ACS Publications. Most Trusted. Most Cited. Most Read
Enlightening the Path to Protein Engineering: Chemoselective Turn-On Probes for High-Throughput Screening of Enzymatic Activity
My Activity
  • Open Access
Review

Enlightening the Path to Protein Engineering: Chemoselective Turn-On Probes for High-Throughput Screening of Enzymatic Activity
Click to copy article linkArticle link copied!

Open PDF

Chemical Reviews

Cite this: Chem. Rev. 2023, 123, 6, 2832–2901
Click to copy citationCitation copied!
https://doi.org/10.1021/acs.chemrev.2c00304
Published February 28, 2023

Copyright © 2023 The Authors. Published by American Chemical Society. This publication is licensed under

CC-BY 4.0 .

Abstract

Click to copy section linkSection link copied!

Many successful stories in enzyme engineering are based on the creation of randomized diversity in large mutant libraries, containing millions to billions of enzyme variants. Methods that enabled their evaluation with high throughput are dominated by spectroscopic techniques due to their high speed and sensitivity. A large proportion of studies relies on fluorogenic substrates that mimic the chemical properties of the target or coupled enzymatic assays with an optical read-out that assesses the desired catalytic efficiency indirectly. The most reliable hits, however, are achieved by screening for conversions of the starting material to the desired product. For this purpose, functional group assays offer a general approach to achieve a fast, optical read-out. They use the chemoselectivity, differences in electronic and steric properties of various functional groups, to reduce the number of false-positive results and the analytical noise stemming from enzymatic background activities. This review summarizes the developments and use of functional group probes for chemoselective derivatizations, with a clear focus on screening for enzymatic activity in protein engineering.

This publication is licensed under

CC-BY 4.0 .
  • cc licence
  • by licence
Copyright © 2023 The Authors. Published by American Chemical Society

1. Introduction

Click to copy section linkSection link copied!

Protein engineering involves the design and creation of proteins with specific, desired functionalities. (1) For applications in synthetic chemistry, these efforts generally involve changing the substrate specificity, increasing the enzymatic activity, inverting and/or improving stereoselectivity, and increasing the stability of the biocatalysts to solvent, pH, temperature, proteolytic agents, etc. (2) All modifications aim to complement or even completely replace reactions that use hazardous or environmentally unfavorable reagents or conditions. The demands of synthetic chemistry, however, do not necessarily align with nature’s repertoire of enzymatically catalyzed reactions. (3) Consequently, tools for the improvement of existing and the development of new enzymatic catalysts are needed.
In recent years, it has become possible to use rational structure-based and de novo design to create new active enzymes. These new methods have been demonstrated to work in improving catalytic performance, broadening substrate range, increasing or inverting existing stereoselectivity, and in creating “new-to-nature” functionality (4) to obtain products with high value for industrial applications. (5) In contrast to specific design, methods that test all (or almost all members) of a given population (“screening”) are typically used when there is no prior knowledge of a protein structure or when rational design seems to only deliver incremental improvements. (6−8)
These screening methods are often subsumed under the term “directed evolution,” pioneered in the 1990s and awarded with the 2018 Nobel Prize in Chemistry. (9) Random mutagenesis (by error-prone PCR or gene shuffling) enables the researcher to explore uncharted territory of chemical reactions by unguided substitutions of the genetic code, thus causing exchanges of single amino acids or regions of an enzyme. Goldsmith and Tawfik found a correlation between increases in activity and the number of mutations in the protein; large improvements (>103-fold activity of the wild type) required an average of five mutations per 10-fold improvement in catalytic efficiency. (10) A specific example: To cover the whole sequence space of a protein with 300 amino acids by using random mutations at five random positions, 1.9.7 × 1012 variants would have to be assessed. (11) Such large numbers of experiments practically exclude the use of chromatographic methods (e.g., GC, HPLC) and can only be managed by spectroscopic techniques, using chromo- or fluorogenic reporters, with short analysis times, low detection limits, and high sensitivity (Figure 1). (6)
Many studies rely on chromo- or fluorogenic substrates that mimic the true target or on enzyme-coupled chromo- or fluorogenic assays for the detection of a wide range of catalytic activities. (12) The major drawback of these methods is that they generate indirect experimental evidence of the desired activity but no analytical signal from the relevant transformation. The most reliable hits are achieved by detecting the target product.
Functional group assays (FGA) use the functionality of a target molecule as an analytical handle. They thereby reduce the number of false-positive results, and the analytical noise from enzymatic background activities. (13,14) In this review, we summarize the development and use of functional group probes for the chemoselective detection of a broad range of target molecules in physiological media (e.g., aqueous buffered solutions, cell culture media). We approach the topic primarily in the context of protein engineering, focusing on the applicability of those tools to the development of new biocatalysts. With this summary, we wish to target two fields that are heavily involved in this area of research: biochemistry and synthetic organic chemistry. We aim to equip the biochemist with a toolset to pursue the discovery of valuable biocatalysts, and we hope to motivate organic chemists to identify and tackle underdeveloped areas in the field of assay development. The review is structured from a chemist’s point of view, ordered by functional groups typically encountered in molecules of interest in academic and industrial research and development: amines, alcohols, aldehydes, ketones, phenols, epoxides, diols, thiols, etc. This classification is intended to enable easy access to information and offer chemical solutions to challenges in screening for enzymatic activity. Compatible buffer ranges, spectral regions of the respective assay product, and the commercial availability of the presented probes are additionally highlighted.

2. Assays in High-Throughput Screening

Click to copy section linkSection link copied!

High-throughput screening is essentially a technical method in which every detail matters. If a good photometric assay is available, a library of enzymes based on a single site-saturation mutagenesis (e.g., using the NNK codon) can be easily screened for activity in a single 96-well microtiter plate. A combinatorial library of two residues (using NNK codons, as before) requires about 4000 clones (for statistical reasons), or 10 384-well plates. (15) When there are even more amino acids in the enzyme that need to be mutated and there is no better way of designing variants than combinatorics, the number of variants to be screened rapidly exceeds the (human) capacity for handling plates and performing manipulations, with or without automation. Miniaturization can offer a significant improvement to throughput, with certain implications on the applied downstream analytics used for the high-throughput screening (HTS; 104–105 variants to analyze), and the development of novel (ultra) high-throughput (uHTS; 106–108 variants) screening methods becomes critical.
There are various scalable methods to obtain experimental information about structure–property relationships in enzymes. Generally, the methods with the highest throughput are in vitro binding assays. (16) The data they produce is sometimes rich enough to complete even complicated scientific tasks, such as solving a protein structure from a deep mutational scan. They are not suitable when the goal is to optimize a biocatalyst for conditions relevant in an industrial process.
Most approaches rely on the spectroscopic detection of target or reporter molecules from absorbance or fluorescence measurements in the visual range, or close to it. Whenever a chromatographic separation is included before detection, throughput drops greatly. Fluorescence-activated cell sorting (FACS) can be used if (i) the reaction is run in cells, (ii) the reporter or product is fluorescent, and (iii) it does not permeate between cells. Recent developments in microfluidic fluorescence-activated droplet sorting (FADS) circumvent the issue of cell permeability by compartmentalizing cells in picoliter-sized water-in-oil droplets but in turn require analytes to be hydrophilic to prevent cross-talk between droplets; applications are currently limited to very polar fluorophores. (17,18) Only few metabolites of interest have useful absorbance or fluorescence, and thus artificial fluorogenic substrates are needed. These artificial probes are often custom-designed and synthesized for a single application only.
High-throughput screening is still in its infancy and largely an academic pursuit where new approaches are being developed for linking enzymatic activity to a spectroscopic signal (Figure 2). While UV absorbance-based assays can detect a broad range of compounds, their sensitivity is usually lower than fluorescence-based ones. Compounds in crude cell lysates might interfere with the absorptive detection of the analyte of choice. Therefore, ratiometric “turn-on” probes, linking product formation to a proportional increase of a ratio of fluorescence signals, constitute the majority in the vast landscape of reporters (Figure 2). (13,19,20) The increase in fluorescence intensity can result from various molecular implementations: (i) the inherent fluorescence of the desired product, formed by the enzymatic reaction (Figure 2a), (ii) a coupled and strongly correlated process that produces a fluorescent molecule, e.g., a cofactor recycling system (Figure 2b), (iii) enhanced transcription of fluorescent biosensors (Figure 2c), (iv) the concomitant formation of fluorescent metabolites (Figure 2d), (v) by induction of Förster resonance energy transfer (FRET) or suppression of photoinduced electron transfer (PET) processes (Figure 2e), or (vi) a tandem reaction producing a fluorescent molecule in a cascade, downstream of the targeted reaction (Figure 2g).
Not all the modes to generate a fluorescent signal listed above are equally applicable to any problem at hand. Given that many relevant products are not or only weakly fluorescent, direct monitoring of their formation is often not feasible (Figure 2a). Not all enzymatic transformations offer possibilities for indirect read-outs, e.g., from cofactor recycling (Figure 2b,d). Many protein engineering efforts thus use analytical tools based on fluorogenic substrates that mimic the chemical properties of the target compound to some degree while still being easy to synthesize. (12) Screening for enzymatic activity with such mimics, however, bears the risk of identifying biocatalysts that are better at producing the mimic and not necessarily the target product. An aphorism capturing this issue, nicknamed as the First Law of Directed Evolution, that is often quoted, says: “You get what you screen for.” (21,22) Assays based on mimics might be useful for proof-of-concept studies or the detection of new catalytic activity; they are not powerful in the optimization of biocatalysts.
Functional group probes (Figure 2f) use chemoselectivity, a result of steric and electronic properties of a functional group, as an analytical handle. This simplification, reducing the analytical complexity to a distinct interconversion of one functional group to another, often enables a clear discrimination of the desired enzymatic activity, even in the presence of substrates and interfering molecules from cellular background. Functional group probes offer an attractive option to reduce the number of false-positive results and the analytical noise from enzymatic or cellular background activities. (13,14) Over the last decades, many functional group sensors have been developed, often benefiting from research in bioconjugation chemistry and vice versa. Among these examples, many have proven to be applicable in protein engineering, when HTS was used to test large libraries of enzyme variants.

3. About High Throughput

Click to copy section linkSection link copied!

In any screening campaign, every manipulation decreases throughput. For instance, if one has to start with plating out cells after transformation, an agar-plate assay will likely have a higher throughput than one based on microtiter plates; (23) or, transferring lysate from a deep-well plate (for cell growth) to an optical plate (for reading) halves throughput by doubling the number of plates to be handled. As a rule of thumb, throughput doubles when the analysis is performed in the same plate as the cultivation of the microorganism producing the enzyme. That also requires the assay reagents to not damage the DNA, which encodes the sequence of a potential hit.

Figure 1

Figure 1. Change of analytical approach depending on screening scale and mutant library size: Increase in library size urges a shift from chromatographic methods to spectroscopic detection for (ultra) high-throughput screenings: (u)HTS. TLC: thin-layer chromatography; GC: gas chromatography; HPLC: high-performance liquid chromatography; (μ)MS: (microscale) mass spectrometry. Created with BioRender.com.

Assays where reagents for detection at the end can be added from the beginning of the reaction (e.g., in a single master mix with the substrate) are better than assays with multiple components that need to be added in a certain order. (24) There are limitations to such straightforward optimization of protocols. For instance, some reagents react with substrates or enzymes, inactivating them, or are unstable in water in general (e.g., fluorescamine 18 (25)) or at certain pH (e.g., DTNB 232 (26)). They can only be added at a later stage, when the chosen reaction time has passed or even just before analysis. The duration should be chosen in consideration of the dynamic range of the detection method without extra dilution (or concentration). Variations between wells on microtiter plates (in growth rates and/or expression levels, even for the same variant) are often large and significant, resulting in cases where relative specific activities cannot be estimated by merely quantifying the product. Continuous kinetic measurements are unsuitable because every moment one sample stays in the reader longer than necessary corresponds to other variants not being screened. Normalization to the actual enzyme content (e.g., using tagged-protein complementation assays like the split-GFP (27) or split-luciferase assays (28)) is possible, but it can be expensive and time-consuming─it is therefore rarely done.
The power of HTS lies in producing information by interrogating many variants, not scrutinizing few. For the purpose of discovery, end-point measurements often provide information as good as kinetic measurements. For this reason, HTS is typically stronger in the discovery of new activities or the early rounds of improvement than in improving enzymes with substantial activity. This trend is a conundrum because initial improvements in activity are often simpler to achieve than further optimization: quantitative performance of the assay has to become more granular, more sensitive, and more reliable at later stages than in the beginning. Driven by this increasing challenge, HTS methods are used for tasks at which they do not excel by design, and work-arounds are introduced (e.g., use of lower substrate concentrations, subsequent dilution steps). The dynamic range of assay systems thus needs to match the aim of the study (discovery or optimization).
The requirements for HTS-capable assay development are high. Ideally, an assay makes as few mechanical and fluidic manipulations as possible, use stable compounds, is sensitive, selective, simple, and has a large dynamic range. Any assay that can run at room temperature is simpler than one requiring heating or refrigeration. High temperatures, extreme pH, corrosive, or toxic chemicals and solvents (especially, volatiles and halogenated ones) should be avoided, as they can make assays incompatible with common equipment (plastic multiwell microtiter plates and their readers). It is critical that assays work under aqueous conditions because crude cell lysates are prepared using aqueous buffers. Therefore, this review focuses mainly on probes used to analyze aqueous samples.

4. Sensitivity and Selectivity

Click to copy section linkSection link copied!

The best functional group probes are those where fluorescence solely depends on the concentration of the analyte. Such a simple relation can be typically observed in systems with only one enzyme and substrate present in the reaction. Many assays, however, require additional components (e.g., more than one substrate or enzyme, cofactors, activators or inhibitors, stabilizers, surfactants, cosolvents, detection reagents, etc.), making the selective detection of the main product challenging.

Figure 2

Figure 2. Overview of several ratiometric absorption or fluorescence turn-on strategies applied in directed enzyme evolution in biological systems: (a) direct detection of activity by functional group transformations or cleavage reaction of the target substrate, (b) indirect detection of conversion via dye cofactor recycling with a concomitant increase in signal, (c) triggered enhanced transcription of fluorescent biosensors, (d) formation of the dye in a downstream reaction upon cleavage of a metabolic reporter, detection via suitable recognition domain or subsequent reaction with a secondary chromo-/fluorogenic substrate, (e) induction of Förster resonance energy transfer (FRET) or suppression of photoinduced electron transfer (PET), (f) functional group transformation is detected via subsequent reaction with functional group selective assay component, (g) tandem reaction cascade of a model-substrate (mimic) unmasks chromo-/fluorophore.

Crude cell lysates, which are the preferred formulation of enzymes used in HTS, compromise the detection because they introduce many intracellular metabolites in high concentrations, often much higher than the concentration of the product. For instance, in Escherichia coli, metabolites present in the mM range include several amino acids (glutamate, at 96 mM, aspartate, valine, glutamine, alanine, and citrulline), several nucleotide phosphates or thiols (e.g., glutathione, at 17 mM, or coenzyme A, at 1.4 mM). (29) The high concentrations of these and other metabolites complicates the detection of low enzyme activities with rather nonspecific colorimetric probes. For example, the derivatization of primary amines using o-phthaladehyde (OPA, 22) in the presence of 2-mercaptoethanol (ME) (24,30) may be influenced by endogenous thiols because evolution of the signal is dependent on the thiol coupling partner.
Buffers too can obfuscate the results of an assay when they contain constituents that might interfere with the chemoselectivity of the probe. For instance, Good’s buffers contain amines, which may therefore compete with the detection of an amino functionality in the product.
Functional group assays work best when the functional group is being created from another in the reaction (i.e., a functional group interconversion, FGI). For example, carboxylate reductases produce aldehydes from carboxylic acids, amidases produce amines from amides, and thioesterases produce thiols from thioesters. In such straightforward cases, probes only need to discriminate one functional group from another and not between two of the same type. Hence, many probes are useful for FGI screening. (31)
In many other reactions (e.g., aminotransferases, isomerases, etc.), functional groups are not interconverted but moved around or between molecules, making chemoselectivity essential and challenging. Probes that cannot differentiate between substrate and product bearing the same functional group (e.g., by forming conjugates with different spectral properties from each) would require separation prior to detection (e.g., extraction or chromatography). Because the substrate is usually present in high concentration to saturate the enzyme (typically, ∼ 100-fold the KM value), specificity for the product by ways of vastly different reaction rate constants is important (because with similar constants, specifically detecting a product at low concentration, is noisy against the background from the substrate).
Examples for the power of chemoselective probes are (i) o-diacetylbenzene (DAB, 26), which can discriminate amines in amino acids from primary amines, used in HTS of amino acid decarboxylases, (32,33) (ii) β-aryloxy fluorescein aldehyde 38 reported by Leslie et al., (34) which undergoes accelerated β-elimination by secondary amines (34) and is valuable for detecting reactions that convert primary to secondary amines (e.g., methyltransferases).
Manipulating the reaction conditions can also be used to change selectivity of an assay by changing the reactivity of the target. For example, fluorescamine 18, a probe for amines, can under acidic conditions be used for the selective detection of aromatic amines in the presence of aliphatic amines because the pKa values of these types of amines are sufficiently different. (35)
Because the selectivity profile of most reported assays is not extensively characterized (even for the small number of commercially available probes), it is often unclear to protein engineers which reagent might be the most suitable to detect the product of interest and many need to be tested. Additionally, many studies on assay development add the analyte in large excess over the detection probe. While this approach is advantageous to show the optical performance of the assay, it does not guarantee its practicability for screenings. It would add great value if research groups developing new assays characterized the selectivity of a new probe in commonly used reaction media toward various derivatives of the same substance or reactivity class (e.g., nucleophiles, electrophiles, oxidants) and (cellular) interfering molecules.
The goal of this review is to present protein engineers, biochemists, and chemists with an overview of available options for the detection of compounds of interest. For synthetic chemists, we provide a summary of known starting points for the development of even more selective probes.

5. Principles for Designing Functional Group Probes

Click to copy section linkSection link copied!

The detection of organic functional groups is widespread in chemistry. Thin-layer chromatography, (36) which is taught in beginner chemistry laboratory courses, often uses chemoselective, chromogenic stains based on reactions with functional groups. Historically, the discovery and development of staining agents improved the characterization of new compounds via derivatization and enabled reliable and fast reaction analytics. Many selective reagents enjoyed great popularity, including Ehrlich’s reagent 1, (37) 2,4-dinitrophenylhydrazine (DNPH) 109, (38) phosphomolybdic acid, (39) Fe(III)Cl3/hydroxylamine, (40) Hanessian’s cerium stain, (41) Simon’s reagent, (42) among others, and are still used for illicit drug screenings (Scheme 1).

Scheme 1

Scheme 1. Representative Reagents Typically Used in Thin-Layer Chromatography (TLC) Staining Solutions
Only a few of these classic probes have been applied in protein engineering, so far, likely because of the drawbacks that many of these agents entail: they are largely incompatible with biological systems, have poor selectivity or sensitivity, and inadequate optical properties. Nevertheless, some emerged as the structural or conceptual basis for many probes that followed.
Designing chemoselective probes has several requirements to enable their general application: (14) (i) Most importantly, the probe must react selectively with the target analyte in the presence of substrates and other background (from cells or the medium). Chemoselectivity has to be sufficiently high to promote the preferential reaction with the desired analyte even in the presence of molecules with similar functional groups. (ii) Probes should react fast under actual screening conditions. A HTS assay should not be limited in throughput by the reaction rate of the probe. (iii) If cells need to be kept alive (e.g., for cofactor recycling), probes have to be nontoxic.
Two options commonly serve as the starting point for initial development: (i) the discovery of a suitable, highly selective chromo- or fluorogenic reaction and subsequent refinement of the chemical probe’s spectral properties, or (ii) the incorporation of a chemical sensing mechanism into an already established chromophore which causes a desired optical read-out.
Detecting targets at low concentration with a short lifetime brings particularly stringent requirements for selectivity and reactivity. Probes that meet these requirements can enable special methods, such as single-cell droplet assays, or HTS campaigns that search for the proverbial “needle in the haystack.”
In addition to defining the chemical properties of a probe, the choice of the chromogenic platform is also essential. Knowledge about the genesis of and control over the optical signal is necessary for the rational design of good probes. Spectroscopic properties of organic molecules at visible and near-visible wavelengths are closely associated with their electronic (delocalized) structures. (43) Extension of a conjugated π–electron system upon reaction is a common element in designing chromogenic probes, which generally leads to a bathochromic shift in the absorption maximum. It does not simultaneously guarantee the emission of a fluorescent signal, which is only observed for a small fraction of absorbers due to the action of prominent quenching mechanisms, such as internal conversion (IC) or intersystem crossing (ISC). (44) Fluorescence is favored over absorption as the signal-generating mode due to a better signal-to-noise ratio and thus lower detection limits. Generally, it is easier to see a strong signal against a weak background than to look for weakening of a strong background.
Most turn-on fluorescent probes are based on a change of the electronic structure of the chromogenic dye, (45,46) triggered by changes to the size of a delocalized electron system (e.g., degree of conjugation of π–electron systems) or to the distribution of electrons in the system (by the introduction, manipulation, or removal of electron-donating or withdrawing groups). The resulting changes in absorption coefficients, quantum yields, and/or Stokes shifts can be measured and, if chemoselective, related to changes of the analytical target. Other mechanisms of fluorescent assays to generate a signal include: suppression of a photoinduced electron transfer (PET), (47) enabling a fluorescence resonance energy transfer (FRET) between donor and acceptor fluorophores, (48) modulating an intramolecular charge transfer (ICT), (49) or enabling an excited-state intramolecular proton transfer (ESIPT). (50) Recent developments, such as aggregation-induced emission (AIE), (51) twisted intramolecular charge transfer (TICT), (52) electronic energy transfer (EET), (53) or C═N isomerization, have not yet found widespread application in the design of reaction-based turn-on probes.
There are excellent reviews on general design principles of fluorescent chemosensors, with a focus on chemoselective bioimaging of endogenous metabolites, (14) fluorescent probes in redox biology applications, (54) imaging of reactive oxygen, nitrogen, or sulfur species (ROS, NOS, RSS), (55,56) anions and cations, (57) screening for defined enzymatic functionalities, (13,58) or the high-throughput aspect of selected methods. (7) Here, we will elaborate on reaction-based probes in the context of protein engineering. We outline the rational design of most known functional-group probes; some of them have not (yet) been used in protein engineering.

6. Chromo- and Fluorogenic Probes for the Detection of Amines and Amino Acids

Click to copy section linkSection link copied!

Amines are a highly valuable class of compounds used in the chemical, pharmaceutical, and agrochemical industries. They are frequently used as intermediates for a vast number of bioactive compounds, especially in their optically pure form. (59) Efficient methods for their enantioselective preparation typically require sophisticated catalyst design or the use of transition metals. (60,61) Hence, chemists have turned their attention to enzymes, which have since been established as an excellent approach for the preparation and use of amines as building blocks. (364)

Figure 3

Figure 3. (top) Summary of general properties of amines, reactivity trends toward amine-selective probes, and expected cross-reactivity. (bottom) Compatible pH range and buffers in reported amine-selective assays. Bold numbers indicate the most commonly used pH values.

To assess the progress of transformations that move amino groups intermolecularly, a probe with a significant difference in reactivity between substrate and product is required. Large differences often result from an increase of substitution and steric bulk. Primary amines and ammonia have the highest reactivity and thus can be efficiently discriminated from secondary amines and most amino acids. Tertiary amines are mostly unreactive toward conjugation due to the decreased steric accessibility. (663)
Amidases have the biggest potential to benefit from these chemical probes because they release amino compounds as products. The strong nucleophilicity of this functional group creates possibilities for electrophilic reactions in addition or substitution reactions. Cross-reactivity in biocatalytic studies can be expected with other nucleophiles present (e.g., thiols, cyanides, and, if not targeted themselves, amino acids). The rate of those side reactions depends on the actual substrate, the conjugation partner, pH, temperature, and buffer composition.
Amines are generally mildly basic, and reactions with aliphatic amines are thus confined to a pH range of approximately 8–10 to avoid deactivation by protonation. Aromatic amines can be selectively targeted in neutral or slightly acidic regimes because their pKa values are generally lower than those of aliphatic analogues (4.5 vs 9). (664)
Amine-containing buffers (e.g., Tris; tris(hydroxymethyl)aminomethane) must be avoided when using amine-reactive probes because the rate of the reaction with the buffer would greatly exceed that with amino groups in the product (Figure 3).
Commonly used enzymes to yield amine-containing compounds, include amidases, (62) lipases, (63,64) transaminases (TA), (65) reductive aminases (RA), (66,67) imine reductases (IRED), (68) monoamine oxidases (MAO), (69) and lyases. (70) Several studies have described the potential of these catalysts (Scheme 2). In a seminal study, Savile et al. applied substrate walking, modeling, and a mutational approach to establish a highly potent transaminase for the synthesis of the antidiabetic compound sitagliptin 3 (Scheme 2a). (71) Their best variant included 27 mutations and converted the ketone 2 to the chiral amine 3 at 200 g/L substrate loading in 50% DMSO, with a 92% assay yield and full enantioselectivity. Challenged with the sterically substantially hindered bicyclic amine 5, Weiß et al. applied protein engineering for the optimization of a fold class I ((S)-selective amine transaminase). (72) Saturation mutagenesis and two rounds of directed evolution facilitated the kinetic resolution of the racemic substrate 4 in excellent yield and full stereocontrol (Scheme 2b). Interestingly, reactions performed with the prochiral ketone as starting material determined the mutant to be unsuitable for direct asymmetric synthesis. Machine-directed evolution was applied by Novartis researchers as a potent enzyme engineering strategy by employing a moderately stereoselective imine reductase as the model system. (73) One screening cycle already yielded a library of highly active mutants with a dramatically shifted activity distribution compared to that of the traditional directed evolution methodology. Using their most selective and active mutant, the authors synthesized the H4 receptor antagonist ZPL389 7 in 72% yield with >99% ee (Scheme 2c). Smart rational engineering of amidase AJ270 from Rhodococcus erythropolis by the group of Wang enabled the desymmetrization of heterocyclic meso-dicarboxamides by generating only 10 variants. (74) Using their biocatalytic platform, they enabled access to both antipodes of O,N-heterocyclic and carbocyclic dicarboxamides 9 in excellent yields and full stereocontrol (Scheme 2d). Another protein engineering example for amidases was delivered by the group of McLeish. (75) A comparison of several members of the enzyme family inferred that residues involved in substrate recognition were not conserved. Substitution of position G202, crucial to substrate binding, to slightly bulkier alanine or valine residues, prompted a dramatic decrease in the kcat/KM value of mandelamide 10a with concomitant increase in lactamide 10b activity, effectively transforming the biocatalyst into an aliphatic amidase (Scheme 2e). Given the scope of this review, however, only a subset of the stated biocatalysts can be efficiently targeted using the presented methodology.

Scheme 2

Scheme 2. Representative Selection of Enzymatic Transformations for the Production of Amine-Containing Products

6.1. Nucleophilic Aromatic Substitution (SNAr)

SNAr reactions of para-substituted electron-poor aromatic systems enable the emergence of an optical signal (Figure 4). The change in absorbance (or fluorescence) is commonly explained by the formation of electronic “push-pull” structures: The incorporation of electron-donating groups (EDGs) by reaction with the amine and the pre-existing electron-withdrawing groups (EWGs) in the para position enable ICT processes. (76)

Figure 4

Figure 4. Summary of chromo- or fluorogenic turn-on probes for the selective detection of amines based on nucleophilic aromatic substitution (SNAr) with aromatic molecules containing electron-withdrawing groups (EWG). Abs: wavelength typically used for UV absorbance measurements. Ex or em: wavelengths typically used for excitation and emission in fluorescence measurements. The wavelengths of probes in this section are shown on the strip at the bottom representing the visible spectrum.

The most common leaving groups are halides (Cl, F) and sulfonates (SO3); fluorides are most reactive. Historically, this concept was used by Sanger for the detection of amines in insulin (Sanger’s reagent: 2,4-dinitrofluorobenzene, DNFB). (78) Nitrated aryl systems have high absorption coefficients and led to the development of nitrobenzoxadiazole (NBD) reagents. Ghosh et al. were the first to report the use of NBD-Cl 12 as a fluorogenic nucleofuge and noticed different optical signatures depending on the reaction partner. Whereas primary and secondary aliphatic amines formed highly fluorescent products, reactions with O, S, and N-Ar nucleophiles only gave weak fluorescence. Tertiary amines were found to be unreactive. (79) Watabe et al. switched to the fluoro-derivative of NBD, which is 50–100-fold more reactive than NBD-Cl, for the detection of hydroxyproline and proline in blood plasma. (80) Recent studies used NBD-Cl 12 for the detection of l-ornithine (81) and NBD-F 13 for the derivatization of Alendronate. (82) In a recent study, Annenkov et al. showed that NBD-Cl can be a viable derivatizing agent even for tertiary amines at room temperature, albeit in dioxane. (77) The substituted products are believed to undergo subsequent Hofmann elimination or substitution by the expelled chloride ion and form a mixture of highly fluorescent NBD adducts. Application in aqueous systems was not investigated.
NBD-fluorophores, with excitation maxima between 460 and 490 nm, are often incompatible with other green-fluorescent reporters, such as the broadly used green fluorescent protein (GFP) with an excitation maximum at 488 nm. To circumvent this overlap and thus improve their usefulness in in vivo studies, Benson et al. substituted the oxygen in the ring with a quaternary carbon (SCOTfluors 14a14e, Figure 5). These probes showed tunable emissions in the near-infrared (NIR) spectrum under physiological conditions and were applied for in vivo cell imaging. (83)

Figure 5

Figure 5. Structures and spectral properties of SCOTfluors 14ae. aValues determined in EtOH. bQY (quantum yield) in dioxane using acridine orange, fluorescein, rhodamine 101, and Cy5 as standards. Reproduced with permission from ref (83). Copyright 2019 Wiley-VCH under Creative Commons Attribution 4.0 International License https://creativecommons.org/licenses/by/4.0/.

Another historical example of an SNAr reagent is 2,4,6-trinitrobenzenesulfonic acid (TNBS 15), which was first described by Okuyama and Satake in 1960. (84) Formation of aniline derivatives upon reactions with aliphatic amines can be traced by observing intense orange bands. Even though it is explosive, TNBS has still found broad use in protein modification of terminal amino acids because it selectively reacts with primary amines. (85−88) Recently, Olivero et al. used the reagent in combination with 3-(4-carboxybenzoyl)quinoline-2-carboxaldehyde (CBQCA, 24) (cf. section 6.3) for the quantification of proteins bound to nanoparticles. (89) With the emergence of new fluorogenic scaffolds, Kim et al. incorporated the amino-responsive function into the BODIPY fluorophore. (90) Upon reaction with lysine residues, the excitation wavelength of the BODIPY derivative 16 shifts from 525 to 409 nm and could be used for fluorescent labeling under mild conditions (37 °C, 2 h) (Figure 6).

Figure 6

Figure 6. Labeling scheme of lysine residues in protein with 16. (a) Normalized absorbance spectra (dashed-lines) and fluorescence emission spectra (solid-lines) of 16 (1 μM) and lysozyme labeled with 16 (5 μg/mL lysozyme, 2 h incubation at 37 °C), measured in HEPES buffer (10 mM, pH 7.4, 0.1% dimethyl sulfoxide). The emission spectra were obtained with excitation at 485 and 375 nm, respectively. The inset photo was taken under UV light (365 nm). (b) A plot of time-dependent (0–4 h) fluorescence intensity at 409 nm. The excitation wavelength was 375 nm. Inset box: lysozyme activity calculated using an assay kit at 2 h incubation point. Means and standard deviations were calculated from triplicate measurements. Reproduced with permission from ref (90). Copyright 2017 Korean Chemical Society; Wiley-VCH.

6.2. Michael Addition (Type) Reactions

As functional group probes, Michael acceptors make use of the nucleophilicity of amines in neutral and basic aqueous solutions. Most probes of this type (Figure 7) undergo a Michael addition and subsequent β-elimination: upon 1,4-conjugate addition of the amine, the leaving group is expelled, forming a stable enamine.

Figure 7

Figure 7. Summary of chromo- or fluorogenic turn-on probes for the selective detection of amines based on Michael addition type reactions. Abs: wavelength typically used for UV absorbance measurements. Ex or em: wavelengths typically used for excitation and emission in fluorescence measurements. The wavelengths of probes in this section are shown on the strip at the bottom representing the visible spectrum.

A prominent example is epicocconone 17 (Figure 8), a yellow dye isolated from the fungus Epicoccum nigrum. (91) Elucidation of the structure enabled its application for the detection and staining of primary amines. (92) The conjugate that forms upon reaction with lysine has a high molar absorptivity (104 M–1cm–1), strong orange to red fluorescence (λem = 610 nm), and a large Stokes shift (100 nm). The stability of the enamine depends on the pH and is highest at 2.4. (92) The derivatization is reversible at neutral pH, which enabled the analysis of proteins by staining gels after electrophoresis. (93) Substitution of the heptatriene chain in 17, especially with a p-chloropyridyl functionality, increases fluorescence yields without impacting the performance in staining. (94)

Figure 8

Figure 8. Epicocconone 17 is the active ingredient in Deep Purple Total Protein Stain and is responsible for the apparent noncovalent staining of proteins in polyacrylamide gels and electroblots. The reaction of 17 with amines has shown that 17 reacts reversibly with primary amines to produce a highly fluorescent enamine that is readily hydrolyzed by base or strong acid. Such conditions are used in postelectrophoretic analysis, such as peptide mass fingerprinting or Edman degradation. Reproduced with permission from ref (92). Copyright 2005 American Chemical Society.

Fluorescamine 18, an intrinsically nonfluorescent reagent, is another commonly used probe for the selective derivatization of primary amines. (25) Although secondary amines and alcohols react with the probe, they do not form fluorescent products. Furthermore, the reaction with alcohols is reversible and thus tolerated in amine screenings. (95) Unfortunately, due to the poor hydrolytic stability of the reagent, an excess must always be used to achieve reliable derivatization, (96,97) but because the hydrolyzed product of 18 is not fluorescent, this assay enjoys high popularity. The maximum in reactivity and fluorescence appears at pH 9, but assays performed at pH 6–8 still provide acceptable results with aliphatic amines.
Acetic acid can be used as a reaction medium to quantify aromatic amines, as shown by Rinde and Troll. (35) Aromatic amines remain unprotonated even in this acidic milieu, making selective detection in the presence of aliphatic amines possible. Two recent studies applied this reagent in HTS: Murugayah et al. used 18 to detect activity of N-acyl-l-homoserine lactone (AHL) acylase, (98) and Ashby et al. used it to detect the interaction between proteins and nanoparticles (changes in the structure of the studied proteins caused a change in fluorescence, which made conformational changes upon interactions with nanoparticles detectable). (99)
1,2-Naphthoquinone-4-sulfonate (NQS), also called Folin’s reagent 19, is one of the most commonly utilized chromogenic reagents for the determination of illicit drugs and pharmaceuticals containing aliphatic or aromatic amines. (100) The reaction of NQS with amines forms an orange dye, which is stable over a broad pH range (2–11). The rate of the reaction is independent of the pH but impacted by the amine substrate. (101) Most assays are complete in less than 20 min (at pH 9–10). (102) Besides amines, it is also possible to detect sulfamides and tetrazoles.
Ylidenemalononitriles (compounds 20ac) are another useful group of compounds for amine assays, as shown by Longstreet et al. (103) They are weakly fluorescent, but upon reaction with (some) primary amines form cyclic amidines with a 900-fold increase in emission intensity. Several derivatives were synthesized (Figure 9, 20a20c), spanning the spectral excitation (λex = 390–510 nm) and emission windows (λem = 500–555 nm). Their use was demonstrated in the derivatization of a human transferrin glycoprotein at pH 7, albeit in 50% DMSO.

Figure 9

Figure 9. Normalized emission spectra of ylidenemalononitrile enamines 20ac in CH2Cl2 at room temperature. Insets show photographs of 1 mM solutions of 20a–c in CH2Cl2 irradiated at 365 nm. Reproduced with permission from ref (103). Copyright 2014 American Chemical Society.

A recent example of a red-emitting probe, PMB 21, was reported by Sathiskumar and Easwaramoorthi. (104) In this compound, phenothiazine and 1,3-dimethylbarbituric acid are connected via a methylene bridge. Fluorogenic properties of the probe are observed upon reaction with several primary aliphatic and electron-rich aniline systems. The described mechanism suggests that addition of the amine leads to the cleavage of the barbituric acid moiety and the formation of phenothiazine-conjugated imines. Higher substituted amines and deactivated aromatic rings did not give any signal. The strong dependence of the fluorescence emission on the reaction partner can be explained by large differences in ICT processes in the products. Compound 21 was used for the detection of 3-amino-l-tyrosine in buffered medium.

6.3. Isoindole Formation

Another approach for the detection of amines is entrapping them in an isoindole structure (Figure 10). In a double-condensation reaction, o-phthalaldehydes, such as compound 22, and amines form highly reactive o-quinoid structures. (24,30) Initial attempts to form isoindolines used reducing agents, e.g., 2-mercaptoethanol (ME) or bisulfite, (105) to “stabilize” these products. Analysis of the highly fluorescent conjugates, however, revealed that 1-substituted isoindoles are formed, their emission depending on the structures of the parent dialdehyde and trapping nucleophiles. Practical trapping agents include aliphatic thiols (e.g. ME, cysteine) sulfites, or cyanide. In consequence, this multicomponent system can also be used for the detection of thiols or cyanides. Especially the electron-withdrawing CN-substituent is believed to stabilize the resulting conjugates. (106) The addition sequence of the reagents is crucial for the intensity of the fluorescence emission. (24) For instance, addition of the thiol before the amine was beneficial (but not necessary) for strong fluorescence. Amino acid conjugates gave a strong fluorescent signal in the range of pH 7–12. The reactions, however, were typically performed above pH 8 to enable a broad spectrum of amines to be detected. With thiols as nucleophiles, a substantial amount of thiolate (pKa ∼ 11) is required. (30) Generally, sterically accessible amines react fastest with the isoindole probes, reducing the derivatization time to several minutes for simple primary amines. (111,112) The fluorescence emission, however, depends on multiple factors, including pH, the trapping nucleophile, the detectable amine, and even temperature. Ammonia generates significant background emission: its conjugates are 100–1000-fold more emissive than those of amines. (107)

Figure 10

Figure 10. Summary of chromo- or fluorogenic turn-on probes for the selective detection of amines based on isoindole formation with o-phthalaldehyde derivatives and o-diacetylbenzene. Ex or em: wavelengths typically used for excitation and emission in fluorescence measurements. The wavelengths of probes in this section are shown on the strip at the bottom representing the visible spectrum.

Enlarging the π–electron system, e.g., with naphthalene as in NDA 23, or analogues, (108,109) is generally accompanied by a large red-shift and increase in sensitivity. A systematic study by de Montigny et al. (106) revealed a strong dependence of the Stokes shift and the quantum yield of the conjugates on the trapping nucleophile. While reactions with 22 were ideally performed with ME, 23 was shown to be practically nonfluorescent with this thiol donor. Cyanide can be used with both dialdehydes, but premixed solutions of 23 with cyanide were prone to decomposition; they had to be prepared separately. Reactions with amino acids were shown to proceed fastest between pH 8.5 and 10, with a maximum at ∼9.5 (23/CN detecting alanine).
CBQCA 24 (89,665) and 5-furoylquinoline-3-carboxaldehyde (FQ, 25) (666−668) are additional analogues for the same class. FQ is several orders of magnitude more sensitive than its parent compound OPA 22, or fluorescamine 18, in the quantification of bovine serum albumin (BSA), but long incubation periods are necessary (90 min for FQ, compared to 30 min for 22). Compound 24 has been used to quantify proteins bound to nanoparticles in HTS by the group of De Crescenzo. (89) Hapuarachchi and Aspinwall used a combination of 22 and SAMSA-F, a thiol-bearing fluorescein derivative for the labeling of primary amines and amino acids. (110) Although their method required the subsequent separation of the tagged amines from the unreacted probe by capillary electrophoresis, it showed how mechanistic understanding can be used in design. Recently, Medici et al. showed that the ketone analogue of 22, o-diacetylbenzene (DAB, 26), can efficiently differentiate between primary amines and their corresponding amino acids, thus enabling HTS of amino acid decarboxylases. (32) As with 22, ME was used as the trapping nucleophile. Given that the resulting isoindole does not condense to a fully aromatic product, a study by Choi et al. attempted to elucidate the structure of the actual fluorophore. (33)
The presented probes have been used for detection of amines, amino acids, and even ammonia as nitrogen sources in cellular systems. Examples using compound 22 include screening for activity of: (i) catalase, measured via the H2O2 triggered depletion of GSH, (113) (ii) nitrilase, by detecting ammonia, (114) (iii) aminotransferase, (115) by ME-catalyzed formation of fluorescent pyrazino-diisoindole-diones 27 (Scheme 3). Unexpectedly, ME was not incorporated into the product but was nevertheless needed as a catalyst. The fluorophore exhibited a blue-shifted excitation maximum at 410 nm, only observed with vicinal diamines; other diamines did not show measurable fluorescence upon excitation at this wavelength.

Scheme 3

Scheme 3. Reaction of 22 with Vicinal Diamine Resulting in the Formation of Pyrazino-diisoindole-dione 27

6.4. Amide Formation via Activated Ester Cleavage

Activated esters are the synthetic chemist’s mimic of nature’s highly reactive thioester or acyl phosphates used for selective amidations. In recent years, many probes were developed by incorporating metastable leaving groups, such as mixed anhydrides, NHS esters (N-hydroxysuccinimide), (116) pentafluorophenyl (PFP)-, (117) or thiophenols, (118) into established fluorescent molecules (e.g., fluorescein, (119) coumarin, (120) BODIPY, (121)etc.; Figure 11). (122) The majority of such designs modify the molecular structure of the dye in a region distant from the fluorophore core, which typically does not significantly change the spectral properties of the dye and thus requires the separation of the conjugate from any unreacted probe after the reaction. (123−126) Conversely, the incorporation of the reactive handle directly into the fluorophore modulates the fluorescent emission and led to the development of reactive turn-on probes.

Figure 11

Figure 11. Summary of chromo- or fluorogenic turn-on probes for the selective detection of amines based on amide formation using activated esters. Ex or em: wavelengths typically used for excitation and emission in fluorescence measurements. The wavelengths of probes in this section are shown on the strip at the bottom representing the visible spectrum.

Typically, the turn-on effect is achieved by liberating fluorogenic carboxy or hydroxy moieties from their initial state, when they are trapped as reactive anhydrides, carbonates, or labile esters. Two examples for ratiometric turn-on probes: (i) 6-(dimethylamino)naphtho[2,3-c]furan-1,3-dione (ANH, 28) (127) and derivatives thereof (128) as well as isatoic anhydride analogues. (129,130) ANH has great potential as an environment- and solvent-sensitive turn-on fluorophore, but it is only weakly fluorescent in water. An increase in fluorescence was only observed upon conjugation to apolar protein domains. To tackle the challenge of poor aqueous solubility of a majority of apolar turn-on-based (NHS-) probes, the group of Ogle derivatized fluorogenic isatoic anhydride 29 with a quaternary ammonium handle. (130) This permanently charged group substantially increased the solubility in water, no longer requiring cosolvents for conjugation, e.g., with BSA. They furthermore demonstrated the efficiency of their probe using a rapid-injection NMR experiment: a 10-fold excess of n-butylamine completely reacted with their probe in less than 5 s.
All reagents of this type, however, were unstable toward hydrolysis, with half-lives of 16 h (pH 7) and 2 h (pH 8.4). This reactivity is generally expected of any activated ester, as hydrolysis under basic conditions always competes with the desired reaction.
Similar to the well-established methods for amine conjugation using cleavable moieties, the group of Kim adapted their meso-carboxyBODIPY 30 probe and achieved a remarkable turn-on effect upon reaction of its activated acid, (131) initially with the NHS- 30a and PFP 30b esters (Figure 12) and later also with thioester 30c. (132) Using their PFP derivative 30b, they showed a 3000-fold increase in fluorescence upon reaction with MeNH2 under physiological conditions in less than 5 min. A similar rate was observed for simple primary aliphatic mono- and diamines. An increase in steric bulk caused a substantial decrease in reactivity and required reaction times of 60 min to reach similar turn-on factors (e.g., 900-fold with cyclohexylamine). Tertiary and aromatic amines were unreactive. Compound 30a produced a strong fluorescent signal upon conjugation to lysine or arginine after 5 min under physiological conditions. The other canonical amino acids only gave a weak signal, likely due to a steric effect. Compound 30b was not tested with amino acids.

Figure 12

Figure 12. (top) Synthesis of fluorogenic meso-active-ester-BODIPYs 30a/30b and their conversion to fluorescent meso-amide-BODIPY products upon reaction with amines. (bottom) Absorption (a) and emission (b) spectra of 30b (10 μM) upon treatment with MeNH2 (20 μM) in CH3CN as a function of time (0–5 min) at 25 °C. λex = 470 nm. (inset) Relative fluorescence intensity at 546 nm as a function of incubation time, in the absence (red) and the presence (green) of MeNH2. (c) Relative fluorescence intensity at 546 nm as a function of [MeNH2] (0–50 μM). (inset) Calibration curve of fluorescence intensity at 546 nm vs [MeNH2] (0–6 μM). (d) Photographs of 30b (20 μM) upon addition of MeNH2 (left to right: 0, 3, 5, 10, 20 μM) under ambient light (left) and 365 nm UV irradiation (right). Incubation time = 5 min. Reproduced with permission from ref (131). Copyright 2020 American Chemical Society.

Figure 13

Figure 13. Summary of chromo- or fluorogenic turn-on probes for the selective detection of amines based on pyridinium salt formation with pyrylium salts. Ex or em: wavelengths typically used for excitation and emission in fluorescence measurements. The wavelengths of probes in this section are shown on the strip at the bottom representing the visible spectrum.

NHS-derivative 30a has remarkable hydrolytic stability at pH 7.4, with a half-life of 2.5 d at 25 °C. Interestingly, changing its reactive moiety to a thioester caused a large decrease in reactivity: It would only react with cysteine, subsequently undergoing a reaction cascade known as native chemical ligation (NCL; intermolecular conversion of a labile thioester to a stable amide) and eventually showing turn-on behavior. As with the 8-SMe–BODIPY 16 (cf. section 6.1), only amino or amide-conjugates gave the desired increase in fluorescence; esters or thiol ethers remained nonemitters. The importance of the correct placement of amino groups on the BODIPY core was highlighted in a paper by Esnal et al. (133) Other dyes bearing such reactive groups could potentially give a similar turn-on behavior upon cleavage of the fluorescence-concealing group, but they have yet to be explored (Figure 14). (134)

6.5. Pyrylium Salt Formation

A fascinating approach for the selective detection of amines builds on initial reports by Katritzky (135) from 1984. The reaction of his, from then on eponymous, pyrylium salts with amines, forming stable pyridinium salts, was taken up by Wetzl et al. (136) Twenty years later, their group modified the pyrylium moieties with color-forming p-anilino substituents (31 and 32), which resulted in a new class of chromogenic labels, the so-called “Chameleon” labels (Figure 13). The class, although not limited to pyrylium salts, generally shows strong bathochromic shifts upon conjugation. Due to their straightforward synthesis from the corresponding aldehydes and methyl pyrylium salts, they have already been applied in HTS as chromogenic reagents for the detection of biogenic amines (Figure 14). (137,138)

Figure 14

Figure 14. Image of a Py-1 31 sensing microplate reacted with various concentrations of histamine (the numbers indicate the concentrations of histamine in μg mL–1). Reproduced with permission from ref (137). Copyright 2011 The Royal Society of Chemistry.

Figure 15

Figure 15. Colors under daylight and the colors of fluorescence in solution of pH 7.0 under a UV lamp (365 nm) of label 31 and of some other Py labels when conjugated to serum albumin at 37 °C. Note that label 33 does not react but remains blue (at any concentration). Reproduced with permission from ref (141). Copyright 2011 The Royal Society of Chemistry.

Although Py-1 31 is the most widely used derivative, (138−140) many other spectral bands are addressable by modification of the aromatic core as exemplified by the benzothiazole derivative 33 (Figure 15). The reactivity of those salts could furthermore be fine-tuned by modifying the substituents of the pyrylium moiety. (141) Primary amines and amino acids (or proteins containing terminal amino groups) could be efficiently labeled at physiological conditions in less than 30 min. Anilines required reactions to be run at 50 °C and in organic solvents. Reactions with secondary amines or ammonia did not form colored products and thus did not interfere with the assay. These probes were stable in acidic solution (pH < 5) but decomposed quickly even at slightly basic pH. The hydrolysis products, however, also did not interfere with the optical read-out. (142)

Figure 16

Figure 16. Summary of chromo- or fluorogenic turn-on probes for the selective detection of amines based on the formation of thioureas or 2-aminothiazoles from isothiocyanates. Ex or em: wavelengths typically used for excitation and emission in fluorescence measurements. The wavelengths of probes in this section are shown on the strip at the bottom representing the visible spectrum.

6.6. Thiourea or 2-Aminothiazole Formation

Isothiocyanates (ITCs) have long known to be reactive toward a great number of nucleophiles and have been used for the synthesis of stable thiocarbamates and thioureas (Figure 16). (143,144) Fluorescein isothiocyanate (FITC) is used for protein labeling, relying on the strong nucleophilicity of amines (compared to alcohols or water). It has disadvantages similar to the active esters of fluorescein discussed above: relatively small changes in the electronic structures upon conjugation require separation of the tagged amines from the unreacted label. (145,146) Thus, only a few examples of efficient use as turn-on probes have been reported.

Figure 17

Figure 17. (top) Fluorescence spectra of unbound 9-ITPPo 35 (blue) and its covalent adduct ATAZPo (red) with BSA (a) and gold nanoclusters (AuNCs) (b). (bottom) Absorbance of a mixture of 35 (15 μM) and n-butylamine (3 mM) in toluene: Structures, time curves of concentration, and spectral profiles of the reactant, the intermediate, and the product (from left to right). Reproduced with permission from ref (148). Copyright 2015 The Royal Society of Chemistry.

The group of Belfield, experienced in the synthesis of fluorene-based dyes, reported the synthesis and simple modification of isothiocyanate dyes, such as compound 34, that exhibited a small hypsochromic shift upon conjugation with an amine. (147) Using their probe, they performed the exemplary derivatization of reelin, a large extracellular matrix glycoprotein, at physiological pH. Although in these experiments the unreacted fluorophore was removed after the reaction, they could show a large increase in quantum yield upon thiourea formation, in DMSO.
Recently, Planas et al. used another prominent scaffold, porphycene, as a fluorogenic sensor. (148) They modified it (9ITPPo 35; Figure 17) to give a 70 nm Stokes shift to the near-IR upon conjugation, enabling spectral discrimination from the unreacted probe. Isolation of the reaction products of 35 with several primary, secondary, and even aromatic amines revealed that the underlying cause for the color change was a cyclization of the initially formed thiourea into a 2-aminothiazole fragment. Although synthetic methods for this type of cyclization are known, (149,150) the spontaneous cyclization, particularly for these scaffolds, had previously not been reported. (151) For a proof of concept, derivatizations of BSA, and amino-functionalized gold nanoclusters were performed in carbonate buffer at pH 9.2 and in EtOH. Although the first step (formation of thiourea) was completed in less than 15 min, the fluorescence continued to shift significantly due to the rate-limiting cyclization, which required prolonged reaction times. Biocompatible auxiliaries to the cyclization might thus lead to the development of new turn-on methods based on these ITC scaffolds.

Figure 18

Figure 18. Summary of chromo- or fluorogenic turn-on probes for the selective detection of amines based on hemiaminal, aminal, or Schiff’s base formation. Abs: wavelengths typically used for UV absorbance measurements. Ex or em: wavelengths typically used for excitation and emission in fluorescence measurements. The wavelengths of probes in this section are shown on the strip at the bottom representing the visible spectrum.

6.7. Hemiaminal, Aminal, or Schiff’s Base Formation

Another approach for the detection of amines is based on the disruption of electron-withdrawing properties of carbonyl, imine, or imino groups (Figure 18). When amines add to such functional groups and form stable hemiaminals, they tend to induce a hypsochromic shift of conjugated electron systems. (282) In some cases, further elimination (of water, alcohols, or amines) to imines or iminium salts is favored, and these transformations have also been shown to result in changed spectral features.

Figure 19

Figure 19. Amine-sensing panels of 37 (10 μmol L−1) in water: (1) 37; (2) butylamine; (3) tert-butylamine; (4) benzylamine; (5) cyclohexylamine; (6) ethylene diamine; (7) 1,3-diaminopropane; (8) cadaverine; (9) morpholine; (10) ephedrine; (11) 4-aminopyridine; (12) ethanolamine. Reproduced with permission form ref (154). Copyright 2012 Wiley-VCH.

The first turn-on probes based on the formation of Schiff’s bases were reported by the group of Glass in 2003: (152) They used coumarin aldehydes, such as compound 36, for the detection of amino acids and primary amines. By mimicking chelation-based metal sensors, based on the protonation of the formed imine (Scheme 4), they achieved a 26-fold increase in fluorescence at pH 7.4 (buffered) with glycine as the target. Water-soluble analogues containing aliphatic tertiary amine groups as well as carboxylic acid functionalities were also synthesized but not yet spectroscopically characterized. Alkylation of the 4-position of the coumarin was found to be necessary for a large change in absorption: the imine adopts a trans-configuration due to steric repulsion.

Scheme 4

Scheme 4. Sensing Mechanism of 36: Chelation of the Formed Iminium Species by the Coumarin Moiety Triggers a Turn-on Response (152)
A second example emerged from the experience with cross-shaped (“cruciform”) fluorophores by the group of Bunz. (153) Based on initial designs of such dyes, Kumpf and Bunz synthesized the water-soluble, aldehyde-appended distyrylbenzene 37, which showed strong turn-on behavior in aqueous systems. (154) Their reaction with a broad set of (biogenic) primary and secondary amines as well as amino acids at μM concentrations yielded different Stokes shifts upon the formation of hemiaminal or imine structures, with emissions in the range from 450 to 520 nm. Among the tested amino acids, only cysteine, arginine, and lysine gave an increase in fluorescence. The signal strength was strongly dependent on the amine. Whereas morpholine exhibited merely a 10-fold turn-on factor, the reaction with benzylamine or ethylenediamine caused a ∼200–300-fold increase in fluorescence. Compound 37 and its aminal and imine products were stable in neutral aqueous solution.
A recent example of the use of Schiff’s base formation was reported by Leslie et al. (Scheme 5). (34) This study, aiming to optimize their previously reported chemodosimeter for the detection of ozone, revealed that the intermediately formed β-aryloxy fluorescein aldehyde 38 undergoes β-elimination, which is rate-determining and catalyzed by amines. The addition of pyrrolidine increased the rate 30-fold and drove the reaction to completion in less than 3 min; the uncatalyzed reaction took 1.5 h. From a selection of cyclic secondary amines, they identified azetidine as the best catalyst. Anilines were excluded from the reaction screening due to their known reactivity with air and ozone. The authors speculated that the efficiency of the catalyst was primarily dependent on steric effects and the pKa of the amine. The rate enhancement was greatest at pH 6 to 7, and the elimination product of 38 showed the highest fluorescence between pH 5 to 9. However, due to the focus of the study on detecting ozone, use of 38 as an amine-selective probe was not investigated.

Scheme 5

Scheme 5. Ozone Sensing Mechanisms Employing 38. (34)
Another reagent relying on the nucleophilic attack of the amine onto an electrophilic center was developed by Wu et al. (669) Their synthetically readily accessible iminium dye BI-CA-ID 39 effectively recognized primary amines upon reaction by ratiometric discoloration of the dye solution and emergence of a blue-shifted fluorescence emission with a peak at 382 nm. The reaction with primary and secondary amines was reported to be instantaneous. Completely discoloring of sample solutions at 40 μM in less than 1 min (containing 800 μM amine) in 50% ACN (acetonitrile)/water mixtures was observed. Even tertiary amines were reported to be viable for the assay, when allowing reaction times of up to 10 min. Aromatic amines were reported to be completely unreactive toward the probe.
Trifluoroacetyl probes (e.g., 40 and 41) are another class of compounds not yet proven to be applicable in aqueous systems but with frequent occurrence in literature. Although we generally limit this review to reactive probes for aqueous solutions, these compounds deserve an honorable mention. Initial reports by Mohr already exemplified the usefulness of these dyes as potent fluorogenic sensors for amines in organic solvents. (155−157) At the core, stilbene or azobenzene motifs are responsible for the spectral features of the dyes; they have since then been modified into even larger delocalized structures. (158) Their kinetics and thermodynamics have been studied, (159) they have been turned into star-shaped, tripodal fluorophores with variable reactive trifluoroacetyl reactive sites, (160) and incorporated into a BINOL-scaffold for the detection of diamines. (161−163) A demonstration of their viability in aqueous systems is yet to be published.

7. Chromo- and Fluorogenic Probes for Alcohols

Click to copy section linkSection link copied!

Alcohols are crucial building blocks for the manufacturing of agrochemicals, flavor and fragrance compounds, and pharmaceutical drugs, especially chiral alcohols in their optically pure form. Typical methods to prepare them enantioselectively typically require the use of expensive, complicated chiral catalysts or transition metals, which can be toxic. (164−166) Biocatalytic approaches are considered an effective and green alternative due to their mild reaction conditions and remarkable enantioselectivity. A recent review by Chen and de Souza highlighted the recent developments in the field. (167) Although a large variety of biocatalysts is available for the formation of chiral alcohols, only few are used in combination with functional group probes. One possible explanation is the relatively low reactivity of alcohols, which impedes selective targeting, especially in an environment where many competitive, strong nucleophiles are present. Below, we highlight a few pertinent examples, chosen from the many success stories of applications and optimization of such enzymes.

Figure 20

Figure 20. Summary of general properties of alcohols and expected cross-reactivity (bottom).

Nature offers a broad repertoire of catalysts for the synthesis of chiral alcohols (Table 1). (168) Using directed evolution, Zhang et al. achieved the reduction of 1-(4′-chlorophenyl)-1-(pyridine-2′-yl)-methyl ketone (CPMK, 42; Scheme 6a). (670) They identified a crucial mutation (S273A) among a library of 2000 clones using the colorimetric DNPH 109 assay to quantify unconverted starting material. In an attempt to identify new enzymes based on sequence similarities, the group of Pohl discovered a new (R)-selective HNL from Arabidopsis thaliana. (170) This enzyme converted a wide range of benzaldehydes, and aliphatic aldehydes and ketones, to the corresponding cyanohydrins and was established as an alternative to other known (R)-selective HNLs. For instance, the enzyme formed the cyanohydrin of 2-hexanone 44 with 95% ee (48% conversion; Scheme 6b).
Table 1. List of Enzyme Classes for the Enantioselective Synthesis of Chiral Alcohols
catalystactivity
alcohol dehydrogenases (ADH) (171) and ketoreductases (KRED) (172)asymmetric reduction of carbonyl compounds
α-ketoacid carboligases (173) or aldolases (174)decarboxylative or regular aldol reactions
hydroxynitrile lyases (HNL) (175)addition of cyanide to aldehydes forming cyanohydrins
(Michael) hydratases (176)addition of water to unactivated or α,β-unsaturated compounds
epoxide hydrolases (177)enantioselective hydrolysis of epoxides affording vicinal diols
oxidases (178)oxidative functionalization of C–H bonds
haloacid dehalogenases, (179) haloalkane dehalogenases (180) or halohydrin dehalogenases (181)hydrolysis of 2-haloalkanoic acids, halogenated aliphatics, or vicinal haloalcohols
lipases (182) or esterases (183)(dynamic) kinetic resolution of racemic mixtures

Scheme 6

Scheme 6. Representative Selection of Enzymatic Transformations for the Production of Alcohol-Containing Products
Chen et al. tackled the challenging enzymatic Michael addition of water motivated by initial reports for the formation of (S)-3-hydroxy-3-methylfuranone 47 from 3-methylfuran-2(5H)-one 46 using Rhodococcus rhodochrous ATCC 17895. (184) They compared six closely related Rhodococcus strains and found similar activities for the transformation with all six bacteria. Optimization of the process parameters of their initially studied strain enabled the formation of 47 in 82% isolated yield and 95% ee on gram scale (Scheme 6c).
Lehwald et al. reported the first asymmetric intermolecular carboligation of an aldehyde and a ketone using a ThDP-dependent flavoenzyme from Yersinia pseudotuberculosis (Scheme 6d). (185) Their catalytic system overcame the known hurdle of low electrophilicity and increased steric bulk of ketones (e.g., 48), which typically hinders aldol reactions with these substrates. The carboligase is promiscuous regarding the substrate, accepting a broad range of cyclic and open-chain ketones, diketones, and α- or β-ketoesters.
An unusual example for the synthesis of chiral alcohols was reported by Hammer et al. (186) They exploited the highly Brønsted-acidic active site of a squalene hopene cyclase (SHC) from Alicyclobacillus acidocaldarius and enabled a stereoselective Prins reaction of (R)-citronellal 50. A single point mutation in the active site (I261A) facilitated the stereospecific cyclization to (−)-iso-isopulegol 51, among the four possible products (Scheme 6e).
Enzymatic late-stage modifications are a flourishing trend in total synthesis. Biocatalysts are uncontested in their selectivity, especially in transformations of highly decorated compounds. The group of Stoltz used an enzymatic approach for the late-stage oxidation of 52 at the C-7 position and screened a small library of cytochrome P450 catalysts. (187) A double mutant was shown to selectively oxidize the intermediate 52 to the corresponding alcohol 53, which was subsequently treated with Dess–Martin periodinane to form the desired natural product Nigelladine A 54 (Scheme 6f).
Many protein engineering efforts rely on indirect detection methods using coupled enzymatic reactions, summarized below. The most widely applied protocol oxidizes the produced alcohols with appropriately chosen alcohol dehydrogenases (ADHs) or alcohol oxidases (AOs), generating NADH, which can be quantified via the reduction of MTT, 3-(4,5-dimethyl-2-thiazolyl)-2,5-diphenyl-2H-tetrazolium bromide 55, forming purple-colored formazan 56abs,max = 550 nm) (Scheme 7). (188−190)

Scheme 7

Scheme 7. MTT Assay: Detection of NAD(P)H or “Reducing Equivalents” via the Reduction of MTT 55 to the Purple Formazan 56
The oxidized products can also be directly detected using carbonyl-selective assays (cf. sections 9 and 10). Chromogenic Purpald 113 has already been used for this purpose by detecting formaldehyde (FA), for example, in the directed evolution of a terpene synthase (191) and in engineering a CYP153A35 oxidase to enhance its ω-hydroxylation activity toward palmitic acid. (192)
Luciferase can be used to further oxidize aldehydes (produced via ADH catalysis from the target alcohols) to quantify alcohol production by the emergence of a blue-green luminescence (λem,max = 490 nm). (193,194)
The drawbacks and limitations of these enzymatically coupled detection methods have motivated the development of alcohol-sensitive probes despite the known issues with selectivity in relevant reaction environments. These probes rely either on an approach similar to that described for detecting amines (nucleophilic attack on electrophilic carbonyl species, cf. section 6.7) or on coordinating metals. Unfortunately, these methods have been found broadly incompatible with aqueous media or poorly sensitive, so far impeding their application in protein engineering. The competition with other nucleophilic species, biogenic amines, but mostly water, creates too much noise in detection. Separation of the analytes from competing nucleophiles (e.g., extraction into an organic solvent) might ultimately be necessary to overcome these limitations (Figure 20).

7.1. Hemiacetal Formation

The reversible formation of hemiacetals from aldehydes or trifluoroacetyl groups can be used to detect alcohols, as the addition of the alcohol changes the spectral properties of these probes (Figure 21). This mechanism is identical for amino and carboxylate probes and will create background signals in the presence of these substance classes (cf. sections 6.7 and 11.3).

Figure 21

Figure 21. Summary of chromo- or fluorogenic turn-on probes for the selective detection of alcohols based on hemiacetal formation. Abs: wavelengths typically used for UV absorbance measurements. Ex or em: wavelengths typically used for excitation and emission in fluorescence measurements. The wavelengths of probes in this section are shown on the strip at the bottom representing the visible spectrum.

The group of Cosa applied this method with their meso-BODIPY scaffolds 57, which become fluorescent upon nucleophilic addition of primary alcohols (Figure 22). (195,196) Conversely, amines formed nonfluorescent imines. Incorporation of electron-withdrawing substituents (e.g., CN) onto the BODIPY core and the use of catalytic amounts of p-toluenesulfonic acid strongly improved the reaction rate. Still, the reaction with water was 10 times faster, and assays had to be performed in dry organic media.

Figure 22

Figure 22. (a) Fluorescence–time trajectories following the reaction of 4.5 μM of compound 57 (R = Cl, R′′ = H) and 1.09 M methanol (MeOH), ethanol (EtOH), butanol (BuOH), ethylene glycol (EtyGly), water (H2O), or ethanedithiol at 21 °C in ACN supplemented with 0.333 mM of p-TsOH. (b) Apparent rate constants for various nucleophiles tested were obtained from fitting the intensity–time trajectories in (a). Reproduced with permission from ref (195). Copyright 2017 American Chemical Society.

Sawminathan and Iyer designed the polarity-sensitive fluorescent dye 58, based on an imidazole core functionalized with a conjugated push–pull system. (197) Addition of MeOH and EtOH triggered a hypsochromic shift of the fluorescence maximum, enabling the ratiometric detection of these alcohols in ACN. Compound 58 was applied for the determination of MeOH in biodiesel and used to detect vapors with paper-based test strips (Figure 23).

Figure 23

Figure 23. Photographs showing the color changes of sensor 58 before and after the addition of methanol vapor. The images were taken under UV irradiation (365 nm). Reproduced with permission from ref (197). Copyright 2021 Royal Society of Chemistry and the Centre National de la Recherche Scientifique.

Mohr et al. extensively studied solvatochromic trifluoroacetyl stilbenes (e.g., compound 59), (156,198,199) which are used for the detection of amines (cf. section 6.7) and found a hypsochromic shift of absorption upon hemiacetal formation in many apolar and polar organic solvents. The reaction with EtOH could be either followed by an increase in absorbance at 360 nm or as fluorogenic turn-off probe (λex,max = 450 nm, λem,max = 570 nm) in buffered media. The fluorescent effect was only observed at concentrations of EtOH above 10 v/v%.
Takano et al. developed the trifluoroacetyl compound BC-1 60, synthesized from a lyophilized bacteriochlorin derivative. (200) This alcohol-sensitive probe fluoresces upon excitation either at 431 or 731 nm, circumventing potential optical interference from the sample. The probe was adsorbed on PVC (because the study was focused on developing alcohol-responsive membranes), giving a 2.5-fold increase in fluorescence with 25 v/v% EtOH in buffered solutions at pH 4–8.

7.2. Coordination to Metals and Chelation

Alcohols can be detected using reactions where they act as ligands, are coordinated to metals, or are chelated by conformationally rigid organic structures (Figure 24). Zhao et al. developed the probe ZR1 61, which tautomerizes upon coordination to an alcohol, (201) causing an increase in ICT and concomitant increase in purple color and fluorescence (at 605 nm). The reaction is selective for MeOH; other alcohols did not react. This selectivity was demonstrated by titrating an ethanolic solution of the probe with methanol. The presence of water was problematic: concentrations as low as 15% eradicated the signal, making an application in aqueous environments impossible.

Figure 24

Figure 24. Summary of chromo- or fluorogenic turn-on probes for the selective detection of alcohols based on the coordination to chelation centers. Ex or em: wavelengths typically used for excitation and emission in fluorescence measurements. The wavelengths of probes in this section are shown in the strip at the bottom representing the visible spectrum.

A probe developed by Roy et al. uses an aluminum species for the selective detection of MeOH. (202) Displacement of coordinating ethoxide on PPY 62 by MeOH caused a change in geometry from the tetrahedral to the octahedral complex, suppressing intramolecular PET and causing an approximately 7-fold increase in fluorescence at 405 nm. A small percentage of water (5 v/v%) was necessary to observe the emission. Higher concentrations than 5% were detrimental, leading to complete annihilation of the signal at 75% and above.
The group of Wang developed a so-called donor−π-system–acceptor (D−π–A) probe TTO 63, using click chemistry to form the triazole as a coplanar π–electron bridge between the sensing units. (203) This approach enabled the facile synthesis of several structural analogues. The emissive properties of the probe were strongly dependent on the solvent and were shown to be selective for MeOH; there was only weak fluorescence with EtOH and none with any higher alcohols. Similar to 62, an increase in water concentration was detrimental to the observed signal intensity.

7.3. Ring Opening of 3-Aminorhodanine

A new strategy for turn-on probes responsive to MeOH was reported by Kumar et al. They synthesized Schiff’s bases of 3-aminorhodanine with aldehyde-bearing fluorophores, which undergo ring-opening methanolysis at the imide C–N bond (Figure 25). The probes are typically nonemissive due to inhibited isomerization of the C═N bond and additionally active PET from the lone pair of the aldimine N atom to the fluorophore. The initially studied coumarin probe RS 64 was proven to be unreactive toward water and alcohols, including EtOH, nPrOH, and nBuOH; other nucleophiles were not tested. (204)

Figure 25

Figure 25. Summary of chromo- or fluorogenic turn-on probes for the selective detection of alcohols based on the ring-opening of 3-aminorhodanine derivatives. Ex or em: wavelengths typically used for excitation and emission in fluorescence measurements. The wavelengths of probes in this section are shown in the strip at the bottom representing the visible spectrum.

The function of the probe was demonstrated in aqueous and ethanolic solutions after 15 min reaction times, with a limit of detection (LOD) of 0.042 wt % MeOH in water. Later, the authors reported a hydroxynaphthalene analogue NRSB 65, (205) which had similar functionality and reactivity trends, except for EtOH (also detected). Incubation with increasing MeOH concentrations (0–10 wt %) in ACN demonstrated a potential for colorimetric detection with the naked eye (LOD: 0.46 wt %; Figure 26).

Figure 26

Figure 26. (a) The visible color change of 65 (5 mM) in different solvents: (i) hexane, (ii) toluene, (iii) cyclohexane, (iv) ethyl acetate, (v) chloroform, (vi) dichloromethane, (vii) tetrahydrofuran, (viii) dioxane, (ix) ACN, (x) MeOH, (xi) EtOH, (xii) iPrOH, and (xiii) nBuOH. (b) The absorption spectra of 65 in different solvents, (c) time-dependent absorption spectra (0–30 min) of 5 mM 65 in MeOH; inset: plot of optical density at 470 nm vs time. (d,e) Absorption spectra of 65 (5 mM) at different ratios of ACN:EtOH (0–10%) and ACN:MeOH (0–10%), respectively. Reproduced with permission from ref (205). Copyright 2019 Royal Society of Chemistry and the Centre National de la Recherche Scientifique.

The conjugate of 65 with MeOH was only weakly fluorescent. The addition of Al3+ to the ring-opened product greatly increased the emission due to chelation-enhanced fluorescence (CHEF) by inhibiting cistrans isomerization as well as PET. Hence, both MeOH and Al3+ are necessary for a strong turn-on response.

8. Chromo- And Fluorogenic Probes for Diols

Click to copy section linkSection link copied!

Diols are found in many relevant compounds, including pharmaceuticals (206,207) and natural products. (208,209) They are used as precursors (206,207,210−213) in the synthesis of sulfites, sulfates, epoxides, and carbonyl compounds. (214,215) Cyclohexadiene diols have been frequently used as starting material for carbohydrates, alkaloids, prostaglandins, and terpenes. (216) The reactivity of diols is similar to that of alcohols: They can react as nucleophiles forming ethers, or esters with a carbonyl donor (Figure 27). Furthermore, they can be oxidized to carbonyl compounds or undergo dehydration or condensation reactions used in polyester synthesis. (215,217,218)

Figure 27

Figure 27. (top) Summary of general properties of diol sensors and expected cross-reactivity. (bottom) Compatible pH range and buffers in reported diol-selective assays. Bold numbers indicate the most commonly used pH values.

Common enzymatic approaches to generate vicinal diols use epoxide hydrolases, (219) dicarbonyl reductases, (207) or dioxygenases. (220,221) Zhao et al. reported the discovery of several microbial epoxide hydrolases (EH) that enabled the diastereoselective synthesis of (R,R)-diol 67 or (S,S)-diol 69via desymmetrization of various aliphatic and aromatic meso-epoxides 66 and 68 (Scheme 8a,b). (222) Makarova et al. designed four different synthetic routes for ent-oxycodone 72. In all approaches, the initial synthetic step employed a toluene dioxygenase (TDO) to obtain a diene-diol 71 from a whole-cell fermentation with E. coli (Scheme 8c). (223)

Scheme 8

Scheme 8. Representative Selection of Enzymatic Transformations for the Production of Diol-Containing Products
Enzymatic procedures yielding 1,3-diols include the enzymatic reduction of 1,3-diketones with Pichia farinosa IAM 4682, (224) aldol additions with aldolases, (225) or the reduction of hydroxy ketones with ketoreductases. (226) Wu et al. employed a diketoreductase from Acinetobacter baylyi for the stereoselective double reduction of ethyl 3,5-diketo-6-benzyloxy hexanoate 73 to ethyl 3R,5S-dihydroxy-6-benzyloxy hexanoate 74, which is used in the synthesis of statin drugs (Scheme 8d). They optimized the whole-cell catalytic system coexpressing diketoreductase and glucose dehydrogenase to work without added cofactors at a high substrate concentration (15-fold increase from base level). (227) Deoxyribose 5-phosphate aldolase (DERA) from Shewanella halifaxensis was used to synthesize the 1,3-diol 76a, an intermediate in the biocatalytic synthesis of islatravir 76b (Scheme 8e). In a nine-enzyme cascade, the product could be obtained in a single aqueous solution without isolation steps in 51% yield. (228)
Numerous diol-sensing assays employ boronic acid-based probes and make use of the formation of fluorescent cyclic boronate esters as a read-out. Works in this area have been reviewed extensively. (209,229−233) Other approaches are based on indirect detection, such as converting diols to aldehydes and their subsequent detection with aldehyde-sensing assays. It remains challenging to detect diols selectively. Particularly, the discrimination between carbohydrates with minor differences in reactivity and stereochemistry is problematic. (234)

8.1. Esterification of Boronic Acids

One approach to detect 1,2- or 1,3-cis-diols is to use boronic acid derivatives: They react reversibly in aqueous solution, forming cyclic boronates (Figure 28). (235) Since initial reports by Czarnik et al. and Shinkai et al., various boronic acid-based fluorescent probes have been developed. (236,237) In 1992, Yoon and Czarnik were the first to describe the use of anthranylboronic acid as a fluorescent probe for polyols in water. (237) In general, fluorescent boronic acid-based probes consist of a fluorophore as a signal unit and a boronic acid as a recognition site. (238) Probes with various fluorophores, including naphthalene, anthracene, quinoline, coumarin, cyanine, and BODIPY, have been developed. (229,230) Upon binding of the diol, a change in fluorescence based on processes, like PET, ICT, or FRET, can be observed. (209,230,233,239)

Figure 28

Figure 28. Summary of fluorogenic turn-on probes for the selective detection of diols based on the formation of fluorescent boronate esters. Ex or em: wavelengths typically used for excitation and emission in fluorescence measurements. The wavelengths of probes in this section are shown in the strip at the bottom representing the visible spectrum.

The interaction between boronic acids and diols is pH-dependent. Boronic acids (neutral) are trigonal planar, whereas the boronates (anionic) are tetrahedral; their esters behave analogously. Usually, the pKa values of the boronate esters are lower than those of boronic acids. (233)
In general, equilibrium constants of ester formation with neutral boronic acids are lower than with boronates. Therefore, common hypotheses are that (i) diol binding is favored at a higher pH value when mainly anionic boronate is present and (ii) that boronic acids with low pKa values tend to have high affinities for diols, and (iii) that a pH value higher than the pKa of the boronic acid is favorable for diol binding. There are reports contradicting these statements, indicating that not only the pKa of the boronic acid but also the pKa of the diol needs to be considered. (233,240)
Steric hindrance, buffer composition, and the solvent play a crucial role as well. Generally, diols with a small distance between their oxygen atoms bind to boronic acids with higher affinity, quantified by a small (and constrained) dihedral angle in the diol and consequently an O–B–O bond angle that favors sp3 hybridization of the tetrahedral molecule.
Boronic acids are Lewis acids and might interact with buffer components in that way, particularly if the buffer contains Lewis bases, e.g., phosphate or chloride, influencing how the probe reacts with water or diols. (233,241,242,240) Cross-reactivities might arise with functional groups such as amino alcohols, cyanides, alcohols, fluorides, α-hydroxy acids, or α-amino acids. (239,241)
Many boronic acid-based probes have been developed for various structurally different saccharides. Gao et al. developed the naphthalene- and ICT-based probe 4-(dimethylamino)naphthalene-1-boronic acid (4-DMANBA, 77) to detect saccharides in phosphate buffer at pH 7.4. The amino group acted as a donor and the boron atom as an acceptor in the ICT process. A 41-fold increase in emission intensity at 445 nm was observed upon adding fructose, which was the largest reported increase with an ICT-based probe at that time. Besides fructose, also galactose, glucose, sorbitol, and tagatose showed fluorescent responses. Moreover, the water-soluble probe could be readily synthesized in one step. (243) The authors observed that a change in the substitution pattern strongly influenced the fluorescence properties. The derivative 6-DMANBA 79 proved to be a turn-off fluorescent probe, whereas 5-DMANBA 78 is a ratiometric probe showing large changes in fluorescence intensity at 513 and 433 nm upon binding to diols. (244,245)
Zhang et al. described several water-soluble naphthalene-based probes for saccharides and observed that 5-(monomethylamino)-naphthalene-1-boronic acid (5-MMANBA, 80) and 5-aminonaphthalene-1-boronic acid (5-ANBA, 81) showed an even higher increase in emission intensity (71-fold and 66-fold, respectively) upon the addition of fructose than previously reported structural analogues. (246) From their studies of carbohydrates on the surfaces of cells, Yang et al. identified quinoline fluorophores as particularly stable and water-soluble analogues. For example, 8-quinolineboronic acid (8-QBA, 82) is a useful probe for the detection of carbohydrates in aqueous solutions at physiological pH. At a pH < 5, background fluorescence was observed. Addition of fructose led to a 47-fold fluorescence increase at pH 7.5. A fluorescence enhancement was also observed with galactose, tagatose, and arabinose. (247) The structural analogue 5-QBA 83 exhibited a similar enhancement of fluorescence. Its binding constants were higher than 82, and the probe could be applied from pH 3.5 and higher. (248)
Cheng et al. developed various fluorescent isoquinolinylboronic acid-based probes (IQBA) 8486, some with opposite changes in fluorescence intensity depending on whether fructose or glucose was added. Weak binding of these probes to cis-cyclohexanediol and binding of some probes with methyl α-d-glucopyranose as a model glycoside was observed. (249)
Elfeky reported a probe, 4-(5-dimethylamino-naphthalene-1-sulfonamido)-3-(4-iodophenyl)butanoic acid (DNSBA 87), which, when combined with a diol quencher, showed fluorescence recovery in the presence of various saccharides and nucleosides at pH 8.2. In contrast, the binding of the hydroxylated acids citric acid, tartaric acid, and glucuronic acid prompted a decrease in fluorescence intensity. (238)
Near-infrared (NIR) probes have several advantages over emitters in the visible range, particularly no interference from the autofluorescence of biomolecules. Samaniego Lopez et al. developed a water-soluble NIR turn-on probe 88, which consisted of a tricarbocyanine, a piperazine unit as a linker, and a boronic acid moiety. They observed that the selectivity was pH-dependent: at pH 7.4, the probe had a high affinity for fructose and sorbitol but not for galactose, glucose, or mannose. Increasing the pH from neutral to basic increased the fluorescence. The probe also showed a fluorescent response upon binding to mucin, human serum albumin, and asialofetuin. (250)
Wang et al. reported the fluorescent, rhodamine B-based probe ATP-Red 1 89, which rapidly and selectively reported intracellular ATP levels using a multisite-binding strategy. In the presence of ATP, the boronic acid formed an ester with the ribose moiety of ATP, and π–π stacking between the xanthene and the adenine moieties and electrostatic interactions of the amino group with the phosphate groups took place. (251)
Huang et al. introduced a concept to improve selectivity, based on ensembles of boronic acid receptors (90). Interactions of π–electron systems and the pyridinium cation and interactions between the diols and the boronic acids led to the formation of conjugates with d-fructose, (unordered 1:1 stoichiometry) and with d-glucose (ordered 2:1 stoichiometry). The ordered complex gave strong fluorescence at 510 nm, but the unordered was only weakly fluorescent at 380 and 400 nm. Binding of the probe to d-glucose was approximately four times stronger than with d-fructose. To improve this ratio, Huang et al. added phenylboronic acid, which binds to fructose 40 times stronger than to glucose, to improve the selectivity for glucose. (252)
Another concept for carbohydrate detection with boronic acid derivatives uses the combination of a cationic viologen compound linked to a boronic acid moiety and a fluorescent anionic dye. The fluorescence of the dye is quenched by coordination to the viologen compound and restored in the presence of saccharides. Vilozny et al. applied a system composed of the boronic acid-appended viologens 3,3′-o-BBV 91/4,4′-o-BBV 92 and commercially available, fluorescent 8-hydroxypyrene-1,3,6-trisulfonic acid trisodium salt (HPTS, 93) for enzymatic assays of sucrose phosphorylase (E.C. 2.4.1.7) and phosphoglucomutase (PGM, E.C. 5.4.2.2). (253) The binding of the enzymatic products, fructose and glucose-6-phosphate, to the boronic acid moieties led to the recovery of fluorescence, which allowed the successful HTS of PGM inhibitors and the determination of enzyme kinetics (Figure 29). (253) Sharrett et al. showed that instead of HPTS 93, aminopyrene trisulfonic acid (APTS) 94 could also be used, which had constant fluorescence intensity from pH 4 to 10 and could be readily decorated with handles for polymerization. They applied it for the continuous detection of glucose (in the mM range) by immobilization of the APTS monomers in hydroxyethyl methacrylate hydrogels. (254)

Figure 29

Figure 29. Sensing mechanism for the selective detection of fructose or glucose-6-phosphate by 91 or 92 and viologen HPTS 93. Reproduced with permission from ref (253). Copyright 2009 Elsevier BV.

Differentiation between diol-containing compounds remains a major challenge in the development of boronic acid-based probes. Choi et al. sought a rapid method to screen for activity in P450 mutants and, among other approaches, conducted preliminary experiments with probe 82. They found the assay incompatible with solid agar plates and not fast or sensitive enough to detect cis-diols in the nM range in solution. High noise from other diol-containing compounds caused many false-positive identifications of mutants. (23)
One approach to achieve strong binding affinity and high selectivity is to use probes with multiple recognition elements, for instance, by including several boronic acid moieties or other functional groups. (229,255,256) Using this principle, fluorescent probes for the detection of diol-containing molecules, such as glucosamine, (257) digoxin, (258) rifampicin, (258) erythromycin, (258) and d-glucose, (259) have been developed.
As previously mentioned, chiral diols are part of many intermediates, pharmaceuticals, and catalysts. Fast and straightforward (probe-based) methods to determine the enantiomeric or diastereomeric excess (ee, de) and the absolute configuration are desirable. Shcherbakova et al. developed a technique that enabled the determination of ee/de, absolute configuration, and yield of 1,2- or 1,3-diols in high throughput. The assay is based on the self-assembly of 1,2- or 1,3-diols with enantiopure tryptophan methyl ester (or tryptophanol, 95) and 2-formylphenylboronic acid (2-FPBA, 96), resulting in diastereomeric boronates that exhibit different fluorogenic properties depending on the absolute configuration. This approach was applied in determining the optical purity of chiral diols such as atorvastatin. (260,261)

8.2. Oxidative Cleavage

An indirect approach to detect diols is converting them to carbonyl compounds by oxidative cleavage with metaperiodate 97 and subsequently using a carbonyl-specific assay (Figure 30) (cf. section 9). The reaction of 1,2-diols with periodic acid or periodates leads to the formation of a cyclic intermediate, which subsequently collapses to give ketones and/or aldehydes. (217) Doderer et al. used imine-based probe 123 to detect the formed aldehydes colorimetrically. With this approach, they determined the activity of epoxide hydrolase from Streptomyces antibioticus Tü4 toward styrene oxide and 1-butene oxide. (262) Other assays are based on cleavage using stoichiometric amounts of periodate and subsequent titration of the remaining unreacted oxidant with chromogenic or fluorogenic reagents. (263,264) Wahler and Reymond demonstrated that adrenaline 98 could be used this way: it reacted rapidly with the excess of sodium periodate, forming the red dye adrenochrome 99 (Scheme 9a). (263)

Figure 30

Figure 30. Selective detection of diols based on the cleavage reaction by metaperiodate 97 and subsequent application of aldehyde sensing methodology (for aldehyde-selective probes, see section 9).

Scheme 9

Scheme 9. (a) Oxidation of Adrenaline 98 by Metaperiodate 97, Forming Red-Colored Adrenochrome 99, Which Allows for Determination of Excess 97 (Back Titration of Diol); (263) (b) Structure of Carboxyfluorescein, Which Can Also Be Used as a 97 Indicator; The Mechanism for Its Oxidation Is Not Known
Kirschner and Bornscheuer used the above approach in the screening of Baeyer–Villiger monooxygenase mutants for enhanced activity and enantioselectivity. The reactions formed acetic acid esters, which were hydrolyzed by an esterase, yielding a 1,2-diol that was detected. (265) Apart from 1,2 diols, this so-called adrenaline test can also be applied with 1,2-amino alcohols, 1,2-diamines, or α-hydroxy ketones. It has been used to assay lipases, esterases, phytases, and epoxide hydrolases. (263,266,267)
Titration of residual periodate was also efficiently performed with carboxyfluorescein 100 (Scheme 9b), which exhibited a decrease in fluorescence upon its reaction with residual periodate. (268) Sialic acids in glycoproteins can also be detected and quantified using periodate, as shown by Matsuno and Suzuki. They converted the released formaldehyde to a fluorescent dihydropyridine derivative with acetoacetanilide 135 (cf. section 9.4) and ammonia. N-acetylneuraminic acid, N-acetylneuraminyllactose, sorbitol, and galactitol all gave enhanced fluorescence, but d-glucose and methyl α-d-glucoside did not; this effect was explained by their slow oxidation at 0 °C. (269)
Recently, Preston-Herrera et al. developed a HTS assay for Rieske dioxygenase activity. These enzymes catalyze the stereo- and regioselective dihydroxylation of aromatic compounds. The assay was based on oxidizing enzymatically formed cis-diols to the dialdehydes with sodium metaperiodate and the subsequent reaction with fluoresceinamine 127 (cf. section 9.2), which led to a fluorescence enhancement at 520 nm (Figure 31). It was applied in HTS of toluene dioxygenase variants to broaden their substrate scope. (270)

Figure 31

Figure 31. (a) Coupled reactions employed by the fluorescence-based assay system for the detection of cis-diol metabolites using fluoresceinamine 127. (b) Concentration-dependent fluorescence response of the assay to the presence of cis-diol metabolites. All studies were performed in triplicate; all measurements demonstrated standard deviation <5%; fluorescence responses normalized to negative control ([II0]/I0). Reproduced with permission from ref (270). Copyright 2021 The Royal Society of Chemistry.

