Home / Publications / Polyfluorinated benzoic acids as promising reagents for organic synthesis and medicinal chemistry

Polyfluorinated benzoic acids as promising reagents for organic synthesis and medicinal chemistry

Share
Cite this
GOST
Cite
GOST copy
Evgeny V. Shchegol'kov et al. Polyfluorinated benzoic acids as promising reagents for organic synthesis and medicinal chemistry // Russian Chemical Reviews. 2024. Vol. 93. No. 8. RCR5131
GOST all authors (up to 50) copy
Evgeny V. Shchegol'kov, Yanina V. Burgart, Irina V. Shchur, Victor I. Saloutin Polyfluorinated benzoic acids as promising reagents for organic synthesis and medicinal chemistry // Russian Chemical Reviews. 2024. Vol. 93. No. 8. RCR5131
 | 
RIS
Cite
RIS copy
TY - GENERIC
DO - 10.59761/RCR5131
UR - https://rcr.colab.ws/publications/10.59761/RCR5131
TI - Polyfluorinated benzoic acids as promising reagents for organic synthesis and medicinal chemistry
T2 - Russian Chemical Reviews
PB - Autonomous Non-profit Organization Editorial Board of the journal Uspekhi Khimii
AU - Shchegol'kov, Evgeny V.
AU - Burgart, Yanina V.
AU - Shchur, Irina V.
AU - Saloutin, Victor I.
PY - 2024
SP - RCR5131
IS - 8
VL - 93
ER -
 | 
BibTeX
Cite
BibTeX copy
@misc{2024_Shchegol'kov,
author = {Evgeny V. Shchegol'kov and Yanina V. Burgart and Irina V. Shchur and Victor I. Saloutin},
title = {Polyfluorinated benzoic acids as promising reagents for organic synthesis and medicinal chemistry},
month = {aug},
year = {2024}
}
 | 
MLA
Cite
MLA copy
Shchegol'kov, Evgeny V., et al. “Polyfluorinated benzoic acids as promising reagents for organic synthesis and medicinal chemistry.” Russian Chemical Reviews, vol. 93, no. 8, Aug. 2024, p. RCR5131. https://rcr.colab.ws/publications/10.59761/RCR5131.
Publication views
52

Keywords

biological activity
chemical transformations
polyfluorobenzoic acids
polyfluorosalicylic acids

Abstract

Polyfluoroaromatic compounds occupy a special place in organic synthesis due to a wide range of their chemical transformations and unique biological properties. The introduction of the carboxyl function into polyfluoroarenes allows further diversification of the chemistry of these compounds. This review summarizes data on the chemical transformations of polyfluorobenzoic acids, including derivatives of polyfluorosalicylic acids. The reactions of esterification, amidation, reduction, decarboxylation, metal-catalyzed decarboxylative cross-coupling, C–H functionalization, reductive defluorination, nucleophilic aromatic substitution, heterocyclization and complex formation are considered. Reactivity features of polyfluorobenzoates in comparison to their non-fluorinated counterparts are highlighted. The potential for practical applications of polyfluorobenzoic acid derivatives, primarily as biologically active compounds, is presented.

The bibliography includes 300 references.

1. Introduction

Benzoic acid and its derivatives are widely used as stabilizers, catalysts and reagents in the production of coolants, solvents, plastics, textiles, pesticides, paper, dyes, pharmaceuticals, as well as preservatives and flavourings for food and cosmetics[1, 2]. A separate area of application for such compounds is pharmaceuticals. This is because benzoic acid per se occurs naturally in many plants and is a by-product in the biosynthesis of various secondary metabolites[3, 4]. Benzoic acid and its sodium salt are part of many pharmaceutical preparations and mixtures. For example, acerbine is used as an antiseptic, amsterol and thermopsis are used to treat coughs, the combined preparation caffeine – sodium benzoate is a psychostimulant, etc.

The significance of substituted benzoic acids and their derivatives for medical applications is invaluable. Commercially important benzamide drugs include the analgesics ethenzamide and procainamide, the antidepressant moclobemide, the antiemetics bromopride and cisapride, the neuroleptics sulpiride and tiapride, the diagnostic agent 4-aminohippuric acid, the cytostatics imatinib and procarbazine, and the antiparkinsonian drug raclopride. The group of pharmacologically important drugs that are esters of benzoic acids is even larger and more diverse in both composition and use.

2-Hydroxybenzoic acid and its derivatives, commonly known as salicylates, deserve special attention as they are among the longest-lasting drugs on the pharmaceutical market. Aspirin, methyl salicylate, trolamine salicylate, salsalate, acelisin, etc.[5] are entered in the register of clinically used non-steroidal anti-inflammatory drugs. By modifying the structure of salicylic acid, it has been possible to tune the biological action of its derivatives, which have shown new types of activity. The introduction of an amino group at positions 4 and 3 led to the antituberculosis drug PASA (p-aminosalicylic acid)[6] and the antiulcer drug mesalazine[7, 8], respectively. The importance of salicylic acid as a phenolic phytohormone that regulates plant physiological processes[9] should not be overlooked, and its derivatives are used to reduce the incidence of viral, bacterial and fungal infections in various crops[10-13].

Due to their wide demand, benzoic acid derivatives, including salicylates, occupy leading positions in organic synthesis as building blocks for the preparation of a variety of open-chain[5, 14-17], as well as carbo- and heterocyclic compounds[18-21]. In addition, benzoates and salicylates are widely used as ligands in coordination chemistry[22, 23] and as analytical reagents[24-26].

Polyfluorobenzoic acids can also play an important role in solving various applied problems, as the introduction of fluorine atoms into organic compounds is effectively and successfully used in chemistry to modulate the physicochemical and biological properties of the resulting products. Due to specific possibilities of chemical modification, fluorinecontaining compounds, especially polyfluoroaromatic compounds, occupy an important place in organic synthesis[27-29]; they are in great demand as reagents for the preparation of valuable organic products for use in technology, medicine and agriculture[30]. The enormous potential of such derivatives in targeted drug development is confirmed by the fact that ~30% of drugs contain at least one fluorine atom in their molecules, often as a substituent in the aromatic ring[31-39]. Examples include fluoroquinolone antibiotics (levofloxacin, ciprofloxacin, sparfloxacin, etc.), antifungal agents (fluconazole, voriconazole) and hypoglycemic agents (sitagliptin) (Figure 1). Salicylate derivatives are not neglected either. For example, diflunisal, which contains fluorine, has twice the anti-inflammatory effect of aspirin and is less irritating to the mucous membranes of the gastrointestinal tract[40]. Fluoroaromatic products are also widely used in agrochemicals: for example, diflufenican, flutriafol, flusilazole are used as herbicides, and polyfluorobenzoyl ureas (diflubenzuron, chlorfluazuron, teflubenzuron, flufenoxuron) and ethoxazole are used as insecticides[41, 42].

Figure 1
Polyfluoroarene-based pharmaceutical drugs and insecticides.

In addition, polyfluoroarenes have been used to obtain liquid crystals[43] and polymers, including dendrimers[44, 45], which can be used to fabricate matrices for organic light-emitting diodes (OLEDs) (Figure 2).

Figure 2
Polyfluoroarene-based functional materials.

Given the practical importance of polyfluorobenzoic acids and the great prospects for their chemical modification, there was a need to address these aspects in a single review, as no data on these compounds have been summarized and analyzed to date. The present review aims to fill this gap by presenting the chemical properties and assessing the potential applications of polyfluorobenzoic acids and their derivatives, including polyfluorosalicylic acid analogues, mainly in the field of medicinal chemistry. From a synthetic point of view, polyfluorobenzoic acids have clear potential for modifications, since in addition to the transformations of the carboxyl group and the aromatic core, which are also inherent in their nonfluorinated analogues, they are characterized by reactions involving the polyfluoroaromatic ring due to the activation of the C – F bond[28].

2. Reactions of polyfluorinated acids involving carbonyl group

2.1. Synthesis of esters and amides

The most common ways to modify polyfluorobenzoic acids are the preparation of esters and amides therefrom. This is due not only to the relative simplicity of such reactions, but also to their great potential for the synthesis of a variety of biologically active products. The esters are mainly prepared using the classical esterification reaction by heating the reagents in excess alcohol in the presence of an acid catalyst[46-54]. Polyfluorobenzoic acid amides are preferably obtained directly from acids using peptide synthesis reagents (RPS) such as PyBOB (benzotriazol1-yloxytripyrrolidinophosphonium hexafluorophosphate), HATU (hexafluorophosphate azabenzotriazole tetramethyluronium), HOBt (N-hydroxybenzotriazole), DCC (1,3-dicyclohexylcarbodiimide) and others[55-60] or from their acid chlorides[60-67] in the presence of organic bases (pyridine, triethylamine, diisopropylethylamine (DIPEA), etc.) or without them (Scheme 1).

Scheme 1

The pool of alcohols and amines involved in such transformations is quite large, so we have confined ourselves to referring to current publications on the subject. Other methods for the synthesis of esters and amides and for the preparation of compounds of practical importance are discussed in more detail below. For example, 2-methoxy-2-oxo-1-phenylethyl pentafluorobenzoate was obtained in moderate yield by the reaction of pentafluorobenzoic acid and methyl phenyldiazoacetate in the presence of tropylium tetrafluoroborate (Trop) (Scheme 2)[68].

Scheme 2

The use of alkylating agents such as diazomethane[69], methyl iodide[70, 71], dimethyl sulfate[72] or ethyl iodide[73] in reactions with polyfluorosalicylic acids provided an approach to alkyl 2-alkoxypolyfluorobenzoates (Scheme 3). Partial hydrolysis of combined diesters was used to obtain ortho-alkoxy-substituted polyfluorobenzoic acids. For example, alkali treatment of methyl 2-methoxy-3,5,6-trifluorobenzoate and ethyl 2-ethoxy3,5-difluorobenzoate gives 2-methoxy-3,5,6-trifluorobenzoic acid[74] and 2-ethoxy-3,5-difluorobenzoic acid[73], respectively.

Scheme 3

Polyfluorosalicylic acids readily form acetylpolyfluorosalicylic acids under the action of acetic anhydride (Scheme 4; T is heating)[51, 75].

Scheme 4

In vivo tests have revealed a moderate anti-inflammatory effect of fluorinated analogues of aspirin and methyl 3,4,5-trifluorosalicylate[51]. In addition, fluorinated acetylsalicylic acids have a marked analgesic effect, superior not only to aspirin but also to diclofenac.

Hexadecanyl- and 10-chlorodecanyl-substituted tetrafluorosalicylates, synthesized from tetrafluorosalicylic acid and the appropriate alcohols by boiling in benzene over Amberlite IR120 ion exchange resin (Scheme 5), have properties that allow them to be considered as potential antifriction additives for advanced motor oils[76].

Scheme 5

The reaction of polyfluorobenzoic acids with [3-bromo-1-(3chloro-2-pyridyl)-1H-pyrazol-5-yl]methanol in the presence of DCC gave [3-bromo-1-(3-chloro-2-pyridyl)-1H-pyrazol-5-yl]methylpolyfluorobenzoates (Scheme 6), the difluoro derivative of which has insecticidal activity against the cabbage pests Plutella xylostella and Spodoptera frugiperda[77].

Scheme 6

Among the O-substituted esters of difluorosalicylic acid, biologically active compounds have been identified such as 6-[(4,6-dimethoxypyrimidin-2-yl)oxy]-2,3-difluorobenzoic acid and its methyl ester (1), which have herbicidal activity[78], and methyl 2-[phenyl(pyrrolidin-3-yl)methoxy]-3,5-difluorobenzoate 2, which are serotonin and noradrenaline reuptake inhibitors[79].

Structures 1, 2

Selective alkylation of 4,6-dihydroxy-2,3,5-trifluorobenzoic acid via the 4-hydroxyl group was carried out. For this purpose, the dihydroxy acid was first treated with pentafluorophenyltrifluoroacetate to afford pentafluorophenyl ester, which was further reacted with paraformaldehyde[80] or acetaldehyde[81] in the presence of 1,4-diazabicyclo[2.2.2]octane (DABCO) to give 7-hydroxy-5,6,8-trifluoro-4H-benzo[d][1,3]dioxin-4-ones (Scheme 7). The latter compounds were coupled with O-(4methoxytetrahydropyran-4-yl)-2'-deoxyuridine under Mitsunobu conditions to give O-(5-hydroxy-4-carboxy-2,3,6-trifluorophenyl)-2'-deoxyuridine in moderate yield (see Scheme 7, path a)[82, 83]. Treatment of benzo[d][1,3]dioxin-4ones with alkyl bromides and subsequent passing the intermediates through an ion exchange resin column furnished the target 4-alkoxy-substituted 3,5,6-trifluorosalicylic acids (path b)[80, 81]. O-(5-Hydroxy-4-carboxy-2,3,6-trifluorophenyl)-2'-deoxyuridine and similar acids containing geranyl- or farnesyl-substituted units showed inhibitory activity against Ras protein farnesyl transferase and geranylgeranyl transferase I, one of the key enzymes involved in proliferative processes[80, 82].

Scheme 7

An original method for the synthesis of 6-(6-hydroxy-3,5difluorobenzamido)hexanoic acid has been developed[84]. First, 3,5-difluorosalicylic acid was refluxed in xylene in the presence of a small excess of acetic anhydride, resulting in oligosalicylate which was reacted with aminocaproic acid without isolation (Scheme 8).

Scheme 8

Rearrangements of nitrone derivatives have also been used to prepare polyfluorosalicylic acid amides. For example, α-pentafluorophenyl-N-phenylnitrones form tetrafluorosalicylic acid arylamides on heating in DMF under acidic conditions[85] or on refluxing in toluene[86] (Scheme 9).

Scheme 9

The EDC-mediated condensation of 2-methoxymethoxy-3,6-difluorobenzoic acid with 3-amino-2-ethyl-N-(3-fluorophenylethyl)but-2-enamide in the presence of hydroxybenzotriazole was used to synthesize amide A, the cyclization of which in the presence of alkali gave 2-(2-hydroxy-3,6-difluorophenyl-6-methyl-3-[2-(3-fluorophenyl)ethyl]-5-ethyl-4(3H)-pyrimidone (Scheme 10)[87].

Scheme 10

The methyl esters of perfluoro-3,4,5-trifluorosalicylic acid have been shown to react with ammonia in methanol on heating to give the corresponding amides, which are converted to acids when heated in hydrochloric acid solution (Scheme 11)[88]. 3,4,5-Trifluorosalicylic acid amide in the in vivo tests showed a pronounced analgesic effect in the hot plate test along with low acute toxicity.

Scheme 11

In contrast to reactions with ammonia, polyfluorosalicylate esters form stable salts in reactions with diethylamine (Scheme 12)[89]. At the same time, polyfluorosalicylic acid ethyl esters give N,N-diethylammonium 2-hydroxypolyfluorobenzoates and methyltetrafluorosalicylate gives N,N-diethylammonium 6-methoxycarbonyl-2,3,4-trifluorophenolate. However, methyl tetrafluorosalicylate forms the same product as its ethyl counterpart.

Scheme 12

The reaction of 3,4-difluorosalicylic acid methyl ester with 2,2-diphenylethylamine at 110 °C yields N-(2,2-diphenylethyl)2-hydroxy-3,4-difluorobenzamide (Scheme 13), which selectively inhibits dihydroorotate dehydrogenase, an enzyme of the malaria parasite Plasmodium falciparum, at micromolar concentrations[90].

Scheme 13

Based on 4,5-difluorosalicylic acid and 2-methoxy-4,5-difluorobenzoic acid, conjugates with tacrine linked via an amidoalkylene spacer have been obtained (Scheme 14). Such products are able to inhibit acetyl-(AChE) and butyrylcholinesterase at micromolar concentrations and block AChE-induced β-amyloid aggregation, as well as have antioxidant and metal chelating properties, allowing them to be considered as promising multitarget agents for the treatment of Alzheimer’s disease[91].

Scheme 14

By the reaction of tetrafluorophthalic anhydride with primary (cyclo)alkylamines[92] or with aminobarbituric acid derivatives[93, 94], a series of 2,3,4,5-tetrafluorobenzamides with antiangiogenic properties have been obtained (Scheme 15).

Scheme 15

A large number of publications, most of them patents, are devoted to polyfluorosalicylate amides, which have different biological effects depending on the variation of the amide moiety. For example, 3,4-difluorosalicylic acid amide (3) has been patented as an antagonist of M 3-muscarinic receptors, which is promising for the development of a drug for the treatment of asthma and chronic obstructive pulmonary disease[95, 96]. The patent[97] describes (±) N-(2-hydroxy-5-oxo-7oxabicyclo[4.1.0]hept-3-en-3-yl)-2-methoxy-3,4-difluorobenzamide (4), which is capable of inhibiting the transcription factor NF-κB, dysregulation of which triggers inflammatory processes, autoimmune diseases, viral infections and cancer. 2-(6-Hydroxy-2,4-difluorobenzamido)-5-(morpholin-4-ylsulfonyl)-thiophene-3-carboxamide (5) has been proposed as a therapeutic agent for the treatment of cystic fibrosis[98]. Amides 6 showed antiproliferative activity against 22RV1 and LNaP95 cancer cells causing prostatitis[62, 63]. N-[3-Chloro-4-(4-chlorophenoxy)phenyl]-2-hydroxy-3,5-difluorobenzamide (7) showed properties as an inhibitor of human pancreatic lipase[58], the same compound is claimed in a patent[99] as a potential treatment for dengue fever.

