Active ester-based peptide bond formation and its application in peptide synthesis

Jinhua Yang b, Huanan Huang c and Junfeng Zhao *a
aKey Laboratory of Molecular Target & Clinical Pharmacology and the NMPA & State Key Laboratory of Respiratory Disease, School of Pharmaceutical Sciences & The Fifth Affiliated Hospital, Guangzhou Medical University, Guangzhou 511436, P. R. China. E-mail: zhaojf@gzhmu.edu.cn
bCollege of Chemistry and Chemical Engineering, Hubei Key Laboratory of Pollutant Analysis and Reuse Technology, Hubei Normal University, Huangshi 435002, P. R. China
cCollege of Chemistry and Chemical Engineering, Jiujiang University, Jiujiang 332005, P. R. China

Received 24th October 2022 , Accepted 29th December 2022

First published on 12th January 2023


Abstract

Peptide bonds, the amide bonds formed between α-amino acids, constitute the backbone of peptides and proteins. Although peptide bonds can be constructed efficiently under mild reaction conditions in nature, their chemical approach imposes great challenges on chemists. The notorious racemization/epimerization issue caused by the “overactivation” of the carboxylic groups by using conventional coupling reagents always complicated peptide synthesis and purification. Alternatively, active esters with well-defined structures and moderate reactivity, formed in situ or prior prepared, play a crucial role in suppressing racemization/epimerization. Herein, we systematically and critically summarized the pros and cons of using active esters in peptide syntheses with a hope of stimulating novel innovations for disruptive peptide synthesis strategies.


1. Introduction

The fundamental chemistry of protein and peptide synthesis is amide bond formation, which itself is an essential basic reaction of organic chemistry. As a ubiquitous functional group, the amide bond is widely found in polymers, pharmaceuticals, cosmetics, agrochemicals, materials, and other fine chemicals. The Comprehensive Medicinal Chemistry database shows that amide bonds are found in up to 25% of currently marketed drugs and two-thirds of drug candidates.1 Moreover, the acylation of an amine is engaged in approximately 16% of reactions in the synthesis of pharmaceuticals.2 As a consequence, a broad range of amide bond formation strategies, e.g., acyl halides,3 acyl azides,4 and anhydrides,5 have been developed.6–9 However, most of these methods are not amenable to peptide bond formation because of the limited availability of the corresponding chiral starting materials. The most reliable and widely used method for peptide synthesis is still the coupling reagent-mediated dehydration condensation between the carboxylic group of the Nα-protected α-amino acid and the amino group of the other α-amino acid. More than hundreds of coupling reagents have been developed and commercialized over the past decades.10–13 Many thanks to Merrifield for his great contribution of inventing solid-phase peptide synthesis (SPPS), which has revolutionized peptide synthesis and plays a mainstay role for peptide synthesis by combining with coupling reagents and peptide synthesizers. However, SPPS is far from perfect. Particularly, SPPS has been criticized because of its low atom-economy in the context of environmentally sustainable development.14 In addition, extensive reverse-phase high-performance liquid chromatography (HPLC) purification is endlessly necessary because the overactivation of conventional coupling reagents results in epimers and other side products which have similar properties to that of the target peptides. During the past two decades, peptide drug discovery has witnessed a rapid renaissance with innovations in peptide drug formulation and delivery technologies. With the growing request for peptide therapeutics, highly efficient and environmentally friendly peptide manufacturing processes are in great demand. To highlight the importance and urgency of this topic, general methods for sustainable peptide bond formation have twice been identified as one of the top ten challenges in the green chemistry research area of synthetic organic chemistry over the past two decades.15 It is noted that these problems could not be addressed by optimization of current peptide synthesis strategies and technologies, which were developed in the 1950–1980s when the green chemistry concept was not established yet. The loss of chiral integrity of α-amino acids not only requires extensive HPLC purification but also results in the tremendous loss of target peptides, which is an enduring and notorious issue encountered by practitioners. The use of an extra racemization suppressor, which can form a transient active ester intermediate in situ during the coupling process, is an efficient strategy to suppress racemization/epimerization. Although the aminolysis reaction of common esters cannot take place spontaneously,16 active esters with a good leaving group by taking advantage of a thermodynamically favorable process are susceptible to the attack of an amine. In fact, the active ester strategy with structure well-defined stable esters as the key intermediates was an effective method widely used for peptide synthesis before the advent of coupling reagents (Scheme 1).17–20 For example, the first chemical total synthesis of oxytocin was accomplished by employing the p-nitrophenyl (PNP) esters of α-amino acids.21 However, the prior preparation of active esters requires not only extra steps but also a stoichiometric coupling reagent, a hydroxyl nucleophile, and even a racemization suppressor. In terms of step- and atom-economy as well as operational convenience, there is no advantage for the conventional two-step active ester method compared with coupling reagent-mediated one-pot peptide bond formation. Thus, the active ester method gradually faded with the rapid development of coupling reagents, particularly the bifunctional peptide coupling reagents which play dual roles in activating the carboxylic acid and suppressing racemization. However, after decades of optimization, current peptide synthesis strategies with conventional coupling reagents are reaching a very high standard and their inherent limits. In this context, the active ester strategies employing the structure well-defined active esters would provide a promising solution for the predicament of conventional coupling reagent-mediated peptide synthesis if their disadvantages could be addressed. We herein provide a historical overview and critical analysis of the pros and cons of using active esters for peptide syntheses as well as the state-of-the-art of active ester methods, with a hope of inspiring innovations in developing novel active esters for peptide synthesis.
image file: d2qo01686a-s1.tif
Scheme 1 Timeline for peptide synthesis using active ester methods discussed in this review.

The stable active esters with well-defined structures will be the focus of this review. However, to maintain the integrity of the overview of the role played by active esters in peptide synthesis, some transient active esters will also be discussed when necessary. It should be noted that transient active esters involved in conventional coupling reagents, such as carbodiimides,22 phosphonium salts,23 aminium/uronium salts,24 and bifunctional coupling reagents [O-(benzotriazol-1-yl)-1,1,3,3-tetramethyluro-nium hexafluoro-phosphate (HBTU),24O-(7-azabenzotriazol-1-yl)-1,1,3,3-tetramethyl-uronium hexafluorophosphate (HATU),25 benzo-triazol-1-yloxytris(dimethylamino)phosphonium hexafluorophosphate (BOP),26 benzotriazol-1-yloxytri(pyrrolidino) phosphonium hexafluorophosphate (PyBop),27 3-(diethoxyphosphoryloxy)-1,2,3-benzotriazin-4(3H)-one (DEPBT),28 diphenylphosphoryl azide (DPPA),29,30etc.], mediated peptide bond formation, native chemical ligation (NCL),31–35 imine/hemiaminal-type aldehyde capture ligation,36 Ser/Thr ligation,37,38 peptide thioester,39–44 and peptide hydrazides,45,46 which have been extensively reviewed elsewhere, will not be covered. In addition, enzyme-catalyzed peptide syntheses by employing amino/peptide esters are also beyond the scope of this review.47–50 Interested readers are referred to several excellent reviews on these exciting topics.3,6,10,11,33,51,52 We will focus on the stable active esters with an emphasis on recently developed novel active esters for peptide bond formation. The active esters will be divided into three typical categories in terms of their preparation strategies, i.e., formed in situ with the assistance of a catalyst, synthesized from carboxylic acids and nucleophilic hydroxyl components in the presence of activating reagents, and formed spontaneously from addition reactions of coupling reagents and carboxylic acids directly.

2. Transient active esters

Protein biosyntheses in ribosomes involve enzyme-catalyzed aminolysis of the aminoacyl-tRNA (the tRNA ester of proteinogenic amino acids) for peptide elongation (Scheme 2). Similar to nature, organic chemists have also taken the aminolysis of esters as an effective strategy for synthesizing amides and peptides,53 and a variety of active esters have been exploited for peptide bond formation because of their moderate reactivity, which is beneficial to avoid racemization/epimerization and other undesired side reactions.
image file: d2qo01686a-s2.tif
Scheme 2 Protein synthesis in ribosomes.

2.1 Common esters activated in situ by a Lewis acid or base catalyst

Alkyl esters are commonly inert and can only be attacked by strong nucleophiles like hydroxylamine and hydrazine.16 Until 1902, in his seminal contribution to peptide synthesis, Fischer reported that the aminolysis of ethoxycarbonyl diglycine ethyl ester (Eoc-Gly-Gly-OEt) with leucine ethyl ester (H-Leu-OEt) could be adopted for Eoc-Gly-Gly-Leu-OEt synthesis (Scheme 3) under heating conditions.54 Nevertheless, low reactivity toward aminolysis restricted their practical application. Catalyst or heating is crucial to promote the aminolysis of Nα-protected amino acid esters.
image file: d2qo01686a-s3.tif
Scheme 3 Aminolysis of ethyl ester for peptide bond formation.

A series of catalysts have been developed to facilitate the aminolysis of alkyl esters. Transition metal complexes were initially chosen as well-designed catalysts to promote the aminolysis of α-amino acid esters.55–57 In 1967, Buckingham58 and Collman59 independently reported cobalt(III) complex-catalyzed peptide bond formation via the aminolysis of amino acid esters. Subsequently, Nakahara et al. discovered a copper(II) complex-catalyzed peptide bond formation reaction.60 In 1970, Yamada et al. reported another copper(II)-catalyzed peptide bond-forming reaction.61 Unfortunately, such copper(II) catalysis resulted in the polymerization of amino esters and gave peptide mixtures (Scheme 4).


image file: d2qo01686a-s4.tif
Scheme 4 Cobalt(III) and copper(II)-catalyzed peptide formation.

In 2012, Ohshima and Mashima reported that NaOMe was an efficient catalyst for peptide bond formation. However, an extra addition of acidic alcohol or phenol was required to alleviate the basic reaction conditions, which caused severe racemization. Furthermore, a high reaction temperature and prolonged reaction time restricted its application in peptide synthesis (Scheme 5).62


image file: d2qo01686a-s5.tif
Scheme 5 Peptide coupling reaction catalyzed by NaOMe (adapted with permission from ref. 62, Chem. Commun., 2012, 48, 5434. Copyright (2012) Royal Society of Chemistry).

In 2016, Tsuji and Yamamoto developed a novel strategy for the chemoselective synthesis of amides using a tantalum-catalyzed hydroxy-directed amidation of β-hydroxycarboxylic acid esters (Scheme 6).63 The β-hydroxy group initiated the site-selective activation of the carbonyl group via coordinating with the tantalum catalyst. The reaction proceeded with high chemoselectivity, obtaining the corresponding β-hydroxyamides in moderate to high yields. This strategy could also be extended to the synthesis of dipeptide derivatives. However, a hydroxy directing group at the β-position of the carbonyl group is necessary. In this regard, this method is limited to specific amino acids such as threonine (Thr) and serine (Ser). To circumvent such limitations, an oxime functionality was introduced to the α-position of the ester group. By doing so, the hydroxy-directed strategy has evolved into a solvent-free amidation of N-hydroxyimino esters by using a niobium catalyst (Scheme 7).64 Theoretically, the following reduction of the C–N double bond and consequent cleavage of the N–OH bond would offer peptide with a free N-terminal amino group that could be used for the next peptide bond formation. Unfortunately, the unstable N-hydroxyimino esters were immediately transformed into stable isoxazolones.64 Afterwards, the third-generation Lewis-acid catalyzed substrate-directed peptide bond formation was reported to resolve the limited substrate scope. The carboxyl groups of methyl ester were “remotely” activated via coordination between the protecting group and the Ta catalyst (Scheme 8, eqn (1) and (2)).65–67 This reaction system has a broad functional group tolerance and is amenable not only for α-amino acid esters but also β-homoamino acid esters.


image file: d2qo01686a-s6.tif
Scheme 6 Hydroxy-directed synthesis of dipeptides (adapted with permission from ref. 63, J. Am. Chem. Soc., 2016, 138, 14218. Copyright (2016) American Chemical Society).

image file: d2qo01686a-s7.tif
Scheme 7 N-Hydroxyimino ester-directed synthesis of peptide derivatives (adapted with permission from ref. 64, ACS Catal., 2018, 8, 2181. Copyright (2018) American Chemical Society).

image file: d2qo01686a-s8.tif
Scheme 8 Substrate-directed Lewis-acid-catalyzed peptide synthesis.

2.2 Alkyl ester intermediate bearing electron-withdrawing group

In 2014, Nguyen and Lyons established a novel tropylium-based reagent to activate carboxylic acids for amidation via the dearomatization/rearomatization of 1,1-dichlorocycloheptatriene (TropCl2) systems.68 The TropCl2 could conveniently and reversibly convert to the positively charged aromatic system, which could react with the carboxylic acid to offer an active ester (Scheme 9).68,69 Aminolysis of such tropylium esters furnished amides or peptides upon releasing a better leaving group tropone. The catalytic version of this strategy has also been achieved by employing stoichiometric oxalyl chloride which regenerated the TropCl2in situ. Although a low level of racemization was observed when it was used for peptide bond formation, its further application in long peptide synthesis deserved further exploration.
image file: d2qo01686a-s9.tif
Scheme 9 TropCl2-mediated peptide formation.

A 4-dimethylamino-pyridine (DMAP)-alkyl halide charge-transfer complex was designed by Eichen and Szpilman group as a novel coupling reagent.70 This in situ formed reagent reacted with carboxylic acids to afford the hemiaminal active esters (Scheme 10), which could couple with the amines or amino acid esters to form amides or peptides, respectively.70 The reaction proceeds smoothly with common Nα-protected amino acids without any detectable racemization/epimerization. A series of dipeptides and tripeptides have been constructed using this strategy. However, the loading of DMAP (10 equiv.) and CBrCl3 (10 equiv.) for the generation of the charge-transfer complex is relatively high.


image file: d2qo01686a-s10.tif
Scheme 10 Peptide synthesis using hemiaminal active esters (adapted with permission from ref. 70, Angew. Chem., Int. Ed., 2021, 60, 12406. Copyright (2021) Wiley-VCH).

Very recently, Wang and Sun have innovatively introduced a recyclable coupling reagent 5-nitro-4,6-dithiocyanatopyrimidine (NDTP) for amide and peptide bond formation under mild conditions.71 The mechanism indicated that NDTP activated the carboxylic acid via acyl thiocyanide, which could be viewed as an active thioester bearing an electron-withdrawing group (EWG). Various racemization-free dipeptides could be prepared from protected amino acids and methyl esters of amino acid hydrochloride by using NDTP as the coupling reagent (Scheme 11). The subsequent experiment indicated that NDTP was practical as well for SPPS. This novel coupling reagent might open a new avenue for peptide synthesis.


image file: d2qo01686a-s11.tif
Scheme 11 NDTP-mediated peptide synthesis (adapted with permission from ref. 71, Org. Lett., 2022, 24, 1169. Copyright (2022) American Chemical Society).

2.3 Silicon reagent-promoted peptide bond formation

In 2011, Liebeskind and coworkers reported a novel approach for in situ activation of thioacids via the silatropic switch process. The proposed mechanism indicated that the sulfur atom of the thioacid was kinetically trapped by the trimethylsilyl group to form the S-trimethylsilylthiol esters, which could spontaneously transform into O-silylthionoesters via a thermodynamically driven tautomerization. The silylation of thioacid was accomplished smoothly by treating thioacid with bistrimethylsilylacetamide (BSA). Interestingly, the aminolysis of the O-silylthionoesters afforded an amide bond exclusively rather than a thioamide bond (Scheme 12). It could be used for peptide bond formation with little racemization, which was positively influenced by the nature of the silylating agent.72
image file: d2qo01686a-s12.tif
Scheme 12 BSA-mediated peptide synthesis.

