Skip to main content
NIHPA Author Manuscripts logoLink to NIHPA Author Manuscripts
. Author manuscript; available in PMC: 2018 Aug 9.
Published in final edited form as: Chem Rev. 2017 Jun 22;117(15):10358–10376. doi: 10.1021/acs.chemrev.7b00090

Oximes and Hydrazones in Bioconjugation: Mechanism and Catalysis

Dominik K Kölmel 1, Eric T Kool 1,
PMCID: PMC5580355  NIHMSID: NIHMS898404  PMID: 28640998

Abstract

The formation of oximes and hydrazones is employed in numerous scientific fields as a simple and versatile conjugation strategy. This imine-forming reaction is applied in fields as diverse as polymer chemistry, biomaterials and hydrogels, dynamic combinatorial chemistry, organic synthesis, and chemical biology. Here we outline chemical developments in this field, with special focus on the past ∼10 years of developments. Recent strategies for installing reactive carbonyl groups and α-nucleophiles into biomolecules are described. The basic chemical properties of reactants and products in this reaction are then reviewed, with an eye to understanding the reaction's mechanism and how reactant structure controls rates and equilibria in the process. Recent work that has uncovered structural features and new mechanisms for speeding the reaction, sometimes by orders of magnitude, is discussed. We describe recent studies that have identified especially fast reacting aldehyde/ketone substrates and structural effects that lead to rapid-reacting α-nucleophiles as well. Among the most effective new strategies has been the development of substituents near the reactive aldehyde group that either transfer protons at the transition state or trap the initially formed tetrahedral intermediates. In addition, the recent development of efficient nucleophilic catalysts for the reaction is outlined, improving greatly upon aniline, the classical catalyst for imine formation. A number of uses of such second- and third-generation catalysts in bioconjugation and in cellular applications are highlighted. While formation of hydrazone and oxime has been traditionally regarded as being limited by slow rates, developments in the past 5 years have resulted in completely overturning this limitation; indeed, the reaction is now one of the fastest and most versatile reactions available for conjugations of biomolecules and biomaterials.

Graphical abstract

graphic file with name nihms898404u1.jpg

1. Introduction

In recent years, bioconjugation reactions have become an indispensable tool for studying and manipulating biomolecules in vitro and in vivo. Reactions that allow for covalent tagging of biomolecules in their native environment with high selectivity and specificity are of particular interest, since this can yield new insights into cellular processes.1,2 Among many other labeling reactions that can be done under physiological conditions,3,4 the formation of oximes and hydrazones is nowadays commonly used to facilely link biomolecules to various probes. Generally, oximes 4 and hydrazones 5 can be easily formed from the corresponding α-effect nucleophile (alkoxylamine 1 or hydrazine 2, respectively) and a carbonyl compound (aldehyde or ketone 3, Scheme 1) with water being the sole byproduct.

Scheme 1. Formation of Oximes 4 or Hydrazones 5 from Alkoxyamines 1 or Hydrazines 2 and Aldehydes or Ketones 3.

Scheme 1

Notably, those two reactions are considerably older than many other bioconjugation reactions. As early as 1882, oximes and their formation have been intensively studied by Meyer and Janny.58 Shortly later, the term hydrazone was coined by Fischer in 1888.9 Due to the simplicity of these venerable ligation reactions, hydrazones and oximes have also had a pervasive influence on numerous other research fields ever since. They have been extensively used for 18F-labeling of peptides and proteins,10 syntheses of molecular switches, metallo-assemblies and sensors,11,12 treatment of organophosphorus poisoning,13,14 decoration of nanoparticles,15,16preparation of palladacycle precatalysts,17,18 syntheses of alicycles and heterocycles, 1923 and derivatization of carbohydrates for mass spectrometric analysis.24 Other specific aspects and applications of oximes and hydrazones that have previously been reviewed include their biological activities,13,25 their utility as valuable synthetic intermediates,2630 the multifunctionalization of biological scaffolds,31 and the preparation of protein conjugates.32

This review will highlight the most recent strategies to improve oxime/hydrazone bioconjugation reactions with a particular emphasis on approaches that enable rapid conjugation under physiological conditions (see the Supporting Information for a selection of detailed, exemplary oxime/hydrazone bioconjugation protocols). After a brief account of the strengths and features of these linker moieties, synthetic methods that incorporate α-effect nucleophiles or carbonyl groups into diverse classes of biomolecules will be described. The mechanism of the formation of oximes and hydrazones will be explained in detail. The stability of those conjugates will subsequently be critically analyzed, since they can theoretically undergo hydrolysis in aqueous solution. Finally, the latest advances in the designs of efficient bifunctional catalysts as well as aldehydes and α-effect nucleophiles with intrinsically high reactivity will be examined.

2. Utility of Oxime/Hydrazone Ligation

Ideally, ligation reactions have to meet certain requirements for greatest utility in bioconjugations. The new linkage should preferably be stable across a broad range of biological environments and should be small in size to minimize its steric hindrance. Furthermore, the ligation should proceed with high selectivity and specificity, which means that the two bond-forming moieties should be rather scarce among naturally occurring molecules and not too reactive to avoid random labeling. Moreover, the reaction should be bioorthogonal; 1,33,34 i.e., the conjugation can be conducted in living systems without interfering with biological processes or metabolites.

Oxime and hydrazone bond formation meets several of these requirements and is therefore frequently applied for bioconjugations. The oxime/hydrazone tether is very small in size, consisting only of three non-hydrogen atoms (C=N–X with X = O or NH), and thus constitutes a minimal perturbation of the native biomolecule. Notably, the stability of this linkage must be regarded with caution. Although Scheme 1 depicts the oxime/hydrazone formation as a straightforward condensation, the reaction is in fact reversible and the conjugates can undergo hydrolysis in aqueous media (see section 4 for the detailed mechanism). However, the hydrolytic stability can be effectively tailored by making structural changes to the linking moiety, for example, by employing oximes, which generally display much higher stability than hydrazones.35 In addition, the pH of the solution also has a marked influence on the conjugates' stability. As a consequence, the half-life of oxime and hydrazone conjugates can span several orders of magnitude. Hence, the reversibility of the oxime/hydrazone conjugation can be exploited as a powerful feature for the controlled release of biologically active molecules. The diverse stability factors will be discussed in greater detail in section 5.

Oxime- and hydrazone-based conjugations are occasionally listed as bioorthogonal reactions.3642 However, the bioorthogonality of these ligations is not unconditional and must be treated with caution. Although α-effect nucleophiles are rarely found in nature,43 aldehydes and ketones are common. Importantly, the hemiacetal of every carbohydrate is in equilibrium with the respective carbonyl compound, which makes many of them amenable to reactions with α-nucleophiles.4446 Furthermore, numerous aldehydes and ketones (e.g., ketone bodies)47 are generated through normal and pathogenic metabolism.48 Specific conditions, like oxidative stress,4951 are also known to trigger the production of a variety of aldehydes. Consequently, the concentration of cellular carbonyl compounds can span several orders of magnitude, from low nanomolar to high micromolar.52

To gauge their bioorthogonality, the reactivity of both reaction partners must be scrutinized as well. Due to the α-effect,5355 hydrazines and alkoxyamines are strongly nucleophilic and readily react with a wide range of electrophiles. Aldehydes and ketones, on the other hand, can form imines with various amines, albeit in a reversible manner, which also explains their potential cytotoxicity.56

Although the oxime/hydrazone conjugation is not strictly bioorthogonal, it can be effectively bioorthogonal, if highly reactive electrophiles and aldehydes/ketones are only present at low concentration.15 Fortunately, this requirement is sufficiently met in many biological environments, which renders this reaction truly versatile.

3. Functionalization of Biomolecules For Oxime/Hydrazone Ligation

In order to undergo an oxime/hydrazone conjugation reaction, the respective biomolecules must contain either a carbonyl moiety or an α-nucleophile. The former can be found in some biologically relevant molecules, most notably reducing sugars, but typically methods that install either of these two reactive groups are prerequisite. The following section will give a brief overview of strategies that allow for the site-selective decoration of different biomolecules with the respective functional groups.

3.1. Functionalization with Aldehyde and Ketone Moieties

An old but still widely used method for the functionalization of biomolecules with carbonyl groups is the oxidative cleavage of vicinal diols with sodium periodate (Scheme 2).57 This strategy can be used for facile modification of the glycoproteins of the glycocalyx by oxidizing sialic acid residues.58

Scheme 2. Periodate-Mediated Oxidative Cleavage of Vicinal Diols 6 or 1,2-Amino Alcohols 7.

Scheme 2

Nucleic acids that display 3′-ribonucleotides (12) are likewise known to undergo oxidation to yield the respective dialdehyde 13, which exists in its hydrated 1,4-dioxane form 14 in aqueous solution (Scheme 3).59,60 In addition, 5-formylcytosine (5-fC) and 5-formyluracil (5-fU) can be used as building blocks for the site-specific aldehyde labeling of synthetic DNA. The latter is more reactive, which allows for selective tagging of DNA that contains both of these naturally occurring modifications.61

Scheme 3. Periodate-Mediated Oxidative Cleavage of 3′-Ribonucleotide 12 and the Equilibrium between the Resulting Dialdehyde 13 and 1,4-Dioxane 14.

Scheme 3

Similarly, 1,2-amino alcohols 7 are susceptible to oxidative cleavage as well (see Scheme 2), which can be exploited for the selective labeling of peptides with N-terminal serine or threonine.62 As an extension, the unnatural amino acids 1519 and the 3′-appendix 20 have been developed, which can be incorporated into synthetic peptides or nucleic acids to enable site-selective side chain or backbone cleavage with periodate under mild conditions (Figure 1).6366

Figure 1.

Figure 1

Generic structures of 1,2-amino alcohol-containing peptides 1519 and nucleic acid 20, which can be cleaved with periodate.

A wide variety of ketone-containing amino acid congeners have been introduced into peptides and proteins (Figure 2). They can be installed site-specifically either by classical solid-phase peptide synthesis67 or, as an alternative, by using the biosynthetic machinery via amber-stop codon suppression.6873

Figure 2.

Figure 2

Selection of ketone-containing unnatural amino acids 2127, which have been incorporated into proteins.

Furthermore, Francis and co-workers have demonstrated that proteins can be modified postsynthetically with a ketone via azo coupling with keto-diazonium ion 29 (Scheme 4).74,75 Importantly, this modification targets exclusively the electron-rich aromatic ring of tyrosine residues.

Scheme 4. Site-Selective Modification of Tyrosine Residues by Azo Coupling with Diazonium Ion 29.

Scheme 4

Cysteine-containing peptides and proteins can be readily functionalized with the enzyme protein farnesyltransferase (PFTase) and the farnesyl pyrophosphate derivative 32 (Scheme 5).76,77

Scheme 5. Enzymatic Aldehyde-Modification of Cysteine Residues 31 with Pyrophosphate 32 and Protein Farnesyltransferase (PFTase)76.

Scheme 5

In that method, the cysteine residue is required to be positioned at the C-terminus as a part of a tetrapeptide sequence termed the CAAX-box. In addition, Bertozzi and co-workers have shown that proteins can be equipped with an aldehyde moiety in living cells by recruiting the formylglycine-generating enzyme (FGE) to the 6-amino acid tag LCTPSR.78 Concomitantly, methods for the selective functionalization of the N-terminus of peptides and proteins have been devised. A biomimetic approach enables N-terminal transamination with pyridoxal phosphate (35) as the carbonyl oxygen donor (Scheme 6).79 This remarkable selectivity over lysine chains arises from the increased acidity of the N-terminal α-C–H bond due to the adjacent carbonyl moiety, which facilitates the isomerization from imine 36 to 37.

Scheme 6. Pyridoxal Phosphate (35)-Mediated Oxidation of N-Terminal Amino Acids 34 via Biomimetic Transamination79.

Scheme 6

The same result can be obtained by site-specific oxidation of N-terminal amine 34 with oxone, yielding oxime 40, which can subsequently undergo exchange with other alkoxyamines (Scheme 7).80

Scheme 7. Selective Oxidation of N-Terminal Amines 34 with Oxone80.

Scheme 7

3.2. Functionalization with Alkoxyamine, Hydrazide, and Hydrazine Moieties

Due to the fact that most biomolecules can be readily functionalized with aldehyde or ketone moieties,81 fewer methods for the implementation of α-effect nucleophiles have been developed. Thus, the introduction of α-nucleophiles into biomolecules is typically limited to synthetic nucleic acids and peptides. Several phosphoramidite derivatives have been applied for the 5′-modification of synthetic nucleic acids (Figure 3).8286

Figure 3.

Figure 3

Selection of α-effect nucleophile-containing phosphoramidites 4248 for 5′-modification of nucleic acids. Trt = trityl; Mmt = 4-methoxytrityl; Dmt = 4,4′-dimethoxytrityl; NPhth = phthalimidyl.

At the same time, the building blocks 4953 have been designed for internal modification of nucleic acid strands (Figure 4). The α-nucleophile can be tethered to the 1′-position either as a replacement (49)87 or as a modification of the nucleobase (50 and 51).88,89 Furthermore, ribonucleotide building blocks 52 and 53 can be obtained by functionalization of the 2′-hydroxy function.89,90

Figure 4.

Figure 4

Examples of alkoxyamine-containing phosphoramidites 49-53 for nucleic acid synthesis. Dmt = 4,4′-dimethoxytrityl; NPhth = phthalimidyl; Bz = benzoyl.

