Next Article in Journal
Chemical Diversity and Biological Activity of Secondary Metabolites Isolated from Indonesian Marine Invertebrates
Next Article in Special Issue
Mono- and Diamination of 4,6-Dichloropyrimidine, 2,6-Dichloropyrazine and 1,3-Dichloroisoquinoline with Adamantane-Containing Amines
Previous Article in Journal
An RP-LC-UV-TWIMS-HRMS and Chemometric Approach to Differentiate between Momordicabalsamina Chemotypes from Three Different Geographical Locations in Limpopo Province of South Africa
Previous Article in Special Issue
Synthesis of Novel Tryptamine Derivatives and Their Biological Activity as Antitumor Agents
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Communication

Practical Synthesis of 2-Iodosobenzoic Acid (IBA) without Contamination by Hazardous 2-Iodoxybenzoic Acid (IBX) under Mild Conditions

1
Department of Medical Bioscience, Nagahama Institute of Bio-Science and Technology, 1266, Tamuracho Nagahama-shi, Shiga 526-0829, Japan
2
College of Pharmaceutical Sciences, Ritsumeikan University, 1-1-1 Nojihigashi, Kusatsu, Shiga 525-8577, Japan
3
Faculty of Pharmacy, Meijo University, 150 Yagotoyama, Tempaku-ku, Nagoya 468-8503, Japan
*
Authors to whom correspondence should be addressed.
Molecules 2021, 26(7), 1897; https://doi.org/10.3390/molecules26071897
Submission received: 28 February 2021 / Revised: 16 March 2021 / Accepted: 24 March 2021 / Published: 27 March 2021

Abstract

:
We report a convenient and practical method for the preparation of nonexplosive cyclic hypervalent iodine(III) oxidants as efficient organocatalysts and reagents for various reactions using Oxone® in aqueous solution under mild conditions at room temperature. The thus obtained 2-iodosobenzoic acids (IBAs) could be used as precursors of other cyclic organoiodine(III) derivatives by the solvolytic derivatization of the hydroxy group under mild conditions of 80 °C or lower temperature. These sequential procedures are highly reliable to selectively afford cyclic hypervalent iodine compounds in excellent yields without contamination by hazardous pentavalent iodine(III) compound.

1. Introduction

Cyclic hypervalent iodine reagents, such as 2-iodosobenzoic acid (IBA) and 2-iodoxybenzoic acid (IBX) are nonmetallic green oxidants with excellent recyclability [1,2,3]. IBA and IBX can be regenerated from 2-iodobenzoic acid (2-IB) without requiring an external ligand except for water in this reoxidation step. This is because the carboxy group adjacent to the iodine atom serves as an endogenous ligand. Recently, IBA, a representative trivalent cyclic hypervalent iodine oxidant, has been used as a catalyst and reagent in various reactions, i.e., decarboxylative alkynylation [4,5], decarboxylative acylarylation [6], oxyalkenylation [7], oxyarylation [8], oxidative C–H arylation [9], C–H hydroxylation [10], C-H oxidation [11,12], ring-opening hydrazination [13], and asymmetric intramolecular α-cyclopropanation [14]. IBA derivatives containing OAc [15,16,17,18,19,20,21,22,23,24,25,26,27,28], OMe [29,30,31,32], OTs [33,34,35], OTf [36,37], Cl [38,39,40,41,42], F [43,44,45], CN [46], N3 [47,48,49,50,51,52,53,54], CF3 [55,56], OCOCF3 [57], alkynyl [58,59,60,61,62] ligands instead of the hydroxy group have also found application in various reactions (Figure 1).
Although IBAs can be prepared from 2-IBs by existing methods (Figure 2) [63,64,65,66], the development of a safer and more efficient method for their synthesis is highly desirable. As shown in Figure 2, IBAs can be further oxidized to pentavalent cyclic hypervalent IBXs [67], which need to be prevented for the preparation of IBAs [68,69,70], mainly due to the explosive nature of IBXs on heating and impact, while IBXs are useful in some small-scale reactions [71,72,73,74,75,76]. Thus, contamination by IBX in the IBA products should be avoided for long-term safe storage or large-scale use.
In recent decades, many reactions using Oxone®, which is an inexpensive and commercially available stable triple salt (2KHSO5/KHSO4/K2SO4), have been developed for practical synthetic purposes [77]. In particular, the use of Oxone® as a re-oxidant for pentavalent hypervalent iodine reagents is drawing attention for catalytic oxidation reactions [78,79,80,81,82,83,84,85,86,87,88,89,90,91,92]. The reaction systems for alcohol oxidations [78,79,80,81,82,83,84] involving in situ generated active hypervalent iodine(V) species are optimized on the basis of the preparative conditions of IBX from 2-IB at 70 °C [93]. Meanwhile, oxidative lactonizations from modified 2-IBs using Oxone® occur at room temperature [85,86,87]. In this context, the generation of nonexplosive trivalent cyclic hypervalent iodines, i.e., IBA and its analogs, using Oxone® can be expected to provide a convenient and safe synthetic procedure; however, to best of our knowledge, the selective preparation of IBAs using Oxone® has not been reported so far.
Recently, we reported that IBAs generated in a reaction system containing 2-IB and Oxone® play a catalytic role in the selective oxidation of alkoxybenzenes to p-quinones [94]. This resulted in the development of the practical method herein reported for the selective preparation of IBAs under mild conditions.

2. Results and Discussion

2.1. Selective Synthesis of IBA and Its Analogs

We started our investigation on the selective preparation of IBAs by evaluating the solvent effects on the oxidation of 2-IB 1a using 1.0 equivalent of Oxone® to obtain IBA 2a as a representative compound, and the results are summarized in Figure 3. First, the reaction in water led to the successful production of IBA 2a in 82% yield (Figure 3, entry 2), whereas IBA 2a was not produced in organic solvent in the absence of water (no reaction because Oxone® was not dissolved) (see entry 1). This result indicates that water plays an essential role in the formation of IBA. Therefore, we assumed that an aqueous system similar to the selective formation of p-quinone from alkoxybenzenes catalyzed by 2-IB 1a with Oxone® [94] could be suitable for the present reaction. We then investigated in detail the effect of a series of organic solvents on the aqueous preparation of IBA 2a using Oxone®.
Various water-miscible organic solvents were investigated to dissolve 2-IB 1a in this reaction. The preparation of IBA 2a using acetonitrile (MeCN) in aqueous condition (Figure 3, entry 3) was similar to that performed in the absence of organic solvents (Figure 3, entry 2). Tetrahydrofuran (THF), dioxane, benzene and N,N-dimethylformamide (DMF) were also examined, finding that the use of highly polar dioxane and DMF led to excellent yields of IBA 2a (Figure 3, entries 5 and 7), whereas benzene, the least polar solvent among these aprotic solvents, significantly reduced the yield of the desired product (Figure 3, entry 6). The reason for this very low yield IBA formation was interpreted as being due to that benzene forms a two-phase system and interferes with the dissolution of 2-IB into water. Protic solvents such as MeOH, EtOH and 2,2,2-trifuoroethanol (TFE) gave IBA 2a in high yields; however, they also worked as a ligand for IBA, causing the formation of very small amounts of ligand-exchanged byproducts 3ac (Figure 3, entries 8–10). The white solid IBA 2a obtained after the water and acetone washings did not contain any other byproducts. Although this result indicated that protic organic solvents were not suitable for the selective preparation of IBAs, it also revealed that the IBA hydroxyl group could undergo substitution reactions under mild conditions (vide infra). The yields indicated in Figure 3 are almost equal to the conversion of 2-IB 1a.
Next, we investigated the substrate scope for the synthesis of IBAs using Oxone® under aqueous conditions with MeCN, and the results are shown in Figure 4. MeCN was used as a component of the solvent to dissolve substrates 1. By oxidation of 5-substituted 2-IBs, IBAs 2bd containing fluoro-, chloro-, and bromo-substituents were smoothly obtained in excellent yields from the corresponding halo-substituted 2-IBs 1bd. From 2-IBs 1ej with electron-donating groups such as methyl-, methoxy-, and acethoxy-substituents (1eg) and electron-withdrawing groups such as trifluoromethyl-, nitro- and cyano-substituents (1hj), the desired IBAs 2ej were also produced in good yields. However, 2-IB bearing a hydroxy-substituent 1k afforded the desired product 2k in a moderate yield under the same conditions. In the oxidation of 4-substituted 2-IBs, fluoro-, chloro-, bromo-, trifluoromethyl-, and carboxy-substituted IBAs 2lp were obtained in excellent yields from the corresponding 2-IBs 1lp. In addition, the oxidized products of 4,5-disubstituted 2-IBs containing difluoro-substituents 2q and dimethoxy-substituents 2r were obtained in high yields. Meanwhile, with regard to 3-substituted 2-IBs, the reaction of methyl-substituted 1t with a slight excess of Oxone® afforded the expected IBA 2t in a good yield, whereas the yield of bromo-substituted IBA 2s was lower even at the elevated temperature and in the presence of a large excess of Oxone®. Steric effects are probably important in the formation of the cyclic λ3-iodanes. Indeed, the presence of a substituent at the ortho position of the iodine atom (3-position) interfered in the synthesis of the corresponding product for 3-bromo-substituted 2-IB 1s. In the case of 6-substituted 2-IBs, fluoro-substituted IBA 2u and methyl-substituted 2v were obtained in good yields. Finally, the reaction of 3-iodonaphthalene-2-carboxylic acid 3 under the present conditions led to the expected tricyclic hypervalent iodine compound 4 in an excellent yield (Scheme 1).

2.2. IBAs Synthesis Using Ferric Effect

As mentioned in Section 2.1, our present method can selectively afford trivalent cyclic hypervalent iodine IBA at room temperature without contamination by pentavalent iodine byproduct. The mild conditions used contributed favorably to this product selectivity. Interestingly, we further found that iron ion in tap water (TW), which contained iron ion (5.8 μM or less), contribute to the IBA formation, whereas calcium and magnesium ions as main minerals in TW do not affect the selectivity. Indeed, IBA 2a was selectively produced from 2-IB 1a by heating even at 100 °C in DW containing 5 mol% FeCl3 (Scheme 2, left). On the other hand, IBX was instead formed as a main product in the absence of FeCl3 in deionized water (DW) [93]. Other ferric salts such as Fe(NO3), Fe(OTf)3, and FeSO4 had similar effects. In addition, it was found that pentavalent IBX 5 was converted to IBA 2a in the presence of a catalytic amount of FeCl3 at 100 °C (Scheme 2, right), while the formation of unidentified high- and low-polar decomposition products were detected in the water and the acetone washing solution, respectively. Here, 2-IB 1a was not produced. These results would indicate that overoxidation of IBA 2a to pentavalent IBX 5 was strongly prevented by ferric salts. Thus, the effect of the metal ion in the decomposition of hazardous IBX 5 is also a significant key factor to ensure the safety for our trivalent cyclic hypervalent iodine synthesis under heating conditions.
The reaction time in the synthesis of IBAs was significantly shortened by heating. In the investigation of the heating conditions for the synthesis of IBA 2a in 0.2 M 2-IB 1a in the presence of 2.5 mol% FeCl3 for 10 min, the required amount of Oxone® and the reaction temperature were thus optimized to 60 °C (Figure 5a) and 1.0 equivalent (Figure 5b), respectively. The yield of IBA 2a was very sensitive to the reaction time, which dropped from 83% for 10 min to 70% for 1 h. IBA may be decomposed to small molecules in the presence of excess Oxone®; it has been reported that Oxone® causes oxidative cleavage of the aromatic ring [95]. Without Oxone®, we also confirmed that IBA 2a was hardly decomposed under the conditions of Scheme 2 in the presence of 1.0 equivalent of H2SO4 and 10 mol% FeCl3 at 100 °C for 10 min, while only 64% of IBA 2a was recovered by replacing H2SO4 in the presence of Oxone® under the same conditions. Thus, the excess uses of Oxone® and performing the reaction at high temperature would decrease the IBA yield as shown in Figure 5a,b.
This optimized heating method could be applied to the synthesis of IBAs 2aj (Scheme 3). 2-IB 1a as well as the substrates 1bd and 1gj that are tolerable to over-oxidation at this temperature were successfully converted to the desired IBAs 2ad and 2gj in high yields. However, the transformation of 2-IB having an electron-rich functional group, i.e., methoxy-substituted 2-IB 1f, resulted in low yield of the corresponding IBA 2f due to the formation of 2-carboxy-p-benzoquinone by the oxidation with Oxone®. Therefore, in order to apply the heating conditions, the stability of the product to oxidation must be considered.