Structures 3 – 16

A number of polyfluorosalicylic acid amides 8 have been patented as anticoagulants[100]. Alen et al.[101] propose to use the sepiapterin reductase inhibitor (6-azaspiro[3.4]octan-6-yl) (6-hydroxy-2,4-difluorophenyl)methanone (9) to create new potent analgesics. Polyfluorinated N-([4-oxo-6-fluoro-7(cyclohexylamino)-1-cyclopentyl-1,4-dihydroquinolin-3-yl] methyl)-2-hydroxybenzamides 10 have been claimed as inhibitors of platelet aggregation and P2Y12 receptors[102]. Salicylamide 11, which contains a polynuclear heterocyclic moiety, has been shown to be an effective inhibitor of B-cell lymphoma protein 6 (BCL-6)[57].

Polyfluorobenzoic acid amides also proved to be a convenient platform for the generation of promising physiologically active compounds. For example, N-{4-[3-(1-methyl-1H-imidazol-2yl)-3-oxoprop-1-en-1-yl]phenyl}-2,3,4,5,6-pentafluorobenzamide 12 has anticancer activity against HCT-116 and NCI-H522 cell lines[103], and polyfluoro-substituted amides 13 bearing a coumarin residue 67 showed activity against the MCF-7, T47D, MDA-MB-231 and SkBr3 cancer cell lines in the micromolar concentration range. N-[4-(Hydroxycarbamoyl)benzyl]-2,3,4,5,6-pentafluorobenzamide (14) was found to be a selective inhibitor of histone deacetylase 6 (HDAC6), which causes hematological malignancies. N-(5-Sulfamoyl-2,3-dihydro-1H-inden-2-yl)-2,3,4,5,6-pentafluorobenzamide (15) has moderate anticonvulsant activity[104], while pentafluorobenzamide 16 with a polynuclear amine moiety has shown activity against MCF-7 breast cancer cell cultures and the adriamycin-resistant MCF-7/ADR cell line[105].

The development of biologically active amides based on substituted polyfluorosalicylic acids proved promising. For example, [2-(4-bromo-2-fluorobenzylcarbamoyl)-4,5-difluorophenoxy)]acetic acid (17) was patented as a potent inhibitor of human aldose reductase 2 (hALR2), an enzyme responsible for complications in diabetes mellitus[106]. N-(3-Carbamoyl-4-fluorophenyl)-6-[2-methoxy-4-(trifluoromethoxy)phenoxy]-2,3,4-trifluorobenzamide 18 was found to be a Nav1.8 sodium channel blocker[107]. (S)-2-tert-Butoxy-2-[6-methyl-1-(2-methoxy-3,6-difluorobenzoyl)-4-(p-methylphenyl)-2,3-dihydroindol-5-yl]acetic acid (19) and (S)-2-tert-butoxy-2-[6-methyl-1-(2-methoxy-3,4-difluorobenzoyl)-4-(5-methyl-8-fluorochroman-6-yl)-2,3-dihydroindol-5-yl]acetic acid (20) showed properties of potent inhibitors of human immunodeficiency virus (HIV-1)[108].

Structures 17 – 20

To summarize, the above data indicate the prospects for the development of new methods for the synthesis of amides and esters of polyfluorobenzoic and polyfluorosalicylic acids as privileged scaffolds for medicinal chemistry.

2.2. Reduction reactions

The reduction of the carboxyl function in polyfluorobenzoic acids is used to obtain polyfluoroaryl-substituted benzyl alcohols. A number of methods are reported for the synthesis of pentafluorobenzyl alcohol by reduction of the carboxyl group in pentafluorobenzoic acid with zinc(II) borohydride, generated in situ from zinc(II) chloride and sodium borohydride[109], using a catalytic system consisting of zinc(II) acetate, phenyl silane and N-methylmorpholine[110], and in the presence of titanium(IV) chloride and borazane[111] (Scheme 16). Using lithium aluminium hydride, this process is also accompanied by reductive defluorination to give 2,3,5,6-tetrafluorobenzyl alcohol[109].

Scheme 16

The carboxyl group in polyfluorobenzoic acids can be sequentially converted into alcohol and aldehyde functions. Thus, reduction of the carboxyl substituent in difluorosalicylic acids with LiAlH4 (see [112, 113]) or BH3 · Me2S (see [114]) affords 3,5-difluoro-, 3,4-difluoro- and 4,5-difluoro-2-(hydroxymethyl)phenols (Scheme 17). Oxidation of alcohols with pyridinium chlorochromate (PCC) or 2,3-dichloro-5,6-dicyano-1,4-benzoquinone (DDQ) opened the way to difluorosalicylic aldehydes[113, 114].

Scheme 17

2.3. Decarboxylation reactions

Decarboxylation of polyfluorobenzoic acids is used to produce polyfluorobenzenes. For example, it has been shown that polyfluorobenzoic acids can be decarboxylated by heating in a sealed ampoule in the presence of Cu2O and 1,10-phenanthroline (phen) (Scheme 18)[115].

Scheme 18

The synthesis of polyfluorobenzenes is also accessible by heating polyfluorobenzoic acids in a toluene – water mixture in the presence of the catalytic system [(COD)Rh(OH)2 – 1,3-bis(diphenylphosphino)propane (DPPP) and a base (NaOH) (Scheme 19)[116].

Scheme 19

The use of polyfluorosalicylic acids in decarboxylation reactions is an effective approach to polyfluorophenols which are difficult to obtain. Thermal decarboxylation of perfluoro(see [117]) and 3,5,6-trifluorosalicylic acids[117, 118] is carried out in a sealed ampoule at temperatures of 270 – 300 °C for 3 – 7 h (Scheme 20) to give 2,3,4,5-tetrafluoro-2,4,5-trifluorophenols.

Scheme 20

Depending on the conditions, decarboxylation of 2,6-dimethoxy-3,4,5-trifluorobenzoic acid affords 4,5,6-trifluororesorcin or its dimethyl ester (Scheme 21)[119].

Scheme 21

2.4. Metal-catalyzed decarboxylative cross-coupling reactions

Modern approaches to the modification of polyfluorobenzoic acids or their salts, which have been actively developed in recent years, are metal-catalyzed decarboxylative cross-coupling reactions used for the synthesis of biaryl or aryl-hetaryl derivatives. A process has been proposed for the preparation of a variety of biaryl compounds (Scheme 22), based on the reaction of potassium polyfluorobenzoates with (het)aryl halides or triflates in the presence of the catalytic system Pd(OAc)2 (0.01 – 0.04 equiv.) – PCy3 or P(o-Tol)3 (0.02 – 0.08 equiv.)[120]. The same research group[121] showed that cheaper copper(I) iodide could be used in this reaction. However, (het)aryl chlorides could not be involved in the cross-coupling under these conditions, so this process was successfully carried out for (het)aryl iodides and bromides in the presence of phen as a ligand. Notably, the highest yields of products were reached with pentafluorobenzoic acid, whereas a decrease in the number of fluorine atoms often led to a decrease in the yield of the target biaryls, down to trace amounts.

Scheme 22

Copper(I) iodide immobilized on a solid support (mesoporous MCM-41) and functionalized with the phen ligand was proposed as a renewable catalyst for the cross-coupling of potassium salts of polyfluorobenzoic acids with aryl halides (Scheme 23)[122]. This method not only allows the catalyst to be recovered many times without a significant decrease in the yield of the target product, but also increases the yield of biaryls based on di- or trifluorobenzoic acids.

Scheme 23

For the cross-coupling of potassium polyfluorobenzoates with (het)aryl halides or pseudohalides, the Ni(COD)2 catalyst was also used in the presence of Bu3tP · HBF4, 1,1'-(diphenylphosphino)ferrocene (DPPF) or 2-[2-(dicyclohexylphosphino)phenyl]-1-methyl-1H-indol (CM-Phos) as phosphine ligands (Scheme 24)[123].

Scheme 24

The Ni(COD)2 catalyst was also utilized in the synthesis of biaryls or diarylmethanes via the reaction of potassium polyfluorobenzoates with arylpivalates, arylsulfamates or arylmethylpivalates, but in this case it was necessary to add zinc(II) acetate to the reaction mixture (Scheme 25)[124]. It was also found that in the absence of the second aryl reagent, decarboxylation of potassium pentafluorobenzoate gives pentafluorobenzene.

Scheme 25

It has been shown[125] that zinc salts of polyfluorobenzoic acids can be effectively used in the decarboxylative crosscoupling, and the reaction with (poly)aryl bromides or nonaflates runs in the presence of the catalyst Pd(PPh3)4 under milder conditions (heating up to 60 °C) (Scheme 26). This approach provided the preparation of a large number of bis- and polyaryl derivatives (Figure 3).

Figure 3
Examplary polyaryl structures obtained by the cross-coupling reaction (see Scheme 26).

Scheme 26

Polyfluorobenzoic acids per se have also been involved in decarboxylative cross-coupling reactions, but these have been carried out under Suzuki conditions using arylboronic acids, palladium(II) trifluoroacetate as catalyst and silver(II) carbonate as the base[126]. At the same time, the replacement of the catalyst with copper(II) triflate in the reaction of 4-methoxy-2,6-difluorobenzoic acid with phenylboronic acid gave O-phenylsubstituted 4-methoxy-2,6-difluorobenzoate (Scheme 27).

Scheme 27

In addition, polyfluorobenzoic acids can undergo Pdcatalyzed C – H functionalization with (het)arenes (Scheme 28) such as benzene and its substituted analogues[127, 128], (benzo)thiazoles[127], 4-methylfuran[129], 4-substituted thiophenes[128, 130] and indole[131]. The reaction requires the presence of a silver(II) carbonate base and sometimes the addition of a phosphonium ligand. As a result of these processes, the range of known polyfluorinated biaryl and aryl-hetaryl derivatives has been greatly expanded.

Scheme 28

The use of N-methylpyridinium iodide and its benzoanalogues in the C – H functionalization of polyfluorobenzoic acids promoted by CuBr2 and the phase transfer catalyst tetra-n-butylammonium bromide (TBAB) in the synthesis of 2-polyfluorophenylpyridines, 2,6-difluorophenylpyridine and 1-polyfluorophenyl(iso)quinolines has been described (Scheme 29)[132].

Scheme 29

It has been shown that polyfluorobenzoic acids or their salts can undergo decarboxylative cross-coupling with (benz)oxazoles in the presence of a nickel catalyst, NiCl2(PCy3)[133] or Ni(OTf)2 (see [134]), to give 2-polyfluorophenyl(benz)oxazoles (Scheme 30).

Scheme 30

Polyfluorobenzoic acids were found to undergo self-coupling with double decarboxylation in the presence of palladium[135] or copper[136] catalysts to furnish symmetrical polyfluorobiaryls (Scheme 31).

Scheme 31

In addition, polyfluorobenzoic acids can undergo crosscoupling with other (het)arylcarboxylic acids under the action of the PdCl2 – PPh3 catalytic system in the presence of silver(II) carbonate, leading to asymmetric biaryl or aryl-hetaryl derivatives (Scheme 32)[135]. The moderate yields of the target products are explained by the formation of symmetric biaryl byproducts.

Scheme 32

The pool of reagents subjected to cross-coupling reactions with polyfluorobenzoic acids and their potassium salts has been extended by the use of styrenes. Such transformations were carried out under the catalysis of Pd(O2CCF3)2 in the presence of Ag2CO3 to produce styryl-polyfluorobenzenes (Scheme 33)[137]. It should be noted that these reactions are sometimes accompanied by decarboxylation of the starting polyfluorobenzoic acids, which decreases the yield of the target products.

Scheme 33

1,2,3,4,5-Pentafluorostilbene was obtained from potassium pentafluorobenzoate and (2-bromoethyl)benzene in the presence of the CuI – phen catalytic system (Scheme 34)[121].

Scheme 34

The outcome of the reaction of polyfluorobenzoic acids with α,β-unsaturated carbonyl compounds in the presence of the catalyst [(COD)Rh(OH)]2 and NaOH base depends on the phosphonium ligand used. Thus, in the case of (4R,5R)-(–)-4,5-bis(diphenylphosphanylmethyl)-2,2-dimethyl-1,3-dioxolane (DIOP), mainly alkyl 3-(polyfluorophenyl)prop-2-enoates or 3-(polyfluorophenyl)prop-2-enamides are formed[116], whereas the use of 2,2'-bis(diphenylphosphino)-1,1'-binaphthyl (BINAP) gives alkyl-3-(polyfluorophenyl)propanoates or 3-(polyfluorophenyl)propanamides (Scheme 35)[138].

Scheme 35

It was shown that polyfluorobenzoic acids react with allyl acetate when heated in a mixture of DMSO – dioxane in the presence of catalysts Pd(OAc)2, Cu2O and oxidant Ag2CO3 to form a mixture of 3-(polyfluorophenyl)allyl acetate (as E,Z-isomers) and 2-(polyfluorophenyl)allyl acetate in a ratio of ca. 20 : 1 (Scheme 36)[139].

Scheme 36

However, attempts to use butadiene in the cross-coupling reaction with polyfluorobenzoic acids in the presence of the catalytic system [Rh(COD)Cl]2–tri(2-furyl) phosphine (TFP) resulted only in the formation of esters, (but-3-en-2-yl)polyfluorobenzoates (Scheme 37)[140].

Scheme 37

Summarizing the information given in this Section, it can be concluded that metal-catalyzed cross-coupling reactions of polyfluorobenzoic acids provide access to a diversity of (het)aryl-containing polyfluoroarenes.

2.5. Other decarboxylative reactions

In addition to cross-coupling, polyfluorobenzoic acids are known to undergo decarboxylative thiolation, halogenation and acylation reactions and also combinations of these transformations with cross-coupling.

In the reaction of pentafluorobenzoic acid and thiophenol mediated by the CuI – phen system, in addition to thiolation, nucleophilic substitution of the fluorine atom in the para position occurs to afford (perfluoro-1,4-phenylene) bis(phenylsulfane) (Scheme 38)[141].

Scheme 38

The potassium salts of polyfluorobenzoic acids undergo a decarboxylative acylation with (het)aroyl fluorides to give the corresponding biaryl ketones or aryl-hetaryl ketones (Scheme 39)[142], without the need for catalysts and/or bases.

Scheme 39

Polyfluorobenzoic acids undergo halogenation along with decarboxylation by sequential treatment first with a Bu3t PAuCl complex in the presence of Ag2O and then with halosuccinimides[143] using TBAB[144] or iodine[145] under the action of a base (K3PO4) (Scheme 40).

Scheme 40

It was shown that consecutive reactions of iodination of 2,6-difluorobenzoic acid with iodine and cross-coupling of the intermediate with pentafluorobenzoic acid under the action of the CuI – phen catalytic system give 2,2',3,4,5,6,6,6'-heptafluoro-1,1'-biphenyl (Scheme 41)[145].

Scheme 41

An original synthetic approach to fluorinated biaryls, styrylpolyfluorobenzenes and arylethynylpentafluorobenzenes based on pentafluorobenzoic esters of ketoximes by cleavage of O – N, С–С bonds and subsequent arylation followed by decarboxylation has been proposed (Scheme 42)[146].

Scheme 42

Metal-catalyzed reactions of pentafluorobenzoyl oximes, in which the polyfluorobenzoic acid residue acts as an auxiliary group to generate the iminyl radical, have been reported (Scheme 43). The subsequent transformations include several directions, which are discussed in detail in the review articles: for example, aza-Heck cyclization[147-156], as well as functionalization of C(sp3) – C(sp3) and C(sp3) – H bonds[156-159]. The realization of transformations involving pentafluorobenzoate as the leaving group opens up possibilities for the synthesis of a wide range of open-chain and heterocyclic compounds.

Scheme 43

To conclude, the analysis of the data considered in this Section shows that the decarboxylation of polyfluorobenzoic acids can be used to introduce new groups of different nature into the molecule.

3. Transformations of polyfluorobenzoic acids involving benzene ring

3.1. С – Н Functionalization reactions

Partially fluorinated benzoic acids and their derivatives can undergo C – H functionalization reactions that are also typical of non-fluorinated analogues. Thus, a method has been proposed for the selective Pd-catalyzed ortho-hydroxylation of difluorobenzoic acids with hydrogen peroxide in the presence of [2-methyl-2-(6-oxo-1,6-dihydropyridin-2-yl]propanoic acid (L3) to give the corresponding difluorosalicylic acids (Scheme 44)[113].

Scheme 44

A series of substituted biaryls have been synthesized by the C – H functionalization reaction of di- and trifluorobenzamides with aryl bromides, using palladium(II) chloride as a catalyst (Scheme 45)[160]. It should be noted that the arylation proceeds regioselectively, exclusively to the ortho position relative to the two fluorine atoms in the benzene ring.

Scheme 45

3,5-Difluorobenzamide reacts with 2-bromobenzyl bromides in the presence of palladium(II) acetate to give 2-(2-bromobenzyl)-3,5-difluorobenzamides (Scheme 46)[161].

Scheme 46

It has been shown that N-disubstituted di(trifluorobenzamides) can also undergo a Pd-catalyzed C – H functionalization with ethyl acrylate to yield ethyl 3-{3-[(4-nitrophenylsulfonyl) (2-cyanophenethyl)carbamoyl]-2,4-difluorophenyl}acrylates 21 (Scheme 47)[162].

Scheme 47

3.2. Reductive hydrodefluorination reactions

Methods for the selective reductive hydrodefluorination at the para-position of polyfluorobenzoic acids by treatment with zinc in liquid[163, 164] or aqueous ammonia[165] have been reported (Scheme 48). The use of the reductive sodium – liquid ammonia system in this reaction produces a mixture of (fluoro)benzoic acids[163, 164].

