Next Article in Journal
Talaromyces sp. Ethyl Acetate Crude Extract as Potential Mosquitocide to Control Culex pipiens quinquefasciatus
Previous Article in Journal
Mathematical Modelling and Optimization for Facile Synthesis of Structured Activated Carbon (ACs) from Adansonia kilima (Baobab) Wood Chips Integrating Microwave-Assisted Pyrolysis for the Elimination of Lead (II) Cations from Wastewater Effluents
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Drug Discovery Based on Fluorine-Containing Glycomimetics

1
Department of Pharmacy, Changzhi Medical College, No. 161, Jiefang East Street, Changzhi 046012, China
2
State Key Laboratory of Natural and Biomimetic Drugs, School of Pharmaceutical Sciences, Peking University, Xue Yuan Rd. No. 38, Beijing 100191, China
*
Author to whom correspondence should be addressed.
These authors contributed equally to this work.
Molecules 2023, 28(18), 6641; https://doi.org/10.3390/molecules28186641
Submission received: 31 August 2023 / Revised: 12 September 2023 / Accepted: 13 September 2023 / Published: 15 September 2023
(This article belongs to the Section Medicinal Chemistry)

Abstract

:
Glycomimetics, which are synthetic molecules designed to mimic the structures and functions of natural carbohydrates, have been developed to overcome the limitations associated with natural carbohydrates. The fluorination of carbohydrates has emerged as a promising solution to dramatically enhance the metabolic stability, bioavailability, and protein-binding affinity of natural carbohydrates. In this review, the fluorination methods used to prepare the fluorinated carbohydrates, the effects of fluorination on the physical, chemical, and biological characteristics of natural sugars, and the biological activities of fluorinated sugars are presented.

1. Introduction

Carbohydrates play major structural, physical, and biological roles in organisms. They often form complex and diverse glycans on the cell surface by interacting with protein and lipid scaffolds. The glycan molecules bind to proteins and actively participate in various biological processes, such as embryogenesis, adhesion, immunity, inflammation, cancer metastasis, and host–pathogen interactions [1,2,3]. The composition of glycans changes during cell differentiation and tissue development. Disease states and the degree of inflammation also influence the compositional changes. These changes can be attributed to the altered degree of the expression of glycosidase and/or glycosyltransferase (GT) in cells. The generation of this dynamic expression is an integral part of the cell-to-cell communication process, and it can potentially provide new targets for disease treatment. Carbohydrate–protein binding is achieved by exploiting low-energy interactions (such as hydrogen bonding, salt bridges, and metal chelation), which cannot compensate for the high enthalpy cost of the desolvation of polar substrates and shallow protein-binding sites. The high polar surface area of natural carbohydrates hinders the passive penetration of the molecules into the intestinal membrane, making them non-bioavailable orally. It has also been observed that natural carbohydrates exhibit inherently poor pharmacokinetic properties, limiting their therapeutic potential [4,5].
Natural carbohydrates can be modified to improve their drug-like properties, and “carbohydrate-like compounds” can be designed to mimic the structure and function of natural carbohydrates to improve their affinity, pharmacokinetic properties, and bioavailability [6,7]. Fluorine, which exhibits unique properties, has been widely used in drug design and development. Its high electronegativity makes it a powerful tool for modulating the pKa and electron density of the groups present in its proximity, and it can serve as a considerable element to control molecular conformation [8,9,10]. Pharmaceuticals containing fluorine exhibit the property of tunable lipophilicity and increased metabolic stability. It is noteworthy that the oxidative metabolism of these molecules can be prevented. The strategy of the fluorination of carbohydrates has long been used to investigate protein-carbohydrate and carbohydrate-carbohydrate interactions. This method has been used to investigate the contributions of individual sugar alcohol groups [11,12,13] or design mechanism-based inhibitors [14,15]. The excellent NMR (nuclear magnetic resonance) properties of the 19F nucleus have been analyzed to study protein-carbohydrate binding at the molecular level [16,17,18,19,20,21,22,23,24]. The method of fluoridation is also used in drug discovery programs and to develop synthetic carbohydrate vaccines [25,26,27,28,29,30]. Fluorination could optimize the physicochemical properties, absorption ability, and distribution properties of natural sugars, and the introduction of fluorine can also help improve the binding affinity and pharmacokinetic properties of natural sugars [31,32]. This review aims to provide up-to-date information on the synthesis method, properties, and biological applications of fluorinated carbohydrates.

2. Synthesis of Fluorine-Containing Glycomimetics

In essence, the fluorine-based carbohydrate modification method is followed to substitute various atoms on the molecular skeleton with fluorine atoms or fluorine-containing groups such as polyfluorene. The fluorinating reagents used for fluorination include deoxyfluorination agents: DAST (diethylaminosulfur trifluoride, 1) [33,34]; nucleophilic fluorination agents: TASF (tris(dimethylamino)sulfonium difluorotrimethylsilicate, 2) [35]; TBAF (tertbutylammonium fluoride, 3) [36]; MFn (Metal fluorides, 4) [37]; KHF2 (5) [38]; anhydrous hydrogen fluoride system (6) [39]; electrophilic fluorination agents: SelectFluor (1-chloromethyl-4-fluoro-1,4-diazoniabicyclo[2.2.2]octane bis(tetrafluoroborate), also abbreviated as F-TEDA-BF4, 7) [40]; NFSI (N-fluorobenzenesulfonimide, 8) [41]; etc. (Figure 1). With the use of these fluorinating reagents above, the reactions for fluorination of sugars could be divided into nucleophilic reactions, electrophilic addition reactions, radical reactions, metal-catalyzed reactions, and de novo synthesis with fluorine-containing blocks. In this subsection, the synthesis methods will be described in detail according to the reaction types.

2.1. Fluorination of Carbohydrates via Nucleophilic Fluorination

A large number of fluorine-containing carbohydrates are synthesized via nucleophilic reactions with fluoride ions. Depending on the reaction condition, fluoride ions can act as good nucleophilic reagents in polar non-protonic solvents [42]. Currently, fluorinated reagents such as DAST (diethylaminosulfur trifluoride), TASF (tris(dimethylamino)sulfonium difluorotrimethylsilicate), KHF2, etc., are mainly used to provide nucleophilic fluorine atoms, which subsequently replace the hydroxyl or sulfonate groups on the sugar ring to obtain fluorinated sugars. These are the most popular methods for the preparation of fluorinated carbohydrates [43,44,45,46] (Figure 2a–c). The fluorination of the equatorial hydroxyl group obtained 10 with fluorine nucleophilic substituted in the axial position, which undergoes an SN2 reaction mechanism. The stereo-configuration of the fluorination product is generally flipped from the starting material.
While reagents like DAST and TASF have been traditionally used in fluorination reactions, researchers have encountered certain drawbacks associated with their use. For example, protecting group migration can be a significant issue in some fluorination reactions, leading to reduced yields or even the absence of the desired target product [47]. DFMBA is considered milder compared to reagents like DAST and exhibits improved compatibility with various protecting groups [48,49], such as acetyl, methyl, silyl, benzyl, etc. (Figure 2d). In addition to the above-mentioned organic fluorinating agents, other inorganic fluorine-containing compounds, such as hydrogen fluoride, TBAF, CsF, AgF, iodine–fluorine compounds, etc., can also provide nucleophilic fluoride ions [45]. Cheeseman’s group successfully synthesized fluorinated furanose 18 using CsF as the fluorinating agent. This method allowed for the introduction of fluorine-18 into the substrate. After deprotection, the resulting product 19 was found to be suitable as a contrast agent for breast cancer (Figure 2e).
Polyfluorinated carbohydrates have gained attention due to their unique properties and potential applications in medicinal chemistry. The positional combinations for deoxy polyfluorination of pentoses and hexoses have been extensively studied in the context of nucleophilic fluorination reactions. Epoxide opening with fluoride sources is a versatile method for introducing fluorine in a regioselective manner. Thus, 2,3,4-trideoxy-2,3,4-trifluoro-D-glucopyranose (24) can be synthesized selectively, utilizing the regioselective opening of the epoxide, displacement of the triflate intermediate using Et3N·3HF, and subsequent nucleophilic fluorination with DAST (Figure 3). The synthesis of polyfluorinated carbohydrates, including ketosugars (such as sialic acids), aminosugars, and nucleosides, with controlled fluorine introduction at specific positions has garnered significant attention in recent years. This area of research has been recently reviewed [50].
The nucleophilic reaction of carbonyl-containing glycans with difluoroalkylation reagents produces difluorinated modification of glycans. This type of reaction was first reported by Bobek et al. in 1977 [51]. They utilized a DAST reagent to modify furanose 25 by introducing a difluoromethyl group, resulting in the formation of product 26 (Figure 4a). However, the yield of this reaction was low, and its applicability was limited, particularly for pyranose. In 2009, Jiménez-Barbero et al. [52] successfully synthesized difluorine-modified carboglycosides 28 using the Deoxo-Fluor reagent (Figure 4b). CBr2F2 is another commonly used reagent for difluorination, often in conjunction with additional phosphorus-containing small molecules to generate the active intermediate carbene. Slawin’s group [53] employed this method to successfully synthesize compound 30 (Figure 4c), which, upon reduction, produced the extracyclic difluoromethyl glycosides 31.
In recent years, newly developed difluoroalkylation reagents have emerged as effective tools for the difluorinated substitution modification of sugars. One commonly used and cost-effective difluoroalkylation reagent is BrCF2CO2Et [54]. Quirion’s group [55] utilized the Reformatsky reaction to efficiently and smoothly introduce the CF2CO2Et moiety into lactone 32, yielding the exocyclic difluorocarboside 33 (Figure 5a). This reaction exhibits high stereoselectivity and directly forms the β-mannose carboglycoside bond. Another reagent frequently employed for introducing 2-(ethoxycarbonyl)difluoromethyl is difluoroenol silyl ether. Quirion’s group [56], in 2009, employed this reagent in conjunction with boron trifluoride ether as a catalyst to couple with compound 34, resulting in the formation of exocyclic difluorocarboside 35 (Figure 5b).
The trifluoromethylation of carbohydrates can be achieved via the nucleophilic addition of trifluoromethyl groups to aldehydes or ketones. A commonly employed nucleophilic trifluoromethyl reagent is TMSCF3. This reagent enables the direct introduction of trifluoromethyl to the carbonyl carbon in the sugar structure. For instance, TMSCF3 was used to treat lactone 36, leading to the formation of compound 37 (Figure 6a) [57]. Moreover, TMSCF3 can also react with other carbonyl-containing sugars and subsequently undergo the Barton–McCombie dehydroxylation reaction to produce the corresponding trifluoromethyl-modified carbohydrates (Figure 6b,c) [58].

2.2. Fluorination of Carbohydrates via Electrophilic Addition Reactions

In contrast to nucleophilic reactions, electrophilic addition reactions predominantly involve glycals. The commonly utilized reagents for these reactions are N-F reagents and F-O reagents. N-F reagents are well established, commercially available, and have proven their usefulness in numerous important reactions [59]. Established and readily available N-F reagents have demonstrated their utility in a variety of important reactions [59]. For instance, the Selectfluor reagent has proven to be a highly efficient and mild reagent for fluorination, surpassing the effectiveness of the DAST reagent. In 1997, Wong’s group [60] discovered that the Selectfluor reagent could be used for the one-pot synthesis of 2-fluoroglycosides (Figure 7a). The resulting product, 2-fluoro-2-deoxyglycoside 46, was utilized as a probe to investigate the catalytic mechanisms of glycosidases and glycosyltransferases. In that same year, Dax’s group [61] successfully achieved electrophilic addition to glycals using another N-F reagent known as N-fluorobenzenesulfonamide (Figure 7b).
In addition to N-F-type fluorination reagents, certain O-F-type reagents can also be employed in the electrophilic addition reactions of glycals. In 1970, Adamson’s group [62] first synthesized 2-Fluoro-2-deoxy-D-glucose 49 using CF3OF (Figure 8). Since then, efforts have been made to improve this method. In 2006, Schrobilgen’s group [63] achieved the introduction of a fluorine atom into C-2 of glucal 34 using AcO18F (Figure 8). Interestingly, contrary to their expectations, they discovered that the resulting product was not 2-fluoro-2-deoxy-D-glucose but rather 2-fluoro-2-deoxy-D-allose. The researchers proposed that the acetyl groups at positions C-3 and C-4 were involved in the reaction, giving rise to an intermediate that caused a configuration change in the hydroxyl group at position C-3. Leveraging this method, they successfully synthesized 2-fluoro[18F]-2-deoxy-D-allose 50.

