Skip to content
Publicly Available Published by De Gruyter February 18, 2021

Fluorine analogs of dicamba and tricamba herbicides; synthesis and their pesticidal activity

  • Bogumiła Huras , Jerzy Zakrzewski , Krzysztof Żelechowski , Anna Kiełczewska , Maria Krawczyk , Jarosław Hupko and Katarzyna Jaszczuk

Abstract

Fluorine analogs of the dicamba and tricamba herbicides were synthesized. Their herbicide activities were compared with the activities of the pattern herbicides dicamba and tricamba.

1 Introduction

The investigations of the properties of fluorine analogs of organic molecules are the result of the activity and specific physicochemical properties of compounds containing fluorine atoms.

The fluorine atom is known to imitate a hydrogen atom [1]. Structural similarity and analogous biological effects of two functional groups (known as bioisosterism) are used especially for the replacement of hydrogen by fluorine. Such a replacement is considered as the most popular monovalent bioisosteric substitution [2]. A similar size of both atoms is the most common reason for substituting hydrogen by fluorine [1], [2], [3], [4]. The strong inductive electron-withdrawing effect of the fluorine atom can significantly affect the reactivity and stability of functional groups as well as the reactivity of neighboring reaction centers [1]. In one of our previous works, attempts were made to exchange the hydrogen atoms in the methylene group of the phenoxyacetic acid herbicides 2,4-dichlorophenoxyacetic acid (2,4-D) and 2-methyl-4-chlorophenoxyacetic acid (MCPA) with fluorine atoms [5].

The second popular idea of converting one set of atoms into another one is to examine various atoms belonging to the same group of the periodic table of elements – exchange halogen atoms with other halogens or an oxygen atom with another chalcogen: sulfur, selenium or tellurium.

The exchange of entire fragments can also be seen in the literature. The analogs of phenoxyacetic acids as potential candidates for compounds showing an herbicidal activity were suggested. Instead of chlorophenoxy fragments (2,4-dichlorophenoxy, 2-methyl-4-chlorophenoxy, 2,4,5-trichlorophenoxy), eugenol, thymol, carvacrol, vanillin, ferulic acid [6], [, 7], gallic acid [8] and piperidine derivatives [9] have been proposed.

The purpose of this study is to exchange the chlorine atoms in the chlorinated herbicides dicamba (Scheme 1) and tricamba (Scheme 2) with fluorine, in order to examine the herbicidal and fungistatic activity of the fluorine analogs and to compare it with the activity of the parent herbicides.

Scheme 1: Synthesis of fluorine analog 3 of dicamba.
Scheme 1:

Synthesis of fluorine analog 3 of dicamba.

Scheme 2: Synthesis of fluorine analog 11 of tricamba.
Scheme 2:

Synthesis of fluorine analog 11 of tricamba.

2 Results and discussion

The syntheses of the fluorine analog of the herbicide dicamba (2-methoxy-3,6-difluorobenzoic acid (3)) and the fluorine analog of the herbicide tricamba (2,3,5-trifluoro-6-methoxybenzoic acid (11)) are presented in Schemes 1 and 2, respectively.

The fluorine analog of the herbicide dicamba (2-methoxy-3,6-difluorobenzoic acid (3)) was synthesized by the lithiation of 1,4-difluoro-2-methoxybenzene (2,5-difluoroanisole (2)) with lithium diisopropylamide (LDA), followed by the carboxylation with the gaseous carbon dioxide [10] (Scheme 1).

A fluorine analog of the herbicide tricamba (2,3,5-trifluoro-6-methoxybenzoic acid (11)) was synthesized by the lithiation of 1,2,4-trifluoro-5-(methoxymethoxy)benzene (9) with n-buthyllithium followed by carboxylation with gaseous carbon dioxide. The salicylic acid derivative 2-hydroxy-3,5,6-trifluorobenzoic acid (3,5,6-trifluorosalicylic acid (10)) thus obtained was then methylated with dimethyl sulfate to the target 2,3,5-trifluoro-6-methoxybenzoic acid (11) (Scheme 2).

The lithiation of 1,2,4-trifluoro-5-methoxybenzene (2.4.5-trifluoroanisole (6)) with either LDA or n-buthyllithium followed by the carboxylation with the gaseous carbon dioxide lead to the regioisomeric (in relation to the tricamba scaffold) compound 3-methoxy-2,5,6-trifluorobenzoic acid (7) (Scheme 2).

The observed difference between the regiodirected lithiation of 6 (ArOCH3) and 9 (ArOCH2OCH3, ArOMOM) is a result consistent with the work of Schlosser [11], and in accordance with the known ortho-directing ability of the OMOM substituent during lithiation processes [12], [13], [14], [15], [16], [17], [18].

The methylation with dimethyl sulfate of the compound bearing both carboxyl and hydroxyl groups (10 → 11) yielded a mixture of the desired ether and the methyl ester of the carboxylic acid. To avoid the unnecessary separation of both ether and ester, dimethyl sulfate was used in excess. The crude product was routinely treated with sodium hydroxide solution to hydrolyze the undesired ester.

It is noteworthy to mention, that the synthesis of the compounds 10 and 11 has been reported using a different starting material. The elimination of one fluorine atom from the benzene ring containing four fluorine atoms was proposed. The reaction of 2,3,5,6-tetrafluorobenzoic acid with magnesium methoxide gave 11 (no experimental data due to the co-presence of dimethoxide side-product). Further demethylation and hydrolysis using HBr of the corresponding acid chloride yielded well characterized 10 [19].

For additional confirmation of the structure of the obtained anizole derivatives 3 and 7, they were demethylated using boron tribromide yielding the corresponding hydroxyl derivatives 4 and 8.

The structures of the synthesized compounds were confirmed by spectroscopic methods. The structures of the starting phenolic ethers (2, 6, 9) were confirmed and characterized using MS and IR spectra. The structures of the carboxylated derivatives, compounds 3, 4, 7, 8, 10, and 11, were confirmed as characterized by MS, 1H, 13C, 19F NMR and IR spectra. The new compounds 7 and 11 were additionally characterized by HRMS measurements. The mass spectra of the synthesized compounds show the parent signals with great intensity (100% for 7 and 8).

The regioselectivity of the compounds 7, 8, 10, 11 was confirmed by the analysis of the coupling pattern and coupling constants of the hydrogen atom with the fluorine ones.

The 1H NMR of the hydrogen atom signal in the benzene ring are presented in Figure 1a (the fluorine analog of tricamba 11 and the corresponding salicylic acid 10) and in Figure 1b (the isomeric compounds 7 and 8.).

Figure 1a: 1H NMR: the hydrogen atom signal in the benzene ring for the fluorine analog of tricamba 11 (td, J = 10.2, 7.5 Hz), and the corresponding salicylic acid 10 (td, J = 10.5, 7.2 Hz).
Figure 1a:

1H NMR: the hydrogen atom signal in the benzene ring for the fluorine analog of tricamba 11 (td, J = 10.2, 7.5 Hz), and the corresponding salicylic acid 10 (td, J = 10.5, 7.2 Hz).