2.3. IBA Derivatives

As previously mentioned, when IBA 2a was synthesized in an aqueous solution with alcohols, alkoxy-substituted derivatives 3ac were obtained as byproducts by substitution of the hydroxyl ligand of IBA 2a (see Figure 3), implying the potential of the solvolytic ligand exchange of IBA 2a under mild conditions. For the ligand derivatization of IBAs, the water molecule is an obstacle because the ligand exchanges of the IBA hydroxy group are reversible. Thus, molecular sieves with a pore diameter of 3 Å (MS3Å) was used for the solvolytic functionalization of IBA 2a in dehydrated protic solvent (Figure 6). The quantitative derivatization to benziodoxole methoxide (IB-OMe) 6a was achieved by heating IBA 2a at 60 °C in MeOH (Figure 6, entry 1). Upon treatment at 80 °C, benziodoxole ethoxide (IB-OEt) 6b and benziodoxole 2,2,2-trifluoroethoxide (IB-OCH2CF3) 6c were also produced in high yields by the ligand exchange reaction with EtOH and TFE, respectively (Figure 6, entries 2 and 3). Benziodoxole n-propoxide (IB-OnPr) 6d was obtained in 98% yield using nPrOH at 70 °C (Figure 6, entry 4), and benziodoxole isopropoxide (IB-OiPr) 6e was produced in 52% yield at 60 °C in the presence of iPrOH (Figure 6, entry 5). In the cases of IB-OnPr 6d and IB-OiPr 6e, the temperature control was essential to suppress the formation of a 2-IB-IBA condensate as a byproduct; here, the formation of 2-IB 1a can be explained in terms of the alcohol oxidation by IBA. It is known that secondary alcohols are readily oxidized by IBA [83]. No unwanted byproduct was found during the transformation to benziodoxole hexafluoroisopropoxide (IB-OCH(CF3)2) 6f using hexafluoroisopropanol (HFIP) at 80 °C (Figure 6, entry 6), which is most likely due to the stability of HFIP against oxidation. Indeed, the condensate between 2-IB and IBA appeared during the reaction for benziodoxole n-buthoxide (IB-OnBu) 6g using an oxidizable primary alcohol, nBuOH, at 80 °C, whereas such byproduct was not observed in the synthesis of benziodoxole tert-buthoxide (IB-OtBu) 6h using tBuOH as a solvent inert to oxidation. Nevertheless, IB-OnBu 6g could be selectively obtained by heat treatment at 60 °C without the formation of the condensate.
Using AcOH as a solvent, the solvolytic method was further applied to the synthesis of benziodoxole acetate (IB-OAc) 7a and its analogs (R-IB-OAc) 7bi from the corresponding IBAs 2ai (Scheme 4). IB-OAc 7a was easily produced in good yield by ligand exchange of IBA 2a with AcOH at room temperature. Similarly, these transformations successfully afforded R-IB-OAc 7bd containing fluoro-, chloro-, and bromo-substituents; 7eg with electron-donating methyl-, methoxy-, and acethoxy-groups; and 7h and 7i bearing an electron-withdrawing trifluoromethyl- and nitro-substituent, respectively.

3. Materials and Methods

3.1. General Information

Substrates 1i [96], 1k [97], 1n [98], 1o [98], 1p [99], 1q [100], 1s [101], 1t [98], 1v [102], and 3 [65] were prepared by Sandmeyer reaction of the corresponding anthranilic acids. Substrate 1g [84] was synthesized by acetylation of compound 1k. Substrate 1j [103] is derived from 5-bromo-anthranilic acid methyl ester. 1H, 13C, and 19F nuclear magnetic resonance (NMR) spectra were recorded on ECS 400 and ECX 500 NMR spectrometers (JEOL Ltd., Tokyo, Japan) using deuterated dimethyl sulfoxide (DMSO-d6) or chloroform (CDCl3) as a solvent. Chemical shifts (δ) are reported in parts per million (ppm) relative to tetramethylsilane (δ = 0 ppm) as an internal standard for 1H and 13C NMR spectra and hexafluoroacetone (δ = −84.6 ppm) as an internal standard for 19F NMR spectra. Coupling constants (J) are reported in Hertz (Hz), and the multiplicity is reported according to the following convention: singlet (s), doublet (d), double doublet (dd), double double doublet (ddd), double triplet (dt), triplet (t), triple doublet (td), quartet (q), quintet (quin), sextet (sext), septet (sep), and multiplet (m). Data are reported as follows: Chemical shift (number of protons, multiplicity, coupling constants). Infrared (IR) spectra were recorded on a JASCO FT/IR-4200 spectrometer (JASCO Co., Tokyo, Japan) on diffuse reflectance method using KBr powder. Absorptions are expressed in reciprocal centimeter (cm−1). High resolution mass spectra (HRMS) obtained by the direct analysis in real time (DART) method were recorded on a Thermo Scientific Exactive Plus Orbitrap (Thermo Fisher Scientific, Inc., Waltham, MA, USA).

3.2. Synthesis of IBA Analogues

3.2.1. General Procedure for the Synthesis of IBAs 2av and 4

To a solution of 2-IBs (1.0 mmol) in MeCN (5 mL) was added Oxone® (738 mg, 1.2 mmol) and H2O (TW for Figure 3; Figure 4, Scheme 1 or DW for Scheme 2; Scheme 3, 5 mL). After the mixture was stirred at room temperature for the appropriate time (see Figure 4 and Scheme 1), the product was filtered under reduced pressure. The residue was washed with water and acetone to obtain the corresponding IBAs 2av and 4 (see Supplementary Materials for 1H NMR spectroscopic data) as a white powder.

3.2.2. 1-Hydroxy-1λ3-benzo[d][1,2]iodaoxol-3(1H)-one (2a)

1H NMR (400 MHz, DMSO-d6): δ 7.72 (1H, td, J = 7.3, 0.9 Hz, H5), 7.86 (1H, d, J = 8.2 Hz, H3), 7.97 (1H, ddd, J = 8.7, 7.4, 1.8 Hz, H4), 8.03 (1H, dd, J = 7.8, 1.4 Hz, H6), 8.08 (1H, s) ppm. 13C NMR (100 MHz, DMSO-d6): δ 120.3 (C2), 126.2 (C3), 130.3 (C5), 131.0 (C1), 131.4 (C6), 134.4 (C4), 167.7 (COOH) ppm. IR (ATR, KBr): ν 2936 (OH), 1616 (C=O), 1566 (C=O) cm−1. Mp: 243–244 °C. 1H and 13C NMR data are consistent with those reported in the literature [33].

3.2.3. 5-Fluoro-1-hydroxy-1λ3-benzo[d][1,2]iodaoxol-3(1H)-one (2b)

1H NMR (500 MHz, DMSO-d6): δ 7.76 (1H, dd, J = 8.3, 2.6 Hz), 7.79–7.88 (2H, m), 8.21 (1H, s) ppm. 13C NMR (125 MHz, DMSO-d6): δ 114.2, 117.3 (d, J = 22.7 Hz), 121.7 (d, J = 23.8 Hz), 128.3 (d, J = 8.4 Hz), 134.1 (d, J = 7.2 Hz), 163.9 (d, J = 246.8 Hz), 166.4 (d, J = 2.4 Hz) ppm. 19F NMR (470 MHz, DMSO-d6): δ −116.2 (dt, J = 5.7, 8.6 Hz) ppm. IR (ATR, KBr): ν 2904 (OH), 1635 (C=O), 1577 (C=O) cm−1. Mp: 241–242 °C. 1H and 13C NMR data are consistent with those reported in the literature [104].

3.2.4. 5-Chloro-1-hydroxy-1λ3-benzo[d][1,2]iodaoxol-3(1H)-one (2c)

1H NMR (400 MHz, DMSO-d6): δ 7.81 (1H, d, J = 8.7 Hz), 7.95 (1H, d, J = 2.3 Hz), 8.03 (1H, dd, J = 8.7, 2.3 Hz), 8.28 (1H, s) ppm. 13C NMR (100 MHz, DMSO-d6): δ 118.6, 128.1, 130.3, 133.5, 134.0, 135.8, 166.3 ppm. IR (ATR, KBr): ν 2905 (OH), 1624 (C=O), 1560 (C=O) cm−1. Mp: 294–295 °C. 1H and 13C NMR data are consistent with those reported in the literature [105].

3.2.5. 5-Bromo-1-hydroxy-1λ3-benzo[d][1,2]iodaoxol-3(1H)-one (2d)

1H NMR (400 MHz, DMSO-d6): δ 7.74 (1H, d, J = 8.7 Hz), 8.07 (1H, d, J = 2.3 Hz), 8.15 (1H, dd, J = 8.7, 2.3 Hz), 8.27 (1H, s) ppm. 13C NMR (100 MHz, DMSO-d6): δ 119.5, 124.2, 128.3, 133.3, 133.7, 136.8, 166.2 ppm. IR (ATR, KBr): ν 2884 (OH), 1617 (C=O), 1557 (C=O) cm−1. Mp: 236–238 °C. 1H and 13C NMR data are consistent with those reported in the literature [106].

3.2.6. 1-Hydroxy-5-methyl-1λ3-benzo[d][1,2]iodaoxol-3(1H)-one (2e)

1H NMR (400 MHz, DMSO-d6): δ 2.48 (3H, s), 7.70 (1H, d, J = 8.2 Hz), 7.79 (1H, dd, J = 8.7, 1.8 Hz), 7.85 (1H, s), 8.01 (1H, s) ppm. 13C NMR (100 MHz, DMSO-d6): δ 20.1, 116.7, 125.9, 131.3, 131.4, 135.2, 140.4, 167.7 ppm. IR (ATR, KBr): ν 3054 (OH), 1622 (C=O), 1569 (C=O) cm−1. Mp: 212–214 °C. 1H and 13C NMR data are consistent with those reported in the literature [105].

3.2.7. 1-Hydroxy-5-methoxy-1λ3-benzo[d][1,2]iodaoxol-3(1H)-one (2f)

1H NMR (400 MHz, DMSO-d6): δ 3.89 (3H, s), 7.52 (1H, d, J = 2.7 Hz), 7.55 (1H, dd, J = 8.7, 2.8 Hz), 7.67 (1H, d, J = 9.2 Hz), 8.04 (1H, s) ppm. 13C NMR (100 MHz, DMSO-d6): δ 55.8, 108.9, 114.8, 121.5, 127.0, 132.9, 161.4, 167.4 ppm. IR (ATR, KBr): ν 2953 (OH), 1620 (C=O), 1577 (C=O) cm−1. Mp: 217–218 °C. 1H and 13C NMR data are consistent with those reported in the literature [105].