Scheme 48

The esters and amides of pentafluorobenzoic acid undergo selective monohydrodefluorination at para position in DMSO when treated with sodium borohydride (Scheme 49)[166].

Scheme 49

A similar process for esters has also been carried out under photocatalytic conditions in the presence of Ir(ppy)3 complex (ppy is 2-phenylpyridine)[167] but in this case, the yields of products were somewhat lower. Methyl 4-amino-2,3,5,6-tetrafluorobenzoate is characterized by selective dehydrodefluorination of the ortho-fluorine atom to give methyl 4-amino-3,5,6-trifluorobenzoate[167].

Treatment of pentafluobenzoic acid solution with zinc in the presence of 0.01 equiv. of the NiCl2 – 2,2'-bipyridine (bipy) system leads to selective reductive hydrodefluorination of the ortho-fluorine atom and formation of 2,3,4,5-tetrafluorobenzoic acid. With 0.05 equiv. of this system, hydrodefluorination of both ortho-fluorine atoms takes place to give 3,4,5-tetrafluorobenzoic acid (Scheme 50)[168]. It should be noted that the regioselectivity of the process reduces dramatically when pentafluorobenzoic acid ethyl ester or amide is involved in this reaction, since incomplete conversion of the starting materials is observed, and in both cases a mixture of ortho- and para- defluorinated products is formed.

Scheme 50

The selective monohydrodefluorination of the ortho-fluorine atom in pentafluorobenzoic acid and in 2,6-difluorobenzoic acid in THF by heating in the presence of 0.2 equiv. of ytterbium(II) complex YbCp2(DME) (Cp is cyclopentadienyl, DME is 1,2-dimethoxyethane), 2 equiv. of magnesium and 0.02 equiv. of iodine has been described (Scheme 51)[169].

Scheme 51

By varying the reaction conditions, including the amount of a base (DIPEA) and the reaction temperature, Kharbanda and Weaver 115 succeeded in carrying out mono-, di- and trihydrodefluorination in methyl pentafluorobenzoate to obtain the corresponding methyl esters of 2,3,5,6-tetrafluoro-, 2,3,5-trifluoro-3,5-difluorobenzoic acids (Scheme 52). Methyl 2,3,5,6-tetrafluorobenzoate was further subjected to Pdcatalyzed C – H functionalization mediated with 1-iodo-3,5-dimethyl- or (3-chloroprop-1-en-1-yl)benzenes to give the corresponding 4-substituted products.

Scheme 52

It should be noted that the transformations of arylamides of pentafluorobenzoic acids under similar conditions are less selective because the reaction produces mixtures of mono- and dihydrodefluorinated derivatives 22 – 26. The ratio of these products depends not only on the amount of DIPEA and water, but also on the structure of the aryl moiety (Scheme 53)[115].

Scheme 53

Under similar conditions, a photocatalytic cross-coupling of polyfluorobenzoate esters with (het)arenes was carried out without isolating the hydrodefluorinated intermediate, resulting in the synthesis of biaryl derivatives and 4-(pyrrol-3-yl)substituted methyl esters of 2,3,5,6-tetrafluorobenzoic acid (Scheme 54)[170].

Scheme 54

To summarize, the conditions for selective para- and ortho- hydrodefluorination of polyfluorobenzoic acids, their esters and amides are described, and the possibility of polyhydrodefluorination reactions is also demonstrated.

3.3. Nucleophilic aromatic substitution

For the modification of polyfluorobenzoic acids, their esters and amides, the nucleophilic aromatic substitution of fluorine atoms (SNArF) reactions, which allow the introduction of various moieties, including pharmacophore ones, into the polyfluorobenzoate molecule, have great potential.

It has been shown that the fluorine atom in the para position of pentafluorobenzoic acid or its esters is easily substituted under the action of various nucleophilic reagents such as sodium methoxide, sodium methanethiolate, hydrazine[171], sodium hydrosulfide[172], mercaptoethanol[173], imidazole[174] (Scheme 55). However, the reaction of pentafluorobenzoic acid with an ethanol solution of methylamine or dimethylamine produces a mixture of products from which para- and ortho-N-(di)methylsubstituted tetrafluorobenzoic acids could be isolated[171], with the starting acid not reacting with aqueous ammonia.

Scheme 55

Reactions of polyfluorobenzoic acids via the SNArF mechanism are actively used to introduce the hydroxyl function, and the search for selective methods of para- and ortho- hydroxylation has become the main trend of such studies. Of particular interest is the development of approaches to o-hydroxypolyfluorobenzoic acids as fluorinated analogues of salicylic acid.

Fluorine atoms in polyfluorobenzoic acids can be easily replaced with the OH group in 1,3-dimethylimidazolidin-2-one (DMI) upon heating with alkali[175-177]. Not only the number of fluorine atoms in the benzene ring, but also their position affects the process selectivity. Thus, when 2,4-difluoro-2,3,4-trifluorobenzoic acids are treated with sodium hydroxide, the content of the resulting ortho-hydroxy isomers in the mixture reaches 99 and 99.8%, respectively, and their isolated yield is almost quantitative (95%) (Scheme 56, Table 1)[177]. Under similar conditions, in 2,4,6-trifluorobenzoic acid, only one of the ortho-fluorine atoms is predominantly substituted and the ratio of ortho- to para-isomers reaches 97 : 3. The reaction of 2,4,5-trifluorobenzoic acid with alkali is less selective, the ratio of ortho- to para-isomers being 72 : 28. The authors of patents[178, 179] claim that, under similar conditions, 2,4,5-trifluorobenzoic acid is selectively converted to 4,5-difluorosalicylic acid, although no convincing evidence of the structure of the product is provided, except for the 1H NMR spectrum, which is not very informative in this case. If the number of fluorine atoms is increased to 4 or 5, the process becomes even less selective, since 2,3,4,5-tetrafluorobenzoic acid in this reaction gives a mixture of products in the ratio 67 : 33, whereas the reaction of pentafluorobenzoic acid with NaOH furnishes only a hard-to-separate mixture of products[177] (see Scheme 56 and Table 1). It should be noted that the 3,4,5-trifluoroand tetrafluorosalicylic acids thus synthesized could not be isolated in pure form.

Scheme 56

Table 1
\[ \]
Products of hydroxylation of fluorobenzoic acids (see Scheme 56).
(1)

A similar method is used for the preparation of the isomeric 3,4-difluoro- (see [55, 57, 180-182]) and 2,4-difluorosalicylic acids[183].

3,5-Difluorosalicylic acid was synthesized by the reaction between 2,3,5-trifluorobenzoic acid and alkali on heating in DMSO (Scheme 57)[70, 172].

Scheme 57

It should be noted that this acid can be used as an additive to herbicides such as paraquat and atrazine to increase their efficacy against tobacco mosaic virus and reduce their toxic effect on tobacco[184-188].

Our research group[119, 189] found a selective ortho-substitution of the fluorine atom in pentafluorobenzoic acid under the action of magnesium methoxide. As a result, 2-methoxytetrafluorobenzoic acid is formed, which is hydrolyzed in situ to tetrafluorosalicylic acid in 42% yield (Scheme 58). The proposed mechanism of methoxylation involves the initial formation of intermediate I through the addition of magnesium methoxide to the carboxyl group. Due to the specific coordinating ability of the magnesium atom, the fluoride ion is eliminated on heating and the methoxide anion is subsequently attached to the ortho-position of intermediate II[119]. For pentafluoro (see [190]) and 2,3,5,6-tetrafluorobenzoic acids[51], the optimum conditions for the selective substitution of one of the ortho-fluorine atoms were chosen including the use of 2.5 equiv. of Mg(OMe)2 in boiling toluene and 3.5 equiv. of Mg(OMe)2 when heating in diglyme. The mixtures of the resulting methoxy-substituted acids II were treated with PCl5, which made possible to separate the corresponding 2-methoxypolyfluorobenzoic acid chlorides by distillation. Treatment of the latter with 48% HBr solution gave tetrafluoro- and 3,5,6-trifluorosalicylic acids (see Scheme 58). Hydrolysis of 2-methoxy-3,4,5,6-tetrafluorobenzoic acid chloride with 20% NaOH solution at room temperature afforded 2-methoxytetrafluorobenzoic acid in individual form[190].

Scheme 58

When pentafluorobenzoic acid is reacted with 12 equiv. of magnesium methoxide on heating in toluene or diglyme, both ortho-fluorine atoms are substituted to give 2,6-dimethoxy-3,4,5-trifluorobenzoic acid (see Scheme 58).

In the case of polyfluorobenzoic acids with only one ortho- fluorine atom, the synthesis is not complicated by the formation of disubstitution by-products, and the mono-ortho-methoxylation proceeds selectively, providing virtually quantitative yields of the target polyfluorosalicylic acids (Scheme 59)[190].

Scheme 59

In our opinion, this method is the most promising approach to polyfluorosalicylic acids, especially to those containing more than three fluorine atoms in the benzene ring, because of the simplicity of the equipment and the availability of starting compounds. This approach allows not only to obtain high purity polyfluorosalicylic acids in good yields, but also to produce them for biological studies. It should be noted that tetrafluoro-, 3,4-difluoro- and 4,5-difluorosalicylic acid[190] has a pronounced antimycobacterial activity against Mycobactérium tuberculosis H37Rv, M. avium, M. terrae and multidrug-resistant (MDR) strains, with a minimum inhibitory concentration (MIC) ranging from 0.7 to 1.5 mg ml –1. Moreover, tetrafluorosalicylic acid has a weak antifungal activity against pathogenic fungi of the genus Trichophiton (MIC = 25 – 50 mg ml–1). Tetrafluoro-, 3,4,5-trifluoro-, 3,5,6-trifluorosalicylic acids and 2-methoxy derivatives of tetrafluoro- and 3,4,5-trifluorosalicylic acids showed anti-inflammatory activity in the carrageenan-induced paw edema test in vivo comparable to that of aspirin[51]. The same compounds and 4,5-difluorosalicylic acid showed analgesic activity in the hot plate test. In vitro tests have shown that their mechanism of action as potential non-steroidal anti-inflammatory agents is due to inhibition of cyclooxygenase-1 (COX-1).

In studying the reactions of ethyl pentafluorobenzoate with sodium ethoxide at room temperature, it was found that the selectivity of the fluorine atom substitution can be adjusted by varying the ratio of ethanol to diethyl ether (Scheme 60)[191]. It should be noted that the reaction of pentafluorobenzoic acid methyl ester with sodium methoxide is less selective, since even in excess of methanol a 53 : 48 mixture of ortho- and para- methoxy-substituted isomers is formed[192].

Scheme 60

Methyl 4-methyl-2,3,5,6-tetrafluorobenzoate reacts with sodium methoxide to give a 6-methoxy derivative (Scheme 61)[193, 194].The authors report high yields of the ortho- substituted ester and deny the formation of by-products, but the patent[193] uses the preparative chromatography to purify the final product, which casts doubt on the above data. At the same time, it was found that when the methyl ester of 4-benzylamino-2,3,5,6-tetrafluorobenzoic acid was treated with sodium methoxide, even at room temperature, both ortho-fluorine atoms were readily substituted to give a dimethoxy ester (see Scheme 61)[195, 196].

Scheme 61

Benzyl pentafluorobenzoate can undergo double nucleophilic substitution of ortho- and para-fluorine atoms under the action of benzyl alcohol in THF in the presence of potassium tert-butoxide to give a 2,4-disubstituted ether but this process is not selective and is accompanied by the formation of a 2,4,6-trisubstituted product (Scheme 62)[81]. The same compounds can be obtained by the analogous reaction of the benzyl 4-benzyloxy-2,3,5,6-tetrafluorobenzoate[80]. Hydrogenolysis of benzyl 2,4-bis(benzyloxy)-3,5,6-trifluorobenzoate gives 4,6-dihydroxy-2,3,5-trifluorobenzoic acid[80, 81].

Scheme 62

Under the action of sodium azide, the fluorine atom in the para position of pentafluorobenzoic acid esters is easily replaced by the azide group upon heating in aqueous acetone (Scheme 63)[197]. The introduction of the azide function is promising for subsequent modifications since it can be selectively reduced to the amino group with tin(II) chloride[198, 199] or copper(II) sulfate in the presence of sodium ascorbate[200] to give 4-amino-2,3,5,6-tetrafluorobenzoic acid or its methyl ester. 4-Azido-2,3,5,6-tetrafluorobenzoic acid[201] and its methyl ester[202] undergo a Staudinger reaction with triarylphosphines to give 4-[(triarylphosphoranylidene)amino] derivatives. For 4-azido-substituted tetrafluorobenzoates, azide-alkyne addition reactions have been studied, leading to the synthesis of a series of 4-(1,2,3-triazol-1-yl)-2,3,5,6-tetrafluorobenzoic acids[203-210]. In addition, transformations of methyl 4-azidotetrafluorobenzothiazole in [3+2] cycloaddition reactions with aldehydes[211, 212] and thioacetic acid (TAA)[213] to give amides, and with enamines[214, 215] to give amidines have been described (see Scheme 63).

Scheme 63

As shown above, esters of polyfluorosalicylic acids react with ammonia to give the corresponding amides (see Scheme 11). In this context, the synthesis of 4-aminopolyfluorosalicylic acids[88] is achieved through the preliminary preparation of 4-azidopolyfluorosalicylates by substitution of the fluorine atom at the activated 4-position with sodium azide. The reduction of the azide function was carried out under the action of zinc metal in the presence of NH4Cl in refluxing methanol to give methyl 4-amino-2-hydroxy-2,3,5-trifluoro(or 3,5-difluoro)-benzoates, the hydrolysis of which afforded the corresponding benzoic acids (Scheme 64). 4-Amino-3,5-difluorosalicylic acid, a fluorinated analogue of the well-known anti-tuberculosis drug PASA, has been shown to inhibit the growth of M. tuberculosis H37Rv, M. avium and M. terrae strains and to inhibit the growth of MDR strains (MIC = 1.5 mg ml–1), whereas PASA in the similar concentration only inhibits the growth of Mycobacterium tuberculosis H37Rv[88].

Scheme 64

(Poly)fluorobenzoic acids with different heterocyclic substituents have been obtained by SNArF reactions. Studying the reaction of polyfluorosalicylic acid esters with cyclic amines showed that the outcome of the transformation depends on the substrate structure, the temperature regime, the nature of the solvent and the amine moiety[53]. For example, reactions esters of tetrafluoro- and 3,4,5-trifluorosalicylic acids with morpholine or N-methylpiperazine in DMSO at room temperature lead to 4-substituted products, methyl 4-(morpholin-4-yl)polyfluoromethyl- and methyl 4-(N-methylpiperazin-1-yl)tetrafluorosalicylates (Scheme 65, path 1). When similar processes are carried out by heating in DMSO or acetonitrile in the presence of DIPEA, not only substitution of the fluorine atom at the 4-position occurs, but also hydrolysis of the ester functionality, affording N-cycloamino-substituted polyfluorosalicylic acids MTFA, MDFA, PTFA, PDFA (path 2). Reactions of polyfluorosalicylic acid esters with morpholine, carried out at a lower temperature (50 °C), provided the selective preparation of 4-(morpholin-1-yl)polyfluorosalicylamides (path 3). 2,4,5-Trifluoro-3-(morpholin-4-yl)phenol was obtained by treating the corresponding acid or amide with 48% HBr solution.

Scheme 65

Methyl polyfluorosalicylates react with pyrrolidine on heating in DMSO or acetonitrile in the presence of DIPEA to give 4-(pyrrolidin-1-yl)polyfluorosalicylic acids[53] as a result of substitution of the fluorine atom at the 4-position and hydrolysis of the ester group (Scheme 66). In contrast to this transformation, the same esters only react with L-proline when heated with DIPEA to give methyl 4-(proline-1-yl)polyfluorosalicylates.

Scheme 66

Substitution to the morpholine moiety was also carried out for methyl 2-methoxy-3,4,5-trifluorobenzoate when the reaction was carried out in refluxing acetonitrile, giving methyl 2-methoxy-4-(morpholin-4-yl)-3,5-difluorobenzoate (Scheme 67)[53]. Methyl 2-methoxy-4,5-difluorobenzoate, treated with tert-butylpiperazine-1-carboxylate on heating in DMF in the presence of K2CO3, also readily undergoes substitution of the fluorine atom in the para position to give tert-butyl 4-(5-methoxy-4-methoxycarbonyl-2-fluorophenyl)piperazine1-carboxylate[216].

Scheme 67

Ethyl 4,5-difluorosalicylate reacts with morpholine in DMSO at room temperature with substitution of the fluorine atom in the 4-position to give ethyl 2-hydroxy-4-(morpholin-4-yl)-5-fluorobenzoate (Scheme 68, path 1), and on heating, the ester moiety is also involved in this reaction in addition to this process. As a result, [2-hydroxy-4-(morpholin-4-yl)-5-fluorophenyl](morpholin-1-yl)methanone (path 2) was isolated, the hydrolysis of which in 48% HBr solution leads to decarboxylation and the formation of 3-(morpholin-4-yl)-4-fluorophenol[53]. When heated with N-methylpiperazin in DMSO, this ester forms 2-hydroxy4-(N-methylpiperazin-1-yl)-3-fluorobenzoic acid (PMFA) (path 3).

