer lengths and diameters. Highlighted Synthetic Procedures α-(Triisopropylsilyl-acetylene)-ω-bromo-poly(2-aminoethyl

methacrylate-co-2-

hydroxyethyl methacrylate-co-fluorescein O-methacrylate) (TIPS-P(AMA-co-HEMA)). A mixture of 90 % methanol with water was purged with nitrogen gas for one hour. Excess 2hydroxyethyl methacrylate (HEMA) was passed through basic alumina to remove inhibitor, while the 2-aminoethyl methacrylate (AMA-HCl) was purified from its inhibitor during its synthesis. To a 25 mL Schlenk tube was added CuBr2 (22.5 mg, 0.100 mmol, 5 Eq), 100 µL of 90 % MeOH, TPMA (35.1 mg, 0.121 mmol, 6 Eq), and HEMA (294.0 µL, 315.5 mg, 2.424 mmol, 120 Eq). The mixture was allowed to stir at room temperature for 15 minutes to allow the copper complex to fully form. To the lime green solution was added 90 % MeOH (965 µL), 3-(triisopropylsilyl)-2-propynyl-2-bromo-2-methylpropanoate initiator (7.3 mg from a 100 mg/mL stock solution, 20. µmol, 1 Eq), fluorescein O-methacrylate (8.0 mg, 20.2 µmol, 1 Eq), and AMA-HCl (100. mg, 0.606 mmol, 30 Eq). In a 50 mL two neck roundbottom flask was prepared excess L-ascorbic acid solution (40 mg/mL in 90 % MeOH). Each flask was placed under nitrogen atmosphere with four freeze-pump-thaw cycles. The reaction mixture was allowed to stir at 30 oC, and once the temperature had equilibrated, the reaction was started by injecting L-ascorbic acid solution (178 µL, 7.12 mg, 40.4 µmol, 2 Eq). The lime green solution was covered in foil and allowed to continue stirring at 30 oC for 22 hours, at which point the reaction was stopped by exposing to air. Reaction progress was monitored by 1H NMR, and

ACS Paragon Plus Environment

25

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

Page 26 of 39

stopped at 88.3 % conversion. The following day, an increase in viscosity had caused all stirring to stop. Polymer solutions were recollected with water and dialyzed in 8-10 kDa MWCO Spectra/Por RC dialysis tubing against 0.1 mM acetic acid for one day with frequent buffer exchanges. Polyethyleneimine (acetate salt) was added to dialysis buffer overnight (600 mg/L, 50-100 kDa) to reduce swelling. The final deep highlighter-yellow dialyte was concentrated under reduced pressure and lyophilized to yield 364.8 mg of deep yellow solid. The 30 kDa, 92 kDa, and 101 kDa polymer samples were obtained by modifying the molar ratios of initiator:AMA:HEMA:fluorescein to match 1:30:120:1, 1:60:240:2, and 1:90:360:3 respectively. Batches were prepared with equal amounts of initiator, and the amount of 90 % MeOH added was altered maintain equal molar concentrations of all non-monomer reagents across batches. Polymer dn/dc of 0.0802 mL/g in DMF (10 mM LiBr) was obtained during GPC by the detector dRI signal assuming 100 % recovery of all polymer. Polymer dn/dc of 0.1620 +/- 0.0174 mL/g (95 % confidence) in buffer (pH 3.3, 1 M HOAc, 0.3 M NaH2PO4, 24.5 oC) was determined by making five polymer stock solutions of known concentrations in buffer, dialyzing each for 2.5 hours against a common beaker of buffer to reach Donnan equilibrium, gravimetrically correcting the concentration by accounting for swelling, measuring refractive index by refractometry, and fitting a line (r2 = 0.9745). When such measures were not taken, refractive index values showed extreme deviations from linearity (Figure S54). 1H NMR (500 MHz, D2O): δ 8.20-6.77 (8.20, 7.90, 7.47, 7.27, 6.84, 6.77, s, 10H, fluorescein), 4.23 (s, 2H, PAMA CH2 on ester), 4.13 (s, 8H PHEMA CH2 on ester), 3.86 (s, 8H, PHEMA CH2 on OH), 3.28 (s, 2H, PAMA CH2 on NH3), 2.01-1.69 (s, 10H, methacryl CH2), 1.92 (s, 3H, acetate CH3), 1.12-0.94 (m, 15H, methacryl CH3) ppm. 13C NMR (125 MHz, D2O): δ 179.8 (ester), 67.7, 67.0, 59.2, 53.5 (methacryl quaternary C), 45.2 and 44.7 (methacryl CH2), 23.27, 16.8, and 9.74 ppm. 1H NMR