9. Chromo- and Fluorogenic Probes for Aldehydes

Click to copy section linkSection link copied!

Aldehydes play an essential role in many areas of modern civilization, including their widespread use as food and flavor ingredients (e.g., nutty, fatty, fruity, or grassy notes), as synthetic precursors for fine chemicals and pharmaceuticals, as well as their use as valuable building blocks for the synthesis of macromolecules in the rubber and plastics industry. (671,672) Although aldehydes are highly reactive and often toxic, they are ubiquitous as metabolic intermediates of natural compounds, drugs, and xenobiotics. Volatile aldehydes, such as formaldehyde, acetaldehyde, and acrolein, are prevalent in the exhaust of internal combustion engines, in tobacco smoke, and as industrial pollutants. (634,673)

Figure 32

Figure 32. (top) Summary of general properties of aldehydes, reactivity trends toward aldehyde-selective probes, as well as expected cross-reactivity during screenings. (bottom) Compatible pH range and buffers in reported aldehyde-selective assays. Bold numbers indicate the most commonly used pH values.

Most microorganisms use aldo-keto reductases (AKRs) or alcohol dehydrogenases (ADHs) to eliminate aldehydes, limit their accumulation, and reduce their toxicity. (674) This homeostatic machinery complicates the biocatalytic synthesis of aldehydes in whole cells. Nevertheless, the drawbacks of typical approaches of organic synthesis to make aldehydes (toxicity of metals; additional steps required for reduction to alcohols or protection; unforgiving, narrow reaction parameters) have motivated the development of enzymatic transformations in recent years. By repurposing natural enzymatic pathways, a broad set of biocatalysts has been identified and applied in the biotransformation of suitable substrates to aldehydes (Scheme 10). These pathways include reduction of carboxylic acids by carboxylic acid reductases (CAR), (271) decarboxylation of α-ketoacids by decarboxylases (KDC), (272) reduction of activated esters by acyl-ACP reductases or acyl-CoA reductases, (273) or the well-studied oxidation of alcohols by alcohol dehydrogenases (ADH) (274) and alcohol oxidases (AO). (275) These biocatalysts now rival classic methods from organic chemistry in their performance, convenience, and safety.

Scheme 10

Scheme 10. Selection of Enzymatic Transformations for the Production of Aldehyde-Containing Products
While screening for activity in allylic oxidation, the group of Kroutil found an alkene oxidase that is capable of the biocatalytic equivalent of ozonolysis, converting substituted styrene derivatives 101 to the corresponding benzaldehydes 102. (276) Employing a different biocatalytic route toward aromatic aldehydes, Schwendenwein et al. optimized the promiscuous CAR from Nocardia iowensis to improve its activity for sterically demanding 2-substituted benzoic acid derivatives 103. (277) Screening 6000 mutants using a selective ABAO aldehyde assay (126) identified a single-point mutation (Q283P) which led to a 3–7-fold increase in activity toward o-substituted substrates.
Soh et al. used a directed evolution strategy to optimize the decarboxylation of 2-ketoisovalerate 105 catalyzed by a KDC at 60 °C. Facilitating the reaction at this temperature would enable the production of isobutyraldehyde 106 during the decomposition of cellulose in lignocellulolytic thermophiles. Random mutagenesis combined with further modeling was used to identify a mutant with a 4-fold increased half-life at the desired temperature. (278)
Improvements in catalytic utility of alcohol oxidases were achieved by tackling the problem of poor O2 solubility in aqueous media, which limits mass transfer and results in low reaction rates. By running reactions in continuous flow, the group of Hollmann demonstrated the oxidation of trans-hex-2-enol 107 to trans-hex-2-enal 108 with turnover numbers >3 × 105 using an aryl alcohol oxidase from Pleurotus eryngii (PeAAOx). (279)
The high reactivity of aldehydes, higher than many other electrophilic species, makes the detection of these compounds in physiological environments relatively simple. Their reactive nature also means that unwanted reactions occur readily, such as oxidations, reductions, and condensation reactions with cellular material. (675) Often, their volatility requires fast-reacting probes to achieve efficient detection. Additionally, formation of geminal diols at non-neutral pH needs to be taken into account; this is important, especially for the detection of formaldehyde, which exists predominantly in its hydrated form in water. (676) Most applied ratiometric assays rely on the reaction of amines or amino-derivatives with aldehydes, forming imines, hydrazones, or oximes. These primary conjugates are often converted further to stable heterocycles via intermolecular reactions, irreversible rearrangements, or oxidations to prevent their hydrolysis. Among these conjugates, hydrazones or oximes are inherently more stable due to the inductive effect of the adjacent nitrogen or oxygen atoms. (340) Because aldehydes are much more reactive than other carbonyls, only minimal cross-reactivity can be expected with ketones, esters, or amide derivatives. Aromatic or sterically demanding aldehydes react more slowly than aliphatic ones; electron-donating effects can additionally reduce their electrophilicity. The low reactivity can be compensated by using highly nucleophilic (electron-rich) reagents, which sometimes are prone to oxidation (Figure 32).

9.1. Hydrazone Formation

The most widely applied method to detect aldehydes is based on the formation of stable hydrazones (Figure 33). Historically, hydrazines were used to identify (potentially liquid) carbonyl species by derivatization to solid hydrazones, which typically have well-defined and thus extensively catalogued melting points. (280) The drive toward developing fast analytical tools has repurposed these compounds for use in spectroscopic detection systems, (38) which have continuously evolved since their initial application as TLC staining solutions.

Figure 33

Figure 33. Summary of chromo- and fluorogenic turn-on probes for the selective detection of aldehydes based on the formation of hydrazone adducts. Abs: wavelengths typically used for UV absorbance measurements. Ex or em: wavelengths typically used for excitation and emission in fluorescence measurements. The wavelengths of probes in this section are shown in the strip at the bottom representing the visible spectrum.

An example of a reagent used for selective carbonyl detection is dinitrophenylhydrazine (DNPH, 109), also called “Brady’s reagent”. Since its first mention in 1926, (281) it has been widely used to detect the carbonylation of proteins. It requires a strong acid as a catalyst to form hydrazones, which limits the application of such reagents for detection in situ. Unreacted probe has to be separated because of the small Stokes shift. (283) However, deprotonation of the resulting hydrazones with strong bases typically causes an 80 nm bathochromic shift, separating the signal from background created by the unreacted reagent. (38) Solutions of 109 are generally prepared in dilute phosphoric acid (284) or dilute HCl. (285) This method was used to quantify carbonyls in oxidized proteins (284) and to determine the concentration of benzaldehyde and phenylpyruvate in the optimization of the stereoselectivity of a threonine aldolase. (286) The reagent was also used in a HTS of variants of mandelate racemase to detect oxidized α-hydroxy acids. (285,287)
The use of 109 is limited to absorbance in the UV, so its fluorescent benzoxadiazole (BD) analogues have been widely adopted as fluorogenic reporters for aldehydes and ketones; (288) they emerged from the fluorogenic amino-selective NBD-Cl 12. (289,290) Substituting the electron-withdrawing nitro-group with more soluble sulfonamides (ABD-H 111, DBD-H 112) increases reaction rates with carbonyl compounds while maintaining a similar spectral profile. The cleanest reactions with aldehydes were found with NBD-H 110 (commercially available as hydrazine adduct), whereas the sulfonamides reacted more cleanly with ketones.
Generally, the substituted hydrazines presented here are synthesized from commercially available chlorides or fluorides by nucleophilic aromatic substitution with hydrazine. Similar to 109, assays are typically performed in acidic (288,289) solution or with a high excess of the reagent under physiological conditions (pH 7.5 in PBS buffer). (291)
Anilines are known to catalyze the conjugation: (292−294) their low pKa values (Figure 34) make them nucleophilic even at pH 5–7, forming highly reactive imines as intermediates and thus activating the carbonyl group. These species undergo fast conjugations at physiological pH. Aniline derivatives bearing hydrogen-bond donors in o-position were found to additionally boost the reaction by iminium formation. These “catalysts” are typically used in excess to the applied probes to compensate for the slow kinetics caused by the poor solubility and reactivity of assay probes. Furthermore, a high salt concentration was found to be beneficial for the conjugation of such systems. (295)

Figure 34

Figure 34. Plot of the nucleophilicity parameters N of amines versus pKaH in water. (296) Reproduced with permission from refs. (297and298). Copyright 2007 American Chemical Society.

NBD-H 110 is the most prominent representative of this class of probes. It has been used in HTS to evaluate esterase reactions with vinyl esters as acyl donors in organic solvents. (291) The release of acetaldehyde upon esterification was tracked efficiently using this assay. The sulfonamide analogues (111, 112) have yet to be used in HTS. Nevertheless, they have seen use in the chromatographic separation of unsaturated aliphatic aldehydes with low molecular mass. (162,299)
A long known reagent that is still in regular use is 4-amino-3-hydrazino-5-mercapto-1,2,4-triazole (Purpald, 113). (300,301) This chromogenic probe reacts with aliphatic and aromatic aldehydes at room temperature to form tetrazinanes, which quickly oxidize to the violet-colored tetrazines 115 (Scheme 11). The coloration requires the oxidation of the tetrazinanes, which can be promoted with H2O2, dilute periodate, or atmospheric oxygen. Vigorous shaking of the sample to enable aeration is often sufficient. (302) The assay is performed under strongly basic conditions (e.g., 0.5–2 M NaOH). (302,286,303)

Scheme 11

Scheme 11. Sensing Mechanism of Purpald 113via the Intermediate Formation of Cyclic Tetrazinanes and Subsequent Oxidation to Fully Aromatized Violet-Colored Tetrazines 115
Purpald has been extensively used in off-line HTS and protein engineering studies. Lauchli et al. used the reagent for the directed evolution of a terpene synthase: Non-natural isoprenoids released methanol upon cyclization by enzymatic catalysis, (191) which was oxidized to formaldehyde using an alcohol oxidase and subsequently detected using 113 at 550 nm. Another example, from the group of Arnold, used 113 in protein engineering of a P450 monooxygenase. (304) Directed evolution transformed a known epoxidase from the rhodobacterium Labrenzia aggregata (P450LA1) into an efficient anti-Markovnikov alkene oxidase. The aldehydes produced in the reaction were detected using 113.
The group of Fraaije engineered an alcohol oxidase fromPhanerochaete chrysosporium (PcAOX) to enhance the oxidation of glycerol. (305) The cavity-enlarging mutation F101S, which increased kcat to 3 s–1, was identified using an assay based on 113.
Another widely used hydrazine probe for the detection of aldehydes is 3-methyl-2-benzothiazolone hydrazone (MBTH) 114. (306) A mechanistic study revealed that two equivalents of the reagent are required to form the blue, cationic dye: one molecule reacts with the aldehyde (116a) and the other is oxidized (116b). They further react to form the highly conjugated system 117, a dark-blue dye with an absorption maximum at 635 nm (Scheme 12).

Scheme 12

Scheme 12. Sensing Mechanism for MBTH 114
Concluding from the mechanistic analysis, the chromogenic reaction only happens when there is an excess of 114 that has not reacted with an aldehyde and is available for oxidation. An excess of aldehyde might thus not be detectable. (307) Contrary to 113, the reaction of 114 is performed at acidic pH (0.1 M HCl) and with ferric chloride as the oxidant, forming the blue dye. Modifications to the assay include the addition of sulfamic acid, which prevented the emergence of turbidity during the oxidation process. (308)
In a comparison of three chromogenic reagents for the detection of formaldehyde (Purpald 113, Nash reagent 132 (cf. section 9.4), and MBTH 114), 114 proved to be the only dye that formed stable adducts and prevented further oxidation toward formic acid in the screening of an alcohol oxidase (AO). (303) Even though 114 has not yet found widespread use in protein engineering, it has been applied to determine the concentration of reducing sugars (309,310) and of total aldehyde content in fuel ethanol samples. (311)
Recently, the group of Bane synthesized a dye with a benzocoumarin core and a hydrazine handle. This reagent, BzCH 118, was used for the detection of carbonylation in A549 lung carcinoma cells. (677) The quantification of the carbonyl content in the cells was performed in HEPES buffer for 4.5 h or in growth medium for 30–60 min at room temperature. Upon reaction with an aldehyde, the emission shifted from 430 to 550 nm due to disaggregation of the fluorophores. The conjugated hydrazones additionally exhibited an exceptionally large Stokes shift (195 nm) with no meaningful overlap between its absorption and emission spectra. The emission of the probe was shown to be stable from pH 4 to 8.
Methyl 5-MeO-N-aminoanthranilate 119 is a fluorogenic probe for aldehydes developed by Suchý et al., based on a modified anthranilic acid. It was used to measure the concentration of aldehydes in cells. (312) The reaction with their test substate (malondialdehyde) immediately led to the formation of a blue-green fluorescent species (λem = 480 nm), followed by a rapid decay of the signal (50% within 5 min; Figure 35). Still, initial enhancement of fluorescence with aliphatic aldehydes was between 5- and 20-fold.

Figure 35

Figure 35. (a) 119 as a fluorogenic probe for the detection of MDA. (a) Malondialdehyde (MDA) was added to a 25 μM PBS solution of 119 (pH 7.2) under illumination at 365 nm and photographed at indicated time points. (b) (left) Emission spectra of 1 μM 119 (black) or 119 after 1 min incubation with 5 equiv MDA at 37 °C and excited at 310 nm (blue) or 375 nm (cyan). (right) Formation and rapid decay of short-lived intermediate formed in the reaction between 119 and MDA followed by fluorescence spectroscopy with λexem = 375/480 nm. Kinetics of 25 μM 119 only (squares) and 119 + MDA (circles) are shown, with three replicates of 119 + MDA shown (light, medium, and dark cyan). (c) 119 shows broad specificity to aliphatic aldehydes. The enhancement of fluorescence relative to 119 only was recorded with λexem 375/480 nm following 1 min of incubation of 1 μM solutions of 119 in PBS (pH 7.2) with 5 equiv of MDA (black bar), simple aliphatic aldehydes (red bars), dialdehydes, and complex aldehydes (blue bars), ketones (purple bars), glutathione (yellow bar), and biogenic metal cations (orange bars). Reproduced with permission from ref (312). Copyright 2019 The Royal Society of Chemistry.

In PBS buffer, the reagent cyclized to blue fluorescent indazoles, which would not react any further; no such reaction occurred in DMSO. Because the desired reaction is much faster than the side reaction in PBS (e.g., 10-fold faster with hexanal), the probe could still be used to study the peroxidation of lipids in HEK293 cells incubated with diethyl maleate.
The group of Kool developed the fluorogenic hydrazine derivative “DarkZone” 120, preloaded with an aromatic azoaldehyde that quenches fluorescence. (313) Upon reaction with the target aldehyde, the quencher is liberated and the desired turn-on is achieved. A variety of other fluorogenic dyes could be incorporated into the “DarkZone” approach (e.g., BODIPY, DEAC, Cy3, fluorescein, TAMRA). Only reactions with aliphatic aldehydes produced a strong signal because bisaryl hydrazone products (similarly to the structure of the original probe) also caused a quenching of the fluorescence emission. To facilitate conjugation at physiological pH, the authors tested various anilines and identified 5-methoxyanthranilic acid as a nontoxic and efficient catalyst for the transformation. Those tests were performed with a large excess of catalyst and analyte (dye/aldehyde/aniline catalyst in a ratio of 1:4000:10000) in phosphate-buffered systems at pH 7, resulting in an enhancement of up to 30-fold. To demonstrate the applicability of their system, aldehyde production from ethanol in K562 cells was monitored using their acetylated fluorescein analogue, which resulted in a 2.5-fold increase in fluorescence after incubation of the cells for 24 h.
Liu et al. (314) developed hydrazine dyes based on the squaraine (SQ) scaffold, such as 121, which have narrow absorbance and fluorescence spectra in the near-infrared region. (315) The cyclobutene ring is electron-deficient and thus susceptible to nucleophilic attack; this property is extensively used in colorimetric and fluorescent probes for the detection of nucleophiles. (316) The loss of π-conjugation upon the addition of various amines or hydrazine bleaches the blue color. Subsequent addition of aliphatic or aromatic aldehydes reverts the optical properties of the parent SQs, most prominently observed with the SQ–hydrazine system (Figure 36). Using the water-soluble SQOH–hydrazine 121, addition of an excess of formaldehyde was proven to completely scavenge the hydrazine, accompanied by a 23-fold increase in absorption and 86% recovery of the emission intensity of the original squaric acid dye in water.

Figure 36

Figure 36. (top) Scheme of the reaction mechanism of colorimetric and fluorescent detection of aldehydes based on the SQs–N2H4 adduct 121. (bottom) The changes in UV–vis absorption of dye 121 (6.0 mM) upon addition of N2H4 (3.0 mM) and subsequent addition of formaldehyde (10.0 mM). Reproduced with permission from ref (314). Copyright 2019 Elsevier BV.

9.2. Imine Formation (and Trapping)

This group of reagents uses changes of spectral properties upon formation of Schiff’s bases from amines with aldehydes. Imines are less stable in water than hydrazones. Therefore, most of these probes irreversibly entrap the initially formed imine with intra- or intermolecular nucleophiles (Figure 37). The earliest examples of this class of compounds include pararosaniline 122, its methylated analogues rosaniline (also Schiff’s reagent, 123), magenta II 124, and “new fuchsin” 125.

Figure 37

Figure 37. Summary of chromo- and fluorogenic turn-on probes for the selective detection of aldehydes based on the formation of imines (Schiff’s bases). Subsequent entrapment of these electrophilic imines by intra- or intermolecular nucleophiles makes the reaction irreversible. Abs: wavelengths typically used for UV Absorbance measurements. Ex or em: wavelengths typically used for excitation and emission in fluorescence measurements. The wavelengths of probes in this section are shown in the strip at the bottom representing the visible spectrum.

Methylation has a large effect on the absorption spectra after conjugation to formaldehyde. (317) Robins et al. used NMR spectroscopy to study the colored species and the kinetics of its formation. (318) They found that “careful calibration is required” to obtain accurate results when quantifying aldehydes, with a complicated dependency on the concentrations of the dye, the sulfurous acid (or SO2) used as the nucleophile, and the aldehyde.
Despite these issues, the reagent has been used to determine o-dealkylation activity of 4′,7-dihydroxyisoflavone (daidzein) hydroxylase by detection of the purple formaldehyde adduct using an in vivo assay on LB agar plates (Figure 38). (23)

Figure 38

Figure 38. (a) The color of the aldehyde-sensing probe 122 changes to purple when treated with formaldehyde (detection limit: μM concentrations of formaldehyde in the in vitro system). (b) Color changes on solid agar plate from O-dealkylation activity. Aldehyde appears to diffuse out of the cells; cell color changes are observed in the presence of 122. Reproduced with permission from ref (23). Copyright 2013 Wiley-VCH.

Amino benzamidoxime (ABAO, 126) reagents are another class of probes using an initial Schiff base formation for the selective detection of carbonyl compounds. According to the postulated mechanism, the intermediately formed imine is trapped by the nitrogen of the oxime. The resulting quinazoline oxides show a substantial shift in absorbance from 360 to 405 nm and strong fluorescence at 490 nm. Kitov et al. (319) were the first to report an in-depth analysis of these detection systems and found a linear free energy relationship (LFER) with the pKa of the respective anilinium species (Figure 39). A p-nitro substituent led to a loss of activity while a p-methoxy group increased the conjugation speed.

Figure 39

Figure 39. Relationship between the measured reaction rates and pKa of the corresponding protonated anilines for seven ABAO analogues 126a–g. (i) Structures of ABAOs 126 and corresponding protonated anilines. (ii) Scatter plot showing relationship between log k and pKa. (iii) Reaction rates measured by NMR in CD3COONa buffer (100 mM) at pD 4.5.(iv) Structures and reaction rates for 126f and 126g. Basicity of aniline nitrogen in 2-aminopyridine / 126f was estimated using ab initio calculations. Reproduced with permission from ref (319). Copyright 2014 American Chemical Society.

Rudroff et al. used ABAO for a systematic mutation study of a carboxylic acid reductase from Nocardia iowensis (CARNI). (277,320) Using the 3-methoxy substituted MeO-126 reagent as a chromogenic aldehyde probe, 6000 mutants of CARNI mutants created using error-prone PCR (epPCR) were screened for improved activity with a panel of 20 aromatic aldehydes. The best mutant (Q283P) had a 9-fold improvement in KM and worked even with typically unreactive o-substituted benzoic acids. In a similar fashion, the ABAO assay was used in screening a library of 598 clones of Mycobacterium marinum (CARMM) to improve the production of piperonal in Pichia pastoris. (321) The assay is typically performed under acidic conditions (pH 4.5) and was also shown to work at physiological pH, but with lower activity. (322)
Xing et al. demonstrated the use of 5-aminofluorescein 127 as potent fluorogenic probe for the detection of a broad range of aliphatic aldehydes. (323) Using the restored fluorescence emission of the dye upon Schiff base formation, they used 127 to detect microbial oxidation of primary alcohols by the ADH NCIMB 621 from Gluconobacter oxidans at pH 6.0 at mM substrate loadings. The detection was improved by using a biphasic mixture of water and isooctane, which accumulated hydrophobic aldehydes in the organic phase. Building upon these findings, Preston-Herrera et al. applied probe 127 in a HTS assay for activity of a Rieske toluene dioxygenase. (270) They converted the cis-diol metabolites to the corresponding dialdehydes by oxidation with sodium metaperiodate, followed by detection of the dialdehydes with 127 (cf. section 8) (Figure 31). The authors showed that Tris buffer was detrimental for the oxidation due to possible reaction of metaperiodate with the hydroxy groups of the buffer. Reactions performed in HEPES or phosphate buffers worked well, with the highest rates at pH 6.5, which was attributed to acid catalysis of the imine formation. A 10–30-fold enhancement of fluorescence emission was observed using the optimal conditions after 5 h. They then used the assay to evaluate mutants from a site-saturation mutagenesis, identifying the critical active site residue L272 and subsequently applying the assay for the screening of a broad range of toluene derivatives.
Li et al. recently reported the development of the rhodamine derivative dRB-EDA 128, which undergoes ring-opening of its deoxylactam moiety upon reaction with an aldehyde (Scheme 13). (324) Cleavage of the C–N bond restores conjugation and causes an increase in fluorescence (only shown in DMF). Solutions of 128 (1 mg mL–1) exposed to low mM concentrations of formaldehyde showed a strong increase in absorption at 560 nm and fluorescence at 590 nm after 90 min (Figure 40).

Scheme 13

Scheme 13. Sensing Mechanism for dRB-EDA 128: Schiff’s Base Formation with Formaldehyde Triggers the Ring-Opening of the Deoxylactam Moiety

Figure 40

Figure 40. Chromogenic detection of formaldehyde with 128. (a) UV–absorption spectra of 128 in the presence of formaldehyde (concentration: 0, 3.5, 7, 10.5, 14, 17.5, and 35 mM, from bottom to top). The inset shows the titration curve by absorbance at 560 nm. (b) Visual detection of formaldehyde (0–7 mM as indicated) with 128 in DMF after incubation at room temperature for 90 min. Reproduced with permission from ref (324). Copyright 2011 The Royal Society of Chemistry.

The detection of simple aliphatic and aromatic aldehydes, such as 4-hydroxybenzaldehyde and hexanaldehyde, was shown. The assay was furthermore applied for the staining of oxidized sialoproteins on the cell surface of L929 cells, which were then analyzed by confocal fluorescence microscopy.

9.3. Imidazole Formation

A frequent design for sensitive assays uses the formation of stable heterocycles that irreversibly trap the analyte while concomitantly enlarging a delocalized electron system. This method has been applied for the detection of aldehydes using a combination of 1,2-diketones and suitable ammonia donors, forming imidazoles (Figure 41). Variations of the long known colored 2,4,5-triphenylimidazole (trivial name: lophine) formed by the condensation of benzil, benzaldehyde, and ammonia (325) have been analyzed as potential chemiluminogens. (326) Two particularly promising derivatives emerged from these studies: 2,2′-furil 129 and 9,10-phenanthrenequinone 130. Both probes were used as derivatization reagents, with sensitivity down to the nM range, for the monitoring of aliphatic aldehydes in human serum. (327,328) Derivatizations typically required incubation at 100 °C for 30 min in concentrated ammonium acetate solutions, making an in situ analysis of aldehydes impossible.

Figure 41

Figure 41. Summary of fluorogenic turn-on probes for the selective detection of aldehydes based on the formation of imidazole structures. Ex or em: wavelengths typically used for excitation and emission in fluorescence measurements. The wavelengths of probes in this section are shown in the strip at the bottom representing the visible spectrum.

The aromatic diamine analogue 1,2-diamino-4,5-dimethoxybenzene (DDB, 131) has been reported as an effective reagent for the fluorometric detection of aromatic aldehydes, (329) acrolein, (330) and methylglyoxal. (331) Derivatizations either required strongly basic conditions at room temperature or heating to 80 °C in DMSO. Its application in situ or in protein engineering has not yet been reported.

9.4. 1,4-Dihydropyridine Formation

Aldehydes and ammonia condense with 1,3-diketones to form chromogenic 1,4-dihydropyridines, a reaction also used for the Hantzsch pyridine synthesis (Figure 42). (332) The yellow-colored products (333) often have chromogenic and fluorescent properties, depending on the 1,3-diketone. (334) The simplest diketone, acetylacetone 132, can be used, either in its ketone form (“Nash reagent”) or condensed with ammonia (Fluoral-P), for the detection of aldehydes at room temperature in 2 M ammonium acetate buffer at pH 6. (335,336) Conjugates with formaldehyde showed particularly strong fluorescence, other aliphatic and aromatic aldehydes gave only weak fluorescence in solution (and could only be detected with sensitive detectors).

Figure 42

Figure 42. Summary of fluorogenic turn-on probes for the selective detection of aldehydes based on the formation of 1,4-dihydropyridines. Ex or em: wavelengths typically used for excitation and emission in fluorescence measurements. The wavelengths of probes in this section are shown in the strip at the bottom representing the visible spectrum.

Sawicki and Carnes hypothesized that the rigidity of the fluorophore might influence signal strength (334) and tested 1,3-cyclohexanedione 133 and dimedone 134, resulting in a 100-fold higher fluorescence than with acetylacetone when detecting substituted aldehyde analytes. The increased steric bulk of these sensitive reagents required that the reaction be incubated at 100 °C for 5 min to reach completion. Furthermore, deprotonation of the products using tetraethylammonium hydroxide resulted in a bathochromic shift of approximately 60 nm (λem = 460 nm, λex = 520 nm).
These assays were used to detect aliphatic and aromatic aldehydes in blood plasma or seawater. (337,338) Anthon and Barrett compared three assays used to detect formaldehyde, all based on the AO-catalyzed oxidation of methanol: the Nash reagent 132, Purpald 113, and MBTH 114. Based on absorbance measurements, the Nash reagent had the lowest sensitivity. (303)
Li et al. used acetoacetanilide (AAA, 135) for the detection of formaldehyde, addressing the problem of low reactivity with sterically demanding diketone reagents. Quantification of μM concentrations was possible at room temperature in 10 min in 4 M ammonium acetate buffer at pH 7.5. The conjugation was selective for formaldehyde; even acetaldehyde did not react with the probe. (339)

9.5. Oxime Formation and Fragmentation

Similar to hydrazine reagents, oxyamines react rapidly with carbonyl compounds, forming stable oximes even at low concentrations. (340,341) These conjugates are stable toward hydrolysis under physiological conditions and have been used for the chemoselective ligation of macromolecular fragments. (342,343) The group of Reymond observed the fragmentation of the oxime bond at pH > 11, similar to the Kemp reaction: (344) Base-induced β-elimination of the aldoximes forms a nitrile and a phenolate (Figure 43). Synthetic modification of known nitrophenol or umbelliferone chromophores with the reactive oxyamine moiety enabled the base-induced release of the dyes (136 and 137), (345) and thus spectroscopic or fluorometric measurement of aldehyde conjugation. The same reaction was catalyzed by BSA at neutral pH, where electron-rich aromatic aldehyde and ketone conjugates remained stable.

Figure 43

Figure 43. Summary of chromo- and fluorogenic turn-on probes for the selective detection of aldehydes based on the formation of unstable oximes. Fragmentation of these intermediates leads to the liberation of colored phenols. Abs: wavelengths typically used for UV absorbance measurements. Ex or em: wavelengths typically used for excitation and emission in fluorescence measurements. The wavelengths of probes in this section are shown in the strip at the bottom representing the visible spectrum.

The unspecific hydrolysis of the oxyamine still proceeded with a significant rate, making the detection of aldehydes problematic. The authors thus opted for a sequential approach: formation of the oxime was stopped by adding acetone in excess and any unreacted probe was thus converted into a stable ketoxime. Only then was the fragmentation of the aldoxime initiated by the addition of base.
This assay was used in screening for the release of formaldehyde by lipase-catalyzed hydrolysis of pivaloyloxymethyl ethers, but a low signal-to-noise ratio caused by the spontaneous hydrolysis of the oxyamine made the system impractical. A similar observation was made by the group of Roelfes during their investigations toward designer enzymes catalyzing hydrazone and oxime bond formation. (346) Contrary to the aforementioned report, base-induced fragmentation of their oxyamine-modified NBD–benzaldehyde adduct 138 caused the formation of the corresponding aldoxime and phenolate. The formation of p-nitrophenolate derivative NBD-O was used for the quantification of the catalytic efficiency of their mutants (Scheme 14). At 25 °C and pH 7.4 the system was able to discriminate the enzymatically catalyzed process against background hydrolysis of the unreacted probe.

Scheme 14

Scheme 14. Enzyme-Catalyzed Formation of the 138–Benzaldehyde Adduct Could Be Efficiently Traced by Spontaneous Decomposition Forming Colored NBD-O

9.6. Aza-Cope Reaction

A recently reported approach for the detection of formaldehyde is based on mechanistic investigations from 1950 by Horowitz and Geissman. (347) Their attempts to achieve double alkylation of allylamines with formaldehyde under acid catalysis remained unsuccessful and only formed cleavage products, which were recently shown to result from a 2-aza-Cope sigmatropic rearrangement (Figure 44). This finding led to the development of a diverse set of probes based on the same mechanism. Although this class of compounds has sparked growing interest over the last years, their reactivity limits applications to the detection of formaldehyde. Recent reviews summarize these and other approaches, including the detection method presented in this section. (357,358)

Figure 44

Figure 44. Summary of fluorogenic turn-on probes for the selective detection of formaldehyde based on the 2-aza-Cope sigmatropic rearrangement of initially formed Schiff’s bases. Fragmentation of these intermediates leads to the liberation of colored phenols. Ex or em: wavelengths typically used for excitation and emission in fluorescence measurements. The wavelengths of probes in this section are shown in the strip at the bottom representing the visible spectrum.

The first studies were both based on silicon rhodamine dyes but relied on different fluorogenic mechanisms. One approach built on the finding that spirocyclization renders rhodamine dyes weakly emissive at physiologically relevant pH. (348) Probe FAP-1 139 undergoes ring opening after an aza-Cope rearrangement. Subsequent hydrolysis of the resulting imine yields an aldehyde that can not re-(spiro-)cyclize. (349) This assay could detect formaldehyde at 100 μM concentration by an ∼8-fold increase in fluorescence (1 h, pH 7.4). The other approach uses the same rearrangement to liberate a p-nitrobenzyl group, which is known to quench fluorescence through a donor-excited PET process. (350) The assay using probe FP1 140 could detect formaldehyde at 250 μM concentration by a ∼7-fold increase in fluorescence (3 h, pH 7.4).
Building upon these initial reports, the group of Chang developed their assay toward a general 2-aza-Cope reaction trigger with a self-immolative β-elimination linker. They created the homoallylamine probe 141 that, in response to formaldehyde, unmasked fluorogenic phenols on a fluorophore. (351) By decoration of four different phenol-bearing fluorophores (coumarin, carbofluorescein, resorfurin, and rhodol), they achieved a 2–10-fold increase in fluorescence upon formaldehyde conjugation at 100 μM concentration (2 h, pH 7.4, 25 °C).
Du et al. recently investigated a broad set of N-substituted homoallylamines 142 as probes for formaldehyde. (352)N-p-methoxybenzyl homoallylamine was found, by experiment and computation, to have the highest reactivity in the 2-aza-Cope reaction with formaldehyde. Using similar self-immolative linkers as above, they synthesized five silenced phenol-bearing fluorophores 142, with excitation wavelengths ranging from 445 to 706 nm. Exposure to formaldehyde at 500 μM concentrations under physiological conditions prompted a turn-on factor of 5–125. All probes were successfully applied in the visualization of changes of intracellular formaldehyde in living cells (Figure 45). (352)

Figure 45

Figure 45. Confocal fluorescence images of exogenous formaldehyde (FA) with FFP fluorescent probes 142 in HEK293T cells. (A–D) Cells were loaded with (a) FFP511 (Tokyo Green fluorophore) (20 μM), (b) FFP551 (142) (10 μM), (c) FFP585 (resorufin fluorophore) (5 μM), or FFP706 (hemicyanine fluorophore) (2 μM) for 30 min and treated with FA of varying concentrations (0–1 mM) for 60 min before being imaged by confocal fluorescence microscopy. Scale bars represent 50 μm. Reproduced with permission from ref (352). Copyright 2021 The Royal Society of Chemistry.

Making these reagents typically requires great synthetic effort, which is impractical for biochemists. Consequently, many probes based on simple synthetic modifications of known fluorophores have been developed in recent years. For instance, He et al. and Xu et al. reported the naphthalene-based probe RFFP/AENO 143, which can be synthesized in a one-pot reaction from commercially available 6-hydroxy-2-naphthaldehyde. (353,354) The use of the reagent as turn-on probe with a 100-fold enhancement of fluorescence within 2.5 h for 1 mM formaldehyde was reported (Figure 46). (354)

Figure 46

Figure 46. Fluorescent responses of 143 (10 μM) in DMF/PBS solution (v/v = 1/4, pH 7.4, 10 mM) upon addition of various small molecular species (5.0 mM). λexem = 390/513 nm; slits, 2.5/2.5 nm. Each spectrum was recorded after 3 h at 37 °C. Error bars are ±1SD, n = 3. Inset: (a) The color and (b) fluorescence images of 143 (10 μM) in the presence of various small molecular species (5.0 mM). (354) Reproduced with permission from ref (354). Copyright 2016 Elsevier BV.

The groups of Lin and Peng reported the synthesis of turn-on probes PDB-FA 144a and BD-CHO 144b, based on NBD as a fluorophore, (355,356) in a three-step procedure from commercial 4-chlorobenzoxadiazole. The probes showed 55-fold (PDB-FA) and 255-fold enhancement (BD-CHO) upon exposure to excess formaldehyde after 2 h at pH 7.4. (357,358)

10. Chromo- and Fluorogenic Probes for Ketones

Click to copy section linkSection link copied!

Ketones are widespread in nature. This electrophilic functional group is present, for instance, in the carbohydrate family of ketoses, as a polar element in important steroids (e.g., testosterone, progesterone), in fragrance compounds derived from terpenoid units, (359) or as the structurally important methyl ketone motif (360) (e.g., camphor, carvone, pulegone, 2-heptanone). Ketones are also found as biological intermediates and energy storage units in important regulatory systems of cells (Krebs cycle, (361) fatty acid synthesis, (362) ketogenesis (363)).

Figure 47

Figure 47. (top) Summary of general properties of ketones, reactivity trends toward ketone-selective probes as well as expected cross-reactivity during screenings. (bottom) Compatible pH range and buffers in reported ketone-selective assays. Bold pH numbers indicate the most commonly used pH values.

In industry, ketones are widely used in bulk as solvents (e.g., acetone or methyl ethyl ketone, MEK) or as synthetic intermediates. They have found broad use in biocatalysis as starting materials for the synthesis of chiral alcohol and amines (364) using transaminases (TA), alcohol dehydrogenases (ADH), (365) Baeyer–Villiger monooxygenases (BVMO), (366) or aldolases. (367)
They can be synthesized biochemically from β-ketoacids by decarboxylation using carboxy lyases, (368) using aldolases, (369) from carbinols using TDP-dependent lyases, (370) or using acyl transferases (ATase). (371) These biocatalysts have already been demonstrated as good alternatives in organic synthesis. The group of Hammer recently optimized a ketone synthase for the regioselective anti-Markovnikov oxidation of aromatic alkenes. Starting from a P450 wild-type, 12 rounds of evolution identified 18 mutations that were beneficial for the stereoselective oxidation of styrene derivatives 145 (Scheme 15a). (372)

Scheme 15

Scheme 15. Representative Selection of Enzymatic Transformations for the Production of Carboxylic Acid-Containing Products
Many applications of ketone synthesis use alcohol dehydrogenases as mild oxidative catalysts. Zhang et al. applied the highly regio- and enantioselective ADH from Bacillus subtilis BDHA for the oxidation of racemic trans-cyclic vicinal diols 147 (Scheme 15b) (373) to (R)-α-hydroxy ketones 148 and (S,S)-cyclic diols 149 in >99%ee at 50% conversion in one pot.
The group of Kroutil identified an ADH from Sphingobium yanoikuyae for the opposite purpose, oxidizing the aliphatic alcohols 150 to the corresponding ketones 151 without stereopreference and without the need for a huge excess of ketones as hydrogen acceptors for cofactor recycling (Scheme 15c). (374) The researchers found that α-Cl, -F, or -MeO substituted ketones were not oxidized, hence limiting the necessity for organic recycling reagents to one molar equivalent only.
González-Granda et al. reported the oxidation of propargylic alcohols 152 using the catalytic system composed of a laccase from Trametes versicolor, the oxy-radical (2,2,6,6-tetramethylpiperidin-1-yl)oxyl (TEMPO), and oxygen as terminal oxidant. (375) Under these mild conditions, the products were obtained in good yields (Scheme 15d).
The group of Kroutil was the first to report an enzymatic Friedel–Crafts acylation of activated arenes 154, a reaction that is typically in the realm of classical organic chemistry, as well as the first biocatalytic equivalent to the Fries rearrangement (the intramolecular analogue of the Friedel–Crafts reaction). (371) They reported conversions of up to 95%, using readily available isopropenyl acetate as acetyl donor and without any CoA-activated reagents (Scheme 15e).
Ketones are much less reactive than most aldehydes due to an increase in steric bulk and the diminished electrophilicity caused by the electron-donating effect of the attached alkyl groups (Figure 47). This relatively low activity makes specific derivatization of ketones challenging; reactivity and selectivity of reagents are inversely correlated. With certain exceptions, approaches are similar to those introduced in the section on aldehydes (cf. section 9).
Ketones react slowly in conjugations under physiological conditions, (293) particularly at low concentration. Most assays require harsh reaction conditions that are incompatible with biological systems. Only a few studies have addressed the challenge of developing selective conjugations for ketones, so far, likely for the following reasons: (i) Aldehydes are generally more important in synthesis due to their broader versatility as chemical building blocks, (ii) It is difficult to match the rates of aldehyde conjugations (under both abiological and physiological conditions). Ketones are mostly used as negative examples to highlight the specificity of carbonyl probes for aldehyde conjugations.

10.1. Hydrazone Formation

The most promising probes to detect ketones use the fast reaction of carbonyl compounds with hydrazines forming stable hydrazones (Figure 48). In principle, the approach is the same as for aldehydes, summarized above (cf. section 9.1), but only a small number of assays showed useful reactivity toward ketones. The long-known reagent DNPH 109 has been used by Yang et al. as an at-line analytical tool and in HTS for the activity of mandelate racemase variants by detection of oxidized α-hydroxy acids. (287) Enantioselective mandelate dehydrogenase D-MDH was used to selectively oxidize the desired enantiomer, which was converted to the hydrazone using 109 and detected at 450 nm after basification with aqueous NaOH (Figure 49).

Figure 48

Figure 48. Summary of chromo- and fluorogenic turn-on probes for the selective detection of ketones based on the formation of hydrazone adducts. Abs: wavelengths typically used for UV absorbance measurements. Ex or em: wavelengths typically used for excitation and emission in fluorescence measurements. The wavelengths of probes in this section are shown in the strip at the bottom representing the visible spectrum.

Figure 49

Figure 49. Variation in color of reaction mixtures containing different concentrations of 109 and the validation of the developed HTS method by chiral HPLC analysis. Control, no 109 was added into the reaction mixture; 1–10×, different concentrations of 109 were added into the reaction mixture (1×, 8.4 mg/mL). Reproduced with permission from ref (287). Copyright 2017 Springer-Verlag.

The BD-probes (NBD-H 110, ABD-H 111. DBD-H 112) form fluorescent derivatives with aliphatic and aromatic ketones at room temperature, but only in ACN. (288) Among these compounds, 110 is the most broadly applied for conjugations with aldehydes and ketones that work even under physiological conditions. (291,376) Crisalli and Kool used anilines to catalyze conjugations at neutral pH. Reactions with aromatic ketones were improved using these catalysts, but they still proceeded slower than reactions with aldehydes. (294,377) Wang et al. showed that increased salt concentrations were beneficial for conjugation rates at physiological pH. (295)
Among the already described fluorogenic hydrazines (cf. section 9.1), methyl 5-MeO-N-aminoanthranilate 119 was tested as a probe for aliphatic ketones and showed a small (unquantified) increase in fluorescence emission upon conjugation. (312)

10.2. Aldol Reaction

When ketones tautomerize to enols, they can participate in aldol reactions (Figure 50). Similar to the hemiaminal formation used for the detection of amines (cf.section 6.7), electron-withdrawing carbonyl groups can be used to trigger a hypsochromic shift of the absorption maximum upon aldol reactions. This method has been explored by the group of Barbas and Tanaka. (378−380) They investigated different designs (compounds 156158) to detect acetone, which required activation via enamines formation using catalytic amounts of pyrrolidine, arginine, glycine, proline, or the aldolase antibody 38C2. A 1000-fold excess of the substrate was required to afford useful conversions. The triazine aldehyde 156 and the phenanthrene aldehyde 157 showed similar excitation and emission properties (approximately 100–1000-fold turn-on) from pH 5 to 9. Exchange of the linked amide aryl-moieties to the benzimidazole analogue 158 red-shifted the excitation maximum from 260 to 315 nm. In turn, it also lowered the sensitivity of the assay (3–20-fold turn-on) due to the parent aldehydes becoming stronger emitters.

Figure 50

Figure 50. Summary of fluorogenic turn-on probes for the selective detection of acetone based on aldol reactions. Ex or em: wavelengths typically used for excitation and emission in fluorescence measurements. The wavelengths of probes in this section are shown in the strip at the bottom representing the visible spectrum.

10.3. Michael Addition

The activation via enamine formation can be also used with Michael acceptor probes in 1,4-additions. The benzimidazole 159 and 6-aminoquinoline 160 have useful turn-on properties upon conjugation to ketones (Figure 51). (379) Exposition of fluorogenic 6-aminoquinoline probe 160 to acetone additionally triggers an immediate cyclization by aldol-condensation with the terminal methyl ketone (Scheme 16). Due to a strong spectral overlap between probe and conjugate, their application is problematic. Both probes are slower than the previously reported aldol sensors (cf. section 10.2) and also require a 1000-fold excess of acetone. Due to the unattractive properties of the screening system, no additional studies were undertaken.

Figure 51

Figure 51. Summary of fluorogenic turn-on probes for the selective detection of acetone based on a Michael addition approach. Ex or em: wavelengths typically used for excitation and emission in fluorescence measurements. The wavelengths of probes in this section are shown in the strip at the bottom representing the visible spectrum.

Scheme 16

Scheme 16. Michael Addition of Acetone onto 159 Triggers a Subsequent Aldol Reaction Forming Cyclic Hydroxyketones

10.4. Other Methods

Two additional methods for the detection of ketones are (i) the Janovsky reaction, and (ii) reaction with ABAO 126 (Figure 52).

Figure 52

Figure 52. (top) Chromogenic and fluorogenic detection of ketones bearing an enolizable position via the formation of the purple-colored Janovsky complexes with 3,5-dinitrobenzoic acid 161 under strongly basic conditions. (bottom) Fluorogenic detection of ketones via the formation of fluorescent ABAO-adducts (cf. section 9.2). A methoxy substitution of the ABAO core 126 is necessary to increase the reactivity of the aniline NH2 position.

The Janovsky reaction requires analytes to form reactive enolates under strongly basic conditions, which then form stable, purple-colored σ-complexes with 3,5-dinitrobenzoic acid 161. This method was used to rapidly evaluate 11 substrates for Baeyer–Villiger monooxygenases by monitoring the decrease of absorbance at 550 nm. (381)
Reaction with ABAO, initially suspected to be unreactive toward ketones, emerged as an effective probe for aliphatic and aromatic ketones. (322) The methoxy analogue MeO-126 reduced long reaction times with its reactivity enhanced by the electron-donating effect of the methoxy group. A set of 24 ketones was tested at pH 5.0 using an equimolar amount of ABAO at 10 mM concentration. Fluorometric evaluation of the reaction after 30 min revealed a ∼2–20-fold turn-on. The assay was applied in enzyme mining and directed evolution of alcohol dehydrogenases (ADHs) for the catalytic oxidation of rhododendrol to raspberry ketone (p-hydroxybenzyl acetone).

11. Chromo- and Fluorogenic Probes for Carboxylic Acids

Click to copy section linkSection link copied!

Carboxylic acids are found in many compounds in nature, (382) pharmaceuticals, (383) and food constituents. (384) Many structures bearing this functional group are involved in essential physiological processes, such as prostaglandins regulating inflammation and blood pressure (385) or citric acid and succinic acid in the Krebs cycle. (386) Fatty acids are the building blocks of lipids and are involved in energy metabolism, (387) whereas α-amino acids are the building blocks for peptides and proteins. (388)

Figure 53

Figure 53. (top) Summary of general properties of carboxylic acid sensors as well as expected cross-reactivity during screenings. (bottom) Compatible pH range and buffers in reported carboxylic acid-selective assays. Bold pH numbers indicate the most commonly used pH values.

Various biocatalytic methods for the synthesis of carboxylic acids have been developed, complementing nonenzymatic approaches. (389) These methods include the hydrolysis of nitriles by nitrilases, (390−392) the conversion of aldehydes by hydroxynitrile lyases with subsequent hydrolysis, (393) the hydrolysis of esters by esterases, (394) and the oxidation of alcohols or aldehydes by dehydrogenases and oxidases. (395)
Glieder et al. reported a hydroxynitrile lyase mutant from Prunus amygdalus for the enantioselective synthesis of α-hydroxyacids by the addition of HCN to aldehydes and the subsequent hydrolysis of the intermediately formed cyanohydrins. The enzyme was stable under acidic conditions and enabled the conversion of 2-chlorobenzaldehyde 162 to (R)-2-(2-chlorophenyl)-2-hydroxyacetonitrile 163 at pH 3.4 with 96% yield and 97%ee. Hydrolysis of the intermediate yielded the enantiopure α-hydroxyacid 164 (Scheme 17a). (393)

Scheme 17

Scheme 17. Representative Selection of Enzymatic Transformations for the Production of Carboxylic Acid-Containing Products
DeSantis et al. identified a nitrilase mutant for production of (R)-4-cyano-3-hydroxybutyric acid 166, an intermediate in the synthesis of the drug Lipitor, by directed evolution. The hydroxybutyric acid 166 could be obtained in 96% yield with 98.5%ee and a volumetric productivity of 619 g L–1d–1 (Scheme 17b). (396)
The reactivity of carboxylic acids is mainly determined by their acidic hydroxy group and carbonyl functionality. Under physiological and basic conditions, carboxylic acids (pKa ∼ 3–5 (397,398)) are deprotonated, negatively charged carboxylates, which render nucleophilic attacks at the carbonyl center impossible (Figure 53). (399) Also, protonated carboxylic acids are much less reactive than derivatives like acyl halides or anhydrides, which generally makes an activation for derivatizations necessary. Therefore, detection methods based on the conversion of an acid to a fluorescent derivative (e.g., hydrazide, hydroxamate) usually involve activation as the first step (e.g., with dicyclohexylcarbodiimide DCC, 167). Many approaches for carboxylic acids involve monitoring of pH shifts by pH indicators (404) and the enzymatic conversion of acids with NADH. (405)
Other methods exploit the properties of carboxylic acids and carboxylates to form intermolecular hydrogen bonds, which enable their detection as adducts with probes that exhibit hydrogen bond donating groups, such as amides, thioureas, or guanidinium moieties. (401,402) Some carboxylate probes are based on the coordination of the anion to a metal ion, e.g., Zn2+ or Cu2+. (403)
A challenge in sensing carboxylic acid is the selective detection and the differentiation of acids. (401,406,407) One method to improve selectivity is introducing multiple recognition sites at the probe. (403,407) Many of the assays require nonaqueous conditions or apolar/aprotic solvents (e.g., chloroform) to prevent competitive interactions between the solvent and the binding sites of the probe. (408−411)

11.1. Activation and Nucleophilic Trapping

Carboxylic acids can be detected by converting them to acid hydrazides with hydrazine derivatives (Figure 54). (412,413) Miwa et al. reported a colorimetric method using 2-nitrophenylhydrazine 168 in ethanolic solution in the presence of 167 as a coupling agent and catalytic amounts of pyridine. The resulting acid hydrazides had a violet color upon basification with KOH, and their absorbance could be measured at about 550 nm. (414)

Figure 54

Figure 54. Summary of chromogenic turn-on probes for the selective detection of carboxylic acids based on the entrapment of activated carboxylic acid species. Abs: wavelengths typically used for UV absorbance measurements. The wavelengths of probes in this section are shown in the strip at the bottom representing the visible spectrum.

Another colorimetric method uses derivatization of activated carboxylic acids to hydroxamic acids with hydroxylamine 169, followed by the formation of a colored ferric hydroxamate (Figure 54). (415,416) Based on work by Kasai et al. (417) and Takeuchi et al., (418) He et al. reported a spectrophotometric HTS assay to detect various aliphatic and aromatic carboxylic acids in aqueous solutions at room temperature by converting them to ferric hydroxamates. The carboxylic acids were converted to hydroxamic acids with hydroxylamine perchlorate 169·ClO4 in the presence of 167. Purple ferric hydroxamates were formed upon the addition of an acidic solution of ferric perchlorate. The absorbance of the ferric hydroxamate was measured at 520 nm. He et al. applied this method to assay the nitrile-hydrolyzing enzymes Rhodococcus erythropolis CGMCC 1.2362 and Alcaligenes sp. ECU0401. (419)

Figure 55

Figure 55. Summary of fluorogenic turn-on probes for the selective detection of carboxylic acids based on the coordination to metal centers or hydrogen bond donors. Ex or em: wavelengths typically used for excitation and emission in fluorescence measurements. The wavelengths of probes in this section are shown in the strip at the bottom representing the visible spectrum.

11.2. Coordination to Metals and Hydrogen Bond Donors

Formation of complexes with a probe via hydrogen bonding or electrostatic interactions can be used to detect carboxylic acids and carboxylates (Figure 55). (403,420−422) These interactions can induce fluorescence by electronic changes of the probe: Nonbonding electron pairs of amines or adjacent electron-donating groups are known to quench fluorescence by PET processes, which can be restored by coordination of the amines to carboxylic acids. (408,422) Similarly, quenching has been observed by binding of amine probes to Cu2+ ions. By coordination and thus scavenging of the Cu2+ ions by carboxylates, the probe’s fluorescence emission can be effectively restored. (411,423) In some cases, hydrogen bonding or electrostatic interactions of carboxylate and probe lead to a fluorescence response by the displacement of a noncovalently bound fluorescent indicator from the probe (indicator-receptor-complex) by the carboxylate which results in the release of the fluorescent dye. (406,409)

Figure 56

Figure 56. Summary of fluorogenic turn-on probes for the selective detection of carboxylic acids based on nucleophilic addition onto trifluoroacetyl groups. Ex or em: wavelengths typically used for excitation and emission in fluorescence measurements. The wavelengths of probes in this section are shown in the strip at the bottom representing the visible spectrum.

Cabell et al. developed probe 170 for the detection of citrate in HEPES buffer (pH 7.4), enabling the quantification of citrate in beverages at μM concentration. Probe 170 uses a 1,10-phenanthroline moiety as a fluorophore and guanidinium moieties as recognition sites ensuring effective hydrogen bonding to carboxylates in water. Fluorescence is quenched while phenanthroline binds Cu2+. When tribasic citrate binds to Cu2+ and the two guanidinium groups of the probe, electron transfer from the metal to the fluorophore changes as the metal becomes more electron-rich, which leads to an increase in fluorescence at 365 nm. (423)
Xu et al. developed the Eu3+ complex 171, based on a phenanthroline scaffold with chiral amino alcohol groups as ligands that enabled the selective detection of malate anions in aqueous solutions (0.01 M HEPES, pH 7.4, with 0.3% DMSO). Binding of Eu3+ suppresses the fluorescence at about 380 nm, which gets restored when malate forms a complex with the ligand. Other carboxylates, such as alanine, phenylalanine, mandelate, and tartrate, did not show significant responses. The authors assumed that the selectivity for malate arose from the preorganized structure of 171 and the complementary shapes of probe and anion. The enantiomers of malate gave different responses: 487% enhancement with l-malate and 332% with d-malate. (424)
Burguete et al. reported a detection method for citrate based on the fluorescent anthracene-probe 172 in methanolic aqueous solution (70%, pH 7.5–8.0). Their probe consisted of a nonfluorescent Cu2+ complex of two aminoanthracene moieties, each coupled with an l-phenylalanine, and enabled the detection of citrate in the μM range. The addition of citrate led to a displacement of the anthracene moieties, which resulted in a more than 15-fold increase in fluorescence at 420 nm. Mandelate, tartrate, and malate gave small increases in fluorescence (∼10–15%) compared to citrate; succinate, fumarate, lactate, acetate, and glutarate did not produce any increase. Only l-glutamate, which is similar to citrate in the number of potential coordination sites (two carboxylates and one amino group), showed a similar increase in fluorescence (40%). (425)
Boiocchi et al. developed a probe based on the indicator-displacement approach to detect dicarboxylates in water. The binding of the cryptate complex 175 to the fluorescent indicator carboxy-rhodamine 174via the coordination of the two Cu2+ ions to two carboxylates led to the quenching of the indicator’s (174) fluorescence. A displacement of 174 by selected dicarboxylates restored fluorescence at λem, max = 570 nm. Only 1,4-, 1,5-, and 1,6-dicarboxylates terephthalate, glutarate, and adipate that all have an appropriate spacial distance between their COO groups (irrespective of their carbon chain length) could coordinate to both Cu2+ ions and displace the indicator. Dicarboxylates with shorter or longer distances (e.g., succinate, pimelate, phthalate, or isophthalate) led to no or only minor fluorescence emission. In the case of the very small oxalate anion, a coordination of two molecules to bind both Cu2+ also resulted in the displacement of the indicator. (409)
Amino acids are known to form stable complexes with Cu2+ ions. (406) Klein and Reymond reported the Cu2+-based probe 176 that was used in HTS of the hydrolysis of N-acetyl-l-methionine by acylase I and of l-leucinamide by aminopeptidase. The complex of the quinacridone-derived ligand and Cu2+, gave no fluorescent signal. Due to the weak chelating effect of this macrocycle, amino acids competed with the ligand and displaced it. The fluorescence of the released quinacridone-derived ligand was measured at 584 nm. The low solubility of 176 in aqueous buffer required the use of organic cosolvent (∼40%), lowering enzymatic activity; alternatively, the probe and solvent could be added after the enzymatic reaction has completed. (426)
In later reports, substitution of the quinacridone-derived ligand with the commercial fluorescein-based ligand calcein 177 was proven to be beneficial due to its increased solubility in aqueous buffers. The probe showed enhanced fluorescence at 530 nm with various amino acids, such as valine, alanine, methionine, leucine, serine, histidine, N-acetylmethionine, and glycine, at μM concentrations. The assay worked at pH 6–10 in various buffers (bis-tris, phosphate, borate, Tris) and mixed solvent systems (0–50% organic), including EtOH, MeOH, ACN, and DMSO. Strongly coordinating buffers (e.g., bis-tris-propane) are incompatible because they bind to metal ions. The assay does not work at low pH as the protonated carboxyl groups of 177 do not coordinate well. Probe 177 enabled HTS for activity of proteases, such as trypsin, chymotrypsin, and subtilisin. (427) Xu et al. validated this approach and used the Cu2+–calcein complex for HTS of more than 2500 leucine dehydrogenase mutants with 2-oxobutyric acid as substrate. They detected 11 mutants with a higher activity than the wild type. (428)

11.3. Nucleophilic Addition onto Trifluoroacetyl Groups

Carboxylates can react with trifluoroacetyl ketones in a nucleophilic addition (Figure 56). As in sections 7.1 and 6.7, addition disrupts electron-withdrawing effect of the carbonyl group on conjugated systems and leads to changes in fluorescence. (429,430) Kim and Ahn developed trifluoroacetophenone derivatives 178 with a terthiophene or a pentathiophene moiety as fluorophores and a trifluoroacetyl group as recognition site for carboxylates. Fluorescence of the terthiophene probe rose 120-fold in the presence of acetate in ACN. Other anions, such as phosphate, fluoride, and cyanide, also enhanced the fluorescence. (430)
The trifluoroacetyl carbonyl moiety can also function as a recognition site for amines (cf. section 6.7), as shown by Ryu et al. in the development of probe 179, which uses two trifluoroacetyl groups for the detection of α-amino carboxylates. Binding of the amine and the carboxylate group to both recognition sites resulted in the formation of a bridged adduct accompanied by an increase in fluorescence. The addition of glycinate resulted in a 110-fold enhancement of fluorescence in ACN. Other α-amino carboxylates caused similar emissions; the α-substituents did not affect the formation of the adduct. Probe 179 discriminated α-amino carboxylates against β- and γ-carboxylates. (429)

12. Chromo- and Fluorogenic Probes for Carboxylic Acid Esters

Click to copy section linkSection link copied!

Carboxylic acid esters are widely found in nature, (431) pharmaceuticals, (432,433) fine chemicals, (434) agrochemicals, (435) and the food industry. (436) Most are long-chained triglycerides in fats and oils and are of central importance in energy metabolism. (387,437) Short-chained derivatives are typically found in fruit and flower fragrances (e.g., benzyl acetate in jasmine (438) or isoamyl acetate in banana (439)). Due to their flavor-carrying and fragrant nature, they are valuable ingredients in cosmetics, food, and beverages. (440) Moreover, they are important building blocks for polymers, (441) lubricants, (442) plasticizers, (443) and surfactants. (444) Enzymatic approaches for the synthesis of carboxylic acid esters include esterifications and transesterifications catalyzed by esterases and lipases. (445−447)

Figure 57

Figure 57. (top) Summary of general properties of carboxylic acid ester sensors as well as expected cross-reactivity during screenings. Due to the lack of direct reactive probes applicable in aqueous environments, no compatible buffers and pH ranges can be stated.

Lozano et al. reported a biocatalytic approach for the synthesis of aliphatic esters 182 by esterification of C1–C4 carboxylic acids 180 with fragrant alcohols 181 (isoamyl alcohol, citronellol, geraniol, and nerol) using an immobilized lipase from Candida antarctica (Novozym 435) in the ionic liquid N,N′,N″,N′′′-hexadecyltrimethylammonium bis(trifluoromethylsulfonyl)imide (Scheme 18). The authors optimized the workup procedure and achieved yields >78% for all esters. Geranyl valerate was obtained with the highest product concentration of 0.757 g mL–1. (448)

Scheme 18

Scheme 18. Representative Selection of Enzymatic Transformations for the Production of Carboxylic Acid Ester-Containing Products
Carboxylic esters are compounds derived from carboxylic acids and alcohols. The electrophilic carbonyl center can react by substitution of the alcohol group with many nucleophiles, such as amines, alkoxylates, and hydroxylamines. Under basic or acidic conditions, esters can also hydrolyze. (400,449) The volatility of many carboxylic acid esters must be considered in quantitative measurements (450) (Figure 57).
The only direct colorimetric approach for the detection of carboxylic acid esters is based on the conversion to hydroxamic acids with hydroxylamine hydrochloride 169 in the presence of a base (Figure 58). (451−453) The subsequent addition of ferric chloride under acidic conditions results in the formation of a ferric hydroxamate complex, which has a typical magenta color. (451,452) We described a similar principle for the detection of carboxylic acids (see above).

Figure 58

Figure 58. Chromogenic detection of carboxylic acid esters via the formation of the hydroxamic acids and subsequent complexation with Fe3+ resulting in purple-colored ferric hydroxamates.

Esters are inherently reactive toward hydroxylamine (cf. section 11) and do not need activation. Buckles and Thelen tested the scope and limitations of this approach with various aromatic and aliphatic esters. Besides simple carboxylic esters, also polymer esters and glycerides of fatty acids were detected. Carbonates, chloroformates, sulfonates, urethanes, and esters of inorganic acids gave negative results. Coloration was also observed in the presence of other electrophiles, including aldehydes, trihalomethyl compounds, imides, nitriles, and amides. (451) Although only certain amides were shown to give a positive colorimetric response, an extraction step might be useful before analysis to reduce false-positive hits from endogenous proteins.
Based on the same concept, Löbs et al. developed a method for the rapid screening of ethyl acetate production by four different strains of Kluyveromyces marxianus growing on various C5, C6, and C12 carbon sources. The formed esters were extracted with hexane and subsequently converted to ferric hydroxamates at room temperature. The absorbance was measured at 520 nm. All strains exhibited the highest ethyl acetate production employing the C6 sugars glucose and fructose. The assay was also validated for ethyl butyrate, ethyl hexanoate, isoamyl acetate, and ethyl octanoate. (454)

13. Chromo- and Fluorogenic Probes for Epoxides

Click to copy section linkSection link copied!

Epoxides react with various nucleophiles, oxidizing, and reducing agents due to its high polarity and strain of the three-membered ring. Nucleophilic ring-opening reactions, leading to 1,2-disubstituted products, are of particular synthetic interest. (455,456) Possible products include: amino alcohols, glycol ethers, alcohols, hydroxynitriles, and chlorohydrins. (457)

Figure 59

Figure 59. (top) Summary of general properties of epoxide sensors as well as expected cross-reactivity during screenings. (bottom) Compatible pH range and buffers in reported epoxide-selective assays. Bold pH numbers indicate the most commonly used pH values.

Epoxides are used as industrial intermediates in the production of pharmaceuticals, (458−461) soaps, (460) lubricants, (459) cosmetics, (459,460) and surfactants. (459) Enantiopure epoxides are particularly valuable chiral functional groups for the production of more complex compounds. (462−464) Simple epoxides, such as ethylene oxide or propylene oxide, are of interest for the production of ethylene glycols, ethanolamines, ethylene glycol ethers, and propylene glycols, polyether polyols, and propylene glycol ethers. (465,466) Enzymatic methods for epoxide synthesis include the oxidation of alkenes by monooxygenases (MO), the conversion of halohydrins by halohydrin epoxidases, and the use of epoxide hydrolases. (214,467)
Halohydrin dehalogenases and epoxidases catalyze the reversible enantioselective dehalogenation of vicinal halohydrins 183 to the corresponding epoxides. Ma et al. employed halohydrin dehalogenase (HHDH, EC 3.8.1.5) in the second step of a two-step biocatalytic process to synthesize hydroxynitriles, which were used as intermediates in an environmentally friendly synthesis of atorvastatin. They exploited the promiscuitiy of halohydrin dehalogenases, accepting cyanide as a nucleophile: after formation of epoxide 184, cyanide readily opened the ring and gave the corresponding β-hydroxynitrile 185 (Scheme 19a). Iterations of DNA shuffling increased the enzyme activity >2500-fold over the wild type, and ethyl (R)-4-cyano-3-hydroxybutanoate 185 was obtained in 92% yield. (468)

Scheme 19

Scheme 19. Representative Selection of Enzymatic Transformations for the Production of Epoxide-Containing Products
Kong et al. created a library of variants of the epoxide hydrolase from Bacillus megaterium (BmEH) to expand the substrate scope from phenyl glycidyl ether toward bulkier epoxides. By mutation of two active site residues, variants with an improved activity (6–340-fold) for nine β-blocker precursors were discovered. Resolution of the racemic epoxide precursors 186 yielded the (S)-epoxides 187 in 37–45% yield and 96.6–99.5%ee (Scheme 19b). (469)
Lin et al. reported a method for the oxidative dearomatization of indoles yielding asymmetric furoindolines using the flavin-dependent monooxygenase TsrE. The enzyme catalyzed the 2,3-epoxidation of 2-methyl-3-indole-3-acetic acid 188 and the subsequent epoxide opening of 189 in the synthesis of (3aS,8aR)-3a-hydroxy-8a-methyl oxofuroindoline 190, which was obtained in 84% yield and 98%ee (Scheme 19c). (470)
Assays for epoxides are usually based on nucleophilic ring-opening reactions, and detection often happens indirectly. They are mostly based on the conversion of epoxides to aldehydes, thiols, or diols and the application of aldehyde-, thiol-, or diol- detection methods. Detection might be complicated by instability (e.g., to hydrolysis), side reactions, (464,471) and volatility. (456,465,472) Epoxides can undergo reactions with hemoglobin or nucleic acids and are therefore associated with mutagenicity, carcinogenicity, and genotoxicity (473) (Figure 59).

13.1. Nucleophilic Ring-Opening by Pyridine Derivatives

Various epoxide assays are based on ring-opening reactions by nucleophilic pyridine derivates, resulting in colorimetric or fluorometric signals (Figure 60). The commercially available reagent 4-(p-nitrobenzyl)pyridine (NBP, 191) is often applied to detect epoxides. (459,461,473−476) The nucleophilic substitution reaction of 191 with epoxides results in an uncolored intermediate, which can subsequently be converted to the blue-colored dihydropyridine using base. Changes in absorption in the range of 560–620 nm have been used for read-outs. (458,459,473,475) Zocher et al. developed a HTS assay with 191 for microtiter plates to screen for activity of an epoxide hydrolase. (458) Alcalde, Farinas, and Arnold also used an NBP-based HTS assay for directed evolution to improve styrene epoxidation by cytochrome P450 BM-3 139-3. (459) Probe 191 can also be used to detect organophosphorus compounds, ethylenimines, and other alkylating reagents, and side reactions with these compound classes might interfere with the detection of the epoxides. (461)

Figure 60

Figure 60. Summary of chromo- and fluorogenic turn-on probes for the selective detection of epoxides based on ring-opening reactions by nucleophilic pyridine derivatives. Abs: wavelengths typically used for UV absorbance measurements. Ex or em: wavelengths typically used for excitation and emission in fluorescence measurements. The wavelengths of probes in this section are shown in the strip at the bottom representing the visible spectrum.

Because 191 is insoluble in water, (473) Nelis and Sinsheimer described a similar fluorometric detection method with high sensitivity and solubility in aqueous systems. Their approach was based on the alkylation of nicotinamide 192 by aliphatic epoxides (Scheme 20). The alkylated amides 194 were converted to dihydropyridine derivatives 195 with acetophenone 193 under basic conditions, followed by a cyclization triggered by formic acid, which resulted in the formation of fluorescent products 196. This method is not specific for epoxides; it detects various electrophilic compounds. Fluorescence of NADP+, itself an alkylated nicotinamide, might interfere, too. (474)

Scheme 20

Scheme 20. Sensing Mechanism for Epoxides Employing the Combination of Nicotinamide 192 and Acetophenone 193

13.2. Indirect Detection

In many cases, the detection of epoxides is done indirectly by conversion to other functional groups and subsequently detecting those (Figure 61). Sano and Takitani developed a two-stage approach to detect epoxides in ethanolic aqueous solutions. The epoxides were converted to the respective S-alkylated derivatives by ring-opening with sodium sulfide. The resulting thiolates could be efficiently labeled by using a system known for fluorogenic isoindole formation employing o-phthalaldehyde 22 and taurine (cf. section 6.3). (477)

Figure 61

Figure 61. Selective detection of epoxides based on indirect detection methods after initial nucleophilic ring-opening reactions with hydroxide or bisulfide. The resulting intermediates can be detected with thiol sensing assays (cf. section 15), diol sensing assays (cf. section 8), or by using a similar method discussed for the indirect detection of diols by oxidative cleavage with metaperiodate 97 and subsequent detection using aldehyde-sensing assays (cf. section 9).

Another approach for the indirect detection of epoxides is based on their conversion to diols and the subsequent application of diol-sensitive assays. Doderer et al. described a colorimetric HTA that could be applied for all epoxides with at least one hydrogen substituent. The epoxides were converted to the diols using epoxide hydrolase, which were subsequently cleaved by periodate 97 to the corresponding aldehydes and/or ketones. The formed aldehydes could then be detected by Schiff’s reagent 123. The absorption of the resulting magenta dye was measured at 560 nm. (262) Doderer and Schmid reported another concept for diols based on oxidative cleavage with periodate to the corresponding carbonyls. Titration of the residual periodate 97 with carboxyfluorescein resulted in a detectable decrease in fluorescence. (268) Some of the carbonyl species resulting from diol cleavage already have a strong UV absorbance, allowing their direct detection. Mateo et al. used this property for the determination of styrene oxide by detecting benzaldehyde at 290 nm after hydrolysis and cleavage. They used this method for the fast determination of activity of epoxide hydrolase. (481)
Halohydrin dehalogenases are commonly used as enzymatic catalysts to obtain epoxides from aliphatic or aromatic halohydrins. This and the reverse reaction is accompanied by a change in pH, which enables the indirect detection of epoxides using pH indicators, such as phenol red or bromothymol blue. (478,479) Other assays make use of the direct change in UV absorption or fluorescence upon the conversion of epoxides to their corresponding diols. (472,480) Wixtrom and Hammock reported two such spectrophotometric assays. (472)

14. Chromo- and Fluorogenic Probes for Phenols

Click to copy section linkSection link copied!

Phenols and their derivatives are an important structural motif ubiquitous in nature in the form of catechols, quinones, lignans, flavones, or in polymeric form in lignin, flavolan, or tannins, among others. These hydroxylated aromatic structures are also crucial in industry (epoxide and phenolic resins, azo dyes, explosives) and as biologically active ingredients (drugs, herbicides). (482)

Figure 62

Figure 62. (top) Summary of general properties of phenol sensors as well as expected cross-reactivity during screenings. (bottom): Compatible pH range and buffers in reported phenol-selective assays. Bold pH numbers indicate the most commonly used pH values.

Synthesis of hydroxylated aromatic compounds (483) typically requires harsh conditions, which has motivated a shift to enzymatic production of this class of compounds. (484)
With the emergence of novel biocatalysts, namely mono- (MO), di- (DO), or peroxygenases (PO) as well as tyrosinases (TYR), which enable aromatic hydroxylation, the scientific community has strived to discover and develop improved enzymatic catalysts. (484) Lan Tee and Schwaneberg summarized the developments for directed evolutions of oxygenases, its success stories, and challenges in a recent review. (485)
Among many examples for the success of these transformations (Scheme 21), Molina-Espeja et al. applied directed evolution strategies for the optimization of an unspecific peroxygenase from Agrocybe aegerita (AaeUPO1), engineered for the synthesis of 1-naphthol 198. (486,678) The resulting double mutant (G241D-R257K) enabled the regioselective synthesis with a total turnover number of 50 000. The wild-type has also been efficiently applied for the hydroxylation of acetanilide 199 (among other relevant precursors), resulting in the selective production of acetaminophen (paracetamol) 200. (487) The group of Arnold evolved a toluene dioxygenase to improve the turnover of 4-picoline 201, resulting in the discovery of the single-point mutant R459L with 5.6-times higher activity than the wild type. (488)

Scheme 21

Scheme 21. Representative Selection of Enzymatic Transformations for the Production of Phenol-Containing Products
Phenols are more acidic than alcohols because the anion formed in deprotonation is stabilized by the conjugated electron system of the aryl. The delocalized charge increases the nucleophilicity and the antioxidant properties of phenols (stabilization of radicals compared to benzene). (489) Their qualities as antioxidants are often used as a means to quantifying phenol content in terms of total antioxidant capability (AOC). (490)
Several assays have been developed in the food and live sciences due to the necessity for the assessment of antioxidant properties: the Folin–Ciocalteu (FC) test, (491−493) ABTS (2,2′-azino-bis(3-ethylbenzothiazoline-6-sulfonic acid) radical scavenging, (494) ORAC (oxygen radical absorbance capacity), (495) FRAP (ferric-reducing antioxidant power), (496) CuPRAC (cupric ion reducing antioxidant capacity), (496) TEAC (trolox equivalent antioxidant capacity), (497) and DPPH (1,1-diphenyl-2-picrylhydrazyl) radical scavenging. (494) None of these established methods are recognized as universal, (496,498,499) and the correct selection for the study of AOC remains difficult.
Reducing power of the matrix can be problematic in detecting phenols, particularly in biological media with many reducing compounds (e.g., ascorbic acid, sulphites, reducing sugars). (500) Hence, reaction-based methods that integrate the phenol into a chromo- or fluorogenic product have proven superior in sensitivity. (501) Due to the direct conjugation of the phenol structures into the final chromophores, large differences in absorbance can be expected for each analyte (Figure 62).

14.1. Electrophilic Aromatic Substitution (SEAr) with Activated Amino Groups

Most methods rely on trapping the nucleophilic phenol or phenolate with a positively charged probe in an electrophilic aromatic substitution (Figure 63). The subsequent oxidation of the dienone typically triggers the large change in color (and/or intensity) and increases the stability of the products (iminoquinones). The most widely applied method uses commercially available 4-aminoantipyrine (4-AAP, 203), initially reported by Emerson in 1943. (502) Its oxidative coupling to phenols produces an intensely amber to red colored complex, (490,503,504) which is most sensitive at pH 9–10. (503,505) Appropriate buffering is necessary to reduce noise, which was typically caused by a distinct drop in pH upon the addition of reagents. (505,506)

Figure 63

Figure 63. Summary of chromogenic turn-on probes for the selective detection of phenols based on the electrophilic aromatic substitution with oxidized amino derivatives. Abs: wavelengths typically used for UV absorbance measurements. Ex or em: wavelengths typically used for excitation and emission in fluorescence measurements. The wavelengths of probes in this section are shown in the strip at the bottom representing the visible spectrum. Alk = alkyl, Ar = aryl.

Typically, K3[Fe(CN)6] or K2S2O8 are used as oxidants, although recently systems consisting of iron(III)-octacarboxyphthalocyanine and t-BuOOH have been demonstrated as fast, water-soluble oxidants. (507) Synthetic optimizations of the parent structures have been proposed with the exchange of alkyl and aryl-side substituents, with little improvement over the 4-AAP assay. (506,508) Only phenols with hydrogen, halides, or alkoxy substituents in para-position are detectable. (506) This assay system has been used for the colorimetric determination of phenolic compounds in pharmaceutical preparations, (503) tetracyclines, (509) and many other examples. (505,510)
Probe 203 was also used in directed evolution of hydroxylases toward aromatic and heterocyclic compounds (501,511) and esterases by the group of Schwaneberg. (512) Reliable detection of phenolic substrates at 50–800 μM was enabled after 30 min incubation. Choi et al. applied the probe in HTS and directed evolution of a tyrosine phenol-lyase. (513)
Probe 204 (“Gibbs reagent”) is a prominent example in this category. It is already preoxidized (aniline is trapped as chloroimide), thus does not require an oxidant. (514) The chloroimide is rapidly hydrolyzed at pH > 8.5, yielding the free imine. Both forms react exclusively at the para-position of the phenol ring. As with 203, only hydrogen, halide, or alkoxy substituents are tolerated. (515,516) Even though decomposition of the quinoneimine was shown to occur in alkaline medium, the formation of the product is favored, and the reaction is not hampered by prolonged reaction times. Full conversion of phenol was reported using a 10-fold excess at pH 10 after approximately 10–20 min at μM concentrations. (517)
204 was used by Ang et al. for the directed evolution of an aniline dioxygenase for bioremediation purposes, targeting the resulting catechol derivatives. (518) Arnold’s group used it in HTS of dioxygenase activity using solid-phase digital imaging (486) and for the evolution of a toluene dioxygenase toward 4-picoline (4-methylpyridine). (488) Here, the initially formed cis-dihydrodiol product was not detected and deemed to be unstable. Its spontaneous rearomatization by dehydration, forming the 3-hydroxy-4-picoline 202, allowed the detection with 204 without using dihydrodiol dehydrogenase or decreasing of pH, which is typically required for this transformation. Bornscheuer et al. also used the reagent to test the selectivity of 28 lipases and esterases in the hydrolysis of butanoates of ortho-, meta-, or para-substituted phenols. (519)
Taking advantage of the well-studied spectroscopic properties of phenol blue, Altahir et al. developed a phenol probe based on p-amino-N,N-dimethylaniline 205. (520) Using a similar protocol as for the 4-AAP assay, the two-electron oxidation of the aniline with K3[Fe(CN)6] under basic conditions enabled the detection of eugenol contained in personal-care products.
MBTH 114, known as a tool to detect aldehydes (cf. section 9.1), has also been efficiently used for the derivatization of phenols. (521) Da Silva Granja et al. evaluated 114 among other methods for the determination of total phenol content (203, FC assay, ABTS assay, DPPH 109 assay) and found it to be equally useful. (500) The reaction with 114 works in dilute H2SO4 (500,521) and at at pH 8–9, (522,523) contrary to the typically neutral to basic conditions required by reagents 203 and 204. Oxidations are typically performed with K3[Fe(CN)6], FeCl3, or K2S2O8. Oxidative coupling mediated by laccase or horseradish peroxidase have also been reported. (524,525) Reactions with phenols are not limited to the para-position, as ortho-positions have also been successfully derivatized. Side reactions can be expected with amines, electron-rich aromatic compounds, and aldehydes (e.g., reducing sugars). Nolan and O’Connor applied the assay for the quantification of a toluene-4-monooxygenase activity producing para-substituted phenols. These were detected after tyrosinase-assisted transformation of the products to 4-substituted catechols. (526)

14.2. Azo Coupling

Phenols can be detected using azo-coupling reactions (Figure 64). Herein, aromatic diazonium salts react as strong electrophiles with activated arenes in electrophilic aromatic substitutions at the ortho- or para-unsubstituted positions. (527) The resulting azo compounds often have vivid colors due to their large conjugated systems and are widely used as dyes and pigments. (528,529)

Figure 64

Figure 64. Summary of chromogenic turn-on probes for the selective detection of phenols based on the formation of azo-dyes. Abs: wavelengths typically used for UV absorbance measurements. The wavelengths of probes in this section are shown in the strip at the bottom representing the visible spectrum.

Urbányi and Mollica tested 27 differently substituted aniline derivatives and established a strong dependency of the sensitivity of these probes on the functionality and substitution pattern of their aromatic systems. Electron-poor diazonium salts were shown to form more intensely colored products after fixed reaction times (6 min). Although no underlying kinetic study was undertaken and single points were used as read-outs, the authors concluded that electron-poor systems are the better analytical probes for rapid screenings. (530)
Diazotized sulfanilic acid 206 (Pauly’s or Ehrlich’s diazo reagent) has been used as early as 1883 by Ehrlich for the detection of bilirubin in urine samples (531) and has found many applications in the detection of phenolic derivatives since then. (532−535) Its electron-poor character, however, makes it unstable; fresh preparation before use is necessary.
The emergence of commercially available, stabilized diazonium salts, either by decoration of the aromatic core with electron-donating functionalities or complexation with suitable anions, such as zinc chloride or tetrafluoroborates, enabled broad usage of these reagents (“Fast” salts) for the detection of phenols in food, water, and biological samples. (535−538) The high stability of electron-rich arenes has been systematically studied by Schotten et al. (539) by comparing differently substituted anilines. The formation of reversibly cleavable triazenes with secondary amines was reported to further increase stability.
Currently, many differently substituted diazonium salts are commercially available, with a broad scope of applications in different buffer systems. Johnston and Ashford studied the nine most widely used “Fast” reagents for the quantification of activity of hog liver esterase, using α-naphthyl acetate as substrate. (540) Fast Red GG 207 (p-nitrobenzenediazonium tetrafluoroborate) and Fast Violet B 208 emerged as the best combination of reagents to cover a combined pH range of 3–9.5. Coupling reactions with these probes were reported to be complete in under 5 s at approximately 1 mM concentrations using 5-fold excess naphthol, allowing the determination of initial rates.
It has been reported for 207 (but not for 208) that enzymatic activity might be inhibited by certain diazonium salts or their stabilizers. (541) Purification of commercial material or validation of activity by an orthogonal method might prove useful. Lugg tested combinations of 24 airborne phenols and anilines with 25 Fast salts and thus showed the differences in spectral responses of these assays. (542) The author concluded that probe 210 has universal applicability and a low background. The salt has also been used to quantify phenol in air and urine (543) and o-phenylphenol in aqueous samples. (544)
Structural analogues with red-shifted absorption maxima (Fast Blue RR 209, BB 210, and B 211) have also been used to detect phenol in samples. (535,536) Probe 211 was used by the group of Wood for the directed evolution of a toluene o-monooxygenase. (545) Among other protocols for HTS for aromatic hydroxylations using 203 and 204, Otey and Joern presented a concise summary for the detection of phenols, also using 208 with a large set of differently substituted phenols. (510)

14.3. Mannich Type Reaction

A new approach for a fluorogenic detection of phenols is based on Mannich-type reactions of activated arenes (Betti reaction; Figure 65). (546,547) The group of Tanaka reported the reaction of phenols and peptides bearing tyrosine units with cyclic imines to proceed without the use of catalysts from pH 2–10. (548)

Figure 65

Figure 65. Summary of fluorogenic turn-on probes for the selective detection of phenol based on Mannich-type reactions. Ex or em: wavelengths typically used for excitation and emission in fluorescence measurements. The wavelengths of probes in this section are shown in the strip at the bottom representing the visible spectrum.

They designed several fluorogenic reporters 212215 based on this mechanism. (679) Substitution patterns of the attacking arene are crucial: only ortho-phenol substituted product gave the desired increase in fluorescence. The authors hypothesized that intramolecular interaction between the phenolic hydroxyl group and the amine, or π–π stacking interactions between the phenol and fluorophore aryl moieties, were essential for fluorescence to arise. Using methyl 4-hydroxyphenylacetate and their coumarin probe 212, they demonstrated an approximately 100-fold increase in fluorescence after 70 min at pH 7. The slow reaction required a 1000-fold excess of phenol at μM concentrations. No further investigations and optimization of the reaction conditions were undertaken to test the limits of the system.

15. Chromo- and Fluorogenic Probes for Thiols

Click to copy section linkSection link copied!

Thiols are essential components of biological systems. (549) Compounds such as glutathione (GSH), cysteine (Cys), homocysteine (Hcy), and coenzyme A are involved in various biological processes including regulation of the oxidation–reduction states of the cells, (550) protein synthesis, (551) signal transduction, (552) detoxification of xenobiotics, (553) fatty acid biosynthesis, (554) and posttranslational modifications. (555) Aberrations of thiol levels in cells are considered to be related to diseases such as Parkinson’s disease, (556) Alzheimer’s disease, (557) osteoporosis, (558) and cancer. (559) Therefore, the development of selective and sensitive detection methods for thiols has become an important research field in clinical diagnosis. (556,559)
Other thiols, e.g., thiophenols, are used in the agrochemical industry for pesticide syntheses, (560,561) in the pharmaceutical industry, (562) and in the food industry as flavoring ingredients (e.g., 2-ethylthiophenol with a roasted, smokey odor). (563)
Thiols are nucleophilic and react with electrophiles. In this regard, their reactivity is highly correlated with their pKa value, as the compounds become stronger nucleophiles when deprotonated. A typical pKa of aromatic thiols is ∼6.5, and ∼8.5 for aliphatic thiols. (564,565) Fluorescence-based assays can be applied for thiol detection and quantification, in addition to chromatography, UV/vis spectroscopy, potentiometry, and mass spectrometry. (566−568) Many of the fluorescence assays use the nucleophilicity of the thiols in reactions such as the Michael addition, cleavage of benzenesulfonamides and benzenesulfonates, cyclization with an aldehydes, cleavage of (O-, S-, Se-)phenyl ethers, cleavage of disulfides, or make use of the high affinity of thiols toward metals. (249,564,569)
Potential challenges in the development/application of fluorescent probes for thiols include cross-reactivity with other nucleophiles (e.g., cyanides, (570,571) amines, and amino acids), (572) requirements for a narrow pH range, (573) and slow reaction rates under physiological conditions (Figure 66). Thiols occur in a large range of concentrations in cells, making selective detection even more difficult; Bennett et al. quantified metabolite concentrations in E. coli and found GSH at 17 mM and Hcy at only 0.37 mM. (29) Additionally, selectivity between thiols can be an issue (e.g., GSH vs Cys/Hcy, or the distinction between Cys/Hcy, GSH, and H2S). (563,565,574−576) Various probes thus require substrates with an amino group in addition to the thiol functionality to produce a color/fluorescence response. In the following section, those probes are labeled with a blue “NH2-required”.

Figure 66

Figure 66. (top) Summary of general properties of thiols, reactivity trends toward thiol-selective probes, and expected cross-reactivity. (bottom) Compatible pH range and buffers in reported thiol-selective assays. Bold pH numbers indicate the most commonly used pH values.

These probes discriminate GSH and Cys/Hcy based on a reaction sequence: nucleophilic substitution reaction or conjugate addition, and subsequent intramolecular thiol to amine rearrangements or cyclizations. These S–N rearrangements for Cys/Hcy are favored compared to GSH due to lower energy transition states.
Some assays that discriminate Cys/Hcy against GSH are based on the cyclization of Cys and Hcy with aldehyde-containing probes, forming stable five-and six-membered rings. Most development is focused on detecting the thiols already mentioned (Cys, Hcy, GSH, and coenzyme A). Hence many assays were only tested with these substrates.