Scheme 68

The reaction of methyl 3,4-difluorosalicylate with morpholine in DMSO at room temperature led to the corresponding unsubstituted amide (Scheme 69, path 1), and after heating a mixture of this amide with 4-(morpholin-4-yl)-substituted amide in a ratio of 1 : 0.4 was isolated (path 2)[53]. When methyl 3,4-difluorosalicylate reacts with N-methylpiperazine in acetonitrile under heating, the main process is hydrolysis of the ester functionality to 3,4-difluorosalicylic acid, and the content of the substitution product in the reaction mixture is as low as 10% (path 3).

Scheme 69

The reaction of ethyl 2-methoxy-3,5-difluorobenzoate with morpholine when heated in DMSO proceeded with low substrate conversion, and the yields of ethyl 2-methoxy-3-(morpholin-4-yl)-5-fluorobenzoate and 3-fluorobenzoate were only 1.6 and 3.9%, respectively (Scheme 70)[217].

Scheme 70

From the studies presented in this Section, it can be concluded that the presence of a fluorine atom in the para position of the benzene ring with respect to the ester moiety is necessary for efficient nucleophilic substitution in polyfluorosalicylates.

The study of the biological properties of cycloalkylamine derivatives has shown that the introduction of heterocyclic amines into polyfluorosalicylic acids is an effective way of reducing their toxicity. At the same time, N-piperazine derivatives PTFA, PDFA (see Scheme 65), PMFA (see Scheme 68) have significant analgesic activity, and their morpholinecontaining analogues MTFA, MDFA (see Scheme 65) combine moderate anti-inflammatory and analgesic properties. In addition, MTFA and PTFA acids have shown the ability to inhibit the growth of M. tuberculosis H37Rv strain at MIC = 6.25 mg ml–1.

The substitution of the ortho-bromine atom in 2-bromopolyfluorobenzoic acids under the action of 2-methoxy-4-(trifluoromethoxy)phenols in the presence of a base (Cs2CO3) and a catalyst (CuI) on heating in toluene at 100 °C in a sealed ampoule has been described (Scheme 71)[218]. This gives o-phenyl-substituted acids. Difluoro-2-(3,5-dichlorophenoxy) benzoic acids were obtained in a similar way but using Cu(MeCN)4PF6 as the catalyst. Such acids have herbicidal activity against a number of weeds such as Alopecurus myosuroides, Avena fatua, Echinochloa crus-galli, Setaria faberi, Helianthus annuus and Ipomoea hederacea[219].

Scheme 71

Considering the data presented in this Section, it can be concluded that SNArF reactions are a convenient tool for the synthesis of polyfluorobenzoic acids modified on the aromatic ring by various O- and N-containing units.

3.4. Photocatalytic alkylation via the C – F bond activation

Recently, photocatalytic alkylation of polyfluoroarenes due to activation of the C – F bond have been widely studied[216], but similar transformations with polyfluorobenzoic acid derivatives are poorly described in the literature. For example, polyfluorobenzoic acid esters react with N-methylaniline under photocatalytic conditions in the presence of Ir(ppy)2(dtbbpy)PF6 (dtbbpy is 4,4'-di-tert-butyl-2,2'-dipyridyl) to give 4-(anilinomethyl)polyfluorobenzoic acid esters (Scheme 72)[220].

Scheme 72

Esters and amides of polyfluorobenzoic acids were found to react with alcohols at the para-position of the aryl ring to afford 4-substituted polyfluorobenzoates and polyfluorobenzamides (Scheme 73). This process is also carried out under photocatalytic conditions under visible light irradiation (blue LEDs) in the presence of the catalyst 4-CzIPN[221].

Scheme 73

Methyl pentafluorobenzoate reacts with alkenes under photocatalytic conditions in the presence of fac-Ir(ppy)3 to give methyl esters of 4-alkyl-substituted tetrafluorobenzoic acids (Scheme 74a), with the methyl esters of 4-substituted 3,5,6-trifluorobenzoic acids characterized by substitution at the ortho-fluorine atom (see Scheme 74b)[222].

Scheme 74

4. Synthesis of heterocyclic compounds

The ability of polyfluorobenzoic acids and their derivatives to undergo SNArF reactions is often exploited to construct heterocyclic cores, involving the carboxyl function and the adjacent ortho-fluorine atom in such transformations.

Polyfluorobenzoyl chlorides, which can be cyclized under the action of various dinucleophiles to form a diversity of heterocyclic structures, have proved to be highly reactive bifunctional building blocks[28, 223, 224]. It has been shown that cyclization of polyfluorobenzoyl chlorides with N,N-dinucleophiles such as alkylisothiourea[223, 224] and diphenylsubstituted guanidine[223] provides fluorine-containing 2-alkylthio-2-iminoquinazolin-4-ones (Scheme 75).

Scheme 75

The direction of cyclization of polyfluorobenzoyl chlorides with 2-aminoazaheterocycles depends on the process conditions[28, 223, 224]. Thus, [b]-annulated quinazolinones are obtained in dichloromethane at room temperature in the presence of DIPEA, whereas their [a]-annulated analogues are formed under more harsh conditions, e.g., in refluxing acetonitrile, toluene or DMF using 1,8-diazobicyclo[5.4.0]undec-7-en (DBU) as the base (Scheme 76).

Scheme 76

Cyclization of polyfluorobenzoyl chlorides with C,N-dinucleophiles (2-cyanomethylpyridine, 2-cyanomethylquinoline, 2-cyanomethyl-2-benzoylmethylbenzimidazoles) produced pyrido-, quinolino- and benzimidazo[1,2-a]quinolones (Scheme 77).

Scheme 77

Using S,N-dinucleophiles in reactions with polyfluorobenzoyl chlorides opens the way to the preparation of benzannuated thiazines[223, 224]. In the case of thioamides, [1,3]benzothiazin-4ones are formed, while thioureas afford a mixture of [1,3]benzothiazin-4-ones and 2-thioquinazolin-4-ones. Polyfluorobenzoyl chlorides react with (benz)imidazol-2thiones or 1H-perimidine-2-thiones to give polyfluorinated (benz)imidazo[2,1-b][1,3]benzothiazin-4-ones or tetrafluorobenzo[5,6][1,3]thiazino[3,2-a]perimidin-13-ones (Scheme 78).

Scheme 78

The use of C,O-dinucleophiles in cyclization with polyfluorobenzoyl chlorides is a convenient approach to constructing polyfluorochromone scaffolds. Thus, polyfluorochromone-3-carboxamides have been obtained by the reaction of these substrates with acetoacetamides[223]. To involve 1,3-dicarbonyl compounds (acetoacetic acid ester[223, 225], acetylacetone[226-228], 1,3-diphenylpropane-1,3-dione[228], ethyl 3-aryl-3-oxopropionate[229], diethyl malonate[230] and others) in this process, their activation by magnesium alkylate is necessary. Such transformations result in polyfluorochromones and flavones functionalized at the 3-position by various carbonyl derivatives have been synthesized (Scheme 79).

Scheme 79

The reaction of polyfluorobenzoic acid chlorides with 2-hydroxyacetophenone gives 2-aroyloxyacetophenones, which further furnish 1,3-diaryl-1,3-diketones via the Baker-Venkataraman rearrangement. The acid- and base-catalyzed cyclization of fluorine-containing 1,3-diaryl-1,3-diketones has provided alternative approaches to flavones containing fluorine atoms in both the A and B rings (Scheme 80)[231]. Among the polyfluorinated 1,3-diaryl-1,3-diketones, compounds with promising antimicrobial properties against influenza virus type A (H1N1) and fungistatic activity against a number of dermatophytes Trichophyton tonsurans, T. interdigitale, Epidermophyton floccosum, Microsporum canis with MIC < 1.56 mg ml–1 were found. However, flavones with the fluorinesubstituted B-ring almost completely lose their antimycotic activity, whereas flavones bearing fluorine atoms in the A-ring inhibit the growth of two fungal strains, T. tonsurans and E. floccosum[231].

Scheme 80

Salicylic acids and their derivatives, including fluorinated analogues, can serve as reagents for the synthesis of various heterocyclic molecules due to the convenient ortho-arrangement of two functional groups (carboxyl and hydroxyl)[232-235]. First of all, polyfluorosalicylates are able to form oxaheterocycles. For example, polyfluorinated benzodioxin-4-ones[80, 81] have been obtained on the basis of pentafluorophenyl ester (see Scheme 7). Notably, this transformation is used to protect the ortho-hydroxy group, e.g. to carry out a selective o-alkylation at the para position of the benzene ring.

Alkylation of the ethyl ester of tetrafluorosalicylic acid with bromoacetate in acetone on heating gives ethyl 3,4,5,6-tetrafluoro-2-(2-ethoxy-2-oxoethoxy)benzoate, which is readily cyclized by the action of sodium hydride to 3-oxo-4,5,6,7-tetrafluoro-2-ethoxycarbonylbenzo[b]furan. Subsequent heating of the product in an alcoholic alkali solution leads to decarboxylation to give compound 27 (Scheme 81)[47, 48].

Scheme 81

A large series of works has been devoted to the cyclization of 6-amino derivatives of polyfluorosalicylic acid esters containing a protected hydroxyl group with aryl ketones promoted by sodium hydride on refluxing in 1,4-dioxane (or toluene). As a result of such reactions, a wide range of flavones 28 (> 250 examples) has been obtained in which the substituents in positions 2, 5 and 7 vary (Scheme 82)[236-241]. Interest in compounds of this group is due to their antibacterial and antitumour properties.

Scheme 82

An alternative approach to the flavone scaffold was proposed in a study[242]. It was shown that 3,5-difluorosalicylic acid can react with 1-[(methylsulfonyl)ethynyl]-4-methylbenzene with the assistance of the copper(II) acetate–hexafluorophosphate (benzotriazol-1-yloxy)tris(dimethylamino)phosphonium (BOP) catalytic system in the presence of DMAP as the base on heating in nitromethane to give 2-methyl-3-(p-tolyl-4-sulfonyl)-6,8-difluorochromen-4-one (29) (Scheme 83)

Scheme 83

The reaction of 2-methoxy-3,4,5,6-tetrafluorobenzoic acid chloride with acetoacetic ester in the presence of magnesium methoxide gives ethyl 2-(2-methoxy-3,4,5,6-tetrafluorobenzoyl)-3-oxobutanoate, which undergoes cyclization on heating to give 2-methyl-5-methoxy-6,7,8-trifluoro-3-ethoxycarbonyl-4H-1,4-dihydrobenzopyran-4-one, the cyclization taking place by intramolecular substitution of the fluorine atom by the hydroxyl group on the methyl substituent (Scheme 84)[243, 244].

Scheme 84

Cyclization of 1,3,5-trihydroxybenzene with 5,6-difluorosalicylic acid on heating in methanesulfonic acid in the presence of phosphoric anhydride provided an access to 6,8-dihydroxy-1,2-difluoro-9H-xanthen-9-one (30), which has anticancer activity (Scheme 85)[245].

Scheme 85

An example of the formation of an oxaheterocycle from difluorosalicylic acid with retention of the carboxyl group is also described. For this purpose, 2-methylallyl 2-(2-methylallyloxy)-5,6-difluorobenzoate, obtained from 2-(2-methylallyloxy)-5,6-difluorobenzoic acid by alkylation with 2-methyl-3-chloropropene, is subjected to a Claisen rearrangement on heating in NMP. This affords (2-methylallyl)-2-hydroxy-3-(2-methylallyl)-5,6-difluorobenzoate, which converts into 2,2-dimethyl-5,6-difluoro-3H-benzofuran-7-carboxylic acid (31) on refluxing in 96% formic acid (Scheme 86)[246, 247].

Scheme 86

The use of nitrogen-containing reagents in cyclizations with polyfluorosalicylic acid chlorides opens the way to the synthesis of azaheterocycles. Based on transformations of 2-methoxy-polyfluorobenzoic acid chlorides, analogues of known fluoroquinolone antibacterial agents have been synthesized (Scheme 87)[248-251]. First, the reaction of acid chlorides with monomalonic esters activated via a lithiation step gives ethyl 3-(2-methoxypolyfluorophenyl)-3-oxopropanoates. Their sequential treatment with triethyl orthoformate and ethylamine or cyclic amine gives aminomethylidene-substituted products, which on further cyclization and hydrolysis of affords quinolones 32. Substitution of the 7-fluorine atom in the latter compound with pyrrolidine or piperazine derivatives gives the target fluoroquinolones 33.

Scheme 87

The patent[252] teaches a similar process for the synthesis of fluorine-containing 7-amino-5-hydroxy-1-(oxetan-3-yl)-4-oxo-1,4-dihydroquinoline-3-carboxylic acids from 2-benzyloxypolyfluorobenzoic acid chlorides (Scheme 88).

Scheme 88

An approach to the synthesis of levofloxacin via the reaction of polyfluorobenzoic acid chlorides with ethyl (S,Z)-3-[(1-acetoxypropan-2-yl)amino]acrylate has been proposed[253-255]. It leads to the initial formation of an intermediate amide, whose further cyclizes under the action of N,O-bis(trimethylsilyl) acetamide (BSA) to give the quinolone 34. Subsequent hydrolysis and intramolecular substitution of the 8-fluorine atom in compound 34 gives (S)-3-methyl-7-oxo-9,10-difluoro-2,3-dihydro-7H-[1,4]oxazino[2,3,4-ij]quinoline-6-carboxylic acid (35) (Scheme 89), where substitution of the 10-fluorine atom with the N-methylpiperazine moiety furnishes levofloxacin.

Scheme 89

Patent[256] describes the carbonyl diimidazole (CDI)-mediated cyclization of 3,4-difluorosalicylic acid in a sealed ampoule to 6,7-difluorobenzo[c]isoxazole-3-(1H)-one in 36% yield (Scheme 90). A similar product was obtained in higher yield in the reaction between the methyl ester of 3,4-difluorosalicylic acid and aqueous hydroxylamine[52, 181, 257, 258] or the ethyl ester of the same acid with hydroxylamine hydrochloride[259] followed by treatment of the intermediates with CDI.

Scheme 90

The synthesis of 3-aryl-6,7-difluoro-2,3-dihydrobenzo[e][1,3]oxazin-4-ones via cyclization of N-aryl-2-hydroxy-4,5-difluorobenzamides obtained by amidation of 4,5-difluorosalicylic acid followed by the reaction of the intermediate amide with paraform in the presence of TsOH in refluxing toluene has been patented (Scheme 91)[178, 179, 260].

Scheme 91

The reaction of methyl ester of 3,6-difluorosalicylic acid with 5-nitro-2,4-dichloropyridine leads to methyl-2-[(5-nitro-2-chloro-4-pyridyl)oxy]-3,6-difluorobenzoate, which, after reduction of the nitro group in refluxing toluene, undergoes cyclization in the presence of TsOH to give 6,9-difluoro-3-chlorobenzo[f]pyrido[4,3-b][1,4]oxazepine-10-(11H)-one (Scheme 92)[261].

Scheme 92

Treatment of 2-{[(2R)-4-(tert-butoxycarbonyl)piperazin-2-yl]methoxy}-4-iodo-5,6-difluoro-3-chlorobenzoic acid with HATU (hexafluorophosphate-1-[bis(dimethylamino) methylene]-1H-1,2,3-triazolo[4,5-b]pyridinium-3-oxide) in the presence of a base (DIPEA) in DMF at room temperature induces cyclization to give tert-butyl-(12aR)-9-iodo-6-oxo-7,8-difluoro-10-chloro-3,4,12,12,12a-tetrahydro-6H-pyrazino[2,1-c][1,4]benzoxazepine-2-(1H)-carboxylate (Scheme 93)[262].

Scheme 93

The possibility of 3,5-difluorosalicylic acid to undergo heterocyclization through the carboxyl function alone, while retaining the ortho-hydroxyl substituent, has also been demonstrated. Such a reaction was carried out with o-aminophenol under microwave (MW) irradiation at 250 °C to yield 2-(benzoxazol-2-yl)-4,6-difluorophenol (Scheme 94)[263].

Scheme 94

There are examples of the synthesis of macrocyclic systems based on polyfluorosalicylic acids. Firstly, 3,5-difluoro-5,6-difluorosalicylic acids are alkylated with allyl bromide to give esters 36, which react with amine 37 in the presence of HATU to give amides 38. Promoted with the Grubbs catalyst (benzylidene-[1,3-bis(2,4,6-trimethylphenyl)-2-imidazolidinylidene]dichloro(tricyclohexylphosphino) ruthenium), the said amides are cyclized to macrocyclic derivatives 39 (Scheme 95). Macrocycles 39 in nanomolar concentrations are potent inhibitors of the cathepsin L cysteine protease, which may be promising for the treatment of a wide range of diseases such as diabetes, atherosclerosis, cancer, pancreatitis, etc[264].

Scheme 95

To obtain a macrocycle with a triple bond, Giroud et al.[265] proposed an alternative strategy (see Scheme 95). The sequence of transformations starts with the alkylation of tert-butyl 3,5-difluorosalicylate with 1,4-dibromobut-2-yne to give the compound 40, which is treated with phenol 41 and aminocyclopropane 42 to afford a diether 43. The latter was cyclized to macrocycle 44 under the action of HATU. In the same study, another type of activity of this compound was observed, namely a high inhibitory effect against the African trypanosomiasis pathogen Trypanosoma brucei (half-maximal inhibitory concentration (IC50) is 8 nM, selectivity index (SI) is 5400).