Based on Lewis acid-catalyzed peptide bond formation, a silicon-mediated approach to peptides via an in situ-generated silyl ester was reported by Yamamoto and Muramatsu to circumvent the use of rare and expensive metals.73 The HSi[OCH(CF3)2]3 reacted with N-protected amino acids to afford the active silyl esters in situ. Such esters then reacted with the amino acid esters or N-silylated amino acid esters to furnish the peptides (Scheme 13).73 This strategy has also been extended to the unprotected or partially protected amino acids using two silylating reagents with different reactivities (Scheme 14).74 In such a one-pot peptide bond formation strategy, dipeptides and tripeptides could be obtained smoothly using unprotected amino acids as the acyl donor. Recently, a novel type of peptide bond formation utilizing imidazolylsilane derivatives and unprotected amino acids has been reported by Hattori and Yamamoto (Scheme 15).75 The diimidazolylsilane derivatives are effective reagents to construct five-membered ring active esters, which react with amino acid tert-butyl (tBu) esters smoothly to offer the desired dipeptides.


image file: d2qo01686a-s13.tif
Scheme 13 Hydrosilane-mediated peptide synthesis (adapted with permission from ref. 73, ACS Catal., 2020, 10, 9594. Copyright (2020) American Chemical Society).

image file: d2qo01686a-s14.tif
Scheme 14 Peptide synthesis by transient masking with silylating reagents (adapted with permission from ref. 74, J. Am. Chem. Soc., 2021, 143, 6792. Copyright (2021) American Chemical Society).

image file: d2qo01686a-s15.tif
Scheme 15 Diimidazolylsilane derivative-mediated peptide synthesis (adapted with permission from ref. 75, J. Am. Chem. Soc., 2022, 144, 1758. Copyright (2022) American Chemical Society).

Furthermore, deprotection of the tBu ester at the C-terminus affords silicon-containing fused cyclic dipeptides. Remarkably, these fused cyclic dipeptides are sufficiently stable to be isolated with silica gel chromatography and handled under atmospheric conditions. In addition, the fused cyclic dipeptides turn out to be perfect active esters to react with both nucleophiles (elongation at the C-terminus) and electrophiles (elongation at the N-terminus). Besides unnecessary purification of the in situ formed silyl active ester intermediates, the silyl active ester strategies possess other features, including high functional group tolerance, convenient operation, and offering peptides with high optical purity. In addition, by employing unprotected amino acids as the starting material, these silyl active ester strategies merged the protection and activation of amino acids in one operation, and thus shortened the tedious procedure and remarkably reduced chemical waste as well as the time, energy, chemicals, and solvents, making this protocol a potential paradigm for bioactive peptide synthesis. However, the application of this method for larger peptide syntheses and SPPS needs to be further explored.

2.4 Boron ester promoted peptide bond formation

The first boron reagent for promoting the coupling of Nα-protected amino acid and α-amino ester for peptide bond formation was reported in 1970.76 However, the efficient boronic acid catalyzed amidation was disclosed in 1996 by Yamamoto.77 The original mechanistic proposal suggested that the carboxylic acid was activated by boronic acid via an in situ-formed “monomeric” mixed anhydride (heteroatom active ester) (Scheme 16). Later, Whiting78 and Ishihara79 revised the mechanism and suggested a feasible “diboroxane” mixed anhydride intermediate based on spectroscopic observations. After the first report of arylboronic acid-catalyzed amide bond formation, various modified aromatic aryl boronic acids, including (2-(thiophen-2-ylmethyl)phenyl)boronic acid,80 5-methoxy-2-iodophenylboronic acid,81 and 3,5-bis(trifluoromethyl)phenyl-boronic acid,79 have been developed for the catalytic peptide syntheses (Scheme 16).6,9,82 Although important progress has been achieved in the boronic acid-catalyzed peptide synthesis, the reactions still suffer from inefficient catalysis. Recently, Takemoto and co-workers reported that the formation of the “diboroxane” mixed anhydride intermediate might be hampered, preventing the further development of aryl boronic acids as catalysts for peptide bond formation.83 To circumvent this dilemma, gem-diboronic acid (gem-DBA, Scheme 17)83 and diboronic acid anhydride (DBAA)84 with a bidentate B–O–B skeleton activation of α-amino acids were designed as peptide condensation catalysts. These reagents are compatible with various functionalized α-amino acids, albeit with a slight racemization.
image file: d2qo01686a-s16.tif
Scheme 16 Boron acid-catalyzed peptide bond formation.

image file: d2qo01686a-s17.tif
Scheme 17 gem-DBA-mediated peptide synthesis (adapted with permission from ref. 83, ACS Catal., 2019, 10, 683. Copyright (2019) American Chemical Society).

2.5 Organophosphorus ester intermediate-promoted peptide bond formation

Phosphorus species have been widely adopted as coupling reagents since the systematic study by Castro, who first introduced BOP as a peptide coupling reagent.26 In seeking new coupling reagents, Ni and co-workers recently reported an efficient racemization-free amidation and esterification reaction by employing a Ph3PO(cat.)/(COCl)2 system.85 The in situ-generated acyl phosphonium salt served as an active ester to afford amides and esters in the presence of the corresponding amines and alcohols (Scheme 18). The reaction was easy to handle and could be used to synthesize dipeptides without racemization. The success of the 100 mmol reaction showcased a great potential for scale application. Thanks to the recent advancements in photoredox catalysis, a photocatalytic redox condensation strategy for amide/peptide synthesis was designed by Zhao and co-workers using a trivalent phosphine (PPh3) as the reductant.86 During the reaction, PPh3 was transformed into an acyloxyphosphonium species, which was readily attacked by amines (Scheme 19). The compatibility with functionalized amino acids and assembly of peptides through SPPS, especially the utility of a continuous-flow photoreactor, illustrated its potential application.
image file: d2qo01686a-s18.tif
Scheme 18 Ph3PO(cat.)/(COCl)2-mediated peptide synthesis (adapted with permission from ref. 85, Org. Lett., 2021, 23, 7497. Copyright (2021) American Chemical Society).

image file: d2qo01686a-s19.tif
Scheme 19 Photocatalytic redox condensation for peptide synthesis (adapted with permission from ref. 86, Angew. Chem., Int. Ed., 2022, 61, e202112668. Copyright (2022) Wiley-VCH).

3 The prior prepared active esters

Although unactivated alkyl esters of Nα-protected amino acids could be employed for peptide bond formation via in situ activation by catalysts or additives, these methods continue to face serious problems, including thermodynamic and kinetic barriers. In addition, the scope of these methods has been mostly limited to simple dipeptide or tripeptide substrates. Thus, it is necessary to explore their applications in the synthesis of larger peptides, especially in SPPS. Current mainstream peptide syntheses are accomplished by converting the –OH groups of the amino acids into better leaving groups, namely activated carboxylic acid derivatives. Although hundreds of activating reagents have been developed to realize this purpose, these activated carboxylic acid derivatives tend to undergo undesired side reactions such as racemization/epimerization due to their overactivation.3,6,7,10,11,51 Therefore, additives are always necessary to convert the higher active intermediate into the mild acyl donors, whereby racemization/epimerization could be suppressed or avoided (Scheme 20). Among them, several strong hydroxyl nucleophiles were identified as effective additives to convert the highly reactive acyl donor species into active esters in situ or prior prepared. The classical prior prepared active esters, including EWG-substituted alkyl esters, lactones, phenyl ester derivatives, and hydroxamic active esters, have been widely used for peptide synthesis because they offered an opportunity for addressing the notorious racemization/epimerization issue.
image file: d2qo01686a-s20.tif
Scheme 20 Prior prepared active ester.

3.1 Alkyl esters activated by electron-withdrawing substituents

Once an EWG is incorporated into an alkyl moiety, the alkyl esters will be activated and can be attacked by amine to execute the amide bond formation. The acylation efficiency of methyl esters substituted with one or more EWGs was thoroughly examined by Schwyzer et al.,87 among which cyanomethyl ester was identified as an effective acylating reagent for peptide synthesis (Scheme 21).87–90 Although the racemization could be avoided using the cyanomethyl ester strategy,90 their applications in peptide manufacture are limited because of their low reactivity.91
image file: d2qo01686a-s21.tif
Scheme 21 Peptide synthesis using cyanomethyl esters.

3.2 Lactones

β-Lactones formed from the intramolecular esterification of the β-hydroxy carboxylic acid are more reactive than common alkyl esters because of the strain of the four-membered ring. Thus, Ser-β-lactone has been used to prepare dipeptides (Scheme 22, eqn (1)).92 It is known that the sulfonium derivatives of methionine easily undergo elimination of the MeS-functional group with concomitant formation of homoserine (Hse)-γ-lactone.92 Using this method, γ-lactones have been prepared by cyanogen bromide cleavage with methionine-containing peptides.93 However, owing to their low reactivity toward nucleophiles, ring-opening of the γ-lactones was only observed in the intramolecular acylation case (Scheme 22, eqn (2)).94 Bifunctional reagents have been employed to construct five-membered heterocyclic lactones not only for protecting the amino group but also for activating the carboxyl group.95 Such lactones can be attacked by a sort of nucleophile, in particular the amino group, to construct an amide bond. Among these bifunctional reagents, hexafluoroacetone (HFA) derivatives represented the most successful one for constructing tailor-made building blocks for peptide, depsipeptide, and diketopiperazine syntheses (Scheme 23).96–99 An impressive HFA strategy was reported to give a two-step synthesis of the sweetener aspartame, which typically took at least four steps by conventional methods (Scheme 23).100 Similar to the silyl active ester strategies, this method avoided the protection and deprotection steps, which were necessary for conventional peptide synthesis strategies and thus were promising for developing atom-economical peptide synthesis strategies. However, the elongation of peptide chains using HFA-amino acids was accompanied by the formation of undesired diketopiperazines (DKPs).98
image file: d2qo01686a-s22.tif
Scheme 22 Peptide synthesis using the lactone method.

image file: d2qo01686a-s23.tif
Scheme 23 Examples of the application of HFA derivatives.

3.3 Active aryl esters

A systematic study of the aminolysis of esters RCOOR′ indicated that the reactivity and electronic effects of the R and R′ groups are densely interwoven.101 In light of this, attention has been paid to ester derivatives with π-electron systems, which displayed greater reactivity than alkyl esters by several magnitudes. The first attempt to boost ester reactivity was realized by using thiophenyl esters (Scheme 24).102 The amino acid thioesters and sodium salt of unprotected amino acids were boiled in methanol for several hours to furnish the sodium salt of the dipeptides with concurrently release of the thiophenol. Inspired by this, various electron-deficient thiophenyl esters and other chalcogen-substituted phenol esters were studied.103–107 Such modifications resulted in higher aminolysis reactivity than that of the corresponding phenyl esters. However, the released chalcogen-substituted phenol derivatives such as thiophenols are unpleasant to handle because of their malodorous smell.
image file: d2qo01686a-s24.tif
Scheme 24 Use of amino acid thioester for peptide synthesis.

It was soon recognized that the aromatic ring rather than the chalcogen atom is responsible for activation. Bodanszky embarked on a systematic investigation of substituted phenyl esters, in which electron-withdrawing substituents were installed in the phenyl group to further increase its leaving ability.108 Preliminary studies involved four nitrophenol esters with different substitution styles, whose reactivities were far greater than that of the phenyl ester and even the thiophenyl ester (Scheme 25, 25-1–25-4).103,108 The 2,4-dinitrophenyl esters (Scheme 25, 25–4) provide excessive reactivity, rendering themselves susceptible to hydrolysis. Compared with acyl chlorides, mixed anhydrides, acid azides, and coupling reagent-mediated amide bond formation, nitrophenyl esters could efficiently avoid racemization/epimerization and other undesired side reactions. Meanwhile, the stability of the nitrophenyl esters allowed for preparation and storage on a large scale. These active esters proved to be preferable agents for a stepwise approach to peptides in the 1950s. p-Nitrophenyl (PNP) esters of proteinogenic amino acids were identified as the optimal building blocks for synthesizing long peptide chains such as oxytocin (Scheme 26),21,109,110 gramicidin-S,111 secretin,112 and the B-chain of insulin in practice.113


image file: d2qo01686a-s25.tif
Scheme 25 Active aryl esters for peptide bond formation.

image file: d2qo01686a-s26.tif
Scheme 26 Total synthesis of oxytocin using PNP esters.

The positive feedback of using electron-deficient phenyl esters to enhance the reactivity of aryl esters stimulated further investigations on aryl esters carrying various electron-withdrawing substituents. Boissonnas and co-workers114 employed the more electron-deficient phenyl esters such as 2,4,6-trichloro- and pentachlorophenyl (PCP) esters (Scheme 25, 25-5, 25-6), which were highly compatible with N-benzyloxycarbonyl (Cbz) and tBu protecting groups for peptide synthesis. Tripeptides could also be constructed via chain elongation from either the N- or C-terminus using PCP esters, further highlighting the flexibility of the active ester strategy (Scheme 27).114–116


image file: d2qo01686a-s27.tif
Scheme 27 Synthesis of a tripeptide using PCP esters.

Meanwhile, the impact of halogen substituents on the substituted phenol esters was explored. Pless and Boissonnas conducted an extensive investigation regarding the reaction rates and dissociation constants of diversely substituted phenols.114 Their study suggested that the pKa of phenol derivatives did impact the reactivity of the aryl esters. Moreover, the steric hindrance caused by substituents offset the electron-withdrawing effects. The steric hindrance effects of the phenyl esters are ranked in the order of Cl < Br < I. The reactivity of monochloro aryl esters with different substitution positions on the phenyl ring is ranked in the order of p < m < o. Regardless of the steric hindrance effects, the electronic effects of multiple halogen substituents are superimposed to enhance the electrophilicity of the ester carbonyl group. This investigation led to the identification of 2,4,5-trichlorophenyl esters (Scheme 25, 25-7), which resolved the steric issues of the 2,4,6-trichloro- and PCP esters as the optimal active ester for peptide synthesis. However, it remains limited in practice due to its slow ammonolysis rate, especially in SPPS.114,117 A variety of aryl esters have been examined to gain additional insight,118 among which the pentafluorophenyl (PFP) esters (Scheme 25, 25-8) appeared to be of significant advance.119–121 The van der Waals radius of fluorine is smaller than that of chlorine and slightly larger than that of hydrogen, yet its electronegativity far surpasses chlorine and hydrogen atoms. Given a better electron-withdrawing effect delivered by the fluorine atom without causing bulkiness, PFP esters are efficient even in a crowded environment.120,121 The advances of these active esters in solution-phase peptide synthesis prompted chemists to explore their applicability in SPPS.122 A decapeptide ACP(65–74) and a dodecapeptide were prepared via SPPS with Fmoc-amino-acid PFP esters as the building blocks (Scheme 28).123,124


image file: d2qo01686a-s28.tif
Scheme 28 Solid phase synthesis of ACP(65–74) using PFP esters.