3.2.8. 1-Hydroxy-3-oxo-1,3-dihydro-1λ3-benzo[d][1,2]iodaoxol-5-yl acetate (2g)

1H NMR (500 MHz, DMSO-d6): δ 2.33 (3H, s), 7.74 (1H, dd, J = 8.6, 2.3 Hz), 7.77 (1H, d, J = 2.3 Hz), 7.84 (1H, d, J = 8.6 Hz), 8.16 (1H, s) ppm. 13C NMR (125 MHz, DMSO-d6): δ 20.8, 116.1, 124.2, 127.4, 127.9, 133.0, 152.5, 166.8, 169.0 ppm. IR (ATR, KBr): ν 2891 (OH), 1759 (C=O), 1604 (C=O), 1559 (C=O) cm−1. Mp: 207–208 °C. HRMS (DART, m/z) calcd for C9H8IO5 [M + H]+: 322.9411; found: 322.9413.

3.2.9. 1-Hydroxy-5-(trifluoromethyl)-1λ3-benzo[d][1,2]iodaoxol-3(1H)-one (2h)

1H NMR (500 MHz, DMSO-d6): δ 8.08 (1H, d, J = 8.0 Hz), 8.21 (1H, s), 8.33 (1H, d, J = 8.1 Hz), 8.38 (1H, s) ppm. 13C NMR (125 MHz, DMSO-d6): δ 123.4 (q, J = 271.0 Hz), 125.5, 127.1 (d, J = 3.6 Hz), 127.9, 130.6 (d, J = 2.4 Hz), 131.6 (q, J = 32.6 Hz), 132.9, 166.3 ppm. 19F NMR (470 MHz, DMSO-d6): δ −64.5 ppm. IR (ATR, KBr): ν 2854 (OH), 1597 (C=O), 1559 (C=O) cm−1. Mp: 233–235 °C. HRMS (DART, m/z) calcd for C8H5F3IO3 [M + H]+: 332.9230; found: 332.9227.

3.2.10. 1-Hydroxy-5-nitro-1λ3-benzo[d][1,2]iodaoxol-3(1H)-one (2i)

1H NMR (400 MHz, DMSO-d6): δ 8.10 (1H, d, J = 8.7 Hz), 8.54 (1H, s), 8.57 (1H, d, J = 2.3 Hz), 8.73 (1H, dd, J = 8.7, 2.8 Hz) ppm. 13C NMR (100 MHz, DMSO-d6): δ 124.8, 127.7, 128.1, 128.2, 133.4, 149.7, 165.9 ppm. IR (ATR, KBr): ν 2834 (OH), 1617 (C=O), 1572 (C=O), 1541 (C=O) cm−1. Mp: 214–216 °C. 1H and 13C NMR data are consistent with those reported in the literature [107].

3.2.11. 1-Hydroxy-3-oxo-1,3-dihydro-1λ3-benzo[d][1,2]iodaoxole-5-carbonitrile (2j)

1H NMR (500 MHz, DMSO-d6): δ 8.01 (1H, d, J = 8.6 Hz), 8.32–8.41 (3H, m) ppm. 13C NMR (125 MHz, DMSO-d6): δ 113.5, 117.3, 126.4, 127.7, 132.9, 134.2, 137.0, 166.0 ppm. IR (ATR, KBr): ν 2903 (OH), 1625 (C=O), 1582 (C=O), 1561 (C=O) cm−1. Mp: 234–236 °C. HRMS (DART, m/z) calcd for C8H5INO3 [M + H]+: 289.9309; found: 289.9310.

3.2.12. 1,5-Dihydroxy-1λ3-benzo[d][1,2]iodaoxol-3(1H)-one (2k)

1H NMR (400 MHz, DMSO-d6): δ 7.36 (1H, dd, J = 8.7, 2.7 Hz), 7.40 (1H, d, J = 2.3 Hz), 7.57 (1H, d, J = 9.2 Hz), 7.94 (1H, s) ppm. 13C NMR (100 MHz, DMSO-d6): δ 106.6, 117.1, 122.0, 127.0, 132.8, 159.7, 167.6 ppm. IR (ATR, KBr): ν 3447 (OH), 3234 (OH), 1576 (C=O) cm−1. Mp: 230–232 °C. HRMS (DART, m/z) calcd for C7H6IO4 [M + H]+: 280.9305; found: 280.9304.

3.2.13. 6-Fluoro-1-hydroxy-1λ3-benzo[d][1,2]iodaoxol-3(1H)-one (2l)

1H NMR (500 MHz, DMSO-d6): δ 7.53–7.60 (2H, m), 8.01 (1H, dd, J = 8.0, 5.2 Hz), 8.21 (1H, s) ppm. 13C NMR (125 MHz, DMSO-d6): δ 113.5 (d, J = 27.4 Hz), 118.0 (d, J = 22.7 Hz), 122.8 (d, J = 8.4 Hz), 128.3, 132.9 (d, J = 8.4 Hz), 166.0 (d, J = 254.0 Hz), 166.7 ppm. 19F NMR (470 MHz, DMSO-d6): δ −109.0 (dt, J = 5.8, 8.6 Hz) ppm. IR (ATR, KBr): ν 3091 (OH), 1636 (C=O), 1586 (C=O) cm−1. Mp: 206–208 °C. 1H and 13C NMR data are consistent with those reported in the literature [105].

3.2.14. 6-Chloro-1-hydroxy-1λ3-benzo[d][1,2]iodaoxol-3(1H)-one (2m)

1H NMR (400 MHz, DMSO-d6): δ 7.75 (1H, d, J = 1.8 Hz), 7.78 (1H, dd, J = 7.8, 1.8 Hz), 7.96 (1H, d, J = 8.2, 2.3 Hz), 8.27 (1H, s) ppm. 13C NMR (100 MHz, DMSO-d6): δ 122.1, 125.7, 130.6, 130.7, 132.2, 139.3, 166.7 ppm. IR (ATR, KBr): ν 2854 (OH), 1607 (C=O), 1557 (C=O) cm−1. Mp: 212–214 °C. 1H NMR data is consistent with those reported in the literature [94].

3.2.15. 6-Bromo-1-hydroxy-1λ3-benzo[d][1,2]iodaoxol-3(1H)-one (2n)

1H NMR (500 MHz, DMSO-d6): δ 7.80–7.98 (3H, m), 8.23 (1H, s) ppm. 13C NMR (125 MHz, DMSO-d6): δ 122.1, 127.9, 128.5, 131.0, 132.5, 133.5, 166.8 ppm. IR (ATR, KBr): ν 2844 (OH), 1602 (C=O), 1556 (C=O) cm−1. Mp: 222–224 °C. HRMS (DART, m/z) calcd for C7H5BrIO3 [M + H]+: 342.8461; found: 342.8460.

3.2.16. 1-Hydroxy-6-(trifluoromethyl)-1λ3-benzo[d][1,2]iodaoxol-3(1H)-one (2o)

1H NMR (400 MHz, DMSO-d6): δ 8.06 (1H, s), 8.10 (1H, d, J = 8.1 Hz), 8.20 (1H, d, J = 8.0 Hz), 8.38 (1H, s) ppm. 13C NMR (100 MHz, DMSO-d6): δ 121.7, 123.1 (d, J = 3.6 Hz), 123.4 (d, J = 270.7 Hz), 127.6 (d, J = 3.6 Hz), 131.9, 133.9 (q, J = 32.2 Hz), 135.4, 166.4 ppm. 19F NMR (370 MHz, DMSO-d6): δ −64.6 ppm. IR (ATR, KBr): ν 2871 (OH), 1616 (C=O), 1560 (C=O) cm−1. Mp: 216–217 °C. 1H and 13C NMR data are consistent with those reported in the literature [12].

3.2.17. 1-Hydroxy-3-oxo-1,3-dihydro-1λ3-benzo[d][1,2]iodaoxole-6-carboxylic acid (2p)

1H NMR (500 MHz, DMSO-d6): δ ppm. 13C NMR (125 MHz, DMSO-d6): δ ppm. IR (ATR, KBr): ν 2832 (OH), 1704 (C=O), 1616 (C=O), 1558 (C=O) cm−1. Mp: 291–293 °C. HRMS (DART, m/z) calcd for C8H6IO5 [M + H]+: 308.9254; found: 308.9252.

3.2.18. 5,6-Difluoro-1-hydroxy-1λ3-benzo[d][1,2]iodaoxol-3(1H)-one (2q)

1H NMR (500 MHz, DMSO-d6): δ 7.73 (1H, dd, J = 9.2, 6.9 Hz), 7.97 (1H, dd, J = 9.7, 7.4 Hz), 8.37 (1H, s) ppm. 13C NMR (125 MHz, DMSO-d6): δ 115.5 (d, J = 22.7 Hz), 119.2 (d, J = 19.1 Hz), 129.16 (d, J = 2.4 Hz), 129.20 (d, J = 3.6 Hz), 151.3 (dd, J = 263.5, 13.1 Hz), 153.7 (dd, J = 256.4, 14.3 Hz), 165.9 ppm. 19F NMR (470 MHz, DMSO-d6): δ −138.7–138.5 (m), −132.6–132.4 (m) ppm. IR (ATR, KBr): ν 2895 (OH), 1624 (C=O), 1591 (C=O) cm−1. Mp: 201–203 °C. HRMS (DART, m/z) calcd for C7H4F2IO3 [M + H]+: 300.9168; found: 300.9170.

3.2.19. 1-Hydroxy-5,6-dimethoxy-1λ3-benzo[d][1,2]iodaoxol-3(1H)-one (2r)

1H NMR (500 MHz, DMSO-d6): δ 3.89 (6H, s), 7.24 (1H, s), 7.46 (1H, s), 7.95 ppm. 13C NMR (125 MHz, DMSO-d6): δ 55.9, 56.0, 107.4, 110.7, 112.4, 123.9, 150.6, 154.1, 167.8 ppm. IR (ATR, KBr): ν 3016 (OH), 1592 (C=O), 1559 (C=O) cm−1. Mp: 201–203 °C. 1H and 13C NMR data are consistent with those reported in the literature [46].

3.2.20. 7-Bromo-1-hydroxy-1λ3-benzo[d][1,2]iodaoxol-3(1H)-one (2s)

1H NMR (500 MHz, DMSO-d6): δ 7.60 (1H, t, J = 7.7 Hz), 7.97 (1H, dd, J = 7.7, 1.5 Hz), 8.02 (1H, dd, J = 7.5, 1.2 Hz) ppm. 13C NMR (125 MHz, DMSO-d6): δ 119.6, 130.0, 133.1, 135.4, 140.5, 146.1, 167.0 ppm. IR (ATR, KBr): ν 3273 (OH), 1647 (C=O) cm−1. Mp: 154–155 °C. HRMS (DART, m/z) calcd for C7H5BrIO3 [M + H]+: 342.8461; found: 342.8462.

3.2.21. 1-Hydroxy-7-methyl-1λ3-benzo[d][1,2]iodaoxol-3(1H)-one (2t)

1H NMR (400 MHz, DMSO-d6): δ 2.79 (3H, s), 7.57–7.73 (2H, m), 7.90 (1H, d, J = 6.9 Hz) ppm. 13C NMR (100 MHz, DMSO-d6): δ 19.6, 128.7, 132.0, 132.6, 137.9, 139.1, 147.4, 167.9 ppm. IR (ATR, KBr): ν 1672 (C=O) cm−1. Mp: 164–166 °C. 1H NMR data are consistent with those reported in the literature [106].