2.3. Fluorination of Carbohydrates via Radical Reactions

In addition to nucleophilic and electrophilic reactions, radical reactions have also been employed in the fluorination of sugars. In 2020, Li’s group [64] successfully introduced a fluorine atom at the C-5 position of compound 56 using an Ag(II)-initiated radical reaction (Figure 9). The authors observed that the carbon atom at the C-6 position underwent an elimination reaction. They proposed a potential reaction mechanism as follows: Ag(II) initiates a radical reaction, leading to the transformation of the C-6 hydroxyl group of the sugar into the oxygen radical intermediate 57. Subsequently, intermediate 58 is formed via β-scission, followed by the fluorination of intermediate 58 to yield the final product 59. By utilizing this strategy, the group successfully synthesized a range of rare sugars and accomplished the total synthesis of Clostridium jejuni tetrasaccharides [65].
Based on the previous work, Li’s group further advanced the method for decarboxylative fluorination of uronic acids, employing a silver radical pathway [66] (Figure 10a). This reaction demonstrates superior substrate scope and compatibility with functional groups compared to dehydroxymethylative fluorination. Notably, it successfully transforms sugars containing benzyl groups and a hexp-(1-4)-hexp moiety. In a recent development, Li’s group established a protocol for the organophotocatalytic synthesis of glycosyl fluorides without the addition of silver species. This method relies on 9-mesityl-10-methyl-acridinium (Mes-Acr+)-mediated oxidative decarboxylative fluorination of uronic acids [67] (Figure 10b). The potential of this method is promising in the synthesis of fluorinated nucleosides.
Difluorinated sugars can also be synthesized via the radical addition to glycals. Commonly employed radical initiation conditions include BEt3 and metal catalysis, and others. Quirion’s group [68] achieved the successful introduction of 2-(ethoxycarbonyl)difluoromethyl into glycal 64 using BEt3 as a radical initiator, resulting in the formation of CF2-glycoside 65 (Figure 11).
Radical addition reactions can also be employed in the preparation of trifluoromethyl-modified sugars. In 2015, Ye’s group [69] developed a photocatalytic C-H trifluoromethylation method using glycal 66 as substrates and Umemoto reagent as a trifluoromethyl source. This method exhibits mild reaction conditions, excellent compatibility with various protecting groups (such as acetyl, benzyl, and p-methoxybenzyl), and applicability to glycals featuring different types of sugars (such as glucose, galactose, rhamnose, arabinose, and lactose) (Figure 12).
In 2018, Vincent’s group [70] successfully accomplished the C-H trifluoromethylation of exo-glycal 72 using photoredox and copper catalysis methods. Two different trifluoromethyl sources, A and B, were employed under distinct reaction conditions. The method displayed excellent selectivity in producing Z-configured trifluoromethylation product 73 (Figure 13) while maintaining good compatibility with various protecting groups, such as benzyl, silyl, and acetyl.
In 2021, Ye’s group [71] used electrochemical methods for the first time in the trifluoromethyl modification of glycoconjugates. The authors used glycal 66 as the substrate, a cheap and readily available sodium trifluoromethanesulfonate reagent as the trifluoromethyl source, and added MnBr2 as the redox medium, and the reaction was carried out in acetonitrile solution. The method is mild and suitable for a wide range of glycoconjugate substrates protected by different protecting groups (Figure 14).
In 2020, Postigo’s group [72] utilized glycal 83 as a substrate, combined with iodinated perfluoroalkanes as a fluorine source, and employed Ir[dF(CF3)PPy]2(dtbPy)PF6 as a photocatalyst to achieve the synthesis of perfluoroalkyl-modified glycal 84 under blue light conditions (Figure 15).

2.4. Fluorination of Carbohydrates via Metal-Catalyzed Coupling Reactions

Pannecoucke’s group [73,74] developed a copper-catalyzed method for the synthesis of difluoroalkylated glycals. The researchers synthesized a series of difluoroalkylation-modified glycals using Cu(PF6)(MeCN)4 as the catalyst, BrCF2CO2Et as the difluoroalkylation reagent, and 1,10-phenanthroline as the ligand (Figure 16). This method exhibited good compatibility with various protecting groups (benzyl, acetyl, methyl, etc.) and could be applied to different types of sugars (glucose, galactose, rhamnose, arabinose, etc.). The proposed reaction mechanism suggests that, initially, copper(I) Cu(PF6)(MeCN)4 undergoes oxidative addition with BrCF2CO2Et to form compound 93, resulting in the trivalent copper species. Compound 93 then undergoes an addition reaction with glycal 66, selectively adding to the electron-rich C-2 position, generating the oxonium intermediate 94. Subsequently, the base removes a proton to form compound 95, which is, ultimately, reduced and eliminated to yield the desired product 89.
The introduction of trifluoromethyl to the glycal was also achieved via metal-catalyzed coupling reactions. Boutureira’s group [75] utilized a 2-iodoglycal 96 as the substrate and CuCF3 as the trifluoromethyl source to successfully obtain the trifluoromethyl-modified glycal 97 in a satisfactory yield upon heating at 50 °C (Figure 17). However, this method required the pre-preparation of 2-iodoglycal, making it somewhat tedious and limited in scope. Alternatively, metal-catalyzed coupling reactions were employed to introduce perfluoroalkyl groups into glycal. Boutureira’s group [76] utilized 2-iodoglycal 96 as the substrate and CuC2F5 as the perfluoroalkyl source, achieving the synthesis of perfluoroalkyl-modified glycal 98 with higher yields under heating at 50 °C.

2.5. Fluorination of Carbohydrates via Building Block Strategy

Fluorine-modified sugars can be synthesized using a fluorine-containing building block strategy. In 2007, Mootoo’s group [77] successfully synthesized monofluorinated glycosides using compounds 107 as the initial materials in a four-step tandem reaction, as shown in Figure 18: esterification, Takai reaction, cyclization reaction, and borohydride oxidation reaction. Through variation in the R group, a wide range of fluorinated carbohydrates with different substituents can be obtained.
In addition to the methods mentioned above, fluorocarbohydrates can be obtained by modifying the backbone of the raw material using fluorinated reagents. Figure 19a shows the introduction of difluoroalkyl into the precursor 111 via the Reformatsky reaction, followed by ring closure to achieve difluoro-C-glycosides 113 [78]. Alternatively, the difluoroalkylated product 116 can be obtained by initially introducing two fluorine atoms at the C-2 position of the sugar ring, followed by a one-step ring closure (Figure 19b) [79].
The synthesis of saccharides modified with trifluoromethyl groups using non-sugar compounds as starting materials was initially achieved by the Kobayashi group [80]. They employed the trifluoromethyl-containing compound 117 as the starting material and generated the intermediate product 118 via a Lewis acid-catalyzed aldol condensation reaction. The carbonyl group in 118 was subsequently reduced, yielding the reduced compound. Furthermore, the double bond in the product was oxidized using a dilute potassium permanganate solution to produce the dihydroxy compound 120 (Figure 20a). In a related study, Burger’s group [81] utilized methyl trifluoropyruvate 122 as a source of trifluoromethyl groups. This compound was catalytically coupled with compound 121 using titanium trichloride, leading to the formation of compound 123. Subsequently, compound 123 underwent a ring-closing reaction under acidic conditions, resulting in the formation of compound 124. Finally, compound 125 was synthesized in a single vessel via a reductive reaction (Figure 20b).
The synthesis of sugars modified with polyfluorinated groups was achieved using non-sugar compounds that contained polyfluorinated modifications. In 2013, Vincent’s group [82] utilized compound 126, which featured a CF2CF2 structure, as the starting material. Through a ring-closing reaction, they successfully obtained the final product 128 (Figure 21).
The above section discusses the chemical synthesis of fluorine-containing glycomimetics. Additionally, fluorinated carbohydrates can be enzymatically produced from chemically fluorinated sugars. Enzymes such as kinases, phosphorylases, and nucleotidyltransferases can mediate this process. A review paper [83] provides a comprehensive overview of the current state of the art in the chemoenzymatic production of fluorinated carbohydrates. Due to space limitations, the detailed content of this review will not be discussed here.

3. Impact of Fluorination on the Physical Properties of Carbohydrates

The size of fluorine atoms is comparable to that of oxygen atoms, but they have a higher ionization potential. This higher ionization potential makes fluorine atoms less capable of accepting hydrogen bonds compared to oxygen atoms. As a result, the C-F group is more hydrophobic than the alcohol group. Introducing fluorine atoms into molecules can aid in enhancing biomolecular recognition via the process of hydrophobic desolvation. Carbohydrates have a strong hydrogen-bonding network with solvent molecules, and their low binding affinity can be attributed to the high energy required for desolvation. Deoxyfluorination, the process of removing hydroxyl groups and replacing them with fluorine atoms in carbohydrates, can improve the binding affinity of these molecules towards target units. The high electronegativity of fluorine limits the polarizability of the C-F bond, reduces the solvation of the ligand, increases hydrophobicity, decreases the penalty of desolvation, and enhances binding affinity. Additionally, polar C-F bonds maintain necessary electrostatic interactions between ligands and binding sites such as cations and dipoles.
The C-F bond (1.35 Å) is longer than the C-H bond (1.09 Å) and shorter than the C-OH bond (1.43 Å). CF2 is considered a more favorable option compared to CHF for replacing CHOH, as it has lower steric and desolvation costs. Monodeoxyfluorinated sugar mimics showed improved affinity for various targets after modification, and a significantly high affinity was observed when the compounds were modified with polydeoxyfluoride. It has been observed that hexose analogs containing multiple fluorine substitutes exhibit stronger binding compared to normal substrates. For example, 2-deoxy-2,2-difluoro-D-glucose showed preferential binding to yeast hexokinase, and 2-deoxy-2-fluoro-α-D-glucose (129) displayed enhanced inhibition of glycogen phosphorylase [12] (Figure 22a). Another example is hexa-fluorinated-glucose derivatives such as 1-hydroxy-5-hydroxymethyl-2,2,3,3,4,4-hexafluorane (132), which exhibited approximately ten times higher permeability across red blood cell membranes compared to 3-deoxy-3-fluoro-D-glucose (133), indicating enhanced affinity for the transporter [9]. These results indicate the importance of the polar hydrophobicity of fluoro-sugars.

3.1. The Impact on Lipophilicity

The lipophilicity of a compound is determined using its partition coefficient (P), which is measured between 1-octanol and water and expressed as logP. Deoxyfluorination induces increased hydrophobicity, resulting in a significant rise in logP. The magnitude of this change is dependent on the position and number of fluorine atoms integrated into the molecule. The Linclau group employed a 19F NMR-based approach to determine logP values for various deoxyfluorinated carbohydrate derivatives (Figure 23) [84,85]. Generally, each deoxyfluorination leads to a logP increase of approximately 1 unit. The site of fluorination (e.g., 6-deoxy-6-fluoro-glucose, logP = −2.36) and the stereochemistry of the carbohydrate (e.g., 2-deoxy-2-fluoro-galactose, logP = −2.37) also influence the logP values. The lipophilicity of 2,3,4-trideoxy-2,3,4-trifluoro-monosaccharide derivatives can vary considerably [31], and the all-cis derivative exhibits the lowest hydrophobicity, likely due to its large molecular dipoles.
St-Gelais and colleagues recently conducted a comprehensive investigation into the lipophilicity of monofluorinated, difluorinated, and trifluorinated glucopyranose analogs, with modifications at positions C-2, C-3, C-4, and C-6. This study complements the previous work by Linclau and colleagues, who also explored the lipophilicity of fluorinated carbohydrates. In general, analogs bearing fluorine atoms at C-6 exhibit the highest hydrophilicity, while compounds with vicinal polyfluorinated motifs display the greatest lipophilicity. This observation can potentially be attributed to the considerable reduction in the polar surface area of the molecules [86]. The calculated solvation-free energy demonstrates a good correlation with the experimentally determined log P values. Although the lipophilicity conferred by polyfluorination on sugars has the potential to improve affinity and membrane permeability, an excessive degree of lipophilicity can compromise the efficacy and safety of drugs, leading to increased attrition rates [87,88]. Therefore, striking an appropriate balance is crucial.

3.2. The Impact on Conformation

The small van der Waals radius and high electronegativity of the fluorine atom result in the formation of a C–F bond with a strong dipole moment (μ C–F = 1.41 D) and a low-lying C–F σ* orbital available for hyperconjugation. These factors have various effects on the conformation of the molecules. The preference for certain conformations can be exploited to stabilize a desired binding conformation and improve affinity. However, it can also inadvertently favor an undesired conformation, which negatively affects binding [89,90].

3.2.1. The Impact on the Chair Conformation of Pyranose

The conformation of fluorinated carbohydrate derivatives has been extensively studied using X-ray diffraction and NMR techniques [31]. In general, the introduction of an ectopic F atom affects the 4C1/1C4 equilibrium, and it has been observed that fluorination does not result in a significant degree of distortion in the pyranose conformation, even when the compounds are multifluorinated (e.g., 139) (Figure 24). The review article [31] published by Linucau et al. provides an excellent summary of the different ways in which fluoridation affects carbohydrate structure and conformation. Analysis of the crystal structures of free trideoxytrifluorinated sugars such as glucose 140 [91,92] and the corresponding altrose and galactose molecules (all β-anomers) reveals that these molecules present 4C1-conformations. However, slight distortions in the structures are possible when 1,3-coaxial C-F bonds are present in the molecules, as shown for 141 [93,94]. Hexafluoropyranose 142 presents the 4C1-conformation, as the repulsive forces between the two F atoms cause them to move away from each other. Conformational analysis of α-and β-difluoro-glucose, galactose, and mannose derivatives reveals that the CF2 unit in the ring does not significantly influence the properties of the 4C1 chair. The corresponding alkyl bioisostere of a CF3 group has been studied, and it has been reported that based on the size and shape of the group, ethyl or an isopropyl group is the closest match [95,96]. In pentopyranosyl fluorides, the strong anomeric effect involving the anomeric C–F bond promotes chair inversion. This phenomenon has also been observed in the case of the triacetylated β-D-xylopyranosyl fluoride 143, where ring inversion results in the generation of a structure with three axial ester groups.