15.1. Nucleophilic Aromatic Substitution (SNAr)

Thiolates readily react with electron-deficient centers via nucleophilic aromatic substitution (SNAr). Electron-withdrawing groups at the aromatic ring provide the needed electron deficiency for the substitution reaction. Common leaving groups for the thiol-mediated SNAr include phenylsulfonyl moieties, halogenides, O-phenyl ethers, and S-phenyl ethers (Figure 67). (577)

Figure 67

Figure 67. Summary of fluorogenic turn-on probes for the selective detection of thiols based on nucleophilic aromatic substitution reactions (SNAr) with aromatic systems containing electron-withdrawing groups (EWG). Ex or em: wavelengths typically used for excitation and emission in fluorescence measurements. The wavelengths of probes in this section are shown in the strip at the bottom representing the visible spectrum.

In 1968, Ghosh and Whitehouse observed that benzofurazans, e.g., 4-chloro-7-nitro-2,1,3-benzofurazan (NBD-Cl, 12), showed a fluorescent response upon reaction with amino groups. They also noticed that thiol derivatives were slightly fluorescent. (79) Since then, 12 and (aminosulfonyl)-7-fluoro-2,1,3-benzoxadiazole (ABD-F, 216), which was developed some years later by Toyo’oka and Imai, (578) have been applied for the detection of thiols and activity screenings of carnitine palmitoyltransferase, citrate synthase, and acetylcholinesterase. In these cases, the commercially available probes showed a fluorescent signal upon reaction with the enzymatically released thiols. (579,580)
Niu et al. showed that 12 could be used for the selective and rapid detection of Cys/Hcy in the presence of GSH under mild conditions in an aqueous acetonitrile (25% ACN). The selectivity for of Cys/Hcy is potentially linked to an intramolecular displacement of the thiolate, leading to the formation of the N-substituted NBD derivative with strong fluorescence. In contrast, the S-substituted derivative with GSH only gave negligible fluorescence. The observation that the S-substituted and not the N-substituted derivative was obtained in the case of GSH was explained by comparing transition states (showing an unfavorable macrocyclic transition state with GSH, and favorable five- or six-membered rings as transition states with Cys and Hcy). An possible explanation for the faster response with Cys than Hcy could be that the formation of five-membered rings is faster than six-membered ones (Scheme 22). (581)

Scheme 22

Scheme 22. Proposed Mechanism for the Reaction of NBD-Cl 12 with Cys, Hcy, and GSH (581)
The ratiometric BODIPY probe (BODIPY-PPh3,217) was used for the selective detection of GSH over Cys/Hcy by nucleophilic substitution of chlorine with the thiolate. Distinction between GSH and Cys/Hcy was possible because GSH gave a highly fluorescent, sulfur-substituted conjugate, but Cys/Hcy formed weakly fluorescent, amino-substituted BODIPY derivatives upon rearrangement. (582) By adding a triphenylphosphonium moiety, it was possible to target the probe at mitochondria. (583) Hu et al. made use of thiols’ ability to cleave phenyl ethers and developed a nonfluorescent probe with an NBD scaffold (NR-NBD, 218) that enabled rapid discrimination (within a few minutes) between Cys/Hcy and GSH under aqueous conditions by using two excitation wavelengths, accompanied by a color change from blue to green. (584) He et al. enhanced this idea and were the first that expanded the functionality with the probe HMN 219, that simultaneously distinguished between Cys/Hcy, GSH, and hydrogen sulfide (Figure 68). (565)

Figure 68

Figure 68. Design of the probe HMN 219 and the proposed fluorescence signal changes in response to individual or sequential detection of thiols with three well-defined emission bands. Reproduced with permission from ref (565). Copyright 2017 Royal Society of Chemistry.

S-Phenyl ethers and Se-phenyl ethers can also be used as leaving groups. The bond length of the carbon-heteroatom bond increases from oxyethers over thioethers to selenoethers, with a concomitant decrease in stability. The most stable bond (oxyethers) also makes the reaction rate of O-phenyl ether-based probes generally slow. (575,585,586) However, rates can be increased by adding electron-withdrawing groups at the aromatic ring, for example, in nitrophenyl-O-ethers. (565,586)
In 2015, Kim et al. developed probe 220 for the selective and fast detection of GSH under physiological conditions. A phenylselenide linked to a coumarin probe was used as a leaving group. The electrophilicity of the probe was increased by installing two carbonyl groups adjacent to the reaction site, enabling a fast substitution reaction with the nucleophilic thiol. Subsequently, an intramolecular cyclization forming an iminium group was observed. With 220, detection of GSH in an aqueous DMSO solution was possible within seconds. (587) Later, the authors presented modifications of the probe by substituting the diethylamino group with an N-heterocyclic ring and introduction of an ethyl butanoate group. They envisioned that by restricting the free rotation of the amino group and the additional ethyl butanoate functionality, the low quantum yield and the poor water solubility of 220 would be improved. Although probe 221 (DACP-2) showed a slower response (about 10 min) than 220, it indeed was better soluble in water and had a higher quantum yield (0.48 vs <0.001). It enabled the selective detection of GSH and Cys/Hcy. (588)

15.2. DNBS Cleavage via Nucleophilic Aromatic Substitution (SNAr)

DNBS (2,4-dinitrobenzene sulfonyl) is a commonly used building block in thiol-sensitive probes. Upon reaction with thiolates, the benzenesulfonamide or benzenesulfonate bond is cleaved, SO2 is released, and fluorescence is restored (Figure 69). (589−593) In 2005, Maeda et al. developed the DNBS-based probe BESThio 222, which showed a high increase in fluorescence within minutes upon reaction with thiols under mild aqueous conditions (pH 7.4, 37 °C). At the same time, no fluorescence signal was observed with amines, such as ethylamine or piperidine, or alcohols, such as catechol or glucose, at pH 5.8–7.4. The authors demonstrated that their probe could be used as an alternative to the commonly used colorimetric Ellman’s reagent 232 and applied it in the fast screening of ChE (cholinesterase) inhibitors. Furthermore, they showed that their probe could detect selenols and distinguish between thiols and selenols by adjusting the pH from 7.4 to 5.8 (Figure 70). (594,589)

Figure 69

Figure 69. Summary of fluorogenic turn-on probes for the selective detection of thiols based on 2,4-dinitrobenzene sulfonyl (DNBS) cleavage by nucleophilic aromatic substitution reactions (SNAr) with aromatic systems containing electron-withdrawing groups (EWG). Ex or em: wavelengths typically used for excitation and emission in fluorescence measurements. The wavelengths of probes in this section are shown in the strip at the bottom representing the visible spectrum.

Figure 70

Figure 70. Fluorescent probe BESThio 222 and its discrimination between selenols and thiols. Reproduced with permission from ref (594). Copyright 2006 Wiley-VCH.

Figure 71

Figure 71. Summary of fluorogenic turn-on probes for the selective detection of thiols based on disulfide bond cleavage. Abs: wavelengths typically used for UV absorbance measurements. Ex or em: wavelengths typically used for excitation and emission in fluorescence measurements. The wavelengths of probes in this section are shown in the strip at the bottom representing the visible spectrum.

Peng et al. investigated methods to detect the activity of β-lactamase and reported an indirect detection approach based on aggregation-induced emission (AIE) and excited-state intramolecular proton transfer (ESIPT). (595) Fluorophores that exhibit AIE fluoresce strongly in the solid state or as aggregates in solution. The AIE and ESIPT characteristics of salicylaldehyde azine-based fluorophores depend on their hydroxyl groups. In probe DNBS-CSA 223, one hydroxyl groups was blocked by a 2,4-dinitrobenzenesulfonate group. The reaction of 223 with thiols led to their deprotection, which recovered fluorescence by AIE and ESIPT.
The indirect detection method by Peng et al. consisted of three steps. After the enzymatic cleavage of the β-lactam ring of the cefazolin substrate, a spontaneous elimination yielded a thiol, which was detected with DNBS-CSA by releasing CSA (salicylaldehyde azine derivative) and thereby enabling AIE and ESIPT fluorescence. This approach enabled the detection of β-lactamase activity under mild conditions in solution (detection limit 0.5 mU mL–1) and on paper-based test strips. (595)
Yuan et al. described the NIR-fluorescent probe NIR-Thiol 224, with DNBS as a recognition moiety, for the detection and imaging of thiols in living mice. The probe does not fluoresce by itself and showed a fast response to μM concentrations of thiols. The addition of Cys in PBS buffer/ACN (v/v = 7:3; pH 7.4) resulted in a 50-fold increase in fluorescence, reaching the maximum within 60 s. Probe 224 was proven to be stable under assay conditions (24 h) and could be stored as a solid for more than one year. (596)
Challenges in clinical diagnostics motivated the development of fast fluorescent probes for thiols. (579,597−599) Lin et al. developed the NIR DNBS-based probe Hcyc-NO 225 that allowed the almost instantaneous detection (<5 s) of thiols relevant for evaluating cellular oxidative stress, including GSH, Cys, and Hcy. The assay works under mild, aqueous conditions (PBS/DMSO 8:2, pH 7.4) and showed detection limits of 0.08 μM (Cys), 0.20 μM (Hcy), and 0.11 μM (GSH). Although probe 225 readily reacts with thiols from pH 4–8, it was observed that more basic conditions resulted in increased fluorescence. (600)
Lu et al. reported a far-red fluorescent probe based on the iminobenzo[g]coumarin dye and DNBS as recognition site (Cou-DNBS, 226), which reacted quickly even with low amounts (nM) of thiols under basic or neutral conditions. It was used for monitoring intracellular fluctuations in thiol levels, demonstrated by incubating HeLa cells that had been pretreated with H2O2 with the probe and imaging the development of thiol concentration over time. Fluorescence increased from zero (or insignificant noise) to strong signals within 20 min, indicating thiol recovery. (597)
Development of probes that are specific to a particular thiol is challenging. Thiophenols are toxic to aquatic organisms and animals and harmful to human health after prolonged exposure. Therefore, probes that can detect thiophenols and differentiate between those and other thiols are of interest. (590,601−603) Jiang et al. developed the essentially nonfluorescent NBD-Cl 12 modified probe 227, with a benzenesulfonamide reaction site. The selectivity for thiophenols was achieved by relying on the pKa differences between aliphatic thiols (∼8.5) and thiophenols (∼6.5). Thus, thiophenolates could rapidly cleave the benzenesulfonamide bond under neutral conditions, whereas the aliphatic thiols remained in the less reactive neutral form. No fluorescence response was observed for aliphatic thiols, such as t-butyl thioalcohol, GSH, and Cys, or other nucleophiles (e.g., NaCN, BnNH2). The water-soluble probe was synthesized in two steps from 12 with moderate yield and gave a more than 50-fold increase in fluorescence intensity within 10 min of addition of thiophenol or derivatives (two equivalents), such as 4-methyl-, 4-methoxy, or 4-chlorothiophenol. A lower fluorescence enhancement was measured with 4-nitrothiophenol, explained by the strong electron-withdrawing properties of the nitro-group. (590)
The same concept for discrimination between aliphatic and aromatic thiols was utilized by Xiong et al. They developed the fluorogenic and chromogenic probe SQ-DNBS 228, with a rapid response (≤10 min) and high selectivity toward thiophenols. Their squaraine-based probe with DNBS as the reactive site was used to detect and quantify thiophenols in water samples. The far-red/NIR probe showed a high sensitivity (LOD: 9.9 nM), good photostability (>1 h under constant irradiation), and could be applied under neutral or slightly basic aqueous conditions without the need for a cosolvent. Furthermore, the color change from pink to blue upon the addition of thiophenols enabled the convenient detection by the naked eye. (604)
Lanthanide complexes can be useful to eliminate background fluorescence in biological samples emanating from flavins, porphyrins, or nicotinamide adenine dinucleotides. Whereas the biogenic fluorescent signals typically have short decay times (<10 ns), the photoluminescence (PL) of lanthanide complexes (e.g., with Eu3+ or Tb3+) takes 10–1000 μs to decay (1000 to 100 000 longer lasting). By starting the measurement only after the autofluorescence has already decayed, a better signal-to-noise ratio can be obtained. With the DNBS-based Eu3+/Tb3+ complex NSTTA-Ln3+229, Dai et al. quantified thiols at μM concentration in aqueous samples and living cells, using this time-gated acquisition. (592,605,606)
Shang et al. and Fan et al. synthesized two-photon fluorescent DNBS probes. The probe DNEPI 230, by Fan et al., exhibited a fast response (about 2 min) and a high selectivity for Cys over other thiols like Hcy and GSH. (569,607) The probe FR-TP 231, developed by Shang et al., showed high selectivity toward thiophenols, a 155-fold fluorescence enhancement with an emission in the far-red region, and a rapid response (complete within minutes). Probe 231 enabled the detection of thiophenol poisoning in an animal model for thiophenol inhalation, using a two-photon fluorescence microscope.

15.3. Disulfide Cleavage

Thiols can cleave disulfide bonds, which can be used to detect them (Figure 71). (591) Chromogenic DTNB 232 (“Ellman’s reagent”) has been frequently used to quantify thiols and enzymes. (608−610) When thiols react with 232, the disulfide bond is cleaved, and chromogenic 5-thio-2-nitrobenzoate with an absorption maximum at 412 nm is released. Instability at pH > 8 creates issues with high background signal and interferences with absorption from other biomolecules (e.g., hemoglobin) and led to the development of other reagents. (609,611−615)
Based on the work by Adamczyk and Grote, Pullela et al. synthesized fluorescein- and rhodamine-containing probes (DSSA, 233) with a disulfide linker that could be used for detecting thiols in vivo and in vitro in the mM range under aqueous conditions. (616)
Cao et al. synthesized the ratiometric fluorescent probe Ratio-HPSSC 234, based on the tetrakis(4-hydroxyphenyl)porphyrin-coumarin scaffold to detect thiols under neutral conditions in ethanolic (50%) aqueous buffer solutions. A change in emission color from red to blue enabled the visual detection of thiols, with a detection limit for Cys at 0.73 μM. It was shown that the probe was suitable for imaging thiols in living cells. (617)
Wang et al. developed the ratiometric probe 235, with a disulfide bond, which linked a BODIPY moiety as FRET donor and rhodamine as FRET acceptor. It showed a high sensitivity for GSH and enabled detection down to the low mM range. (618)
Chen et al. pursued a different strategy using fluorescent carbon nanoparticles, so-called carbon quantum dots, which exhibit beneficial properties including tunable fluorescence emissions, easy functionalization, and high photochemical stability. They reported a nanoswitch based on fluorescent carbon nanoparticles, which had thiol groups covalently attached at their surface (Figure 72). These quantum dots showed a green fluorescence that vanished when the dots aggregated due to disulfide bond formation. In the presence of thiols, the disulfide bonds were cleaved and fluorescence was restored. This process proved to be reversible as the disulfide linkage and reaggregation could be triggered by H2O2. This concept was applied for evaluating the activity of butyrylcholinesterase (BChE) and for screening butyrylcholinesterase inhibitors using butyrylthiocholine iodide as a substrate. Conversion of the substrate yielded butyrate and thiocholine, which reacted with the quantum dots and thereby restored the signal. The assay could also be used to quantify the activity of AChE by changing the substrate to acetylthiocholine chloride. (611,619) Gong et al. developed another method to determine the activity of BChE and test its inhibitors. Their disulfide containing probe (“FRET chemical probe” 236) enabled measuring BChE activity in serum directly without interference from GSH, explained by the probe’s steric hindrance. (612)

Figure 72

Figure 72. Schematic illustration of redox-controlled fluorescent nanoswitch and detection strategy for BChE activity based on thiol-triggered disulfide cleavage on fluorescent carbon nanoparticles. Reproduced with permission from ref (611). Copyright 2018 American Chemical Society.

15.4. Diselenide Cleavage

Diselenides (R-Se-Se-R) can be used instead of disulfides as a reactive center to detect thiols. (Figure 73). Diselenide bonds have lower bond energies and are cleaved faster by the nucleophilic attack of thiols than disulfides. (620,621) An notable toxicity of organoselenium compound should be expected due to their oxidizing as well as inhibitory properties for certain thiol-containing enzymes. (622)

Figure 73

Figure 73. Summary of fluorogenic turn-on probes for the selective detection of thiols based on diselenide bond cleavage. Ex or em: wavelengths typically used for excitation and emission in fluorescence measurements. The wavelengths of probes in this section are shown in the strip at the bottom representing the visible spectrum.

Lou et al. developed the diselenide probe FSeSeF 237, with two fluorescein moieties as signal reporters that were linked by a diselenide bond. Upon cleavage of the diselenide bond, the two fluorescein-containing moieties are released, resulting in strong fluorescence. Probe 237 showed high selectivity for thiols and rapid response (within milliseconds for GSH in PBS buffer pH 7.4). The reaction is reversible in the presence of H2O2, and the probe was thus used for the detection of redox changes mediated by thiols and ROS. (623) Han et al. designed the diselenide-based fluorescent probe Cyo-Dise 238, which showed high selectivity for GSH and could detect it within seconds in HEPES buffer (pH 7.4) by fluorescent response, and the color change from green to red. The ratiometric fluorescent probe 238 was applied for evaluating changes in GSH levels in HepG2 and HL-7702 cells under hypothermic and hyperthermic conditions and in HepG2 and HepG2/DDP xenografts on nude mice. (624)

15.5. Selenium–Nitrogen Bond Cleavage

Thiols are sufficiently nucleophilic to cleave selenium–nitrogen bonds. Based on this reactivity, thiol-detecting probes containing selenium–nitrogen linkages have been developed (Figure 74). In 2007, Tang et al. designed the weakly fluorescent rhodamine-based probe 239 with a Se–N bond, which releases the strongly fluorescent dye rhodamine 6G upon cleavage. They demonstrated that probe 239 could be used for the quantitative detection of GSH (LOD: 1.4 nM) in PBS buffer (pH 7.4). Interestingly, a higher sensitivity was observed for the protein-bound thiols in glutathione reductase or metallothionein than for free thiols, such as Cys, 2-mercaptoethanol, or dithiothreitol. (625)

Figure 74

Figure 74. Summary of fluorogenic turn-on probes for the selective detection of thiols based on selenium–nitrogen bond cleavage. Ex or em: wavelengths typically used for excitation and emission in fluorescence measurements. The wavelengths of probes in this section are shown in the strip at the bottom representing the visible spectrum.in the strip at the bottom representing the visible spectrum.

The high noise and only moderate enhancement of fluorescence (about 6-fold) prompted Tang et al. to develop a better rhodamine-based probe, Rh-Se-2 240. Structural modification prevented an opening of the spirocycle under screening conditions, thereby reducing noise. Probe 240 reacted fast (within minutes) and selectively with thiols, with a >170-fold enhancement in fluorescence in aqueous solution. In addition to the turn-on response in fluorescence, the color changed from light yellow to yellow-green. (626)
The cyanine-dye-based NIR probes Cy-TfSe 241 and Cy-NiSe 242 developed by Wang et al. showed a fast signal response with thiols in PBS buffer at physiological pH. Fluorescence of these dyes was stable from pH 4.0 to 8.6. (627) Tian et al. designed the colorimetric and fluorescent probe 243, with two Se–N bonds that allowed differentiation of Cys, Hcy, and GSH (Figure 75), whereas GSH and Hcy only cleaved one Se–N bond, Cys cleaved both. Differences in the absorption and fluorescence spectra allowed distinguishing between GSH, Hcy, and Cys simultaneously. Determination of the kinetic constants and equilibrium constants showed that the reactivity of Cys toward the probe is much higher than that of GSH and Hcy, which could explain that only the presence of Cys led to the cleavage of both bonds. (628)

Figure 75

Figure 75. Structure and color changes of 243 in the presence of amino acids (1, Cys; 2, Hcy; 3, GSH; 4, none; 5, other natural amino acids). Reproduced with permission from ref (628) . Copyright 2018 Elsevier BV.

15.6. Cyclization with Aldehydes

1,2- and 1,3-Aminothiols can form five- or six-membered rings with aldehydes. Based on this mechanism, fluorescent probes with aldehyde moieties for the selective detection of cysteine and homocysteine have been reported (Figure 76). (572,629,630) In 2004, Rusin et al. developed the xanthene-based probe 244, bearing an aldehyde for colorimetric and turn-off fluorometric detection of Cys and Hcy in alkaline aqueous solutions (pH 9.5). (629) The requirement for basic conditions prompted the development of other thiol detecting probes that could be used under physiological conditions. Lee et al. reported the coumarin-based probe 245 with a salicylaldehyde moiety, which showed high selectivity for Cys/Hcy and could be used at neutral pH. (631)

Figure 76

Figure 76. Summary of fluorogenic turn-on probes for the selective detection of thiols based on the cyclization with aldehydes. A subsequent cyclization via an additional NH2 moiety limits the application to analytes containing both functional groups. Abs: wavelengths typically used for UV absorbance measurements. Ex or em: wavelengths typically used for excitation and emission in fluorescence measurements. The wavelengths of probes in this section are shown in the strip at the bottom representing the visible spectrum.

15.7. Conjugate Addition (1,4-Addition)

Thiols readily undergo 1,4-additions to α,β-unsaturated carbonyl compounds due to their soft nucleophilicity. They are softer nucleophiles than amines and alcohols. (632−635) Maleimides are frequently used as reactive centers for thiol-sensitive probes because they react selectively with thiols and quench fluorescence based on n−π* transitions or PET (Figure 77). Upon addition of the thiol, the quenching process is suppressed and fluorescence recovered. (632)

Figure 77

Figure 77. Summary of fluorogenic turn-on probes for the selective detection of thiols-based conjugate additions. Ex or em: wavelengths typically used for excitation and emission in fluorescence measurements. The wavelengths of probes in this section are shown in the strip at the bottom representing the visible spectrum.

Commercially available N-(4-(7-diethylamino-4-methylcoumarin-3-yl)phenyl)maleimide (CPM, 246) is often used to detect thiols. (636,637) Aharoni et al. applied 246 in a HTS of serum paraoxonase (PON1) mutants to find variants with improved thiolactonase activity. The gene variants were transformed into E. coli, and the single cells were encapsulated in w/o/w (water-in-oil-in-water) droplets together with the substrate thiobutyrolactone and 246. The droplets were sorted using FACS. With this method, they were able to discover variants with ∼100-fold higher catalytic efficiency than the wild-type PON1. (638)
Marcella and Barb presented a FabD (malonyl-coenzyme A transacylase) and a FabH (ketoacyl synthase III) assay that used 246 to detect the enzymatically released coenzyme A. Interfering thiol-containing holo-ACP in the FabD assay was removed by precipitation with trichloroacetic acid before the probe was added. Fluorescence was measured at 470 nm. They applied the assays for the rapid determination of FabH and FabD activity. (639)
Niu et al. used 246 in an assay for HTS of potent and selective hCBS (human cystathionine β-synthase) inhibitors derived from natural compounds. The assay was based on the conversion of methylcysteine by cystathionine β-synthase (CBS), leading to the release of methanethiol, which was subsequently detected by 246 at physiological pH (Figure 78). With this approach, 6491 compounds were tested, and several compounds that inhibited hCBS activity were discovered, including sikokianin C, which exhibited cytotoxic activity against HT29 cancer cells with an IC50 value (half-maximal inhibitory concentration) of 1.5 uM. The assay also could be used for monitoring hCSE (human cystathionine γ-lyase) activity. (637)

Figure 78

Figure 78. Illustration of the high-throughput assays for cystathionine β-synthase (CBS). (a) Methanethiol (CH3SH) generation catalyzed by CBS using methylcysteine (MCys) as a substrate was determined with the CPM 246. (b) The CBS activity was monitored using 10 mM MCys as the substrate in the presence of 15 μM 246 and 2 μg of hCBS in a 200 μL reaction mixture containing 50 mM HEPES, pH 7.4. The fluorescence of the reaction mixture at 460 nm (λex = 400 nm) was monitored for 600 s. Reproduced with permission from ref (637). Copyright 2017 Royal Society of Chemistry.

Yi et al. designed the coumarin-based probe CME 247 with a maleimide moiety that showed a rapid response with high selectivity and sensitivity for thiols (<1 min for Cys) in Tris-HCl buffer (pH 7.4). It had a 470-fold increase in quantum yield upon conjugation with GSH. The assay’s sensitivity enabled the determination of GSH at concentrations as low as 0.5 nM. The water-soluble 247 could be efficiently synthesized in three steps and is stable for months in solution and in pure form. Yi et al. successfully used it to develop a high-throughput fluorescence assay for glutathione reductase. (640)
Gao et al. compared the performance of various fluorescent probes in assaying the activity of a histone acetyltransferase and concluded that probes 246 and 247 were well suited in a high-throughput format (fast reaction, fluorescence enhancement of >250-fold with coenzyme A). (555)
Zhou et al. developed the cationic probe 248, consisting of a coumarin core linked to a methylpyridinium moiety by an unsaturated ketone. It enabled the detection of Cys at nM concentration under physiological conditions, based on a new assisted electrostatic attraction strategy, and showed a fast fluorescent and colorimetric response. They assumed that the pyridinium group confers high solubility in water, high reactivity with thiols (due to its electron-withdrawing properties), a fast response (due to electrostatic interactions), and good cell permeability. The nonfluorescent and water-soluble 248 showed a 148-fold increase in fluorescence emission upon the addition of Cys and reached the maximum fluorescence intensity within 1 min. Hcy and GSH reacted much slower with the probe than Cys, explained by the lower steric hindrance, pKa, and electrostatic attraction of Cys. (641)
Langmuir et al. developed several maleimide-containing probes (including 249, 250) that almost instantaneously reacted with thiols, such as ME, GSH, and N-acetyl-l-cysteine, under physiological pH. Upon addition of a thiol to the maleimide, the initially weak fluorescence of the probes was strongly enhanced. (636)
Li et al. reported on the development of the NIR probe NIR-BODIPY-Ac 251 for the selective detection of Cys in the presence of GSH and Hcy under neutral aqueous conditions. It consisted of an acrylate group as a reactive site and an indolium-BODIPY as NIR fluorophore. Upon conjugate addition of Cys/Hcy to the acrylate, an intramolecular cyclization led to the subsequent elimination of a cyclic amide (Scheme 23). This release prompted a strong increase in fluorescence emission. The probe is highly selective for Cys in the presence of Hcy, explained by the formation of a kinetically more favored, seven-membered ring with Cys. The probe was used to detect exogenous and endogenous Cys in in vitro and in vivo. (642)

Scheme 23

Scheme 23. Proposed Response Mechanism for NIR-BODIPY-Ac 251 to Cys or Hcy

15.8. Sulfur–Metal Interaction

Many assays rely on the high affinity of thiols toward metals and various probes that build upon these interactions with Ag, Cu, Hg, and Au, have been reported (Figure 79). (643−645) Apart from concepts based on displacement, assays also use oxidation and reduction of metals. (646−649) Jung et al. and Wang et al. developed probes 252 and 253 with Cu2+ ensembles for the detection of biogenic thiols (Figure 80). Both probes showed a fast response and a high selectivity toward thiols, in the presence of amino acids, in aqueous solution under neutral or alkaline conditions. It was inferred that in the presence of thiols, a decomplexation of Cu2+ from the iminofluorescein-Cu2+252 probe or the iminocoumarin-Cu2+ probe 253 took place, followed by hydrolysis of the probe to yield the fluorescent fluoresceinaldehyde or coumarinaldehyde. (650,651)

Figure 79

Figure 79. Summary of fluorogenic turn-on probes for the selective detection of thiols-based sulfur–metal interactions. Ex or em: wavelengths typically used for excitation and emission in fluorescence measurements. The wavelengths of probes in this section are shown in the strip at the bottom representing the visible spectrum.

Figure 80

Figure 80. Schematic illustration of the thiol detecting chemodosimetric mechanism iminocoumarin-Cu2+ probe 253 in aqueous media. Adapted with permission from ref (651). Copyright 2011 The Royal Society of Chemistry.

Many assays based on thiol–metal interactions use metal nanoparticles or nanoclusters. Based on silver nanoparticles and fluorescent carbon dots, a colorimetric and fluorescent assay for the activity of AChE with acetylthiocholine as a substrate was developed by Zhao et al. (652) The fluorescence of the carbon dots was quenched by the absorption of the dispersed nanoparticles (inner filter effect). The high affinity of the formed thiocholine for the silver nanoparticles and electrostatic interactions led to an aggregation, reducing inner filter effect and leading to a change of fluorescence (LOD: 0.016 mU mL–1) and color (LOD: 0.021 mU mL–1).
Polyethyleneimine-protected copper nanoclusters, or a water-soluble perylene derivative in combination with Cu2+, were used for the detection of thiols, measuring the activity of AChE, and for screening inhibitors. (653,654) The mechanism for thiols was based on quenching the strong fluorescence of the polyethyleneimine-protected nanoclusters by binding of Cu2+ to the amino groups of polyethyleneimine. In the presence of thiols, however, the high affinity of Cu2+ to thiols led to the formation of thiol–Cu2+ complexes instead, which resulted in an increased fluorescent signal. (653) Similarly, adding Cu2+ quenched the fluorescence of the perylene diimide derivative 254 by binding of Cu2+ to the carboxylic groups. Capture of the Cu2+ ions coordinated to the probe by thiols enhanced the fluorescence. (654) Both probes enabled the determination of AChE activity with a detection limit of 1.38 mU mL–1 and 1.78 mU mL–1, respectively. (653,654)
Jiang et al. developed a sensitive, photoluminescent probe based on cytidine-stabilized Au nanoclusters 255 for rapid measurement of glutathione reductase activity and screening of its inhibitors. Although the photoluminescence of metal nanoclusters is often quenched by thiols, their probe showed enhanced photoluminescence. The authors speculated that the increase might result from the formation of smaller, fluorescent Au species (including Au2) etched by thiols. The probe enabled them to detect GSH at nM concentrations (LOD 2.0 nM). (169)
A displacement assay based on BODIPY-Au nanoparticles 256 was reported by Xu et al. The BODIPY chromophores on the Au nanoparticle (AuNP) surfaces could be displaced by thiols, thereby restoring the fluorescence of the BODIPY molecules (Figure 81). It showed a fast response for Cys and Hcy (∼2 min) in PBS buffer (pH 7.4), high selectivity and sensitivity (LOD: 30 nM for Cys) and was used for imaging thiols in living HeLa cells. (655)

Figure 81

Figure 81. Schematic representation of the thiol sensor 256 based on modulation of the fluorescence quenching of the BODIPY chromophore by AuNPs. Reproduced with permission from ref (655) . Copyright 2016 Elsevier BV.

15.9. Further Reaction Types

There are other types of thiol-sensing probes, based on reaction types such as 1,2 additions to meso-vinyl BODIPY derivatives, (656) the addition of thiols to squaraine-like compounds, (570) thiol-meditated reduction of quinone, (657) thiol–halogen nucleophilic substitution reactions, (581,658−660) or transesterification of thioesters. (661) Chen et al. developed an NIR fluorescent probe for thiols (CHMC-thiol) that consisted of two reaction sites: the 2,4-dinitrobenzenesulfonate moiety was chosen as a high-sensitivity reaction site, and the less reactive chloro group was chosen as a low-sensitivity site. Low concentrations of cysteine (0–50 μM) led to the cleavage of the benzenesulfonate moiety and resulted in an increase in fluorescence at 680 nm. In the presence of high concentrations of cysteine (50–500 uM), the chlorine was substituted, followed by an intramolecular rearrangement to yield the amino product, and a ratiometric fluorescence response was obtained. This approach enabled the sensing of multiple concentration ranges via different fluorescence signals. (662)

16. Concluding Remarks

Click to copy section linkSection link copied!

Enzymes have intrigued chemists to unleash their catalytic utility ever since their discovery. They provide an efficient and environmentally benign alternative to conventional organic–synthetic methodology due to their unparalleled selectivity. The surge in recent identifications and the potential for optimizing catalytically active proteins promoted their use in an industrial context. In particular, computationally designed proteins and directed evolution allowed chemists to explore uncharted territory in catalysis.
Directed evolution, however, requires sophisticated screening tools that can cope with the immense numbers of variants created in randomized approaches. Libraries containing millions or billions of variants need to be assessed in a time-effective way while maintaining accuracy in the screening process. The number of potential screening events thus imposes constraints in time and matter on the underlying analytical procedures. Every mechanical manipulation has an impact on throughput, making a direct detection in crude cell lysates the optimal choice. Molecular tools thus need to function in complex environments and reduce errors stemming from reactions with endogenous compounds to a minimum.
Fluorogenic substrates that mimic the natural or desired substrate offer a viable alternative to search for catalytic functionality but decouple the analytical signal from the synthetically relevant transformation. Functional group assays offer a more generalized approach for an optical read-out. By selective interaction with defined structural properties of analytes, these tools generate powerful handles to address the issues outlined above.
Finding and adapting these handles for compatibility to physiological conditions are the biggest challenges in the field of assay development, clearly shown in the distribution of probes presented here for the different substance classes. Some functional groups are more easily detected because of high reactivity (e.g., aldehydes, amines), while others have large numbers of reactive probes due to a strong scientific or applied interest (e.g., thiols). In general, assays that work under “real” screening conditions enable applications in HTS and are thus more likely to be used frequently.
Several limitations of functional group probes exist that limit the number of tools for detecting other (not highly reactive) substance classes in situ: low solubility in water, incompatibility with water, or high/low pH, slow reaction kinetics, especially at low product concentrations, insufficient selectivity, or undesirable optical properties. Nevertheless, there are designs that address all these disadvantages: Addition of PEG groups to 37 or incorporation of charged residues like quaternary ammonium or pyridinium groups to known fluorogenic dyes, such as 29 and 248, have been used to increase their water solubility. Kinetical hurdles of ABAO reagents 126 were overcome by understanding the underlying reaction mechanism. Implementation of a methoxy group into the scaffold rendered the aldehyde selective tools also suitable for more challenging ketone detection. Selectivity issues for efficient discrimination between the GSH, Cys, and H2S were effectively overcome by designing the fluorogenic probe HMN 219, which contained multiple reactive sites. The optical properties of pyrylium probes (cf. section 6.5) were also adjusted by modifications of their fluorophore, essentially making the assays applicable over the whole visible spectrum.
Understanding the mechanisms for the genesis of the optical signal and preserving structural features critical to the desired reactivity aids the development of good probes. Although some probes seem to be applicable for a broad set of analytes (e.g., nucleophiles targeting electrophilic reactive sites), the exact selectivity of most assays is generally not (extensively) characterized. It might thus prove useful for protein engineers to test a range of different assays.
We would like to encourage research groups working on new assays to characterize the selectivity toward various derivatives of the same substance or reactivity class (e.g., nucleophiles, electrophiles, oxidants) and with expected interference from cellular components. Only a small number of probes are commercially available or easily accessible by synthesis, particularly for some functional groups (e.g., alcohols, esters). We hope that this overview of available methods motivates the reader to tackle the remaining challenges, helping protein engineers to advance in their pursuit for new and better catalytic activity.

Author Information

Click to copy section linkSection link copied!

  • Corresponding Author
  • Authors
    • Sebastian Hecko - Institute of Applied Synthetic Chemistry, OC-163, TU Wien, Getreidemarkt 9, 1060 Vienna, AustriaOrcidhttps://orcid.org/0000-0002-6210-4425
    • Astrid Schiefer - Institute of Applied Synthetic Chemistry, OC-163, TU Wien, Getreidemarkt 9, 1060 Vienna, AustriaOrcidhttps://orcid.org/0000-0002-3591-8765
    • Christoffel P. S. Badenhorst - Institute of Biochemistry, Dept. of Biotechnology & Enzyme Catalysis, University of Greifswald, Felix-Hausdorff-Str. 4, 17489 Greifswald, GermanyOrcidhttps://orcid.org/0000-0002-5874-4577
    • Michael J. Fink - Department of Chemistry and Chemical Biology, Harvard University, 12 Oxford St, Cambridge, Massachusetts 02138, United StatesOrcidhttps://orcid.org/0000-0003-0023-8767
    • Marko D. Mihovilovic - Institute of Applied Synthetic Chemistry, OC-163, TU Wien, Getreidemarkt 9, 1060 Vienna, Austria
    • Uwe T. Bornscheuer - Institute of Biochemistry, Dept. of Biotechnology & Enzyme Catalysis, University of Greifswald, Felix-Hausdorff-Str. 4, 17489 Greifswald, GermanyOrcidhttps://orcid.org/0000-0003-0685-2696
  • Author Contributions

    CRediT: Sebastian Hecko data curation, visualization, writing-original draft, writing-review & editing; Astrid Schiefer data curation, visualization, writing-original draft, writing-review & editing; Christoffel P. S. Badenhorst writing-original draft, writing-review & editing; Marko D. Mihovilovic writing-review & editing; Uwe T. Bornscheuer validation, writing-original draft, writing-review & editing; Michael J. Fink validation, writing-review & editing; Florian Rudroff conceptualization, funding acquisition, project administration, resources, supervision, writing-review & editing.

  • Funding

    APC Funding Statement: Open Access is funded by the Austrian Science Fund (FWF).

  • Notes
    The authors declare no competing financial interest.

Biographies

Click to copy section linkSection link copied!

Sebastian Hecko received his B.Sc. and Dipl.Ing. in Technical Chemistry from the TU Wien. He then joined the joint research group of Prof. Rudroff and Prof. Mihovilovic at TU Wien, where he obtained his Ph.D. degree in Applied Synthetic Chemistry, developing a high-throughput screening platform for the detection of carbonyl compounds in fluorescence-activated droplet sorting systems. His current research as postdoctoral researcher focuses on the synthesis and application of ultrahigh-throughput assay systems for the directed evolution of carbonyl producing enzymes.

Astrid Schiefer is currently a Ph.D. student in the research group Bioorganic Synthetic Chemistry at TU Wien. She studied Technical Chemistry at TU Wien and received her master’s degree in 2021. During her master’s thesis, she gained experience in assay development for the application in fluorescence-activated droplet sorting. Her current research focuses on enzyme-assisted methods for C═C bond cleavage.

Chris P. S. Badenhorst received a B.Sc. in Biochemistry and Microbiology from the University of South Africa. He then joined Prof. Albie van Dijk’s laboratory at the Potchefstroom Campus of the North West University, where he obtained his B.Sc. Honours, M.Sc., and Ph.D. degrees in Biochemistry with a focus on molecular biology and human metabolism. He was subsequently awarded an Alexander von Humboldt Research fellowship to join Uwe Bornscheuer’s group, where he has been working on a variety of enzyme classes and the development of new ultrahigh-throughput screening methods.

Michael J. Fink works in applied and translational research using chemistry, biotechnology, material science, and informatics. In 2019, he cofounded the startup company Datacule (together with George M. Whitesides) to develop new solutions for digital data storage based on his research at the Department of Chemistry and Chemical Biology, Harvard University (2015–2020). His research interests include information in molecular systems, microbial biotechnology, health-and-disease-related microbiology, forensic analysis, and the functional design of molecules for applications in medicine and public health. He received his undergraduate and graduate education in chemical engineering from TU Wien in Austria (B.Sc. 2008, M.Sc. 2009, Ph.D. 2013).

Marko D. Mihovilovic graduated in Technical Chemistry at TU Wien in 1993, also receiving his doctorate from the same university in 1996, in the field of Organic Synthetic Chemistry. Postdoc placements as an Erwin Schrödinger scholarship holder then followed at the University of New Brunswick (Canada) and the University of Florida (USA) in the fields of biocatalysis and molecular biology. Returning to TU Wien, he set up his own research group in 1999. He completed his habilitation in 2003 in bio-organic chemistry and was appointed Associate Professor in 2004. Marko Mihovilovic has been Head of the Institute for Applied Synthetic Chemistry at TU Wien since 2013 and was appointed as Full Professor for Bioorganic Synthetic Chemistry in February 2014.

Uwe T. Bornscheuer studied Chemistry and received his Ph.D. in 1993 at Hannover University followed by a postdoc at Nagoya University (Japan). In 1998, he completed his Habilitation at Stuttgart University about the use of lipases and esterases in organic synthesis. He has been Professor at the Institute of Biochemistry at Greifswald University since 1999. In 2022, he received, beside other awards, the Enzyme Engineering Award. His current research interests are on the discovery and engineering of enzymes from various classes for applications in organic synthesis, for flavors and fragrances, in lipid modification, and the degradation of plastics or complex marine polysaccharides.

Florian Rudroff obtained his diploma and Ph.D. degree at TU Wien in Organic Chemistry. Afterward, he was awarded an “Erwin Schrödinger fellowship” and went for a postdoctoral stay with the group of Prof. Uwe Sauer at ETH Zürich in the field of Systems Biology. In 2011, he returned to the TU Wien and started his independent scientific career in Systems Biocatalysis. In December 2017, he became Assistant Professor and finished his Habilitation in Bioorganic Chemistry at the Institute of Applied Synthetic Chemistry, TU Wien. In 2020, he was appointed as Associate Professor. His main research interests are enzyme cascade catalysis, photoredox biocatalysis (e.g., CO2 utilization by cyanobacteria), protein engineering, the development of high-throughput platforms (e.g., fluorescence-activated droplet sorting, FADS), and organic synthesis.

Acknowledgments

Click to copy section linkSection link copied!

This work has been financially supported by the Austrian Science Fund (FWF Project P33687-N) and TU Wien.

Abbreviations

Click to copy section linkSection link copied!

AA

amino acid

AAOx

aryl alcohol oxidase

AaeAPO

aromatic peroxygenase (Agrocybe aegerita)

AaeUPO

unspecific peroxygenase (Agrocybe aegerita)

ABAO

amino benzamidoxime

ABTS

2,2′-azino-bis(3-ethylbenzothiazoline-6 sulfonic acid)

AChE

acetylcholinesterase

ACN

acetonitrile

ACP

acyl carrier protein

ADH

alcohol dehydrogenase

AHL

amidohydrolase

AIE

aggregation-induced emission

AKR

aldo-keto reductase

Alk

alkyl

AOC

antioxidant capability

AO/AOx

alcohol oxidase

Ar

aryl

ATA

amine transaminase

ATase

acyl transferase

AtHNL

hydroxynitrile lyase (Arabidopsis thaliana)

ATP

adenosine triphosphate

AuNCs

gold nanoclusters

AuNPs

gold nanoparticles

BChE

butyrylcholinesterase

BD

benzoxadiazole

BDHA

alcohol dehydrogenase (Bacillus subtilis)

BINOL

(1,1′-binaphthalene)-2,2′-diol

BmEH

epoxide hydrolase (Bacillus megaterium)

BnNH2

benzylamine

BphA

biphenyl dioxygenase

BSA

bovine serum albumin

(n)BuOH

butanol

BVMO

Baeyer–Villiger monooxygenase

CAR

carboxylic acid reductase

CBS

cystathionine β-synthase

CFE

cell-free extract

ChE

cholinesterase

CHEF

chelation-enhanced fluorescence

CoA/CoASH

coenzyme A

CQD

carbon quantum dot

CS

chitosan

CSA

salicylaldehyde azine derivative

CuPRAC

cupric ion reducing antioxidant capacity

Cys

cysteine

DCC

dicyclohexylcarbodiimide

de

diastereomeric excess

DERA

deoxyribose 5-phosphate aldolase

DMF

N,N-dimethylformamide

DMSO

dimethylsulfoxide

DNA

deoxyribonucleic acid

DNBS

2,4-dinitrophenyl sulfonyl

DNPH

2,4-dinitrophenylhydrazine

DO

dioxygenase

DPPH

1,1-diphenyl-2-picrylhydrazyl

EDG

electron-donating group

ee

enantiomeric excess

EET

electronic energy transfer

EH

epoxide hydrolase

epPCR

error-prone PCR

ESIPT

excited-state intramolecular proton transfer

EtOH

ethanol

EtyGly

ethylene glycol

EWG

electron-withdrawing groups

ε

molar absorption coefficient

FA

formaldehyde

FabD

malonyl-coenzyme A transacylase

FabH

ketoacyl synthase III

FACS

fluorescence-activated cell sorting

FADS

fluorescence-activated droplet sorting

FC test

Folin–Ciocalteau test

FGA

functional group assays

FGI

functional group interconversion

FITC

fluorescein isothiocyanate

FRAP

ferric-reducing antioxidant power

FRET

Förster resonance energy transfer

GC

gas chromatography

GDH

glutamate dehydrogenase

GFP

green fluorescent protein

GSH

glutathione

hCBS

human cystathionine β-synthase

hCES

human cystathionine γ-lyase

Hcy

homocysteine

HHDH

halohydrin dehalogenase

HNL

hxdroxynitrile lyases

HPLC

high-performance liquid chromatography

HTA

high-throughput assay

HPTS

8-hydroxypyrene-1,3,6-trisulfonic acid trisodium salt

HTS

high-throughput screening

IC

internal conversion

ICT

intramolecular charge transfer

iPrNH2

isopropylamine

IR

infrared

IRED

Imine reductase

ISC

intersystem crossing

ITC

isothiocyanate

KDC

keto acid decarboxylase

KIVD

2-ketoisovalerate decarboxylase

KM

Michaelis–Menten constant

KpADH

diaromatic ketone reductase

KRED

ketoreductase

KS

ketone synthase

LDH

lactate dehydrogenase

LFER

linear free energy relationship

LG

leaving group

LOD

limit of detection

λabs/abs

absorbance wavelength [nm]

λem/em

emission wavelength [nm]

λex/ex

excitation wavelength [nm]

MAH

mandelamide hydrolase

MAO

monoamine oxidase

MCys

methylcysteine

MDA

malondialdehyde

MDH

mandelate dehydrogenase

ME

mercaptoethanol

MEK

methyl ethyl ketone

MeNH2

methylamine

MeNH3Cl

methylammonium chloride

MeOH

methanol

MO

monooxygenase

MS

mass spectrometry

μMS

microscale mass spectrometry

NAD+/NADH

nicotinamide adenine dinucleotide

NADP+/NADPH

nicotinamide adenine dinucleotide phosphate

NBD

nitrobenzoxadiazole

NCL

native chemical ligation

NHS

N-hydroxysuccinimide

NIR

near-infrared

NOS

reactive nitrogen species

Nuc

nucleophile

ORAC

oxygen radical absorbance capacity

PBS

phosphate-buffered saline

PcAOX

alcohol oxidase (Phanerochaete chrysosporium)

PCR

polymerase chain reaction

PEG

polyethylene glycol

PET

photoinduced electron transfer

PFP

pentafluorophenyl

PGM

phosphoglucomutase

PL

photoluminescence

PLP

pyridoxal phosphate

PMA

phosphomolybdic acid

PO

peroxygenase

(n)PrOH

propanol

PVC

polyvinyl chloride

Φ, QY

quantum yield

RA

reductive aminase

ROS

reactive oxygen species

RSS

reactive sulfur species

SHC

squalene hopene cyclase

SNAr

nucleophilic aromatic substitution

SQ

squaraine

TA

transaminase

t-BuOH

tert-butyl alcohol

TDO

toluene dioxygenase

TDP

thiamine diphosphate

TEAC

trolox equivalent antioxidant capacity

TEMPO

(2,2,6,6-tetramethylpiperidin-1-yl)oxyl

ThDP

thiamine diphosphate

TICT

twisted intramolecular charge transfer

TLC

thin-layer chromatography

p-TsOH

p-toluenesulfonic acid

TsrE

monooxygenase

TTN

total turnover number

TYR

tyrosinase

uHTS

ultra-high-throughput screening

UV

ultraviolet

VIS

visible

w/o/w

water-in-oil-in-water

YerE

ThDP-dependent flavoenzyme (Yersinia pseudotuberculosis)

References

Click to copy section linkSection link copied!

This article references 679 other publications.