Similar enhancements in reactivity are achievable using EWGs, other than nitro groups and halogens, such as cyano (Scheme 25, 25-9)125 and sulfonyl groups (Scheme 25, 25-10).126,127 Aromatic heterocycles had also proved to be viable substitutes for the phenyl ring of phenol ester. The electronegative nitrogen and oxygen atoms in these heteroaromatic rings may diminish the electron density at the ester carbonyl group, thus prompting its reactivity toward nucleophiles. The esters of 5-hydroxypyrazole (Scheme 25, 25-11)128 and 2-hydroxypyridine (Scheme 25, 25-12)129 had been employed in peptide syntheses. Although the application of active aryl esters in peptide syntheses prospered for a long time and most aryl esters could be purified and stored, and PFP esters were even commercially available, the preparation of these esters using standard ester-formation strategies such as dicyclohexylcarbodiimide (DCC)-mediated coupling was tedious. Thus, the aryl active esters were only used in some special cases. The in situ-formed benzyne trapping strategy may mitigate this dilemma in the future.130

3.4 Hydroxamic active esters

In 1961, a well-known type of active ester was proposed by Nefkens and Tesser, in which N-hydroxyphthalimide rather than the phenol derivatives was used as the hydroxyl component of the active ester (Scheme 29, 29-1).131 The aminolysis of N-hydroxyphthalimide esters took place rapidly. Being insoluble in water, the released N-hydroxyphthalimide can be removed by a basic solution treatment. However, serious racemization makes this method inapplicable for peptide segment condensation.132 More robust N-hydroxy-succinimide (HOSu) esters (Scheme 29, 29-2) were soon designed to mitigate such defects.133 The HOSu esters of proteinogenic amino acids displayed satisfactory properties in peptide synthesis owing to their higher reactivity, easy preparation, high stability against hydrolysis, and compatibility with aqueous reaction conditions. Furthermore, the released HOSu is water-soluble and can be straightforwardly removed at the purification stage (Scheme 30).134 The HOSu esters have also been widely adopted to modify lysine residues in proteins135 and living cells.136 Despite their extraordinary potential, Gross and Bilk's study demonstrated that these esters suffered from a side reaction through Lossen rearrangement (Scheme 31).11,137–140 To address this issue, Fujino et al. introduced the N-hydroxy-5-norbornene-endo-2,3-dicarboxyimide (HONB) esters (Scheme 29, 29-3) containing a rigid backbone, which is beneficial for suppressing the formation of the β-alanine-type by-product. The successful application of HONB esters in various peptide syntheses prompted the synthesis of porcine or bovine luteinizing hormone-releasing hormone (LH-RH, Scheme 32).141 This strategy showcased a high degree of flexibility. Both C to N and N to C elongation strategies are compatible.141
image file: d2qo01686a-s29.tif
Scheme 29 Hydroxamic active esters used in peptide syntheses.

image file: d2qo01686a-s30.tif
Scheme 30 Peptide synthesis using the HOSu ester method.

image file: d2qo01686a-s31.tif
Scheme 31 The mechanism for the side reaction during the Fmoc protection of amino acids with Fmoc-OSu (adapted with permission from ref. 11, Chem. Rev., 2011, 111, 6557. Copyright (2011) American Chemical Society).

image file: d2qo01686a-s32.tif
Scheme 32 Synthesis of LH-RH using HONB esters.

The former works focused on the N,N-diacylhydroxylamine derivatives, but these active esters typically suffered from poor solubility, side reactions, and tedious preparation. The O-acyl hydroxylamine esters were recommended as potential substitutes to circumvent these issues. The investigation of O-acyl hydroxylamine derivatives indicated that the neighbor nitrogen atom enables the hydroxyl group to be readily acylated; more importantly, the electronegativity of the nitrogen atom could effectively activate the acylated derivatives as well.142,143 Therefore, N-monoacylhydroxyl-amine was employed to construct the activated ester of N-protected amino acids. 2-Pyridone derivatives prepared by Paquette were the first batch of N-monoacylhydroxylamine active esters, which coupled with amino esters to generate dipeptides (Scheme 29, 29-4).144 In 1965, Bittner and co-workers reported a kind of novel activated amino acid ester via the esterification of N,N-diethylhydroxylamine and N-carboxy-α-amino acid anhydrides (Scheme 29, 29-5).145

However, the reactivity of this kind of ester was relatively low, “some days” were required for completing the aminolysis reaction. Coincidentally, Young et al. developed similar 1-hydroxypiperidine (PIP) esters for peptide synthesis (Scheme 29, 29-6),146 which were stable intermediates and could be activated by adding a certain amount of acetic acid. In this case, the N,N-dialkylhydroxylamine esters afforded acidic conditions for peptide synthesis, avoiding the risk of base-induced racemization/epimerization (Scheme 33).147 In addition, the aminolysis of PIP esters is revealed to be accelerated by the neighboring nitrogen atom, which can form a hydrogen bond with the incoming amine and accept a proton to boost the subsequent aminolysis step.148


image file: d2qo01686a-s33.tif
Scheme 33 Synthesis of peptides using PIP esters.

In 1970, more than 30 other N-hydroxy amino acid esters were prepared by König and Geiger, among which the amino acid esters of 1-hydroxybenzotriazole (HOBt) were favorable acylating and racemization-resistant active esters (Scheme 29, 29-7).149 The preformed HOBt esters displayed a striking advance in suppressing racemization during peptide fragment condensation (FC) in solution.149,150 Later, the preactivation of peptide fragments in the presence of HOBt with DCC as the coupling reagent was successfully expanded to solid-phase fragment condensation (SPFC);151 however, a flaw of the DCC/HOBt method is that the excess activated peptide fragments cannot be recovered unchanged. Racemization/epimerization contaminates unreacted and potentially valuable excess fragments.151 Shortly thereafter, different HOBt derivatives were introduced.152–155 Those 1-hydroxy-7-azabenzotriazole-based (HOAt, Scheme 29, 29-8) active esters were recognized to be superior to others, and accelerated coupling reactions with reduced loss of chiral integrity in both solution- and solid-phase peptide synthesis.156 The mechanistic study suggests that the nitrogen atom located at the 7-position of the 7-aza benzotriazole improves its leaving ability by taking advantage of the neighboring group effect, which is generally acknowledged to enhance reactivity and reduce the risk of losing chiral integrity.153 Though HOBt and its derivatives have been used widely as racemization suppressors or peptide coupling reagents, potentially explosive properties limit their applications.157

3,4-Dihydro-3-hydroxy-4-oxo-1,2,3-benzotriazine (HODhbt), a six-membered analogous of HOBt, is not only a superior racemization suppressor but also a better additive to improve the coupling efficacy (Scheme 29, 29-9).149 It had also been used in SPPS (Scheme 34).158–160 Inspired by the earlier studies of HOAt, Carpino and co-workers synthesized aza-derivatives (HODhat and HODhad) of HODhbt (Scheme 29, 29-10, 29-11).161 Their experimental results showed that an unexpected peptide chain termination occurred during the peptide synthesis, which might be attributed to o-azidobenzoic acid ester, a by-product produced in HODhbt ester preparation, although HODhat esters are slightly more reactive than the HOAt esters.161 To reduce the impact of this by-product, the DCC used in the preparation of these esters has to be controlled rigorously, and the crude products should be recrystallized carefully.158,159,162


image file: d2qo01686a-s34.tif
Scheme 34 Synthesis of endoplasmin(1–18) using HODhbt esters.

The additive ethyl 2-cyano-2-(hydroxyimino)acetate (Oxyma), which was first reported as a racemization suppressor for peptide synthesis by Itoh in 1973,163 was reinvestigated by El-Faham and Albericio to overcome the limitations of benzotriazole-based and HODhbt active esters (Scheme 29, 29-12).164 Oxyma displays greater superiority in suppressing racemization/epimerization and improving coupling efficiency than HOBt in both automated and manual peptide synthesis. Moreover, the excellent thermal stability of Oxyma makes it a promising candidate with lower thermal risk.

4. Active esters formed upon reaction with activating reagents

As described in the third part, the preparations of conventional active esters required not only stoichiometric coupling reagents but also stoichiometric hydroxyl nucleophiles, such as HOPFP, HOSu, and HOAt. Most active esters necessitate extra preparation steps rendered such methods neither atom- nor step-economical. In addition, the risk of explosion,157 life-threatening anaphylaxis,165 or generating toxic HCN gas166 was always intertwined with these active esters. Fortunately, the in situ-formed active esters demonstrated an excellent ability to alleviate the above problems and have emerged as potentially green and ideal substitutes. The advantages and disadvantages of the in situ-formed active esters are discussed in detail in this section (Scheme 35).
image file: d2qo01686a-s35.tif
Scheme 35 Active esters formed upon reaction with activating reagents.

4.1 Alkoxyacetylenes

Methoxyacetylene, a coupling agent for peptide synthesis, was developed by Arens and co-workers in 1955, almost at the same time as that of DCC (Scheme 36).167 Mechanistically, methoxyacetylene first reacts with a carboxylic acid, producing acyl enol esters, which undergo aminolysis or alcoholysis to produce the corresponding amides or esters, respectively.168–175 These active acyl enol esters are quite attractive for peptide synthesis because the only by-product of the aminolysis reaction is ethyl acetate, which is harmless, volatile, and can be removed easily. However, long reaction times were required for carboxylic acids with moderate acidity. Although transition metal catalysts such as Hg(II),170 Ru(II),173 Ag(I),176 and Au(I)177 complexes are helpful for facilitating the activation of carboxylic acids, the low thermal stability of alkoxyacetylene and the serious racemization that occurred during the coupling circumscribed its further application in peptide synthesis.178 Recently, Breinbauer and co-workers have demonstrated that enol esters of peptide fragments prepared via the Ru(II)-catalyzed addition of peptide acid and alkyne could be used as an acyl donor for enzyme-catalyzed racemization-free peptide fragment condensation.179 Inspired by such work, Gooßen and co-workers disclosed that acetylene, in the presence of Ru(II) catalyst, could also be used as a coupling reagent for amide bond formation (Scheme 37).178 The only by-product was the versatile acetaldehyde. However, remarkable racemization occurred when it was used for peptide synthesis. As the smallest coupling reagent, acetylene would become an ideal coupling reagent if its flaws could be resolved.
image file: d2qo01686a-s36.tif
Scheme 36 Methoxyacetylene-mediated peptide bond formation.

image file: d2qo01686a-s37.tif
Scheme 37 Ruthenium-catalyzed acetylene or alkoxyacetylene-mediated peptide synthesis (adapted with permission from ref. 178, Nat. Commun., 2016, 7, 11732. Copyright (2016) Springer Nature).

4.2 α-Chlorobenzaldoxime and imidoyl halide

In 1964, Widdowson and Sutherland reported the synthesis of O,N-dibenzoylhydroxylamine esters from α-chlorobenzaldoxime and silver benzoate without the use of other reagents.180 The nitrile oxide liberated by the α-chlorobenzaldoxime reacts with the carboxylic acid spontaneously to yield the desired active esters (Scheme 38, eqn (1)). Subsequently, various nitrile oxide derivatives were evaluated for peptide synthesis.181 Among these, N-trimethylacetyl-hydroxylamine proved to be the optimal choice for preparing active esters of proteinogenic amino acids (Scheme 38).182 This method was applied for the synthesis of di- and tripeptides but was not amenable for larger peptides due to the competing Lossen rearrangement reaction, which gave isocyanate by-product.183,184 Similarly, the imidoyl halide reagents underwent rapid unimolecular ionization to give the nitrilium ions, which react rapidly with carboxylate ions to afford stable O-acylisoamide active esters (Scheme 38, eqn (2)).185 The aminolysis of these active esters completed rapidly. The potential racemization of the α-amino acid could be suppressed by adjusting the pH ≤ 7 or modifying the substituent of imidoyl halide reagents.
image file: d2qo01686a-s38.tif
Scheme 38 Peptide syntheses using N-monoacylhydroxylamine esters.

4.3 Woodward reagents

In 1961, Woodward and Olofson reinvestigated the reactions between isoxazolium salts and carboxylic acids,186–188 with a hope of developing a coupling reagent that could compete with DCC. After extensive studies, N-ethyl-S-phenylisoxazolium-3′-sulfonate (Woodward's Reagent K) was identified as an optimal choice.189 Mechanistically, the hydrogen atom at C-3 of isoxazolium 39-1 is first abstracted by a base, opening the ring in a concerted manner to offer α-ketoketenimine 39-2 (Scheme 39). Then an N-protected amino acid is added to such a reactive intermediate, forming an unstable adduct (39-4) that rapidly rearranges to enol ester 39-5. As an appealing acylating agent with high reactivity, the enol ester 39-5 enables spontaneous aminolysis to furnish a peptide bond (Scheme 39). Despite the high efficacy of the first-generation Woodward coupling reagent and the water-solubility of the by-product, a competitive rearrangement of the enol ester results in the formation of the imide, which is a less reactive acylating agent and challenging to remove.186 To address this issue, a bulky substituent was installed on the nitrogen atom to suppress such uninvited rearrangement.190,191 The enol ester derived from 39-1b was sufficiently stable for storage and resistant to rearrangement. However, the formation of enol esters from 39-1b was relatively slow and suffered from side reactions in polar solvents.190,191 Later, N-alkylbenzisoxazolium cation was used to circumvent this issue, in which the enol esters derived from these reagents were stable and did not rearrange.192–194 Unfortunately, this approach suffered from racemization and low reactivity during the aminolysis step, thus eventually resulting in Woodward reagents being abandoned by the peptide community.
image file: d2qo01686a-s39.tif
Scheme 39 Peptide synthesis using Woodward coupling reagents.

4.4 Triazine-based coupling reagents (TBCRs)

In 1994, to rationalize the drastically high reactivity of their triazine-based coupling reagents (TBCRs),195 Kamiński coined a new term “superactive ester” to differentiate it from classic “active ester” with different rate-limiting steps for aminolysis.196 In general, the aminolysis of active ester proceeds via a stepwise mechanism involving a nucleophilic attack of an amine on the ester carbonyl group to furnish a tetrahedral intermediate (TI) and a subsequent expulsion of a leaving group (Scheme 40).197,198 The rate-limiting step of aminolysis reaction is closely related to the basicity of the amines and the reactivity of active esters. This relationship can be measured by the Brønsted parameter βnuc (defined as the slope in a plot of log k vs. pKa).196 For a large number of active esters, βnuc = 0.9 ± 0.2 has been assigned to the aminolysis reactions with a rate-limiting breakdown of TI, while βnuc = 0.2 ± 0.1 for the others with a rate-limiting attack of the amine on the ester carbonyl group (Scheme 40).196,199 To differentiate these two types of active ester, Kamiński classified esters bearing a good leaving group with a βnuc = 0.2 ± 0.1 as “superactive esters”.199 As shown in Fig. 1, lowering the energy barrier in TS2 formation below that of the TS1 required the leaving group to participate in an additional energetically favored process, such as rearrangement, to offer further stabilization synchronized with its departure. Accordingly, the “superactive ester” has the merits of coupling smoothly with not only strong nucleophilic amines but also moderate basic amino components. 2-Alkylacetyloxy-4,6-dimethoxy-1,3,5-triazine ester formation is likely to proceed in a way that the tertiary amine is substituted for the chlorine atom and forms a quaternary ammonium species (Scheme 41). This step proceeded with extreme sensitivity toward the steric hindrance of the amine. The “superactive ester” is then afforded by the SNAr reaction of the quaternary ammonium species and a carboxylic acid, in which the steric bulkiness of the acid is compatible (Scheme 41).196,200 Unfortunately, the triazine reagents are irritating to the eyes and nose; thus, proper protection is necessary during the operation.201,202 In addition, the tertiary amine for the activation of carboxylic acids by CDMT is essential, and only certain tertiary amines, such as N-methylpiperidine (NMM), are able to activate the reaction.203 The formation rate of triazinylammonium salts is dependent on the steric bulkiness of the N-substituents. The success of this two-step amide bond formation procedure heavily relied on the completion of the carboxylic acid activation step.201,203 To resolve this issue, Kamiński et al. integrated the CDMT–NMM system to obtain 4-(4,6-dimethoxy-1,3,5-triazin-2-yl)-4-methylmorpholinium chloride (DMTMM).203 However, the initial DMTMM suffered from a demethylation side reaction in aprotic solvents during storage.201,203 The chloride anion was thus switched to the non-nucleophilic tetrafluoroborate anion to block the demethylation.199 This tetrafluoroborate salt coupling reagent (DMTMMBF4) was widely adopted in the synthesis of dipeptides derived from natural and unnatural hindered amino acids. Although TBCRs have been demonstrated to be efficient in assembling peptides, head-to-tail peptide cyclization (Scheme 42), and SPPS (Scheme 43), the indispensable presence of base increases the epimerization risk.199
image file: d2qo01686a-s40.tif
Scheme 40 Mechanism of the aminolysis of an active ester and Brønsted relation for amide bond formation (shadowed area) with a classic active ester (left panel) or “superactive ester” (right panel) (adapted with permission from ref. 199, J. Am. Chem. Soc., 2005, 127, 16912. Copyright (2005) American Chemical Society).

image file: d2qo01686a-f1.tif
Fig. 1 Energetic profile for the aminolysis reaction of a classic active ester (left, rate-limiting TI breakdown) and “superactive ester” (right, fast TI breakdown, rate-limiting attack of amine on an acylating reagent) (adapted with permission from ref. 199, J. Am. Chem. Soc., 2005, 127, 16912. Copyright (2005) American Chemical Society).

image file: d2qo01686a-s41.tif
Scheme 41 Mechanism of carboxylic acid activation with 2-chloro-4,6-dimethoxy-1,3,5-triazine (CDMT) (adapted with permission from ref. 199, J. Am. Chem. Soc., 2005, 127, 16912. Copyright (2005) American Chemical Society).

image file: d2qo01686a-s42.tif
Scheme 42 Synthesis of peptides using the “superactive ester” method.

image file: d2qo01686a-s43.tif
Scheme 43 Solid phase synthesis of ACP(65–74) with DMTMMBF4 as the coupling reagent.