Figure 1b: 1H NMR: the hydrogen atom signal in the benzene ring for the isomeric compounds 7 (dt, J = 11.1, 7.5 Hz) and 8 (dt, J = 11.7, 7.95 Hz).
Figure 1b:

1H NMR: the hydrogen atom signal in the benzene ring for the isomeric compounds 7 (dt, J = 11.1, 7.5 Hz) and 8 (dt, J = 11.7, 7.95 Hz).

Comparison of the coupling constants of the hydrogen atom with the fluorine atoms for the pairs 11 and 10, as well as for 7 and 8 confirms the position of the hydrogen atom in the ring. For the compounds 7 and 8, the hydrogen atom is coupled with one fluorine atom in the ortho position (d, J ∼ 11 Hz) and two fluorine atoms in the meta position (t, J ∼ 7.5 Hz) This leads to a doublet of triplets – the pattern presented in Figure 1b.

For the compounds 10, 11 the situation is opposite. A hydrogen atom in the ring is coupled with two fluorine atoms in the ortho postion (t, J ∼ 11 Hz) and one fluorine atom in the meta position (d, J ∼ 7.5 Hz). This leads to a triplet of doublets – the pattern presented in Figure 1a.

The herbicidal activity of the obtained compounds was tested against the following species of weeds: cleavers (Galium aparine), pale smartweed (Polygonum lapathifolium), common poppy (Papaver rhoeas), barnyard grass (Echinochloa crusgali), gallant soldier (Galinsoga parviflora), fathen (Chenopodium album), ribwort (Plantago lanceolata), black mustard (Brassica nigra), red-root amaranth (Amaranthus retroflexus), common chickweed (Stellaria media). The results are presented in Table 1.

Table 1:

Herbicidal activity of the fluorine analog of dicamba 3 and the fluorine analog of tricamba 11 against weeds (pre-emergent/post-emergent activities).a

CompoundDose (g ha−1)Cleavers (Galium aparine)Pale smartweed (Polygonum lapathifolium)Common poppy (Papaver rhoeas)Barnyard grass (Echinochloa crusgali)Gallant soldier (Galinsoga parviflora)Fathen (Chenopodium album)Ribwort (Plantago lanceolata)Black mustard (Brassica nigra)Red-root amaranth (Amaranthus retroflexus)Common chickweed (Stellaria media)
20002010090010010040554040
551001000100100557010070
3000000000000
2070700100100407010040
1500000000000
10404007070204010020
2000100100100100100100100100100100
10010010095100100100100100100
30060401000401000706065
100100100010010070100100100
15040207002070040200
901001000100100407010070
2000402070040700602040
4040400404020404040
3000000000000
0000000000
200010010070070100209010070
951001002010010090100100100
3000000000000
60100550100100406510055
  1. aResults (degree [in %] of a visual observation of a phytotoxicity weed damage for a weed species used against control, dose 2 kg ha−1) are expressed as a fraction: a numerator means a result (in %) of a pre-emergence (via soil) treatment, a denominator means a result (in %) of a post-emergence (on leaves) treatment.

The fluorine analog of dicamba (2-methoxy-3,6-difluorobenzoic acid (3)) showed herbicidal activity slightly weaker than dicamba itself (2-methoxy-3,6-dichlorobenzoic acid). Dicamba itself destroyed eight weed species respectively with >90% efficiency. Compound 3 did not reveal pre-emergent activity.

The fluorine analog of tricamba (2,3,5-trifluoro-6-methoxybenzoic acid (11)) showed no herbicidal activity as opposed to the tricamba itself (2,3,5-trichloro-6 methoxybenzoic acid), which destroyed four weed species with >90% efficiency.

None of the tested salicylic acids 3,6-difluoro-2-hydroxybenzoic acid (3,6-difluorosalicylic acid (4)), 3,6-dichloro-2-hydroxybenzoic acid (3,6-dichlorosalicylic acid), 2,3,5-trifluoro-6-hydroxybenzoic acid (3,5,6-trifluorosalicylic acid (10)), as well as 2,3,5-trichloro-6-hydroxybenzoic acid (3,5,6-trichlorosalicylic acid) showed herbicidal activity.

Fungistatic activity of the obtained compounds was tested against the following phytopathogenic strains: Alternaria alternata, Botrrytis cinerea, Fusarium culmorum, Phytophtora cactorum, Rhizoctonia solani, Phytophtora infestans. The results are presented in Table 2.

Table 2:

Fungistatic activity of fluorine analogs of dicamba 3 and 4, fluorine analogs of tricamba 10 and 11, as well as isomeric derivatives 7 and 8; against phytopathogenic fungi; concentration: 200 mg L−1; solvent: acetone.a

CompoundFormulaAlternaria alternataBotrytis cinereaFusarium culmorumPhytophtora cactorumRhizoctonia solaniPhytophtora infestans
406.327.1004.2
3000002.1
3,6-Dichloro-2-hydroxybenzoic acid (3,6-dichloro-salicylic acid)11.96.316.7000
Dicamba000000
808.226.27.225.50
70028.60012.5
1006.322.9000
114.806.3006.2
2,3,5-trichloro-6-hydroxybenzoic acid (3,5,6-trichloro-salicylic acid)16.741.722.921.2012.5
Tricamba0016.7004.2
  1. aThe results are expressed as a percentage of linear growth reduction of a fungus colony; percentage of linear growth reduction = ([colony diameter of a control plate − colony diameter of a tested plate]/[colony diameter of a control plate]) × 100.

The fluorine analog of dicamba (3,6-difluoro-2-methoxybenzoic acid (3)) as well as dicamba itself showed no fungicidal activity. The fluorine analog of tricamba (2,3,5-trifluoro-6-methoxybenzoic acid (11)), tricamba itself (2,3,5-trichloro-6-methoxybenzoic acid) and all tested salicylic acids 3,6-difluoro-2-hydroxybenzoic acid (3,6-difluorosalicylic acid (4)), 3,6-dichloro-2-hydroxybenzoic acid (3,6-dichlorosalicylic acid), 2,3,5-trifluoro-6-hydroxybenzoic acid (3,5,6-trifluorosalicylic acid (10)) as well as 2,3,5-trichloro-6-hydroxybenzoic acid (3,5,6-trichlorosalicylic acid) showed weak fungicidal activity against individual tested strains of phytopathogenic fungi.

The isomeric compounds 2,3,6-trifluoro-5-methoxybenzoic acid (7) and 2,3,6-trifluoro-5-hydroxybenzoic acid (8) showed but weak fungistatic activity.

3 Conclusions

The chlorine atoms in two herbicides 2,5-dichloro-6-methoxybenzoic acid (dicamba) and 2,3,5-trichloro-6-methoxybenzoic acid (tricamba) were replaced with fluorine atoms. The fluorine analogs of dicamba and tricamba 2,5-difluoro-6-methoxybenzoic acid (3) and 2,3,5-trifluoro-6-methoxybenzoic acid (11), respectively, were characterized by spectroscopic methods (MS, HRMS, 1H NMR, 13C NMR, 19F NMR, IR), which fully confirmed their structures. The fluorine analogs 3 and 11 were evaluated for their herbicidal and fungistatic activities which revealed a weaker herbicidal and fungistatic activity than the parent herbicides dicamba and tricamba.