ACS Paragon Plus Environment

26

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

Bioconjugate Chemistry

(500 MHz, pyridine-d5): δ 5.50 (br s, 14H, H2O/OH), 4.48 (s, 2H, PAMA CH2 on ester), 4.39 (s, 8H PHEMA CH2 on ester), 4.27 (s, 1H), 4.21 (s, 1H), 4.09 (s, 8H, PHEMA CH2 on OH), 3.06 (s, 2H, PAMA CH2 on NH3), 2.37-2.25 (s, 10H, methacryl CH2), 2.15 (s, 3H, acetate CH3), 1.851.53 (s, 15H, methacryl CH3) ppm. Small peaks visible from 7-8 ppm omitted due to potential confusion with solvent

13

C satelites. FTIR-ATR (neat): 3256.1 (O–H), 2995.8, 2949.6 (C–H),

2884.3, 1717.7 (C=O), 1642.2, 1563.2, 1481.4, 1451.4, 1405.2, 1391.6, 1334.4, 1271.7, 1246.9, 1152.6 (C–O), 1075.8 (C–O), 1021.2, 969.3, 947.5, 901.3, and 852.2 cm-1. For molecular weight and size statistics determined by GPC, SLS, and DLS, see Table 1. Protection of the alkyne group was verified by diagnostic SDS-PAGE (Figure 2). α-(Acetylene)-ω-bromo-poly(2-aminoethyl

methacrylate-co-2-hydroxyethyl

methacrylate) (alkyne-P(AMA-co-HEMA)). To a 1.5 mL Eppendorf centrifuge tube was dissolved TIPS-P(AMA-co-HEMA) (150. mg, ~1.5 µmol, 1 Eq) into 320 µL methanol. In darkness was added 380 µL of 10 mg/mL AgF in methanol solution (3.80 mg, 30 µmol, ~1545 Eq per TIPS group). With the minimal light in the room, it was possible to see orange precipitate immediately form and then qiuickly redissolve. Argon was blown into the tube. It was capped and sealed with parafilm, foiled, and allowed to mix at room temperature by inversion for 24 hours. 1 M HCl was added (45 µL, 45 µmol, 1.5 Eq per AgF) and no precipitate was observed as it was mixed at room temperature for 30 minutes by inversion. The sample was dialyzed in darkness with 25 kDa MWCO Spectra/por RC dialysis tubing against 0.1 mM acetic acid for a day. The samples were then lyophilized to yield 149.4-155.3 mg fluffy intense yellow solid, corresponding to quantitative recovery. 1H NMR (500 MHz, D2O) and FTIR (neat) of these compounds were identical to their precursor polymers. Alkyne deprotection was verified by a diagnostic click reaction on an SDS-PAGE gel (Figure 2).

ACS Paragon Plus Environment

27

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

α-(ELP)-ω-bromo-poly(2-aminoethyl

methacrylate-co-2-hydroxyethyl

Page 28 of 39

methacrylate)

(ELP-P(AMA-co-HEMA)). The stoichiometry for this reaction was altered for each polymer to maintain constant molar concentrations of all reagents by using with different alkyne mass concentrations (according to P(AMA-co-HEMA) Mn’s determined by DMF GPC prior to diazotransfer of 37.0 kDa, 73.1 kDa, and 104.9 kDa for the reported 30 kDa, 92 kDa, and 101 kDa samples respectively). This procedure for 92 kDa polymer is representative for all samples. To a 1.5 mL Eppendorf centrifuge tube was added alkyne-P(AMA-co-HEMA) (16.0 mg, 0.219 µmol, 1 Eq), azido-ELP (35.8 mg, 2.19 µmol, 10 Eq), H2O (463 µL), 2 M NEt3 HOAc buffer (109 µL, pH 7.1, 10x concentrated), and pre-mixed CuSO4-THPTA solution (10 mM in H2O, 328 µL, 3.28 µmol, 15 Eq). The contents of the vial were mixed by inversion into a homogenous, deep highlighter-yellow solution, followed by addition of L-ascorbic acid solution (578 µg, 3 mg/mL in H2O, 192.5 µL, 3.279 µmol, 15 Eq), leading to a total solution volume of 1093 µL. Argon was blown into the vial before sealing with parafilm and shaking briefly. The unchanged, homogenous solution was covered in foil and allowed to mix by inversion for 23 hours. To the homogenous yellow solution was added (NH4)2SO4 solution (400 mg/mL in H2O) to a final concentration of 66.6 mg/mL (NH4)2SO4, instantly forming a large amount of white precipitate. The tube was then spun at 10,000 rcf for 10 minutes, and the lime supernatant was collected. The pellet was redissolved into 833 µL of H2O, precipitated with 166 µL (NH4)2SO4 solution, spun down, and the lime supernatant was collected twice more for a total of three precipitations. The combined lime supernatants were dialyzed against 0.1 mM acetic acid in 25 kMWCO Spectra/Por RC dialysis tubing for one day with two buffer exchanges. The dialyte was transferred with H2O and lyophilized to yield 19.5 mg deep yellow fluffy solid. Due to material constraints, the pellet, consisting mostly of protein, was also purified by dialysis

