Skip to main content
NIHPA Author Manuscripts logoLink to NIHPA Author Manuscripts
. Author manuscript; available in PMC: 2022 Sep 16.
Published in final edited form as: Chem Catal. 2021 Jun 28;1(4):870–884. doi: 10.1016/j.checat.2021.05.016

Organometallic AlaM Reagents for Umpolung Peptide Diversification

Feng Zhu 1,2,3, Wyatt C Powell 1,3, Ruiheng Jing 1, Maciej A Walczak 1,*
PMCID: PMC8562471  NIHMSID: NIHMS1711918  PMID: 34738092

Abstract

Selective modifications of peptides and proteins have emerged as a promising strategy to develop novel mechanistic probes and prepare compounds with translational potentials. Here, we report alanine carbastannatranes AlaSn as a universal synthon in various C-C and C-heteroatom bond-forming reactions. These reagents are compatible with peptide manipulation techniques and can undergo chemoselective conjugation in minutes when promoted by Pd(0). Despite their increased nucleophilicity and propensity to transfer the alkyl group, C(sp3)-C(sp2) coupling with AlaSn can be accomplished at room temperature under buffered conditions (pH 6.5–8.5). We also show that AlaSn can be easily transformed into several canonical L- and D-amino acids in arylation, acylation, and etherification reactions. Furthermore, AlaSn can partake in macrocyclizations exemplified by the synthesis of medium size cyclic peptides with various topologies. Taken together, metalated alanine AlaSn demonstrates unparalleled scope and represents a new type of umpolung reagents suitable for structure-activity relationship studies and peptide diversification.

eTOC

AlaM amino acids with a carbastannatrane group installed at the β position are umpolung reagents that can be engaged in chemoselective Pd- and Cu-catalyzed cross-coupling reactions. Despite their ability to undergo facile transmetalation, they are compatible with common peptide synthesis protocols and operate under aqueous conditions. AlaM building blocks could be incorporated into oligopeptides and produced “mutated” amino acids with aryl, acyl, and thioether functionalities as well as cyclic structures.

Graphical Abstract

graphic file with name nihms-1711918-f0001.jpg

INTRODUCTION

Site- and chemoselective modification of proteins and peptides is becoming recognized as an important tool for probing structure-function relationship and accessing new therapeutic leads.14 Significant advances have been made over the past decade to modify peptides using heteroatom conjugation with cysteine/selenocysteine,58 serine,9 lysine,10 or methionine,11 site specific C-H functionalization of aromatic rings in tryptophan, histidine, or phenylalanine,1214 radical functionalization,15,16 and decarboxylative couplings of C-terminal amino acids or side chains functionalities in aspartic and glutamic acids.1721 In addition to these chemical diversification strategies,22 natural peptides with posttranslational modifications are gaining increasing importance due to their translational potential.2327 Among peptides of ribosomal origin,28 lantipeptides (exemplified by nisin 1, Scheme 1) form a subset of polycyclic natural products featuring a thioether linkage in the form of meso-lanthionine (Lan, 2) and 3-methyllanthionine (MeLan, 3).29 In the same category, tryptothionine cross-linked toxic peptides such as actin-binding phalloidin 430,31 and RNA polymerase II inhibitor α-amanitin32,33 from the Amanita phalloides mushroom constitute another class of thioether modifications. Furthermore, variations at the aryl groups resulting from oxidative dimerization of tyrosine and hydroxyphenylglycine such as arylomycins (5)3439 or oxidative cleavage of the indole ring in tryptophan (L-kynurenine in lipopeptide daptomycin 6)40 give rise to agents with promising antibacterial activities. The unique structural modifications contribute to the diversity of peptides but also represent a unique synthetic challenge. One strategy that has attracted considerable attention are conjugate additions to dehydroalanine (Dha) 8 readily generated from cysteine 7 (Scheme 1B).4144 Due to its polarization, the β-carbon in Dha can accept both radical and anionic reactants offering a broad scope of peptide modifications, and an array of radical and nucleophiles were used to prepare protein conjugates achieving divergent late-stage modifications.41,45,46 However, the stereochemistry at the resulting α-carbon is difficult to control,47,48 and only a handful of examples such as nucleophilic addition of dehydroalanine within a complex environment of proteins,42,43 Rh-catalyzed tandem 1,4-addition/stereoselective protonation,4952 and Friedel–Crafts conjugate addition53 are known. Complementary to C-C bond forming processes, enantioselective organocatalytic addition of aryl or benzyl thiols to α-aminoacrylates proceeded in moderate to good enantioselectivities.54,55

Scheme 1.

Scheme 1.

Peptide post-translational modifications and chemical methods for their installation.

To address the above limitations, we envisioned that reversal of polarity at the amino acid β carbon represents a promising yet unexplored approach (Scheme 1B). This strategy calls for generation of metalated alanine AlaM 10 that could be engaged in reactions with electrophilic partners. In addition to addressing the concerns of epimerization, AlaM 10 constitutes a universal synthon as 15 of out 20 canonical amino acids can be directly derived from this building block through C-C, C-O, or C-S cross-coupling reactions. Furthermore, a broad selection of coupling partners can vastly increase amino acid diversity and provide access to topologically unique structures such as lantipeptides.

In designing a new method based on AlaM, two critical considerations need to be addressed: (a) formation and stability of AlaM and (b) efficiency of the potential transmetalation step that can control (and ultimately limit) the compatibility of the protocol with complex systems. These two aspects reduce the reaction discovery process to identification of a suitable metal in AlaM while maintaining the amine and carboxylate groups intact for broad synthetic utility. Catalytic metalation of the methyl group in alanine has been achieved through directed C-H activation,5658 but these conditions (high temperatures, pure organic solvents, and specialized directing groups) may be incompatible with complex peptides, proteins, and even some functional groups found in common amino acids. Alternatively, AlaM can be used stoichiometrically as a stable reagent, and previous attempts to realize this strategy utilized organolithium,59 organozinc,6066 organonickel,67,68 organogermanium,6971 and organoboron58,59,68,7286 reagents. Although some of these compounds could be successfully engaged in downstream applications, only single amino acid derivatives were used and their instability under aqueous conditions render them suboptimal for a widespread use.

In line with our interest in glycoconjugate synthesis via cross-coupling with anomeric nucleophiles,8791 we hypothesized that a stable stannane could be installed at the β carbon in alanine. Tetraalkylstannanes are generally considered poorly nucleophilic, but selective transfer of alkyl groups can be achieved using carbastannatranes9296 leading us to propose amino acids with the general formula AlaSn 11 as competent reagents for umpolung functionalization. Carbastannatranes are significantly less toxic than tetraalkyl stannanes97 and compatible with aqueous and buffered conditions (see below). By-products of reactions with carbastannatranes are in general polar and poorly soluble in organic solvents, thus simple flash purification is sufficient to obtain products in high purity. Coordination of the nitrogen atom improves their reactivity and determines a selective transfer of one alkyl group. Herein, we reported a novel strategy for the late-stage modification of peptides with AlaSn carbastannatrane amino acid synthons. This protocol exhibits high chemoselectivity compared to other heteroatom-based nucleophiles and conjugation with a variety of electrophiles was achieved through C-C, C-S, C-Se bond-forming processes. All of these protocols are operational under mild “biological” conditions (aqueous buffers, high dilution, and room temperatures) and can be applied to complex substrates.

RESULTS AND DISCUSSION

A. C(sp3)-C(sp2) Arylation.

At the outset of our studies we investigated protocols for the synthesis of AlaM amino acids (Scheme 2). AlaSn derivatives 11a and 11b were prepared in a reaction of β-iodoalanine 12c and 12d with zinc followed by quenching with 5-chloro-1-aza-5-stannabicyclo[3.3.3]undecane 13 (48–64%). Methyl esters 11 were easily synthesized on a multigram scale and could be converted into acids via saponification (NaOH, LiOH, Me3SnOH, or TMSOK). The Fmoc group in 11b can be removed under standard deprotection conditions (piperidine, DBU, Et2NH) without loss of the carbastannatrane group. Free amines and carboxylic acids of 11 can be also engaged in amide couplings without epimerization in either component, and these reagents are stable in water and various buffered solutions (pH 6.5–8.5) for at least 24 h at room temperature. We also note that S-phenylthioester of Boc-AlaSn can participate in native chemical ligation with L-cysteine, therefore free thiols remain compatible with activated carbastannatranes.