4 Experimental part

4.1 General

2,5-Difluorophenol, 2,4,5-trifluorophenol, dimethyl sulfate, boron tribromide, as well as n-butyllithium and LDA solutions were commercially available and used as received.

Chloromethylmethyl ether (MOMCl) was synthesized according to ref. [20]. Gaseous HCl was passed through a mixture of methanol (44 g, 1.375 mol, 55.6 mL) and 37.5% aqueous solution of formaldehyde (90 g, 1.125 mol) for 6 h in accord with the literature procedure. MOMCl (49.13 g, 54.25%, b.p. 55–60 °C/760 mmHg) (lit. 55–60 °C/760 mmHg [20]) was obtained.

TLC was carried out on silica gel Merck Alurolle 5562 or Alufolien 5554; typical mobile phases: hexane-ethyl acetate (9:1); benzene-ethyl acetate (9:1); benzene-methanol (9:1). TLC visualization was achieved using UV 254 nm light and/or I2 vapor. Column chromatography was performed on silica gel 0.040–0.063 mm, 230–400 mesh: Merck 1.09385.1000 or Zeochem 60 hyd. mobile phases: benzene, benzene-ethyl acetate (95:5, 9:1, 4:1). EI MS data (70 eV) were recorded on an AMD 604 and Agilent Technologies 5975 B mass spectrometers. EI HR MS data were recorded by using an AMD 604 mass spectrometer. IR spectra were recorded on an FT/IR Jasco 420 spectrophotometer. 1H, 13C and 19F NMR data were collected using a Varian UNITYplus 300 or 500 spectrophotometers (at 200 or 500 MHz for 1H, respectively).

4.2 1,4-Difluoro-2-methoxybenzene (2,5-difluoroanisole) (2)

A mixture of 2,5-difluorophenol (1, 11.96 g, 0.092 mol), triturated anhydrous potassium carbonate (15.25 g, 0.11 mol), dimethyl sulfate (13.92 g, 10.45 mL, 0.11 mol) and acetone (240 mL) was stirred at reflux for 2 h. After cooling, the precipitate was filtered off and washed with acetone (10 mL). The filtrate was evaporated. Water (150 mL) was added to the residue (16.5 g) and the mixture was extracted with diethyl ether. The ether layer was washed with water, ammonium hydroxide (15%, 150 mL, stirring for 1.5 h), and with water to pH 7–8 again. The ether layer was dried over anhydrous magnesium sulfate. Filtration of the drying agent and evaporation of the solvent gave the crude product (14.4 g) which was distilled under reduced pressure to give 2.5-difluoroanisole (2, 10.7 g, 81.4%, colorless oil, b.p. 59–60 °C/15 mm Hg (lit. 72–73 °C/25 mm Hg [21]). – MS (EI, 70 eV): m/z (%) = 144 (96) [M]+, 129 (47), 115 (12), 114 (15), 113 (11), 101 (100), 96 (4), 95 (6), 81 (10), 75 (15), 63 (19), 57 (11), 51 (10), 50 (6). – IR (cm−1, film): ν = 3100, 3010, 2980, 2946, 2920, 2890, 2850, 1626, 1514, 1453, 1420, 1325, 1290, 1250, 1210, 1192, 1150, 1101, 1101, 1031, 950, 836, 786, 715, 608.

4.3 3,6-Difluoro-2-methoxybenzoic acid (3)

1,4-Difluoro-2-methoxybenzene (2,5-difluoroanisole, 2, 5.56 g, 0.039 mol) in anhydrous THF (15 mL) was cooled to −60 °C. A solution of LDA 1.5 m (30.5 mL, 0.045 mol) was added with a syringe via septum in portions (9, 4, 4, 6, 7.5 mL). Because the temperature of the reaction mixture rose rapidly, adding of LDA was carried out at such a rate that the temperature of the reaction mixture did not increase above −50 °C. After adding LDA, the reaction mixture was stirred for 30 min at −70 °C. Gaseous carbon dioxide was bubbled through the reaction mixture for 1 h. The temperature was maintained at −78 to −50 °C. The reaction mixture was allowed to slowly warm to 0 °C and then was carefully acidified with 30% H2SO4 to pH 3–4. The aqueous layer was extracted with ether (2 × 50 mL). The organic layer was concentrated. The residue was dissolved in 2 m NaOH (50 mL) and washed with hexane to remove unreacted 2,5-difluoroanisole (2). The aqueous layer was acidified with 30% H2SO4 to pH 3–4 and left for 12 h in a refrigerator. The precipitate (7.14 g) was dissolved in ethyl acetate. The solution was dried over anhydrous magnesium sulfate, and after removing the drying agent, concentrated. Crude 3,6-difluoro-2-methoxybenzoic acid (3, 5.43 g) was purified by a double column chromatography (eluent: hexane-ethyl acetate 4:1). 3,6-Difluoro-2-methoxybenzoic acid (3) was obtained as a colorless solid, 3.87 g, 53.3%, m.p. 81–82 °C (lit. 82–83 °C [10]). – MS (EI, 70 eV): m/z (%) = 188 (56) [M]+, 171 (22), 170 (18), 169 (8), 159 (32), 156 (34), 141 (100), 128 (29), 115 (13), 114 (7), 113 (15), 112 (14), 101 (21), 100 (20), 99 (12). – 1H NMR (300 MHz, CDCl3): δ = 12.27 (1 H, s, COOH), 7.20 (1H, td, J = 10.0, 5.1 Hz, Ph), 6.84 (1H, td, J = 8.85, 3.6 Hz, Ph), 4.06 (3H, s, OMe). – 13C NMR (75.5 MHz, CDCl3): δ = 62.30 (d, J = 6.3 Hz, OCH3), 110.83 (dd, J = 24.1, 7.4 Hz), 116.02 (d, J = 17.5 Hz), 119.71 (dd, J = 21.8, 10.3 Hz), 146.38 (dd, J = 13.7, 5.6 Hz), 151.55 (dd, J = 244.8, 3.5 Hz), 155.81 (dd, J = 251.8, 2.4 Hz), 168.28 (s, COOH). –19F NMR (282 MHz, CDCl3): δ = −116.96 (m), −133.94 (tm, J = 13.8 Hz). – IR (cm−1, KBr): ν = 430, 2962, 2680, 2580, 1705, 1622, 1489, 1429, 1320, 1291, 1239, 1059, 969, 913, 819, 764.

