Next Article in Journal
Production of the Anti-Inflammatory Compound 6-O-Palmitoyl-3-O-β-D-glucopyranosylcampesterol by Callus Cultures of Lopezia racemosa Cav. (Onagraceae)
Next Article in Special Issue
Cycloaddition of 1,3-Butadiynes: Efficient Synthesis of Carbo- and Heterocycles
Previous Article in Journal
Potential Anti-Inflammatory Effects of the Hydrophilic Fraction of Pomegranate (Punica granatum L.) Seed Oil on Breast Cancer Cell Lines
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Three-Dimensional Heterocycles: New Uracil-Based Structures Obtained by Nucleophilic Substitution at the sp2 Carbon of Bromoisoxazoline

Dipartimento di Chimica, Univesità degli Studi di Pavia, Viale Taramelli 12, 27100 Pavia, Italy
*
Author to whom correspondence should be addressed.
Molecules 2014, 19(6), 8661-8678; https://doi.org/10.3390/molecules19068661
Submission received: 7 May 2014 / Revised: 4 June 2014 / Accepted: 13 June 2014 / Published: 24 June 2014
(This article belongs to the Special Issue Cycloaddition Chemistry)

Abstract

:
The regioisomeric cycloadducts of bromonitrile oxide and N-benzoyl-2,3-oxaza-norborn-5-ene were easily prepared and elaborated into a novel class of uracil-based scaffolds. The key-synthetic step is the nucleophilic substitution at the sp2 carbon atom of the bromoisoxazoline three-dimensional heterocycles. The protocol to perform the nucleophilic substitution of uracil anions was optimized and adapted to the steric requirements of the substrates. A library of pyrimidine derivatives was prepared in very good yields and the products were fully characterized. They are proposed as nucleoside analogues and as synthons for β-turn motifs within PNA structures.

Graphical Abstract

1. Introduction

A novel approach to useful precursors for the synthesis of isoxazoline-carbocyclic nucleosides was recently detailed starting from the readily available N-benzoyl-2,3-oxazanorborn-5-ene (1) and introducing a versatile functional group via a condensed bromoisoxazoline ring. These structures can be prepared through the efficient and exo-selective 1,3-dipolar cycloaddition reaction of bromonitrile oxide 3 (in-situ generated from the corresponding bromoxime 2) to the 2,3-oxazanorbornene (1) affording the regioisomeric cycloadducts 4a,b in good yields (Scheme 1) [1].
Scheme 1. Cycloaddition reaction of bromonitrile oxide (3) to the N-benzoyl-2,3-oxazanorborn-5-ene (1).
Scheme 1. Cycloaddition reaction of bromonitrile oxide (3) to the N-benzoyl-2,3-oxazanorborn-5-ene (1).
Molecules 19 08661 g003
In previous studies the regioisomeric cycloadducts 4 were used as starting materials for the synthesis of aminols containing the β-hydroxynitrile functionalities derived from the isoxazoline ring opening [2]. They served as intermediates in the linear construction of uracil-substituted nucleosidic structures [3,4].
Nitrosocarbonyl (R-CONO) chemistry represents the key strategy of the reported syntheses because of the high dienophilic power of these fleeting intermediates and the synthetic potential of their hetero- Diels-Alder (HDA) cycloadducts of type 1. Traditionally, nitrosocarbonyls are generated through periodate oxidation of hydroxamic acids and instantly trapped by dienes as reported by Kirby [5]. Recently, we developed two alternative entries to nitrosocarbonyls through the mild oxidation of nitrile oxides with N-methylmorpholine-N-oxide (NMO) [6,7,8,9,10].
In the planned efforts to expand upon the versatility of 1[1,2,3,4,5,6,7,8,9,10], we were interested in varying the synthetic applications of the regioisomeric three-dimensional cycloadducts of the bromonitrile oxide to the oxazanorbornene 1. The presence of the bromine atom on the nitrile oxide moiety suggests the exploration of the nucleophilic substitution (SN) reaction on a sp2 carbon atom [11]. We wish to report here a study regarding the reactivity of the regioisomeric cycloadducts 4a,b in the presence of the in-situ generated anions of uracil bases (Scheme 2). The methodologies applied are discussed in the light of the reaction mechanism
Scheme 2. Synthetic strategy towards uracil derivatives.
Scheme 2. Synthetic strategy towards uracil derivatives.
Molecules 19 08661 g004
The study also aimed to ascertain the scope and eventual limitations of this type of chemistry upon variation of the experimental conditions and the feasibility of the planned synthetic elaborations towards the preparation of novel β-turn inducers containing nucleobases for their insertion in Peptide Nucleic Acids (PNA) as well as the synthesis of new nucleoside analogues [12,13].
Scheme 3. Synthesis of uracil compounds (5a,b).
Scheme 3. Synthesis of uracil compounds (5a,b).
Molecules 19 08661 g005

