Next Article in Journal
Ranolazine Interacts Antagonistically with Some Classical Antiepileptic Drugs—An Isobolographic Analysis
Next Article in Special Issue
5-Chloroisoxazoles: A Versatile Starting Material for the Preparation of Amides, Anhydrides, Esters, and Thioesters of 2H-Azirine-2-carboxylic Acids
Previous Article in Journal
Crystallization and Structural Properties of Oleogel-Based Margarine
Previous Article in Special Issue
Design, Synthesis and Assay of Novel Methylxanthine–Alkynylmethylamine Derivatives as Acetylcholinesterase Inhibitors
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

N,N′-Di-Boc-2H-Isoindole-2-carboxamidine—First Guanidine-Substituted Isoindole

Laboratory for Physical-Organic Chemistry, Division of Organic Chemistry and Biochemistry, Ruđer Bošković Institute, 10001 Zagreb, Croatia
*
Author to whom correspondence should be addressed.
Molecules 2022, 27(24), 8954; https://doi.org/10.3390/molecules27248954
Submission received: 2 December 2022 / Revised: 10 December 2022 / Accepted: 11 December 2022 / Published: 15 December 2022
(This article belongs to the Special Issue Chemistry of Nitrogen Heterocyclic Compounds)

Abstract

:
Synthesis of N,N′-Di-Boc-2H-isoindole-2-carboxamidine, the first representative of isoindoles containing guanidine functionality, was carried out. The cycloaddition reactivity of this new Diels–Alder heterodiene was studied and the title compound was employed as a cycloaddition delivery reagent for guanidine functionality. Higher reactivity was found in comparison with the corresponding pyrrole derivative. Substitution with fluorine or guanidine functionality does not change the reactivities of isoindoles, and these findings are in good accord with computational results.

Graphical Abstract

1. Introduction

Guanidines are a class of nitrogen-containing molecules with very interesting physico-chemical properties, [1] especially very high basicity [2] and biological activity [3]. Hence, various aspects of guanidine chemistry were extensively studied computationally and experimentally, including their synthesis. The most viable synthetic routes towards polycyclic, complex organic molecules include the introduction of guanidine functionality at the later stages of multi-step synthesis, such as in tetrodotoxin synthesis [4,5]. One of the most efficient ways for the construction of polycyclic molecules is Diels–Alder (DA) cycloaddition; however, the cycloaddition approach to polycycles containing guanidine functionality has been rarely utilized. Diels–Alder cycloadditions, which involve diene or dienophile partners possessing guanidine functionality, were summarized in Figure 1 [6,7] which also depicts guanidine delivery cycloaddition reagents [8,9]. All of these are unsymmetrical diene molecules, and their cycloaddition reactions lead to the formation of unsymmetrical products.
For the purpose of our studies towards the synthesis of polycyclic molecules, symmetrical diene reagents were required, and, in view of earlier studies [10,11], pyrrole and isoindole N-carboxamidine derivatives 1 and 2 were selected. These molecules possess C2v symmetry [12] and upon cycloaddition form guanidine functionality which is ‘protected’ in pyrrole moiety. Whereas pyrrole-2-carboxamidine 1 is a known compound, its DA cycloaddition properties were not reported. Only rhodium-catalyzed [4+3] cycloaddition was utilized to prepare tropane bicyclo [3.2.1] octane skeleton 3, where pyrrole acted as a dipolarophile partner (Figure 1) [13]. Corresponding isoindole-2-carboxamidines have not been synthetized previously.
The objective of this work is to prepare isoindole-2-carboxamidine 2, the first representative of an isoindole guanidine cycloaddition delivery reagent, and assess its cycloaddition properties experimentally and computationally. Our preliminary DFT computational study revealed that this approach is feasible and both pyrrole and isoindole dienes are predicted to have sufficient reactivity [14].
Figure 1. Used guanidines in Diels–Alder reactions and 2-carboxamidine dienes described in this work [6,7,13].
Figure 1. Used guanidines in Diels–Alder reactions and 2-carboxamidine dienes described in this work [6,7,13].
Molecules 27 08954 g001

2. Results and Discussion

2.1. Cycloadditions of 1-Carboxamidine Pyrrole

The 1-(N′,N′-Di-Boc)pyrrole carboxamidine 1 was prepared according to the literature starting from 3-pyrroline [15,16]. Pyrrole 1 showed poor reactivity in cycloaddition reactions when acting as 1,3-diene (Scheme 1). For instance, the thermal reaction of 1 with N-methylmaleimide did not provide the expected cycloadduct 5. In order to increase its reactivity, an extremely high-pressure technique was employed [17]. The pressurization at 10 kbar, for 2 days, in dichloromethane resulted in the formation of 5 in 51% yield. It was found that product 5 was unstable in CHCl3 solution (also in solid state) and quickly cycloreverses back to reactants. This behavior could explain the failure of thermal conditions. An alternative way to increase dienes’ reactivity is to employ more reactive dienophiles. Thermal and high-pressure reactions with naphthoquinone did not provide conclusive evidence for the formation of a cycloadduct. Equally unsuccessful were reactions of mechanochemically solid state in-situ-generated imide 8 [18] in a ball mill, due to harsh conditions for the guanidine moiety of 1. Reaction with related anhydride 10 [19] also provided a complex reaction mixture.