3.2.2. The Impact on the Conformation of the Glycosidic Bond

In O-glycosides 144, stereoelectronic stabilization caused by an overlap between the lone pairs of electrons on the anomeric oxygen and the σ*C1–O1 orbital (the exo-anomeric effect) and steric repulsion result in the generation of a specific exo conformation of the glycosidic bond (Figure 25a). Solitary pairs on the outer ring atoms are absent in CH2-glycosides, and the polarity of the C1–C2 and C1–C5a units are comparable. The conformation distribution is influenced by the minimum spatial repulsion force. However, the formation of CF2-glycosides introduces the possibility of the generation of the fluoro-gauche effect. This stereoelectronic effect (Figure 25b) is based on the hyperconjugation of the C–F bond to the C1–H1 and C1–C2 bonds and is closely related to the ectopic effect that influences the conformation of fluorinated molecules [97]. Figure 25b presents the schematic representation of the two conformations, 145 and 146, that allow the generation of the most stable hyperconjugated states. The evaluations of the ectopic effects exerted by CHF- and CF2-glycosides indicate the predominance of non-natural conformations. The CF2-glycosidic linkage has been incorporated into several different glycomimetics structures. The CF2-linked α-(2,3)-sialylgalactose compound has been used as a non-hydrolyzable mimic of sialylated gangliosides (such as GM3 and GM4). A CF2-linked GM4 derivative exhibited inhibitory activity against NEU2 (IC50 = 754 mm) and NEU4 (IC50 = 930 mm). It was observed that this compound successfully inhibited the proliferation of human lymphocytes [98].
In carbasugars, C1–C2 and C1–C5a have the same polarity; therefore, no preferential conformation is observed. Leclerc et al. proposed a CF2 analog of carbasugars in which the CF2 group replaced the endocyclic oxygen atom to restore the exo conformation attributable to a polarization of the C1–CF2 bond [99] (Figure 26). Carbasugars lack a hemiacetal structure, making these derivatives significantly less susceptible to hydrolysis than derivatives with an exocyclic CF2 group replaced. Studies have shown that the replacement of the endocyclic oxygen atom with a CF2 group in maltose restored stereoelectronic stabilization caused by the lone pairs of electrons on the anomeric oxygen and the σ*C1–O1 orbital. The replacement resulted in the preferential generation of the exo conformation [100]. The electron-absorbing CF2 group replaces the endocyclic CH2 unit and reduces the electron density on the endocyclic-O atom, thus reducing the degree of the endo-anomeric effect. This suggests that electron-absorbing substituents can reduce the strength of the spatial torsion interaction and help maintain a natural conformational equilibrium. Fluorinated carbasugars have the added advantage of increasing the overall lipophilicity, but this property must be carefully evaluated as they can also influence the acidity of neighboring functional groups and potentially generate novel O–H···F intramolecular interactions or 1,3-diaxial steric clashes.
The conformational mobility of the iduronic acid (L-IdoA) moiety in heparin oligosaccharides, both in free and bound states, has been studied. Distinct conformations of the L-IdoA ring are recognized depending on the protein receptor [101]. For instance, antithrombin III (AT-III) recognizes the skew boat conformer [102], while the L-IdoA rings of a heparin hexasaccharide maintain the chair–skew boat flexibility when bound to the fibroblast growth factor-1 (FGF-1) unit [103]. Fluorination can significantly affect the biological activity of IdoA-based glycomimetics, and this can be attributed to the conformational plasticity of L-IdoA that mediates receptor binding. Gem-difluorocarbasugar analogs based on L-Ido and Glc (control) have been studied using a combination of NMR spectroscopy and computational techniques. The results demonstrate that the fluorocarbasugar L-IdoA analog closely mimics the conformational plasticity of native Ido, whereas the non-fluorinated carbasugar was found to be incapable of retaining this conformational flexibility [104].

3.2.3. The Impact on the Conformation of the Exocyclic Hydroxymethyl Group

The C4 configuration plays a significant role in determining the exocyclic conformation of the C5–C6 units, and conformations characterized via a regimented arrangement of C–O bonds are usually unfavorable, and this is expected to be similar in the case of deoxyfluorination. The preferred conformation of the heterocyclic fluoromethyl groups in 6-deoxy-6-fluoroglucose is shown in Figure 27. The gt conformation is the most abundant for the galactose derivative 149, and the tg conformation is present in a small amount in this system. The gt conformation is proposed to be the dominant species in solution for the 4,6-dideoxy-4,6-difluorinated GalNAc derivative 150 [105], while the gg conformation is the predominant configuration of the 4,6-difluorinated glucose derivative 151 [106]. Interestingly, the crystal structure of the 2,3,4,6-tetrafluorinated galactose derivative 152 presents the gg conformation. Therefore, it was proposed that the gt conformation was present in the solution phase [90].

3.3. The Impact on the Stability of the Glycosidic Bond

The high electronegativity and inductive ability of fluorine make it difficult for fluorinated carbohydrate derivatives to form oxocarbenium moieties [106]. This was verified in hydrolysis studies of glycosyl phosphates (where the observed order of phosphate cleavage was native, 157 > 6-fluoro, 156 > 3-fluoro, 155 > 4-fluoro, 154 > 2-fluoro, 153 (Figure 28) and dinitrophenylglycosides. However, the order was reversed (2-deoxy, 160 > 4-deoxy, 159 > 3-deoxy, 158 > native, 157) in a series of deoxygenated alkyl and aryl glucoside mimetics [107,108,109]. As both enzymatic and acid-catalyzed glycosidic cleavage typically proceeds through an oxocarbenium-like intermediate, all positions of the sugar molecule were fluorinated to improve hydrolytic stability and/or prevent enzymatic cleavage. Fluorinated glycomimetics are often used as excellent inhibitors of enzymes, and this is important for improving drug bioavailability.
The destabilizing effect on the process of formation of the oxocarbenium transition state has been widely exploited for designing inhibitors for GTs [110,111,112,113,114] and glycosidases. 2,4-dinitrophenyl-2-deoxy-2-fluoro-β-glucopyranoside 161 (Figure 29a) was the first example of a fluorinated glycoside inhibitor reported by the Withers laboratory [115]. A number of differently configured fluorinated glycosides have since been synthesized, which are hydrolyzed via the target enzyme in a manner similar to that of a natural substrate. The presence of the electron-withdrawing fluorine substituent at C-2 destabilizes both the glycosylation and the deglycosylation transition states. As a result, the rates with which the enzyme–substrate intermediate is formed as well as hydrolyzed are decreased. To compensate for the overall deactivation of the inhibitor, an activated anomeric leaving group (such as dinitro or trinitrophenate or fluoride) is usually introduced to accelerate the glycosylation step, resulting in the accumulation of covalent glycosyl–enzyme intermediates, thereby inhibiting the enzyme. These inhibitors have been shown to inhibit various types of retaining glycosidases [116,117,118,119]. The success of 2-fluoro-2-deoxyglucose as the most utilized radiotracer used for positron emission tomography (PET) can be attributed to its resistance to hydrolysis. This helps the molecule avoid metabolic degradation and enables tissue accumulation. The effects of fluorination on anomeric reactivity are presented for fluorinated maltoses 163 [120] and 164 [121] (Figure 29b). The GlgE1 maltosyl transferase from Streptomyces coelicor was studied to investigate these molecules as mechanism-based enzyme inhibitors. It was observed that compound 163 formed covalent intermediates by reacting with relevant nucleophilic Asp residues [120]. The enzyme was able to effect hydrolysis. This could only be prevented under conditions of an E423A point mutation. However, trifluorinated maltose 164 exhibited increased deactivation properties resulting from C2-difluorination, which made the compound unreactive to the enzymes [121]. It is interesting that 164 can crystallize with GlgE1 in the complex.
Phlorizin is a natural product isolated from apple tree bark with inhibitory activity against the sodium-glucose linked transporters (SGLT1 and SGLT2) (Figure 30a) and has been studied as a therapy for type II diabetes mellitus. Phlorizin has been associated with rapid hydrolytic degradation, and there is much interest in achieving selective SGLT2 inhibition. 2-Deoxy-2-fluoro derivatives of Phlorizin analogs 165 were evaluated in efforts to improve its metabolic stability and improved resistance to hydrolysis and were also observed to improve selectivity for SGLT2 (albeit with lower affinity) [122], which made it a better substrate for metabolic enzymes. In another study, fluorinated sialic acid analogue (P-3Fax-Neu5Ac) 166 (Figure 30b) was used to inhibit sialyltransferase activity since its inherent electron-withdrawing capability significantly increases the electrophilicity of the C-2 atom and destabilizes the oxocarbenium transition state. This destabilization hinders its formation, significantly slowing the enzymatic rate and effectively reducing tumor growth in vivo [123,124].

4. Application

4.1. Fluorine-Containing Sugars as Glycosyltransferase Inhibitors

4.1.1. Fluorosugars as in Cellulose Glycosyltransferase Inhibitors

Glycosyltransferase (GT) participates in the glycosylation of glycosylated receptor substrates and helps in assembling oligosaccharides. Nucleotide–sugar units are used as the activated sugar donor substrates during the process. The inhibition of GT has been a focus of research due to its biological relevance [125,126,127,128,129].
The fluorinated sialic acid analog, which acts as a transition-state inhibitor for sialic transferase, is a GT inhibitor [110]. Although it cannot cross cell membranes due to its high hydrophilicity, some of the precursors (167) can be transported across the membranes into the cytoplasm and enzymatically converted to fluorinated sugar-nucleotides, resulting in the in situ production of GT inhibitors (Figure 31a). For example, peracetylated 3-fluorosialic acid 168 (Figure 31b) [130] acts as a specific inhibitor for sialic transferase, and it can penetrate cell membranes. Studies have shown that this inhibitor can hinder the process of tumor growth and metastasis in vitro and in vivo [123,124,131]. A second-generation pre-fluorination inhibitor 169 with a carbamate group at C-5 was synthesized. This compound exhibited effective and increased SialT inhibitory activity, and this could be attributed to the efficient intracellular conversion of the molecules to their active nucleotide sugar forms [132].
These results indicate that 168 is an effective sugar-like agent, and the peracetylation form of the compound does not affect the process of sialylation. It can be inferred that the fluorine atoms in the inhibitor structure impart the inhibitory effect to the molecule. The results suggest that fluorinated sialic acid analogs inhibit the process of sialacidification and are potential tools for anticancer therapy [124].
Linclau and coworkers thoroughly investigated the process of formation of 2-deoxy-2-fluoro-fucose (GDP 170, Figure 32) in cellulose [110,130,133]. The results revealed that the accumulation of nucleotide sugar 170 not only inhibited the activity of FucTs but also shut down the biosynthesis of GDP-fucose (natural FucT donor) [130].
Recently, efforts have been made to produce non-fucosylated recombinant antibodies in CHO cells by blocking fucosylation. This was achieved by inhibiting the de novo biosynthesis of GDP fucose via the hydrolysis of petrosterol, which resulted in the transformation of peracetylated fucostatin to GDP fucostatin [134]. Non-fucosylated therapeutic antibodies are reported to exhibit better antibody-dependent cytotoxicity and improved in vivo efficacy. This blocking strategy was also used to generate low-fucosylated monoclonal antibodies from murine hybridoma cells [135]. Monofluoride and difluoride galactoside 171 (Figure 33) inhibited the biosynthesis of UDP-galactose in epithelial cells and fibroblasts, resulting in the production of metabolically induced galactose phenotypes.
Recently, a monofluorinated analog of 2,4-diacetamide-2,4,6-trideoxygalactose, which is typically found only in bacteria, was found to inhibit the ability of H. pylori to synthesize glycoproteins significantly. This resulted in a reduction in its growth and motility and a decrease in the extent of biofilm formation [136]. The proposed mode of action involves the in situ generation of glycosyltransferase inhibitors.

4.1.2. Fluorine-Containing Sugars as Glycosidase Inhibitors

Glycosidase is a family of hundreds of enzymes that hydrolyze glycosidic bonds, forming monosaccharides in living organisms. The dysfunction of glycosidase is directly related to the etiology of diseases such as diabetes. Therefore, glycosidase inhibitors are potential drugs for treating these diseases [137]. Enzyme-catalyzed hydrolysis with glycosidase retention occurs via the production of glycosylase intermediates with oxacarbenium ion-like transition states. When the highly electronegative fluorine atom replaces the hydroxyl group in the molecules (especially the group at the C-2 site) of the carbohydrate ring, the fluorine atom inhibits the formation and hydrolysis of intermediates. This can be attributed to the inducing effects exerted and the instability of the oxacarbenium ion-like transition states [138]. This results in a significantly reduced rate of glycosylation and deglycosylation. In vivo studies have demonstrated the inhibitory effect of 2-deoxy-2-fluoro-β-glucopyranoside and flumannoside on β-glucosidase and β-mannosidase [139].
2-Fluoro-deoxysugars, particularly fluoro-sialosides represented as 2,3-difluorosialoside (DFSA, trypanosoma Cruzi trans-sialase (TcTs) inhibitors) were used to study the mechanism of action of glycosylated hydrolases [140,141]. In 2013, Wong and co-workers developed DFSA 172 (Figure 33a), which contained acetylene as functional groups at five sites. It was found to be an irreversible inhibitor of viral, bacterial, and human sialidase [142]. Withers et al. studied a group of DFSAs (173; Figure 33a) as compounds with antiviral activity. They also tested the compounds for anti-influenza activity. DFSAs form a potent class of inhibitors that act through the instantaneous formation of covalent intermediates. As an enzyme, they exhibit high efficiency, and the degree of antiviral activity of these compounds is comparable to or higher than that of Zanamivir [143,144]. Notably, tyrosine residues in DFSAs 172174 played a catalytic role as a nucleophile. Since the discovery of nojirimycin 175 and its derivatives 176 (Figure 33b) [145,146], nitrogen-containing carbohydrate analogs have played important roles as inhibitors of glycoside hydrolase. Given the activity of iminosugars, many research groups have been trying to synthesize fluorinated analogs of iminosugars like 177 to achieve stronger and highly selective inhibition properties, which were summarized in previously conducted reviews [147,148,149,150].