Synthetic peptides can be easily equipped with a C-terminal hydrazide moiety by cleaving the respective solid-supported peptide 54 from the Wang resin with hydrazine (Scheme 8).91 Alternatively, the peptidyl hydrazide 55 can be directly synthesized on the corresponding supports 56 or 57.92,93 The final cleavage can be achieved either photolytically or under acidic conditions, respectively.

Scheme 8. Solid-Phase Synthesis of Peptidyl Hydrazide 55 via Cleavage from Resin 54, 56, or 57.

Scheme 8

N-Terminally decorated peptides can be generated by using N,N,N′-tris(Boc)hydrazinoacetic acid for the last coupling step of their solid-supported assembly.94 Additionally, unnatural amino acids 5864 have been created to generate synthetic peptides with internal α-nucleophiles (Figure 5).64,9598

Figure 5.

Figure 5

A selection of α-effect nucleophile-containing unnatural amino acids (5864) that have been incorporated into proteins.64,9598

4. Mechanism of Oxime/Hydrazone Formation

In order to develop and apply oxime/hydrazone formation to its full potential, it is crucial to understand the underlying mechanism. As shown in sections below, this knowledge helps users understand the equilibria and rates involved and is empowering for future catalyst development and for designing fast-reacting substrates.

In the 1930s, seminal works on the formation and hydrolysis of semicarbazones brought first insights into the mechanism of this reaction.99,100 This was followed by important mechanistic studies by Jencks in the 1960s.101 An important finding was that this conjugation is fully reversible. The following mechanistic reflections will outline the formation of oximes 4 and hydrazones 5; the hydrolysis reaction is obtained by reversing the order of the reaction steps.

The reaction commences by a proton-catalyzed attack of the α-effect nucleophile 1 or 2 on the carbonyl carbon atom of electrophile 3 (Scheme 9). Upon proton transfer, the hemiaminal 67 or 68—the tetrahedral intermediate—is obtained. This hemiaminal can undergo dehydration via protonation of the hydroxyl function and subsequent elimination of water. The resulting protonated intermediate is represented by two resonance forms (71/72). The final deprotonation yields oxime 4 or hydrazone 5.

Scheme 9. Standard Mechanism for the Formation of Oximes 4 or Hydrazones 5 from Carbonyl Compound 3 and Alkoxyamines 1 or Hydrazines 2, Respectively.

Scheme 9

Typically, this reaction is under general acid catalysis.99,100,102 Jencks and co-worker proposed the transition state 73 for the acid-catalyzed nucleophilic attack on the carbonyl carbon atom, which leads to the protonated intermediate 65 or 66 (Scheme 10).103 According to their detailed studies on semicarbazone formation, the proton transfer from a general acid to the carbonyl oxygen atom occurs concurrently with the attack of the nucleophilic reagent. Thereby, the termolecular complex 73/74 would likely arise from two consecutive bimolecular reactions, rather than a ternary collision. Facilitated by hydrogen bonding, the carbonyl compound 3 can exist in a pre-equilibrium complex with the general acid, which would be followed by the C–N bond-forming attack of the α-nucleophile 1 or 2.

Scheme 10. Proposed Transition State 73/74, Which Leads to the Formation of the Protonated Intermediate 65 or 66103 a.

Scheme 10

aHA = generic acid.

Attack of strong α-nucleophiles on aldehydes and ketones is a fast reaction and is not rate-limiting in the large majority of cases. In the pH range from ca. 3 to 7, the acid-catalyzed dehydration of the tetrahedral intermediate 67/68 is typically the rate-determining step.103 This is supported by the observation that the formation of oximes and hydrazones becomes concentration-independent if the excess of α-nucleophile is sufficiently high.104 More supporting evidence is provided by density functional theory (DFT) calculations.105

Although the reaction rate for hydrazone/oxime formation can be greatly accelerated under general acid catalysis, the reaction slows down again if the pH of the reaction medium is too low. This is due to the fact that the α-nucleophile 1 or 2 is in equilibrium with up to two different protonation states (Scheme 11). Hydrazines 2 can undergo protonation at either of their nitrogen atoms, forming hydrazinium ion 76 or 78, respectively.106 For alkoxyamines 1, the protonation occurs exclusively at the terminal nitrogen atom to yield alkoxyammonium ion 77. Owing to the low basicity of oxygen, the formation of oxonium ion 75 is of no importance. Since these protonated species do not lead to product formation, due to their low nucleophilicity, the reaction slows at more acidic pH (typically under pH 3).107 Under such conditions, the attack of the α-effect nucleophile 1 or 2 will become rate-determining.103 Consequently, this reaction is fastest when the acidity of the solution strikes a balance between fast acid-catalyzed dehydration of hemiaminal 67 or 68 and negligible formation of unreactive protonated α-effect nucleophiles 7678.108 For the formation of oximes and hydrazones, a pH of ca. 4.5 is typically advantageous.109,110 However, many biological applications require this ligation to proceed under physiological conditions, which is challenging due to the slow reaction rate at neutral pH and the low concentrations of the reaction partners.111 Thus, the development of efficient catalysts and fast-reacting substrates, which could lead to more rapid bond formation under biological conditions, has been a strong focus of recent work (see sections 6 and 7).

Scheme 11. Equilibrium between the α-Nucleophiles 1 and 2 and the Respective Protonated Species 75–78.

Scheme 11

5. Stability Factors

Among commonly used bioconjugation reactions, a unique feature of hydrazones and oximes is that the stability of the linkage can be fine-tuned by virtue of the general reversibility of this ligation. Since the hydrolysis of oximes and hydrazones can be obtained by reversing the order of the reaction steps (Scheme 9), it can be easily seen that the back-reaction is initiated by protonation of the imine nitrogen.107 Mechanistic studies by Kalia and Raines showed that the negative inductive effect of the group X attached to the imine-forming nitrogen directly influences the stability of the respective conjugate.35 If X is an electronegative heteroatom, the basicity of the imine nitrogen is diminished. This explains a commonly observed trend: imines (X = CH2) hydrolyze readily under aqueous conditions, whereas oximes and hydrazones are much more stable due to the negative inductive effect of the additional heteroatom (X = O or NH), with the former being the most stable conjugate in this series due to the high electronegativity of oxygen [χp(O) = 3.5 vs χp(N) = 3.0]. This reversibility can be an appealing feature of the oxime/hydrazone conjugation, which enables valuable applications, like the controlled release of small molecules from suitable drug delivery platforms.112114 Due to their inherently greater stability, oximes are typically preferred in bioconjugation reactions if a more stable linkage is required, while the more labile hydrazones are employed for the controlled release of biologically active molecules. Moreover, in principle one can use a catalyst to “switch on” this reversal in an otherwise relatively stable linkage, thus allowing the controlled breaking of the bond.

The following two subsections will discuss intrinsic and extrinsic factors that govern this bond stability.

5.1. Intrinsic Stability Factors

As outlined above, the intrinsic stability of oxime and hydrazone conjugates is directly influenced by steric and, more importantly, electronic factors in the vicinity of the linkage. In general, conjugates obtained from ketones have a higher stability than the respective aldehyde derivatives.115 The equilibrium constants of hydrazones are typically in the range of 104–106 and >108 M−1 for oximes.116 Figure 6 depicts the equilibrium constants of some prototypical oximes, hydrazones, and semicarbazones for comparison.117119

Figure 6.

Figure 6

Equilibrium constants of some prototypical oximes, hydrazones, and semicarbazones. Keq = c(conjugate)/[c(aldehyde) × c(α-nucleophile)].

The thermodynamic stability of oximes formed from different aldehydes and ketones increases in the following order: acetone < cyclohexanone ∼ furfural ∼ benzaldehyde < pyruvic acid (Figure 7).108 Aromatic aldehydes and derivatives of α-oxo acids are thus frequently used for bioconjugations.31,200

Figure 7.

Figure 7

Stability trend in a series of oximes 8791.108

The substituents at the α-nucleophile moiety can also have a distinct influence on the stability. In a series of isostructural conjugates, the first-order rate constant for the hydrolysis of oxime 95 was 160-fold lower than semicarbazone 94, 300-fold lower than acetylhydrazone 93, and 600-fold lower than methylhydrazone 92 (Figure 8).35 Hydrazones bearing two electron-withdrawing groups were found to be especially labile and hydrolyzed rapidly even at neutral pH.118

Figure 8.

Figure 8

Comparison between the kinetic stabilities in a series of isostructural conjugates 9295.35 The relative first-order rate constants krel for the hydrolysis are given [krel(oxime 95) = 1].

If required, the stability of oxime and hydrazone conjugates can be enhanced by reducing their C=N double bond, e.g. with sodium cyanoborohydride.121,122 Kalia and Raines made the interesting observation that trimethylhydrazonium ions 96 are exceptionally stable and even exceed the hydrolytic stability of oximes.35 In this case, the intrinsically charged trimethylammonium group inductively disfavors the protonation of the imine nitrogen (Scheme 12). However, those conjugates are not readily accessible in a biological environment, because the quaternization involves the strong and toxic electrophile methyl iodide.

Scheme 12. Proposed Mechanism for the Hydrolysis of Trimethylhydrazonium Ions 9635a.

Scheme 12

aProtonation of the imine nitrogen atom is disfavored due to the presence of an additional positive charge.

Simanek and co-workers have investigated the hydrolytic stability of a novel series of hydrazones 99, which were formed from s-triazinylhydrazines and various aldehydes/ketones (Figure 9).123 At pH >5, those conjugates had a lower stability compared to structurally analogous acetylhydrazones 100.

Figure 9.

Figure 9

Comparison of hydrolytic stability between s-triazinylhydrazones 99 and acetylhydrazones 100 at different pH values.123

Interestingly, this stability trend reversed around pH 5. The authors noted that this observation can be explained by the protonation of the s-triazinyl moiety (pKa ∼ 5), which subsequently disfavors the required protonation of the imine nitrogen due to the presence of a delocalized positive charge (Scheme 13).

Scheme 13. Protonation of the s-Triazinyl Hydrazone 99, Which Lowers the Basicity of the Imine Nitrogen Due to the Adjacent Delocalized Positive Charge123.

Scheme 13

Bertozzi and co-workers have recently presented a variation of the classical oxime formation, termed Pictet–Spengler ligation, which yields hydrolytically stable conjugates (Scheme 14).124 In analogy to the Pictet–Spengler reaction, this conjugation involves the indolyl-substituted nucleophile 102, which reacts with aldehyde 11 to form oxyiminium ion 103. Ensuing intramolecular C–C bond formation results in the tricyclic heterocycle 104. This Pictet–Spengler ligation was also applied for the conjugation of isostructural hydrazine nucleophiles.125

Scheme 14. Pictet–Spengler Ligation between Alkoxyamine 102 and Aldehyde 11124.

Scheme 14

Recently, Gillingham and co-workers found that aromatic aldehydes/ketones 105, which feature a boronyl moiety in the ortho position, will yield 4,3-borazaroisoquinolines (BIQs) 107 upon reaction with hydrazines 106 (Scheme 15).126 Most notably, those BIQs proved to form irreversibly and thus they were found to be indefinitely stable in aqueous solution. This approach could be further developed into a fluorogenic reaction by using 2-hydrazinylphenol as a substrate, which results in the formation of highly emissive tetracyclic BIQ 108.

Scheme 15. Formation of 4,3-Borazaroisoquinolines (BIQs) 107 from Aromatic Boronic Acids 105 and Hydrazines 106126a.

Scheme 15

aBIQ 108 with extended delocalized π-electron system displays fluorescence emission.

Another irreversible condensation takes place between aldehydes 11 and α-aminooxy acetohydrazides 109 (Scheme 16).127

Scheme 16. Synthesis of Stable 1,2,4-Oxadiazinan-5-ones 110 from Aldehydes 11 and α-Aminooxy Acetohydrazides 109127.

Scheme 16

5.2. Extrinsic Stability Factors

The stability of the oxime/hydrazone linkage is also influenced by some external factors. As expected, the hydrolysis and transimination of hydrazones and oximes are significantly faster under acidic conditions and at elevated temperature.48,128131 Senter and co-workers studied the release of the antimitotic agent auristatin E from a monoclonal antibody conjugate and found that the half-life time t1/2 of the respective hydrazone linkage dropped from t1/2 = 183 h at pH 7.2 to t1/2 = 4.4 h at pH 5.132 This can be a very useful feature for targeted drug delivery. Hydrazone linkers have been shown to release their covalently bound payload primarily in the acidic environment of specific organelles, like the endosomes (pH 5.5–6.2) or lysosomes (pH 4.5–5.0).133 For in vivo applications of oximes and hydrazones, it is also important to consider the composition of the biological medium they are exposed to. It was shown that various aroylhydrazones rapidly decomposed in plasma, whereas they were relatively stable in phosphate-buffered saline (PBS).134 Low molecular weight compounds (<30 kDa) and, to a lesser extent, plasma proteins (e.g., albumin) were identified as the origin for this marked stability difference. This is in line with previous reports that proteinogenic amino acids can catalyze hydrazone hydrolysis in cell culture medium and serum.135

6. Catalysts For Oxime/Hydrazone Formation

Since reactants for oxime and hydrazone formation may exist at low concentrations in some applications, the rates of these reactions, and finding mechanisms to enhance these rates, are important from a practical standpoint. For example, it has been pointed out3 that biomolecular conjugation and/or labeling is often carried out under conditions in which the biomolecule quantity and concentration are limiting. Indeed, it is common to work with biomolecules at low micromolar concentrations or below. In such situations, reactions must proceed at an appreciably rapid inherent rate in order to produce high reaction yields.