4.5 Isonitriles

While probing preferable methods for the synthesis of larger size peptides and glycopeptides, Danishefsky and his co-workers reinvestigated the reactivities of isonitriles and developed a two-component amide bond formation strategy by taking advantage of the condensation of a carboxylic acid and an isonitrile.204–207 The amide bond formation was accomplished by the reaction of carboxylic acid and isonitrile to generate a high-energy formimidate carboxylate mixed anhydride (FCMA), which undergoes a 1,3-ON acyl transfer to deliver an N-formyl moiety to the nitrogen atom.208 Being reversible, the formation of oxy-FCMA is disfavored under usual reaction conditions. Only high temperatures with microwave irradiation can trigger the 1,3-O → N acyl transfer of oxy-FCMA. Conversely, the generation of the thio-FCMA and the corresponding SN rearrangement is more rapid and facile than that of the oxy-FCMA counterpart (Scheme 44, type I). The glycopeptide N-glycosyl asparagine was synthesized based on the type I method.209 Given the sluggishness and inefficiency of the type II intermolecular nucleophilic reaction (Scheme 44, type II), the FCMA, as an active ester, is more likely to undergo an intramolecular acyl transfer in oxyacid cases. The reaction of the thio-FCMA proceeds competitively between the intramolecular 1,3-SN acyl transfer and the intermolecular acylation. Later, sterically demanding groups were introduced to the isonitriles, which were beneficial to effectively suppress the type I intramolecular SN acyl migration pathway while favoring the type II bimolecular acylation (Scheme 44).210 Founded on hindered isonitriles (e.g., tert-butylisonitrile), thio-FCMA intermediates, as powerful acyl donors, accommodated a broad range of substrates, including bulky thioacids, secondary amines, and alcohols.210 However, expanding this coupling method to even a dipeptidyl thioacid failed to provide satisfactory results. The dipeptide thio-FCMA underwent rapid conversion to the non-productive oxazolone intermediate.211 To remedy this situation, HOBt was added to the reaction mixture to convert the active thio-FCMA intermediate into an HOBt ester.211 Although the pathway was not definitively proved in the solution phase, the mechanism was well elucidated under SPPS conditions.212 Both the type I and II methodologies have been exploited for simple amide synthesis and peptide fragment condensations,211,212 as well as the synthesis of natural products (Scheme 44)213 such as cyclosporine A209 and human granulocyte colony-stimulating factor (G-CSF).214
image file: d2qo01686a-s44.tif
Scheme 44 Coupling mechanism for isonitrile reagents and applications to the synthesis of G-CSF and cyclosporine A (adapted with permission from ref. 209, J. Am. Chem. Soc., 2010, 132, 4098. Copyright (2010) American Chemical Society).

4.6 Ynamides

Given the low reactivity of the C–C triple bond of alkoxyacetylenes, Viehe and co-workers designed an ynamine coupling reagent by replacing the oxygen atom of alkoxyacetylenes with a stronger electron-donating nitrogen atom with the hope of addressing the low reactivity of alkoxyacetylenes (Scheme 45, a).215–217 By doing so, the reactivity of the alkynyl-based coupling reagents increased dramatically, thus making the transition metal catalyst used for alkoxyacetylenes not necessary for ynamine coupling reagents. However, ynamines were too reactive to handle. Despite continued efforts, ynamines have proved to be of little utility as peptide coupling reagents due to their poor thermal stability, moisture sensitivity, and high susceptibility to racemization during peptide bond formation.217–219 The high sensitivity and reactivity of ynamines are attributed to the strong delocalization ability of the nitrogen atom lone pair toward the alkynyl C–C triple bond. To retain the advantages while mitigating the instability of ynamine coupling reagents, Gais and Neuenschwander, independently, anchored an EWG to the opposite side of the C–C triple bond of ynamine by taking advantage of a “push–pull” effect. Indeed, the “push–pull” effect worked well to increase the stability of ynamine (Scheme 45, b).220–222 The mechanism study implies that the carboxyl group is initially added to the C–C triple bond of the ynamine (Scheme 45, 45-3). This primary adduct is unstable and rapidly rearranges to the active ester 45-4.223,224 Although the “push–pull” effect stabilizes ynamine coupling reagents remarkably and demonstrates broad compatibility with amino acid side chain functional groups,223 ynamine coupling reagents were finally abandoned by the peptide community, which might be attributed to their difficult preparation and handling, and racemization/epimerization occurred during peptide bond formation.
image file: d2qo01686a-s45.tif
Scheme 45 The ynamine- or ynamide-mediated amide bond formation.

Ynamides, a special class of ynamines bearing an EWG on the nitrogen atom, have attracted much attention over the past three decades.225,226 The EWG attaching to the nitrogen atom diminished the electron-donating ability of the nitrogen atom toward the alkynyl moiety. Thus, these electron-deficient ynamines (ynamides) display remarkable stability compared with ynamines. With this knowledge in mind, Zhao and his co-workers successfully developed a class of novel ynamide coupling reagents.227 The EWG on the nitrogen atom plays a crucial role in attenuating the basicity of the ynamide, thus enabling the ynamide as a racemization-free coupling reagent for peptide bond formation.227 Among the ynamides screened, N-methylynemethylsulfonamide (MYMsA) and N-methylynetoluene-sulfonamide (MYTsA) displayed superior reactivity in facilitating amide and peptide bond formation in a racemization/epimerization-free manner.227,228 Ynamide coupling reagent-mediated amide bond formation proceeded via a two-step process with α-acyloxyenamides as stable active ester intermediates.229 The addition of carboxylic acids to the ynamide proceeded smoothly to deliver the α-acyloxyenamides in quantitative yields. It should be noted that the α-acyloxyenamides are stable and can be purified and stored for months without any deterioration. Upon treatment with an amino component, the aminolysis of α-acyloxyenamides afforded the corresponding amides in excellent yields. Both the formation and the aminolysis of the α-acyloxyenamides proceeded spontaneously under very mild reaction conditions, thus enabling a convenient two-step, one-pot protocol. The α-acyloxyenamide active esters are compatible with amino acid side chain functional groups, including -OH, -SH, -CONH2, and indole -NH (Scheme 46). With this strategy, the full scope of proteinogenic α-amino acids could act as both acyl and amine components for racemization/epimerization-free peptide synthesis, and peptide segments could also condensate smoothly. Although the ynamide coupling reagent proved to be a promising racemization-free coupling reagent, the moderate reactivity of α-acyloxyenamide active esters plagued their application potential. To overcome the inherent limitation, recently, a systematic mechanistic study including kinetics studies, Brønsted-type structure–reactivity study, and density functional theory (DFT) calculation had been performed. The X-ray crystallography analysis of the α-acyloxyenamides of pivalic acid esters indicated that the C(O)-OR bond length between the acyl moiety and the leaving group of the α-acyloxyenamide was longer than that of the common alkyl esters (1.355 Å vs. 1.340 Å, Table 1). This hinted that the acyl group of α-acyloxyenamide was more electrophilic and favorable for splitting. The Brønsted parameter βnuc = 0.23 suggested that the α-acyloxyenamides belong to a “superactive ester”, for which the formation of the tetrahedral intermediate is the rate-limiting step of the aminolysis step. In addition, the energetically favored keto–enol tautomerization of the leaving enolate decreased the kinetic barrier for the breakdown of the tetrahedral intermediate.230 The DFT calculations elucidated that the H2O-aided hydrogen transfer proceeded via a favorable six-membered-ring geometry. The energy barrier of the intermolecular hydrogen-bonding model was 7.0 kcal mol−1 lower than that of the model without H2O. The kinetics studies, Brønsted-type structure–reactivity relationship study, and DFT calculation results were consistent and implied that polar solvents were favored for the aminolysis reaction. Based on the mechanistic studies, H2O or a H2O/DMSO mixture was identified as a better solvent, and could drastically shorten the aminolysis reaction time from 24 h to 1–2 h. This significant improvement successfully expanded the applicability scope of ynamide coupling reagent to peptide fragment condensation (Scheme 47), head-to-tail cyclization (Scheme 48), and SPPS as well (Scheme 49).230


image file: d2qo01686a-s46.tif
Scheme 46 Peptide synthesis using the MYTsA reagent (adapted with permission from ref. 227, J. Am. Chem. Soc., 2016, 138, 13135. Copyright (2016) American Chemical Society).

image file: d2qo01686a-s47.tif
Scheme 47 MYMsA-mediated peptide fragment condensation (adapted with permission from ref. 230, Angew. Chem., Int. Ed., 2022, 61, e202212247. Copyright (2022) Wiley-VCH).

image file: d2qo01686a-s48.tif
Scheme 48 MYMsA-mediated peptides head-to-tail cyclization (adapted with permission from ref. 230, Angew. Chem., Int. Ed., 2022, 61, e202212247. Copyright (2022) Wiley-VCH).

image file: d2qo01686a-s49.tif
Scheme 49 Solid-phase synthesis of ACP(65–74) with α-acyloxyenamide esters (adapted with permission from ref. 230, Angew. Chem., Int. Ed., 2022, 61, e202212247. Copyright (2022) Wiley-VCH).
Table 1 Geometry of the ester linkage: C(O)-OR and C(O)O-R bond lengths of selected esters (adapted with permission from ref. 230, Angew. Chem., Int. Ed., 2022, 61, e202212247. Copyright (2022) Wiley-VCH)
image file: d2qo01686a-u1.tif


Zhao's subsequent investigation suggested that ynamide could also be used as a coupling reagent for thioamide bond formation, which was highly prone to racemization/epimerization, by employing monothiocarboxylic acid as the thioacyl donor.231 The novel thioacylating reagent α-thioacyloxyenamides could be obtained selectively from the addition reaction of monothiocarboxylic amino acids and ynamides. Notably, α-thioacyloxyenamides acted as active thiocarbonyl esters to furnish thioamide bonds when treated with various primary and secondary amines under mild reaction conditions. Most importantly, the α-thioacyloxyenamides of the proteinogenic amino acids could be used to construct thiopeptide bonds without any detectable racemization/epimerization, a notorious issue that plagues the application of thioamide-substituted peptides in protein chemical biology (Scheme 50). Trithioamide-substituted Leu-enkephalin and closthioamide were constructed smoothly using this strategy.231 α-Thioacyloxyenamides could also be used as efficient thioacylating reagents for site-specific incorporation of a thioamide bond into a growing peptide backbone via SPPS. Except for His, the monothiocarboxylic acids derived from 19 of 20 proteinogenic α-amino acids worked well to afford the thioamide-substituted peptide with excellent efficiency. This robust protocol could introduce continuous multi-thioamide bonds to the peptides on SPPS, thus offering a practical approach for thioamide-substituted peptides (Scheme 51).232


image file: d2qo01686a-s50.tif
Scheme 50 Thioamide-substituted peptide synthesis using the MYTsA reagent (adapted with permission from ref. 231, Angew. Chem., Int. Ed., 2019, 58, 1382. Copyright (2019) Wiley-VCH).

image file: d2qo01686a-s51.tif
Scheme 51 Synthesis of thioamide-substituted peptides in SPPS (adapted with permission from ref. 232, J. Org. Chem., 2020, 85, 1484. Copyright (2020) American Chemical Society).

4.7 Allenones

Although efficient for solution-phase peptide synthesis, ynamide coupling reagents are not attractive for standard SPPS because of their moderate reactivity. In their efforts to search for more reactive vinyl esters, Zhao and co-workers discovered that α-carbonyl vinyl esters derived from allenone are promising candidates.233 A common feature of DCC,22 diphenylketenimine,234 Woodward's Reagent K,186 1,3-dipole resonance structures of ethoxyacetylene235 and ynamide coupling reagent227 was that all of them contained an sp-hybridized electron-deficient carbon center. Inspired by such features and based on their success with the ynamide coupling reagent, they envisioned that it would be feasible to design a more reactive coupling reagent from allene derivatives whose linear structure also contains an sp carbon center. In addition, there is no basic center in the vinyl ester intermediates, which might be beneficial for avoiding racemization/epimerization. After extensive studies, Zhao and co-workers disclosed that allenones are the optimal choice.233 Allenone activated carboxylic acids via a spontaneous addition/rearrangement cascade reaction to furnish active vinyl ester intermediates, which underwent rapid aminolysis reaction with primary or secondary amines (Scheme 52). Compared with the aminolysis reaction of the α-acyloxyenamides, which required 2–20 h, the vinyl esters derived from allenone proceeded faster and reached completion in 20 min – 2 h with excellent yields. As shown in Scheme 52, broad substrate scopes of both the electrophilic carboxy and nucleophilic amine partners were observed. The functional groups in the side chains of Ser and Thr (–OH), Asn and Gln (–CONH2), and Trp (NH) were all tolerated in the aminolysis reaction. Not only was this strategy effective for the synthesis of simple amides and dipeptides, but it was also amenable for peptide fragment condensation (Scheme 52). Similar to that of ynamide coupling reagents, no racemization/epimerization was detected for allenone-mediated peptide bond formation. Easy preparation and rapid aminolysis make the α-carbonyl vinyl esters appealing for SPPS. Using proteinogenic amino acid-based α-carbonyl vinyl esters as the building blocks, the difficult peptide ACP(65–74) could be obtained via SPPS with an excellent crude purity of 98% after cleavage from the resin (Scheme 53). However, with the use of the same loading of amino acids, a crude purity of 86%, 77%, or 60% was obtained when PyBOP, HBTU, or PFP ester was employed as the coupling reagent or active ester, respectively. The superiority of the allenone coupling reagent in suppressing the racemization/epimerization of peptide acids was further demonstrated in carfilzomib synthesis (Scheme 54). The total synthesis of carfilzomib was accomplished via a N to C peptide elongation strategy, which was seldom used for peptide synthesis due to the severe epimerization. The superiority of allenone in suppressing the racemization/epimerization of peptide acids afforded an opportunity for peptide syntheses, which were incompatible with the classical C to N peptide elongation mode. Stability, without racemization/epimerization, and few undesired side reactions unambiguously illustrated that the α-carbonyl vinyl esters originating from allenone and proteinogenic amino acids would come into the spotlight as an attractive alternative for conventional peptide syntheses.
image file: d2qo01686a-s52.tif
Scheme 52 Allenone-mediated peptide bond formation (adapted with permission from ref. 233, J. Am. Chem. Soc., 2021, 143, 10374. Copyright (2021) American Chemical Society).

image file: d2qo01686a-s53.tif
Scheme 53 Solid-phase synthesis of ACP(65–74) using α-carbonyl vinyl esters as the building blocks (adapted with permission from ref. 233, J. Am. Chem. Soc., 2021, 143, 10374. Copyright (2021) American Chemical Society).

image file: d2qo01686a-s54.tif
Scheme 54 Synthesis of carfilzomib with allenone (52-2) as the coupling reagent (adapted with permission from ref. 233, J. Am. Chem. Soc., 2021, 143, 10374. Copyright (2021) American Chemical Society).