ACS Paragon Plus Environment

28

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

Bioconjugate Chemistry

against 0.1 mM acetic acid in 3.5 kMWCO Snakeskin RC dialysis tubing for 1 day with three buffer exchanges, recollected with H2O, and lyophilized to yield 28.0 mg of pale beige fluffy solid as the final product. The pellet’s purity was verified quickly by FTIR (neat) to show it predominantly consisted of protein. The pellet was then reused for a second batch of proteinpolymer conjugate while taking precautions to never potentially cross-contaminate different molecular weight backbones by keeping those sample pellets separate. FTIR (neat) of supernatant samples matched a combination of ELP and polymer precursors. FTIR (neat) of pellets matched that of ELP precursors. Functionalization was confirmed by diagnostic SDSPAGE (Figure 2). More thorough removal of ELP impurities from the polymers was performed during the following diazotransfer step. Diazotransfer Reaction for α-(ELP)-ω-bromo-poly(2-azidoethyl methacrylate-co-2hydroxyethyl methacrylate) (ELP-P(azido-AMA-co-HEMA)). This procedure for 108 kDa ELP-P(AMA-co-HEMA) is representative, as stoichiometry was adjusted to maintain constant molar concentrations and equivalents of amine across samples. To a 50 mL roundbottom flask was added 7.68 mL of K2CO3 (10 mg/mL in 70 % MeOH, 76.1 mg, 556 µmol, 15 Eq), 12.9 mL of 70 % methanol, and imidazole-1-sulfonyl azide hydrogen sulfate (1.00 mL of 100 mg/mL in 70 % MeOH, 100. mg, 370 µmol, 10 Eq). The reaction solution became finely cloudy white with the precipitation of potassium bicarbonate. After switching to incandescent lighting, 92.3 µL of CuSO4 pentahydrate (10.0 mg/mL in H2O, 923 µg, 3.70 µmol, 0.1 Eq) and ELP-P(AMA-coHEMA) (32.2 mg in 2 mL 70 % MeOH, 37.0 µmol amine, 1 Eq) were added to the reaction mixture, causing it to become turbid yellow. The roundbottom flask was capped with a septum, covered with foil, and stirred at room temperature for 22 hours, ending with a finely turbid green-yellow solution. The reaction was stopped by the addition of 1 M acetic acid (1.11 mL,

ACS Paragon Plus Environment

29

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

Page 30 of 39

30 Eq), causing the yellow color to immediately fade and precipitate white solid. After stirring for 15 minutes at room temperature, the mixture was concentrated under reduced pressure, and transferred to 25 kDa MWCO SpectraPor RC dialysis tubing with 2 mL DMF to dissolve the precipitate. The sample was dialyzed against pH 4 acetic acid (0.1 mM) for a day, yielding an agglomerated beige precipitate and clear solution. The final dialyte was spun down at 2,000 rcf for 10 minutes, and the clear supernatant was decanted. The hybrid-rich pellet was dried by lyophalization for 23.2 mg pale beige solid. The combined dry masses of the pellet and supernatant matched quantitative recovery, with the supernatant (ELP impurity) corresponding to roughly 20-30 % of the input polymer. FTIR-ATR and 1H NMR were consistent with a combination of P(azido-AMA-co-HEMA) and ELP. Ratios of ELP to synthetic polymer by 1HNMR and FTIR were equal across all samples. Presence of azide is confirmed by FTIR and disappearance of the precursor amine peak (8.16 ppm, 1H NMR). For P(azido-AMA-co-HEMA) samples, diagnostic SDS-PAGE gave variable results for the same sample due to solubility issues. 1

H NMR (500 MHz, DMSO-d6): δ 8.21-6.62 (8.04, 7.97, 7.76, 6.90, and 6.62 ppm, s, ELP amide

N-H and fluorescein), 4.81 (s, 4H, PHEMA O-H), 4.31 (s, ELP alpha CH), 4.03 (s, 2H, PAMA CH2 on ester), 3.90 (s, 8H PHEMA CH2 on ester), 3.58 (s, 10H, PHEMA CH2 on OH and PAMA CH2 on N3), 3.38 (s, ill-resolved, PAMA CH2 on azide), 1.83 (s, 8H, PHEMA methacryl CH2), 1.52 (s, 1H, PAMA methacryl CH2), 1.24 (s, 3H, residual acetate and/or acetic acid CH3, integral corresponds to >93 % conversion), and 0.96-0.79 ppm (m, 15H, methacryl CH3). 3.38 ppm (s, PAMA CH2 on azide) was not resolvable on most samples due to overlap with water. FTIR-ATR (neat): 3319.2 (O–H), 2958.3 (C–H), 2884.0, 2107.4 (N=N=N), 1722.8 (C=O), 1652.8 (amide I), 1526.2 (amide II), 1449.0, 1389.5, 1245.2, 1152.5 (C–O), 1074.5 (C–O), 1022.2, 964.4, 989.3, 851.3, and 746.7 cm-1.