4.4 3,6-Difluoro-2-hydroxybenzoic acid (3,6-difluoro-salicylic acid) (4)

To a solution of 3,6-difluoro-2-methoxybenzoic acid (3, 0.5 g, 0.0027 mol) in anhydrous dichloromethane (20 mL), 1.0 m solution of boron tribromide (5.4 mL, 2 eq.) in dichloromethane was added with magnetic stirring, using a syringe, at 0 °C, under argon atmosphere. The reaction mixture was stirred for 2 h at 0 °C, then was allowed to slowly warm to room temperature and the stirring was continued for 12 h. In order to quench the reaction, methanol (approx. 5 mL) was added dropwise at 0 °C. During the methanol addition, the temperature rose to 17 °C. Water was added to the reaction mixture. The layers were separated. The organic layer was washed with brine and dried over anhydrous sodium sulfate. Filtration of the drying agent and concentration gave the crude product (0.5 g), which was purified by column chromatography with hexane-ethyl acetate 20:1, 9:1, 4:1 as eluents. The fractions containing the product were combined and concentrated. The residue was washed with hexane to give 3,6-difluoro-2-hydroxybenzoic acid (4, 0.2 g, 42.6%, HPLC: 99.95%, m.p.: 154–155 °C (lit. 154–155 °C [11])). – MS (EI, 70 eV): m/z (%) = 174 (38) [M]+, 156 (100), 128 (54), 100 (44), 81 (12). – 1H NMR (300 MHz, acetone-d6): δ = 11.58 (bs), 7.41 (1H, ddd, J = 10.35, 9.2, 4.8 Hz), 6.70 (1H, ddd, J = 10.35, 9.2, 3.6 Hz). – 13C NMR (75.5 MHz, acetone-d6): δ = 170.99 (t, J = 3.28 Hz, COOH), 158.90 (dd, J = 256.2, 2.76 Hz), 152.15 (dd, 1 J = 4.65, 4.38 Hz), 148.63 (dd, J = 240.28, 3.66 Hz), 122.04 (dd, J = 19.78, 11.55 Hz), 106.45 (dd, J = 25.82, 6.72 Hz), 105.23 (dd, J = 15.44, 3.05 Hz). – 19F NMR (282.5 MHz, acetone-d6): δ = −111 (ddd, J = 17.37, 10.45, 4.7 Hz), −142.12 (ddd, J = 17.44 Hz, 10.38, 3.53 Hz). – IR (cm−1, KBr): ν = 3427, 3100, 1715, 1662, 1640, 1495, 1445, 1249, 1181, 1035, 810, 760.

4.5 1,2,4-Trifluoro-5-methoxybenzene (2.4.5-trifluoroanisole) (6)

A mixture of 2,4,5-trifluorophenol (5, 5.03 g, 0.034 mol), triturated anhydrous potassium carbonate (8.28 g, 0.06 mol), dimethyl sulfate (5.68 mL, 0.06 mol, 7.56 g), and acetone (90 mL) was stirred at reflux for 2 h. After cooling, the precipitate was filtered and washed with acetone (10 mL). The filtrate was evaporated. Water (55 mL) was added to the residue (7.6 g) and the mixture was extracted with diethyl ether. The ether layer was washed with water, ammonium hydroxide (15%, 55 mL, stirring for 1.5 h) and with water to pH 7–8 again. The ether layer was dried over anhydrous magnesium sulfate. Filtration of the drying agent and evaporation of the solvent afforded the crude 1,2,4-trifluoro-5-methoxybenzene (6, 3.84 g, 69.7%, GC: 97.8%, colorless solid, m.p. 37–38 °C (lit. m.p. 35–37 °C [22], 36–37 °C [23])) which was used without purification for the next step. – GC/MS: GC: t.r. = 6.164 min. – MS (EI, 70 eV): m/z (%) = 163(8), 162 (94) [M]+,148 (5), 147 (71), 131 (6), 120 (6), 119 (100), 81 (16), 69 (9). – IR (cm−1, KBr): ν = , 2962, 1640, 1529, 1261, 1096, 1020, 802.

4.6 2,3,6-Trifluoro-5-methoxybenzoic acid (7)

4.6.1 Lithiation with n-butyllithium

2,4,5-Trifluoroanisole (6, 2.9 g, 0.018 mol) was placed in anhydrous THF (36 mL) under argon. The solution was cooled to −65 °C. 2.5 m n-butyllithium solution (8 mL, 0.02 mol) was added in portions with a syringe through a septum. The temperature of the reaction mixture rose to −60 °C. n-Butyllithium solution were added dropwise in two portions (2 × 4 mL). After completion of the addition, the reaction mixture was stirred for 0.5 h at −65 °C. Gaseous carbon dioxide was passed through the reaction mixture for 1 h at approx. −64 °C. A copious amount of white precipitate was observed. The reaction mixture was allowed to warm to 0 °C, and hydrolyzed with 30% H2SO4 (5 mL) to pH about two. Layers were separated. The aqueous layer was washed with diethyl ether (2 × 20 mL). The combined organic layers were washed with water (2 × 20 mL) and dried over anhydrous sodium sulfate (TLC control of 7 benzene-acetone 1:1). After evaporation the crude 7 was obtained (6.6 g). The crude 2,3,5-trifluoro-6-methoxybenzoic acid 7 was purified by column chromatography. The crude product was dissolved in ethyl acetate, silica gel (20 g) was added and the mixture was evaporated to dryness. The product applied on the gel was placed on the top of a chromatography column. Elution with hexane-ethyl acetate 9:1 (100 mL), 4:1 (200 mL), 1:1 (100 mL), acetone gave the purified 2,3,5-trifluoro-6-methoxybenzoic acid (7, 2.61 g, 70.4%, colorless solid, m.p. 130–133 °C).