Many hydrazone and oxime formation reactions proceed with second-order rate constants of 0.01 M−1 s−1 or below at neutral pH, which is slower than many of the faster cycloaddition-based bioconjugation reactions.136139 This is particularly true of oxime and acylhydrazone formation, which (in the absence of catalysts or specialized reactants) typically proceeds more slowly than hydrazone formation. This was recently illustrated by Bode and co-workers, who compared the second-order rate constants of several widely used chemo-selective ligation reactions (Figure 10).140 The kinetic studies were carried out with functionalized heptapeptides, which were reacted with azo dyes featuring the complementary reactive group. The two reactants of each conjugation were used in a 1:1 ratio. As expected, at neutral pH the formation of oxime 111 was significantly slower than other bioconjugation reactions. Notably, the analogous Pictet–Spengler conjugate 112 forms 10 times faster than oxime 111.

Figure 10.

Figure 10

Comparison of reaction conditions and second-order rate constants of several bioconjugation reactions. The kinetic studies were carried out with a 1:1 stoichiometry of the respective reactants. TBTA = tris[(1-benzyl-1H-1,2,3-triazol-4-yl)methyl]amine.140

To overcome those shortcomings, there has been considerable interest of late in developing catalysts to speed hydrazone and oxime formation reactions and in developing fast-reacting substrates. Both of those strategies lead to greatly improved reaction rates (Figure 11).141,142 The former of these topics is discussed here and the latter in section 7.

Figure 11.

Figure 11

Apparent second-order rate constants for the formation of hydrazones 116118. The apparent second-order rate constants were derived from pseudo-first-order kinetic measurements, where one the reaction partners was present in 50-fold excess. Conditions: buffer (pH 7.4)/DMF (10:1), room temperature.141,142

It is important to note that catalysts for the bond-forming reaction also by definition catalyze the back-reaction as well. This allows them to be useful in promoting equilibration among different hydrazones in exchange reactions, some of which are mentioned below. Hydrazone-based dynamic combinatorial libraries can thus greatly benefit from the application of appropriate catalysts to achieve thermodynamic control under mild conditions.143,144 Second, catalysts for imine formation can be useful not just for hydrazone formation, but also in related imine hydrolysis reactions, such as removal of formaldehyde adducts from biomolecules of formalin-fixed tissue specimens, also described below.

In the search for improved catalysts and new catalytic frameworks, it is useful to have in hand simple assays of hydrazone/oxime formation with easily measured signals. To that end, multiple color-changing hydrazine substrates 119121 have been described,142,145,146 and the FRET-based fluorescence probe 122 for evaluating catalysts was recently reported as well (Figure 12).147

Figure 12.

Figure 12

Molecular structures of α-nucleophiles that either undergo color change (119 and 120) or interact with a fluorescently labeled substrate (121 and 122) upon hydrazone/oxime formation.142,145147

6.1. Monofunctional Catalysts

Jencks was a pioneer in the study of the catalysis of hydrazone (semicarbazone) formation.101 He documented the rate-determining step of the reaction, namely, breakdown of the tetrahedral intermediate (see Scheme 9), and reported on mechanisms of catalysis as well. In 1962, he described the general acid catalysis of semicarbazone formation using simple organic acids of varying pKa.103 Although this established that proton donation can speed the reaction, the studies were carried out at acidic pH (in the range 2–4) and so may not be highly relevant to current interests in biological application.

Also importantly, Jencks early on documented the fact that aniline acts as a nucleophilic catalyst for semicarbazone formation at pH 2–4.148 At 4 mM (pH 2.6) aniline enhanced the rate by 3.5-fold. Substituted anilines were also tested. Jencks was the first to establish that the mechanism occurred via nucleophilic catalysis (Scheme 17). In this mechanism, aniline (123) adds first to the carbonyl, producing the first tetrahedral intermediate, 124, which breaks down to form the catalyst imine intermediate 125. This then is subject to attack by the α-effect nucleophile 1/2, resulting in the second tetrahedral intermediate, 126/127, which then breaks down by eliminating the catalyst 123 to form the ultimate product 4/5. Although this pathway requires more steps and involves more intermediates than the uncatalyzed reaction, the energy of the highest transition state barrier is lower than that of the uncatalyzed reaction, thus speeding the reaction overall. Aniline (123) acts as a successful catalyst because it has a nucleophilic amine, adding rapidly to the carbonyl. In addition, because the lone pair electrons of the amine are delocalized into its aromatic ring, it behaves as an effective leaving group. The rate-determining step of the catalyzed reaction is still a breakdown of a tetrahedral intermediate, but this proceeds more favorably in the catalyzed reaction.

Scheme 17. Generally Recognized Mechanism for the Aniline-Catalyzed Formation of Oximes 4 or Hydrazones 5.

Scheme 17

In the same study, Jencks reported that glycine also acts as a nucleophilic catalyst for the reaction, albeit less effectively. More recently, Cid reported another simple amine, pyrrolidine, as a catalyst for oxime and hydrazone formation. The amine was employed at relatively high concentrations, and no reaction rates were reported.149 Subsequent studies (described below) have consistently found aromatic amines (aniline derivatives) to have more favorable catalytic properties than aliphatic examples.

Subsequent to Jencks' early study, many laboratories have investigated arylamines (anilines) as catalysts, but with a trend toward biological applications and to pH values closer to neutral. An important study by Dawson and co-workers documented the use of aniline (123) both at pH 4.5 and 7.0 to speed oxime ligations involving modified peptides.150 The catalyst was employed at 10 or 100 mM. At the high aniline concentration, reaction rates were increased by a factor of up to 40 at neutral pH and even more (400-fold) at pH 4.5. Neutral-pH reactions with lower catalyst concentration were not described, possibly because of the slow reaction rate.

Subsequent to Dawson's initial study, several additional reports have appeared further documenting catalysis by the parent compound aniline and its use in a number of applications. Dawson's own laboratory used aniline (10 mM, pH 5.7) to catalyze hydrazone formation in dynamic covalent chemistry of peptides.151 In a more recent study, Varghese and co-workers used a combination of aniline and acetate catalysis to enhance coupling of aminooxy compounds to ketones.152 Meijler and co-workers described aminooxy-substituted BOD-IPY dye labeling of LasR receptor protein in bacteria with aniline (1 mM) catalysis.153 Moving to eukaryotic cells, Dawson, Paulson, and their co-workers employed 10 mM aniline in the cell-surface labeling of sialylated glycoproteins, with little toxicity observed after 90 min incubation with K88 or CHO cells (Figure 13).58 Finally, Domaille and Cha described the tethering of aniline on DNA and the use of DNA duplex formation to increase its local concentration relative to an acylhydrazide group during its reaction with nitrobenzaldehyde.154

Figure 13.

Figure 13

Confocal microscopy images of untreated and periodate-treated (1 mM at 4 °C for 5 min) CHO cells subjected to ligation with alkoxyamine-functionalized biotin (100 μM) in the absence or presence of aniline (10 mM) at 4 °C for 90 min. Cells were permeabilized and stained with dye–streptavidin conjugate (green) and DAPI (blue, nuclei). Scale bars are 15 μm. Reprinted in part with permission from ref 58. Copyright 2009 Nature Publishing Group.

It is now clear that substitutions on the aryl ring of aniline can enhance its catalytic efficiency. Dawson and co-workers' early study150 tested p-methoxyaniline (128, 100 mM) at neutral pH due to its higher pKa (5.3 vs 4.6 of aniline). As pointed out, enhanced basicity should promote protonation of the aniline Schiff base intermediate. Dawson did not directly compare aniline (123) to its p-methoxy analogue 128, however. Jensen and co-workers later confirmed its superiority at neutral pH in catalyzing oxime formation with carbohydrates, although no rate comparisons were given.155 A recent study of hydrazone exchange reactions in the detection of biological aldehydes showed that p-methoxyaniline (128, 10 mM) enhanced the exchange rate by a factor of 2.6-fold relative to aniline at pH 7.48 Because of its enhanced performance, p-methoxyaniline (128, 100 mM) was employed in the pH 7 spin labeling of a human sulfite oxidase enzyme.156

Recent studies have explored numerous multisubstituted aniline derivatives in catalysis of oxime and hydrazone formation (Figure 14), and some of these have improved performance over both aniline (123) and p-methoxyaniline (128). 4-Aminophenylalanine (129) was shown to catalyze the reaction between formyl amino acids and hydrazine-substituted fluorophores, albeit at a rate slower than that of aniline.157 More recently, Yuen et al. tested 16 different catalysts for hydrazone exchange reactions (5 mM, pH 7) and found that both 5-aminoindole (130) and 2,4-dimethoxyaniline (131, 2,4-DMA) were superior to 4-methoxyaniline (128), besting the parent aniline by nearly a factor of 3 in rate.48 Interestingly, 2,4-dimethoxyaniline (131) was far superior to the 3,5-dimethoxy isomer 132, by a factor of over 50, suggesting the importance of ortho and para electron donation enhancing the basicity of the amine. Notably, 2,4-DMA (131, 10 mM) showed little toxicity to human cells and was employed by Yuen et al. to perform intracellular hydrazone exchange reactions, using a dark hydrazone (“DarkZone”) dye to label and image cellular aldehydes (Figure 15).48

Figure 14.

Figure 14

Selection of substituted anilines 128135, which have been shown to be efficient catalysts for the formation of oximes/hydrazones.

Figure 15.

Figure 15

Intracellular imaging of acetaldehyde by hydrazone exchange catalyzed by the cell-permeable catalyst 2,4-DMA. HeLa cells were incubated with 20 μM of a fluorogenic hydrazone dye, which undergoes hydrazone exchange with acetaldehyde (varying concentrations, 0.1–2 mM), and 10 mM of 2,4-DMA (131) for 1 h. Scale bars are 20 μm. Reprinted in part with permission from ref 48. Copyright 2016 American Chemical Society.

A number of research groups have observed good catalytic activity in phenylenediamines, albeit with some drawbacks in stability and toxicity. Distefano and co-workers found that m-phenylenediamine (133) was approximately 2-fold faster than aniline (123) as a catalyst of oxime formation and offered the advantage of greater solubility, allowing the application in bioconjugations at up to 900 mM catalyst concentration.158 The same group went on to describe protocols for protein labeling using phenylenediamines.159 Baca and co-workers extended this finding by testing 12 different aniline and phenylenediamine derivatives and found p-phenylenediamine (134) to be superior to aniline (123) over a range of pH values, with a 19-fold faster rate than aniline at oxime formation measured at neutral pH.160 Yuen et al. found p-phenylenediamine (134) to be nearly 5-fold faster than aniline (123) at hydrazone exchange reactions.48 A drawback of p-phenylenediamine (134) and similar compounds is their oxidative instability, which results in their conversion to p-benzoquinones in air. Indeed, commercial samples are often brown in color, and reactions containing them turn brown over time,48 likely because of this reaction. Moreover, p-phenylenediamine (134) was shown to cause ∼60% cellular toxicity at 10 mM after only 1 h exposure.48 Meta-substituted isomers are less likely to suffer the oxidative instability; an example of such a catalyst is 3,5-diaminobenzoic acid (135, 3,5-DABA), which was found by Crisalli and Kool to be an effective catalyst of hydrazone and oxime formation, yielding reaction rates ∼3-fold greater than that of aniline (123) at pH 7.4.161

6.2. Bifunctional Catalysts

Developing catalysts with higher activity remains attractive not only for enhancing reaction rates further, but also for their utility at low concentrations, suppressing potential toxicity and background reaction effects that might occur at high concentrations. In this light, Crisalli and Kool in 2013 described the discovery of bifunctional catalysis in hydrazone and oxime formation.161 This bifunctionality was defined as two nearby functional groups playing a direct role in catalysis. Over 20 potential catalysts containing a nucleophilic amine were studied, and anthranilic acids were found to have the highest activity of those tested with an aryl aldehyde substrate. Controls replacing the carboxy group with a nitrile or amide group showed no activity, suggesting a mechanism involving general acid catalysis by the o-carboxy group at the transition state 136 (Scheme 18). Testing of varied anthranilic acids showed that substitution by electron-donating groups on the aryl ring enhanced catalysis further at neutral pH, consistent with the need for a more basic and nucleophilic amine group. 5-Methoxyanthranilic acid (137, 5MA) was shown to have activity 6-fold greater than aniline (123) in hydrazone formation at pH 7.4 using only 1 mM catalyst (Figure 16).161

Scheme 18. Proposed Transition State 136, Which Leads to an Activated Imine Intermediate162.

Scheme 18

Figure 16.

Figure 16

Conversion of color-changing hydrazine 119 with 4-nitrobenzaldehyde (150) to the corresponding hydrazone in the presence of selected catalysts. Reprinted with permission from ref 161. Copyright 2013 American Chemical Society.