2.2. Synthesis of Isoindoles and Their Reactivity

Further increase in the reactivity of pyrroles could be achieved by the addition of a benzene ring, i.e., to use isoindole derivatives as dienes. Synthesis of isoindole precursor 20 in four reaction steps is shown in Scheme 2. It follows the already established synthesis of 7-azabenzonorbornadiene 18, and subsequent guanylation with N,N′-Di-Boc-1H-pyrazole-1-carboxamidine provided 20 in 62% yield. Alternatively, 20 could be prepared by in-situ-generated benzyne cycloaddition with pyrrole 1 (84%). Preparation of a nitro derivative of 20 was achieved by in-situ generation of 4-nitro benzyne from iodonium salt 22 [20] and its reaction with pyrrole 1, which provided cycloadduct 23 in 32% yield. These reactions show that the reactivity of pyrrole-1-carboxamidine 1 could be increased by the presence of a highly reactive dienophile such as arynes.
Warrener’s cycloaddition/elimination/cycloreversion method employing bis(2-pyridyl)-sym-1,2,4,5-tetrazine 24 [21,22] was used for the generation of isoindole 2 (Scheme 3). The formation of 2 was confirmed by 1H NMR spectroscopy, by spectrum recorded 30 min after the addition of 24 to a solution of 20 in an NMR tube (Figure 2). The most characteristic signals which indicate the presence of 2 are a singlet of H1,3 appearing at δ 7.57, whereas aromatic multiplets of H4,7 and H5,6 are found at δ 7.36 and 6.85. However, trapping experiments offer indirect but more solid evidence of its formation.
Scheme 4 summarizes the cycloaddition properties of isoindole-2-carboxamidine 2. When 2 was generated in the presence of dienophiles N-methylmaleimide, dimethylacetylenedicarboxylate (DMAD), and benzoquinone, corresponding cycloadducts 26, 28, and 29 were obtained (in 91, 80 and 77% yields, respectively). In variance, norbornenes 10, 20, 30, and 31 did not react or afforded intractable mixtures, regardless of reaction conditions (thermal or high pressure). The endo-adduct 26 was solely formed, as shown by the single methyl resonance at δ 2.28 in the 1H NMR spectrum (see Supplementary Materials), and the endo-configuration is proven by the shielding of the N-methyl protons by the ring current effect of the aromatic ring [22]. Furthermore, the exo-protons are multiplets, is characteristic of the endo-adducts of isoindoles [23,24]. This endo-stereospecificity is similar to N-benzyl-isoindole cycloaddition [25] and in variance with maleic anhydride reactions of isoindoles where exo/endo mixtures were formed, [24,26] whereas the outcome of cycloadditions of 2-substituted isoindoles with tolyl-maleimide was not specified [27].
An interesting feature of the 1H NMR spectra of cycloaddition products 26 and 28 is the broadness of bridgehead signals at 20 °C. Recording the spectra at 50 °C led to the sharpening of the signal, whereas cooling down to 5 °C gives two sets of bridgehead signals, which are associated with nitrogen inversion [28]. The N-inversion barrier in 26 is estimated to be low, 13.5 kcal mol−1 in deuterated chloroform. A similar broadness of bridgehead protons was observed for pyrrole cycloadduct 5; however, this adduct is thermally unstable and quickly cycloreverses.
In continuation, the electronics of isoindoles were altered by fluorine substituents on the aromatic ring and positioning of the guanidine functionality. Tetrafluoro isoindole precursor 37 was prepared in 35% yield by mechanochemical guanylation [29] of the known 7-azabenzonorbornadiene 36 [30,31] (Scheme 5). It was found that fluorine substitution did not have noticeable effects on the cycloaddition reactivity of isoindole. When tetrafluoroisoindole 38 was trapped in a tetrazine reaction with 37, the endo-cycloadduct 39 was obtained in 11% yield, while, similarly to isoindole 2, tetrafluoro derivative 38 also did not react with its precursor 37.
Until now, isoindoles substituted on an aromatic ring with the nitrogen atom have been known only with the nitro group, [32,33,34,35] and we prepared the first example of a guanidine aromatic-ring-substituted isoindole. The synthetic route for the introduction of guanidine functionality at position 5 of the isoindole ring in 45 is depicted in Scheme 6. In-situ-generated 4-nitro-benzyne was reacted with 1-benzyloxycarbonyl pyrrole 41 to afford the known cycloadduct 42 [23]. The nitro group was reduced by Al/Hg and the amine 43 was obtained in 66% yield. Guanylation in solution led to the formation of isoindole precursor 44 in 86% yield. This compound was treated with tetrazine 24 in chloroform and intermediate isoindole 45 was trapped as N-methylmaleimide cycloadduct 46 (66%). The change in the position of guanidine functionality and N-CBz substitution did not increase the cycloaddition reactivity of isoindole. Analogously to isoindoles 5 and 38, in the case of 45, a reaction with 44 as a dienophile was not observed. These results indicate the similar cycloaddition reactivity of all three investigated guanidine isoindoles.
Previous density functional theory (DFT) calculations B3LYP/6-31G(d) predict that activation energies (Ea) for reactions of pyrrole and isoindole-2-carboxamidine with DMAD are 32.37 and 23.17 kcal mol−1, respectively, indicating that the amidine substitution decreases Ea by 4–5 kcal mol−1 in comparison to parent unsubstituted dienes, whereas Boc protection of amidinopyrrole causes a further drop in Ea by 2.5 kcal mol−1. Now, these theoretical predictions are supplemented with the M062X/6-311+G** calculations [23] of the reaction of acetylene with pyrrole and isoindoles. All located transition states possess structures resembling the synchronous concerted mechanism of Diels–Alder reactions, such as the one illustrated in Figure 3. Computed activation-free energies (ΔG values) are given in Figure 3 and reveal similar predictions to the previously obtained B3LYP calculations. Firstly, N-substitution with amidine lowers ΔG by 1.5–2.3 kcal mol−1. The largest difference in ΔG values was obtained when pyrrole was fused with a benzene ring in isoindoles, which is in qualitative accordance with published AM1 results [23]. The position of an amidine (guanidine) substituent and the addition of fluorine atoms has only a marginal effect on the ΔG values, with differences in the reactivity of three experimentally studied isoindoles within 0.54 kcal mol−1. These predictions are in full accordance with almost identical experimentally observed reactivities of three isoindoles.