4.6.2 Lithiation with lithium diisopropyl amide (LDA)

2,4,5-Trifluoroanisole (6, 4.05 g, 0.025 mol) was placed in anhydrous THF (20 mL) under argon. The solution was cooled to −70 °C. A 1.5 m LDA solution (approx. 20 mL, 0.029 mol) was added in portions with a syringe through a septum. The temperature of the reaction mixture rose to −66 °C. LDA solution was added dropwise in five portions (5 × 4 mL). After completion of the addition, the reaction mixture was stirred for 0.5 h at −72 °C. Gaseous carbon dioxide was passed through the reaction mixture for 1 h. Initially, the temperature of the reaction mixture rapidly increased to −45 °C, then it maintained at approx. −70 °C. A copious amount of white precipitate was observed. The reaction mixture was allowed to warm to 0 °C, and hydrolyzed with 30% H2SO4 to pH 3–4 (10 mL). After the addition of the first 2 mL of H2SO4, the reaction mixture thickened and a copious amount of CO2 was evolved. Diethyl ether was added (approx. 15 mL). The presence of 7 was controlled with TLC benzene-acetone 1:1). The organic layer was extracted with water (30 mL). The aqueous layer was additionally acidified (30% H2SO4, 1 mL) and extracted with ether (20 mL). The combined ether extracts were dried over anhydrous sodium sulfate. After evaporation the crude product 7 was obtained (4.23 g). The crude 7 was purified by double column chromatography. The crude product was dissolved in ethyl acetate, silica gel (20 g) was added and the mixture evaporated to dryness. The product applied on the gel, was placed on the top of a chromatography column. Elution with hexane (150 mL), hexane-ethyl acetate 9:1 (100 mL), 4:1 (200 mL), 1:1 (100 mL), gave pre-purified product (3.52 g) which was subjected to a second chromatography. The second chromatography gave finally 2,3,5-trifluoro-6-methoxybenzoic acid (7, 3.0 g, light-creamy solid, 58.3%, GC: 99.73%, HPLC: 99.5%, m.p. 130–135 °C, dec.). – MS (EI, 70 eV): m/z (%) = 207 (12), 206 (100) [M]+, 191 (16), 189 (14), 171 (33), 161 (11), 147 (22), 146 (11), 143 (36), 119 (32), 118 (15), 115 (9), 99 (27), 81 (9), 80 (9), 68 (9). – HRMS (EI, 70 eV): m/z = 206.0190 (calcd. 206.0191 for C8H5F3O3, [M]+). – 1H NMR (300 MHz, CDCl3): δ = 10.738 (1H, bs, COOH), 7.00 (1H, dt, J = 11.1, 7.5 Hz, Har), 3.90 (1H, s, OCH3). – 13C NMR (75.5 MHz, CDCl3): δ = 165.66 (s, COOH), 146.40 (ddd, J = 253.831, 7.32, 4.15 Hz), 146.22 (dd, J = 246.55, 4.33 Hz), 144.31 (ddd, J = 11.91, 7.87, 3.83 Hz), 141.98 (ddd, J = 253.79, 15.41, 4.49 Hz), 111.47 (t, J = 15.18 Hz), 105.79 (dd, J = 22.42, 2.72 Hz), 57.14 (s, OCH3). – 19F NMR (282.5 MHz, CDCl3): δ = −135.43 (ddd, J = 13.56, 7.49, 6.215), −139.20 (ddd, J = 21.89, 13.42, 11.3 Hz), −145.35 (dt, J = 22.32, 6.50 Hz). – IR (cm−1, KBr): ν = 3111, 2690, 1711, 1610, 1499, 1453, 1412, 1378, 1299, 1209, 1038, 954, 906, 849, 694.

4.7 2,3,6-Trifluoro-5-hydroxybenzoic acid (8)

To a solution of 2,3,6-trifluoro-5-methoxybenzoic acid (7, 1.0 g, 0.0049 mol) in anhydrous dichloromethane (20 mL), 1.0 m solution of boron tribromide (9.7 mL, 2 eq.) in dichloromethane was added with magnetic stirring using a syringe, at 0 °C under argon atmosphere. The reaction mixture was stirred for 2 h at 0 °C, then was allowed to slowly warm to room temperature and the stirring was continued for 12 h. In order to quench the reaction, methanol (approx. 5 mL) was added dropwise at 0 °C. During the methanol addition, the temperature rose to 17 °C. Water was added to the reaction mixture. The layers were separated. The organic layer was washed with brine and dried over anhydrous sodium sulfate. Filtration of the drying agent and concentration gave the crude product (0.36 g, m.p. 97–98 °C). The aqueous layer was extracted with ethyl acetate. After evaporation of the ethyl acetate, an additional 0.82 g of the product was obtained (0.42 g after washing with hexane and drying). The combined crude products were purified by column chromatography with hexane-ethyl acetate 95:5, 9:1, 4:1. The fractions containing the product were combined and concentrated. The residue was washed with hexane to give 2,3,6-trifluoro-5-hydroxybenzoic acid (0.26 g, m.p. 134–137 °C and 0.13 g, m.p. 138–142 °C (lit. 136–137 °C [11]). Yield of 8 in total 0.39 g, 41.4%. – MS (EI, 70 eV): m/z (%) = 193 (8), 192 (100) [M]+, 175 (39), 149 (8), 148 (20), 147 (25), 128 (8), 119 (23), 100 (42), 99 (30), 75 (9), 69 (12). – 1H NMR (300 MHz, CDCl3): δ = 10.02 (2H, bs, COOH, OH), 7.00 (1H, dt, J = 11.7, 7.95 Hz, Har). – 13C NMR (75.5 MHz, acetone-d6): δ = 161.73 (d, J = 13.25 Hz, COOH), 147.25 (ddd, J = 243.19, 13.55, 3.78 Hz, C–F), 145.39 (ddd, J = 246.92, 4.57, 3.25 Hz, C–F), 142.68 (ddd, J = 14.80, 9.74, 3.47 Hz), 141.43 (ddd, J = 246.53, 15.50, 5.51 Hz, C–F), 114.21 (t, J = 17.86 Hz), 108.76 (dd, J = 21.71, 3.59 Hz). – 19F NMR (282.5 MHz, acetone-d6): δ = −141.08 (ddd, J = 13.56, 7.91, 5.79 Hz), −142.58 (ddd, J = 22.11, 13.28, 11.51 Hz), −150.45 (ddd, J = 22.18, 7.77, 5.93 Hz). – IR (cm−1, KBr): ν = 3228, 1700, 1650, 1620, 1499, 1416, 1271, 1170, 946, 707.

4.8 1,2,4-Trifluoro-5-(methoxymethoxy)benzene (9)

2,4,5-Trifluorophenol (5 g, 0.031 mol), and N-ethyl-diisopropylamine (5.1 mL, 0.034 mol) in methylene chloride (23 mL) were placed in a reaction flask. The reaction mixture was cooled to 0 °C and MOMCl (2.6 mL, 0.034 mol) was added dropwise at 0 °C. The reaction mixture was stirred for 2 h at room temperature. The reaction mixture was poured into 3 m sodium hydroxide (60 mL) and extracted with diethyl ether. The organic layer was dried over anhydrous magnesium sulfate, and concentrated. The oily residue was obtained (5.68 g) and distilled under reduced pressure. 1,2,4-Trifluoro-5-(methoxymethoxy) benzene was obtained (3.67 g, 56.6%, GC/MS 97.5%, b.p. 70–72 °C/10 mm Hg (lit. 61–62 °C/8 mmHg, [11])). – nD20 = 1.4475 (lit. nD20 = 1.4407 [11]). – MS (EI, 70 eV): m/z (%) = 192 (17) [M]+, 162 (4), 161 (12), 148 (5), 147 (8),133 (4), 131 (11), 119 (20), 99 (4), 81 (13), 45 (100). – IR (cm−1, film): ν = 3800, 2963, 2840, 1523, 1431, 1218, 1157, 1076, 981, 853, 688.