2. Results and Discussion

The bromonitrile oxide 3 is generated in situ in an AcOEt/NaHCO3 solution from the corresponding bromoxime 2 [14], which is easily prepared by treating glyoxylic acid with hydroxylamine hydrochloride in the presence of bromine at 5 °C in dichloromethane [15]. Using N-benzoyl-2,3-oxazanorborn-5-ene (1) as dipolarophile, the 1,3-dipolar cycloaddition reaction smoothly occurs affording the regioisomeric compounds 4a,b in 45% and 44% yields, respectively, which are easily isolated upon chromatographic separation (Scheme 3).
The structures of cycloadducts 4a,b were assigned based on their analytical and spectroscopic data as well as upon comparison with previously prepared authentic samples [1]. The known exceptionally high reactivity and exclusive exo-selectivity of the cycloaddition reaction of 1 with the bromonitrile oxide 3 is akin to that of the related parent carbocyclic norbornene, which adds all types of reactants only on the exo face [16]. The origin of the norbornene selectivity is usually attributed to the relief of strain [17], geometric deformation of the double bond (pyramidalization due to torsional [18] and hyperconjugative effects [19]) as well as favourable staggering effects in the exo attack [20]. This fact pushed us to explore the insertion of a heterobase of choice anti-related to the oxaza-moiety of the cycloadducts 4a,b through an sp2 SN reaction as pivotal synthetic step of our strategy towards nucleoside analogues and β-turn motifs.
The sp2 SN reaction conditions and the synthetic method were first optimized using as benchmark reaction that between the two regioisomeric cycloadducts 4a,b and to uracil as heterobase (Scheme 4). In the search for the best reaction conditions, we took advantage of previously reported results on an sp2 SN reaction conducted on a bromoisoxazoline [21]. Upon addition of the regioisomeric cycloadducts 4a,b to an anhydrous DMF suspension of the uracil anion, generated by treatment with NaH, under stirring at 95 °C, poor results were obtained in terms of final products 5a,b (10%–12% yields, respectively) after 3 days of reaction. Moreover, great difficulties were encountered in the work-up procedures (emulsions) for the isolation of the products.
The reference authors themselves report the various efforts done to find the proper experimental conditions to perform the sp2 SN reaction. However they did not detail the experimental procedure but indicated DMF as the best solvent to run the reactions along with the temperature and reaction time employed [21]. The choice of different solvents such as anhydrous THF or MeCN was soon abandoned since just the starting materials were recovered after the work-up and no reaction occurred. Then we tried to change base, temperature and reaction time. The sp2 SN reaction failed when NaOH was used as well as running the reactions both at room temperature or upon heating, even after several days under stirring. Decomposition of the starting materials was also observed under microwave heating (MW) (DMF/NaH at 160 °C for 30 min).
We attributed the low yields of the adducts 5a,b to the low concentration of the nucleophile. In order to have a high conversion of the uracil base into the corresponding anion, we used NaH 95% instead of the standard 70%–80% and performed the nucleophile in-situ generation under vacuum in a Schlenk tube. In an anhydrous Schlenk tube an excess of NaH 95% (2.1 equivalents) is suspended in anhydrous DMF (20 mg/mL) and vacuum is applied to get rid of the air. The uracil base is added portionwise and the suspension is left under stirring at room temperature for 30 min.
Scheme 4. Synthesis of uracil derivatives (6a,b) and thymine compounds (7a,b).
Scheme 4. Synthesis of uracil derivatives (6a,b) and thymine compounds (7a,b).
Molecules 19 08661 g006
Finally, 0.8 equivalents of the regioisomeric cycloadducts 4a,b were added and the mixture was heated at 95 °C under stirring for 24 h. The reaction was quenched by adjusting the pH to 7 by means of addition of aqueous 3:1 NH4Cl/NaCl and the water phase was extracted with dichloromethane (DCM). The adducts 5a,b were purified by column chromatography under medium pressure (MPLC) and the yields were 64% and 50% for 5a and 5b, respectively (yields determined by HPLC-RP C-18 analysis on crude extracted solutions). The isolated yields were 44% and 30% for 5a and 5b of 98.5% purity after recrystallization, respectively.
The assigned structures of adducts 5a,b rely upon the corresponding analytical and spectroscopic data. In particular, the vinyl protons of the uracil moiety in 5a are found at δ 5.79 (d, J = 8 Hz, =CH-C=O) and at δ 7.83 (d, J = 8 Hz, N-CH=) while the singlet corresponding to the NH group is found at δ 11.70. Similarly, in compound 5b, the vinyl protons of the uracil moiety are found at δ 5.79 (d, J = 8 Hz, =CH-C=O) and at δ 7.85 (d, J = 8 Hz, N-CH=) while the NH group singlet is found at δ 11.72. All the other signals of the three-dimensional isoxazoline-norbornane moieties are in the expected range and clearly support the reported structures (Scheme 3). A definitive confirmation of the structural assignments given above came from the X-ray structure of compound 5a whose ORTEP view is shown in Figure 1. In the oxazanorbornane ring of the compound 5a the sum of the angles at N8 345.0(2)° is consistent with the sp3 hybridization; the 2pz lone pair takes no part in π–bonding with the C=O group, indeed the bond distance N8-C18, 1.376(3) Å is longer than the shortened partial double bond (1.352(5) Å). The height of the pyramid with the nitrogen atom at the apex and the three atoms connected to it at the base is 0.230(3) Å. The atoms N8 C18 O18 C19 are almost coplanar: the maximum deviation from least–squares plane is 0.028(2) Å; the angle between the carbonyl group and the moiety C1 O9 N8 C7 is 36.1(2)°. Bond lengths in oxazanorbornane ring C-C (range 1.523(4)–1.534.1(3) Å) and N-C 1.477(3) Å are comparable with the values of the literature [22]. The six-membered ring of the oxazanorbornane is puckered with deviations from the least-squares plane in the range −0.569(3) + 0.305(2) Å. These ring exhibit a boat conformation with the parameters [23] Q = 0.985(2), φ = 60.4(1)°, θ = 90.3(1)°, (ideal conformation: φ = 60.0°, θ = 90°, respectively).
Figure 1. ORTEP plot of compound 5a with atom labeling (ellipsoid at 20% probability).
Figure 1. ORTEP plot of compound 5a with atom labeling (ellipsoid at 20% probability).
Molecules 19 08661 g001
In the isoxazoline rings of compound 5a the angle at N4 109.1(2)° is consistent with the sp2 hybridization of N4, whereby the 2pz lone pair takes part in the π–bonding within a part of the heterocyclic ring; indeed the bond distance N4-C5 1.270(3) Å shows a double–bond character. The internal angles at the nitrogen atoms O3-N4-C5 109.1(2)° in 5a differ by 3.9°, justified by the different type of nitrogen. The deviation of atoms from the least–squares plane of isoxazoline rings are in the ranges −0.013(2) + 0.012(2) Å with a T conformations; the puckering parameters are Q = 0.020(3), φ = 132.2(2)° in 5a (ideal value φ = 126°).
The pyrimidine ring in compound 5a attached to the isoxazoline is nearly planar with a maximum deviation of −0.016(3) Å. This ring is tilted (with respect to the isoxazoline ring) by 5.3(1)° implying only a small conjugation. The orientation of the pyrimidine ring, due to the steric crowding the attached atoms is confirmed by the torsion angles N4-C5-N11-C12 −7.3(3)°, 12.2(4)°.