3.2.22. 4-Fluoro-1-hydroxy-1λ3-benzo[d][1,2]iodaoxol-3(1H)-one (2u)

1H NMR (400 MHz, DMSO-d6): δ 7.51 (1H, dd, J = 10.1, 8.2 Hz), 7.71 (1H, d, J = 7.7 Hz), 7.90 (1H, td, J = 8.2, 4.6 Hz), 8.24 (1H, s) ppm. 13C NMR (100 MHz, DMSO-d6): δ 118.6 (d, J = 22.0 Hz), 119.2 (d, J = 11.5 Hz), 122.5 (d, J = 3.8 Hz), 123.2, 134.3 (d, J = 8.6 Hz), 163.8 (d, J = 4.8 Hz), 163.8 (d, J = 264.4 Hz) ppm. 19F NMR (375 MHz, DMSO-d6): δ −114.7 (dd, J = 15.2, 4.9 Hz) ppm. IR (ATR, KBr): ν 3091 (OH), 1636 (C=O), 1586 (C=O) cm−1. Mp: 213–214 °C. 1H and 13C NMR data are consistent with those reported in the literature [106,108].

3.2.23. 1-Hydroxy-4-methyl-1λ3-benzo[d][1,2]iodaoxol-3(1H)-one (2v)

1H NMR (400 MHz, DMSO-d6): δ 2.70 (3H, s), 7.48–7.55 (1H, m), 7.72–7.80 (2H, m), 7.94 (1H, s) ppm. 13C NMR (100 MHz, DMSO-d6): δ 20.4, 122.2, 124.2, 128.2, 133.3, 133.4, 144.3, 168.0 ppm. IR (ATR, KBr): ν 2926 (OH), 1625 (C=O), 1584 (C=O) cm−1. Mp: 212–213 °C. 1H and 13C NMR data are consistent with those reported in the literature [106].

3.2.24. 1-Hydroxy-1λ3-naphtho[2,3-d][1,2]iodaoxol-3(1H)-one (4)

1H NMR (400 MHz, DMSO-d6): δ 7.76 (2H, m), 8.14–8.33 (2H, m), 8.29 (1H, d, J = 8.2 Hz), 8.39 (1H, s), 8.69 (1H, s) ppm. 13C NMR (100 MHz, DMSO-d6): δ 115.9, 126.3, 127.8, 127.9, 128.1, 128.9, 129.3, 131.7, 132.8, 135.8, 167.7 ppm. IR (ATR, KBr): ν 3053 (OH), 1698 (C=O), 1607 (C=O), 1559 (C=O) cm−1. Mp: 164–165 °C. 1H and 13C NMR data are consistent with those reported in the literature [107].

3.3. Synthesis of Benziodoxole Alkoxides

3.3.1. General Procedure for the Synthesis of Benziodoxole Alkoxides (6)

To a suspension of IBA 2a (264 mg, 1.0 mmol) in an appropriate alcohol (10 mL) was added MS3Å (1 g). After the mixture was stirred under the appropriate conditions (see Figure 6), MS3Å was filtered using CH2Cl2, and the solvents were then removed by evaporation. The residue was washed with hexane and filtered to remove the corresponding alcohol completely. The residue was dissolved with CH2Cl2 and the extract was then filtered through filter paper to remove unreacted substrate. Removal of the solvent by evaporation gave the corresponding benziodoxole alkoxides 6ah as a white powder.

3.3.2. 1-Methoxy-1λ3-benzo[d][1,2]iodaoxol-3(1H)-one (6a)

1H NMR (500 MHz, CDCl3): δ 4.29 (3H, s), 7.70 (1H, t, J = 7.7 Hz), 7.78 (1H, d, J = 8.1 Hz), 7.91 (1H, ddd, J = 8.6, 6.9, 1.2 Hz), 8.28 (1H, dd, J = 7.5, 1.2 Hz) ppm. 13C NMR (125 MHz, CDCl3): δ 62.3, 118.6, 126.0, 130.6, 131.0, 132.9, 135.1, 168.0 ppm. IR (ATR, KBr): ν 1653 (C=O) cm−1. Mp: 161–163 °C. 1H and 13C NMR data are consistent with those reported in the literature [28,109].

3.3.3. 1-Ethoxy-1λ3-benzo[d][1,2]iodaoxol-3(1H)-one (6b)

1H NMR (500 MHz, CDCl3): δ 1.35 (3H, t, J = 6.9 Hz), 4.30 (2H, q, J = 6.9 Hz), 7.70 (1H, t, J = 7.4 Hz), 7.79 (1H, d, J = 8.9 Hz), 7.89 (1H, td, J = 8.6, 1.8 Hz), 8.28 (1H, dd, J = 7.2, 2.0 Hz) ppm. 13C NMR (125 MHz, CDCl3): δ 19.0, 69.9, 118.8, 125.9, 130.7, 131.0, 132.9, 135.0, 168.0 ppm. IR (ATR, KBr): ν 1655 (C=O) cm−1. Mp: 123–125 °C. 1H and 13C NMR data are consistent with those reported in the literature [109].

3.3.4. 1-(2,2,2-trifluoroethoxy)-1λ3-benzo[d][1,2]iodaoxol-3(1H)-one (6c)

1H NMR (500 MHz, CDCl3): δ 4.52 (2H, q, J = 8.6 Hz), 7.74 (1H, t, J = 7.5 Hz), 7.86 (1H, d, J = 8.0 Hz), 7.97 (1H, ddd, J = 8.0, 6.9, 1.2 Hz), 8.27 (1H, dd, J = 7.5, 1.2 Hz) ppm. 13C NMR (125 MHz, CDCl3): δ 69.7 (q, J = 34.2 Hz), 119.0, 123.5 (q, J = 278.2 Hz), 126.5, 129.5, 131.4, 133.2, 135.9, 167.9 ppm. 19F NMR (470 MHz, CDCl3): δ −77.2 (q, J = 9.1 Hz) ppm. IR (ATR, KBr): ν 1646 (C=O) cm−1. Mp: 139–141 °C. HRMS (DART, m/z) calcd for C9H7F3IO3 [M + H]+: 346.9386; found: 346.9384.

3.3.5. 1-Propoxy-1λ3-benzo[d][1,2]iodaoxol-3(1H)-one (6d)

1H NMR (500 MHz, CDCl3): δ 1.02 (3H, t, J = 7.5 Hz), 1.72 (2H, sext, J = 7.1 Hz), 4.20 (2H, t, J = 6.6 Hz), 7.70 (1H, t, J = 7.2 Hz), 7.79 (1H, d, J = 8.6 Hz), 7.89 (1H, t, J = 7.2 Hz), 8.28 (1H, d, J = 7.5 Hz) ppm. 13C NMR (125 MHz, CDCl3): δ 10.1, 26.5, 76.0, 118.9, 125.9, 130.7, 130.9, 132.8, 135.0, 167.9 ppm. IR (ATR, KBr): ν 1651 (C=O) cm−1. Mp: 146–148 °C. HRMS (DART, m/z) calcd for C10H12IO3 [M + H]+: 306.9826; found: 306.9823.

3.3.6. 1-Isopropoxy-1λ3-benzo[d][1,2]iodaoxol-3(1H)-one (6e)

1H NMR (500 MHz, CDCl3): δ 1.36 (6H, d, J = 6.3 Hz), 4.33 (1H, sep, J = 6.1 Hz), 7.69 (1H, t, J = 7.4 Hz), 7.82 (1H, d, J = 7.5 Hz), 7.88 (1H, td, J = 7.8, 1.5 Hz), 8.28 (1H, dd, J = 7.5, 1.7 Hz) ppm. 13C NMR (125 MHz, CDCl3): δ 25.2, 75.6, 119.1, 126.0, 130.8, 130.9, 132.7, 134.8, 168.0 ppm. IR (ATR, KBr): ν 1653 (C=O) cm−1. Mp: 253–254 °C. 1H and 13C NMR data are consistent with those reported in the literature [109].

3.3.7. 1-((1,1,1,3,3,3-hexafluoropropan-2-yl)oxy)-1λ3-benzo[d][1,2]iodaoxol-3(1H)-one (6f)

1H NMR (500 MHz, CDCl3): δ 4.80 (1H, sep, J = 5.8 Hz), 7.74 (1H, ddd, J = 8.0, 7.5, 1.4 Hz), 7.97–8.04 (2H, m), 8.24 (1H, dd, J = 7.4, 1.2 Hz) ppm. 13C NMR (125 MHz, CDCl3): δ 76.1 (quin, J = 32.5 Hz), 119.4, 122.2 (q, J = 283.8 Hz), 127.3, 128.5, 131.6, 133.3, 136.4, 168.1 ppm. 19F NMR (470 MHz, CDCl3): δ −76.0 (d, J = 5.7 Hz) ppm. IR (ATR, KBr): ν 1661 (C=O) cm−1. Mp: 148–149 °C. HRMS (DART, m/z) calcd for C10H6F6IO3 [M + H]+: 414.9260; found: 414.9258.

3.3.8. 1-Butoxy-1λ3-benzo[d][1,2]iodaoxol-3(1H)-one (6g)

1H NMR (500 MHz, CDCl3): δ 0.98 (3H, t, J = 7.5 Hz), 1.46 (2H, sext, J = 7.4 Hz), 1.68 (2H, quin, J = 7.2 Hz), 4.24 (2H, t, J = 6.6 Hz), 7.70 (1H, t, J = 7.5 Hz), 7.78 (1H, d, J = 8.0 Hz), 7.89 (1H, t, J = 7.5 Hz), 8.28 (1H, d, J = 7.5 Hz) ppm. 13C NMR (125 MHz, CDCl3): δ 13.9, 19.0, 35.4, 74.2, 118.9, 125.9, 130.7, 131.0, 132.9, 135.0, 167.9 ppm. IR (ATR, KBr): ν 1650 (C=O) cm−1. Mp: 143–144 °C. 1H and 13C NMR data are consistent with those reported in the literature [110].

3.3.9. 1-(tert-Butoxy)-1λ3-benzo[d][1,2]iodaoxol-3(1H)-one (6h)

1H NMR (500 MHz, CDCl3): δ 1.41 (9H, s), 7.67 (1H, ddd, J = 8.1, 7.5, 1.8 Hz), 7.83–7.91 (2H,m), 8.26 (1H, dd, J = 8.1, 1.8 Hz) ppm. 13C NMR (125 MHz, CDCl3): δ 30.4, 78.6, 119.6, 126.2, 130.8, 131.0, 132.4, 134.7, 168.0 ppm. IR (ATR, KBr): ν 1659 (C=O) cm−1. Decomp: 265 °C. HRMS (DART, m/z) calcd for C11H13IO3 [M + H]+: 320.9982; found: 320.9980.

3.4. Synthesis of Benziodoxole Acetates

3.4.1. General Procedure for the Synthesis of Benziodoxole Acetates (7)

MS3Å (0.5 g) was added to a suspension of IBAs 2ai (0.50 mmol) in AcOH (5 mL), and the mixture was stirred under the appropriate conditions (see Scheme 4). Then, MS3Å was filtered using CH2Cl2, and the solvents were removed by evaporation. The residue was washed with ether and filtered to remove AcOH completely. The resulting residue was dissolved with CH2Cl2, and the extract was then filtered through filter paper to remove unreacted substrate. After solvent removal by evaporation, the corresponding benziodoxole acetates 7ai were obtained as a white powder.

3.4.2. 3-Oxo-1λ3-benzo[d][1,2]iodaoxol-1(3H)-yl acetate (7a)

1H NMR (500 MHz, CDCl3): δ 2.27 (3H, s), 7.72 (1H, td, J = 7.5, 1.2 Hz), 7.94 (1H, ddd, J = 8.6, 6.9, 1.2 Hz), 8.01 (1H, d, J = 8.6 Hz), 8.25 (1H, dd, J = 7.5, 1.2 Hz) ppm. 13C NMR (125 MHz, CDCl3): δ 20.2, 118.3, 128.9, 129.2, 131.2, 133.1, 136.1, 168.1, 176.3 ppm. IR (ATR, KBr): ν 1684 (C=O) cm−1. Mp: 220–222 °C. 1H and 13C NMR data are consistent with those reported in the literature [109].