  1. 1
    Li, C.; Zhang, R.; Wang, J.; Wilson, L. M.; Yan, Y. Protein Engineering for Improving and Diversifying Natural Product Biosynthesis. Trends Biotechnol. 2020, 38 (7), 72944,  DOI: 10.1016/j.tibtech.2019.12.008
  2. 2
    Kazlauskas, R. Engineering More Stable Proteins. Chem. Soc. Rev. 2018, 47 (24), 902645,  DOI: 10.1039/C8CS00014J
  3. 3
    Behrens, G. A.; Hummel, A.; Padhi, S. K.; Schätzle, S.; Bornscheuer, U. T. Discovery and Protein Engineering of Biocatalysts for Organic Synthesis. Adv. Synth. Catal. 2011, 353 (13), 2191215,  DOI: 10.1002/adsc.201100446
  4. 4
    Liu, Z.; Arnold, F. H. New-to-Nature Chemistry from Old Protein Machinery: Carbene and Nitrene Transferases. Curr. Opin. Biotechnol. 2021, 69, 4351,  DOI: 10.1016/j.copbio.2020.12.005
  5. 5
    Steiner, K.; Schwab, H. Recent Advances in Rational Approaches for Enzyme Engineering. Comput. Struct. Biotechnol. J. 2012, 2 (3), e201209010  DOI: 10.5936/csbj.201209010
  6. 6
    Zeymer, C.; Hilvert, D. Directed Evolution of Protein Catalysts. Annu. Rev. Biochem. 2018, 87 (1), 13157,  DOI: 10.1146/annurev-biochem-062917-012034
  7. 7
    Markel, U.; Essani, K. D.; Besirlioglu, V.; Schiffels, J.; Streit, W. R.; Schwaneberg, U. Advances in Ultrahigh-Throughput Screening for Directed Enzyme Evolution. Chem. Soc. Rev. 2020, 49 (1), 23362,  DOI: 10.1039/C8CS00981C
  8. 8
    Arnold, F. H.; Volkov, A. A. Directed Evolution of Biocatalysts. Curr. Opin. Chem. Biol. 1999, 3 (1), 549,  DOI: 10.1016/S1367-5931(99)80010-6
  9. 9
    Arnold, F. H.; Smith, G. P.; Winter, G. P. Directed Evolution-Bringing the Power of Evolution to the Laboratory: 2018 Nobel Prize in Chemistry. Curr. Sci. 2018, 115 (9), 1627
  10. 10
    Goldsmith, M.; Tawfik, D. S. Enzyme Engineering: Reaching the Maximal Catalytic Efficiency Peak. Curr. Opin. Struct. Biol. 2017, 47, 14050,  DOI: 10.1016/j.sbi.2017.09.002
  11. 11
    Currin, A.; Swainston, N.; Day, P. J.; Kell, D. B. Synthetic Biology for the Directed Evolution of Protein Biocatalysts: Navigating Sequence Space Intelligently. Chem. Soc. Rev. 2015, 44 (5), 1172239,  DOI: 10.1039/C4CS00351A
  12. 12
    Reymond, J. L. Enzyme Assays; Wiley, 2006.
  13. 13
    Singh, H.; Tiwari, K.; Tiwari, R.; Pramanik, S. K.; Das, A. Small Molecule as Fluorescent Probes for Monitoring Intracellular Enzymatic Transformations. Chem. Rev. 2019, 119 (22), 1171860,  DOI: 10.1021/acs.chemrev.9b00379
  14. 14
    Chan, J.; Dodani, S. C.; Chang, C. J. Reaction-Based Small-Molecule Fluorescent Probes for Chemoselective Bioimaging. Nat. Chem. 2012, 4 (12), 97384,  DOI: 10.1038/nchem.1500
  15. 15
    Reetz, M. T.; Carballeira, J. D. Iterative Saturation Mutagenesis (ISM) for Rapid Directed Evolution of Functional Enzymes. Nat. Protoc. 2007, 2 (4), 891903,  DOI: 10.1038/nprot.2007.72
  16. 16
    Lapetina, S.; Gil-Henn, H. A Guide to Simple, Direct, and Quantitative in Vitro Binding Assays. J. Biol. Methods 2017, 4 (1), e62, DOI: 10.14440/jbm.2017.161 .
  17. 17
    Major, J. Challenges and Opportunities in High Throughput Screening: Implications for New Technologies. J. Biomol. Screen. 1998, 3 (1), 137,  DOI: 10.1177/108705719800300102
  18. 18
    Neun, S.; Kaminski, T. S.; Hollfelder, F. Single-Cell Activity Screening in Microfluidic Droplets. Methods Enzymol. 2019, 628 (5), 95112
  19. 19
    Zhang, J.; Chai, X.; He, X.-P.; Kim, H.-J.; Yoon, J.; Tian, H. Fluorogenic Probes for Disease-Relevant Enzymes. Chem. Soc. Rev. 2019, 48 (2), 683722,  DOI: 10.1039/C7CS00907K
  20. 20
    Kaur, K.; Saini, R.; Kumar, A.; Luxami, V.; Kaur, N.; Singh, P.; Kumar, S. Chemodosimeters: An Approach for Detection and Estimation of Biologically and Medically Relevant Metal Ions, Anions and Thiols. Coord. Chem. Rev. 2012, 256 (17), 19922028,  DOI: 10.1016/j.ccr.2012.04.013
  21. 21
    You, L.; Arnold, F. H. Directed Evolution of Subtilisin E in Bacillus Subtilis to Enhance Total Activity in Aqueous Dimethylformamide. Protein Eng. 1996, 9 (1), 7783,  DOI: 10.1093/protein/9.1.77
  22. 22
    Schmidt-Dannert, C.; Arnold, F. H. Directed Evolution of Industrial Enzymes. Trends Biotechnol. 1999, 17 (4), 1356,  DOI: 10.1016/S0167-7799(98)01283-9
  23. 23
    Choi, K.-Y.; Jung, E.-O.; Yun, H.; Yang, Y.-H.; Kazlauskas, R. J.; Kim, B.-G. Development of Colorimetric HTS Assay of Cytochrome P450 for Ortho-Specific Hydroxylation, and Engineering of CYP102D1 with Enhanced Catalytic Activity and Regioselectivity. ChemBioChem. 2013, 14 (10), 12318,  DOI: 10.1002/cbic.201300212
  24. 24
    Roth, M. Fluorescence Reaction for Amino Acids. Anal. Chem. 1971, 43 (7), 8802,  DOI: 10.1021/ac60302a020
  25. 25
    Weigele, M.; DeBernardo, S. L.; Tengi, J. P.; Leimgruber, W. Novel Reagent for the Fluorometric Assay of Primary Amines. J. Am. Chem. Soc. 1972, 94 (16), 59278,  DOI: 10.1021/ja00771a084
  26. 26
    Fellner, M.; Doughty, L. M.; Jameson, G. N.; Wilbanks, S. M. A Chromogenic Assay of Substrate Depletion by Thiol Dioxygenases. Anal. Biochem. 2014, 459, 5660,  DOI: 10.1016/j.ab.2014.05.008
  27. 27
    Santos-Aberturas, J.; Dörr, M.; Waldo, G. S.; Bornscheuer, U. T. In-Depth High-Throughput Screening of Protein Engineering Libraries by Split-GFP Direct Crude Cell Extract Data Normalization. Chem. Biol. 2015, 22 (10), 140614,  DOI: 10.1016/j.chembiol.2015.08.014
  28. 28
    Azad, T.; Tashakor, A.; Hosseinkhani, S. Split-Luciferase Complementary Assay: Applications, Recent Developments, and Future Perspectives. Anal. Bioanal. Chem. 2014, 406 (23), 554160,  DOI: 10.1007/s00216-014-7980-8
  29. 29
    Bennett, B. D.; Kimball, E. H.; Gao, M.; Osterhout, R.; Van Dien, S. J.; Rabinowitz, J. D. Absolute Metabolite Concentrations and Implied Enzyme Active Site Occupancy in Escherichia Coli. Nat. Chem. Biol. 2009, 5 (8), 5939,  DOI: 10.1038/nchembio.186
  30. 30
    Zuman, P. Reactions of Orthophthalaldehyde with Nucleophiles. Chem. Rev. 2004, 104 (7), 321738,  DOI: 10.1021/cr0304424
  31. 31
    Henke, E.; Bornscheuer, U. T. Fluorophoric Assay for the High-Throughput Determination of Amidase Activity. Anal. Chem. 2003, 75 (2), 25560,  DOI: 10.1021/ac0258610
  32. 32
    Médici, R.; de María, P. D.; Otten, L. G.; Straathof, A. J. J. A High-Throughput Screening Assay for Amino Acid Decarboxylase Activity. Adv. Synth. Catal. 2011, 353 (13), 236976,  DOI: 10.1002/adsc.201100386
  33. 33
    Choi, J.-Y.; Black, R.; Lee, H.; Di Giovanni, J.; Murphy, R. C.; Ben Mamoun, C.; Voelker, D. R. An Improved and Highly Selective Fluorescence Assay for Measuring Phosphatidylserine Decarboxylase Activity. J. Biol. Chem. 2020, 295, 9211,  DOI: 10.1074/jbc.RA120.013421
  34. 34
    Leslie, A. K.; Li, D.; Koide, K. Amine-Promoted β-Elimination of a β-Aryloxy Aldehyde for Fluorogenic Chemodosimeters. J. Org. Chem. 2011, 76 (16), 68605,  DOI: 10.1021/jo200947e
  35. 35
    Rinde, E.; Troll, W. Colorimetric Assay for Aromatic Amines. Anal. Chem. 1976, 48 (3), 5424,  DOI: 10.1021/ac60367a036
  36. 36
    Spangenberg, B.; Poole, C. F.; Weins, C. Quantitative Thin-Layer Chromatography: A Practical Survey; Springer Science & Business Media, 2011.
  37. 37
    Ehmann, A. The Van Urk-Salkowski Reagent ─ a Sensitive and Specific Chromogenic Reagent for Silica Gel Thin-Layer Chromatographic Detection and Identification of Indole Derivatives. J. Chromatogr. A 1977, 132 (2), 26776,  DOI: 10.1016/S0021-9673(00)89300-0
  38. 38
    Jones, L. A.; Holmes, J. C.; Seligman, R. B. Spectrophotometric Studies of Some 2,4-Dinitrophenylhydrazones. Anal. Chem. 1956, 28 (2), 1918,  DOI: 10.1021/ac60110a013
  39. 39
    Zarzycki, P.; Bartoszuk, M.; Radziwon, A. Optimization of Tlc Detection by Phosphomolybdic Acid Staining for Robust Quantification of Cholesterol and Bile Acids. J. Planar Chromatogr. - Mod. TLC 2006, 19 (107), 527,  DOI: 10.1556/JPC.19.2006.1.9
  40. 40
    Bergmann, F. Colorimetric Determination of Amides as Hydroxamic Acids. Anal. Chem. 1952, 24 (8), 13679,  DOI: 10.1021/ac60068a033
  41. 41
    Gayathri, T. Stains for Developing TLC Plates. Int. J. Adv. Res. Sci. Eng. Technol. 2014, 5 (11), 7983
  42. 42
    Leeuwenkamp, O. R.; van Bennekom, W. P.; van der Mark, E. J.; Bult, A. Nitroprusside, Antihypertensive Drug and Analytical Reagent. Review of (Photo)stability, Pharmacology and Analytical Properties. Pharm. Weekbl. Sci. Ed. 1984, 6 (4), 129140,  DOI: 10.1007/BF01954040
  43. 43
    Lavis, L. D.; Raines, R. T. Bright Building Blocks for Chemical Biology. ACS Chem. Biol. 2014, 9 (4), 85566,  DOI: 10.1021/cb500078u
  44. 44
    Bowen, E. J. Fluorescence and Fluorescence Quenching. Q. Rev. Chem. Soc. 1947, 1 (1), 115,  DOI: 10.1039/qr9470100001
  45. 45
    Fu, Y.; Finney, N. S. Small-Molecule Fluorescent Probes and Their Design. RSC Adv. 2018, 8 (51), 2905161,  DOI: 10.1039/C8RA02297F
  46. 46
    Grimm, J. B.; Heckman, L. M.; Lavis, L. D. The Chemistry of Small-Molecule Fluorogenic Probes. Prog. Mol. Biol.Transl. Sci. 2013, 113 (1), 134
  47. 47
    Daly, B.; Ling, J.; De Silva, A. P. Current Developments in Fluorescent PET (Photoinduced Electron Transfer) Sensors and Switches. Chem. Soc. Rev. 2015, 44 (13), 420311,  DOI: 10.1039/C4CS00334A
  48. 48
    Algar, W. R.; Hildebrandt, N.; Vogel, S. S.; Medintz, I. L. Fret as a Biomolecular Research Tool ─ Understanding Its Potential While Avoiding Pitfalls. Nat. Methods 2019, 16 (9), 81529,  DOI: 10.1038/s41592-019-0530-8
  49. 49
    Misra, R.; Bhattacharyya, S. P. Intramolecular Charge Transfer: Theory and Applications; Wiley VCH, 2018 DOI: 10.1002/9783527801916
  50. 50
    Sedgwick, A. C.; Wu, L.; Han, H.-H.; Bull, S. D.; He, X.-P.; James, T. D.; Sessler, J. L.; Tang, B. Z.; Tian, H.; Yoon, J. Excited-State Intramolecular Proton-Transfer (ESIPT) Based Fluorescence Sensors and Imaging Agents. Chem. Soc. Rev. 2018, 47 (23), 884280,  DOI: 10.1039/C8CS00185E
  51. 51
    Hong, Y.; Lam, J. W.; Tang, B. Z. Aggregation-Induced Emission. Chem. Soc. Rev. 2011, 40 (11), 536188,  DOI: 10.1039/c1cs15113d
  52. 52
    Grabowski, Z. R.; Rotkiewicz, K.; Rettig, W. Structural Changes Accompanying Intramolecular Electron Transfer: Focus on Twisted Intramolecular Charge-Transfer States and Structures. Chem. Rev. 2003, 103 (10), 38994032,  DOI: 10.1021/cr940745l
  53. 53
    Tanner, P. A.; Zhou, L.; Duan, C.; Wong, K.-L. Misconceptions in Electronic Energy Transfer: Bridging the Gap between Chemistry and Physics. Chem. Soc. Rev. 2018, 47 (14), 523465,  DOI: 10.1039/C8CS00002F
  54. 54
    Jiang, X.; Wang, L.; Carroll, S. L.; Chen, J.; Wang, M. C.; Wang, J. Challenges and Opportunities for Small-Molecule Fluorescent Probes in Redox Biology Applications. Antioxid. Redox Signal. 2018, 29 (6), 51840,  DOI: 10.1089/ars.2017.7491
  55. 55
    Wu, L.; Sedgwick, A. C.; Sun, X.; Bull, S. D.; He, X. P.; James, T. D. Reaction-Based Fluorescent Probes for the Detection and Imaging of Reactive Oxygen, Nitrogen, and Sulfur Species. Acc. Chem. Res. 2019, 52 (9), 258297,  DOI: 10.1021/acs.accounts.9b00302
  56. 56
    Jiao, X.; Li, Y.; Niu, J.; Xie, X.; Wang, X.; Tang, B. Small-Molecule Fluorescent Probes for Imaging and Detection of Reactive Oxygen, Nitrogen, and Sulfur Species in Biological Systems. Anal. Chem. 2018, 90 (1), 53355,  DOI: 10.1021/acs.analchem.7b04234
  57. 57
    Yang, Y.; Zhao, Q.; Feng, W.; Li, F. Luminescent Chemodosimeters for Bioimaging. Chem. Rev. 2013, 113 (1), 192270,  DOI: 10.1021/cr2004103
  58. 58
    Chyan, W.; Raines, R. T. Enzyme-Activated Fluorogenic Probes for Live-Cell and in Vivo Imaging. ACS Chem. Biol. 2018, 13 (7), 181023,  DOI: 10.1021/acschembio.8b00371
  59. 59
    Lawrence, S. A. Amines: Synthesis, Properties and Applications; Cambridge University Press, 2004.
  60. 60
    Yin, Q.; Shi, Y.; Wang, J.; Zhang, X. Direct Catalytic Asymmetric Synthesis of α-Chiral Primary Amines. Chem. Soc. Rev. 2020, 49 (17), 614153,  DOI: 10.1039/C9CS00921C
  61. 61
    Cabré, A.; Verdaguer, X.; Riera, A. Recent Advances in the Enantioselective Synthesis of Chiral Amines Via Transition Metal-Catalyzed Asymmetric Hydrogenation. Chem. Rev. 2022, 122 (1), 269339,  DOI: 10.1021/acs.chemrev.1c00496
  62. 62
    Wu, Z.; Liu, C.; Zhang, Z.; Zheng, R.; Zheng, Y. Amidase as a Versatile Tool in Amide-Bond Cleavage: From Molecular Features to Biotechnological Applications. Biotechnol. Adv. 2020, 43, 107574,  DOI: 10.1016/j.biotechadv.2020.107574
  63. 63
    Breuer, M.; Ditrich, K.; Habicher, T.; Hauer, B.; Keßeler, M.; Stürmer, R.; Zelinski, T. Industrial Methods for the Production of Optically Active Intermediates. Angew. Chem., Int. Ed. 2004, 43 (7), 788824,  DOI: 10.1002/anie.200300599
  64. 64
    Balkenhohl, F.; Ditrich, K.; Hauer, B.; Ladner, W. Optisch Aktive Amine Durch Lipase-Katalysierte Methoxyacetylierung. J. Prakt. Chem. Chem. Ztg. 1997, 339 (1), 3814,  DOI: 10.1002/prac.19973390166
  65. 65
    Gomm, A.; O’Reilly, E. Transaminases for Chiral Amine Synthesis. Curr. Opin. Chem. Biol. 2018, 43, 10612,  DOI: 10.1016/j.cbpa.2017.12.007
  66. 66
    Abrahamson, M. J.; Vázquez-Figueroa, E.; Woodall, N. B.; Moore, J. C.; Bommarius, A. S. Development of an Amine Dehydrogenase for Synthesis of Chiral Amines. Angew. Chem., Int. Ed. 2012, 51 (16), 396972,  DOI: 10.1002/anie.201107813
  67. 67
    Li, B.-B.; Zhang, J.; Chen, F.-F.; Chen, Q.; Xu, J.-H.; Zheng, G.-W. Direct Reductive Amination of Ketones with Amines by Reductive Aminases. Green Synth. Catal. 2021, 2 (4), 3459,  DOI: 10.1016/j.gresc.2021.08.005
  68. 68
    Lenz, M.; Borlinghaus, N.; Weinmann, L.; Nestl, B. M. Recent Advances in Imine Reductase-Catalyzed Reactions. World J. Microbiol. Biotechnol. 2017, 33 (11), 199,  DOI: 10.1007/s11274-017-2365-8
  69. 69
    Gaweska, H.; Fitzpatrick, P. F. Structures and Mechanism of the Monoamine Oxidase Family. Biomol Concepts 2011, 2 (5), 36577,  DOI: 10.1515/BMC.2011.030
  70. 70
    Parmeggiani, F.; Weise, N. J.; Ahmed, S. T.; Turner, N. J. Synthetic and Therapeutic Applications of Ammonia-Lyases and Aminomutases. Chem. Rev. 2018, 118 (1), 73118,  DOI: 10.1021/acs.chemrev.6b00824
  71. 71
    Savile, C. K.; Janey, J. M.; Mundorff, E. C.; Moore, J. C.; Tam, S.; Jarvis, W. R.; Colbeck, J. C.; Krebber, A.; Fleitz, F. J.; Brands, J. Biocatalytic Asymmetric Synthesis of Chiral Amines from Ketones Applied to Sitagliptin Manufacture. Science 2010, 329 (5989), 3059,  DOI: 10.1126/science.1188934
  72. 72
    Weiß, M. S.; Pavlidis, I. V.; Spurr, P.; Hanlon, S. P.; Wirz, B.; Iding, H.; Bornscheuer, U. T. Protein-Engineering of an Amine Transaminase for the Stereoselective Synthesis of a Pharmaceutically Relevant Bicyclic Amine. Org. Biomol. Chem. 2016, 14 (43), 1024954,  DOI: 10.1039/C6OB02139E
  73. 73
    Ma, E. J.; Siirola, E.; Moore, C.; Kummer, A.; Stoeckli, M.; Faller, M.; Bouquet, C.; Eggimann, F.; Ligibel, M.; Huynh, D. Machine-Directed Evolution of an Imine Reductase for Activity and Stereoselectivity. ACS Catal. 2021, 11 (20), 1243345,  DOI: 10.1021/acscatal.1c02786
  74. 74
    Ao, Y.-F.; Hu, H.-J.; Zhao, C.-X.; Chen, P.; Huang, T.; Chen, H.; Wang, Q.-Q.; Wang, D.-X.; Wang, M.-X. Reversal and Amplification of the Enantioselectivity of Biocatalytic Desymmetrization toward meso Heterocyclic Dicarboxamides Enabled by Rational Engineering of Amidase. ACS Catal. 2021, 11 (12), 69007,  DOI: 10.1021/acscatal.1c01220
  75. 75
    Wang, P. F.; Yep, A.; Kenyon, G. L.; McLeish, M. J. Using Directed Evolution to Probe the Substrate Specificity of Mandelamide Hydrolase. Protein Eng. Des. Sel. 2009, 22 (2), 10310,  DOI: 10.1093/protein/gzn073
  76. 76
    Thooft, A. M.; Cassaidy, K.; VanVeller, B. A Small Push-Pull Fluorophore for Turn-On Fluorescence. J. Org. Chem. 2017, 82 (17), 88427,  DOI: 10.1021/acs.joc.7b00939
  77. 77
    Annenkov, V. V.; Verkhozina, O. N.; Shishlyannikova, T. A.; Danilovtseva, E. N. Application of 4-Chloro-7-Nitrobenzo-2-Oxa-1,3-Diazole in Analysis: Fluorescent Dyes and Unexpected Reaction with Tertiary Amines. Anal. Biochem. 2015, 486, 513,  DOI: 10.1016/j.ab.2015.06.025
  78. 78
    Sanger, F. The Free Amino Groups of Insulin. Biochem. J. 1945, 39 (5), 50715,  DOI: 10.1042/bj0390507
  79. 79
    Ghosh, P. B.; Whitehouse, M. W. 7-Chloro-4-Nitrobenzo-2-Oxa-1,3-Diazole: A New Fluorigenic Reagent for Amino Acids and Other Amines. Biochem. J. 1968, 108 (1), 1556,  DOI: 10.1042/bj1080155
  80. 80
    Watanabe, Y.; Imai, K. High-Performance Liquid Chromatography and Sensitive Detection of Amino Acids Derivatized with 7-Fluoro-4-Nitrobenzo-2-Oxa-1,3-Diazole. Anal. Biochem. 1981, 116 (2), 4712,  DOI: 10.1016/0003-2697(81)90390-0
  81. 81
    Aly, H.; El-Shafie, A. S.; El-Azazy, M. Utilization of 7-Chloro-4-Nitrobenzo-2-Oxa-1,3-Diazole (NBD-Cl) for Spectrochemical Determination of L-Ornithine: A Multivariate Optimization-Assisted Approach. RSC Adv. 2019, 9 (38), 2210615,  DOI: 10.1039/C9RA03311D
  82. 82
    Walash, M. I.; Metwally, M. E. S.; Eid, M.; El-Shaheny, R. N. Validated Spectrophotometric Methods for Determination of Alendronate Sodium in Tablets through Nucleophilic Aromatic Substitution Reactions. Chem. Cent. J. 2012, 6, 25,  DOI: 10.1186/1752-153X-6-25
  83. 83
    Benson, S.; Fernandez, A.; Barth, N. D.; de Moliner, F.; Horrocks, M. H.; Herrington, C. S.; Abad, J. L.; Delgado, A.; Kelly, L.; Chang, Z. SCOTfluors: Small, Conjugatable, Orthogonal, and Tunable Fluorophores for in Vivo Imaging of Cell Metabolism. Angew. Chem., Int. Ed. 2019, 58 (21), 69115,  DOI: 10.1002/anie.201900465
  84. 84
    Satake, K.; Okuyama, T.; Ohashi, M.; Shinoda, T. The Spectrophotometric Determination of Amine, Amino Acid and Peptide with 2,4,6-Trinitrobenzene 1-Sulfonic Acid. J. Biochem. 1960, 47 (5), 65460,  DOI: 10.1093/oxfordjournals.jbchem.a127107
  85. 85
    Means, G. E.; Congdon, W. I.; Bender, M. L. Reactions of 2,4,6-Trinitrobenzenesulfonate Ion with Amines and Hydroxide Ion. Biochemistry 1972, 11 (19), 356471,  DOI: 10.1021/bi00769a011
  86. 86
    Snyder, S. L.; Sobocinski, P. Z. An Improved 2,4,6-Trinitrobenzenesulfonic Acid Method for the Determination of Amines. Anal. Biochem. 1975, 64 (1), 2848,  DOI: 10.1016/0003-2697(75)90431-5
  87. 87
    Qi, X.-Y.; Keyhani, N. O.; Lee, Y. C. Spectrophotometric Determination of Hydrazine, Hydrazides, and Their Mixtures with Trinitrobenzenesulfonic Acid. Anal. Biochem. 1988, 175 (1), 13944,  DOI: 10.1016/0003-2697(88)90371-5
  88. 88
    Morçöl, T.; Subramanian, A.; Velander, W. H. Dot-Blot Analysis of the Degree of Covalent Modification of Proteins and Antibodies at Amino Groups. J. Immunol. Methods 1997, 203 (1), 4553,  DOI: 10.1016/S0022-1759(97)00013-6
  89. 89
    Oliverio, R.; Liberelle, B.; Murschel, F.; Garcia-Ac, A.; Banquy, X.; De Crescenzo, G. Versatile and High-Throughput Strategy for the Quantification of Proteins Bound to Nanoparticles. ACS Appl. Nano Mater. 2020, 3 (10), 10497507,  DOI: 10.1021/acsanm.0c02414
  90. 90
    Kim, D.; Cho, S. W.; Jun, Y. W.; Ahn, K. H. Fluorescent Labeling of Lysine Residues in Protein Using 8-Thiomethyl-BODIPY. Bull. Korean Chem. Soc. 2017, 38 (9), 9956,  DOI: 10.1002/bkcs.11213
  91. 91
    Bell, P. J. L.; Karuso, P. Epicocconone, a Novel Fluorescent Compound from the Fungus Epicoccum Nigrum. J. Am. Chem. Soc. 2003, 125 (31), 93045,  DOI: 10.1021/ja035496+
  92. 92
    Coghlan, D. R.; Mackintosh, J. A.; Karuso, P. Mechanism of Reversible Fluorescent Staining of Protein with Epicocconone. Org. Lett. 2005, 7 (12), 24014,  DOI: 10.1021/ol050665b
  93. 93
    Peixoto, P. A.; Boulangé, A.; Ball, M.; Naudin, B.; Alle, T.; Cosette, P.; Karuso, P.; Franck, X. Design and Synthesis of Epicocconone Analogues with Improved Fluorescence Properties. J. Am. Chem. Soc. 2014, 136 (43), 1524856,  DOI: 10.1021/ja506914p
  94. 94
    Chatterjee, S.; Karuso, P.; Boulangé, A.; Franck, X.; Datta, A. Excited State Dynamics of Brightly Fluorescent Second Generation Epicocconone Analogues. J. Phys. Chem. B 2015, 119 (20), 6295303,  DOI: 10.1021/acs.jpcb.5b02190
  95. 95
    Castell, J. V.; Cervera, M.; Marco, R. A Convenient Micromethod for the Assay of Primary Amines and Proteins with Fluorescamine. A Reexamination of the Conditions of Reaction. Anal. Biochem. 1979, 99 (2), 37991,  DOI: 10.1016/S0003-2697(79)80022-6
  96. 96
    Funk, G. M.; Hunt, C. E.; Epps, D. E.; Brown, P. K. Use of a Rapid and Highly Sensitive Fluorescamine-Based Procedure for the Assay of Plasma Lipoproteins. J. Lipid Res. 1986, 27, 792,  DOI: 10.1016/S0022-2275(20)38803-9
  97. 97
    Udenfriend, S.; Stein, S.; Böhlen, P.; Dairman, W.; Leimgruber, W.; Weigele, M. Fluorescamine: A Reagent for Assay of Amino Acids, Peptides, Proteins, and Primary Amines in the Picomole Range. Science 1972, 178 (4063), 871,  DOI: 10.1126/science.178.4063.871
  98. 98
    Murugayah, S. A.; Warring, S. L.; Gerth, M. L. Optimisation of a High-Throughput Fluorescamine Assay for Detection of N-Acyl-L-Homoserine Lactone Acylase Activity. Anal. Biochem. 2019, 566, 102,  DOI: 10.1016/j.ab.2018.10.029
  99. 99
    Ashby, J.; Duan, Y.; Ligans, E.; Tamsi, M.; Zhong, W. High-Throughput Profiling of Nanoparticle-Protein Interactions by Fluorescamine Labeling. Anal. Chem. 2015, 87 (4), 22139,  DOI: 10.1021/ac5036814
  100. 100
    Elbashir, A. A.; Ahmed, A. A.; Ali Ahmed, S. M.; Aboul-Enein, H. Y. 1,2-Naphthoquinone-4-Sulphonic Acid Sodium Salt (NQS) as an Analytical Reagent for the Determination of Pharmaceutical Amine by Spectrophotometry. Appl. Spectrosc. Rev. 2012, 47 (3), 21932,  DOI: 10.1080/05704928.2011.639107
  101. 101
    Rosenblatt, D. H.; Hlinka, P.; Epstein, J. Use of 1,2-Naphthoquinone-4-Sulfonate for Estimation of Ethylenimine and Primary Amines. Anal. Chem. 1955, 27 (8), 12903,  DOI: 10.1021/ac60104a024
  102. 102
    Darwish, I. A.; Abdine, H. H.; Amer, S. M.; Al-Rayes, L. I. Simple Spectrophotometric Method for Determination of Paroxetine in Tablets Using 1,2-Naphthoquinone-4-Sulphonate as a Chromogenic Reagent. Int. J. Anal. Chem. 2009, 2009, 237601,  DOI: 10.1155/2009/237601
  103. 103
    Longstreet, A. R.; Jo, M.; Chandler, R. R.; Hanson, K.; Zhan, N.; Hrudka, J. J.; Mattoussi, H.; Shatruk, M.; McQuade, D. T. Ylidenemalononitrile Enamines as Fluorescent “Turn-On” Indicators for Primary Amines. J. Am. Chem. Soc. 2014, 136 (44), 154936,  DOI: 10.1021/ja509058u
  104. 104
    Sathiskumar, U.; Easwaramoorthi, S. Red-Emitting Ratiometric Fluorescence Chemodosimeter for the Discriminative Detection of Aromatic and Aliphatic Amines. ChemistrySelect 2019, 4 (25), 748694,  DOI: 10.1002/slct.201901254
  105. 105
    Turiák, G.; Volicer, L. Stability of o-Phthalaldehyde─Sulfite Derivatives of Amino Acids and Their Methyl Esters: Electrochemical and Chromatographic Properties. J. Chromatogr. A 1994, 668 (2), 3239,  DOI: 10.1016/0021-9673(94)80121-5
  106. 106
    De Montigny, P.; Stobaugh, J. F.; Givens, R. S.; Carlson, R. G.; Srinivasachar, K.; Sternson, L. A.; Higuchi, T. Naphthalene-2,3-Dicarboxyaldehyde/Cyanide Ion: A Rationally Designed Fluorogenic Reagent for Primary Amines. Anal. Chem. 1987, 59 (8), 1096101,  DOI: 10.1021/ac00135a007
  107. 107
    Mroz, E. A.; Roman, R. J.; Lechene, C. Fluorescence Assay for Picomole Quantities of Ammonia. Kidney Int. 1982, 21 (3), 5247,  DOI: 10.1038/ki.1982.56
  108. 108
    Carlson, R. G.; Srinivasachar, K.; Givens, R. S.; Matuszewski, B. K. New Derivatizing Agents for Amino Acids and Peptides. 1. Facile Synthesis of N-Substituted 1-Cyanobenz[f]Isoindoles and Their Spectroscopic Properties. J. Org. Chem. 1986, 51 (21), 397883,  DOI: 10.1021/jo00371a013
  109. 109
    Bantan-Polak, T.; Kassai, M.; Grant, K. B. A Comparison of Fluorescamine and Naphthalene-2,3-Dicarboxaldehyde Fluorogenic Reagents for Microplate-Based Detection of Amino Acids. Anal. Biochem. 2001, 297 (2), 12836,  DOI: 10.1006/abio.2001.5338
  110. 110
    Hapuarachchi, S.; Aspinwall, C. A. Design, Characterization, and Utilization of a Fast Fluorescence Derivatization Reaction Utilizing o-Phthaldialdehyde Coupled with Fluorescent Thiols. Electrophoresis 2007, 28 (7), 11006,  DOI: 10.1002/elps.200600567
  111. 111
    Zhang, L.-Y.; Liu, Y.-M.; Wang, Z.-L.; Cheng, J.-K. Capillary Zone Electrophoresis with Pre-Column NDA Derivatization and Amperometric Detection for the Analysis of Four Aliphatic Diamines. Anal. Chim. Acta 2004, 508 (2), 1415,  DOI: 10.1016/j.aca.2003.11.075
  112. 112
    Fekete, A.; Lahaniatis, M.; Lintelmann, J.; Schmitt-Kopplin, P. Determination of Aliphatic Low-Molecular-Weight and Biogenic Amines by Capillary Zone Electrophoresis. In Capillary Electrophoresis: Methods and Protocols; Humana Press: Totowa, NJ, 2008; pp 6591. DOI: 10.1007/978-1-59745-376-9_4
  113. 113
    Böhmer, A.; Jordan, J.; Tsikas, D. High-Performance Liquid Chromatography Ultraviolet Assay for Human Erythrocytic Catalase Activity by Measuring Glutathione as o-Phthalaldehyde Derivative. Anal. Biochem. 2011, 410 (2), 296303,  DOI: 10.1016/j.ab.2010.11.026
  114. 114
    Black, G. W.; Brown, N. L.; Perry, J. J. B.; Randall, P. D.; Turnbull, G.; Zhang, M. A High-Throughput Screening Method for Determining the Substrate Scope of Nitrilases. Chem. Commun. 2015, 51 (13), 26602,  DOI: 10.1039/C4CC06021K
  115. 115
    Mann, S.; Eveleigh, L.; Lequin, O.; Ploux, O. A Microplate Fluorescence Assay for DAPA Aminotransferase by Detection of the Vicinal Diamine 7,8-Diaminopelargonic Acid. Anal. Biochem. 2013, 432 (2), 906,  DOI: 10.1016/j.ab.2012.09.038
  116. 116
    Kalkhof, S.; Sinz, A. Chances and Pitfalls of Chemical Cross-Linking with Amine-Reactive N-Hydroxysuccinimide Esters. Anal. Bioanal. Chem. 2008, 392 (1), 30512,  DOI: 10.1007/s00216-008-2231-5
  117. 117
    Gayo, L. M.; Suto, M. J. Use of Pentafluorophenyl Esters for One-Pot Protection/Activation of Amino and Thiol Carboxylic Acids. Tetrahedron Lett. 1996, 37 (28), 49158,  DOI: 10.1016/0040-4039(96)00985-9
  118. 118
    Fang, L.; Demee, M.; Sierra, T.; Kshirsagar, T.; Celebi, A. A.; Yan, B. Kinetics Study of Amine Cleavage Reactions of Various Resin-Bound Thiophenol Esters from Marshall Linker. J. Comb. Chem. 2002, 4 (4), 3628,  DOI: 10.1021/cc020010r
  119. 119
    Lavis, L. D. Teaching Old Dyes New Tricks: Biological Probes Built from Fluoresceins and Rhodamines. Annu. Rev. Biochem. 2017, 86, 82543,  DOI: 10.1146/annurev-biochem-061516-044839
  120. 120
    Pereira, A.; Martins, S.; Caldeira, A. T. Phytochemicals in Human Health, Coumarins as Fluorescent Labels of Biomolecules; IntechOpen, 2019.
  121. 121
    Boens, N.; Leen, V.; Dehaen, W. Fluorescent Indicators Based on BODIPY. Chem. Soc. Rev. 2012, 41 (3), 113072,  DOI: 10.1039/C1CS15132K
  122. 122
    Han, S.-Y.; Kim, Y.-A. Recent Development of Peptide Coupling Reagents in Organic Synthesis. Tetrahedron 2004, 60 (11), 244767,  DOI: 10.1016/j.tet.2004.01.020
  123. 123
    Karongo, R.; Ge, M.; Horak, J.; Gross, H.; Kohout, M.; Lindner, W.; Lämmerhofer, M. Rapid Enantioselective Amino Acid Analysis by Ultra-High Performance Liquid Chromatography-Mass Spectrometry Combining 6-Aminoquinolyl-N-Hydroxysuccinimidyl Carbamate Derivatization with Core-Shell Quinine Carbamate Anion Exchanger Separation. J. Chromatogr. Open 2021, 1, 100004,  DOI: 10.1016/j.jcoa.2021.100004
  124. 124
    Cao, L.; Wang, H.; Zhang, H. Analytical Potential of 6-Oxy-(N-Succinimidyl Acetate)-9-(2’-Methoxycarbonyl) Fluorescein for the Determination of Amino Compounds by Capillary Electrophoresis with Laser-Induced Fluorescence Detection. Electrophoresis 2005, 26 (10), 195462,  DOI: 10.1002/elps.200410227
  125. 125
    Xia, L.-J.; Guo, X.-F.; Ji, Y.; Chen, L.; Wang, H. A Long-Wavelength Fluorescent Probe for Amino Compounds and Its Application in the Determination of Aliphatic Amines. Anal. Methods 2018, 10 (26), 318896,  DOI: 10.1039/C8AY00706C
  126. 126
    Gao, P. F.; Guo, X. F.; Wang, H.; Zhang, H. S. Determination of Trace Biogenic Amines with 1,3,5,7-Tetramethyl-8-(N-Hydroxysuccinimidyl Butyric Ester)-Difluoroboradiaza-s-Indacene Derivatization Using High-Performance Liquid Chromatography and Fluorescence Detection. J. Sep. Sci. 2011, 34 (12), 138390,  DOI: 10.1002/jssc.201100120
  127. 127
    Mallick, S.; Chandra, F.; Koner, A. L. A Ratiometric Fluorescent Probe for Detection of Biogenic Primary Amines with Nanomolar Sensitivity. Analyst 2016, 141 (3), 82731,  DOI: 10.1039/C5AN01911G
  128. 128
    Vázquez, M. E.; Blanco, J. B.; Imperiali, B. Photophysics and Biological Applications of the Environment-Sensitive Fluorophore 6-N,N-Dimethylamino-2,3-Naphthalimide. J. Am. Chem. Soc. 2005, 127 (4), 13006,  DOI: 10.1021/ja0449168
  129. 129
    Anumula, K. R.; Schulz, R. P.; Back, N. Fluorescent N-Methylanthranilyl (Mantyl) Tag for Peptides: Its Application in Subpicomole Determination of Kinins. Peptides 1992, 13 (4), 6639,  DOI: 10.1016/0196-9781(92)90170-8
  130. 130
    Fessler, A.; Garmon, C.; Heavey, T.; Fowler, A.; Ogle, C. Water-Soluble and UV Traceable Isatoic Anhydride-Based Reagents for Bioconjugation. Org. Biomol. Chem. 2017, 15 (45), 9599602,  DOI: 10.1039/C7OB02377D
  131. 131
    Jeon, S.; Kim, T.-I.; Jin, H.; Lee, U.; Bae, J.; Bouffard, J.; Kim, Y. Amine-Reactive Activated Esters of meso-CarboxyBODIPY: Fluorogenic Assays and Labeling of Amines, Amino Acids, and Proteins. J. Am. Chem. Soc. 2020, 142 (20), 92319,  DOI: 10.1021/jacs.9b13982
  132. 132
    Lee, U.; Kim, T.-I.; Jeon, S.; Luo, Y.; Cho, S.; Bae, J.; Kim, Y. Native Chemical Ligation-Based Fluorescent Probes for Cysteine and Aminopeptidase N Using meso-Thioester-BODIPY. Eur. J. Chem. 2021, 27 (49), 1254551,  DOI: 10.1002/chem.202101990
  133. 133
    Esnal, I.; Bañuelos, J.; López Arbeloa, I.; Costela, A.; Garcia-Moreno, I.; Garzón, M.; Agarrabeitia, A. R.; José Ortiz, M. Nitro and Amino BODIPYs: Crucial Substituents to Modulate Their Photonic Behavior. RSC Adv. 2013, 3 (5), 154756,  DOI: 10.1039/C2RA22916A
  134. 134
    Liu, E. Y.; Jung, S.; Weitz, D. A.; Yi, H.; Choi, C.-H. High-Throughput Double Emulsion-Based Microfluidic Production of Hydrogel Microspheres with Tunable Chemical Functionalities toward Biomolecular Conjugation. Lab Chip 2018, 18 (2), 32334,  DOI: 10.1039/C7LC01088E
  135. 135
    Katritzky, A. R.; Marson, C. M. Pyrylium Mediated Transformations of Primary Amino Groups into Other Functional Groups. New Synthetic Methods (41). Angew. Chem., Int. Ed. Engl. 1984, 23 (6), 4209,  DOI: 10.1002/anie.198404201
  136. 136
    Wetzl, B. K.; Yarmoluk, S. M.; Craig, D. B.; Wolfbeis, O. S. Chameleon Labels for Staining and Quantifying Proteins. Angew. Chem., Int. Ed. Engl. 2004, 43 (40), 54002,  DOI: 10.1002/anie.200460508
  137. 137
    Azab, H. A.; El-Korashy, S. A.; Anwar, Z. M.; Khairy, G. M.; Steiner, M.-S.; Duerkop, A. High-Throughput Sensing Microtiter Plate for Determination of Biogenic Amines in Seafood Using Fluorescence or Eye-Vision. Analyst 2011, 136 (21), 44929,  DOI: 10.1039/c1an15049a
  138. 138
    Azab, H. A.; El-Korashy, S. A.; Anwar, Z. M.; Khairy, G. M.; Duerkop, A. Reactivity of a Luminescent “Off-On” Pyrylium Dye toward Various Classes of Amines and Its Use in a Fluorescence Sensor Microtiter Plate for Environmental Samples. J. Photochem. Photobiol. A: Chem. 2012, 243, 416,  DOI: 10.1016/j.jphotochem.2012.05.029
  139. 139
    Bayer, M.; König, S. Pyrylium-Based Dye and Charge Tagging in Proteomics. Electrophoresis 2016, 37 (22), 29538,  DOI: 10.1002/elps.201600318
  140. 140
    Turner, E. H.; Dickerson, J. A.; Ramsay, L. M.; Swearingen, K. E.; Wojcik, R.; Dovichi, N. J. Reaction of Fluorogenic Reagents with Proteins: III. Spectroscopic and Electrophoretic Behavior of Proteins Labeled with Chromeo P503. J. Chromatogr. A 2008, 1194 (2), 2536,  DOI: 10.1016/j.chroma.2008.04.046
  141. 141
    Wolfbeis, O. S. Fluorescent Chameleon Labels for Bioconjugation and Imaging of Proteins, Nucleic Acids, Biogenic Amines and Surface Amino Groups. A Review. Methods Appl. Fluoresc. 2021, 9 (4), 042001,  DOI: 10.1088/2050-6120/ac1a0a
  142. 142
    Höfelschweiger, B. K., The Pyrylium Dyes: a New Class of Biolabels. Synthesis, Spectroscopy, and Application as Labels and in General Protein Assay , 2005.
  143. 143
    Walter, W.; Bode, K. D. Syntheses of Thiocarbamates. Angew. Chem., Int. Ed. Engl. 1967, 6 (4), 28193,  DOI: 10.1002/anie.196702811
  144. 144
    Maddani, M. R.; Prabhu, K. R. A Concise Synthesis of Substituted Thiourea Derivatives in Aqueous Medium. J. Org. Chem. 2010, 75 (7), 232732,  DOI: 10.1021/jo1001593
  145. 145
    Reetz, M. T.; Kühling, K. M.; Deege, A.; Hinrichs, H.; Belder, D. Super-High-Throughput Screening of Enantioselective Catalysts by Using Capillary Array Electrophoresis. Angew. Chem., Int. Ed. 2000, 39 (21), 38913,  DOI: 10.1002/1521-3773(20001103)39:21<3891::AID-ANIE3891>3.0.CO;2-1
  146. 146
    Dabek-Zlotorzynska, E.; Maruszak, W. Determination of Dimethylamine and Other Low-Molecular-Mass Amines Using Capillary Electrophoresis with Laser-Induced Fluorescence Detection. J. Chromatogr. B Biomed. Appl. 1998, 714 (1), 7785,  DOI: 10.1016/S0378-4347(98)00070-X
  147. 147
    Morales, A. R.; Schafer-Hales, K. J.; Marcus, A. I.; Belfield, K. D. Amine-Reactive Fluorene Probes: Synthesis, Optical Characterization, Bioconjugation, and Two-Photon Fluorescence Imaging. Bioconj. Chem. 2008, 19 (12), 255967,  DOI: 10.1021/bc800415t
  148. 148
    Planas, O.; Gallavardin, T.; Nonell, S. A Novel Fluoro-Chromogenic Click Reaction for the Labelling of Proteins and Nanoparticles with near-IR Theranostic Agents. Chem. Commun. 2015, 51 (26), 55869,  DOI: 10.1039/C4CC09070E
  149. 149
    Wang, R.; Yang, W.-j.; Yue, L.; Pan, W.; Zeng, H.-y. DDQ-Promoted C-S Bond Formation: Synthesis of 2-Aminobenzothiazole Derivatives under Transition-Metal-, Ligand-, and Base-Free Conditions. Synlett 2012, 23 (11), 16438,  DOI: 10.1055/s-0031-1291159
  150. 150
    Guo, Y.-J.; Tang, R.-Y.; Zhong, P.; Li, J.-H. Copper-Catalyzed Tandem Reactions of 2-Halobenzenamines with Isothiocyanates under Ligand- and Base-Free Conditions. Tetrahedron Lett. 2010, 51 (4), 64952,  DOI: 10.1016/j.tetlet.2009.11.086
  151. 151
    Sutton, J. M.; Clarke, O. J.; Fernandez, N.; Boyle, R. W. Porphyrin, Chlorin, and Bacteriochlorin Isothiocyanates: Useful Reagents for the Synthesis of Photoactive Bioconjugates. Bioconj. Chem. 2002, 13 (2), 24963,  DOI: 10.1021/bc015547x
  152. 152
    Feuster, E. K.; Glass, T. E. Detection of Amines and Unprotected Amino Acids in Aqueous Conditions by Formation of Highly Fluorescent Iminium Ions. J. Am. Chem. Soc. 2003, 125 (52), 161745,  DOI: 10.1021/ja036434m
  153. 153
    Patze, C.; Broedner, K.; Rominger, F.; Trapp, O.; Bunz, U. H. F. Aldehyde Cruciforms: Dosimeters for Primary and Secondary Amines. Eur. J. Chem. 2011, 17 (49), 137205,  DOI: 10.1002/chem.201101871
  154. 154
    Kumpf, J.; Bunz, U. H. F. Aldehyde-Appended Distyrylbenzenes: Amine Recognition in Water. Eur. J. Chem. 2012, 18 (29), 89214,  DOI: 10.1002/chem.201200930
  155. 155
    Mohr, G. J.; Demuth, C.; Spichiger-Keller, U. E. Application of Chromogenic and Fluorogenic Reactands in the Optical Sensing of Dissolved Aliphatic Amines. Anal. Chem. 1998, 70 (18), 386873,  DOI: 10.1021/ac980279q
  156. 156
    Mohr, G. J. Tailoring the Sensitivity and Spectral Properties of a Chromoreactand for the Detection of Amines and Alcohols. Anal. Chim. Acta 2004, 508 (2), 2337,  DOI: 10.1016/j.aca.2003.12.005
  157. 157
    Mohr, G. J. Chromo- and Fluororeactands: Indicators for Detection of Neutral Analytes by Using Reversible Covalent-Bond Chemistry. Eur. J. Chem. 2004, 10 (5), 108290,  DOI: 10.1002/chem.200305524
  158. 158
    Mohr, G. J.; Grummt, U.-W. Comparison of Trifluoroacetyl Monostyryl and Distyryl Dyes: Effects of Chromophore Elongation on the Spectral Properties and Chemical Reactivity. J. Fluoresc. 2006, 16 (2), 18590,  DOI: 10.1007/s10895-005-0047-7
  159. 159
    Mertz, E.; Beil, J. B.; Zimmerman, S. C. Kinetics and Thermodynamics of Amine and Diamine Signaling by a Trifluoroacetyl Azobenzene Reporter Group. Org. Lett. 2003, 5 (17), 312730,  DOI: 10.1021/ol0351605
  160. 160
    Körsten, S.; Mohr, G. J. Star-Shaped Tripodal Chemosensors for the Detection of Aliphatic Amines. Eur. J. Chem. 2011, 17 (3), 96975,  DOI: 10.1002/chem.201000787
  161. 161
    Xu, Y.; Yu, S.; Wang, Y.; Hu, L.; Zhao, F.; Chen, X.; Li, Y.; Yu, X.; Pu, L. Ratiometric Fluorescence Sensors for 1,2-Diamines Based on Trifluoromethyl Ketones. Eur. J. Org. Chem. 2016, 2016 (35), 586875,  DOI: 10.1002/ejoc.201601157
  162. 162
    Ali, M. F. B.; Kishikawa, N.; Ohyama, K.; Mohamed, H. A.-M.; Abdel-Wadood, H. M.; Mahmoud, A. M.; Imazato, T.; Ueki, Y.; Wada, M.; Kuroda, N. Chromatographic Determination of Low-Molecular Mass Unsaturated Aliphatic Aldehydes with Peroxyoxalate Chemiluminescence Detection after Fluorescence Labeling with 4-(N,N-Dimethylaminosulfonyl)-7-Hydrazino-2,1,3-Benzoxadiazole. J. Chromatogr. B 2014, 953–954, 14752,  DOI: 10.1016/j.jchromb.2014.02.009
  163. 163
    Xu, Y.; Yu, S.; Chen, Q.; Chen, X.; Li, Y.; Yu, X.; Pu, L. Fluorescent Recognition of 1,2-Diamines by a 1,1′-Binaphthyl-Based Trifluoromethyl Ketone. Eur. J. Chem. 2016, 22 (34), 120617,  DOI: 10.1002/chem.201601540
  164. 164
    Collados, J. F.; Solà, R.; Harutyunyan, S. R.; Maciá, B. Catalytic Synthesis of Enantiopure Chiral Alcohols Via Addition of Grignard Reagents to Carbonyl Compounds. ACS Catal. 2016, 6 (3), 195270,  DOI: 10.1021/acscatal.5b02832
  165. 165
    Malic, N.; Moorhoff, C.; Sage, V.; Saylik, D.; Teoh, E.; Scott, J. L.; Strauss, C. R. Toward Preparative Resolution of Chiral Alcohols by an Organic Chemical Method. New J. Chem. 2010, 34 (3), 398402,  DOI: 10.1039/b9nj00768g
  166. 166
    Rong, J.; Pellegrini, T.; Harutyunyan, S. R. Synthesis of Chiral Tertiary Alcohols by CuI-Catalyzed Enantioselective Addition of Organomagnesium Reagents to Ketones. Eur. J. Chem. 2016, 22 (11), 355870,  DOI: 10.1002/chem.201503412
  167. 167
    Chen, B.-S.; Ribeiro de Souza, F. Z. Enzymatic Synthesis of Enantiopure Alcohols: Current State and Perspectives. RSC Adv. 2019, 9 (4), 210215,  DOI: 10.1039/C8RA09004A
  168. 168
    Kourist, R.; Dominguez de Maria, P.; Bornscheuer, U. T. Enzymatic Synthesis of Optically Active Tertiary Alcohols: Expanding the Biocatalysis Toolbox. ChemBioChem. 2008, 9 (4), 4918,  DOI: 10.1002/cbic.200700688
  169. 169
    Jiang, H.; Su, X.; Zhang, Y.; Zhou, J.; Fang, D.; Wang, X. Unexpected Thiols Triggering Photoluminescent Enhancement of Cytidine Stabilized Au Nanoclusters for Sensitive Assays of Glutathione Reductase and Its Inhibitors Screening. Anal. Chem. 2016, 88 (9), 476671,  DOI: 10.1021/acs.analchem.6b00112
  170. 170
    Andexer, J.; von Langermann, J.; Mell, A.; Bocola, M.; Kragl, U.; Eggert, T.; Pohl, M. An R-Selective Hydroxynitrile Lyase from Arabidopsis Thaliana with an α/β-Hydrolase Fold. Angew. Chem., Int. Ed. 2007, 46 (45), 867981,  DOI: 10.1002/anie.200701455
  171. 171
    Zheng, Y.-G.; Yin, H.-H.; Yu, D.-F.; Chen, X.; Tang, X.-L.; Zhang, X.-J.; Xue, Y.-P.; Wang, Y.-J.; Liu, Z.-Q. Recent Advances in Biotechnological Applications of Alcohol Dehydrogenases. Appl. Microbiol. Biotechnol. 2017, 101 (3), 9871001,  DOI: 10.1007/s00253-016-8083-6
  172. 172
    Huisman, G. W.; Liang, J.; Krebber, A. Practical Chiral Alcohol Manufacture Using Ketoreductases. Curr. Opin. Chem. Biol. 2010, 14 (2), 1229,  DOI: 10.1016/j.cbpa.2009.12.003
  173. 173
    Iding, H.; Siegert, P.; Mesch, K.; Pohl, M. Application of α-Keto Acid Decarboxylases in Biotransformations. Biochim. Biophys. Acta 1998, 1385 (2), 30722,  DOI: 10.1016/S0167-4838(98)00076-4
  174. 174
    Lee, S.-H.; Yeom, S.-J.; Kim, S.-E.; Oh, D.-K. Development of Aldolase-Based Catalysts for the Synthesis of Organic Chemicals. Trends Biotechnol. 2022, 40, 306,  DOI: 10.1016/j.tibtech.2021.08.001
  175. 175
    Bracco, P.; Busch, H.; von Langermann, J.; Hanefeld, U. Enantioselective Synthesis of Cyanohydrins Catalysed by Hydroxynitrile Lyases - a Review. Org. Biomol. Chem. 2016, 14 (27), 637589,  DOI: 10.1039/C6OB00934D
  176. 176
    Chen, B.-S.; Otten, L. G.; Hanefeld, U. Stereochemistry of Enzymatic Water Addition to C=C Bonds. Biotechnol. Adv. 2015, 33 (5), 52646,  DOI: 10.1016/j.biotechadv.2015.01.007
  177. 177
    Archelas, A.; Furstoss, R. Synthetic Applications of Epoxide Hydrolases. Curr. Opin. Chem. Biol. 2001, 5 (2), 1129,  DOI: 10.1016/S1367-5931(00)00179-4
  178. 178
    Freakley, S. J.; Kochius, S.; van Marwijk, J.; Fenner, C.; Lewis, R. J.; Baldenius, K.; Marais, S. S.; Opperman, D. J.; Harrison, S. T. L.; Alcalde, M.; Smit, M. S.; Hutchings, G. J. A Chemo-Enzymatic Oxidation Cascade to Activate C-H Bonds with in Situ Generated H2O2. Nat. Commun. 2019, 10 (1), 4178,  DOI: 10.1038/s41467-019-12120-w
  179. 179
    Wang, Y.; Xiang, Q.; Zhou, Q.; Xu, J.; Pei, D. Mini Review: Advances in 2-Haloacid Dehalogenases. Front. Microbiol. 2021, 12, 758886,  DOI: 10.3389/fmicb.2021.758886
  180. 180
    Koudelakova, T.; Bidmanova, S.; Dvorak, P.; Pavelka, A.; Chaloupkova, R.; Prokop, Z.; Damborsky, J. Haloalkane Dehalogenases: Biotechnological Applications. Biotechnol. J. 2013, 8 (1), 3245,  DOI: 10.1002/biot.201100486
  181. 181
    Schallmey, A.; Schallmey, M. Recent Advances on Halohydrin Dehalogenases-from Enzyme Identification to Novel Biocatalytic Applications. Appl. Microbiol. Biotechnol. 2016, 100 (18), 782739,  DOI: 10.1007/s00253-016-7750-y
  182. 182
    Belafriekh, A.; Secundo, F.; Serra, S.; Djeghaba, Z. Enantioselective Enzymatic Resolution of Racemic Alcohols by Lipases in Green Organic Solvents. Tetrahedron: Asymmetry 2017, 28 (3), 4738,  DOI: 10.1016/j.tetasy.2017.02.004
  183. 183
    Bustos-Jaimes, I.; Hummel, W.; Eggert, T.; Bogo, E.; Puls, M.; Weckbecker, A.; Jaeger, K.-E. A High-Throughput Screening Method for Chiral Alcohols and Its Application to Determine Enantioselectivity of Lipases and Esterases. ChemCatChem. 2009, 1 (4), 4458,  DOI: 10.1002/cctc.200900190
  184. 184
    Chen, B.-S.; Resch, V.; Otten, L. G.; Hanefeld, U. Enantioselective Michael Addition of Water. Eur. J. Chem. 2015, 21 (7), 302030,  DOI: 10.1002/chem.201405579
  185. 185
    Lehwald, P.; Richter, M.; Röhr, C.; Liu, H.-w.; Müller, M. Enantioselective Intermolecular Aldehyde-Ketone Cross-Coupling through an Enzymatic Carboligation Reaction. Angew. Chem., Int. Ed. 2010, 49 (13), 238992,  DOI: 10.1002/anie.200906181
  186. 186
    Hammer, S. C.; Marjanovic, A.; Dominicus, J. M.; Nestl, B. M.; Hauer, B. Squalene Hopene Cyclases Are Protonases for Stereoselective Brønsted Acid Catalysis. Nat. Chem. Biol. 2015, 11 (2), 1216,  DOI: 10.1038/nchembio.1719
  187. 187
    Loskot, S. A.; Romney, D. K.; Arnold, F. H.; Stoltz, B. M. Enantioselective Total Synthesis of Nigelladine a Via Late-Stage C-H Oxidation Enabled by an Engineered P450 Enzyme. J. Am. Chem. Soc. 2017, 139 (30), 101969,  DOI: 10.1021/jacs.7b05196
  188. 188
    Zanon, J. P.; Peres, M. F. S.; Gattás, E. A. L. Colorimetric Assay of Ethanol Using Alcohol Dehydrogenase from Dry Baker’s Yeast. Enzyme Microb. Technol. 2007, 40 (3), 46670,  DOI: 10.1016/j.enzmictec.2006.07.029
  189. 189
    Baker, J. L.; Faustoferri, R. C.; Quivey, R. G., Jr. A Modified Chromogenic Assay for Determination of the Ratio of Free Intracellular NAD(+)/NADH in Streptococcus Mutans. Bio Protoc. 2016, 6 (16), e1902,  DOI: 10.21769/BioProtoc.1902
  190. 190
    Yang, Y.; Liu, J.; Li, Z. Engineering of P450pyr Hydroxylase for the Highly Regio- and Enantioselective Subterminal Hydroxylation of Alkanes. Angew. Chem., Int. Ed. 2014, 53 (12), 31204,  DOI: 10.1002/anie.201311091
  191. 191
    Lauchli, R.; Rabe, K. S.; Kalbarczyk, K. Z.; Tata, A.; Heel, T.; Kitto, R. Z.; Arnold, F. H. High-Throughput Screening for Terpene-Synthase-Cyclization Activity and Directed Evolution of a Terpene Synthase. Angew. Chem., Int. Ed. 2013, 52 (21), 55714,  DOI: 10.1002/anie.201301362
  192. 192
    Jung, E.; Park, B. G.; Yoo, H.-W.; Kim, J.; Choi, K.-Y.; Kim, B.-G. Semi-Rational Engineering of CYP153A35 to Enhance Ω-Hydroxylation Activity toward Palmitic Acid. Appl. Microbiol. Biotechnol. 2018, 102 (1), 26977,  DOI: 10.1007/s00253-017-8584-y
  193. 193
    Francisco, W. A.; Abu-Soud, H. M.; Baldwin, T. O.; Raushel, F. M. Interaction of Bacterial Luciferase with Aldehyde Substrates and Inhibitors. J. Biol. Chem. 1993, 268 (33), 2473441,  DOI: 10.1016/S0021-9258(19)74526-8
  194. 194
    Minak-Bernero, V.; Bare, R. E.; Haith, C. E.; Grossman, M. J. Detection of Alkanes, Alcohols, and Aldehydes Using Bioluminescence. Biotechnol. Bioeng. 2004, 87 (2), 1707,  DOI: 10.1002/bit.20089
  195. 195
    Greene, L. E.; Lincoln, R.; Krumova, K.; Cosa, G. Development of a Fluorogenic Reactivity Palette for the Study of Nucleophilic Addition Reactions Based on meso-Formyl BODIPY Dyes. ACS Omega 2017, 2 (12), 861824,  DOI: 10.1021/acsomega.7b01795
  196. 196
    Lincoln, R.; Greene, L. E.; Bain, C.; Flores-Rizo, J. O.; Bohle, D. S.; Cosa, G. When Push Comes to Shove: Unravelling the Mechanism and Scope of Nonemissive meso-Unsaturated BODIPY Dyes. J. Phys. Chem. B 2015, 119 (13), 475865,  DOI: 10.1021/acs.jpcb.5b02080
  197. 197
    Sawminathan, S.; Iyer, S. K. A New Imidazole Based Phenanthridine Probe for Ratiometric Fluorescence Monitoring of Methanol in Biodiesel. New J. Chem. 2021, 45 (13), 603341,  DOI: 10.1039/D0NJ06252A
  198. 198
    Mohr, G. J.; Spichiger-Keller, U. E. Novel Fluorescent Sensor Membranes for Alcohols Based on p-N,N-Dioctylamino-4′-Trifluoroacetylstilbene. Anal. Chim. Acta 1997, 351 (1), 18996,  DOI: 10.1016/S0003-2670(97)00365-6
  199. 199
    Mohr, G. J.; Lehmann, F.; Grummt, U.-W.; Spichiger-Keller, U. E. Fluorescent Ligands for Optical Sensing of Alcohols: Synthesis and Characterisation of p-N,N-Dialkylamino-Trifluoroacetylstilbenes. Anal. Chim. Acta 1997, 344 (3), 21525,  DOI: 10.1016/S0003-2670(97)00113-X
  200. 200
    Takano, K.; Sasaki, S.-i.; Citterio, D.; Tamiaki, H.; Suzuki, K. An Oxo-Bacteriochlorin Derivative for Long-Wavelength Fluorescence Ratiometric Alcohol Sensing. Analyst 2010, 135 (9), 23349,  DOI: 10.1039/c0an00173b
  201. 201
    Zhao, M.; Yue, Y.; Liu, C.; Hui, P.; He, S.; Zhao, L.; Zeng, X. A Colorimetric and Fluorometric Dual-Modal Sensor for Methanol Based on a Functionalized Pentacenequinone Derivative. Chem. Commun. 2018, 54 (60), 833942,  DOI: 10.1039/C8CC04515A
  202. 202
    Roy, S.; Das, S.; Ray, A.; Parui, P. P. Fluorometric Trace Methanol Detection in Ethanol and Isopropanol in a Water Medium for Application in Alcoholic Beverages and Hand Sanitizers. RSC Adv. 2021, 11 (48), 30093101,  DOI: 10.1039/D1RA05201B
  203. 203
    Wu, Z.; Fu, X.; Wang, Y. Click Synthesis of a Triphenylamine-Based Fluorescent Methanol Probe with a Unique D-Π-a Structure. Sens. Actuators B Chem. 2017, 245, 40613,  DOI: 10.1016/j.snb.2017.01.164
  204. 204
    Kumar, V.; Kumar, A.; Diwan, U.; Singh, M. K.; Upadhyay, K. K. A Radical Approach for Fluorescent Turn ‘On’ Detection, Differentiation and Bioimaging of Methanol. Org. Biomol. Chem. 2015, 13 (33), 88226,  DOI: 10.1039/C5OB01333J
  205. 205
    Kumar, V.; Kundu, S.; Sk, B.; Patra, A. A Naked-Eye Colorimetric Sensor for Methanol and ‘Turn-On’ Fluorescence Detection of Al3+. New J. Chem. 2019, 43 (47), 185829,  DOI: 10.1039/C9NJ04688G
  206. 206
    Patel, R. N. Synthesis of Chiral Pharmaceutical Intermediates by Biocatalysis. Coord. Chem. Rev. 2008, 252 (5), 659701,  DOI: 10.1016/j.ccr.2007.10.031
  207. 207
    Chen, Y.; Chen, C.; Wu, X. Dicarbonyl Reduction by Single Enzyme for the Preparation of Chiral Diols. Chem. Soc. Rev. 2012, 41 (5), 174253,  DOI: 10.1039/c1cs15230k
  208. 208
    Ramasastry, S. S. V.; Zhang, H.; Tanaka, F.; Barbas, C. F. Direct Catalytic Asymmetric Synthesis of anti-1,2-Amino Alcohols and syn-1,2-Diols through Organocatalytic anti-Mannich and syn-Aldol Reactions. J. Am. Chem. Soc. 2007, 129 (2), 2889,  DOI: 10.1021/ja0677012
  209. 209
    Guo, Z.; Shin, I.; Yoon, J. Recognition and Sensing of Various Species Using Boronic Acid Derivatives. Chem. Commun. 2012, 48 (48), 595667,  DOI: 10.1039/c2cc31985c
  210. 210
    Hudlicky, T.; Thorpe, A. J. Current Status and Future Perspectives of Cyclohexadiene-cis-Diols in Organic Synthesis: Versatile Intermediates in the Concise Design of Natural Products. Chem. Commun. 1996, 17 (17), 19932000,  DOI: 10.1039/cc9960001993
  211. 211
    Wingstrand, M. J.; Madsen, C. M.; Clausen, M. H. Rapid Synthesis of Macrocycles from Diol Precursors. Tetrahedron Lett. 2009, 50 (6), 6935,  DOI: 10.1016/j.tetlet.2008.11.100
  212. 212
    Tang, X.; Wei, J.; Ding, N.; Sun, Y.; Zeng, X.; Hu, L.; Liu, S.; Lei, T.; Lin, L. Chemoselective Hydrogenation of Biomass Derived 5-Hydroxymethylfurfural to Diols: Key Intermediates for Sustainable Chemicals, Materials and Fuels. Renew. Sust. Energy Rev. 2017, 77, 28796,  DOI: 10.1016/j.rser.2017.04.013
  213. 213
    Song, J.; Shao, P.-L.; Wang, J.; Huang, F.; Zhang, X. Asymmetric Hydrogenation of 1,4-Diketones: Facile Synthesis of Enantiopure 1,4-Diarylbutane-1,4-Diols. Chem. Commun. 2021, 58 (2), 2625,  DOI: 10.1039/D1CC05359K
  214. 214
    Faber, K.; Mischitz, M.; Kroutil, W. Microbial Epoxide Hydrolases. Acta Chem. Scand. 1996, 50, 249,  DOI: 10.3891/acta.chem.scand.50-0249
  215. 215
    Ouellette, R. J.; Rawn, J. D. Alcohols: Reactions and Synthesis. In Organic Chemistry, 2nd ed.; Academic Press, 2018; Chapter 16, pp 463505.
  216. 216
    Rinner, U. Chiral Pool Synthesis. In Chiral Pool Syntheses from cis-Cyclohexadiene Diols, Comprehensive Chirality; Elsevier: Amsterdam, 2012; Chapter 2.9, pp 24067.
  217. 217
    Bruice, P. Y. Organic Chemistry; 7. ed.; Pearson Education: Harlow, 2014.
  218. 218
    Gubbels, E.; Heitz, T.; Yamamoto, M.; Chilekar, V.; Zarbakhsh, S.; Gepraegs, M.; Köpnick, H.; Schmidt, M.; Brügging, W.; Rüter, J. Polyesters. Ullmann’s Encyclopedia of Industrial Chemistry Wiley VCH, 2018; 130 DOI: 10.1002/14356007.a21_227.pub2 .
  219. 219
    Steinreiber, A.; Faber, K. Microbial Epoxide Hydrolases for Preparative Biotransformations. Curr. Opin. Biotechnol. 2001, 12 (6), 5528,  DOI: 10.1016/S0958-1669(01)00262-2
  220. 220
    Gibson, D. T.; Parales, R. E. Aromatic Hydrocarbon Dioxygenases in Environmental Biotechnology. Curr. Opin. Biotechnol. 2000, 11 (3), 23643,  DOI: 10.1016/S0958-1669(00)00090-2
  221. 221
    Boyd, D. R.; Sharma, N. D.; Allen, C. C. R. Aromatic Dioxygenases: Molecular Biocatalysis and Applications. Curr. Opin. Biotechnol. 2001, 12 (6), 56473,  DOI: 10.1016/S0958-1669(01)00264-6
  222. 222
    Zhao, L.; Han, B.; Huang, Z.; Miller, M.; Huang, H.; Malashock, D. S.; Zhu, Z.; Milan, A.; Robertson, D. E.; Weiner, D. P. Epoxide Hydrolase-Catalyzed Enantioselective Synthesis of Chiral 1,2-Diols Via Desymmetrization of meso-Epoxides. J. Am. Chem. Soc. 2004, 126 (36), 111567,  DOI: 10.1021/ja0466210
  223. 223
    Makarova, M.; Endoma-Arias, M. A. A.; Dela Paz, H. E.; Simionescu, R.; Hudlicky, T. Chemoenzymatic Total Synthesis of ent-Oxycodone: Second-, Third-, and Fourth-Generation Strategies. J. Am. Chem. Soc. 2019, 141 (27), 10883904,  DOI: 10.1021/jacs.9b05033
  224. 224
    Ikeda, H.; Sato, E.; Sugai, T.; Ohta, H. Yeast-Mediated Synthesis of Optically Active Diols with C2-Symmetry and (R)-4-Pentanolide. Tetrahedron 1996, 52 (24), 811322,  DOI: 10.1016/0040-4020(96)00373-0
  225. 225
    Gijsen, H. J. M.; Wong, C.-H. Unprecedented Asymmetric Aldol Reactions with Three Aldehyde Substrates Catalyzed by 2-Deoxyribose-5-Phosphate Aldolase. J. Am. Chem. Soc. 1994, 116 (18), 84223,  DOI: 10.1021/ja00097a082
  226. 226
    Müller, M. Chemoenzymatic Synthesis of Building Blocks for Statin Side Chains. Angew. Chem., Int. Ed. 2005, 44 (3), 3625,  DOI: 10.1002/anie.200460852
  227. 227
    Wu, X.; Jiang, J.; Chen, Y. Correlation between Intracellular Cofactor Concentrations and Biocatalytic Efficiency: Coexpression of Diketoreductase and Glucose Dehydrogenase for the Preparation of Chiral Diol for Statin Drugs. ACS Catal. 2011, 1 (12), 16614,  DOI: 10.1021/cs200408y
  228. 228
    Huffman, M. A.; Fryszkowska, A.; Alvizo, O.; Borra-Garske, M.; Campos, K. R.; Canada, K. A.; Devine, P. N.; Duan, D.; Forstater, J. H.; Grosser, S. T. Design of an in Vitro Biocatalytic Cascade for the Manufacture of Islatravir. Science 2019, 366 (6470), 12559,  DOI: 10.1126/science.aay8484
  229. 229
    Bian, Z.; Liu, A.; Li, Y.; Fang, G.; Yao, Q.; Zhang, G.; Wu, Z. Boronic Acid Sensors with Double Recognition Sites: A Review. Analyst 2020, 145 (3), 71944,  DOI: 10.1039/C9AN00741E
  230. 230
    Cao, H.; Heagy, M. D. Fluorescent Chemosensors for Carbohydrates: A Decade’s Worth of Bright Spies for Saccharides in Review. J. Fluoresc. 2004, 14 (5), 56984,  DOI: 10.1023/B:JOFL.0000039344.34642.4c
  231. 231
    Fang, G.; Wang, H.; Bian, Z.; Sun, J.; Liu, A.; Fang, H.; Liu, B.; Yao, Q.; Wu, Z. Recent Development of Boronic Acid-Based Fluorescent Sensors. RSC Adv. 2018, 8 (51), 2940027,  DOI: 10.1039/C8RA04503H
  232. 232
    Wang, W.; Gao, X.; Wang, B. Boronic Acid-Based Sensors. Curr. Org. Chem. 2002, 6 (14), 1285317,  DOI: 10.2174/1385272023373446
  233. 233
    Wu, X.; Li, Z.; Chen, X.-X.; Fossey, J. S.; James, T. D.; Jiang, Y.-B. Selective Sensing of Saccharides Using Simple Boronic Acids and Their Aggregates. Chem. Soc. Rev. 2013, 42 (20), 803248,  DOI: 10.1039/c3cs60148j
  234. 234
    Arnaud, J.; Audfray, A.; Imberty, A. Binding Sugars: From Natural Lectins to Synthetic Receptors and Engineered Neolectins. Chem. Soc. Rev. 2013, 42 (11), 4798813,  DOI: 10.1039/c2cs35435g
  235. 235
    Wu, X.; Chen, X.-X.; Jiang, Y.-B. Recent Advances in Boronic Acid-Based Optical Chemosensors. Analyst 2017, 142 (9), 140314,  DOI: 10.1039/C7AN00439G
  236. 236
    Kondo, K.; Shiomi, Y.; Saisho, M.; Harada, T.; Shinkai, S. Specific Complexation of Disaccharides with Diphenyl-3,3′-Diboronic Acid That Can Be Detected by Circular Dichroism. Tetrahedron 1992, 48 (38), 823952,  DOI: 10.1016/S0040-4020(01)80492-0
  237. 237
    Yoon, J.; Czarnik, A. W. Fluorescent Chemosensors of Carbohydrates. A Means of Chemically Communicating the Binding of Polyols in Water Based on Chelation-Enhanced Quenching. J. Am. Chem. Soc. 1992, 114 (14), 58745,  DOI: 10.1021/ja00040a067
  238. 238
    Elfeky, S. Novel Bronic Acid-Based Fluorescent Sensor for Sugars and Nucleosides. Curr. Org. Synth. 2011, 8 (6), 872,  DOI: 10.2174/1570179411108060872
  239. 239
    James, T. D.; Shinkai, S. Artificial Receptors as Chemosensors for Carbohydrates. Host-Guest Chemistry 2002, 218, 159200,  DOI: 10.1007/3-540-45010-6_6
  240. 240
    Springsteen, G.; Wang, B. A Detailed Examination of Boronic Acid-Diol Complexation. Tetrahedron 2002, 58 (26), 5291300,  DOI: 10.1016/S0040-4020(02)00489-1
  241. 241
    Jin, S.; Cheng, Y.; Reid, S.; Li, M.; Wang, B. Carbohydrate Recognition by Boronolectins, Small Molecules, and Lectins. Med. Res. Rev. 2009, 30, 171257,  DOI: 10.1002/med.20155
  242. 242
    Yan, J.; Fang, H.; Wang, B. Boronolectins and Fluorescent Boronolectins: An Examination of the Detailed Chemistry Issues Important for the Design. Med. Res. Rev. 2005, 25 (5), 490520,  DOI: 10.1002/med.20038
  243. 243
    Gao, X.; Zhang, Y.; Wang, B. New Boronic Acid Fluorescent Reporter Compounds. 2. A Naphthalene-Based On-Off Sensor Functional at Physiological pH. Org. Lett. 2003, 5 (24), 46158,  DOI: 10.1021/ol035783i
  244. 244
    Gao, X.; Zhang, Y.; Wang, B. A Highly Fluorescent Water-Soluble Boronic Acid Reporter for Saccharide Sensing That Shows Ratiometric UV Changes and Significant Fluorescence Changes. Tetrahedron 2005, 61 (38), 91117,  DOI: 10.1016/j.tet.2005.07.035
  245. 245
    Gao, X.; Zhang, Y.; Wang, B. Naphthalene-Based Water-Soluble Fluorescent Boronic Acid Isomers Suitable for Ratiometric and Off-On Sensing of Saccharides at Physiological pH. New J. Chem. 2005, 29 (4), 57986,  DOI: 10.1039/b413376e
  246. 246
    Zhang, Y.; Gao, X.; Hardcastle, K.; Wang, B. Water-Soluble Fluorescent Boronic Acid Compounds for Saccharide Sensing: Substituent Effects on Their Fluorescence Properties. Eur. J. Chem. 2006, 12 (5), 137784,  DOI: 10.1002/chem.200500982
  247. 247
    Yang, W.; Yan, J.; Springsteen, G.; Deeter, S.; Wang, B. A Novel Type of Fluorescent Boronic Acid That Shows Large Fluorescence Intensity Changes Upon Binding with a Carbohydrate in Aqueous Solution at Physiological pH. Bioorg. Med. Chem. Lett. 2003, 13 (6), 101922,  DOI: 10.1016/S0960-894X(03)00086-6
  248. 248
    Yang, W.; Lin, L.; Wang, B. A New Type of Boronic Acid Fluorescent Reporter Compound for Sugar Recognition. Tetrahedron Lett. 2005, 46 (46), 79814,  DOI: 10.1016/j.tetlet.2005.09.074
  249. 249
    Cheng, Y.; Ni, N.; Yang, W.; Wang, B. A New Class of Fluorescent Boronic Acids That Have Extraordinarily High Affinities for Diols in Aqueous Solution at Physiological pH. Eur. J. Chem. 2010, 16 (45), 1352838,  DOI: 10.1002/chem.201000637
  250. 250
    Samaniego Lopez, C.; Lago Huvelle, M. A.; Uhrig, M. L.; Coluccio Leskow, F.; Spagnuolo, C. C. Recognition of Saccharides in the NIR Region with a Novel Fluorogenic Boronolectin: In Vitro and Live Cell Labeling. Chem. Commun. 2015, 51 (23), 48958,  DOI: 10.1039/C4CC10425K
  251. 251
    Wang, L.; Yuan, L.; Zeng, X.; Peng, J.; Ni, Y.; Er, J. C.; Xu, W.; Agrawalla, B. K.; Su, D.; Kim, B. A Multisite-Binding Switchable Fluorescent Probe for Monitoring Mitochondrial ATP Level Fluctuation in Live Cells. Angew. Chem., Int. Ed. 2016, 55 (5), 17736,  DOI: 10.1002/anie.201510003
  252. 252
    Huang, Y.-J.; Ouyang, W.-J.; Wu, X.; Li, Z.; Fossey, J. S.; James, T. D.; Jiang, Y.-B. Glucose Sensing Via Aggregation and the Use of “Knock-Out” Binding to Improve Selectivity. J. Am. Chem. Soc. 2013, 135 (5), 17003,  DOI: 10.1021/ja311442x
  253. 253
    Vilozny, B.; Schiller, A.; Wessling, R. A.; Singaram, B. Enzyme Assays with Boronic Acid Appended Bipyridinium Salts. Anal. Chim. Acta 2009, 649 (2), 24651,  DOI: 10.1016/j.aca.2009.07.032
  254. 254
    Sharrett, Z.; Gamsey, S.; Hirayama, L.; Vilozny, B.; Suri, J. T.; Wessling, R. A.; Singaram, B. Exploring the Use of APTS as a Fluorescent Reporter Dye for Continuous Glucose Sensing. Org. Biomol. Chem. 2009, 7 (7), 146170,  DOI: 10.1039/b821934f
  255. 255
    James, T. D.; Sandanayake, K. R. A. S.; Shinkai, S. A Glucose-Selective Molecular Fluorescence Sensor. Angew. Chem., Int. Ed. 1994, 33 (21), 22079,  DOI: 10.1002/anie.199422071
  256. 256
    James, T. D.; Sandanayake, K. R. A. S.; Iguchi, R.; Shinkai, S. Novel Saccharide-Photoinduced Electron Transfer Sensors Based on the Interaction of Boronic Acid and Amine. J. Chem. Soc., Chem. Commun. 1995, 117 (35), 89827,  DOI: 10.1021/ja00140a013
  257. 257
    Tran, T. M.; Alan, Y.; Glass, T. E. A Highly Selective Fluorescent Sensor for Glucosamine. Chem. Commun. 2015, 51 (37), 79158,  DOI: 10.1039/C5CC00415B
  258. 258
    Rout, B.; Unger, L.; Armony, G.; Iron, M. A.; Margulies, D. Medication Detection by a Combinatorial Fluorescent Molecular Sensor. Angew. Chem., Int. Ed. 2012, 51 (50), 1247781,  DOI: 10.1002/anie.201206374
  259. 259
    Liu, Y.; Deng, C.; Tang, L.; Qin, A.; Hu, R.; Sun, J. Z.; Tang, B. Z. Specific Detection of D-Glucose by a Tetraphenylethene-Based Fluorescent Sensor. J. Am. Chem. Soc. 2011, 133 (4), 6603,  DOI: 10.1021/ja107086y
  260. 260
    Shcherbakova, E. G.; Brega, V.; Lynch, V. M.; James, T. D.; Anzenbacher, P., Jr. High-Throughput Assay for Enantiomeric Excess Determination in 1,2- and 1,3-Diols and Direct Asymmetric Reaction Screening. Eur. J. Chem. 2017, 23 (42), 102229,  DOI: 10.1002/chem.201701923
  261. 261
    Shcherbakova, E. G.; James, T. D.; Anzenbacher, P. High-Throughput Assay for Determining Enantiomeric Excess of Chiral Diols, Amino Alcohols, and Amines and for Direct Asymmetric Reaction Screening. Nat. Protoc. 2020, 15 (7), 220329,  DOI: 10.1038/s41596-020-0329-1
  262. 262
    Doderer, K.; Lutz-Wahl, S.; Hauer, B.; Schmid, R. D. Spectrophotometric Assay for Epoxide Hydrolase Activity toward any Epoxide. Anal. Biochem. 2003, 321 (1), 1314,  DOI: 10.1016/S0003-2697(03)00399-3
  263. 263
    Wahler, D.; Reymond, J.-L. The Adrenaline Test for Enzymes. Angew. Chem., Int. Ed. 2002, 41 (7), 122932,  DOI: 10.1002/1521-3773(20020402)41:7<1229::AID-ANIE1229>3.0.CO;2-5
  264. 264
    Jie, N.; Yang, D.; Zhang, Q.; Yang, J.; Song, Z. Fluorometric Determination of Periodate with Thiamine and Its Application to the Determination of Ethylene Glycol and Glycerol. Anal. Chim. Acta 1998, 359 (1), 8792,  DOI: 10.1016/S0003-2670(97)00657-0
  265. 265
    Kirschner, A.; Bornscheuer, U. T. Directed Evolution of a Baeyer-Villiger Monooxygenase to Enhance Enantioselectivity. Appl. Microbiol. Biotechnol. 2008, 81 (3), 46572,  DOI: 10.1007/s00253-008-1646-4
  266. 266
    Wahler, D.; Boujard, O.; Lefèvre, F.; Reymond, J.-L. Adrenaline Profiling of Lipases and Esterases with 1,2-Diol and Carbohydrate Acetates. Tetrahedron 2004, 60 (3), 70310,  DOI: 10.1016/j.tet.2003.11.059
  267. 267
    Fluxá, V. S.; Wahler, D.; Reymond, J.-L. Enzyme Assay and Activity Fingerprinting of Hydrolases with the Red-Chromogenic Adrenaline Test. Nat. Protoc. 2008, 3 (8), 12707,  DOI: 10.1038/nprot.2008.106
  268. 268
    Doderer, K.; Schmid, R. D. Fluorometric Assay for Determining Epoxide Hydrolase Activity. Biotechnol. Lett. 2004, 26 (10), 8359,  DOI: 10.1023/B:BILE.0000025887.36874.33
  269. 269
    Matsuno, K.; Suzuki, S. Simple Fluorimetric Method for Quantification of Sialic Acids in Glycoproteins. Anal. Biochem. 2008, 375 (1), 539,  DOI: 10.1016/j.ab.2008.01.002
  270. 270
    Preston-Herrera, C.; Jackson, A. S.; Bachmann, B. O.; Froese, J. T. Development and Application of a High Throughput Assay System for the Detection of Rieske Dioxygenase Activity. Org. Biomol. Chem. 2021, 19 (4), 77584,  DOI: 10.1039/D0OB02412K
  271. 271
    Winkler, M. Carboxylic Acid Reductase Enzymes (CARS). Curr. Opin. Chem. Biol. 2018, 43, 239,  DOI: 10.1016/j.cbpa.2017.10.006
  272. 272
    Sutiono, S.; Carsten, J.; Sieber, V. Structure-Guided Engineering of α-Keto Acid Decarboxylase for the Production of Higher Alcohols at Elevated Temperature. ChemSusChem 2018, 11 (18), 333544,  DOI: 10.1002/cssc.201800944
  273. 273
    Greenhalgh, J. C.; Fahlberg, S. A.; Pfleger, B. F.; Romero, P. A. Machine Learning-Guided Acyl-ACP Reductase Engineering for Improved in Vivo Fatty Alcohol Production. Nat. Commun. 2021, 12, 5825,  DOI: 10.1038/s41467-021-25831-w
  274. 274
    Sellés Vidal, L.; Kelly, C. L.; Mordaka, P. M.; Heap, J. T. Review of NAD(P)H-Dependent Oxidoreductases: Properties, Engineering and Application. Biochim. Biophys. Acta 2018, 1866 (2), 32747,  DOI: 10.1016/j.bbapap.2017.11.005
  275. 275
    Goswami, P.; Chinnadayyala, S. S. R.; Chakraborty, M.; Kumar, A. K.; Kakoti, A. An Overview on Alcohol Oxidases and Their Potential Applications. Appl. Microbiol. Biotechnol. 2013, 97 (10), 425975,  DOI: 10.1007/s00253-013-4842-9
  276. 276
    Mang, H.; Gross, J.; Lara, M.; Goessler, C.; Schoemaker, H. E.; Guebitz, G. M.; Kroutil, W. Biocatalytic Single-Step Alkene Cleavage from Aryl Alkenes: An Enzymatic Equivalent to Reductive Ozonization. Angew. Chem., Int. Ed. 2006, 45 (31), 52013,  DOI: 10.1002/anie.200601574
  277. 277
    Schwendenwein, D.; Ressmann, A. K.; Doerr, M.; Höhne, M.; Bornscheuer, U. T.; Mihovilovic, M. D.; Rudroff, F.; Winkler, M. Random Mutagenesis-Driven Improvement of Carboxylate Reductase Activity Using an Amino Benzamidoxime-Mediated High-Throughput Assay. Adv. Synth. Catal. 2019, 361, 2544,  DOI: 10.1002/adsc.201900155
  278. 278
    Soh, L. M. J.; Mak, W. S.; Lin, P. P.; Mi, L.; Chen, F. Y. H.; Damoiseaux, R.; Siegel, J. B.; Liao, J. C. Engineering a Thermostable Keto Acid Decarboxylase Using Directed Evolution and Computationally Directed Protein Design. ACS Synth. Biol. 2017, 6 (4), 6108,  DOI: 10.1021/acssynbio.6b00240
  279. 279
    Van Schie, M. M. C. H.; Pedroso De Almeida, T.; Laudadio, G.; Tieves, F.; Fernández-Fueyo, E.; Noël, T.; Arends, I. W. C. E.; Hollmann, F. Biocatalytic Synthesis of the Green Note trans-2-Hexenal in a Continuous-Flow Microreactor. Beilstein J. Org. Chem. 2018, 14, 697703,  DOI: 10.3762/bjoc.14.58
  280. 280
    Allen, C. F. The Identification of Carbonyl Compounds by Use of 2,4-Dinitrophenylhydrazine. J. Am. Chem. Soc. 1930, 52 (7), 29559,  DOI: 10.1021/ja01370a058
  281. 281
    Brady, O. L.; Elsmie, G. V. The Use of 2,4-Dinitrophenylhydrazine as a Reagent for Aldehydes and Ketones. Analyst 1926, 51 (599), 778,  DOI: 10.1039/an9265100077
  282. 282
    Bohlmann, F. Konstitution Und Lichtabsorption, I. Mitteil.: Carbonyl-Derivate. Chem. Ber. 1951, 84 (5–6), 490504,  DOI: 10.1002/cber.19510840515
  283. 283
    Georgiou, C. D.; Zisimopoulos, D.; Argyropoulou, V.; Kalaitzopoulou, E.; Salachas, G.; Grune, T. Protein and Cell Wall Polysaccharide Carbonyl Determination by a Neutral pH 2,4-Dinitrophenylhydrazine-Based Photometric Assay. Redox Biol. 2018, 17, 12842,  DOI: 10.1016/j.redox.2018.04.010
  284. 284
    Mesquita, C. S.; Oliveira, R.; Bento, F.; Geraldo, D.; Rodrigues, J. V.; Marcos, J. C. Simplified 2,4-Dinitrophenylhydrazine Spectrophotometric Assay for Quantification of Carbonyls in Oxidized Proteins. Anal. Biochem. 2014, 458, 6971,  DOI: 10.1016/j.ab.2014.04.034
  285. 285
    Xue, Y.-P.; Wang, W.; Wang, Y.-J.; Liu, Z.-Q.; Zheng, Y.-G.; Shen, Y.-C. Isolation of Enantioselective α-Hydroxyacid Dehydrogenases Based on a High-Throughput Screening Method. Bioprocess Biosyst. Eng. 2012, 35 (9), 151522,  DOI: 10.1007/s00449-012-0741-1
  286. 286
    Chen, Q.; Chen, X.; Feng, J.; Wu, Q.; Zhu, D.; Ma, Y. Improving and Inverting Cβ-Stereoselectivity of Threonine Aldolase Via Substrate-Binding-Guided Mutagenesis and a Stepwise Visual Screening. ACS Catal. 2019, 9 (5), 44629,  DOI: 10.1021/acscatal.9b00859
  287. 287
    Yang, C.; Ye, L.; Gu, J.; Yang, X.; Li, A.; Yu, H. Directed Evolution of Mandelate Racemase by a Novel High-Throughput Screening Method. Appl. Microbiol. Biotechnol. 2017, 101 (3), 106372,  DOI: 10.1007/s00253-016-7790-3
  288. 288
    Uzu, S.; Kanda, S.; Imai, K.; Nakashima, K.; Akiyama, S. Fluorogenic Reagents: 4-Aminosulphonyl-7-Hydrazino-2,1,3-Benzoxadiazole, 4-(N,N-Dimethylaminosulphonyl)-7-Hydrazino-2,1,3-Benzoxadiazole and 4-Hydrazino-7-Nitro-2,1,3-Benzoxadiazole Hydrazine for Aldehydes and Ketones. Analyst 1990, 115 (11), 147782,  DOI: 10.1039/an9901501477
  289. 289
    Uchiyama, S.; Santa, T.; Okiyama, N.; Fukushima, T.; Imai, K. Fluorogenic and Fluorescent Labeling Reagents with a Benzofurazan Skeleton. Biomed. Chromatogr. 2001, 15 (5), 295318,  DOI: 10.1002/bmc.75
  290. 290
    Frizon, T. E. A.; Vieira, A. A.; da Silva, F. N.; Saba, S.; Farias, G.; de Souza, B.; Zapp, E.; Lôpo, M. N.; Braga, H. d. C.; Grillo, F. Synthesis of 2,1,3-Benzoxadiazole Derivatives as New Fluorophores─Combined Experimental, Optical, Electro, and Theoretical Study. Front. Chem. 2020, 8, 360,  DOI: 10.3389/fchem.2020.00360
  291. 291
    Konarzycka-Bessler, M.; Bornscheuer, U. T. A High-Throughput-Screening Method for Determining the Synthetic Activity of Hydrolases. Angew. Chem., Int. Ed. 2003, 42 (12), 141820,  DOI: 10.1002/anie.200390365
  292. 292
    Jencks, W. P. Progress in Physical Organic Chemistry; Index to Reviews, Symposia Volumes and Monographs in Organic Chemistry; Cohen, S. G., Streitwieser, A., Jr., Taft, R. W., Eds.; John Wiley & Sons, 1964; Vol. 2, p 110 DOI: 10.1002/9780470171813
  293. 293
    Larsen, D.; Pittelkow, M.; Karmakar, S.; Kool, E. T. New Organocatalyst Scaffolds with High Activity in Promoting Hydrazone and Oxime Formation at Neutral pH. Org. Lett. 2015, 17 (2), 2747,  DOI: 10.1021/ol503372j
  294. 294
    Crisalli, P.; Kool, E. T. Importance of Ortho Proton Donors in Catalysis of Hydrazone Formation. Org. Lett. 2013, 15 (7), 16469,  DOI: 10.1021/ol400427x
  295. 295
    Wang, S.; Nawale, G. N.; Kadekar, S.; Oommen, O. P.; Jena, N. K.; Chakraborty, S.; Hilborn, J.; Varghese, O. P. Saline Accelerates Oxime Reaction with Aldehyde and Keto Substrates at Physiological pH. Sci. Rep. 2018, 8, 2193,  DOI: 10.1038/s41598-018-20735-0
  296. 296
    Mayr, H.; Bug, T.; Gotta, M. F.; Hering, N.; Irrgang, B.; Janker, B.; Kempf, B.; Loos, R.; Ofial, A. R.; Remennikov, G. Reference Scales for the Characterization of Cationic Electrophiles and Neutral Nucleophiles. J. Am. Chem. Soc. 2001, 123 (39), 950012,  DOI: 10.1021/ja010890y
  297. 297
    Brotzel, F.; Chu, Y. C.; Mayr, H. Nucleophilicities of Primary and Secondary Amines in Water. J. Org. Chem. 2007, 72 (10), 367988,  DOI: 10.1021/jo062586z
  298. 298
    Brotzel, F., lmu, 2008.
  299. 299
    Nakashima, K.; Hidaka, Y.; Yoshida, T.; Kuroda, N.; Akiyama, S. High-Performance Liquid Chromatographic Determination of Short-Chain Aliphatic Aldehydes Using 4-(N,N-Dimethylaminosulphonyl)-7-Hydrazino-2,1,3-Benzoxadiazole as a Fluorescence Reagent. J. Chromatogr. B Biomed. Appl. 1994, 661 (2), 20510,  DOI: 10.1016/0378-4347(94)00359-9
  300. 300
    Jacobsen, N. W.; Dickinson, R. G. Spectrometric Assay of Aldehydes as 6-Mercapto-3-Substituted-s-Trizolo (4,3-b)-Tetrazines. Anal. Chem. 1974, 46 (2), 2989,  DOI: 10.1021/ac60338a039
  301. 301
    Dickinson, R. G.; Jacobsen, N. W. A New Sensitive and Specific Test for the Detection of Aldehydes: Formation of 6-Mercapto-3-Substituted-S-Triazolo[4,3-b]-s-Tetrazines. J. Chem. Soc. D 1970, (24), 171920,  DOI: 10.1039/c29700001719
  302. 302
    Quesenberry, M. S.; Lee, Y. C. A Rapid Formaldehyde Assay Using Purpald Reagent: Application under Periodation Conditions. Anal. Biochem. 1996, 234 (1), 505,  DOI: 10.1006/abio.1996.0048
  303. 303
    Anthon, G. E.; Barrett, D. M. Comparison of Three Colorimetric Reagents in the Determination of Methanol with Alcohol Oxidase. Application to the Assay of Pectin Methylesterase. J. Agric. Food Chem. 2004, 52 (12), 374953,  DOI: 10.1021/jf035284w
  304. 304
    Hammer, S. C.; Kubik, G.; Watkins, E.; Huang, S.; Minges, H.; Arnold, F. H. Anti-Markovnikov Alkene Oxidation by Metal-Oxo-Mediated Enzyme Catalysis. Science 2017, 358 (6360), 2158,  DOI: 10.1126/science.aao1482
  305. 305
    Nguyen, Q.-T.; Romero, E.; Dijkman, W. P.; de Vasconcellos, S. P.; Binda, C.; Mattevi, A.; Fraaije, M. W. Structure-Based Engineering of Phanerochaete Chrysosporium Alcohol Oxidase for Enhanced Oxidative Power toward Glycerol. Biochemistry 2018, 57 (43), 620918,  DOI: 10.1021/acs.biochem.8b00918
  306. 306
    Sawicki, E.; Hauser, T. R.; Stanley, T. W.; Elbert, W. The 3-Methyl-2-Benzothiazolone Hydrazone Test. Sensitive New Methods for the Detection, Rapid Estimation, and Determination of Aliphatic Aldehydes. Anal. Chem. 1961, 33 (1), 936,  DOI: 10.1021/ac60169a028
  307. 307
    Zhu, P.; Lin, S.-M. MBTH for aliphatic aldehyde measurement, U.S. Patent US7112448B2, September 26, 2006.
  308. 308
    Hauser, T. R.; Cummins, R. L. Increasing Sensitivity of 3-Methyl-2-Benzothiazolone Hydrozone Test for Analysis of Aliphatic Aldehydes in Air. Anal. Chem. 1964, 36 (3), 67981,  DOI: 10.1021/ac60209a067
  309. 309
    Dean Pakulski, J.; Benner, R. An Improved Method for the Hydrolysis and MBTH Analysis of Dissolved and Particulate Carbohydrates in Seawater. Mar. Chem. 1992, 40 (3), 14360,  DOI: 10.1016/0304-4203(92)90020-B
  310. 310
    Gomez, L. D.; Whitehead, C.; Barakate, A.; Halpin, C.; McQueen-Mason, S. J. Automated Saccharification Assay for Determination of Digestibility in Plant Materials. Biotechnol. Biofuels 2010, 3, 23,  DOI: 10.1186/1754-6834-3-23
  311. 311
    Oliveira, F. S. d.; Leite, B. C. O.; Andrade, M. V. A. S. d.; Korn, M. Determination of Total Aldehydes in Fuel Ethanol by MBTH Method: Sequential Injection Analysis. J. Braz. Chem. Soc. 2005, 16, 8792,  DOI: 10.1590/S0103-50532005000100013
  312. 312
    Suchý, M.; Lazurko, C.; Kirby, A.; Dang, T.; Liu, G.; Shuhendler, A. J. Methyl 5-MeO-N-Aminoanthranilate, a Minimalist Fluorogenic Probe for Sensing Cellular Aldehydic Load. Org. Biomol. Chem. 2019, 17 (7), 184353,  DOI: 10.1039/C8OB02255K
  313. 313
    Yuen, L. H.; Saxena, N. S.; Park, H. S.; Weinberg, K.; Kool, E. T. Dark Hydrazone Fluorescence Labeling Agents Enable Imaging of Cellular Aldehydic Load. ACS Chem. Biol. 2016, 11 (8), 23129,  DOI: 10.1021/acschembio.6b00269
  314. 314
    Liu, T.; Yang, L.; Zhang, J.; Liu, K.; Ding, L.; Peng, H.; Belfield, K. D.; Fang, Y. Squaraine-Hydrazine Adducts for Fast and Colorimetric Detection of Aldehydes in Aqueous Media. Sens. Actuators B Chem. 2019, 292, 8893,  DOI: 10.1016/j.snb.2019.04.138
  315. 315
    Liu, T.; Liu, X.; Wang, W.; Luo, Z.; Liu, M.; Zou, S.; Sissa, C.; Painelli, A.; Zhang, Y.; Vengris, M. Systematic Molecular Engineering of a Series of Aniline-Based Squaraine Dyes and Their Structure-Related Properties. J. Phys. Chem. C 2018, 122 (7), 39944008,  DOI: 10.1021/acs.jpcc.7b11997
  316. 316
    Xia, G.; Wang, H. Squaraine Dyes: The Hierarchical Synthesis and Its Application in Optical Detection. J. Photochem. Photobiol. C: Photochem. Rev. 2017, 31, 84113,  DOI: 10.1016/j.jphotochemrev.2017.03.001
  317. 317
    Lichtenstein, S. J.; Nettleton, G. S. Effects of Fuchsin Variants in Aldehyde Fuchsin Staining. J. Histochem. Cytochem. 1980, 28 (7), 6838,  DOI: 10.1177/28.7.6156202
  318. 318
    Robins, J. H.; Abrams, G. D.; Pincock, J. A. The Structure of Schiff Reagent Aldehyde Adducts and the Mechanism of the Schiff Reaction as Determined by Nuclear Magnetic Resonance Spectroscopy. Can. J. Chem. 1980, 58 (4), 33947,  DOI: 10.1139/v80-055
  319. 319
    Kitov, P. I.; Vinals, D. F.; Ng, S.; Tjhung, K. F.; Derda, R. Rapid, Hydrolytically Stable Modification of Aldehyde-Terminated Proteins and Phage Libraries. J. Am. Chem. Soc. 2014, 136 (23), 814952,  DOI: 10.1021/ja5023909
  320. 320
    Ressmann, A. K.; Schwendenwein, D.; Leonhartsberger, S.; Mihovilovic, M. D.; Bornscheuer, U. T.; Winkler, M.; Rudroff, F. Substrate-Independent High-Throughput Assay for the Quantification of Aldehydes. Adv. Synth. Catal. 2019, 361, 2538,  DOI: 10.1002/adsc.201900154
  321. 321
    Horvat, M.; Larch, T.-S.; Rudroff, F.; Winkler, M. Amino Benzamidoxime (ABAO)-Based Assay to Identify Efficient Aldehyde-Producing Pichia Pastoris Clones. Adv. Synth. Catal. 2020, 362 (21), 46739,  DOI: 10.1002/adsc.202000558
  322. 322
    Mei, Z.; Zhang, K.; Qu, G.; Li, J.-K.; Liu, B.; Ma, J.-A.; Tu, R.; Sun, Z. High-Throughput Fluorescence Assay for Ketone Detection and Its Applications in Enzyme Mining and Protein Engineering. ACS Omega 2020, 5 (23), 1358894,  DOI: 10.1021/acsomega.0c00245
  323. 323
    Xing, Y.; Wang, S.; Mao, X.; Zhao, X.; Wei, D. An Easy and Efficient Fluorescent Method for Detecting Aldehydes and Its Application in Biotransformation. J. Fluoresc. 2011, 21 (2), 58794,  DOI: 10.1007/s10895-010-0746-6
  324. 324
    Li, Z.; Xue, Z.; Wu, Z.; Han, J.; Han, S. Chromo-Fluorogenic Detection of Aldehydes with a Rhodamine Based Sensor Featuring an Intramolecular Deoxylactam. Org. Biomol. Chem. 2011, 9 (22), 76524,  DOI: 10.1039/c1ob06448g
  325. 325
    Davidson, D.; Weiss, M.; Jelling, M. The Action of Ammonia on Benzil. J. Org. Chem. 1937, 02 (4), 31927,  DOI: 10.1021/jo01227a004
  326. 326
    Nakashima, K.; Yamasaki, H.; Kuroda, N.; Akiyama, S. Evaluation of Lophine Derivatives as Chemiluminogens by a Flow-Injection Method. Anal. Chim. Acta 1995, 303 (1), 1037,  DOI: 10.1016/0003-2670(94)00360-X
  327. 327
    El-Maghrabey, M. H.; Kishikawa, N.; Ohyama, K.; Kuroda, N. Analytical Method for Lipoperoxidation Relevant Reactive Aldehydes in Human Sera by High-Performance Liquid Chromatography-Fluorescence Detection. Anal. Biochem. 2014, 464, 3642,  DOI: 10.1016/j.ab.2014.07.002
  328. 328
    Fathy Bakr Ali, M.; Kishikawa, N.; Ohyama, K.; Abdel-Mageed Mohamed, H.; Mohamed Abdel-Wadood, H.; Mohamed Mohamed, A.; Kuroda, N. Chromatographic Determination of Aliphatic Aldehydes in Human Serum after Pre-Column Derivatization Using 2,2′-Furil, a Novel Fluorogenic Reagent. J. Chromatogr. A 2013, 1300, 199203,  DOI: 10.1016/j.chroma.2013.03.033
  329. 329
    Nakamura, M.; Toda, M.; Saito, H.; Ohkura, Y. Fluorimetric Determination of Aromatic Aldehydes with 4,5-Dimethoxy-1,2-Diaminobenzene. Anal. Chim. Acta 1982, 134, 3945,  DOI: 10.1016/S0003-2670(01)84175-1
  330. 330
    Imazato, T.; Kanematsu, M.; Kishikawa, N.; Ohyama, K.; Hino, T.; Ueki, Y.; Maehata, E.; Kuroda, N. Determination of Acrolein in Serum by High-Performance Liquid Chromatography with Fluorescence Detection after Pre-Column Fluorogenic Derivatization Using 1,2-Diamino-4,5-Dimethoxybenzene. Biomed. Chromatogr. 2015, 29 (9), 13048,  DOI: 10.1002/bmc.3422
  331. 331
    McLellan, A. C.; Phillips, S. A.; Thornalley, P. J. The Assay of Methylglyoxal in Biological Systems Byderivatization with 1,2-Diamino-4,5-Dimethoxybenzene. Anal. Biochem. 1992, 206 (1), 1723,  DOI: 10.1016/S0003-2697(05)80005-3
  332. 332
    Hantzsch, A. Condensationsprodukte Aus Aldehydammoniak Und Ketonartigen Verbindungen. Ber. Dtsch. Chem. Ges. 1881, 14 (2), 16378,  DOI: 10.1002/cber.18810140214
  333. 333
    Belman, S. The Fluorimetric Determination of Formaldehyde. Anal. Chim. Acta 1963, 29, 1206,  DOI: 10.1016/S0003-2670(00)88591-8
  334. 334
    Sawicki, E.; Carnes, R. Spectrophotofluorimetric Determination of Aldehydes with Dimedone and Other Reagents. Mikrochim. Acta 1968, 56 (1), 14859,  DOI: 10.1007/BF01216118
  335. 335
    Compton, B. J.; Purdy, W. C. The Mechanism of the Reaction of the Nash and the Sawicki Aldehyde Reagent. Can. J. Chem. 1980, 58 (21), 220711,  DOI: 10.1139/v80-355
  336. 336
    Compton, B. J.; Purdy, W. C. Fluoral-P, a Member of a Selective Family of Reagents for Aldehydes. Anal. Chim. Acta 1980, 119 (2), 34957,  DOI: 10.1016/S0003-2670(01)93636-0
  337. 337
    Mopper, K.; Stahovec, W. L.; Johnson, L. Trace Analysis of Aldehydes by Reversed-Phase High-Performance Liquid Chromatography and Precolumn Fluorigenic Labeling with 5,5-Dimethyl-1,3-Cyclohexanedione. J. Chromatogr. A 1983, 256, 24352,  DOI: 10.1016/S0021-9673(01)88237-6
  338. 338
    Al-Moniee, M. A.; Koopal, C.; Akmal, N.; van Veen, S.; Zhu, X.; Sanders, P. F.; Sanders, P. F.; Al-Abeedi, F. N.; Amer, A. M. Applicability of Dimedone Assays for the Development of Online Aldehyde Sensor in Seawater Flooding Systems. J. Sens. Sci. Technol. 2016, 6, 101,  DOI: 10.4236/jst.2016.64008
  339. 339
    Li, Q.; Oshima, M.; Motomizu, S. Flow-Injection Spectrofluorometric Determination of Trace Amounts of Formaldehyde in Water after Derivatization with Acetoacetanilide. Talanta 2007, 72 (5), 167580,  DOI: 10.1016/j.talanta.2007.01.054
  340. 340