Subsequent work in multiple laboratories has confirmed the superior catalysis of anthranilates over anilines. McKay and Finn described the testing of polyvalent and dendrimer-based anthranilic acid catalysts, confirming superior activity over anilines in dendrimer labeling.163 Langenhan and co-workers used mono- and bifunctional aniline derivatives for fluorescent conjugation of aldose sugars, and confirmed 5MA as the most effective catalyst among 10 tested.164

Crisalli and Kool followed up on their earlier work to investigate the mechanism of bifunctional catalysis in hydrazone and oxime formation, resulting in the design of further improved catalysts.142 To test their hypothesis that proton donors ortho to the nucleophilic amine group of aniline provide general acid/base catalysis during the reaction, they synthesized a series of catalysts having ortho groups of varied pKa (Figure 17). They found that the two acidic groups (tetrazole 138 and phosphonate 139) having a pKa closest to neutral (i.e., the solution pH) were the most effective catalysts. This is strong evidence of the proposed general acid/base mechanism of catalysis (Scheme 18).

Figure 17.

Figure 17

Comparison of relative reaction rates in a series of bifunctional catalysts 137139 and aniline (123).142,161 The relative reaction rates refer to the hydrazone formation between the chromogenic, aromatic hydrazine 119 and 4-nitrobenzaldehyde (150) (uncatalyzed reaction krel = 1).

This mechanistic study resulted in new phosphonate and tetrazole catalysts having high activity with aryl and alkyl aldehyde substrates. For example, the methyl-substituted phosphanilate catalyst 139 (4-methyl-2-phosphonoaniline, 4MPA) gave an 83-fold rate enhancement over the uncatalyzed reaction at only 1 mM catalyst concentration (pH 7.4).142 One notable exception to the effective catalysis was a bulky aryl ketone substrate, for which aniline was the superior catalyst. It was speculated that the added bulk of the bifunctional catalysts hinders the transition state for the reaction. Hydrazone and oxime formation with aryl ketones (an indeed, ketones in general) remains a challenge for improved catalyst design.

In an effort to identify new structural motifs for bifunctional catalysts of hydrazone and oxime formation, a recent study by Larsen et al. measured 28 different bifunctional catalyst candidates having potential proton donors near a nucleophilic amino group. The study resulted in the identification of two effective new catalyst scaffolds (Figure 18): 2-aminophenols and 2-(aminomethyl)-benzimidazoles.162 Tuning the pKa in these systems resulted in improved catalysts with considerably greater activity than aniline for a range of substrates. To evaluate the scope of catalysis, the study compared rates of hydrazone formation for 10 varied carbonyl substrates. Notably, at 1 mM, aniline (123) showed only modest activity for aldehyde substrates and no activity at all with four different ketone substrates. Some of the new bifunctional amino-methylbenzimidazole catalysts showed measurable and significant activity (up to 10-fold rate enhancement) at 1 mM, even with ketone substrates. With aldehyde substrates, the phosphanilate 4MPA (139) described above was superior to all catalysts. This latter compound appears to be the most effective existing catalyst to date for common aldehyde substrates in hydrazone and oxime formation.

Figure 18.

Figure 18

New bifunctional 2-aminophenol 140142 and 2-(aminomethyl)benzimidazole 143145 catalyst scaffolds discovered by Larsen et al.162 The relative reaction rates refer to the hydrazone formation between phenylhydrazine (168) and 4-chlorobenzaldehyde (uncatalyzed reaction krel = 1).

Bifunctional catalysts were used recently in a related transimination reaction, namely, reversal of formaldehyde adducts on RNA preserved with formalin. 4MPA (139) was used successfully to enhance the removal of hemiaminal adducts and aminal cross-links from RNA in vitro and enabled the recovery of considerably greater quantities of RNA amplification signals from preserved tissue.165

7. Discovery and Design of Fast-Reacting Substrates

Since hydrazone formation (and particularly oxime formation) has historically been regarded as slower than ideal for some applications, it is of interest to study how structural variations in the reacting substrates affect their reaction rates. Knowledge gained from such studies can result in the design of improved substrates that react at much greater rates.

7.1. Fast-Reacting Aldehydes and Ketones

In early work, Wolfenden and Jencks noted that aryl aldehydes react somewhat more rapidly in semicarbazone formation when ortho groups (e.g., methoxy, chloro) were present.166 At the time, a mechanistic reason for this was not proffered. Nevertheless, Wang and Li recently took advantage of this reactivity in documenting acylhydrazone formation of peptides with o-halo-substituted benzaldehydes (Figure 19).167 The o-chloro aldehydes 146 gave the highest reaction yields among the halogens, and a similarly derivatized version of a fluorescent label was shown to react favorably. Rates were not reported in the study. In another application of the ortho-effect, Wang and Canary studied pyridoxal phosphoramide aldehyde substrates 147 (Figure 19) and showed that they react particularly rapidly,42 likely due both to the electron deficiency of the pyridine (see below) and to the o-hydroxyl group. A rate constant of 10 M−1 s−1 was documented at pH 7.4, and applications to protein conjugation were reported.

Figure 19.

Figure 19

Examples of aromatic aldehydes 146148, which exhibit greater reactivity to α-effect nucleophiles than benzaldehyde (151) due to the presence of an ortho substituent.

A 2016 report by Bane and co-workers used an o-phosphate group to strongly enhance reaction rates for the conjugation of aryl aldehyde 148 with a hydrazine-substituted coumarin dye.40 A second-order reaction rate constant of up to 17 M−1 s−1 was documented at pH 7, giving a reaction rate approximately 20-fold faster than with an o-hydroxy group. The authors cited intramolecular general acid catalysis by the phosphate group, as proposed earlier,142 as the mechanism of rate enhancement.

In a general effort to explore the effects of structure on reactivity, Kool et al. carried out kinetics studies of 50 different structurally varied aldehydes and ketones reacting with arylhydrazines.111 Surveying the structural effects on rates, one general conclusion was that electron-deficient carbonyl groups react somewhat more rapidly on average; for example, 4-nitrobenzaldehyde (150) reacted 4.5-fold more rapidly than the 4-methoxy derivative 149, and trifluoroacetone (156) was an especially fast-reacting ketone (Figure 20). Steric effects were moderate to small for hydrazone formation, and while some bulky aldehydes (pivaldehyde, 157) showed some apparent steric inhibition, highly bulky ketones (di-tert-butyl ketone, 155) reacted as fast as nonbulky ones (e.g., 2-butanone, 154). A major effect was found in comparing alkyl systems to aryl examples; alkyl substrates were found to react much more rapidly. For example, butyraldehyde (158) reacted 65-fold faster than benzaldehyde (151). Another clear and strong effect was the greater reactivity of aldehydes over ketones: for example, butyraldehyde (158) reacted 44-fold faster than 2-butanone (154).

Figure 20.

Figure 20

Relative hydrazone formation rates for a selected set of aldehydes and ketones 149158 at pH 7.111 The relative reaction rates refer to the hydrazone formation between the corresponding carbonyl compound and phenylhydrazine (168) [krel(aldehyde 149) = 1].

Finally, the authors noted that aldehydes with nearby basic groups (2-formylpyridine 152, 8-formylquinoline 153) reacted up to 8-fold faster than control compounds lacking the basic groups. This was hypothesized to be due to internal acid–base catalysis at the reaction transition state. A similar effect was seen in hydrazines as well (see section 7.2).

Continuing the exploration of the importance of ortho-substitution in enhanced reaction rates, Gillingham and co-workers in 2014 documented that phthalaldehyde (159) is an exceptionally fast reactant in oxime formation.83 Mechanistic studies showed that the mechanism is changed from standard oxime formation, as the first-formed tetrahedral intermediate is trapped intramolecularly to form an isoindoline(bis)-hemi-aminal heterocycle 160 (Scheme 19). This heterocyclic intermediate then slowly rearranges to oxime 161 (Figure 21). Because the rapid first step joins the reactants it can be considered the “ligation” step, and the rate constant for this step is high, on the order of 500 M−1 s−1, which is faster than the large majority of bioorthogonal “click”-type reactions used in bioconjugations.

Scheme 19. Conjugation of Aromatic Aldehydes 159 and 105, Which Display Exceptionally Fast Reaction Rates Due to the Presence of an Electrophilic Group That Stabilizes the Tetrahedral Intermediate.

Scheme 19

Figure 21.

Figure 21

Time course NMR analysis for the conjugation of phthalaldehyde (159) and O-benzylhydroxylamine. The isoindoline intermediate 160 forms rapidly, whereas the oxime 161 is generated rather slowly. Reprinted in part with permission from ref 83. Copyright 2014 Wiley-VCH.

Gillingham and co-workers subsequently modified this intramolecular trapping mechanism by replacing the o-aldehyde trap with a boronic acid group.126 In this design, reactants are again rapidly trapped at the tetrahedral intermediate stage 162/ 163, with the oxyanion coordinated to boron, and then more slowly rearrange to oximes 164 (Scheme 19). This strategy is an important new reaction for bioconjugation development. Exceptional rate constants of 104 M−1 s−1 have been measured, rendering this among the fastest of “click”-type conjugation reactions. Subsequent to Gillingham's report, studies from the Gao and Bane laboratories also described boron-substituted aldehyde and ketone variants and applications.126,168172

7.2. Fast-Reacting Alkoxyamines and Hydrazines

It has long been known that, among α-nucleophiles, rates of reaction vary strongly with the electron deficiency of the non-nucleophilic atom. Thus, alkoxyamines react to form imines (i.e., oximes) more slowly than acylhydrazines (forming acylhydrazones), which in turn react more slowly than hydrazines. For example, relative rates of similar alkoxyamines and hydrazines have been measured to differ by a factor of ca. 20.141 The reason for this slow reactivity of aminooxy compounds is that the rate of the rate-limiting step, breakdown of the tetrahedral intermediate, relies on the position of pre-equilibrium leading to it. Lower nucleophilicity pushes the equilibrium further toward the side of starting materials, lowering the concentration of tetrahedral intermediate and thus the rate of imine formation.

Beyond this general observation, little work had been done prior to 2014 on structural influences on α-effect nucleophile activity with carbonyl substrates. To address this, Kool et al. measured kinetics of hydrazone formation for 20 varied hydrazines in the reaction with 2-formylpyridine (152) at pH 7.4.141 Notably, rates varied over a factor of 23-fold across the range of structures tested (see examples in Figure 22). Relatively slow were electron-deficient hydrazines such as pentafluorophenylhydrazine (166) and acylhydrazines 167. 2-Hydrazinopyridine (169), a very common α-effect nucleophile generally thought to be efficient, was only 1.2-fold faster than phenylhydrazine (168). However, two hydrazines with acid/ base groups near the nucleophilic site were found to be much more reactive: namely, o-carboxyphenylhydrazine (170, OCPH) and dimethylaminoethylhydrazine (171, DMAEH). These fast alpha nucleophiles (FAN) showed 3- and 6-fold faster rates (respectively) than phenylhydrazine (168) at pH 7.141 Combining the DMAE-substituted hydrazine with a fast-reacting aldehyde produced rapid reactions (rate constant of 10 M−1 s−1) that exceed most cycloaddition “click”-type reactions in rate (Figure 23). The same DMAE substitution on an alkoxyamine also produced enhanced reaction rates as compared with a control alkoxyamine.141

Figure 22.

Figure 22

Effect of hydrazine structure on rate of hydrazone formation.141 The relative reaction rates refer to the hydrazone formation between the corresponding hydrazine 166171 and 2-formylpyridine (152) [krel(hydrazine 166) = 1].

Figure 23.

Figure 23

Time course plots for the hydrazone formation of 2-formylpyridine (152) and DMAEH (171, green trace) or ethylhydrazine (blue trace). Reprinted in part with permission from ref 141. Copyright 2014 American Chemical Society.

A special class of reactive hydrazines is found in indolylmethylhydrazines that undergo Pictet–Spengler ligation (see section 5.1).124 Rates for ligation are fastest at lower pH values, but are as high as 4.2 M−1 s−1 at pH 6.0, as reported by the groups of Bertozzi and Rabuka.124,125 However, the rates for the reaction drop as neutral pH is approached.

8. Conclusions

The mechanistic studies carried out by Jencks in the 1960s established the basic mechanism of imine formation by α-effect nucleophiles. That early work, carried out chiefly at low pH in organic and partially organic solvents, was important for laying the groundwork for modern research in the field. However, modern research in the field has been focused largely on the reaction performed in water, and at pH values closer to neutral, with biological relevance in mind. Indeed, while Jencks worked only with small molecules, it is now common to work with peptides, proteins, and nucleic acids as reacting substrates and to carry out reactions not only in solution but also in living cells. It seems highly likely that this trend to biological and biomedical relevance will continue.

Before the last 5 years, the rates of hydrazone and oxime formation were widely regarded as slow as compared with modern biofunctionalization chemistries, including strain-enhanced cycloaddition reactions. However, due to recent studies, new mechanistic insights into substrate reactivity and catalysis have led to the identification of reactant features that considerably enhance these rates. This work has led to the identification of fast-reacting aldehydes with o-proton donors and highly reactive hydrazines with proton donors that aid in the breakdown of the tetrahedral intermediate. The use of such reactive, self-catalyzing substrates has resulted in reaction rates (rate constants on the order of 10 M−1 s−1 at neutral pH) that exceed those of many modern cycloaddition reactions based on strained alkynes. Thus, slow reaction rates should no longer be considered a limitation of hydrazone and oxime bond formation.