Scheme 2.

Scheme 2.

Synthesis of AlaSn reagents.

Having access to the key building blocks, we next prepared model dipeptide 14a and used it in optimization studies geared toward C(sp3)-C(sp2) cross-coupling (Table 1). The initial evaluations using Pd2(dba)3 (5 mol%) and JackiePhos (20 mol%)98,99 as the catalytic system with CuCl (1.5 equiv), KF (2.0 equiv), and 4-phenylbromobenzene 15a (1.5 equiv) in 1,4-dioxane proved to be quite effective and afforded biphenyl peptide 16a in 88% yield (entry 1). We note that no C-N cross-coupling by-products of tryptophan and 4-bromodiphenyl was observed. Several control experiments established that the Pd catalyst and CuCl were indispensable for the success of this reaction, but absence of KF had no significant effect on the reaction yield (entries 2–4). When 1,4-dioxane was replaced with DMF or MeCN as alternative solvents, the yields of the desired product 16a were reduced to 46% and 70%, respectively. Moreover, our attempts to use other mono- and bidentate phosphines such as PPh3, dppf, AdBrettPhos, or tBuBrettPhos proved ineffective and the yields were consistently lower than for JackiePhos (for details, Table S1). Further reduction of the amount of CuCl to 50 mol% led to little improvement (entries 7 and 8). To develop mild bioconjugation conditions, we ultimately found that the C-C cross-coupling worked well at 23°C (entries 9–11). Furthermore, to our delight, we established that dipeptide 14a was compatible with co-solvent systems of MeCN or t-BuOH and water with 70% isolated yield of 16a obtained by tuning the amount of nucleophile (entries 12–14). To further demonstrate the mildness of the new protocol, we employed phosphate buffers with near-neutral pH that are relevant to bioconjugation of peptides and proteins.100 The desired peptide was also obtained in good yield (66%−70%) when phosphate buffers within the range of pH 6.5–8.5 were used. Of note is the fact that the cross-coupling reactions can be completed in 15 minutes (0.005 M) with 83% isolated yield of 16a (for details, Schemes S1 and S2). The high chemoselectivity and mild conditions (room temperature, aqueous buffers, and short reaction time) make this method suitable for the late-stage modification of complex oligopeptides.

Table 1.

Reaction development of AlaM Aryl Cross-Coupling.

graphic file with name nihms-1711918-t0007.jpg
Entry CuCl (equiv) KF (equiv) Solvent Temperature [°C] Yield [%]

1 1.50 2 dioxane 100 88
2 1.50 - dioxane 100 84
3 - 2 dioxane 100 n.d.
4a 1.50 2 dioxane 100 n.d.
5 1.50 2 DMF 100 46
6 1.50 2 MeCN 100 70
7 0.50 2 dioxane 100 95
8 0.50 - dioxane 100 92
9 0.50 - dioxane 60 97
10 0.50 - dioxane 40 95
11 0.50 - dioxane 23 90
12 0.50 - MeCN:H2O (3:1) 23 54
13b 0.50 - MeCN:H2O (3:1) 23 70
14b 0.50 - t-BuOH:H2O (3:1) 23 70

14a (0.100 mmol, 1 equiv), 4-bromodiphenyl (1.50 equiv), Pd2(dba)3 (2.5 – 5.0 mol%), JackiePhos (10 – 20 mol%), CuCI (0.50 – 1.50 equiv), KF (2.00 equiv), solvent (2.00 mL), 100 °C, 24 h, isolated yields.

a

Pd2(dba)3 was not used.

b

14a (0.150 mmol, 1.5 equiv) was used. dba = dibenzylideneacetone.

With the optimized conditions in hand, we next evaluated the generality of C(sp3)-C(sp2) cross-coupling method (Scheme 3). A wide range of electrophiles with different functional groups could be successfully transformed into arylalanine derivatives (Scheme 3A). In addition of aryl halides (PhCl, PhBr, and PhI), oxygen-based partners such as PhOTf are also viable under the standard conditions resulting in the preparation of L-phenylalanine 16b (L-AlaML-Phe mutation). Notably, substituents such as ester (16c), cyano (16d), trifluoromethyl (16e), pyridyl (16f), furyl (16g) and thienyl (16h) groups were tolerated without significant variation in yield (57–87%) delivering the targeted products in excellent chemoselectivities. We were pleased to find that alkenyl and benzyl bromides are suitable for the cross-coupling under the general conditions delivering L-phenylallylglycine 16i (80%) and L-homoalanine derivative 17a (70%). It is worth pointing out that AlaSn is compatible with free phenols and indole derivatives resulting in a conversion of L-AlaSn into L-tyrosine (17b, 87%) and L-tryptophan (17c, 88%).

Scheme 3. Scope of AlaM Arylation.

Scheme 3.

General reaction conditions: 14 (0.100 mmol, 1.0 equiv), electrophile reagent (1.5 equiv), Pd2(dba)3 (5.0 mol%), JackiePhos (20 mol%), CuCl (50 mol%), 1,4-dioxane (2.00 mL), 100 °C, 24 h, isolated yields. a14a or 14b (0.150 mmol, 1.5 equiv) was used. bKF (0.200 mmol, 2.0 equiv) was used. c14b (0.250 mmol, 2.5 equiv), CuCl (1 equiv), 37 °C, and 48 h were used. d14b (2.5 equiv), CuCl (1 equiv), 90 °C, and 48 h were used. e23 °C, 48 h was used; fArg side chain protected with Cbz and removed with 10% Pd/C in MeOH/EtOAc (1:1), H2 (1 atm). Nap = 2-naphthyl.

We next applied the AlaM cross-coupling protocol to peptide conjugation with small bioactive molecules. These studies were inspired by the previous work on direct attachment of cytotoxic payloads to antibodies as well as modifications of cyclic peptides with low molecular-weight iron chelators exemplifying only selected strategies to overcome target selectivity and poor cellular permeability by site-selective modifications.101 Several complex substrates including commercially available pharmaceuticals and other biologically active molecules (16j-16l, 17d, and 17e) shown in Scheme 3A demonstrate that late-stage functionalization can be advantageous for the preparation of new scaffolds derived from phenylalanine (16j), BODIPY dye (17d), lipid-lowering drug fenofibrate (17e), antidepressant moclobemide (16k), and anti-inflammatory drug indomethacin (16l) all achieved by coupling with dipeptide stannanes 14. The installation of a fluorescence imaging probe such as BODIPY (17d) is of particular significance102 because it complements nucleophilic cysteine arylation methods previously described to attached BODIPY to peptides103 and avoids the use nitrogen protecting groups required to direct CH activation in the earlier attempts to install fluorescent dyes.104,105 High chemoselectivity was also observed in the reactions with aromatic chlorides (16k, 16l, 17e). A series of substituents such as methyl, methoxy, chloro, carbonyl, and amido groups were tolerated. We note that the cross-coupling protocol can be easily extended to double coupling (17f and 17g) in excellent yields.

The overall success of AlaSn cross-coupling relies on the compatibility of the optimized conditions with common functional groups present in peptides and proteins. Since carbastannatranes are stable under typical amidation conditions (as shown here in the preparation of several AlaSn-containing peptides), the next task was to evaluate the Pd-catalyzed protocols. As shown in Scheme 3B, potentially detrimental functionalities such as thioethers (18a), primary alcohols (18b), amides (18c), azides (18d), and guanidine in arginine (18e) were compatible with Pd(0) and JackiePhos.