    Kolmel, D. K.; Kool, E. T. Oximes and Hydrazones in Bioconjugation: Mechanism and Catalysis. Chem. Rev. 2017, 117 (15), 1035876,  DOI: 10.1021/acs.chemrev.7b00090
  341. 341
    Kalia, J.; Raines, R. T. Hydrolytic Stability of Hydrazones and Oximes. Angew. Chem., Int. Ed. 2008, 47 (39), 75236,  DOI: 10.1002/anie.200802651
  342. 342
    Forget, D.; Boturyn, D.; Defrancq, E.; Lhomme, J.; Dumy, P. Highly Efficient Synthesis of Peptide-Oligonucleotide Conjugates: Chemoselective Oxime and Thiazolidine Formation. Eur. J. Chem. 2001, 7 (18), 397684,  DOI: 10.1002/1521-3765(20010917)7:18<3976::AID-CHEM3976>3.0.CO;2-X
  343. 343
    Shao, J.; Tam, J. P. Unprotected Peptides as Building Blocks for the Synthesis of Peptide Dendrimers with Oxime, Hydrazone, and Thiazolidine Linkages. J. Am. Chem. Soc. 1995, 117 (14), 38939,  DOI: 10.1021/ja00119a001
  344. 344
    Kemp, D. S.; Woodward, R. B. The N-Ethylbenzisoxazolium Cation─I: Preparation and Reactions with Nucleophilic Species. Tetrahedron 1965, 21 (11), 301935,  DOI: 10.1016/S0040-4020(01)96921-2
  345. 345
    Salahuddin, S.; Renaudet, O.; Reymond, J.-L. Aldehyde Detection by Chromogenic/Fluorogenic Oxime Bond Fragmentation. Org. Biomol. Chem. 2004, 2 (10), 14715,  DOI: 10.1039/b400314d
  346. 346
    Drienovská, I.; Mayer, C.; Dulson, C.; Roelfes, G. A Designer Enzyme for Hydrazone and Oxime Formation Featuring an Unnatural Catalytic Aniline Residue. Nat. Chem. 2018, 10 (9), 94652,  DOI: 10.1038/s41557-018-0082-z
  347. 347
    Horowitz, R. M.; Geissman, T. A. A Cleavage Reaction of α-Allylbenzylamines. J. Am. Chem. Soc. 1950, 72 (4), 151822,  DOI: 10.1021/ja01160a025
  348. 348
    Sakabe, M.; Asanuma, D.; Kamiya, M.; Iwatate, R. J.; Hanaoka, K.; Terai, T.; Nagano, T.; Urano, Y. Rational Design of Highly Sensitive Fluorescence Probes for Protease and Glycosidase Based on Precisely Controlled Spirocyclization. J. Am. Chem. Soc. 2013, 135 (1), 40914,  DOI: 10.1021/ja309688m
  349. 349
    Brewer, T. F.; Chang, C. J. An Aza-Cope Reactivity-Based Fluorescent Probe for Imaging Formaldehyde in Living Cells. J. Am. Chem. Soc. 2015, 137 (34), 108869,  DOI: 10.1021/jacs.5b05340
  350. 350
    Roth, A.; Li, H.; Anorma, C.; Chan, J. A Reaction-Based Fluorescent Probe for Imaging of Formaldehyde in Living Cells. J. Am. Chem. Soc. 2015, 137 (34), 108903,  DOI: 10.1021/jacs.5b05339
  351. 351
    Bruemmer, K. J.; Walvoord, R. R.; Brewer, T. F.; Burgos-Barragan, G.; Wit, N.; Pontel, L. B.; Patel, K. J.; Chang, C. J. Development of a General Aza-Cope Reaction Trigger Applied to Fluorescence Imaging of Formaldehyde in Living Cells. J. Am. Chem. Soc. 2017, 139 (15), 533850,  DOI: 10.1021/jacs.6b12460
  352. 352
    Du, Y.; Zhang, Y.; Huang, M.; Wang, S.; Wang, J.; Liao, K.; Wu, X.; Zhou, Q.; Zhang, X.; Wu, Y.-D. Systematic Investigation of the Aza-Cope Reaction for Fluorescence Imaging of Formaldehyde in Vitro and in Vivo. Chem. Sci. 2021, 12 (41), 1385769,  DOI: 10.1039/D1SC04387K
  353. 353
    He, L.; Yang, X.; Liu, Y.; Kong, X.; Lin, W. A Ratiometric Fluorescent Formaldehyde Probe for Bioimaging Applications. Chem. Commun. 2016, 52 (21), 402932,  DOI: 10.1039/C5CC09796G
  354. 354
    Xu, J.; Zhang, Y.; Zeng, L.; Liu, J.; Kinsella, J. M.; Sheng, R. A Simple Naphthalene-Based Fluorescent Probe for High Selective Detection of Formaldehyde in Toffees and HeLa Cells Via Aza-Cope Reaction. Talanta 2016, 160, 64552,  DOI: 10.1016/j.talanta.2016.08.010
  355. 355
    Yang, X.; He, L.; Xu, K.; Yang, Y.; Lin, W. The Development of an Ict-Based Formaldehyde-Responsive Fluorescence Turn-On Probe with a High Signal-to-Noise Ratio. New J. Chem. 2018, 42 (15), 123614,  DOI: 10.1039/C8NJ02467G
  356. 356
    Yang, M.; Fan, J.; Du, J.; Long, S.; Wang, J.; Peng, X. Imaging of Formaldehyde in Live Cells and Daphnia Magna Via Aza-Cope Reaction Utilizing Fluorescence Probe with Large Stokes Shifts. Front. Chem. 2018, 6, 488,  DOI: 10.3389/fchem.2018.00488
  357. 357
    Bruemmer, K. J.; Brewer, T. F.; Chang, C. J. Fluorescent Probes for Imaging Formaldehyde in Biological Systems. Curr. Opin. Chem. Biol. 2017, 39, 1723,  DOI: 10.1016/j.cbpa.2017.04.010
  358. 358
    Liu, X.; Li, N.; Li, M.; Chen, H.; Zhang, N.; Wang, Y.; Zheng, K. Recent Progress in Fluorescent Probes for Detection of Carbonyl Species: Formaldehyde, Carbon Monoxide and Phosgene. Coord. Chem. Rev. 2020, 404, 213109,  DOI: 10.1016/j.ccr.2019.213109
  359. 359
    Hanson, J. R. Terpenoids and Steroids, 7 ed.; Royal Society of Chemistry, 2007.
  360. 360
    Forney, F. W.; Markovetz, A. J. The Biology of Methyl Ketones. J. Lipid Res. 1971, 12 (4), 38395,  DOI: 10.1016/S0022-2275(20)39487-6
  361. 361
    McNally, M. A.; Hartman, A. L. Ketone Bodies in Epilepsy. J. Neurochem. 2012, 121 (1), 2835,  DOI: 10.1111/j.1471-4159.2012.07670.x
  362. 362
    Wakil, S. J.; Stoops, J. K.; Joshi, V. C. Fatty Acid Synthesis and Its Regulation. Annu. Rev. Biochem. 1983, 52 (1), 53779,  DOI: 10.1146/annurev.bi.52.070183.002541
  363. 363
    Dhillon, K. K.; Gupta, S. Biochemistry, Ketogenesis; StatPearls: Treasure Island, FL, 2021.
  364. 364
    Mutti, F. G.; Knaus, T. Enzymes Applied to the Synthesis of Amines. Biocatalysis for Practitioners 2021, 14380,  DOI: 10.1002/9783527824465.ch6
  365. 365
    Kroutil, W.; Mang, H.; Edegger, K.; Faber, K. Recent Advances in the Biocatalytic Reduction of Ketones and Oxidation of sec-Alcohols. Curr. Opin. Chem. Biol. 2004, 8 (2), 1206,  DOI: 10.1016/j.cbpa.2004.02.005
  366. 366
    Leisch, H.; Morley, K.; Lau, P. C. K. Baeyer- Villiger Monooxygenases: More Than Just Green Chemistry. Chem. Rev. 2011, 111 (7), 4165222,  DOI: 10.1021/cr1003437
  367. 367
    Dean, S. M.; Greenberg, W. A.; Wong, C. H. Recent Advances in Aldolase-Catalyzed Asymmetric Synthesis. Adv. Synth. Catal. 2007, 349 (8–9), 130820,  DOI: 10.1002/adsc.200700115
  368. 368
    Walsh, C. T. Biologically Generated Carbon Dioxide: Nature’s Versatile Chemical Strategies for Carboxy Lyases. Nat. Prod. Rep. 2020, 37 (1), 10035,  DOI: 10.1039/C9NP00015A
  369. 369
    Liu, M.; Wei, D.; Wen, Z.; Wang, J.-b. Progress in Stereoselective Construction of C-C Bonds Enabled by Aldolases and Hydroxynitrile Lyases. Front. Bioeng. Biotechnol. 2021, 9, 653682,  DOI: 10.3389/fbioe.2021.653682
  370. 370
    Müller, M.; Gocke, D.; Pohl, M. Thiamin Diphosphate in Biological Chemistry: Exploitation of Diverse Thiamin Diphosphate-Dependent Enzymes for Asymmetric Chemoenzymatic Synthesis. FEBS J. 2009, 276 (11), 2894904,  DOI: 10.1111/j.1742-4658.2009.07017.x
  371. 371
    Schmidt, N. G.; Pavkov-Keller, T.; Richter, N.; Wiltschi, B.; Gruber, K.; Kroutil, W. Biocatalytic Friedel-Crafts Acylation and Fries Reaction. Angew. Chem., Int. Ed. 2017, 56 (26), 76159,  DOI: 10.1002/anie.201703270
  372. 372
    Gergel, S.; Soler, J.; Klein, A.; Schülke, K.; Hauer, B.; Garcia-Borràs, M.; Hammer, S. Directed Evolution of a Ketone Synthase for Efficient and Highly Selective Functionalization of Internal Alkenes by Accessing Reactive Carbocation Intermediates. ChemRxiv 2022. dp94p DOI: 10.26434/chemrxiv-2022-dp94p
  373. 373
    Zhang, J.; Xu, T.; Li, Z. Enantioselective Biooxidation of Racemic trans-Cyclic Vicinal Diols: One-Pot Synthesis of Both Enantiopure (S,S)-Cyclic Vicinal Diols and (R)-α-Hydroxy Ketones. Adv. Synth. Catal. 2013, 355 (16), 314753,  DOI: 10.1002/adsc.201300301
  374. 374
    Lavandera, I.; Kern, A.; Resch, V.; Ferreira-Silva, B.; Glieder, A.; Fabian, W. M. F.; de Wildeman, S.; Kroutil, W. One-Way Biohydrogen Transfer for Oxidation of sec-Alcohols. Org. Lett. 2008, 10 (11), 21558,  DOI: 10.1021/ol800549f
  375. 375
    González-Granda, S.; Méndez-Sánchez, D.; Lavandera, I.; Gotor-Fernández, V. Laccase-Mediated Oxidations of Propargylic Alcohols. Application in the Deracemization of 1-Arylprop-2-Yn-1-Ols in Combination with Alcohol Dehydrogenases. ChemCatChem. 2020, 12 (2), 5207,  DOI: 10.1002/cctc.201901543
  376. 376
    Key, J. A.; Li, C.; Cairo, C. W. Detection of Cellular Sialic Acid Content Using Nitrobenzoxadiazole Carbonyl-Reactive Chromophores. Bioconj. Chem. 2012, 23 (3), 36371,  DOI: 10.1021/bc200276k
  377. 377
    Crisalli, P.; Kool, E. T. Water-Soluble Organocatalysts for Hydrazone and Oxime Formation. J. Org. Chem. 2013, 78 (3), 11849,  DOI: 10.1021/jo302746p
  378. 378
    Guo, H.-M.; Tanaka, F. A Fluorogenic Aldehyde Bearing a 1,2,3-Triazole Moiety for Monitoring the Progress of Aldol Reactions. J. Org. Chem. 2009, 74 (6), 241724,  DOI: 10.1021/jo900013w
  379. 379
    Tanaka, F.; Thayumanavan, R.; Barbas, C. F. Fluorescent Detection of Carbon-Carbon Bond Formation. J. Am. Chem. Soc. 2003, 125 (28), 85238,  DOI: 10.1021/ja034069t
  380. 380
    Tanaka, F.; Mase, N.; Barbas, C. F. Design and Use of Fluorogenic Aldehydes for Monitoring the Progress of Aldehyde Transformations. J. Am. Chem. Soc. 2004, 126 (12), 36923,  DOI: 10.1021/ja049641a
  381. 381
    Linares-Pastén, J. A.; Chávez-Lizárraga, G.; Villagomez, R.; Mamo, G.; Hatti-Kaul, R. A Method for Rapid Screening of Ketone Biotransformations: Detection of Whole Cell Baeyer-Villiger Monooxygenase Activity. Enzyme Microb. Technol. 2012, 50 (2), 1016,  DOI: 10.1016/j.enzmictec.2011.10.004
  382. 382
    Georgiana Ileana, B.; Gabriel Lucian, R. Carboxylic Acid - Key Role in Life Sciences; IntechOpen, 2018.
  383. 383
    Brune, K.; Patrignani, P. New Insights into the Use of Currently Available Non-Steroidal Anti-Inflammatory Drugs. J. Pain Res. 2015, 8, 10518,  DOI: 10.2147/JPR.S75160
  384. 384
    Murali, N.; Srinivas, K.; Ahring, B. K. Biochemical Production and Separation of Carboxylic Acids for Biorefinery Applications. Fermentation 2017, 3 (2), 22,  DOI: 10.3390/fermentation3020022
  385. 385
    Simmons, D. L.; Botting, R. M.; Hla, T. Cyclooxygenase Isozymes: The Biology of Prostaglandin Synthesis and Inhibition. Pharmacol. Rev. 2004, 56 (3), 387437,  DOI: 10.1124/pr.56.3.3
  386. 386
    Kumar, P.; Dubey, K. K. Citric Acid Cycle Regulation: Back Bone for Secondary Metabolite Production. In New and Future Developments in Microbial Biotechnology and Bioengineering; Elsevier: Amsterdam, 2019; Chapter 13, pp 16581.
  387. 387
    Nakamura, M. T.; Yudell, B. E.; Loor, J. J. Regulation of Energy Metabolism by Long-Chain Fatty Acids. Prog. Lipid Res. 2014, 53, 12444,  DOI: 10.1016/j.plipres.2013.12.001
  388. 388
    Barrett, G. C.; Elmore, D. T. Amino Acids and Peptides; Cambridge University Press: Cambridge, 1998 DOI: 10.1017/CBO9781139163828 .
  389. 389
    Ogliaruso, M. A.; Wolfe, J. F. Synthesis of Carboxylic Acids. Esters and Their Derivatives; Wiley Online Library, 1979; Vol. 267 DOI: 10.1002/9780470771587.ch7 .
  390. 390
    Wang, H.; Fan, H.; Sun, H.; Zhao, L.; Wei, D. Process Development for the Production of (R)-(−)-Mandelic Acid by Recombinant Escherichia Coli Cells Harboring Nitrilase from Burkholderia Cenocepacia J2315. Org. Process Res. Dev. 2015, 19 (12), 20126,  DOI: 10.1021/acs.oprd.5b00269
  391. 391
    Martínková, L.; Rucká, L.; Nešvera, J.; Pátek, M. Recent Advances and Challenges in the Heterologous Production of Microbial Nitrilases for Biocatalytic Applications. World J. Microbiol. Biotechnol. 2017, 33, 8,  DOI: 10.1007/s11274-016-2173-6
  392. 392
    Xu, Z.; Xiong, N.; Zou, S.-P.; Liu, Y.-X.; liu, Z.-Q.; Xue, Y.-P.; Zheng, Y.-G. Highly Efficient Conversion of 1-Cyanocycloalkaneacetonitrile Using a “Super Nitrilase Mutant”. Bioprocess Biosyst. Eng. 2019, 42 (3), 45563,  DOI: 10.1007/s00449-018-2049-2
  393. 393
    Glieder, A.; Weis, R.; Skranc, W.; Poechlauer, P.; Dreveny, I.; Majer, S.; Wubbolts, M.; Schwab, H.; Gruber, K. Comprehensive Step-by-Step Engineering of an (R)-Hydroxynitrile Lyase for Large-Scale Asymmetric Synthesis. Angew. Chem., Int. Ed. 2003, 42 (39), 48158,  DOI: 10.1002/anie.200352141
  394. 394
    Jiang, M.; Liu, Y.-J.; Liu, G.-S.; Zhou, W.; Shen, Y.-L.; Gao, B.; Wang, F.-Q.; Wei, D.-Z. Sequential Resolution of (S) and (R)-6-Fluoro-Chroman-2-Carboxylic Acid by Two Esterases in Turn. Green Chem. 2022, 24 (8), 323542,  DOI: 10.1039/D1GC04512A
  395. 395
    Dijkman, W. P.; Groothuis, D. E.; Fraaije, M. W. Enzyme-Catalyzed Oxidation of 5-Hydroxymethylfurfural to Furan-2,5-Dicarboxylic Acid. Angew. Chem., Int. Ed. 2014, 53 (25), 65158,  DOI: 10.1002/anie.201402904
  396. 396
    DeSantis, G.; Wong, K.; Farwell, B.; Chatman, K.; Zhu, Z.; Tomlinson, G.; Huang, H.; Tan, X.; Bibbs, L.; Chen, P. Creation of a Productive, Highly Enantioselective Nitrilase through Gene Site Saturation Mutagenesis (GSSM). J. Am. Chem. Soc. 2003, 125 (38), 114767,  DOI: 10.1021/ja035742h
  397. 397
    Franz, R. G. Comparisons of pKa and Log P Values of Some Carboxylic and Phosphonic Acids: Synthesis and Measurement. AAPS PharmSci. 2001, 3 (2), 1,  DOI: 10.1208/ps030210
  398. 398
    Toth, A. M.; Liptak, M. D.; Phillips, D. L.; Shields, G. C. Accurate Relative Pka Calculations for Carboxylic Acids Using Complete Basis Set and Gaussian-N Models Combined with Continuum Solvation Methods. J. Chem. Phys. 2001, 114 (10), 4595606,  DOI: 10.1063/1.1337862
  399. 399
    Vollhardt, K. P. C. Organische Chemie, 3rd ed.; Wiley-VCH: Weinheim, 2000.
  400. 400
    Bruice, P. Y. Organic Chemistry, 5 ed.; Pearson Prentice Hall: Upper Saddle River, NJ, 2007.
  401. 401
    Descalzo, A. B.; Rurack, K.; Weisshoff, H.; Martínez-Máñez, R.; Marcos, M. D.; Amorós, P.; Hoffmann, K.; Soto, J. Rational Design of a Chromo- and Fluorogenic Hybrid Chemosensor Material for the Detection of Long-Chain Carboxylates. J. Am. Chem. Soc. 2005, 127 (1), 184200,  DOI: 10.1021/ja045683n
  402. 402
    Reddy G, U.; Lo, R.; Roy, S.; Banerjee, T.; Ganguly, B.; Das, A. A New Receptor with a FRET Based Fluorescence Response for Selective Recognition of Fumaric and Maleic Acids in Aqueous Medium. Chem. Commun. 2013, 49 (84), 9818,  DOI: 10.1039/c3cc45051a
  403. 403
    Fitzmaurice, R. J.; Kyne, G. M.; Douheret, D.; Kilburn, J. D. Synthetic Receptors for Carboxylic Acids and Carboxylates. J. Chem. Soc., Perkin Trans. 1 2002, (7), 84164,  DOI: 10.1039/b009041g
  404. 404
    Janes, L. E.; Löwendahl, A. C.; Kazlauskas, R. J. Quantitative Screening of Hydrolase Libraries Using pH Indicators: Identifying Active and Enantioselective Hydrolases. Eur. J. Chem. 1998, 4 (11), 232431,  DOI: 10.1002/(SICI)1521-3765(19981102)4:11<2324::AID-CHEM2324>3.0.CO;2-I
  405. 405
    Baumann, M.; Stürmer, R.; Bornscheuer, U. T. A High-Throughput-Screening Method for the Identification of Active and Enantioselective Hydrolases. Angew. Chem., Int. Ed. 2001, 40 (22), 42014,  DOI: 10.1002/1521-3773(20011119)40:22<4201::AID-ANIE4201>3.0.CO;2-V
  406. 406
    Wang, J.; Liu, H.-B.; Tong, Z.; Ha, C.-S. Fluorescent/Luminescent Detection of Natural Amino Acids by Organometallic Systems. Coord. Chem. Rev. 2015, 303, 13984,  DOI: 10.1016/j.ccr.2015.05.008
  407. 407
    Gong, R.; Mu, H.; Sun, Y.; Fang, X.; Xue, P.; Fu, E. The First Fluorescent Sensor for Medium-Chain Fatty Acids in Water: Design, Synthesis and Sensing Properties of an Organic-Inorganic Hybrid Material. J. Mater. Chem. B 2013, 1 (15), 203847,  DOI: 10.1039/c3tb00355h
  408. 408
    Kusukawa, T.; Tanaka, S.; Inoue, K. Fluorescent Detection of Amidinium-Carboxylate and Amidinium Formation Using a 1,8-Diphenylnaphthalene-Based Diamidine: Dicarboxylic Acid Recognition with High Fluorescence Efficiency. Tetrahedron 2014, 70 (26), 404956,  DOI: 10.1016/j.tet.2014.03.036
  409. 409
    Boiocchi, M.; Bonizzoni, M.; Fabbrizzi, L.; Piovani, G.; Taglietti, A. A Dimetallic Cage with a Long Ellipsoidal Cavity for the Fluorescent Detection of Dicarboxylate Anions in Water. Angew. Chem., Int. Ed. 2004, 43 (29), 384752,  DOI: 10.1002/anie.200460036
  410. 410
    Metzger, A.; Lynch, V. M.; Anslyn, E. V. A Synthetic Receptor Selective for Citrate. Angew. Chem., Int. Ed. 1997, 36 (8), 8625,  DOI: 10.1002/anie.199708621
  411. 411
    Yang, X.; Liu, X.; Shen, K.; Zhu, C.; Cheng, Y. A Chiral Perazamacrocyclic Fluorescent Sensor for Cascade Recognition of Cu(II) and the Unmodified α-Amino Acids in Protic Solutions. Org. Lett. 2011, 13 (13), 35103,  DOI: 10.1021/ol2013268
  412. 412
    Görög, S.; Laukó, A.; Rényei, M.; Hegedüs, B. New Derivatization Reactions in Pharmaceutical Analysis. J. Pharm. Biomed. Anal. 1983, 1 (4), 497506,  DOI: 10.1016/0731-7085(83)80063-6
  413. 413
    Munson, J. W.; Bilous, R. Colorimetric Determination of Aliphatic Acids. J. Pharm. Sci. 1977, 66 (10), 14035,  DOI: 10.1002/jps.2600661013
  414. 414
    Miwa, H.; Yamamoto, M.; Momose, T. Colorimetric Detection and Determination of Carboxylic Acids with 2-Nitrophenylhydrazine Hydrochloride. Chem. Pharm. Bull. (Tokyo) 1980, 28 (2), 599605,  DOI: 10.1248/cpb.28.599
  415. 415
    Hill, U. T. Colorimetric Determination of Fatty Acids and Esters. Ind. Eng. Chem. Anal. Ed. 1946, 18 (5), 3179,  DOI: 10.1021/i560153a017
  416. 416
    Montgomery, H. A. C.; Dymock, J. F.; Thom, N. S. The Rapid Colorimetric Determination of Organic Acids and Their Salts in Sewage-Sludge Liquor. Analyst 1962, 87 (1041), 94955,  DOI: 10.1039/an9628700949
  417. 417
    Kasai, Y.; Tanimura, T.; Tamura, Z. J. A. C. Spectrophotometric Determination of Carboxylic Acids by the Formation of Hydroxamic Acids with Dicyclohexylcarbodiimide. Anal. Chem. 1975, 47 (1), 347,  DOI: 10.1021/ac60351a047
  418. 418
    Takeuchi, T.; Horikawa, R.; Tanimura, T. Spectrophotometry Determination of Carboxylic Acids by Ferric Hydroxamate Formation with Water-Soluble Carbodiimide. Anal. Lett. 1980, 13 (7), 6039,  DOI: 10.1080/00032718008077690
  419. 419
    He, Y.-C.; Ma, C.-L.; Xu, J.-H.; Zhou, L. A High-Throughput Screening Strategy for Nitrile-Hydrolyzing Enzymes Based on Ferric Hydroxamate Spectrophotometry. Appl. Microbiol. Biotechnol. 2011, 89 (3), 81723,  DOI: 10.1007/s00253-010-2977-5
  420. 420
    Goswami, S.; Hazra, A.; Das, M. K. Selenodiazole-Fused Diacetamidopyrimidine, a Selective Fluorescence Sensor for Aliphatic Monocarboxylates. Tetrahedron Lett. 2010, 51 (25), 33203,  DOI: 10.1016/j.tetlet.2010.04.085
  421. 421
    Goswami, S.; Hazra, A.; Chakrabarty, R.; Fun, H.-K. Recognition of Carboxylate Anions and Carboxylic Acids by Selenium-Based New Chromogenic Fluorescent Sensor: A Remarkable Fluorescence Enhancement of Hindered Carboxylates. Org. Lett. 2009, 11 (19), 43503,  DOI: 10.1021/ol901737s
  422. 422
    Galindo, F.; Becerril, J.; Isabel Burguete, M.; Luis, S. V.; Vigara, L. Synthesis and Study of a Cyclophane Displaying Dual Fluorescence Emission: A Novel Ratiometric Sensor for Carboxylic Acids in Organic Medium. Tetrahedron Lett. 2004, 45 (8), 165962,  DOI: 10.1016/j.tetlet.2003.12.116
  423. 423
    Cabell, L. A.; Best, M. D.; Lavigne, J. J.; Schneider, S. E.; Perreault, D. M.; Monahan, M.-K.; Anslyn, E. V. Metal Triggered Fluorescence Sensing of Citrate Using a Synthetic Receptor. J. Chem. Soc., Perkin trans. II 2001, (3), 31523,  DOI: 10.1039/b008694k
  424. 424
    Xu, K.-x.; Xie, X.-m.; Kong, H.-j.; Li, P.; Zhang, J.-l.; Pang, X.-b. Selective Fluorescent Sensors for Malate Anion Using the Complex of Phenanthroline-Based Eu(III) in Aqueous Solution. Sens. Actuators B Chem. 2014, 201, 1317,  DOI: 10.1016/j.snb.2014.04.086
  425. 425
    Burguete, M. I.; Galindo, F.; Luis, S. V.; Vigara, L. A Turn-On Fluorescent Indicator for Citrate with Micromolar Sensitivity. Dalton Trans. 2007, (36), 402733,  DOI: 10.1039/b711139h
  426. 426
    Klein, G.; Reymond, J.-L. An Enzyme Assay Using pM. Angew. Chem., Int. Ed. 2001, 40 (9), 17713,  DOI: 10.1002/1521-3773(20010504)40:9<1771::AID-ANIE17710>3.0.CO;2-M
  427. 427
    Dean, K. E. S.; Klein, G.; Renaudet, O.; Reymond, J.-L. A Green Fluorescent Chemosensor for Amino Acids Provides a Versatile High-Throughput Screening (HTS) Assay for Proteases. Bioorg. Med. Chem. Lett. 2003, 13 (10), 16536,  DOI: 10.1016/S0960-894X(03)00280-4
  428. 428
    Xu, J.-M.; Fu, F.-T.; Hu, H.-F.; Zheng, Y.-G. A High-Throughput Screening Method for Amino Acid Dehydrogenase. Anal. Biochem. 2016, 495, 2931,  DOI: 10.1016/j.ab.2015.11.012
  429. 429
    Ryu, D.; Park, E.; Kim, D.-S.; Yan, S.; Lee, J. Y.; Chang, B.-Y.; Ahn, K. H. A Rational Approach to Fluorescence “Turn-On” Sensing of α-Amino-Carboxylates. J. Am. Chem. Soc. 2008, 130 (8), 23945,  DOI: 10.1021/ja078308e
  430. 430
    Kim, D.-S.; Ahn, K. H. Fluorescence “Turn-On” Sensing of Carboxylate Anions with Oligothiophene-Based o-(Carboxamido)Trifluoroacetophenones. J. Org. Chem. 2008, 73 (17), 68314,  DOI: 10.1021/jo801178y
  431. 431
    Schrader, J.; Etschmann, M. M. W.; Sell, D.; Hilmer, J. M.; Rabenhorst, J. Applied Biocatalysis for the Synthesis of Natural Flavour Compounds - Current Industrial Processes and Future Prospects. Biotechnol. Lett. 2004, 26 (6), 46372,  DOI: 10.1023/B:BILE.0000019576.80594.0e
  432. 432
    Gotor, V. Biocatalysis Applied to the Preparation of Pharmaceuticals. Org. Process Res. Dev. 2002, 6 (4), 4206,  DOI: 10.1021/op020008o
  433. 433
    SÁ, A. G. A.; Meneses, A. C. d.; Araújo, P. H. H. d.; Oliveira, D. d. A Review on Enzymatic Synthesis of Aromatic Esters Used as Flavor Ingredients for Food, Cosmetics and Pharmaceuticals Industries. Trends Food Sci. Technol. 2017, 69, 95105,  DOI: 10.1016/j.tifs.2017.09.004
  434. 434
    Panke, S.; Held, M.; Wubbolts, M. Trends and Innovations in Industrial Biocatalysis for the Production of Fine Chemicals. Curr. Opin. Biotechnol. 2004, 15 (4), 2729,  DOI: 10.1016/j.copbio.2004.06.011
  435. 435
    Joshi, A.; Singh, S.; Iqbal, Z.; De, S. R. CO Free Esterifications of (Hetero)Arenes Via Transition-Metal-Catalyzed Chelation-Induced C-H Activation: Recent Updates. Tetrahedron 2022, 104, 132601,  DOI: 10.1016/j.tet.2021.132601
  436. 436
    Larios, A.; García, H. S.; Oliart, R. M.; Valerio-Alfaro, G. Synthesis of Flavor and Fragrance Esters Using Candida Antarctica Lipase. Appl. Microbiol. Biotechnol. 2004, 65 (4), 3736,  DOI: 10.1007/s00253-004-1602-x
  437. 437
    Smith, G. A. Fatty Acid, Methyl Ester, and Vegetable Oil Ethoxylates, Biobased Surfactants; AOCS Press, 2019; pp 287301.
  438. 438
    Panten, J.; Surburg, H. Flavors and Fragrances, 4. Natural Raw Materials, Ullmann’s Encyclopedia of Industrial Chemistry , 2011; pp 158.
  439. 439
    Johannes, P.; Surburg, H. Flavors and Fragrances, 2. Aliphatic Compounds. In Ullmann’s Encyclopedia of Industrial Chemistry , 2015; pp 155.
  440. 440
    Berger, R. G. Biotechnology of Flavours─the Next Generation. Biotechnol. Lett. 2009, 31 (11), 1651,  DOI: 10.1007/s10529-009-0083-5
  441. 441
    Gunatillake, P. A.; Adhikari, R.; Gadegaard, N. Biodegradable Synthetic Polymers for Tissue Engineering. Eur. Cell. Mater. 2003, 5 (1), 116,  DOI: 10.22203/eCM.v005a01
  442. 442
    Kodali, D. R. High Performance Ester Lubricants from Natural Oils. Ind. Lubr. Tribol. 2002, 54 (4), 16570,  DOI: 10.1108/00368790210431718
  443. 443
    Chen, J.; Wang, Y.; Huang, J.; Li, K.; Nie, X. Synthesis of Tung-Oil-Based Triglycidyl Ester Plasticizer and Its Effects on Poly(Vinyl Chloride) Soft Films. ACS Sustain. Chem. Eng. 2018, 6 (1), 64251,  DOI: 10.1021/acssuschemeng.7b02989
  444. 444
    Gumel, A. M.; Annuar, M. S. M.; Heidelberg, T.; Chisti, Y. Lipase Mediated Synthesis of Sugar Fatty Acid Esters. Process Biochem. 2011, 46 (11), 207990,  DOI: 10.1016/j.procbio.2011.07.021
  445. 445
    Badgujar, K. C.; Bhanage, B. M. Application of Lipase Immobilized on the Biocompatible Ternary Blend Polymer Matrix for Synthesis of Citronellyl Acetate in Non-Aqueous Media: Kinetic Modelling Study. Enzyme Microb. Technol. 2014, 57, 1625,  DOI: 10.1016/j.enzmictec.2014.01.006
  446. 446
    Stergiou, P.-Y.; Foukis, A.; Filippou, M.; Koukouritaki, M.; Parapouli, M.; Theodorou, L. G.; Hatziloukas, E.; Afendra, A.; Pandey, A.; Papamichael, E. M. Advances in Lipase-Catalyzed Esterification Reactions. Biotechnol. Adv. 2013, 31 (8), 184659,  DOI: 10.1016/j.biotechadv.2013.08.006
  447. 447
    Gao, W.; Wu, K.; Chen, L.; Fan, H.; Zhao, Z.; Gao, B.; Wang, H.; Wei, D. A Novel Esterase from a Marine Mud Metagenomic Library for Biocatalytic Synthesis of Short-Chain Flavor Esters. Microbial Cell Factories 2016, 15, 41,  DOI: 10.1186/s12934-016-0435-5
  448. 448
    Lozano, P.; Bernal, J. M.; Navarro, A. A Clean Enzymatic Process for Producing Flavour Esters by Direct Esterification in Switchable Ionic Liquid/Solid Phases. Green Chem. 2012, 14 (11), 302633,  DOI: 10.1039/c2gc36081k
  449. 449
    Otera, J.; Nishikido, J. Esterification: Methods, Reactions, and Applications; John Wiley & Sons, 2009. DOI: 10.1002/9783527627622
  450. 450
    Beekwilder, J.; Alvarez-Huerta, M.; Neef, E.; Verstappen, F. W. A.; Bouwmeester, H. J.; Aharoni, A. Functional Characterization of Enzymes Forming Volatile Esters from Strawberry and Banana. Plant Physiol. 2004, 135 (4), 186578,  DOI: 10.1104/pp.104.042580
  451. 451
    Buckles, R. E.; Thelen, C. Qualitative Determination of Carboxylic Esters. Anal. Chem. 1950, 22 (5), 6768,  DOI: 10.1021/ac60041a016
  452. 452
    Stern, I.; Shapiro, B. A Rapid and Simple Method for the Determination of Esterified Fatty Acids and for Total Fatty Acids in Blood. J. Clin. Pathol. 1953, 6 (2), 15860,  DOI: 10.1136/jcp.6.2.158
  453. 453
    Davidson, D. Hydroxamic Acids in Qualitative Organic Analysis. J. Chem. Educ. 1940, 17 (2), 81,  DOI: 10.1021/ed017p81
  454. 454
    Löbs, A.-K.; Lin, J.-L.; Cook, M.; Wheeldon, I. High Throughput, Colorimetric Screening of Microbial Ester Biosynthesis Reveals High Ethyl Acetate Production from Kluyveromyces Marxianus on C5, C6, and C12 Carbon Sources. Biotechnol. J. 2016, 11 (10), 127481,  DOI: 10.1002/biot.201600060
  455. 455
    Smith, J. G. Synthetically Useful Reactions of Epoxides. Synthesis 1984, 1984, 629656,  DOI: 10.1055/s-1984-30921
  456. 456
    Yudin, A. K. Aziridines and Epoxides in Organic Synthesis; John Wiley & Sons, 2006.
  457. 457
    Kotik, M.; Archelas, A.; Wohlgemuth, R. Epoxide Hydrolases and Their Application in Organic Synthesis. Curr. Org. Chem. 2012, 16 (4), 45182,  DOI: 10.2174/138527212799499840
  458. 458
    Zocher, F.; Enzelberger, M. M.; Bornscheuer, U. T.; Hauer, B. D.; Schmid, R. D. A Colorimetric Assay Suitable for Screening Epoxide Hydrolase Activity. Anal. Chim. Acta 1999, 391 (3), 345351,  DOI: 10.1016/S0003-2670(99)00216-0
  459. 459
    Alcalde, M.; Farinas, E. T.; Arnold, F. H. Colorimetric High-Throughput Assay for Alkene Epoxidation Catalyzed by Cytochrome P450 BM-3 Variant 139–3. J. Biomol. Screen. 2004, 9 (2), 1416,  DOI: 10.1177/1087057103261913
  460. 460
    Van Damme, F.; Oomens, A. C. Determination of Residual Free Epoxide in Polyether Polyols by Derivatization with Diethylammonium N,N-Diethyldithiocarbamate and Liquid Chromatography. J. Chromatogr. A 1995, 696 (1), 417,  DOI: 10.1016/0021-9673(94)01270-O
  461. 461
    Hammock, L. G.; Hammock, B. D.; Casida, J. E. Detection and Analysis of Epoxides with 4-(p-Nitrobenzyl)-Pyridine. Bull. Environ. Contam. Toxicol. 1974, 12 (6), 75964,  DOI: 10.1007/BF01685927
  462. 462
    Pedragosa-Moreau, S.; Morisseau, C.; Baratti, J.; Zylber, J.; Archelas, A.; Furstoss, R. Microbiological Transformations 37. An Enantioconvergent Synthesis of the β-Blocker (R)-NiféNalol Using a Combined Chemoenzymatic Approach. Tetrahedron 1997, 53 (28), 970714,  DOI: 10.1016/S0040-4020(97)00639-X
  463. 463
    Finney, N. S. Enantioselective Epoxide Hydrolysis: Catalysis Involving Microbes, Mammals and Metals. J. Biol. Chem. 1998, 5 (4), R73R9,  DOI: 10.1016/S1074-5521(98)90630-5
  464. 464
    Choi, W. J.; Choi, C. Y. Production of Chiral Epoxides: Epoxide Hydrolase-Catalyzed Enantioselective Hydrolysis. Biotechnol. Bioprocess Eng. 2005, 10 (3), 167,  DOI: 10.1007/BF02932009
  465. 465
    Baer, H.; Bergamo, M.; Forlin, A.; Pottenger, L. H.; Lindner, J. Propylene Oxide. In Ullmann’s Encyclopedia of Industrial Chemistry: Wiley-VCH, 2000.
  466. 466
    Rebsdat, S.; Mayer, D. Ethylene Oxide. In Ullmann’s Encyclopedia of Industrial Chemistry; Wiley-VCH, 2001.
  467. 467
    Archelas, A.; Furstoss, R. Synthesis of Enantiopure Epoxides through Biocatalytic Approaches. Annu. Rev. Microbiol. 1997, 51 (1), 491525,  DOI: 10.1146/annurev.micro.51.1.491
  468. 468
    Ma, S. K.; Gruber, J.; Davis, C.; Newman, L.; Gray, D.; Wang, A.; Grate, J.; Huisman, G. W.; Sheldon, R. A. A Green-by-Design Biocatalytic Process for Atorvastatin Intermediate. Green Chem. 2010, 12 (1), 816,  DOI: 10.1039/B919115C
  469. 469
    Kong, X.-D.; Ma, Q.; Zhou, J.; Zeng, B.-B.; Xu, J.-H. A Smart Library of Epoxide Hydrolase Variants and the Top Hits for Synthesis of (S)-β-Blocker Precursors. Angew. Chem., Int. Ed. Engl. 2014, 53 (26), 66414,  DOI: 10.1002/anie.201402653
  470. 470
    Lin, Z.; Xue, Y.; Liang, X.-W.; Wang, J.; Lin, S.; Tao, J.; You, S.-L.; Liu, W. Oxidative Indole Dearomatization for Asymmetric Furoindoline Synthesis by a Flavin-Dependent Monooxygenase Involved in the Biosynthesis of Bicyclic Thiopeptide Thiostrepton. Angew. Chem. 2021, 60 (15), 84015,  DOI: 10.1002/anie.202013174
  471. 471
    Agarwal, S. C.; Van Duuren, B. L.; Solomon, J. J.; Kline, S. A. Reaction of Epoxides with 4-Nitrothiophenol. Its Possible Application for Trapping and Characterization of Epoxides. Environ. Sci. Technol. 1980, 14 (10), 124953,  DOI: 10.1021/es60170a012
  472. 472
    Wixtrom, R. N.; Hammock, B. D. Continuous Spectrophotometric Assays for Cytosolic Epoxide Hydrolase. Anal. Biochem. 1988, 174 (1), 2919,  DOI: 10.1016/0003-2697(88)90548-9
  473. 473
    González-Pérez, M.; Gómez-Bombarelli, R.; Arenas-Valgañón, J.; Pérez-Prior, M. T.; García-Santos, M. P.; Calle, E.; Casado, J. Connecting the Chemical and Biological Reactivity of Epoxides. Chem. Res. Toxicol. 2012, 25 (12), 275562,  DOI: 10.1021/tx300389z
  474. 474
    Nelis, H. J. C. F.; Sinsheimer, J. E. A Sensitive Fluorimetric Procedure for the Determination of Aliphatic Epoxides under Physiological Conditions. Anal. Biochem. 1981, 115 (1), 1517,  DOI: 10.1016/0003-2697(81)90538-8
  475. 475
    Rink, R.; Fennema, M.; Smids, M.; Dehmel, U.; Janssen, D. B. Primary Structure and Catalytic Mechanism of the Epoxide Hydrolase from Agrobacterium Radiobacter Ad1. J. Biol. Chem. 1997, 272 (23), 146507,  DOI: 10.1074/jbc.272.23.14650
  476. 476
    Elder, D. P.; Snodin, D.; Teasdale, A. Analytical Approaches for the Detection of Epoxides and Hydroperoxides in Active Pharmaceutical Ingredients, Drug Products and Herbals. J. Pharm. Biomed. Anal. 2010, 51 (5), 101523,  DOI: 10.1016/j.jpba.2009.11.023
  477. 477
    Sano, A.; Takitani, S. Spectrofluorometric Determination of Common Epoxides with Sodium Sulfide and o-Phthalaldehyde and Taurine Reagents. Anal. Chem. 1985, 57 (8), 168790,  DOI: 10.1021/ac00285a039
  478. 478
    Tang, L.; Li, Y.; Wang, X. A High-Throughput Colorimetric Assay for Screening Halohydrin Dehalogenase Saturation Mutagenesis Libraries. J. Biotechnol. 2010, 147 (3), 1648,  DOI: 10.1016/j.jbiotec.2010.04.002
  479. 479
    Gul, I.; Bogale, T. F.; Deng, J.; Chen, Y.; Fang, R.; Feng, J.; Tang, L. Enzyme-Based Detection of Epoxides Using Colorimetric Assay Integrated with Smartphone Imaging. Biotechnol. Appl. Biochem. 2020, 67 (4), 68592,  DOI: 10.1002/bab.1898
  480. 480
    Dansette, P. M.; DuBois, G. C.; Jerina, D. M. Continuous Fluorometric Assay of Epoxide Hydrase Activity. Anal. Biochem. 1979, 97 (2), 3405,  DOI: 10.1016/0003-2697(79)90083-6
  481. 481
    Mateo, C.; Archelas, A.; Furstoss, R. A Spectrophotometric Assay for Measuring and Detecting an Epoxide Hydrolase Activity. Anal. Biochem. 2003, 314 (1), 13541,  DOI: 10.1016/S0003-2697(02)00646-2
  482. 482
    Michałowicz, J.; Duda, W. Phenols - Sources and Toxicity. Poish J. Environ. Stud. 2007, 16 (3), 347362
  483. 483
    Schmidt, R. J. Industrial Catalytic Processes─Phenol Production. Appl. Catal. A: Gen. 2005, 280 (1), 89103,  DOI: 10.1016/j.apcata.2004.08.030
  484. 484
    Ullrich, R.; Hofrichter, M. Enzymatic Hydroxylation of Aromatic Compounds. Cell. Mol. Life Sci. 2007, 64 (3), 27193,  DOI: 10.1007/s00018-007-6362-1
  485. 485
    Lan Tee, K.; Schwaneberg, U. Directed Evolution of Oxygenases: Screening Systems, Success Stories and Challenges. Comb. Chem. High Throughput Screen. 2007, 10 (3), 197217,  DOI: 10.2174/138620707780126723
  486. 486
    Joern, J. M.; Sakamoto, T.; Arisawa, A.; Arnold, F. H. A Versatile High Throughput Screen for Dioxygenase Activity Using Solid-Phase Digital Imaging. J. Biomol. Screen. 2001, 6 (4), 21923,  DOI: 10.1177/108705710100600403
  487. 487
    Poraj-Kobielska, M.; Kinne, M.; Ullrich, R.; Scheibner, K.; Kayser, G.; Hammel, K. E.; Hofrichter, M. Preparation of Human Drug Metabolites Using Fungal Peroxygenases. Biochem. Pharmacol. 2011, 82 (7), 78996,  DOI: 10.1016/j.bcp.2011.06.020
  488. 488
    Sakamoto, T.; Joern, J. M.; Arisawa, A.; Arnold, F. H. Laboratory Evolution of Toluene Dioxygenase to Accept 4-Picoline as a Substrate. Appl. Environ. Microbiol. 2001, 67 (9), 3882,  DOI: 10.1128/AEM.67.9.3882-3887.2001
  489. 489
    Foti, M. C. Antioxidant Properties of Phenols. J. Pharm. Pharmacol. 2010, 59 (12), 167385,  DOI: 10.1211/jpp.59.12.0010
  490. 490
    del Rosario Garcia-Mateos, M.; Palma-Tenango, M.; Soto-Hernandez, M. Phenolic Compounds-Biological Activity; IntechOpen, 2017.
  491. 491
    Ainsworth, E. A.; Gillespie, K. M. Estimation of Total Phenolic Content and Other Oxidation Substrates in Plant Tissues Using Folin-Ciocalteu Reagent. Nat. Protoc. 2007, 2 (4), 8757,  DOI: 10.1038/nprot.2007.102
  492. 492
    Singleton, V. L.; Orthofer, R.; Lamuela-Raventós, R. M. Analysis of Total Phenols and Other Oxidation Substrates and Antioxidants by Means of Folin–Ciocalteu Reagent. Methods in Enzymology; Academic Press, 1999; Vol. 299; p 152178. DOI: 10.1016/S0076-6879(99)99017-1
  493. 493
    Platzer, M.; Kiese, S.; Herfellner, T.; Schweiggert-Weisz, U.; Eisner, P. How Does the Phenol Structure Influence the Results of the Folin-Ciocalteu Assay?. Antioxidants 2021, 10 (5), 811,  DOI: 10.3390/antiox10050811
  494. 494
    Miliauskas, G.; Venskutonis, P. R.; van Beek, T. A. Screening of Radical Scavenging Activity of Some Medicinal and Aromatic Plant Extracts. Food Chem. 2004, 85 (2), 2317,  DOI: 10.1016/j.foodchem.2003.05.007
  495. 495
    Cao, G.; Alessio, H. M.; Cutler, R. G. Oxygen-Radical Absorbance Capacity Assay for Antioxidants. Free Radical Biol. Med. 1993, 14 (3), 30311,  DOI: 10.1016/0891-5849(93)90027-R
  496. 496
    Prior, R. L.; Wu, X.; Schaich, K. Standardized Methods for the Determination of Antioxidant Capacity and Phenolics in Foods and Dietary Supplements. J. Agric. Food Chem. 2005, 53 (10), 4290302,  DOI: 10.1021/jf0502698
  497. 497
    van den Berg, R.; Haenen, G. R. M. M.; van den Berg, H.; Bast, A. Applicability of an Improved Trolox Equivalent Antioxidant Capacity (TEAC) Assay for Evaluation of Antioxidant Capacity Measurements of Mixtures. Food Chem. 1999, 66 (4), 5117,  DOI: 10.1016/S0308-8146(99)00089-8
  498. 498
    Dudonné, S.; Vitrac, X.; Coutière, P.; Woillez, M.; Mérillon, J.-M. Comparative Study of Antioxidant Properties and Total Phenolic Content of 30 Plant Extracts of Industrial Interest Using DPPH, ABTS, FRAP, SOD, and ORAC Assays. J. Agric. Food Chem. 2009, 57 (5), 176874,  DOI: 10.1021/jf803011r
  499. 499
    Thaipong, K.; Boonprakob, U.; Crosby, K.; Cisneros-Zevallos, L.; Hawkins Byrne, D. Comparison of ABTS, DPPH, FRAP, and ORAC Assays for Estimating Antioxidant Activity from Guava Fruit Extracts. J. Food Compos. Anal. 2006, 19 (6), 66975,  DOI: 10.1016/j.jfca.2006.01.003
  500. 500
    da Silva Granja, B.; Honório de Mendonça Filho, J. R.; de Souza Oliveira, W.; Caldas Santos, J. C. Exploring MBTH as a Spectrophotometric Probe for the Determination of Total Phenolic Compounds in Beverage Samples. Anal. Methods 2018, 10 (19), 2197204,  DOI: 10.1039/C8AY00464A
  501. 501
    Wong, T. S.; Wu, N.; Roccatano, D.; Zacharias, M.; Schwaneberg, U. Sensitive Assay for Laboratory Evolution of Hydroxylases toward Aromatic and Heterocyclic Compounds. J. Biomol. Screen. 2005, 10 (3), 24652,  DOI: 10.1177/1087057104273336
  502. 502
    Emerson, E. The Condensation of Aminoantipyrine. II. A New Color Test for Phenolic Compounds. J. Org. Chem. 1943, 8 (5), 417428,  DOI: 10.1021/jo01193a004
  503. 503
    Johnson, C. A.; Savidge, R. A. The Determination of Phenolic Compounds in Pharmaceutical Preparations Using 4-Aminohenazone. J. Pharm. Pharmacol. 2011, 10, 171T81T,  DOI: 10.1111/j.2042-7158.1958.tb10396.x
  504. 504
    Varvounis, G.; Fiamegos, Y.; Pilidis, G. Pyrazol-3-Ones. Part III: Reactivity of the Ring Substituents. In Advances in Heterocyclic Chemistry; Academic Press, 2007; Vol. 95; p 27141
  505. 505
    Dannis, M. Determination of Phenols by the Amino-Antipyrine Method. Sewage Ind. Waste 1951, 23, 15161522
  506. 506
    Fiamegos, Y. C.; Stalikas, C. D.; Pilidis, G. A.; Karayannis, M. I. Synthesis and Analytical Applications of 4-Aminopyrazolone Derivatives as Chromogenic Agents for the Spectrophotometric Determination of Phenols. Anal. Chim. Acta 2000, 403 (1), 31523,  DOI: 10.1016/S0003-2670(99)00644-3
  507. 507
    Li, D.; Ge, S.; Huang, J.; Gong, J.; Yan, P.; Lu, W.; Tiana, G.; Ding, L. Fast Chromogenic Identification of Phenolic Pollutants Via Homogeneous Oxidation with t-BuOOH in the Presence of Iron (III) Octacarboxyphthalocyanine. Catal. Commun. 2014, 45, 959,  DOI: 10.1016/j.catcom.2013.10.038
  508. 508
    Fiamegos, Y.; Stalikas, C. D.; Pilidis, G. A. 4-Aminoantipyrine Spectrophotometric Method of Phenol Analysis: Study of the Reaction Products Via Liquid Chromatography with Diode-Array and Mass Spectrometric Detection. Anal. Chim. Acta 2002, 467 (1–2), 10514,  DOI: 10.1016/S0003-2670(02)00072-7
  509. 509
    Ayad, M.; El-Sadek, M.; Mostaffa, S. 4-Aminoantipyrine as an Analytical Reagent for the Colorimetric Determination of Tetracycline and Oxytetracycline. Anal. Lett. 1986, 19 (21–22), 216981,  DOI: 10.1080/00032718608080875
  510. 510
    Otey, C. R.; Joern, J. M. High-Throughput Screen for Aromatic Hydroxylation, Directed Enzyme Evolution: Screening and Selection Methods; Humana Press: Totowa, NJ, 2003; p 1418.
  511. 511
    Santos, G. d. A.; Dhoke, G. V.; Davari, M. D.; Ruff, A. J.; Schwaneberg, U. Directed Evolution of P450 BM3 Towards Functionalization of Aromatic O-Heterocycles. Int. J. Mol. Sci. 2019, 20 (13), 3353,  DOI: 10.3390/ijms20133353
  512. 512
    Lülsdorf, N.; Vojcic, L.; Hellmuth, H.; Weber, T. T.; Mußmann, N.; Martinez, R.; Schwaneberg, U. A First Continuous 4-Aminoantipyrine (4-AAP)-Based Screening System for Directed Esterase Evolution. Appl. Microbiol. Biotechnol. 2015, 99 (12), 523746,  DOI: 10.1007/s00253-015-6612-3
  513. 513
    Choi, S.-L.; Rha, E.-G.; Kim, D.-Y.; Song, J.-J.; Hong, S.-P.; Sung, M.-H.; Lee, S.-G. High Throughput Screening and Directed Evolution of Tyrosine Phenol-Lyase. Microbiol. Biotechnol. Lett. 2006, 34, 5862
  514. 514
    Quintana, M. G.; Didion, C.; Dalton, H. Colorimetric Method for a Rapid Detection of Oxygenated Aromatic Biotransformation Products. Biotechnol. Technol. 1997, 11 (8), 5857,  DOI: 10.1023/A:1018499024466
  515. 515
    Dacre, J. C. Nonspecificity of the Gibbs Reaction. Anal. Chem. 1971, 43 (4), 58991,  DOI: 10.1021/ac60299a015
  516. 516
    Josephy, P. D.; Van Damme, A. Reaction of Gibbs Reagent with Para-Substituted Phenols. Anal. Chem. 1984, 56 (4), 8134,  DOI: 10.1021/ac00268a052
  517. 517
    Svobodová, D.; Křenek, P.; Fraenkl, M.; Gasparič, J. The Colour Reaction of Phenols with the Gibbs Reagent. Microchim. Acta 1978, 70 (3), 197211,  DOI: 10.1007/BF01201610
  518. 518
    Ang, E. L.; Obbard, J. P.; Zhao, H. Directed Evolution of Aniline Dioxygenase for Enhanced Bioremediation of Aromatic Amines. Appl. Microbiol. Biotechnol. 2009, 81 (6), 106370,  DOI: 10.1007/s00253-008-1710-0
  519. 519
    Bornscheuer, U. T.; Ordoñez, G. R.; Hidalgo, A.; Gollin, A.; Lyon, J.; Hitchman, T. S.; Weiner, D. P. Selectivity of Lipases and Esterases Towards Phenol Esters. J. Mol. Catal. B: Enzym. 2005, 36 (1), 813,  DOI: 10.1016/j.molcatb.2005.07.004
  520. 520
    Altahir, B. M.; Abdulazeez, O.; Dikran, S. B.; Taylor, K. E. Determination of Eugenol in Personal-Care Products by Dispersive Liquid-Liquid Microextraction Followed by Spectrophotometry Using p-Amino-N,N-Dimethylaniline as a Derivatizing Agent. Indones. J. Chem. 2020, 21, 192,  DOI: 10.22146/ijc.56198
  521. 521
    Gasparic, J.; Svobodova, D.; Pospisilova, M. Investigation of the Colour Reaction of Phenols with the MBTH Reagent. Identification of Organic Compounds. LXXXVI. Microchim. Acta 1977, 67, 24150,  DOI: 10.1007/BF01213034
  522. 522
    Pospíšilová, M.; Polášek, M.; Svobodová, D. Spectrophotometric Study of Reactions of Substituted Phenols with MBTH in Alkaline Medium: The Effect of Phenol Structure on the Formation of Analytically Useful Coloured Products. Microchimica Acta 1998, 129 (3), 2018,  DOI: 10.1007/BF01244742
  523. 523
    Kamata, E. Color Reactions of 3-Methyl-2-Benzothiazolone Hydrazone with Phenol Derivatives. I. The Spectrophotometric Microdetermination of Catechol, Hydroquinone and Resorcinol. Bull. Chem. Soc. Jpn. 1964, 37 (11), 16747,  DOI: 10.1246/bcsj.37.1674
  524. 524
    Setti, L.; Giuliani, S.; Spinozzi, G.; Pifferi, P. G. Laccase Catalyzed-Oxidative Coupling of 3-Methyl 2-Benzothiazolinone Hydrazone and Methoxyphenols. Enzyme Microb. Technol. 1999, 25 (3), 2859,  DOI: 10.1016/S0141-0229(99)00059-9
  525. 525
    Setti, L.; Scali, S.; Angeli, I. D.; Pifferi, P. G. Horseradish Peroxidase-Catalyzed Oxidative Coupling of 3-Methyl 2-Benzothiazolinone Hydrazone and Methoxyphenols. Enzyme Microb. Technol. 1998, 22 (8), 65661,  DOI: 10.1016/S0141-0229(97)00259-7
  526. 526
    Nolan, L. C.; O’Connor, K. E. A Spectrophotometric Method for the Quantification of an Enzyme Activity Producing 4-Substituted Phenols: Determination of Toluene-4-Monooxygenase Activity. Anal. Biochem. 2005, 344 (2), 22431,  DOI: 10.1016/j.ab.2005.05.032
  527. 527
    Regitz, M. Diazo Compounds: Properties and Synthesis; Elsevier, 2012.
  528. 528
    Chudgar, R. J. Azo Dyes; Kirk–Othmer Encyclopedia of Chemical Technology, 2000.
  529. 529
    de Keijzer, M. The Delight of Modern Organic Pigment Creations, Issues in Contemporary Oil Paint; Springer International Publishing: Cham, 2014; p 4573.
  530. 530
    Urbányi, T.; Mollica, J. A. Potential Diazo Reagents for Colorimetric Determination. J. Pharm. Sci. 1968, 57 (7), 125760,  DOI: 10.1002/jps.2600570746
  531. 531
    Ehrlich, P. Uber Eine Neue Harnprobe. Dtsch. Med. Wochenschr. 1883, 9 (1), 11,  DOI: 10.1055/s-0029-1196984
  532. 532
    Hassan, S. M.; Walash, M. I.; El-Sayed, S. M.; Abou Ouf, A. M. Colorimetric Determination of Certain Phenol Derivatives in Pharmaceutical Preparations. J. Assoc. Off. Anal. Chem. 1981, 64 (6), 14425,  DOI: 10.1093/jaoac/64.6.1442
  533. 533
    Swaminathan, M. Chemical Estimation of Vitamin B6 in Foods by Means of the Diazo Reaction and the Phenol Reagent. Nature 1940, 145 (3681), 780,  DOI: 10.1038/145780a0
  534. 534
    Whitlock, L. R.; Siggia, S.; Smola, J. E. Spectrophotometric Analysis of Phenols and of Sulfonates by Formation of an Azo Dye. Anal. Chem. 1972, 44 (3), 5326,  DOI: 10.1021/ac60311a021
  535. 535
    Abdel Azeem, S. M.; Al Mohesen, I. A.; Ibrahim, A. M. H. Analysis of Total Phenolic Compounds in Tea and Fruits Using Diazotized Aminobenzenes Colorimetric Spots. Food Chem. 2020, 332, 127392,  DOI: 10.1016/j.foodchem.2020.127392
  536. 536
    Lester, G. E.; Lewers, K. S.; Medina, M. B.; Saftner, R. A. Comparative Analysis of Strawberry Total Phenolics Via Fast Blue BB Vs. Folin-Ciocalteu: Assay Interference by Ascorbic Acid. J. Food Compos. Anal. 2012, 27 (1), 1027,  DOI: 10.1016/j.jfca.2012.05.003
  537. 537
    Pompella, A.; Comporti, M. The Use of 3-Hydroxy-2-Naphthoic Acid Hydrazide and Fast Blue B for the Histochemical Detection of Lipid Peroxidation in Animal Tissues─a Microphotometric Study. Histochemistry 1991, 95, 255262,  DOI: 10.1007/BF00744997
  538. 538
    Tłuścik, F.; Kazubek, A.; Mejbaum-Katzenellenbogen, W. Alkylresorcinols in Rye (Secale Cereale L.) Grains. Vi. Colorimetric Micromethod for the Determination of Alkylresorcinols with the Use of Diazonium Salt, Fast Blue B. Acta Soc. Bot. Polym. 1981, 50, 645,  DOI: 10.5586/asbp.1981.086
  539. 539
    Schotten, C.; Leprevost, S. K.; Yong, L. M.; Hughes, C. E.; Harris, K. D. M.; Browne, D. L. Comparison of the Thermal Stabilities of Diazonium Salts and Their Corresponding Triazenes. Org. Process Res. Dev. 2020, 24 (10), 233641,  DOI: 10.1021/acs.oprd.0c00162
  540. 540
    Johnston, K. J.; Ashford, A. E. A Simultaneous-Coupling Azo Dye Method for the Quantitative Assay of Esterase Using α-Naphthyl Acetate as Substrate. Histochem. J. 1980, 12 (2), 22134,  DOI: 10.1007/BF01024552
  541. 541
    Pearse, A. G. E. Histochemistry, Theoretical and Applied; Harcourt Brace/Churchill Livingstone, 1968.
  542. 542
    Lugg, G. A. Stabilized Diazonium Salts as Analytical Reagents for the Determination of Air-Borne Phenols and Amines. Anal. Chem. 1963, 35 (7), 899904,  DOI: 10.1021/ac60200a039
  543. 543
    Buchwald, H. The Colorimetric Determination of Phenol in Air and Urine with a Stabilized Diazonium Salt. Ann. Occup. Hyg. 1966, 9, 714,  DOI: 10.1093/annhyg/9.1.7
  544. 544
    Harvey, D.; Penketh, G. E. The Determination of Small Amounts of o-Phenylphenol. Analyst 1957, 82 (976), 498503,  DOI: 10.1039/an9578200498
  545. 545
    Canada, K. A.; Iwashita, S.; Shim, H.; Wood, T. K. Directed Evolution of Toluene ortho-Monooxygenase for Enhanced 1-Naphthol Synthesis and Chlorinated Ethene Degradation. J. Bacteriol. 2002, 184 (2), 3449,  DOI: 10.1128/JB.184.2.344-349.2002
  546. 546
    Joshi, N. S.; Whitaker, L. R.; Francis, M. B. A Three-Component Mannich-Type Reaction for Selective Tyrosine Bioconjugation. J. Am. Chem. Soc. 2004, 126 (49), 159423,  DOI: 10.1021/ja0439017
  547. 547
    Cardellicchio, C.; Capozzi, M. A. M.; Naso, F. The Betti Base: The Awakening of a Sleeping Beauty. Tetrahedron: Asymmetry 2010, 21 (5), 50717,  DOI: 10.1016/j.tetasy.2010.03.020
  548. 548
    Minakawa, M.; Guo, H.-M.; Tanaka, F. Imines That React with Phenols in Water over a Wide pH Range. J. Org. Chem. 2008, 73 (21), 866972,  DOI: 10.1021/jo8017389
  549. 549
    Kanaoka, Y. Organic Fluorescence Reagents in the Study of Enzymes and Proteins. Angew. Chem., Int. Ed. 1977, 16 (3), 13747,  DOI: 10.1002/anie.197701371
  550. 550
    Davis, W.; Ronai, Z.; Tew, K. D. Cellular Thiols and Reactive Oxygen Species in Drug-Induced Apoptosis. J. Pharmacol. Exp. Ther. 2001, 296, 16
  551. 551
    Ball, R. O.; Courtney-Martin, G.; Pencharz, P. B. The in Vivo Sparing of Methionine by Cysteine in Sulfur Amino Acid Requirements in Animal Models and Adult Humans. J. Nutr. 2006, 136 (6), 1682S93S,  DOI: 10.1093/jn/136.6.1682S
  552. 552
    Mishanina, T. V.; Libiad, M.; Banerjee, R. Biogenesis of Reactive Sulfur Species for Signaling by Hydrogen Sulfide Oxidation Pathways. Nat. Chem. Biol. 2015, 11 (7), 45764,  DOI: 10.1038/nchembio.1834
  553. 553
    Coleman, J.; Blake-Kalff, M.; Davies, E. Detoxification of Xenobiotics by Plants: Chemical Modification and Vacuolar Compartmentation. Trends Plant Sci. 1997, 2 (4), 14451,  DOI: 10.1016/S1360-1385(97)01019-4
  554. 554
    Tsay, J. T.; Oh, W.; Larson, T. J.; Jackowski, S.; Rock, C. O. Isolation and Characterization of the Beta-Ketoacyl-Acyl Carrier Protein Synthase III Gene (FabH) from Escherichia Coli K-12. J. Biol. Chem. 1992, 267 (10), 680714,  DOI: 10.1016/S0021-9258(19)50498-7
  555. 555
    Gao, T.; Yang, C.; Zheng, Y. G. Comparative Studies of Thiol-Sensitive Fluorogenic Probes for HAT Assays. Anal. Bioanal. Chem. 2013, 405 (4), 136171,  DOI: 10.1007/s00216-012-6522-5
  556. 556
    Tang, Y.; Jin, L.; Yin, B. A Dual-Selective Fluorescent Probe for GSH and Cys Detection: Emission and pH Dependent Selectivity. Anal. Chim. Acta 2017, 993, 8795,  DOI: 10.1016/j.aca.2017.09.028
  557. 557
    Zaric, B. L.; Obradovic, M.; Bajic, V.; Haidara, M. A.; Jovanovic, M.; Isenovic, E. R. Homocysteine and Hyperhomocysteinaemia. Curr. Med. Chem. 2019, 26 (16), 294861,  DOI: 10.2174/0929867325666180313105949
  558. 558
    van Meurs, J.; Dhonukshe-Rutten, R. A.; Pluijm, S. M.; Van Der Klift, M.; De Jonge, R.; Lindemans, J.; De Groot, L. C.; Hofman, A.; Witteman, J. C.; Van Leeuwen, J. P.; Breteler, M. M.; Lips, P.; Pols, H. A.; Uitterlinden, A. G. Homocysteine Levels and the Risk of Osteoporotic Fracture. N. Engl. J. Med. 2004, 350, 203341,  DOI: 10.1056/NEJMoa032546
  559. 559
    Lee, D.; Kim, G.; Yin, J.; Yoon, J. An Aryl-Thioether Substituted Nitrobenzothiadiazole Probe for the Selective Detection of Cysteine and Homocysteine. Chem. Commun. 2015, 51 (30), 651820,  DOI: 10.1039/C5CC01071C
  560. 560
    Giannoulis, K. M.; Giokas, D. L.; Tsogas, G. Z.; Vlessidis, A. G. Ligand-Free Gold Nanoparticles as Colorimetric Probes for the Non-Destructive Determination of Total Dithiocarbamate Pesticides after Solid Phase Extraction. Talanta 2014, 119, 27683,  DOI: 10.1016/j.talanta.2013.10.063
  561. 561
    Munday, R. Toxicity of Thiols and Disulphides: Involvement of Free-Radical Species. Free Radical Biol. Med. 1989, 7 (6), 65973,  DOI: 10.1016/0891-5849(89)90147-0
  562. 562
    Cazzola, M.; Calzetta, L.; Page, C.; Rogliani, P.; Matera, M. G. Thiol-Based Drugs in Pulmonary Medicine: Much More Than Mucolytics. Trends Pharmacol. Sci. 2019, 40 (7), 45263,  DOI: 10.1016/j.tips.2019.04.015
  563. 563
    Wang, H.; Wu, X.; Yang, S.; Tian, H.; Liu, Y.; Sun, B. A Rapid and Visible Colorimetric Fluorescent Probe for Benzenethiol Flavor Detection. Food Chem. 2019, 286, 3228,  DOI: 10.1016/j.foodchem.2019.02.033
  564. 564
    Liu, X.-L.; Duan, X.-Y.; Du, X.-J.; Song, Q.-H. Quinolinium-Based Fluorescent Probes for the Detection of Thiophenols in Environmental Samples and Living Cells. Chem.─Asian J. 2012, 7, 26962702,  DOI: 10.1002/asia.201200594
  565. 565
    He, L.; Yang, X.; Xu, K.; Kong, X.; Lin, W. A Multi-Signal Fluorescent Probe for Simultaneously Distinguishing and Sequentially Sensing Cysteine/Homocysteine, Glutathione, and Hydrogen Sulfide in Living Cells. Chem. Sci. 2017, 8 (9), 625765,  DOI: 10.1039/C7SC00423K
  566. 566
    Wang, N.; Chen, M.; Gao, J.; Ji, X.; He, J.; Zhang, J.; Zhao, W. A Series of BODIPY-Based Probes for the Detection of Cysteine and Homocysteine in Living Cells. Talanta 2019, 195, 2819,  DOI: 10.1016/j.talanta.2018.11.066
  567. 567
    Yao, Z.; Ge, W.; Guo, M.; Xiao, K.; Qiao, Y.; Cao, Z.; Wu, H.-C. Ultrasensitive Detection of Thiophenol Based on a Water-Soluble Pyrenyl Probe. Talanta 2018, 185, 14650,  DOI: 10.1016/j.talanta.2018.03.068
  568. 568
    Zhang, S.; Cai, F.; Hou, B.; Chen, H.; Gao, C.; Shen, X.-c.; Liang, H. Constructing a Far-Red to near-Infrared Fluorescent Probe for Highly Specific Detection of Cysteine and Its Bioimaging Applications in Living Cells and Zebrafish. New J. Chem. 2019, 43 (17), 6696701,  DOI: 10.1039/C9NJ00260J
  569. 569
    Fan, L.; Zhang, W.; Wang, X.; Dong, W.; Tong, Y.; Dong, C.; Shuang, S. A Two-Photon Ratiometric Fluorescent Probe for Highly Selective Sensing of Mitochondrial Cysteine in Live Cells. Analyst 2019, 144 (2), 43947,  DOI: 10.1039/C8AN01908H
  570. 570
    Ros-Lis, J. V.; García, B.; Jiménez, D.; Martínez-Máñez, R.; Sancenón, F.; Soto, J.; Gonzalvo, F.; Valldecabres, M. C. Squaraines as Fluoro-Chromogenic Probes for Thiol-Containing Compounds and Their Application to the Detection of Biorelevant Thiols. J. Am. Chem. Soc. 2004, 126 (13), 40645,  DOI: 10.1021/ja031987i
  571. 571
    Zhang, X.; Li, C.; Cheng, X.; Wang, X.; Zhang, B. A near-Infrared Croconium Dye-Based Colorimetric Chemodosimeter for Biological Thiols and Cyanide Anion. Sens. Actuators B Chem. 2008, 129 (1), 1527,  DOI: 10.1016/j.snb.2007.07.094
  572. 572
    Tanaka, F.; Mase, N.; Barbaa, C. F., III Determination of Cysteine Concentration by Fluorescence Increase: Reaction of Cysteine with a Fluorogenic Aldehyde. Chem. Commun. 2004, 2004, 1762,  DOI: 10.1039/b405642f
  573. 573
    Jung, H. S.; Ko, K. C.; Kim, G.-H.; Lee, A.-R.; Na, Y.-C.; Kang, C.; Lee, J. Y.; Kim, J. S. Coumarin-Based Thiol Chemosensor: Synthesis, Turn-On Mechanism, and Its Biological Application. Org. Lett. 2011, 13 (6), 1498501,  DOI: 10.1021/ol2001864
  574. 574
    Zhang, Y.; Shao, X.; Wang, Y.; Pan, F.; Kang, R.; Peng, F.; Huang, Z.; Zhang, W.; Zhao, W. Dual Emission Channels for Sensitive Discrimination of Cys/Hcy and GSH in Plasma and Cells. Chem. Commun. 2015, 51 (20), 42458,  DOI: 10.1039/C4CC08687B
  575. 575
    Niu, L.-Y.; Guan, Y.-S.; Chen, Y.-Z.; Wu, L.-Z.; Tung, C.-H.; Yang, Q.-Z. A Turn-On Fluorescent Sensor for the Discrimination of Cystein from Homocystein and Glutathione. Chem. Commun. 2013, 49 (13), 12946,  DOI: 10.1039/c2cc38429a
  576. 576
    Chen, W.; Luo, H.; Liu, X.; Foley, J. W.; Song, X. Broadly Applicable Strategy for the Fluorescence Based Detection and Differentiation of Glutathione and Cysteine/Homocysteine: Demonstration in Vitro and in Vivo. Anal. Chem. 2016, 88 (7), 363846,  DOI: 10.1021/acs.analchem.5b04333
  577. 577
    Mikaliunaite, L.; Green, D. B. Using a 3-Hydroxyflavone Derivative as a Fluorescent Probe for the Indirect Determination of Aminothiols Separated by Ion-Pair HPLC. Anal. Methods 2021, 13 (26), 291525,  DOI: 10.1039/D1AY00499A
  578. 578
    Toyooka, T.; Imai, K. New Fluorogenic Reagent Having Halogenobenzofurazan Structure for Thiols: 4-(Aminosulfonyl)-7-Fluoro-2,1,3-Benzoxadiazole. Anal. Chem. 1984, 56 (13), 24614,  DOI: 10.1021/ac00277a044
  579. 579
    Zhao, L.; Zhao, L.; Zhang, C.; Li, Y. A Naked-Eye Visible and Fluorescence ″Turn-On″ Probe for Acetylcholinesterase Assay and the Discrimination of Biothiols. J. Appl. Spectrosc. 2018, 85 (3), 43744,  DOI: 10.1007/s10812-018-0669-6
  580. 580
    Hassett, R. P.; Crockett, E. L. Endpoint Fluorometric Assays for Determining Activities of Carnitine Palmitoyltransferase and Citrate Synthase. Anal. Biochem. 2000, 287 (1), 1769,  DOI: 10.1006/abio.2000.4799
  581. 581
    Niu, L.-Y.; Zheng, H.-R.; Chen, Y.-Z.; Wu, L.-Z.; Tung, C.-H.; Yang, Q.-Z. Fluorescent Sensors for Selective Detection of Thiols: Expanding the Intramolecular Displacement Based Mechanism to New Chromophores. Analyst 2014, 139 (6), 138995,  DOI: 10.1039/c3an01849k
  582. 582
    Niu, L.-Y.; Guan, Y.-S.; Chen, Y.-Z.; Wu, L.-Z.; Tung, C.-H.; Yang, Q.-Z. BODIPY-Based Ratiometric Fluorescent Sensor for Highly Selective Detection of Glutathione over Cysteine and Homocysteine. J. Am. Chem. Soc. 2012, 134 (46), 1892831,  DOI: 10.1021/ja309079f
  583. 583
    Liu, X.-L.; Niu, L.-Y.; Chen, Y.-Z.; Zheng, M.-L.; Yang, Y.; Yang, Q.-Z. A Mitochondria-Targeting Fluorescent Probe for the Selective Detection of Glutathione in Living Cells. Org. Biomol. Chem. 2017, 15 (5), 10725,  DOI: 10.1039/C6OB02407F
  584. 584
    Hu, Q.; Yu, C.; Xia, X.; Zeng, F.; Wu, S. A Fluorescent Probe for Simultaneous Discrimination of GSH and Cys/Hcy in Human Serum Samples Via Distinctly-Separated Emissions with Independent Excitations. Biosens. Bioelectron. 2016, 81, 3418,  DOI: 10.1016/j.bios.2016.03.011
  585. 585
    Zhang, J.; Ji, X.; Ren, H.; Zhou, J.; Chen, Z.; Dong, X.; Zhao, W. meso-Heteroaryl BODIPY Dyes as Dual-Responsive Fluorescent Probes for Discrimination of Cys from Hcy and GSH. Sens. Actuators B Chem. 2018, 260, 8619,  DOI: 10.1016/j.snb.2018.01.016
  586. 586
    Liu, Y.; Yu, Y.; Zhao, Q.; Tang, C.; Zhang, H.; Qin, Y.; Feng, X.; Zhang, J. Fluorescent Probes Based on Nucleophilic Aromatic Substitution Reactions for Reactive Sulfur and Selenium Species: Recent Progress, Applications, and Design Strategies. Coord. Chem. Rev. 2021, 427, 213601,  DOI: 10.1016/j.ccr.2020.213601
  587. 587
    Kim, Y.; Mulay, S. V.; Choi, M.; Yu, S. B.; Jon, S.; Churchill, D. G. Exceptional Time Response, Stability and Selectivity in Doubly-Activated Phenyl Selenium-Based Glutathione-Selective Platform. Chem. Sci. 2015, 6, 5435,  DOI: 10.1039/C5SC02090E
  588. 588
    Mulay, S. V.; Kim, Y.; Choi, M.; Lee, D. Y.; Choi, J.; Lee, Y.; Jon, S.; Churchill, D. G. Enhanced Doubly Activated Dual Emission Fluorescent Probes for Selective Imaging of Glutathione or Cysteine in Living Systems. Anal. Chem. 2018, 90 (4), 264854,  DOI: 10.1021/acs.analchem.7b04375
  589. 589
    Maeda, H.; Matsuno, H.; Ushida, M.; Katayama, K.; Saeki, K.; Itoh, N. 2,4-Dinitrobenzenesulfonyl Fluoresceins as Fluorescent Alternatives to Ellman’s Reagent in Thiol-Quantification Enzyme Assays. Angew. Chem., Int. Ed. 2005, 44 (19), 29225,  DOI: 10.1002/anie.200500114
  590. 590
    Jiang, W.; Fu, Q.; Fan, H.; Ho, J.; Wang, W. A Highly Selective Fluorescent Probe for Thiophenols. Angew. Chem., Int. Ed. 2007, 46 (44), 84458,  DOI: 10.1002/anie.200702271
  591. 591
    Wang, S.; Huang, Y.; Guan, X. Fluorescent Probes for Live Cell Thiol Detection. Molecules 2021, 26 (12), 3575,  DOI: 10.3390/molecules26123575
  592. 592
    Dai, Z.; Tian, L.; Ye, Z.; Song, B.; Zhang, R.; Yuan, J. A Lanthanide Complex-Based Ratiometric Luminescence Probe for Time-Gated Luminescence Detection of Intracellular Thiols. Anal. Chem. 2013, 85 (23), 1165864,  DOI: 10.1021/ac403370g
  593. 593
    Rong, X.; Xu, Z.-Y.; Yan, J.-W.; Meng, Z.-Z.; Zhu, B.; Zhang, L. Nile-Red-Based Fluorescence Probe for Selective Detection of Biothiols, Computational Study, and Application in Cell Imaging. Molecules 2020, 25 (20), 4718,  DOI: 10.3390/molecules25204718
  594. 594
    Maeda, H.; Katayama, K.; Matsuno, H.; Uno, T. 3′-(2,4-Dinitrobenzenesulfonyl)-2′,7′-Dimethylfluorescein as a Fluorescent Probe for Selenols. Angew. Chem., Int. Ed. 2006, 45 (11), 18103,  DOI: 10.1002/anie.200504299
  595. 595
    Peng, L.; Xiao, L.; Ding, Y.; Xiang, Y.; Tong, A. A Simple Design of Fluorescent Probes for Indirect Detection of β-Lactamase Based on AIE and ESIPT Processes. J. Mater. Chem. B 2018, 6 (23), 39226,  DOI: 10.1039/C8TB00414E
  596. 596
    Yuan, L.; Lin, W.; Zhao, S.; Gao, W.; Chen, B.; He, L.; Zhu, S. A Unique Approach to Development of Near-Infrared Fluorescent Sensors for in Vivo Imaging. J. Am. Chem. Soc. 2012, 134 (32), 1351023,  DOI: 10.1021/ja305802v
  597. 597
    Lu, Z.; Sun, X.; Wang, M.; Wang, H.; Fan, C.; Lin, W. Rational Design of a Far-Red Fluorescent Probe for Endogenous Biothiol Imbalance Induced by Hydrogen Peroxide in Living Cells and Mice. Bioorg. Chem. 2020, 103, 104173,  DOI: 10.1016/j.bioorg.2020.104173
  598. 598
    Kand, D.; Kalle, A. M.; Varma, S. J.; Talukdar, P. A Chromenoquinoline-Based Fluorescent Off-On Thiol Probe for Bioimaging. Chem. Commun. 2012, 48 (21), 27224,  DOI: 10.1039/c2cc16593g
  599. 599
    Liu, T.; Huo, F.; Li, J.; Chao, J.; Zhang, Y.; Yin, C. A Fast Response and High Sensitivity Thiol Fluorescent Probe in Living Cells. Sens. Actuators B Chem. 2016, 232, 61924,  DOI: 10.1016/j.snb.2016.04.014
  600. 600
    Lin, X.; Hu, Y.; Yang, D.; Chen, B. Cyanine-Coumarin Composite NIR Dye Based Instantaneous-Response Probe for Biothiols Detection and Oxidative Stress Assessment of Mitochondria. Dyes Pigm. 2020, 174, 107956,  DOI: 10.1016/j.dyepig.2019.107956
  601. 601
    Sun, Q.; Yang, S.-H.; Wu, L.; Yang, W.-C.; Yang, G.-F. A Highly Sensitive and Selective Fluorescent Probe for Thiophenol Designed Via a Twist-Blockage Strategy. Anal. Chem. 2016, 88 (4), 226672,  DOI: 10.1021/acs.analchem.5b04029
  602. 602
    Jiang, W.; Cao, Y.; Liu, Y.; Wang, W. Rational Design of a Highly Selective and Sensitive Fluorescent PET Probe for Discrimination of Thiophenols and Aliphatic Thiols. Chem. Commun. 2010, 46 (11), 19446,  DOI: 10.1039/B926070F
  603. 603
    Li, Y.; Su, W.; Zhou, Z.; Huang, Z.; Wu, C.; Yin, P.; Li, H.; Zhang, Y. A Dual-Response near-Infrared Fluorescent Probe for Rapid Detecting Thiophenol and Its Application in Water Samples and Bio-Imaging. Talanta 2019, 199, 35560,  DOI: 10.1016/j.talanta.2019.02.022
  604. 604
    Xiong, L.; Ma, J.; Huang, Y.; Wang, Z.; Lu, Z. Highly Sensitive Squaraine-Based Water-Soluble Far-Red/near-Infrared Chromofluorogenic Thiophenol Probe. ACS Sens. 2017, 2 (4), 599605,  DOI: 10.1021/acssensors.7b00151
  605. 605
    Song, B.; Wang, G.; Tan, M.; Yuan, J. A Europium(III) Complex as an Efficient Singlet Oxygen Luminescence Probe. J. Am. Chem. Soc. 2006, 128 (41), 1344250,  DOI: 10.1021/ja062990f
  606. 606
    Bjartell, A.; Laine, S.; Pettersson, K.; Nilsson, E.; Lövgren, T.; Lilja, H. Time-Resolved Fluorescence in Immunocytochemical Detection of Prostate-Specific Antigen in Prostatic Tissue Sections. Histochem. J. 1999, 31 (1), 4552,  DOI: 10.1023/A:1003504115690
  607. 607
    Shang, H.; Chen, H.; Tang, Y.; Ma, Y.; Lin, W. Development of a Two-Photon Fluorescent Turn-On Probe with Far-Red Emission for Thiophenols and Its Bioimaging Application in Living Tissues. Biosens. Bioelectron. 2017, 95, 816,  DOI: 10.1016/j.bios.2017.04.017
  608. 608
    Fellner, M.; Doughty, L. M.; Jameson, G. N. L.; Wilbanks, S. M. A Chromogenic Assay of Substrate Depletion by Thiol Dioxygenases. Anal. Biochem. 2014, 459, 5660,  DOI: 10.1016/j.ab.2014.05.008
  609. 609
    Zhu, J.; Dhimitruka, I.; Pei, D. 5-(2-Aminoethyl)Dithio-2-Nitrobenzoate as a More Base-Stable Alternative to Ellman’s Reagent. Org. Lett. 2004, 6 (21), 380912,  DOI: 10.1021/ol048404+
  610. 610
    Li, Y.; Lopez, P.; Durand, P.; Ouazzani, J.; Badet, B.; Badet-Denisot, M.-A. An Enzyme-Coupled Assay for Amidotransferase Activity of Glucosamine-6-Phosphate Synthase. Anal. Biochem. 2007, 370 (2), 1426,  DOI: 10.1016/j.ab.2007.07.031
  611. 611
    Chen, G.; Feng, H.; Jiang, X.; Xu, J.; Pan, S.; Qian, Z. Redox-Controlled Fluorescent Nanoswitch Based on Reversible Disulfide and Its Application in Butyrylcholinesterase Activity Assay. Anal. Chem. 2018, 90 (3), 164351,  DOI: 10.1021/acs.analchem.7b02976
  612. 612
    Gong, M.-M.; Dai, C.-Y.; Severance, S.; Hwang, C.-C.; Fang, B.-K.; Lin, H.-B.; Huang, C.-H.; Ong, C.-W.; Wang, J.-J.; Lee, P.-L. A Bioorthogonally Synthesized and Disulfide-Containing Fluorescence Turn-On Chemical Probe for Measurements of Butyrylcholinesterase Activity and Inhibition in the Presence of Physiological Glutathione. Catalysts 2020, 10 (10), 1169,  DOI: 10.3390/catal10101169
  613. 613
    Peng, H.; Chen, W.; Cheng, Y.; Hakuna, L.; Strongin, R.; Wang, B. Thiol Reactive Probes and Chemosensors. Sensors 2012, 12 (11), 1590746,  DOI: 10.3390/s121115907
  614. 614
    Worek, F.; Mast, U.; Kiderlen, D.; Diepold, C.; Eyer, P. Improved Determination of Acetylcholinesterase Activity in Human Whole Blood. Clin. Chim. Acta 1999, 288 (1), 7390,  DOI: 10.1016/S0009-8981(99)00144-8
  615. 615
    Eyer, P.; Worek, F.; Kiderlen, D.; Sinko, G.; Stuglin, A.; Simeon-Rudolf, V.; Reiner, E. Molar Absorption Coefficients for the Reduced Ellman Reagent: Reassessment. Anal. Biochem. 2003, 312 (2), 2247,  DOI: 10.1016/S0003-2697(02)00506-7
  616. 616
    Pullela, P. K.; Chiku, T.; Carvan, M. J.; Sem, D. S. Fluorescence-Based Detection of Thiols in Vitro and in Vivo Using Dithiol Probes. Anal. Biochem. 2006, 352 (2), 26573,  DOI: 10.1016/j.ab.2006.01.047
  617. 617
    Cao, X.; Lin, W.; Yu, Q. A Ratiometric Fluorescent Probe for Thiols Based on a Tetrakis(4-Hydroxyphenyl)Porphyrin-Coumarin Scaffold. J. Org. Chem. 2011, 76 (18), 742330,  DOI: 10.1021/jo201199k
  618. 618
    Wang, L.; Wang, J.; Xia, S.; Wang, X.; Yu, Y.; Zhou, H.; Liu, H. A FRET-Based near-Infrared Ratiometric Fluorescent Probe for Detection of Mitochondria Biothiol. Talanta 2020, 219, 121296,  DOI: 10.1016/j.talanta.2020.121296
  619. 619
    Lim, S. Y.; Shen, W.; Gao, Z. Carbon Quantum Dots and Their Applications. Chem. Soc. Rev. 2015, 44 (1), 36281,  DOI: 10.1039/C4CS00269E
  620. 620
    Steinmann, D.; Nauser, T.; Koppenol, W. H. Selenium and Sulfur in Exchange Reactions: A Comparative Study. J. Org. Chem. 2010, 75 (19), 66969,  DOI: 10.1021/jo1011569
  621. 621
    Mafireyi, T. J.; Laws, M.; Bassett, J. W.; Cassidy, P. B.; Escobedo, J. O.; Strongin, R. M. A Diselenide Turn-On Fluorescent Probe for the Detection of Thioredoxin Reductase. Angew. Chem., Int. Ed. 2020, 59 (35), 1514751,  DOI: 10.1002/anie.202004094
  622. 622
    Nogueira, C. W.; Barbosa, N. V.; Rocha, J. B. Toxicology and Pharmacology of Synthetic Organoselenium Compounds: An Update. Arch. Toxicol. 2021, 95 (4), 1179226,  DOI: 10.1007/s00204-021-03003-5
  623. 623
    Lou, Z.; Li, P.; Sun, X.; Yang, S.; Wang, B.; Han, K. A Fluorescent Probe for Rapid Detection of Thiols and Imaging of Thiols Reducing Repair and H2O2 Oxidative Stress Cycles in Living Cells. Chem. Commun. 2013, 49 (4), 3913,  DOI: 10.1039/C2CC36839K
  624. 624
    Han, X.; Song, X.; Yu, F.; Chen, L. A Ratiometric Fluorescent Probe for Imaging and Quantifying Anti-Apoptotic Effects of GSH under Temperature Stress. Chem. Sci. 2017, 8 (10), 69917002,  DOI: 10.1039/C7SC02888A
  625. 625
    Tang, B.; Xing, Y.; Li, P.; Zhang, N.; Yu, F.; Yang, G. A Rhodamine-Based Fluorescent Probe Containing a Se-N Bond for Detecting Thiols and Its Application in Living Cells. J. Am. Chem. Soc. 2007, 129 (38), 116667,  DOI: 10.1021/ja072572q
  626. 626
    Tang, B.; Yin, L.; Wang, X.; Chen, Z.; Tong, L.; Xu, K. A Fast-Response, Highly Sensitive and Specific Organoselenium Fluorescent Probe for Thiols and Its Application in Bioimaging. Chem. Commun. 2009, (35), 52935,  DOI: 10.1039/b909542j
  627. 627
    Wang, R.; Chen, L.; Liu, P.; Zhang, Q.; Wang, Y. Sensitive near-Infrared Fluorescent Probes for Thiols Based on Se-N Bond Cleavage: Imaging in Living Cells and Tissues. Eur. J. Chem. 2012, 18, 11343,  DOI: 10.1002/chem.201200671
  628. 628
    Tian, Y.; Zhu, B.; Yang, W.; Jing, J.; Zhang, X. A Fluorescent Probe for Differentiating Cys, Hcy and GSH Via a Stepwise Interaction. Sens. Actuators B Chem. 2018, 262, 3459,  DOI: 10.1016/j.snb.2018.01.181
  629. 629
    Rusin, O.; St Luce, N. N.; Agbaria, R. A.; Escobedo, J. O.; Jiang, S.; Warner, I. M.; Dawan, F. B.; Lian, K.; Strongin, R. M. Visual Detection of Cysteine and Homocysteine. J. Am. Chem. Soc. 2004, 126 (2), 438,  DOI: 10.1021/ja036297t
  630. 630
    Lim, S.; Escobedo, J. O.; Lowry, M.; Xu, X.; Strongin, R. Selective Fluorescence Detection of Cysteine and N-Terminal Cysteine Peptide Residues. Chem. Commun. 2010, 46 (31), 57079,  DOI: 10.1039/c0cc01398f
  631. 631
    Lee, K.-S.; Kim, T.-K.; Lee, J. H.; Kim, H.-J.; Hong, J.-I. Fluorescence Turn-On Probe for Homocysteine and Cysteine in Water. Chem. Commun. 2008, (46), 61735,  DOI: 10.1039/b814581d
  632. 632
    Eun Jun, M.; Roy, B.; Han Ahn, K. “Turn-On” Fluorescent Sensing with “Reactive” Probes. Chem. Commun. 2011, 47 (27), 7583601,  DOI: 10.1039/c1cc00014d
  633. 633
    Hoyle, C. E.; Lowe, A. B.; Bowman, C. N. Thiol-Click Chemistry: A Multifaceted Toolbox for Small Molecule and Polymer Synthesis. Chem. Soc. Rev. 2010, 39 (4), 135587,  DOI: 10.1039/b901979k
  634. 634
    LoPachin, R. M.; Gavin, T. Molecular Mechanisms of Aldehyde Toxicity: A Chemical Perspective. Chem. Res. Toxicol. 2014, 27 (7), 108191,  DOI: 10.1021/tx5001046
  635. 635
    Chan, J. W.; Hoyle, C. E.; Lowe, A. B.; Bowman, M. Nucleophile-Initiated Thiol-Michael Reactions: Effect of Organocatalyst, Thiol, and Ene. Macromolecules 2010, 43 (15), 63818,  DOI: 10.1021/ma101069c
  636. 636
    Langmuir, M. E.; Yang, J.-R.; Moussa, A. M.; Laura, R.; LeCompte, K. A. New Naphthopyranone Based Fluorescent Thiol Probes. Tetrahedron Lett. 1995, 36 (23), 398992,  DOI: 10.1016/0040-4039(95)00695-9
  637. 637
    Niu, W.; Wu, P.; Chen, F.; Wang, J.; Shang, X.; Xu, C. Discovery of Selective Cystathionine β-Synthase Inhibitors by High-Throughput Screening with a Fluorescent Thiol Probe. MedChemComm 2017, 8 (1), 198201,  DOI: 10.1039/C6MD00493H
  638. 638
    Aharoni, A.; Amitai, G.; Bernath, K.; Magdassi, S.; Tawfik, D. S. High-Throughput Screening of Enzyme Libraries: Thiolactonases Evolved by Fluorescence-Activated Sorting of Single Cells in Emulsion Compartments. Chem. Biol. 2005, 12 (12), 12819,  DOI: 10.1016/j.chembiol.2005.09.012
  639. 639
    Barb, A. W.; Marcella, A. M. A Rapid Fluorometric Assay for the S-Malonyltransacylase FabD and Other Sulfhydryl Utilizing Enzymes. J. Biol. Methods 2016, 3 (4), e53  DOI: 10.14440/jbm.2016.144
  640. 640
    Yi, L.; Li, H.; Sun, L.; Liu, L.; Zhang, C.; Xi, Z. A Highly Sensitive Fluorescence Probe for Fast Thiol-Quantification Assay of Glutathione Reductase. Angew. Chem., Int. Ed. 2009, 48 (22), 40347,  DOI: 10.1002/anie.200805693
  641. 641
    Zhou, X.; Jin, X.; Sun, G.; Li, D.; Wu, X. A Cysteine Probe with High Selectivity and Sensitivity Promoted by Response-Assisted Electrostatic Attraction. Chem. Commun. 2012, 48 (70), 87935,  DOI: 10.1039/c2cc33971d
  642. 642
    Li, S.-J.; Fu, Y.-J.; Li, C.-Y.; Li, Y.-F.; Yi, L.-H.; Ou-Yang, J. A near-Infrared Fluorescent Probe Based on BODIPY Derivative with High Quantum Yield for Selective Detection of Exogenous and Endogenous Cysteine in Biological Samples. Anal. Chim. Acta 2017, 994, 7381,  DOI: 10.1016/j.aca.2017.09.031
  643. 643
    Li, R.; Lei, C.; Zhao, X.-E.; Gao, Y.; Gao, H.; Zhu, S.; Wang, H. A Label-Free Fluorimetric Detection of Biothiols Based on the Oxidase-Like Activity of Ag+ Ions. Spectrochim. Acta - A: Mol. Biomol. Spectrosc. 2018, 188, 205,  DOI: 10.1016/j.saa.2017.06.056
  644. 644
    Mandani, S.; Sharma, B.; Dey, D.; Sarma, T. K. Carbon Nanodots as Ligand Exchange Probes in Au@C-Dot Nanobeacons for Fluorescent Turn-On Detection of Biothiols. Nanoscale 2015, 7 (5), 18028,  DOI: 10.1039/C4NR05424E
  645. 645
    Fu, Y.; Li, H.; Hu, W.; Zhu, D. Fluorescence Probes for Thiol-Containing Amino Acids and Peptides in Aqueous Solution. Chem. Commun. 2005, (25), 318991,  DOI: 10.1039/b503772g
  646. 646
    Nie, L.; Ma, H.; Sun, M.; Li, X.; Su, M.; Liang, S. Direct Chemiluminescence Determination of Cysteine in Human Serum Using Quinine-Ce(IV) System. Talanta 2003, 59 (5), 95964,  DOI: 10.1016/S0039-9140(02)00649-5
  647. 647
    Wang, S.; Ma, H.; Li, J.; Chen, X.; Bao, Z.; Sun, S. Direct Determination of Reduced Glutathione in Biological Fluids by Ce(Iv)-Quinine Chemiluminescence. Talanta 2006, 70 (3), 51821,  DOI: 10.1016/j.talanta.2005.12.052
  648. 648
    Zou, T.; Lum, C. T.; Chui, S. S.-Y.; Che, C.-M. Gold(III) Complexes Containing N-Heterocyclic Carbene Ligands: Thiol “Switch-On” Fluorescent Probes and Anti-Cancer Agents. Angew. Chem., Int. Ed. 2013, 52 (10), 29303,  DOI: 10.1002/anie.201209787
  649. 649
    Han, G.-C.; Peng, Y.; Hao, Y.-Q.; Liu, Y.-N.; Zhou, F. Spectrofluorimetric Determination of Total Free Thiols Based on Formation of Complexes of Ce(III) with Disulfide Bonds. Anal. Chim. Acta 2010, 659 (1), 23842,  DOI: 10.1016/j.aca.2009.11.057
  650. 650
    Wang, H.; Zhou, G.; Chen, X. An Iminofluorescein-Cu2+ Ensemble Probe for Selective Detection of Thiols. Sens. Actuators B Chem. 2013, 176, 698703,  DOI: 10.1016/j.snb.2012.10.006
  651. 651
    Jung, H. S.; Han, J. H.; Habata, Y.; Kang, C.; Kim, J. S. An Iminocoumarin-Cu(II) Ensemble-Based Chemodosimeter toward Thiols. Chem. Commun. 2011, 47 (18), 51424,  DOI: 10.1039/c1cc10672d
  652. 652
    Zhao, D.; Chen, C.; Sun, J.; Yang, X. Carbon Dots-Assisted Colorimetric and Fluorometric Dual-Mode Protocol for Acetylcholinesterase Activity and Inhibitors Screening Based on the Inner Filter Effect of Silver Nanoparticles. Analyst 2016, 141 (11), 32808,  DOI: 10.1039/C6AN00514D
  653. 653
    Yang, J.; Song, N.; Lv, X.; Jia, Q. UV-Light-Induced Synthesis of PEI-CuNCs Based on Cu2+-Quenched Fluorescence Turn-On Assay for Sensitive Detection of Biothiols, Acetylcholinesterase Activity and Inhibitor. Sens. Actuators B Chem. 2018, 259, 22632,  DOI: 10.1016/j.snb.2017.12.045
  654. 654
    Chen, Y.; Liu, W.; Zhang, B.; Suo, Z.; Xing, F.; Feng, L. Sensitive and Reversible Perylene Derivative-Based Fluorescent Probe for Acetylcholinesterase Activity Monitoring and Its Inhibitor. Anal. Biochem. 2020, 607, 113835,  DOI: 10.1016/j.ab.2020.113835
  655. 655
    Xu, J.; Yu, H.; Hu, Y.; Chen, M.; Shao, S. A Gold Nanoparticle-Based Fluorescence Sensor for High Sensitive and Selective Detection of Thiols in Living Cells. Biosens. Bioelectron. 2016, 75, 17,  DOI: 10.1016/j.bios.2015.08.007
  656. 656
    Mu, H.; Miki, K.; Kubo, T.; Otsuka, K.; Ohe, K. Substituted meso-Vinyl-BODIPY as Thiol-Selective Fluorogenic Probes for Sensing Unfolded Proteins in the Endoplasmic Reticulum. Chem. Commun. 2021, 57 (14), 181821,  DOI: 10.1039/D0CC08160D
  657. 657
    Huang, S.-T.; Ting, K.-N.; Wang, K.-L. Development of a Long-Wavelength Fluorescent Probe Based on Quinone-Methide-Type Reaction to Detect Physiologically Significant Thiols. Anal. Chim. Acta 2008, 620 (1), 1206,  DOI: 10.1016/j.aca.2008.05.006
  658. 658
    Kosower, N. S.; Kosower, E. M.; Newton, G. L.; Ranney, H. M. Bimane Fluorescent Labels: Labeling of Normal Human Red Cells under Physiological Conditions. Proc. Natl. Acad. Sci. U.S.A 1979, 76 (7), 33826,  DOI: 10.1073/pnas.76.7.3382
  659. 659
    Qi, S.; Liu, W.; Zhang, P.; Wu, J.; Zhang, H.; Ren, H.; Ge, J.; Wang, P. A Colorimetric and Ratiometric Fluorescent Probe for Highly Selective Detection of Glutathione in the Mitochondria of Living Cells. Sens. Actuators B Chem. 2018, 270, 45965,  DOI: 10.1016/j.snb.2018.05.017
  660. 660
    Li, X.; Huo, F.; Yue, Y.; Zhang, Y.; Yin, C. A Coumarin-Based “Off-On” Sensor for Fluorescence Selectivily Discriminating GSH from Cys/Hcy and Its Bioimaging in Living Cells. Sens. Actuators B Chem. 2017, 253, 429,  DOI: 10.1016/j.snb.2017.06.120
  661. 661
    Long, L.; Lin, W.; Chen, B.; Gao, W.; Yuan, L. Construction of a FRET-Based Ratiometric Fluorescent Thiol Probe. Chem. Commun. 2011, 47 (3), 8935,  DOI: 10.1039/C0CC03806G
  662. 662
    Chen, H.; Tang, Y.; Ren, M.; Lin, W. Single near-Infrared Fluorescent Probe with High- and Low-Sensitivity Sites for Sensing Different Concentration Ranges of Biological Thiols with Distinct Modes of Fluorescence Signals. Chem. Sci. 2016, 7 (3), 1896903,  DOI: 10.1039/C5SC03591K
  663. 663
    Hall, H. K., Jr. Correlation of the Nucleophilic Reactivity of Aliphatic Amines. J. Org. Chem. 1964, 29 (12), 35393544,  DOI: 10.1021/jo01035a024
  664. 664
    Inoue, N. Selective Detection of Carcinogenic Aromatic Diamines in Aqueous Solutions Using 4-(N-Chloroformylmethyl-N-Methylamino)-7-Nitro-2,1,3-Benzoxadiazole (NBD-COCl) by HPLC. Anal. Sci. 2017, 33 (12), 13751380
  665. 665
    You, W. W.; Haugland, R. P.; Ryan, D. K.; Haugland, R. P. 3-(4-Carboxybenzoyl)quinoline-2-carboxaldehyde, a Reagent with Broad Dynamic Range for the Assay of Proteins and Lipoproteins in Solution. Anal. Biochem. 1997, 244 (2), 277,  DOI: 10.1006/abio.1996.9920
  666. 666
    Wu, J.; Chen, Z.; Dovichi, N. J Reaction Rate, Activation Energy, and Detection Limit for the Reaction of 5-Furoylquinoline-3-Carboxaldehyde with Neurotransmitters in Artificial Cerebrospinal Fluid. J. Chromatogr. B Biomed. Appl. 2000, 741 (1), 85,  DOI: 10.1016/S0378-4347(99)00540-X
  667. 667
    Pinto, D.; Arriaga, E. A.; Schoenherr, R. M.; Chou, S. S.-H.; Dovichi, N. J. Kinetics and Apparent Activation Energy of the Reaction of the Fluorogenic Reagent 5-Furoylquinoline-3-Carboxaldehyde with Ovalbumin. J. Chromatogr. B 2003, 793 (1), 107,  DOI: 10.1016/S1570-0232(03)00368-4
  668. 668
    Arrell, M. S.; Kalman, F. Estimation of Protein Concentration at High Sensitivity Using SDS-Capillary Gel Electrophoresis-Laser Induced Fluorescence Detection with 3-(2-Furoyl)quinoline-2-Carboxaldehyde Protein Labeling. Electrophoresis 2016, 37 (22), 2913,  DOI: 10.1002/elps.201600246
  669. 669
    Wu, D.; Liu, Y.; Zheng, F.; Rong, S.-Q.; Yang, T.; Zhao, Y.-K.; Yang, R.-W.; Zou, P.; Wang, G.-T. Detection of Organic Amines Using a Ratiometric Chromophoric Fluorescent Probe with a Significant Emission Shift. J. Chem. Res. 2020, 44, 475481,  DOI: 10.1177/1747519820902944
  670. 670
    Zhang, J.; Zhou, J.; Xu, G.; Ni, Y. Stereodivergent Evolution of KpADH for the Asymmetric Reduction of Diaryl Ketones with Para-Substituents. Mol. Catal. 2022, 524, 112315,  DOI: 10.1016/j.mcat.2022.112315
  671. 671
    Ribeaucourt, D.; Bissaro, B.; Lambert, F.; Lafond, M.; Berrin, J.-G. Biocatalytic Oxidation of Fatty Alcohols into Aldehydes for the Flavors and Fragrances Industry. Biotechnol. Adv. 2022, 56, 107787,  DOI: 10.1016/j.biotechadv.2021.107787
  672. 672
    O'Brien, P. J.; Siraki, A. G.; Shangari, N. Aldehyde Sources, Metabolism, Molecular Toxicity Mechanisms, and Possible Effects on Human Health. Crit. Rev. Toxicol. 2005, 35 (7), 609662,  DOI: 10.1080/10408440591002183
  673. 673
    Corley, R. A.; Kabilan, S.; Kuprat, A. P.; Carson, J. P.; Jacob, R. E.; Minard, K. R.; Teeguarden, J. G.; Timchalk, C.; Pipavath, S.; Glenny, R.; Einstein, D. R. Comparative Risks of Aldehyde Constituents in Cigarette Smoke Using Transient Computational Fluid Dynamics/Physiologically Based Pharmacokinetic Models of the Rat and Human Respiratory Tracts. Toxicol. Sci. 2015, 146 (1), 6588,  DOI: 10.1093/toxsci/kfv071
  674. 674
    Ahmed Laskar, A.; Younus, H. Aldehyde Toxicity and Metabolism: The Role of Aldehyde Dehydrogenases in Detoxification, Drug Resistance and Carcinogenesis. Drug Metab. Rev. 2019, 51 (1), 4264,  DOI: 10.1080/03602532.2018.1555587
  675. 675
    Fritz, K. S.; Petersen, D. R. An Overview of the Chemistry and Biology of Reactive Aldehydes. Free Radical Biol. Med. 2013, 59, 8591,  DOI: 10.1016/j.freeradbiomed.2012.06.025
  676. 676
    Urbansky, E. T. Carbinolamines and Geminal Diols in Aqueous Environmental Organic Chemistry. J. Chem. Ed. 2000, 77 (12), 1644,  DOI: 10.1021/ed077p1644
  677. 677
    Mukherjee, K.; Chio, T. I.; Gu, H.; Banerjee, A.; Sorrentino, A. M.; Sackett, D. L.; Bane, S. L. Benzocoumarin Hydrazine: A Large Stokes Shift Fluorogenic Sensor for Detecting Carbonyls in Isolated Biomolecules and in Live Cells. ACS Sens. 2017, 2 (1), 128134,  DOI: 10.1021/acssensors.6b00616
  678. 678
    Molina-Espeja, P.; Canellas, M.; Plou, F. J.; Hofrichter, M.; Lucas, F.; Guallar, V.; Alcalde, M. Synthesis of 1-Naphthol by a Natural Peroxygenase Engineered by Directed Evolution. ChemBioChem 2016, 17 (4), 341349,  DOI: 10.1002/cbic.201500493
  679. 679
    Guo, H.-M.; Minakawa, M.; Tanaka, F. Fluorogenic Imines for Fluorescent Detection of Mannich-Type Reactions of Phenols in Water. J. Org. Chem. 2008, 73 (10), 39643966,  DOI: 10.1021/jo8003293