Another important advance in the past 5 years has been the development of specialized aldehydes with trapping groups nearby, resulting in stable conjugation and, in some cases, very rapid rates. Important examples include the dialdehydes, the o-boron-substituted aldehydes, and the indole methyl “Pictet– Spengler” hydrazines. Rate constants in some of the orthotrapping examples are 500 M−1 s−1, rivaling the fastest bioconjugation reactions known.

Such fast-reacting aldehydes and α-effect nucleophiles can show great utility in bioconjugation reactions and are rapid enough to proceed conveniently without the use of a catalyst. However, catalysis can also achieve high reaction rates and offers other advantages as well. Unlike fast-reacting substrates, catalysts offer temporal control: a reaction can be sped up at a desired time, using a catalyst as a switch. Moreover, catalysts speed both forward and reverse reactions, thus allowing them to be useful in reversal of conjugation (bond-breaking reactions) and in speeding equilibration in dynamic combinatorial chemistry applications. Moreover, catalysts can be highly effective for substrates where specialized fast-reacting carbonyl and α-nucleophile structures are unavailable or impractical.

Following up on Jencks' early observation of nucleophilic catalysis by aniline, work in the past decade has produced dozens of new catalysts for hydrazone and oxime formation. Although all of the most effective catalysts to date make use of a nucleophilic amine group, added acid/base functional groups and varied chemical scaffolds have improved catalyst efficiency and versatility over earlier anilines. Some of the best new catalysts are more efficient than aniline by 1–2 orders of magnitude and are at the same time less toxic and more soluble. Important examples of the most highly effective newer catalysts include the anthranilates (such as 5MA, 137) and the phosphanilates (including 4MPA, 139). More work is needed to develop efficient catalysts for ketone substrates, but the new aminomethylbenzimidazole catalyst scaffolds 143145 show promise for their advantages in promoting reactions with ketone substrates.

Dominik K Kolmel received his Ph.D. from the Karlsruhe Institute of Technology (KIT) in 2013 for his work on the synthesis of fluorescently labeled cell-penetrating peptoids under the supervision of Prof. Stefan Bräse. He then joined the group of Prof. Eric T. Kool at Stanford University in 2014 for his postdoctoral studies, where he focused on the synthesis of DNA-based polyfluorophores and new substrates for oxime/hydrazine bioconjugation. In 2016, he moved to a postdoctoral position with Prof. David W. C. MacMillan at Princeton University. In the MacMillan group, he is engaged in the development of novel chemical reactions using a combination of photoredox and transition-metal catalysis for the formation of C–C bonds.

Eric T. Kool received his Ph.D. training in chemistry at Columbia University and followed with postdoctoral work at the California Institute of Technology. He began his independent career at the University of Rochester, and in 1999, he moved to Stanford University, where he is the George and Hilda Daubert Professor of Chemistry. His interests lie in the development and applications of molecular tools for the study of biological systems, with special focus on nucleic acids.

Supplementary Material

supporting

Acknowledgments

This work was supported by the U.S. National Institutes of Health (GM110050, GM068122, and GM067201). We also acknowledge the German National Academy of Sciences Leopoldina for a fellowship to D.K.K. (grant LPDS 2013-15).

Abbreviations

2,4-DMA

2,4-dimethoxyaniline

3,5-DABA

3,5-diaminobenzoic acid

5-fC

5-formylcytosine

5-fU

5-formyluracil

5MA

5-methoxyanthranilic acid

4MPA

4-methyl-2-phosphonoaniline

Ac

acetyl

Alk

alkyl residue

Ar

aromatic residue

BIQ

4,3-borazaroisoqunioline

Boc

tert-butyloxycarbonyl

Bu

butyl

Bz

benzoyl

c

concentration

CHO

Chinese hamster ovary

DAPI

4′,6-diamidino-2-phenylindole

DFT

density functional theory

DMAEH

dimethylaminoethylhydrazine

Dmt

4,4′-dimethoxytrityl

DNA

deoxyribonucleic acid

FAN

fast alpha nucleophiles

FGE

formylglycine-generating enzyme

HA

generic acid

k2(app)

apparent second-order rate constant

Keq

equilibrium constant

krel

relative rate constant

Me

methyl

Mmt

4-methoxytrityl

NMR

nuclear magnetic resonance

NPhth

phthalimidyl

OCPH

o-carboxyphenylhydrazine

PFTase

protein farnesyltransferase

PBS

phosphate-buffered saline

Ph

phenyl

RNA

ribonucleic acid

rt

room temperature

TBTA

tris[(1-benzyl-1H-1,2,3-triazol-4-yl)methyl]amine

t-Bu

tert-butyl

Trt

trityl

χp

electronegativity (Pauling scale)

Footnotes

Supporting Information: The Supporting Information is available free of charge on the ACS Publications website at DOI: 10.1021/acs.chem-rev.7b00090. Selected protocols for the oxime/hydrazone conjugation with biomolecules (PDF)

Notes: The authors declare no competing financial interest.