B. Alanine Acylation.

We next evaluated the generality of our approach in alanine acylations that introduce a carbonyl functionality at the β-methylene position (Scheme 4A). In addition to direct conversion of AlaM into aspartic acid and asparagine, β-amino acid ketones represent an important class of bioactive peptides.106,107 The synthesis of amino ketones from α-amino acid derivatives either with organometallic reagents108110 or via Friedel−Crafts acylation111 were described, but the catalytic reactions targeting carboxylic acid side groups of amino acids to obtain amino ketones are rare. To the best of our knowledge, only one palladium-catalyzed Suzuki−Miyaura reaction of phenyl esters of aspartic acid with aryl boronic acids was reported.112 Enantioselective synthesis of side chain amino ketone derivatives by an NHC-catalyzed intermolecular Stetter reaction of aromatic aldehydes and methyl 2-acetamidoacrylate were developed, but electron-rich alkyl aldehydes were not compatible with these conditions.113 Collectively, the lack of general methods for side chain acylation represents an opportunity to develop new synthetic strategies, and AlaM are suitable for this study because a large collection of potential acyl donors is known.

Scheme 4. Scope of AlaM Acylation and Etherification.

Scheme 4.

General reaction conditions for AlaM acylation: 14a or 14b (0.100 mmol, 1 equiv), electrophile (1.5 equiv), Pd2(dba)3 (5.0 mol%), JackiePhos (20 mol%), CuCl (50 mol%), 1,4-dioxane (2.00 mL), 100 °C, 24 h, isolated yields, 14a was used for 19a-19d and 19k-19n; 14b was used for 19e-19j and 19o. atert-Butyl ester of 14b and Fmoc-protected anthranilic acid S-phenyl ester were used for cross-coupling, then 20% piperidine in CH2Cl2 was used for aniline deprotection. btert-Butyl ester of 14b was used for cross-coupling, then LiOH·H2O was used for hydrolysis. ctert-Butyl ester of 14b was used for cross-coupling, then saturated solution of ammonia in methanol was used. dPd2(dba)3 (10 mol%), dppp (25 mol%), CuCl (3 equiv), 1,4-dioxane (2.00 mL), 110 °C, 24 h. e36 h reaction time. General reaction conditions for AlaM C-S/C-Se cross-coupling: 14a (0.100 mmol, 1 equiv), electrophile reagents (1.5 equiv), CuCl (50 mol%), 1,4-dioxane (2.00 mL), 23 °C, 48 h, isolated yields. f100 °C, 24 h. g14a (0.100 mmol, 1 equiv), diselenide glycosyl donor (0.75 equiv), 100 °C, 24 h, under air. dppp = 1,3-bis(diphenylphosphino)propane.

The scope of the acylation reaction with dipeptide carbastannatranes 14 was tested using various thioesters derived from C(sp2 ) and C(sp3) carboxylic acids (Scheme 4A). Thioesters represent a compromise between reactivity of the acyl donor, stability, and the ease of preparation. Furthermore, their properties can be matched with the reactivity of the nucleophile by changing the electronics of the thiolate leaving group. However, in our studies we found that thiophenyl group is sufficiently activated to serve as a general acyl donor in all reactions described here.

After surveying several palladium pre-catalysts and phosphine ligands, we found that Pd2(dba)3, JackiePhos, and CuCl are the optimal combination for a broad collection of aryl and alkyl carboxylic acid thioesters. S-Phenyl thioesters 20 were readily converted to the corresponding ketones in moderate to excellent yields whereas S-alkyl thioesters resulted in ~20% lower yields. Aromatic carboxylic acid thioesters such as S-phenyl benzothioate (19a) and S-phenyl thiophene-2-carbothioate (19b) performed well despite the potential issues with catalyst deactivation by the resultant thiophenolate. To our delight, S-phenyl thioesters with alkyl side chains were also viable substrates as demonstrated by a smooth conversion alanine-derived ester (19c), fatty acid donor (19d), or PEG-derived amino acid (19e). Notably, we were unable to detect any loss of stereochemical integrity at the α-position or loss of CO for alkyl and aryl substrates.

The acylation protocol allows for a direct conversion of AlaM into naturally occurring amino acids. For example, a reaction of (2-aminophenyl)acetic acid thioester with the model peptide 14b afforded a metabolite amino acid kynurenine 19f typically introduced into the peptide via ozonolysis of tryptophan.114 Similarly, when 14a was treated with iPrSCOSePh as the acyl electrophile followed by basic hydrolysis (LiOH, H2O), aspartic acid 19g was obtained in 79%. In this reaction the C-Se bond underwent preferential cleavage, and the potentially problematic second activation of the intermediate thioester was suppressed by maintaining a slight excess of the electrophile (1.5 equiv). Furthermore, treatment of thioester intermediate 19i with NH3 in MeOH afforded asparagine 19h. Thioester 19i can be isolated if needed (72%) and can serve as a competent acyl donor for downstream functionalizations. Similarly, N-linked asparagine derivatives can be introduced into peptides if N,N-diaryl thiocarbamates are used (19j).

To further demonstrate the practicality of the AlaSn acylation as a tool for site-selective conjugation, we converted several bioactive small molecule carboxylic acids into thioesters and engaged them in C-C couplings. These reactions included derivatives of ibuprofen (19k), probalan (19l), indometacin (19m), D-homoglucuronic acid (19n), and D-(+)-biotin (19o) used here as examples of functional group compatibility and high chemoselectivity.

C. Inverse (Seleno)Cysteine Arylation and Alkylation.

In the course of the method development, we turned to reactions that give rise to (seleno)cysteine-modified peptides (Scheme 4B). Cysteine arylations have received considerable attention as a means to perform site-selective conjugation complementing thiol alkylations or Michael additions.5, 100,115 In these protocols, the nucleophilic cysteine thiol was modified with organopalladium/organogold reagents,116 boronic acids,117 or diazonium salts.118 As a complementary strategy representing an inverse approach, we envisioned that AlaSn could be used to introduce aryl (seleno)cysteine with redox-neutral electrophiles such as N-sulfenylsuccinimides 22a or diselenides 22b (Scheme 4B). We found that the cross-coupling of AlaSn could be catalyzed by CuCl (50 mol%) with no additional activators since the AlaSn nucleophiles are sufficiently activated to undergo transmetalation. Other Cu(I) sources such as CuBr or CuI were less efficient in promoting this transformation, an observation consistent with our prior work that underscored the importance of the halide counterion. Substrates such as aryl (21a, 21b) and alkyl N-sulfenylsuccinimides (21c) were converted into thioethers at room temperature or selenides at 100 °C. A direct coupling of cysteine N-sulfenylsuccinimide dipeptide generated S-linked lanthionine 21d in 71% without epimerization at the α-carbonyl. This strategy is complementary to the earlier synthetic studies that relied on nucleophilic substitution of β-haloalanine with free cysteine.119 This example further demonstrates that oligopeptides can be efficiently coupled without detrimental formation of Dha that frequently competes with substitutions of β-haloalanine electrophiles. These results led us then to extend the scope of C-heteroatom cross-couplings with symmetrical D-glucose diselenide (21e) and N-sulfenylsuccinimidate donors (21f), resulting in 68% and 78%, respectively, with retention of anomeric configuration for both examples. This strategy, which represents an umpolung approach to glycodiversification, can be used in the preparation of (seleno)cysteine-modified peptides.8890

D. Peptide macrocyclization.

Complementary to intermolecular transformations, we were intrigued by the possibility of engaging AlaSn in cyclizations with properly functionalized electrophiles (Scheme 5A and 5B). Because cyclic peptides are a promising scaffold for the development of drug candidates due to their ability to bind to a wide range of target molecules and proteolytic stability, research in synthetic methodology for peptide cyclization focus on side chain cyclizations and, more recently, biosynthetic engineering.120122 Among those, methods that can selectively connect the aromatic ring in the form cyclophane-type frameworks can facilitate the discovery of novel bioactive compounds.13,123,124 The rigid, planar, and hydrophobic aromatic rings support the cyclic structures, can be fully fitted into the main skeleton of the cyclic peptide molecule, and are amenable to structural modifications. The non-canonical aryl linkers can stabilize secondary structures and promote hydrogen bonding that can be beneficial for optimizing membrane permeability and bioavailability. Inspired by these novel functions of cyclic peptides, we wondered whether our methods could be employed to generate similar structures via intramolecular C(sp3)-C(sp2) reactions. We pursued two cyclization strategies that were dictated by the availably of the electrophilic components and their ease of introduction into a peptide: (a) reactions at the acyl side chains of Asp (23a) and Glu (23b) in the form of a thioester that furnished 7- and 8-membered ketones 24 and 25 in 31–48% yield, and (b) couplings of phenylalanine functionalized with a halogen handle at the ortho- (26a, 26b) and para- (26c) positions leading the formation for 8- (27), 11- (28), and 13- (29) membered rings in 53–60%. The reactions with thioesters represent a rare example of carbonylative cyclization in peptide scaffold 24 and 25 and introduce a novel ketone linker. Similarly, arylation reactions with a phenylalanine electrophile produced a strained para-cyclophane structure 29 formed through a unique cyclization strategy.