4.9 2,3,5-trifluoro-6-hydroxybenzoic acid (3,5,6-trifluorosalicylic acid) (10)

1,2,4-Trifluoro-5-(methoxymethoxy)benzene (9, 1.8 g, 0.0094 mol) was placed in anhydrous diethyl ether (30 mL). under argon. The solution was cooled to −64 °C. 2.5 m n-butyllithium solution (0.0104 mol, 4.5 mL) was added in portions with a syringe through septum. The temperature of the reaction mixture rose to −61 °C. n-Butyllithium solution were added dropwise in five portions (4 × 1, 1 × 0.5 mL). After completion of the addition, the reaction mixture was stirred for 4 h at −65 °C. Gaseous carbon dioxide was passed through the reaction mixture for 1 h. The reaction mixture initially rose to −45 °C, then it maintained at approx. −65 °C. A copious amount of white precipitate was observed. The reaction mixture was allowed to warm to 0 °C, Water (10 mL) was added carefully. Aqueous layer was extracted with ether (after evaporation of the ether layer, residue was approx. 0.135 g). The aqueous layer was acidified with concentrated HCl to pH 1, stirred for 5 min. and extracted with ether. The combined ether layers were dried over anhydrous sodium sulfate. After evaporation of the solvent, a solid residue (1.49 g) was obtained. In order to remove the OMOM protection, the residue was dissolved in 6 mL of diethyl ether. A 2 n HCl (6 mL) was added and stirred at room temperature for 7 h (TLC control: benzene-acetone, 1:1). Diethyl ether layer was evaporated. Concentrated HCl (0.2 mL) was added to the residue and the mixture was stirred at room temperature for 12 h. The mixture was extracted with methylene chloride. Drying of the organic layer over anhydrous sodium sulfate and evaporation gave 2,3,5-trifluoro-6-hydroxybenzoic acid (10, 1.03 g, 57.2%, colorless solid, m.p. 170–178 °C, which was subjected without further purification to the methylation to 11. For analytical purposes, a sample of about 0.25 g was subjected to purification by column chromatography. The crude acid 10 was dissolved in ethyl acetate, silica gel (1 g) was added and evaporated to dryness. The sample applied on the gel was placed on the top of a chromatography column. Elution with hexane, hexane-ethyl acetate 9:1, 4:1, 1:1, ethyl acetate, methanol gave the product 10 (0.17 g) which was stirred with hexane at room temperature. The solid was filtered off. The filtrate was concentrated. 2,3,5-Trifluoro-6-hydroxybenzoic acid was obtained (10, 0.13 g, colorless solid, m.p. 173–174 °C. An analytical sample purified by column chromatography had an m.p. of 178–180 °C (lit. 173 °C [24]; 166–168 °C [11]; 165–167 °C [19]). – MS (EI, 70 eV): m/z (%) = 192 (39) [M]+, 175 (20), 174 (100), 147 (8), 146 (50), 119 (21), 118 (65), 99 (27), 75 (8), 69 (8), 68 (8), 45 (8). – 1H NMR (300 MHz, acetone-d6): δ = 11.39 (1H, bs, COOH), 7.61 (1H, td, J = 10.5, 7.2 Hz, Har). – 13C NMR (75.5 MHz, acetone-d6): δ = 170.03 (dd, J = 6.64, 3.32 Hz, COOH), 147.81 (dt, J = 14.345, 2.53 Hz), 147.26 (ddd, J = 243.808, 9.7, 3.7 Hz), 146.91 (ddd, J = 256.9, 14.25, 4.1 Hz), 143.10 (ddd, J = 239.66, 14.9, 10.7 Hz), 111.69 (td, J = 23.2, 1.58 Hz), 106.13 (dd, J = 11.627, 3.775 Hz). – 19F NMR (282.5 MHz, acetone-d6): δ = −138.50 (ddd, J = 22.18, 15.11, 7.2 Hz), −139.49 (ddd, J = 15.1, 10.6, 2.0 Hz), −148.89 (ddd, J = 22.0, 10.5, 2.26 Hz). – IR (cm−1, KBr): ν = 3440, 3100, 1682, 1644, 1496, 1445, 1269, 1178, 956, 714.

4.10 2,3,5-Trifluoro-6-methoxybenzoic acid (11)

To 2,3,5-trifluoro-6-hydroxybenzoic acid (10, 0.7 g, 0.0036 mol), acetone (12 mL), triturated anhydrous potassium carbonate (1.24 g, 0.009 mol), dimethyl sulfate (0.85 mL, 1.134 g, 0.009 mol) was added and refluxed for 7.5 and stirred 12 h at room temperature. The precipitate was then filtered off, the filtrate was concentrated and the crude methyl 2,3,5-trifluoro-2-methoxybenzoate (0.9 g) was obtained and subjected to basic hydrolysis with 40% potassium hydroxide (0.51 mL) in methanol (1 mL). The reaction mixture was stirred under reflux for 3 h. TLC analysis showed only partial hydrolysis (pH approx.7). Additional amount of 40% potassium hydroxide (0.3 mL) and methanol (1 mL) were added and the hydrolysis was continued for 1 h. TLC analysis showed complete hydrolysis of the ester. Methanol was evaporated, the residue was acidified with concentrated HCl. The product was extracted with ethyl acetate (3 × 15 mL) and methylene chloride (3 × 15 mL). Drying over anhydrous magnesium sulfate and concentration afforded 2,3,5-trifluoro-6-methoxybenzoic acid (11, 0.7 g, colorless solid, 93.2%, m.p. 86–89 °C). For analytical purposes a sample of approx. 0.25 g was subjected to purify by column chromatography. The crude 2,3,5-trifluoro-6-methoxybenzoic acid (11) was dissolved in warm benzene and applied to the top of the chromatography column. Elution with benzene, benzene-ethyl acetate 4:1, hexane-ethyl acetate 1:1 gave 2,3,5-trifluoro-6-methoxybenzoic acid (11, colorless solid, 0.22 g, m.p. 87–89 °C). – MS (EI, 70 eV): m/z (%) = 207 (6), 206 (62) [M]+, 189 (16), 188 (12), 177 (28), 175 (12), 174 (37), 160 (28), 159 (100), 149 (8), 147 (18), 146 (33), 143 (14), 135 (10), 133 (16), 131 (18), 130 (17), 119 (54), 118 (33), 113 (13), 99 (34), 81 (26), 80 (12), 75 (15), 69 (11), 68 (13), 45 (9). – HRMS (EI, 70 eV): m/z = 206.0193 (calcd 206.0191 for C8H5F3O3, [M]+). – 1H NMR (300 MHz, CDCl3): δ = 11.84 (1H, s, COOH), 7.14 (1H, td, J = 10.2, 7.5 Hz, Har), 4.02 (3H, d, J = 1.5 Hz, OCH3). – 13C NMR (75.5 MHz, CDCl3): δ = 166.91 (t, J = 3.29 Hz, COOH), 150.71 (ddd, J = 248.72, 9.11, 3.6 Hz), 145.70 (ddd, J = 248.94, 14.13, 10.5 Hz), 144.41 (ddd, J = 254.93, 14.8, 4.0 Hz), 142.15 (dt, J = 13.44, 3.7 Hz), 117.22 (dd, J = 13.70, 2.68 Hz), 108.71 (ddd, J = 24.59, 21.86, 1.27 Hz), 62.78 (dd, J = 5.51, 0.53 Hz, OCH3). – 19F NMR (282.5 MHz, CDCl3): δ = −130.1 (t, J = 12 Hz), −138.42 (dd, J = 22.18, 9.8 Hz), −140.54 (ddd, J = 21.89, 14.13, 7.49 Hz). IR (cm−1, KBr): ν = 3430, 3073, 1713, 1498, 1410, 1304, 1006, 868, 725.