Based on difluorosalicylic acid esters, benzoxaborinines have been synthesized (Scheme 96)[266, 267], with an inhibitory affect towards β-lactamases. Thus, esters 45 are treated with alkene 46 to give an alkenyl derivative 47 which is then reacted with borane 48 to give compound 49. Subsequent cyclization and deprotection leads to 2-hydroxy-6,7-difluoro-3,4-dihydro-2H-benzo[e][1,2]oxaborinin-8-carboxylic acid 50[251].

Scheme 96

The same ester 45, when reacted with alkene 51 with subsequent bromination, gives a bromoalkenyl derivative 52, which reacts with diborane 53 to furnish the product 54 (see Scheme 96). Further transformations lead to the disodium salt of 2,2-dihydroxy-5,6-difluoro-1a,7b-dihydro-1H-cyclopropa[c][1,2]benzoxaborinin-4-carboxylic acid (55)[267].

Reactions of polyfluorosalicylic acid esters 56 with diborane 57 or borane 58 give derivatives 59 and/or 60 (Scheme 97). Ester 59, after elongation of the alkyl chain length is converted to compound 60, from which a chlorine-containing intermediate 61 is obtained by the Matteson reaction. Further reaction with 5-amino-1,3,4-thiadiazol-2-thiol 62 followed by cyclization and deprotection gives difluoro-containing 3-[(5-amino-1,3,4-thiadiazol-2-yl)thio]-2-hydroxy-3,4-dihydro-2H-benzo[e][1,2]oxaborinin-8-carboxylic acids 63[268]. The same reaction when using amide 64 in the final step gives 3-amido-2-hydroxy-3,4-dihydro-2H-benzo[e][1,2]oxaborinin-8-carboxylic acids 65 (see Scheme 97)[269-272].

Scheme 97

Patent[273] describes a process for the preparation of new tetracycline antibiotics based on 2-methoxy-4,5-difluorobenzoic acid (Scheme 98). First, ortho-methylation of this acid, esterification and addition of benzyl protection of the hydroxyl group were carried out to obtain phenyl 6-benzyloxy-2-methyl-3,4-difluorobenzoate (66). Next, intermediate 67 was synthesized from ester 66 by introducing an aminomethylidene residue at the 5-position. The subsequent reaction sequence involved the reaction of compound 67 with cyclic enone (CEN), introduction of an amine functionality, hydrolysis and deprotection, affording the target tetracycles 68.

Scheme 98

Summarizing the above results, it can be concluded that polyfluorobenzoic acids have a great potential for converting into heterocyclic structures, including biologically active ones.

5. Synthesis of metal complexes

Polyfluorobenzoic acids are actively used in coordination chemistry to form metal complexes (Figure 4). Fluorine atoms often make a significant contribution to intermolecular interactions, including πF – πF and π–πF stacking, formation of fluorine-hydrogen bonds, etc., imparting practically useful properties to the complexes. Complexes based on polyfluorobenzoates have been the subject of a number of reviews discussing, for example, the structure and photoluminescent properties of copper(I)[274] and lanthanide complexes[275, 276]. Complexes of polyfluorobenzoates with copper(II)[277-279], nickel(II)[280], cobalt(II)[281], uranium(VI)[282], cadmium(II) and zinc(II)[283] have been described. Polyfluorobenzoic acids also find application in the construction of metal-organic frameworks (MOFs)[284, 285]. Metal complexes based on polyfluorosalicylic acids are less well known than metal polyfluorobenzoates and will therefore be given special attention in this review.

Figure 4

When studying the properties of polyfluorosalicylic acids, it was found that the number of fluorine atoms has practically no effect on the value of their ionisation constants, since the pKa values measured for polyfluorosalicylic acids in a DMSO – water mixture (1 : 1) are in the range of 2.9 – 3.5. Interestingly, the effect of fluorine atoms on the acidity of the compounds was negligible, as the pKa of salicylic acid is 3.05[286]. For 2-methoxy-3,4,5-trifluorobenzoic acid, the pKa increases to 4.75, and for methyl 2-hydroxy-3,4,5-trifluorosalicylate, the pKa = 8.0.

Polyfluorosalicylic acids and esters, as well as their nonfluorinated analogues[287-290], readily undergo salt formation reactions with metal compounds. Thus, when these acids react with sodium hydroxide in acetonitrile at room temperature, sodium salts are formed (Scheme 99)[51]. Reactions of tetrafluoroand trifluorosalicylic acids with copper(II) chloride or acetates of copper(II), cobalt(II) or manganese(II) give the corresponding salts of the general formula [MII(L4–2H)(H2O)2] · nH2O, where L4 is polyfluorosalicylic acid[286, 291]. When manganese(II) chloride reacts with tetrafluorosalicylic acid, a complex [MnII(L4–2H)(H2O)Cl] is formed in which the ratio of metal cation to organic anion is also 1 : 1, but a chlorine anion remains in the coordination sphere of the metal. It should be noted that unlike polyfluorinated analogues, salicylic acid (Sal) reacts with copper(II) and cobalt(II) acetates to form complexes [MII(Sal – H)2][292], bearing two anions of deprotonated acid. In contrast to acids, methyl trifluorosalicylate binds CoII and CuII ions to form bis(salicylate) complexes [M(L5 – H)2]• nH2O, where L5 is methyl trifluorosalicylate (Scheme 100)[293].

Scheme 99

Scheme 100

Metal salicylates are convenient reagents for the preparation of heteroligand metal complexes[294-297], and their polyfluorosalicylate analogues have also been used for these purposes. For example, heating polyfluorosalicylate salts of copper(II) with triphenylphosphine or pyridine in ethanol, gave rise to polyfluorosalicylato-κ2O-bis(triphenylphosphates-κ2P) and (polyfluorosalicylato-κO-pyridines-κN) of copper(II) (Scheme 101)[286].

Scheme 101

The salts of polyfluorosalicylic acids and bipy, 5-aryl-2,2'-bipyridines or phen gave rise to a series of neutral heteroligand metal complexes, the structure of which depends on the nature of the nitrogen-containing ligand and the coordinating metal ion (Scheme 102)[286, 291, 298, 299].

Scheme 102

Copper(II) or cobalt(II) bis(methyltrifluorosalicylate) (L5) reacts with bipy or phen in refluxing methanol to give heteroligand complexes 69, 70 with the general formula [M(L5 – H)2bipy(or phen)] (Scheme 103)[293]. Similar reactions of cobalt(II) bis(methyltrifluorosalicylate) with bipy or phen in ethanol lead to transesterification of the methoxyl group of the salicylate ligand to the ethoxyl group yielding metal complexes 71. The reverse transesterification is also possible: refluxing complexes 71 in methanol gives the coordination compounds 70. However, the same direct reaction in refluxing ethanol is accompanied by hydrolysis of the ester moiety to afford complexes 72.

Scheme 103

Refluxing copper(II) complexes 69 in HFIP promotes hydrolysis of the ester group and loss of one of the trifluorosalicylate units, resulting in solvated metal complexes 73 (see Scheme 103). The latter can also be obtained by treating the complexes 72 with hexafluoroisopropyl alcohol. On the contrary, cobalt(II) complexes 70 form solvates 74 with retaining of the basic structure when heated in HFIP[300].

Metal complexes prepared from polyfluorosalicylic acids and 2,2’-bipyridine-type ligands have potent antimycotic activity against a wide range of pathogenic dermatophytes of the genus Trichophiton and yeast-like microorganisms Candida albicans, as well as antimicrobial activity against Neisseria gonorrhoeae, Staphylococcus aureus, St. aureus MRSA and their single-species biofilms[286, 291, 298, 299].

Therefore, the practical potential of metal complex compounds comprising polyfluorobenzoate anion for the design of modern materials and complexes based on polyfluorosalicylates for the development of new effective drugs is obvious.

6. Conclusion

Summarizing the literature data presented in this review, it is possible to note the significant progress that has been made recently in the chemistry of polyfluorobenzoic acids. The practical need for polyfluorobenzoate derivatives, including polyfluorosalicylates, is urgent, particularly in terms of creating biologically active compounds. These are mainly amides of such acids, as evidenced by numerous patents in this field. Therefore, we can confidently speak about the prospects of searching for new physiologically active substances based on polyfluorobenzoic or polyfluorosalicylic acid scaffold, which are intended for the development of modern pharmaceutical drugs. These include antibacterial, antiparasitic, antimycotic, antitumour and inhibitors of specific receptors and enzymes responsible for the development of a wide range of diseases. Not to mention the successful use of polyfluorobenzoates and salicylates as agrochemicals, particularly herbicides and fungicides.

In terms of chemical transformations, polyfluorobenzoic acids and their derivatives have a greater synthetic potential than their non-fluorinated analogues, since the reactions involve not only the carboxyl group to furnish ethers and esters, amides, diaryl and hetaryl derivatives by cross-coupling accompanied by decarboxylation, but also the fluoroaromatic core, with elimination of fluorine atoms. Nucleophilic aromatic substitution can be used to introduce various moieties, including pharmacophoric ones, into the aromatic ring. In addition, polyfluorobenzoates and salicylates have proven to be convenient building blocks for the construction of benzannulated heterocycles such as quinazoline, quinolone, benzoxazine, benzoxazole, benzofuran, chromone, flavone, etc.

The specificity of the reactivity of polyfluorobenzoates and salicylates lies in their complexing properties, which makes such reagents of interest to specialists in coordination and analytical chemistry. However, it should be noted that there are no studies on the design of new functional materials based on polyfluorosalicylates for the fabrication of magnetic and photosensitive devices, as well as sorption and catalytic systems.

It is therefore expected that the chemistry of polyfluorobenzoates (and especially polyfluorosalicylates) will continue to develop as these compounds combine the availability required for modern reagents, a wide range of chemical transformations and practical applications.

7. List of abbreviations

AChE — acetylcholinesterase,
B — base,
BINAP — 2,2'-bis(diphenylphosphino)-1,1'-binaphthyl,
bipy — 2,2'-bipyridine,
Boc — tert-butoxycarbonyl,
BOP — (benzotriazol-1-yloxy)tris(dimethylamino)phosphonium hexafluorophasphate,
BSA — N,O-bis(trimethylsilyl)acetamide,
CDI — carbonyl diimidazole,
CMPhos — 2-[2-(dicyclohexylphosphino)phenyl]-1-methyl-1H-indol,
COD — cycloocta-1,5-diene,
COX-1 — cyclooxygenase-1,
Cp — cyclopentadienyl,
Cy — cyclohexyl,
DABCO — 1,4-diazabicyclo[2.2.2]octane,
DBU — 1,8-diazobicyclo[5.4.0]undec-7-en,
DCC — 1,3-dicyclohexylcarbodiimide,
DDQ — 2,3-dichloro-5,6-dicyano-1,4-benzoquinone,
DIAD — diisopropylazodicarboxylate,
DIOP — (4R,5R)-(–)-4,5-bis(diphenylphosphanylmethyl)2,2-dimethyl-1,3-dioxolane,
DIPEA — diisopropylethylamine,
DMA — N,N-dimethylacetamide,
DMAP — 4-dimethylaminopyridine,
DME — dimethoxyethane,
DMFDMA — N,N-dimethylformamide dimethyl acetal,
DMI — 1,3-dimethylimidazolidin-2-one,
DPPF — 1,1'-(diphenylphosphino)ferrocene,
dtbbpy — 4,4'-di-tert-butyl-2,2'-dipyridyl,
EDC — 1-(3-dimethylaminopropyl)-3-ethylcarbodiimide,
FG — functional group,
HATU — 1-[bis(dimethylamino)methylene]-1H-1,2,3-triazolo[4,5-b]pyridinium 3-oxide hexafluorophosphate,
HFIP — hexafluoroisopropyl alcohol,
HOBt — N-hydroxybenzotriazole,
IC50 — half-maximal inhibitory concentration,
LDA — lithium diisopropylamide,
LHMDS — lithium bis(trimethylsilyl)amide,
MCM-41 — mesoporous material (Mobil Composition of Matter No 41),
MDR — multidrug-resistance,
MIC — minimum inhibitory concentration,
MOF — metal-organic frameworks,
MS — molecular sieves,
MW — microwave irradiation,
NMP — N-methyl-2-pyrrolidone,
Ns — 4-nitrobenzenesulfonyl,
OTf — triflate (trifluoromethanesulfonate),
PASA— p-amino salycilic acid,
PCC — pyridinium chlorochromate,
PG — protecting group,
phen — 1,10-phenanthroline,
Piv — pivaloyl,
ppy — 2-phenylpyridine,
Py — pyridine,
PyBOB — benzotriazole(pyrrolidino)phosphonium hexafluorophosphate,
rr — regioisomeric ratio,
Sаl — salicylic acid,
SI — selectivity index,
T — heating,
TAA — thioacetic acid,
TBAB — tetra-n-butylammonium bromide,
TFA — trifluoroacetic acid,
TFP — tri(2-furyl) phosphine,
TMEDA — N,N,N,N-tetramethylethylenediamine,
Tol — tolyl,
Trop — tropylium,
Ts — p-toluenesulfonyl (tosyl).

Acknowledgements

This review was financially supported by the Russian Science Foundation (Project No. 24-23-00062, https://rscf. ru/project/24-23-00062/).