Scheme 5. Scope of Late Stage Functionalization.

Scheme 5.

General reaction conditions for intramolecular acylation and aryl cross-coupling: 23 or 26 (0.100 mmol, 1 equiv), Pd2(dba)3 (5.0 mol%), JackiePhos (20 mol%), CuCl (1 equiv), 1,4-dioxane (50.0 mL), isolated yields; a. 90 °C and 48 h were used; b. room temperature and 72 h were used. Reaction conditions for late-stage peptide functionalization: c. 30 (0.010 mmol), aryl bromide (0.100 mmol), Pd2(dba)3 (0.050 mmol), JackiePhos (0.200 mmol), CuCl (0.100 mmol), MeCN:Buffer pH 7.5 (1:1, 2.00 mL), 37 °C, 1 h; d. 30 (0.010 mmol), S-phenyl naphthalene-2-carbothioate (0.100 mmol), Pd2(dba)3 (0.05 mmol), JackiePhos (0.200 mmol), CuCl (0.100 mmol), 1,4-dioxane (1.00 mL), 37 °C, 4 h; e. 30 (0.010 mmol), 1-((4-methoxyphenyl)thio)pyrrolidine-2,5-dione (0.100 mmol), CuCl (0.200 mmol), MeCN:CH2Cl2 (1:1, 2.00 mL), 37 °C, 2 h; f. 31 (0.010 mmol), tert-butyl (2-((2,5-dioxopyrrolidin-1-yl)thio)ethyl)carbamate (0.100 mmol), CuCl (0.200 mmol), 1,4-dioxane:CH2Cl2 (1:1, 2.00 mL), 37 °C, 12 h; g. 30 (0.010 mmol), CuCl (0.100 mmol), 1,2-diphenyldiselenide (0.100 mmol), MeCN:Buffer pH 7.5 (1:1, 2.00 mL), 37 °C, 1 h.

E. Oligopeptide functionalization.

To further demonstrate the utility of all coupling methods in a relevant peptide example, we assembled via automated solid-support peptide synthesis gramicidin S oligopeptide 30 with one position mutated into D-AlaSn (Scheme 5C). This linear peptide was used to compare side-by-side all reactions developed earlier but in a more complex setting. Consistent with the results described earlier, both arylation and acylation reactions provided the C-C coupling products 31a31d in good yields, and heteroaromatic substrates, such as 3-bromoquinoline (31c), can be accomplished in acceptable 45% yield. Notably, low temperature (≤37°C), close to neutral pH buffers were optimal for these reactions. Furthermore, thioetherifications performed in variable yields (31e and 31f, 31–84%), and introduction of selenocysteine proceeded in a somewhat moderate yield (31g, 65%) but in excellent chemoselectivity.

CONCLUSIONS

It is becoming abundantly clear that polarity reversal applied to biomolecule functionalization offers an unprecedented opportunity to access new reactivity and explore novel chemical space. Here, we demonstrated that a stable nucleophile installed at the β-carbon in AlaM can serve as an efficient synthon for divergent synthesis of modified peptides. This strategy capitalizes on transmetalation of primary carbastannatranes embedded in a peptide chain that could be coupled with aryl, acyl, and chalcogen-based electrophiles even at ambient conditions and in aqueous solutions. As we showcased these reactions in the synthesis of several high value structures, late-stage functionalization and cyclization reactions stand out due to their potential to streamline discovery of new biomaterials, therapeutics, and probes. It is also conceivable that the presented collection of methods can be integrated with the emerging technologies in peptide manipulation such as encoded libraries and direct bioconjugation.

EXPERIMENTAL PROCEDURES

Resource availability

Lead contact

Further information and requests for resources should be directed to and will be fulfilled by the lead contact, Maciej Walczak (maciej.walczak@colorado.edu).

Materials availability

All unique reagents generated in this study will be made available on request to the lead contact.

Data and code availability

There is no dataset or code associated with this paper. Full experimental procedures are provided in the supplemental information.

Supplementary Material

1

Bigger picture.

Over the last few years various strategies have emerged to functionalize peptides and proteins by site-selective modifications. While most of these approaches capitalize on inherent nucleophilicity of heteroatoms or conjugate additions to dehydroalanine, reversal of polarity has not been studied extensively. Our work introduces umpolung reagents AlaM in the form of carbastannatrane amino acids that can be engaged in several C-C, C-S, and C-Se bond forming reactions. To demonstrate this powerful approach, we optimized inter- and intramolecular cross-couplings and applied these protocols to oligopeptide substrates. We also highlight the fact that AlaM reagents are compatible with aqueous conditions and operate at room temperature without compromising chemoselectivity and yield. The manuscript describes a conceptual departure from known strategies in peptide functionalization and sets the stage for future work to access peptide-based structures with novel topologies.

Highlights.

  • Alanine with β-carbastannatrane group are umpolung peptide reagents

  • AlaSn reagents undergo chemo- and regioselective C-C, C-S, and C-Se cross-couplings

  • Inter- and intramolecular reactions produced amino acids with unnatural side chains

ACKNOWLEDGEMENTS

This work was supported by National Science Foundation (CAREER Award CHE-1753225) and National Institutes of Health (R21GM138808). We thank Dr. Xuan Wang for the synthesis of 5-chloro-1-aza-5-stannabicyclo[3.3.3]undecane.

Footnotes

Declaration of interests. The authors declare no competing interests.

Publisher's Disclaimer: This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to our customers we are providing this early version of the manuscript. The manuscript will undergo copyediting, typesetting, and review of the resulting proof before it is published in its final form. Please note that during the production process errors may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain.