References

  • 1.Sletten EM, Bertozzi CR. Bioorthogonal Chemistry: Fishing for Selectivity in a Sea of Functionality. Angew Chem, Int Ed. 2009;48:6974–6998. doi: 10.1002/anie.200900942. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 2.Takaoka Y, Ojida A, Hamachi I. Protein Organic Chemistry and Applications for Labeling and Engineering in Live-Cell Systems. Angew Chem, Int Ed. 2013;52:4088–4106. doi: 10.1002/anie.201207089. [DOI] [PubMed] [Google Scholar]
  • 3.Kalia J, Raines RT. Advances in Bioconjugation. Curr Org Chem. 2010;14:138–147. doi: 10.2174/138527210790069839. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 4.Tang W, Becker ML. “Click” reactions: a versatile toolbox for the synthesis of peptide-conjugates. Chem Soc Rev. 2014;43:7013–7039. doi: 10.1039/c4cs00139g. [DOI] [PubMed] [Google Scholar]
  • 5.Meyer V, Janny A. Ueber stickstoffhaltige Acetonderivate. Ber Dtsch Chem Ges. 1882;15:1164–1167. [Google Scholar]
  • 6.Meyer V, Janny A. Ueber die Einwirkung von Hydroxylamin auf Aceton. Ber Dtsch Chem Ges. 1882;15:1324–1326. [Google Scholar]
  • 7.Meyer V, Janny A. Ueber eine neue Bildungsweise der α-Nitrosopropionsaure und die Wirkungsweise des Hydroxylamins. Ber Dtsch Chem Ges. 1882;15:1525–1529. [Google Scholar]
  • 8.Janny A. Ueber die Acetoxime. Ber Dtsch Chem Ges. 1882;15:2778–2783. [Google Scholar]
  • 9.Fischer E. Ueber die Hydrazone. Ber Dtsch Chem Ges. 1888;21:984–988. [Google Scholar]
  • 10.Li XG, Haaparanta M, Solin O. Oxime formation for fluorine-18 labeling of peptides and proteins for positron emissiontomography (PET) imaging: A review. J Fluorine Chem. 2012;143:49–56. [Google Scholar]
  • 11.Su X, Aprahamian I. Hydrazone-based switches, metallo-assemblies and sensors. Chem Soc Rev. 2014;43:1963–1981. doi: 10.1039/c3cs60385g. [DOI] [PubMed] [Google Scholar]
  • 12.Tatum LA, Su X, Aprahamian I. Simple Hydrazone BuildingBlocks for Complicated Functional Materials. Acc Chem Res. 2014;47:2141–2149. doi: 10.1021/ar500111f. [DOI] [PubMed] [Google Scholar]
  • 13.Worek F, Thiermann H, Wille T. Oximes in organo-phosphate poisoning: 60 years of hope and despair. Chem-Biol Interact. 2016;259:93–98. doi: 10.1016/j.cbi.2016.04.032. [DOI] [PubMed] [Google Scholar]
  • 14.Voicu VA, Bajgar J, Medvedovici A, Radulescu FS, Miron DS. Pharmacokincetics and pharmacodynamics of some oximes and associated therapeutic consequences: a critical review. J Appl Toxicol. 2010;30:719–729. doi: 10.1002/jat.1561. [DOI] [PubMed] [Google Scholar]
  • 15.Algar WR, Prasuhn DE, Stewart MH, Jennings TL, Blanco-Canosa JB, Dawson PE, Medintz IL. The Controlled Display of Biomolecules on Nanoparticles: A Challenge Suited to Bioorthogonal Chemistry. Bioconjugate Chem. 2011;22:825–858. doi: 10.1021/bc200065z. [DOI] [PubMed] [Google Scholar]
  • 16.della Sala F, Kay ER. Reversible Control of Nanoparticle Functionalization and Physicochemical Properties by Dynamic Covalent Exchange. Angew Chem, Int Ed. 2015;54:4187–4191. doi: 10.1002/anie.201409602. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 17.Alonso DA, Nájera C. Oxime-derived palladacycles as source of palladium nanoparticles. Chem Soc Rev. 2010;39:2891–2902. doi: 10.1039/b821314n. [DOI] [PubMed] [Google Scholar]
  • 18.Nájera C. Oxime-Derived Palladacycles: Applications in Catalysis. ChemCatChem. 2016;8:1865–1881. [Google Scholar]
  • 19.Walton JC. The Oxime Portmanteau Motif: Released Heteroradicals Undergo Incisive EPR Interrogation and Deliver Diverse Heterocycles. Acc Chem Res. 2014;47:1406–1416. doi: 10.1021/ar500017f. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 20.Dang TT, Dang TT, Langer P. Synthesis of Functionalized Heterocycles by Cyclization Reactions of Oxime and Hydrazone 1,4-Dianions. Synlett. 2011;2011:2633–2642. [Google Scholar]
  • 21.Abele E, Abele R. Recent Advances in the Synthesis of Heterocycles from Oximes. Curr Org Synth. 2014;11:403–428. [Google Scholar]
  • 22.Walton JC. Functionalised Oximes: Emergent Precursors for Carbon-, Nitrogen- and Oxygen-Centered Radicals. Molecules. 2016;21:63. doi: 10.3390/molecules21010063. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 23.Huang H, Cai J, Deng GJ. O-Acyl oximes: versatile building blocks for N-heterocycle formation in recent transition metal catalysis. Org Biomol Chem. 2016;14:1519–1530. doi: 10.1039/c5ob02417j. [DOI] [PubMed] [Google Scholar]
  • 24.Lattová E, Perreault H. The usefulness of hydrazine derivatives for mass spectrometric analysis of carbohydrates. Mass Spectrom Rev. 2013;32:366–385. doi: 10.1002/mas.21367. [DOI] [PubMed] [Google Scholar]
  • 25.Narang R, Narasimhan B, Sharma S. A Review on Biological Activities and Chemical Synthesis of Hydrazide Derivatives. Curr Med Chem. 2012;19:569–612. doi: 10.2174/092986712798918789. [DOI] [PubMed] [Google Scholar]
  • 26.Lazny R, Nodzewska A. N,N-Dialkylhydrazones in Organic Synthesis. From Simple N,N-Dimethylhydrazones to Supported Chiral Auxiliaries. Chem Rev. 2010;110:1386–1434. doi: 10.1021/cr900067y. [DOI] [PubMed] [Google Scholar]
  • 27.Barluenga J, Valdés C. Tosylhydrazones: New Uses for Classical Reagents in Palladium-Catalyzed Cross-Coupling and Metal-Free Reactions. Angew Chem, Int Ed. 2011;50:7486–7500. doi: 10.1002/anie.201007961. [DOI] [PubMed] [Google Scholar]
  • 28.Sukhorukov AY, Ioffe SL. Chemistry of Six-Membered Cyclic Oxime Ethers. Application in the Synthesis of Bioactive Compounds. Chem Rev. 2011;111:5004–5041. doi: 10.1021/cr100292w. [DOI] [PubMed] [Google Scholar]
  • 29.Kobayashi S, Mori Y, Fossey JS, Salter MM. Catalytic Enantioselective Formation of C-C Bonds by Addition to Imines and Hydrazones: A Ten-Year Update. Chem Rev. 2011;111:2626–2704. doi: 10.1021/cr100204f. [DOI] [PubMed] [Google Scholar]
  • 30.Cho H. Rearrangement of oximes and hydroxylamines with aluminum reductants. Tetrahedron. 2014;70:3527–3544. [Google Scholar]
  • 31.Ulrich S, Boturyn D, Marra A, Renaudet O, Dumy P. Oxime Ligation: A Chemoselective Click-Type Reaction for AccessingMultifunctional Biomolecular Constructs. Chem - Eur J. 2014;20:34–41. doi: 10.1002/chem.201302426. [DOI] [PubMed] [Google Scholar]
  • 32.Agten SM, Dawson PE, Hackeng TM. Oxime conjugation in protein chemistry: from carbonyl incorporation to nucleophilic catalysis. J Pept Sci. 2016;22:271–279. doi: 10.1002/psc.2874. [DOI] [PubMed] [Google Scholar]
  • 33.Sletten EM, Bertozzi CR. From Mechanism to Mouse: A Tale of Two Bioorthogonal Reactions. Acc Chem Res. 2011;44:666–676. doi: 10.1021/ar200148z. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 34.Ramil CP, Lin Q. Bioorthogonal chemistry: strategies and recent developments. Chem Commun. 2013;49:11007–11022. doi: 10.1039/c3cc44272a. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 35.Kalia J, Raines RT. Hydrolytic Stability of Hydrazones and Oximes. Angew Chem, Int Ed. 2008;47:7523–7526. doi: 10.1002/anie.200802651. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 36.Tang L, Yin Q, Xu Y, Zhou Q, Cai K, Yen J, Dobrucki LW, Cheng J. Bioorthogonal oxime ligation mediated in vivo cancer targeting. Chem Sci. 2015;6:2182–2186. doi: 10.1039/c5sc00063g. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 37.Stanley M, Virdee S. Genetically Directed Production of Recombinant, Isosteric and Nonhydrolysable Ubiquitin Conjugates. ChemBioChem. 2016;17:1472–1480. doi: 10.1002/cbic.201600138. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 38.Pfeiffer H, Sowik T, Schatzschneider U. Bioorthogonal oxime ligation of a Mo(CO)4(N-N) CO-releasing molecule (CORM) to a TGF β-binding peptide. J Organomet Chem. 2013;734:17–24. [Google Scholar]
  • 39.Pandey AK, Naduthambi D, Thomas KM, Zondlo NJ. Proline Editing: A General and Practical Approach to the Synthesis of Functionally and Structurally Diverse Peptides. Analysis of Steric versus Stereoelectronic Effects of 4-Substituted Prolines on Conformation within Peptides. J Am Chem Soc. 2013;135:4333–4363. doi: 10.1021/ja3109664. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 40.Dilek O, Sorrentino AM, Bane S. Intramolecular Catalysis of Hydrazone Formation of Aryl-Aldehydes via ortho-Phosphate Proton Exchange. Synlett. 2016;27:1335–1338. doi: 10.1055/s-0035-1561387. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 41.Lee JH, Domaille DW, Noh H, Oh T, Choi C, Jin S, Cha JN. High-Yielding and Photolabile Approaches to the Covalent Attachment of Biomolecules to Surfaces via Hydrazone Chemistry. Langmuir. 2014;30:8452–8460. doi: 10.1021/la500744s. [DOI] [PubMed] [Google Scholar]
  • 42.Wang X, Canary JW. Rapid Catalyst-Free Hydrazone Ligation: Protein-Pyridoxal Phosphoramides. Bioconjugate Chem. 2012;23:2329–2334. doi: 10.1021/bc300430k. [DOI] [PubMed] [Google Scholar]
  • 43.Le Goff G, Ouazzani J. Natural hydrazine-containing compounds: Biosynthesis, isolation, biological activities and synthesis. Bioorg Med Chem. 2014;22:6529–6544. doi: 10.1016/j.bmc.2014.10.011. [DOI] [PubMed] [Google Scholar]
  • 44.Clo E, Blixt O, Jensen KJ. Chemoselective Reagents for Covalent Capture and Display of Glycans in Microarrays. Eur J Org Chem. 2010;2010:540–554. [Google Scholar]
  • 45.Gudmundsdottir AV, Paul CE, Nitz M. Stability studies of hydrazide and hydroxylamine-based glycoconjugates in aqueous solution. Carbohydr Res. 2009;344:278–284. doi: 10.1016/j.carres.2008.11.007. [DOI] [PubMed] [Google Scholar]
  • 46.Prudden AR, Chinoy ZS, Wolfert MA, Boons GJ. A multifunctional anomeric linker for the chemoenzymatic synthesis of complex oligosaccharides. Chem Commun. 2014;50:7132–7135. doi: 10.1039/c4cc02222j. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 47.Laffel L. Ketone Bodies: a Review of Physiology, Pathophysi-ology and Application of Monitoring to Diabetes. Diabetes/Metab Res Rev. 1999;15:412–426. doi: 10.1002/(sici)1520-7560(199911/12)15:6<412::aid-dmrr72>3.0.co;2-8. [DOI] [PubMed] [Google Scholar]
  • 48.Yuen LH, Saxena NS, Park HS, Weinberg K, Kool ET. Dark Hydrazone Fluorescence Labeling Agents Enable Imaging of Cellular Aldehydic Load. ACS Chem Biol. 2016;11:2312–2319. doi: 10.1021/acschembio.6b00269. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 49.Esterbauer H, Schaur RJ, Zollner H. Chemistry andbiochemistry of 4-hydroxynonenal, malonaldehyde and relatedaldehydes. Free Radical Biol Med. 1991;11:81–128. doi: 10.1016/0891-5849(91)90192-6. [DOI] [PubMed] [Google Scholar]
  • 50.Kikuchi S, Shinpo K, Moriwaka F, Makita Z, Miyata T, Tashiro K. Neurotoxicity of methylglyoxal and 3-deoxyglucosone on cultured cortical neurons: Synergism between glycation and oxidative stress, possibly involved in neurodegenerative diseases. J Neurosci Res. 1999;57:280–289. doi: 10.1002/(SICI)1097-4547(19990715)57:2<280::AID-JNR14>3.0.CO;2-U. [DOI] [PubMed] [Google Scholar]
  • 51.Anderson MM, Hazen SL, Hsu FF, Heinecke JW. Human neutrophils employ the myleoperoxidase-hydrogen peroxide-chloride system to convert hydroxy-amino acids into glycoaldehyde, 2-hydroxypropanal, and acrolein. A mechanism for the generation of highly reactive alpha-hydroxy and alpha,beta-unsaturated aldehydes by phagocytes at sites of inflammation. J Clin Invest. 1997;99:424–432. doi: 10.1172/JCI119176. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 52.Dalleau S, Baradat M, Gueraud F, Huc L. Cell death and diseases related to oxidative stress: 4-hydroxynonenal (HNE) in the balance. Cell Death Differ. 2013;20:1615–1630. doi: 10.1038/cdd.2013.138. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 53.Fina NJ, Edwards JO. The alpha effect. A review. Int J Chem Kinet. 1973;5:1–26. [Google Scholar]
  • 54.Buncel E, Um IH. The α-effect and its modulation by solvent. Tetrahedron. 2004;60:7801–7825. [Google Scholar]
  • 55.Ren Y, Yamataka H. G2(+) Investigation on the α-Effect in the SN2 Reactions at Saturated Carbon. Chem - Eur J. 2007;13:677–682. doi: 10.1002/chem.200600203. [DOI] [PubMed] [Google Scholar]
  • 56.Lo Pachin RM, Gavin T. Molecular Mechanisms of Aldehyde Toxicity: A Chemical Perspective. Chem Res Toxicol. 2014;27:1081–1091. doi: 10.1021/tx5001046. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 57.Sudalai A, Khenkin A, Neumann R. Sodium periodatemediated oxidative transformations in organic synthesis. Org Biomol Chem. 2015;13:4374–4394. doi: 10.1039/c5ob00238a. [DOI] [PubMed] [Google Scholar]
  • 58.Zeng Y, Ramya TNC, Dirksen A, Dawson PE, Paulson JC. High-efficiency labeling of sialylated glycoproteins on living cells. Nat Methods. 2009;6:207–209. doi: 10.1038/nmeth.1305. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 59.Hunt JA. Terminal-Sequence Studies of High-Molecular-Weight Ribonucleic Acid. Biochem J. 1965;95:541–551. doi: 10.1042/bj0950541. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 60.Hansske F, Sprinzl M, Cramer F. Reaction of the Ribose Moiety of Adenosine and AMP with Periodate and Carboxylic Acid Hydrazides. Bioorg Chem. 1974;3:367–376. [Google Scholar]