In order to ascertain the eventual limitations of this chemistry we extended the functionalization to the entire family of the halogenated uracils and thymine. Scheme 4 reports the reactions and the structures of the halogenated-uracil adducts 6a,b (AD) and the thymine adducts 7a,b obtained by applying the above reported sp2 SN reaction protocol. The positive results indicate that the methodology set-up for the uracil base can be easily transported to the other pyrimidine-type heterobases without any variation or adjustment when other derivatives are employed. In fact, as we can see from the data of Table 1, the overall yields are good, with small variations when moving from one substituent to another.
Table 1. Yields, physical data, relevant IR bands and diagnostic 1H-NMR signals of compounds 6a, b (AD) and 7a,b.
Table 1. Yields, physical data, relevant IR bands and diagnostic 1H-NMR signals of compounds 6a, b (AD) and 7a,b.
EntryAdduct 6aYield (%)m.p. (°C) #IR (cm−1) νNHνC=O1H-NMR N-CH=CX(δ, DMSO- d6) J (Hz)
1A46193–198339417338.24 (d)7
2B53180–189336017348.23 (s)
3C5696–100320017348.27 (s)
4D57110–115339416328.23 (s)
6b
5A6495–100315717238.24 (d)7
6B7698–105317817328.24 (s)
7C49104–106312117328.29 (s)
8D57162–164316017208.24 (s)
97a45215–217317617157.69 (d)1.81 (Me)
107b55220–222314917107.69 (d)1.82 (Me)
# From diisopropyl ether/ethanol.
The structures of the adducts 6a,b (AD) and 7a,b rely upon the corresponding analytical and spectroscopic data. Table 1 also reports the physical data (mps) and the frequencies of the IR bands corresponding to the uracil moieties inserted upon sp2 SN reactions as well as the diagnostic 1H-NMR signals confirming the insertion of the uracil rings onto the tricyclic structures.
Belonging the heterobases to the same family of the pyrimidine heterocycles, the changes in the NMR signals relative to the isoxazoline-norbornane moiety are negligible while the presence of the attached uracil units is clearly indicated by the vinylic protons as well as the NH groups found in the δ 12.05–12.25 range. A definitive confirmation of the structural assignments given above came from the X-ray structure of compound 7a whose ORTEP view is shown in Figure 2.
Figure 2. ORTEP plot of compound 7a with atom labeling (ellipsoid at 20% probability).
Figure 2. ORTEP plot of compound 7a with atom labeling (ellipsoid at 20% probability).
Molecules 19 08661 g002
In the oxazanorbornane ring of the compound 7a the sum of the angles at N8 339.4(3)° is consistent with the sp3 hybridization; the 2pz lone pair takes no part in π–bonding with the C=O group, indeed the bond distance N8-C18 1.377(5) Å is longer than the shortened partial double bond (1.352(5) Å). The height of the pyramid with the nitrogen atom at the apex and the three atoms connected to it at the base is 0.249(3) Å. The atoms N8 C18 O18 C19 are almost coplanar: the maximum deviation from least–squares plane is 0.039(4) Å; the angle between the carbonyl group and the moiety C1 O9 N8 C7 is 40.9(2)° in compounds 7a. Bond lengths in oxazanorbornane rings C-C (range 1.507(5)–1.540(5) Å) and N—C 1.487(4) Å in compound 7a, are comparable with the values of the literature [22]. The six–membered ring of the oxazanorbornane is puckered with deviations from the least–squares plane in the range −0.306(5) +0.575(5) Å. This ring exhibits a boat conformation with the parameters [23] Q = 0.982(3)°, φ = 60.8(3)° and θ = 90.0(3)° (ideal conformation: φ = 60.0°, θ = 90°, respectively).
In the isoxazoline rings of the compounds 7a the angles at N4 109.9(3)° is consistent with the sp2 hybridization of N4, whereby the 2pz lone pair takes part in the π–bonding within a part of the heterocyclic ring; indeed the bond distances N4-C5 1.272(4) Å shows a double–bond character. The internal angles at the nitrogen atoms O3-N4-C5 109.9(3)° and C7-N8-O9 105.3(3)° in 7a differ by 4.6°, justified by the different type of nitrogen. The deviation of atoms from the least–squares plane of isoxazoline ring is in the ranges −0.032(3) + 0.031(3) Å, with a T conformations; the puckering parameters are Q = 0.055(4), φ = 126.9(3)° (ideal value φ = 126°).
The pyrimidine ring in compound 7a attached to the isoxazoline is nearly planar with a maximum deviations of −0.061(4) Å; this rings is tilted (with respect to the isoxazoline rings) by 22.1(2)° respectively implying only a small conjugation. The orientation of the pyrimidine ring, due to the steric crowding of the attached atoms is confirmed by the torsion angles N4-C5-N11-C16 175.8(2)°, −165.0(4)°.
Finally, we have also scaled-up the protocol (from 250 mg to 2.5 g) performing the reactions with a larger amount of starting materials in order to have the required quantities of all the reported compounds to carry on with the synthetic elaborations towards the final targets. Yields did not change significantly (±2%) remaining in the reported range and the work-up procedures worked nicely, allowing for the easy isolation of the desired compounds.
The synthetic pathway shown in the Scheme 3 and Scheme 4 represent the pivotal step as well as the most delicate and somewhat difficult in the approach to our final goals. In fact, the first attempts to perform the sp2 SN reaction failed and the method had to be properly set-up to be adapted to our tricyclic substrates.
The SN reaction on a sp2 carbon atom is a transformation of relevance to organic synthesis, industrial applications as well as biochemistry [24]. Recent computational studies shed some light on the sp2 SN reaction in comparison with the mechanism followed by the nucleophilic aromatic substitution, SNAr, and the nucleophilic aliphatic substitution, SN2 [25].
Contrary to the aliphatic version of the SN reaction, the sp2 SN reaction on vinyl (V) carbons can follow different pathways. A common one is the Addition-Elimination route (A-ER); if electron-withdrawing groups are located on the α-carbon atom, a stable anion represents the intermediate of the reaction and the successive elimination affords the substitution product.
The presence of a nitrogen atom, belonging to the isoxazoline moiety, in the α position with respect to the =C-Br sp2 carbon centre can suggest the A-ER as the operating mechanism. The nitrogen atom helps the addition step allowing to stabilize the negative charge generated in the intermediate that however suffers somewhat of the increased steric requirements determined by the hybridation change of the C-Br carbon atom (Scheme 5) [26].
We believe that stereochemical requirements and nucleophile concentration have a consequence in the experimental conditions we have previously detailed. For this reason the generation under vacuum of the anion of the uracil bases can be considered a success from the experimental point of view. Temperature is also important; lower temperatures do not allowed for obtaining the products in good yields while higher temperature resulted in the decomposition of the reagents.
Scheme 5. Nucleophile addition orientation.
Scheme 5. Nucleophile addition orientation.
Molecules 19 08661 g007
As final remark, we wish to point out that we have detailed the methodology to promote a high yield process to substitute the bromine in regioisomeric bromoisoxazoline derivatives to insert pyrimidine heterobases on a three-dimensional heterocyclic system. The protocol has to be adapted and properly tuned for the purine-type heterobases, which suffer from the known possibility to give regioisomeric adducts at the N7 and N9 nitrogen atoms. Preliminary results indicate that mixtures of compounds are obtained although in lower yields and for this reason some changes in the method had to be done to make the protocol of general application.