4.2. Synthetic Vaccines and Therapeutic Antibodies

Synthetic vaccines containing natural sugar antigens face the problem of enzymatic degradation in vivo. However, this issue can be overcome using fluoridation. Fluorine has unique properties that can enhance ligand binding affinity and improve immunogenicity. The similar size of fluorine and hydrogen allows for the elicitation of cross-reactive antibodies, while the super-hydrophobicity of fluorocarbons and the unique induction effect and “polar hydrophobicity” of fluorine are often utilized to enhance the binding affinity of the molecules. The process of antigen fluorination has become an effective means of improving the affinity of TCR without significantly interfering with antigen composition or structure.
Epithelial tumor cells are often present as truncated and/or hyper-sialylated glycan in glycoproteins. These glycan structures are often the target of anticancer vaccines. However, these exhibit poor metabolic stability. Hoffmann-Roder and Kunz’s groups reported that fluorinated Thomsen–Friedenreich (TF) conjugates, such as 178 and 179 (Figure 34), generated a strong immune response in mice, and their antibodies were shown to bind strongly to epithelial tumor cells (MCF-7) [25,26,27,151,152,153]. Subsequently, it was found that immunization against a wider range of TF-MUC analogs fluorinated at the 6- and 2- positions produced antisera that exhibited little difference in their binding ability to bind to fluorinated and nonfluorinated antigens.
Our research group conducted a systematic and thorough investigation into the structural modification of STn antigens [154,155,156]. Three N-acetyl-fluoro-modified derivatives (Figure 35) were conjugated with keyhole limpet hemocyanin (KLH), and these generated a high titer immune response. The degree of response was 3–5 times higher than the degree of response generated via the natural STn–KLH system. These N-acetyl-fluoro-modified STn derivatives were site-modified onto the MUC1 glycopeptide moiety and conjugated with KLH. The results from immunological experiments revealed that the structurally modified glycoantigen could potentially improve the immunogenicity of the MUC1 glycopeptides [157]. Other TACAs, such as Tn [158], TF [159], GM3 [160], and KH-1 [161], were also fluorinated and conjugated with protein carriers to develop vaccines, which exhibited strong immune cross-reaction.
Globo H is a tumor-associated carbohydrate antigen (TACA), and clinical trials have been conducted to evaluate vaccines based on Globo H for cancers, including breast, ovarian, and prostate cancer [162]. This vaccine typically induces higher IgM titers (compared to IgG titers). A deoxyfluoro derivative of Globo H was developed that elicited IgG antibodies with titers comparable to those elicited by native Globo H to address this issue. This derivative also recognized Globo H, its related epitopes (stage-specific embryonic antigens 3 and 4), and Globo H-expressing tumor cells [162], leading to complement-dependent cell cytotoxicity.
The process of fluoridation is also being utilized to develop HIV vaccines. Efforts have been made to cluster sugar antigens to enhance their immunogenicity and affinity for 2G12. For example, Danishefsky [163] linked oligomannosaccharide Man9(GlcNAc)2 to OMPC via a cyclic peptide skeleton, which resulted in the formation of a sugar complex that could strongly react with 2G12 antibodies [164]. However, this complex did not induce the production of the corresponding bnAbs units in guinea pigs and rhesus monkeys [165]. Studies have shown that inoculating different mixtures of immunogens with the same epitope can improve the antibody’s affinity toward antigen epitopes. Based on this, our research group developed a highly convergent and effective synthesis strategy for Man5 and its monofluoro-modified, trifluoro-modified, and S-linked analogs, which could be combined with CRM197 to develop an anti-HIV vaccine candidate [166]. The results revealed that the modified analogs elicited strong antibody responses. However, the induced antibodies were unable to bind to the native gp120 antigen. This could be potentially attributed to immunotolerance mechanisms that suppress the generation of immune response for Man5-related structures and the conformation of sugar epitopes on synthetic glycoconjugates. These conformations differed from the conformations of natural sugar epitopes on gp120. The results provide significant insights into the process of development of HIV vaccines.

5. Conclusions

Carbohydrates are widely found in nature and have significant involvement in physiological processes. Therefore, the development of methods to control these processes is of great interest for therapeutic applications. However, progress in the field of carbohydrate-based drugs has been limited. This can be attributed to their high polarity, resulting in low binding affinities, as well as their poor pharmacokinetic properties, such as short residence time and rapid renal excretion. In recent years, fluorine has been introduced into more than one-third of newly developed small-molecule drugs. This has provided valuable insights into the design of glycomimetics as a potential solution to address the challenges associated with carbohydrate-based drugs. Despite there being growing evidence of the environmental accumulation of certain per- and polyfluorinated alkyl substance materials and further concerns for their bioaccumulation and potential toxicological effects [167], fluorine modification continues to be irreplaceably important in new drug research and development.
This review focuses on the synthesis of fluoro-sugars and explores the effects of fluorination on their properties and biological activities. Extensive research has been conducted on various methods for synthesizing fluorinated glycomimetics, including nucleophilic substitution, electrophilic addition reactions, and radical fluorination. However, there remains a need for the development of milder, stereoselective, and regioselective fluorine substitution methods. Substituting fluorine for the hydroxyl group can maintain certain electrostatic interactions while reducing overall polarity through increased hydrophobicity. This modification has demonstrated significant utility in the development of glycosidase inhibitors, synthetic vaccines, and therapeutic antibodies. It enhances residence time and binding affinity, thereby improving the efficacy of these compounds.
Due to the extensive body of research on glycomimetics and the acknowledged significance of carbohydrates in transportation and delivery, there exists considerable potential to investigate the possible biological activities of fluoro-sugars, provided that appropriate fluorinated modifications are undertaken. Furthermore, given the intensive research on glycan synthesis and biological functions, the development of fluorinated glycomimetics is anticipated to enhance the application of carbohydrates in diagnostic, therapeutic, and vaccine research.

Author Contributions

Conceptualization, D.X.; methodology, X.W. and P.W.; software, D.X.; validation, D.X. and F.L.; formal analysis, D.X.; investigation, X.W. and P.W.; resources, D.X.; data curation, F.L.; writing—original draft preparation, X.W. and P.W.; writing—review and editing, X.W. and P.W.; visualization, P.W.; supervision, D.X. and X.Y.; project administration, D.X.; funding acquisition, D.X. and F.L. All authors have read and agreed to the published version of the manuscript.

Funding

This work was financially supported by the National Natural Science Foundation of China (Nos. 21977003, 22007003).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

Sample Availability

Samples of the compounds are not available from the authors.