References

1.
Benzoic acid and its derivatives as naturally occurring compounds in foods and as additives: Uses, exposure, and controversy
del Olmo A., Calzada J., Nuñez M.
Critical Reviews in Food Science and Nutrition, Taylor & Francis, 2015
2.
Antimicrobial Food Additives. Characteristics, Uses, Effects. (2nd Edn). (Berlin, Heidelberg: Springer)
E.Lück, M.Jager
1997
3.
Completion of the core β-oxidative pathway of benzoic acid biosynthesis in plants
Qualley A.V., Widhalm J.R., Adebesin F., Kish C.M., Dudareva N.
Proceedings of the National Academy of Sciences of the United States of America, Proceedings of the National Academy of Sciences (PNAS), 2012
4.
Variations on a theme: synthesis and modification of plant benzoic acids
Wildermuth M.C.
Current Opinion in Plant Biology, Elsevier, 2006
5.
Salicylic acid derivatives: synthesis, features and usage as therapeutic tools
Ekinci D., Şentürk M., Küfrevioğlu Ö.İ.
Expert Opinion on Therapeutic Patents, Taylor & Francis, 2011
6.
Para-aminosalicylic acid: the return of an old friend
Donald P.R., Diacon A.H.
The Lancet Infectious Diseases, Elsevier, 2015
7.
Systematic review: the use of mesalazine in inflammatory bowel disease
BERGMAN R., PARKES M.
Alimentary Pharmacology and Therapeutics, Wiley, 2006
8.
Systematic review: safety of mesalazine in ulcerative colitis
Sehgal P., Colombel J.-., Aboubakr A., Narula N.
Alimentary Pharmacology and Therapeutics, Wiley, 2018
9.
Martindale: The Complete Drug Reference. (36th Edn). (London, Chicago: Pharmaceutical Press)
S.C.Sweetman
2009
10.
Koo Y.M., Heo A.Y., Choi H.W.
Plant Pathology Journal, Korean Society of Plant Pathology, 2020
11.
Salicylic Acid, a Multifaceted Hormone to Combat Disease
Vlot A.C., Dempsey D.A., Klessig D.F.
Annual Review of Phytopathology, Annual Reviews, 2009
12.
Salicylic acid-induced accumulation of biochemical components associated with resistance against Xanthomonas oryzae pv. oryzae in rice
Le Thanh T., Thumanu K., Wongkaew S., Boonkerd N., Teaumroong N., Phansak P., Buensanteai N.
Journal of Plant Interactions, Taylor & Francis, 2017
14.
Salicylic Acid: Synthetic Strategies and Their Biological Activities
Rosheen, Sharma S., Utreja D.
ChemistrySelect, Wiley, 2023
15.
Recent Progress in Decarboxylative Oxidative Cross-Coupling for Biaryl Synthesis
Perry G.J., Larrosa I.
European Journal of Organic Chemistry, Wiley, 2017
16.
Decarboxylative Halogenation of Organic Compounds
Varenikov A., Shapiro E., Gandelman M.
Chemical Reviews, American Chemical Society (ACS), 2020
17.
Recent developments in decarboxylative cross-coupling reactions between carboxylic acids and N–H compounds
Arshadi S., Ebrahimiasl S., Hosseinian A., Monfared A., Vessally E.
RSC Advances, Royal Society of Chemistry (RSC), 2019
19.
Synthesis of Heterocycles from 2‐Acylbenzoic Acids
Liu Y., Li C., Yang G.
European Journal of Organic Chemistry, Wiley, 2023
20.
Recent advances in synthesis of isocoumarins: An overview
Gogoi N., Parhi R., Tripathi R.K., Pachuau L., Kaishap P.P.
Tetrahedron, Elsevier, 2024
21.
Oxidative coupling of benzoic acids with alkynes: Catalyst design and selectivity
Loginov D.A., Konoplev V.E.
Journal of Organometallic Chemistry, Elsevier, 2018
23.
Hydrogen bonding in tungsten (VI) salicylate free acids
Baroni T.E., Bembenek S., Heppert J.A., Hodel R.R., Laird B.B., Morton M.D., Barnes D.L., Takusagawa F.
Coordination Chemistry Reviews, Elsevier, 1998
24.
Analytical challenges and advantages of using flow-based methodologies for ammonia determination in estuarine and marine waters
Šraj L.O., Almeida M.I., Swearer S.E., Kolev S.D., McKelvie I.D.
TrAC - Trends in Analytical Chemistry, Elsevier, 2014
25.
The case for the use of unrefined natural reagents in analytical chemistry—A green chemical perspective
Grudpan K., Hartwell S.K., Lapanantnoppakhun S., McKelvie I.
Analytical Methods, Royal Society of Chemistry (RSC), 2010
26.
Development of analytical methods for ammonium determination in seawater over the last two decades
Zhu Y., Chen J., Yuan D., Yang Z., Shi X., Li H., Jin H., Ran L.
TrAC - Trends in Analytical Chemistry, Elsevier, 2019
27.
Functionalization of Fluorinated Molecules by Transition-Metal-Mediated C–F Bond Activation To Access Fluorinated Building Blocks
Ahrens T., Kohlmann J., Ahrens M., Braun T.
Chemical Reviews, American Chemical Society (ACS), 2014
28.
C−F Bond Activation in Organic Synthesis
Amii H., Uneyama K.
Chemical Reviews, American Chemical Society (ACS), 2009
30.
Organofluorine chemistry: promising growth areas and challenges
Politanskaya L.V., Selivanova G.A., Panteleeva E.V., Tretyakov E.V., Platonov V.E., Nikul’shin P.V., Vinogradov A.S., Zonov Y.V., Karpov V.M., Mezhenkova T.V., Vasilyev A.V., Koldobskii A.B., Shilova O.S., Morozova S.M., Burgart Y.V., et. al.
Russian Chemical Reviews, Autonomous Non-profit Organization Editorial Board of the journal Uspekhi Khimii, 2019
31.
Recent advances (1995–2005) in fluorinated pharmaceuticals based on natural products
Bégué J., Bonnet-Delpon D.
Journal of Fluorine Chemistry, Elsevier, 2006
32.
The Many Roles for Fluorine in Medicinal Chemistry
Hagmann W.K.
Journal of Medicinal Chemistry, American Chemical Society (ACS), 2008
33.
Chemical Aspects of Human and Environmental Overload with Fluorine
Han J., Kiss L., Mei H., Remete A.M., Ponikvar-Svet M., Sedgwick D.M., Roman R., Fustero S., Moriwaki H., Soloshonok V.A.
Chemical Reviews, American Chemical Society (ACS), 2021
34.
Next generation organofluorine containing blockbuster drugs
Han J., Remete A.M., Dobson L.S., Kiss L., Izawa K., Moriwaki H., Soloshonok V.A., O’Hagan D.
Journal of Fluorine Chemistry, Elsevier, 2020
35.
Fluorine in medicinal chemistry: A review of anti-cancer agents
Isanbor C., O’Hagan D.
Journal of Fluorine Chemistry, Elsevier, 2006
37.
Fluorine in health care: Organofluorine containing blockbuster drugs
38.
Fluorine in Pharmaceutical Industry: Fluorine-Containing Drugs Introduced to the Market in the Last Decade (2001–2011)
Wang J., Sánchez-Roselló M., Aceña J.L., del Pozo C., Sorochinsky A.E., Fustero S., Soloshonok V.A., Liu H.
Chemical Reviews, American Chemical Society (ACS), 2013
39.
Next Generation of Fluorine-Containing Pharmaceuticals, Compounds Currently in Phase II-III Clinical Trials of Major Pharmaceutical Companies: New Structural Trends and Therapeutic Areas.
Zhou Y., Wang J., Gu Z., Wang S., Zhu W., Aceña J.L., Soloshonok V.A., Izawa K., Liu H.
Chemical Reviews, American Chemical Society (ACS), 2016
40.
Aza analogs of 5-(p-fluorophenyl)salicylic acid
Walford G.L., Jones H., Shen T.
Journal of Medicinal Chemistry, American Chemical Society (ACS), 1971
44.
Fluorinated dendrimers
Caminade A.
Current Opinion in Colloid and Interface Science, Elsevier, 2003
45.
Synthesis, Characterization, and Electron-Transport Property of Perfluorinated Phenylene Dendrimers
Sakamoto Y., Suzuki T., Miura A., Fujikawa H., Tokito S., Taga Y.
Journal of the American Chemical Society, American Chemical Society (ACS), 2000
46.
Phenoxide-directed ortho lithiation
Posner G.H., Canella K.A.
Journal of the American Chemical Society, American Chemical Society (ACS), 1985
47.
Partially fluorinated heterocyclic compounds. Part III. The preparation of 4,5,6,7-tetrafluorobenzo[b]furan
Brooke G.M., Furniss B.S.
Journal of the Chemical Society C Organic, Royal Society of Chemistry (RSC), 1967
48.
Patent GB 1227352
1971
50.
An Improved Method for the Preparation OF 3,5-Difluorosalicylaldehyde and 3,5-Difluorosalicylic Acid#
Weidner-Wells M.A., Fraga-Spano S.A.
Synthetic Communications, Taylor & Francis, 1996
51.
Polyfluorinated salicylic acid derivatives as analogs of known drugs: Synthesis, molecular docking and biological evaluation.
Shchegol'kov E.V., Shchur I.V., Burgart Y.V., Saloutin V.I., Trefilova A.N., Ljushina G.A., Solodnikov S.Y., Markova L.N., Maslova V.V., Krasnykh O.P., Borisevich S.S., Khursan S.L.
Bioorganic and Medicinal Chemistry, Elsevier, 2017
52.
Patent WO 2010043893
2010
53.
Synthesis and Biological Activity of 4-Cycloaminopolyfluorosalicylic Acids
Shchur I.V., Shchegolkov E.V., Burgart Y.V., Triandafilova G.A., Maslova V.V., Solodnikov S.Y., Krasnykh O.P., Borisevich S.S., Khursan S.L., Saloutin V.I.
ChemistrySelect, Wiley, 2019
54.
Patent WO 2011121105
2011
55.
Patent WO 2019119145
2019
56.
Patent US 20210269454
2021
57.
Discovery of OICR12694: A Novel, Potent, Selective, and Orally Biovailable BCL6 BTB Inhibitor
Mamai A., Chau A.M., Wilson B.J., Watson I.D., Joseph B.B., Subramanian P.R., Morshed M.M., Morin J.A., Prakesh M.A., Lu T., Connolly P., Kuntz D.A., Pomroy N.C., Poda G., Nguyen K., et. al.
ACS Medicinal Chemistry Letters, American Chemical Society (ACS), 2023
58.
Design, synthesis and biological evaluation of salicylanilides as novel allosteric inhibitors of human pancreatic lipase
Zhao Y., Zhang M., Hou X., Han J., Qin X., Yang Y., Song Y., Liu Z., Zhang Y., Xu Z., Jia Q., Li Y., Chen K., Li B., Zhu W., et. al.
Bioorganic and Medicinal Chemistry, Elsevier, 2023
59.
Anti-parallel sheet structures of side-chain-free γ-, δ-, and ε-dipeptides stabilized by benzene–pentafluorobenzene stacking
Wang J., Xu J., Wang D., Wang H., Li Z., Zhang D.
CrystEngComm, Royal Society of Chemistry (RSC), 2014
60.
Design, synthesis and insecticidal activity of novel anthranilic diamides with benzyl sulfide scaffold
61.
Patent WO 2006120178
2006
62.
Patent WO 2018098270
2018
63.
Design, Synthesis, and Evaluation of Niclosamide Analogs as Therapeutic Agents for Enzalutamide-Resistant Prostate Cancer
Kang B., Mottamal M., Zhong Q., Bratton M., Zhang C., Guo S., Hossain A., Ma P., Zhang Q., Wang G., Payton-Stewart F.
Pharmaceuticals, Multidisciplinary Digital Publishing Institute (MDPI), 2023
64.
PTG-0861: A novel HDAC6-selective inhibitor as a therapeutic strategy in acute myeloid leukaemia
Gawel J.M., Shouksmith A.E., Raouf Y.S., Nawar N., Toutah K., Bukhari S., Manaswiyoungkul P., Olaoye O.O., Israelian J., Radu T.B., Cabral A.D., Sina D., Sedighi A., de Araujo E.D., Gunning P.T., et. al.
European Journal of Medicinal Chemistry, Elsevier, 2020
65.
Redox-responsive conformational alteration of aromatic amides bearing N-quinonyl system
Okamoto I., Takahashi Y., Sawamura M., Matsumura M., Masu H., Katagiri K., Azumaya I., Nishino M., Kohama Y., Morita N., Tamura O., Kagechika H., Tanatani A.
Tetrahedron, Elsevier, 2012
66.
Neuroactive steroids with perfluorobenzoyl group
Černý I., Buděšínský M., Pouzar V., Vyklický V., Krausová B., Vyklický L.
Steroids, Elsevier, 2012
67.
Synthesis and biological evaluation of fluorinated N -benzoyl and N -phenylacetoyl derivatives of 3-(4-aminophenyl)-coumarin-7- O -sulfamate as steroid sulfatase inhibitors
Daśko M., Przybyłowska M., Rachon J., Masłyk M., Kubiński K., Misiak M., Składanowski A., Demkowicz S.
European Journal of Medicinal Chemistry, Elsevier, 2017
68.
Tropylium-Catalyzed O–H Insertion Reactions of Diazoalkanes with Carboxylic Acids
Empel C., Nguyen T.V., Koenigs R.M.
Organic Letters, American Chemical Society (ACS), 2021
69.
Partially fluorinated heterocyclic compounds. Part VI. Some reactions of 4,5,6,7-tetrafluorobenzo[b]furan
Brooke G.M., Furniss B.S., Musgrave W.K.
Journal of the Chemical Society C Organic, Royal Society of Chemistry (RSC), 1968
70.
Patent WO 2017011725
2017
71.
Discovery of HTL6641, a dual orexin receptor antagonist with differentiated pharmacodynamic properties
Christopher J.A., Aves S.J., Brown J., Errey J.C., Klair S.S., Langmead C.J., Mace O.J., Mould R., Patel J.C., Tehan B.G., Zhukov A., Marshall F.H., Congreve M.
MedChemComm, Royal Society of Chemistry (RSC), 2015
72.
Novel fluorinated pyrrolomycins as potent anti-staphylococcal biofilm agents: Design, synthesis, pharmacokinetics and antibacterial activities
Yang Z., Liu Y., Ahn J., Qiao Z., Endres J.L., Gautam N., Huang Y., Li J., Zheng J., Alnouti Y., Bayles K.W., Li R.
European Journal of Medicinal Chemistry, Elsevier, 2016
73.
Patent WO 2013091773
2013
74.
Fluorine analogs of dicamba and tricamba herbicides; synthesis and their pesticidal activity
Huras B., Zakrzewski J., Żelechowski K., Kiełczewska A., Krawczyk M., Hupko J., Jaszczuk K.
Zeitschrift fur Naturforschung - Section B Journal of Chemical Sciences, Walter de Gruyter, 2021
75.
Some properties of polyfluorinated chromones
Prudchenko A.T., Vovdenko L.P., Barkhash V.A., Vorozhtsov N.N.
Chemistry of Heterocyclic Compounds, Springer Nature, 1968
77.
Design, synthesis, insecticidal activities, and molecular docking of novel pyridylpyrazolo carboxylate derivatives
Yang S., Peng H., Zhu J., Zhao C., Xu H.
Journal of Heterocyclic Chemistry, Wiley, 2022
78.
Patent EP 0227415
1987
79.
Patent WO 2010011811
2010
81.
Polymer-assisted solution-phase library synthesis of 4-alkoxy-2-hydroxy-3,5,6-trifluorobenzoic acids
Hardcastle I.R., Moreno Barber A., Marriott J.H., Jarman M.
Tetrahedron Letters, Elsevier, 2001
82.
Patent WO 9740006
1997
83.
SYNTHESIS OF CERTAIN 2′-DEOXYURIDINE DERIVATIVES CONTAINING SUBSTITUTED PHENOXY GROUPS ATTACHED TO C-5′; EVALUATION AS POTENTIAL dUTP ANALOGUES
Marriott J.H., Aherne G.W., Hardcastle A., Jarman M.
Nucleosides, Nucleotides and Nucleic Acids, Taylor & Francis, 2001
84.
Use of Oligosalicylates in the Preparation of Phenolic Amido Acids
Gschneidner D., Corvino J., Freeman J., O'Toole D., Shields L., Wang E.
Synthetic Communications, Taylor & Francis, 2005
85.
Intramolecular nucleophilic substitution of fluorine in α-pentafluorophenyl-N-phenylnitrone
Petrenko N.I., Gerasimova T.N.
Russian Chemical Bulletin, Springer Nature, 1987
86.
2,3,4,5-Tetrafluoro-N-(4-fluorophenyl)-6-hydroxybenzamide: an Example of a Combined Inter- and Intramolecular O.H.O Bifurcated Hydrogen Bond
Banks R.E., DuBoisson R.A., Pritchard R.G., Tipping A.E.
Acta Crystallographica Section C Crystal Structure Communications, International Union of Crystallography (IUCr), 1995
87.
Patent WO 2007062370
2007
88.
Synthesis and biological activity of polyfluorinated p-aminosalicylic acids and their amides
Burgart Y.V., Shchur I.V., Shchegolkov E.V., Saloutin V.I.
Mendeleev Communications, Elsevier, 2020
89.
Esters of polyfluorosalicylic acids in reactions with amines
Shchur I.V., Shchegolkov E.V., Burgart Y.V., Saloutin V.I.
AIP Conference Proceedings, American Institute of Physics (AIP), 2022
90.
N-Substituted salicylamides as selective malaria parasite dihydroorotate dehydrogenase inhibitors
Fritzson I., Bedingfield P.T., Sundin A.P., McConkey G., Nilsson U.J.
MedChemComm, Royal Society of Chemistry (RSC), 2011
91.
Conjugates of Tacrine and Salicylic Acid Derivatives as New Promising Multitarget Agents for Alzheimer’s Disease
Makhaeva G.F., Kovaleva N.V., Rudakova E.V., Boltneva N.P., Grishchenko M.V., Lushchekina S.V., Astakhova T.Y., Serebryakova O.G., Timokhina E.N., Zhilina E.F., Shchegolkov E.V., Ulitko M.V., Radchenko E.V., Palyulin V.A., Burgart Y.V., et. al.
International Journal of Molecular Sciences, Multidisciplinary Digital Publishing Institute (MDPI), 2023
92.
Synthesis, Structural Characterization, and Antiangiogenic Activity of Polyfluorinated Benzamides
Steinebach C., Ambrożak A., Dosa S., Beedie S.L., Strope J.D., Schnakenburg G., Figg W.D., Gütschow M.
ChemMedChem, Wiley, 2018
93.