Inspired by the allenone coupling reagent, Li et al. reported that propargyl sulfoniums, which have similar behavior to that of allenones under basic conditions, could also act as a coupling reagent.236 They had been successfully applied to peptide synthesis and protein amidation. The mechanistic study indicated that the active thioesters prepared by the addition of thioacids and propargyl sulfoniums are the key intermediates, which could be applied in SPPS and protein labeling studies (Scheme 55).


image file: d2qo01686a-s55.tif
Scheme 55 Sulfonium-promoted peptide and protein amidation (adapted with permission from ref. 236, Org. Lett., 2022, 24, 581. Copyright (2022) American Chemical Society).

5. Conclusion

In conclusion, we have systematically summarized the origin, development, pros and cons of using active ester strategies for peptide synthesis. The significant advantage of an active ester strategy is their ability to suppress racemization/epimerization during peptide bond formation. In addition, active esters are commonly inert to the side chains of amino acids such as –OH, –SH, –COOH, –CONH2, and indole NH, making the least protection strategy feasible. However, the passion for using active esters faded with the development of coupling reagents because of their extra preparation step, additional operation and chemicals, moderate reactivity, and long aminolysis reaction time. In the past two decades, the growing request for peptides, the stringent requirement of sustainable development, and the inherent limits and drawbacks of conventional coupling reagent-mediated peptide synthesis, in turn, promoted a reinvestigation of active ester strategies. The in situ-formed transient active esters with the assistance of catalysts illustrated the benefit of active esters for peptide synthesis methodology innovations, albeit its application in relatively larger peptide synthesis as well as SPPS is waiting to be addressed. The prior prepared stable active esters, represented by PFP esters and HOSu esters, with well-defined structures, had always played a major part in peptide synthesis. However, nowadays they are only used occasionally for some special cases because of their tedious preparation and long aminolysis reaction time. The in situ-formed active esters from the addition reaction between coupling reagents and carboxylic acids would be an attractive and practical alternative because their concise preparation avoided the use of extra hydroxyl nucleophiles. For examples, coupling reagents such as alkoxyacetylenes, Woodward reagents, TBCR, and isonitriles facilitated peptide bond formation via the in situ-formed active esters with crucial intermediates. Unfortunately, active esters derived from these coupling reagents are not stable enough, resulting in complex reaction systems, racemization/epimerization, and other undesired side reactions. Gratifyingly, recently developed ynamide and allenone coupling reagents overcome the disadvantages mentioned above because they could afford bench-stable active vinyl esters upon activation of carboxylic acids. These stable vinyl esters could be used to construct amide bonds in a two-step, one-pot manner or a step-by-step manner with great flexibility. The energy released in the tautomerization of the enolate leaving group of the vinyl esters into acylate drove the aminolysis reaction to complete in a reasonably short time. More importantly, no racemization/epimerization was detected when they were used for peptide bond formation, enabling ynamides and allenones to be successfully used for dipeptide synthesis, peptide fragment condensation, peptide head-to-tail cyclization, and SPPS. Importantly, both ynamide and allenone coupling reagents are not only effective for conventional C to N peptide synthesis but also amenable for N to C peptide synthesis, which generally cannot be achieved by conventional coupling reagents, thus offering infinite possibilities for peptide synthesis. The design concept of ynamide and allenone coupling reagents with stable vinyl esters as active intermediates integrated the advantages of traditional coupling reagents and active esters for racemization-free peptide bond formation while avoiding their disadvantages. Owing to their ready preparation, stability, superiority in addressing racemization, flexibility and convenience for operation, the broad application of active esters in both SPPS and liquid-phase peptide synthesis (the third wave of peptide synthesis)237 can be expected in the future. It is foreseeable that this review will inspire new ideas about the design of novel active esters and spur the future application of active ester strategies for peptide synthesis and other coupling reactions.

Author contributions

Junfeng Zhao conceived the idea and directed the literature survey. Jinhua Yang and Huanan Huang composed the manuscript.

Conflicts of interest

There are no conflicts to declare.

Acknowledgements

This work was supported by the National Natural Science Foundation of China (22277015, 22067008), and the Natural Science Foundation of Jiangxi Province (20202ACBL203004).