3. Experimental

3.1. General Information

All melting points are uncorrected. Elemental analyses were done on a C. Erba 1106 elemental analyzer available in our Department. IR spectra (nujol mulls) were recorded on an FT-IR Perkin-Elmer RX-1. 1H- and 13C-NMR spectra were recorded on a Bruker AVANCE 300 in DMSO-d6. Chemical shifts are expressed in ppm (δ) from internal tetramethylsilane. Column chromatography: silica gel 60 (0.063–0.200 mm; Merck, Darmstadt, Germany); eluant was a cyclohexane/ethyl acetate gradient from 9:1 to 5:5. For MPLC a Biotage FMP apparatus equipped with KP-SIL columns, eluant cyclohexane/ethyl acetate from 9:1 to 7:3 was used. TLC was performed on Merck silica gel F254 plates. The identification of samples from different experiments was secured by mixed mps and superimposable IR spectra. Glyoxylic acid (Sigma-Aldrich, St. Louis, MO, USA) was used to prepare the bromoxime 2 according to the reported method [15]. N-Benzoyl-2-oxa-3-azanorborn-5-ene (1) was prepared according to the reported method [1,2,3,4,5,6,7,8,9,10]. Anhydrous DMF was purchased from Sigma-Aldrich. Other chemicals and solvents were purchased from Sigma-Aldrich and Carlo Erba Reagenti (Rodano - MI, Italy) and used without any further purification.

3.2. Cycloaddition of Bromonitrile Oxide 3 to N-Benzoyl-2,3-oxazanorborn-5-ene (1)

To a stirred solution of N-benzoyl-2,3-oxazanorborn-5-ene (1, 5.28 g, 26.24 mmol) in EtOAc (200 mL) NaHCO3 (3.93 g, 47.2 mmol) was added. To this suspension bromoxime 2 (7.98 g, 39.4 mmol) dissolved in EtOAc (50 mL) were added dropwise. The reaction proceeded under stirring for 48 h at room temperature. Upon filtration and evaporation of the solvent, an oily residue is obtained and submitted to chromatographic separation, allowing for the isolation of the regioisomeric cycloadducts 4a,b. The compounds were found to be identical to previously prepared authentic samples [1].