3.4.3. 5-Fluoro-3-oxo-1λ3-benzo[d][1,2]iodaoxol-1(3H)-yl acetate (7b)

1H NMR (500 MHz, CDCl3): δ 2.26 (3H, s), 7.64 (1H, ddd, J = 8.6, 7.7, 2.9 Hz), 7.94–8.00 (2H, m) ppm. 13C NMR (125 MHz, CDCl3): δ 20.2, 111.3, 120.0 (d, J = 23.8 Hz), 123.7 (d, J = 22.7 Hz), 131.0 (d, J = 8.3 Hz), 131.8 (d, J = 7.2 Hz), 165.0 (d, J = 252.8 Hz), 166.8, 176.4 ppm. 19F NMR (470 MHz, CDCl3): δ −110.8 (td, J = 7.2, 4.3 Hz) ppm. IR (ATR, KBr): ν 1696 (C=O) cm−1. Mp: 225–226 °C. 1H and 13C NMR data are consistent with those reported in the literature [111].

3.4.4. 5-Chloro-3-oxo-1λ3-benzo[d][1,2]iodaoxol-1(3H)-yl acetate (7c)

1H NMR (500 MHz, CDCl3): δ 2.27 (3H, s), 7.87 (1H, dd, J = 8.6, 2.3 Hz), 7.93 (1H, d, J = 9.2 Hz), 8.22 (1H, d, J = 1.7 Hz) ppm. 13C NMR (125 MHz, CDCl3): δ 20.2, 115.5, 130.4, 130.9, 133.0, 136.0, 138.8, 166.7, 176.4 ppm. IR (ATR, KBr): ν 1698 (C=O) cm−1. Mp: 244–245 °C. 1H and 13C NMR data are consistent with those reported in the literature [105].

3.4.5. 5-Bromo-3-oxo-1λ3-benzo[d][1,2]iodaoxol-1(3H)-yl acetate (7d)

1H NMR (500 MHz, CDCl3): δ 2.27 (3H, s), 7.85 (1H, d, J = 9.2 Hz), 8.01 (1H, dd, J = 8.6, 2.3 Hz), 8.37 (1H, d, J = 1.8 Hz) ppm. 13C NMR (125 MHz, CDCl3): δ 20.2, 116.5, 126.6, 130.7, 131.0, 136.1, 138.9, 166.6, 176.4 ppm. IR (ATR, KBr): ν 1680 (C=O) cm−1. Mp: 226–228 °C. 1H and 13C NMR data are consistent with those reported in the literature [112].

3.4.6. 5-Methyl-3-oxo-1λ3-benzo[d][1,2]iodaoxol-1(3H)-yl acetate (7e)

1H NMR (500 MHz, CDCl3): δ 2.25 (3H, s), 2.56 (3H, s), 7.73 (1H, dd, J = 8.6, 1.7 Hz), 7.84 (1H, d, J = 8.6 Hz), 8.07 (1H, d, J = 1.8 Hz) ppm. 13C NMR (125 MHz, CDCl3): δ 20.3, 20.8, 114.6, 128.9, 133.6, 137.1, 142.3, 168.3, 176.4 ppm. IR (ATR, KBr): ν 1682 (C=O), 1659 (C=O) cm−1. Mp: 215–217 °C. 1H and 13C NMR data are consistent with those reported in the literature [105].

3.4.7. 5-Methoxy-3-oxo-1λ3-benzo[d][1,2]iodaoxol-1(3H)-yl acetate (7f)

1H NMR (500 MHz, CDCl3): δ 2.25 (3H, s), 3.94 (3H, s), 7.46 (1H, dd, J = 9.2, 2.9 Hz), 7.74 (1H, d, J = 2.9 Hz), 7.81 (1H, d, J = 8.6 Hz) ppm. 13C NMR (125 MHz, CDCl3): δ 20.3, 56.2, 106.8, 115.9, 124.5, 129.7, 130.6, 162.7, 168.1, 176.4 ppm. IR (ATR, KBr): ν 1697 (C=O), 1681 (C=O), 1656 (C=O) cm−1. Mp: 207–209 °C. 1H and 13C NMR data are consistent with those reported in the literature [105].

3.4.8. 3-Oxo-1λ3-benzo[d][1,2]iodaoxole-1,5(3H)-diyl diacetate (7g)

1H NMR (500 MHz, CDCl3): δ 2.26 (3H, s), 2.37 (3H, s), 7.68 (1H, dd, J = 8.9, 2.6 Hz), 7.96–8.02 (2H, m) ppm. 13C NMR (125 MHz, CDCl3): δ 20.3, 21.1, 113.6, 126.2, 129.8, 130.3, 130.9, 153.6, 167.2, 168.7, 176.5 ppm. IR (ATR, KBr): ν 1690 (C=O) cm−1. Mp: 153–154 °C. HRMS (DART, m/z) calcd for C11H10IO6 [M + H]+: 364.9517; found: 364.9518.

3.4.9. 3-Oxo-5-(trifluoromethyl)-1λ3-benzo[d][1,2]iodaoxol-1(3H)-yl acetate (7h)

1H NMR (500 MHz, CDCl3): δ 2.29 (3H, s), 8.14 (1H, dd, J = 8.6, 1.7 Hz), 8.20 (1H, d, J = 8.6 Hz), 8.52 (1H, d, J = 1.7 Hz) ppm. 13C NMR (125 MHz, CDCl3): δ 20.2, 121.9, 122.8 (q, J = 271.4 Hz), 130.1 (d, J = 3.6 Hz), 130.4 (d, J = 9.5 Hz), 132.4 (d, J = 2.4 Hz), 134.5 (q, J = 33.8 Hz), 166.7, 176.5 ppm. 19F NMR (470 MHz, CDCl3): δ −64.9 ppm. IR (ATR, KBr): ν 1692 (C=O), 1647 (C=O) cm−1. Mp: 212–213 °C. HRMS (DART, m/z) calcd for C10H7F3IO4 [M + H]+: 374.9336; found: 374.9334.

3.4.10. 5-Nitro-3-oxo-1λ3-benzo[d][1,2]iodaoxol-1(3H)-yl acetate (7i)

1H NMR (500 MHz, CDCl3): δ 2.30 (3H, s), 8.27 (1H, d, J = 9.2 Hz), 8.71 (1H, dd, J = 9.0, 2.5 Hz), 9.04 (1H, d, J = 2.3 Hz) ppm. 13C NMR (125 MHz, CDCl3): δ 20.2, 124.1, 127.6, 129.8, 131.0, 131.5,150.8, 165.7, 176.6 ppm. IR (ATR, KBr): ν 1705 (C=O), 1665 (C=O) cm−1. Mp: 209–210 °C. 1H and 13C NMR data are consistent with those reported in the literature [107].

4. Conclusions

We have presented a practical synthetic method for IBA from 2-IB without contamination by hazardous pentavalent IBX using cost-effective Oxone® in aqueous solution. This highly safe, convenient method operates under mild conditions such as room temperature, which contrasts with traditional method using reflux conditions and expensive NaIO4 in AcOH solution. The use of mild conditions circumvents the problem of the formation of byproducts such as potentially explosive pentavalent cyclic hypervalent iodine compound, i.e., IBX; the contamination of IBX into IBA is generally not desired for safety reasons. The reaction time can be shortened by heating; in this case, addition of a ferric salt in our reaction system can effectively suppress the formation of IBX as byproducts. In addition, a convenient derivatization of the hydroxy group of IBAs by solvolytic treatment is presented. These derivatizations were generally achieved under mild conditions below 80 °C. Our methods, which do not require any chromatography technique, can be performed safely and would be suitable for large-scale synthesis.

Supplementary Materials

Supplementary materials are available online, 1H NMR spectroscopic data for the compounds 2av and 4.

Author Contributions

H.C. found the selective reaction to obtain IBAs using Oxone® and the solvolytic functionalization for IBA and drafted the manuscript; N.K., H.Y., and N.T. also contributed to the experiments; T.D. directed this study as a project and finalized the manuscript with critical discussion. All authors have read and agreed to the published version of the manuscript.

Funding

This work was financially supported by a Grant-in-Aid for Scientific Research (C) (JSPS KAKENHI Grant Number JP19K05466) from JSPS and the Ritsumeikan Global Innovation Research Organization (R-GIRO) project. H.C. also acknowledges support from a Grant-in-Aid for Early-Career Scientists (JSPS KAKENHI Grant Number 20K15103) from JSPS.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data presented in this study are available in the communication.

Conflicts of Interest

The authors declare no conflict of interest.

Sample Availability

Samples of the compounds are not available from the authors.