3. Materials and Methods

3.1. General

Solvents and chemicals were obtained from Tokyo Chemical Industry (Tokyo, Japan) and Sigma Aldrich (Burlington, VT, USA). Kemika (Zagreb, Croatia), Sigma Aldrich, and VWR Chemicals (Radnor, PA, USA) supplied the solvents, which were used without further purification, unless otherwise stated. The NMR spectra were recorded on Bruker Avance 300 MHz and Bruker Avance 600 MHz spectrometers in deuterated solvents. Chemical shifts (δ) are given in ppm using tetramethylsilane (TMS) as an internal standard, whereas coupling constants (J) are expressed in Hertz (Hz). The following abbreviations were used to describe multiplicity in the 1H spectra: (s) singlet; (d) doublet; (dd) doublet of doublets; (t) triplet; (m) multiplet; (brs) broad signal. Fourier Transform Infrared Attenuated Total Reflection PerkinElmer UATR Two Spectrometer (range 400–4000 cm−1) was used to record infrared spectra (FTIR-ATR). Milling reactions were carried out in Retsch MM400 vibrational mill (frequency 30 Hz), using stainless steel (SS) vials (10 mL) and one 12 mm size SS milling ball. High-pressure reactions were performed in Teflon vials (V = 1.5 mL) using a high-pressure-piston cylinder apparatus (Unipress, Polish Academy of Sciences), and pentane as a pressure-transmitting liquid. Thin-layer chromatography (TLC) was performed on silica-gel plates (silica gel 60 F254, Merck), whereas silica gel (Silica gel 60, 0.063–0.200 mm, Merck, Darmstadt, Germany) was used for column chromatography. High-resolution mass spectra (HRMS) were recorded on Agilent 6550 Series Accurate-Mass-Quadrupole Time-of-Flight (Q-TOF) Agilent 1290 Infinity II instrument.

3.2. Synthesis of Cycloadduct 5

Pyrrole 1 (20 mg, 0.065 mmol) and N-methylmaleimide (11 mg, 0.098 mmol) were dissolved in CH2Cl2 (1 mL) and the solution was subjected to high pressure at 10 kbar for 48 h at room temperature. The reaction mixture was evaporated and purified by column chromatography, starting with petroleum ether/EtOAc mixture from 5:1 to 2.5:1. Two fractions were isolated, pyrrole 1 (11 mg) and product 5 as colorless solid (14 mg, 51%).
1H NMR (CDCl3), δ/ppm: 1.49 (s, 18H, t-Bu), 2.84 (s, 3H, NCH3), 3.69 (dd, 1H, J = 3.3, 1.5 Hz, exo-H), 5.31 (brs, 2H, N bridge), 6.39 (brs, 2H, C=CH), 10.60 (brs, 1H, NH),
FTIR-ATR νmax/cm−1: 2979, 1700 (C=O), 1600 (C=O), 1275, 1121.

3.3. Synthesis of 20

Benzonorbornadiene 18 (25 mg, 0.175 mmol) and N,N′-Di-Boc-1H-pyrazole-1-carboxamidine (49 mg, 0.158 mmol) were dissolved in CHCl3 (1 mL) and stirred at room temperature for 7 days. The reaction mixture was purified by radial chromatography using CH2Cl2. Product 20 was isolated as a white solid (42 mg, 62%).
m.p. 158–160 °C,
1H NMR (CDCl3), δ/ppm: 1.49 (s, 9H, t-Bu), 1.50 (s, 9H, t-Bu), 5.84 (brs, 2H, N bridge), 6.98 (dd, 1H, J = 5.2, 3.2 Hz, Ar), 7.02 (d, 2H, J = 2.6 Hz, C=CH), 7.27 (dd, 1H, J = 5.2, 3.2 Hz, Ar), 10.63 (brs, 1H, NH),
13C NMR (CDCl3), δ/ppm: 28.1 (t-Bu), 28.2 (t-Bu), 65.8 (N bridge), 68.2 (N bridge), 79.8, 82.2, 120.9 (C=C), 125.3, 144.1, 147.5, 150.2 (C=O), 151.2 (C=N), 162.8 (C=O),
FTIR-ATR νmax/cm−1: 2977, 1756 (C=O), 1636 (C=O),
HRMS-MALDI found: 386.2087, calculated for C21H28N3O4 [MH]+: 386.2080.

3.4. Synthesis of Cycloadduct 23

Under argon, dry toluene (1 mL) was added to pyrrole 1 (20 mg, 0.064 mmol) and iodonium salt 22 (35 mg, 0.064 mmol). LiHDMS solution in toluene (64 μL, 0.064 mmol, 1 M) was added dropwise and the resulting mixture was stirred for 1 h at room temperature. The reaction was quenched with saturated NH4Cl solution (5 mL) and extracted with EtOAc (3 × 10 mL), and combined extracts were dried with Na2SO4 and evaporated. The crude mixture was purified by radial chromatography using petroleum ether/diethyl ether 5:1 and gradually increasing polarity to 1:1. Product 23 was isolated as a yellow solid (9 mg, 32%).
m.p. 88–90 °C,
1H NMR (CDCl3), δ/ppm: 1.50 (s, 18H, t-Bu), 5.93 (brs, 2H, N bridge), 7.05 (brs, 2H, C=CH), 7.39 (d, 1H, J = 7.9 Hz, Ar), 7.99 (dd, 1H, J = 7.9, 1.9 Hz, Ar), 8.08 (d, 1H, J = 1.9 Hz, Ar), 10.68 (brs, 1H, NH),
13C NMR (CDCl3), δ/ppm δ/ppm: 28.0 (t-Bu), 28.1 (t-Bu), 80.3 (N bridge), 82.5, 82.6, 82.7 (N bridge), 122.4 (C=C), 145.9, 149.6, 149.8, 150.0, 150.1, 151.9, 155.0 (C=N), 162.5 (C=O), 162.6 (C=O),
FTIR-ATR νmax/cm−1: 2979, 1754 (C=O), 1683 (C=O),
HRMS-MALDI found: 431.1942, calculated for C21H27N4O6 [MH]+: 431.1931.