3.3. General Procedure for the Synthesis of Compounds 5a,b, 6a,b (AD) and 7a,b

In a dry Schlenk tube an excess of 95% NaH (2.1 equivalents) is suspended in anhydrous DMF (20 mg/mL) and vacuum is applied to get rid of the air inside the suspension. The dry uracil bases are added portionwise and vacuum is applied after every addition keeping watch on the gas evolution. The suspension is left under stirring at room temperature for 30 min to complete the nucleophile generation. Finally, 0.8 equivalents of the regioisomeric cycloadducts 4a,b were added in one portion and again vacuum applied. The mixtures were heated at 95 °C under vigorous stirring for 24 h. After this period of time, the reactions were quenched by pouring the solution in the minimum amount of ice (50 g/0.8 equivs. of 4a,b), adjusting the pH to 7 by means of solid NH4Cl/NaCl 3:1 and the water phase extracted with dichloromethane (DCM, 3 × 50 mL). The desired adducts 5a,b, 6a,b (AD) and 7a,b were purified by column chromatography under medium pressure (MPLC).
1-(5-Benzoyl-4,5,7,7a-tetrahydro-3aH-4,7-methanoisoxazolo[4,5-d][1,2]oxazin-3-yl)pyrimidine-2,4(1H,3H)-dione (5a): 0.35 g (64%), m.p. 225 °C (dec.), colorless crystals from diisopropyl ether/ethanol. IR: νNH 3248 cm−1, νC=O 1731 cm−1, νC=N 1694 cm−1. 1H-NMR (δ): 3.47 (m, 2H, CH2), 4.55 (d, 1H J = 8 Hz, H4isox), 5.06 (s, 1H, CH-N), 5.13 (d, 1H, J = 8 Hz, H5isox), 5.31 (s, 1H, HC-O), 5.79 (d, 1H, J = 8 Hz, =CH-C=O), 7.44 (m, 2H, arom.) 7.57 (m, 1H, arom.), 7.69 (m, 2H, arom.). 7.83 (d, 1H, J = 8 Hz, N-CH=), 11.70 (s, 1H, NH). 13C-NMR (δ): 30.7, 32.8, 53.8, 79.6, 84.6, 103.9, 128.3, 128.6, 131.9, 132.7, 141.6, 148.8, 153.2, 163.1, 169.9. Elemental analysis: cald. for C17H14N4O5 (MW = 354.32) C 57.63, H 3.98, N 15.81. Found C C 57.60, H 3.97, N 15.80.
1-(6-Benzoyl-4,6,7,7a-tetrahydro-3aH-4,7-methanoisoxazolo[5,4-d][1,2]oxazin-3-yl)pyrimidine-2,4(1H,3H)-dione (5b): 0.27 g (50%), m.p. > 240 °C (dec.), colorless crystals from diisopropyl ether/ethanol. IR: νNH 3200 cm−1, νC=O 1716 cm−1, νC=N 1646 cm−1. 1H-NMR (δ): 2.04 (m, 2H, CH2), 4.39 (d, 1H J = 8 Hz, H4isox), 5.00 (s, 1H, CH-N), 5.28 (m, 1H + 1H, H5isox and HC-O), 5.79 (d, 1H, J 8 Hz, =CH-C=O), 7.49 (m, 2H, arom.), 7.57 (m, 1H, arom.), 7.69 (m, 2H, arom.), 7.85 (d, 1H, J = 8 Hz, N-CH=), 11.72 (s, 1H, NH). 13C-NMR (δ): 32.6, 55.0, 60.9, 61.2, 80.1, 83.9, 103.9, 121.4, 128.3, 128.5, 131.8, 133.0, 141.2, 141.5, 148.8, 152.9, 163.0. Elemental analysis: cald. for C17H14N4O5 (MW = 354.32) C 57.63, H 3.98, N 15.81. Found C C 57.64, H 4.00, N 15.81.
1-(5-Benzoyl-4,5,7,7a-tetrahydro-3aH-4,7-methanoisoxazolo[4,5-d][1,2]oxazin-3-yl)-5-fluoropyrimidine-2,4(1H,3H)-dione (6aA): 0.27 g (46%), m.p. 193–198 °C, colorless crystals from diisopropyl ether/ethanol. IR: νNH 3394 cm−1, νC=O 1733 cm−1, νC=N 1654 cm−1. 1H-NMR (δ): 1.97 (m, 2H, CH2), 4.57 (d, 1H J = 8 Hz, H4isox), 5.07 (s, 1H, CH-N), 5.15 (d, 1H, J = 8 Hz, H5isox), 5.34 (s, 1H, HC-O), 7.47 (m, 2H, arom.), 7.57 (m, 1H, arom.), 7.69 (m, 2H, arom.), 8.24 (d, 1H, J = 7 Hz, N-CH=), 12.22 (s, 1H, NH). 13C-NMR (δ): 32.7, 39.8, 53.6, 79.6, 84.7, 90.0, 127.4, 128.2, 128.3, 128.7, 131.2, 131.8, 132.8, 147.6, 152.9, 157.4, 170.0. Elemental analysis: cald. for C17H13FN4O5 (MW = 372.31) C 54.84, H 3.52, N 15.05. Found C 54.85, H 3.50, N 15.06.
1-(6-Benzoyl-4,6,7,7a-tetrahydro-3aH-4,7-methanoisoxazolo[5,4-d][1,2]oxazin-3-yl)-5-fluoropyrimidine-2,4(1H,3H)-dione (6bA): 0.37 g (64%), m.p. 95–100 °C, colorless crystals from diisopropyl ether/ethanol. IR: νNH 3157 cm−1, νC=O 1723 cm−1, νC=N 1648 cm−1. 1H-NMR (δ): 2.04 (s, 2H, CH2), 4.39 (d, 1H J = 8 Hz, H4isox), 5.00 (s, 1H, CH-N), 5.28 (m, 1H + 1H, H5isox and HC-O), 7.47 (m, 2H, arom.), 7.56 (m, 1H, arom.), 7.71 (m, 2H, arom.), 8.24 (d, 1H, J = 7 Hz, N-CH=), 12.25 (s, 1H, NH). 13C-NMR (δ): 32.6, 54.8, 56.6, 80.0, 84.0, 125.8, 127.0, 127.4, 128.2, 128.3, 128.5, 131.2, 131.8, 133.0, 147.6, 152.5, 168.9. Elemental analysis: cald. for C17H13FN4O5 (MW = 372.31) C 54.84, H 3.52, N 15.05. Found C 54.83, H 3.51, N 15.02.
1-(5-Benzoyl-4,5,7,7a-tetrahydro-3aH-4,7-methanoisoxazolo[4,5-d][1,2]oxazin-3-yl)-5-chloropyrimidine-2,4(1H,3H)-dione (6aB): 0.32 g (53%), m.p. 189–196 °C, colorless crystals from diisopropyl ether/ethanol. IR: νNH 3360 cm−1, νC=O 1734 cm−1, νC=N 1703 cm−1. 1H-NMR (δ): 1.96 (m, 2H, CH2), 4.55 (d, 1H J = 8 Hz, H4isox), 5.08 (s, 1H, CH-N), 5.17 (d, 1H, J = 8 Hz, H5isox), 5.36 ( s, 1H, HC-O), 7.45 (m, 2H, arom.), 7.57 (m, 1H, arom.), 7.73 (m, 2H, arom.), 8.23 (s, 1H, N-CH=), 12.23 (s, 1H, NH). 13C-NMR (δ): 32.8, 41.2, 53.4, 56.3, 84.8, 109.8, 127.4, 128.2, 128.7, 131.9, 132.7, 138.7, 148.2, 159.0, 172.4. Elemental analysis: cald. for C17H13ClN4O5 (MW = 388.76) C 52.52, H 3.37, N 14.41. Found C 52.54, H 3.35, N 14.40.