References

  1. Dube, D.H.; Bertozzi, C.R. Glycans in cancer and inflammation--potential for therapeutics and diagnostics. Nat. Rev. Drug Discov. 2005, 4, 477–488. [Google Scholar] [CrossRef] [PubMed]
  2. Taylor, S.L.; McGuckin, M.A.; Wesselingh, S.; Rogers, G.B. Infection’s sweet tooth: How glycans mediate infection and disease susceptibility. Trends Microbiol. 2018, 26, 92–101. [Google Scholar] [CrossRef] [PubMed]
  3. Fuster, M.M.; Esko, J.D. The sweet and sour of cancer: Glycans as novel therapeutic targets. Nat. Rev. Cancer 2005, 5, 526–542. [Google Scholar] [CrossRef] [PubMed]
  4. Ernst, B.; Magnani, J.L. From carbohydrate leads to glycomimetic drugs. Nat. Rev. Drug Discov. 2009, 8, 661–677. [Google Scholar] [CrossRef] [PubMed]
  5. Magnani, J.L. Glycomimetic drugs—A new source of therapeutic opportunities. Discov. Med. 2009, 8, 247–252. [Google Scholar] [PubMed]
  6. Hevey, R. Bioisosteres of Carbohydrate Functional Groups in Glycomimetic Design. Biomimetics 2019, 4, 53. [Google Scholar] [CrossRef] [PubMed]
  7. Hevey, R. Strategies for the Development of Glycomimetic Drug Candidates. Pharmaceuticals 2019, 12, 55. [Google Scholar] [CrossRef] [PubMed]
  8. Shah, P.; Westwell, A.D. The role of fluorine in medicinal chemistry. J. Enzyme. Inhib. Med. Chem. 2007, 22, 527–540. [Google Scholar] [CrossRef]
  9. Biffinger, J.C.; Kim, H.W.; DiMagno, S.G. The polar hydrophobicity of fluorinated compounds. ChemBioChem 2004, 5, 622–627. [Google Scholar] [CrossRef]
  10. Alabugin. Stereoelectronic effects and general trends in hyperconjugative acceptor ability of σ bonds. J. Am. Chem. Soc. 2002, 124, 3175–3185. [Google Scholar] [CrossRef]
  11. Glaudemans, C.P.; Kovac, P. Probing the combining site of monoclonal IgA J539 using deoxyfluoro- and other galactosides as ligands. Mol. Immunol. 1985, 22, 651–653. [Google Scholar] [CrossRef] [PubMed]
  12. Street, I.P.; Armstrong, C.R.; Withers, S.G. Hydrogen bonding and specificity. Fluorodeoxy sugars as probes of hydrogen bonding in the glycogen phosphorylase-glucose complex. Bio Chem. Eur. J. 1986, 25, 6021–6027. [Google Scholar] [CrossRef] [PubMed]
  13. Glaudemans, C.P.J. Mapping of subsites of monoclonal, anti-carbohydrate antibodies using deoxy and deoxyfluoro sugars. Chem. Rev. 2002, 91, 25–33. [Google Scholar] [CrossRef]
  14. Tysoe, C.; Withers, S.G. Fluorinated mechanism-based inhibitors: Common themes and recent developments. Curr. Top. Med. Chem. 2014, 14, 865–874. [Google Scholar] [CrossRef] [PubMed]
  15. Rempel, B.P.; Withers, S.G. Covalent inhibitors of glycosidases and their applications in biochemistry and biology. Glycobiology 2008, 18, 570–586. [Google Scholar] [CrossRef] [PubMed]
  16. Arda, A.; Jimenez-Barbero, J. The recognition of glycans by protein receptors. Insights from NMR spectroscopy. Chem. Commun. 2018, 54, 4761–4769. [Google Scholar] [CrossRef] [PubMed]
  17. Tengel, T.; Fex, T.; Emtenas, H.; Almqvist, F.; Sethson, I.; Kihlberg, J. Use of 19F NMR spectroscopy to screen chemical libraries for ligands that bind to proteins. Org. Biomol. Chem. 2004, 2, 725–731. [Google Scholar] [CrossRef]
  18. Diercks, T.; Ribeiro, J.P.; Canada, F.J.; Andre, S.; Jimenez-Barbero, J.; Gabius, H.J. Fluorinated carbohydrates as lectin ligands: Versatile sensors in 19F-detected saturation transfer difference NMR spectroscopy. Chem. Eur. J. 2009, 15, 5666–5668. [Google Scholar] [CrossRef]
  19. André, S.; Cañada, F.J.; Shiao, T.C.; Largartera, L.; Diercks, T.; Bergeron-Brlek, M.; Biari, K.; Papadopoulos, A.; Ribeiro, J.P.; Touaibia, M.; et al. Fluorinated carbohydrates as lectin ligands: Biorelevant sensors with capacity to monitor anomer affinity in 19F-NMR-based inhibitor screening. Eur. J. Org. Chem. 2012, 2012, 4354–4364. [Google Scholar] [CrossRef]
  20. Matei, E.; Andre, S.; Glinschert, A.; Infantino, A.S.; Oscarson, S.; Gabius, H.J.; Gronenborn, A.M. Fluorinated carbohydrates as lectin ligands: Dissecting glycan-cyanovirin interactions by using 19F NMR spectroscopy. Chem. Eur. J. 2013, 19, 5364–5374. [Google Scholar] [CrossRef]
  21. Ribeiro, J.P.; Diercks, T.; Jimenez-Barbero, J.; Andre, S.; Gabius, H.J.; Canada, F.J. Fluorinated carbohydrates as lectin ligands: 19F-based direct STD monitoring for detection of anomeric selectivity. Biomolecules 2015, 5, 3177–3192. [Google Scholar] [CrossRef] [PubMed]
  22. Diercks, T.; Infantino, A.S.; Unione, L.; Jimenez-Barbero, J.; Oscarson, S.; Gabius, H.J. Fluorinated carbohydrates as lectin ligands: Synthesis of OH/F-Substituted N-Glycan core trimannoside and epitope mapping by 2D STD-TOCSY 19F-NMR spectroscopy. Chem. Eur. J. 2018, 24, 15761–15765. [Google Scholar] [CrossRef] [PubMed]
  23. Valverde, P.; Quintana, J.I.; Santos, J.I.; Arda, A.; Jimenez-Barbero, J. Novel NMR avenues to explore the conformation and interactions of glycans. ACS Omega 2019, 4, 13618–13630. [Google Scholar] [CrossRef] [PubMed]
  24. Martinez, J.D.; Manzano, A.I.; Calvino, E.; Diego, A.; Rodriguez de Francisco, B.; Romano, C.; Oscarson, S.; Millet, O.; Gabius, H.J.; Jimenez-Barbero, J.; et al. Fluorinated carbohydrates as lectin ligands: Simultaneous screening of a monosaccharide library and chemical mapping by 19F NMR spectroscopy. J. Org. Chem. 2020, 85, 16072–16081. [Google Scholar] [CrossRef] [PubMed]
  25. Hoffmann-Roder, A.; Kaiser, A.; Wagner, S.; Gaidzik, N.; Kowalczyk, D.; Westerlind, U.; Gerlitzki, B.; Schmitt, E.; Kunz, H. Synthetic antitumor vaccines from tetanus toxoid conjugates of MUC1 glycopeptides with the Thomsen-Friedenreich antigen and a fluorine-substituted analogue. Angew. Chem. Int. Ed. 2010, 49, 8498–8503. [Google Scholar] [CrossRef] [PubMed]
  26. Hoffmann-Roder, A.; Johannes, M. Synthesis of a MUC1-glycopeptide-BSA conjugate vaccine bearing the 3′-deoxy-3′-fluoro-Thomsen-Friedenreich antigen. Chem. Commun. 2011, 47, 9903–9905. [Google Scholar] [CrossRef] [PubMed]
  27. Oberbillig, T.; Mersch, C.; Wagner, S.; Hoffmann-Roder, A. Antibody recognition of fluorinated MUC1 glycopeptide antigens. Chem. Commun. 2012, 48, 1487–1489. [Google Scholar] [CrossRef]
  28. Wei, M.M.; Wang, Y.S.; Ye, X.S. Carbohydrate-based vaccines for oncotherapy. Med. Res. Rev. 2018, 38, 1003–1026. [Google Scholar] [CrossRef]
  29. Anderluh, M.; Berti, F.; Bzducha-Wrobel, A.; Chiodo, F.; Colombo, C.; Compostella, F.; Durlik, K.; Ferhati, X.; Holmdahl, R.; Jovanovic, D.; et al. Recent advances on smart glycoconjugate vaccines in infections and cancer. FEBS J. 2022, 289, 4251–4303. [Google Scholar] [CrossRef]
  30. Mettu, R.; Chen, C.Y.; Wu, C.Y. Synthetic carbohydrate-based vaccines: Challenges and opportunities. J. Biomed. Sci. 2020, 27, 9. [Google Scholar] [CrossRef]
  31. Linclau, B.; Arda, A.; Reichardt, N.C.; Sollogoub, M.; Unione, L.; Vincent, S.P.; Jimenez-Barbero, J. Fluorinated carbohydrates as chemical probes for molecular recognition studies. Current status and perspectives. Chem. Soc. Rev. 2020, 49, 3863–3888. [Google Scholar] [CrossRef] [PubMed]
  32. Hevey, R. The role of fluorine in glycomimetic drug design. Chem. Eur. J. 2021, 27, 2240–2253. [Google Scholar] [CrossRef] [PubMed]
  33. Markovski, L.N.; Pashinnik, V.E. Applications of dialkylaminosulfur trifluorides for the syntheses of acid fluorides. Synthesis 1975, 12, 801–802. [Google Scholar] [CrossRef]
  34. Middleton, W.J. New fluorinating reagents. Dialkylaminosulfur fluorides. J. Org. Chem. 1975, 40, 574–578. [Google Scholar] [CrossRef]
  35. Doboszewski, B.; Hay, G.W.; Szarek, W.A. The rapid synthesis of deoxyfluoro sugars using tris(dimethylamino)sulfonium difluorotrimethylsilicate (TASF). Can. J. Chem. 1987, 65, 412–419. [Google Scholar] [CrossRef]
  36. Card, P.J.; Hitz, W.D. Synthesis of 1′-deoxy-1′-fluorosucrose via sucrose synthetase mediated coupling of 1-deoxy-1-fluorofructose with uridine diphosphate glucose. J. Am. Chem. Soc. 1984, 106, 5348–5350. [Google Scholar] [CrossRef]
  37. Masataka, Y. Methods of synthesis of glycosyl fluorides. Carbohydr. Res. 2000, 327, 5–14. [Google Scholar]
  38. Takahashi, Y.; Tsuneda, S.; Tsuchiya, T.; Koyama, Y.; Umezawa, S. Synthesis of 4’-deoxy-4’-fluorokanamycin A and B. Carbohydr. Res. 1992, 232, 89–105. [Google Scholar] [CrossRef]
  39. Card, P.J. Synthesis of Fluorinated Carbohydrates. J. Carbohydr. Chem. 2007, 4, 451–487. [Google Scholar] [CrossRef]
  40. Nyffeler, P.T.; Duron, S.G.; Burkart, M.D.; Vincent, S.P.; Wong, C.H. Selectfluor: Mechanistic insight and applications. Angew. Chem. Int. Ed. 2005, 44, 192–212. [Google Scholar] [CrossRef]
  41. Cen, Y.; Sauve, A.A. Diastereocontrolled electrophilic fluorinations of 2-deoxyribonolactone: Syntheses of all corresponding 2-deoxy-2-fluorolactones and 2’-deoxy-2’-fluoro-NAD+s. J. Org. Chem. 2009, 74, 5779–5789. [Google Scholar] [CrossRef] [PubMed]
  42. Liu, P.; Sharon, A.; Chu, C.K. Fluorinated nucleosides: Synthesis and biological implication. J. Fluorine Chem. 2008, 129, 743–766. [Google Scholar] [CrossRef] [PubMed]
  43. Herdewijn, P.; Aerschot, A.V.; Kerremans, L. Synthesis of nucleosides fluorinated in the sugar moiety. The application of diethylaminosulfur trifluoride to the synthesis of fluorinated nucleosides. Nucleosides Nucleotides 1989, 8, 65–96. [Google Scholar] [CrossRef]
  44. Uhrig, M.L.; Lantaño, B.; Postigo, A. Synthetic strategies for fluorination of carbohydrates. Org. Biomol. Chem. 2019, 17, 5173–5189. [Google Scholar] [CrossRef] [PubMed]
  45. Dax, K.; Albert, M.; Ortner, J.; Paul, B.J. Synthesis of deoxyfluoro sugars from carbohydrate precursors. Carbohydr. Res. 2000, 327, 47–86. [Google Scholar] [CrossRef] [PubMed]
  46. Miethchen, R. Modified natural substances—Fluorinated and fluoroalkylated monosaccharides and inositols. J. Fluorine Chem. 2004, 125, 895–901. [Google Scholar] [CrossRef]
  47. Lloyd, A.; Coe, P.; Walker, R.; Howarth, O. Some intramolecular rearrangements when pentofuranoses are treated with diethylaminosulphur trifluoride (DAST). J. Fluorine Chem. 1993, 60, 239–250. [Google Scholar] [CrossRef]
  48. Dmowski, W.; Kamiński, M. Dialkyl-α,α-difluorobenzylamines and dialkyl(trifluoromethyl)-amines-novel fluorinating reagents. J. Fluorine Chem. 1983, 23, 219–228. [Google Scholar] [CrossRef]
  49. Kobayashi, S.; Yoneda, A.; Fukuhara, T.; Hara, S. Selective synthesis of fluorinated carbohydrates using N,N-diethyl-α,α-difluoro-(m-methylbenzyl) amine. Tetrahedron Lett. 2004, 45, 1287–1289. [Google Scholar] [CrossRef]
  50. Huonnic, K.; Linclau, B. The synthesis and glycoside formation of polyfluorinated carbohydrates. Chem. Rev. 2022, 122, 15503–15602. [Google Scholar] [CrossRef]
  51. Sharma, R.; Kavai, I.; Fu, Y.; Bobek, M. Synthesis of gem-difluorosaccharides. Tetrahedron Lett. 1977, 18, 3433–3436. [Google Scholar] [CrossRef]
  52. Kolympadi, M.; Fontanella, M.; Venturi, C.; André, S.; Gabius, H.J.; Jiménez-Barbero, J.; Vogel, P. Synthesis and conformational analysis of (α-D-galactosyl) phenylmethane and α-, β-difluoromethane analogues: Interactions with the plant lectin viscumin. Chem. Eur. J. 2009, 15, 2861–2873. [Google Scholar] [CrossRef]
  53. Houlton, J.S.; Motherwell, W.B.; Ross, B.C.; Tozer, M.J.; Williams, D.J.; Slawin, A.M. A convenient strategy for replacement of the anomeric hydroxyl group by difluoromethyl functionality in carbohydrate derivatives. Tetrahedron 1993, 49, 8087–8106. [Google Scholar] [CrossRef]
  54. Leclerc, E.; Pannecoucke, X. Synthetic efforts towards the synthesis of fluorinated C-glycosidic analogues of α-galactosylceramides. C. R. Chim. 2012, 15, 57–67. [Google Scholar] [CrossRef]
  55. Cuenca, A.B.; D’Hooge, F.; Gouge, V.; Castelot-Deliencourt, G.; Oulyadi, H.; Leclerc, E.; Jubault, P.; Pannecoucke, X.; Quirion, J.C. Addition of ethyl bromodifluoroacetate to lactones: Reactivity and stereoselectivity. Synlett 2005, 17, 2627–2630. [Google Scholar] [CrossRef]
  56. Poulain, F.; Leclerc, E.; Quirion, J.C. Approaches to the synthesis of CF2-analogues of 2-deoxy-2-aminoglycosides. Tetrahedron Lett. 2009, 50, 1803–1805. [Google Scholar] [CrossRef]
  57. Munier, P.; Giudicelli, M.B.; Picq, D.; Anker, D. Synthese des 5-Desoxy-5,5,5-trifluoro-D-et-LPentofuranoses. J. Carbohydr. Chem. 1996, 15, 739–762. [Google Scholar] [CrossRef]
  58. Lavaire, S.; Plantier-Royon, R.; Portella, C. Synthesis of 3-deoxy-3-C-trifluoromethyl-D-ribose from D-xylose or D-glucose. J. Carbohydr. Chem. 1996, 15, 361–370. [Google Scholar] [CrossRef]
  59. Lal, G.S.; Pez, G.P.; Syvret, R.G. Electrophilic NF fluorinating agents. Chem. Rev. 1996, 96, 1737–1756. [Google Scholar] [CrossRef]
  60. Burkart, M.D.; Zhang, Z.Y.; Hung, S.C.; Wong, C.H. A new method for the synthesis of fluorocarbohydrates and glycosides using selectfluor. J. Am. Chem. Soc. 1997, 119, 11743–11746. [Google Scholar] [CrossRef]