3.5. Synthesis of Cycloadduct 26

Isoindole precursor 20 (98 mg, 0.25 mmol), bis(2-pyridyl)-sym-1,2,4,5-tetrazine 24 (59 mg, 0.25 mmol) and N-methylmaleimide (28 mg, 0.25 mmol) were dissolved in CHCl3 (2.5 mL) and stirred overnight at room temperature. The reaction mixture was purified by radial chromatography using CH2Cl2/MeOH mixture 99:1. Cycloadduct 26 was isolated as a white solid (107 mg, 91%).
m.p. 120–122 °C,
1H NMR (CDCl3), δ/ppm: 1.50 (s, 18H, t-Bu), 2.28 (s, 3H, NCH3), 3.88 (d, 1H, J = 1.9 Hz, exo-H), 5.76 (brs, 2H, N bridge), 7.21 (dd, 1H, J = 5.5, 3.1 Hz, Ar), 7.28 (dd, 1H, J = 5.5, 3.1 Hz, Ar), 10.60 (brs, 1H, NH),
13C NMR (CDCl3), δ/ppm: 23.9 (exo-H), 28.1 (t-Bu), 47.2 (NCH3), 62.6 (N bridge), 80.4, 82.8, 121.6, 126.0, 139.3, 150.2 (C=O), 152.6, (C=O), 162.6 (C=N), 174.5 (C=O),
FTIR-ATR νmax/cm−1: 2975, 1751 (C=O), 1704 (C=O), 1650 (C=O),
HRMS-MALDI found: 471.2255, calculated for C24H31N4O6 [MH]+: 471.2244.

3.6. Synthesis of Cycloadduct 28

Isoindole precursor 20 (50 mg, 0.13 mmol), bis(2-pyridyl)-sym-1,2,4,5-tetrazine 24 (31 mg, 0.13 mmol), and DMAD (8.0 µL, 0.065 mmol) were dissolved in CHCl3 (1 mL) and stirred overnight at room temperature. The reaction mixture was purified by radial chromatography using CH2Cl2. Cycloadduct 28 was isolated as an off-white solid (52 mg, 80%).
m.p. 70–72 °C,
1H NMR (CDCl3), δ/ppm: 1.49 (s, 18H, t-Bu), 3.80 (s, 6H, OCH3), 6.12 (brs, 2H, N bridge), 7.07 (dd, 1H, J = 5.4, 3.1 Hz, Ar), 7.43 (dd, 1H, J = 5.4, 3.1 Hz, Ar), 10.61 (brs, 1H, NH),
FTIR-ATR νmax/cm−1: 2979, 1755 (C=O), 1722 (C=O),
13C NMR (CDCl3), δ/ppm: 28.1 (t-Bu), 52.4 (N bridge), 109.8 (C=C), 122.2, 126.2, 128.9, 139.1, 142.7 (C=O), 145.0 (C=O), 162.4 (C=O), 162.7 (C=N),
HRMS-MALDI found: 502.2197, calculated for C25H32N3O8 [MH]+: 502.2189.

3.7. Synthesis of Cycloadduct 29

Isoindole precursor 20 (50 mg, 0.13 mmol), bis(2-pyridyl)-sym-1,2,4,5-tetrazine 24 (31 mg, 0.13 mmol) and naphthoquinone monohydrate (23 mg, 0.13 mmol) were dissolved in CHCl3 (1 mL) and stirred overnight at room temperature. The reaction mixture was purified by radial chromatography using CH2Cl2. Cycloadduct 29 was isolated as a brown solid (52 mg, 77%).
m.p. 95–97 °C,
1H NMR (CDCl3), δ/ppm: 1.52 (s, 18H, t-Bu), 4.01 (s, 2H, exo-H), 5.92 (brs, 2H, N bridge), 6.87 (dd, 1H, J = 6.9, 3.2 Hz, Ar), 7.10 (dd, 1H, J = 6.9, 3.2 Hz, Ar), 7.52 (dd, 1H, J = 5.9, 3.3 Hz, Ar), 7.77 (dd, 1H, J = 5.9, 3.3 Hz, Ar), 10.58 (brs, 1H, NH),
13C NMR (CDCl3), δ/ppm: 28.2 (t-Bu), 49.6 (exo-H), 66.8 (N bridge), 80.2, 82.8, 121.8, 126.6, 127.4, 134.1, 134.6, 140.6, 150.2 (C=O), 152.3, (C=O), 162.6 (C=N), 194.4 (C=O),
FTIR-ATR νmax/cm−1: 2980, 1732 (C=O), 1677 (C=O),
HRMS-MALDI found: 518.2295, calculated for C29H32N3O6 [MH]+: 518.2291.

3.8. Synthesis of 37

Tetrafluoroazabenzonorbornadiene 36 (215 mg, 1.0 mmol) and N,N′-Di-Boc-1H-pyrazole-1-carboxamidine (310 mg, 1.0 mmol) were grinded in a ball mill for 2 h. The reaction mixture was purified by radial chromatography using CH2Cl2/hexane mixture 70:30. Cycloadduct 37 was isolated as a brown solid (158 mg, 35%).
m.p. 84–86 °C,
1H NMR (CDCl3), δ/ppm: 1.50 (s, 18H, t-Bu), 6.10 (brs, 2H, N bridge), 7.06 (brs, 2H, C=CH), 10.61 (brs, 1H, NH),
13C NMR (CDCl3), δ/ppm: 28.1 (t-Bu), 62.9 (N bridge), 65.4(N bridge), 80.5, 82.8, 129.2 (d, 4JCF = 17 Hz, C=C), 138.0 (t, 2JCF = 16.1 Hz, Ar), 139.6 (t, 2JCF = 16.1 Hz, Ar), 141.2 (m, Ar), 144.1 (m, Ar), 149.9 (C=O), 151.6 (C=O), 162.4 (C=N),
FTIR-ATR νmax/cm−1: 2981, 1760 (C=O), 1653 (C=O),
HRMS-MALDI found: 458.1716, calculated for C21H24F4N3O4 [MH]+: 458.1703.