1-(6-Benzoyl-4,6,7,7a-tetrahydro-3aH-4,7-methanoisoxazolo[5,4-d][1,2]oxazin-3-yl)-5-chloropyrimidine-2,4(1H,3H)-dione (6bB): 0.46 g (76%), m.p. 98–105 °C, colorless crystals from diisopropyl ether/ethanol. IR: νNH 3178 cm−1, νC=O 1732 cm−1, νC=N 1708 cm−1. 1H-NMR (δ): 2.07 (m, 2H, CH2), 4.38 (d, 1H J = 9 Hz, H4isox), 5.00 (s, 1H, CH-N), 5.28 ( s, 1H, HC-O), 5.32 (d, 1H, J = 9 Hz, H5isox), 7.46 (m, 2H, arom.), 7.58 (m, 1H, arom.), 7.70 (m, 2H, arom.), 8.24 (s, 1H, N-CH=), 12.25 (s, 1H, NH). 13C-NMR (δ): 32.7, 35.8, 54.6, 79.8, 84.1, 109.8, 127.4, 128.2, 128.3, 128.5, 131.8, 133.0, 138.6, 148.2, 152.5, 159.0, 162.3, 168.9. Elemental analysis: cald. for C17H13ClN4O5 (MW = 388.76) C 52.52, H 3.37, N 14.41. Found C 52.50, H 3.36, N 14.42.
1-(5-Benzoyl-4,5,7,7a-tetrahydro-3aH-4,7-methanoisoxazolo[4,5-d][1,2]oxazin-3-yl)-5-bromopyrimidine-2,4(1H,3H)-dione (6aC): 0.38 g (56%), m.p. 96–100 °C, colorless crystals from diisopropyl ether/ethanol. IR: νNH 3200 cm−1, νC=O 1734 cm−1, νC=N 1621 cm−1. 1H-NMR (δ): 1.95 (m, 2H, CH2), 4.54 (d, 1H J = 10 Hz, H4isox), 5.08 (s, 1H, CH-N), 5.16 (d, 1H, J = 10 Hz, H5isox), 5.36 (s, 1H, HC-O), 7.47 (m, 2H, arom.), 7.57 (m, 1H, arom.), 7.68 (m, 2H, arom.), 8.27 (s, 1H, N-CH=), 12.18 (s, 1H, NH). 13C-NMR (δ): 30.7, 32.8, 53.4, 79.6, 84.8, 98.6, 128.2, 128.7, 131.8, 132.8, 140.9, 148.4, 152.7, 159.2. Elemental analysis: cald. for C17H13BrN4O5 (MW = 433.21) C 47.13, H 3.02, N 12.93. Found C 47.10, H 3.01, N 12.95.
1-(6-Benzoyl-4,6,7,7a-tetrahydro-3aH-4,7-methanoisoxazolo[5,4-d][1,2]oxazin-3-yl)-5-bromopyrimidine-2,4(1H,3H)-dione (6bC): 0.33 g (49%), m.p. 104–106 °C, colorless crystals from diisopropyl ether/ethanol. IR: νNH 3121 cm−1, νC=O 1732 cm−1, νC=N 1702 cm−1. 1H-NMR (δ): 2.07 (m, 2H, CH2), 4.37 (d, 1H J = 9 Hz, H4isox), 5.00 (s, 1H, CH-N), 5.28 (s, 1H, HC-O), 5.31 (d, 1H, J = 9 Hz, H5isox), 7.49 (m, 2H, arom.), 7.57 (m, 1H, arom.), 7.69 (m, 2H, arom.), 8.29 (s, 1H, N-CH=), 12.20 (s, 1H, NH). 13C-NMR (δ): 26.3, 32.7, 54.6, 79.8, 84.1, 98.6, 128.3, 128.5, 131.7, 132.4, 133.0, 140.9, 148.4, 152.4, 159.1, 165.7, 168.9. Elemental analysis: cald. for C17H13BrN4O5 (MW = 433.21) C 47.13, H 3.02, N 12.93. Found C 47.14, H 3.03, N 12.95.
1-(5-Benzoyl-4,5,7,7a-tetrahydro-3aH-4,7-methanoisoxazolo[4,5-d][1,2]oxazin-3-yl)-5-iodopyrimidine-2,4(1H,3H)-dione (6aD): 0.42 g (57%), m.p. 110–115 °C, colorless crystals from diisopropyl ether/ethanol. IR: νNH 3394 cm−1, νC=O 1732 cm−1, νC=N 1654 cm−1. 1H-NMR (δ): 1.95 (m, 2H, CH2), 4.53 (d, 1H J = 9 Hz, H4 isox), 5.07 (s, 1H, CH-N), 5.15 (d, 1H, J = 9 Hz, H5 isox), 5.34 (s, 1H, HC-O), 7.46 (m, 2H, arom.), 7.57 (m, 1H, arom.), 7.67 (m, 2H, arom.), 8.23 (s, 1H, N-CH=), 12.04 (s, 1H, NH). 13C-NMR (δ): 26.3, 32.8, 53.4, 73.0, 79.6, 74.7, 127.4 128.2, 128.6, 128.7, 131.8, 132.8, 145.3, 148.8, 152.6, 166.0. Elemental analysis: cald. for C17H13IN4O5 (MW = 480.21) C 42.52, H 2.73, N 11.67. Found C 42.50, H 2.75, N 11.66.
1-(6-Benzoyl-4,6,7,7a-tetrahydro-3aH-4,7-methanoisoxazolo[5,4-d][1,2]oxazin-3-yl)-5-iodopyrimidine-2,4(1H,3H)-dione (6bD): 0.42 g (57%), m.p. 162–164 °C, colorless crystals from diisopropyl ether/ethanol. IR: νNH 3160 cm−1, νC=O 1720 cm−1, νC=N 1654 cm−1. 1H-NMR (δ): 2.05 (m, 2H, CH2), 4.35 (d, 1H J = 9 Hz, H4isox), 4.98 (s, 1H, CH-N), 5.26 ( s, 1H, HC-O), 5.28 (d, 1H, J = 9 Hz, H5isox), 7.52 (m, 3H, arom.), 7.68 (m, 2H, arom.), 8.24 (s, 1H, N-CH=), 12.05 (s, 1H, NH). 13C-NMR (δ): 32.7, 39.8, 54.7, 73.0, 79.9, 84.2, 127.4, 128.3, 128.4, 129.2, 131.7, 132.7, 142.1, 145.3, 148.8, 152.4, 160.7. Elemental analysis: cald. for C17H13IN4O5 (MW = 480.21) C 42.52, H 2.73, N 11.67. Found C 42.51, H 2.76, N 11.68.
1-(5-Benzoyl-4,5,7,7a-tetrahydro-3aH-4,7-methanoisoxazolo[4,5-d][1,2]oxazin-3-yl)-5-methylpyrimidine-2,4(1H,3H)-dione (7a): 0.26 g (45%), m.p. 215–217 °C, colorless crystals from diisopropyl ether/ethanol. IR: νNH 3176 cm−1, νC=O 1715 cm−1, νC=N 1674 cm−1. 1H-NMR (δ): 1.81 (d, 3H, J = 1 Hz, CH3), 1.92 (s, 2H, CH2), 4.58 (d, 1H J = 8 Hz, H4isox), 5.07 (s, 1H, CH-N), 5.12 (d, 1H, J = 8 Hz, H5isox), 5.30 (s, 1H, HC-O), 7.51 (m, 3H, arom.), 7.69 (d, 1H, J = 1 Hz, N-CH=), 7.73 (m, 2H, arom.), 11.70 (s, 1H, NH). 13C-NMR (δ): 11.9, 32.9, 54.1, 60.1, 79.8, 84.6, 111.8, 128.3, 128.8, 131.9, 133.1, 137.0, 148.9, 153.3, 163.8, 170.0. Elemental analysis: cald. for C18H16N4O5 (MW = 368.34) C 58.69, H 4.38, N 15.21. Found C 58.70, H 4.39, N 15.19.
1-(6-Benzoyl-4,6,7,7a-tetrahydro-3aH-4,7-methanoisoxazolo[5,4-d][1,2]oxazin-3-yl)-5-methylpyrimidine-2,4(1H,3H)-dione (7b): 0.31 g (55%), m.p. 220–222 °C, colorless crystals from diisopropyl ether/ethanol. IR: νNH 3149 cm−1, νC=O 1710 cm−1, νC=N 1698 cm−1. 1H-NMR (δ): 1.82 (d, 3H, J = 1 Hz, CH3), 2.04 (m, 2H, CH2), 4.41 (d, 1H J = 8 Hz, H4isox), 4.99 (s, 1H, CH-N), 5.25 (d, 1H, J = 8 Hz, H5isox), 5.27 (s, 1H, HC-O), 7.51 (m, 3H, arom.), 7.69 (d, 1H, J = 1 Hz, N-CH=), 7.73 (m, 2H, arom.), 11.63 (s, 1H, NH). 13C-NMR (δ): 11.9, 32.6, 55.1, 61.2, 80.1, 83.7, 111.7, 128.3, 128.5, 131.8, 130.0, 136.8, 148.8, 152.9, 163.7, 168.9. Elemental analysis: cald. for C18H16N4O5 (MW = 368.34) C 58.69, H 4.38, N 15.21. Found C 58.71, H 4.37, N 15.20.