ACS Paragon Plus Environment

30

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

Bioconjugate Chemistry

Grafting of alkyne-terminated POEGMA to poly(2-azidoethyl methacrylate-co-2hydroxyethyl

methacrylate)

(P((AMA-g-POEGMA)-co-HEMA)).

This

procedure

is

representative for the preparation of all brush polymers regardless of backbone length and ELPincorporation. P(azido-AMA-co-HEMA) (5.0 mg, 5.04 µmol azide, 1 Eq) was dissolved into 900 µL DMF and 100 µL H2O, a process which took a few minutes for ELP-lacking polymers and several hours for ELP-containing polymers. To a separate 50 mL roundbottom flask was added CuSO4 pentahydrate (10 mg/mL in H2O from ground solid, 1.89 mL, 18.9 mg, 75.6 µmol, 15 Eq) and THPTA (30 mg/mL in H2O, 1.46 mL, 43.7 mg, 101 µmol, 20 Eq), producing a medium intensity sapphire-blue colored solution. To the mixture was added 2 M NEt3 HOAc (1.74 mL, pH 7.1, 10x concentation), 9.4 mL methanol, and 370 µL H2O. After degassing with argon for 2 hours, to the mixture was added P(azido-AMA-co-HEMA) solution, and alkynePOEGMA (491.8 mg, 25.19 µmol, 5 Eq) dissolved in 1 mL methanol. Argon was blown into the reaction flask before finally adding sodium L-ascorbate (10 mg/mL in H2O, 1.50 mL, 15.0 mg, 75.6 µmol, 15 Eq), causing the mixture to quickly turn pale turquoise. The final reaction medium was thus 17.4 mL of 0.2 M NEt3 HOAc in 60 % (v/v) methanol (40 % water) with trace DMF. The flask was septum-sealed, covered in foil, and allowed to stir at room temperature for 22 hours. The resulting pale yellow mixture was exposed to air, concentrated under reduced pressure, and combined with 1 mL of 0.1 M EDTA (disodium salt), causing the solution to turn medium intensity green-yellow. The solution was dialyzed in 25 kMWCO Spectra/Por RC dialysis tubing against 0.1 mM EDTA (disodium salt) for 5 hours, 0.1 mM acetic acid overnight, then MilliQ H2O for 6 hours with several buffer exchanges. The dialyte was concentrated under reduced pressure, transferred to a glass centrifuge tube with water, and lyophilized. To the medium intensity yellow goo was added 600 µL methanol to redissolve, then adding 12 mL

ACS Paragon Plus Environment

31

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

Page 32 of 39

diethyl ether, causing the solution to immediately turn cloudy white. The tube was spun down, the supernatant was carefully removed by pipette, and the small amount of residual diethyl ether and methanol was removed by airstream. The process of redissolving and precipitating was repeated three more times with 1.0 mL, 1.0 mL, and 1.4 mL of methanol in each consecutive cycle. We have found the separation of bristle from the brush to very critically depend on the amount of methanol added, requiring roughly 200 µL of methanol for 4 kDa POEGMA to 1.4 mL methanol for 20 kDa POEGMA (data not shown). Generally, the polymer is repeatedly precipitated until both a sufficient mass is obtained and GPC (THF) indicates that all bristle has been removed (Figure 3). 1H NMR (500 MHz, D2O) was identical to that of alkyne-POEGMA precursors with an additional peak at 8.21 ppm, corresponding to triazole C–H. FTIR-ATR (neat) was identical to alkyne-POEGMA. NMR molecular weight statistics were calculated based on data from 1H NMR in pyridine-d5. For consistency, all integral regions were precisely matched across polymer samples. Bristle degree of polymerization (Xn) was calculated from the ratio between the POEGMA terminal methoxy CH3 (3.4 ppm) and triazole CH (8.4 ppm). The product of grafting efficiency and bristle degree of polymerization was calculated from then ratio of PHEMA CH2 (4.12 ppm) and POEGMA terminal methoxy CH3 (3.4 ppm). Grafting efficiency was then calculated by division of the NMR-derived values. 1H NMR (500 MHz, pyridine-d5): δ 8.41 (s, triazole C–H), 6.54 (s, annihilated by D2O shake, PHEMA OH), 5.65 (s, 2H), 4.42 (s, POEGMA CH2 on ester, PAMA CH2 on ester, and PHEMA CH2 on ester), 4.12 (s, PHEMA CH2 on O–H), 3.82 (s, POEGMA bulk CH2), 3.64 (s, POEGMA last CH2), 3.40 (s, POEGMA terminal OCH3), 2.27 (s, methacryl CH2), and 1.53-1.44 ppm (s, methacryl CH3). Diagnostic Click Test Reaction. Fresh stock aqueous solution of 3 mg/mL L-ascorbic acid and 1.3 mM solution of the polymer to be tested was made the day of the test reaction. To a