  • 61.Hardisty RE, Kawasaki F, Sahakyan AB, Balasubramanian S. Selective Chemical Labeling of Natural T Modifications in DNA. J Am Chem Soc. 2015;137:9270–9272. doi: 10.1021/jacs.5b03730. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 62.Chelius D, Shaler TA. Capture of Peptides with N-Terminal Serine and Threonine: A Sequence-Specific Chemical Method for Peptide Mixture Simplification. Bioconjugate Chem. 2003;14:205–211. doi: 10.1021/bc025605u. [DOI] [PubMed] [Google Scholar]
  • 63.Amore A, Wals K, Koekoek E, Hoppes R, Toebes M, Schumacher TNM, Rodenko B, Ovaa H. Development of a Hypersensitive Periodate-Cleavable Amino Acid that is Methionine-and Disulfide-Compatible and its Application in MHC Exchange Reagents for T Cell Characterisation. ChemBioChem. 2013;14:123–131. doi: 10.1002/cbic.201200540. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 64.Haney CM, Loch MT, Horne WS. Promoting peptide α-helix formation with dynamic covalent oxime side-chain cross-links. Chem Commun. 2011;47:10915–10917. doi: 10.1039/c1cc12010g. [DOI] [PubMed] [Google Scholar]
  • 65.Haney CM, Horne WS. Oxime Side-Chain Cross-Links in an α-Helical Coiled-Coil Protein: Structure, Thermodynamics, and Folding-Templated Synthesis of Bicyclic Species. Chem - Eur J. 2013;19:11342–11351. doi: 10.1002/chem.201300506. [DOI] [PubMed] [Google Scholar]
  • 66.Forget D, Renaudet O, Boturyn D, Defrancq E, Dumy P. 3′-Oligonucleotides conjugation via chemoselective oxime bond formation. Tetrahedron Lett. 2001;42:9171–9174. [Google Scholar]
  • 67.Choudhary A, Kamer KJ, Shoulders MD, Raines RT. 4-Ketoproline: An Electrophile Proline Analog for Bioconjugation. Biopolymers. 2015;104:110–115. doi: 10.1002/bip.22620. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 68.Spicer CD, Davis BG. Selective chemical proteinmodification. Nat Commun. 2014;5:4740. doi: 10.1038/ncomms5740. [DOI] [PubMed] [Google Scholar]
  • 69.Cornish VW, Hahn KM, Schultz PG. Site-Specific ProteinModification Using a Ketone Handle. J Am Chem Soc. 1996;118:8150–8151. [Google Scholar]
  • 70.Wang L, Zhang Z, Brock A, Schultz PG. Addition of theketo functional group to the genetic code of Escherichia coli. Proc Natl Acad Sci U S A. 2003;100:56–61. doi: 10.1073/pnas.0234824100. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 71.Zhang Z, Smith BAC, Wang L, Brock A, Cho C, Schultz PG. A New Strategy for the Site-Specific Modification of Proteins in Vivo. Biochemistry. 2003;42:6735–6746. doi: 10.1021/bi0300231. [DOI] [PubMed] [Google Scholar]
  • 72.Zeng H, Xie J, Schultz PG. Genetic introduction of a diketone-containing amino acid into proteins. Bioorg Med Chem Lett. 2006;16:5356–5359. doi: 10.1016/j.bmcl.2006.07.094. [DOI] [PubMed] [Google Scholar]
  • 73.Huang Y, Wan W, Russell WK, Pai PJ, Wang Z, Russell DH, Liu W. Genetic incorporation of an aliphatic keto-containing amino acid into proteins for their site-specific modification. Bioorg Med Chem Lett. 2010;20:878–880. doi: 10.1016/j.bmcl.2009.12.077. [DOI] [PubMed] [Google Scholar]
  • 74.Schlick TL, Ding Z, Kovacs EW, Francis MB. Dual-Surface Modification of the Tobacco Mosaic Virus. J Am Chem Soc. 2005;127:3718–3723. doi: 10.1021/ja046239n. [DOI] [PubMed] [Google Scholar]
  • 75.Gavrilyuk J, Ban H, Nagano M, Hakamata W, Barbas CF., III Formylbenzene Diazonium Hexafluorophosphate Reagent for Tyrosine-Selective Modification of Proteins and the Introduction of a Bioorthogonal Aldehyde. Bioconjugate Chem. 2012;23:2321–2328. doi: 10.1021/bc300410p. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 76.Rashidian M, Dozier JK, Lenevich S, Distefano MD. Selective labeling of polypeptides using protein farnesyltransferase via rapid oxime ligation. Chem Commun. 2010;46:8998–9000. doi: 10.1039/c0cc03305g. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 77.Dozier JK, Khatwani SL, Wollack JW, Wang YC, Schmidt-Dannert C, Distefano MD. Engineering Protein Farnesyltransferase for Enzymatic Protein Labeling Applications. Bioconjugate Chem. 2014;25:1203–1212. doi: 10.1021/bc500240p. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 78.Carrico IS, Carlson BL, Bertozzi CR. Introducing genetically encoded aldehydes into proteins. Nat Chem Biol. 2007;3:321–322. doi: 10.1038/nchembio878. [DOI] [PubMed] [Google Scholar]
  • 79.Gilmore JM, Scheck RA, Esser-Kahn AP, Joshi NS, Francis MB. N-Terminal Protein Modification through a Biomimetic Transamination Reaction. Angew Chem, Int Ed. 2006;45:5307–5311. doi: 10.1002/anie.200600368. [DOI] [PubMed] [Google Scholar]
  • 80.Kung KKY, Wong KF, Leung KC, Wong MK. N-terminal α-amino group modification of peptides by an oxime formation-exchange reaction sequence. Chem Commun. 2013;49:6888–6890. doi: 10.1039/c3cc42261e. [DOI] [PubMed] [Google Scholar]
  • 81.Spears RJ, Fascione MA. Site-selective incorporation and ligation of protein aldehydes. Org Biomol Chem. 2016;14:7622–7638. doi: 10.1039/c6ob00778c. [DOI] [PubMed] [Google Scholar]
  • 82.Raddatz S, Mueller-Ibeler J, Kluge J, Wäβ L, Burdinski G, Havens JR, Onofrey TJ, Wang D, Schweitzer M. Hydrazide oligonucleotides: new chemical modification for chip array attachment and conjugation. Nucleic Acids Res. 2002;30:4793–4802. doi: 10.1093/nar/gkf594. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 83.Schmidt P, Zhou L, Tishinov K, Zimmermann K, Gillingham D. Dialdehydes Lead to Exceptionally Fast Bioconjuga-tions at Neutral pH by Virtue of a Cyclic Intermediate. Angew Chem. 2014;126:11108–11111. doi: 10.1002/anie.201406132. [DOI] [PubMed] [Google Scholar]
  • 84.Lönnberg T, Suzuki Y, Komiyama M. Prompt site-selective DNA hydrolysis by Ce(IV)-EDTA using oligonucleotide multi-phosphate conjugates. Org Biomol Chem. 2008;6:3580–3587. doi: 10.1039/b807789d. [DOI] [PubMed] [Google Scholar]
  • 85.Forget D, Boturyn D, Defrancq E, Lhomme J, Dumy P. Highly Efficient Synthesis of Peptide-Oligonucleotide Conjugates: Chemoselective Oxime and Thiazolidine Formation. Chem - Eur J. 2001;7:3976–3984. doi: 10.1002/1521-3765(20010917)7:18<3976::aid-chem3976>3.0.co;2-x. [DOI] [PubMed] [Google Scholar]
  • 86.Cebon B, Lambert JN, Leung D, Mackie H, McCluskey KL, Nguyen X, Tassone C. New DNA Modification Strategies Involving Oxime Formation. Aust J Chem. 2000;53:333–339. [Google Scholar]
  • 87.Karskela M, Helkearo M, Virta P, Lönnberg H. Synthesis of Oligonucleotide Glycoconjugates Using Sequential Click and Oximation Ligations. Bioconjugate Chem. 2010;21:748–755. doi: 10.1021/bc900529g. [DOI] [PubMed] [Google Scholar]
  • 88.Abendroth F, Seitz O. Double-Clicking Peptides onto Phosphorothioate Oligonucleotides: Combining Two Proapoptotic Agents in One Molecule. Angew Chem, Int Ed. 2014;53:10504–10509. doi: 10.1002/anie.201406674. [DOI] [PubMed] [Google Scholar]
  • 89.Salo H, Virta P, Hakala H, Prakash TP, Kawasaki AM, Manoharan M, Lönnberg H. Aminooxy Functionalized Oligonucleo-tides: Preparation, On-Support Derivatization, and Postsynthetic Attachment to Polymer Support. Bioconjugate Chem. 1999;10:815–823. doi: 10.1021/bc990021m. [DOI] [PubMed] [Google Scholar]
  • 90.Kawasaki AM, Casper MD, Prakash TP, Manalili S, Sasmor H, Manoharan M, Cook PD. Synthesis, Hybridization, and Nuclease Resistance Properties of 2′-O-Aminooxyethyl (2′-O-AOE) Modified Oligonucleotides. Tetrahedron Lett. 1999;40:661–664. [Google Scholar]
  • 91.Bello C, Kikul F, Becker CFW. Efficient generation of peptide hydrazides via direct hydrazinolysis of Peptidyl-Wang-TentaGel resins. J Pept Sci. 2015;21:201–207. doi: 10.1002/psc.2747. [DOI] [PubMed] [Google Scholar]
  • 92.Qvortrup K, Komnatnyy VV, Nielsen TE. A Photolabile Linker for the Solid-Phase Synthesis of Peptide Hydrazides and Heterocycles. Org Lett. 2014;16:4782–4785. doi: 10.1021/ol502219s. [DOI] [PubMed] [Google Scholar]
  • 93.Chelushkin PS, Polyanichko KV, Leko MV, Dorosh MY, Bruckdorfer T, Burov SV. Convenient method of peptide hydrazide synthesis using a new hydrazone resin. Tetrahedron Lett. 2015;56:619–622. [Google Scholar]
  • 94.Margathe JF, Iturrioz X, Regenass P, Karpenko IA, Humbert N, de Rocquigny H, Hibert M, Llorens-Cortes C, Bonnet D. Convenient Access to Fluorescent Probes by Chemo-selective Acylation of Hydrazinopeptides: Application to the Synthesisof the First Far-Red Ligand for Apelin Receptor Imaging. Chem - Eur J. 2016;22:1399–1405. doi: 10.1002/chem.201503630. [DOI] [PubMed] [Google Scholar]
  • 95.Itoh Y, Aihara K, Mellini P, Tojo T, Ota Y, Tsumoto H, Solomon VR, Zhan P, Suzuki M, Ogasawara D, et al. Identification of SNAIL1 Peptide-Based Irreversible Lysine-Specific Demethylase 1-Selective Inactivators. J Med Chem. 2016;59:1531–1544. doi: 10.1021/acs.jmedchem.5b01323. [DOI] [PubMed] [Google Scholar]
  • 96.Oh JE, Lee KH. Synthesis of Novel Unnatural Amino Acids as a Building Block and Its Incorporation into an Antimicrobial Peptide. Bioorg Med Chem. 1999;7:2985–2990. doi: 10.1016/s0968-0896(99)00247-3. [DOI] [PubMed] [Google Scholar]
  • 97.Haney CM, Horne WS. Dynamic covalent side-chain crosslinks via intermolecular oxime or hydrazone formation from bifunctional peptides and simple organic linkers. J Pept Sci. 2014;20:108–114. doi: 10.1002/psc.2596. [DOI] [PubMed] [Google Scholar]
  • 98.Peluso S, Imperiali B. Asparagine surrogates for the assembly of N-linked glycopeptide mimetics by chemoselective ligation. Tetrahedron Lett. 2001;42:2085–2087. [Google Scholar]
  • 99.Conant JB, Bartlett PD. A Quantitative Study of Semicarbazone Formation. J Am Chem Soc. 1932;54:2881–2899. [Google Scholar]
  • 100.Westheimer FH. Semicarbazone Formation in Sixty Per Cent. Methyl Cellosolve. J Am Chem Soc. 1934;56:1962–1965. [Google Scholar]
  • 101.Jencks WP. Mechanism and Catalysis of Simple Carbonyl Group Reactions. In: Cohen SG, Streitwieser A Jr, Taft RW, editors. Progress in Physical Organic Chemistry. Vol. 2. John Wiley & Sons, Inc; Hoboken, NJ: 1964. pp. 63–128. [Google Scholar]
  • 102.Stempel GH, Jr, Schaffel GS. A Comparative Study of the Kinetics and Mechanisms of Formation of the Phenzlhydrazone, Semicarbazone and Oxime of d-Carvone. J Am Chem Soc. 1944;66:1158–1161. [Google Scholar]
  • 103.Cordes EH, Jencks WP. General Acid Catalysis of Semicarbazone Formation. J Am Chem Soc. 1962;84:4319–4328. [Google Scholar]
  • 104.Calzadilla M, Malpica A, Cordova T. Effect of structure on reactivity in oxime formation of benzaldehydes. J Phys Org Chem. 1999;12:708–712. [Google Scholar]
  • 105.Yildiz I. A DFT Approach to the Mechanistic Study of Hydrazone Hydrolysis. J Phys Chem A. 2016;120:3683–3692. doi: 10.1021/acs.jpca.6b02882. [DOI] [PubMed] [Google Scholar]
  • 106.Bagno A, Menna E, Mezzina E, Scorrano G, Spinelli D. Site of Protonation of Alkyl- and Arylhydrazines Probed by 14N, 15N,and 13C NMR Relaxation and Quantum Chemical Calculations. J Phys Chem A. 1998;102:2888–2892. [Google Scholar]
  • 107.Egberink H, van Heerden C. The Mechanism of the Formation and Hydrolysis of Cyclohexanone Oxime in Aqueous Solutions. Anal Chim Acta. 1980;118:359–368. [Google Scholar]
  • 108.Jencks WP. Studies on the Mechanism of Oxime and Semicarbazone Formation. J Am Chem Soc. 1959;81:475–481. [Google Scholar]
  • 109.King TP, Zhao SW, Lam T. Preparation of Protein Conjugates via Intermolecular Hydrazone Linkage. Biochemistry. 1986;25:5774–5779. doi: 10.1021/bi00367a064. [DOI] [PubMed] [Google Scholar]
  • 110.Agten SM, Suylen DPL, Hackeng TM. Oxime Catalysis by Freezing. Bioconjugate Chem. 2016;27:42–46. doi: 10.1021/acs.bioconjchem.5b00611. [DOI] [PubMed] [Google Scholar]
  • 111.Kool ET, Park DH, Crisalli P. Fast Hydrazone Reactants: Electronic and Acid/Base Effects Strongly Influence Rate at Biological pH. J Am Chem Soc. 2013;135:17663–17666. doi: 10.1021/ja407407h. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 112.Deygen IM, Seidl C, Kölmel DK, Bednarek C, Heissler S, Kudryashova EV, Bräse S, Schepers U. Novel Prodrug of Doxorubicin Modified by Stearoylspermine Encapsulated into PEG-Chitosan-Stabilized Liposomes. Langmuir. 2016;32:10861–10869. doi: 10.1021/acs.langmuir.6b01023. [DOI] [PubMed] [Google Scholar]
  • 113.Matson JB, Stupp SI. Drug release from hydrazone-containing peptide amphiphiles. Chem Commun. 2011;47:7962–7964. doi: 10.1039/c1cc12570b. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 114.Brudno Y, Silva EA, Kearney CJ, Lewin SA, Miller A, Martinick KD, Aizenberg M, Mooney DJ. Refilling drug deliverydepots through the blood. Proc Natl Acad Sci U S A. 2014;111:12722–12727. doi: 10.1073/pnas.1413027111. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 115.Hermanson GT. Bioconjugate Techniques. 3rd. Academic Press; 2013. Hydrazine and Hydrazide Derivatives; p. 251. [Google Scholar]
  • 116.Dirksen A, Dawson PE. Rapid Oxime and Hydrazone Ligations with Aromatic Aldehydes for Biomolecular Labeling. Bioconjugate Chem. 2008;19:2543–2548. doi: 10.1021/bc800310p. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 117.Schenk GH. Organic Functional Group Analysis: Theory and Development. Pergamon Press; 1968. [Google Scholar]
  • 118.Nguyen R, Huc I. Optimizing the reversibility of hydrazone formation for dynamic combinatorial chemistry. Chem Commun. 2003:942–943. doi: 10.1039/b211645f. [DOI] [PubMed] [Google Scholar]
  • 119.Palling DJ, Hollocher TC. Equilibrium Constant for Pyruvic Oxime Formation. J Org Chem. 1984;49:388–390. [Google Scholar]