3.4. X-ray Crystallographic Analysis of Compounds 5a and 7a

Unit cell dimensions for compounds 5a and 7a were obtained by least-squares fitting of 2θ values for 25 reflections, with an Enraf–Nonius CAD4 diffractometer and graphite-monochromated Mo–Kα radiation at the Centro Grandi Strumenti (CGS) of the University of Pavia, Italy. The structure was solved by direct methods, and the E-map correctly revealed the non-hydrogen atoms in the molecules. The positions of the hydrogen atoms were located by difference Fourier synthesis, compared with those calculated from the geometry of the molecules and refined isotropically in the subsequent least-squares refinement. The program SHELXL [27] was used to solve the structure. The ORTEP program [28] was used for molecular graphics. The ORTEP plots of 8-benzoyl-5-uracil-3,9-dioxa-4,8-diaza-[5.2.1.02.5]-dec-4-ene (5a) and 8-benzoyl-5-thymine-3,9-dioxa-4,8-diaza-[5.2.1.02.5]dec-4-ene (7a) are shown in Figure 1 and Figure 2 and the corresponding bond lengths and angles are given in Table 2.
A total 2,893 reflections for 5a and 1,828 for 7a was explored with the ω/2θ scan tecnique. Correction was applied for Lorentz and polarization. An approximate scale and a mean thermal factor of 2.867 Å2 and 3.417 Å2 were determined by Wilson statistic [29] for 5a and 7a. The structure was solved by direct methods and E-map correctly revealed all the non-hydrogen atoms in the molecules. The position of the hydrogen atoms was checked in a final difference Fourier-map and refined isotropically in the subsequent least- squares refinement. The experimental crystal X-ray data are summarized in Table 3 and Table 4, respectively. CCDC deposition numbers: (5a) 987413 and (7a) 987414.
Table 2. Selected bond lengths (Å) and angles (deg) (with Esd’s in Parentheses). Compounds 5a 7a 5a 7a.
Table 2. Selected bond lengths (Å) and angles (deg) (with Esd’s in Parentheses). Compounds 5a 7a 5a 7a.
C1—C21.523(4) 1.507(6)C5—C61.499(3) 1.498(5)
C1—O91.458(3) 1.456(5)C5—N111.417(3) 1.408(4)
C1—C101.501(4) 1.497(6)C6—C71.538(4) 1.533(5)
C2—O31.452(3) 1.453(4)C7—N81.477(3) 1.487(5)
C2—C61.534(3) 1.540(5)C7—C101.517(3) 1.520(6)
O3—N41.420(2) 1.402(4)N8—O91.450(2) 1.455(4)
N4—C51.270(3) 1.272(4)N8—C181.376(3) 1.377(5)
Bond Angles (°)
C2—C1—O9104.8(2) 104.3(2)C2—C6—C5100.1(2) 99.8(3)
C2—C1—C10103.9(2) 104.2(4)C2—C6—C7101.9(2) 101.7(3)
O9—C1—C10103.3(2) 103.8(4)C6—C7—N8104.0(2) 104.2(3)
C1—C2—C6102.6(2) 102.9(3)C6—C7—C10103.2(2) 102.7(4)
O3—C2—C6105.8(2) 105.5(3)N8—C7—C10101.6(2) 101.2(3)
C2—O3—N4109.4(2) 109.4(3)C7—N8—O9105.2(2) 105.3(3)
O3—N4—C5109.1(2) 109.9(3)C7—N8—C18126.9(2) 124.1(3)
N4—C5—C6115.6(2) 115.1(3)O9—N8—C18112.9(2) 110.0(3)
N4—C5—N11118.1(2) 116.4(3)C1—O9—N8103.7(2) 103.4(3)
C6—C5—N11126.4(2) 128.3(3)C1—C10—C792.7(2) 92.9(4)
Table 3. Crystal data, data collection and structure refinement (compound 5a).
Table 3. Crystal data, data collection and structure refinement (compound 5a).
Empirical FormulaC17H14N4O5
Formula weight354.32
Crystal size, mm0.56 × 0.42 × 0.18
Temperature, K293
Crystal systemOrthorhombic
Space groupP bca
a, Å9.977(4)
b, Å16.774(3)
c, Å19.649(3)
α90
β90
γ90
V, Å33288(1)
Z8
Dcalcd, g·cm−31.431
Absorption coeff., µ, mm−10.108
Diffractometer/scanEnraf–Nonius CAD–4, θ/2θ
λ, Å0.71073
F(000)1472
Range (°) for data2.0 < θ > 25
Index ranges 0 < h> 11, 0 < k> 19, 0 < l > 23
No. of reflects. measd2893
No. of unique reflects1824
Correction appliedLorentz–polarization
Refinement methodFull–matrix least–squares
No. of variables291
Goodness–of–fit (2893)0.900
R1 (I) > 2 σ (I), (1824)0.0403
R1 (2883)0.0785
(∆ρ) max, min, eÅ−30.124, −0.151
Table 4. Crystal data, data collection and structure refinement (compound 7a).
Table 4. Crystal data, data collection and structure refinement (compound 7a).
Empirical FormulaC20H22N4O6
Formula weight414.417
Crystal size, mm0.50 × 0.385 × 0.14
Temperature, K293
Crystal systemMonoclinic
Space groupP 21/n
a, Å11.991(3)
b, Å11.4437(8)
c, Å14.63(1)
α90.0
β101.34(4)
γ90.0
V, Å31968.3(14)
Z4
Dcalcd, g·cm−31.3985
Absorption coeff., µ, mm−10.1051
Diffractometer/scanEnraf–Nonius CAD–4, θ/2 θ
RadiationMoKα
λ, Å0.71073
F(000)872
Range (°) for data2.0 < θ > 25.0
Index ranges−11 < h > 11, 0 < k > 10, 0 < l > 14
No. of reflects. measd1828
No. of unique reflects1266
Correction appliedLorentz–polarization
Refinement methodFull–matrix least–squares
No. of variables359
Goodness–of–fit (1828)1.024
R1 (I) > 2 σ (I), (1266)0.0414
R1 (1828)0.0728
(∆ρ) max, min, eÅ−30.119, −0.129

4. Conclusions

The regioisomeric cycloadducts of the bromonitrile oxide to the N-benzoyl-2,3-oxazanorborn-5-ene were easily prepared and elaborated into a novel class of uracil-based scaffolds. The key synthetic step is represented by the nucleophilic substitution at the sp2 carbon atom of the bromoisoxazoline three-dimensional heterocycles. The protocol to perform the nucleophilic substitution of uracil anions was revised, and optimized as well as adapted to the steric requirements of the substrates.
Regarding the uracil derivatives of type 5, 6 and 7 the synthetic elaboration will continue with standard hydrolysis for the detachment of the benzoyl group and hydrogenolysis of the resulting N-O bond to give the desired aminols. These latter represent a novel class of nucleoside analogues [21] to be tested against a variety of viruses or used as synthons for the preparation of β-turn inducers containing nucleobases for their insertion in PNAs [12,30,31].