Cited By

Click to copy section linkSection link copied!
Citation Statements
Explore this article's citation statements on scite.ai

This article is cited by 24 publications.

  1. Jiali Wang, Manwen Wang, Anil Ravi, Seyed-Omar Zaraei, Bilal O. Alkubaisi, Yingjie Wen, Lingda Zeng, Mohammed I. El-Gamal, Hanhong Xu. Inhibition of Glucosinolate Sulfatases to Combat Plutella xylostella: Development of Novel Phenylcarboxamide Derivatives for Plant-Integrated Pest Management. Journal of Agricultural and Food Chemistry 2025, 73 (19) , 12061-12071. https://doi.org/10.1021/acs.jafc.5c01873
  2. Matthew Penner, Oskar James Klein, Maximilian Gantz, Friederike E. H. Nintzel, Anne-Cathrin Prowald, Sally Boss, Paul Barker, Paul Dupree, Florian Hollfelder. Fluorogenic, Subsingle-Turnover Monitoring of Enzymatic Reactions Involving NAD(P)H Provides a Generalized Platform for Directed Ultrahigh-Throughput Evolution of Biocatalysts in Microdroplets. Journal of the American Chemical Society 2025, 147 (13) , 10903-10915. https://doi.org/10.1021/jacs.4c11804
  3. Tomohiro Umeno, Lisa Muroi, Yuto Kayama, Kazuteru Usui, Koichi Hamada, Akihiro Mizutani, Satoru Karasawa. Naphthyridine-Based Electron Push–Pull-Type Amine-Reactive Fluorescent Probe for Sensing Amines and Proteins in Aqueous Media. Bioconjugate Chemistry 2023, 34 (8) , 1439-1446. https://doi.org/10.1021/acs.bioconjchem.3c00220
  4. Ming Liu, Jingpi Gao, Yang Zhang, Xin Zhou, Yu Wang, Li Wu, Zhiyuan Tian, Jian-Hong Tang. Recent advances in bioresponsive macrocyclic gadolinium( iii ) complexes for MR imaging and therapy. Dalton Transactions 2025, 54 (17) , 6741-6777. https://doi.org/10.1039/D5DT00191A
  5. Rahul Vikram Singh. Immobilization of Nitrilase: Driving Efficiency and Sustainability in Pharmaceutical Synthesis. ChemBioEng Reviews 2025, 12 (2) https://doi.org/10.1002/cben.70003
  6. Muhammad Israr, Santosh Kumar Padhi, Yi Zhou, Shuke Wu. Recent Advances in Protein Engineering and Synthetic Applications of Amino Acid Transaminases. ChemCatChem 2025, 17 (8) https://doi.org/10.1002/cctc.202401952
  7. Ali Foroutan Kalourazi, Seyed Amirabbas Nazemi, Ajmal Roshan Unniram Parambil, Manuel Ferrer, S. Shirin Shahangian, Patrick Shahgaldian. Exploring the Potential of Various Cyclodextrin‐Based Derivatives in Enzyme Supramolecular Engineering. ChemBioChem 2025, 26 (4) https://doi.org/10.1002/cbic.202400840
  8. Xinyang Liu, Lu Chen, Lin Li, Yiqi Yan, Han Zhang. Research Progress in In Vitro Screening Techniques for Natural Antithrombotic Medicines. Pharmaceuticals 2025, 18 (2) , 137. https://doi.org/10.3390/ph18020137
  9. Arpita Devi, Abhilekha Borah, Sayanika Saikia, Lakshi Saikia, Magdi E.A. Zaki, Kusum K. Bania. Impact of halloysite nanotube on the catalytic activity of transition metal oxides in benzyl alcohol oxidation. Inorganic Chemistry Communications 2025, 123 , 113962. https://doi.org/10.1016/j.inoche.2025.113962
  10. Yuanbo Wang, Jie Zhou, Canhao Qiu, Lei Wang, Shiyue Zheng, Junqian Peng, Sheng Lu, Fang Wang, Ziyi Yu, Weiliang Dong, Min Jiang, Xiaoqiang Chen. Fluorescence‐activated screening of polyester‐depolymerizing enzymes based on pseudo‐ PET polythioester plastics. AIChE Journal 2024, 70 (11) https://doi.org/10.1002/aic.18539
  11. Isabella Tavernaro, Anna Matiushkina, Kai Simon Rother, Celina Mating, Ute Resch-Genger. Exploring the potential of simple automation concepts for quantifying functional groups on nanomaterials with optical assays. Nano Research 2024, 17 (11) , 10119-10126. https://doi.org/10.1007/s12274-024-6970-1
  12. Jihang Zhang, Jiale Chen, Yu Sha, Jiawei Deng, Jinglan Wu, Pengpeng Yang, Fengxia Zou, Hanjie Ying, Wei Zhuang. Water-mediated active conformational transitions of lipase on organic solvent interfaces. International Journal of Biological Macromolecules 2024, 277 , 134056. https://doi.org/10.1016/j.ijbiomac.2024.134056
  13. Shakeel Ahmed Mohammed, Shahbaz Aman, Bharat Singh. Unveiling the positive impacts of the genus Rhodococcus on plant and environmental health. Journal of Experimental Biology and Agricultural Sciences 2024, 12 (4) , 557-572. https://doi.org/10.18006/2024.12(4).557.572
  14. Zhi Zou, Bradley Higginson, Thomas R. Ward. Creation and optimization of artificial metalloenzymes: Harnessing the power of directed evolution and beyond. Chem 2024, 10 (8) , 2373-2389. https://doi.org/10.1016/j.chempr.2024.07.007
  15. Chao Zhong, Bernd Nidetzky. Bottom‐Up Synthesized Glucan Materials: Opportunities from Applied Biocatalysis. Advanced Materials 2024, 36 (27) https://doi.org/10.1002/adma.202400436
  16. Jingkai Yang, Ziyi Xu, Le Yu, Bingyun Wang, Ruibin Hu, Jiahu Tang, Jiahui Lv, Hongjun Xiao, Xuan Tan, Guanghui Wang, Jia‐Xin Li, Ying Liu, Pan‐Lin Shao, Bo Zhang. Organic Fluorophores with Large Stokes Shift for the Visualization of Rapid Protein and Nucleic Acid Assays. Angewandte Chemie 2024, 136 (17) https://doi.org/10.1002/ange.202318800
  17. Jingkai Yang, Ziyi Xu, Le Yu, Bingyun Wang, Ruibin Hu, Jiahu Tang, Jiahui Lv, Hongjun Xiao, Xuan Tan, Guanghui Wang, Jia‐Xin Li, Ying Liu, Pan‐Lin Shao, Bo Zhang. Organic Fluorophores with Large Stokes Shift for the Visualization of Rapid Protein and Nucleic Acid Assays. Angewandte Chemie International Edition 2024, 63 (17) https://doi.org/10.1002/anie.202318800
  18. Caroline Paul, Ulf Hanefeld, Frank Hollmann, Ge Qu, Bo Yuan, Zhoutong Sun. Enzyme engineering for biocatalysis. Molecular Catalysis 2024, 555 , 113874. https://doi.org/10.1016/j.mcat.2024.113874
  19. Manfred T. Reetz, Ge Qu, Zhoutong Sun. Engineered enzymes for the synthesis of pharmaceuticals and other high-value products. Nature Synthesis 2024, 3 (1) , 19-32. https://doi.org/10.1038/s44160-023-00417-0
  20. Marian J. Menke, Pascal Schneider, Christoffel P. S. Badenhorst, Andreas Kunzendorf, Florian Heinz, Mark Dörr, Martin A. Hayes, Uwe T. Bornscheuer. Ein universeller, kontinuierlicher Assay für SAM‐abhängige Methyltransferasen. Angewandte Chemie 2023, 135 (51) https://doi.org/10.1002/ange.202313912
  21. Marian J. Menke, Pascal Schneider, Christoffel P. S. Badenhorst, Andreas Kunzendorf, Florian Heinz, Mark Dörr, Martin A. Hayes, Uwe T. Bornscheuer. A Universal, Continuous Assay for SAM‐dependent Methyltransferases. Angewandte Chemie International Edition 2023, 62 (51) https://doi.org/10.1002/anie.202313912
  22. Yan Liu, Bingya Wang, Ji-Ting Hou, Peng Xie, Weiyi Li, Shan Wang. Molecular engineering and bioimaging applications of C2-alkenyl indole dyes with tunable emission wavelengths covering visible to NIR light. Bioorganic Chemistry 2023, 141 , 106905. https://doi.org/10.1016/j.bioorg.2023.106905
  23. Xianzhu Luo, Cuiling Zhang, Zihang Yu, Shihui Wen, Yuezhong Xian. Recent advances in responsive lanthanide-doped luminescence nanoprobes in the near-infrared-II window. TrAC Trends in Analytical Chemistry 2023, 169 , 117368. https://doi.org/10.1016/j.trac.2023.117368
  24. Nicoletta Cascelli, Vicente Gotor-Fernández, Iván Lavandera, Giovanni Sannia, Vincenzo Lettera. Spectrophotometric Assay for the Detection of 2,5-Diformylfuran and Its Validation through Laccase-Mediated Oxidation of 5-Hydroxymethylfurfural. International Journal of Molecular Sciences 2023, 24 (23) , 16861. https://doi.org/10.3390/ijms242316861

Chemical Reviews

Cite this: Chem. Rev. 2023, 123, 6, 2832–2901
Click to copy citationCitation copied!
https://doi.org/10.1021/acs.chemrev.2c00304
Published February 28, 2023

Copyright © 2023 The Authors. Published by American Chemical Society. This publication is licensed under

CC-BY 4.0 .

Article Views

12k

Altmetric

-

Citations

Learn about these metrics

Article Views are the COUNTER-compliant sum of full text article downloads since November 2008 (both PDF and HTML) across all institutions and individuals. These metrics are regularly updated to reflect usage leading up to the last few days.

Citations are the number of other articles citing this article, calculated by Crossref and updated daily. Find more information about Crossref citation counts.

The Altmetric Attention Score is a quantitative measure of the attention that a research article has received online. Clicking on the donut icon will load a page at altmetric.com with additional details about the score and the social media presence for the given article. Find more information on the Altmetric Attention Score and how the score is calculated.

  • Abstract

    Figure 1

    Figure 1. Change of analytical approach depending on screening scale and mutant library size: Increase in library size urges a shift from chromatographic methods to spectroscopic detection for (ultra) high-throughput screenings: (u)HTS. TLC: thin-layer chromatography; GC: gas chromatography; HPLC: high-performance liquid chromatography; (μ)MS: (microscale) mass spectrometry. Created with BioRender.com.

    Figure 2

    Figure 2. Overview of several ratiometric absorption or fluorescence turn-on strategies applied in directed enzyme evolution in biological systems: (a) direct detection of activity by functional group transformations or cleavage reaction of the target substrate, (b) indirect detection of conversion via dye cofactor recycling with a concomitant increase in signal, (c) triggered enhanced transcription of fluorescent biosensors, (d) formation of the dye in a downstream reaction upon cleavage of a metabolic reporter, detection via suitable recognition domain or subsequent reaction with a secondary chromo-/fluorogenic substrate, (e) induction of Förster resonance energy transfer (FRET) or suppression of photoinduced electron transfer (PET), (f) functional group transformation is detected via subsequent reaction with functional group selective assay component, (g) tandem reaction cascade of a model-substrate (mimic) unmasks chromo-/fluorophore.

    Scheme 1

    Scheme 1. Representative Reagents Typically Used in Thin-Layer Chromatography (TLC) Staining Solutions

    Figure 3

    Figure 3. (top) Summary of general properties of amines, reactivity trends toward amine-selective probes, and expected cross-reactivity. (bottom) Compatible pH range and buffers in reported amine-selective assays. Bold numbers indicate the most commonly used pH values.

    Scheme 2

    Scheme 2. Representative Selection of Enzymatic Transformations for the Production of Amine-Containing Products

    Figure 4

    Figure 4. Summary of chromo- or fluorogenic turn-on probes for the selective detection of amines based on nucleophilic aromatic substitution (SNAr) with aromatic molecules containing electron-withdrawing groups (EWG). Abs: wavelength typically used for UV absorbance measurements. Ex or em: wavelengths typically used for excitation and emission in fluorescence measurements. The wavelengths of probes in this section are shown on the strip at the bottom representing the visible spectrum.

    Figure 5

    Figure 5. Structures and spectral properties of SCOTfluors 14ae. aValues determined in EtOH. bQY (quantum yield) in dioxane using acridine orange, fluorescein, rhodamine 101, and Cy5 as standards. Reproduced with permission from ref (83). Copyright 2019 Wiley-VCH under Creative Commons Attribution 4.0 International License https://creativecommons.org/licenses/by/4.0/.

    Figure 6

    Figure 6. Labeling scheme of lysine residues in protein with 16. (a) Normalized absorbance spectra (dashed-lines) and fluorescence emission spectra (solid-lines) of 16 (1 μM) and lysozyme labeled with 16 (5 μg/mL lysozyme, 2 h incubation at 37 °C), measured in HEPES buffer (10 mM, pH 7.4, 0.1% dimethyl sulfoxide). The emission spectra were obtained with excitation at 485 and 375 nm, respectively. The inset photo was taken under UV light (365 nm). (b) A plot of time-dependent (0–4 h) fluorescence intensity at 409 nm. The excitation wavelength was 375 nm. Inset box: lysozyme activity calculated using an assay kit at 2 h incubation point. Means and standard deviations were calculated from triplicate measurements. Reproduced with permission from ref (90). Copyright 2017 Korean Chemical Society; Wiley-VCH.

    Figure 7

    Figure 7. Summary of chromo- or fluorogenic turn-on probes for the selective detection of amines based on Michael addition type reactions. Abs: wavelength typically used for UV absorbance measurements. Ex or em: wavelengths typically used for excitation and emission in fluorescence measurements. The wavelengths of probes in this section are shown on the strip at the bottom representing the visible spectrum.

    Figure 8

    Figure 8. Epicocconone 17 is the active ingredient in Deep Purple Total Protein Stain and is responsible for the apparent noncovalent staining of proteins in polyacrylamide gels and electroblots. The reaction of 17 with amines has shown that 17 reacts reversibly with primary amines to produce a highly fluorescent enamine that is readily hydrolyzed by base or strong acid. Such conditions are used in postelectrophoretic analysis, such as peptide mass fingerprinting or Edman degradation. Reproduced with permission from ref (92). Copyright 2005 American Chemical Society.

    Figure 9

    Figure 9. Normalized emission spectra of ylidenemalononitrile enamines 20ac in CH2Cl2 at room temperature. Insets show photographs of 1 mM solutions of 20a–c in CH2Cl2 irradiated at 365 nm. Reproduced with permission from ref (103). Copyright 2014 American Chemical Society.

    Figure 10

    Figure 10. Summary of chromo- or fluorogenic turn-on probes for the selective detection of amines based on isoindole formation with o-phthalaldehyde derivatives and o-diacetylbenzene. Ex or em: wavelengths typically used for excitation and emission in fluorescence measurements. The wavelengths of probes in this section are shown on the strip at the bottom representing the visible spectrum.

    Scheme 3

    Scheme 3. Reaction of 22 with Vicinal Diamine Resulting in the Formation of Pyrazino-diisoindole-dione 27

    Figure 11

    Figure 11. Summary of chromo- or fluorogenic turn-on probes for the selective detection of amines based on amide formation using activated esters. Ex or em: wavelengths typically used for excitation and emission in fluorescence measurements. The wavelengths of probes in this section are shown on the strip at the bottom representing the visible spectrum.

    Figure 12

    Figure 12. (top) Synthesis of fluorogenic meso-active-ester-BODIPYs 30a/30b and their conversion to fluorescent meso-amide-BODIPY products upon reaction with amines. (bottom) Absorption (a) and emission (b) spectra of 30b (10 μM) upon treatment with MeNH2 (20 μM) in CH3CN as a function of time (0–5 min) at 25 °C. λex = 470 nm. (inset) Relative fluorescence intensity at 546 nm as a function of incubation time, in the absence (red) and the presence (green) of MeNH2. (c) Relative fluorescence intensity at 546 nm as a function of [MeNH2] (0–50 μM). (inset) Calibration curve of fluorescence intensity at 546 nm vs [MeNH2] (0–6 μM). (d) Photographs of 30b (20 μM) upon addition of MeNH2 (left to right: 0, 3, 5, 10, 20 μM) under ambient light (left) and 365 nm UV irradiation (right). Incubation time = 5 min. Reproduced with permission from ref (131). Copyright 2020 American Chemical Society.

    Figure 13

    Figure 13. Summary of chromo- or fluorogenic turn-on probes for the selective detection of amines based on pyridinium salt formation with pyrylium salts. Ex or em: wavelengths typically used for excitation and emission in fluorescence measurements. The wavelengths of probes in this section are shown on the strip at the bottom representing the visible spectrum.

    Figure 14

    Figure 14. Image of a Py-1 31 sensing microplate reacted with various concentrations of histamine (the numbers indicate the concentrations of histamine in μg mL–1). Reproduced with permission from ref (137). Copyright 2011 The Royal Society of Chemistry.

    Figure 15

    Figure 15. Colors under daylight and the colors of fluorescence in solution of pH 7.0 under a UV lamp (365 nm) of label 31 and of some other Py labels when conjugated to serum albumin at 37 °C. Note that label 33 does not react but remains blue (at any concentration). Reproduced with permission from ref (141). Copyright 2011 The Royal Society of Chemistry.

    Figure 16

    Figure 16. Summary of chromo- or fluorogenic turn-on probes for the selective detection of amines based on the formation of thioureas or 2-aminothiazoles from isothiocyanates. Ex or em: wavelengths typically used for excitation and emission in fluorescence measurements. The wavelengths of probes in this section are shown on the strip at the bottom representing the visible spectrum.

    Figure 17

    Figure 17. (top) Fluorescence spectra of unbound 9-ITPPo 35 (blue) and its covalent adduct ATAZPo (red) with BSA (a) and gold nanoclusters (AuNCs) (b). (bottom) Absorbance of a mixture of 35 (15 μM) and n-butylamine (3 mM) in toluene: Structures, time curves of concentration, and spectral profiles of the reactant, the intermediate, and the product (from left to right). Reproduced with permission from ref (148). Copyright 2015 The Royal Society of Chemistry.

    Figure 18

    Figure 18. Summary of chromo- or fluorogenic turn-on probes for the selective detection of amines based on hemiaminal, aminal, or Schiff’s base formation. Abs: wavelengths typically used for UV absorbance measurements. Ex or em: wavelengths typically used for excitation and emission in fluorescence measurements. The wavelengths of probes in this section are shown on the strip at the bottom representing the visible spectrum.

    Figure 19

    Figure 19. Amine-sensing panels of 37 (10 μmol L−1) in water: (1) 37; (2) butylamine; (3) tert-butylamine; (4) benzylamine; (5) cyclohexylamine; (6) ethylene diamine; (7) 1,3-diaminopropane; (8) cadaverine; (9) morpholine; (10) ephedrine; (11) 4-aminopyridine; (12) ethanolamine. Reproduced with permission form ref (154). Copyright 2012 Wiley-VCH.

    Scheme 4

    Scheme 4. Sensing Mechanism of 36: Chelation of the Formed Iminium Species by the Coumarin Moiety Triggers a Turn-on Response (152)

    Scheme 5

    Scheme 5. Ozone Sensing Mechanisms Employing 38. (34)

    Figure 20

    Figure 20. Summary of general properties of alcohols and expected cross-reactivity (bottom).

    Scheme 6

    Scheme 6. Representative Selection of Enzymatic Transformations for the Production of Alcohol-Containing Products

    Scheme 7

    Scheme 7. MTT Assay: Detection of NAD(P)H or “Reducing Equivalents” via the Reduction of MTT 55 to the Purple Formazan 56

    Figure 21

    Figure 21. Summary of chromo- or fluorogenic turn-on probes for the selective detection of alcohols based on hemiacetal formation. Abs: wavelengths typically used for UV absorbance measurements. Ex or em: wavelengths typically used for excitation and emission in fluorescence measurements. The wavelengths of probes in this section are shown on the strip at the bottom representing the visible spectrum.

    Figure 22

    Figure 22. (a) Fluorescence–time trajectories following the reaction of 4.5 μM of compound 57 (R = Cl, R′′ = H) and 1.09 M methanol (MeOH), ethanol (EtOH), butanol (BuOH), ethylene glycol (EtyGly), water (H2O), or ethanedithiol at 21 °C in ACN supplemented with 0.333 mM of p-TsOH. (b) Apparent rate constants for various nucleophiles tested were obtained from fitting the intensity–time trajectories in (a). Reproduced with permission from ref (195). Copyright 2017 American Chemical Society.

    Figure 23

    Figure 23. Photographs showing the color changes of sensor 58 before and after the addition of methanol vapor. The images were taken under UV irradiation (365 nm). Reproduced with permission from ref (197). Copyright 2021 Royal Society of Chemistry and the Centre National de la Recherche Scientifique.

    Figure 24

    Figure 24. Summary of chromo- or fluorogenic turn-on probes for the selective detection of alcohols based on the coordination to chelation centers. Ex or em: wavelengths typically used for excitation and emission in fluorescence measurements. The wavelengths of probes in this section are shown in the strip at the bottom representing the visible spectrum.

    Figure 25

    Figure 25. Summary of chromo- or fluorogenic turn-on probes for the selective detection of alcohols based on the ring-opening of 3-aminorhodanine derivatives. Ex or em: wavelengths typically used for excitation and emission in fluorescence measurements. The wavelengths of probes in this section are shown in the strip at the bottom representing the visible spectrum.

    Figure 26

    Figure 26. (a) The visible color change of 65 (5 mM) in different solvents: (i) hexane, (ii) toluene, (iii) cyclohexane, (iv) ethyl acetate, (v) chloroform, (vi) dichloromethane, (vii) tetrahydrofuran, (viii) dioxane, (ix) ACN, (x) MeOH, (xi) EtOH, (xii) iPrOH, and (xiii) nBuOH. (b) The absorption spectra of 65 in different solvents, (c) time-dependent absorption spectra (0–30 min) of 5 mM 65 in MeOH; inset: plot of optical density at 470 nm vs time. (d,e) Absorption spectra of 65 (5 mM) at different ratios of ACN:EtOH (0–10%) and ACN:MeOH (0–10%), respectively. Reproduced with permission from ref (205). Copyright 2019 Royal Society of Chemistry and the Centre National de la Recherche Scientifique.

    Figure 27

    Figure 27. (top) Summary of general properties of diol sensors and expected cross-reactivity. (bottom) Compatible pH range and buffers in reported diol-selective assays. Bold numbers indicate the most commonly used pH values.

    Scheme 8

    Scheme 8. Representative Selection of Enzymatic Transformations for the Production of Diol-Containing Products

    Figure 28

    Figure 28. Summary of fluorogenic turn-on probes for the selective detection of diols based on the formation of fluorescent boronate esters. Ex or em: wavelengths typically used for excitation and emission in fluorescence measurements. The wavelengths of probes in this section are shown in the strip at the bottom representing the visible spectrum.

    Figure 29

    Figure 29. Sensing mechanism for the selective detection of fructose or glucose-6-phosphate by 91 or 92 and viologen HPTS 93. Reproduced with permission from ref (253). Copyright 2009 Elsevier BV.

    Figure 30

    Figure 30. Selective detection of diols based on the cleavage reaction by metaperiodate 97 and subsequent application of aldehyde sensing methodology (for aldehyde-selective probes, see section 9).

    Scheme 9

    Scheme 9. (a) Oxidation of Adrenaline 98 by Metaperiodate 97, Forming Red-Colored Adrenochrome 99, Which Allows for Determination of Excess 97 (Back Titration of Diol); (263) (b) Structure of Carboxyfluorescein, Which Can Also Be Used as a 97 Indicator; The Mechanism for Its Oxidation Is Not Known

    Figure 31

    Figure 31. (a) Coupled reactions employed by the fluorescence-based assay system for the detection of cis-diol metabolites using fluoresceinamine 127. (b) Concentration-dependent fluorescence response of the assay to the presence of cis-diol metabolites. All studies were performed in triplicate; all measurements demonstrated standard deviation <5%; fluorescence responses normalized to negative control ([II0]/I0). Reproduced with permission from ref (270). Copyright 2021 The Royal Society of Chemistry.

    Figure 32

    Figure 32. (top) Summary of general properties of aldehydes, reactivity trends toward aldehyde-selective probes, as well as expected cross-reactivity during screenings. (bottom) Compatible pH range and buffers in reported aldehyde-selective assays. Bold numbers indicate the most commonly used pH values.

    Scheme 10

    Scheme 10. Selection of Enzymatic Transformations for the Production of Aldehyde-Containing Products

    Figure 33

    Figure 33. Summary of chromo- and fluorogenic turn-on probes for the selective detection of aldehydes based on the formation of hydrazone adducts. Abs: wavelengths typically used for UV absorbance measurements. Ex or em: wavelengths typically used for excitation and emission in fluorescence measurements. The wavelengths of probes in this section are shown in the strip at the bottom representing the visible spectrum.

    Figure 34

    Figure 34. Plot of the nucleophilicity parameters N of amines versus pKaH in water. (296) Reproduced with permission from refs. (297and298). Copyright 2007 American Chemical Society.

    Scheme 11

    Scheme 11. Sensing Mechanism of Purpald 113via the Intermediate Formation of Cyclic Tetrazinanes and Subsequent Oxidation to Fully Aromatized Violet-Colored Tetrazines 115

    Scheme 12

    Scheme 12. Sensing Mechanism for MBTH 114

    Figure 35

    Figure 35. (a) 119 as a fluorogenic probe for the detection of MDA. (a) Malondialdehyde (MDA) was added to a 25 μM PBS solution of 119 (pH 7.2) under illumination at 365 nm and photographed at indicated time points. (b) (left) Emission spectra of 1 μM 119 (black) or 119 after 1 min incubation with 5 equiv MDA at 37 °C and excited at 310 nm (blue) or 375 nm (cyan). (right) Formation and rapid decay of short-lived intermediate formed in the reaction between 119 and MDA followed by fluorescence spectroscopy with λexem = 375/480 nm. Kinetics of 25 μM 119 only (squares) and 119 + MDA (circles) are shown, with three replicates of 119 + MDA shown (light, medium, and dark cyan). (c) 119 shows broad specificity to aliphatic aldehydes. The enhancement of fluorescence relative to 119 only was recorded with λexem 375/480 nm following 1 min of incubation of 1 μM solutions of 119 in PBS (pH 7.2) with 5 equiv of MDA (black bar), simple aliphatic aldehydes (red bars), dialdehydes, and complex aldehydes (blue bars), ketones (purple bars), glutathione (yellow bar), and biogenic metal cations (orange bars). Reproduced with permission from ref (312). Copyright 2019 The Royal Society of Chemistry.

    Figure 36

    Figure 36. (top) Scheme of the reaction mechanism of colorimetric and fluorescent detection of aldehydes based on the SQs–N2H4 adduct 121. (bottom) The changes in UV–vis absorption of dye 121 (6.0 mM) upon addition of N2H4 (3.0 mM) and subsequent addition of formaldehyde (10.0 mM). Reproduced with permission from ref (314). Copyright 2019 Elsevier BV.

    Figure 37

    Figure 37. Summary of chromo- and fluorogenic turn-on probes for the selective detection of aldehydes based on the formation of imines (Schiff’s bases). Subsequent entrapment of these electrophilic imines by intra- or intermolecular nucleophiles makes the reaction irreversible. Abs: wavelengths typically used for UV Absorbance measurements. Ex or em: wavelengths typically used for excitation and emission in fluorescence measurements. The wavelengths of probes in this section are shown in the strip at the bottom representing the visible spectrum.

    Figure 38

    Figure 38. (a) The color of the aldehyde-sensing probe 122 changes to purple when treated with formaldehyde (detection limit: μM concentrations of formaldehyde in the in vitro system). (b) Color changes on solid agar plate from O-dealkylation activity. Aldehyde appears to diffuse out of the cells; cell color changes are observed in the presence of 122. Reproduced with permission from ref (23). Copyright 2013 Wiley-VCH.

    Figure 39

    Figure 39. Relationship between the measured reaction rates and pKa of the corresponding protonated anilines for seven ABAO analogues 126a–g. (i) Structures of ABAOs 126 and corresponding protonated anilines. (ii) Scatter plot showing relationship between log k and pKa. (iii) Reaction rates measured by NMR in CD3COONa buffer (100 mM) at pD 4.5.(iv) Structures and reaction rates for 126f and 126g. Basicity of aniline nitrogen in 2-aminopyridine / 126f was estimated using ab initio calculations. Reproduced with permission from ref (319). Copyright 2014 American Chemical Society.

    Scheme 13

    Scheme 13. Sensing Mechanism for dRB-EDA 128: Schiff’s Base Formation with Formaldehyde Triggers the Ring-Opening of the Deoxylactam Moiety

    Figure 40

    Figure 40. Chromogenic detection of formaldehyde with 128. (a) UV–absorption spectra of 128 in the presence of formaldehyde (concentration: 0, 3.5, 7, 10.5, 14, 17.5, and 35 mM, from bottom to top). The inset shows the titration curve by absorbance at 560 nm. (b) Visual detection of formaldehyde (0–7 mM as indicated) with 128 in DMF after incubation at room temperature for 90 min. Reproduced with permission from ref (324). Copyright 2011 The Royal Society of Chemistry.

    Figure 41

    Figure 41. Summary of fluorogenic turn-on probes for the selective detection of aldehydes based on the formation of imidazole structures. Ex or em: wavelengths typically used for excitation and emission in fluorescence measurements. The wavelengths of probes in this section are shown in the strip at the bottom representing the visible spectrum.

    Figure 42

    Figure 42. Summary of fluorogenic turn-on probes for the selective detection of aldehydes based on the formation of 1,4-dihydropyridines. Ex or em: wavelengths typically used for excitation and emission in fluorescence measurements. The wavelengths of probes in this section are shown in the strip at the bottom representing the visible spectrum.

    Figure 43

    Figure 43. Summary of chromo- and fluorogenic turn-on probes for the selective detection of aldehydes based on the formation of unstable oximes. Fragmentation of these intermediates leads to the liberation of colored phenols. Abs: wavelengths typically used for UV absorbance measurements. Ex or em: wavelengths typically used for excitation and emission in fluorescence measurements. The wavelengths of probes in this section are shown in the strip at the bottom representing the visible spectrum.

    Scheme 14

    Scheme 14. Enzyme-Catalyzed Formation of the 138–Benzaldehyde Adduct Could Be Efficiently Traced by Spontaneous Decomposition Forming Colored NBD-O

    Figure 44

    Figure 44. Summary of fluorogenic turn-on probes for the selective detection of formaldehyde based on the 2-aza-Cope sigmatropic rearrangement of initially formed Schiff’s bases. Fragmentation of these intermediates leads to the liberation of colored phenols. Ex or em: wavelengths typically used for excitation and emission in fluorescence measurements. The wavelengths of probes in this section are shown in the strip at the bottom representing the visible spectrum.

    Figure 45

    Figure 45. Confocal fluorescence images of exogenous formaldehyde (FA) with FFP fluorescent probes 142 in HEK293T cells. (A–D) Cells were loaded with (a) FFP511 (Tokyo Green fluorophore) (20 μM), (b) FFP551 (142) (10 μM), (c) FFP585 (resorufin fluorophore) (5 μM), or FFP706 (hemicyanine fluorophore) (2 μM) for 30 min and treated with FA of varying concentrations (0–1 mM) for 60 min before being imaged by confocal fluorescence microscopy. Scale bars represent 50 μm. Reproduced with permission from ref (352). Copyright 2021 The Royal Society of Chemistry.

    Figure 46

    Figure 46. Fluorescent responses of 143 (10 μM) in DMF/PBS solution (v/v = 1/4, pH 7.4, 10 mM) upon addition of various small molecular species (5.0 mM). λexem = 390/513 nm; slits, 2.5/2.5 nm. Each spectrum was recorded after 3 h at 37 °C. Error bars are ±1SD, n = 3. Inset: (a) The color and (b) fluorescence images of 143 (10 μM) in the presence of various small molecular species (5.0 mM). (354) Reproduced with permission from ref (354). Copyright 2016 Elsevier BV.

    Figure 47

    Figure 47. (top) Summary of general properties of ketones, reactivity trends toward ketone-selective probes as well as expected cross-reactivity during screenings. (bottom) Compatible pH range and buffers in reported ketone-selective assays. Bold pH numbers indicate the most commonly used pH values.

    Scheme 15

    Scheme 15. Representative Selection of Enzymatic Transformations for the Production of Carboxylic Acid-Containing Products

    Figure 48

    Figure 48. Summary of chromo- and fluorogenic turn-on probes for the selective detection of ketones based on the formation of hydrazone adducts. Abs: wavelengths typically used for UV absorbance measurements. Ex or em: wavelengths typically used for excitation and emission in fluorescence measurements. The wavelengths of probes in this section are shown in the strip at the bottom representing the visible spectrum.

    Figure 49

    Figure 49. Variation in color of reaction mixtures containing different concentrations of 109 and the validation of the developed HTS method by chiral HPLC analysis. Control, no 109 was added into the reaction mixture; 1–10×, different concentrations of 109 were added into the reaction mixture (1×, 8.4 mg/mL). Reproduced with permission from ref (287). Copyright 2017 Springer-Verlag.

    Figure 50

    Figure 50. Summary of fluorogenic turn-on probes for the selective detection of acetone based on aldol reactions. Ex or em: wavelengths typically used for excitation and emission in fluorescence measurements. The wavelengths of probes in this section are shown in the strip at the bottom representing the visible spectrum.

    Figure 51

    Figure 51. Summary of fluorogenic turn-on probes for the selective detection of acetone based on a Michael addition approach. Ex or em: wavelengths typically used for excitation and emission in fluorescence measurements. The wavelengths of probes in this section are shown in the strip at the bottom representing the visible spectrum.

    Scheme 16

    Scheme 16. Michael Addition of Acetone onto 159 Triggers a Subsequent Aldol Reaction Forming Cyclic Hydroxyketones

    Figure 52

    Figure 52. (top) Chromogenic and fluorogenic detection of ketones bearing an enolizable position via the formation of the purple-colored Janovsky complexes with 3,5-dinitrobenzoic acid 161 under strongly basic conditions. (bottom) Fluorogenic detection of ketones via the formation of fluorescent ABAO-adducts (cf. section 9.2). A methoxy substitution of the ABAO core 126 is necessary to increase the reactivity of the aniline NH2 position.

    Figure 53

    Figure 53. (top) Summary of general properties of carboxylic acid sensors as well as expected cross-reactivity during screenings. (bottom) Compatible pH range and buffers in reported carboxylic acid-selective assays. Bold pH numbers indicate the most commonly used pH values.

    Scheme 17

    Scheme 17. Representative Selection of Enzymatic Transformations for the Production of Carboxylic Acid-Containing Products

    Figure 54

    Figure 54. Summary of chromogenic turn-on probes for the selective detection of carboxylic acids based on the entrapment of activated carboxylic acid species. Abs: wavelengths typically used for UV absorbance measurements. The wavelengths of probes in this section are shown in the strip at the bottom representing the visible spectrum.

    Figure 55

    Figure 55. Summary of fluorogenic turn-on probes for the selective detection of carboxylic acids based on the coordination to metal centers or hydrogen bond donors. Ex or em: wavelengths typically used for excitation and emission in fluorescence measurements. The wavelengths of probes in this section are shown in the strip at the bottom representing the visible spectrum.

    Figure 56

    Figure 56. Summary of fluorogenic turn-on probes for the selective detection of carboxylic acids based on nucleophilic addition onto trifluoroacetyl groups. Ex or em: wavelengths typically used for excitation and emission in fluorescence measurements. The wavelengths of probes in this section are shown in the strip at the bottom representing the visible spectrum.

    Figure 57

    Figure 57. (top) Summary of general properties of carboxylic acid ester sensors as well as expected cross-reactivity during screenings. Due to the lack of direct reactive probes applicable in aqueous environments, no compatible buffers and pH ranges can be stated.

    Scheme 18

    Scheme 18. Representative Selection of Enzymatic Transformations for the Production of Carboxylic Acid Ester-Containing Products

    Figure 58

    Figure 58. Chromogenic detection of carboxylic acid esters via the formation of the hydroxamic acids and subsequent complexation with Fe3+ resulting in purple-colored ferric hydroxamates.

    Figure 59

    Figure 59. (top) Summary of general properties of epoxide sensors as well as expected cross-reactivity during screenings. (bottom) Compatible pH range and buffers in reported epoxide-selective assays. Bold pH numbers indicate the most commonly used pH values.

    Scheme 19

    Scheme 19. Representative Selection of Enzymatic Transformations for the Production of Epoxide-Containing Products

    Figure 60

    Figure 60. Summary of chromo- and fluorogenic turn-on probes for the selective detection of epoxides based on ring-opening reactions by nucleophilic pyridine derivatives. Abs: wavelengths typically used for UV absorbance measurements. Ex or em: wavelengths typically used for excitation and emission in fluorescence measurements. The wavelengths of probes in this section are shown in the strip at the bottom representing the visible spectrum.

    Scheme 20

    Scheme 20. Sensing Mechanism for Epoxides Employing the Combination of Nicotinamide 192 and Acetophenone 193

    Figure 61

    Figure 61. Selective detection of epoxides based on indirect detection methods after initial nucleophilic ring-opening reactions with hydroxide or bisulfide. The resulting intermediates can be detected with thiol sensing assays (cf. section 15), diol sensing assays (cf. section 8), or by using a similar method discussed for the indirect detection of diols by oxidative cleavage with metaperiodate 97 and subsequent detection using aldehyde-sensing assays (cf. section 9).

    Figure 62

    Figure 62. (top) Summary of general properties of phenol sensors as well as expected cross-reactivity during screenings. (bottom): Compatible pH range and buffers in reported phenol-selective assays. Bold pH numbers indicate the most commonly used pH values.

    Scheme 21

    Scheme 21. Representative Selection of Enzymatic Transformations for the Production of Phenol-Containing Products

    Figure 63

    Figure 63. Summary of chromogenic turn-on probes for the selective detection of phenols based on the electrophilic aromatic substitution with oxidized amino derivatives. Abs: wavelengths typically used for UV absorbance measurements. Ex or em: wavelengths typically used for excitation and emission in fluorescence measurements. The wavelengths of probes in this section are shown in the strip at the bottom representing the visible spectrum. Alk = alkyl, Ar = aryl.

    Figure 64

    Figure 64. Summary of chromogenic turn-on probes for the selective detection of phenols based on the formation of azo-dyes. Abs: wavelengths typically used for UV absorbance measurements. The wavelengths of probes in this section are shown in the strip at the bottom representing the visible spectrum.

    Figure 65

    Figure 65. Summary of fluorogenic turn-on probes for the selective detection of phenol based on Mannich-type reactions. Ex or em: wavelengths typically used for excitation and emission in fluorescence measurements. The wavelengths of probes in this section are shown in the strip at the bottom representing the visible spectrum.

    Figure 66

    Figure 66. (top) Summary of general properties of thiols, reactivity trends toward thiol-selective probes, and expected cross-reactivity. (bottom) Compatible pH range and buffers in reported thiol-selective assays. Bold pH numbers indicate the most commonly used pH values.

    Figure 67

    Figure 67. Summary of fluorogenic turn-on probes for the selective detection of thiols based on nucleophilic aromatic substitution reactions (SNAr) with aromatic systems containing electron-withdrawing groups (EWG). Ex or em: wavelengths typically used for excitation and emission in fluorescence measurements. The wavelengths of probes in this section are shown in the strip at the bottom representing the visible spectrum.

    Scheme 22

    Scheme 22. Proposed Mechanism for the Reaction of NBD-Cl 12 with Cys, Hcy, and GSH (581)

    Figure 68

    Figure 68. Design of the probe HMN 219 and the proposed fluorescence signal changes in response to individual or sequential detection of thiols with three well-defined emission bands. Reproduced with permission from ref (565). Copyright 2017 Royal Society of Chemistry.

    Figure 69

    Figure 69. Summary of fluorogenic turn-on probes for the selective detection of thiols based on 2,4-dinitrobenzene sulfonyl (DNBS) cleavage by nucleophilic aromatic substitution reactions (SNAr) with aromatic systems containing electron-withdrawing groups (EWG). Ex or em: wavelengths typically used for excitation and emission in fluorescence measurements. The wavelengths of probes in this section are shown in the strip at the bottom representing the visible spectrum.

    Figure 70

    Figure 70. Fluorescent probe BESThio 222 and its discrimination between selenols and thiols. Reproduced with permission from ref (594). Copyright 2006 Wiley-VCH.

    Figure 71

    Figure 71. Summary of fluorogenic turn-on probes for the selective detection of thiols based on disulfide bond cleavage. Abs: wavelengths typically used for UV absorbance measurements. Ex or em: wavelengths typically used for excitation and emission in fluorescence measurements. The wavelengths of probes in this section are shown in the strip at the bottom representing the visible spectrum.

    Figure 72

    Figure 72. Schematic illustration of redox-controlled fluorescent nanoswitch and detection strategy for BChE activity based on thiol-triggered disulfide cleavage on fluorescent carbon nanoparticles. Reproduced with permission from ref (611). Copyright 2018 American Chemical Society.

    Figure 73

    Figure 73. Summary of fluorogenic turn-on probes for the selective detection of thiols based on diselenide bond cleavage. Ex or em: wavelengths typically used for excitation and emission in fluorescence measurements. The wavelengths of probes in this section are shown in the strip at the bottom representing the visible spectrum.

    Figure 74

    Figure 74. Summary of fluorogenic turn-on probes for the selective detection of thiols based on selenium–nitrogen bond cleavage. Ex or em: wavelengths typically used for excitation and emission in fluorescence measurements. The wavelengths of probes in this section are shown in the strip at the bottom representing the visible spectrum.in the strip at the bottom representing the visible spectrum.

    Figure 75

    Figure 75. Structure and color changes of 243 in the presence of amino acids (1, Cys; 2, Hcy; 3, GSH; 4, none; 5, other natural amino acids). Reproduced with permission from ref (628) . Copyright 2018 Elsevier BV.

    Figure 76

    Figure 76. Summary of fluorogenic turn-on probes for the selective detection of thiols based on the cyclization with aldehydes. A subsequent cyclization via an additional NH2 moiety limits the application to analytes containing both functional groups. Abs: wavelengths typically used for UV absorbance measurements. Ex or em: wavelengths typically used for excitation and emission in fluorescence measurements. The wavelengths of probes in this section are shown in the strip at the bottom representing the visible spectrum.

    Figure 77

    Figure 77. Summary of fluorogenic turn-on probes for the selective detection of thiols-based conjugate additions. Ex or em: wavelengths typically used for excitation and emission in fluorescence measurements. The wavelengths of probes in this section are shown in the strip at the bottom representing the visible spectrum.

    Figure 78

    Figure 78. Illustration of the high-throughput assays for cystathionine β-synthase (CBS). (a) Methanethiol (CH3SH) generation catalyzed by CBS using methylcysteine (MCys) as a substrate was determined with the CPM 246. (b) The CBS activity was monitored using 10 mM MCys as the substrate in the presence of 15 μM 246 and 2 μg of hCBS in a 200 μL reaction mixture containing 50 mM HEPES, pH 7.4. The fluorescence of the reaction mixture at 460 nm (λex = 400 nm) was monitored for 600 s. Reproduced with permission from ref (637). Copyright 2017 Royal Society of Chemistry.

    Scheme 23

    Scheme 23. Proposed Response Mechanism for NIR-BODIPY-Ac 251 to Cys or Hcy

    Figure 79

    Figure 79. Summary of fluorogenic turn-on probes for the selective detection of thiols-based sulfur–metal interactions. Ex or em: wavelengths typically used for excitation and emission in fluorescence measurements. The wavelengths of probes in this section are shown in the strip at the bottom representing the visible spectrum.

    Figure 80

    Figure 80. Schematic illustration of the thiol detecting chemodosimetric mechanism iminocoumarin-Cu2+ probe 253 in aqueous media. Adapted with permission from ref (651). Copyright 2011 The Royal Society of Chemistry.

    Figure 81

    Figure 81. Schematic representation of the thiol sensor 256 based on modulation of the fluorescence quenching of the BODIPY chromophore by AuNPs. Reproduced with permission from ref (655) . Copyright 2016 Elsevier BV.

  • References


    This article references 679 other publications.