ACS Paragon Plus Environment

32

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

Bioconjugate Chemistry

0.5 mL plastic centrifuge tube was added 3 µL of test polymer solution (4 nmol of azide/alkyne), 0.8 µL of 10 mM Cy3 dye in DMSO (8 nmol, azide or alkyne terminated), 6.0 µL of 10 mM CuSO4 solution with 10 mM THPTA (60 nmol, pre-mixed), 2.0 µL of pH 7 2.0 M NEt3 HOAc buffer, and 4.7 µL water. The final buffer conditions were 0.2 M NEt3 HOAc pH 7. For test reactions involving azide backbones, all samples to be tested on the same gel were reacted using 40 % (v/v) ethanol in place of water. The 20 µL of magenta solution was combined by light centrifugation and mixed by pipet. The reaction pot was sealed and allowed to react at room temperature for overnight. A gel loading sample was prepared by mixing 2 µL reaction mixture, 10 µL of 2x Laemli buffer/glycerol in a 4:1 v/v mixture, and 8 µL of water. 10 µL of gel loading sample was run on a 4-20 % polyacrylamide gel alongside a protein ladder and positive control click reaction mixture (using 10 kDa PEG azide or 20 kDa PEG alkyne) at 120 V for an hour. The gels were fluorescently imaged for the presence of Cy3, and total Cy3 fluorescent intensity was determined using ImageJ. Afterward, gels with amine-containing polymer samples were stained using Coomassie brilliant blue for 10 minutes, and subsequently destained in water overnight while changing the water four times. The stained gel was then imaged to corroborate the migration and shape of the polymer band in the Cy3 scan and verify presence of loaded polymer.

ASSOCIATED CONTENT Supporting Information. Detailed synthetic procedures, characterization of all prepared compounds, and an extended discussion are available in supporting information. The following

ACS Paragon Plus Environment

33

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

Page 34 of 39

files are available free of charge. Synthetic procedures and spectra (PDF) AUTHOR INFORMATION Corresponding Authors *Email: [email protected] (S.Z.) *Email: [email protected] (L.A.N.) Present Addresses Author Contributions The manuscript was written through contributions of all authors. All authors have given approval to the final version of the manuscript. Luis Navarro prepared and characterized all small molecule reagents, synthetic polymers, and bioconjugates. Daniel French prepared unfunctionalized ELPs used in this study. Stefan Zauscher aided in development. Funding Sources Funding was provided by the National Science Foundation (NSF) Triangle Materials Research Science and Engineering Center (MRSEC) (DMR-1121107). We are also grateful for generous support from Prof. Farshid Guilak. Notes The authors declare no competing financial interests. ACKNOWLEDGMENT We are grateful to Yeongun Ko in Prof. Jan Genzer’s group (North Carolina State University) for performing DMF GPC of backbone polymer samples. We also thank Meredith Barbee in Prof.

ACS Paragon Plus Environment

34

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

Bioconjugate Chemistry

Steven Craig’s group (Duke University) for aid in performing THF GPC of POEGMA-based polymers. We thank Prof. Ashutosh Chilkoti’s laboratory for supplying plasmids for preparing ELP and offering use of their light scattering and biosynthesis equipment. We thank Peter Silinski for acquiring and processing ESI-TOF data. ABBREVIATIONS Cy3, cyanine3 (dye); ELP, elastin-like polypeptide (with the inclusion of a collagen-binding peptide, model protein); ISA, imidazole-1-sulfonyl azide (hydrogen sulfate salt, diazotransfer reagent); PAMA, poly(2-aminoethyl methacrylate) (amine-rich backbone); PHEMA, poly(2hydroxyethyl methacrylate) (backbone spacer); POEGMA, poly(oligoethylene glycol methacrylate) (monomer Mn = 300, bristle); TIPS, triisopropylsilyl (alkyne protecting group). References (1) (2)

(3)

(4)

(5)

(6) (7)

(8)