References

  1. H. Lundberg, F. Tinnis, N. Selander and H. Adolfsson, Catalytic amide formation from non-activated carboxylic acids and amines, Chem. Soc. Rev., 2014, 43, 2714 RSC .
  2. S. D. Roughley and A. M. Jordan, The medicinal chemist's toolbox: an analysis of reactions used in the pursuit of drug candidates, J. Med. Chem., 2011, 54, 3451 CrossRef CAS PubMed .
  3. C. A. G. N. Montalbetti and V. Falque, Amide bond formation and peptide coupling, Tetrahedron, 2005, 61, 10827 CrossRef CAS .
  4. T. Curtius, Synthetische Versuche mit Hippurazid, Ber. Dtsch. Chem. Ges., 1902, 35, 3226 CrossRef CAS .
  5. J. R. Vaughan and R. L. Osato, Preparation of peptides using mixed carboxylic acid anhydrides, J. Am. Chem. Soc., 1951, 73, 5553 CrossRef CAS .
  6. R. M. de Figueiredo, J. S. Suppo and J. M. Campagne, Nonclassical routes for amide bond formation, Chem. Rev., 2016, 116, 12029 CrossRef CAS PubMed .
  7. V. R. Pattabiraman and J. W. Bode, Rethinking amide bond synthesis, Nature, 2011, 480, 471 CrossRef CAS PubMed .
  8. M. T. Sabatini, L. T. Boulton, H. F. Sneddon and T. D. Sheppard, A green chemistry perspective on catalytic amide bond formation, Nat. Catal., 2019, 2, 10 CrossRef CAS .
  9. M. Todorovic and D. M. Perrin, Recent developments in catalytic amide bond formation, Pept. Sci., 2020, 112, 24210 Search PubMed .
  10. S.-Y. Han and Y.-A. Kim, Recent development of peptide coupling reagents in organic synthesis, Tetrahedron, 2004, 60, 2447 CrossRef CAS .
  11. A. El-Faham and F. Albericio, Peptide coupling reagents, more than a letter soup, Chem. Rev., 2011, 111, 6557 CrossRef CAS PubMed .
  12. F. Albericio and A. El-Faham, Choosing the right coupling reagent for peptides: a twenty-five-year journey, Org. Process Res. Dev., 2018, 22, 760 CrossRef CAS .
  13. K. Hollanders, B. Maes and S. Ballet, A new wave of amide bond formations for peptide synthesis, Synthesis, 2019, 2261 CAS .
  14. Z. P. Gates and N. Hartrampf, Flow-based SPPS for protein synthesis: a perspective, Pept. Sci., 2020, 112, e24198 CAS .
  15. M. C. Bryan, P. J. Dunn, D. Entwistle, F. Gallou, S. G. Koenig, J. D. Hayler, M. R. Hickey, S. Hughes, M. E. Kopach, G. Moine, P. Richardson, F. Roschangar, A. Steven and F. J. Weiberth, Key green chemistry research areas from a pharmaceutical manufacturers’ perspective revisited, Green Chem., 2018, 20, 5082 RSC .
  16. S. De Sarkar, S. Grimme and A. Studer, NHC catalyzed oxidations of aldehydes to esters: chemoselective acylation of alcohols in presence of amines, J. Am. Chem. Soc., 2010, 132, 1190 CrossRef CAS PubMed .
  17. M. Bodanszky and Y. S. Klausner, in The chemistry of polypeptides: essays in honor of dr. leonidas zervas, ed. P. G. Katsoyannis, Springer US, Boston, MA, 1973, pp. 21 Search PubMed .
  18. M. Bodanszky, in Major methods of peptide bond formation, ed. E. Gross and J. Meienhofer, Academic Press, 1979, vol. 1, pp. 105 Search PubMed .
  19. M. Bodanszky and M. A. Bednarek, Active esters in solid-phase peptide synthesis, J. Protein Chem., 1989, 8, 461 CrossRef CAS PubMed .
  20. M. Bodanszky, in Principles of Peptide Synthesis, Springer-Verlag, Berlin, 1993, p. 9 Search PubMed .
  21. M. Bodanszky and V. Du Vigneaud, Synthesis of oxytocin by the nitrophenyl ester Method, Nature, 1959, 183, 1324 CrossRef CAS PubMed .
  22. J. C. Sheehan and G. P. Hess, A new method of forming peptide bonds, J. Am. Chem. Soc., 1955, 77, 1067 CrossRef CAS .
  23. A. J. Bates, I. J. Galpin, A. Hallett, D. Hudson, G. W. Kenner, R. Ramage and R. C. Sheppard, A New Reagent for Polypeptide Synthesis: μ-Oxo-bis-[tris-(dimethylamino)-phosphonium]-bis-tetrafluoroborate, Helv. Chim. Acta, 1975, 58, 688 CrossRef CAS PubMed .
  24. V. Dourtoglou, J.-C. Ziegler and B. Gross, L'hexafluorophosphate de O-benzotriazolyl-N,N-tetramethyluronium: Un reactif de couplage peptidique nouveau et efficace, Tetrahedron Lett., 1978, 19, 1269 CrossRef .
  25. L. A. Carpino, A. El-Faham and F. Albericio, Racemization studies during solid-phase peptide synthesis using azabenzotriazole-based coupling reagents, Tetrahedron Lett., 1994, 35, 2279 CrossRef CAS .
  26. B. Castro, J.-R. Dormoy, B. Dourtoglou, G. Evin, C. Selve and J.-C. Ziegler, Peptide coupling reagents VI1. A novelcheaper preparation of benzotriazolyloxytris[dimethylamino] phosphonium hexafluorophosphate (BOP reagent), Synthesis, 1976, 751 CrossRef CAS .
  27. J. Coste, D. Le-Nguyen and B. Castro, PyBOP®: A new peptide coupling reagent devoid of toxic by-product, Tetrahedron Lett., 1990, 31, 205 CrossRef CAS .
  28. C.-X. Fan, X.-L. Hao and Y.-H. Ye, A novel organophosphorus compound as a coupling reagent for peptide synthesis, Synth. Commun., 1996, 26, 1455 CrossRef CAS .
  29. T. Shioiri, K. Ninomiya and S. Yamada, Diphenylphosphoryl azide. New convenient reagent for a modified Curtius reaction and for peptide synthesis, J. Am. Chem. Soc., 1972, 94, 6203 CrossRef CAS PubMed .
  30. T. Shioiri, K. Ishihara and M. Matsugi, Cutting edge of diphenyl phosphorazidate (DPPA) as a synthetic reagent - A fifty-year odyssey, Org. Chem. Front., 2022, 9, 3360 RSC .
  31. P. E. Dawson, T. W. Muir, I. Clark-Lewis and S. B. Kent, Synthesis of proteins by native chemical ligation, Science, 1994, 266, 776 CrossRef CAS PubMed .
  32. Q. Wan and S. J. Danishefsky, Free-radical-based, specific desulfurization of cysteine: a powerful advance in the synthesis of polypeptides and glycopolypeptides, Angew. Chem., Int. Ed., 2007, 46, 9248 CrossRef CAS PubMed .
  33. J. Yang and J. Zhao, Recent developments in peptide ligation independent of amino acid side-chain functional group, Sci. China: Chem., 2018, 61, 97 CrossRef CAS .
  34. A. C. Conibear, E. E. Watson, R. J. Payne and C. F. W. Becker, Native chemical ligation in protein synthesis and semi-synthesis, Chem. Soc. Rev., 2018, 47, 9046 RSC .
  35. H. Yin, M. Zheng, H. Chen, S. Wang, Q. Zhou, Q. Zhang and P. Wang, Stereoselective and divergent construction of β-thiolated/selenolated amino acids via photoredox- catalyzed asymmetric giese reaction, J. Am. Chem. Soc., 2020, 142, 14201 CrossRef CAS PubMed .
  36. C.-F. Liu and J. P. Tam, Chemical ligation approach to form a peptide bond between unprotected peptide segments. Concept and model study, J. Am. Chem. Soc., 1994, 116, 4149 CrossRef CAS .
  37. X. Li, H. Y. Lam, Y. Zhang and C. K. Chan, Salicylaldehyde ester-induced chemoselective peptide ligations: enabling generation of natural peptidic linkages at the serine/threonine sites, Org. Lett., 2010, 12, 1724 CrossRef CAS PubMed .
  38. H. Liu and X. Li, Serine/Threonine ligation: origin, mechanistic aspects, and applications, Acc. Chem. Res., 2018, 51, 1643 CrossRef CAS PubMed .
  39. B. Yan, W. Shi, L. Ye and L. Liu, Acyl donors for native chemical ligation, Curr. Opin. Chem. Biol., 2018, 46, 33 CrossRef CAS PubMed .
  40. S. Leleu, M. Penhoat, A. Bouet, G. Dupas, C. Papamicael, F. Marsais and V. Levacher, Amine capture strategy for peptide bond formation by means of quinolinium thioester salts, J. Am. Chem. Soc., 2005, 127, 15668 CrossRef CAS PubMed .
  41. X. Wang, Y. Zhao, J. Yang, Y. Li, Y. Luo, M. Xu and J. Zhao, Ynamide-mediated thioester synthesis, J. Org. Chem., 2021, 86, 18265 CrossRef CAS PubMed .
  42. J. S. Zheng, S. Tang, Y. K. Qi, Z. P. Wang and L. Liu, Chemical synthesis of proteins using peptide hydrazides as thioester surrogates, Nat. Protoc., 2013, 8, 2483 CrossRef CAS .
  43. H. Li and S. Dong, Recent advances in the preparation of Fmoc-SPPS-based peptide thioester and its surrogates for NCL-type reactions, Sci. China: Chem., 2017, 60, 201 CrossRef CAS .
  44. Y. Gui, L. Qiu, Y. Li, H. Li and S. Dong, Internal activation of peptidyl prolyl thioesters in native chemical ligation, J. Am. Chem. Soc., 2016, 138, 4890 CrossRef CAS PubMed .
  45. G. M. Fang, Y. M. Li, F. Shen, Y. C. Huang, J. B. Li, Y. Lin, H. K. Cui and L. Liu, Protein chemical synthesis by ligation of peptide hydrazides, Angew. Chem., Int. Ed., 2011, 50, 7645 CrossRef CAS PubMed .
  46. J. S. Zheng, S. Tang, Y. C. Huang and L. Liu, Development of new thioester equivalents for protein chemical synthesis, Acc. Chem. Res., 2013, 46, 2475 CrossRef CAS PubMed .
  47. F. Bordusa, Proteases in organic synthesis, Chem. Rev., 2002, 102, 4817 CrossRef CAS PubMed .
  48. R. E. Thompson and T. W. Muir, Chemoenzymatic semisynthesis of proteins, Chem. Rev., 2020, 120, 3051 CrossRef CAS PubMed .
  49. A. M. Weeks and J. A. Wells, Subtiligase-catalyzed peptide ligation, Chem. Rev., 2020, 120, 3127 CrossRef CAS PubMed .
  50. S. Xu, Z. Zhao and J. Zhao, Recent advances in enzyme-mediated peptide ligation, Chin. Chem. Lett., 2018, 29, 1009 CrossRef CAS .
  51. E. Valeur and M. Bradley, Amide bond formation: beyond the myth of coupling reagents, Chem. Soc. Rev., 2009, 38, 606 RSC .
  52. P.-Q. Huang, Organic named reactions, reagents and rules, 2nd ed, Chemical Industry press, BeiJing, 2019 Search PubMed .
  53. M. Kazemi, F. Himo and J. Åqvist, Peptide release on the ribosome involves substrate-assisted base catalysis, ACS Catal., 2016, 6, 8432 CrossRef CAS .
  54. E. Fischer, Ueber einige derivate des glykocolls, alanins und leucins, Ber. Dtsch. Chem. Ges., 1902, 35, 1095 CrossRef CAS .
  55. M. D. Alexander and D. H. Busch, Preparation and characterization of cobalt(III) complexes of glycine esters, Inorg. Chem., 1966, 5, 602 CrossRef CAS .
  56. M. D. Alexander and D. H. Busch, Reactions of coordinated ligands. XIII. cobalt(III)-promoted hydrolysis of glycine esters1a, J. Am. Chem. Soc., 1966, 88, 1130 CrossRef CAS .
  57. H. Kroll, The participation of heavy metal ions in the hydrolysis of amino acid esters, J. Am. Chem. Soc., 1952, 74, 2036 CrossRef CAS .
  58. D. A. Buckingham, L. G. Marzilli and A. M. Sargeson, Peptide bond formation and subsequent hydrolysis at a cobalt(III) center, J. Am. Chem. Soc., 1967, 89, 2772 CrossRef CAS .
  59. J. P. Collman and E. Kimura, The formation of peptide bonds in the coordination sphere of cobalt(III), J. Am. Chem. Soc., 1967, 89, 6096 CrossRef CAS .
  60. T. Higashiyama, H. Ueda and A. Nakahara, The formation of peptide bonds at the coordination sites around copper(II), Bull. Chem. Soc. Jpn., 1968, 41, 3028 CrossRef CAS .
  61. S.-i. Yamada, S. Terashima and M. Wagatsuma, A simple and novel peptide bond formation by copper(II) complex, Tetrahedron Lett., 1970, 11, 1501 CrossRef PubMed .
  62. T. Ohshima, Y. Hayashi, K. Agura, Y. Fujii, A. Yoshiyama and K. Mashima, Sodium methoxide: a simple but highly efficient catalyst for the direct amidation of esters, Chem. Commun., 2012, 48, 5434 RSC .
  63. H. Tsuji and H. Yamamoto, Hydroxy-directed amidation of carboxylic acid esters using a tantalum alkoxide catalyst, J. Am. Chem. Soc., 2016, 138, 14218 CrossRef CAS PubMed .
  64. W. Muramatsu, H. Tsuji and H. Yamamoto, Catalytic peptide synthesis: amidation of N-hydroxyimino esters, ACS Catal., 2018, 8, 2181 CrossRef CAS .
  65. W. Muramatsu, T. Hattori and H. Yamamoto, Substrate-directed lewis-acid catalysis for peptide synthesis, J. Am. Chem. Soc., 2019, 141, 12288 CrossRef CAS PubMed .
  66. W. Muramatsu and H. Yamamoto, Tantalum-catalyzed amidation of amino acid homologues, J. Am. Chem. Soc., 2019, 141, 18926 CrossRef CAS PubMed .
  67. W. Muramatsu, T. Hattori and H. Yamamoto, Amide bond formation: beyond the dilemma between activation and racemisation, Chem. Commun., 2021, 57, 6346 RSC .
  68. T. V. Nguyen and D. J. Lyons, A novel aromatic carbocation-based coupling reagent for esterification and amidation reactions, Chem. Commun., 2015, 51, 3131 RSC .
  69. T. V. Nguyen and A. Bekensir, Aromatic cation activation: Nucleophilic substitution of alcohols and carboxylic acids, Org. Lett., 2014, 16, 1720 CrossRef CAS PubMed .
  70. A. K. Mishra, G. Parvari, S. K. Santra, A. Bazylevich, O. Dorfman, J. Rahamim, Y. Eichen and A. M. Szpilman, Solar and visible light assisted peptide coupling, Angew. Chem., Int. Ed., 2021, 60, 12406 CrossRef CAS .
  71. Y. Li, J. Li, G. Bao, C. Yu, Y. Liu, Z. He, P. Wang, W. Ma, J. Xie, W. Sun and R. Wang, NDTP mediated direct rapid amide and peptide synthesis without epimerization, Org. Lett., 2022, 24, 1169 CrossRef CAS PubMed .
  72. W. Wu, Z. Zhang and L. S. Liebeskind, In situ carboxyl activation using a silatropic switch: a new approach to amide and peptide constructions, J. Am. Chem. Soc., 2011, 133, 14256 CrossRef CAS PubMed .
  73. W. Muramatsu, C. Manthena, E. Nakashima and H. Yamamoto, Peptide bond-forming reaction via amino acid silyl esters: New catalytic reactivity of an aminosilane, ACS Catal., 2020, 10, 9594 CrossRef CAS .
  74. W. Muramatsu and H. Yamamoto, Peptide bond formation of amino acids by transient masking with silylating reagents, J. Am. Chem. Soc., 2021, 143, 6792 CrossRef CAS PubMed .
  75. T. Hattori and H. Yamamoto, Synthesis of silacyclic dipeptides: peptide elongation at both N- and C-termini of dipeptide, J. Am. Chem. Soc., 2022, 144, 1758 CrossRef CAS PubMed .
  76. A. Pelter, T. E. Levitt and P. Nelsoni, Some amide forming reactions involving boron reagents, Tetrahedron, 1970, 26, 1539 CrossRef CAS .
  77. K. Ishihara, S. Ohara and H. Yamamoto, 3,4,5-Trifluorobenzeneboronic acid as an extremely active amidation catalyst, J. Org. Chem., 1996, 61, 4196 CrossRef CAS PubMed .
  78. S. Arkhipenko, M. T. Sabatini, A. S. Batsanov, V. Karaluka, T. D. Sheppard, H. S. Rzepa and A. Whiting, Mechanistic insights into boron-catalysed direct amidation reactions, Chem. Sci., 2018, 9, 1058 RSC .
  79. K. Wang, Y. Lu and K. Ishihara, The ortho-substituent on 2,4-bis(trifluoromethyl)phenylboronic acid catalyzed dehydrative condensation between carboxylic acids and amines, Chem. Commun., 2018, 54, 5410 RSC .
  80. T. Mohy El Dine, W. Erb, Y. Berhault, J. Rouden and J. Blanchet, Catalytic chemical amide synthesis at room temperature: one more step toward peptide synthesis, J. Org. Chem., 2015, 80, 4532 CrossRef CAS PubMed .
  81. N. Gernigon, R. M. Al-Zoubi and D. G. Hall, Direct amidation of carboxylic acids catalyzed by ortho-iodo arylboronic acids: catalyst optimization, scope, and preliminary mechanistic study supporting a peculiar halogen acceleration effect, J. Org. Chem., 2012, 77, 8386 CrossRef CAS PubMed .
  82. H. Noda, M. Furutachi, Y. Asada, M. Shibasaki and N. Kumagai, Unique physicochemical and catalytic properties dictated by the B3NO2 ring system, Nat. Chem., 2017, 9, 571 CrossRef CAS PubMed .
  83. K. Michigami, T. Sakaguchi and Y. Takemoto, Catalytic dehydrative peptide synthesis with gem-diboronic acids, ACS Catal., 2019, 10, 683 CrossRef .
  84. M. Koshizuka, K. Makino and N. Shimada, Diboronic acid anhydride-catalyzed direct peptide bond formation enabled by hydroxy-directed dehydrative condensation, Org. Lett., 2020, 22, 8658 CrossRef CAS PubMed .