  61. Albert, M.; Dax, K.; Ortner, J. A novel direct route to 2-deoxy-2-fluoro-aldoses and their corresponding derivatives. Tetrahedron 1998, 54, 4839–4848. [Google Scholar] [CrossRef]
  62. Adamson, J.; Foster, A.B.; Hall, L.D.; Johnson, R.N.; Hesse, R.H. Fluorinated carbohydrates: Part III. 2-deoxy-2-fluoro-D-glucose and 2-deoxy-2-fluoro-D-mannose. Carbohydr. Res. 1970, 15, 351–359. [Google Scholar] [CrossRef]
  63. Ashique, R.; Chirakal, R.V.; Hughes, D.W.; Schrobilgen, G.J. Two-step regio-and stereoselective syntheses of 19F-and 18F-2-deoxy-2-(R)-fluoro-β-D-allose. Carbohydr. Res. 2006, 341, 457–466. [Google Scholar] [CrossRef] [PubMed]
  64. Zhou, X.; Ding, H.; Chen, P.; Liu, L.; Sun, Q.; Wang, X.; Wang, P.; Lv, Z.; Li, M. Radical dehydroxymethylative fluorination of carbohydrates and divergent transformations of the resulting reverse glycosyl fluorides. Angew. Chem. Int. Ed. 2020, 59, 4138–4144. [Google Scholar] [CrossRef] [PubMed]
  65. Li, T.; Wang, J.; Zhu, X.; Zhou, X.; Sun, S.; Wang, P.; Cao, H.; Yu, G.; Li, M. Synthesis of Rare 6-Deoxy-D-/L-heptopyranosyl fluorides: Assembly of a hexasaccharide corresponding to campylobacter jejuni strain CG8486 capsular polysaccharide. J. Am. Chem. Soc. 2021, 143, 11171–11179. [Google Scholar] [CrossRef] [PubMed]
  66. Chen, P.; Wang, P.; Long, Q.; Ding, H.; Cheng, G.; Li, T.; Li, M. Synthesis of reverse glycosyl fluorides and rare glycosyl fluorides enabled by radical decarboxylative fluorination of uronic acids. Org. Lett. 2020, 22, 9325–9330. [Google Scholar] [CrossRef] [PubMed]
  67. Ding, H.; Yan, N.; Wang, P.; Song, N.; Sun, Q.; Li, T.; Li, M. Synthesis of reverse glycosyl fluorides via organophotocatalytic decarboxylative fluorination of uronic acids. Org. Chem. Front. 2022, 9, 2808–2814. [Google Scholar] [CrossRef]
  68. Moreno, B.; Quehen, C.; Rose-Helene, M.; Leclerc, E.; Quirion, J.C. Addition of difluoromethyl radicals to glycals: A new route to α-CF2-D-glycosides. Org. Lett. 2007, 9, 2477–2480. [Google Scholar] [CrossRef]
  69. Wang, B.; Xiong, D.C.; Ye, X.S. Direct C-H trifluoromethylation of glycals by photoredox catalysis. Org. Lett. 2015, 17, 5698–5701. [Google Scholar] [CrossRef]
  70. Frédéric, C.J.M.; Cornil, J.; Vandamme, M.; Dumitrescu, L.; Tikad, A.; Robiette, R.; Vincent, S.P. Highly (Z)-diastereoselective synthesis of trifluoromethylated exo-glycals via photoredox and copper catalysis. Org. Lett. 2018, 20, 6769–6773. [Google Scholar] [CrossRef]
  71. Liu, M.; Luo, Z.X.; Li, T.; Xiong, D.C.; Ye, X.S. Electrochemical trifluoromethylation of glycals. J. Org. Chem. 2021, 86, 16187–16194. [Google Scholar] [CrossRef] [PubMed]
  72. Flores, E.W.M.; Uhrig, M.L.; Postigo, A. Photocatalyzed reductive fluoroalkylation of 2-acetoxyglycals towards the stereoselective synthesis of α-1-fluoroalkyl-C-glycosyl derivatives. Org. Biomol. Chem. 2020, 18, 8724–8734. [Google Scholar] [CrossRef] [PubMed]
  73. Belhomme, M.C.; Dru, D.; Xiong, H.Y.; Cahard, D.; Besset, T.; Poisson, T.; Pannecoucke, X. Coppermediated direct functionalization of unsaturated C–C bonds with ethyl bromo (difluoro) acetate: A straightforward access to highly valuable difluoromethylated alkenes. Synthesis 2014, 46, 1859–1870. [Google Scholar]
  74. Belhomme, M.C.; Poisson, T.; Pannecoucke, X. Copper catalyzed β-difluoroacetylation of dihydropyrans and glycals by means of direct C-H functionalization. Org. Lett. 2013, 15, 3428–3431. [Google Scholar] [CrossRef] [PubMed]
  75. Mestre, J.; Lishchynskyi, A.; Castillon, S.; Boutureira, O. Trifluoromethylation of electron-rich alkenyl iodides with fluoroform-derived “ligandless” CuCF3. J. Org. Chem. 2018, 83, 8150–8160. [Google Scholar] [CrossRef] [PubMed]
  76. Mestre, J.; Castillón, S.; Boutureira, O. “Ligandless” pentafluoroethylation of unactivated (hetero) aryl and alkenyl halides enabled by the controlled self-condensation of TMSCF3-derived CuCF3. J. Org. Chem. 2019, 84, 15087–15097. [Google Scholar] [CrossRef] [PubMed]
  77. Tony, K.A.; Denton, R.W.; Dilhas, A.; Jiménez-Barbero, J.; Mootoo, D.R. Synthesis of β-C-galactopyranosides with fluorine on the pseudoanomeric substituent. Org. Lett. 2007, 9, 1441–1444. [Google Scholar] [CrossRef] [PubMed]
  78. Stéphane, M.; François, D.; Sathunuru, R.; Christian, F.; Xavier, P.; Jean-Charles, Q. Synthesis of new α- and β-gem-difluoromethylene C-glycosides in the galactose and glucose series. Tetrahedron Lett. 2001, 42, 5879–5882. [Google Scholar]
  79. Li, Y.; Drew, M.G.; Welchman, E.V.; Shirvastava, R.K.; Jiang, S.; Valentine, R.; Singh, G. Syntheses of ethyl 3-deoxy-3, 3-difluoro-D-arabino-heptulosonate and analogues. Tetrahedron 2004, 60, 6523–6531. [Google Scholar] [CrossRef]
  80. Kawada, K.; Kitagawa, O.; Taguchi, T.; Hanzawa, Y.; Kobayashi, Y.; Iitaka, Y. Studies on organic fluorine compounds. XLVII. synthesis of trifluoromethylated sugars through the Aldol Reaction of 2-trimethylsilyloxy-4-trifluoromethylfuran. Chem. Pharm. Bull. 1986, 17, 4216–4222. [Google Scholar] [CrossRef]
  81. Eilitz, U.; Bottcher, C.; Sieler, J.; Gockel, S.; Haas, A.; Burger, K. Synthesis of 2-C-trifluoromethyl substituted D-ribose. Tetrahedron 2001, 57, 3921–3925. [Google Scholar] [CrossRef]
  82. N’Go, I.; Golten, S.; Arda, A.; Canada, J.; Jimenez-Barbero, J.; Linclau, B.; Vincent, S.P. Tetrafluorination of sugars as strategy for enhancing protein-carbohydrate affinity: Application to UDP-Galp mutase inhibition. Chem. Eur. J. 2014, 20, 106–112. [Google Scholar] [CrossRef] [PubMed]
  83. Council, C.E.; Kilpin, K.J.; Gusthart, J.S.; Allman, S.A.; Linclau, B.; Lee, S.S. Enzymatic glycosylation involving fluorinated carbohydrates. Org. Biomol. Chem. 2020, 18, 3423–3451. [Google Scholar] [CrossRef] [PubMed]
  84. Linclau, B.; Wang, Z.; Compain, G.; Paumelle, V.; Fontenelle, C.Q.; Wells, N.; Weymouth-Wilson, A. Investigating the influence of (deoxy)fluorination on the lipophilicity of non-UV-active fluorinated alkanols and carbohydrates by a new log P determination method. Angew. Chem. Int. Ed. 2016, 55, 674–678. [Google Scholar] [CrossRef] [PubMed]
  85. Denavit, V.; Lainé, D.; St-Gelais, J.; Johnson, P.A.; Giguère, D. A Chiron approach towards the stereoselective synthesis of polyfluorinated carbohydrates. Nat. Commun. 2018, 9, 4721. [Google Scholar] [CrossRef] [PubMed]
  86. St-Gelais, J.; Cote, E.; Laine, D.; Johnson, P.A.; Giguere, D. Addressing the structural complexity of fluorinated glucose analogues: Insight into lipophilicities and solvation effects. Chem. Eur. J. 2020, 26, 13499–13506. [Google Scholar] [CrossRef] [PubMed]
  87. Tarcsay, A.; Keseru, G.M. Contributions of molecular properties to drug promiscuity. J. Med. Chem. 2013, 56, 1789–1795. [Google Scholar] [CrossRef]
  88. Gleeson, M.P.; Hersey, A.; Montanari, D.; Overington, J. Probing the links between in vitro potency, ADMET and physicochemical parameters. Nat. Rev. Drug Discov. 2011, 10, 197–208. [Google Scholar] [CrossRef]
  89. Huchet, Q.A.; Kuhn, B.; Wagner, B.; Kratochwil, N.A.; Fischer, H.; Kansy, M.; Zimmerli, D.; Carreira, E.M.; Muller, K. Fluorination Patterning: A study of structural motifs that impact physicochemical properties of relevance to drug discovery. J. Med. Chem. 2015, 58, 9041–9060. [Google Scholar] [CrossRef]
  90. Scheidt, F.; Selter, P.; Santschi, N.; Holland, M.C.; Dudenko, D.V.; Daniliuc, C.; Muck-Lichtenfeld, C.; Hansen, M.R.; Gilmour, R. Emulating Natural Product Conformation by Cooperative, Non-Covalent Fluorine Interactions. Chem. Eur. J. 2017, 23, 6142–6149. [Google Scholar] [CrossRef]
  91. Bresciani, S.; Lebl, T.; Slawin, A.M.; O’Hagan, D. Fluorosugars: Synthesis of the 2,3,4-trideoxy-2,3,4-trifluoro hexose analogues of D-glucose and D-altrose and assessment of their erythrocyte transmembrane transport. Chem. Commun. 2010, 46, 5434–5436. [Google Scholar] [CrossRef] [PubMed]
  92. Quiquempoix, L.; Wang, Z.; Graton, J.; Latchem, P.G.; Light, M.; Le Questel, J.Y.; Linclau, B. Synthesis of 2,3,4-Trideoxy-2,3,4-trifluoroglucose. J. Org. Chem. 2019, 84, 5899–5906. [Google Scholar] [CrossRef] [PubMed]
  93. Linclau, B.; Golten, S.; Light, M.; Sebban, M.; Oulyadi, H. The conformation of tetrafluorinated methyl galactoside anomers: Crystallographic and NMR studies. Carbohydr. Res. 2011, 346, 1129–1139. [Google Scholar] [CrossRef] [PubMed]
  94. Golten, S.; Fontenelle, C.Q.; Timofte, R.S.; Bailac, L.; Light, M.; Sebban, M.; Oulyadi, H.; Linclau, B. Enantioselective synthesis of dideoxy-tetrafluorinated hexoses. J. Org. Chem. 2016, 81, 4434–4453. [Google Scholar] [CrossRef] [PubMed]
  95. Muller, K.; Faeh, C.; Diederich, F. Fluorine in pharmaceuticals: Looking beyond intuition. Science 2007, 317, 1881–1886. [Google Scholar] [CrossRef] [PubMed]
  96. Jagodzinska, M.; Huguenot, F.; Candiani, G.; Zanda, M. Assessing the bioisosterism of the trifluoromethyl group with a protease probe. ChemMedChem. 2009, 4, 49–51. [Google Scholar] [CrossRef] [PubMed]
  97. Hunter, L. The C-F bond as a conformational tool in organic and biological chemistry. Beilstein J. Org. Chem. 2010, 6, 38. [Google Scholar] [CrossRef] [PubMed]
  98. Hirai, G.; Watanabe, T.; Yamaguchi, K.; Miyagi, T.; Sodeoka, M. Stereocontrolled and convergent entry to CF2-sialosides: Synthesis of CF2-linked ganglioside GM4. J. Am. Chem. Soc. 2007, 129, 15420–15421. [Google Scholar] [CrossRef]
  99. Leclerc, E.; Pannecoucke, X.; Etheve-Quelquejeu, M.; Sollogoub, M. Fluoro-C-glycosides and fluoro-carbasugars, hydrolytically stable and synthetically challenging glycomimetics. Chem. Soc. Rev. 2013, 42, 4270–4283. [Google Scholar] [CrossRef]
  100. Xu, B.; Unione, L.; Sardinha, J.; Wu, S.; Etheve-Quelquejeu, M.; Pilar Rauter, A.; Bleriot, Y.; Zhang, Y.; Martin-Santamaria, S.; Diaz, D.; et al. gem-Difluorocarbadisaccharides: Restoring the exo-anomeric effect. Angew. Chem. Int. Ed. 2014, 53, 9597–9602. [Google Scholar] [CrossRef]
  101. Turnbull, J.; Powell, A.; Guimond, S. Heparan sulfate: Decoding a dynamic multifunctional cell regulator. Trends Cell Biol. 2001, 11, 75–82. [Google Scholar] [CrossRef] [PubMed]
  102. Li, W.; Johnson, D.J.; Esmon, C.T.; Huntington, J.A. Structure of the antithrombin-thrombin-heparin ternary complex reveals the antithrombotic mechanism of heparin. Nat. Struct. Mol. Biol. 2004, 11, 857–862. [Google Scholar] [CrossRef] [PubMed]
  103. Canales, A.; Lozano, R.; Lopez-Mendez, B.; Angulo, J.; Ojeda, R.; Nieto, P.M.; Martin-Lomas, M.; Gimenez-Gallego, G.; Jimenez-Barbero, J. Solution NMR structure of a human FGF-1 monomer, activated by a hexasaccharide heparin-analogue. FEBS J. 2006, 273, 4716–4727. [Google Scholar] [CrossRef] [PubMed]
  104. Sawen, E.; Roslund, M.U.; Cumpstey, I.; Widmalm, G. Synthesis and conformational analysis of carbasugar bioisosteres of alpha-L-iduronic acid and its methyl glycoside. Carbohydr. Res. 2010, 345, 984–993. [Google Scholar] [CrossRef] [PubMed]
  105. Sharma, M.; Bernacki, R.J.; Paul, B.; Korytnyk, W. Fluorinated carbohydrates as potential plasma membrane modifiers. Synthesis of 4- and 6-fluoro derivatives of 2-acetamido-2-deoxy-D-hexopyranoses. Carbohydr. Res. 1990, 198, 205–221. [Google Scholar] [CrossRef] [PubMed]
  106. Lebedel, L.; Arda, A.; Martin, A.; Desire, J.; Mingot, A.; Aufiero, M.; Aiguabella Font, N.; Gilmour, R.; Jimenez-Barbero, J.; Bleriot, Y.; et al. Structural and computational analysis of 2-halogeno-glycosyl cations in the presence of a superacid: An expansive platform. Angew. Chem. Int. Ed. 2019, 58, 13758–13762. [Google Scholar] [CrossRef] [PubMed]
  107. Withers, S.G.; MacLennan, D.J.; Street, I.P. The synthesis and hydrolysis of a series of deoxyfluoro-D-glucopyranosyl phosphates. Carbohydr. Res. 1986, 154, 127–144. [Google Scholar] [CrossRef]
  108. Withers, S.G.; Percival, M.D. and Street, I.P. The synthesis and hydrolysis of a series of deoxy- and deoxyfluoro-α-D-“glucopyranosyl” phosphates. Carbohydr. Res. 1989, 187, 43. [Google Scholar] [CrossRef]
  109. Namchuk, M.N.; McCarter, J.D.; Becalski, A.; Andrews, T.; Withers, S.G. The role of sugar substituents in glycoside hydrolysis. J. Am. Chem. Soc. 2000, 122, 1270–1277. [Google Scholar] [CrossRef]
  110. Burkart, M.D.; Vincent, S.P.; Duffels, A.; Murray, B.W.; Ley, S.V.; Wong, C.H. Chemo-enzymatic synthesis of fluorinated sugar nucleotide: Useful mechanistic probes for glycosyltransferases. Bioorg. Med. Chem. 2000, 8, 1937–1946. [Google Scholar] [CrossRef]
  111. Nishimura, S.; Hato, M.; Hyugaji, S.; Feng, F.; Amano, M. Glycomics for drug discovery: Metabolic perturbation in androgen-independent prostate cancer cells induced by unnatural hexosamine mimics. Angew. Chem. Int. Ed. 2012, 51, 3386–3390. [Google Scholar] [CrossRef] [PubMed]