3.9. Synthesis of Cycloadduct 39

Isoindole precursor 38 (15 mg, 0.033 mmol), bis(2-pyridyl)-sym-1,2,4,5-tetrazine 24 (8 mg, 0.033 mmol) and N-methylmaleimide (4 mg, 0.036 mmol) were dissolved in CHCl3 (0.5 mL) and stirred overnight at 60 °C. The reaction mixture was purified by radial chromatography using CH2Cl2. Cycloadduct 39 was isolated as a white solid (2 mg, 11%).
m.p. 108–110 °C,
1H NMR (CDCl3), δ/ppm: 1.49 (s, 9H, t-Bu), 1.52 (s, 9H, t-Bu), 2.56 (s, 3H, NCH3), 3.96 (dd, 2H, J = 3.8, 1.8 Hz, exo-H), 5.99 (brs, 2H, N bridge), 10.59 (brs, 1H, NH),
13C NMR (CDCl3), δ/ppm: 24.6 (NCH3), 28.0 (t-Bu), 47.1, 60.0 (N bridge), 65.2 (N bridge), 81.1, 83.5, (140.3, 146.6, 149.9), 152.5 (C=O), 162.1 (C=N), 173.2 (C=O),
FTIR-ATR νmax/cm−1: 2978, 1741 (C=O), 1709 (C=O), 1662 (C=O),
HRMS-MALDI found: 543.1874, calculated for C24H27F4N4O6 [MH]+: 543.1867.

3.10. Synthesis of 42

Under argon, dry toluene (11 mL) was added to Cbz-pyrrole 41 (200 mg, 1.0 mmol) and iodonium salt 22 (539 mg, 1.0 mmol). LiHDMS solution in toluene (1.0 mL, 1.0 mmol, 1 M) was added dropwise and the resulting mixture was stirred for 1 h at room temperature. The reaction was quenched with saturated NH4Cl solution (40 mL) and extracted with EtOAc (3x30 mL); combined extracts were dried with Na2SO4 and evaporated. The crude mixture was purified by radial chromatography using CH2Cl2 and gradually increasing polarity with MeOH. Product 42 was isolated as a viscous yellow oil (128 mg, 40%).
1H NMR (CDCl3), δ/ppm: 5.06 (s, 2H, CH2), 5.67 (brs, 2H, N bridge), 7.02 (d, 2H, J = 10.1 Hz, C=CH), 7.23–7.25 (m, 2H, Ar), 7.30–7.37 (m, 4H, Ar), 7.95 (dd, 1H, J = 7.9, 1.8 Hz, Ar), 8.03 (brs, 1H, Ar),
13C NMR (CDCl3), δ/ppm: 66.1 (N bridge), 67.0 (OCH2), 67.6 (N bridge), 122.2 (C=C), 127.9, 128.2, 128.3, 128.6, 128.7, 128.8, 128.9, 129.0, 135.8, 145.7, 155.0 (C=O),
FTIR-ATR νmax/cm−1: 2955, 1708 (C=O), 1517 (N-O), 1323 (N-O),
HRMS-MALDI found: 323.1038, calculated for C18H15N2O4 [MH]+: 323.1032.

3.11. Synthesis of 43

Azabenzonorbornadiene 42 (89 mg, 0.28 mmol) was dissolved in THF/H2O mixture (40 mL, 10% H2O) and heated to 60 °C. Aluminium amalgam was prepared by immersing aluminium foil (400 mg) in a solution of HgCl2 (500 mg) in water (50 mL) for 1 min, followed by washing in ethanol (50 mL) and diethyl ether (50 mL). Amalgam was added to the solution and the mixture was continuously heated for 1 h, filtered through Celite and washed with THF. The filtrate was evaporated and purified by radial chromatography using CH2Cl2 and gradually increasing polarity with MeOH. Product 43 was isolated as a viscous brown oil (53 mg, 66%).
1H NMR (CDCl3), δ/ppm: 3.56 (brs, 2H, NH2), 5.06 (s, 2H, CH2), 5.49 (d, 2H, J = 6.6 Hz, N bridge), 6.22 (dd, 1H, J = 7.6, 2.0 Hz, Ar), 6.69 (brs, 1H), 6.85–6.95 (m, 3H, C=C, Ar), 7.24–7.33 (m, 5H, Ar),
13C NMR (CDCl3), δ/ppm: 65.8 (N bridge), 67.2(N bridge), 67.3 (OCH2),112.7 (C=C), 121.4 (C=C), 127.7, 127.8, 128.1, 128.2, 128.5, 128.6, 130.1, 130.2, 136.1, 136.2, 155.2 (C=O),
FTIR-ATR νmax/cm−1: 3361 (N-H2), 1699 (C=O),
HRMS-MALDI found: 293.1292, calculated for C18H17N2O2 [MH]+: 293.1290.