  • 120.El-Mahdi O, Melnyk O. α-Oxo Aldehyde or Glyoxylyl GroupChemistry in Peptide Bioconjugation. Bioconjugate Chem. 2013;24:735–765. doi: 10.1021/bc300516f. [DOI] [PubMed] [Google Scholar]
  • 121.Liu F, Stephen AG, Adamson CS, Gousset K, Aman MJ, Freed EO, Fisher RJ, Burke TR., Jr Hydrazone- and Hydrazide-Containing N-Substituted Glycine as Peptoid Surrogatesfor Expedited Library Synthesis: Application to the Preparation of Tsg101-Directed HIV-1 Budding Antagonists. Org Lett. 2006;8:5165–5168. doi: 10.1021/ol0622211. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 122.Vanderesse R, Thevenet L, Marraud M, Boggetto N, Reboud M, Corbier C. α-Aminoxy Acids as Building Blocks for the Oxime and Hydroxylamine Pseudopeptide Links. Application to the Synthesis of Human Elastase Inhibitors. J Pept Sci. 2003;9:282–299. doi: 10.1002/psc.452. [DOI] [PubMed] [Google Scholar]
  • 123.Ji K, Lee C, Janesko BG, Simanek EE. Triazine-Substituted and Acyl Hydrazones: Experiment and Computation Reveal a Stability Inversion at Low pH. Mol Pharmaceutics. 2015;12:2924–2927. doi: 10.1021/acs.molpharmaceut.5b00205. [DOI] [PubMed] [Google Scholar]
  • 124.Agarwal P, van der Weijden J, Sletten EM, Rabuka D, Bertozzi CR. A Pictet-Spengler ligation for protein chemical modification. Proc Natl Acad Sci U S A. 2013;110:46–51. doi: 10.1073/pnas.1213186110. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 125.Agarwal P, Kudirka R, Albers AE, Barfield RM, de Hart GW, Drake PM, Jones LC, Rabuka D. Hydrazino-Pictet-Spengler Ligation as a Biocompatible Method for the Generation of Stable Protein Conjugates. Bioconjugate Chem. 2013;24:846–851. doi: 10.1021/bc400042a. [DOI] [PubMed] [Google Scholar]
  • 126.Stress CJ, Schmidt PJ, Gillingham DG. Comparison of boron-assisted oxime and hydrazone formation leads to the discovery of a fluorogenic variant. Org Biomol Chem. 2016;14:5529–5533. doi: 10.1039/c6ob00168h. [DOI] [PubMed] [Google Scholar]
  • 127.Trindade AF, Bode JW. Irreversible Conjugation of Aldehydes in Water To Form Stable 1,2,4-Oxadiazinan-5-ones. Org Lett. 2016;18:4210–4213. doi: 10.1021/acs.orglett.6b01889. [DOI] [PubMed] [Google Scholar]
  • 128.Cousins GRL, Furlan RLE, Ng YF, Redman JE, Sanders JKM. Identification and Isolation of a Receptor for N-Methyl Alkylammonium Salts: Molecular Amplification in a Pseudo-peptide Dynamic Combinatorial Library. Angew Chem, Int Ed. 2001;40:423–428. doi: 10.1002/1521-3773(20010119)40:2<423::AID-ANIE423>3.0.CO;2-6. [DOI] [PubMed] [Google Scholar]
  • 129.Cousins GRL, Poulsen SA, Sanders JKM. Dynamic combinatorial libraries of pseudo-peptide hydrazone macrocycles. Chem Commun. 1999:1575–1576. [Google Scholar]
  • 130.Polyakov VA, Nelen MI, Nazarpack-Kandlousy N, Ryabov AD, Eliseev AV. Imine exchange in O-aryl and O-alkyloximes as a base reaction for aqueous ‘dynamic’ combinatorial libraries. A kinetic and thermodynamic study. J Phys Org Chem. 1999;12:357–363. [Google Scholar]
  • 131.Simpson MG, Pittelkow M, Watson SP, Sanders JKM. Dynamic combinatorial chemistry with hydrazones: libraries incorporating heterocyclic and steroidal motifs. Org Biomol Chem. 2010;8:1181–1187. doi: 10.1039/b917146k. [DOI] [PubMed] [Google Scholar]
  • 132.Doronina SO, Toki BE, Torgov MY, Mendelsohn BA, Cerveny CG, Chace DF, DeBlanc RL, Gearing RP, Bovee TD, Siegall CB, Senter PD, et al. Development of potent monoclonal antibody auristatin conjugates for cancer therapy. Nat Biotechnol. 2003;21:778–784. doi: 10.1038/nbt832. [DOI] [PubMed] [Google Scholar]
  • 133.Bouchard H, Viskov C, Garcia-Echeverria C. Antibody—drug conjugates—A new wave of cancer drugs. Bioorg Med Chem Lett. 2014;24:5357–5363. doi: 10.1016/j.bmcl.2014.10.021. [DOI] [PubMed] [Google Scholar]
  • 134.Kovarikova P, Mrkvickova Z, Klimes J. Investigation of the stability of aromatic hydrazones in plasma and related biological material. J Pharm Biomed Anal. 2008;47:360–370. doi: 10.1016/j.jpba.2008.01.011. [DOI] [PubMed] [Google Scholar]
  • 135.Buss JL, Ponka P. Hydrolysis of pyridoxal isonicotinoyl hydrazone and its analogs. Biochim Biophys Acta, Gen Subj. 2003;1619:177–186. doi: 10.1016/s0304-4165(02)00478-6. [DOI] [PubMed] [Google Scholar]
  • 136.Agard NJ, Prescher JA, Bertozzi CR. A Strain-Promoted [3 + 2] Azide-Alkyne Cycloaddition for Covalent Modification of Biomolecules in Living Systems. J Am Chem Soc. 2004;126:15046–15047. doi: 10.1021/ja044996f. [DOI] [PubMed] [Google Scholar]
  • 137.Baskin JM, Prescher JA, Laughlin ST, Agard NJ, Chang PV, Miller IA, Lo A, Codelli JA, Bertozzi CR. Copper-free click chemistry for dynamic in vivo imaging. Proc Natl Acad Sci U S A. 2007;104:16793–16797. doi: 10.1073/pnas.0707090104. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 138.Jewett JC, Bertozzi CR. Cu-free click cycloaddition reactions in chemical biology. Chem Soc Rev. 2010;39:1272–1279. doi: 10.1039/b901970g. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 139.Knall AC, Slugovc C. Inverse electron demand Diels-Alder (iEDDA)-initiated conjugation: a (high) potential click chemistry scheme. Chem Soc Rev. 2013;42:5131–5142. doi: 10.1039/c3cs60049a. [DOI] [PubMed] [Google Scholar]
  • 140.Saito F, Noda H, Bode JW. Critical Evaluation and Rate Constants of Chemoselective Ligation Reactions for Stoichiometric Conjugations in Water. ACS Chem Biol. 2015;10:1026–1033. doi: 10.1021/cb5006728. [DOI] [PubMed] [Google Scholar]
  • 141.Kool ET, Crisalli P, Chan KM. Fast Alpha Nucleophiles: Structures that Undergo Rapid Hydrazone/Oxime Formation at Neutral pH. Org Lett. 2014;16:1454–1457. doi: 10.1021/ol500262y. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 142.Crisalli P, Kool ET. Importance of ortho Proton Donors in Catalysis of Hydrazone Formation. Org Lett. 2013;15:1646–1649. doi: 10.1021/ol400427x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 143.Beeren SR, Pittelkow M, Sanders JKM. From static to dynamic: escaping kinetic traps in hydrazone-based dynamic combinatorial libraries. Chem Commun. 2011;47:7359–7361. doi: 10.1039/c1cc12268a. [DOI] [PubMed] [Google Scholar]
  • 144.Lascano S, Zhang KD, Wehlauch R, Gademann K, Sakai N, Matile S. The third orthogonal dynamic covalent bond. Chem Sci. 2016;7:4720–4724. doi: 10.1039/c6sc01133k. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 145.Crisalli P, Hernández AR, Kool ET. Fluorescence Quenchers for Hydrazone and Oxime Orthogonal Bioconjugation. Bioconjugate Chem. 2012;23:1969–1980. doi: 10.1021/bc300344b. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 146.Gan Ja, Tian H, Wang Q, Chen K. Red luminescence of novel naphthalimides in solid state. Proc SPIE. 2001;4464:359–365. [Google Scholar]
  • 147.Herbst E, Shabat D. FRET-based cyanine probes for monitoring ligation reactions and their applications to mechanistic studies and catalyst screening. Org Biomol Chem. 2016;14:3715–3728. doi: 10.1039/c5ob02127h. [DOI] [PubMed] [Google Scholar]
  • 148.Cordes EH, Jencks WP. Nucleophilic Catalysis of Semicarbazone Formation by Anilines. J Am Chem Soc. 1962;84:826–831. [Google Scholar]
  • 149.Morales S, Acena JL, García Ruano JL, Cid MB. Sustainable Synthesis of Oximes, Hydrazones, and Thiosemicarba-zones under Mid Organocatalyzed Reaction Conditions. J Org Chem. 2016;81:10016–10022. doi: 10.1021/acs.joc.6b01912. [DOI] [PubMed] [Google Scholar]
  • 150.Dirksen A, Hackeng TM, Dawson PE. Nucleophilic Catalysis of Oxime Ligation. Angew Chem, Int Ed. 2006;45:7581–7584. doi: 10.1002/anie.200602877. [DOI] [PubMed] [Google Scholar]
  • 151.Dirksen A, Dirksen S, Hackeng TM, Dawson PE. Nucleophilic Catalysis of Hydrazone Formation and Transimination: Implications for Dynamic Covalent Chemistry. J Am Chem Soc. 2006;128:15602–15603. doi: 10.1021/ja067189k. [DOI] [PubMed] [Google Scholar]
  • 152.Wang S, Gurav D, Oommen OP, Varghese OP. Insights into the Mechanism and Catalysis of Oxime Coupling Chemistry at Physiological pH. Chem - Eur J. 2015;21:5980–5985. doi: 10.1002/chem.201406458. [DOI] [PubMed] [Google Scholar]
  • 153.Rayo J, Amara N, Krief P, Meijler MM. Live Cell Labeling of Native Intracellular Bacterial Receptors Using Aniline-Catalyzed Oxime Ligation. J Am Chem Soc. 2011;133:7469–7475. doi: 10.1021/ja200455d. [DOI] [PubMed] [Google Scholar]
  • 154.Domaille DW, Cha JN. Aniline-terminated DNA catalyzes rapid DNA-hydrazone formation at physiological pH. Chem Commun. 2014;50:3831–3833. doi: 10.1039/c4cc00292j. [DOI] [PubMed] [Google Scholar]
  • 155.Thygesen MB, Munch H, Sauer J, Cló E, Jorgensen MR, Hindsgaul O, Jensen KJ. Nucleophilic Catalysis of Carbohydrate Oxime Formation by Anilines. J Org Chem. 2010;75:1752–1755. doi: 10.1021/jo902425v. [DOI] [PubMed] [Google Scholar]
  • 156.Hahn A, Reschke S, Leimkuhler S, Risse T. Ketoxime Coupling of p-Acetylphenylalanine at Neutral pH for Site-Directed Spin Labeling of Human Sulfite Oxidase. J Phys Chem B. 2014;118:7077–7084. doi: 10.1021/jp503471j. [DOI] [PubMed] [Google Scholar]
  • 157.Blanden AR, Mukherjee K, Dilek O, Loew M, Bane SL. 4-Aminophenylalanine as a Biocompatible Nucleophilic Catalyst for Hydrazone Ligations at Low Temperature and Neutral pH. Bioconjugate Chem. 2011;22:1954–1961. doi: 10.1021/bc2001566. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 158.Rashidian M, Mahmoodi MM, Shah R, Dozier JK, Wagner CR, Distefano MD. A Highly Efficient Catalyst for OximeLigation and Hydrazone-Oxime Exchange Suitable for Bioconjugation. Bioconjugate Chem. 2013;24:333–342. doi: 10.1021/bc3004167. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 159.Mahmoodi MM, Rashidian M, Zhang Y, Distefano MD. Application of meta- and para-Phenylenediamine as Enhanced Oxime Ligation Catalysts for Protein Labeling, PEGylation, Immobilization, and Release. Curr Protoc Protein Sci. 2015;79:15.4.1–15.4.28. doi: 10.1002/0471140864.ps1504s79. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 160.Wendeler M, Grinberg L, Wang X, Dawson PE, Baca M. Enhanced Catalysis of Oxime-Based Bioconjugations by Substituted Anilines. Bioconjugate Chem. 2014;25:93–101. doi: 10.1021/bc400380f. [DOI] [PubMed] [Google Scholar]
  • 161.Crisalli P, Kool ET. Water-Soluble Organocatalysts for Hydrazone and Oxime Formation. J Org Chem. 2013;78:1184–1189. doi: 10.1021/jo302746p. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 162.Larsen D, Pittelkow M, Karmakar S, Kool ET. New Organocatalyst Scaffolds with High Activity in Promoting Hydrazone and Oxime Formation at Neutral pH. Org Lett. 2015;17:274–277. doi: 10.1021/ol503372j. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 163.McKay CS, Finn MG. Polyvalent Catalysts Operating on Polyvalent Substrates: A Model for Surface-Controlled Reactivity. Angew Chem, Int Ed. 2016;55:12643–12649. doi: 10.1002/anie.201602797. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 164.Loskot SA, Zhang J, Langenhan JM. Nucleophilic Catalysis of MeON-Neoglycoside Formation by Aniline Derivatives. J Org Chem. 2013;78:12189–12193. doi: 10.1021/jo401688p. [DOI] [PubMed] [Google Scholar]
  • 165.Karmakar S, Harcourt EM, Hewings DS, Scherer F, Lovejoy AF, Kurtz DM, Ehrenschwender T, Barandun LJ, Roost C, Alizadeh AA, Kool ET. Organocatalytic removal of formaldehyde adducts from RNA and DNA bases. Nat Chem. 2015;7:752–758. doi: 10.1038/nchem.2307. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 166.Wolfenden R, Jencks WPJ. The Effect of o-Substituents on Benzaldehyde Semicarbazone Formation. J Am Chem Soc. 1961;83:2763–2768. [Google Scholar]
  • 167.Xu Y, Xu L, Xia Y, Guan CJ, Guo QX, Fu Y, Wang C, Li YM. Fast and catalyst-free hydrazone ligation via ortho-halo-substituted benzaldehydes for protein C-terminal labeling at neutral pH. Chem Commun. 2015;51:13189–13192. doi: 10.1039/c5cc04382d. [DOI] [PubMed] [Google Scholar]
  • 168.Schmidt P, Stress C, Gillingham D. Boronic acids facilitate rapid oxime condensations at neutral pH. Chem Sci. 2015;6:3329–3333. doi: 10.1039/c5sc00921a. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 169.Bandyopadhyay A, Gao J. Iminoboronate Formation Leads to Fast and Reversible Conjugation Chemistry of α-Nucleophiles at Neutral pH. Chem - Eur J. 2015;21:14748–14752. doi: 10.1002/chem.201502077. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 170.Dilek O, Lei Z, Mukherjee K, Bane S. Rapid formation of a stable boron–nitrogen heterocycle in dilute, neutral aqueous solution for bioorthogonal coupling reactions. Chem Commun. 2015;51:16992–16695. doi: 10.1039/c5cc07453c. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 171.Gillingham D. The role of boronic acids in accelerating condensation reactions of α-effect amines with carbonyls. Org Biomol Chem. 2016;14:7606–7609. doi: 10.1039/c6ob01193d. [DOI] [PubMed] [Google Scholar]
  • 172.Bandyopadhyay A, Cambray S, Gao J. Fast Diazaborine Formation of Semicarbazide Enables Facile Labeling of Bacterial Pathogens. J Am Chem Soc. 2017;139:871–878. doi: 10.1021/jacs.6b11115. [DOI] [PMC free article] [PubMed] [Google Scholar]

Associated Data

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

Supplementary Materials

supporting

RESOURCES