REFERENCES

  • 1.Craik DJ, Fairlie DP, Liras S, and Price D (2013). The future of peptide-based drugs. Chem. Biol. Drug Design 81, 136–147. [DOI] [PubMed] [Google Scholar]
  • 2.Boutureira O, and Bernardes G.a.J. (2015). Advances in chemical protein modification. Chem. Rev 115, 2174–2195. [DOI] [PubMed] [Google Scholar]
  • 3.Isenegger PG, and Davis BG (2019). Concepts of Catalysis in Site-Selective Protein Modifications. J. Am. Chem. Soc. 141, 8005–8013. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 4.Hoyt EA, Cal PMSD, Oliveira BL, and Bernardes GJL (2019). Contemporary approaches to site-selective protein modification. Nat. Rev. Chem. 3, 147–171. [Google Scholar]
  • 5.Cohen DT, Zhang C, Pentelute BL, and Buchwald SL (2015). An Umpolung Approach for the Chemoselective Arylation of Selenocysteine in Unprotected Peptides. J. Am. Chem. Soc. 137, 9784–9787. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 6.Vinogradova EV, Zhang C, Spokoyny AM, Pentelute BL, and Buchwald SL (2015). Organometallic palladium reagents for cysteine bioconjugation. Nature 526, 687. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 7.Zhang C, Dai P, Vinogradov AA, Gates ZP, and Pentelute BL (2018). Site-Selective Cysteine–Cyclooctyne Conjugation. Angew. Chem. Int. Ed. 57, 6459–6463. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 8.Cohen DT, Zhang C, Fadzen CM, Mijalis AJ, Hie L, Johnson KD, Shriver Z, Plante O, Miller SJ, and Buchwald SL (2019). A chemoselective strategy for late-stage functionalization of complex small molecules with polypeptides and proteins. Nat. Chem. 11, 78–85. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 9.Vantourout JC, Adusumalli SR, Knouse KW, Flood DT, Ramirez A, Padial NM, Istrate A, Maziarz K, deGruyter JN, Merchant RR, et al. (2020). Serine-Selective Bioconjugation. J. Am. Chem. Soc. 142, 17236–17242. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 10.Lee HG, Lautrette G, Pentelute BL, and Buchwald SL (2017). Palladium-mediated arylation of lysine in unprotected peptides. Angew. Chem. Int. Ed. 56, 3177–3181. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 11.Taylor MT, Nelson JE, Suero MG, and Gaunt MJ (2018). A protein functionalization platform based on selective reactions at methionine residues. Nature 562, 563–568. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 12.Koniev O, and Wagner A (2015). Developments and recent advancements in the field of endogenous amino acid selective bond forming reactions for bioconjugation. Chem. Soc. Rev 44, 5495–5551. [DOI] [PubMed] [Google Scholar]
  • 13.Zhang X, Lu G, Sun M, Mahankali M, Ma Y, Zhang M, Hua W, Hu Y, Wang Q, and Chen J (2018). A general strategy for synthesis of cyclophane-braced peptide macrocycles via palladium-catalysed intramolecular sp 3 C− H arylation. Nat. Chem 10, 540–548. [DOI] [PubMed] [Google Scholar]
  • 14.Gruß H, and Sewald N (2020). Late-Stage Diversification of Tryptophan-Derived Biomolecules. Chem. Eur. J 26, 5328–5340. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 15.Ichiishi N, Caldwell JP, Lin M, Zhong W, Zhu X, Streckfuss E, Kim H-Y, Parish CA, and Krska SW (2018). Protecting group free radical C–H trifluoromethylation of peptides. Chem. Sci 9, 4168–4175. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 16.Imiołek M, Karunanithy G, Ng W-L, Baldwin AJ, Gouverneur V.r., and Davis BG (2018). Selective radical trifluoromethylation of native residues in proteins. J. Am. Chem. Soc 140, 1568–1571. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 17.Qin T, Cornella J, Li C, Malins LR, Edwards JT, Kawamura S, Maxwell BD, Eastgate MD, and Baran PS (2016). A general alkyl-alkyl cross-coupling enabled by redox-active esters and alkylzinc reagents. Science 352, 801–805. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 18.Qin T, Malins LR, Edwards JT, Merchant RR, Novak AJ, Zhong JZ, Mills RB, Yan M, Yuan C, and Eastgate MD (2017). Nickel-catalyzed Barton decarboxylation and Giese reactions: a practical take on classic transforms. Angew. Chem. Int. Ed 56, 260–265. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 19.McCarver SJ, Qiao JX, Carpenter J, Borzilleri RM, Poss MA, Eastgate MD, Miller MM, and MacMillan DW (2017). Decarboxylative peptide macrocyclization through photoredox catalysis. Angew. Chem. Int. Ed 56, 728–732. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 20.Malins LR (2018). Decarboxylative couplings as versatile tools for late-stage peptide modifications. Pept. Sci 110, e24049. [DOI] [PubMed] [Google Scholar]
  • 21.Bloom S, Liu C, Kölmel DK, Qiao JX, Zhang Y, Poss MA, Ewing WR, and MacMillan DW (2018). Decarboxylative alkylation for site-selective bioconjugation of native proteins via oxidation potentials. Nat. Chem. 10, 205. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 22.Keskin O, Tuncbag N, and Gursoy A (2016). Predicting protein–protein interactions from the molecular to the proteome level. Chem. Rev. 116, 4884–4909. [DOI] [PubMed] [Google Scholar]
  • 23.Deyle K, Kong X-D, and Heinis C (2017). Phage Selection of Cyclic Peptides for Application in Research and Drug Development. Acc. Chem. Res 50, 1866–1874. [DOI] [PubMed] [Google Scholar]
  • 24.Loktev A, Haberkorn U, and Mier W (2017). Multicyclic Peptides as Scaffolds for the Development of Tumor Targeting Agents. Current medicinal chemistry 24, 2141–2155. [DOI] [PubMed] [Google Scholar]
  • 25.Ong YS, Gao L, Kalesh KA, Yu Z, Wang J, Liu C, Li Y, Sun H, and Lee SS (2017). Recent Advances in Synthesis and Identification of Cyclic Peptides for Bioapplications. Current topics in medicinal chemistry 17, 2302–2318. [DOI] [PubMed] [Google Scholar]
  • 26.Rubin SJS, and Qvit N (2018). Backbone-Cyclized Peptides: A Critical Review. Current topics in medicinal chemistry 18, 526–555. [DOI] [PubMed] [Google Scholar]
  • 27.Vinogradov AA, Yin Y, and Suga H (2019). Macrocyclic Peptides as Drug Candidates: Recent Progress and Remaining Challenges. J. Am. Chem. Soc 141, 4167–4181. [DOI] [PubMed] [Google Scholar]
  • 28.Arnison PG, Bibb MJ, Bierbaum G, Bowers AA, Bugni TS, Bulaj G, Camarero JA, Campopiano DJ, Challis GL, Clardy J, et al. (2013). Ribosomally synthesized and post-translationally modified peptide natural products: overview and recommendations for a universal nomenclature. Nat. Prod. Rep 30, 108–160. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 29.Knerr PJ, and Donk WAv.d. (2012). Discovery, Biosynthesis, and Engineering of Lantipeptides. Ann. Rev. Biochem 81, 479–505. [DOI] [PubMed] [Google Scholar]
  • 30.Yao G, Joswig J-O, Keller BG, and Süssmuth RD (2019). Total Synthesis of the Death Cap Toxin Phalloidin: Atropoisomer Selectivity Explained by Molecular-Dynamics Simulations. Chem. Eur. J 25, 8030–8034. [DOI] [PubMed] [Google Scholar]
  • 31.Blanc A, Todorovic M, and Perrin DM (2019). Solid-phase synthesis of a novel phalloidin analog with on-bead and off-bead actin-binding activity. Chem. Commun 55, 385–388. [DOI] [PubMed] [Google Scholar]