3.12. Synthesis of Guanidine 44

Azabenzonorbornadiene 43 (167 mg, 0.572 mmol) and N,N′-Di-Boc-1H-pyrazole-1-carboxamidine (177 mg, 0.572 mmol) were dissolved in CHCl3 (5 mL) and stirred at room temperature for 2 days. The crude mixture was purified by radial chromatography using petroleum ether and gradually increasing polarity with CH2Cl2. Guanidine 44 was isolated as a viscous yellow oil (262 mg, 86%).
1H NMR (CDCl3), δ/ppm: 1.50 (s, 9H, t-Bu), 1.53 (s, 9H, t-Bu), 5.06 (d, 2H, J = 13.8 Hz, CH2), 5.56 (d, 2H, J = 16.1 Hz, N bridge), 6.92 (brs, 1H, C=CH), 6.97 (brs, 1H, C=CH), 7.05–7.21 (m, 3H, Ar), 7.26–7.34 (m, 5H, Ar), 10.29 (brs, 1H; NH), 11.64 (brs, 1H, NH),
13C NMR (CDCl3), δ/ppm: 28.1 (t-Bu), 28.2 (t-Bu), 65.9 (N bridge), 66.4 (N bridge), 67.2 (OCH2), 116.5 (C=C), 118.3, 120.9 (C=C), 127.8, 128.0, 128.5, 129.0, 133.9, 136.3, 142.8, 144.3, 149.2, 153.3 (C=O), 153.6 (C=O), 155.1 (C=N), 163.5 (C=O),
FTIR-ATR νmax/cm−1: 2978, 1713 (C=O),
HRMS-MALDI found: 535.2572, calculated for C29H35N4O6 [MH]+: 535.2557.

3.13. Synthesis of Cycloadduct 46

Isoindole precursor 44 (15 mg, 0.028 mmol), bis(2-pyridyl)-sym-1,2,4,5-tetrazine 24 (7.0 mg, 0.029 mmol) and N-methylmaleimide (3.0 mg, 0.027 mmol) were dissolved in CHCl3 (0.5 mL) and stirred overnight at room temperature. The reaction mixture was purified by radial chromatography using CH2Cl2. Cycloadduct 46 was isolated as a white solid (11 mg, 66%).
m.p. 100–102 °C,
1H NMR (CDCl3), δ/ppm: 1.49 (s, 9H, t-Bu), 1.53 (s, 9H, t-Bu), 2.36 (s, 3H, NCH3), 3.69 (brs, 2H, exo-H), 5.08 (brs, 2H, CH2), 5.52 (d, 1H, J = 4.4 Hz, N bridge), 5.56 (d, 1H, J = 4.4 Hz, N bridge), 7.21 (d, 1H, J = 8.0 Hz, Ar), 7.31–7.38 (m, 5H, Ar), 7.49 (brs, 2H, Ar), 10.29 (brs, 1H, NH), 11.58 (brs, 1H, NH),
13C NMR (CDCl3), δ/ppm: 28.1 (t-Bu), 28.2 (t-Bu), 45.6 (exo-H), 46.6 (exo-H), 62.4 (CH2), 62.8 (N bridge), 67.8 (N bridge), 79.8, 83.9, 116.2, 122.1, 128.1, 128.4, 128.5, 128.6, 135.7. 135.9, 136.8. 140.5, 153.3 (C=O), 153.6 (C=O), 154.6 (C=O), 163.4 (C=N), 174.1 (C=O), 174.2 (C=O),
FTIR-ATR νmax/cm−1: 2980, 1704 (C=O), 1634 (C=O),
HRMS-MALDI found: 620.2730, calculated for C32H38N5O8 [MH]+: 620.2720.

4. Conclusions

Novel isoindoles possessing guanidine substituents were synthesized and their cycloaddition reactivity was explored. These show higher reactivity than the corresponding pyrrole-1-carboxamidine and lead to the formation of polycyclic structures which incorporate guanidine functionality into the 7-azanobornene skeleton. New isoindoles were reactive towards dienophiles possessing electron-withdrawing groups (N-methylmaleimide, DMAD, naphthoquinone) and highly reactive arynes, whereas reactions with 7-azabenzonorbornadiene did not occur. In addition, 2H-isoindole-2-carboxamidine, its tetrafluoro counterpart and 4-guanidino isoindole showed similar reactivity. Experimentally observed reactivities are in good accord with theoretical predictions obtained at the M062X/6-311+G** level. Only small (about 0.5 kcal mol−1) differences in energy barriers for cycloaddition reactions of isoindoles were predicted.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/molecules27248954/s1, 1H and 13C NMR and IR spectra for the synthesized compounds.

Author Contributions

Conceptualization, D.M.; methodology, P.Š. and A.B.; investigation, P.Š. and A.B.; resources, D.M.; writing—original draft preparation, D.M.; writing—review and editing, D.M., P.Š. and A.B.; supervision, D.M.; project administration, D.M.; funding acquisition, D.M. All authors have read and agreed to the published version of the manuscript.