  112. Marathe, D.D.; Buffone, A.J.; Chandrasekaran, E.V.; Xue, J.; Locke, R.D.; Nasirikenari, M.; Lau, J.T.; Matta, K.L.; Neelamegham, S. Fluorinated per-acetylated GalNAc metabolically alters glycan structures on leukocyte PSGL-1 and reduces cell binding to selectins. Blood 2010, 115, 1303–1312. [Google Scholar] [CrossRef] [PubMed]
  113. van Wijk, X.M.; Lawrence, R.; Thijssen, V.L.; van den Broek, S.A.; Troost, R.; van Scherpenzeel, M.; Naidu, N.; Oosterhof, A.; Griffioen, A.W.; Lefeber, D.J.; et al. A common sugar-nucleotide-mediated mechanism of inhibition of (glycosamino)glycan biosynthesis, as evidenced by 6F-GalNAc (Ac3). FASEB J. 2015, 29, 2993–3002. [Google Scholar] [CrossRef] [PubMed]
  114. Barthel, S.R.; Antonopoulos, A.; Cedeno-Laurent, F.; Schaffer, L.; Hernandez, G.; Patil, S.A.; North, S.J.; Dell, A.; Matta, K.L.; Neelamegham, S.; et al. Peracetylated 4-fluoro-glucosamine reduces the content and repertoire of N- and O-glycans without direct incorporation. J. Biol. Chem. 2011, 286, 21717–21731. [Google Scholar] [CrossRef] [PubMed]
  115. Withers, S.G.; Street, I.P.; Bird, P.; Dolphin, D.H. 2-Deoxy-2-fluoroglucosides: A novel class of mechanism-based glucosidase inhibitors. J. Am. Chem. Soc. 1987, 109, 7530–7531. [Google Scholar] [CrossRef]
  116. Braun, C.; Brayer, G.D.; Withers, S.G. Mechanism-based inhibition of yeast alpha-glucosidase and human pancreatic alpha-amylase by a new class of inhibitors. 2-Deoxy-2,2-difluoro-alpha-glycosides. J. Biol. Chem. 1995, 270, 26778–26781. [Google Scholar] [CrossRef] [PubMed]
  117. Ly, H.D.; Howard, S.; Shum, K.; He, S.; Zhu, A.; Withers, S.G. The synthesis, testing and use of 5-fluoro-alpha-D-galactosyl fluoride to trap an intermediate on green coffee bean alpha-galactosidase and identify the catalytic nucleophile. Carbohydr. Res. 2000, 329, 539–547. [Google Scholar] [CrossRef]
  118. Amaya, M.F.; Watts, A.G.; Damager, I.; Wehenkel, A.; Nguyen, T.; Buschiazzo, A.; Paris, G.; Frasch, A.C.; Withers, S.G.; Alzari, P.M. Structural Insights into the Catalytic Mechanism of Trypanosoma cruzi trans-Sialidase. Structure 2004, 12, 775–784. [Google Scholar] [CrossRef]
  119. Vocadlo, D.J.; Mayer, C.; He, S.; Withers, S.G. Mechanism of action and identification of Asp242 as the catalytic nucleophile of Vibrio furnisii N-acetyl-beta-D-glucosaminidase using 2-acetamido-2-deoxy-5-fluoro-alpha-L-idopyranosyl fluoride. BioChem. Eur. J. 2000, 39, 117–126. [Google Scholar] [CrossRef]
  120. Syson, K.; Stevenson, C.E.M.; Rashid, A.M.; Saalbach, G.; Tang, M.; Tuukkanen, A.; Svergun, D.I.; Withers, S.G.; Lawson, D.M.; Bornemann, S. Structural Insight into How Streptomyces coelicolor Maltosyl Transferase GlgE Binds α-Maltose 1-Phosphate and Forms a Maltosyl-enzyme Intermediate. BioChem. Eur. J. 2014, 53, 2494–2504. [Google Scholar] [CrossRef]
  121. Thanna, S.; Lindenberger, J.J.; Gaitonde, V.V.; Ronning, D.R.; Sucheck, S.J. Synthesis of 2-deoxy-2,2-difluoro-α-maltosyl fluoride and its X-ray structure in complex with Streptomyces coelicolor GlgEI-V279S. Org. Biomol. Chem. 2015, 13, 7542–7550. [Google Scholar] [CrossRef] [PubMed]
  122. Sadurní, A.; Kehr, G.; Ahlqvist, M.; Wernevik, J.; Sjögren, H.P.; Kankkonen, C.; Knerr, L.; Gilmour, R. Fluorine-Directed Glycosylation Enables the Stereocontrolled Synthesis of Selective SGLT2 Inhibitors for Type II Diabetes. Eur. J. Org. Chem. 2018, 24, 2832–2836. [Google Scholar] [CrossRef] [PubMed]
  123. Bull, C.; Boltje, T.J.; van Dinther, E.A.; Peters, T.; de Graaf, A.M.; Leusen, J.H.; Kreutz, M.; Figdor, C.G.; den Brok, M.H.; Adema, G.J. Targeted delivery of a sialic acid-blocking glycomimetic to cancer cells inhibits metastatic spread. ACS Nano 2015, 9, 733–745. [Google Scholar] [CrossRef] [PubMed]
  124. Bull, C.; Boltje, T.J.; Wassink, M.; de Graaf, A.M.; van Delft, F.L.; den Brok, M.H.; Adema, G.J. Targeting aberrant sialylation in cancer cells using a fluorinated sialic acid analog impairs adhesion, migration, and in vivo tumor growth. Mol. Cancer Ther. 2013, 12, 1935–1946. [Google Scholar] [CrossRef] [PubMed]
  125. Philippe, C.; Olivier, R.M. Carbohydrate mimetics-based glycosyltransferase inhibitors. Bioorg. Med. Chem. 2001, 9, 3077–3092. [Google Scholar]
  126. Conforti, I.; Marra, A. Iminosugars as glycosyltransferase inhibitors. Org. Biomol. Chem. 2021, 19, 5439–5475. [Google Scholar] [CrossRef] [PubMed]
  127. Tu, Z.; Lin, Y.N.; Lin, C.H. Development of fucosyltransferase and fucosidase inhibitors. Chem. Soc. Rev. 2013, 42, 4459–4475. [Google Scholar] [CrossRef]
  128. Wang, L.; Liu, Y.; Wu, L.; Sun, X.L. Sialyltransferase inhibition and recent advances. Biochim. Biophys. Acta 2016, 1864, 143–153. [Google Scholar] [CrossRef]
  129. Breton, C.; Snajdrova, L.; Jeanneau, C.; Koca, J.; Imberty, A. Structures and mechanisms of glycosyltransferases. Glycobiology 2006, 16, 29–37. [Google Scholar] [CrossRef]
  130. Rillahan, C.D.; Antonopoulos, A.; Lefort, C.T.; Sonon, R.; Azadi, P.; Ley, K.; Dell, A.; Haslam, S.M.; Paulson, J.C. Global metabolic inhibitors of sialyl- and fucosyltransferases remodel the glycome. Nat. Chem. Biol. 2012, 8, 661–668. [Google Scholar] [CrossRef]
  131. Bull, C.; Boltje, T.J.; Balneger, N.; Weischer, S.M.; Wassink, M.; van Gemst, J.J.; Bloemendal, V.R.; Boon, L.; van der Vlag, J.; Heise, T.; et al. Sialic acid blockade suppresses tumor growth by enhancing T-cell-mediated tumor immunity. Cancer Res. 2018, 78, 3574–3588. [Google Scholar] [CrossRef] [PubMed]
  132. Heise, T.; Pijnenborg, J.F.A.; Bull, C.; van Hilten, N.; Kers-Rebel, E.D.; Balneger, N.; Elferink, H.; Adema, G.J.; Boltje, T.J. Potent metabolic sialylation inhibitors based on C-5-modified fluorinated sialic acids. J. Med. Chem. 2019, 62, 1014–1021. [Google Scholar] [CrossRef] [PubMed]
  133. Murray, B.W.; Wittmann, V.; Burkart, M.D.; Hung, S.C.; Wong, C.H. Mechanism of human alpha-1,3-fucosyltransferase V: Glycosidic cleavage occurs prior to nucleophilic attack. BioChem. Eur. J. 1997, 36, 823–831. [Google Scholar] [CrossRef] [PubMed]
  134. Achmatowicz, M.M.; Allen, J.G.; Bio, M.M.; Bartberger, M.D.; Borths, C.J.; Colyer, J.T.; Crockett, R.D.; Hwang, T.L.; Koek, J.N.; Osgood, S.A.; et al. Telescoped process to manufacture 6,6,6-trifluorofucose via diastereoselective transfer hydrogenation: Scalable access to an inhibitor of fucosylation utilized in monoclonal antibody production. J. Org. Chem. 2016, 81, 4736–4743. [Google Scholar] [CrossRef]
  135. McKenzie, N.C.; Scott, N.E.; John, A.; White, J.M.; Goddard-Borger, E.D. Synthesis and use of 6,6,6-trifluoro-L-fucose to block core-fucosylation in hybridoma cell lines. Carbohydr. Res. 2018, 465, 4–9. [Google Scholar] [CrossRef] [PubMed]
  136. Williams, D.A.; Pradhan, K.; Paul, A.; Olin, I.R.; Tuck, O.T.; Moulton, K.D.; Kulkarni, S.S.; Dube, D.H. Metabolic inhibitors of bacterial glycan biosynthesis. Chem. Sci. 2020, 11, 1761–1774. [Google Scholar] [CrossRef] [PubMed]
  137. Liu, Z.; Ma, S. Recent advances in synthetic alpha-glucosidase inhibitors. ChemMedChem 2017, 12, 819–829. [Google Scholar] [CrossRef]
  138. Williams, S.J.; Withers, S.G. Glycosyl fluorides in enzymatic reactions. Carbohydr. Res. 2000, 327, 27–46. [Google Scholar] [CrossRef]
  139. McCarter, J.D.; Adam, M.J.; Hartman, N.G.; Withers, S.G. In vivo inhibition of beta-glucosidase and beta-mannosidase activity in rats by 2-deoxy-2-fluoro-beta-glycosyl fluorides and recovery of activity in vivo and in vitro. Biochem. J. 1994, 301, 343–348. [Google Scholar] [CrossRef]
  140. Watts, A.G.; Damager, I.; Amaya, M.L.; Buschiazzo, A.; Alzari, P.; Frasch, A.C.; Withers, S.G. Trypanosoma cruzi trans-sialidase operates through a covalent sialyl-enzyme intermediate: Tyrosine is the catalytic nucleophile. J. Am. Chem. Soc. 2003, 125, 7532–7533. [Google Scholar] [CrossRef]
  141. Buchini, S.; Buschiazzo, A.; Withers, S.G. A new generation of specific Trypanosoma cruzi trans-sialidase inhibitors. Angew. Chem. Int. Ed. 2008, 47, 2700–2703. [Google Scholar] [CrossRef] [PubMed]
  142. Tsai, C.S.; Yen, H.Y.; Lin, M.I.; Tsai, T.I.; Wang, S.Y.; Huang, W.I.; Hsu, T.L.; Cheng, Y.S.; Fang, J.M.; Wong, C.H. Cell-permeable probe for identification and imaging of sialidases. Proc. Natl. Acad. Sci. USA 2013, 110, 2466–2471. [Google Scholar] [CrossRef] [PubMed]
  143. Weck, S.; Robinson, K.; Smith, M.R.; Withers, S.G. Understanding viral neuraminidase inhibition by substituted difluorosialic acids. Chem. Commun. 2015, 51, 2933–2935. [Google Scholar] [CrossRef] [PubMed]
  144. Kim, J.H.; Resende, R.; Wennekes, T.; Chen, H.M.; Bance, N.; Buchini, S.; Watts, A.G.; Pilling, P.; Streltsov, V.A.; Petric, M.; et al. Mechanism-based covalent neuraminidase inhibitors with broad-spectrum influenza antiviral activity. Science 2013, 340, 71–75. [Google Scholar] [CrossRef] [PubMed]
  145. Yumoto, H.; Yumoto, H. Nojirimycin, a new antibiotic. II. Isolation, characterization and biological activity. Antibiotics 1967, 20, 66–71. [Google Scholar]
  146. Inouye, S.; Tsuruoka, T.; Ito, T.; Niida, T. Structure and synthesis of nojirimycin. Tetrahedron 1968, 24, 2125–2144. [Google Scholar] [CrossRef] [PubMed]
  147. Monika, B.M.; Anna, S.; Bartosz, M. Design, properties and applications of fluorinated and fluoroalkylated N-containing monosaccharides and their analogues. J. Fluorine Chem. 2019, 227, 109364. [Google Scholar]
  148. Uhrig, M.L.; Mora Flores, E.W.; Postigo, A. Approaches to the synthesis of perfluoroalkyl-modified carbohydrates and derivatives: Thiosugars, iminosugars, and tetrahydro(thio)pyrans. Chem. Eur. J. 2021, 27, 7813–7825. [Google Scholar] [CrossRef]
  149. Stütz, A.E. (Ed.) Glycosidase Inhibitors: Nojirimycin and Beyond; Wiley-VCH: Weinheim, Germany, 1999. [Google Scholar]
  150. Compain, P.; Martin, O.R. (Eds.) Iminosugars, from Synthesis to Therapeutic Applications; Wiley-VCH: Weinheim, Germany, 2007. [Google Scholar]
  151. Wagner, S.; Mersch, C.; Hoffmann-Roder, A. Fluorinated glycosyl amino acids for mucin-like glycopeptide antigen analogues. Chem. Eur. J. 2010, 16, 7319–7930. [Google Scholar] [CrossRef]
  152. Johannes, M.; Oberbillig, T.; Hoffmann-Roder, A. Synthesis of fluorinated Thomsen-Friedenreich antigens: Direct deoxyfluorination of alphaGalNAc-threonine tert-butyl esters. Org. Biomol. Chem. 2011, 9, 5541–5546. [Google Scholar] [CrossRef]
  153. Johannes, M.; Reindl, M.; Gerlitzki, B.; Schmitt, E.; Hoffmann-Roder, A. Synthesis and biological evaluation of a novel MUC1 glycopeptide conjugate vaccine candidate comprising a 4’-deoxy-4’-fluoro-Thomsen-Friedenreich epitope. Beilstein J. Org. Chem. 2015, 11, 155–161. [Google Scholar] [CrossRef] [PubMed]
  154. Yang, F.; Zheng, X.J.; Huo, C.X.; Wang, Y.; Zhang, Y.; Ye, X.S. Enhancement of the immunogenicity of synthetic carbohydrate vaccines by chemical modifications of STn antigen. ACS Chem. Biol. 2011, 6, 252–259. [Google Scholar] [CrossRef] [PubMed]
  155. Song, C.; Zheng, X.J.; Liu, C.C.; Zhou, Y.; Ye, X.S. A cancer vaccine based on fluorine-modified sialyl-Tn induces robust immune responses in a murine model. Oncotarget 2017, 8, 47330–47343. [Google Scholar] [CrossRef] [PubMed]
  156. Li, B.H.; Yao, W.L.; Yang, H.; Wu, C.Y.; Xiong, D.C.; Yin, Y.X.; Ye, X.S. Total synthesis of tumor-associated KH-1 antigen core nonasaccharide via photo-induced glycosylation. Org. Chem. Front. 2020, 7, 1255–1259. [Google Scholar] [CrossRef]
  157. Xiao, A.; Zheng, X.J.; Song, C.; Gui, Y.; Huo, C.X.; Ye, X.S. Correction: Synthesis and immunological evaluation of MUC1 glycopeptide conjugates bearing N-acetyl modified STn derivatives as anticancer vaccines. Org. Biomol. Chem. 2017, 15, 2120–2121. [Google Scholar] [CrossRef] [PubMed]
  158. Song, C.; Sun, S.; Huo, C.X.; Li, Q.; Zheng, X.J.; Tai, G.; Zhou, Y.; Ye, X.S. Synthesis and immunological evaluation of N-acyl modified Tn analogues as anticancer vaccine candidates. Bioorg. Med. Chem. 2016, 24, 915–920. [Google Scholar] [CrossRef] [PubMed]
  159. Sun, S.; Zheng, X.J.; Huo, C.X.; Song, C.; Li, Q.; Ye, X.S. Synthesis and Evaluation of Glycoconjugates Comprising N-Acyl-Modified Thomsen-Friedenreich Antigens as Anticancer Vaccines. ChemMedChem 2016, 11, 1090–1096. [Google Scholar] [CrossRef]
  160. Zheng, X.J.; Yang, F.; Zheng, M.; Huo, C.X.; Zhang, Y.; Ye, X.S. Improvement of the immune efficacy of carbohydrate vaccines by chemical modification on the GM3 antigen. Org. Biomol. Chem. 2015, 13, 6399–6406. [Google Scholar] [CrossRef]
  161. Liu, Y.; Li, B.; Zheng, X.; Xiong, D.; Ye, X. Cancer vaccines based on fluorine-modified KH-1 elicit robust immune response. Molecules 2023, 28, 1934. [Google Scholar] [CrossRef]