Han, L., Grodzinsky, A. J., and Ortiz, C. (2011) Nanomechanics of the Cartilage Extracellular Matrix. Annu Rev Mater Res 41, 133-168. Rhee, D. K., Marcelino, J., Baker, M., Gong, Y., Smits, P., Lefebvre, V., Jay, G. D., Stewart, M., Wang, H., Warman, M. L., et al. (2005) The secreted glycoprotein lubricin protects cartilage surfaces and inhibits synovial cell overgrowth. Journal of Clinical Investigation 115, 622-631. Coles, J. M., Zhang, L., Blum, J. J., Warman, M. L., Jay, G. D., Guilak, F., and Zauscher, S. (2010) Loss of cartilage structure, stiffness, and frictional properties in mice lacking PRG4. Arthritis Rheum 62, 1666-1674. Cui, Z., Xu, C., Li, X., Song, J., and Yu, B. (2015) Treatment with recombinant lubricin attenuates osteoarthritis by positive feedback loop between articular cartilage and subchondral bone in ovariectomized rats. Bone 74, 37-47. Sharma, S., Panitch, A., and Neu, C. P. (2013) Incorporation of an aggrecan mimic prevents proteolytic degradation of anisotropic cartilage analogs. Acta Biomater 9, 46184625. Bernhard, J. C., and Panitch, A. (2012) Synthesis and characterization of an aggrecan mimic. Acta Biomater 8, 1543-50. Ostedgaard, L. S., Moninger, T. O., McMenimen, J. D., Sawin, N. M., Parker, C. P., Thornell, I. M., Powers, L. S., Ganswmer, N. D., Bouzek, k. D., Cook, D. P., et al. (2017) Gel-forming mucins form distinct morphologic structures in airways. PNAS 114, 68426847. Sterner, O., Karageoraki, C., Zurcher, M., Zurcher, S., Scales, C. W., Fadli, Z., Spencer, N. D., and Tosatti, S. G. P. (2017) Reducing Friction in the Eye: A Comparative Study of

ACS Paragon Plus Environment

35

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

(9) (10) (11)

(12)

(13)

(14)

(15)

(16)

(17) (18)

(19) (20)

(21) (22)

(23)

(24)

Page 36 of 39

Lubrication by Surface-Anchored Synthetic and Natural Ocular Mucin Analogues. Applied Materials and Interfaces 2017, 20150-20160. Bansil, R., Stanley, E., and LaMont, J. T. (1995) Mucin biophysics. Annu Rev Physiol 57, 635-657. Jay, G. D., Harris, D. A., and Cha, C. J. (2001) Boundary lubrication by lubricin is mediated by O-linked beta(1-3)Gal-GalNAc oligosaccharides. Glycoconj J 18, 807-815. Flannery, C. R., Zollner, R., Corcoran, C., Jones, A. R., Root, A., Rivera-Bermudez, M. A., Blanchet, T., Gleghorn, J. P., Bonassar, L. J., Bendele, A. M., et al. (2009) Prevention of cartilage degeneration in a rat model of osteoarthritis by intraarticular treatment with recombinant lubricin. Arthritis Rheum 60, 840-847. Abubacker, S., Dorosz, S. G., Ponjevic, D., Jay, G. D., Matyas, J. R., and Schmidt, T. A. (2016) Full-Length Recombinant Human Proteoglycan 4 Interacts with Hyaluronan to Provide Cartilage Boundary Lubrication. Ann Biomed Eng 44, 1128-1137. Gleghorn, J. P., Jones, A. R., Flannery, C. R., and Bonassar, L. J. (2009) Boundary mode lubrication of articular cartilage by recombinant human lubricin. J Orthop Res 27, 771777. Banquy, X., Burdynska, J., Lee, D. W., Matyjaszewski, K., and Israelachvili, J. (2014) Bioinspired bottle-brush polymer exhibits low friction and Amontons-like behavior. J Am Chem Soc 136, 6199-6202. Lawrence, A., Xu, X., Bible, M. D., Calve, S., Neu, C. P., and Panitch, A. (2015) Synthesis and characterization of a lubricin mimic (mLub) to reduce friction and adhesion on the articular cartilage surface. Biomaterials 73, 42-50. Prudnikova, K., Yucha, R. W., Patel, P., Kriete, A. S., Han, L., Penn, L. S., and Marcolongo, M. S. (2017) Biomimetic Proteoglycans Mimic Macromolecular Architecture and Water Uptake of Natural Proteoglycans. Biomacromolecules 18, 17131723. Kramer, J. R., Onoa, B., Bustamante, C., and Bertozzi, C. R. (2015) Chemically tunable mucin chimeras assembled on living cells. PNAS 112, 12574-12579. Hokajo, T. T., K.; Nakamura, Y.; Norisuye, T. (2001) Solution Properties of Polymacromonomers Consisting of Polystyrene v. Effect of Side Chain Length on Chain Stiffness. Polymer Journal 33, 481-485. Gauge, A., and Pakula, T. (1995) Static Properties of Noninteracting Comb Polymers in Dense and Dilute Media. A Monte Carlo Study. Macromolecules 28, 190-196. Rathgeber, S., Pakula, T., Wilk, A., Matyjaszewski, K., and Beers, K. L. (2005) On the shape of bottle-brush macromolecules: systematic variation of architectural parameters. J Chem Phys 122, 124904. Paturej, J., Sheiko, S. S., Panyukov, S., and Rubinstein, M. (2016) Molecular structure of bottlebrush polymers in melts. Sci Adv 2, e1601478. Wintermantel, M., Gerle, M., Fischer, K., Schmidt, M., Wataoka, I., Urakawa, H., Kajiwara, K., and Tsukahara, Y. (1996) Molecular bottlebrushes. Macromolecules 29, 978-983. Daniel, W. F., Burdynska, J., Vatankhah-Varnoosfaderani, M., Matyjaszewski, K., Paturej, J., Rubinstein, M., Dobrynin, A. V., and Sheiko, S. S. (2016) Solvent-free, supersoft and superelastic bottlebrush melts and networks. Nat Mater 15, 183-189. McKay, C. S., and Finn, M. G. (2014) Click chemistry in complex mixtures: bioorthogonal bioconjugation. Chem Biol 21, 1075-1101.