References

  1. Zhdankin, V.V. Organoiodine(V) reagents in organic synthesis. J. Org. Chem. 2011, 76, 1185–1197. [Google Scholar] [CrossRef]
  2. Singh, F.V.; Wirth, T. Hypervalent iodine-catalyzed oxidative functionalizations including stereoselective reactions. Chem. Asian J. 2014, 9, 950–971. [Google Scholar] [CrossRef] [PubMed]
  3. Yoshimura, A.; Zhdankin, V.V. Advances in synthetic applications of hypervalent iodine compounds. Chem. Rev. 2016, 116, 3328–3435. [Google Scholar] [CrossRef] [PubMed]
  4. Tan, H.; Ji, W.; Wang, L. Sunlight-driven decarboxylative alkynylation of α-keto acids with bromoacetylenes by hypervalent iodine reagent catalysis: A facile approach to ynones. Angew. Chem. Int. Ed. 2015, 54, 8374–8377. [Google Scholar] [CrossRef] [PubMed]
  5. Wang, D.; Zhang, L.; Luo, S. Enantioselective decarboxylative α-alkynylation of β-ketocarbonyls via a catalytic α-imino radical Intermediate. Org. Lett. 2017, 19, 4924–4927. [Google Scholar] [CrossRef]
  6. Yang, S.; Tan, H.; Ji, W.; Zhang, X.; Li, P. Visible light-induced decarboxylative acylarylation of phenyl propiolates with α-oxocarboxylic acids to coumarins catalyzed by hypervalent iodine reagents under transition metal-free conditions. Adv. Synth. Catal. 2017, 359, 443–453. [Google Scholar] [CrossRef]
  7. Sun, X.; Liu, T.; Yang, Y.-T.; Gu, Y.-J.; Liu, Y.W.; Ji, Y.G.; Luo, K.; Zhu, J.; Wu, L. Visible-light-promoted regio- and stereoselective oxyalkenylation of phosphinyl allenes. Adv. Synth. Catal. 2020, 362, 2701–2708. [Google Scholar] [CrossRef]
  8. Ball, L.T.; Lloyd-Jones, G.C.; Russell, C.A. Gold-catalysed oxyarylation of styrenes and mono- and gem-disubstituted olefins facilitated by an iodine(III) oxidant. Chem. Eur. J. 2012, 18, 2931–2937. [Google Scholar] [CrossRef] [PubMed]
  9. Hata, K.; Segawa, Y.; Itami, K. Pyridylidene ligand facilitates gold-catalyzed oxidative C–H arylation of heterocycles. Beilstein. J. Org. Chem. 2015, 11, 2737–2746. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  10. Hu, J.; Lan, T.; Sun, Y.; Chen, H.; Yao, J.; Rao, Y. Unactivated C(sp3)–H hydroxylation through palladium catalysis with H2O as the oxygen source. Chem. Commun. 2015, 51, 14929–14932. [Google Scholar] [CrossRef] [PubMed]
  11. Bindu, V.H.; Parvathaneni, S.P.; Rao, V.J. Iodosobenzoic acid (IBA) catalysed benzylic and aromatic C–H oxidations. Catal. Lett. 2017, 147, 1434–1440. [Google Scholar] [CrossRef]
  12. Li, G.-X.; Morales-Rivera, C.A.; Gao, F.; Wang, Y.; He, G.; Liu, P.; Chen, G. A unified photoredox-catalysis strategy for C(sp3)–H hydroxylation and amidation using hypervalent iodine. Chem. Sci. 2017, 8, 7180–7185. [Google Scholar] [CrossRef] [Green Version]
  13. Wang, D.; Ren, R.; Zhu, C. Manganese-promoted ring-opening hydrazination of cyclobutanols: Synthesis of alkyl hydrazines. J. Org. Chem. 2016, 81, 8043–8049. [Google Scholar] [CrossRef]
  14. Ye, L.; Gu, Q.-S.; Tian, Y.; Meng, X.; Chen, G.-C.; Liu, X.-Y. Radical asymmetric intramolecular α-cyclopropanation of aldehydes towards bicyclo[3.1.0]hexanes containing vicinal all-carbon quaternary stereocenters. Nat. Commun. 2018, 9, 1–13. [Google Scholar] [CrossRef] [Green Version]
  15. Amos, S.G.E.; Nicolai, S.; Waser, J. Photocatalytic umpolung of N- and O-substituted alkenes for the synthesis of 1,2-amino alcohols and diols. Chem. Sci. 2020, 11, 11274–11279. [Google Scholar] [CrossRef]
  16. He, X.-K.; Lu, J.; Zhang, A.-J.; Zhang, Q.-Q.; Xu, G.-Y.; Xuan, J. Bi-OAc-accelerated C3-H alkylation of quinoxaline-2(1H)-ones under visible-light irradiation. Org. Lett. 2020, 22, 5984–5989. [Google Scholar] [CrossRef]
  17. Li, J.; Liu, Z.; Wu, S.; Chen, Y. Acyl radical smiles rearrangement to construct hydroxybenzophenones by photoredox catalysis. Org. Lett. 2019, 21, 2077–2080. [Google Scholar] [CrossRef] [PubMed]
  18. Fang, Z.; Wang, Y.; Wang, Y. Synthesis of 4-iodoisoquinolin-1(2H)-ones by a dirhodium (II)-catalyzed 1,4-bisfunctionalization of isoquinolinium iodide salts. Org. Lett. 2019, 21, 434–438. [Google Scholar] [CrossRef] [PubMed]
  19. Wu, L.; Hao, Y.; Lin, Y.; Wang, Q. Visible-light-induced intramolecular sp3-C-H oxidation of 2-alkyl-substituted benzamides for the synthesis of functionalized iminoisobenzofurans. Chem. Commun. 2019, 55, 13908–13911. [Google Scholar] [CrossRef] [PubMed]
  20. Wang, N.; Gu, Q.-S.; Li, Z.-L.; Li, Z.; Guo, Y.-L.; Guo, Z.; Liu, X.-Y. Direct photocatalytic synthesis of medium-sized lactams by C−C bond cleavage. Angew. Chem. Int. Ed. 2018, 57, 14225–14229. [Google Scholar] [CrossRef] [PubMed]
  21. Pawar, G.G.; Robert, F.; Grau, E.; Cramail, H.; Landais, Y. Visible-light photocatalyzed oxidative decarboxylation of oxamic acids: A green route to urethanes and ureas. Chem. Commun. 2018, 54, 9337–9340. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  22. Jia, K.; Zhang, F.; Huang, H.; Chen, Y. Visible-light-induced alkoxyl radical generation enables selective C(sp3)–C(sp3) bond cleavage and functionalizations. J. Am. Chem. Soc. 2016, 138, 1514–1517. [Google Scholar] [CrossRef]
  23. Li, G.-X.; Morales-Rivera, C.A.; Wang, Y.; Gao, F.; He, G.; Liu, P.; Chen, G. Photoredox-mediated Minisci C–H alkylation of N-heteroarenes using boronic acids and hypervalent iodine. Chem. Sci. 2016, 7, 6407–6412. [Google Scholar] [CrossRef] [Green Version]
  24. Ji, W.; Tan, H.; Wang, M.; Li, P.; Wang, L. Photocatalyst-free hypervalent iodine reagent catalyzed decarboxylative acylarylation of acrylamides with α-oxocarboxylic acids driven by visible-light irradiation. Chem. Commun. 2016, 52, 1462–1465. [Google Scholar] [CrossRef] [PubMed]
  25. Dai, J.-J.; Zhang, W.-M.; Shu, Y.-J.; Sun, Y.-Y.; Xu, J.; Feng, Y.-S.; Xu, H.-J. Deboronative cyanation of potassium alkyltrifluoroborates via photoredox catalysis. Chem. Commun. 2016, 52, 6793–6796. [Google Scholar] [CrossRef] [PubMed]
  26. Huang, H.; Zhang, G.; Chen, Y. Dual hypervalent iodine (III) reagents and photoredox catalysis enable decarboxylative ynonylation under mild conditions. Angew. Chem. Int. Ed. 2015, 54, 7872–7876. [Google Scholar] [CrossRef] [PubMed]
  27. Zhang, D.; Gao, F.; Nian, Y.; Zhou, Y.; Jiang, H.; Liu, H. Palladium-catalyzed picolinamide-directed coupling of C(sp2)–H and C(sp2)–H: A straightforward approach to quinolinone and pyridone scaffolds. Chem. Commun. 2015, 51, 7509–7511. [Google Scholar] [CrossRef]
  28. Shan, G.; Yang, X.; Zong, Y.; Rao, Y. An efficient palladium-catalyzed C-H alkoxylation of unactivated methylene and methyl groups with cyclic hypervalent iodine (I3+) oxidants. Angew. Chem. Int. Ed. 2013, 52, 13606–13610. [Google Scholar] [CrossRef]
  29. Zong, Y.; Rao, Y. Developing Pd (II) catalyzed double sp3 C–H alkoxylation for synthesis of symmetric and unsymmetric acetals. Org. Lett. 2014, 16, 5278–5281. [Google Scholar] [CrossRef] [PubMed]
  30. Sun, X.; Yao, X.; Zhang, C.; Rao, Y. Pd (ii) catalyzed ortho C–H iodination of phenylcarbamates at room temperature using cyclic hypervalent iodine reagents. Chem. Commun. 2015, 51, 10014–10017. [Google Scholar] [CrossRef]
  31. Yu, Q.-Y.; Zhong, H.-M.; Sun, W.-W.; Zhang, S.-J.; Cao, P.; Dong, X.-P.; Qin, H.-B.; Liu, J.-K.; Wu, B. Palladium-catalyzed C(sp3)−H functionalization at the C3 position of l-pipecolinic acid derivatives. Asian J. Org. Chem. 2016, 5, 608–612. [Google Scholar] [CrossRef]
  32. Yang, B.; Xu, X.-H.; Qing, F.-L. Synthesis of difluoroalkylated arenes by hydroaryldifluoromethylation of alkenes with α,α-difluoroarylacetic acids under photoredox catalysis. Org. Lett. 2016, 18, 5956–5959. [Google Scholar] [CrossRef]
  33. Nappi, M.; He, C.; Whitehurst, W.G.; Chappell, B.G.N.; Gaunt, M.J. Selective reductive elimination at alkyl palladium(IV) by dissociative ligand ionization: Catalytic C(sp3)−H amination to azetidines. Angew. Chem. Int. Ed. 2018, 57, 3178–3182. [Google Scholar] [CrossRef] [PubMed]
  34. Muraki, T.; Togo, H.; Yokoyama, M. Synthetic use of 1-(p-toluenesulfonyloxy)-1,2-benziodoxol-3(1H)-one: Iodination of aromatic rings. Synlett 1998, 3, 286–288. [Google Scholar] [CrossRef]
  35. Muraki, T.; Togo, H.; Yokoyama, M. Reactivity and synthetic utility of 1-(arenesulfonyloxy)benziodoxolones. J. Org. Chem. 1999, 64, 2883–2889. [Google Scholar] [CrossRef]
  36. Harper, M.J.; Emmett, E.J.; Bower, J.F.; Russell, C.A. Oxidative 1,2-difunctionalization of ethylene via gold-catalyzed oxyarylation. J. Am. Chem. Soc. 2017, 139, 12386–12389. [Google Scholar] [CrossRef] [Green Version]
  37. Yoshimura, A.; Nguyen, K.C.; Klasen, S.C.; Postnikov, P.S.; Yusubov, M.S.; Saito, A.; Nemykin, V.N.; Zhdankin, V.V. Hypervalent iodine-catalyzed synthesis of 1,2,4-oxadiazoles from aldoximes and nitriles. Asian J. Org. Chem. 2016, 5, 1128–1133. [Google Scholar] [CrossRef]
  38. Vinayak, B.; Ravindrakumar, P.V.; Ramana, D.V.; Chandrasekharam, M. Revisiting 1-chloro-1,2-benziodoxol-3-one: Efficient ortho-chlorination of aryls under aqueous conditions. New J. Chem. 2018, 42, 8953–8959. [Google Scholar] [CrossRef]
  39. Parvathaneni, S.P.; Perumgani, P.C. Regioselective chlorination of aryl C−H bonds with hypervalent iodine(III) reagent 1-chloro-1,2-benziodoxol-3-one. Asian J. Org. Chem. 2018, 7, 324–327. [Google Scholar] [CrossRef]
  40. Jia, Y.; Chen, L.; Zhang, H.; Zheng, Y.; Jiang, Z.-X.; Yang, Z. Electrophilic chloro(ω-alkoxy)lation of alkenes employing 1-chloro-1,2-benziodoxol-3-one: Facile synthesis of β-chloroethers. Org. Biomol. Chem. 2018, 16, 7203–7213. [Google Scholar] [CrossRef]
  41. Wang, M.; Zhang, Y.; Wang, T.; Wang, C.; Xue, D.; Xiao, J. Story of an age-old reagent: An electrophilic chlorination of arenes and heterocycles by 1-chloro-1,2-benziodoxol-3-one. Org. Lett. 2016, 18, 1976–1979. [Google Scholar] [CrossRef] [PubMed]