4.11 2,3,5-Trichloro-6-methoxybenzoic acid (tricamba)

To 2,3,5-trichloro-6-hydroxybenzoic acid (4.8 g, 0.02 mol), acetone (67 mL), triturated anhydrous potassium carbonate (6.9 g, 0.05 mol), dimethyl sulfate (4.7 mL, 6.3 g, 0.05 mol) was added and refluxed for 7.5 h. The precipitate was filtered off, the filtrate was concentrated and the crude methyl 2,3,5-trichloro-2 methoxybenzoate (6.03 g) was obtained and subjected to basic hydrolysis. with 40% sodium hydroxide (3.1 mL) in methanol (5.6 mL). The reaction mixture was stirred under reflux. After about 1 h the reaction mixture solidified. Water (about 5 mL) was added and heating was continued for 5 h under reflux. Methanol was evaporated, the residue was acidified with concentrated HCl. The product was extracted with ethyl acetate. After drying over anhydrous magnesium sulfate and concentration, the residue (8.13 g) was purified by column chromatography. Elution with hexane, hexane-ethyl acetate 4:1, 1:1, acetone gave 4.1 g of a colorless solid, which was washed with hexane and dried. 2,3,5-Trichloro-6-methoxybenzoic acid was obtained (3.0 g, 58.8%, GC 99.5%, HPLC 98.3%, m.p. 135–138 °C (lit. 137–139 °C [25])) – GC-MS (EI, 70 eV): m/z (%) = 258 (29), 256 (92), 254 (95) [M]+, 241 (19), 239 (43), 23 (15), 237 (39), 236 (12), 227 (29), 226 (18), 225 (37), 224 (39), 222 (33), 213 (18), 212 (15), 211 (51), 210 (40), 209 (100), 208 (410, 207 (92), 196 (25), 194 (26), 185 (22), 183 (39), 181 (29), 176 (19), 175 (31), 169 (30), 167 (30), 149 (17), 147 (31), 146 (17), 145 (25), 143 (17), 133 (35), 131 (47), 111 (150, 109 (27), 108 (27), 107 (17), 97 (24), 96 (37), 74 (180), 73 (22), 61 (34). – 1H NMR (300 MHz, CDCl3): δ = 10.19 (1H, bs, COOH), 7.60 (1H, s, Ph), 3.98 (s, 3H, OCH3). – 13C NMR (75.5 MHz, CDCl3): δ = 168.94 (s, COOH), 152.34 (s), 132.33 (s), 130.54 (s), 129.22 (s), 128.07 (s), 127.40 (s), 62.62 (s, OCH3). – IR (cm−1, KBr): ν = 2949, 1713, 1649, 1461, 1421, 1243, 1136, 1014, 875, 685.

4.12 Herbicidal bioassay, pre- and post-emergence experiments

Herbicidal activity was evaluated using different weed species in pot experiments under controlled conditions. Polyethylene pots, 3.5 L capacity, were filled with 0.75 kg of soil (physicochemical characteristic: sandy clay, pH/KCl/6.7; organic matter 2.8%) and were wetted with water. Seeds of the weed species were planted in soil (0.5 cm depth). Plants were grown to the two-leaf stage under normal glasshouse propagation conditions (temperature, 20 ± 5 °C; lighting, 14 h photoperiod of daylight supplemented by lamps, 400 W). The pots were watered overhead. The test compounds were dissolved in an appropriate volume of acetone-water solution (1:3) with the addition of Tween 20 (0.05% v/v) to give the required dose of the tested substance. All treatments were applied as a pre-emergence or post-emergence spray at a volume rate of 300 L ha−1 using track laboratory sprayer (nozzle TeeJet60, pressure 0.2 MPa). There were three replicate pots per treatment arranged in a randomized block design. Pots after spraying were transferred to the growth chamber (temperature, day/night 20/15 °C; lighting, 16 h photoperiod, white fluorescent tubes giving 200 μmol m−2 s−2 PAR [photosynthetic active radiation]). A visual assessment of phytotoxicity separately for each species was made (18 or 25 days after treatment for pre- and post-emergence experiments) as a percentage compared to the untreated plants. Results (degree [%] of a visual observation of a phytotoxicity weed damage for a weed species used against control in doses: 2000, 300, 150 g ha−1) are expressed as a fraction: a numerator means a result (in %) of a pre-emergence (via soil) treatment, a denominator means a result (in %) of a post-emergence (on leaves) treatment.

4.13 Fungistatic bioassay in vitro

Fungitoxicity of the tested compounds against phytopathogenic fungi was assessed in vitro using agar growth medium poison technique. PDA media in 100 mm Petri plates containing the acetone solutions of the tested compounds in the defined concentrations were infected with agar disks with thin mycelium of fungi cultures, and allowed the solvent to evaporate. Linear growth of each colony was determined after 3–5 days. The effect of each compound on mycelial growth was assessed by calculating the percentage of growth reduction of a fungus colony; percentage of linear growth reduction = ([colony diameter of a control plate − colony diameter of a tested plat]/[colony diameter of a control plate]) × 100.


Corresponding author: Jerzy Zakrzewski, Łukasiewicz Research Network, Institute of Industrial Organic Chemistry, Annopol 6, 03-236Warsaw, Poland, E-mail:

Funding source: Polish Ministry of Science and Higher Education

Award Identifier / Grant number: 4180/E-142/S/2019

  1. Author contributions: All the authors have accepted responsibility for the entire content of this submitted manuscript and approved submission.

  2. Research funding: This work was supported by the Polish Ministry of Science and Higher Education, 4180/E-142/S/2019.

  3. Conflict of interest statement: The authors declare no conflicts of interest regarding this article.

5 Supporting information

MS, 1H, 13C, 19F NMR, and IR spectra are given as supplementary material available online (https://doi.org/10.1515/znb-2020-0179).