ACS Paragon Plus Environment

36

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

Bioconjugate Chemistry

(25) (26) (27)

(28)

(29)

(30)

(31)

(32)

(33)

(34)

(35)

(36) (37)

(38)

(39)

Tornøe, M. M. C. W. (2008) Cu-Catalyzed Azide-Alkyne Cycloaddition. Chem Rev 108, 2952-3015. Engler, A. C., Lee, H. I., and Hammond, P. T. (2009) Highly efficient "grafting onto" a polypeptide backbone using click chemistry. Angew Chem Int Ed Engl 48, 9334-9338. Sheiko, S. S., Sumerlin, B. S., and Matyjaszewski, K. (2008) Cylindrical molecular brushes: Synthesis, characterization, and properties. Progress in Polymer Science 33, 759-785. Yan, Y., Shi, Y., Zhu, W., and Chen, Y. (2013) Highly efficient synthesis of cylindrical polymer brushes with various side chains via click grafting-onto approach. Polymer 54, 5634-5642. Hasneen, A., Han, H. S., and Paik, H.-J. (2009) Synthesis of linear tetrablock quaterpolymers via atom transfer radical polymerization and a click coupling approach. Reactive and Functional Polymers 69, 681-687. Mansfeld, U., Pietsch, C., Hoogenboom, R., Becer, C. R., and Schubert, U. S. (2010) Clickable initiators, monomers and polymers in controlled radical polymerizations – a prospective combination in polymer science. Polymer Chemistry 1, 1560-1598. Leophairatana, P., Samanta, S., De Silva, C. C., and Koberstein, J. T. (2017) Preventing Alkyne-Alkyne (i.e., Glaser) Coupling Associated with the ATRP Synthesis of AlkyneFunctional Polymers/Macromonomers and for Alkynes under Click (i.e., CuAAC) Reaction Conditions. J Am Chem Soc 139, 3756-3766. Fleischmann, S., Komber, H., Appelhans, D., and Voit, B. I. (2007) Synthesis of Functionalized NMP Initiators for Click Chemistry: A Versatile Method for the Preparation of Functionalized Polymers and Block Copolymers. Macromolecular Chemistry and Physics 208, 1050-1060. Opsteen, J. A., and van Hest, J. C. M. (2007) Modular synthesis of ABC type block copolymers by “click” chemistry. Journal of Polymer Science Part A: Polymer Chemistry 45, 2913-2924. O'Reilly, R. K., Joralemon, M. J., Hawker, C. J., and Wooley, K. L. (2006) Fluorogenic 1,3-dipolar cycloaddition within the hydrophobic core of a shell cross-linked nanoparticle. Chemistry 12, 6776-6786. Ladmiral, V., Mantovani, G., Clarkson, G. J., Cauet, S., Irwin, J. L., and Haddleton, D. M. (2006) Synthesis of neoglycopolymers by a combination of "click chemistry" and living radical polymerization. J Am Chem Soc 128, 4823-4830. Gao, H. M., Krzysztof. (2007) Synthesis of Molecular Brushes by “Grafting onto” Method: Combination of ATRP and Click Reactions. J Am Chem Soc 129, 6633-6639. Sumerlin, B. S., Nicolay V. Tsarevsky, Guillaume Louche, Robert Y. Lee, and Krzysztof Matyjaszewski. (2005) Highly Efficient “Click” Functionalization of Poly(3-azidopropyl methacrylate) Prepared by ATRP. Macromolecules 38, 7540-7545. Van Camp, W., Germonpré, V., Mespouille, L., Dubois, P., Goethals, E. J., and Du Prez, F. E. (2007) New poly(acrylic acid) containing segmented copolymer structures by combination of “click” chemistry and atom transfer radical polymerization. Reactive and Functional Polymers 67, 1168-1180. Nyffeler, P. T., Liang, C.-H., Koeller, K. M., and Wong, C.-H. (2002) The Chemistry of Amine−Azide Interconversion: Catalytic Diazotransfer and Regioselective Azide Reduction. Journal of the American Chemical Society 124, 10773-10778.