  42. Egami, H.; Yoneda, T.; Uku, M.; Ide, T.; Kawato, Y.; Hamashima, Y. Difunctionalization of alkenes using 1-chloro-1,2-benziodoxol-3-(1H)-one. J. Org. Chem. 2016, 81, 4020–4030. [Google Scholar] [CrossRef] [PubMed]
  43. Xing, B.; Ni, C.; Hu, J. Hypervalent iodine(III)-catalyzed Balz–Schiemann fluorination under mild conditions. Angew. Chem. Int. Ed. 2018, 57, 9896–9900. [Google Scholar] [CrossRef]
  44. Jiang, X.; Zheng, C.; Lei, L.; Lin, K.; Yu, C. Synthesis of 2-oxindoles from substituted indoles by hypervalent-iodine oxidation. Eur. J. Org. Chem. 2018, 2018, 1437–1442. [Google Scholar] [CrossRef]
  45. Jiang, X.; Li, G.; Yu, C. A Diels-Alder reaction/oxa-Michael addition/acyloin rearrangement cascade on tropolonic substrates. Tetrahedron Lett. 2018, 59, 1506–1510. [Google Scholar] [CrossRef]
  46. Vaillant, F.L.; Wodrich, M.D.; Waser, J. Room temperature decarboxylative cyanation of carboxylic acids using photoredox catalysis and cyanobenziodoxolones: A divergent mechanism compared to alkynylation. Chem. Sci. 2017, 8, 1790–1800. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  47. Li, X.; Qi, X.; Hou, C.; Chen, P.; Liu, G. Palladium(II)-catalyzed enantioselective azidation of unactivated alkenes. Angew. Chem. Int. Ed. 2020, 59, 17239–17244. [Google Scholar] [CrossRef]
  48. Wang, Y.; Wang, N.; Zhao, J.; Sun, M.; You, H.; Fang, F.; Li, Z.-Q. Visible-light-promoted site-specific and diverse functionalization of a C(sp3)-C(sp3) bond adjacent to an arene. ACS Catal. 2020, 10, 6603–6612. [Google Scholar] [CrossRef]
  49. Wang, Y.; Hu, X.; Morales-Rivera, C.A.; Li, G.-X.; Huang, X.; He, G.; Liu, P.; Chen, G. Epimerization of tertiary carbon centers via reversible radical cleavage of unactivated C(sp3)-H bonds. J. Am. Chem. Soc. 2018, 140, 9678–9684. [Google Scholar] [CrossRef]
  50. Shen, K.; Wang, Q. Copper-catalyzed alkene aminoazidation as a rapid entry to 1,2-diamines and installation of an azide reporter onto azaheterocycles. J. Am. Chem. Soc. 2017, 139, 13110–13116. [Google Scholar] [CrossRef] [PubMed]
  51. Li, L.; Li, Z.-L.; Wang, F.-L.; Guo, Z.; Cheng, Y.-F.; Wang, N.; Dong, X.-W.; Fang, C.; Liu, J.; Hou, C.; et al. Radical aryl migration enables diversity-oriented synthesis of structurally diverse medium/macro- or bridged-rings. Nat. Commun. 2016, 7, 13852. [Google Scholar] [CrossRef] [Green Version]
  52. Wang, Y.; Li, G.-X.; Yang, G.; He, G.; Chen, G. A visible-light-promoted radical reaction system for azidation and halogenation of tertiary aliphatic C–H bonds. Chem. Sci. 2016, 7, 2679–2683. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  53. Sharma, A.; Hartwig, J.F. Metal-catalyzed azidation of tertiary C-H bonds suitable for late-stage functionalization. Nature 2015, 600–604. [Google Scholar] [CrossRef] [Green Version]
  54. Fumagalli, G.; Rabet, P.T.G.; Boyd, S.; Greaney, M.F. Three-component azidation of styrene-type double bonds: Light-switchable behavior of a copper photoredox catalyst. Angew. Chem. Int. Ed. 2015, 54, 11481–11484. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  55. Liu, T.; Zhu, J.; Sun, X.; Cheng, L.; Wu, L. I2/TBHP mediated divergent C(sp3)-P cleavage of allenylphosphine oxides: Substituent-controlled regioselectivity. Adv. Synth. Catal. 2019, 361, 3532–3537. [Google Scholar] [CrossRef]
  56. Jiang, H.; He, Y.; Cheng, Y.; Yu, S. Radical alkynyltrifluoromethylation of alkenes initiated by an electron donor–acceptor complex. Org. Lett. 2017, 19, 1240–1243. [Google Scholar] [CrossRef]
  57. Mangaonkar, S.R.; Kole, P.B.; Singh, F.V. Oxidation of organosulfides to organosulfones with trifluoromethyl 3-oxo-1λ3,2-benziodoxole-1(3H)-carboxylate as an oxidant. Synlett 2018, 29, 199–202. [Google Scholar] [CrossRef]
  58. Shen, K.; Wang, Q. Copper-catalyzed aminoalkynylation of alkenes with hypervalent iodine reagents. Chem. Sci. 2017, 8, 8265–8270. [Google Scholar] [CrossRef] [Green Version]
  59. Mukherjee, S.; Garza-Sauchez, R.A.; Tlahuext-Aca, A.; Glorius, F. Alkynylation of Csp2(O)-H bonds enabled by photoredox-mediated hydrogen-atom transfer. Angew. Chem. Int. Ed. 2017, 56, 14723–14726. [Google Scholar] [CrossRef]
  60. Nicolai, S.; Waser, J. Pd(0)-catalyzed oxy- and aminoalkynylation of olefins for the synthesis of tetrahydrofurans and pyrrolidines. Org. Lett. 2011, 13, 6324–6327. [Google Scholar] [CrossRef] [Green Version]
  61. Brand, J.P.; Waser, J. Direct alkynylation of thiophenes: Cooperative activation of TIPS–EBX with gold and Brønsted acids. Angew. Chem. Int. Ed. 2010, 49, 7304–7307. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  62. González, D.F.; Brand, J.P.; Waser, J. Asymmetric synthesis of 4-amino-4H-chromenes by organocatalytic oxa-Michael/aza-Baylis–Hillman tandem reactions. Chem. Eur. J. 2010, 16, 9457–9461. [Google Scholar]
  63. Baker, G.P.; Mann, G.; Sheppard, N.; Tetlow, A.J. The structure of o-iodosobenzoic acid and of certain derivatives. J. Chem. Soc. 1965, 3721–3728. [Google Scholar] [CrossRef]
  64. Kraszkiewicz, L.; Skulski, L. Optimized syntheses of iodylarenes from iodoarenes, with sodium periodate as the oxidant. Part II. Arkivoc 2003, 6, 120–125. [Google Scholar] [CrossRef] [Green Version]
  65. Bosset, C.; Coffinier, R.; Peioto, P.A.; Assal, M.E.; Miqueu, K.; Sotiropoulos, M.; Pouységu, L.; Quideau, S. Asymmetric hydroxylative phenol dearomatization promoted by chiral binaphthylic and biphenylic iodanes. Angew. Chem. Int. Ed. 2014, 53, 9860–9864. [Google Scholar] [CrossRef] [PubMed]
  66. Maity, A.; Hyun, S.-M.; Powers, D.C. Oxidase catalysis via aerobically generated hypervalent iodine intermediates. Nat. Chem. 2018, 10, 200–204. [Google Scholar] [CrossRef]
  67. Chandrasekar, S.; Sekar, G. An efficient synthesis of iminoquinones by a chemoselective domino ortho-hydroxylation/oxidation/imidation sequence of 2-aminoarylketones. Org. Biomol. Chem. 2016, 14, 3053–3060. [Google Scholar] [CrossRef]
  68. Wirth, T. IBX—New reactions with an old reagent. Angew. Chem. Int. Ed. 2001, 40, 2812–2814. [Google Scholar] [CrossRef]
  69. Ladziata, U.; Zhdankin, V.V. Hypervalent iodine(V) reagents in organic synthesis. Arkivoc 2006, 9, 26–58. [Google Scholar] [CrossRef] [Green Version]
  70. Sharma, P.; Kaur, N.; Pareek, A.; Kishore, D. An insight in to general features of IBX (2-iodoxybenzoic acid). Sci. Revs. Chem. Commun. 2013, 3, 16–42. [Google Scholar]
  71. Plumb, J.B.; Harper, D.J. IBX is reported to be a shock-sensitive explosive. Chem. Eng. News 1990, 68, 3. [Google Scholar]
  72. Dess, D.B.; Martin, J.C.J. A useful 12-I-5 triacetoxyperiodinane (the Dess-Martin periodinane) for the selective oxidation of primary or secondary alcohols and a variety of related 12-I-5 species. J. Am. Chem. Soc. 1991, 113, 7277–7287. [Google Scholar] [CrossRef]
  73. Boeckmann, R.K.; Shao, P.; Mullins, J.J. The Dess-Martin periodinane: 1,1,1-triacetoxy-1,1-dihydro-1,2-benziodoxolo-3(1H)-one. Org. Synth. 2000, 77, 141–152. [Google Scholar]
  74. More, J.D.; Finney, N.S. A simple and advantageous protocol for the oxidation of alcohols with o-iodoxybenzoic acid (IBX). Org. Lett. 2002, 4, 3001–3003. [Google Scholar] [CrossRef]
  75. Tiffner, M.; Stockhammer, L.; Schörgenhumer, J.; Röser, K.; Waser, M. Towards an asymmetric organocatalytic α-azidation of β-ketoesters. Molecules 2018, 23, 1142. [Google Scholar] [CrossRef] [Green Version]
  76. Bernini, R.; Barontini, M.; Spatafora, C. New lipophilic piceatannol derivatives exhibiting antioxidant activity prepared by aromatic hydroxylation with 2-iodoxybenzoic acid (IBX). Molecules 2009, 14, 4669–4681. [Google Scholar] [CrossRef] [Green Version]
  77. Hussain, H.; Green, I.R.; Ahmed, I. Journey describing applications of oxone in synthetic chemistry. Chem. Rev. 2013, 113, 3329–3371. [Google Scholar] [CrossRef]
  78. Thottumkara, A.P.; Bowsher, M.S.; Vinod, T.K. In situ generation of o-iodoxybenzoic acid (IBX) and the catalytic use of it in oxidation reactions in the presence of oxone as a co-oxidant. Org. Lett. 2005, 7, 2933–2936. [Google Scholar] [CrossRef] [PubMed]
  79. Uyanik, M.; Akakura, M.; Ishihara, K. 2-Iodoxybenzenesulfonic acid as an extremely active catalyst for the selective oxidation of alcohols to aldehydes, ketones, carboxylic acids, and enones with oxone. J. Am. Chem. Soc. 2009, 131, 251–262. [Google Scholar] [CrossRef]
  80. Miura, T.; Nakashima, K.; Tada, N.; Itoh, A. An effective and catalytic oxidation using recyclable fluorous IBX. Chem. Commun. 2011, 47, 1875–1877. [Google Scholar] [CrossRef]
  81. Moorthy, J.N.; Senapati, K.; Parida, K.N.; Jhulki, S.; Sooraj, K.; Nair, N.N. Twist does a twist to the reactivity: Stoichiometric and catalytic oxidations with twisted tetramethyl-IBX. J. Org. Chem. 2011, 76, 9593–9601. [Google Scholar] [CrossRef] [PubMed]
  82. Ishihara, K.; Uyanik, M.; Crockett, R. 2-Iodoxy-5-methylbenzenesulfonic acid-catalyzed selective oxidation of 4-bromobenzyl alcohol to 4-bromobenzaldehyde or 4-bromobenzoic acid with oxone. Org. Synth. 2012, 89, 105–114. [Google Scholar]
  83. Bikshapathi, R.; Prathima, P.S.; Rao, V.J. Hypervalent iodine catalysis for selective oxidation of Baylis–Hillman adducts via in situ generation of o-iodoxybenzoic acid (IBX) from 2-iodosobenzoic acid (IBA) in the presence of oxone. New J. Chem. 2016, 40, 10300–10304. [Google Scholar] [CrossRef]
  84. Yakura, T.; Fujiwara, T.; Yamada, A.; Nambu, H. 2-Iodo-N-isopropyl-5-methoxybenzamide as a highly reactive and environmentally benign catalyst for alcohol oxidation. Beilstein J. Org. Chem. 2018, 14, 971–978. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  85. Mishra, A.K.; Moorthy, J.N. Mechanochemical catalytic oxidations in the solid state with in situ-generated modified IBX from 3,5-di-tert-butyl-2-iodobenzoic acid (DTB-IA)/Oxone. Org. Chem. Front. 2017, 4, 343–349. [Google Scholar] [CrossRef]