  85. J. W. Ren, M. N. Tong, Y. F. Zhao and F. Ni, Synthesis of dipeptide, amide, and ester without racemization by oxalyl chloride and catalytic triphenylphosphine oxide, Org. Lett., 2021, 23, 7497 CrossRef CAS PubMed .
  86. J. Su, J. N. Mo, X. Chen, A. Umanzor, Z. Zhang, K. N. Houk and J. Zhao, Generation of oxyphosphonium ions by photoredox/cobaloxime catalysis for scalable amide and peptide synthesis in batch and continuous-flow, Angew. Chem., Int. Ed., 2022, 61, e202112668 CAS .
  87. R. Schwyzer, B. Iselin and M. Feurer, Über aktivierte ester. I. Aktivierte ester der hippursäure und ihre umsetzungen mit benzylamin, Helv. Chim. Acta, 1955, 38, 69 CrossRef CAS .
  88. R. Schwyzer, M. Feurer, B. Iselin and H. Kägi, Über aktivierte ester. II. Synthese aktivierter ester von aminosäure-derivaten, Helv. Chim. Acta, 1955, 38, 80 CrossRef CAS .
  89. R. Schwyzer, M. Feurer and B. Iselin, Über aktivierte ester. III. Umsetzungen aktivierter ester von aminosäure- und peptid-derivaten mit aminen und aminosäureestern, Helv. Chim. Acta, 1955, 38, 83 CrossRef CAS .
  90. B. Iselin, M. Feurer and R. Schwyzer, Über aktivierte ester V. verwendung der cyanmethylester-methode zur herstellung von (N-carbobenzoxy-S-benzyl-l-cysteinyl)-l-tyrosyl-l-isoleucin auf verschiedenen wegen, Helv. Chim. Acta, 1955, 38, 1508 CrossRef CAS .
  91. T. Miyazawa, K. Tanaka, E. Ensatsu, R. Yanagihara and T. Yamada, Broadening of the substrate tolerance of α-chymotrypsin by using the carbamoylmethyl ester as an acyl donor in kinetically controlled peptide synthesis, J. Chem. Soc., Perkin Trans. 1, 2001, 87 RSC .
  92. J. C. Sheehan, K. Hasspacher and Y. L. Yeh, Activated cyclic derivatives of amino acids, J. Am. Chem. Soc., 1959, 81, 6086 CrossRef CAS .
  93. E. Gross and B. Witkop, Selective cleavage of the methionyl peptide bonds in ribonuclease with cyanogen bromide1, J. Am. Chem. Soc., 1961, 83, 1510 CrossRef CAS .
  94. D. F. Dyckes, T. Creighton and R. C. Sheppard, Spontaneous re-formation of a broken peptide chain, Nature, 1974, 247, 202 CrossRef CAS PubMed .
  95. J. Spengler, C. Bottcher, F. Albericio and K. Burger, Hexafluoroacetone as protecting and activating reagent: new routes to amino, hydroxy, and mercapto acids and their application for peptide and glyco- and depsipeptide modification, Chem. Rev., 2006, 106, 4728 CrossRef PubMed .
  96. F. Albericio, K. Burger, J. Ruiz-Rodriguez and J. Spengler, A new strategy for solid-phase depsipeptide synthesis using recoverable building blocks, Org. Lett., 2005, 7, 597 CrossRef CAS PubMed .
  97. M. C. Pirrung, F. Zhang, S. Ambadi and T. R. Ibarra-Rivera, Reactive esters in amide ligation with β-Hydroxyamines, Eur. J. Org. Chem., 2012, 4283 CrossRef CAS .
  98. K. Burger, M. Rudolph, E. Windeisen, A. Worku and S. Fehn, Hexafluoraceton als schutzgruppen-und aktivierungsreagenz in der aminosäure-und peptidchemie, 11. Mitt.: Ein neuer präparativ einfacher zugang zu 2,5-dioxopiperazinen und 2,5-dioxomorpholinen, Monatsh. Chem., 1993, 124, 453 CrossRef CAS .
  99. K. Pumpor, J. Spengler, F. Albericio, A. Haas and K. Burger, Incorporation of the α-mercapto acid unit into peptides, Synthesis, 2004, 1088 CAS .
  100. K. Burger and M. Rudolph, Regiospezifische reaktionen mit ω-carboxy-α-ami-nosäuren-eine einfache synthese für aspartam, Chem.-Ztg., 1990, 114, 249 CAS .
  101. M. Gordon, J. G. Miller and A. R. Day, Effect of structure on reactivity; ammonolysis of esters with special reference to the electron release effects of alkyl and aryl groups, J. Am. Chem. Soc., 1948, 70, 1946 CrossRef CAS PubMed .
  102. T. Wieland, W. Schäfer and E. Bokelmann, Über peptidsynthesen v. über eine bequeme darstellungsweise von acylthiophenolen und ihre verwendung zu amid- und peptid-synthesen, Justus Liebigs Ann. Chem., 1951, 573, 99 CrossRef CAS .
  103. J. A. Farringrton, P. J. Hextall, G. W. Kenner and J. M. Turner, 265. Peptides. part VII. the preparation and use of p-nitrophenyl thiolesters, J. Chem. Soc., 1957, 1407 RSC .
  104. G. W. Kenner, P. J. Thomson and J. M. Turner, 835. Peptides. part VIII. cyclic peptides derived from leucine and glycine, J. Chem. Soc., 1958, 4148 RSC .
  105. H.-D. Jakubke, Synthese von aminoacyl- und peptidyl-selenophenolen und deren verwendung zur darstellung linearer und cyclischer peptide, Chem. Ber., 1964, 97, 2816 CrossRef CAS .
  106. H.-D. Jakubke and A. Voigt, Über aktivierte ester, VIII. peptidsynthesen mit halogensubstituierten acylaminosäure- und acylpeptid-[chinolyl-(8)-estern], Chem. Ber., 1966, 99, 2944 CrossRef CAS .
  107. K. Lloyd and G. T. Young, The use of acylamino-acid esters of 2-mercaptopyridine in peptide synthesis, Chem. Commun., 1968, 1400 RSC .
  108. M. BodÁNszky, Synthesis of peptides by aminolysis of nitrophenyl esters, Nature, 1955, 175, 685 CrossRef PubMed .
  109. M. Bodanszky and V. du Vigneaud, A method of synthesis of long peptide chains using a synthesis of oxytocin as an example, J. Am. Chem. Soc., 1959, 81, 5688 CrossRef CAS .
  110. M. Bodanszky and V. du Vigneaud, Synthesis of a biologically active analog of oxytocin, with phenylalanine replacing tyrosine, J. Am. Chem. Soc., 1959, 81, 6072 CrossRef CAS .
  111. R. Schwyzer and P. Sieber, Die synthese von Gramicidin S, Helv. Chim. Acta, 1957, 40, 624 CrossRef CAS .
  112. M. Bodanszky, M. A. Ondetti, S. D. Levine and N. J. Williams, Synthesis of secretin. II. The stepwise approach, J. Am. Chem. Soc., 1967, 89, 6753 CrossRef CAS PubMed .
  113. P. G. Katsoyannis and K. Suzuki, Insulin peptides. III. synthesis of a protected nonapeptide containing the c-terminal sequence of the B-chain of Insulin, J. Am. Chem. Soc., 1962, 84, 1420 CrossRef CAS .
  114. J. Pless and R. A. Boissonnas, Uber die Geschwindigkeit der aminolyse von verschiedenen neuen, aktivierten, N-geschtzten α-aminosure-phenylestern, insbesondere 2,4,5-Trichlorphenylestern, Helv. Chim. Acta, 1963, 46, 1609 CrossRef .
  115. J. Kovacs and A. Kapoor, Synthesis of polypeptides with known repeating sequence of amino acids, J. Am. Chem. Soc., 1965, 87, 118 CrossRef CAS PubMed .
  116. J. Kovacs, R. Giannotti and A. Kapoor, Polypeptides with known repeating sequence of amino acids. synthesis of poly-l-glutamyl-l-alanyl-l-glutamic acid and polyglycylglycyl-l-phenylalanine through pentachlorophenyl active ester, J. Am. Chem. Soc., 1966, 88, 2282 CrossRef CAS .
  117. S. Guttmann and R. A. Boissonnas, Synthèse de la Sér4-oxytocine, de l'Ala4-oxytocine, de la Sér5-oxytocine et de l'Ala5-oxytocine, Helv. Chim. Acta, 1963, 46, 1626 CrossRef CAS .
  118. F. H. C. Stewart, Benzyloxycarbonylamino acid 2,6-dichloro-4-nitrophenyl esters and the reaction of 2,6-dichloro-4-nitrophenol with N,N’-dicyclohexylcarbodiimide, Aust. J. Chem., 1968, 21, 477 CrossRef CAS .
  119. L. Kisfaludy, J. E. Roberts, R. H. Johnson, G. L. Mayers and J. Kovacs, Synthesis of N-carbobenzoxyamino acid and peptide pentafluorophenyl esters as intermediates in peptide synthesis, J. Org. Chem., 1970, 35, 3563 CrossRef CAS PubMed .
  120. J. Kovacs, G. L. Mayers, R. H. Johnson, R. E. Cover and U. R. Ghatak, Rates of racemization and coupling of cysteine active ester derivatives, J. Chem. Soc. D, 1970, 53 RSC .
  121. J. Kovacs, G. L. Mayers, R. H. Johnson, R. E. Cover and U. R. Ghatak, Racemization of amino acid derivatives. Rate of racemization and peptide bond formation of cysteine active esters, J. Org. Chem., 1970, 35, 1810 CrossRef CAS PubMed .
  122. L. Kisfaludy, I. Schőn, T. Szirtes, O. Nyéki and M. Lőw, A novel and rapid peptide synthesis, Tetrahedron Lett., 1974, 15, 1785 CrossRef .
  123. E. Atherton and R. C. Sheppard, Solid phase peptide synthesis using Nα-fluorenylmethoxycarbonylamino acid pentafluorophenyl esters, J. Chem. Soc., Chem. Commun., 1985, 165 RSC .
  124. E. Atherton, L. R. Cameron and R. C. Sheppard, Peptide synthesis, Tetrahedron, 1988, 44, 843 CrossRef CAS .
  125. B. Iselin, W. Rittel, P. Sieber and R. Schwyzer, Neue methoden zur herstellung von carbonsäure-arylestern. über aktivierte ester VIII, Helv. Chim. Acta, 1957, 40, 373 CrossRef CAS .
  126. B. J. Johnson and P. M. Jacobs, A new carboxy-protecting group for peptide synthesis and its direct conversion into an activated ester suitable for peptide formation: 4-(methylthio)phenyl and 4-(methylsulphonyl)phenyl esters, Chem. Commun., 1968, 73 RSC .
  127. B. J. Johnson and E. G. Trask, The 4-(methylthio)phenyl and 4-(methylsulfonyl)phenyl esters in the preparation of peptides and polypeptides. Synthesis of the protected heptapeptide (A82-A88) of bovine chymotrypsinogen A, J. Org. Chem., 1968, 33, 4521 CrossRef CAS PubMed .
  128. G. Losse, K.-H. Hoffmann and G. Hetzer, Peptidsynthesen mit O-[Cbo-Aminoacyl]-oximen und O-[Cbo-Aminoacyl]-pyrazolon-enolen, Justus Liebigs Ann. Chem., 1965, 684, 236 CrossRef .
  129. A. S. Dutta and J. S. Morley, Polypeptides. XII. the preparation of 2-pyridyl esters and their use in peptide synthesis, J. Chem. Soc. C, 1971, 17, 2896 RSC .
  130. J. Zhao, J. Shi and Y. Li, Benzyne-mediated esterification reaction, Org. Lett., 2021, 23, 7274 CrossRef CAS PubMed .
  131. G. H. L. Nefkens and G. I. Tesser, A novel activated ester in peptide syntheses, J. Am. Chem. Soc., 1961, 83, 1263 CrossRef CAS .
  132. G. H. L. Nefkens, G. I. Tesser and R. J. F. Nivard, Synthesis and reactions of esters of N-hydroxyphthalimide and N-protected amino acids, Recl. Trav. Chim. Pays-Bas, 1962, 81, 683 CrossRef CAS .
  133. G. W. Anderson, J. E. Zimmerman and F. M. Callahan, N-Hydroxysuccinimide esters in peptide synthesis, J. Am. Chem. Soc., 1963, 85, 3039 CrossRef CAS .
  134. G. W. Anderson, J. E. Zimmerman and F. M. Callahan, The use of esters of N-hydroxysuccinimide in peptide synthesis, J. Am. Chem. Soc., 1964, 86, 1839 CrossRef CAS .
  135. J. M. Becker, M. Wilchek and E. Katchalski, Irreversible inhibition of biotin transport in yeast by biotinyl-p-nitrophenyl ester, Proc. Natl. Acad. Sci. U. S. A., 1971, 68, 2604 CrossRef CAS PubMed .
  136. J. Becker and M. Wilchek, Inactivation by avidin of biotin-modified bacteriophage, Biochim. Biophys. Acta, 1972, 264, 165 CrossRef CAS PubMed .
  137. W. Lossen, Ueber benzoylderivate des hydroxylamins, Justus Liebigs Ann. Chem., 1872, 161, 347 CrossRef .
  138. H. Gross and L. Bilk, Zur reaktion von N-Hydroxysuccinimid mit dicyclohexylcarbodiimid, Tetrahedron, 1968, 24, 6935 CrossRef CAS .
  139. J. Savrda, An unusual side reaction of 1-succinimidyl esters during peptide synthesis, J. Org. Chem., 1977, 42, 3199 CrossRef CAS PubMed .
  140. A. Isidro-Llobet, X. Just-Baringo, A. Ewenson, M. Alvarez and F. Albericio, Fmoc-2-mercaptobenzothiazole, for the introduction of the Fmoc moiety free of side-reactions, Biopolymers, 2007, 88, 733 CrossRef CAS PubMed .
  141. M. Fujino, S. Kobayashi, M. Obayashi, T. Fukuda and S. Shinagawa, The use of N-hydroxy-5-morbornene-2,3-dicarboximide active esters in peptide synthesis, Chem. Pharm. Bull., 1974, 22, 1857 CrossRef CAS PubMed .
  142. J. O. Edwards and R. G. Pearson, The factors determining nucleophilic reactivities, J. Am. Chem. Soc., 1962, 84, 16 CrossRef CAS .
  143. P. G. Sammes, Studies on some O-acyl N-substituted hydroxylamines, J. Chem. Soc., 1965, 6608 RSC .
  144. L. A. Paquette, Electrophilic additions of acyl and sulfonyl halides to 2-ethoxypyridine 1-oxide. a new class of activated esters and their application to peptide synthesis, J. Am. Chem. Soc., 1965, 87, 5186 CrossRef CAS PubMed .
  145. S. Bittner, Y. Knobler and M. Frankel, Active α-aminoacyl O-derivatives of N-substituted hydroxylamine. O to N migration and a method for peptide synthesis, Tetrahedron Lett., 1965, 6, 95 CrossRef .
  146. S. M. Beaumont, B. O. Handford, J. H. Jones and G. T. Young, The use of esters of N,N-dialkylhydroxylamines in peptide synthesis and as selective acylating agents, Chem. Commun., 1965, 53 RSC .
  147. B. O. Handford, J. H. Jones, G. T. Young and T. F. Johnson, 1260. Amino-acids and peptides. XXIV. The use of esters of 1-hydroxypiperidine and of other N,N-dialkylhydroxylamines in peptide synthesis and as selective acylating agents, J. Chem. Soc., 1965, 6814 RSC .
  148. J. H. Jones and G. T. Young, Amino-acids and peptides. 28. Anchimeric acceleration of the aminolysis of esters. The use of mono-esters of catechol in peptide synthesis, J. Chem. Soc. C, 1968, 4, 436 RSC .
  149. W. König and R. Geiger, Racemisierung bei peptid synthesen, Chem. Ber., 1970, 103, 2024 CrossRef PubMed .
  150. W. König and R. Geiger, Eine neue methode zur synthese von peptiden: aktivierung der carboxylgruppe mit dicyclohexylcarbodiimid und 3-hydroxy-4-oxo-3.4-dihydro-l.2.3-benzotriazin, Chem. Ber., 1970, 103, 788 CrossRef PubMed .
  151. H. Benz, The role of solid-phase fragment condensation (SPFC) in peptide synthesis, Synthesis, 1994, 337 CrossRef CAS .
  152. L. A. Carpino and A. El-Faham, The diisopropylcarbodiimide/1-hydroxy-7-azabenzotriazole system: segment coupling and stepwise peptide assembly, Tetrahedron, 1999, 55, 6813 CrossRef CAS .
  153. L. A. Carpino, H. Imazumi, B. M. Foxman, M. J. Vela, P. Henklein, A. El-Faham, J. Klose and M. Bienert, Comparison of the effects of 5- and 6-HOAt on model peptide coupling reactions relative to the cases for the 4- and 7-isomers, Org. Lett., 2000, 2, 2253 CrossRef CAS PubMed .
  154. A. El-Faham and F. Albericio, Synthesis and application of N-hydroxylamine derivatives as potential replacements for HOBt, Eur. J. Org. Chem., 2009, 1499 CrossRef CAS .
  155. J. C. Spetzler, M. Meldal, J. Felding, P. Vedsø and M. Begtrup, Novel acylation catalysts in peptide synthesis: derivatives of N-hydroxytriazoles and N-hydroxytetrazoles, J. Chem. Soc., Perkin Trans. 1, 1998, 1727 RSC .
  156. L. A. Carpino, 1-Hydroxy-7-azabenzotriazole. An efficient peptide coupling additive, J. Am. Chem. Soc., 1993, 115, 4397 CrossRef CAS .
  157. K. D. Wehrstedt, P. A. Wandrey and D. Heitkamp, Explosive properties of 1-hydroxybenzotriazoles, J. Hazard. Mater., 2005, 126, 1 CrossRef CAS PubMed .
  158. E. Atherton, L. Cameron, M. Meldal and R. C. Sheppard, Self-indicating activated esters for use in solid phase peptide synthesis. Fluorenylmethoxycarbonylamino acid derivatives of 3-hydroxy-4-oxodihydrobenzotriazine, J. Chem. Soc., Chem. Commun., 1986, 1763 RSC .
  159. E. Atherton, J. L. Holder, M. Meldal, R. C. Sheppard and R. M. Valerio, Peptide synthesis. Part 12. 3,4-Dihydro-4-oxo-1,2,3-benzotriazin-3-yl esters of fluorenylmethoxycarbonyl amino acids as self-indicating reagents for solid phase peptide synthesis, J. Chem. Soc., Perkin Trans. 1, 1988, 2887 RSC .