  162. Lee, H.Y.; Chen, C.Y.; Tsai, T.I.; Li, S.T.; Lin, K.H.; Cheng, Y.Y.; Ren, C.T.; Cheng, T.J.; Wu, C.Y.; Wong, C.H. Immunogenicity study of Globo H analogues with modification at the reducing or nonreducing end of the tumor antigen. J. Am. Chem. Soc. 2014, 136, 16844–16853. [Google Scholar] [CrossRef]
  163. Krauss, I.J.; Joyce, J.G.; Finnefrock, A.C.; Song, H.C.; Dudkin, V.Y.; Geng, X.; Warren, J.D.; Chastain, M.; Shiver, J.W.; Danishefsky, S.J. Fully synthetic carbohydrate HIV antigens designed on the logic of the 2G12 antibody. J. Am. Chem. Soc. 2007, 129, 11042–11044. [Google Scholar] [CrossRef]
  164. Astronomo, R.D.; Kaltgrad, E.; Udit, A.K.; Wang, S.K.; Doores, K.J.; Huang, C.Y.; Pantophlet, R.; Paulson, J.C.; Wong, C.H.; Finn, M.G.; et al. Defining criteria for oligomannose immunogens for HIV using icosahedral virus capsid scaffolds. Chem. Biol. 2010, 17, 357–370. [Google Scholar] [CrossRef]
  165. Joyce, J.G.; Krauss, I.J.; Song, H.C.; Opalka, D.W.; Grimm, K.M.; Nahas, D.D.; Esser, M.T.; Hrin, R.; Feng, M.; Dudkin, V.Y.; et al. An oligosaccharide-based HIV-1 2G12 mimotope vaccine induces carbohydrate-specific antibodies that fail to neutralize HIV-1 virions. Proc. Natl. Acad. Sci. USA 2008, 105, 15684–15689. [Google Scholar] [CrossRef]
  166. Liu, C.C.; Huo, C.X.; Zhai, C.; Zheng, X.J.; Xiong, D.C.; Ye, X.S. Synthesis and immunological evaluation of pentamannose-based HIV-1 vaccine candidates. Bioconjug. Chem. 2022, 33, 807–820. [Google Scholar] [CrossRef]
  167. Nicholas, D.T. A Proposal That Would Ban Manufacture, Supply, and Use of All Fluoropolymers and Most Fluorinated Reagents within the Entire EU. Org. Process Res. Dev. 2023, 27, 1422–1426. [Google Scholar]
Figure 1. Typical fluorinating reagents: 1, DAST (diethylaminosulfur trifluoride); 2, TASF ((tris(dimethylamino)sulfonium difluorotrimethylsilicate), 3, TBAF (tertbutylammonium fluoride); 4, Metal fluorides; 5, KHF2; 6, anhydrous hydrogen fluoride system; 7, SelectFluor, (1-chloromethyl-4-fluoro-1,4-diazoniabicyclo[2.2.2]octane bis(tetrafluoroborate) and 8, NFSI (N-fluorobenzenesulfonimide).
Figure 1. Typical fluorinating reagents: 1, DAST (diethylaminosulfur trifluoride); 2, TASF ((tris(dimethylamino)sulfonium difluorotrimethylsilicate), 3, TBAF (tertbutylammonium fluoride); 4, Metal fluorides; 5, KHF2; 6, anhydrous hydrogen fluoride system; 7, SelectFluor, (1-chloromethyl-4-fluoro-1,4-diazoniabicyclo[2.2.2]octane bis(tetrafluoroborate) and 8, NFSI (N-fluorobenzenesulfonimide).
Molecules 28 06641 g001
Figure 2. Nucleophilic fluorination of carbohydrates using different fluorinating reagents: (a) DAST as fluorinated reagents, (b) TASF as fluorinated reagents, (c) KHF2 as fluorinated reagents, (d) DFMBA as fluorinated reagents, (e) Cs18F as fluorinated reagents.
Figure 2. Nucleophilic fluorination of carbohydrates using different fluorinating reagents: (a) DAST as fluorinated reagents, (b) TASF as fluorinated reagents, (c) KHF2 as fluorinated reagents, (d) DFMBA as fluorinated reagents, (e) Cs18F as fluorinated reagents.
Molecules 28 06641 g002
Figure 3. Polyfluorination of carbohydrates via iterative nucleophilic reactions and the synthesis of 2,3,4-trifluoro-glucopyranose.
Figure 3. Polyfluorination of carbohydrates via iterative nucleophilic reactions and the synthesis of 2,3,4-trifluoro-glucopyranose.
Molecules 28 06641 g003
Figure 4. Difluorination of carbohydrates via nucleophilic reactions. (a) nucleophilic difluoroalkylation via DAST, (b) nucleophilic difluoroalkylation via Deoxo-Fluor, (c) nucleophilic difluoroalkylation via CBr2F2.
Figure 4. Difluorination of carbohydrates via nucleophilic reactions. (a) nucleophilic difluoroalkylation via DAST, (b) nucleophilic difluoroalkylation via Deoxo-Fluor, (c) nucleophilic difluoroalkylation via CBr2F2.
Molecules 28 06641 g004
Figure 5. Difluoroalkylation of sugars via nucleophilic reactions. (a), difluoroalkylation via BrCF2CO2Et, (b) difluoroalkylation via CF2CO2Et.
Figure 5. Difluoroalkylation of sugars via nucleophilic reactions. (a), difluoroalkylation via BrCF2CO2Et, (b) difluoroalkylation via CF2CO2Et.
Molecules 28 06641 g005
Figure 6. Trifluoromethylation of sugars via nucleophilic reactions. (a) trifluoromethylation via the nucleophilic addition of TMSCF3 to ketone on C-1, (b) trifluoromethylation of ketone on C-2, (c) trifluoromethylation of ketone on C-3.
Figure 6. Trifluoromethylation of sugars via nucleophilic reactions. (a) trifluoromethylation via the nucleophilic addition of TMSCF3 to ketone on C-1, (b) trifluoromethylation of ketone on C-2, (c) trifluoromethylation of ketone on C-3.
Molecules 28 06641 g006
Figure 7. Monofluorination of glycals using N-F reagents. (a) monofluorination via Selectfluor, (b) monofluorination via N-fluorobenzenesulfonamide.
Figure 7. Monofluorination of glycals using N-F reagents. (a) monofluorination via Selectfluor, (b) monofluorination via N-fluorobenzenesulfonamide.
Molecules 28 06641 g007
Figure 8. Monofluorination of glycals using O-F reagents.
Figure 8. Monofluorination of glycals using O-F reagents.
Molecules 28 06641 g008
Figure 9. Dehydroxymethylative fluorination of sugars via Ag(II)-initiated radical reaction.
Figure 9. Dehydroxymethylative fluorination of sugars via Ag(II)-initiated radical reaction.
Molecules 28 06641 g009
Figure 10. Decarboxylative fluorination of uronic acids via radical reaction. (a) decarboxylative fluorination via silver radical pathway, (b) Mes-Acr+-mediated oxidative decarboxylative fluorination.
Figure 10. Decarboxylative fluorination of uronic acids via radical reaction. (a) decarboxylative fluorination via silver radical pathway, (b) Mes-Acr+-mediated oxidative decarboxylative fluorination.
Molecules 28 06641 g010
Figure 11. Difluoromethylation of glycals via radical reaction.
Figure 11. Difluoromethylation of glycals via radical reaction.
Molecules 28 06641 g011
Figure 12. Trifluoromethylation of glycals using Umemoto reagent.
Figure 12. Trifluoromethylation of glycals using Umemoto reagent.
Molecules 28 06641 g012
Figure 13. Trifluoromethylation of glycals via photoredox and copper catalysis.
Figure 13. Trifluoromethylation of glycals via photoredox and copper catalysis.
Molecules 28 06641 g013
Figure 14. Trifluoromethylation of glycals via electrochemical method.
Figure 14. Trifluoromethylation of glycals via electrochemical method.
Molecules 28 06641 g014
Figure 15. Photocatalyzed reductive fluoroalkylation of 2-acetoxyglycals.
Figure 15. Photocatalyzed reductive fluoroalkylation of 2-acetoxyglycals.
Molecules 28 06641 g015
Figure 16. Copper-catalyzed β-difluoroacetylation of glycals via direct C–H functionalization.
Figure 16. Copper-catalyzed β-difluoroacetylation of glycals via direct C–H functionalization.
Molecules 28 06641 g016
Figure 17. Copper-catalyzed trifluoromethylation of 2-iodoglycals.
Figure 17. Copper-catalyzed trifluoromethylation of 2-iodoglycals.
Molecules 28 06641 g017
Figure 18. Monofluorination of sugars via building block strategy.
Figure 18. Monofluorination of sugars via building block strategy.
Molecules 28 06641 g018
Figure 19. Difluorination of sugars via building block strategy: (a) synthesis of difluoro-C-glycosides 113 via building block strategy, (b) synthesis of difluoroalkylated product 116 via building block strategy.
Figure 19. Difluorination of sugars via building block strategy: (a) synthesis of difluoro-C-glycosides 113 via building block strategy, (b) synthesis of difluoroalkylated product 116 via building block strategy.
Molecules 28 06641 g019
Figure 20. Trifluoromethylation of sugars via building block strategy: (a) synthesis of trifluoromethylated product 120 via building block strategy, (b) synthesis of trifluoromethylated product 125 via building block strategy.
Figure 20. Trifluoromethylation of sugars via building block strategy: (a) synthesis of trifluoromethylated product 120 via building block strategy, (b) synthesis of trifluoromethylated product 125 via building block strategy.
Molecules 28 06641 g020
Figure 21. Polyfluorination of sugars via building block strategy.
Figure 21. Polyfluorination of sugars via building block strategy.
Molecules 28 06641 g021
Figure 22. (a) Inhibitors of glycogen phosphorylase exhibited an enhanced binding affinity when modified with difluorinated derivatives. Ki = inhibition constant values. (b) The polyfluorinated analog exhibits enhanced diffusion characteristics. Pef = efflux permeabilities.
Figure 22. (a) Inhibitors of glycogen phosphorylase exhibited an enhanced binding affinity when modified with difluorinated derivatives. Ki = inhibition constant values. (b) The polyfluorinated analog exhibits enhanced diffusion characteristics. Pef = efflux permeabilities.
Molecules 28 06641 g022
Figure 23. Lipophilicities of fluorinated sugars.
Figure 23. Lipophilicities of fluorinated sugars.
Molecules 28 06641 g023
Figure 24. Ring conformation of fluorinated carbohydrate derivatives.
Figure 24. Ring conformation of fluorinated carbohydrate derivatives.
Molecules 28 06641 g024
Figure 25. The anomeric and fluorine gauche effects determining anomeric conformation. (a) exo conformation of the glycosidic bond in O-glycosides, (b) two conformations allow for the most stabilising hyperconjugation situations in CF2-glycosides.
Figure 25. The anomeric and fluorine gauche effects determining anomeric conformation. (a) exo conformation of the glycosidic bond in O-glycosides, (b) two conformations allow for the most stabilising hyperconjugation situations in CF2-glycosides.
Molecules 28 06641 g025
Figure 26. Schematic representation of the lone pair–s* interactions.
Figure 26. Schematic representation of the lone pair–s* interactions.
Molecules 28 06641 g026
Figure 27. Preferred conformations of the exocyclic fluoromethyl group.
Figure 27. Preferred conformations of the exocyclic fluoromethyl group.
Molecules 28 06641 g027
Figure 28. The relative rates of acid-catalyzed hydrolysis of deoxy- and deoxyfluoro-α-glucopyranoside derivatives.
Figure 28. The relative rates of acid-catalyzed hydrolysis of deoxy- and deoxyfluoro-α-glucopyranoside derivatives.
Molecules 28 06641 g028
Figure 29. Fluorinated maltose derivatives as mechanism-based inhibitors. (a), 2,4-dinitrophenyl-2-deoxy-2-fluoro-glucopyranoside as fluorinated glycoside inhibitor, (b) 2-fluoro-2-deoxyglucose as radiotracer for PET.
Figure 29. Fluorinated maltose derivatives as mechanism-based inhibitors. (a), 2,4-dinitrophenyl-2-deoxy-2-fluoro-glucopyranoside as fluorinated glycoside inhibitor, (b) 2-fluoro-2-deoxyglucose as radiotracer for PET.
Molecules 28 06641 g029
Figure 30. (a) Phlorizin and its 2-deoxyfluoro derivative as SGLT inhibitors; (b) P-3Fax-Neu5Ac inhibiting sialyltransferase activity.
Figure 30. (a) Phlorizin and its 2-deoxyfluoro derivative as SGLT inhibitors; (b) P-3Fax-Neu5Ac inhibiting sialyltransferase activity.
Molecules 28 06641 g030
Figure 31. (a) Fluorinated nucleotide–sugars as enzyme inhibitors or inactivators. (b) 3-Fluoro-sialic acid and its second-generation compounds containing a carbamate group at C-5.
Figure 31. (a) Fluorinated nucleotide–sugars as enzyme inhibitors or inactivators. (b) 3-Fluoro-sialic acid and its second-generation compounds containing a carbamate group at C-5.
Molecules 28 06641 g031
Figure 32. Fucosylation inhibitors and GDP-fucostatin galactose-1-phosphate uridyltransfer-ase inhibitors.
Figure 32. Fucosylation inhibitors and GDP-fucostatin galactose-1-phosphate uridyltransfer-ase inhibitors.
Molecules 28 06641 g032
Figure 33. (a) 2,3-Difluorosialosides as covalent sialidase inhibitors. (b) Nojirimycin, 1-deoxynojirimycin, and fluorinated analogs of iminosugars.
Figure 33. (a) 2,3-Difluorosialosides as covalent sialidase inhibitors. (b) Nojirimycin, 1-deoxynojirimycin, and fluorinated analogs of iminosugars.
Molecules 28 06641 g033
Figure 34. Fluorinated Thomsen–Friedenreich (TF) conjugates.
Figure 34. Fluorinated Thomsen–Friedenreich (TF) conjugates.
Molecules 28 06641 g034
Figure 35. Fluorine-modified STn.
Figure 35. Fluorine-modified STn.
Molecules 28 06641 g035
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Wei, X.; Wang, P.; Liu, F.; Ye, X.; Xiong, D. Drug Discovery Based on Fluorine-Containing Glycomimetics. Molecules 2023, 28, 6641. https://doi.org/10.3390/molecules28186641

AMA Style

Wei X, Wang P, Liu F, Ye X, Xiong D. Drug Discovery Based on Fluorine-Containing Glycomimetics. Molecules. 2023; 28(18):6641. https://doi.org/10.3390/molecules28186641

Chicago/Turabian Style

Wei, Xingxing, Pengyu Wang, Fen Liu, Xinshan Ye, and Decai Xiong. 2023. "Drug Discovery Based on Fluorine-Containing Glycomimetics" Molecules 28, no. 18: 6641. https://doi.org/10.3390/molecules28186641

Article Metrics

Back to TopTop