    1. 1
      Li, C.; Zhang, R.; Wang, J.; Wilson, L. M.; Yan, Y. Protein Engineering for Improving and Diversifying Natural Product Biosynthesis. Trends Biotechnol. 2020, 38 (7), 72944,  DOI: 10.1016/j.tibtech.2019.12.008
    2. 2
      Kazlauskas, R. Engineering More Stable Proteins. Chem. Soc. Rev. 2018, 47 (24), 902645,  DOI: 10.1039/C8CS00014J
    3. 3
      Behrens, G. A.; Hummel, A.; Padhi, S. K.; Schätzle, S.; Bornscheuer, U. T. Discovery and Protein Engineering of Biocatalysts for Organic Synthesis. Adv. Synth. Catal. 2011, 353 (13), 2191215,  DOI: 10.1002/adsc.201100446
    4. 4
      Liu, Z.; Arnold, F. H. New-to-Nature Chemistry from Old Protein Machinery: Carbene and Nitrene Transferases. Curr. Opin. Biotechnol. 2021, 69, 4351,  DOI: 10.1016/j.copbio.2020.12.005
    5. 5
      Steiner, K.; Schwab, H. Recent Advances in Rational Approaches for Enzyme Engineering. Comput. Struct. Biotechnol. J. 2012, 2 (3), e201209010  DOI: 10.5936/csbj.201209010
    6. 6
      Zeymer, C.; Hilvert, D. Directed Evolution of Protein Catalysts. Annu. Rev. Biochem. 2018, 87 (1), 13157,  DOI: 10.1146/annurev-biochem-062917-012034
    7. 7
      Markel, U.; Essani, K. D.; Besirlioglu, V.; Schiffels, J.; Streit, W. R.; Schwaneberg, U. Advances in Ultrahigh-Throughput Screening for Directed Enzyme Evolution. Chem. Soc. Rev. 2020, 49 (1), 23362,  DOI: 10.1039/C8CS00981C
    8. 8
      Arnold, F. H.; Volkov, A. A. Directed Evolution of Biocatalysts. Curr. Opin. Chem. Biol. 1999, 3 (1), 549,  DOI: 10.1016/S1367-5931(99)80010-6
    9. 9
      Arnold, F. H.; Smith, G. P.; Winter, G. P. Directed Evolution-Bringing the Power of Evolution to the Laboratory: 2018 Nobel Prize in Chemistry. Curr. Sci. 2018, 115 (9), 1627
    10. 10
      Goldsmith, M.; Tawfik, D. S. Enzyme Engineering: Reaching the Maximal Catalytic Efficiency Peak. Curr. Opin. Struct. Biol. 2017, 47, 14050,  DOI: 10.1016/j.sbi.2017.09.002
    11. 11
      Currin, A.; Swainston, N.; Day, P. J.; Kell, D. B. Synthetic Biology for the Directed Evolution of Protein Biocatalysts: Navigating Sequence Space Intelligently. Chem. Soc. Rev. 2015, 44 (5), 1172239,  DOI: 10.1039/C4CS00351A
    12. 12
      Reymond, J. L. Enzyme Assays; Wiley, 2006.
    13. 13
      Singh, H.; Tiwari, K.; Tiwari, R.; Pramanik, S. K.; Das, A. Small Molecule as Fluorescent Probes for Monitoring Intracellular Enzymatic Transformations. Chem. Rev. 2019, 119 (22), 1171860,  DOI: 10.1021/acs.chemrev.9b00379
    14. 14
      Chan, J.; Dodani, S. C.; Chang, C. J. Reaction-Based Small-Molecule Fluorescent Probes for Chemoselective Bioimaging. Nat. Chem. 2012, 4 (12), 97384,  DOI: 10.1038/nchem.1500
    15. 15
      Reetz, M. T.; Carballeira, J. D. Iterative Saturation Mutagenesis (ISM) for Rapid Directed Evolution of Functional Enzymes. Nat. Protoc. 2007, 2 (4), 891903,  DOI: 10.1038/nprot.2007.72
    16. 16
      Lapetina, S.; Gil-Henn, H. A Guide to Simple, Direct, and Quantitative in Vitro Binding Assays. J. Biol. Methods 2017, 4 (1), e62, DOI: 10.14440/jbm.2017.161 .
    17. 17
      Major, J. Challenges and Opportunities in High Throughput Screening: Implications for New Technologies. J. Biomol. Screen. 1998, 3 (1), 137,  DOI: 10.1177/108705719800300102
    18. 18
      Neun, S.; Kaminski, T. S.; Hollfelder, F. Single-Cell Activity Screening in Microfluidic Droplets. Methods Enzymol. 2019, 628 (5), 95112
    19. 19
      Zhang, J.; Chai, X.; He, X.-P.; Kim, H.-J.; Yoon, J.; Tian, H. Fluorogenic Probes for Disease-Relevant Enzymes. Chem. Soc. Rev. 2019, 48 (2), 683722,  DOI: 10.1039/C7CS00907K
    20. 20
      Kaur, K.; Saini, R.; Kumar, A.; Luxami, V.; Kaur, N.; Singh, P.; Kumar, S. Chemodosimeters: An Approach for Detection and Estimation of Biologically and Medically Relevant Metal Ions, Anions and Thiols. Coord. Chem. Rev. 2012, 256 (17), 19922028,  DOI: 10.1016/j.ccr.2012.04.013
    21. 21
      You, L.; Arnold, F. H. Directed Evolution of Subtilisin E in Bacillus Subtilis to Enhance Total Activity in Aqueous Dimethylformamide. Protein Eng. 1996, 9 (1), 7783,  DOI: 10.1093/protein/9.1.77
    22. 22
      Schmidt-Dannert, C.; Arnold, F. H. Directed Evolution of Industrial Enzymes. Trends Biotechnol. 1999, 17 (4), 1356,  DOI: 10.1016/S0167-7799(98)01283-9
    23. 23
      Choi, K.-Y.; Jung, E.-O.; Yun, H.; Yang, Y.-H.; Kazlauskas, R. J.; Kim, B.-G. Development of Colorimetric HTS Assay of Cytochrome P450 for Ortho-Specific Hydroxylation, and Engineering of CYP102D1 with Enhanced Catalytic Activity and Regioselectivity. ChemBioChem. 2013, 14 (10), 12318,  DOI: 10.1002/cbic.201300212
    24. 24
      Roth, M. Fluorescence Reaction for Amino Acids. Anal. Chem. 1971, 43 (7), 8802,  DOI: 10.1021/ac60302a020
    25. 25
      Weigele, M.; DeBernardo, S. L.; Tengi, J. P.; Leimgruber, W. Novel Reagent for the Fluorometric Assay of Primary Amines. J. Am. Chem. Soc. 1972, 94 (16), 59278,  DOI: 10.1021/ja00771a084
    26. 26
      Fellner, M.; Doughty, L. M.; Jameson, G. N.; Wilbanks, S. M. A Chromogenic Assay of Substrate Depletion by Thiol Dioxygenases. Anal. Biochem. 2014, 459, 5660,  DOI: 10.1016/j.ab.2014.05.008
    27. 27
      Santos-Aberturas, J.; Dörr, M.; Waldo, G. S.; Bornscheuer, U. T. In-Depth High-Throughput Screening of Protein Engineering Libraries by Split-GFP Direct Crude Cell Extract Data Normalization. Chem. Biol. 2015, 22 (10), 140614,  DOI: 10.1016/j.chembiol.2015.08.014
    28. 28
      Azad, T.; Tashakor, A.; Hosseinkhani, S. Split-Luciferase Complementary Assay: Applications, Recent Developments, and Future Perspectives. Anal. Bioanal. Chem. 2014, 406 (23), 554160,  DOI: 10.1007/s00216-014-7980-8
    29. 29
      Bennett, B. D.; Kimball, E. H.; Gao, M.; Osterhout, R.; Van Dien, S. J.; Rabinowitz, J. D. Absolute Metabolite Concentrations and Implied Enzyme Active Site Occupancy in Escherichia Coli. Nat. Chem. Biol. 2009, 5 (8), 5939,  DOI: 10.1038/nchembio.186
    30. 30
      Zuman, P. Reactions of Orthophthalaldehyde with Nucleophiles. Chem. Rev. 2004, 104 (7), 321738,  DOI: 10.1021/cr0304424
    31. 31
      Henke, E.; Bornscheuer, U. T. Fluorophoric Assay for the High-Throughput Determination of Amidase Activity. Anal. Chem. 2003, 75 (2), 25560,  DOI: 10.1021/ac0258610
    32. 32
      Médici, R.; de María, P. D.; Otten, L. G.; Straathof, A. J. J. A High-Throughput Screening Assay for Amino Acid Decarboxylase Activity. Adv. Synth. Catal. 2011, 353 (13), 236976,  DOI: 10.1002/adsc.201100386
    33. 33
      Choi, J.-Y.; Black, R.; Lee, H.; Di Giovanni, J.; Murphy, R. C.; Ben Mamoun, C.; Voelker, D. R. An Improved and Highly Selective Fluorescence Assay for Measuring Phosphatidylserine Decarboxylase Activity. J. Biol. Chem. 2020, 295, 9211,  DOI: 10.1074/jbc.RA120.013421
    34. 34
      Leslie, A. K.; Li, D.; Koide, K. Amine-Promoted β-Elimination of a β-Aryloxy Aldehyde for Fluorogenic Chemodosimeters. J. Org. Chem. 2011, 76 (16), 68605,  DOI: 10.1021/jo200947e
    35. 35
      Rinde, E.; Troll, W. Colorimetric Assay for Aromatic Amines. Anal. Chem. 1976, 48 (3), 5424,  DOI: 10.1021/ac60367a036
    36. 36
      Spangenberg, B.; Poole, C. F.; Weins, C. Quantitative Thin-Layer Chromatography: A Practical Survey; Springer Science & Business Media, 2011.
    37. 37
      Ehmann, A. The Van Urk-Salkowski Reagent ─ a Sensitive and Specific Chromogenic Reagent for Silica Gel Thin-Layer Chromatographic Detection and Identification of Indole Derivatives. J. Chromatogr. A 1977, 132 (2), 26776,  DOI: 10.1016/S0021-9673(00)89300-0
    38. 38
      Jones, L. A.; Holmes, J. C.; Seligman, R. B. Spectrophotometric Studies of Some 2,4-Dinitrophenylhydrazones. Anal. Chem. 1956, 28 (2), 1918,  DOI: 10.1021/ac60110a013
    39. 39
      Zarzycki, P.; Bartoszuk, M.; Radziwon, A. Optimization of Tlc Detection by Phosphomolybdic Acid Staining for Robust Quantification of Cholesterol and Bile Acids. J. Planar Chromatogr. - Mod. TLC 2006, 19 (107), 527,  DOI: 10.1556/JPC.19.2006.1.9
    40. 40
      Bergmann, F. Colorimetric Determination of Amides as Hydroxamic Acids. Anal. Chem. 1952, 24 (8), 13679,  DOI: 10.1021/ac60068a033
    41. 41
      Gayathri, T. Stains for Developing TLC Plates. Int. J. Adv. Res. Sci. Eng. Technol. 2014, 5 (11), 7983
    42. 42
      Leeuwenkamp, O. R.; van Bennekom, W. P.; van der Mark, E. J.; Bult, A. Nitroprusside, Antihypertensive Drug and Analytical Reagent. Review of (Photo)stability, Pharmacology and Analytical Properties. Pharm. Weekbl. Sci. Ed. 1984, 6 (4), 129140,  DOI: 10.1007/BF01954040
    43. 43
      Lavis, L. D.; Raines, R. T. Bright Building Blocks for Chemical Biology. ACS Chem. Biol. 2014, 9 (4), 85566,  DOI: 10.1021/cb500078u
    44. 44
      Bowen, E. J. Fluorescence and Fluorescence Quenching. Q. Rev. Chem. Soc. 1947, 1 (1), 115,  DOI: 10.1039/qr9470100001
    45. 45
      Fu, Y.; Finney, N. S. Small-Molecule Fluorescent Probes and Their Design. RSC Adv. 2018, 8 (51), 2905161,  DOI: 10.1039/C8RA02297F
    46. 46
      Grimm, J. B.; Heckman, L. M.; Lavis, L. D. The Chemistry of Small-Molecule Fluorogenic Probes. Prog. Mol. Biol.Transl. Sci. 2013, 113 (1), 134
    47. 47
      Daly, B.; Ling, J.; De Silva, A. P. Current Developments in Fluorescent PET (Photoinduced Electron Transfer) Sensors and Switches. Chem. Soc. Rev. 2015, 44 (13), 420311,  DOI: 10.1039/C4CS00334A
    48. 48
      Algar, W. R.; Hildebrandt, N.; Vogel, S. S.; Medintz, I. L. Fret as a Biomolecular Research Tool ─ Understanding Its Potential While Avoiding Pitfalls. Nat. Methods 2019, 16 (9), 81529,  DOI: 10.1038/s41592-019-0530-8
    49. 49
      Misra, R.; Bhattacharyya, S. P. Intramolecular Charge Transfer: Theory and Applications; Wiley VCH, 2018 DOI: 10.1002/9783527801916
    50. 50
      Sedgwick, A. C.; Wu, L.; Han, H.-H.; Bull, S. D.; He, X.-P.; James, T. D.; Sessler, J. L.; Tang, B. Z.; Tian, H.; Yoon, J. Excited-State Intramolecular Proton-Transfer (ESIPT) Based Fluorescence Sensors and Imaging Agents. Chem. Soc. Rev. 2018, 47 (23), 884280,  DOI: 10.1039/C8CS00185E
    51. 51
      Hong, Y.; Lam, J. W.; Tang, B. Z. Aggregation-Induced Emission. Chem. Soc. Rev. 2011, 40 (11), 536188,  DOI: 10.1039/c1cs15113d
    52. 52
      Grabowski, Z. R.; Rotkiewicz, K.; Rettig, W. Structural Changes Accompanying Intramolecular Electron Transfer: Focus on Twisted Intramolecular Charge-Transfer States and Structures. Chem. Rev. 2003, 103 (10), 38994032,  DOI: 10.1021/cr940745l
    53. 53
      Tanner, P. A.; Zhou, L.; Duan, C.; Wong, K.-L. Misconceptions in Electronic Energy Transfer: Bridging the Gap between Chemistry and Physics. Chem. Soc. Rev. 2018, 47 (14), 523465,  DOI: 10.1039/C8CS00002F
    54. 54
      Jiang, X.; Wang, L.; Carroll, S. L.; Chen, J.; Wang, M. C.; Wang, J. Challenges and Opportunities for Small-Molecule Fluorescent Probes in Redox Biology Applications. Antioxid. Redox Signal. 2018, 29 (6), 51840,  DOI: 10.1089/ars.2017.7491
    55. 55
      Wu, L.; Sedgwick, A. C.; Sun, X.; Bull, S. D.; He, X. P.; James, T. D. Reaction-Based Fluorescent Probes for the Detection and Imaging of Reactive Oxygen, Nitrogen, and Sulfur Species. Acc. Chem. Res. 2019, 52 (9), 258297,  DOI: 10.1021/acs.accounts.9b00302
    56. 56
      Jiao, X.; Li, Y.; Niu, J.; Xie, X.; Wang, X.; Tang, B. Small-Molecule Fluorescent Probes for Imaging and Detection of Reactive Oxygen, Nitrogen, and Sulfur Species in Biological Systems. Anal. Chem. 2018, 90 (1), 53355,  DOI: 10.1021/acs.analchem.7b04234
    57. 57
      Yang, Y.; Zhao, Q.; Feng, W.; Li, F. Luminescent Chemodosimeters for Bioimaging. Chem. Rev. 2013, 113 (1), 192270,  DOI: 10.1021/cr2004103
    58. 58
      Chyan, W.; Raines, R. T. Enzyme-Activated Fluorogenic Probes for Live-Cell and in Vivo Imaging. ACS Chem. Biol. 2018, 13 (7), 181023,  DOI: 10.1021/acschembio.8b00371
    59. 59
      Lawrence, S. A. Amines: Synthesis, Properties and Applications; Cambridge University Press, 2004.
    60. 60
      Yin, Q.; Shi, Y.; Wang, J.; Zhang, X. Direct Catalytic Asymmetric Synthesis of α-Chiral Primary Amines. Chem. Soc. Rev. 2020, 49 (17), 614153,  DOI: 10.1039/C9CS00921C
    61. 61
      Cabré, A.; Verdaguer, X.; Riera, A. Recent Advances in the Enantioselective Synthesis of Chiral Amines Via Transition Metal-Catalyzed Asymmetric Hydrogenation. Chem. Rev. 2022, 122 (1), 269339,  DOI: 10.1021/acs.chemrev.1c00496
    62. 62
      Wu, Z.; Liu, C.; Zhang, Z.; Zheng, R.; Zheng, Y. Amidase as a Versatile Tool in Amide-Bond Cleavage: From Molecular Features to Biotechnological Applications. Biotechnol. Adv. 2020, 43, 107574,  DOI: 10.1016/j.biotechadv.2020.107574
    63. 63
      Breuer, M.; Ditrich, K.; Habicher, T.; Hauer, B.; Keßeler, M.; Stürmer, R.; Zelinski, T. Industrial Methods for the Production of Optically Active Intermediates. Angew. Chem., Int. Ed. 2004, 43 (7), 788824,  DOI: 10.1002/anie.200300599
    64. 64
      Balkenhohl, F.; Ditrich, K.; Hauer, B.; Ladner, W. Optisch Aktive Amine Durch Lipase-Katalysierte Methoxyacetylierung. J. Prakt. Chem. Chem. Ztg. 1997, 339 (1), 3814,  DOI: 10.1002/prac.19973390166
    65. 65
      Gomm, A.; O’Reilly, E. Transaminases for Chiral Amine Synthesis. Curr. Opin. Chem. Biol. 2018, 43, 10612,  DOI: 10.1016/j.cbpa.2017.12.007
    66. 66
      Abrahamson, M. J.; Vázquez-Figueroa, E.; Woodall, N. B.; Moore, J. C.; Bommarius, A. S. Development of an Amine Dehydrogenase for Synthesis of Chiral Amines. Angew. Chem., Int. Ed. 2012, 51 (16), 396972,  DOI: 10.1002/anie.201107813
    67. 67
      Li, B.-B.; Zhang, J.; Chen, F.-F.; Chen, Q.; Xu, J.-H.; Zheng, G.-W. Direct Reductive Amination of Ketones with Amines by Reductive Aminases. Green Synth. Catal. 2021, 2 (4), 3459,  DOI: 10.1016/j.gresc.2021.08.005
    68. 68
      Lenz, M.; Borlinghaus, N.; Weinmann, L.; Nestl, B. M. Recent Advances in Imine Reductase-Catalyzed Reactions. World J. Microbiol. Biotechnol. 2017, 33 (11), 199,  DOI: 10.1007/s11274-017-2365-8
    69. 69
      Gaweska, H.; Fitzpatrick, P. F. Structures and Mechanism of the Monoamine Oxidase Family. Biomol Concepts 2011, 2 (5), 36577,  DOI: 10.1515/BMC.2011.030
    70. 70
      Parmeggiani, F.; Weise, N. J.; Ahmed, S. T.; Turner, N. J. Synthetic and Therapeutic Applications of Ammonia-Lyases and Aminomutases. Chem. Rev. 2018, 118 (1), 73118,  DOI: 10.1021/acs.chemrev.6b00824
    71. 71
      Savile, C. K.; Janey, J. M.; Mundorff, E. C.; Moore, J. C.; Tam, S.; Jarvis, W. R.; Colbeck, J. C.; Krebber, A.; Fleitz, F. J.; Brands, J. Biocatalytic Asymmetric Synthesis of Chiral Amines from Ketones Applied to Sitagliptin Manufacture. Science 2010, 329 (5989), 3059,  DOI: 10.1126/science.1188934
    72. 72
      Weiß, M. S.; Pavlidis, I. V.; Spurr, P.; Hanlon, S. P.; Wirz, B.; Iding, H.; Bornscheuer, U. T. Protein-Engineering of an Amine Transaminase for the Stereoselective Synthesis of a Pharmaceutically Relevant Bicyclic Amine. Org. Biomol. Chem. 2016, 14 (43), 1024954,  DOI: 10.1039/C6OB02139E
    73. 73
      Ma, E. J.; Siirola, E.; Moore, C.; Kummer, A.; Stoeckli, M.; Faller, M.; Bouquet, C.; Eggimann, F.; Ligibel, M.; Huynh, D. Machine-Directed Evolution of an Imine Reductase for Activity and Stereoselectivity. ACS Catal. 2021, 11 (20), 1243345,  DOI: 10.1021/acscatal.1c02786
    74. 74
      Ao, Y.-F.; Hu, H.-J.; Zhao, C.-X.; Chen, P.; Huang, T.; Chen, H.; Wang, Q.-Q.; Wang, D.-X.; Wang, M.-X. Reversal and Amplification of the Enantioselectivity of Biocatalytic Desymmetrization toward meso Heterocyclic Dicarboxamides Enabled by Rational Engineering of Amidase. ACS Catal. 2021, 11 (12), 69007,  DOI: 10.1021/acscatal.1c01220
    75. 75
      Wang, P. F.; Yep, A.; Kenyon, G. L.; McLeish, M. J. Using Directed Evolution to Probe the Substrate Specificity of Mandelamide Hydrolase. Protein Eng. Des. Sel. 2009, 22 (2), 10310,  DOI: 10.1093/protein/gzn073
    76. 76
      Thooft, A. M.; Cassaidy, K.; VanVeller, B. A Small Push-Pull Fluorophore for Turn-On Fluorescence. J. Org. Chem. 2017, 82 (17), 88427,  DOI: 10.1021/acs.joc.7b00939
    77. 77
      Annenkov, V. V.; Verkhozina, O. N.; Shishlyannikova, T. A.; Danilovtseva, E. N. Application of 4-Chloro-7-Nitrobenzo-2-Oxa-1,3-Diazole in Analysis: Fluorescent Dyes and Unexpected Reaction with Tertiary Amines. Anal. Biochem. 2015, 486, 513,  DOI: 10.1016/j.ab.2015.06.025
    78. 78
      Sanger, F. The Free Amino Groups of Insulin. Biochem. J. 1945, 39 (5), 50715,  DOI: 10.1042/bj0390507
    79. 79
      Ghosh, P. B.; Whitehouse, M. W. 7-Chloro-4-Nitrobenzo-2-Oxa-1,3-Diazole: A New Fluorigenic Reagent for Amino Acids and Other Amines. Biochem. J. 1968, 108 (1), 1556,  DOI: 10.1042/bj1080155
    80. 80
      Watanabe, Y.; Imai, K. High-Performance Liquid Chromatography and Sensitive Detection of Amino Acids Derivatized with 7-Fluoro-4-Nitrobenzo-2-Oxa-1,3-Diazole. Anal. Biochem. 1981, 116 (2), 4712,  DOI: 10.1016/0003-2697(81)90390-0
    81. 81
      Aly, H.; El-Shafie, A. S.; El-Azazy, M. Utilization of 7-Chloro-4-Nitrobenzo-2-Oxa-1,3-Diazole (NBD-Cl) for Spectrochemical Determination of L-Ornithine: A Multivariate Optimization-Assisted Approach. RSC Adv. 2019, 9 (38), 2210615,  DOI: 10.1039/C9RA03311D
    82. 82
      Walash, M. I.; Metwally, M. E. S.; Eid, M.; El-Shaheny, R. N. Validated Spectrophotometric Methods for Determination of Alendronate Sodium in Tablets through Nucleophilic Aromatic Substitution Reactions. Chem. Cent. J. 2012, 6, 25,  DOI: 10.1186/1752-153X-6-25
    83. 83
      Benson, S.; Fernandez, A.; Barth, N. D.; de Moliner, F.; Horrocks, M. H.; Herrington, C. S.; Abad, J. L.; Delgado, A.; Kelly, L.; Chang, Z. SCOTfluors: Small, Conjugatable, Orthogonal, and Tunable Fluorophores for in Vivo Imaging of Cell Metabolism. Angew. Chem., Int. Ed. 2019, 58 (21), 69115,  DOI: 10.1002/anie.201900465
    84. 84
      Satake, K.; Okuyama, T.; Ohashi, M.; Shinoda, T. The Spectrophotometric Determination of Amine, Amino Acid and Peptide with 2,4,6-Trinitrobenzene 1-Sulfonic Acid. J. Biochem. 1960, 47 (5), 65460,  DOI: 10.1093/oxfordjournals.jbchem.a127107
    85. 85
      Means, G. E.; Congdon, W. I.; Bender, M. L. Reactions of 2,4,6-Trinitrobenzenesulfonate Ion with Amines and Hydroxide Ion. Biochemistry 1972, 11 (19), 356471,  DOI: 10.1021/bi00769a011
    86. 86
      Snyder, S. L.; Sobocinski, P. Z. An Improved 2,4,6-Trinitrobenzenesulfonic Acid Method for the Determination of Amines. Anal. Biochem. 1975, 64 (1), 2848,  DOI: 10.1016/0003-2697(75)90431-5
    87. 87
      Qi, X.-Y.; Keyhani, N. O.; Lee, Y. C. Spectrophotometric Determination of Hydrazine, Hydrazides, and Their Mixtures with Trinitrobenzenesulfonic Acid. Anal. Biochem. 1988, 175 (1), 13944,  DOI: 10.1016/0003-2697(88)90371-5
    88. 88
      Morçöl, T.; Subramanian, A.; Velander, W. H. Dot-Blot Analysis of the Degree of Covalent Modification of Proteins and Antibodies at Amino Groups. J. Immunol. Methods 1997, 203 (1), 4553,  DOI: 10.1016/S0022-1759(97)00013-6
    89. 89
      Oliverio, R.; Liberelle, B.; Murschel, F.; Garcia-Ac, A.; Banquy, X.; De Crescenzo, G. Versatile and High-Throughput Strategy for the Quantification of Proteins Bound to Nanoparticles. ACS Appl. Nano Mater. 2020, 3 (10), 10497507,  DOI: 10.1021/acsanm.0c02414
    90. 90
      Kim, D.; Cho, S. W.; Jun, Y. W.; Ahn, K. H. Fluorescent Labeling of Lysine Residues in Protein Using 8-Thiomethyl-BODIPY. Bull. Korean Chem. Soc. 2017, 38 (9), 9956,  DOI: 10.1002/bkcs.11213
    91. 91
      Bell, P. J. L.; Karuso, P. Epicocconone, a Novel Fluorescent Compound from the Fungus Epicoccum Nigrum. J. Am. Chem. Soc. 2003, 125 (31), 93045,  DOI: 10.1021/ja035496+
    92. 92
      Coghlan, D. R.; Mackintosh, J. A.; Karuso, P. Mechanism of Reversible Fluorescent Staining of Protein with Epicocconone. Org. Lett. 2005, 7 (12), 24014,  DOI: 10.1021/ol050665b
    93. 93
      Peixoto, P. A.; Boulangé, A.; Ball, M.; Naudin, B.; Alle, T.; Cosette, P.; Karuso, P.; Franck, X. Design and Synthesis of Epicocconone Analogues with Improved Fluorescence Properties. J. Am. Chem. Soc. 2014, 136 (43), 1524856,  DOI: 10.1021/ja506914p
    94. 94
      Chatterjee, S.; Karuso, P.; Boulangé, A.; Franck, X.; Datta, A. Excited State Dynamics of Brightly Fluorescent Second Generation Epicocconone Analogues. J. Phys. Chem. B 2015, 119 (20), 6295303,  DOI: 10.1021/acs.jpcb.5b02190
    95. 95
      Castell, J. V.; Cervera, M.; Marco, R. A Convenient Micromethod for the Assay of Primary Amines and Proteins with Fluorescamine. A Reexamination of the Conditions of Reaction. Anal. Biochem. 1979, 99 (2), 37991,  DOI: 10.1016/S0003-2697(79)80022-6
    96. 96
      Funk, G. M.; Hunt, C. E.; Epps, D. E.; Brown, P. K. Use of a Rapid and Highly Sensitive Fluorescamine-Based Procedure for the Assay of Plasma Lipoproteins. J. Lipid Res. 1986, 27, 792,  DOI: 10.1016/S0022-2275(20)38803-9
    97. 97
      Udenfriend, S.; Stein, S.; Böhlen, P.; Dairman, W.; Leimgruber, W.; Weigele, M. Fluorescamine: A Reagent for Assay of Amino Acids, Peptides, Proteins, and Primary Amines in the Picomole Range. Science 1972, 178 (4063), 871,  DOI: 10.1126/science.178.4063.871
    98. 98
      Murugayah, S. A.; Warring, S. L.; Gerth, M. L. Optimisation of a High-Throughput Fluorescamine Assay for Detection of N-Acyl-L-Homoserine Lactone Acylase Activity. Anal. Biochem. 2019, 566, 102,  DOI: 10.1016/j.ab.2018.10.029
    99. 99
      Ashby, J.; Duan, Y.; Ligans, E.; Tamsi, M.; Zhong, W. High-Throughput Profiling of Nanoparticle-Protein Interactions by Fluorescamine Labeling. Anal. Chem. 2015, 87 (4), 22139,  DOI: 10.1021/ac5036814
    100. 100
      Elbashir, A. A.; Ahmed, A. A.; Ali Ahmed, S. M.; Aboul-Enein, H. Y. 1,2-Naphthoquinone-4-Sulphonic Acid Sodium Salt (NQS) as an Analytical Reagent for the Determination of Pharmaceutical Amine by Spectrophotometry. Appl. Spectrosc. Rev. 2012, 47 (3), 21932,  DOI: 10.1080/05704928.2011.639107
    101. 101
      Rosenblatt, D. H.; Hlinka, P.; Epstein, J. Use of 1,2-Naphthoquinone-4-Sulfonate for Estimation of Ethylenimine and Primary Amines. Anal. Chem. 1955, 27 (8), 12903,  DOI: 10.1021/ac60104a024
    102. 102
      Darwish, I. A.; Abdine, H. H.; Amer, S. M.; Al-Rayes, L. I. Simple Spectrophotometric Method for Determination of Paroxetine in Tablets Using 1,2-Naphthoquinone-4-Sulphonate as a Chromogenic Reagent. Int. J. Anal. Chem. 2009, 2009, 237601,  DOI: 10.1155/2009/237601
    103. 103
      Longstreet, A. R.; Jo, M.; Chandler, R. R.; Hanson, K.; Zhan, N.; Hrudka, J. J.; Mattoussi, H.; Shatruk, M.; McQuade, D. T. Ylidenemalononitrile Enamines as Fluorescent “Turn-On” Indicators for Primary Amines. J. Am. Chem. Soc. 2014, 136 (44), 154936,  DOI: 10.1021/ja509058u
    104. 104
      Sathiskumar, U.; Easwaramoorthi, S. Red-Emitting Ratiometric Fluorescence Chemodosimeter for the Discriminative Detection of Aromatic and Aliphatic Amines. ChemistrySelect 2019, 4 (25), 748694,  DOI: 10.1002/slct.201901254
    105. 105
      Turiák, G.; Volicer, L. Stability of o-Phthalaldehyde─Sulfite Derivatives of Amino Acids and Their Methyl Esters: Electrochemical and Chromatographic Properties. J. Chromatogr. A 1994, 668 (2), 3239,  DOI: 10.1016/0021-9673(94)80121-5
    106. 106
      De Montigny, P.; Stobaugh, J. F.; Givens, R. S.; Carlson, R. G.; Srinivasachar, K.; Sternson, L. A.; Higuchi, T. Naphthalene-2,3-Dicarboxyaldehyde/Cyanide Ion: A Rationally Designed Fluorogenic Reagent for Primary Amines. Anal. Chem. 1987, 59 (8), 1096101,  DOI: 10.1021/ac00135a007
    107. 107
      Mroz, E. A.; Roman, R. J.; Lechene, C. Fluorescence Assay for Picomole Quantities of Ammonia. Kidney Int. 1982, 21 (3), 5247,  DOI: 10.1038/ki.1982.56
    108. 108
      Carlson, R. G.; Srinivasachar, K.; Givens, R. S.; Matuszewski, B. K. New Derivatizing Agents for Amino Acids and Peptides. 1. Facile Synthesis of N-Substituted 1-Cyanobenz[f]Isoindoles and Their Spectroscopic Properties. J. Org. Chem. 1986, 51 (21), 397883,  DOI: 10.1021/jo00371a013
    109. 109
      Bantan-Polak, T.; Kassai, M.; Grant, K. B. A Comparison of Fluorescamine and Naphthalene-2,3-Dicarboxaldehyde Fluorogenic Reagents for Microplate-Based Detection of Amino Acids. Anal. Biochem. 2001, 297 (2), 12836,  DOI: 10.1006/abio.2001.5338
    110. 110
      Hapuarachchi, S.; Aspinwall, C. A. Design, Characterization, and Utilization of a Fast Fluorescence Derivatization Reaction Utilizing o-Phthaldialdehyde Coupled with Fluorescent Thiols. Electrophoresis 2007, 28 (7), 11006,  DOI: 10.1002/elps.200600567
    111. 111
      Zhang, L.-Y.; Liu, Y.-M.; Wang, Z.-L.; Cheng, J.-K. Capillary Zone Electrophoresis with Pre-Column NDA Derivatization and Amperometric Detection for the Analysis of Four Aliphatic Diamines. Anal. Chim. Acta 2004, 508 (2), 1415,  DOI: 10.1016/j.aca.2003.11.075
    112. 112
      Fekete, A.; Lahaniatis, M.; Lintelmann, J.; Schmitt-Kopplin, P. Determination of Aliphatic Low-Molecular-Weight and Biogenic Amines by Capillary Zone Electrophoresis. In Capillary Electrophoresis: Methods and Protocols; Humana Press: Totowa, NJ, 2008; pp 6591. DOI: 10.1007/978-1-59745-376-9_4
    113. 113
      Böhmer, A.; Jordan, J.; Tsikas, D. High-Performance Liquid Chromatography Ultraviolet Assay for Human Erythrocytic Catalase Activity by Measuring Glutathione as o-Phthalaldehyde Derivative. Anal. Biochem. 2011, 410 (2), 296303,  DOI: 10.1016/j.ab.2010.11.026
    114. 114
      Black, G. W.; Brown, N. L.; Perry, J. J. B.; Randall, P. D.; Turnbull, G.; Zhang, M. A High-Throughput Screening Method for Determining the Substrate Scope of Nitrilases. Chem. Commun. 2015, 51 (13), 26602,  DOI: 10.1039/C4CC06021K
    115. 115
      Mann, S.; Eveleigh, L.; Lequin, O.; Ploux, O. A Microplate Fluorescence Assay for DAPA Aminotransferase by Detection of the Vicinal Diamine 7,8-Diaminopelargonic Acid. Anal. Biochem. 2013, 432 (2), 906,  DOI: 10.1016/j.ab.2012.09.038
    116. 116
      Kalkhof, S.; Sinz, A. Chances and Pitfalls of Chemical Cross-Linking with Amine-Reactive N-Hydroxysuccinimide Esters. Anal. Bioanal. Chem. 2008, 392 (1), 30512,  DOI: 10.1007/s00216-008-2231-5
    117. 117
      Gayo, L. M.; Suto, M. J. Use of Pentafluorophenyl Esters for One-Pot Protection/Activation of Amino and Thiol Carboxylic Acids. Tetrahedron Lett. 1996, 37 (28), 49158,  DOI: 10.1016/0040-4039(96)00985-9
    118. 118
      Fang, L.; Demee, M.; Sierra, T.; Kshirsagar, T.; Celebi, A. A.; Yan, B. Kinetics Study of Amine Cleavage Reactions of Various Resin-Bound Thiophenol Esters from Marshall Linker. J. Comb. Chem. 2002, 4 (4), 3628,  DOI: 10.1021/cc020010r
    119. 119
      Lavis, L. D. Teaching Old Dyes New Tricks: Biological Probes Built from Fluoresceins and Rhodamines. Annu. Rev. Biochem. 2017, 86, 82543,  DOI: 10.1146/annurev-biochem-061516-044839
    120. 120
      Pereira, A.; Martins, S.; Caldeira, A. T. Phytochemicals in Human Health, Coumarins as Fluorescent Labels of Biomolecules; IntechOpen, 2019.
    121. 121
      Boens, N.; Leen, V.; Dehaen, W. Fluorescent Indicators Based on BODIPY. Chem. Soc. Rev. 2012, 41 (3), 113072,  DOI: 10.1039/C1CS15132K
    122. 122
      Han, S.-Y.; Kim, Y.-A. Recent Development of Peptide Coupling Reagents in Organic Synthesis. Tetrahedron 2004, 60 (11), 244767,  DOI: 10.1016/j.tet.2004.01.020
    123. 123
      Karongo, R.; Ge, M.; Horak, J.; Gross, H.; Kohout, M.; Lindner, W.; Lämmerhofer, M. Rapid Enantioselective Amino Acid Analysis by Ultra-High Performance Liquid Chromatography-Mass Spectrometry Combining 6-Aminoquinolyl-N-Hydroxysuccinimidyl Carbamate Derivatization with Core-Shell Quinine Carbamate Anion Exchanger Separation. J. Chromatogr. Open 2021, 1, 100004,  DOI: 10.1016/j.jcoa.2021.100004
    124. 124
      Cao, L.; Wang, H.; Zhang, H. Analytical Potential of 6-Oxy-(N-Succinimidyl Acetate)-9-(2’-Methoxycarbonyl) Fluorescein for the Determination of Amino Compounds by Capillary Electrophoresis with Laser-Induced Fluorescence Detection. Electrophoresis 2005, 26 (10), 195462,  DOI: 10.1002/elps.200410227
    125. 125
      Xia, L.-J.; Guo, X.-F.; Ji, Y.; Chen, L.; Wang, H. A Long-Wavelength Fluorescent Probe for Amino Compounds and Its Application in the Determination of Aliphatic Amines. Anal. Methods 2018, 10 (26), 318896,  DOI: 10.1039/C8AY00706C
    126. 126
      Gao, P. F.; Guo, X. F.; Wang, H.; Zhang, H. S. Determination of Trace Biogenic Amines with 1,3,5,7-Tetramethyl-8-(N-Hydroxysuccinimidyl Butyric Ester)-Difluoroboradiaza-s-Indacene Derivatization Using High-Performance Liquid Chromatography and Fluorescence Detection. J. Sep. Sci. 2011, 34 (12), 138390,  DOI: 10.1002/jssc.201100120
    127. 127
      Mallick, S.; Chandra, F.; Koner, A. L. A Ratiometric Fluorescent Probe for Detection of Biogenic Primary Amines with Nanomolar Sensitivity. Analyst 2016, 141 (3), 82731,  DOI: 10.1039/C5AN01911G
    128. 128
      Vázquez, M. E.; Blanco, J. B.; Imperiali, B. Photophysics and Biological Applications of the Environment-Sensitive Fluorophore 6-N,N-Dimethylamino-2,3-Naphthalimide. J. Am. Chem. Soc. 2005, 127 (4), 13006,  DOI: 10.1021/ja0449168
    129. 129
      Anumula, K. R.; Schulz, R. P.; Back, N. Fluorescent N-Methylanthranilyl (Mantyl) Tag for Peptides: Its Application in Subpicomole Determination of Kinins. Peptides 1992, 13 (4), 6639,  DOI: 10.1016/0196-9781(92)90170-8
    130. 130
      Fessler, A.; Garmon, C.; Heavey, T.; Fowler, A.; Ogle, C. Water-Soluble and UV Traceable Isatoic Anhydride-Based Reagents for Bioconjugation. Org. Biomol. Chem. 2017, 15 (45), 9599602,  DOI: 10.1039/C7OB02377D
    131. 131
      Jeon, S.; Kim, T.-I.; Jin, H.; Lee, U.; Bae, J.; Bouffard, J.; Kim, Y. Amine-Reactive Activated Esters of meso-CarboxyBODIPY: Fluorogenic Assays and Labeling of Amines, Amino Acids, and Proteins. J. Am. Chem. Soc. 2020, 142 (20), 92319,  DOI: 10.1021/jacs.9b13982
    132. 132
      Lee, U.; Kim, T.-I.; Jeon, S.; Luo, Y.; Cho, S.; Bae, J.; Kim, Y. Native Chemical Ligation-Based Fluorescent Probes for Cysteine and Aminopeptidase N Using meso-Thioester-BODIPY. Eur. J. Chem. 2021, 27 (49), 1254551,  DOI: 10.1002/chem.202101990
    133. 133
      Esnal, I.; Bañuelos, J.; López Arbeloa, I.; Costela, A.; Garcia-Moreno, I.; Garzón, M.; Agarrabeitia, A. R.; José Ortiz, M. Nitro and Amino BODIPYs: Crucial Substituents to Modulate Their Photonic Behavior. RSC Adv. 2013, 3 (5), 154756,  DOI: 10.1039/C2RA22916A
    134. 134
      Liu, E. Y.; Jung, S.; Weitz, D. A.; Yi, H.; Choi, C.-H. High-Throughput Double Emulsion-Based Microfluidic Production of Hydrogel Microspheres with Tunable Chemical Functionalities toward Biomolecular Conjugation. Lab Chip 2018, 18 (2), 32334,  DOI: 10.1039/C7LC01088E
    135. 135
      Katritzky, A. R.; Marson, C. M. Pyrylium Mediated Transformations of Primary Amino Groups into Other Functional Groups. New Synthetic Methods (41). Angew. Chem., Int. Ed. Engl. 1984, 23 (6), 4209,  DOI: 10.1002/anie.198404201
    136. 136
      Wetzl, B. K.; Yarmoluk, S. M.; Craig, D. B.; Wolfbeis, O. S. Chameleon Labels for Staining and Quantifying Proteins. Angew. Chem., Int. Ed. Engl. 2004, 43 (40), 54002,  DOI: 10.1002/anie.200460508
    137. 137
      Azab, H. A.; El-Korashy, S. A.; Anwar, Z. M.; Khairy, G. M.; Steiner, M.-S.; Duerkop, A. High-Throughput Sensing Microtiter Plate for Determination of Biogenic Amines in Seafood Using Fluorescence or Eye-Vision. Analyst 2011, 136 (21), 44929,  DOI: 10.1039/c1an15049a
    138. 138
      Azab, H. A.; El-Korashy, S. A.; Anwar, Z. M.; Khairy, G. M.; Duerkop, A. Reactivity of a Luminescent “Off-On” Pyrylium Dye toward Various Classes of Amines and Its Use in a Fluorescence Sensor Microtiter Plate for Environmental Samples. J. Photochem. Photobiol. A: Chem. 2012, 243, 416,  DOI: 10.1016/j.jphotochem.2012.05.029
    139. 139
      Bayer, M.; König, S. Pyrylium-Based Dye and Charge Tagging in Proteomics. Electrophoresis 2016, 37 (22), 29538,  DOI: 10.1002/elps.201600318
    140. 140
      Turner, E. H.; Dickerson, J. A.; Ramsay, L. M.; Swearingen, K. E.; Wojcik, R.; Dovichi, N. J. Reaction of Fluorogenic Reagents with Proteins: III. Spectroscopic and Electrophoretic Behavior of Proteins Labeled with Chromeo P503. J. Chromatogr. A 2008, 1194 (2), 2536,  DOI: 10.1016/j.chroma.2008.04.046
    141. 141
      Wolfbeis, O. S. Fluorescent Chameleon Labels for Bioconjugation and Imaging of Proteins, Nucleic Acids, Biogenic Amines and Surface Amino Groups. A Review. Methods Appl. Fluoresc. 2021, 9 (4), 042001,  DOI: 10.1088/2050-6120/ac1a0a
    142. 142
      Höfelschweiger, B. K., The Pyrylium Dyes: a New Class of Biolabels. Synthesis, Spectroscopy, and Application as Labels and in General Protein Assay , 2005.
    143. 143
      Walter, W.; Bode, K. D. Syntheses of Thiocarbamates. Angew. Chem., Int. Ed. Engl. 1967, 6 (4), 28193,  DOI: 10.1002/anie.196702811
    144. 144
      Maddani, M. R.; Prabhu, K. R. A Concise Synthesis of Substituted Thiourea Derivatives in Aqueous Medium. J. Org. Chem. 2010, 75 (7), 232732,  DOI: 10.1021/jo1001593
    145. 145
      Reetz, M. T.; Kühling, K. M.; Deege, A.; Hinrichs, H.; Belder, D. Super-High-Throughput Screening of Enantioselective Catalysts by Using Capillary Array Electrophoresis. Angew. Chem., Int. Ed. 2000, 39 (21), 38913,  DOI: 10.1002/1521-3773(20001103)39:21<3891::AID-ANIE3891>3.0.CO;2-1
    146. 146
      Dabek-Zlotorzynska, E.; Maruszak, W. Determination of Dimethylamine and Other Low-Molecular-Mass Amines Using Capillary Electrophoresis with Laser-Induced Fluorescence Detection. J. Chromatogr. B Biomed. Appl. 1998, 714 (1), 7785,  DOI: 10.1016/S0378-4347(98)00070-X
    147. 147
      Morales, A. R.; Schafer-Hales, K. J.; Marcus, A. I.; Belfield, K. D. Amine-Reactive Fluorene Probes: Synthesis, Optical Characterization, Bioconjugation, and Two-Photon Fluorescence Imaging. Bioconj. Chem. 2008, 19 (12), 255967,  DOI: 10.1021/bc800415t
    148. 148
      Planas, O.; Gallavardin, T.; Nonell, S. A Novel Fluoro-Chromogenic Click Reaction for the Labelling of Proteins and Nanoparticles with near-IR Theranostic Agents. Chem. Commun. 2015, 51 (26), 55869,  DOI: 10.1039/C4CC09070E
    149. 149
      Wang, R.; Yang, W.-j.; Yue, L.; Pan, W.; Zeng, H.-y. DDQ-Promoted C-S Bond Formation: Synthesis of 2-Aminobenzothiazole Derivatives under Transition-Metal-, Ligand-, and Base-Free Conditions. Synlett 2012, 23 (11), 16438,  DOI: 10.1055/s-0031-1291159
    150. 150
      Guo, Y.-J.; Tang, R.-Y.; Zhong, P.; Li, J.-H. Copper-Catalyzed Tandem Reactions of 2-Halobenzenamines with Isothiocyanates under Ligand- and Base-Free Conditions. Tetrahedron Lett. 2010, 51 (4), 64952,  DOI: 10.1016/j.tetlet.2009.11.086
    151. 151
      Sutton, J. M.; Clarke, O. J.; Fernandez, N.; Boyle, R. W. Porphyrin, Chlorin, and Bacteriochlorin Isothiocyanates: Useful Reagents for the Synthesis of Photoactive Bioconjugates. Bioconj. Chem. 2002, 13 (2), 24963,  DOI: 10.1021/bc015547x
    152. 152
      Feuster, E. K.; Glass, T. E. Detection of Amines and Unprotected Amino Acids in Aqueous Conditions by Formation of Highly Fluorescent Iminium Ions. J. Am. Chem. Soc. 2003, 125 (52), 161745,  DOI: 10.1021/ja036434m
    153. 153
      Patze, C.; Broedner, K.; Rominger, F.; Trapp, O.; Bunz, U. H. F. Aldehyde Cruciforms: Dosimeters for Primary and Secondary Amines. Eur. J. Chem. 2011, 17 (49), 137205,  DOI: 10.1002/chem.201101871
    154. 154
      Kumpf, J.; Bunz, U. H. F. Aldehyde-Appended Distyrylbenzenes: Amine Recognition in Water. Eur. J. Chem. 2012, 18 (29), 89214,  DOI: 10.1002/chem.201200930
    155. 155
      Mohr, G. J.; Demuth, C.; Spichiger-Keller, U. E. Application of Chromogenic and Fluorogenic Reactands in the Optical Sensing of Dissolved Aliphatic Amines. Anal. Chem. 1998, 70 (18), 386873,  DOI: 10.1021/ac980279q
    156. 156
      Mohr, G. J. Tailoring the Sensitivity and Spectral Properties of a Chromoreactand for the Detection of Amines and Alcohols. Anal. Chim. Acta 2004, 508 (2), 2337,  DOI: 10.1016/j.aca.2003.12.005
    157. 157
      Mohr, G. J. Chromo- and Fluororeactands: Indicators for Detection of Neutral Analytes by Using Reversible Covalent-Bond Chemistry. Eur. J. Chem. 2004, 10 (5), 108290,  DOI: 10.1002/chem.200305524
    158. 158
      Mohr, G. J.; Grummt, U.-W. Comparison of Trifluoroacetyl Monostyryl and Distyryl Dyes: Effects of Chromophore Elongation on the Spectral Properties and Chemical Reactivity. J. Fluoresc. 2006, 16 (2), 18590,  DOI: 10.1007/s10895-005-0047-7
    159. 159
      Mertz, E.; Beil, J. B.; Zimmerman, S. C. Kinetics and Thermodynamics of Amine and Diamine Signaling by a Trifluoroacetyl Azobenzene Reporter Group. Org. Lett. 2003, 5 (17), 312730,  DOI: 10.1021/ol0351605
    160. 160
      Körsten, S.; Mohr, G. J. Star-Shaped Tripodal Chemosensors for the Detection of Aliphatic Amines. Eur. J. Chem. 2011, 17 (3), 96975,  DOI: 10.1002/chem.201000787
    161. 161
      Xu, Y.; Yu, S.; Wang, Y.; Hu, L.; Zhao, F.; Chen, X.; Li, Y.; Yu, X.; Pu, L. Ratiometric Fluorescence Sensors for 1,2-Diamines Based on Trifluoromethyl Ketones. Eur. J. Org. Chem. 2016, 2016 (35), 586875,  DOI: 10.1002/ejoc.201601157
    162. 162
      Ali, M. F. B.; Kishikawa, N.; Ohyama, K.; Mohamed, H. A.-M.; Abdel-Wadood, H. M.; Mahmoud, A. M.; Imazato, T.; Ueki, Y.; Wada, M.; Kuroda, N. Chromatographic Determination of Low-Molecular Mass Unsaturated Aliphatic Aldehydes with Peroxyoxalate Chemiluminescence Detection after Fluorescence Labeling with 4-(N,N-Dimethylaminosulfonyl)-7-Hydrazino-2,1,3-Benzoxadiazole. J. Chromatogr. B 2014, 953–954, 14752,  DOI: 10.1016/j.jchromb.2014.02.009
    163. 163
      Xu, Y.; Yu, S.; Chen, Q.; Chen, X.; Li, Y.; Yu, X.; Pu, L. Fluorescent Recognition of 1,2-Diamines by a 1,1′-Binaphthyl-Based Trifluoromethyl Ketone. Eur. J. Chem. 2016, 22 (34), 120617,  DOI: 10.1002/chem.201601540
    164. 164
      Collados, J. F.; Solà, R.; Harutyunyan, S. R.; Maciá, B. Catalytic Synthesis of Enantiopure Chiral Alcohols Via Addition of Grignard Reagents to Carbonyl Compounds. ACS Catal. 2016, 6 (3), 195270,  DOI: 10.1021/acscatal.5b02832
    165. 165
      Malic, N.; Moorhoff, C.; Sage, V.; Saylik, D.; Teoh, E.; Scott, J. L.; Strauss, C. R. Toward Preparative Resolution of Chiral Alcohols by an Organic Chemical Method. New J. Chem. 2010, 34 (3), 398402,  DOI: 10.1039/b9nj00768g
    166. 166
      Rong, J.; Pellegrini, T.; Harutyunyan, S. R. Synthesis of Chiral Tertiary Alcohols by CuI-Catalyzed Enantioselective Addition of Organomagnesium Reagents to Ketones. Eur. J. Chem. 2016, 22 (11), 355870,  DOI: 10.1002/chem.201503412
    167. 167
      Chen, B.-S.; Ribeiro de Souza, F. Z. Enzymatic Synthesis of Enantiopure Alcohols: Current State and Perspectives. RSC Adv. 2019, 9 (4), 210215,  DOI: 10.1039/C8RA09004A
    168. 168
      Kourist, R.; Dominguez de Maria, P.; Bornscheuer, U. T. Enzymatic Synthesis of Optically Active Tertiary Alcohols: Expanding the Biocatalysis Toolbox. ChemBioChem. 2008, 9 (4), 4918,  DOI: 10.1002/cbic.200700688
    169. 169
      Jiang, H.; Su, X.; Zhang, Y.; Zhou, J.; Fang, D.; Wang, X. Unexpected Thiols Triggering Photoluminescent Enhancement of Cytidine Stabilized Au Nanoclusters for Sensitive Assays of Glutathione Reductase and Its Inhibitors Screening. Anal. Chem. 2016, 88 (9), 476671,  DOI: 10.1021/acs.analchem.6b00112
    170. 170
      Andexer, J.; von Langermann, J.; Mell, A.; Bocola, M.; Kragl, U.; Eggert, T.; Pohl, M. An R-Selective Hydroxynitrile Lyase from Arabidopsis Thaliana with an α/β-Hydrolase Fold. Angew. Chem., Int. Ed. 2007, 46 (45), 867981,  DOI: 10.1002/anie.200701455
    171. 171
      Zheng, Y.-G.; Yin, H.-H.; Yu, D.-F.; Chen, X.; Tang, X.-L.; Zhang, X.-J.; Xue, Y.-P.; Wang, Y.-J.; Liu, Z.-Q. Recent Advances in Biotechnological Applications of Alcohol Dehydrogenases. Appl. Microbiol. Biotechnol. 2017, 101 (3), 9871001,  DOI: 10.1007/s00253-016-8083-6
    172. 172
      Huisman, G. W.; Liang, J.; Krebber, A. Practical Chiral Alcohol Manufacture Using Ketoreductases. Curr. Opin. Chem. Biol. 2010, 14 (2), 1229,  DOI: 10.1016/j.cbpa.2009.12.003
    173. 173
      Iding, H.; Siegert, P.; Mesch, K.; Pohl, M. Application of α-Keto Acid Decarboxylases in Biotransformations. Biochim. Biophys. Acta 1998, 1385 (2), 30722,  DOI: 10.1016/S0167-4838(98)00076-4
    174. 174
      Lee, S.-H.; Yeom, S.-J.; Kim, S.-E.; Oh, D.-K. Development of Aldolase-Based Catalysts for the Synthesis of Organic Chemicals. Trends Biotechnol. 2022, 40, 306,  DOI: 10.1016/j.tibtech.2021.08.001
    175. 175
      Bracco, P.; Busch, H.; von Langermann, J.; Hanefeld, U. Enantioselective Synthesis of Cyanohydrins Catalysed by Hydroxynitrile Lyases - a Review. Org. Biomol. Chem. 2016, 14 (27), 637589,  DOI: 10.1039/C6OB00934D
    176. 176
      Chen, B.-S.; Otten, L. G.; Hanefeld, U. Stereochemistry of Enzymatic Water Addition to C=C Bonds. Biotechnol. Adv. 2015, 33 (5), 52646,  DOI: 10.1016/j.biotechadv.2015.01.007
    177. 177
      Archelas, A.; Furstoss, R. Synthetic Applications of Epoxide Hydrolases. Curr. Opin. Chem. Biol. 2001, 5 (2), 1129,  DOI: 10.1016/S1367-5931(00)00179-4
    178. 178
      Freakley, S. J.; Kochius, S.; van Marwijk, J.; Fenner, C.; Lewis, R. J.; Baldenius, K.; Marais, S. S.; Opperman, D. J.; Harrison, S. T. L.; Alcalde, M.; Smit, M. S.; Hutchings, G. J. A Chemo-Enzymatic Oxidation Cascade to Activate C-H Bonds with in Situ Generated H2O2. Nat. Commun. 2019, 10 (1), 4178,  DOI: 10.1038/s41467-019-12120-w
    179. 179
      Wang, Y.; Xiang, Q.; Zhou, Q.; Xu, J.; Pei, D. Mini Review: Advances in 2-Haloacid Dehalogenases. Front. Microbiol. 2021, 12, 758886,  DOI: 10.3389/fmicb.2021.758886
    180. 180
      Koudelakova, T.; Bidmanova, S.; Dvorak, P.; Pavelka, A.; Chaloupkova, R.; Prokop, Z.; Damborsky, J. Haloalkane Dehalogenases: Biotechnological Applications. Biotechnol. J. 2013, 8 (1), 3245,  DOI: 10.1002/biot.201100486
    181. 181
      Schallmey, A.; Schallmey, M. Recent Advances on Halohydrin Dehalogenases-from Enzyme Identification to Novel Biocatalytic Applications. Appl. Microbiol. Biotechnol. 2016, 100 (18), 782739,  DOI: 10.1007/s00253-016-7750-y
    182. 182
      Belafriekh, A.; Secundo, F.; Serra, S.; Djeghaba, Z. Enantioselective Enzymatic Resolution of Racemic Alcohols by Lipases in Green Organic Solvents. Tetrahedron: Asymmetry 2017, 28 (3), 4738,  DOI: 10.1016/j.tetasy.2017.02.004
    183. 183
      Bustos-Jaimes, I.; Hummel, W.; Eggert, T.; Bogo, E.; Puls, M.; Weckbecker, A.; Jaeger, K.-E. A High-Throughput Screening Method for Chiral Alcohols and Its Application to Determine Enantioselectivity of Lipases and Esterases. ChemCatChem. 2009, 1 (4), 4458,  DOI: 10.1002/cctc.200900190
    184. 184
      Chen, B.-S.; Resch, V.; Otten, L. G.; Hanefeld, U. Enantioselective Michael Addition of Water. Eur. J. Chem. 2015, 21 (7), 302030,  DOI: 10.1002/chem.201405579
    185. 185
      Lehwald, P.; Richter, M.; Röhr, C.; Liu, H.-w.; Müller, M. Enantioselective Intermolecular Aldehyde-Ketone Cross-Coupling through an Enzymatic Carboligation Reaction. Angew. Chem., Int. Ed. 2010, 49 (13), 238992,  DOI: 10.1002/anie.200906181
    186. 186
      Hammer, S. C.; Marjanovic, A.; Dominicus, J. M.; Nestl, B. M.; Hauer, B. Squalene Hopene Cyclases Are Protonases for Stereoselective Brønsted Acid Catalysis. Nat. Chem. Biol. 2015, 11 (2), 1216,  DOI: 10.1038/nchembio.1719
    187. 187
      Loskot, S. A.; Romney, D. K.; Arnold, F. H.; Stoltz, B. M. Enantioselective Total Synthesis of Nigelladine a Via Late-Stage C-H Oxidation Enabled by an Engineered P450 Enzyme. J. Am. Chem. Soc. 2017, 139 (30), 101969,  DOI: 10.1021/jacs.7b05196
    188. 188
      Zanon, J. P.; Peres, M. F. S.; Gattás, E. A. L. Colorimetric Assay of Ethanol Using Alcohol Dehydrogenase from Dry Baker’s Yeast. Enzyme Microb. Technol. 2007, 40 (3), 46670,  DOI: 10.1016/j.enzmictec.2006.07.029
    189. 189
      Baker, J. L.; Faustoferri, R. C.; Quivey, R. G., Jr. A Modified Chromogenic Assay for Determination of the Ratio of Free Intracellular NAD(+)/NADH in Streptococcus Mutans. Bio Protoc. 2016, 6 (16), e1902,  DOI: 10.21769/BioProtoc.1902
    190. 190
      Yang, Y.; Liu, J.; Li, Z. Engineering of P450pyr Hydroxylase for the Highly Regio- and Enantioselective Subterminal Hydroxylation of Alkanes. Angew. Chem., Int. Ed. 2014, 53 (12), 31204,  DOI: 10.1002/anie.201311091
    191. 191
      Lauchli, R.; Rabe, K. S.; Kalbarczyk, K. Z.; Tata, A.; Heel, T.; Kitto, R. Z.; Arnold, F. H. High-Throughput Screening for Terpene-Synthase-Cyclization Activity and Directed Evolution of a Terpene Synthase. Angew. Chem., Int. Ed. 2013, 52 (21), 55714,  DOI: 10.1002/anie.201301362
    192. 192
      Jung, E.; Park, B. G.; Yoo, H.-W.; Kim, J.; Choi, K.-Y.; Kim, B.-G. Semi-Rational Engineering of CYP153A35 to Enhance Ω-Hydroxylation Activity toward Palmitic Acid. Appl. Microbiol. Biotechnol. 2018, 102 (1), 26977,  DOI: 10.1007/s00253-017-8584-y
    193. 193
      Francisco, W. A.; Abu-Soud, H. M.; Baldwin, T. O.; Raushel, F. M. Interaction of Bacterial Luciferase with Aldehyde Substrates and Inhibitors. J. Biol. Chem. 1993, 268 (33), 2473441,  DOI: 10.1016/S0021-9258(19)74526-8
    194. 194
      Minak-Bernero, V.; Bare, R. E.; Haith, C. E.; Grossman, M. J. Detection of Alkanes, Alcohols, and Aldehydes Using Bioluminescence. Biotechnol. Bioeng. 2004, 87 (2), 1707,  DOI: 10.1002/bit.20089
    195. 195
      Greene, L. E.; Lincoln, R.; Krumova, K.; Cosa, G. Development of a Fluorogenic Reactivity Palette for the Study of Nucleophilic Addition Reactions Based on meso-Formyl BODIPY Dyes. ACS Omega 2017, 2 (12), 861824,  DOI: 10.1021/acsomega.7b01795
    196. 196
      Lincoln, R.; Greene, L. E.; Bain, C.; Flores-Rizo, J. O.; Bohle, D. S.; Cosa, G. When Push Comes to Shove: Unravelling the Mechanism and Scope of Nonemissive meso-Unsaturated BODIPY Dyes. J. Phys. Chem. B 2015, 119 (13), 475865,  DOI: 10.1021/acs.jpcb.5b02080
    197. 197
      Sawminathan, S.; Iyer, S. K. A New Imidazole Based Phenanthridine Probe for Ratiometric Fluorescence Monitoring of Methanol in Biodiesel. New J. Chem. 2021, 45 (13), 603341,  DOI: 10.1039/D0NJ06252A
    198. 198
      Mohr, G. J.; Spichiger-Keller, U. E. Novel Fluorescent Sensor Membranes for Alcohols Based on p-N,N-Dioctylamino-4′-Trifluoroacetylstilbene. Anal. Chim. Acta 1997, 351 (1), 18996,  DOI: 10.1016/S0003-2670(97)00365-6
    199. 199
      Mohr, G. J.; Lehmann, F.; Grummt, U.-W.; Spichiger-Keller, U. E. Fluorescent Ligands for Optical Sensing of Alcohols: Synthesis and Characterisation of p-N,N-Dialkylamino-Trifluoroacetylstilbenes. Anal. Chim. Acta 1997, 344 (3), 21525,  DOI: 10.1016/S0003-2670(97)00113-X
    200. 200
      Takano, K.; Sasaki, S.-i.; Citterio, D.; Tamiaki, H.; Suzuki, K. An Oxo-Bacteriochlorin Derivative for Long-Wavelength Fluorescence Ratiometric Alcohol Sensing. Analyst 2010, 135 (9), 23349,  DOI: 10.1039/c0an00173b
    201. 201
      Zhao, M.; Yue, Y.; Liu, C.; Hui, P.; He, S.; Zhao, L.; Zeng, X. A Colorimetric and Fluorometric Dual-Modal Sensor for Methanol Based on a Functionalized Pentacenequinone Derivative. Chem. Commun. 2018, 54 (60), 833942,  DOI: 10.1039/C8CC04515A
    202. 202
      Roy, S.; Das, S.; Ray, A.; Parui, P. P. Fluorometric Trace Methanol Detection in Ethanol and Isopropanol in a Water Medium for Application in Alcoholic Beverages and Hand Sanitizers. RSC Adv. 2021, 11 (48), 30093101,  DOI: 10.1039/D1RA05201B
    203. 203
      Wu, Z.; Fu, X.; Wang, Y. Click Synthesis of a Triphenylamine-Based Fluorescent Methanol Probe with a Unique D-Π-a Structure. Sens. Actuators B Chem. 2017, 245, 40613,  DOI: 10.1016/j.snb.2017.01.164
    204. 204
      Kumar, V.; Kumar, A.; Diwan, U.; Singh, M. K.; Upadhyay, K. K. A Radical Approach for Fluorescent Turn ‘On’ Detection, Differentiation and Bioimaging of Methanol. Org. Biomol. Chem. 2015, 13 (33), 88226,  DOI: 10.1039/C5OB01333J
    205. 205
      Kumar, V.; Kundu, S.; Sk, B.; Patra, A. A Naked-Eye Colorimetric Sensor for Methanol and ‘Turn-On’ Fluorescence Detection of Al3+. New J. Chem. 2019, 43 (47), 185829,  DOI: 10.1039/C9NJ04688G
    206. 206
      Patel, R. N. Synthesis of Chiral Pharmaceutical Intermediates by Biocatalysis. Coord. Chem. Rev. 2008, 252 (5), 659701,  DOI: 10.1016/j.ccr.2007.10.031
    207. 207
      Chen, Y.; Chen, C.; Wu, X. Dicarbonyl Reduction by Single Enzyme for the Preparation of Chiral Diols. Chem. Soc. Rev. 2012, 41 (5), 174253,  DOI: 10.1039/c1cs15230k
    208. 208
      Ramasastry, S. S. V.; Zhang, H.; Tanaka, F.; Barbas, C. F. Direct Catalytic Asymmetric Synthesis of anti-1,2-Amino Alcohols and syn-1,2-Diols through Organocatalytic anti-Mannich and syn-Aldol Reactions. J. Am. Chem. Soc. 2007, 129 (2), 2889,  DOI: 10.1021/ja0677012
    209. 209
      Guo, Z.; Shin, I.; Yoon, J. Recognition and Sensing of Various Species Using Boronic Acid Derivatives. Chem. Commun. 2012, 48 (48), 595667,  DOI: 10.1039/c2cc31985c
    210. 210
      Hudlicky, T.; Thorpe, A. J. Current Status and Future Perspectives of Cyclohexadiene-cis-Diols in Organic Synthesis: Versatile Intermediates in the Concise Design of Natural Products. Chem. Commun. 1996, 17 (17), 19932000,  DOI: 10.1039/cc9960001993
    211. 211
      Wingstrand, M. J.; Madsen, C. M.; Clausen, M. H. Rapid Synthesis of Macrocycles from Diol Precursors. Tetrahedron Lett. 2009, 50 (6), 6935,  DOI: 10.1016/j.tetlet.2008.11.100
    212. 212
      Tang, X.; Wei, J.; Ding, N.; Sun, Y.; Zeng, X.; Hu, L.; Liu, S.; Lei, T.; Lin, L. Chemoselective Hydrogenation of Biomass Derived 5-Hydroxymethylfurfural to Diols: Key Intermediates for Sustainable Chemicals, Materials and Fuels. Renew. Sust. Energy Rev. 2017, 77, 28796,  DOI: 10.1016/j.rser.2017.04.013
    213. 213
      Song, J.; Shao, P.-L.; Wang, J.; Huang, F.; Zhang, X. Asymmetric Hydrogenation of 1,4-Diketones: Facile Synthesis of Enantiopure 1,4-Diarylbutane-1,4-Diols. Chem. Commun. 2021, 58 (2), 2625,  DOI: 10.1039/D1CC05359K
    214. 214
      Faber, K.; Mischitz, M.; Kroutil, W. Microbial Epoxide Hydrolases. Acta Chem. Scand. 1996, 50, 249,  DOI: 10.3891/acta.chem.scand.50-0249
    215. 215
      Ouellette, R. J.; Rawn, J. D. Alcohols: Reactions and Synthesis. In Organic Chemistry, 2nd ed.; Academic Press, 2018; Chapter 16, pp 463505.
    216. 216
      Rinner, U. Chiral Pool Synthesis. In Chiral Pool Syntheses from cis-Cyclohexadiene Diols, Comprehensive Chirality; Elsevier: Amsterdam, 2012; Chapter 2.9, pp 24067.
    217. 217
      Bruice, P. Y. Organic Chemistry; 7. ed.; Pearson Education: Harlow, 2014.
    218. 218
      Gubbels, E.; Heitz, T.; Yamamoto, M.; Chilekar, V.; Zarbakhsh, S.; Gepraegs, M.; Köpnick, H.; Schmidt, M.; Brügging, W.; Rüter, J. Polyesters. Ullmann’s Encyclopedia of Industrial Chemistry Wiley VCH, 2018; 130 DOI: 10.1002/14356007.a21_227.pub2 .
    219. 219
      Steinreiber, A.; Faber, K. Microbial Epoxide Hydrolases for Preparative Biotransformations. Curr. Opin. Biotechnol. 2001, 12 (6), 5528,  DOI: 10.1016/S0958-1669(01)00262-2
    220. 220
      Gibson, D. T.; Parales, R. E. Aromatic Hydrocarbon Dioxygenases in Environmental Biotechnology. Curr. Opin. Biotechnol. 2000, 11 (3), 23643,  DOI: 10.1016/S0958-1669(00)00090-2
    221. 221
      Boyd, D. R.; Sharma, N. D.; Allen, C. C. R. Aromatic Dioxygenases: Molecular Biocatalysis and Applications. Curr. Opin. Biotechnol. 2001, 12 (6), 56473,  DOI: 10.1016/S0958-1669(01)00264-6
    222. 222
      Zhao, L.; Han, B.; Huang, Z.; Miller, M.; Huang, H.; Malashock, D. S.; Zhu, Z.; Milan, A.; Robertson, D. E.; Weiner, D. P. Epoxide Hydrolase-Catalyzed Enantioselective Synthesis of Chiral 1,2-Diols Via Desymmetrization of meso-Epoxides. J. Am. Chem. Soc. 2004, 126 (36), 111567,  DOI: 10.1021/ja0466210
    223. 223
      Makarova, M.; Endoma-Arias, M. A. A.; Dela Paz, H. E.; Simionescu, R.; Hudlicky, T. Chemoenzymatic Total Synthesis of ent-Oxycodone: Second-, Third-, and Fourth-Generation Strategies. J. Am. Chem. Soc. 2019, 141 (27), 10883904,  DOI: 10.1021/jacs.9b05033
    224. 224
      Ikeda, H.; Sato, E.; Sugai, T.; Ohta, H. Yeast-Mediated Synthesis of Optically Active Diols with C2-Symmetry and (R)-4-Pentanolide. Tetrahedron 1996, 52 (24), 811322,  DOI: 10.1016/0040-4020(96)00373-0
    225. 225
      Gijsen, H. J. M.; Wong, C.-H. Unprecedented Asymmetric Aldol Reactions with Three Aldehyde Substrates Catalyzed by 2-Deoxyribose-5-Phosphate Aldolase. J. Am. Chem. Soc. 1994, 116 (18), 84223,  DOI: 10.1021/ja00097a082
    226. 226
      Müller, M. Chemoenzymatic Synthesis of Building Blocks for Statin Side Chains. Angew. Chem., Int. Ed. 2005, 44 (3), 3625,  DOI: 10.1002/anie.200460852
    227. 227
      Wu, X.; Jiang, J.; Chen, Y. Correlation between Intracellular Cofactor Concentrations and Biocatalytic Efficiency: Coexpression of Diketoreductase and Glucose Dehydrogenase for the Preparation of Chiral Diol for Statin Drugs. ACS Catal. 2011, 1 (12), 16614,  DOI: 10.1021/cs200408y
    228. 228
      Huffman, M. A.; Fryszkowska, A.; Alvizo, O.; Borra-Garske, M.; Campos, K. R.; Canada, K. A.; Devine, P. N.; Duan, D.; Forstater, J. H.; Grosser, S. T. Design of an in Vitro Biocatalytic Cascade for the Manufacture of Islatravir. Science 2019, 366 (6470), 12559,  DOI: 10.1126/science.aay8484
    229. 229
      Bian, Z.; Liu, A.; Li, Y.; Fang, G.; Yao, Q.; Zhang, G.; Wu, Z. Boronic Acid Sensors with Double Recognition Sites: A Review. Analyst 2020, 145 (3), 71944,  DOI: 10.1039/C9AN00741E
    230. 230
      Cao, H.; Heagy, M. D. Fluorescent Chemosensors for Carbohydrates: A Decade’s Worth of Bright Spies for Saccharides in Review. J. Fluoresc. 2004, 14 (5), 56984,  DOI: 10.1023/B:JOFL.0000039344.34642.4c
    231. 231
      Fang, G.; Wang, H.; Bian, Z.; Sun, J.; Liu, A.; Fang, H.; Liu, B.; Yao, Q.; Wu, Z. Recent Development of Boronic Acid-Based Fluorescent Sensors. RSC Adv. 2018, 8 (51), 2940027,  DOI: 10.1039/C8RA04503H
    232. 232
      Wang, W.; Gao, X.; Wang, B. Boronic Acid-Based Sensors. Curr. Org. Chem. 2002, 6 (14), 1285317,  DOI: 10.2174/1385272023373446
    233. 233
      Wu, X.; Li, Z.; Chen, X.-X.; Fossey, J. S.; James, T. D.; Jiang, Y.-B. Selective Sensing of Saccharides Using Simple Boronic Acids and Their Aggregates. Chem. Soc. Rev. 2013, 42 (20), 803248,  DOI: 10.1039/c3cs60148j
    234. 234
      Arnaud, J.; Audfray, A.; Imberty, A. Binding Sugars: From Natural Lectins to Synthetic Receptors and Engineered Neolectins. Chem. Soc. Rev. 2013, 42 (11), 4798813,  DOI: 10.1039/c2cs35435g
    235. 235
      Wu, X.; Chen, X.-X.; Jiang, Y.-B. Recent Advances in Boronic Acid-Based Optical Chemosensors. Analyst 2017, 142 (9), 140314,  DOI: 10.1039/C7AN00439G
    236. 236
      Kondo, K.; Shiomi, Y.; Saisho, M.; Harada, T.; Shinkai, S. Specific Complexation of Disaccharides with Diphenyl-3,3′-Diboronic Acid That Can Be Detected by Circular Dichroism. Tetrahedron 1992, 48 (38), 823952,  DOI: 10.1016/S0040-4020(01)80492-0
    237. 237
      Yoon, J.; Czarnik, A. W. Fluorescent Chemosensors of Carbohydrates. A Means of Chemically Communicating the Binding of Polyols in Water Based on Chelation-Enhanced Quenching. J. Am. Chem. Soc. 1992, 114 (14), 58745,  DOI: 10.1021/ja00040a067
    238. 238
      Elfeky, S. Novel Bronic Acid-Based Fluorescent Sensor for Sugars and Nucleosides. Curr. Org. Synth. 2011, 8 (6), 872,  DOI: 10.2174/1570179411108060872
    239. 239
      James, T. D.; Shinkai, S. Artificial Receptors as Chemosensors for Carbohydrates. Host-Guest Chemistry 2002, 218, 159200,  DOI: 10.1007/3-540-45010-6_6
    240. 240
      Springsteen, G.; Wang, B. A Detailed Examination of Boronic Acid-Diol Complexation. Tetrahedron 2002, 58 (26), 5291300,  DOI: 10.1016/S0040-4020(02)00489-1
    241. 241
      Jin, S.; Cheng, Y.; Reid, S.; Li, M.; Wang, B. Carbohydrate Recognition by Boronolectins, Small Molecules, and Lectins. Med. Res. Rev. 2009, 30, 171257,  DOI: 10.1002/med.20155
    242. 242
      Yan, J.; Fang, H.; Wang, B. Boronolectins and Fluorescent Boronolectins: An Examination of the Detailed Chemistry Issues Important for the Design. Med. Res. Rev. 2005, 25 (5), 490520,  DOI: 10.1002/med.20038
    243. 243
      Gao, X.; Zhang, Y.; Wang, B. New Boronic Acid Fluorescent Reporter Compounds. 2. A Naphthalene-Based On-Off Sensor Functional at Physiological pH. Org. Lett. 2003, 5 (24), 46158,  DOI: 10.1021/ol035783i
    244. 244
      Gao, X.; Zhang, Y.; Wang, B. A Highly Fluorescent Water-Soluble Boronic Acid Reporter for Saccharide Sensing That Shows Ratiometric UV Changes and Significant Fluorescence Changes. Tetrahedron 2005, 61 (38), 91117,  DOI: 10.1016/j.tet.2005.07.035
    245. 245
      Gao, X.; Zhang, Y.; Wang, B. Naphthalene-Based Water-Soluble Fluorescent Boronic Acid Isomers Suitable for Ratiometric and Off-On Sensing of Saccharides at Physiological pH. New J. Chem. 2005, 29 (4), 57986,  DOI: 10.1039/b413376e
    246. 246
      Zhang, Y.; Gao, X.; Hardcastle, K.; Wang, B. Water-Soluble Fluorescent Boronic Acid Compounds for Saccharide Sensing: Substituent Effects on Their Fluorescence Properties. Eur. J. Chem. 2006, 12 (5), 137784,  DOI: 10.1002/chem.200500982
    247. 247
      Yang, W.; Yan, J.; Springsteen, G.; Deeter, S.; Wang, B. A Novel Type of Fluorescent Boronic Acid That Shows Large Fluorescence Intensity Changes Upon Binding with a Carbohydrate in Aqueous Solution at Physiological pH. Bioorg. Med. Chem. Lett. 2003, 13 (6), 101922,  DOI: 10.1016/S0960-894X(03)00086-6
    248. 248
      Yang, W.; Lin, L.; Wang, B. A New Type of Boronic Acid Fluorescent Reporter Compound for Sugar Recognition. Tetrahedron Lett. 2005, 46 (46), 79814,  DOI: 10.1016/j.tetlet.2005.09.074
    249. 249
      Cheng, Y.; Ni, N.; Yang, W.; Wang, B. A New Class of Fluorescent Boronic Acids That Have Extraordinarily High Affinities for Diols in Aqueous Solution at Physiological pH. Eur. J. Chem. 2010, 16 (45), 1352838,  DOI: 10.1002/chem.201000637
    250. 250
      Samaniego Lopez, C.; Lago Huvelle, M. A.; Uhrig, M. L.; Coluccio Leskow, F.; Spagnuolo, C. C. Recognition of Saccharides in the NIR Region with a Novel Fluorogenic Boronolectin: In Vitro and Live Cell Labeling. Chem. Commun. 2015, 51 (23), 48958,  DOI: 10.1039/C4CC10425K
    251. 251
      Wang, L.; Yuan, L.; Zeng, X.; Peng, J.; Ni, Y.; Er, J. C.; Xu, W.; Agrawalla, B. K.; Su, D.; Kim, B. A Multisite-Binding Switchable Fluorescent Probe for Monitoring Mitochondrial ATP Level Fluctuation in Live Cells. Angew. Chem., Int. Ed. 2016, 55 (5), 17736,  DOI: 10.1002/anie.201510003
    252. 252
      Huang, Y.-J.; Ouyang, W.-J.; Wu, X.; Li, Z.; Fossey, J. S.; James, T. D.; Jiang, Y.-B. Glucose Sensing Via Aggregation and the Use of “Knock-Out” Binding to Improve Selectivity. J. Am. Chem. Soc. 2013, 135 (5), 17003,  DOI: 10.1021/ja311442x
    253. 253
      Vilozny, B.; Schiller, A.; Wessling, R. A.; Singaram, B. Enzyme Assays with Boronic Acid Appended Bipyridinium Salts. Anal. Chim. Acta 2009, 649 (2), 24651,  DOI: 10.1016/j.aca.2009.07.032
    254. 254
      Sharrett, Z.; Gamsey, S.; Hirayama, L.; Vilozny, B.; Suri, J. T.; Wessling, R. A.; Singaram, B. Exploring the Use of APTS as a Fluorescent Reporter Dye for Continuous Glucose Sensing. Org. Biomol. Chem. 2009, 7 (7), 146170,  DOI: 10.1039/b821934f
    255. 255
      James, T. D.; Sandanayake, K. R. A. S.; Shinkai, S. A Glucose-Selective Molecular Fluorescence Sensor. Angew. Chem., Int. Ed. 1994, 33 (21), 22079,  DOI: 10.1002/anie.199422071
    256. 256
      James, T. D.; Sandanayake, K. R. A. S.; Iguchi, R.; Shinkai, S. Novel Saccharide-Photoinduced Electron Transfer Sensors Based on the Interaction of Boronic Acid and Amine. J. Chem. Soc., Chem. Commun. 1995, 117 (35), 89827,  DOI: 10.1021/ja00140a013
    257. 257
      Tran, T. M.; Alan, Y.; Glass, T. E. A Highly Selective Fluorescent Sensor for Glucosamine. Chem. Commun. 2015, 51 (37), 79158,  DOI: 10.1039/C5CC00415B
    258. 258
      Rout, B.; Unger, L.; Armony, G.; Iron, M. A.; Margulies, D. Medication Detection by a Combinatorial Fluorescent Molecular Sensor. Angew. Chem., Int. Ed. 2012, 51 (50), 1247781,  DOI: 10.1002/anie.201206374
    259. 259
      Liu, Y.; Deng, C.; Tang, L.; Qin, A.; Hu, R.; Sun, J. Z.; Tang, B. Z. Specific Detection of D-Glucose by a Tetraphenylethene-Based Fluorescent Sensor. J. Am. Chem. Soc. 2011, 133 (4), 6603,  DOI: 10.1021/ja107086y
    260. 260
      Shcherbakova, E. G.; Brega, V.; Lynch, V. M.; James, T. D.; Anzenbacher, P., Jr. High-Throughput Assay for Enantiomeric Excess Determination in 1,2- and 1,3-Diols and Direct Asymmetric Reaction Screening. Eur. J. Chem. 2017, 23 (42), 102229,  DOI: 10.1002/chem.201701923
    261. 261
      Shcherbakova, E. G.; James, T. D.; Anzenbacher, P. High-Throughput Assay for Determining Enantiomeric Excess of Chiral Diols, Amino Alcohols, and Amines and for Direct Asymmetric Reaction Screening. Nat. Protoc. 2020, 15 (7), 220329,  DOI: 10.1038/s41596-020-0329-1
    262. 262
      Doderer, K.; Lutz-Wahl, S.; Hauer, B.; Schmid, R. D. Spectrophotometric Assay for Epoxide Hydrolase Activity toward any Epoxide. Anal. Biochem. 2003, 321 (1), 1314,  DOI: 10.1016/S0003-2697(03)00399-3
    263. 263
      Wahler, D.; Reymond, J.-L. The Adrenaline Test for Enzymes. Angew. Chem., Int. Ed. 2002, 41 (7), 122932,  DOI: 10.1002/1521-3773(20020402)41:7<1229::AID-ANIE1229>3.0.CO;2-5
    264. 264
      Jie, N.; Yang, D.; Zhang, Q.; Yang, J.; Song, Z. Fluorometric Determination of Periodate with Thiamine and Its Application to the Determination of Ethylene Glycol and Glycerol. Anal. Chim. Acta 1998, 359 (1), 8792,  DOI: 10.1016/S0003-2670(97)00657-0
    265. 265
      Kirschner, A.; Bornscheuer, U. T. Directed Evolution of a Baeyer-Villiger Monooxygenase to Enhance Enantioselectivity. Appl. Microbiol. Biotechnol. 2008, 81 (3), 46572,  DOI: 10.1007/s00253-008-1646-4
    266. 266
      Wahler, D.; Boujard, O.; Lefèvre, F.; Reymond, J.-L. Adrenaline Profiling of Lipases and Esterases with 1,2-Diol and Carbohydrate Acetates. Tetrahedron 2004, 60 (3), 70310,  DOI: 10.1016/j.tet.2003.11.059
    267. 267
      Fluxá, V. S.; Wahler, D.; Reymond, J.-L. Enzyme Assay and Activity Fingerprinting of Hydrolases with the Red-Chromogenic Adrenaline Test. Nat. Protoc. 2008, 3 (8), 12707,  DOI: 10.1038/nprot.2008.106
    268. 268
      Doderer, K.; Schmid, R. D. Fluorometric Assay for Determining Epoxide Hydrolase Activity. Biotechnol. Lett. 2004, 26 (10), 8359,  DOI: 10.1023/B:BILE.0000025887.36874.33
    269. 269
      Matsuno, K.; Suzuki, S. Simple Fluorimetric Method for Quantification of Sialic Acids in Glycoproteins. Anal. Biochem. 2008, 375 (1), 539,  DOI: 10.1016/j.ab.2008.01.002
    270. 270
      Preston-Herrera, C.; Jackson, A. S.; Bachmann, B. O.; Froese, J. T. Development and Application of a High Throughput Assay System for the Detection of Rieske Dioxygenase Activity. Org. Biomol. Chem. 2021, 19 (4), 77584,  DOI: 10.1039/D0OB02412K
    271. 271
      Winkler, M. Carboxylic Acid Reductase Enzymes (CARS). Curr. Opin. Chem. Biol. 2018, 43, 239,  DOI: 10.1016/j.cbpa.2017.10.006
    272. 272
      Sutiono, S.; Carsten, J.; Sieber, V. Structure-Guided Engineering of α-Keto Acid Decarboxylase for the Production of Higher Alcohols at Elevated Temperature. ChemSusChem 2018, 11 (18), 333544,  DOI: 10.1002/cssc.201800944
    273. 273
      Greenhalgh, J. C.; Fahlberg, S. A.; Pfleger, B. F.; Romero, P. A. Machine Learning-Guided Acyl-ACP Reductase Engineering for Improved in Vivo Fatty Alcohol Production. Nat. Commun. 2021, 12, 5825,  DOI: 10.1038/s41467-021-25831-w
    274. 274
      Sellés Vidal, L.; Kelly, C. L.; Mordaka, P. M.; Heap, J. T. Review of NAD(P)H-Dependent Oxidoreductases: Properties, Engineering and Application. Biochim. Biophys. Acta 2018, 1866 (2), 32747,  DOI: 10.1016/j.bbapap.2017.11.005
    275. 275
      Goswami, P.; Chinnadayyala, S. S. R.; Chakraborty, M.; Kumar, A. K.; Kakoti, A. An Overview on Alcohol Oxidases and Their Potential Applications. Appl. Microbiol. Biotechnol. 2013, 97 (10), 425975,  DOI: 10.1007/s00253-013-4842-9
    276. 276
      Mang, H.; Gross, J.; Lara, M.; Goessler, C.; Schoemaker, H. E.; Guebitz, G. M.; Kroutil, W. Biocatalytic Single-Step Alkene Cleavage from Aryl Alkenes: An Enzymatic Equivalent to Reductive Ozonization. Angew. Chem., Int. Ed. 2006, 45 (31), 52013,  DOI: 10.1002/anie.200601574
    277. 277
      Schwendenwein, D.; Ressmann, A. K.; Doerr, M.; Höhne, M.; Bornscheuer, U. T.; Mihovilovic, M. D.; Rudroff, F.; Winkler, M. Random Mutagenesis-Driven Improvement of Carboxylate Reductase Activity Using an Amino Benzamidoxime-Mediated High-Throughput Assay. Adv. Synth. Catal. 2019, 361, 2544,  DOI: 10.1002/adsc.201900155
    278. 278
      Soh, L. M. J.; Mak, W. S.; Lin, P. P.; Mi, L.; Chen, F. Y. H.; Damoiseaux, R.; Siegel, J. B.; Liao, J. C. Engineering a Thermostable Keto Acid Decarboxylase Using Directed Evolution and Computationally Directed Protein Design. ACS Synth. Biol. 2017, 6 (4), 6108,  DOI: 10.1021/acssynbio.6b00240
    279. 279
      Van Schie, M. M. C. H.; Pedroso De Almeida, T.; Laudadio, G.; Tieves, F.; Fernández-Fueyo, E.; Noël, T.; Arends, I. W. C. E.; Hollmann, F. Biocatalytic Synthesis of the Green Note trans-2-Hexenal in a Continuous-Flow Microreactor. Beilstein J. Org. Chem. 2018, 14, 697703,  DOI: 10.3762/bjoc.14.58
    280. 280
      Allen, C. F. The Identification of Carbonyl Compounds by Use of 2,4-Dinitrophenylhydrazine. J. Am. Chem. Soc. 1930, 52 (7), 29559,  DOI: 10.1021/ja01370a058
    281. 281
      Brady, O. L.; Elsmie, G. V. The Use of 2,4-Dinitrophenylhydrazine as a Reagent for Aldehydes and Ketones. Analyst 1926, 51 (599), 778,  DOI: 10.1039/an9265100077
    282. 282
      Bohlmann, F. Konstitution Und Lichtabsorption, I. Mitteil.: Carbonyl-Derivate. Chem. Ber. 1951, 84 (5–6), 490504,  DOI: 10.1002/cber.19510840515
    283. 283
      Georgiou, C. D.; Zisimopoulos, D.; Argyropoulou, V.; Kalaitzopoulou, E.; Salachas, G.; Grune, T. Protein and Cell Wall Polysaccharide Carbonyl Determination by a Neutral pH 2,4-Dinitrophenylhydrazine-Based Photometric Assay. Redox Biol. 2018, 17, 12842,  DOI: 10.1016/j.redox.2018.04.010
    284. 284
      Mesquita, C. S.; Oliveira, R.; Bento, F.; Geraldo, D.; Rodrigues, J. V.; Marcos, J. C. Simplified 2,4-Dinitrophenylhydrazine Spectrophotometric Assay for Quantification of Carbonyls in Oxidized Proteins. Anal. Biochem. 2014, 458, 6971,  DOI: 10.1016/j.ab.2014.04.034
    285. 285
      Xue, Y.-P.; Wang, W.; Wang, Y.-J.; Liu, Z.-Q.; Zheng, Y.-G.; Shen, Y.-C. Isolation of Enantioselective α-Hydroxyacid Dehydrogenases Based on a High-Throughput Screening Method. Bioprocess Biosyst. Eng. 2012, 35 (9), 151522,  DOI: 10.1007/s00449-012-0741-1
    286. 286
      Chen, Q.; Chen, X.; Feng, J.; Wu, Q.; Zhu, D.; Ma, Y. Improving and Inverting Cβ-Stereoselectivity of Threonine Aldolase Via Substrate-Binding-Guided Mutagenesis and a Stepwise Visual Screening. ACS Catal. 2019, 9 (5), 44629,  DOI: 10.1021/acscatal.9b00859
    287. 287
      Yang, C.; Ye, L.; Gu, J.; Yang, X.; Li, A.; Yu, H. Directed Evolution of Mandelate Racemase by a Novel High-Throughput Screening Method. Appl. Microbiol. Biotechnol. 2017, 101 (3), 106372,  DOI: 10.1007/s00253-016-7790-3
    288. 288
      Uzu, S.; Kanda, S.; Imai, K.; Nakashima, K.; Akiyama, S. Fluorogenic Reagents: 4-Aminosulphonyl-7-Hydrazino-2,1,3-Benzoxadiazole, 4-(N,N-Dimethylaminosulphonyl)-7-Hydrazino-2,1,3-Benzoxadiazole and 4-Hydrazino-7-Nitro-2,1,3-Benzoxadiazole Hydrazine for Aldehydes and Ketones. Analyst 1990, 115 (11), 147782,  DOI: 10.1039/an9901501477
    289. 289
      Uchiyama, S.; Santa, T.; Okiyama, N.; Fukushima, T.; Imai, K. Fluorogenic and Fluorescent Labeling Reagents with a Benzofurazan Skeleton. Biomed. Chromatogr. 2001, 15 (5), 295318,  DOI: 10.1002/bmc.75
    290. 290
      Frizon, T. E. A.; Vieira, A. A.; da Silva, F. N.; Saba, S.; Farias, G.; de Souza, B.; Zapp, E.; Lôpo, M. N.; Braga, H. d. C.; Grillo, F. Synthesis of 2,1,3-Benzoxadiazole Derivatives as New Fluorophores─Combined Experimental, Optical, Electro, and Theoretical Study. Front. Chem. 2020, 8, 360,  DOI: 10.3389/fchem.2020.00360
    291. 291
      Konarzycka-Bessler, M.; Bornscheuer, U. T. A High-Throughput-Screening Method for Determining the Synthetic Activity of Hydrolases. Angew. Chem., Int. Ed. 2003, 42 (12), 141820,  DOI: 10.1002/anie.200390365
    292. 292
      Jencks, W. P. Progress in Physical Organic Chemistry; Index to Reviews, Symposia Volumes and Monographs in Organic Chemistry; Cohen, S. G., Streitwieser, A., Jr., Taft, R. W., Eds.; John Wiley & Sons, 1964; Vol. 2, p 110 DOI: 10.1002/9780470171813
    293. 293
      Larsen, D.; Pittelkow, M.; Karmakar, S.; Kool, E. T. New Organocatalyst Scaffolds with High Activity in Promoting Hydrazone and Oxime Formation at Neutral pH. Org. Lett. 2015, 17 (2), 2747,  DOI: 10.1021/ol503372j
    294. 294
      Crisalli, P.; Kool, E. T. Importance of Ortho Proton Donors in Catalysis of Hydrazone Formation. Org. Lett. 2013, 15 (7), 16469,  DOI: 10.1021/ol400427x
    295. 295
      Wang, S.; Nawale, G. N.; Kadekar, S.; Oommen, O. P.; Jena, N. K.; Chakraborty, S.; Hilborn, J.; Varghese, O. P. Saline Accelerates Oxime Reaction with Aldehyde and Keto Substrates at Physiological pH. Sci. Rep. 2018, 8, 2193,  DOI: 10.1038/s41598-018-20735-0
    296. 296
      Mayr, H.; Bug, T.; Gotta, M. F.; Hering, N.; Irrgang, B.; Janker, B.; Kempf, B.; Loos, R.; Ofial, A. R.; Remennikov, G. Reference Scales for the Characterization of Cationic Electrophiles and Neutral Nucleophiles. J. Am. Chem. Soc. 2001, 123 (39), 950012,  DOI: 10.1021/ja010890y
    297. 297
      Brotzel, F.; Chu, Y. C.; Mayr, H. Nucleophilicities of Primary and Secondary Amines in Water. J. Org. Chem. 2007, 72 (10), 367988,  DOI: 10.1021/jo062586z
    298. 298
      Brotzel, F., lmu, 2008.
    299. 299
      Nakashima, K.; Hidaka, Y.; Yoshida, T.; Kuroda, N.; Akiyama, S. High-Performance Liquid Chromatographic Determination of Short-Chain Aliphatic Aldehydes Using 4-(N,N-Dimethylaminosulphonyl)-7-Hydrazino-2,1,3-Benzoxadiazole as a Fluorescence Reagent. J. Chromatogr. B Biomed. Appl. 1994, 661 (2), 20510,  DOI: 10.1016/0378-4347(94)00359-9
    300. 300
      Jacobsen, N. W.; Dickinson, R. G. Spectrometric Assay of Aldehydes as 6-Mercapto-3-Substituted-s-Trizolo (4,3-b)-Tetrazines. Anal. Chem. 1974, 46 (2), 2989,  DOI: 10.1021/ac60338a039
    301. 301
      Dickinson, R. G.; Jacobsen, N. W. A New Sensitive and Specific Test for the Detection of Aldehydes: Formation of 6-Mercapto-3-Substituted-S-Triazolo[4,3-b]-s-Tetrazines. J. Chem. Soc. D 1970, (24), 171920,  DOI: 10.1039/c29700001719
    302. 302
      Quesenberry, M. S.; Lee, Y. C. A Rapid Formaldehyde Assay Using Purpald Reagent: Application under Periodation Conditions. Anal. Biochem. 1996, 234 (1), 505,  DOI: 10.1006/abio.1996.0048
    303. 303
      Anthon, G. E.; Barrett, D. M. Comparison of Three Colorimetric Reagents in the Determination of Methanol with Alcohol Oxidase. Application to the Assay of Pectin Methylesterase. J. Agric. Food Chem. 2004, 52 (12), 374953,  DOI: 10.1021/jf035284w
    304. 304
      Hammer, S. C.; Kubik, G.; Watkins, E.; Huang, S.; Minges, H.; Arnold, F. H. Anti-Markovnikov Alkene Oxidation by Metal-Oxo-Mediated Enzyme Catalysis. Science 2017, 358 (6360), 2158,  DOI: 10.1126/science.aao1482
    305. 305
      Nguyen, Q.-T.; Romero, E.; Dijkman, W. P.; de Vasconcellos, S. P.; Binda, C.; Mattevi, A.; Fraaije, M. W. Structure-Based Engineering of Phanerochaete Chrysosporium Alcohol Oxidase for Enhanced Oxidative Power toward Glycerol. Biochemistry 2018, 57 (43), 620918,  DOI: 10.1021/acs.biochem.8b00918
    306. 306
      Sawicki, E.; Hauser, T. R.; Stanley, T. W.; Elbert, W. The 3-Methyl-2-Benzothiazolone Hydrazone Test. Sensitive New Methods for the Detection, Rapid Estimation, and Determination of Aliphatic Aldehydes. Anal. Chem. 1961, 33 (1), 936,  DOI: 10.1021/ac60169a028
    307. 307
      Zhu, P.; Lin, S.-M. MBTH for aliphatic aldehyde measurement, U.S. Patent US7112448B2, September 26, 2006.
    308. 308
      Hauser, T. R.; Cummins, R. L. Increasing Sensitivity of 3-Methyl-2-Benzothiazolone Hydrozone Test for Analysis of Aliphatic Aldehydes in Air. Anal. Chem. 1964, 36 (3), 67981,  DOI: 10.1021/ac60209a067
    309. 309
      Dean Pakulski, J.; Benner, R. An Improved Method for the Hydrolysis and MBTH Analysis of Dissolved and Particulate Carbohydrates in Seawater. Mar. Chem. 1992, 40 (3), 14360,  DOI: 10.1016/0304-4203(92)90020-B
    310. 310
      Gomez, L. D.; Whitehead, C.; Barakate, A.; Halpin, C.; McQueen-Mason, S. J. Automated Saccharification Assay for Determination of Digestibility in Plant Materials. Biotechnol. Biofuels 2010, 3, 23,  DOI: 10.1186/1754-6834-3-23
    311. 311
      Oliveira, F. S. d.; Leite, B. C. O.; Andrade, M. V. A. S. d.; Korn, M. Determination of Total Aldehydes in Fuel Ethanol by MBTH Method: Sequential Injection Analysis. J. Braz. Chem. Soc. 2005, 16, 8792,  DOI: 10.1590/S0103-50532005000100013
    312. 312
      Suchý, M.; Lazurko, C.; Kirby, A.; Dang, T.; Liu, G.; Shuhendler, A. J. Methyl 5-MeO-N-Aminoanthranilate, a Minimalist Fluorogenic Probe for Sensing Cellular Aldehydic Load. Org. Biomol. Chem. 2019, 17 (7), 184353,  DOI: 10.1039/C8OB02255K
    313. 313
      Yuen, L. H.; Saxena, N. S.; Park, H. S.; Weinberg, K.; Kool, E. T. Dark Hydrazone Fluorescence Labeling Agents Enable Imaging of Cellular Aldehydic Load. ACS Chem. Biol. 2016, 11 (8), 23129,  DOI: 10.1021/acschembio.6b00269
    314. 314
      Liu, T.; Yang, L.; Zhang, J.; Liu, K.; Ding, L.; Peng, H.; Belfield, K. D.; Fang, Y. Squaraine-Hydrazine Adducts for Fast and Colorimetric Detection of Aldehydes in Aqueous Media. Sens. Actuators B Chem. 2019, 292, 8893,  DOI: 10.1016/j.snb.2019.04.138
    315. 315
      Liu, T.; Liu, X.; Wang, W.; Luo, Z.; Liu, M.; Zou, S.; Sissa, C.; Painelli, A.; Zhang, Y.; Vengris, M. Systematic Molecular Engineering of a Series of Aniline-Based Squaraine Dyes and Their Structure-Related Properties. J. Phys. Chem. C 2018, 122 (7), 39944008,  DOI: 10.1021/acs.jpcc.7b11997
    316. 316
      Xia, G.; Wang, H. Squaraine Dyes: The Hierarchical Synthesis and Its Application in Optical Detection. J. Photochem. Photobiol. C: Photochem. Rev. 2017, 31, 84113,  DOI: 10.1016/j.jphotochemrev.2017.03.001
    317. 317
      Lichtenstein, S. J.; Nettleton, G. S. Effects of Fuchsin Variants in Aldehyde Fuchsin Staining. J. Histochem. Cytochem. 1980, 28 (7), 6838,  DOI: 10.1177/28.7.6156202
    318. 318
      Robins, J. H.; Abrams, G. D.; Pincock, J. A. The Structure of Schiff Reagent Aldehyde Adducts and the Mechanism of the Schiff Reaction as Determined by Nuclear Magnetic Resonance Spectroscopy. Can. J. Chem. 1980, 58 (4), 33947,  DOI: 10.1139/v80-055
    319. 319
      Kitov, P. I.; Vinals, D. F.; Ng, S.; Tjhung, K. F.; Derda, R. Rapid, Hydrolytically Stable Modification of Aldehyde-Terminated Proteins and Phage Libraries. J. Am. Chem. Soc. 2014, 136 (23), 814952,  DOI: 10.1021/ja5023909
    320. 320
      Ressmann, A. K.; Schwendenwein, D.; Leonhartsberger, S.; Mihovilovic, M. D.; Bornscheuer, U. T.; Winkler, M.; Rudroff, F. Substrate-Independent High-Throughput Assay for the Quantification of Aldehydes. Adv. Synth. Catal. 2019, 361, 2538,  DOI: 10.1002/adsc.201900154
    321. 321
      Horvat, M.; Larch, T.-S.; Rudroff, F.; Winkler, M. Amino Benzamidoxime (ABAO)-Based Assay to Identify Efficient Aldehyde-Producing Pichia Pastoris Clones. Adv. Synth. Catal. 2020, 362 (21), 46739,  DOI: 10.1002/adsc.202000558
    322. 322
      Mei, Z.; Zhang, K.; Qu, G.; Li, J.-K.; Liu, B.; Ma, J.-A.; Tu, R.; Sun, Z. High-Throughput Fluorescence Assay for Ketone Detection and Its Applications in Enzyme Mining and Protein Engineering. ACS Omega 2020, 5 (23), 1358894,  DOI: 10.1021/acsomega.0c00245
    323. 323
      Xing, Y.; Wang, S.; Mao, X.; Zhao, X.; Wei, D. An Easy and Efficient Fluorescent Method for Detecting Aldehydes and Its Application in Biotransformation. J. Fluoresc. 2011, 21 (2), 58794,  DOI: 10.1007/s10895-010-0746-6
    324. 324
      Li, Z.; Xue, Z.; Wu, Z.; Han, J.; Han, S. Chromo-Fluorogenic Detection of Aldehydes with a Rhodamine Based Sensor Featuring an Intramolecular Deoxylactam. Org. Biomol. Chem. 2011, 9 (22), 76524,  DOI: 10.1039/c1ob06448g
    325. 325
      Davidson, D.; Weiss, M.; Jelling, M. The Action of Ammonia on Benzil. J. Org. Chem. 1937, 02 (4), 31927,  DOI: 10.1021/jo01227a004
    326. 326
      Nakashima, K.; Yamasaki, H.; Kuroda, N.; Akiyama, S. Evaluation of Lophine Derivatives as Chemiluminogens by a Flow-Injection Method. Anal. Chim. Acta 1995, 303 (1), 1037,  DOI: 10.1016/0003-2670(94)00360-X
    327. 327
      El-Maghrabey, M. H.; Kishikawa, N.; Ohyama, K.; Kuroda, N. Analytical Method for Lipoperoxidation Relevant Reactive Aldehydes in Human Sera by High-Performance Liquid Chromatography-Fluorescence Detection. Anal. Biochem. 2014, 464, 3642,  DOI: 10.1016/j.ab.2014.07.002
    328. 328
      Fathy Bakr Ali, M.; Kishikawa, N.; Ohyama, K.; Abdel-Mageed Mohamed, H.; Mohamed Abdel-Wadood, H.; Mohamed Mohamed, A.; Kuroda, N. Chromatographic Determination of Aliphatic Aldehydes in Human Serum after Pre-Column Derivatization Using 2,2′-Furil, a Novel Fluorogenic Reagent. J. Chromatogr. A 2013, 1300, 199203,  DOI: 10.1016/j.chroma.2013.03.033
    329. 329
      Nakamura, M.; Toda, M.; Saito, H.; Ohkura, Y. Fluorimetric Determination of Aromatic Aldehydes with 4,5-Dimethoxy-1,2-Diaminobenzene. Anal. Chim. Acta 1982, 134, 3945,  DOI: 10.1016/S0003-2670(01)84175-1
    330. 330
      Imazato, T.; Kanematsu, M.; Kishikawa, N.; Ohyama, K.; Hino, T.; Ueki, Y.; Maehata, E.; Kuroda, N. Determination of Acrolein in Serum by High-Performance Liquid Chromatography with Fluorescence Detection after Pre-Column Fluorogenic Derivatization Using 1,2-Diamino-4,5-Dimethoxybenzene. Biomed. Chromatogr. 2015, 29 (9), 13048,  DOI: 10.1002/bmc.3422
    331. 331
      McLellan, A. C.; Phillips, S. A.; Thornalley, P. J. The Assay of Methylglyoxal in Biological Systems Byderivatization with 1,2-Diamino-4,5-Dimethoxybenzene. Anal. Biochem. 1992, 206 (1), 1723,  DOI: 10.1016/S0003-2697(05)80005-3
    332. 332
      Hantzsch, A. Condensationsprodukte Aus Aldehydammoniak Und Ketonartigen Verbindungen. Ber. Dtsch. Chem. Ges. 1881, 14 (2), 16378,  DOI: 10.1002/cber.18810140214
    333. 333
      Belman, S. The Fluorimetric Determination of Formaldehyde. Anal. Chim. Acta 1963, 29, 1206,  DOI: 10.1016/S0003-2670(00)88591-8
    334. 334
      Sawicki, E.; Carnes, R. Spectrophotofluorimetric Determination of Aldehydes with Dimedone and Other Reagents. Mikrochim. Acta 1968, 56 (1), 14859,  DOI: 10.1007/BF01216118
    335. 335
      Compton, B. J.; Purdy, W. C. The Mechanism of the Reaction of the Nash and the Sawicki Aldehyde Reagent. Can. J. Chem. 1980, 58 (21), 220711,  DOI: 10.1139/v80-355
    336. 336
      Compton, B. J.; Purdy, W. C. Fluoral-P, a Member of a Selective Family of Reagents for Aldehydes. Anal. Chim. Acta 1980, 119 (2), 34957,  DOI: 10.1016/S0003-2670(01)93636-0
    337. 337
      Mopper, K.; Stahovec, W. L.; Johnson, L. Trace Analysis of Aldehydes by Reversed-Phase High-Performance Liquid Chromatography and Precolumn Fluorigenic Labeling with 5,5-Dimethyl-1,3-Cyclohexanedione. J. Chromatogr. A 1983, 256, 24352,  DOI: 10.1016/S0021-9673(01)88237-6
    338. 338
      Al-Moniee, M. A.; Koopal, C.; Akmal, N.; van Veen, S.; Zhu, X.; Sanders, P. F.; Sanders, P. F.; Al-Abeedi, F. N.; Amer, A. M. Applicability of Dimedone Assays for the Development of Online Aldehyde Sensor in Seawater Flooding Systems. J. Sens. Sci. Technol. 2016, 6, 101,  DOI: 10.4236/jst.2016.64008
    339. 339
      Li, Q.; Oshima, M.; Motomizu, S. Flow-Injection Spectrofluorometric Determination of Trace Amounts of Formaldehyde in Water after Derivatization with Acetoacetanilide. Talanta 2007, 72 (5), 167580,  DOI: 10.1016/j.talanta.2007.01.054
    340. 340