  160. L. R. Cameron, J. L. Holder, M. Meldal and R. C. Sheppard, Peptide synthesis. Part 13. Feedback control in solid phase synthesis. Use of fluorenylmethoxycarbonyl amino acid 3,4-dihydro-4-oxo-1,2,3-benzotriazin-3-yl esters in a fully automated system, J. Chem. Soc., Perkin Trans. 1, 1988, 2895 RSC .
  161. L. A. Carpino, J. Xia and A. El-Faham, 3-Hydroxy-4-oxo-3,4-dihydro-5-azabenzo-1,2,3-triazene, J. Org. Chem., 2004, 69, 54 CrossRef CAS PubMed .
  162. M. H. Jacobsen, O. Buchardt, T. Engdahl and A. Holm, A new facile one-pot preparation of pentafluorophenyl (Pfp) and 3,4-dihydro-4-oxo-1,2,3-benzotriazine-3-yl (Dhbt) esters of Fmoc amino acids, Tetrahedron Lett., 1991, 32, 6199 CrossRef .
  163. M. Itoh, Peptides. IV. Racemization suppression by the use of ethyl 2-hydroximino-2-cyanoacetate and its amide, Bull. Chem. Soc. Jpn., 1973, 46, 2219 CrossRef CAS .
  164. R. Subiros-Funosas, R. Prohens, R. Barbas, A. El-Faham and F. Albericio, Oxyma: an efficient additive for peptide synthesis to replace the benzotriazole-based HOBt and HOAt with a lower risk of explosion, Chem. – Eur. J., 2009, 15, 9394 CrossRef CAS PubMed .
  165. K. J. McKnelly, W. Sokol and J. S. Nowick, Anaphylaxis Induced by peptide coupling agents: Lessons learned from repeated exposure to HATU, HBTU, and HCTU, J. Org. Chem., 2020, 85, 1764 CrossRef CAS PubMed .
  166. M. Erny, M. Lundqvist, J. H. Rasmussen, O. Ludemann-Hombourger, F. Bihel and J. Pawlas, Minimizing HCN in DIC/Oxyma-mediated amide bond-forming reactions, Org. Process Res. Dev., 2020, 24, 1341 CrossRef CAS .
  167. J. F. Arens, The chemistry of acetylenic ethers XIII: Acetylenic ethers as reagents for the preparation of amides, Recl. Trav. Chim. Pays-Bas, 1955, 74, 769 CrossRef CAS .
  168. R. Broekema, S. van der Werf and J. F. Arens, Chemistry of acetylenic ethers XXX. Additions of trichloroacetic acid, its anhydride and its chloride to ethoxyacetylene, Recl. Trav. Chim. Pays-Bas, 1958, 77, 258 CrossRef CAS .
  169. J. Sheehan and J. Hlavka, A reaction peptide intermediate derived from ethoxyacetylene, J. Org. Chem., 1958, 23, 635 CrossRef CAS .
  170. H. H. Wasserman and P. S. Wharton, 1-Methoxyvinyl esters.1 I. Preparation and properties, J. Am. Chem. Soc., 1960, 82, 661 CrossRef CAS .
  171. H. H. Wasserman and P. S. Wharton, 1-Methoxyvinyl esters.1II.2 An oxygen-18 study of anhydride formation, J. Am. Chem. Soc., 1960, 82, 1411 CrossRef CAS .
  172. H. J. Panneman, A. F. Marx and J. F. Arens, Derivatives of amino-acids and peptides IV: The synthesis of peptides by means of ethoxyethyne, Recl. Trav. Chim. Pays-Bas, 1959, 78, 487 CrossRef CAS .
  173. C. Ruppin, P. H. Dixneuf and S. Lecolier, Regioselective synthesis of isopropenyl esters by ruthenium catalysed addition of N-protected amino-acids to propyne, Tetrahedron Lett., 1988, 29, 5365 CrossRef CAS .
  174. Z. Kabouche, C. Bruneau and P. H. Dixneuf, Enol esters as intermediates for the facile conversion of amino acids into amides and dipeptides, Tetrahedron Lett., 1991, 32, 5359 CrossRef CAS .
  175. F. Weygand and W. Steglich, Peptid-synthesen mit acylaminosäure-vinylestern, Angew. Chem., 1961, 73, 757 CrossRef CAS .
  176. Y. Ishino, I. Nishiguchi, S. Nakao and T. Hirashima, Novel synthesis of enol esters through silver-catalyzed reaction of acetylenic compounds with carboxylic acids, Chem. Lett., 1981, 10, 641 CrossRef .
  177. B. C. Chary and S. Kim, Gold(I)-catalyzed addition of carboxylic acids to alkynes, J. Org. Chem., 2010, 75, 7928 CrossRef CAS PubMed .
  178. T. Krause, S. Baader, B. Erb and L. J. Gooßen, Atom-economic catalytic amide synthesis from amines and carboxylic acids activated in situ with acetylenes, Nat. Commun., 2016, 7, 11732 CrossRef PubMed .
  179. H. Schröder, G. A. Strohmeier, M. Leypold, T. Nuijens, P. J. L. M. Quaedflieg and R. Breinbauer, Racemization-free chemoenzymatic peptide synthesis enabled by the ruthenium-catalyzed synthesis of peptide enol estersviaalkyne-addition and subsequent conversion using alcalase-cross-linked enzyme aggregates, Adv. Synth. Catal., 2013, 355, 1799 CrossRef .
  180. J. K. Sutherland and D. A. Widdowson, The reaction of α-chlorobenzaldoxime with carboxylate ions, J. Chem. Soc., 1964, 4651 CAS .
  181. T. R. Govindachari, K. Nagarajan, S. Rajappa, A. S. Akerkar and V. S. Iyer, Synthesis of O-acylbenzohydroxamic acids and their use in peptide synthesis, Tetrahedron, 1966, 22, 3367 CrossRef CAS .
  182. S. Rajappa, K. Nagarajan and V. S. Iyer, Hydroxamic acids and their derivatives-III, Tetrahedron, 1967, 23, 4805 CrossRef CAS PubMed .
  183. T. R. Govindachari, S. Rajappa, A. S. Akerkar and V. S. Iyer, Hydroxamic acids and their derivatives-IV, Tetrahedron, 1967, 23, 4811 CrossRef CAS PubMed .
  184. D. C. Berndt and H. Shechter, Substituent effects in the Lossen rearrangement of benzoyl acylhydroxamates1a, J. Org. Chem., 1964, 29, 916 CrossRef CAS .
  185. A. F. Hegarty and D. G. Mccarthy, Peptide-synthesis using unprotected amino-acids and novel imidoyl halide reagents, J. Am. Chem. Soc., 1980, 102, 4537 CrossRef CAS .
  186. R. B. Woodward and R. A. Olofson, The reaction of isoxazolium salts with bases, J. Am. Chem. Soc., 1961, 83, 1007 CrossRef CAS .
  187. O. Mumm and G. Münchmeyer, Überführung des oxymethylen-acetophenons in benzoyl-brenztraubensäure und einige neue Derivate, Ber. Dtsch. Chem. Ges., 1910, 43, 3335 CrossRef .
  188. O. Mumm and C. Bergell, Über die freie aceton-oxalsäure und ihre Abkömmlinge, Ber. Dtsch. Chem. Ges., 1912, 45, 3040 CrossRef .
  189. R. B. Woodward, R. A. Olofson and H. Mayer, A new synthesis of peptides, J. Am. Chem. Soc., 1961, 83, 1010 CrossRef CAS .
  190. R. B. Woodward and D. J. Woodman, N-t-butylketoketenimines, J. Am. Chem. Soc., 1966, 88, 3169 CrossRef CAS .
  191. R. B. Woodward and D. J. Woodman, Stable enol esters from N-tert-butyl-5-methylisoxazolium perchlorate, J. Am. Chem. Soc., 1968, 90, 1371 CrossRef CAS .
  192. D. S. Kemp and R. B. Woodward, The N-ethylbenzisoxazolium cation-I, Tetrahedron, 1965, 21, 3019 CrossRef CAS .
  193. D. S. Kemp, The N-ethylbenzisoxazolium cation-II, Tetrahedron, 1967, 23, 2001 CrossRef CAS .
  194. D. S. Kemp and S. W. Chien, A new peptide coupling reagent, J. Am. Chem. Soc., 1967, 89, 2743 CrossRef CAS PubMed .
  195. P. H. Huy and C. Mbouhom, Formamide catalyzed activation of carboxylic acids-versatile and cost-efficient amidation and esterification, Chem. Sci., 2019, 10, 7399 RSC .
  196. Z. J. Kaminski, The concept of superactive esters. Could peptide synthesis be improved by inventing superactive esters?, Int. J. Pept. Protein Res., 1994, 43, 312 CrossRef CAS .
  197. M. L. Bender, Mechanisms of catalysis of nucleophilic reactions of carboxylic acid derivatives, Chem. Rev., 1960, 60, 53 CrossRef CAS .
  198. A. Williams, Concerted mechanisms of acyl group transfer reactions in solution, Acc. Chem. Res., 1989, 22, 387 CrossRef CAS .
  199. Z. J. Kaminski, B. Kolesinska, J. Kolesinska, G. Sabatino, M. Chelli, P. Rovero, M. Blaszczyk, M. L. Glowka and A. M. Papini, N-triazinylammonium tetrafluoroborates. A new generation of efficient coupling reagents useful for peptide synthesis, J. Am. Chem. Soc., 2005, 127, 16912 CrossRef CAS PubMed .
  200. Z. J. Kamiński, Triazine-based condensing reagents, Biopolymers, 2000, 55, 140 CrossRef .
  201. M. Kunishima, C. Kawachi, J. Monta, K. Terao, F. Iwasaki and S. Tani, 4-(4,6-dimethoxy-1,3,5-triazin-2-yl)-4-methyl-morpholinium chloride: an efficient condensing agent leading to the formation of amides and esters, Tetrahedron, 1999, 55, 13159 CrossRef CAS .
  202. Z. J. Kaminski, B. Kolesinska, J. E. Kaminska and J. Gora, A novel generation of coupling reagents. Enantiodifferentiating coupling reagents prepared in situ from 2-chloro-4,6-dimethoxy-1,3,5-triazine (CDMT) and chiral tertiary amines, J. Org. Chem., 2001, 66, 6276 CrossRef CAS PubMed .
  203. Z. J. Kamiński, P. Paneth and J. Rudziński, A study on the activation of carboxylic acids by means of 2-chloro-4,6-dimethoxy-1,3,5-triazine and 2-chloro-4,6-diphenoxy-1,3,5-triazine, J. Org. Chem., 1998, 63, 4248 CrossRef .
  204. X. C. Li and S. J. Danishefsky, New chemistry with old-functional groups: on the reaction of isonitriles with carboxylic acids-a route to various amide types, J. Am. Chem. Soc., 2008, 130, 5446 CrossRef CAS PubMed .
  205. R. M. Wilson, J. L. Stockdill, X. Wu, X. Li, P. A. Vadola, P. K. Park, P. Wang and S. J. Danishefsky, A fascinating journey into history: exploration of the world of isonitriles en route to complex amides, Angew. Chem., Int. Ed., 2012, 51, 2834 CrossRef CAS PubMed .
  206. G. O. Jones, X. Li, A. E. Hayden, K. N. Houk and S. J. Danishefsky, The coupling of isonitriles and carboxylic acids occurring by sequential concerted rearrangement mechanisms, Org. Lett., 2008, 10, 4093 CrossRef CAS PubMed .
  207. X. C. Li, Y. Yuan, W. F. Berkowitz, L. J. Todaro and S. J. Danishefsky, On the two-component microwave-mediated reaction of isonitriles with carboxylic acids: regarding alleged formimidate carboxylate mixed anhydrides, J. Am. Chem. Soc., 2008, 130, 13222 CrossRef CAS PubMed .
  208. X. C. Li, Y. Yuan, C. Kan and S. J. Danishefsky, Addressing mechanistic issues in the coupling of isonitriles and carboxylic acids: potential routes to peptidic constructs, J. Am. Chem. Soc., 2008, 130, 13225 CrossRef CAS PubMed .
  209. X. Y. Wu, J. L. Stockdill, P. Wang and S. J. Danishefsky, Total synthesis of cyclosporine: Access to N-methylated peptides via isonitrile coupling reactions, J. Am. Chem. Soc., 2010, 132, 4098 CrossRef CAS PubMed .
  210. Y. Rao, X. Li and S. J. Danishefsky, Thio FCMA intermediates as strong acyl donors: A general solution to the formation of complex amide bonds, J. Am. Chem. Soc., 2009, 131, 12924 CrossRef CAS PubMed .
  211. P. Wang and S. J. Danishefsky, Promising general solution to the problem of ligating peptides and glycopeptides, J. Am. Chem. Soc., 2010, 132, 17045 CrossRef CAS PubMed .
  212. T. Wang and S. J. Danishefsky, Solid-phase peptide synthesis and solid-phase fragment coupling mediated by isonitriles, Proc. Natl. Acad. Sci. U. S. A., 2013, 110, 11708 CrossRef CAS PubMed .
  213. A. M. Levinson, J. H. McGee, A. G. Roberts, G. S. Creech, T. Wang, M. T. Peterson, R. C. Hendrickson, G. L. Verdine and S. J. Danishefsky, Total chemical synthesis and folding of all-L and all-D variants of oncogenic KRas(G12 V), J. Am. Chem. Soc., 2017, 139, 7632 CrossRef CAS PubMed .
  214. A. G. Roberts, E. V. Johnston, J. H. Shieh, J. P. Sondey, R. C. Hendrickson, M. A. Moore and S. J. Danishefsky, Fully synthetic granulocyte colony-stimulating factor enabled by isonitrile-mediated coupling of large, side-chain-unprotected peptides, J. Am. Chem. Soc., 2015, 137, 13167 CrossRef CAS PubMed .
  215. H. G. Viehe, S. I. Miller and J. I. Dickstein, Alkynylamines from halogenoalkynes and tertiary amines, Angew. Chem., Int. Ed. Engl., 1964, 3, 582 CrossRef .
  216. H. G. Viehe, Synthesis of substituted acetylenic compounds, Angew. Chem., Int. Ed. Engl., 1963, 2, 477 CrossRef .
  217. H. G. Viehe, R. Fuks and M. Reinstein, Addition reactions of alkynylamines, Angew. Chem., Int. Ed. Engl., 1964, 3, 581 CrossRef .
  218. R. Buijle and H. G. Viehe, Peptide syntheses with alkynylamines, Angew. Chem., Int. Ed. Engl., 1964, 3, 582 CrossRef .
  219. F. Weygand, W. König, R. Buyle and H. G. Viehe, Peptidsynthesen mit Inaminen, Chem. Ber., 1965, 98, 3632 CrossRef CAS .
  220. H. J. Gais, K. Hafner and M. Neuenschwander, Acetylene mit elektronendonator und elektronenakzeptorgruppen, Helv. Chim. Acta, 1969, 52, 2641 CrossRef CAS .
  221. U. Lienhard, H.-P. Fahrni and M. Neuenschwander, “Push-Pull”-acetylene für die peptid - synthese, Helv. Chim. Acta, 1978, 61, 1609 CrossRef CAS .
  222. M. Neuenschwander, H.-P. Fahrni and U. Lienhard, “Push-Pull”-Acetylene als hilfsmittel zur synthese von peptiden, Helv. Chim. Acta, 1978, 61, 2437 CrossRef CAS .
  223. H.-J. Gais, 4-Dimethylamino-3-butyn-2-one as activating agent for peptide synthesis, Angew. Chem., Int. Ed. Engl., 1978, 17, 597 CrossRef .
  224. H.-J. Gais and T. Lied, A new method for selective activation of amino, hydroxy, and mercapto carboxylic acids at the carboxyl group: Preparation of thiol and selenol esters, Angew. Chem., Int. Ed. Engl., 1978, 17, 267 CrossRef .
  225. K. A. DeKorver, H. Li, A. G. Lohse, R. Hayashi, Z. Lu, Y. Zhang and R. P. Hsung, Ynamides: a modern functional group for the new millennium, Chem. Rev., 2010, 110, 5064 CrossRef CAS PubMed .
  226. G. Evano, A. Coste and K. Jouvin, Ynamides: versatile tools in organic synthesis, Angew. Chem., Int. Ed., 2010, 49, 2840 CrossRef CAS PubMed .
  227. L. Hu, S. Xu, Z. Zhao, Y. Yang, Z. Peng, M. Yang, C. Wang and J. Zhao, Ynamides as racemization-free coupling reagents for amide and peptide synthesis, J. Am. Chem. Soc., 2016, 138, 13135 CrossRef CAS PubMed .
  228. L. Hu and J. F. Zhao, Ynamide: a new coupling reagent for amide and peptide synthesis, Synlett, 2017, 1663 CAS .
  229. T. Liu, S. Xu and J. Zhao, Recent advances in ynamide coupling reagent, Chin. J. Org. Chem., 2021, 41, 873 CrossRef .
  230. S. Xu, D. Jiang, Z. Peng, L. Hu, T. Liu, L. Zhao and J. Zhao, Ynamide-mediated peptide bond formation: mechanistic study and synthetic applications, Angew. Chem., Int. Ed., 2022, 61, e202212247 CAS .
  231. J. Yang, C. Wang, S. Xu and J. Zhao, Ynamide-mediated thiopeptide synthesis, Angew. Chem., Int. Ed., 2019, 58, 1382 CrossRef CAS PubMed .
  232. J. Yang, C. Wang, C. Yao, C. Chen, Y. Hu, G. He and J. Zhao, Site-specific incorporation of multiple thioamide substitutions into a peptide backbone via solid phase peptide synthesis, J. Org. Chem., 2020, 85, 1484 CrossRef CAS PubMed .
  233. Z. Wang, X. Wang, P. Wang and J. Zhao, Allenone-mediated racemization/epimerization-free peptide bond formation and its application in peptide synthesis, J. Am. Chem. Soc., 2021, 143, 10374 CrossRef CAS PubMed .
  234. C. L. Stevens and M. E. Munk, Nitrogen analogs of ketenes. V.1 formation of the peptide bond, J. Am. Chem. Soc., 1958, 80, 4069 CrossRef CAS .
  235. J. F. Arens and T. Doornbos, Chemistry of acetylenic ethers X. note on the use of ethoxy-acetylene for the preparation of acid anhydrides, Recl. Trav. Chim. Pays-Bas, 1955, 74, 79 CrossRef CAS .
  236. C. Wan, Y. Feng, Z. Hou, C. Lian, L. Zhang, Y. An, J. Sun, D. Yang, C. Jiang, F. Yin, R. Wang and Z. Li, Electrophilic sulfonium-promoted peptide and protein amidation in aqueous media, Org. Lett., 2022, 24, 581 CrossRef CAS PubMed .
  237. A. Sharma, A. Kumar, B. G. de la Torre and F. Albericio, Liquid-phase peptide synthesis (LPPS): a third wave for the preparation of peptides, Chem. Rev., 2022, 122, 13516 CrossRef CAS PubMed .

Footnote

These authors contributed equally.

This journal is © the Partner Organisations 2023