  • 32.Matinkhoo K, Pryyma A, Todorovic M, Patrick BO, and Perrin DM (2018). Synthesis of the Death-Cap Mushroom Toxin α-Amanitin. J. Am. Chem. Soc 140, 6513–6517. [DOI] [PubMed] [Google Scholar]
  • 33.Lutz C, Simon W, Werner-Simon S, Pahl A, and Müller C (2020). Total Synthesis of α- and β-Amanitin. Angew. Chem. Int. Ed. 59, 11390–11393. [DOI] [PubMed] [Google Scholar]
  • 34.Eickhoff H, Jung G, and Rieker A (2001). Oxidative phenol coupling—tyrosine dimers and libraries containing tyrosyl peptide dimers. Tetrahedron 57, 353–364. [Google Scholar]
  • 35.Schimana J, Gebhardt K, Höltzel A, Schmid DG, Süssmuth R, Müller J, Pukall R, and Fiedler HP (2002). Arylomycins A and B, new biaryl-bridged lipopeptide antibiotics produced by Streptomyces sp. Tü 6075. I. Taxonomy, fermentation, isolation and biological activities. J Antibiot (Tokyo) 55, 565–570. [DOI] [PubMed] [Google Scholar]
  • 36.Kulanthaivel P, Kreuzman AJ, Strege MA, Belvo MD, Smitka TA, Clemens M, Swartling JR, Minton KL, Zheng F, Angleton EL, et al. (2004). Novel lipoglycopeptides as inhibitors of bacterial signal peptidase I. The Journal of biological chemistry 279, 36250–36258. [DOI] [PubMed] [Google Scholar]
  • 37.Paetzel M, Goodall JJ, Kania M, Dalbey RE, and Page MG (2004). Crystallographic and biophysical analysis of a bacterial signal peptidase in complex with a lipopeptide-based inhibitor. The Journal of biological chemistry 279, 30781–30790. [DOI] [PubMed] [Google Scholar]
  • 38.Peters DS, Romesberg FE, and Baran PS (2018). Scalable Access to Arylomycins via C–H Functionalization Logic. J. Am. Chem. Soc 140, 2072–2075. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 39.Ben-Lulu M, Gaster E, Libman A, and Pappo D (2020). Synthesis of Biaryl-Bridged Cyclic Peptides via Catalytic Oxidative Cross-Coupling Reactions. Angew. Chem. Int. Ed 59, 4835–4839. [DOI] [PubMed] [Google Scholar]
  • 40.Karas JA, Carter GP, Howden BP, Turner AM, Paulin OKA, Swarbrick JD, Baker MA, Li J, and Velkov T (2020). Structure–Activity Relationships of Daptomycin Lipopeptides. J. Med. Chem [DOI] [PubMed] [Google Scholar]
  • 41.Bernardes GJL, Chalker JM, Errey JC, and Davis BG (2008). Facile Conversion of Cysteine and Alkyl Cysteines to Dehydroalanine on Protein Surfaces: Versatile and Switchable Access to Functionalized Proteins. J. Am. Chem. Soc 130, 5052–5053. [DOI] [PubMed] [Google Scholar]
  • 42.Yang A, Ha S, Ahn J, Kim R, Kim S, Lee Y, Kim J, Söll D, Lee H-Y, and Park H-S (2016). A chemical biology route to site-specific authentic protein modifications. Science 354, 623–626. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 43.Wright TH, Bower BJ, Chalker JM, Bernardes GJL, Wiewiora R, Ng W-L, Raj R, Faulkner S, Vallée MRJ, Phanumartwiwath A, et al. (2016). Posttranslational mutagenesis: A chemical strategy for exploring protein side-chain diversity. Science 354, aag1465. [DOI] [PubMed] [Google Scholar]
  • 44.Josephson B, Fehl C, Isenegger PG, Nadal S, Wright TH, Poh AWJ, Bower BJ, Giltrap AM, Chen L, Batchelor-McAuley C, et al. (2020). Light-driven post-translational installation of reactive protein side chains. Nature 585, 530–537. [DOI] [PubMed] [Google Scholar]
  • 45.Phelan JP, and Ellman JA (2016). Conjugate addition–enantioselective protonation reactions. Beilstein J. Org. Chem. 12, 1203–1228. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 46.Dadová J, Galan SRG, and Davis BG (2018). Synthesis of modified proteins via functionalization of dehydroalanine. Curr. Opin. Chem. Biol. 46, 71–81. [DOI] [PubMed] [Google Scholar]
  • 47.Axon JR, and Beckwith AL (1995). Diastereoselective radical addition to methyleneoxazolidinones: an enantioselective route to α-amino acids. J. Chem. Soc., Chem. Commun, 549–550. [Google Scholar]
  • 48.Aycock R, Vogt D, and Jui NT (2017). A practical and scalable system for heteroaryl amino acid synthesis. Chem. Sci. 8, 7998–8003. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 49.Navarre L, Darses S, and Genet JP (2004). Tandem 1, 4-Addition/Enantioselective Protonation Catalyzed by Rhodium Complexes: Efficient Access to α-Amino Acids. Angew. Chem. Int. Ed. 43, 719–723. [DOI] [PubMed] [Google Scholar]
  • 50.Navarre L, Martinez R, Genet J-P, and Darses S (2008). Access to enantioenriched α-amino esters via rhodium-catalyzed 1, 4-addition/enantioselective protonation. J. Am. Chem. Soc. 130, 6159–6169. [DOI] [PubMed] [Google Scholar]
  • 51.Hargrave JD, Bish G, Köhn GK, and Frost CG (2010). Rhodium-catalysed conjugate addition of arylboronic acids to enantiopure dehydroamino acid derivatives. Org. Biomol. Chem. 8, 5120–5125. [DOI] [PubMed] [Google Scholar]
  • 52.Key HM, and Miller SJ (2017). Site-and stereoselective chemical editing of thiostrepton by Rh-catalyzed conjugate arylation: New analogues and collateral enantioselective synthesis of amino acids. J. Am. Chem. Soc. 139, 15460–15466. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 53.Kieffer ME, Repka LM, and Reisman SE (2012). Enantioselective synthesis of tryptophan derivatives by a tandem Friedel–Crafts conjugate addition/asymmetric protonation reaction. J. Am. Chem. Soc. 134, 5131–5137. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 54.Pracejus H, Wilcke FW, and Hanemann K (1977). Asymmetrisch katalysierte Additionen von Thiolen an α-Aminoacrylsäure-Derivate und Nitroolefine. J. Prakt. Chem. 319, 219–229. [Google Scholar]
  • 55.Leow D, Lin S, Chittimalla SK, Fu X, and Tan CH (2008). Enantioselective protonation catalyzed by a chiral bicyclic guanidine derivative. Angew. Chem. Int. Ed. 47, 5641–5645. [DOI] [PubMed] [Google Scholar]
  • 56.Gong W, Zhang G, Liu T, Giri R, and Yu J-Q (2014). Site-Selective C(sp3)–H Functionalization of Di-, Tri-, and Tetrapeptides at the N-Terminus. J. Am. Chem. Soc. 136, 16940–16946. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 57.He G, Wang B, Nack WA, and Chen G (2016). Syntheses and Transformations of α-Amino Acids via Palladium-Catalyzed Auxiliary-Directed sp3 C–H Functionalization. Acc. Chem. Res. 49, 635–645. [DOI] [PubMed] [Google Scholar]
  • 58.He J, Jiang H, Takise R, Zhu R-Y, Chen G, Dai H-X, Dhar TGM, Shi J, Zhang H, Cheng PTW, and Yu J-Q (2016). Ligand-Promoted Borylation of C(sp3)-H Bonds with Palladium(II) Catalysts. Angew. Chem. Int. Ed. 55, 785–789. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 59.Kenworthy MN, Kilburn JP, and Taylor RJ (2004). Highly functionalized organolithium reagents for enantiomerically pure α-amino acid synthesis. Org. Lett 6, 19–22. [DOI] [PubMed] [Google Scholar]
  • 60.Jackson RF, Wishart N, Wood A, James K, and Wythes MJ (1992). Preparation of enantiomerically pure protected 4-oxo-α-amino acids and 3-aryl-α-amino acids from serine. J. Org. Chem 57, 3397–3404. [Google Scholar]
  • 61.Dexter CS, Jackson RF, and Elliott J (1999). Synthesis of enantiomerically pure β-and γ-amino acid derivatives using functionalized organozinc reagents. J. Org. Chem 64, 7579–7585. [Google Scholar]
  • 62.Collier PN, Campbell AD, Patel I, Raynham TM, and Taylor RJ (2002). Enantiomerically pure α-amino acid synthesis via hydroboration− Suzuki cross-coupling. J. Org. Chem 67, 1802–1815. [DOI] [PubMed] [Google Scholar]