Anticancer Properties of a Novel Class of Tetrafluorinated Thalidomide Analogues
Beedie S.L., Peer C.J., Pisle S., Gardner E.R., Mahony C., Barnett S., Ambrozak A., Gütschow M., Chau C.H., Vargesson N., Figg W.D.
Molecular Cancer Therapeutics, American Association for Cancer Research (AACR), 2015
94.
Synthesis and Antiangiogenic Properties of Tetrafluorophthalimido and Tetrafluorobenzamido Barbituric Acids
Ambrożak A., Steinebach C., Gardner E.R., Beedie S.L., Schnakenburg G., Figg W.D., Gütschow M.
ChemMedChem, Wiley, 2016
95.
Patent WO 2010007552
2010
96.
Patent WO 2010007561
2010
97.
Patent US 20120046357
2012
98.
Patent WO 2014180562
2014
99.
Patent WO 2019222349
2019
100.
Patent JP 2011219368
2011
101.
Fragment-Based Discovery of Novel Potent Sepiapterin Reductase Inhibitors
Alen J., Schade M., Wagener M., Christian F., Nordhoff S., Merla B., Dunkern T.R., Bahrenberg G., Ratcliffe P.
Journal of Medicinal Chemistry, American Chemical Society (ACS), 2019
102.
Patent WO 2007105751
2007
103.
Pharmacophore optimization of imidazole chalcones to modulate microtubule dynamics
Karaj E., Dlamini S., Koranne R., Sindi S.H., Perera L., Taylor W.R., Viranga Tillekeratne L.M.
Bioorganic Chemistry, Elsevier, 2022
104.
Carbonic anhydrase inhibitors. Design of anticonvulsant sulfonamides incorporating indane moieties.
Chazalette C., Masereel B., Rolin S., Thiry A., Scozzafava A., Innocenti A., Supuran C.T.
Bioorganic and Medicinal Chemistry Letters, Elsevier, 2004
105.
Discovery of 2,5-disubstituted furan derivatives featuring a benzamide motif for overcoming P-glycoprotein mediated multidrug resistance in MCF-7/ADR cell
Yang Z., Cai Y., Mao S., Wu Q., Zhu M., Cao X., Wei B., Tian J., Bao X., Ye X., Chen J., Wang S., Yu Y., Zhang H., Sun X., et. al.
European Journal of Medicinal Chemistry, Elsevier, 2023
106.
Patent WO 2001000566
2001
107.
Patent WO 2019014352
2019
108.
Patent WO 2018020357
2018
109.
A new synthetic route to polyfluorobenzyl alcohol
Zhang D., Chen Z., Cai H., Zou X.
Journal of Fluorine Chemistry, Elsevier, 2009
110.
A Safer Reduction of Carboxylic Acids with Titanium Catalysis
Ramachandran P.V., Alawaed A.A., Hamann H.J.
Organic Letters, American Chemical Society (ACS), 2022
111.
In situ silane activation enables catalytic reduction of carboxylic acids
Stoll E.L., Barber T., Hirst D.J., Denton R.M.
Chemical Communications, Royal Society of Chemistry (RSC), 2022
112.
Patent WO 2007053082
2007
113.
Ligand-Enabled C–H Hydroxylation with Aqueous H2O2 at Room Temperature
Li Z., Park H.S., Qiao J.X., Yeung K., Yu J.
Journal of the American Chemical Society, American Chemical Society (ACS), 2022
114.
Patent WO2008077009
2008
117.
Dokl. Akad. Nauk SSSR, 179, 356
A.K.Petrov, B.V.Makarov, G.G.Yakobson
1968
118.
Polyfluorinated heterocyclic compounds
Petrova T.D., Kann L.I., Barkhash V.A., Yakobson G.G.
Chemistry of Heterocyclic Compounds, Springer Nature, 1972
119.
The selective ortho-methoxylation of pentafluorobenzoic acid – a new way to tetrafluorosalicylic acid and its derivatives
Bazyl' I.T., Kisil' S.P., Burgart Y.V., Saloutin V.I., Chupakhin O.N.
Journal of Fluorine Chemistry, Elsevier, 1999
120.
Pd-Catalyzed Decarboxylative Cross Coupling of Potassium Polyfluorobenzoates with Aryl Bromides, Chlorides, and Triflates
Shang R., Xu Q., Jiang Y., Wang Y., Liu L.
Organic Letters, American Chemical Society (ACS), 2010
121.
Copper-Catalyzed Decarboxylative Cross-Coupling of Potassium Polyfluorobenzoates with Aryl Iodides and Bromides
Shang R., Fu Y., Wang Y., Xu Q., Yu H., Liu L.
Angewandte Chemie - International Edition, Wiley, 2009
123.
Nickel-Catalyzed Decarboxylative Cross-Coupling of Perfluorobenzoates with Aryl Halides and Sulfonates
Sardzinski L.W., Wertjes W.C., Schnaith A.M., Kalyani D.
Organic Letters, American Chemical Society (ACS), 2015
124.
Ni-Catalyzed Decarboxylative Cross-Coupling of Potassium Polyfluorobenzoates with Unactivated Phenol and Phenylmethanol Derivatives
Chen Q., Wu A., Qin S., Zeng M., Le Z., Yan Z., Zhang H.
Advanced Synthesis and Catalysis, Wiley, 2018
125.
Mild Pd-Catalyzed Decarboxylative Cross-Coupling of Zinc(II) Polyfluorobenzoates with Aryl Bromides and Nonaflates: Access to Polyfluorinated Biaryls
Wang J., Cui Y., Xie S., Zhang J., Hu D., Shuai S., Zhang C., Ren H.
Organic Letters, American Chemical Society (ACS), 2023
126.
Pd-catalysed decarboxylative Suzuki reactions and orthogonal Cu-based O-arylation of aromatic carboxylic acids
Dai J., Liu J., Luo D., Liu L.
Chemical Communications, Royal Society of Chemistry (RSC), 2011
127.
Pd-Catalyzed Decarboxylative Arylation of Thiazole, Benzoxazole, and Polyfluorobenzene with Substituted Benzoic Acids
Xie K., Yang Z., Zhou X., Li X., Wang S., Tan Z., An X., Guo C.
Organic Letters, American Chemical Society (ACS), 2010
129.
Palladium‐Catalyzed Decarboxylative C–H Bond Arylation of Furans
Pei K., Jie X., Zhao H., Su W.
European Journal of Organic Chemistry, Wiley, 2014
130.
Palladium-Catalyzed Decarboxylative CH Bond Arylation of Thiophenes
Hu P., Zhang M., Jie X., Su W.
Angewandte Chemie - International Edition, Wiley, 2011
131.
A Versatile Catalyst for Intermolecular Direct Arylation of Indoles with Benzoic Acids as Arylating Reagents
Zhou J., Hu P., Zhang M., Huang S., Wang M., Su W.
Chemistry - A European Journal, Wiley, 2010
132.
Selective direct C–H polyfluoroarylation of electron-deficient N-heterocyclic compounds
Li S., Li W., Yang X., Sun R., Tang J., Zheng X., Yuan M., Li R., Chen H., Fu H.
Organic Chemistry Frontiers, Royal Society of Chemistry (RSC), 2020
133.
Nickel-Catalyzed Decarboxylative Arylation of Heteroarenes through sp2C-H Functionalization
Yang K., Wang P., Zhang C., Kadi A.A., Fun H., Zhang Y., Lu H.
European Journal of Organic Chemistry, Wiley, 2014
134.
Nickel-catalyzed decarboxylative arylation of azoles with perfluoro- and nitrobenzoates
Crawford J.M., Shelton K.E., Reeves E.K., Sadarananda B.K., Kalyani D.
Organic Chemistry Frontiers, Royal Society of Chemistry (RSC), 2015
135.
Synthesis of Biaryls by Pd-Catalyzed Decarboxylative Homo- and Heterocoupling of Substituted Benzoic Acids
Xie K., Wang S., Yang Z., Liu J., Wang A., Li X., Tan Z., Guo C., Deng W.
European Journal of Organic Chemistry, Wiley, 2011
137.
Substrate-Dependent Mechanistic Divergence in Decarboxylative Heck Reaction at Room Temperature
Hossian A., Bhunia S.K., Jana R.
Journal of Organic Chemistry, American Chemical Society (ACS), 2016
139.
Pd(ii)-catalyzed decarboxylative allylation and Heck-coupling of arene carboxylates with allylic halides and esters
Wang J., Cui Z., Zhang Y., Li H., Wu L., Liu Z.
Organic and Biomolecular Chemistry, Royal Society of Chemistry (RSC), 2011
140.
Catalytic Addition of Fluorinated Benzoic Acids to Butadiene
Maji T., Maliszewski M.L., Smith M.K., Tunge J.A.
ChemCatChem, Wiley, 2016
141.
Aerobic copper-catalyzed decarboxylative thiolation
Li M., Hoover J.M.
Chemical Communications, Royal Society of Chemistry (RSC), 2016
142.
Decarboxylative Cross-Coupling of Acyl Fluorides with Potassium Perfluorobenzoates
Fu L., Chen Q., Nishihara Y.
Organic Letters, American Chemical Society (ACS), 2020
143.
A Novel Mode of Reactivity for Gold(I): The Decarboxylative Activation of (Hetero)Aromatic Carboxylic Acids
Cornella J., Rosillo-Lopez M., Larrosa I.
Advanced Synthesis and Catalysis, Wiley, 2011
144.
Transition-metal-free decarboxylative bromination of aromatic carboxylic acids
Quibell J., Perry G.J., Cannas D.M., Larrosa I.
Chemical Science, Royal Society of Chemistry (RSC), 2018
145.
Transition-Metal-Free Decarboxylative Iodination: New Routes for Decarboxylative Oxidative Cross-Couplings
Perry G.J., Quibell J.M., Panigrahi A., Larrosa I.
Journal of the American Chemical Society, American Chemical Society (ACS), 2017
146.
Dual ligands relay-promoted transformation of unstrained ketones to polyfluoroarenes and nitriles
Wang Z., Xu H., Zhang X., Wang X., Xu H., Gao H., Dai H.
Science China Chemistry, Springer Nature, 2023
147.
Recent developments in the use of aza-Heck cyclizations for the synthesis of chiral N-heterocycles
Race N.J., Hazelden I.R., Faulkner A., Bower J.F.
Chemical Science, Royal Society of Chemistry (RSC), 2017
148.
Heck‐Like Reactions Involving Heteroatomic Electrophiles
Vulovic B., Watson D.A.
European Journal of Organic Chemistry, Wiley, 2017
149.
Recent Advances in 1,2‐Amino(hetero)arylation of Alkenes
Kwon Y., Wang Q.
Chemistry - An Asian Journal, Wiley, 2022
150.
Activation of N–O σ Bonds with Transition Metals: A Versatile Platform for Organic Synthesis and C–N Bonds Formation
Todorović U., Romero R.M., Anthore-Dalion L.
European Journal of Organic Chemistry, Wiley, 2023
151.
Transition-Metal-Mediated and -Catalyzed C−F Bond Activation by Fluorine Elimination
Fujita T., Fuchibe K., Ichikawa J.
Angewandte Chemie - International Edition, Wiley, 2018
152.
DiverseN-Heterocyclic Ring Systems via Aza-Heck Cyclizations ofN-(Pentafluorobenzoyloxy)sulfonamides
Hazelden I.R., Ma X., Langer T., Bower J.F.
Angewandte Chemie - International Edition, Wiley, 2016
153.
An Umpolung Approach to Alkene Carboamination: Palladium Catalyzed 1,2-Amino-Acylation, -Carboxylation, -Arylation, -Vinylation, and -Alkynylation
Faulkner A., Scott J.S., Bower J.F.
Journal of the American Chemical Society, American Chemical Society (ACS), 2015
154.
O-Acyl oximes: versatile building blocks for N-heterocycle formation in recent transition metal catalysis.
Huang H., Cai J., Deng G.
Organic and Biomolecular Chemistry, Royal Society of Chemistry (RSC), 2016
155.
Cross-Coupling of Heteroatomic Electrophiles.
Korch K.M., Watson D.A.
Chemical Reviews, American Chemical Society (ACS), 2019
156.
Radical reactions enabled by polyfluoroaryl fragments: photocatalysis and beyond
Zubkov M.O., Dilman A.D.
Chemical Society Reviews, Royal Society of Chemistry (RSC), 2024
158.
Recent Developments in Radical Cross‐Coupling of Redox‐Active Cycloketone Oximes
Xiao F., Guo Y., Zeng Y.
Advanced Synthesis and Catalysis, Wiley, 2020
159.
Iron and cobalt catalysis: new perspectives in synthetic radical chemistry
Kyne S.H., Lefèvre G., Ollivier C., Petit M., Ramis Cladera V., Fensterbank L.
Chemical Society Reviews, Royal Society of Chemistry (RSC), 2020
160.
Reactivity of (poly)fluorobenzamides in palladium-catalysed direct arylations
Laidaoui N., He M., El Abed D., Soulé J., Doucet H.
RSC Advances, Royal Society of Chemistry (RSC), 2016
162.
Pd(II)-catalysed meta-C–H functionalizations of benzoic acid derivatives
Li S., Cai L., Ji H., Yang L., Li G.
Nature Communications, Springer Nature, 2016
164.
Fragmentation of Radical Anions of Polyfluorinated Benzoates
Konovalov V.V., Laev S.S., Beregovaya I.V., Shchegoleva L.N., Shteingarts V.D., Tsvetkov Y.D., Bilkis I.
Journal of Physical Chemistry A, American Chemical Society (ACS), 1999
165.
Reductive defluorination of polyfluoroarenes by zinc in aqueous ammonia
Laev S.S., Shteingarts V.D.
Tetrahedron Letters, Elsevier, 1997
166.
Catalyst-Free Hydrodefluorination of Perfluoroarenes with NaBH4
Schoch T.D., Mondal M., Weaver J.D.
Organic Letters, American Chemical Society (ACS), 2021
167.
Photocatalytic Hydrodefluorination: Facile Access to Partially Fluorinated Aromatics
Senaweera S.M., Singh A., Weaver J.D.
Journal of the American Chemical Society, American Chemical Society (ACS), 2014
170.
Dual C–F, C–H Functionalization via Photocatalysis: Access to Multifluorinated Biaryls
Senaweera S., Weaver J.D.
Journal of the American Chemical Society, American Chemical Society (ACS), 2016
171.
1178. Aromatic polyfluoro-compounds. Part XXIX. Nucleophilic replacement reactions of pentafluorobenzoic acid
Burdon J., Hollyhead W.B., Tatlow J.C.
Journal of the Chemical Society (Resumed), Royal Society of Chemistry (RSC), 1965
172.
Synthesis and reactions of 4-sulpho-2,3,5,6,-tetrafluorobenzoic acid
Fielding H.C., Shirley I.M.
Journal of Fluorine Chemistry, Elsevier, 1992
173.
Synthesis of β-functionalized ethyl polyfluoroaryl sulfides, sulfoxides, and sulfones underlain by pentafluorobenzoic acid
Litvak V.V., Kondrat’ev A.S., Shteimgarts V.D.
Russian Journal of Organic Chemistry, Pleiades Publishing, 2009
174.
Nucleophilic substitution of pentafluorobenzes with imidazole
Fujii S., Maki Y., Kimoto H.
Journal of Fluorine Chemistry, Elsevier, 1989
175.
Patent JP 09077716
1997
176.
Patent WO 9817620
1998
178.
Patent WO 2008036843
2008
179.
Patent US 7834023
2010
180.
Patent WO 2019153080
2019
181.
Design, synthesis and anticancer activity of functionalized spiro-quinolines with barbituric and thiobarbituric acids
Bhaskarachar R.K., Revanasiddappa V.G., Hegde S., Balakrishna J.P., Reddy S.Y.
Medicinal Chemistry Research, Springer Nature, 2015
182.
Patent WO 2019051265
2019
183.
3,4-Difluoro-2-hydroxybenzoic acid
Ravi Kiran B., Palakshamurthy B.S., Vijayakumar G.R., Bharath H.S.
Acta Crystallographica Section E Structure Reports Online, International Union of Crystallography (IUCr), 2014
184.
Patent US 20040116293
2004
185.
Patent US 20050050587
2005
186.
Salicylate Activity. 3. Structure Relationship to Systemic Acquired Resistance
Silverman F.P., Petracek P.D., Heiman D.F., Fledderman C.M., Warrior P.
Journal of Agricultural and Food Chemistry, American Chemical Society (ACS), 2005
187.
Salicylate Activity. 2. Potentiation of Atrazine
Silverman F.P., Petracek P.D., Heiman D.F., Ju Z., Fledderman C.M., Warrior P.
Journal of Agricultural and Food Chemistry, American Chemical Society (ACS), 2005
188.
Salicylate Activity. 1. Protection of Plants from Paraquat Injury
Silverman F.P., Petracek P.D., Fledderman C.M., Ju Z., Heiman D.F., Warrior P.
Journal of Agricultural and Food Chemistry, American Chemical Society (ACS), 2005
189.
Zh. Org. Khim., 69, 613
I.T.Bazyl, S.P.Kisil’, Y.V.Burgart, M.I.Kodess, A.G.Gein, V.I.Saloutin
1999
190.
A convenient and efficient approach to polyfluorosalicylic acids and their tuberculostatic activity
Shchegol’kov E.V., Shchur I.V., Burgart Y.V., Saloutin V.I., Solodnikov S.Y., Krasnykh O.P., Kravchenko M.A.
Bioorganic and Medicinal Chemistry Letters, Elsevier, 2016
191.
Zh. Obshch. Khim., 6, 512
L.S.Kobrina, G.G.Furin, G.G.Yakobson
1970
192.
Izv. SO AN SSSR. Ser. Khim. Nauk, 93
L.S.Kobrina, G.G.Furin, V.F.Kollegov, V.S.Chertok, G.G.Yakobson
1972
193.
Patent EP 0266947
1988
194.
Patent US 2019023440
2019
195.
Patent WO 2006038112
2006
196.
Patent WO 2021113462
2021
197.
New reagents for photoaffinity labeling: synthesis and photolysis of functionalized perfluorophenyl azides
Keana J.F., Cai S.X.
Journal of Organic Chemistry, American Chemical Society (ACS), 1990
199.
Extending the series of p-substituted tetrafluorobenzoic acids: synthesis, properties and structure
Zaitsev K.V., Oprunenko Y.F., Lermontova E.K., Churakov A.V.
Journal of Fluorine Chemistry, Elsevier, 2017
200.
Synthesis of bitriazolyl nucleosides and unexpectedly different reactivity of azidotriazole nucleoside isomers in the Huisgen reaction
Xia Y., Li W., Qu F., Fan Z., Liu X., Berro C., Rauzy E., Peng L.
Organic and Biomolecular Chemistry, Royal Society of Chemistry (RSC), 2007
201.
Tetrafluorination of Aromatic Azide Yields a Highly Efficient Staudinger Reaction: Kinetics and Biolabeling
Xie Y., Cheng L., Gao Y., Cai X., Yang X., Yi L., Xi Z.