Funding

This research is funded by the Croatian Science Foundation, grant No. IP-2018-01-3298, Cycloaddition strategies towards polycyclic guanidines (CycloGu).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data are contained within the article and the Supplementary Materials.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Margetić, D. Physico-Chemical Properties of Organosuperbases. In Superbases for Organic Synthesis: Guanidines, Amidines, Phosphazenes and Related Organocatalysts; Ishikawa, T., Ed.; Wiley: Chichester, UK, 2009; Chapter 2; pp. 9–48. [Google Scholar] [CrossRef]
  2. Vazdar, K.; Margetić, D.; Kovačević, B.; Sundermeyer, J.; Leito, I.; Jahn, U. Design of novel uncharged organic superbases: Merging basicity and functionality. Acc. Chem. Res. 2021, 54, 3108–3123. [Google Scholar] [CrossRef] [PubMed]
  3. Saczewski, F.; Balewski, L. Biological activities of guanidine compounds. Expert Opin. Ther. Pat. 2009, 19, 1417–1448. [Google Scholar] [CrossRef] [PubMed]
  4. Nishikawa, T.; Urabe, D.; Isobe, M. An Efficient Total Synthesis of Optically Active Tetrodotoxin. Angew. Chem. Int. Ed. 2004, 43, 4782–4785. [Google Scholar] [CrossRef] [PubMed]
  5. Konrad, D.B.; Rühmann, K.-P.; Ando, H.; Hetzler, B.E.; Strassner, N.; Houk, K.N.; Matsuursa, B.S.; Trauner, D. A concise synthesis of tetrodotoxin. Science 2022, 377, 411–415. [Google Scholar] [CrossRef] [PubMed]
  6. Margetić, D. Cycloadditions of guanidines. In Cycloaddition Reactions: Advances in Research and Applications; Margetić, D., Ed.; Nova Science Publishers: New York, NY, USA, 2019; Chapter 7; pp. 243–280. [Google Scholar]
  7. Briš, A.; Murata, Y.; Hashikawa, Y.; Margetić, D. Utilization of sym-Tetrazines as Guanidine Cycloaddition Delivery Reagents. An Experimental and Computational Study. J. Mol. Struct. 2023, 1272, 134207. [Google Scholar] [CrossRef]
  8. Margetić, D.; Russell, R.A.; Warrener, R.N. Cycloadditions Reagents for Rigidly Attaching the 1,4-Dimethoxynaphthalene Chromophore to Scaffold Alkenes. Org. Lett. 2000, 2, 4003–4006. [Google Scholar] [CrossRef]
  9. Warrener, R.N.; Schultz, A.C.; Houghton, M.A.; Butler, D.N. Rigid molecular racks featuring the 1,10-phenanthroline ligand especially those co-functionalised with redox-active groups or other bidentate ligands. Tetrahedron 1997, 53, 3991–4012. [Google Scholar] [CrossRef]
  10. Malpass, J.R.; Sun, G.; Fawcett, J.; Warrener, R.N. Novel ‘windscreen wiper’ cavity structures formed by the cycloaddition of N-substituted isoindoles onto molrac bis-alkenes. Tetrahedron Lett. 1998, 39, 3083–3086. [Google Scholar] [CrossRef]
  11. Warrener, R.N.; Margetic, D.; Sun, G.; Russell, R.A. Position-Addressable Nano-Scaffolds. I. The Preparation of N,O-, N,C- and N,N-Bridged Sesquinorbornadiene Succinimides as Compact, Highly Functionalized Addressable Building Blocks. Aust. J. Chem. 2003, 56, 263–267. [Google Scholar] [CrossRef]
  12. Donohoe, T.J. Product class 14: 1H- and 2H-isoindoles. Sci. Synth. 2001, 10, 653–692. [Google Scholar] [CrossRef]
  13. Reddy, R.P.; Davies, H.M.L. Asymmetric Synthesis of Tropanes by Rhodium-Catalyzed [4+3] Cycloaddition. J. Am. Chem. Soc. 2007, 129, 10312–10313. [Google Scholar] [CrossRef]
  14. Antol, I.; Barešić, L.; Glasovac, Z.; Margetić, D. Computational Study of Electronic Influence of Guanidine Substitution on Diels-Alder Reactions of Heterocyclic Dienes. Croat. Chem. Acta 2019, 92, 279–286. [Google Scholar] [CrossRef]
  15. Parr, B.T.; Economou, C.; Herzon, S.B. A concise synthesis of (+)-batzelladine B from simple pyrrole-based starting materials. Nature 2015, 525, 507–510. [Google Scholar] [CrossRef] [Green Version]
  16. Economou, C.; Romaire, J.P.; Scott, T.Z.; Parr, B.T.; Herzon, S.B. A convergent approach to batzelladine alkaloids. Total syntheses of (+)-batzelladine E, (−)-dehydrobatzelladine C, and (+)-batzelladine K. Tetrahedron 2018, 74, 3188–3197. [Google Scholar] [CrossRef]
  17. Margetić, D. High Pressure Organic Synthesis; Verlag Walter de Gruyter: Berlin, Germany, 2019; ISBN 978-3-11-055602-5. [Google Scholar] [CrossRef]
  18. Štrbac, P.; Margetić, D. Complementarity of solution and solid state mechanochemical reaction conditions demonstrated by 1,2-debromination of tricyclic imides. Beilstein J. Org. Chem. 2022, 18, 746–753. [Google Scholar] [CrossRef]
  19. Butler, D.N.; Margetić, D.; O’Neill, P.J.C.; Warrener, R.N. Parity Reversal: A New Diels-Alder Strategy for the Synthesis of Sesquinorbornadienes, Including Those with Heterobridges and Those of Unusual Stereochemistry. Synlett 2000, 1, 98–100. [Google Scholar] [CrossRef]
  20. Juršić, B.S. AM1 semiempirical study of benzopyrroles as dienes for Diels-Alder reaction. Can. J. Chem. 1996, 74, 114–120. [Google Scholar] [CrossRef]
  21. Warrener, R.N.; Butler, D.N.; Margetić, D. Preparation of the First Isobenzofuran Containing Two Ring Nitrogens: A New Diels-Alder Diene for the Synthesis of Molecular Scaffolds Containing one or more End-Fused 3,6-di(2-pyridyl)pyridazine Ligands. Aust. J. Chem. 2003, 56, 811–817. [Google Scholar] [CrossRef]