  • 63.Collier PN, Campbell AD, Patel I, and Taylor RJ (2002). Hydroboration–Suzuki cross coupling of unsaturated amino acids; the synthesis of pyrimine derivatives. Tetrahedron 58, 6117–6125. [Google Scholar]
  • 64.Harvey JE, Kenworthy MN, and Taylor RJ (2004). Synthesis of non-proteinogenic phenylalanine analogues by Suzuki cross-coupling of a serine-derived alkyl boronic acid. Tetrahedron Lett. 45, 2467–2471. [Google Scholar]
  • 65.Rilatt I, and Jackson RF (2008). Kinetic Studies on the Stability and Reactivity of β-Amino Alkylzinc Iodides Derived from Amino Acids. J. Org. Chem 73, 8694–8704. [DOI] [PubMed] [Google Scholar]
  • 66.Ross AJ, Lang HL, and Jackson RF (2010). Much improved conditions for the Negishi cross-coupling of iodoalanine derived zinc reagents with aryl halides. J. Org. Chem 75, 245–248. [DOI] [PubMed] [Google Scholar]
  • 67.Castaño AM, and Echavarren AM (1990). Regioselective functionalization of chiral nickelacycles derived from N-protected aspartic and glutamic anhydrides. Tetrahedron Lett. 31, 4783–4786. [Google Scholar]
  • 68.Barfoot CW, Harvey JE, Kenworthy MN, Kilburn JP, Ahmed M, and Taylor RJK (2005). Highly functionalised organolithium and organoboron reagents for the preparation of enantiomerically pure α-amino acids. Tetrahedron 61, 3403–3417. [Google Scholar]
  • 69.Tacke R, Merget M, Bertermann R, Bernd M, Beckers T, and Reissmann T (2000). Syntheses and Properties of Silicon- and Germanium-Containing α-Amino Acids and Peptides:   A Study on C/Si/Ge Bioisosterism. Organometallics 19, 3486–3497. [Google Scholar]
  • 70.Zhou Z-X, Rao W-H, Zeng M-H, and Liu Y-J (2018). Facile synthesis of unnatural β-germyl-α-amino amides via Pd(ii)-catalyzed primary and secondary C(sp3)–H bond germylation. Chem. Commun 54, 14136–14139. [DOI] [PubMed] [Google Scholar]
  • 71.Xu M-Y, Jiang W-T, Li Y, Xu Q-H, Zhou Q-L, Yang S, and Xiao B (2019). Alkyl Carbagermatranes Enable Practical Palladium-Catalyzed sp2–sp3 Cross-Coupling. J. Am. Chem. Soc 141, 7582–7588. [DOI] [PubMed] [Google Scholar]
  • 72.Kinder DH, and Ames MM (1987). Synthesis of 2-amino-3-boronopropionic acid: a boron-containing analog of aspartic acid. J. Org. Chem 52, 2452–2454. [Google Scholar]
  • 73.Hsiao GK, and Hangauer DG (1998). A Facile Synthesis of tert-Butyl 2-[(Benzyloxycarbonyl)amino]-3-(4,4,5,5-tetramethyl-1,3,2-dioxaborolan-2-yl)propionate: An Orthogonally Protected Boronic Acid Analog of Aspartic Acid. Synthesis 1998, 1043–1046. [Google Scholar]
  • 74.Belfaitah A, Isly M, and Carboni B (2004). 1,3-Dipolar cycloadditions of azomethine ylides to alkenylboronic esters. Access to substituted boron analogues of β-proline and 3-hydroxypyrrolidines. Tetrahedron Lett. 45, 1969–1972. [Google Scholar]
  • 75.Reddy VJ, Chandra JS, and Reddy MVR (2007). Concise synthesis of ω-borono-α-amino acids. Org. Biomol. Chem 5, 889–891. [DOI] [PubMed] [Google Scholar]
  • 76.Lee JT, Chen DY, Yang Z, Ramos AD, Hsieh JJD, and Bogyo M (2009). Design, syntheses, and evaluation of Taspase1 inhibitors. Bioorg. Med. Chem. Lett 19, 5086–5090. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 77.López-Pérez A, Segler M, Adrio J, and Carretero JC (2011). Silver-Catalyzed 1,3-Dipolar Cycloaddition of Azomethine Ylides with β-Boryl Acrylates. J. Org. Chem. 76, 1945–1948. [DOI] [PubMed] [Google Scholar]
  • 78.He Z-T, Zhao Y-S, Tian P, Wang C-C, Dong H-Q, and Lin G-Q (2014). Copper-Catalyzed Asymmetric Hydroboration of α-Dehydroamino Acid Derivatives: Facile Synthesis of Chiral β-Hydroxy-α-amino Acids. Org. Lett 16, 1426–1429. [DOI] [PubMed] [Google Scholar]
  • 79.Bartoccini F, Bartolucci S, Lucarini S, and Piersanti G (2015). Synthesis of Boron- and Silicon-Containing Amino Acids through Copper-Catalysed Conjugate Additions to Dehydroalanine Derivatives. Eur. J. Org. Chem 2015, 3352–3360. [Google Scholar]
  • 80.Kubota K, Hayama K, Iwamoto H, and Ito H (2015). Enantioselective Borylative Dearomatization of Indoles through Copper(I) Catalysis. Angew. Chem. Int. Ed 54, 8809–8813. [DOI] [PubMed] [Google Scholar]
  • 81.Xie J-B, Lin S, Qiao S, and Li G (2016). Asymmetric Catalytic Enantio- and Diastereoselective Boron Conjugate Addition Reactions of α-Functionalized α,β-Unsaturated Carbonyl Substrates. Org. Lett 18, 3926–3929. [DOI] [PubMed] [Google Scholar]
  • 82.Bartoccini F, Cannas DM, Fini F, and Piersanti G (2016). Palladium(II)-Catalyzed Cross-Dehydrogenative Coupling (CDC) of N-Phthaloyl Dehydroalanine Esters with Simple Arenes: Stereoselective Synthesis of Z-Dehydrophenylalanine Derivatives. Org. Lett 18, 2762–2765. [DOI] [PubMed] [Google Scholar]
  • 83.Hayama K, Kubota K, Iwamoto H, and Ito H (2017). Copper(I)-catalyzed Diastereoselective Dearomative Carboborylation of Indoles. Chem. Lett 46, 1800–1802. [Google Scholar]
  • 84.Kaldas SJ, Rogova T, Nenajdenko VG, and Yudin AK (2018). Modular Synthesis of β-Amino Boronate Peptidomimetics. J. Org. Chem 83, 7296–7302. [DOI] [PubMed] [Google Scholar]
  • 85.Kim J, Shin M, and Cho SH (2019). Copper-Catalyzed Diastereoselective and Enantioselective Addition of 1,1-Diborylalkanes to Cyclic Ketimines and α-Imino Esters. ACS Catalysis 9, 8503–8508. [Google Scholar]
  • 86.Hayama K, Kojima R, Kubota K, and Ito H (2020). Synthesis of Chiral N-Heterocyclic Allylboronates via the Enantioselective Borylative Dearomatization of Pyrroles. Org. Lett. 22, 739–744. [DOI] [PubMed] [Google Scholar]
  • 87.Zhu F, Rodriguez J, Yang T, Kevlishvili I, Miller E, Yi D, O’Neill S, Rourke MJ, Liu P, and Walczak MA (2017). Glycosyl Cross-Coupling of Anomeric Nucleophiles: Scope, Mechanism, and Applications in the Synthesis of Aryl C-Glycosides. J. Am. Chem. Soc 139, 17908–17922. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 88.Zhu F, Rourke MJ, Yang T, Rodriguez J, and Walczak MA (2016). Highly stereospecific cross-coupling reactions of anomeric stannanes for the synthesis of C-aryl glycosides. J. Am. Chem. Soc. 138, 12049–12052. [DOI] [PubMed] [Google Scholar]
  • 89.Zhu F, Miller E, Zhang S. q., Yi D, O’Neill S, Hong X, and Walczak MA (2018). Stereoretentive C (sp3)–S Cross-Coupling. J. Am. Chem. Soc. 140, 18140–18150. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 90.Zhu F, O’Neill S, Rodriguez J, and Walczak MA (2018). Stereoretentive reactions at the anomeric position: Synthesis of selenoglycosides. Angew. Chem. Int. Ed 57, 7091–7095. [DOI] [PubMed] [Google Scholar]
  • 91.Yang T, Zhu F, and Walczak MA (2018). Stereoselective oxidative glycosylation of anomeric nucleophiles with alcohols and carboxylic acids. Nat. Commun 9, 1–9. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 92.Vedejs E, Haight AR, and Moss WO (1992). Internal coordination at tin promotes selective alkyl transfer in the Stille coupling reaction. J. Am. Chem. Soc 114, 6556–6558. [Google Scholar]
  • 93.Verkade JG (1993). Atranes: new examples with unexpected properties. Acc. Chem. Res 26, 483–489. [Google Scholar]