  86. Jhulki, S.; Seth, S.; Mondal, M.; Moorthy, J.N. Facile organocatalytic domino oxidation of diols to lactones by in situ-generated TetMe-IBX. Tetrahedron 2014, 70, 2286–2293. [Google Scholar] [CrossRef]
  87. Yakura, T.; Horiuchi, Y.; Nishimura, Y.; Yamada, A.; Nambu, H.; Fujiwara, T. Efficient oxidative cleavage of tetrahydrofuran-2-methanols to γ-lactones by a 2-iodobenzamide catalyst in combination with Oxone®. Adv. Synth. Catal. 2016, 358, 869–873. [Google Scholar] [CrossRef]
  88. Yakura, T.; Yamauchi, Y.; Tian, Y.; Omoto, M. Catalytic hypervalent iodine oxidation of p-dialkoxybenzenes to p-quinones using 4-iodophenoxyacetic acid and Oxone®. Chem. Pharm. Bull. 2008, 56, 1632–1634. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  89. Yakura, T.; Tian, Y.; Yamauchi, Y.; Omoto, M.; Konishi, T. Catalytic hypervalent iodine oxidation using 4-iodophenoxyacetic acid and Oxone®: Oxidation of p-alkoxyphenols to p-benzoquinones. Chem. Pharm. Bull. 2009, 57, 252–256. [Google Scholar] [CrossRef] [Green Version]
  90. Zagulyaeva, A.A.; Banek, C.T.; Yusubov, M.S.; Zhdankin, V.V. Hofmann rearrangement of carboxamides mediated by hypervalent iodine species generated in situ from iodobenzene and oxone: Reaction scope and limitations. Org. Lett. 2010, 12, 4644–4647. [Google Scholar] [CrossRef] [PubMed]
  91. Yakura, T.; Omoto, M.; Yamauchi, Y.; Tian, Y.; Ozono, A. Hypervalent iodine oxidation of phenol derivatives using a catalytic amount of 4-iodophenoxyacetic acid and Oxone® as a co-oxidant. Tetrahedron 2010, 66, 5833–5840. [Google Scholar] [CrossRef]
  92. Ren, J.; Lu, L.; Xu, J.; Yu, T.; Zeng, B.-B. Selective oxidation of 1-tetralones to 1,2-naphthoquinones with IBX and to 1,4-naphthoquinones with Oxone® and 2-iodobenzoic acid. Synthesis 2015, 47, 2270–2280. [Google Scholar] [CrossRef]
  93. Frigerio, M.; Santagostino, M.; Sputore, S. A user-friendly entry to 2-iodoxybenzoic acid (IBX). J. Org. Chem. 1999, 64, 4537–4538. [Google Scholar] [CrossRef]
  94. China, H.; Tanihara, K.; Sasa, H.; Kikushima, K.; Dohi, T. Regiodivergent oxidation of alkoxyarenes by hypervalent iodine/oxone® system. Catal. Today 2020, 348, 2–8. [Google Scholar] [CrossRef]
  95. Parida, K.N.; Moorthy, J.N. Synthesis of o-carboxyarylacrylic acids by room temperature oxidative cleavage of hydroxynaphthalenes and higher aromatics with Oxone. J. Org. Chem. 2015, 80, 8354–8360. [Google Scholar] [CrossRef]
  96. Costello, J.P.; Ferreira, E.M. Regioselectivity influences in platinum-catalyzed intramolecular alkyne O-H and N-H additions. Org. Lett. 2019, 21, 9934–9939. [Google Scholar] [CrossRef]
  97. Koguchi, S.; Mihoya, A.; Mimura, M. Alcohol oxidation via recyclable hydrophobic ionic liquid-supported IBX. Tetrahedron 2016, 72, 7633–7637. [Google Scholar] [CrossRef]
  98. Boelke, A.; Kuczmera, T.J.; Caspers, L.D.; Lork, E.; Nachtsheim, B.J. Iodolopyrazolium salts: Synthesis, derivatizations, and applications. Org. Lett. 2020, 22, 7261–7266. [Google Scholar] [CrossRef]
  99. Kalaj, M.; Mimeni, M.R.; Bentz, K.C.; Barcus, K.S.; Palomba, J.M.; Paesani, F.; Cohen, S.M. Halogen bonding in UiO-66 frameworks promotes superior chemical warfare agent simulant degradation. Chem. Commun. 2019, 55, 3481–3484. [Google Scholar] [CrossRef]
  100. Sun, W.; Cama, L.D. Preparation of 6H-benzo[c]chromenes as estrogen receptor modulators. PCT Int. Appl. 2004, WO 2004073612, A2. [Google Scholar]
  101. Caspers, L.D.; Spils, J.; Damrath, M.; Lork, E.; Nachtsheim, B.J. One-pot synthesis and conformational analysis of six-membered cyclic iodonium salts. J. Org. Chem. 2020, 85, 9161–9178. [Google Scholar] [CrossRef] [PubMed]
  102. Miles, K.C.; Le, C.C.; Stambuli, J.P. Direct carbocyclizations of benzoic acids: Catalyst-controlled synthesis of cyclic ketones and the development of tandem aHH (acyl Heck-Heck) reactions. Chem. Eur. J. 2014, 20, 11336–11339. [Google Scholar] [CrossRef]
  103. Hamley, P.; Pimm, A.; Tinker, A. Preparation of spiro[pyridoisoindolequinazoline] derivatives and analogs as nitric oxide synthase inhibitors. PCT Int. Appl. 1999, WO 9901455, A1. [Google Scholar]
  104. Declas, N.; Vaillant, F.L.; Waser, J. Revisiting the Urech synthesis of hydantoins: Direct access to enantiopure 1,5-substituted hydantoins using cyanobenziodoxolone. Org. Lett. 2019, 21, 524–528. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  105. Bertho, S.; Rey-Rodriguez, R.; Colas, C.; Retailleau, P.; Gillaizeau, I. Regio- and stereoselective iron-catalyzed oxyazidation of enamides using a hypervalent iodine reagent. Chem. Eur. J. 2017, 23, 17674–17677. [Google Scholar] [CrossRef] [PubMed]
  106. Lu, B.; Wu, J.; Yoshikai, N. Palladium-catalyzed condensation of N-aryl imines and alkynylbenziodoxolones to form multisubstituted furans. J. Am. Chem. Soc. 2014, 136, 11598–11601. [Google Scholar] [CrossRef]
  107. Chen, M.; Huang, Z.-T.; Zheng, Q.-Y. Organic base-promoted enantioselective electrophilic cyanation of β-keto esters by using chiral phase-transfer catalysts. Org. Biomol. Chem. 2015, 13, 8812–8816. [Google Scholar] [CrossRef]
  108. Hari, D.P.; Waser, J. Enantioselective copper-catalyzed oxy-alkynylation of diazo compounds. J. Am. Chem. Soc. 2017, 139, 8420–8423. [Google Scholar] [CrossRef] [Green Version]
  109. Mocci, F.; Uccheddu, G.; Frongia, A.; Cerioni, G. Solution structure of some λ3 iodanes: An 17O NMR and DFT study. J. Org. Chem. 2007, 72, 4163–4168. [Google Scholar] [CrossRef]
  110. Ito, E.; Fukushima, T.; Kawakami, T.; Murakami, K.; Itami, K. Catalytic dehydrogenative C–H imidation of arenes enabled by photo-generated hole donation to sulfonimide. Chem 2017, 2, 383–392. [Google Scholar] [CrossRef] [Green Version]
  111. Jia, K.; Pan, Y.; Chen, Y. Selective carbonyl-C (sp3) bond cleavage to construct ynamides, ynoates, and ynones by photoredox catalysis. Angew. Chem. Int. Ed. 2017, 56, 2478–2481. [Google Scholar] [CrossRef]
  112. Iinuma, M.; Moriyama, K.; Togo, H. Oxidation of alcohols to aldehydes or ketones with 1-acetoxy-1,2-benziodoxole-3(1H)-one derivatives. Eur. J. Org. Chem. 2014, 2014, 772–780. [Google Scholar] [CrossRef]
Figure 1. 2-Iodosobenzoic acid (IBA) and its precursor 2-iodobenzoic acid (2-IB).
Figure 1. 2-Iodosobenzoic acid (IBA) and its precursor 2-iodobenzoic acid (2-IB).
Molecules 26 01897 g001
Figure 2. Synthesis of 2-iodosobenzoic acids (IBAs) and 2-iodoxybenzoic acids (IBXs) from 2-iodobenzoic acids (2-IBs).
Figure 2. Synthesis of 2-iodosobenzoic acids (IBAs) and 2-iodoxybenzoic acids (IBXs) from 2-iodobenzoic acids (2-IBs).
Molecules 26 01897 g002
Figure 3. Solvent effect on the selective synthesis of IBA 2a from 2-IB 1a using Oxone®.
Figure 3. Solvent effect on the selective synthesis of IBA 2a from 2-IB 1a using Oxone®.
Molecules 26 01897 g003
Figure 4. Substrate scope of the selective preparation of IBAs 2bv using Oxone® under aqueous conditions. (a) This reaction was performed at 30 °C in the presence of 1.2 equivalents of Oxone®. (b) Performed using 5 equivalents of Oxone® at 40 °C.
Figure 4. Substrate scope of the selective preparation of IBAs 2bv using Oxone® under aqueous conditions. (a) This reaction was performed at 30 °C in the presence of 1.2 equivalents of Oxone®. (b) Performed using 5 equivalents of Oxone® at 40 °C.
Molecules 26 01897 g004
Scheme 1. Synthesis of tricyclic hypervalent iodine compound 4.
Scheme 1. Synthesis of tricyclic hypervalent iodine compound 4.
Molecules 26 01897 sch001
Scheme 2. Effect of a ferric salt on the selective synthesis of IBA 2a.
Scheme 2. Effect of a ferric salt on the selective synthesis of IBA 2a.
Molecules 26 01897 sch002
Figure 5. Optimization of (a) temperature using 1.0 equivalent of Oxone® and (b) amount of Oxone® at 60 °C for the synthesis of IBA 2a from 2-IB 1a in the presence of 2.5 mol% FeCl3 (MeCN/DW (1/1)).
Figure 5. Optimization of (a) temperature using 1.0 equivalent of Oxone® and (b) amount of Oxone® at 60 °C for the synthesis of IBA 2a from 2-IB 1a in the presence of 2.5 mol% FeCl3 (MeCN/DW (1/1)).
Molecules 26 01897 g005
Scheme 3. Short-time selective synthesis of IBAs 2a–j in the presence of ferric salt.
Scheme 3. Short-time selective synthesis of IBAs 2a–j in the presence of ferric salt.
Molecules 26 01897 sch003
Figure 6. Benziodoxole alkoxides 6ah by solvolytic functionalization of IBA 2a.
Figure 6. Benziodoxole alkoxides 6ah by solvolytic functionalization of IBA 2a.
Molecules 26 01897 g006
Scheme 4. Transformation of IBAs 2ai to benziodoxole acetates 7ai.
Scheme 4. Transformation of IBAs 2ai to benziodoxole acetates 7ai.
Molecules 26 01897 sch004
Sample Availability
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

China, H.; Kageyama, N.; Yatabe, H.; Takenaga, N.; Dohi, T. Practical Synthesis of 2-Iodosobenzoic Acid (IBA) without Contamination by Hazardous 2-Iodoxybenzoic Acid (IBX) under Mild Conditions. Molecules 2021, 26, 1897. https://doi.org/10.3390/molecules26071897

AMA Style

China H, Kageyama N, Yatabe H, Takenaga N, Dohi T. Practical Synthesis of 2-Iodosobenzoic Acid (IBA) without Contamination by Hazardous 2-Iodoxybenzoic Acid (IBX) under Mild Conditions. Molecules. 2021; 26(7):1897. https://doi.org/10.3390/molecules26071897

Chicago/Turabian Style

China, Hideyasu, Nami Kageyama, Hotaka Yatabe, Naoko Takenaga, and Toshifumi Dohi. 2021. "Practical Synthesis of 2-Iodosobenzoic Acid (IBA) without Contamination by Hazardous 2-Iodoxybenzoic Acid (IBX) under Mild Conditions" Molecules 26, no. 7: 1897. https://doi.org/10.3390/molecules26071897

Article Metrics

Back to TopTop