ACS Paragon Plus Environment

37

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

(40)

(41)

(42)

(43)

(44) (45)

(46)

(47)

(48) (49) (50) (51)

(52)

(53)

Page 38 of 39

Goddard-Borger, E. D., and Stick, R. V. (2007) An efficient, inexpensive, and shelfstable diazotransfer reagent: imidazole-1-sulfonyl azide hydrochloride. Org Lett 9, 37973800. Potter, G. T., Jayson, G. C., Miller, G. J., and Gardiner, J. M. (2016) An Updated Synthesis of the Diazo-Transfer Reagent Imidazole-1-sulfonyl Azide Hydrogen Sulfate. J Org Chem 81, 3443-3446. van Dongen, S. F. M., Teeuwen, R. L. M., Nallani, M., van Berkel, S. S., Cornelissen, J. J. L. M., Nolte, R. J. M., and van Hest, J. C. M. (2009) Single-Step Azide Introduction in Proteins via an Aqueous Diazo Transfer. Bioconjugate Chemistry 20, 20-23. Schoffelen, S., van Eldijk, M. B., Rooijakkers, B., Raijmakers, R., Heck, A. J. R., and van Hest, J. C. M. (2011) Metal-free and pH-controlled introduction of azides in proteins. Chemical Science 2, 701-705. Quiñonez-Angulo, M. P. (2017), Instituto Tecnológico de Tijuana. Smits, F. C. M., Castelijns, W. W. A., and van Hest, J. C. M. (2015) Crosslinked ELPbased nanoparticles, using the strain promoted azide–alkyne cycloaddition. European Polymer Journal 62, 386-393. Pang, Y., Liu, J., Qi, Y., Li, X., and Chilkoti, A. (2016) A Modular Method for the HighYield Synthesis of Site-Specific Protein-Polymer Therapeutics. Angew Chem Int Ed Engl 55, 10296-10300. Golas, P. L., Tsarevsky, N. V., Sumerlin, B. S., and Matyjaszewski, K. (2006) Catalyst Performance in “Click” Coupling Reactions of Polymers Prepared by ATRP: Ligand and Metal Effects. Macromolecules 39, 6451-6457. Presolski, S. I., Hong, V. P., and Finn, M. G. (2011) Copper-Catalyzed Azide-Alkyne Click Chemistry for Bioconjugation. Curr Protoc Chem Biol 3, 153-162. Thompson, K. L., Read, E. S., and Armes, S. P. (2008) Chemical degradation of poly(2aminoethyl methacrylate). Polymer Degradation and Stability 93, 1460-1466. Sumerlin, B. S., and Vogt, A. P. (2010) Macromolecular Engineering through Click Chemistry and Other Efficient Transformations. Macromolecules 43, 1-13. Gerle, M., Fischer, K., Roos, S., Muller, A. H. E., Schmidt, M., Sheiko, S. S., Prokhorova, S., and Moller, M. (1999) Main chain conformation and anomalous elution behavior of cylindrical brushes as revealed by GPC/MALLS, light scattering, and SFM. Macromolecules 32, 2629-2637. Lartia, R., Bonnat, L., Murat, P., and Defrancq, E. (2012) Preparation of azido containing oligonucleotides through diazo transfer reaction. Curr Protoc Nucleic Acid Chem Chapter 4, 4.50.1-4.50.7. McDaniel, J. R., Mackay, J. A., Quiroz, F. G., and Chilkoti, A. (2010) Recursive directional ligation by plasmid reconstruction allows rapid and seamless cloning of oligomeric genes. Biomacromolecules 11, 944-952.

ACS Paragon Plus Environment

38

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

Bioconjugate Chemistry

Table of Contents Graphic NH2 NH2

≡-

NH2

NH2

-N3 ELP

NH2 NH2 NH2

NH2

ISA N3 N3

N3 N3

≡POEGMA

ELP-(PAMA-g-POEGMA) Brush

ACS Paragon Plus Environment

39