  • 94.Theddu N, and Vedejs E (2013). Stille Coupling of an Aziridinyl Stannatrane. J. Org. Chem. 78, 5061–5066. [DOI] [PubMed] [Google Scholar]
  • 95.Srivastav N, Singh R, and Kaur V (2015). Carbastannatranes: a powerful coupling mediators in Stille coupling. RSC Adv. 5, 62202–62213. [Google Scholar]
  • 96.Le Grognec E, Chretien J-M, Zammattio F, and Quintard J-P (2015). Methodologies limiting or avoiding contamination by organotin residues in organic synthesis. Chem. Rev 115, 10207–10260. [DOI] [PubMed] [Google Scholar]
  • 97.Wang C-Y, Derosa J, and Biscoe MR (2015). Configurationally stable, enantioenriched organometallic nucleophiles in stereospecific Pd-catalyzed cross-coupling reactions: an alternative approach to asymmetric synthesis. Chem. Sci 6, 5105–5113. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 98.Hicks JD, Hyde AM, Cuezva AM, and Buchwald SL (2009). Pd-Catalyzed N-Arylation of Secondary Acyclic Amides: Catalyst Development, Scope, and Computational Study. J. Am. Chem. Soc. 131, 16720–16734. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 99.Li L, Wang C-Y, Huang R, and Biscoe MR (2013). Stereoretentive Pd-catalysed Stille cross-coupling reactions of secondary alkyl azastannatranes and aryl halides. Nat. Chem 5, 607–612. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 100.Zhang C, Vinogradova EV, Spokoyny AM, Buchwald SL, and Pentelute BL (2019). Arylation Chemistry for Bioconjugation. Angew. Chem. Int. Ed 58, 4810–4839. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 101.Ghosh M, Miller PA, Möllmann U, Claypool WD, Schroeder VA, Wolter WR, Suckow M, Yu H, Li S, Huang W, et al. (2017). Targeted Antibiotic Delivery: Selective Siderophore Conjugation with Daptomycin Confers Potent Activity against Multidrug Resistant Acinetobacter baumannii Both in Vitro and in Vivo. J. Med. Chem 60, 4577–4583. [DOI] [PubMed] [Google Scholar]
  • 102.Kowada T, Maeda H, and Kikuchi K (2015). BODIPY-based probes for the fluorescence imaging of biomolecules in living cells. Chem. Soc. Rev 44, 4953–4972. [DOI] [PubMed] [Google Scholar]
  • 103.Niu L-Y, Guan Y-S, Chen Y-Z, Wu L-Z, Tung C-H, and Yang Q-Z (2012). BODIPY-Based Ratiometric Fluorescent Sensor for Highly Selective Detection of Glutathione over Cysteine and Homocysteine. J. Am. Chem. Soc. 134, 18928–18931. [DOI] [PubMed] [Google Scholar]
  • 104.Wang W, Lorion MM, Martinazzoli O, and Ackermann L (2018). BODIPY Peptide Labeling by Late-Stage C(sp3)−H Activation. Angew. Chem. Int. Ed 57, 10554–10558. [DOI] [PubMed] [Google Scholar]
  • 105.Guin S, Dolui P, Zhang X, Paul S, Singh VK, Pradhan S, Chandrashekar HB, Anjana SS, Paton RS, and Maiti D (2019). Iterative Arylation of Amino Acids and Aliphatic Amines via δ-C(sp3)−H Activation: Experimental and Computational Exploration. Angew. Chem. Int. Ed 58, 5633–5638. [DOI] [PubMed] [Google Scholar]
  • 106.Goodnick P, Dominguez R, DeVane CL, and Bowden C (1998). Bupropion slow-release response in depression: diagnosis and biochemistry. Biol. Psych. 44, 629–632. [DOI] [PubMed] [Google Scholar]
  • 107.Stone TW, and Darlington LG (2002). Endogenous kynurenines as targets for drug discovery and development. Nat. Rev. Drug Discov 1, 609–620. [DOI] [PubMed] [Google Scholar]
  • 108.Tokuyama H, Yokoshima S, Yamashita T, and Fukuyama T (1998). A novel ketone synthesis by a palladium-catalyzed reaction of thiol esters and organozinc reagents. Tetrahedron Lett. 39, 3189–3192. [Google Scholar]
  • 109.De Luca L, Giacomelli G, and Porcheddu A (2001). A Simple Preparation of Ketones. N-Protected α-Amino Ketones from α-Amino Acids. Org. Lett. 3, 1519–1521. [DOI] [PubMed] [Google Scholar]
  • 110.Florjancic AS, and Sheppard GS (2003). A practical synthesis of α-amino ketones via aryllithium addition to N-Boc-α-amino acids. Synthesis 2003, 1653–1656. [Google Scholar]
  • 111.Buckley III TF, and Rapoport H (1981). α-Amino acids as chiral educts for asymmetric products. Amino acylation with N-acylamino acids. J. Am. Chem. Soc 103, 6157–6163. [Google Scholar]
  • 112.Dardir AH, Hazari N, Miller SJ, and Shugrue CR (2019). Palladium-Catalyzed Suzuki–Miyaura Reactions of Aspartic Acid Derived Phenyl Esters. Org. Lett 21, 5762–5766. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 113.Jousseaume T, Wurz NE, and Glorius F (2011). Highly enantioselective synthesis of α-amino acid derivatives by an NHC-catalyzed intermolecular Stetter reaction. Angew. Chem. Int. Ed. 50, 1410–1414. [DOI] [PubMed] [Google Scholar]
  • 114.Wong CTT, Lam HY, and Li X (2013). Effective synthesis of kynurenine-containing peptides via on-resin ozonolysis of tryptophan residues: synthesis of cyclomontanin B. Org. Biomol. Chem 11, 7616–7620. [DOI] [PubMed] [Google Scholar]
  • 115.Cheng W-M, Lu X, Shi J, and Liu L (2018). Selective modification of natural nucleophilic residues in peptides and proteins using arylpalladium complexes. Org. Chem. Front 5, 3186–3193. [Google Scholar]
  • 116.Kung KK-Y, Ko H-M, Cui J-F, Chong H-C, Leung Y-C, and Wong M-K (2014). Cyclometalated gold(III) complexes for chemoselective cysteine modification via ligand controlled C–S bond-forming reductive elimination. Chem. Commun. 50, 11899–11902. [DOI] [PubMed] [Google Scholar]
  • 117.Ball ZT (2019). Protein Substrates for Reaction Discovery: Site-Selective Modification with Boronic Acid Reagents. Acc. Chem. Res 52, 566–575. [DOI] [PubMed] [Google Scholar]
  • 118.Sengupta S, and Chandrasekaran S (2019). Modifications of amino acids using arenediazonium salts. Org. Biomol. Chem 17, 8308–8329. [DOI] [PubMed] [Google Scholar]
  • 119.Denoël T, Lemaire C, and Luxen A (2018). Progress in Lanthionine and Protected Lanthionine Synthesis. Chem. Eur. J. 24, 15421–15441. [DOI] [PubMed] [Google Scholar]
  • 120.Gulder T, and Baran PS (2012). Strained cyclophane natural products: macrocyclization at its limits. Nat. Prod. Rep 29, 899–934. [DOI] [PubMed] [Google Scholar]
  • 121.Gang D, Kim DW, and Park H-S (2018). Cyclic peptides: Promising scaffolds for biopharmaceuticals. Genes 9, 557. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 122.Chow HY, Zhang Y, Matheson E, and Li X (2019). Ligation technologies for the synthesis of cyclic peptides. Chem. Rev 119, 9971–10001. [DOI] [PubMed] [Google Scholar]
  • 123.Tang J, He Y, Chen H, Sheng W, and Wang H (2017). Synthesis of bioactive and stabilized cyclic peptides by macrocyclization using C(sp3)–H activation. Chem. Sci 8, 4565–4570. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 124.Nguyen TQN, Tooh YW, Sugiyama R, Nguyen TPD, Purushothaman M, Leow LC, Hanif K, Yong RHS, Agatha I, Winnerdy FR, et al. (2020). Post-translational formation of strained cyclophanes in bacteria. Nat. Chem. 12, 1042–1053. [DOI] [PubMed] [Google Scholar]

Associated Data

This section collects any data citations, data availability statements, or supplementary materials included in this article.

Supplementary Materials

1

Data Availability Statement

There is no dataset or code associated with this paper. Full experimental procedures are provided in the supplemental information.

RESOURCES