Chemistry - An Asian Journal, Wiley, 2018
202.
Perfluoroaryl Azide Staudinger Reaction: A Fast and Bioorthogonal Reaction
Sundhoro M., Jeon S., Park J., Ramström O., Yan M.
Angewandte Chemie - International Edition, Wiley, 2017
204.
Building a Cocrystal by Using Supramolecular Synthons for Pressure‐Accelerated Heteromolecular Azide–Alkyne Cycloaddition
Meng X., Chen C., Deng X., Wang Z., Chen Q., Ma Y.
Chemistry - A European Journal, Wiley, 2019
205.
Polyurethane (PU)-derived photoactive and copper-free clickable surface based on perfluorophenyl azide (PFPA) chemistry
Li L., Li J., Kulkarni A., Liu S.
Journal of Materials Chemistry B, Royal Society of Chemistry (RSC), 2013
206.
Substituted 4-(1H-1,2,3-triazol-1-yl)-tetrafluorobenzoates: Selective synthesis and structure
Solodukhin N.N., Borisova N.E., Churakov A.V., Zaitsev K.V.
Journal of Fluorine Chemistry, Elsevier, 2016
207.
Cooperative Binding in a Phosphine Oxide-Based Halogen Bonded Dimer Drives Supramolecular Oligomerization.
Maugeri L., Lébl T., Cordes D.B., Slawin A.M., Philp D.
Journal of Organic Chemistry, American Chemical Society (ACS), 2017
208.
A Potent Halogen-Bonding Donor Motif for Anion Recognition and Anion Template Mechanical Bond Synthesis.
Bunchuay T., Docker A., Martinez‐Martinez A.J., Beer P.D.
Angewandte Chemie - International Edition, Wiley, 2019
209.
Tetra-fluorinated aromatic azide for highly efficient bioconjugation in living cells
Cai X., Wang D., Gao Y., Yi L., Yang X., Xi Z.
RSC Advances, Royal Society of Chemistry (RSC), 2019
211.
N,N-Diethylurea-Catalyzed Amidation between Electron-Deficient Aryl Azides and Phenylacetaldehydes
Xie S., Ramström O., Yan M.
Organic Letters, American Chemical Society (ACS), 2015
212.
Base-catalyzed synthesis of aryl amides from aryl azides and aldehydes
Xie S., Zhang Y., Ramström O., Yan M.
Chemical Science, Royal Society of Chemistry (RSC), 2016
213.
Anilide Formation from Thioacids and Perfluoroaryl Azides
Xie S., Fukumoto R., Ramström O., Yan M.
Journal of Organic Chemistry, American Chemical Society (ACS), 2015
214.
1,3-Dipolar Cycloaddition Reactivities of Perfluorinated Aryl Azides with Enamines and Strained Dipolarophiles
Xie S., Lopez S.A., Ramström O., Yan M., Houk K.N.
Journal of the American Chemical Society, American Chemical Society (ACS), 2015
215.
A versatile catalyst-free perfluoroaryl azide–aldehyde–amine conjugation reaction
Xie S., Zhou J., Chen X., Kong N., Fan Y., Zhang Y., Hammer G., Castner D.G., Ramström O., Yan M.
Materials Chemistry Frontiers, Royal Society of Chemistry (RSC), 2019
216.
Patent WO 2021154669
2021
217.
Patent WO 2008021388
2008
218.
Patent WO 2020146682
2020
219.
A New Class of Diaryl Ether Herbicides: Structure–Activity Relationship Studies Enabled by a Rapid Scaffold Hopping Approach
Kalnmals C.A., Benko Z.L., Hamza A., Bravo-Altamirano K., Siddall T.L., Zielinski M., Takano H.K., Riar D.S., Satchivi N.M., Roth J.J., Church J.B.
Journal of Agricultural and Food Chemistry, American Chemical Society (ACS), 2023
220.
Visible-Light-Induced Selective α-Fluoroarylation of Secondary N-Alkyl Anilines with Polyfluoroarenes via Direct C–H/C–F Coupling
Zhang H., Shao Q., Zhang Q., Xia C., Xu W., Wu M.
Organic Letters, American Chemical Society (ACS), 2023
221.
Visible-Light-Induced Selective Defluoroalkylations of Polyfluoroarenes with Alcohols
Xu W., Shao Q., Xia C., Zhang Q., Xu Y., Liu Y., Wu M.
Chemical Science, Royal Society of Chemistry (RSC), 2023
222.
Photocatalytic C–F alkylation; facile access to multifluorinated arenes
Singh A., Kubik J.J., Weaver J.D.
Chemical Science, Royal Society of Chemistry (RSC), 2015
223.
Stabilization of [2.2]paracyclophane anion as a result of transannular interaction
Nosova E.V., Lipunova G.N., Charushin V.N.
Russian Chemical Bulletin, Springer Nature, 2004
224.
Fluorine-containing quinazolines and their oxa and thia analogues: Synthesis and biological activities
Nosova E.V., Lipunova G.N., Charushin V.N.
Russian Chemical Reviews, Autonomous Non-profit Organization Editorial Board of the journal Uspekhi Khimii, 2009
225.
Features of reactions of polyfluorinated ethyl 4-oxo-2-pnenyl-4H-chromene-3-carboxylates with N-nucleophiles
Shcherbakov K.V., Burgart Y.V., Saloutin V.I.
Russian Journal of Organic Chemistry, Pleiades Publishing, 2013
226.
Novel 3-acetyl-2-methyl-polyfluorochromones in reactions with amines and esters of amino acids
Shcherbakov K.V., Bazhin D.N., Burgart Y.V., Saloutin V.I.
Chemistry of Heterocyclic Compounds, Springer Nature, 2015
227.
Polyfluorine-containing chromen-4-ones: synthesis and transformations
Shcherbakov K.V., Burgart Y.V., Saloutin V.I., Chupakhin O.N.
Russian Chemical Bulletin, Springer Nature, 2016
228.
Transformations of 3-acyl-4H-polyfluorochromen-4-ones under the action of amino acids and biogenic amines
Shcherbakov K.V., Artemyeva M.A., Burgart Y.V., Evstigneeva N.P., Gerasimova N.A., Zilberberg N.V., Kungurov N.V., Saloutin V.I., Chupakhin O.N.
Journal of Fluorine Chemistry, Elsevier, 2019
229.
7-Imidazolyl-substituted 4'-methoxy and 3',4'-dimethoxy-containing polyfluoroflavones as promising antiviral agents
Shcherbakov K.V., Artemyeva M.A., Burgart Y.V., Saloutin V.I., Volobueva A.S., Misiurina M.A., Esaulkova Y.L., Sinegubova E.O., Zarubaev V.V.
Journal of Fluorine Chemistry, Elsevier, 2020
230.
Structure of diethyl (polyfluorobenzoyl)malonates and their thermal intramolecular cyclization
Bazhin D.N., Schegol’kov E.V., Kudyakova Y.S., Burgart Y.V., Saloutin V.I.
Russian Chemical Bulletin, Springer Nature, 2011
231.
The synthesis and biological evaluation of A- and B-ring fluorinated flavones and their key intermediates
Shcherbakov K.V., Panova M.A., Burgart Y.V., Zarubaev V.V., Gerasimova N.A., Evstigneeva N.P., Saloutin V.I.
Journal of Fluorine Chemistry, Elsevier, 2021
233.
An efficient synthesis of 1-H indazoles
Lokhande P.D., Raheem A., Sabale S.T., Chabukswar A.R., Jagdale S.C.
Tetrahedron Letters, Elsevier, 2007
234.
Synthesis of Isoxazole Conjugates of 1,4-Benzodioxane Moiety via Intermolecular 1,3-Dipolar Cycloaddition
Vaidya V.V., Wankhede K.S., Nara S.J., Salunkhe M.M., Trivedi G.K.
Synthetic Communications, Taylor & Francis, 2009
236.
Patent EP 0556720
1993
237.
Patent EP 0638566
1995
238.
Patent US 5539112
1996
239.
Design and Synthesis of Potent Antitumor 5,4‘-Diaminoflavone Derivatives Based on Metabolic Considerations
Akama T., Ishida H., Shida Y., Kimura U., Gomi K., Saito H., Fuse E., Kobayashi S., Yoda N., Kasai M.
Journal of Medicinal Chemistry, American Chemical Society (ACS), 1997
241.
Structure−Activity Relationships of the 7-Substituents of 5,4‘-Diamino-6,8,3‘-trifluoroflavone, a Potent Antitumor Agent
Akama T., Ishida H., Kimura U., Gomi K., Saito H.
Journal of Medicinal Chemistry, American Chemical Society (ACS), 1998
242.
Cu(OAc)2 Mediated Synthesis of 3-Sulfonyl Chromen-4-ones
Chang M., Chen Y., Wang H.
Journal of Organic Chemistry, American Chemical Society (ACS), 2018
243.
New Fluoroaryl-containing β,β'-Dioxoesters in the Synthesis of Fluorobenzopyran-2(4)-ones
Kisil' S.P., Burgart Y.V., Saloutin V.I.
Russian Journal of Organic Chemistry, Pleiades Publishing, 2001
244.
Fluoroaryl containing β,β′-dioxoesters in the synthesis of fluorobenzopyran-4(2)-ones
Kisil’ S.P., Burgart Y.V., Saloutin V.I., Chupakhin O.N.
Journal of Fluorine Chemistry, Elsevier, 2001
245.
Patent CN 108997300
2018
246.
Patent US 20110313003
2011
247.
Patent WO 2011159297
2011
248.
Synthesis and biological activity of 5-amino- and 5-hydroxyquinolones, and the overwhelming influence of the remote N1-substituent in determining the structure-activity relationship
Domagala J.M., Bridges A.J., Culbertson T.P., Gambino L., Hagen S.E., Karrick G., Porter K., Sanchez J.P., Sesnie J.A.
Journal of Medicinal Chemistry, American Chemical Society (ACS), 1991
249.
Patent EP 0255908
1988
250.
Patent EP 0265230
1988
251.
Patent US 5097032
1992
252.
Patent EP 0393400
1990
253.
Patent WO 9823592
1998
254.
Patent WO 2002048113
2002
255.
An Improved Synthesis of Levofloxacin
Kim Y., Bang Kang S., Park S., Hae Kim Y.
Heterocycles, The Japan Institute of Heterocyclic Chemistry, 1997
256.
Patent WO 2003028650
2003
257.
Novel DNA Gyrase Inhibiting Spiropyrimidinetriones with a Benzisoxazole Scaffold: SAR and in Vivo Characterization
Basarab G.S., Brassil P., Doig P., Galullo V., Haimes H.B., Kern G., Kutschke A., McNulty J., Schuck V.J., Stone G., Gowravaram M.
Journal of Medicinal Chemistry, American Chemical Society (ACS), 2014
259.
Design, synthesis and biological evaluation of N-hydroxy-aminobenzyloxyarylamide analogues as novel selective κ opioid receptor antagonists
He G., Song Q., Wang J., Xu A., Peng K., Zhu Q., Xu Y.
Bioorganic and Medicinal Chemistry Letters, Elsevier, 2020
260.
Patent US 20060069093
2006
262.
Patent WO 2020178282
2020
263.
Identification of the 2-Benzoxazol-2-yl-phenol Scaffold as New Hit for JMJD3 Inhibition
Giordano A., Forte G., Terracciano S., Russo A., Sala M., Scala M.C., Johansson C., Oppermann U., Riccio R., Bruno I., Di Micco S.
ACS Medicinal Chemistry Letters, American Chemical Society (ACS), 2019
264.
Patent WO 2013104613
2013
265.
Repurposing a Library of Human Cathepsin L Ligands: Identification of Macrocyclic Lactams as Potent Rhodesain and Trypanosoma brucei Inhibitors
Giroud M., Dietzel U., Anselm L., Banner D., Kuglstatter A., Benz J., Blanc J., Gaufreteau D., Liu H., Lin X., Stich A., Kuhn B., Schuler F., Kaiser M., Brun R., et. al.
Journal of Medicinal Chemistry, American Chemical Society (ACS), 2018
266.
Patent WO 2016149393
2016
267.
Patent WO 2018005662
2018
268.
Patent WO 2016003929
2016
269.
Patent WO 2018218154
2018
270.
Patent WO 2015191907
2015
271.
Patent WO 2015179308
2015
272.
Patent WO 2019226931
2019
273.
Patent WO 2010129057
2010
274.
Structural diversity and photoluminescence of copper(I) carboxylates: From discrete complexes to infinite metal-based wires and helices
275.
On the Structural Features of Substituted Lanthanide Benzoates
Utochnikova V.V., Kalyakina A.S., Solodukhin N.N., Aslandukov A.N.
European Journal of Inorganic Chemistry, Wiley, 2019
277.
Dipole fluctuation and structural phase transition in hydrogen-bonding molecular assemblies of mononuclear CuII complexes with polar fluorobenzoate ligands
Takahashi K., Miyazaki Y., Noro S., Nakano M., Nakamura T., Akutagawa T.
Dalton Transactions, Royal Society of Chemistry (RSC), 2021
279.
Syntheses, Crystal Structures and Thermal Behavior of Five New Complexes Containing 2, 4, 6-Trifluorobenzoate as Ligand
Lamann R., Hülsen M., Dolg M., Ruschewitz U.
Zeitschrift fur Anorganische und Allgemeine Chemie, Wiley, 2012
280.
Crystal structure of catenapoly[diaqua-(μ 24,4′-bipyridine)-κ 2 N:N′)-bis(2,6-difluorobenzoate)-κO)nickel(II)] ethanol monosolvate, C28H30F4N2O8Ni
Hou-Qun Y., Wei X., Chun-Yan H., Guang-Ming B.
Zeitschrift fur Kristallographie - New Crystal Structures, Walter de Gruyter, 2016
282.
Influence of Metal Coordination on the Gas-Phase Chemistry of the Positional Isomers of Fluorobenzoate Complexes
Xiong Z., Yang M., Chen X., Gong Y.
Journal of the American Society for Mass Spectrometry, American Chemical Society (ACS), 2022
283.
Influence of the Fluorinated Aromatic Fragments on the Structures of the Cadmium and Zinc Carboxylate Complexes Using Pentafluorobenzoates and 2,3,4,5-Tetrafluorobenzoates as Examples
Shmelev M.A., Kuznetsova G.N., Dolgushin F.M., Voronina Y.K., Gogoleva N.V., Kiskin M.A., Ivanov V.K., Sidorov A.A., Eremenko I.L.
Russian Journal of Coordination Chemistry/Koordinatsionnaya Khimiya, Pleiades Publishing, 2021
284.
Fluorine-functionalized metal–organic frameworks and porous coordination polymers
285.
Decoration and utilization of a special class of metal–organic frameworks containing the fluorine moiety
Kumar S., Mohan B., Fu C., Gupta V., Ren P.
Coordination Chemistry Reviews, Elsevier, 2023
286.
Metal complexes based on polyfluorosalicylic acids and their antimycotic and antimicrobial activity
Shchur I.V., Shchegolkov E.V., Burgart Y.V., Kozitsina A.N., Ivanova A.V., Alyamovskaya I.S., Evstigneeva N.P., Gerasimova N.A., Ganebnykh I.N., Zilberberg N.V., Kungurov N.V., Saloutin V.I., Chupakhin O.N.
Polyhedron, Elsevier, 2020
287.
Copper complexes of non-steroidal anti-inflammatory drugs: an opportunity yet to be realized
Weder J.E., Dillon C.T., Hambley T.W., Kennedy B.J., Lay P.A., Biffin J.R., Regtop H.L., Davies N.M.
Coordination Chemistry Reviews, Elsevier, 2002
288.
Copper(II) Complexes of Salicylic Acid Combining Superoxide Dismutase Mimetic Properties with DNA Binding and Cleaving Capabilities Display Promising Chemotherapeutic Potential with Fast Acting in Vitro Cytotoxicity against Cisplatin Sensitive and Resistant Cancer Cell Lines
O’Connor M., Kellett A., McCann M., Rosair G., McNamara M., Howe O., Creaven B.S., McClean S., Foltyn-Arfa Kia A., O’Shea D., Devereux M.
Journal of Medicinal Chemistry, American Chemical Society (ACS), 2012
289.
Orally active antioxidative copper(II) aspirinate: synthesis, structure characterization, superoxide scavenging activity, and in vitro and in vivo antioxidative evaluations
290.
Synthesis, crystal structures, and anti-convulsant activities of ternary [ZnII(3,5-diisopropylsalicylate)2], [ZnII(salicylate)2] and [ZnII(aspirinate)2] complexes
Lemoine P., Viossat B., Dung N.H., Tomas A., Morgant G., Greenaway F.T., Sorenson J.R.
Journal of Inorganic Biochemistry, Elsevier, 2004
291.
Patent RU 2706702
2019
292.
A cobalt(ii) bis(salicylate)-based ionic liquid that shows thermoresponsive and selective water coordination
Kohno Y., Cowan M.G., Masuda M., Bhowmick I., Shores M.P., Gin D.L., Noble R.D.
Chemical Communications, Royal Society of Chemistry (RSC), 2014
293.
Copper(II) and cobalt(II) complexes based on methyl trifluorosalicylate and bipyridine-type ligands: Synthesis and their antimicrobial activity
Shchegolkov E.V., Shchur I.V., Burgart Y.V., Slepukhin P.A., Evstigneeva N.P., Gerasimova N.A., Zilberberg N.V., Kungurov N.V., Saloutin V.I., Chupakhin O.N.
Polyhedron, Elsevier, 2021
294.
Synthesis and anti-Candida activity of copper(II) and manganese(II) carboxylate complexes
Geraghty M., Sheridan V., McCann M., Devereux M., McKee V.
Polyhedron, Elsevier, 1999
295.
Synthesis, crystal structure, EPR properties, and anti-convulsant activities of binuclear and mononuclear 1,10-phenanthroline and salicylate ternary copper(II) complexes
Lemoine P., Viossat B., Morgant G., Greenaway F.T., Tomas A., Dung N., Sorenson J.R.
Journal of Inorganic Biochemistry, Elsevier, 2002
297.
C.P.Saxena, S.H.Mishra, P.V.Khadikar
1979
298.
Patent RU 2737435
2020
299.
Promising Antifungal and Antibacterial Agents Based on 5‐Aryl‐2,2′‐bipyridines and Their Heteroligand Salicylate Metal Complexes: Synthesis, Bioevaluation, Molecular Docking
Burgart Y., Shchegolkov E., Shchur I., Kopchuk D., Gerasimova N., Borisevich S., Evstigneeva N., Zyryanov G., Savchuk M., Ulitko M., Zilberberg N., Kungurov N., Saloutin V., Charushin V., Chupakhin O., et. al.
ChemMedChem, Wiley, 2021
300.
Polyfluorosalicylic acids as ligands for the creation of bioactive metal complexes
Shchegolkov E.V., Shchur I.V., Burgart Y.V., Saloutin V.I.
AIP Conference Proceedings, American Institute of Physics (AIP), 2022