      Kolmel, D. K.; Kool, E. T. Oximes and Hydrazones in Bioconjugation: Mechanism and Catalysis. Chem. Rev. 2017, 117 (15), 1035876,  DOI: 10.1021/acs.chemrev.7b00090
    341. 341
      Kalia, J.; Raines, R. T. Hydrolytic Stability of Hydrazones and Oximes. Angew. Chem., Int. Ed. 2008, 47 (39), 75236,  DOI: 10.1002/anie.200802651
    342. 342
      Forget, D.; Boturyn, D.; Defrancq, E.; Lhomme, J.; Dumy, P. Highly Efficient Synthesis of Peptide-Oligonucleotide Conjugates: Chemoselective Oxime and Thiazolidine Formation. Eur. J. Chem. 2001, 7 (18), 397684,  DOI: 10.1002/1521-3765(20010917)7:18<3976::AID-CHEM3976>3.0.CO;2-X
    343. 343
      Shao, J.; Tam, J. P. Unprotected Peptides as Building Blocks for the Synthesis of Peptide Dendrimers with Oxime, Hydrazone, and Thiazolidine Linkages. J. Am. Chem. Soc. 1995, 117 (14), 38939,  DOI: 10.1021/ja00119a001
    344. 344
      Kemp, D. S.; Woodward, R. B. The N-Ethylbenzisoxazolium Cation─I: Preparation and Reactions with Nucleophilic Species. Tetrahedron 1965, 21 (11), 301935,  DOI: 10.1016/S0040-4020(01)96921-2
    345. 345
      Salahuddin, S.; Renaudet, O.; Reymond, J.-L. Aldehyde Detection by Chromogenic/Fluorogenic Oxime Bond Fragmentation. Org. Biomol. Chem. 2004, 2 (10), 14715,  DOI: 10.1039/b400314d
    346. 346
      Drienovská, I.; Mayer, C.; Dulson, C.; Roelfes, G. A Designer Enzyme for Hydrazone and Oxime Formation Featuring an Unnatural Catalytic Aniline Residue. Nat. Chem. 2018, 10 (9), 94652,  DOI: 10.1038/s41557-018-0082-z
    347. 347
      Horowitz, R. M.; Geissman, T. A. A Cleavage Reaction of α-Allylbenzylamines. J. Am. Chem. Soc. 1950, 72 (4), 151822,  DOI: 10.1021/ja01160a025
    348. 348
      Sakabe, M.; Asanuma, D.; Kamiya, M.; Iwatate, R. J.; Hanaoka, K.; Terai, T.; Nagano, T.; Urano, Y. Rational Design of Highly Sensitive Fluorescence Probes for Protease and Glycosidase Based on Precisely Controlled Spirocyclization. J. Am. Chem. Soc. 2013, 135 (1), 40914,  DOI: 10.1021/ja309688m
    349. 349
      Brewer, T. F.; Chang, C. J. An Aza-Cope Reactivity-Based Fluorescent Probe for Imaging Formaldehyde in Living Cells. J. Am. Chem. Soc. 2015, 137 (34), 108869,  DOI: 10.1021/jacs.5b05340
    350. 350
      Roth, A.; Li, H.; Anorma, C.; Chan, J. A Reaction-Based Fluorescent Probe for Imaging of Formaldehyde in Living Cells. J. Am. Chem. Soc. 2015, 137 (34), 108903,  DOI: 10.1021/jacs.5b05339
    351. 351
      Bruemmer, K. J.; Walvoord, R. R.; Brewer, T. F.; Burgos-Barragan, G.; Wit, N.; Pontel, L. B.; Patel, K. J.; Chang, C. J. Development of a General Aza-Cope Reaction Trigger Applied to Fluorescence Imaging of Formaldehyde in Living Cells. J. Am. Chem. Soc. 2017, 139 (15), 533850,  DOI: 10.1021/jacs.6b12460
    352. 352
      Du, Y.; Zhang, Y.; Huang, M.; Wang, S.; Wang, J.; Liao, K.; Wu, X.; Zhou, Q.; Zhang, X.; Wu, Y.-D. Systematic Investigation of the Aza-Cope Reaction for Fluorescence Imaging of Formaldehyde in Vitro and in Vivo. Chem. Sci. 2021, 12 (41), 1385769,  DOI: 10.1039/D1SC04387K
    353. 353
      He, L.; Yang, X.; Liu, Y.; Kong, X.; Lin, W. A Ratiometric Fluorescent Formaldehyde Probe for Bioimaging Applications. Chem. Commun. 2016, 52 (21), 402932,  DOI: 10.1039/C5CC09796G
    354. 354
      Xu, J.; Zhang, Y.; Zeng, L.; Liu, J.; Kinsella, J. M.; Sheng, R. A Simple Naphthalene-Based Fluorescent Probe for High Selective Detection of Formaldehyde in Toffees and HeLa Cells Via Aza-Cope Reaction. Talanta 2016, 160, 64552,  DOI: 10.1016/j.talanta.2016.08.010
    355. 355
      Yang, X.; He, L.; Xu, K.; Yang, Y.; Lin, W. The Development of an Ict-Based Formaldehyde-Responsive Fluorescence Turn-On Probe with a High Signal-to-Noise Ratio. New J. Chem. 2018, 42 (15), 123614,  DOI: 10.1039/C8NJ02467G
    356. 356
      Yang, M.; Fan, J.; Du, J.; Long, S.; Wang, J.; Peng, X. Imaging of Formaldehyde in Live Cells and Daphnia Magna Via Aza-Cope Reaction Utilizing Fluorescence Probe with Large Stokes Shifts. Front. Chem. 2018, 6, 488,  DOI: 10.3389/fchem.2018.00488
    357. 357
      Bruemmer, K. J.; Brewer, T. F.; Chang, C. J. Fluorescent Probes for Imaging Formaldehyde in Biological Systems. Curr. Opin. Chem. Biol. 2017, 39, 1723,  DOI: 10.1016/j.cbpa.2017.04.010
    358. 358
      Liu, X.; Li, N.; Li, M.; Chen, H.; Zhang, N.; Wang, Y.; Zheng, K. Recent Progress in Fluorescent Probes for Detection of Carbonyl Species: Formaldehyde, Carbon Monoxide and Phosgene. Coord. Chem. Rev. 2020, 404, 213109,  DOI: 10.1016/j.ccr.2019.213109
    359. 359
      Hanson, J. R. Terpenoids and Steroids, 7 ed.; Royal Society of Chemistry, 2007.
    360. 360
      Forney, F. W.; Markovetz, A. J. The Biology of Methyl Ketones. J. Lipid Res. 1971, 12 (4), 38395,  DOI: 10.1016/S0022-2275(20)39487-6
    361. 361
      McNally, M. A.; Hartman, A. L. Ketone Bodies in Epilepsy. J. Neurochem. 2012, 121 (1), 2835,  DOI: 10.1111/j.1471-4159.2012.07670.x
    362. 362
      Wakil, S. J.; Stoops, J. K.; Joshi, V. C. Fatty Acid Synthesis and Its Regulation. Annu. Rev. Biochem. 1983, 52 (1), 53779,  DOI: 10.1146/annurev.bi.52.070183.002541
    363. 363
      Dhillon, K. K.; Gupta, S. Biochemistry, Ketogenesis; StatPearls: Treasure Island, FL, 2021.
    364. 364
      Mutti, F. G.; Knaus, T. Enzymes Applied to the Synthesis of Amines. Biocatalysis for Practitioners 2021, 14380,  DOI: 10.1002/9783527824465.ch6
    365. 365
      Kroutil, W.; Mang, H.; Edegger, K.; Faber, K. Recent Advances in the Biocatalytic Reduction of Ketones and Oxidation of sec-Alcohols. Curr. Opin. Chem. Biol. 2004, 8 (2), 1206,  DOI: 10.1016/j.cbpa.2004.02.005
    366. 366
      Leisch, H.; Morley, K.; Lau, P. C. K. Baeyer- Villiger Monooxygenases: More Than Just Green Chemistry. Chem. Rev. 2011, 111 (7), 4165222,  DOI: 10.1021/cr1003437
    367. 367
      Dean, S. M.; Greenberg, W. A.; Wong, C. H. Recent Advances in Aldolase-Catalyzed Asymmetric Synthesis. Adv. Synth. Catal. 2007, 349 (8–9), 130820,  DOI: 10.1002/adsc.200700115
    368. 368
      Walsh, C. T. Biologically Generated Carbon Dioxide: Nature’s Versatile Chemical Strategies for Carboxy Lyases. Nat. Prod. Rep. 2020, 37 (1), 10035,  DOI: 10.1039/C9NP00015A
    369. 369
      Liu, M.; Wei, D.; Wen, Z.; Wang, J.-b. Progress in Stereoselective Construction of C-C Bonds Enabled by Aldolases and Hydroxynitrile Lyases. Front. Bioeng. Biotechnol. 2021, 9, 653682,  DOI: 10.3389/fbioe.2021.653682
    370. 370
      Müller, M.; Gocke, D.; Pohl, M. Thiamin Diphosphate in Biological Chemistry: Exploitation of Diverse Thiamin Diphosphate-Dependent Enzymes for Asymmetric Chemoenzymatic Synthesis. FEBS J. 2009, 276 (11), 2894904,  DOI: 10.1111/j.1742-4658.2009.07017.x
    371. 371
      Schmidt, N. G.; Pavkov-Keller, T.; Richter, N.; Wiltschi, B.; Gruber, K.; Kroutil, W. Biocatalytic Friedel-Crafts Acylation and Fries Reaction. Angew. Chem., Int. Ed. 2017, 56 (26), 76159,  DOI: 10.1002/anie.201703270
    372. 372
      Gergel, S.; Soler, J.; Klein, A.; Schülke, K.; Hauer, B.; Garcia-Borràs, M.; Hammer, S. Directed Evolution of a Ketone Synthase for Efficient and Highly Selective Functionalization of Internal Alkenes by Accessing Reactive Carbocation Intermediates. ChemRxiv 2022. dp94p DOI: 10.26434/chemrxiv-2022-dp94p
    373. 373
      Zhang, J.; Xu, T.; Li, Z. Enantioselective Biooxidation of Racemic trans-Cyclic Vicinal Diols: One-Pot Synthesis of Both Enantiopure (S,S)-Cyclic Vicinal Diols and (R)-α-Hydroxy Ketones. Adv. Synth. Catal. 2013, 355 (16), 314753,  DOI: 10.1002/adsc.201300301
    374. 374
      Lavandera, I.; Kern, A.; Resch, V.; Ferreira-Silva, B.; Glieder, A.; Fabian, W. M. F.; de Wildeman, S.; Kroutil, W. One-Way Biohydrogen Transfer for Oxidation of sec-Alcohols. Org. Lett. 2008, 10 (11), 21558,  DOI: 10.1021/ol800549f
    375. 375
      González-Granda, S.; Méndez-Sánchez, D.; Lavandera, I.; Gotor-Fernández, V. Laccase-Mediated Oxidations of Propargylic Alcohols. Application in the Deracemization of 1-Arylprop-2-Yn-1-Ols in Combination with Alcohol Dehydrogenases. ChemCatChem. 2020, 12 (2), 5207,  DOI: 10.1002/cctc.201901543
    376. 376
      Key, J. A.; Li, C.; Cairo, C. W. Detection of Cellular Sialic Acid Content Using Nitrobenzoxadiazole Carbonyl-Reactive Chromophores. Bioconj. Chem. 2012, 23 (3), 36371,  DOI: 10.1021/bc200276k
    377. 377
      Crisalli, P.; Kool, E. T. Water-Soluble Organocatalysts for Hydrazone and Oxime Formation. J. Org. Chem. 2013, 78 (3), 11849,  DOI: 10.1021/jo302746p
    378. 378
      Guo, H.-M.; Tanaka, F. A Fluorogenic Aldehyde Bearing a 1,2,3-Triazole Moiety for Monitoring the Progress of Aldol Reactions. J. Org. Chem. 2009, 74 (6), 241724,  DOI: 10.1021/jo900013w
    379. 379
      Tanaka, F.; Thayumanavan, R.; Barbas, C. F. Fluorescent Detection of Carbon-Carbon Bond Formation. J. Am. Chem. Soc. 2003, 125 (28), 85238,  DOI: 10.1021/ja034069t
    380. 380
      Tanaka, F.; Mase, N.; Barbas, C. F. Design and Use of Fluorogenic Aldehydes for Monitoring the Progress of Aldehyde Transformations. J. Am. Chem. Soc. 2004, 126 (12), 36923,  DOI: 10.1021/ja049641a
    381. 381
      Linares-Pastén, J. A.; Chávez-Lizárraga, G.; Villagomez, R.; Mamo, G.; Hatti-Kaul, R. A Method for Rapid Screening of Ketone Biotransformations: Detection of Whole Cell Baeyer-Villiger Monooxygenase Activity. Enzyme Microb. Technol. 2012, 50 (2), 1016,  DOI: 10.1016/j.enzmictec.2011.10.004
    382. 382
      Georgiana Ileana, B.; Gabriel Lucian, R. Carboxylic Acid - Key Role in Life Sciences; IntechOpen, 2018.
    383. 383
      Brune, K.; Patrignani, P. New Insights into the Use of Currently Available Non-Steroidal Anti-Inflammatory Drugs. J. Pain Res. 2015, 8, 10518,  DOI: 10.2147/JPR.S75160
    384. 384
      Murali, N.; Srinivas, K.; Ahring, B. K. Biochemical Production and Separation of Carboxylic Acids for Biorefinery Applications. Fermentation 2017, 3 (2), 22,  DOI: 10.3390/fermentation3020022
    385. 385
      Simmons, D. L.; Botting, R. M.; Hla, T. Cyclooxygenase Isozymes: The Biology of Prostaglandin Synthesis and Inhibition. Pharmacol. Rev. 2004, 56 (3), 387437,  DOI: 10.1124/pr.56.3.3
    386. 386
      Kumar, P.; Dubey, K. K. Citric Acid Cycle Regulation: Back Bone for Secondary Metabolite Production. In New and Future Developments in Microbial Biotechnology and Bioengineering; Elsevier: Amsterdam, 2019; Chapter 13, pp 16581.
    387. 387
      Nakamura, M. T.; Yudell, B. E.; Loor, J. J. Regulation of Energy Metabolism by Long-Chain Fatty Acids. Prog. Lipid Res. 2014, 53, 12444,  DOI: 10.1016/j.plipres.2013.12.001
    388. 388
      Barrett, G. C.; Elmore, D. T. Amino Acids and Peptides; Cambridge University Press: Cambridge, 1998 DOI: 10.1017/CBO9781139163828 .
    389. 389
      Ogliaruso, M. A.; Wolfe, J. F. Synthesis of Carboxylic Acids. Esters and Their Derivatives; Wiley Online Library, 1979; Vol. 267 DOI: 10.1002/9780470771587.ch7 .
    390. 390
      Wang, H.; Fan, H.; Sun, H.; Zhao, L.; Wei, D. Process Development for the Production of (R)-(−)-Mandelic Acid by Recombinant Escherichia Coli Cells Harboring Nitrilase from Burkholderia Cenocepacia J2315. Org. Process Res. Dev. 2015, 19 (12), 20126,  DOI: 10.1021/acs.oprd.5b00269
    391. 391
      Martínková, L.; Rucká, L.; Nešvera, J.; Pátek, M. Recent Advances and Challenges in the Heterologous Production of Microbial Nitrilases for Biocatalytic Applications. World J. Microbiol. Biotechnol. 2017, 33, 8,  DOI: 10.1007/s11274-016-2173-6
    392. 392
      Xu, Z.; Xiong, N.; Zou, S.-P.; Liu, Y.-X.; liu, Z.-Q.; Xue, Y.-P.; Zheng, Y.-G. Highly Efficient Conversion of 1-Cyanocycloalkaneacetonitrile Using a “Super Nitrilase Mutant”. Bioprocess Biosyst. Eng. 2019, 42 (3), 45563,  DOI: 10.1007/s00449-018-2049-2
    393. 393
      Glieder, A.; Weis, R.; Skranc, W.; Poechlauer, P.; Dreveny, I.; Majer, S.; Wubbolts, M.; Schwab, H.; Gruber, K. Comprehensive Step-by-Step Engineering of an (R)-Hydroxynitrile Lyase for Large-Scale Asymmetric Synthesis. Angew. Chem., Int. Ed. 2003, 42 (39), 48158,  DOI: 10.1002/anie.200352141
    394. 394
      Jiang, M.; Liu, Y.-J.; Liu, G.-S.; Zhou, W.; Shen, Y.-L.; Gao, B.; Wang, F.-Q.; Wei, D.-Z. Sequential Resolution of (S) and (R)-6-Fluoro-Chroman-2-Carboxylic Acid by Two Esterases in Turn. Green Chem. 2022, 24 (8), 323542,  DOI: 10.1039/D1GC04512A
    395. 395
      Dijkman, W. P.; Groothuis, D. E.; Fraaije, M. W. Enzyme-Catalyzed Oxidation of 5-Hydroxymethylfurfural to Furan-2,5-Dicarboxylic Acid. Angew. Chem., Int. Ed. 2014, 53 (25), 65158,  DOI: 10.1002/anie.201402904
    396. 396
      DeSantis, G.; Wong, K.; Farwell, B.; Chatman, K.; Zhu, Z.; Tomlinson, G.; Huang, H.; Tan, X.; Bibbs, L.; Chen, P. Creation of a Productive, Highly Enantioselective Nitrilase through Gene Site Saturation Mutagenesis (GSSM). J. Am. Chem. Soc. 2003, 125 (38), 114767,  DOI: 10.1021/ja035742h
    397. 397
      Franz, R. G. Comparisons of pKa and Log P Values of Some Carboxylic and Phosphonic Acids: Synthesis and Measurement. AAPS PharmSci. 2001, 3 (2), 1,  DOI: 10.1208/ps030210
    398. 398
      Toth, A. M.; Liptak, M. D.; Phillips, D. L.; Shields, G. C. Accurate Relative Pka Calculations for Carboxylic Acids Using Complete Basis Set and Gaussian-N Models Combined with Continuum Solvation Methods. J. Chem. Phys. 2001, 114 (10), 4595606,  DOI: 10.1063/1.1337862
    399. 399
      Vollhardt, K. P. C. Organische Chemie, 3rd ed.; Wiley-VCH: Weinheim, 2000.
    400. 400
      Bruice, P. Y. Organic Chemistry, 5 ed.; Pearson Prentice Hall: Upper Saddle River, NJ, 2007.
    401. 401
      Descalzo, A. B.; Rurack, K.; Weisshoff, H.; Martínez-Máñez, R.; Marcos, M. D.; Amorós, P.; Hoffmann, K.; Soto, J. Rational Design of a Chromo- and Fluorogenic Hybrid Chemosensor Material for the Detection of Long-Chain Carboxylates. J. Am. Chem. Soc. 2005, 127 (1), 184200,  DOI: 10.1021/ja045683n
    402. 402
      Reddy G, U.; Lo, R.; Roy, S.; Banerjee, T.; Ganguly, B.; Das, A. A New Receptor with a FRET Based Fluorescence Response for Selective Recognition of Fumaric and Maleic Acids in Aqueous Medium. Chem. Commun. 2013, 49 (84), 9818,  DOI: 10.1039/c3cc45051a
    403. 403
      Fitzmaurice, R. J.; Kyne, G. M.; Douheret, D.; Kilburn, J. D. Synthetic Receptors for Carboxylic Acids and Carboxylates. J. Chem. Soc., Perkin Trans. 1 2002, (7), 84164,  DOI: 10.1039/b009041g
    404. 404
      Janes, L. E.; Löwendahl, A. C.; Kazlauskas, R. J. Quantitative Screening of Hydrolase Libraries Using pH Indicators: Identifying Active and Enantioselective Hydrolases. Eur. J. Chem. 1998, 4 (11), 232431,  DOI: 10.1002/(SICI)1521-3765(19981102)4:11<2324::AID-CHEM2324>3.0.CO;2-I
    405. 405
      Baumann, M.; Stürmer, R.; Bornscheuer, U. T. A High-Throughput-Screening Method for the Identification of Active and Enantioselective Hydrolases. Angew. Chem., Int. Ed. 2001, 40 (22), 42014,  DOI: 10.1002/1521-3773(20011119)40:22<4201::AID-ANIE4201>3.0.CO;2-V
    406. 406
      Wang, J.; Liu, H.-B.; Tong, Z.; Ha, C.-S. Fluorescent/Luminescent Detection of Natural Amino Acids by Organometallic Systems. Coord. Chem. Rev. 2015, 303, 13984,  DOI: 10.1016/j.ccr.2015.05.008
    407. 407
      Gong, R.; Mu, H.; Sun, Y.; Fang, X.; Xue, P.; Fu, E. The First Fluorescent Sensor for Medium-Chain Fatty Acids in Water: Design, Synthesis and Sensing Properties of an Organic-Inorganic Hybrid Material. J. Mater. Chem. B 2013, 1 (15), 203847,  DOI: 10.1039/c3tb00355h
    408. 408
      Kusukawa, T.; Tanaka, S.; Inoue, K. Fluorescent Detection of Amidinium-Carboxylate and Amidinium Formation Using a 1,8-Diphenylnaphthalene-Based Diamidine: Dicarboxylic Acid Recognition with High Fluorescence Efficiency. Tetrahedron 2014, 70 (26), 404956,  DOI: 10.1016/j.tet.2014.03.036
    409. 409
      Boiocchi, M.; Bonizzoni, M.; Fabbrizzi, L.; Piovani, G.; Taglietti, A. A Dimetallic Cage with a Long Ellipsoidal Cavity for the Fluorescent Detection of Dicarboxylate Anions in Water. Angew. Chem., Int. Ed. 2004, 43 (29), 384752,  DOI: 10.1002/anie.200460036
    410. 410
      Metzger, A.; Lynch, V. M.; Anslyn, E. V. A Synthetic Receptor Selective for Citrate. Angew. Chem., Int. Ed. 1997, 36 (8), 8625,  DOI: 10.1002/anie.199708621
    411. 411
      Yang, X.; Liu, X.; Shen, K.; Zhu, C.; Cheng, Y. A Chiral Perazamacrocyclic Fluorescent Sensor for Cascade Recognition of Cu(II) and the Unmodified α-Amino Acids in Protic Solutions. Org. Lett. 2011, 13 (13), 35103,  DOI: 10.1021/ol2013268
    412. 412
      Görög, S.; Laukó, A.; Rényei, M.; Hegedüs, B. New Derivatization Reactions in Pharmaceutical Analysis. J. Pharm. Biomed. Anal. 1983, 1 (4), 497506,  DOI: 10.1016/0731-7085(83)80063-6
    413. 413
      Munson, J. W.; Bilous, R. Colorimetric Determination of Aliphatic Acids. J. Pharm. Sci. 1977, 66 (10), 14035,  DOI: 10.1002/jps.2600661013
    414. 414
      Miwa, H.; Yamamoto, M.; Momose, T. Colorimetric Detection and Determination of Carboxylic Acids with 2-Nitrophenylhydrazine Hydrochloride. Chem. Pharm. Bull. (Tokyo) 1980, 28 (2), 599605,  DOI: 10.1248/cpb.28.599
    415. 415
      Hill, U. T. Colorimetric Determination of Fatty Acids and Esters. Ind. Eng. Chem. Anal. Ed. 1946, 18 (5), 3179,  DOI: 10.1021/i560153a017
    416. 416
      Montgomery, H. A. C.; Dymock, J. F.; Thom, N. S. The Rapid Colorimetric Determination of Organic Acids and Their Salts in Sewage-Sludge Liquor. Analyst 1962, 87 (1041), 94955,  DOI: 10.1039/an9628700949
    417. 417
      Kasai, Y.; Tanimura, T.; Tamura, Z. J. A. C. Spectrophotometric Determination of Carboxylic Acids by the Formation of Hydroxamic Acids with Dicyclohexylcarbodiimide. Anal. Chem. 1975, 47 (1), 347,  DOI: 10.1021/ac60351a047
    418. 418
      Takeuchi, T.; Horikawa, R.; Tanimura, T. Spectrophotometry Determination of Carboxylic Acids by Ferric Hydroxamate Formation with Water-Soluble Carbodiimide. Anal. Lett. 1980, 13 (7), 6039,  DOI: 10.1080/00032718008077690
    419. 419
      He, Y.-C.; Ma, C.-L.; Xu, J.-H.; Zhou, L. A High-Throughput Screening Strategy for Nitrile-Hydrolyzing Enzymes Based on Ferric Hydroxamate Spectrophotometry. Appl. Microbiol. Biotechnol. 2011, 89 (3), 81723,  DOI: 10.1007/s00253-010-2977-5
    420. 420
      Goswami, S.; Hazra, A.; Das, M. K. Selenodiazole-Fused Diacetamidopyrimidine, a Selective Fluorescence Sensor for Aliphatic Monocarboxylates. Tetrahedron Lett. 2010, 51 (25), 33203,  DOI: 10.1016/j.tetlet.2010.04.085
    421. 421
      Goswami, S.; Hazra, A.; Chakrabarty, R.; Fun, H.-K. Recognition of Carboxylate Anions and Carboxylic Acids by Selenium-Based New Chromogenic Fluorescent Sensor: A Remarkable Fluorescence Enhancement of Hindered Carboxylates. Org. Lett. 2009, 11 (19), 43503,  DOI: 10.1021/ol901737s
    422. 422
      Galindo, F.; Becerril, J.; Isabel Burguete, M.; Luis, S. V.; Vigara, L. Synthesis and Study of a Cyclophane Displaying Dual Fluorescence Emission: A Novel Ratiometric Sensor for Carboxylic Acids in Organic Medium. Tetrahedron Lett. 2004, 45 (8), 165962,  DOI: 10.1016/j.tetlet.2003.12.116
    423. 423
      Cabell, L. A.; Best, M. D.; Lavigne, J. J.; Schneider, S. E.; Perreault, D. M.; Monahan, M.-K.; Anslyn, E. V. Metal Triggered Fluorescence Sensing of Citrate Using a Synthetic Receptor. J. Chem. Soc., Perkin trans. II 2001, (3), 31523,  DOI: 10.1039/b008694k
    424. 424
      Xu, K.-x.; Xie, X.-m.; Kong, H.-j.; Li, P.; Zhang, J.-l.; Pang, X.-b. Selective Fluorescent Sensors for Malate Anion Using the Complex of Phenanthroline-Based Eu(III) in Aqueous Solution. Sens. Actuators B Chem. 2014, 201, 1317,  DOI: 10.1016/j.snb.2014.04.086
    425. 425
      Burguete, M. I.; Galindo, F.; Luis, S. V.; Vigara, L. A Turn-On Fluorescent Indicator for Citrate with Micromolar Sensitivity. Dalton Trans. 2007, (36), 402733,  DOI: 10.1039/b711139h
    426. 426
      Klein, G.; Reymond, J.-L. An Enzyme Assay Using pM. Angew. Chem., Int. Ed. 2001, 40 (9), 17713,  DOI: 10.1002/1521-3773(20010504)40:9<1771::AID-ANIE17710>3.0.CO;2-M
    427. 427
      Dean, K. E. S.; Klein, G.; Renaudet, O.; Reymond, J.-L. A Green Fluorescent Chemosensor for Amino Acids Provides a Versatile High-Throughput Screening (HTS) Assay for Proteases. Bioorg. Med. Chem. Lett. 2003, 13 (10), 16536,  DOI: 10.1016/S0960-894X(03)00280-4
    428. 428
      Xu, J.-M.; Fu, F.-T.; Hu, H.-F.; Zheng, Y.-G. A High-Throughput Screening Method for Amino Acid Dehydrogenase. Anal. Biochem. 2016, 495, 2931,  DOI: 10.1016/j.ab.2015.11.012
    429. 429
      Ryu, D.; Park, E.; Kim, D.-S.; Yan, S.; Lee, J. Y.; Chang, B.-Y.; Ahn, K. H. A Rational Approach to Fluorescence “Turn-On” Sensing of α-Amino-Carboxylates. J. Am. Chem. Soc. 2008, 130 (8), 23945,  DOI: 10.1021/ja078308e
    430. 430
      Kim, D.-S.; Ahn, K. H. Fluorescence “Turn-On” Sensing of Carboxylate Anions with Oligothiophene-Based o-(Carboxamido)Trifluoroacetophenones. J. Org. Chem. 2008, 73 (17), 68314,  DOI: 10.1021/jo801178y
    431. 431
      Schrader, J.; Etschmann, M. M. W.; Sell, D.; Hilmer, J. M.; Rabenhorst, J. Applied Biocatalysis for the Synthesis of Natural Flavour Compounds - Current Industrial Processes and Future Prospects. Biotechnol. Lett. 2004, 26 (6), 46372,  DOI: 10.1023/B:BILE.0000019576.80594.0e
    432. 432
      Gotor, V. Biocatalysis Applied to the Preparation of Pharmaceuticals. Org. Process Res. Dev. 2002, 6 (4), 4206,  DOI: 10.1021/op020008o
    433. 433
      SÁ, A. G. A.; Meneses, A. C. d.; Araújo, P. H. H. d.; Oliveira, D. d. A Review on Enzymatic Synthesis of Aromatic Esters Used as Flavor Ingredients for Food, Cosmetics and Pharmaceuticals Industries. Trends Food Sci. Technol. 2017, 69, 95105,  DOI: 10.1016/j.tifs.2017.09.004
    434. 434
      Panke, S.; Held, M.; Wubbolts, M. Trends and Innovations in Industrial Biocatalysis for the Production of Fine Chemicals. Curr. Opin. Biotechnol. 2004, 15 (4), 2729,  DOI: 10.1016/j.copbio.2004.06.011
    435. 435
      Joshi, A.; Singh, S.; Iqbal, Z.; De, S. R. CO Free Esterifications of (Hetero)Arenes Via Transition-Metal-Catalyzed Chelation-Induced C-H Activation: Recent Updates. Tetrahedron 2022, 104, 132601,  DOI: 10.1016/j.tet.2021.132601
    436. 436
      Larios, A.; García, H. S.; Oliart, R. M.; Valerio-Alfaro, G. Synthesis of Flavor and Fragrance Esters Using Candida Antarctica Lipase. Appl. Microbiol. Biotechnol. 2004, 65 (4), 3736,  DOI: 10.1007/s00253-004-1602-x
    437. 437
      Smith, G. A. Fatty Acid, Methyl Ester, and Vegetable Oil Ethoxylates, Biobased Surfactants; AOCS Press, 2019; pp 287301.
    438. 438
      Panten, J.; Surburg, H. Flavors and Fragrances, 4. Natural Raw Materials, Ullmann’s Encyclopedia of Industrial Chemistry , 2011; pp 158.
    439. 439
      Johannes, P.; Surburg, H. Flavors and Fragrances, 2. Aliphatic Compounds. In Ullmann’s Encyclopedia of Industrial Chemistry , 2015; pp 155.
    440. 440
      Berger, R. G. Biotechnology of Flavours─the Next Generation. Biotechnol. Lett. 2009, 31 (11), 1651,  DOI: 10.1007/s10529-009-0083-5
    441. 441
      Gunatillake, P. A.; Adhikari, R.; Gadegaard, N. Biodegradable Synthetic Polymers for Tissue Engineering. Eur. Cell. Mater. 2003, 5 (1), 116,  DOI: 10.22203/eCM.v005a01
    442. 442
      Kodali, D. R. High Performance Ester Lubricants from Natural Oils. Ind. Lubr. Tribol. 2002, 54 (4), 16570,  DOI: 10.1108/00368790210431718
    443. 443
      Chen, J.; Wang, Y.; Huang, J.; Li, K.; Nie, X. Synthesis of Tung-Oil-Based Triglycidyl Ester Plasticizer and Its Effects on Poly(Vinyl Chloride) Soft Films. ACS Sustain. Chem. Eng. 2018, 6 (1), 64251,  DOI: 10.1021/acssuschemeng.7b02989
    444. 444
      Gumel, A. M.; Annuar, M. S. M.; Heidelberg, T.; Chisti, Y. Lipase Mediated Synthesis of Sugar Fatty Acid Esters. Process Biochem. 2011, 46 (11), 207990,  DOI: 10.1016/j.procbio.2011.07.021
    445. 445
      Badgujar, K. C.; Bhanage, B. M. Application of Lipase Immobilized on the Biocompatible Ternary Blend Polymer Matrix for Synthesis of Citronellyl Acetate in Non-Aqueous Media: Kinetic Modelling Study. Enzyme Microb. Technol. 2014, 57, 1625,  DOI: 10.1016/j.enzmictec.2014.01.006
    446. 446
      Stergiou, P.-Y.; Foukis, A.; Filippou, M.; Koukouritaki, M.; Parapouli, M.; Theodorou, L. G.; Hatziloukas, E.; Afendra, A.; Pandey, A.; Papamichael, E. M. Advances in Lipase-Catalyzed Esterification Reactions. Biotechnol. Adv. 2013, 31 (8), 184659,  DOI: 10.1016/j.biotechadv.2013.08.006
    447. 447
      Gao, W.; Wu, K.; Chen, L.; Fan, H.; Zhao, Z.; Gao, B.; Wang, H.; Wei, D. A Novel Esterase from a Marine Mud Metagenomic Library for Biocatalytic Synthesis of Short-Chain Flavor Esters. Microbial Cell Factories 2016, 15, 41,  DOI: 10.1186/s12934-016-0435-5
    448. 448
      Lozano, P.; Bernal, J. M.; Navarro, A. A Clean Enzymatic Process for Producing Flavour Esters by Direct Esterification in Switchable Ionic Liquid/Solid Phases. Green Chem. 2012, 14 (11), 302633,  DOI: 10.1039/c2gc36081k
    449. 449
      Otera, J.; Nishikido, J. Esterification: Methods, Reactions, and Applications; John Wiley & Sons, 2009. DOI: 10.1002/9783527627622
    450. 450
      Beekwilder, J.; Alvarez-Huerta, M.; Neef, E.; Verstappen, F. W. A.; Bouwmeester, H. J.; Aharoni, A. Functional Characterization of Enzymes Forming Volatile Esters from Strawberry and Banana. Plant Physiol. 2004, 135 (4), 186578,  DOI: 10.1104/pp.104.042580
    451. 451
      Buckles, R. E.; Thelen, C. Qualitative Determination of Carboxylic Esters. Anal. Chem. 1950, 22 (5), 6768,  DOI: 10.1021/ac60041a016
    452. 452
      Stern, I.; Shapiro, B. A Rapid and Simple Method for the Determination of Esterified Fatty Acids and for Total Fatty Acids in Blood. J. Clin. Pathol. 1953, 6 (2), 15860,  DOI: 10.1136/jcp.6.2.158
    453. 453
      Davidson, D. Hydroxamic Acids in Qualitative Organic Analysis. J. Chem. Educ. 1940, 17 (2), 81,  DOI: 10.1021/ed017p81
    454. 454
      Löbs, A.-K.; Lin, J.-L.; Cook, M.; Wheeldon, I. High Throughput, Colorimetric Screening of Microbial Ester Biosynthesis Reveals High Ethyl Acetate Production from Kluyveromyces Marxianus on C5, C6, and C12 Carbon Sources. Biotechnol. J. 2016, 11 (10), 127481,  DOI: 10.1002/biot.201600060
    455. 455
      Smith, J. G. Synthetically Useful Reactions of Epoxides. Synthesis 1984, 1984, 629656,  DOI: 10.1055/s-1984-30921
    456. 456
      Yudin, A. K. Aziridines and Epoxides in Organic Synthesis; John Wiley & Sons, 2006.
    457. 457
      Kotik, M.; Archelas, A.; Wohlgemuth, R. Epoxide Hydrolases and Their Application in Organic Synthesis. Curr. Org. Chem. 2012, 16 (4), 45182,  DOI: 10.2174/138527212799499840
    458. 458
      Zocher, F.; Enzelberger, M. M.; Bornscheuer, U. T.; Hauer, B. D.; Schmid, R. D. A Colorimetric Assay Suitable for Screening Epoxide Hydrolase Activity. Anal. Chim. Acta 1999, 391 (3), 345351,  DOI: 10.1016/S0003-2670(99)00216-0
    459. 459
      Alcalde, M.; Farinas, E. T.; Arnold, F. H. Colorimetric High-Throughput Assay for Alkene Epoxidation Catalyzed by Cytochrome P450 BM-3 Variant 139–3. J. Biomol. Screen. 2004, 9 (2), 1416,  DOI: 10.1177/1087057103261913
    460. 460
      Van Damme, F.; Oomens, A. C. Determination of Residual Free Epoxide in Polyether Polyols by Derivatization with Diethylammonium N,N-Diethyldithiocarbamate and Liquid Chromatography. J. Chromatogr. A 1995, 696 (1), 417,  DOI: 10.1016/0021-9673(94)01270-O
    461. 461
      Hammock, L. G.; Hammock, B. D.; Casida, J. E. Detection and Analysis of Epoxides with 4-(p-Nitrobenzyl)-Pyridine. Bull. Environ. Contam. Toxicol. 1974, 12 (6), 75964,  DOI: 10.1007/BF01685927
    462. 462
      Pedragosa-Moreau, S.; Morisseau, C.; Baratti, J.; Zylber, J.; Archelas, A.; Furstoss, R. Microbiological Transformations 37. An Enantioconvergent Synthesis of the β-Blocker (R)-NiféNalol Using a Combined Chemoenzymatic Approach. Tetrahedron 1997, 53 (28), 970714,  DOI: 10.1016/S0040-4020(97)00639-X
    463. 463
      Finney, N. S. Enantioselective Epoxide Hydrolysis: Catalysis Involving Microbes, Mammals and Metals. J. Biol. Chem. 1998, 5 (4), R73R9,  DOI: 10.1016/S1074-5521(98)90630-5
    464. 464
      Choi, W. J.; Choi, C. Y. Production of Chiral Epoxides: Epoxide Hydrolase-Catalyzed Enantioselective Hydrolysis. Biotechnol. Bioprocess Eng. 2005, 10 (3), 167,  DOI: 10.1007/BF02932009
    465. 465
      Baer, H.; Bergamo, M.; Forlin, A.; Pottenger, L. H.; Lindner, J. Propylene Oxide. In Ullmann’s Encyclopedia of Industrial Chemistry: Wiley-VCH, 2000.
    466. 466
      Rebsdat, S.; Mayer, D. Ethylene Oxide. In Ullmann’s Encyclopedia of Industrial Chemistry; Wiley-VCH, 2001.
    467. 467
      Archelas, A.; Furstoss, R. Synthesis of Enantiopure Epoxides through Biocatalytic Approaches. Annu. Rev. Microbiol. 1997, 51 (1), 491525,  DOI: 10.1146/annurev.micro.51.1.491
    468. 468
      Ma, S. K.; Gruber, J.; Davis, C.; Newman, L.; Gray, D.; Wang, A.; Grate, J.; Huisman, G. W.; Sheldon, R. A. A Green-by-Design Biocatalytic Process for Atorvastatin Intermediate. Green Chem. 2010, 12 (1), 816,  DOI: 10.1039/B919115C
    469. 469
      Kong, X.-D.; Ma, Q.; Zhou, J.; Zeng, B.-B.; Xu, J.-H. A Smart Library of Epoxide Hydrolase Variants and the Top Hits for Synthesis of (S)-β-Blocker Precursors. Angew. Chem., Int. Ed. Engl. 2014, 53 (26), 66414,  DOI: 10.1002/anie.201402653
    470. 470
      Lin, Z.; Xue, Y.; Liang, X.-W.; Wang, J.; Lin, S.; Tao, J.; You, S.-L.; Liu, W. Oxidative Indole Dearomatization for Asymmetric Furoindoline Synthesis by a Flavin-Dependent Monooxygenase Involved in the Biosynthesis of Bicyclic Thiopeptide Thiostrepton. Angew. Chem. 2021, 60 (15), 84015,  DOI: 10.1002/anie.202013174
    471. 471
      Agarwal, S. C.; Van Duuren, B. L.; Solomon, J. J.; Kline, S. A. Reaction of Epoxides with 4-Nitrothiophenol. Its Possible Application for Trapping and Characterization of Epoxides. Environ. Sci. Technol. 1980, 14 (10), 124953,  DOI: 10.1021/es60170a012
    472. 472
      Wixtrom, R. N.; Hammock, B. D. Continuous Spectrophotometric Assays for Cytosolic Epoxide Hydrolase. Anal. Biochem. 1988, 174 (1), 2919,  DOI: 10.1016/0003-2697(88)90548-9
    473. 473
      González-Pérez, M.; Gómez-Bombarelli, R.; Arenas-Valgañón, J.; Pérez-Prior, M. T.; García-Santos, M. P.; Calle, E.; Casado, J. Connecting the Chemical and Biological Reactivity of Epoxides. Chem. Res. Toxicol. 2012, 25 (12), 275562,  DOI: 10.1021/tx300389z
    474. 474
      Nelis, H. J. C. F.; Sinsheimer, J. E. A Sensitive Fluorimetric Procedure for the Determination of Aliphatic Epoxides under Physiological Conditions. Anal. Biochem. 1981, 115 (1), 1517,  DOI: 10.1016/0003-2697(81)90538-8
    475. 475
      Rink, R.; Fennema, M.; Smids, M.; Dehmel, U.; Janssen, D. B. Primary Structure and Catalytic Mechanism of the Epoxide Hydrolase from Agrobacterium Radiobacter Ad1. J. Biol. Chem. 1997, 272 (23), 146507,  DOI: 10.1074/jbc.272.23.14650
    476. 476
      Elder, D. P.; Snodin, D.; Teasdale, A. Analytical Approaches for the Detection of Epoxides and Hydroperoxides in Active Pharmaceutical Ingredients, Drug Products and Herbals. J. Pharm. Biomed. Anal. 2010, 51 (5), 101523,  DOI: 10.1016/j.jpba.2009.11.023
    477. 477
      Sano, A.; Takitani, S. Spectrofluorometric Determination of Common Epoxides with Sodium Sulfide and o-Phthalaldehyde and Taurine Reagents. Anal. Chem. 1985, 57 (8), 168790,  DOI: 10.1021/ac00285a039
    478. 478
      Tang, L.; Li, Y.; Wang, X. A High-Throughput Colorimetric Assay for Screening Halohydrin Dehalogenase Saturation Mutagenesis Libraries. J. Biotechnol. 2010, 147 (3), 1648,  DOI: 10.1016/j.jbiotec.2010.04.002
    479. 479
      Gul, I.; Bogale, T. F.; Deng, J.; Chen, Y.; Fang, R.; Feng, J.; Tang, L. Enzyme-Based Detection of Epoxides Using Colorimetric Assay Integrated with Smartphone Imaging. Biotechnol. Appl. Biochem. 2020, 67 (4), 68592,  DOI: 10.1002/bab.1898
    480. 480
      Dansette, P. M.; DuBois, G. C.; Jerina, D. M. Continuous Fluorometric Assay of Epoxide Hydrase Activity. Anal. Biochem. 1979, 97 (2), 3405,  DOI: 10.1016/0003-2697(79)90083-6
    481. 481
      Mateo, C.; Archelas, A.; Furstoss, R. A Spectrophotometric Assay for Measuring and Detecting an Epoxide Hydrolase Activity. Anal. Biochem. 2003, 314 (1), 13541,  DOI: 10.1016/S0003-2697(02)00646-2
    482. 482
      Michałowicz, J.; Duda, W. Phenols - Sources and Toxicity. Poish J. Environ. Stud. 2007, 16 (3), 347362
    483. 483
      Schmidt, R. J. Industrial Catalytic Processes─Phenol Production. Appl. Catal. A: Gen. 2005, 280 (1), 89103,  DOI: 10.1016/j.apcata.2004.08.030
    484. 484
      Ullrich, R.; Hofrichter, M. Enzymatic Hydroxylation of Aromatic Compounds. Cell. Mol. Life Sci. 2007, 64 (3), 27193,  DOI: 10.1007/s00018-007-6362-1
    485. 485
      Lan Tee, K.; Schwaneberg, U. Directed Evolution of Oxygenases: Screening Systems, Success Stories and Challenges. Comb. Chem. High Throughput Screen. 2007, 10 (3), 197217,  DOI: 10.2174/138620707780126723
    486. 486
      Joern, J. M.; Sakamoto, T.; Arisawa, A.; Arnold, F. H. A Versatile High Throughput Screen for Dioxygenase Activity Using Solid-Phase Digital Imaging. J. Biomol. Screen. 2001, 6 (4), 21923,  DOI: 10.1177/108705710100600403
    487. 487
      Poraj-Kobielska, M.; Kinne, M.; Ullrich, R.; Scheibner, K.; Kayser, G.; Hammel, K. E.; Hofrichter, M. Preparation of Human Drug Metabolites Using Fungal Peroxygenases. Biochem. Pharmacol. 2011, 82 (7), 78996,  DOI: 10.1016/j.bcp.2011.06.020
    488. 488
      Sakamoto, T.; Joern, J. M.; Arisawa, A.; Arnold, F. H. Laboratory Evolution of Toluene Dioxygenase to Accept 4-Picoline as a Substrate. Appl. Environ. Microbiol. 2001, 67 (9), 3882,  DOI: 10.1128/AEM.67.9.3882-3887.2001
    489. 489
      Foti, M. C. Antioxidant Properties of Phenols. J. Pharm. Pharmacol. 2010, 59 (12), 167385,  DOI: 10.1211/jpp.59.12.0010
    490. 490
      del Rosario Garcia-Mateos, M.; Palma-Tenango, M.; Soto-Hernandez, M. Phenolic Compounds-Biological Activity; IntechOpen, 2017.
    491. 491
      Ainsworth, E. A.; Gillespie, K. M. Estimation of Total Phenolic Content and Other Oxidation Substrates in Plant Tissues Using Folin-Ciocalteu Reagent. Nat. Protoc. 2007, 2 (4), 8757,  DOI: 10.1038/nprot.2007.102
    492. 492
      Singleton, V. L.; Orthofer, R.; Lamuela-Raventós, R. M. Analysis of Total Phenols and Other Oxidation Substrates and Antioxidants by Means of Folin–Ciocalteu Reagent. Methods in Enzymology; Academic Press, 1999; Vol. 299; p 152178. DOI: 10.1016/S0076-6879(99)99017-1
    493. 493
      Platzer, M.; Kiese, S.; Herfellner, T.; Schweiggert-Weisz, U.; Eisner, P. How Does the Phenol Structure Influence the Results of the Folin-Ciocalteu Assay?. Antioxidants 2021, 10 (5), 811,  DOI: 10.3390/antiox10050811
    494. 494
      Miliauskas, G.; Venskutonis, P. R.; van Beek, T. A. Screening of Radical Scavenging Activity of Some Medicinal and Aromatic Plant Extracts. Food Chem. 2004, 85 (2), 2317,  DOI: 10.1016/j.foodchem.2003.05.007
    495. 495
      Cao, G.; Alessio, H. M.; Cutler, R. G. Oxygen-Radical Absorbance Capacity Assay for Antioxidants. Free Radical Biol. Med. 1993, 14 (3), 30311,  DOI: 10.1016/0891-5849(93)90027-R
    496. 496
      Prior, R. L.; Wu, X.; Schaich, K. Standardized Methods for the Determination of Antioxidant Capacity and Phenolics in Foods and Dietary Supplements. J. Agric. Food Chem. 2005, 53 (10), 4290302,  DOI: 10.1021/jf0502698
    497. 497
      van den Berg, R.; Haenen, G. R. M. M.; van den Berg, H.; Bast, A. Applicability of an Improved Trolox Equivalent Antioxidant Capacity (TEAC) Assay for Evaluation of Antioxidant Capacity Measurements of Mixtures. Food Chem. 1999, 66 (4), 5117,  DOI: 10.1016/S0308-8146(99)00089-8
    498. 498
      Dudonné, S.; Vitrac, X.; Coutière, P.; Woillez, M.; Mérillon, J.-M. Comparative Study of Antioxidant Properties and Total Phenolic Content of 30 Plant Extracts of Industrial Interest Using DPPH, ABTS, FRAP, SOD, and ORAC Assays. J. Agric. Food Chem. 2009, 57 (5), 176874,  DOI: 10.1021/jf803011r
    499. 499
      Thaipong, K.; Boonprakob, U.; Crosby, K.; Cisneros-Zevallos, L.; Hawkins Byrne, D. Comparison of ABTS, DPPH, FRAP, and ORAC Assays for Estimating Antioxidant Activity from Guava Fruit Extracts. J. Food Compos. Anal. 2006, 19 (6), 66975,  DOI: 10.1016/j.jfca.2006.01.003
    500. 500
      da Silva Granja, B.; Honório de Mendonça Filho, J. R.; de Souza Oliveira, W.; Caldas Santos, J. C. Exploring MBTH as a Spectrophotometric Probe for the Determination of Total Phenolic Compounds in Beverage Samples. Anal. Methods 2018, 10 (19), 2197204,  DOI: 10.1039/C8AY00464A
    501. 501
      Wong, T. S.; Wu, N.; Roccatano, D.; Zacharias, M.; Schwaneberg, U. Sensitive Assay for Laboratory Evolution of Hydroxylases toward Aromatic and Heterocyclic Compounds. J. Biomol. Screen. 2005, 10 (3), 24652,  DOI: 10.1177/1087057104273336
    502. 502
      Emerson, E. The Condensation of Aminoantipyrine. II. A New Color Test for Phenolic Compounds. J. Org. Chem. 1943, 8 (5), 417428,  DOI: 10.1021/jo01193a004
    503. 503
      Johnson, C. A.; Savidge, R. A. The Determination of Phenolic Compounds in Pharmaceutical Preparations Using 4-Aminohenazone. J. Pharm. Pharmacol. 2011, 10, 171T81T,  DOI: 10.1111/j.2042-7158.1958.tb10396.x
    504. 504
      Varvounis, G.; Fiamegos, Y.; Pilidis, G. Pyrazol-3-Ones. Part III: Reactivity of the Ring Substituents. In Advances in Heterocyclic Chemistry; Academic Press, 2007; Vol. 95; p 27141
    505. 505
      Dannis, M. Determination of Phenols by the Amino-Antipyrine Method. Sewage Ind. Waste 1951, 23, 15161522
    506. 506
      Fiamegos, Y. C.; Stalikas, C. D.; Pilidis, G. A.; Karayannis, M. I. Synthesis and Analytical Applications of 4-Aminopyrazolone Derivatives as Chromogenic Agents for the Spectrophotometric Determination of Phenols. Anal. Chim. Acta 2000, 403 (1), 31523,  DOI: 10.1016/S0003-2670(99)00644-3
    507. 507
      Li, D.; Ge, S.; Huang, J.; Gong, J.; Yan, P.; Lu, W.; Tiana, G.; Ding, L. Fast Chromogenic Identification of Phenolic Pollutants Via Homogeneous Oxidation with t-BuOOH in the Presence of Iron (III) Octacarboxyphthalocyanine. Catal. Commun. 2014, 45, 959,  DOI: 10.1016/j.catcom.2013.10.038
    508. 508
      Fiamegos, Y.; Stalikas, C. D.; Pilidis, G. A. 4-Aminoantipyrine Spectrophotometric Method of Phenol Analysis: Study of the Reaction Products Via Liquid Chromatography with Diode-Array and Mass Spectrometric Detection. Anal. Chim. Acta 2002, 467 (1–2), 10514,  DOI: 10.1016/S0003-2670(02)00072-7
    509. 509
      Ayad, M.; El-Sadek, M.; Mostaffa, S. 4-Aminoantipyrine as an Analytical Reagent for the Colorimetric Determination of Tetracycline and Oxytetracycline. Anal. Lett. 1986, 19 (21–22), 216981,  DOI: 10.1080/00032718608080875
    510. 510
      Otey, C. R.; Joern, J. M. High-Throughput Screen for Aromatic Hydroxylation, Directed Enzyme Evolution: Screening and Selection Methods; Humana Press: Totowa, NJ, 2003; p 1418.
    511. 511
      Santos, G. d. A.; Dhoke, G. V.; Davari, M. D.; Ruff, A. J.; Schwaneberg, U. Directed Evolution of P450 BM3 Towards Functionalization of Aromatic O-Heterocycles. Int. J. Mol. Sci. 2019, 20 (13), 3353,  DOI: 10.3390/ijms20133353
    512. 512
      Lülsdorf, N.; Vojcic, L.; Hellmuth, H.; Weber, T. T.; Mußmann, N.; Martinez, R.; Schwaneberg, U. A First Continuous 4-Aminoantipyrine (4-AAP)-Based Screening System for Directed Esterase Evolution. Appl. Microbiol. Biotechnol. 2015, 99 (12), 523746,  DOI: 10.1007/s00253-015-6612-3
    513. 513
      Choi, S.-L.; Rha, E.-G.; Kim, D.-Y.; Song, J.-J.; Hong, S.-P.; Sung, M.-H.; Lee, S.-G. High Throughput Screening and Directed Evolution of Tyrosine Phenol-Lyase. Microbiol. Biotechnol. Lett. 2006, 34, 5862
    514. 514
      Quintana, M. G.; Didion, C.; Dalton, H. Colorimetric Method for a Rapid Detection of Oxygenated Aromatic Biotransformation Products. Biotechnol. Technol. 1997, 11 (8), 5857,  DOI: 10.1023/A:1018499024466
    515. 515
      Dacre, J. C. Nonspecificity of the Gibbs Reaction. Anal. Chem. 1971, 43 (4), 58991,  DOI: 10.1021/ac60299a015
    516. 516
      Josephy, P. D.; Van Damme, A. Reaction of Gibbs Reagent with Para-Substituted Phenols. Anal. Chem. 1984, 56 (4), 8134,  DOI: 10.1021/ac00268a052
    517. 517
      Svobodová, D.; Křenek, P.; Fraenkl, M.; Gasparič, J. The Colour Reaction of Phenols with the Gibbs Reagent. Microchim. Acta 1978, 70 (3), 197211,  DOI: 10.1007/BF01201610
    518. 518
      Ang, E. L.; Obbard, J. P.; Zhao, H. Directed Evolution of Aniline Dioxygenase for Enhanced Bioremediation of Aromatic Amines. Appl. Microbiol. Biotechnol. 2009, 81 (6), 106370,  DOI: 10.1007/s00253-008-1710-0
    519. 519
      Bornscheuer, U. T.; Ordoñez, G. R.; Hidalgo, A.; Gollin, A.; Lyon, J.; Hitchman, T. S.; Weiner, D. P. Selectivity of Lipases and Esterases Towards Phenol Esters. J. Mol. Catal. B: Enzym. 2005, 36 (1), 813,  DOI: 10.1016/j.molcatb.2005.07.004
    520. 520
      Altahir, B. M.; Abdulazeez, O.; Dikran, S. B.; Taylor, K. E. Determination of Eugenol in Personal-Care Products by Dispersive Liquid-Liquid Microextraction Followed by Spectrophotometry Using p-Amino-N,N-Dimethylaniline as a Derivatizing Agent. Indones. J. Chem. 2020, 21, 192,  DOI: 10.22146/ijc.56198
    521. 521
      Gasparic, J.; Svobodova, D.; Pospisilova, M. Investigation of the Colour Reaction of Phenols with the MBTH Reagent. Identification of Organic Compounds. LXXXVI. Microchim. Acta 1977, 67, 24150,  DOI: 10.1007/BF01213034
    522. 522
      Pospíšilová, M.; Polášek, M.; Svobodová, D. Spectrophotometric Study of Reactions of Substituted Phenols with MBTH in Alkaline Medium: The Effect of Phenol Structure on the Formation of Analytically Useful Coloured Products. Microchimica Acta 1998, 129 (3), 2018,  DOI: 10.1007/BF01244742
    523. 523
      Kamata, E. Color Reactions of 3-Methyl-2-Benzothiazolone Hydrazone with Phenol Derivatives. I. The Spectrophotometric Microdetermination of Catechol, Hydroquinone and Resorcinol. Bull. Chem. Soc. Jpn. 1964, 37 (11), 16747,  DOI: 10.1246/bcsj.37.1674
    524. 524
      Setti, L.; Giuliani, S.; Spinozzi, G.; Pifferi, P. G. Laccase Catalyzed-Oxidative Coupling of 3-Methyl 2-Benzothiazolinone Hydrazone and Methoxyphenols. Enzyme Microb. Technol. 1999, 25 (3), 2859,  DOI: 10.1016/S0141-0229(99)00059-9
    525. 525
      Setti, L.; Scali, S.; Angeli, I. D.; Pifferi, P. G. Horseradish Peroxidase-Catalyzed Oxidative Coupling of 3-Methyl 2-Benzothiazolinone Hydrazone and Methoxyphenols. Enzyme Microb. Technol. 1998, 22 (8), 65661,  DOI: 10.1016/S0141-0229(97)00259-7
    526. 526
      Nolan, L. C.; O’Connor, K. E. A Spectrophotometric Method for the Quantification of an Enzyme Activity Producing 4-Substituted Phenols: Determination of Toluene-4-Monooxygenase Activity. Anal. Biochem. 2005, 344 (2), 22431,  DOI: 10.1016/j.ab.2005.05.032
    527. 527
      Regitz, M. Diazo Compounds: Properties and Synthesis; Elsevier, 2012.
    528. 528
      Chudgar, R. J. Azo Dyes; Kirk–Othmer Encyclopedia of Chemical Technology, 2000.
    529. 529
      de Keijzer, M. The Delight of Modern Organic Pigment Creations, Issues in Contemporary Oil Paint; Springer International Publishing: Cham, 2014; p 4573.
    530. 530
      Urbányi, T.; Mollica, J. A. Potential Diazo Reagents for Colorimetric Determination. J. Pharm. Sci. 1968, 57 (7), 125760,  DOI: 10.1002/jps.2600570746
    531. 531
      Ehrlich, P. Uber Eine Neue Harnprobe. Dtsch. Med. Wochenschr. 1883, 9 (1), 11,  DOI: 10.1055/s-0029-1196984
    532. 532
      Hassan, S. M.; Walash, M. I.; El-Sayed, S. M.; Abou Ouf, A. M. Colorimetric Determination of Certain Phenol Derivatives in Pharmaceutical Preparations. J. Assoc. Off. Anal. Chem. 1981, 64 (6), 14425,  DOI: 10.1093/jaoac/64.6.1442
    533. 533
      Swaminathan, M. Chemical Estimation of Vitamin B6 in Foods by Means of the Diazo Reaction and the Phenol Reagent. Nature 1940, 145 (3681), 780,  DOI: 10.1038/145780a0
    534. 534
      Whitlock, L. R.; Siggia, S.; Smola, J. E. Spectrophotometric Analysis of Phenols and of Sulfonates by Formation of an Azo Dye. Anal. Chem. 1972, 44 (3), 5326,  DOI: 10.1021/ac60311a021
    535. 535
      Abdel Azeem, S. M.; Al Mohesen, I. A.; Ibrahim, A. M. H. Analysis of Total Phenolic Compounds in Tea and Fruits Using Diazotized Aminobenzenes Colorimetric Spots. Food Chem. 2020, 332, 127392,  DOI: 10.1016/j.foodchem.2020.127392
    536. 536
      Lester, G. E.; Lewers, K. S.; Medina, M. B.; Saftner, R. A. Comparative Analysis of Strawberry Total Phenolics Via Fast Blue BB Vs. Folin-Ciocalteu: Assay Interference by Ascorbic Acid. J. Food Compos. Anal. 2012, 27 (1), 1027,  DOI: 10.1016/j.jfca.2012.05.003
    537. 537
      Pompella, A.; Comporti, M. The Use of 3-Hydroxy-2-Naphthoic Acid Hydrazide and Fast Blue B for the Histochemical Detection of Lipid Peroxidation in Animal Tissues─a Microphotometric Study. Histochemistry 1991, 95, 255262,  DOI: 10.1007/BF00744997
    538. 538
      Tłuścik, F.; Kazubek, A.; Mejbaum-Katzenellenbogen, W. Alkylresorcinols in Rye (Secale Cereale L.) Grains. Vi. Colorimetric Micromethod for the Determination of Alkylresorcinols with the Use of Diazonium Salt, Fast Blue B. Acta Soc. Bot. Polym. 1981, 50, 645,  DOI: 10.5586/asbp.1981.086
    539. 539
      Schotten, C.; Leprevost, S. K.; Yong, L. M.; Hughes, C. E.; Harris, K. D. M.; Browne, D. L. Comparison of the Thermal Stabilities of Diazonium Salts and Their Corresponding Triazenes. Org. Process Res. Dev. 2020, 24 (10), 233641,  DOI: 10.1021/acs.oprd.0c00162
    540. 540
      Johnston, K. J.; Ashford, A. E. A Simultaneous-Coupling Azo Dye Method for the Quantitative Assay of Esterase Using α-Naphthyl Acetate as Substrate. Histochem. J. 1980, 12 (2), 22134,  DOI: 10.1007/BF01024552
    541. 541
      Pearse, A. G. E. Histochemistry, Theoretical and Applied; Harcourt Brace/Churchill Livingstone, 1968.
    542. 542
      Lugg, G. A. Stabilized Diazonium Salts as Analytical Reagents for the Determination of Air-Borne Phenols and Amines. Anal. Chem. 1963, 35 (7), 899904,  DOI: 10.1021/ac60200a039
    543. 543
      Buchwald, H. The Colorimetric Determination of Phenol in Air and Urine with a Stabilized Diazonium Salt. Ann. Occup. Hyg. 1966, 9, 714,  DOI: 10.1093/annhyg/9.1.7
    544. 544
      Harvey, D.; Penketh, G. E. The Determination of Small Amounts of o-Phenylphenol. Analyst 1957, 82 (976), 498503,  DOI: 10.1039/an9578200498
    545. 545
      Canada, K. A.; Iwashita, S.; Shim, H.; Wood, T. K. Directed Evolution of Toluene ortho-Monooxygenase for Enhanced 1-Naphthol Synthesis and Chlorinated Ethene Degradation. J. Bacteriol. 2002, 184 (2), 3449,  DOI: 10.1128/JB.184.2.344-349.2002
    546. 546
      Joshi, N. S.; Whitaker, L. R.; Francis, M. B. A Three-Component Mannich-Type Reaction for Selective Tyrosine Bioconjugation. J. Am. Chem. Soc. 2004, 126 (49), 159423,  DOI: 10.1021/ja0439017
    547. 547
      Cardellicchio, C.; Capozzi, M. A. M.; Naso, F. The Betti Base: The Awakening of a Sleeping Beauty. Tetrahedron: Asymmetry 2010, 21 (5), 50717,  DOI: 10.1016/j.tetasy.2010.03.020
    548. 548
      Minakawa, M.; Guo, H.-M.; Tanaka, F. Imines That React with Phenols in Water over a Wide pH Range. J. Org. Chem. 2008, 73 (21), 866972,  DOI: 10.1021/jo8017389
    549. 549
      Kanaoka, Y. Organic Fluorescence Reagents in the Study of Enzymes and Proteins. Angew. Chem., Int. Ed. 1977, 16 (3), 13747,  DOI: 10.1002/anie.197701371
    550. 550
      Davis, W.; Ronai, Z.; Tew, K. D. Cellular Thiols and Reactive Oxygen Species in Drug-Induced Apoptosis. J. Pharmacol. Exp. Ther. 2001, 296, 16
    551. 551
      Ball, R. O.; Courtney-Martin, G.; Pencharz, P. B. The in Vivo Sparing of Methionine by Cysteine in Sulfur Amino Acid Requirements in Animal Models and Adult Humans. J. Nutr. 2006, 136 (6), 1682S93S,  DOI: 10.1093/jn/136.6.1682S
    552. 552
      Mishanina, T. V.; Libiad, M.; Banerjee, R. Biogenesis of Reactive Sulfur Species for Signaling by Hydrogen Sulfide Oxidation Pathways. Nat. Chem. Biol. 2015, 11 (7), 45764,  DOI: 10.1038/nchembio.1834
    553. 553
      Coleman, J.; Blake-Kalff, M.; Davies, E. Detoxification of Xenobiotics by Plants: Chemical Modification and Vacuolar Compartmentation. Trends Plant Sci. 1997, 2 (4), 14451,  DOI: 10.1016/S1360-1385(97)01019-4
    554. 554
      Tsay, J. T.; Oh, W.; Larson, T. J.; Jackowski, S.; Rock, C. O. Isolation and Characterization of the Beta-Ketoacyl-Acyl Carrier Protein Synthase III Gene (FabH) from Escherichia Coli K-12. J. Biol. Chem. 1992, 267 (10), 680714,  DOI: 10.1016/S0021-9258(19)50498-7
    555. 555
      Gao, T.; Yang, C.; Zheng, Y. G. Comparative Studies of Thiol-Sensitive Fluorogenic Probes for HAT Assays. Anal. Bioanal. Chem. 2013, 405 (4), 136171,  DOI: 10.1007/s00216-012-6522-5
    556. 556
      Tang, Y.; Jin, L.; Yin, B. A Dual-Selective Fluorescent Probe for GSH and Cys Detection: Emission and pH Dependent Selectivity. Anal. Chim. Acta 2017, 993, 8795,  DOI: 10.1016/j.aca.2017.09.028
    557. 557
      Zaric, B. L.; Obradovic, M.; Bajic, V.; Haidara, M. A.; Jovanovic, M.; Isenovic, E. R. Homocysteine and Hyperhomocysteinaemia. Curr. Med. Chem. 2019, 26 (16), 294861,  DOI: 10.2174/0929867325666180313105949
    558. 558
      van Meurs, J.; Dhonukshe-Rutten, R. A.; Pluijm, S. M.; Van Der Klift, M.; De Jonge, R.; Lindemans, J.; De Groot, L. C.; Hofman, A.; Witteman, J. C.; Van Leeuwen, J. P.; Breteler, M. M.; Lips, P.; Pols, H. A.; Uitterlinden, A. G. Homocysteine Levels and the Risk of Osteoporotic Fracture. N. Engl. J. Med. 2004, 350, 203341,  DOI: 10.1056/NEJMoa032546
    559. 559
      Lee, D.; Kim, G.; Yin, J.; Yoon, J. An Aryl-Thioether Substituted Nitrobenzothiadiazole Probe for the Selective Detection of Cysteine and Homocysteine. Chem. Commun. 2015, 51 (30), 651820,  DOI: 10.1039/C5CC01071C
    560. 560
      Giannoulis, K. M.; Giokas, D. L.; Tsogas, G. Z.; Vlessidis, A. G. Ligand-Free Gold Nanoparticles as Colorimetric Probes for the Non-Destructive Determination of Total Dithiocarbamate Pesticides after Solid Phase Extraction. Talanta 2014, 119, 27683,  DOI: 10.1016/j.talanta.2013.10.063
    561. 561
      Munday, R. Toxicity of Thiols and Disulphides: Involvement of Free-Radical Species. Free Radical Biol. Med. 1989, 7 (6), 65973,  DOI: 10.1016/0891-5849(89)90147-0
    562. 562
      Cazzola, M.; Calzetta, L.; Page, C.; Rogliani, P.; Matera, M. G. Thiol-Based Drugs in Pulmonary Medicine: Much More Than Mucolytics. Trends Pharmacol. Sci. 2019, 40 (7), 45263,  DOI: 10.1016/j.tips.2019.04.015
    563. 563
      Wang, H.; Wu, X.; Yang, S.; Tian, H.; Liu, Y.; Sun, B. A Rapid and Visible Colorimetric Fluorescent Probe for Benzenethiol Flavor Detection. Food Chem. 2019, 286, 3228,  DOI: 10.1016/j.foodchem.2019.02.033
    564. 564
      Liu, X.-L.; Duan, X.-Y.; Du, X.-J.; Song, Q.-H. Quinolinium-Based Fluorescent Probes for the Detection of Thiophenols in Environmental Samples and Living Cells. Chem.─Asian J. 2012, 7, 26962702,  DOI: 10.1002/asia.201200594
    565. 565
      He, L.; Yang, X.; Xu, K.; Kong, X.; Lin, W. A Multi-Signal Fluorescent Probe for Simultaneously Distinguishing and Sequentially Sensing Cysteine/Homocysteine, Glutathione, and Hydrogen Sulfide in Living Cells. Chem. Sci. 2017, 8 (9), 625765,  DOI: 10.1039/C7SC00423K
    566. 566
      Wang, N.; Chen, M.; Gao, J.; Ji, X.; He, J.; Zhang, J.; Zhao, W. A Series of BODIPY-Based Probes for the Detection of Cysteine and Homocysteine in Living Cells. Talanta 2019, 195, 2819,  DOI: 10.1016/j.talanta.2018.11.066
    567. 567
      Yao, Z.; Ge, W.; Guo, M.; Xiao, K.; Qiao, Y.; Cao, Z.; Wu, H.-C. Ultrasensitive Detection of Thiophenol Based on a Water-Soluble Pyrenyl Probe. Talanta 2018, 185, 14650,  DOI: 10.1016/j.talanta.2018.03.068
    568. 568
      Zhang, S.; Cai, F.; Hou, B.; Chen, H.; Gao, C.; Shen, X.-c.; Liang, H. Constructing a Far-Red to near-Infrared Fluorescent Probe for Highly Specific Detection of Cysteine and Its Bioimaging Applications in Living Cells and Zebrafish. New J. Chem. 2019, 43 (17), 6696701,  DOI: 10.1039/C9NJ00260J
    569. 569
      Fan, L.; Zhang, W.; Wang, X.; Dong, W.; Tong, Y.; Dong, C.; Shuang, S. A Two-Photon Ratiometric Fluorescent Probe for Highly Selective Sensing of Mitochondrial Cysteine in Live Cells. Analyst 2019, 144 (2), 43947,  DOI: 10.1039/C8AN01908H
    570. 570
      Ros-Lis, J. V.; García, B.; Jiménez, D.; Martínez-Máñez, R.; Sancenón, F.; Soto, J.; Gonzalvo, F.; Valldecabres, M. C. Squaraines as Fluoro-Chromogenic Probes for Thiol-Containing Compounds and Their Application to the Detection of Biorelevant Thiols. J. Am. Chem. Soc. 2004, 126 (13), 40645,  DOI: 10.1021/ja031987i
    571. 571
      Zhang, X.; Li, C.; Cheng, X.; Wang, X.; Zhang, B. A near-Infrared Croconium Dye-Based Colorimetric Chemodosimeter for Biological Thiols and Cyanide Anion. Sens. Actuators B Chem. 2008, 129 (1), 1527,  DOI: 10.1016/j.snb.2007.07.094
    572. 572
      Tanaka, F.; Mase, N.; Barbaa, C. F., III Determination of Cysteine Concentration by Fluorescence Increase: Reaction of Cysteine with a Fluorogenic Aldehyde. Chem. Commun. 2004, 2004, 1762,  DOI: 10.1039/b405642f
    573. 573
      Jung, H. S.; Ko, K. C.; Kim, G.-H.; Lee, A.-R.; Na, Y.-C.; Kang, C.; Lee, J. Y.; Kim, J. S. Coumarin-Based Thiol Chemosensor: Synthesis, Turn-On Mechanism, and Its Biological Application. Org. Lett. 2011, 13 (6), 1498501,  DOI: 10.1021/ol2001864
    574. 574
      Zhang, Y.; Shao, X.; Wang, Y.; Pan, F.; Kang, R.; Peng, F.; Huang, Z.; Zhang, W.; Zhao, W. Dual Emission Channels for Sensitive Discrimination of Cys/Hcy and GSH in Plasma and Cells. Chem. Commun. 2015, 51 (20), 42458,  DOI: 10.1039/C4CC08687B
    575. 575
      Niu, L.-Y.; Guan, Y.-S.; Chen, Y.-Z.; Wu, L.-Z.; Tung, C.-H.; Yang, Q.-Z. A Turn-On Fluorescent Sensor for the Discrimination of Cystein from Homocystein and Glutathione. Chem. Commun. 2013, 49 (13), 12946,  DOI: 10.1039/c2cc38429a
    576. 576
      Chen, W.; Luo, H.; Liu, X.; Foley, J. W.; Song, X. Broadly Applicable Strategy for the Fluorescence Based Detection and Differentiation of Glutathione and Cysteine/Homocysteine: Demonstration in Vitro and in Vivo. Anal. Chem. 2016, 88 (7), 363846,  DOI: 10.1021/acs.analchem.5b04333
    577. 577
      Mikaliunaite, L.; Green, D. B. Using a 3-Hydroxyflavone Derivative as a Fluorescent Probe for the Indirect Determination of Aminothiols Separated by Ion-Pair HPLC. Anal. Methods 2021, 13 (26), 291525,  DOI: 10.1039/D1AY00499A
    578. 578
      Toyooka, T.; Imai, K. New Fluorogenic Reagent Having Halogenobenzofurazan Structure for Thiols: 4-(Aminosulfonyl)-7-Fluoro-2,1,3-Benzoxadiazole. Anal. Chem. 1984, 56 (13), 24614,  DOI: 10.1021/ac00277a044
    579. 579
      Zhao, L.; Zhao, L.; Zhang, C.; Li, Y. A Naked-Eye Visible and Fluorescence ″Turn-On″ Probe for Acetylcholinesterase Assay and the Discrimination of Biothiols. J. Appl. Spectrosc. 2018, 85 (3), 43744,  DOI: 10.1007/s10812-018-0669-6
    580. 580
      Hassett, R. P.; Crockett, E. L. Endpoint Fluorometric Assays for Determining Activities of Carnitine Palmitoyltransferase and Citrate Synthase. Anal. Biochem. 2000, 287 (1), 1769,  DOI: 10.1006/abio.2000.4799
    581. 581
      Niu, L.-Y.; Zheng, H.-R.; Chen, Y.-Z.; Wu, L.-Z.; Tung, C.-H.; Yang, Q.-Z. Fluorescent Sensors for Selective Detection of Thiols: Expanding the Intramolecular Displacement Based Mechanism to New Chromophores. Analyst 2014, 139 (6), 138995,  DOI: 10.1039/c3an01849k
    582. 582
      Niu, L.-Y.; Guan, Y.-S.; Chen, Y.-Z.; Wu, L.-Z.; Tung, C.-H.; Yang, Q.-Z. BODIPY-Based Ratiometric Fluorescent Sensor for Highly Selective Detection of Glutathione over Cysteine and Homocysteine. J. Am. Chem. Soc. 2012, 134 (46), 1892831,  DOI: 10.1021/ja309079f
    583. 583
      Liu, X.-L.; Niu, L.-Y.; Chen, Y.-Z.; Zheng, M.-L.; Yang, Y.; Yang, Q.-Z. A Mitochondria-Targeting Fluorescent Probe for the Selective Detection of Glutathione in Living Cells. Org. Biomol. Chem. 2017, 15 (5), 10725,  DOI: 10.1039/C6OB02407F
    584. 584
      Hu, Q.; Yu, C.; Xia, X.; Zeng, F.; Wu, S. A Fluorescent Probe for Simultaneous Discrimination of GSH and Cys/Hcy in Human Serum Samples Via Distinctly-Separated Emissions with Independent Excitations. Biosens. Bioelectron. 2016, 81, 3418,  DOI: 10.1016/j.bios.2016.03.011
    585. 585
      Zhang, J.; Ji, X.; Ren, H.; Zhou, J.; Chen, Z.; Dong, X.; Zhao, W. meso-Heteroaryl BODIPY Dyes as Dual-Responsive Fluorescent Probes for Discrimination of Cys from Hcy and GSH. Sens. Actuators B Chem. 2018, 260, 8619,  DOI: 10.1016/j.snb.2018.01.016
    586. 586
      Liu, Y.; Yu, Y.; Zhao, Q.; Tang, C.; Zhang, H.; Qin, Y.; Feng, X.; Zhang, J. Fluorescent Probes Based on Nucleophilic Aromatic Substitution Reactions for Reactive Sulfur and Selenium Species: Recent Progress, Applications, and Design Strategies. Coord. Chem. Rev. 2021, 427, 213601,  DOI: 10.1016/j.ccr.2020.213601
    587. 587
      Kim, Y.; Mulay, S. V.; Choi, M.; Yu, S. B.; Jon, S.; Churchill, D. G. Exceptional Time Response, Stability and Selectivity in Doubly-Activated Phenyl Selenium-Based Glutathione-Selective Platform. Chem. Sci. 2015, 6, 5435,  DOI: 10.1039/C5SC02090E
    588. 588
      Mulay, S. V.; Kim, Y.; Choi, M.; Lee, D. Y.; Choi, J.; Lee, Y.; Jon, S.; Churchill, D. G. Enhanced Doubly Activated Dual Emission Fluorescent Probes for Selective Imaging of Glutathione or Cysteine in Living Systems. Anal. Chem. 2018, 90 (4), 264854,  DOI: 10.1021/acs.analchem.7b04375
    589. 589
      Maeda, H.; Matsuno, H.; Ushida, M.; Katayama, K.; Saeki, K.; Itoh, N. 2,4-Dinitrobenzenesulfonyl Fluoresceins as Fluorescent Alternatives to Ellman’s Reagent in Thiol-Quantification Enzyme Assays. Angew. Chem., Int. Ed. 2005, 44 (19), 29225,  DOI: 10.1002/anie.200500114
    590. 590
      Jiang, W.; Fu, Q.; Fan, H.; Ho, J.; Wang, W. A Highly Selective Fluorescent Probe for Thiophenols. Angew. Chem., Int. Ed. 2007, 46 (44), 84458,  DOI: 10.1002/anie.200702271
    591. 591
      Wang, S.; Huang, Y.; Guan, X. Fluorescent Probes for Live Cell Thiol Detection. Molecules 2021, 26 (12), 3575,  DOI: 10.3390/molecules26123575
    592. 592
      Dai, Z.; Tian, L.; Ye, Z.; Song, B.; Zhang, R.; Yuan, J. A Lanthanide Complex-Based Ratiometric Luminescence Probe for Time-Gated Luminescence Detection of Intracellular Thiols. Anal. Chem. 2013, 85 (23), 1165864,  DOI: 10.1021/ac403370g
    593. 593
      Rong, X.; Xu, Z.-Y.; Yan, J.-W.; Meng, Z.-Z.; Zhu, B.; Zhang, L. Nile-Red-Based Fluorescence Probe for Selective Detection of Biothiols, Computational Study, and Application in Cell Imaging. Molecules 2020, 25 (20), 4718,  DOI: 10.3390/molecules25204718
    594. 594
      Maeda, H.; Katayama, K.; Matsuno, H.; Uno, T. 3′-(2,4-Dinitrobenzenesulfonyl)-2′,7′-Dimethylfluorescein as a Fluorescent Probe for Selenols. Angew. Chem., Int. Ed. 2006, 45 (11), 18103,  DOI: 10.1002/anie.200504299
    595. 595
      Peng, L.; Xiao, L.; Ding, Y.; Xiang, Y.; Tong, A. A Simple Design of Fluorescent Probes for Indirect Detection of β-Lactamase Based on AIE and ESIPT Processes. J. Mater. Chem. B 2018, 6 (23), 39226,  DOI: 10.1039/C8TB00414E
    596. 596
      Yuan, L.; Lin, W.; Zhao, S.; Gao, W.; Chen, B.; He, L.; Zhu, S. A Unique Approach to Development of Near-Infrared Fluorescent Sensors for in Vivo Imaging. J. Am. Chem. Soc. 2012, 134 (32), 1351023,  DOI: 10.1021/ja305802v
    597. 597
      Lu, Z.; Sun, X.; Wang, M.; Wang, H.; Fan, C.; Lin, W. Rational Design of a Far-Red Fluorescent Probe for Endogenous Biothiol Imbalance Induced by Hydrogen Peroxide in Living Cells and Mice. Bioorg. Chem. 2020, 103, 104173,  DOI: 10.1016/j.bioorg.2020.104173
    598. 598
      Kand, D.; Kalle, A. M.; Varma, S. J.; Talukdar, P. A Chromenoquinoline-Based Fluorescent Off-On Thiol Probe for Bioimaging. Chem. Commun. 2012, 48 (21), 27224,  DOI: 10.1039/c2cc16593g
    599. 599
      Liu, T.; Huo, F.; Li, J.; Chao, J.; Zhang, Y.; Yin, C. A Fast Response and High Sensitivity Thiol Fluorescent Probe in Living Cells. Sens. Actuators B Chem. 2016, 232, 61924,  DOI: 10.1016/j.snb.2016.04.014
    600. 600
      Lin, X.; Hu, Y.; Yang, D.; Chen, B. Cyanine-Coumarin Composite NIR Dye Based Instantaneous-Response Probe for Biothiols Detection and Oxidative Stress Assessment of Mitochondria. Dyes Pigm. 2020, 174, 107956,  DOI: 10.1016/j.dyepig.2019.107956
    601. 601
      Sun, Q.; Yang, S.-H.; Wu, L.; Yang, W.-C.; Yang, G.-F. A Highly Sensitive and Selective Fluorescent Probe for Thiophenol Designed Via a Twist-Blockage Strategy. Anal. Chem. 2016, 88 (4), 226672,  DOI: 10.1021/acs.analchem.5b04029
    602. 602
      Jiang, W.; Cao, Y.; Liu, Y.; Wang, W. Rational Design of a Highly Selective and Sensitive Fluorescent PET Probe for Discrimination of Thiophenols and Aliphatic Thiols. Chem. Commun. 2010, 46 (11), 19446,  DOI: 10.1039/B926070F
    603. 603
      Li, Y.; Su, W.; Zhou, Z.; Huang, Z.; Wu, C.; Yin, P.; Li, H.; Zhang, Y. A Dual-Response near-Infrared Fluorescent Probe for Rapid Detecting Thiophenol and Its Application in Water Samples and Bio-Imaging. Talanta 2019, 199, 35560,  DOI: 10.1016/j.talanta.2019.02.022
    604. 604
      Xiong, L.; Ma, J.; Huang, Y.; Wang, Z.; Lu, Z. Highly Sensitive Squaraine-Based Water-Soluble Far-Red/near-Infrared Chromofluorogenic Thiophenol Probe. ACS Sens. 2017, 2 (4), 599605,  DOI: 10.1021/acssensors.7b00151
    605. 605
      Song, B.; Wang, G.; Tan, M.; Yuan, J. A Europium(III) Complex as an Efficient Singlet Oxygen Luminescence Probe. J. Am. Chem. Soc. 2006, 128 (41), 1344250,  DOI: 10.1021/ja062990f
    606. 606
      Bjartell, A.; Laine, S.; Pettersson, K.; Nilsson, E.; Lövgren, T.; Lilja, H. Time-Resolved Fluorescence in Immunocytochemical Detection of Prostate-Specific Antigen in Prostatic Tissue Sections. Histochem. J. 1999, 31 (1), 4552,  DOI: 10.1023/A:1003504115690
    607. 607
      Shang, H.; Chen, H.; Tang, Y.; Ma, Y.; Lin, W. Development of a Two-Photon Fluorescent Turn-On Probe with Far-Red Emission for Thiophenols and Its Bioimaging Application in Living Tissues. Biosens. Bioelectron. 2017, 95, 816,  DOI: 10.1016/j.bios.2017.04.017
    608. 608
      Fellner, M.; Doughty, L. M.; Jameson, G. N. L.; Wilbanks, S. M. A Chromogenic Assay of Substrate Depletion by Thiol Dioxygenases. Anal. Biochem. 2014, 459, 5660,  DOI: 10.1016/j.ab.2014.05.008
    609. 609
      Zhu, J.; Dhimitruka, I.; Pei, D. 5-(2-Aminoethyl)Dithio-2-Nitrobenzoate as a More Base-Stable Alternative to Ellman’s Reagent. Org. Lett. 2004, 6 (21), 380912,  DOI: 10.1021/ol048404+
    610. 610
      Li, Y.; Lopez, P.; Durand, P.; Ouazzani, J.; Badet, B.; Badet-Denisot, M.-A. An Enzyme-Coupled Assay for Amidotransferase Activity of Glucosamine-6-Phosphate Synthase. Anal. Biochem. 2007, 370 (2), 1426,  DOI: 10.1016/j.ab.2007.07.031
    611. 611
      Chen, G.; Feng, H.; Jiang, X.; Xu, J.; Pan, S.; Qian, Z. Redox-Controlled Fluorescent Nanoswitch Based on Reversible Disulfide and Its Application in Butyrylcholinesterase Activity Assay. Anal. Chem. 2018, 90 (3), 164351,  DOI: 10.1021/acs.analchem.7b02976
    612. 612
      Gong, M.-M.; Dai, C.-Y.; Severance, S.; Hwang, C.-C.; Fang, B.-K.; Lin, H.-B.; Huang, C.-H.; Ong, C.-W.; Wang, J.-J.; Lee, P.-L. A Bioorthogonally Synthesized and Disulfide-Containing Fluorescence Turn-On Chemical Probe for Measurements of Butyrylcholinesterase Activity and Inhibition in the Presence of Physiological Glutathione. Catalysts 2020, 10 (10), 1169,  DOI: 10.3390/catal10101169
    613. 613
      Peng, H.; Chen, W.; Cheng, Y.; Hakuna, L.; Strongin, R.; Wang, B. Thiol Reactive Probes and Chemosensors. Sensors 2012, 12 (11), 1590746,  DOI: 10.3390/s121115907
    614. 614
      Worek, F.; Mast, U.; Kiderlen, D.; Diepold, C.; Eyer, P. Improved Determination of Acetylcholinesterase Activity in Human Whole Blood. Clin. Chim. Acta 1999, 288 (1), 7390,  DOI: 10.1016/S0009-8981(99)00144-8
    615. 615
      Eyer, P.; Worek, F.; Kiderlen, D.; Sinko, G.; Stuglin, A.; Simeon-Rudolf, V.; Reiner, E. Molar Absorption Coefficients for the Reduced Ellman Reagent: Reassessment. Anal. Biochem. 2003, 312 (2), 2247,  DOI: 10.1016/S0003-2697(02)00506-7
    616. 616
      Pullela, P. K.; Chiku, T.; Carvan, M. J.; Sem, D. S. Fluorescence-Based Detection of Thiols in Vitro and in Vivo Using Dithiol Probes. Anal. Biochem. 2006, 352 (2), 26573,  DOI: 10.1016/j.ab.2006.01.047
    617. 617
      Cao, X.; Lin, W.; Yu, Q. A Ratiometric Fluorescent Probe for Thiols Based on a Tetrakis(4-Hydroxyphenyl)Porphyrin-Coumarin Scaffold. J. Org. Chem. 2011, 76 (18), 742330,  DOI: 10.1021/jo201199k
    618. 618
      Wang, L.; Wang, J.; Xia, S.; Wang, X.; Yu, Y.; Zhou, H.; Liu, H. A FRET-Based near-Infrared Ratiometric Fluorescent Probe for Detection of Mitochondria Biothiol. Talanta 2020, 219, 121296,  DOI: 10.1016/j.talanta.2020.121296
    619. 619
      Lim, S. Y.; Shen, W.; Gao, Z. Carbon Quantum Dots and Their Applications. Chem. Soc. Rev. 2015, 44 (1), 36281,  DOI: 10.1039/C4CS00269E
    620. 620
      Steinmann, D.; Nauser, T.; Koppenol, W. H. Selenium and Sulfur in Exchange Reactions: A Comparative Study. J. Org. Chem. 2010, 75 (19), 66969,  DOI: 10.1021/jo1011569
    621. 621
      Mafireyi, T. J.; Laws, M.; Bassett, J. W.; Cassidy, P. B.; Escobedo, J. O.; Strongin, R. M. A Diselenide Turn-On Fluorescent Probe for the Detection of Thioredoxin Reductase. Angew. Chem., Int. Ed. 2020, 59 (35), 1514751,  DOI: 10.1002/anie.202004094
    622. 622
      Nogueira, C. W.; Barbosa, N. V.; Rocha, J. B. Toxicology and Pharmacology of Synthetic Organoselenium Compounds: An Update. Arch. Toxicol. 2021, 95 (4), 1179226,  DOI: 10.1007/s00204-021-03003-5
    623. 623
      Lou, Z.; Li, P.; Sun, X.; Yang, S.; Wang, B.; Han, K. A Fluorescent Probe for Rapid Detection of Thiols and Imaging of Thiols Reducing Repair and H2O2 Oxidative Stress Cycles in Living Cells. Chem. Commun. 2013, 49 (4), 3913,  DOI: 10.1039/C2CC36839K
    624. 624
      Han, X.; Song, X.; Yu, F.; Chen, L. A Ratiometric Fluorescent Probe for Imaging and Quantifying Anti-Apoptotic Effects of GSH under Temperature Stress. Chem. Sci. 2017, 8 (10), 69917002,  DOI: 10.1039/C7SC02888A
    625. 625
      Tang, B.; Xing, Y.; Li, P.; Zhang, N.; Yu, F.; Yang, G. A Rhodamine-Based Fluorescent Probe Containing a Se-N Bond for Detecting Thiols and Its Application in Living Cells. J. Am. Chem. Soc. 2007, 129 (38), 116667,  DOI: 10.1021/ja072572q
    626. 626
      Tang, B.; Yin, L.; Wang, X.; Chen, Z.; Tong, L.; Xu, K. A Fast-Response, Highly Sensitive and Specific Organoselenium Fluorescent Probe for Thiols and Its Application in Bioimaging. Chem. Commun. 2009, (35), 52935,  DOI: 10.1039/b909542j
    627. 627
      Wang, R.; Chen, L.; Liu, P.; Zhang, Q.; Wang, Y. Sensitive near-Infrared Fluorescent Probes for Thiols Based on Se-N Bond Cleavage: Imaging in Living Cells and Tissues. Eur. J. Chem. 2012, 18, 11343,  DOI: 10.1002/chem.201200671
    628. 628
      Tian, Y.; Zhu, B.; Yang, W.; Jing, J.; Zhang, X. A Fluorescent Probe for Differentiating Cys, Hcy and GSH Via a Stepwise Interaction. Sens. Actuators B Chem. 2018, 262, 3459,  DOI: 10.1016/j.snb.2018.01.181
    629. 629
      Rusin, O.; St Luce, N. N.; Agbaria, R. A.; Escobedo, J. O.; Jiang, S.; Warner, I. M.; Dawan, F. B.; Lian, K.; Strongin, R. M. Visual Detection of Cysteine and Homocysteine. J. Am. Chem. Soc. 2004, 126 (2), 438,  DOI: 10.1021/ja036297t
    630. 630
      Lim, S.; Escobedo, J. O.; Lowry, M.; Xu, X.; Strongin, R. Selective Fluorescence Detection of Cysteine and N-Terminal Cysteine Peptide Residues. Chem. Commun. 2010, 46 (31), 57079,  DOI: 10.1039/c0cc01398f
    631. 631
      Lee, K.-S.; Kim, T.-K.; Lee, J. H.; Kim, H.-J.; Hong, J.-I. Fluorescence Turn-On Probe for Homocysteine and Cysteine in Water. Chem. Commun. 2008, (46), 61735,  DOI: 10.1039/b814581d
    632. 632
      Eun Jun, M.; Roy, B.; Han Ahn, K. “Turn-On” Fluorescent Sensing with “Reactive” Probes. Chem. Commun. 2011, 47 (27), 7583601,  DOI: 10.1039/c1cc00014d
    633. 633
      Hoyle, C. E.; Lowe, A. B.; Bowman, C. N. Thiol-Click Chemistry: A Multifaceted Toolbox for Small Molecule and Polymer Synthesis. Chem. Soc. Rev. 2010, 39 (4), 135587,  DOI: 10.1039/b901979k
    634. 634
      LoPachin, R. M.; Gavin, T. Molecular Mechanisms of Aldehyde Toxicity: A Chemical Perspective. Chem. Res. Toxicol. 2014, 27 (7), 108191,  DOI: 10.1021/tx5001046
    635. 635
      Chan, J. W.; Hoyle, C. E.; Lowe, A. B.; Bowman, M. Nucleophile-Initiated Thiol-Michael Reactions: Effect of Organocatalyst, Thiol, and Ene. Macromolecules 2010, 43 (15), 63818,  DOI: 10.1021/ma101069c
    636. 636
      Langmuir, M. E.; Yang, J.-R.; Moussa, A. M.; Laura, R.; LeCompte, K. A. New Naphthopyranone Based Fluorescent Thiol Probes. Tetrahedron Lett. 1995, 36 (23), 398992,  DOI: 10.1016/0040-4039(95)00695-9
    637. 637
      Niu, W.; Wu, P.; Chen, F.; Wang, J.; Shang, X.; Xu, C. Discovery of Selective Cystathionine β-Synthase Inhibitors by High-Throughput Screening with a Fluorescent Thiol Probe. MedChemComm 2017, 8 (1), 198201,  DOI: 10.1039/C6MD00493H
    638. 638
      Aharoni, A.; Amitai, G.; Bernath, K.; Magdassi, S.; Tawfik, D. S. High-Throughput Screening of Enzyme Libraries: Thiolactonases Evolved by Fluorescence-Activated Sorting of Single Cells in Emulsion Compartments. Chem. Biol. 2005, 12 (12), 12819,  DOI: 10.1016/j.chembiol.2005.09.012
    639. 639
      Barb, A. W.; Marcella, A. M. A Rapid Fluorometric Assay for the S-Malonyltransacylase FabD and Other Sulfhydryl Utilizing Enzymes. J. Biol. Methods 2016, 3 (4), e53  DOI: 10.14440/jbm.2016.144
    640. 640
      Yi, L.; Li, H.; Sun, L.; Liu, L.; Zhang, C.; Xi, Z. A Highly Sensitive Fluorescence Probe for Fast Thiol-Quantification Assay of Glutathione Reductase. Angew. Chem., Int. Ed. 2009, 48 (22), 40347,  DOI: 10.1002/anie.200805693
    641. 641
      Zhou, X.; Jin, X.; Sun, G.; Li, D.; Wu, X. A Cysteine Probe with High Selectivity and Sensitivity Promoted by Response-Assisted Electrostatic Attraction. Chem. Commun. 2012, 48 (70), 87935,  DOI: 10.1039/c2cc33971d
    642. 642
      Li, S.-J.; Fu, Y.-J.; Li, C.-Y.; Li, Y.-F.; Yi, L.-H.; Ou-Yang, J. A near-Infrared Fluorescent Probe Based on BODIPY Derivative with High Quantum Yield for Selective Detection of Exogenous and Endogenous Cysteine in Biological Samples. Anal. Chim. Acta 2017, 994, 7381,  DOI: 10.1016/j.aca.2017.09.031
    643. 643
      Li, R.; Lei, C.; Zhao, X.-E.; Gao, Y.; Gao, H.; Zhu, S.; Wang, H. A Label-Free Fluorimetric Detection of Biothiols Based on the Oxidase-Like Activity of Ag+ Ions. Spectrochim. Acta - A: Mol. Biomol. Spectrosc. 2018, 188, 205,  DOI: 10.1016/j.saa.2017.06.056
    644. 644
      Mandani, S.; Sharma, B.; Dey, D.; Sarma, T. K. Carbon Nanodots as Ligand Exchange Probes in Au@C-Dot Nanobeacons for Fluorescent Turn-On Detection of Biothiols. Nanoscale 2015, 7 (5), 18028,  DOI: 10.1039/C4NR05424E
    645. 645
      Fu, Y.; Li, H.; Hu, W.; Zhu, D. Fluorescence Probes for Thiol-Containing Amino Acids and Peptides in Aqueous Solution. Chem. Commun. 2005, (25), 318991,  DOI: 10.1039/b503772g
    646. 646
      Nie, L.; Ma, H.; Sun, M.; Li, X.; Su, M.; Liang, S. Direct Chemiluminescence Determination of Cysteine in Human Serum Using Quinine-Ce(IV) System. Talanta 2003, 59 (5), 95964,  DOI: 10.1016/S0039-9140(02)00649-5
    647. 647
      Wang, S.; Ma, H.; Li, J.; Chen, X.; Bao, Z.; Sun, S. Direct Determination of Reduced Glutathione in Biological Fluids by Ce(Iv)-Quinine Chemiluminescence. Talanta 2006, 70 (3), 51821,  DOI: 10.1016/j.talanta.2005.12.052
    648. 648
      Zou, T.; Lum, C. T.; Chui, S. S.-Y.; Che, C.-M. Gold(III) Complexes Containing N-Heterocyclic Carbene Ligands: Thiol “Switch-On” Fluorescent Probes and Anti-Cancer Agents. Angew. Chem., Int. Ed. 2013, 52 (10), 29303,  DOI: 10.1002/anie.201209787
    649. 649
      Han, G.-C.; Peng, Y.; Hao, Y.-Q.; Liu, Y.-N.; Zhou, F. Spectrofluorimetric Determination of Total Free Thiols Based on Formation of Complexes of Ce(III) with Disulfide Bonds. Anal. Chim. Acta 2010, 659 (1), 23842,  DOI: 10.1016/j.aca.2009.11.057
    650. 650
      Wang, H.; Zhou, G.; Chen, X. An Iminofluorescein-Cu2+ Ensemble Probe for Selective Detection of Thiols. Sens. Actuators B Chem. 2013, 176, 698703,  DOI: 10.1016/j.snb.2012.10.006
    651. 651
      Jung, H. S.; Han, J. H.; Habata, Y.; Kang, C.; Kim, J. S. An Iminocoumarin-Cu(II) Ensemble-Based Chemodosimeter toward Thiols. Chem. Commun. 2011, 47 (18), 51424,  DOI: 10.1039/c1cc10672d
    652. 652
      Zhao, D.; Chen, C.; Sun, J.; Yang, X. Carbon Dots-Assisted Colorimetric and Fluorometric Dual-Mode Protocol for Acetylcholinesterase Activity and Inhibitors Screening Based on the Inner Filter Effect of Silver Nanoparticles. Analyst 2016, 141 (11), 32808,  DOI: 10.1039/C6AN00514D
    653. 653
      Yang, J.; Song, N.; Lv, X.; Jia, Q. UV-Light-Induced Synthesis of PEI-CuNCs Based on Cu2+-Quenched Fluorescence Turn-On Assay for Sensitive Detection of Biothiols, Acetylcholinesterase Activity and Inhibitor. Sens. Actuators B Chem. 2018, 259, 22632,  DOI: 10.1016/j.snb.2017.12.045
    654. 654
      Chen, Y.; Liu, W.; Zhang, B.; Suo, Z.; Xing, F.; Feng, L. Sensitive and Reversible Perylene Derivative-Based Fluorescent Probe for Acetylcholinesterase Activity Monitoring and Its Inhibitor. Anal. Biochem. 2020, 607, 113835,  DOI: 10.1016/j.ab.2020.113835
    655. 655
      Xu, J.; Yu, H.; Hu, Y.; Chen, M.; Shao, S. A Gold Nanoparticle-Based Fluorescence Sensor for High Sensitive and Selective Detection of Thiols in Living Cells. Biosens. Bioelectron. 2016, 75, 17,  DOI: 10.1016/j.bios.2015.08.007
    656. 656
      Mu, H.; Miki, K.; Kubo, T.; Otsuka, K.; Ohe, K. Substituted meso-Vinyl-BODIPY as Thiol-Selective Fluorogenic Probes for Sensing Unfolded Proteins in the Endoplasmic Reticulum. Chem. Commun. 2021, 57 (14), 181821,  DOI: 10.1039/D0CC08160D
    657. 657
      Huang, S.-T.; Ting, K.-N.; Wang, K.-L. Development of a Long-Wavelength Fluorescent Probe Based on Quinone-Methide-Type Reaction to Detect Physiologically Significant Thiols. Anal. Chim. Acta 2008, 620 (1), 1206,  DOI: 10.1016/j.aca.2008.05.006
    658. 658
      Kosower, N. S.; Kosower, E. M.; Newton, G. L.; Ranney, H. M. Bimane Fluorescent Labels: Labeling of Normal Human Red Cells under Physiological Conditions. Proc. Natl. Acad. Sci. U.S.A 1979, 76 (7), 33826,  DOI: 10.1073/pnas.76.7.3382
    659. 659
      Qi, S.; Liu, W.; Zhang, P.; Wu, J.; Zhang, H.; Ren, H.; Ge, J.; Wang, P. A Colorimetric and Ratiometric Fluorescent Probe for Highly Selective Detection of Glutathione in the Mitochondria of Living Cells. Sens. Actuators B Chem. 2018, 270, 45965,  DOI: 10.1016/j.snb.2018.05.017
    660. 660
      Li, X.; Huo, F.; Yue, Y.; Zhang, Y.; Yin, C. A Coumarin-Based “Off-On” Sensor for Fluorescence Selectivily Discriminating GSH from Cys/Hcy and Its Bioimaging in Living Cells. Sens. Actuators B Chem. 2017, 253, 429,  DOI: 10.1016/j.snb.2017.06.120
    661. 661
      Long, L.; Lin, W.; Chen, B.; Gao, W.; Yuan, L. Construction of a FRET-Based Ratiometric Fluorescent Thiol Probe. Chem. Commun. 2011, 47 (3), 8935,  DOI: 10.1039/C0CC03806G
    662. 662
      Chen, H.; Tang, Y.; Ren, M.; Lin, W. Single near-Infrared Fluorescent Probe with High- and Low-Sensitivity Sites for Sensing Different Concentration Ranges of Biological Thiols with Distinct Modes of Fluorescence Signals. Chem. Sci. 2016, 7 (3), 1896903,  DOI: 10.1039/C5SC03591K
    663. 663
      Hall, H. K., Jr. Correlation of the Nucleophilic Reactivity of Aliphatic Amines. J. Org. Chem. 1964, 29 (12), 35393544,  DOI: 10.1021/jo01035a024
    664. 664
      Inoue, N. Selective Detection of Carcinogenic Aromatic Diamines in Aqueous Solutions Using 4-(N-Chloroformylmethyl-N-Methylamino)-7-Nitro-2,1,3-Benzoxadiazole (NBD-COCl) by HPLC. Anal. Sci. 2017, 33 (12), 13751380
    665. 665
      You, W. W.; Haugland, R. P.; Ryan, D. K.; Haugland, R. P. 3-(4-Carboxybenzoyl)quinoline-2-carboxaldehyde, a Reagent with Broad Dynamic Range for the Assay of Proteins and Lipoproteins in Solution. Anal. Biochem. 1997, 244 (2), 277,  DOI: 10.1006/abio.1996.9920
    666. 666
      Wu, J.; Chen, Z.; Dovichi, N. J Reaction Rate, Activation Energy, and Detection Limit for the Reaction of 5-Furoylquinoline-3-Carboxaldehyde with Neurotransmitters in Artificial Cerebrospinal Fluid. J. Chromatogr. B Biomed. Appl. 2000, 741 (1), 85,  DOI: 10.1016/S0378-4347(99)00540-X
    667. 667
      Pinto, D.; Arriaga, E. A.; Schoenherr, R. M.; Chou, S. S.-H.; Dovichi, N. J. Kinetics and Apparent Activation Energy of the Reaction of the Fluorogenic Reagent 5-Furoylquinoline-3-Carboxaldehyde with Ovalbumin. J. Chromatogr. B 2003, 793 (1), 107,  DOI: 10.1016/S1570-0232(03)00368-4
    668. 668
      Arrell, M. S.; Kalman, F. Estimation of Protein Concentration at High Sensitivity Using SDS-Capillary Gel Electrophoresis-Laser Induced Fluorescence Detection with 3-(2-Furoyl)quinoline-2-Carboxaldehyde Protein Labeling. Electrophoresis 2016, 37 (22), 2913,  DOI: 10.1002/elps.201600246
    669. 669
      Wu, D.; Liu, Y.; Zheng, F.; Rong, S.-Q.; Yang, T.; Zhao, Y.-K.; Yang, R.-W.; Zou, P.; Wang, G.-T. Detection of Organic Amines Using a Ratiometric Chromophoric Fluorescent Probe with a Significant Emission Shift. J. Chem. Res. 2020, 44, 475481,  DOI: 10.1177/1747519820902944
    670. 670
      Zhang, J.; Zhou, J.; Xu, G.; Ni, Y. Stereodivergent Evolution of KpADH for the Asymmetric Reduction of Diaryl Ketones with Para-Substituents. Mol. Catal. 2022, 524, 112315,  DOI: 10.1016/j.mcat.2022.112315
    671. 671
      Ribeaucourt, D.; Bissaro, B.; Lambert, F.; Lafond, M.; Berrin, J.-G. Biocatalytic Oxidation of Fatty Alcohols into Aldehydes for the Flavors and Fragrances Industry. Biotechnol. Adv. 2022, 56, 107787,  DOI: 10.1016/j.biotechadv.2021.107787
    672. 672
      O'Brien, P. J.; Siraki, A. G.; Shangari, N. Aldehyde Sources, Metabolism, Molecular Toxicity Mechanisms, and Possible Effects on Human Health. Crit. Rev. Toxicol. 2005, 35 (7), 609662,  DOI: 10.1080/10408440591002183
    673. 673
      Corley, R. A.; Kabilan, S.; Kuprat, A. P.; Carson, J. P.; Jacob, R. E.; Minard, K. R.; Teeguarden, J. G.; Timchalk, C.; Pipavath, S.; Glenny, R.; Einstein, D. R. Comparative Risks of Aldehyde Constituents in Cigarette Smoke Using Transient Computational Fluid Dynamics/Physiologically Based Pharmacokinetic Models of the Rat and Human Respiratory Tracts. Toxicol. Sci. 2015, 146 (1), 6588,  DOI: 10.1093/toxsci/kfv071
    674. 674
      Ahmed Laskar, A.; Younus, H. Aldehyde Toxicity and Metabolism: The Role of Aldehyde Dehydrogenases in Detoxification, Drug Resistance and Carcinogenesis. Drug Metab. Rev. 2019, 51 (1), 4264,  DOI: 10.1080/03602532.2018.1555587
    675. 675
      Fritz, K. S.; Petersen, D. R. An Overview of the Chemistry and Biology of Reactive Aldehydes. Free Radical Biol. Med. 2013, 59, 8591,  DOI: 10.1016/j.freeradbiomed.2012.06.025
    676. 676
      Urbansky, E. T. Carbinolamines and Geminal Diols in Aqueous Environmental Organic Chemistry. J. Chem. Ed. 2000, 77 (12), 1644,  DOI: 10.1021/ed077p1644
    677. 677
      Mukherjee, K.; Chio, T. I.; Gu, H.; Banerjee, A.; Sorrentino, A. M.; Sackett, D. L.; Bane, S. L. Benzocoumarin Hydrazine: A Large Stokes Shift Fluorogenic Sensor for Detecting Carbonyls in Isolated Biomolecules and in Live Cells. ACS Sens. 2017, 2 (1), 128134,  DOI: 10.1021/acssensors.6b00616
    678. 678
      Molina-Espeja, P.; Canellas, M.; Plou, F. J.; Hofrichter, M.; Lucas, F.; Guallar, V.; Alcalde, M. Synthesis of 1-Naphthol by a Natural Peroxygenase Engineered by Directed Evolution. ChemBioChem 2016, 17 (4), 341349,  DOI: 10.1002/cbic.201500493
    679. 679
      Guo, H.-M.; Minakawa, M.; Tanaka, F. Fluorogenic Imines for Fluorescent Detection of Mannich-Type Reactions of Phenols in Water. J. Org. Chem. 2008, 73 (10), 39643966,  DOI: 10.1021/jo8003293