  22. Warrener, R.N. Isolation of isobenzofuran, a stable but highly reactive molecule. J. Am. Chem. Soc. 1971, 93, 2346–2348. [Google Scholar] [CrossRef]
  23. Priestley, G.M.; Warrener, R.N. A new route to isoindole (benzo[c]indole) and its derivatives. Tetrahedron Lett. 1972, 42, 4295–4298. [Google Scholar] [CrossRef]
  24. Kreher, R.P.; Use, G. Untersuchungen zur Chemie von Isoindolen und Isoindoleninen, XXIX. Reaktionen des 2H-Isoindols mit Maleinimiden: Ein einfaches Herstellungsverfahren für 7-Azabicyclo[2.2.1]heptene. Eur. J. Inorg. Chem. 1988, 121, 927–934. [Google Scholar] [CrossRef]
  25. Ohwada, T.; Ishikawa, S.; Mine, Y.; Inami, K.; Yanagimoto, T.; Karaki, F.; Kabasawa, Y.; Otani, Y.; Mochizuki, M. 7-Azabicyclo [2.2.1]heptane as a structural motif to block mutagenicity of nitrosamines. Bioorganic Med. Chem. 2011, 19, 2726–2741. [Google Scholar] [CrossRef] [PubMed]
  26. Kreher, R.P.; Seubert, J.; Kohl, N. Investigations on the Chemistry of Isoindoles and Isoindolenines. Part 26. Simple Methods for the Preparation of 2H-Isoindole. Chem. Ztg. 1987, 111, 349–356. [Google Scholar] [CrossRef]
  27. Kreher, R.P.; Kohl, N. A Rational Synthetic Method for 2H-Isoindoles. Angew. Chem. Int. Ed. 1984, 23, 517–518. [Google Scholar] [CrossRef]
  28. Davies, J.W.; Malpass, J.R.; Moss, R.E. Barriers to inversion at nitrogen in bicyclic amines and hydrazines. Tetrahedron Lett. 1985, 26, 4533–4536. [Google Scholar] [CrossRef]
  29. Đud, M.; Glasovac, Z.; Margetić, D. The utilization of ball-milling in synthesis of aryl guanidines through guanidinylation and N-Boc-deprotection sequence. Tetrahedron 2019, 75, 109–115. [Google Scholar] [CrossRef]
  30. Hewson, M.J.C.; Schmutzler, R. Phosphorus-fluorine chemistry part XLIII. Pyrrole-substituted fluorophosphoranes. Phosphorus Sulfur Relat. Elem. 1980, 8, 9–26. [Google Scholar] [CrossRef]
  31. Davies, J.W.; Durrant, M.L.; Walker, M.P.; Belkacemi, D.; Malpass, J.R. Preparation and spectroscopic studies of the 1,4-dihydro- 1,4-iminonaphthalene (7-azabenzonorbornadiene) ring system. Tetrahedron 1992, 48, 861–884. [Google Scholar] [CrossRef]
  32. Wojciechowski, K. Synthesis of 4-nitro-2H-isoindole derivatives. Liebigs Ann. Chem. 1991, 1991, 831–832. [Google Scholar] [CrossRef]
  33. Murashima, T.; Tamai, R.; Nishi, K.; Nomura, K.; Fujita, K.; Uno, H.; Ono, N. Synthesis and X-ray structure of stable 2H-isoindoles. J. Chem. Soc. Perkin Trans. 2000, 6, 995–998. [Google Scholar] [CrossRef]
  34. Use, G.; Kreher, R. Studies of the chemistry of isoindoles and isoindolenines 18. 2-tert-butyl-5-nitro-2H-isoindole. Preparation and reactions. Chem. Ztg. 1982, 106, 143–144. [Google Scholar] [CrossRef]
  35. Lin, C.; Zhen, L.; Cheng, Y.; Du, H.-J.; Zhao, H.; Wen, X.; Kong, L.-Y.; Xu, Q.-L.; Sun, H. Visible-Light Induced Isoindoles Formation To Trigger Intermolecular Diels-Alder Reactions in the Presence of Air. Org. Lett. 2015, 17, 2684–2687. [Google Scholar] [CrossRef]
Scheme 1. Cycloaddition reactions of pyrrole 1.
Scheme 1. Cycloaddition reactions of pyrrole 1.
Molecules 27 08954 sch001
Scheme 2. Synthesis of isoindole precursors 20 and 23.
Scheme 2. Synthesis of isoindole precursors 20 and 23.
Molecules 27 08954 sch002
Scheme 3. Generation of isoindole 2 using tetrazine method.
Scheme 3. Generation of isoindole 2 using tetrazine method.
Molecules 27 08954 sch003
Figure 2. Section of 1H NMR spectra (in CDCl3) with assigned peaks of isoindole 2.
Figure 2. Section of 1H NMR spectra (in CDCl3) with assigned peaks of isoindole 2.
Molecules 27 08954 g002
Scheme 4. Cycloaddition reactions of isoindole 2.
Scheme 4. Cycloaddition reactions of isoindole 2.
Molecules 27 08954 sch004
Scheme 5. Synthesis and reactions of tetrafluoro isoindole 38.
Scheme 5. Synthesis and reactions of tetrafluoro isoindole 38.
Molecules 27 08954 sch005
Scheme 6. Synthesis and reactions of isoindole 45.
Scheme 6. Synthesis and reactions of isoindole 45.
Molecules 27 08954 sch006
Figure 3. Activation-free energies of the Diels–Alder reaction of pyrrole and isoindoles with acetylene and optimized transition state structure of the reaction with isoindole-2-carboxamidine as calculated by M062X/6-311+G**.
Figure 3. Activation-free energies of the Diels–Alder reaction of pyrrole and isoindoles with acetylene and optimized transition state structure of the reaction with isoindole-2-carboxamidine as calculated by M062X/6-311+G**.
Molecules 27 08954 g003
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Štrbac, P.; Briš, A.; Margetić, D. N,N′-Di-Boc-2H-Isoindole-2-carboxamidine—First Guanidine-Substituted Isoindole. Molecules 2022, 27, 8954. https://doi.org/10.3390/molecules27248954

AMA Style

Štrbac P, Briš A, Margetić D. N,N′-Di-Boc-2H-Isoindole-2-carboxamidine—First Guanidine-Substituted Isoindole. Molecules. 2022; 27(24):8954. https://doi.org/10.3390/molecules27248954

Chicago/Turabian Style

Štrbac, Petar, Anamarija Briš, and Davor Margetić. 2022. "N,N′-Di-Boc-2H-Isoindole-2-carboxamidine—First Guanidine-Substituted Isoindole" Molecules 27, no. 24: 8954. https://doi.org/10.3390/molecules27248954

Article Metrics

Back to TopTop