References

1. Filler, R. Biologically-active fluorochemicals. J. Fluor. Chem. 1986, 33, 361–375; https://doi.org/10.1016/s0022-1139(00)85281-1.Search in Google Scholar

2. Foye, W. O., Lemke, T. L., Williams, D. A. Foye’s Principles of Medicinal Chemistry, 8th ed.; Lippincott Williams & Wilkins: Philadelphia, 2008; p 50.Search in Google Scholar

3. Ando, A., Kumadaki, I. Progress on the syntheses of fluorine analogs of natural porphyrins potentially useful for the diagnosis and therapy of certain cancers. J. Fluor. Chem. 1999, 100, 135–146; https://doi.org/10.1016/s0022-1139(99)00197-9.Search in Google Scholar

4. Biffinger, J. C., Kim, H. W., DiMagno, S. G. The polar hydrophobicity of fluorinated compounds. Chembiochem 2004, 5, 622–627; https://doi.org/10.1002/cbic.200300910.Search in Google Scholar

5. Huras, B., Zakrzewski, J., Kiełczewska, A., Krawczyk, M. Herbicidal and fungistatic properties of fluorine analogs of phenoxyacetic herbicides. J. Fluor. Chem. 2017, 202, 76–81; https://doi.org/10.1016/j.jfluchem.2017.08.013.Search in Google Scholar

6. de Assis Alves, T., Fontes Pinheiro, P., Praça-Fontesa, M. M., Fonseca Andrade-Vieira, L., Barelo Corrêa, K., de Assis Alves, T., da Cruz, F. A., Lacerda Júnior, V., Ferreira, A., Bastos Soares, T. C. Toxicity of thymol, carvacrol and their respective phenoxyacetic acids in Lactuca sativa and Sorghum bicolor. Ind. Crop. Prod. 2018, 114, 59–67; https://doi.org/10.1016/j.indcrop.2018.01.071.Search in Google Scholar

7. Coats, J. R., Klimavicz, J. S., Norris, E. J., Bessette, S. M., Lindsay, A. D., Kittrich Corp., Iowa State University Research Foundation, Inc.). Monoterpenoid/ phenylpropanoid-containing compounds and methods of their making and use as herbicides. WO2018/39390. International publication day: March 01, 2018.Search in Google Scholar

8. Manos-Turvey, A., Bulloch, E. M. M., Rutledge, P. J., Baker, E. N., Lott, J. S., Payne, R. J. Inhibition studies of Mycobacterium tuberculosis salicylate synthase (MbtI). ChemMedChem 2010, 5, 1067–1079; https://doi.org/10.1002/cmdc.201000137.Search in Google Scholar

9. Zakrzewski, J., Krawczyk, J. Reactions of nitroxides XIV. Analogs of phenoxy carboxylic herbicides based on the piperidine scaffold; unexpected fungicidal activity of the 2-(1-oxyl-2,2,6,6-tetramethylpiperidin-4-yl)oxybutanoic acid. Heterocycl. Commun. 2014, 20, 89–91; https://doi.org/10.1515/hc-2013-0169.Search in Google Scholar

10. Luliński, S., Serwatowski, J., Zaczek, A. Long-range effects in the metalation/boronation of functionalized, 1,4-dihalobenzenes. Eur. J. Org Chem. 2006, 5167–5173.10.1002/ejoc.200600541Search in Google Scholar

11. Marzi, E., Górecka, J., Schlosser, M. The regioexhaustive functionalization of difluorophenols and trifluorophenols through organometallic intermediates. Synthesis 2004, 1609–1618.10.1055/s-2004-829110Search in Google Scholar

12. Snieckus, V. Directed ortho metalation. Tertiary amide and O-carbamate directors in synthetic strategies for polysubstituted aromatics. Chem. Rev. 1990, 90, 879–933; https://doi.org/10.1021/cr00104a001.Search in Google Scholar

13. Winkle, M. R., Ronald, R. C. Regioselective metalation reactions of some substituted (Methoxymethoxy)arenes. J. Org. Chem. 1982, 47, 2101–2108; https://doi.org/10.1021/jo00132a022.Search in Google Scholar

14. Ronald, R. C., Winkle, M. R. Regioselective metallations of (methoxymethoxy)arenes. Tetrahedron 1983, 39, 2031–2042; https://doi.org/10.1016/s0040-4020(01)91921-0.Search in Google Scholar

15. Katritzky, A. R., Lang, H., Lan, X. A new and efficient conversion of alkylphenols into the corresponding substituted salicylic acids. Synth. Commun. 1993, 23, 1175–1182; https://doi.org/10.1080/00397919308018594.Search in Google Scholar

16. Comins, D. L., Saha, J. K. Asymmetric synthesis of a key camptothecin intermediate from 2-fluoropyridine. Tetrahedron Lett. 1995, 36, 7995–7998.10.1016/0040-4039(95)01665-5Search in Google Scholar

17. Simas, A. B. C., Coelho, A. L., Costa, P. R. R. Regioselective lithiation of resorcinol derivatives: synthesis of mono O-MOM- and O-benzylresorcinols prenylated at C-2 or C-4 positions. Synthesis 1999, 1017–1021; https://doi.org/10.1055/s-1999-3487.Search in Google Scholar

18. Lau, S. Y. W., Keay, B. A. A highly efficient strategy for the synthesis of 3-substituted salicylic acids by either directed ortho-lithiation or halogen-metal exchange of substituted mom protected phenols followed by carboxylation. Can. J. Chem. 2001, 79, 1541–1545; https://doi.org/10.1139/cjc-79-11-1541.Search in Google Scholar

19. Shchegol’kov, E. V., Shchur, I. V., Burgart, Y. V., Saloutin, V. I., Trefilova, A. N., Ljushina, G. A., Solodnikov, S. Yu., Markova, L. N., Maslova, V. V., Krasnykh, O. P., Borisevich, S. S., Khursan, S. L. Polyfluorinated salicylic acid derivatives as analogs of known drugs: synthesis, molecular docking and biological evaluation. Bioorg. Med. Chem. 2017, 25, 91–99.10.1016/j.bmc.2016.10.014Search in Google Scholar PubMed

20. Marvel, C. S., Porter, P. K. Monochloromethyl Ether. Organic Syntheses, Coll, Vol. 1; John Wiley & Sons: New York, 1941; p 377.Search in Google Scholar

21. Ladd, D. L., Weinstock, J. Improved synthesis of fluoroveratroles and, fluorophenethylamines via organolithium reagents. J. Org. Chem. 1981, 46, 203–206; https://doi.org/10.1021/jo00314a054.Search in Google Scholar

22. Burdon, J., Hollyhead, W. B. Aromatic polyfluoro-compounds. Part XXVI. Nucleophilic replacement reactions of the tetrafluorobenzenes. J. Chem. Soc. 1965, 6326–6328; https://doi.org/10.1039/jr9650006326.Search in Google Scholar

23. Shtark, A. A., Tchuikova, T. V., Shteingarts, V. D. J. Org. Chem. USSR 1984, 20, 960–967.Search in Google Scholar

24. Petrova, T. D., Kann, L. I., Barkhash, V. A., Yakobson, G. G. Polyfluorinated heterocyclic compounds. Chem. Heterocycl. Compd. 1969, 5, 574–575.10.1007/BF00957359Search in Google Scholar

25. Eckstein, Z., Alster, K., Domańska, H. Porównanie metod otrzymywania kwasu 3,6-dwuchloro-2-metoksybenzoesowego (dikamby). Przem. Chem. 1979, 58, 533–536.Search in Google Scholar


Supplementary Material

The online version of this article offers supplementary material (https://doi.org/10.1515/znb-2020-0179).


Received: 2020-10-21
Accepted: 2020-12-20
Published Online: 2021-02-18
Published in Print: 2021-04-27

© 2021 Walter de Gruyter GmbH, Berlin/Boston

Downloaded on 25.4.2024 from https://www.degruyter.com/document/doi/10.1515/znb-2020-0179/html
Scroll to top button