Supplementary Materials

Supplementary materials can be accessed at: https://www.mdpi.com/1420-3049/19/6/8661/s1.

Acknowledgments

Financial support by the University of Pavia, MIUR (PRIN 2011, CUP: F11J12000210001) is gratefully acknowledged. Thanks are also due to the Consorzio Interuniversitario Nazionale Metodologie e Processi Innovativi di Sintesi (CINMPIS) for research support. We also thank the Fondazione Banca del Monte di Lombardia (Progetto Professionalità I. Becchi 2013) for a grant to MGM. COST Action CM1004 “Synthetic Probes for Chemical Proteomics and Elucidation of Biosynthetic Pathways” is gratefully acknowledged.

Author Contributions

The listed authors contributed to this work as described in the following. Misal G. Memeo gave the concepts of work and interpreted the results. Francesco Lapolla carried out of the synthetic work and interpreted the results. Bruna Bovio performed the X-Ray analyses and cooperated in the preparation of the manuscript. Paolo Quadrelli prepared the manuscript and contributed with valuable discussions as well as scientific input and funding of Spectra for Substances. All authors read and approved the final manuscript.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Savion, M.; Memeo, M.G.; Bovio, B.; Grazioso, G.; Legnani, L.; Quadrelli, P. Synthesis and molecular modeling of novel dihydroxycyclopentane-carbonitrile nor-nucleosides by bromonitrile oxide 1,3-dipolar cycloaddition. Tetrahedron 2012, 68, 1845–1852. [Google Scholar] [CrossRef]
  2. Quadrelli, P.; Bovio, B.; Piccinini, A.; Caramella, P.; de Sarlo, F.; Machetti, F. Conversion of a nitrosocarbonyl hetero Diels-Alder cycloadduct to useful isoxazoline-carbocyclic aminols. Tetrahedron 2009, 65, 10679–10686. [Google Scholar] [CrossRef]
  3. Bodnar, B.S.; Miller, M.J. The nitrosocarbonyl hetero-diels–alder reaction as a useful tool for organic syntheses. Angew. Chem. Int. Ed. 2011, 50, 5630–5647. [Google Scholar] [CrossRef]
  4. Scagnelli, L.; Memeo, M.G.; Carosso, S.; Bovio, B.; Quadrelli, P. Syntheses of New Carbanucleosides by Pericyclic Reactions. Eur. J. Org. Chem. 2013, 2013, 3835–3846. [Google Scholar] [CrossRef]
  5. Kirby, G.W. Electrophilic C-Nitroso-compomds. Chem. Soc. Rev. 1977, 6, 1–24. [Google Scholar] [CrossRef]
  6. Quadrelli, P.; Mella, M.; Gamba Invernizzi, A.; Caramella, P. The mild oxidation of nitrile oxides affords a convenient entry to nitrosocarbonyl intermediates, versatile tools in organic synthesis. Tetrahedron 1999, 55, 10497–10510. [Google Scholar] [CrossRef]
  7. Quadrelli, P.; Mella, M.; Paganoni, P.; Caramella, P. Cycloadditions of nitrile oxides to the highly reactive N-Acyl-2-oxa-3-azanorborn-5-enes afford versatile cycloadducts and a convenient entry to highly functionalized derivatives. Eur. J. Org. Chem. 2000, 14, 2613–2620. [Google Scholar]
  8. Quadrelli, P.; Scrocchi, R.; Caramella, P.; Rescifina, A.; Piperno, A. From cyclopentadiene to isoxazoline–carbocyclic nucleosides: A rapid access to biological molecules through nitrosocarbonyl chemistry. Tetrahedron 2004, 60, 3643–3651. [Google Scholar] [CrossRef]
  9. Quadrelli, P.; Mella, M.; Carosso, S.; Bovio, B.; Caramella, P. A straightforward synthesis of isoxazoline-based carbocyclic nucleosides from 1,3-cyclohexadiene through nitrosocarbonyl chemistry. Eur. J. Org. Chem. 2007, 36, 6003–6015. [Google Scholar]
  10. Moggio, Y.; Legnani, L.; Bovio, B.; Memeo, M.G.; Quadrelli, P. Synthesis of novel anthracene derivatives of isoxazolino-carbocyclic nucleoside analogues. Tetrahedron 2012, 68, 1384–1392. [Google Scholar] [CrossRef]
  11. Miyauchi, H.; Chiba, S.; Fukamizu, K.; Ando, K.; Narasaka, K. Synthesis of hetero- and carbocycles by nucleophilic substitution at sp2 carbon. Tetrahedron 2007, 63, 5940–5953. [Google Scholar] [CrossRef]
  12. North, M. Incorporation of conformationally constrained β-amino acids into peptides. J. Pept. Sci. 2000, 6, 301–313. [Google Scholar] [CrossRef]
  13. Memeo, M.G.; Bovio, B.; Quadrelli, P. RuO 4-catalyzed oxidation reactions of isoxazolino-2-azanorbornane derivatives: A short-cut synthesis of tricyclic lactams and peptidomimetic γ-amino acids. Tetrahedron 2011, 67, 1907–1914. [Google Scholar] [CrossRef]
  14. Caldirola, P.; Ciancaglione, M.; de Amici, M.; de Micheli, C. Conversion of isoxazolines to β-hydroxy esters. Synthesis of 2-deoxy-D-ribose. Tetrahedron Lett. 1986, 27, 4647–4650. [Google Scholar] [CrossRef]
  15. Pinto, A.; Conti, P.; Tamborini, L.; de Micheli, C. A novel simplified synthesis of acivicin. Tetrahedron 2009, 20, 508–511. [Google Scholar] [CrossRef]
  16. Freeman, J. Possible criteria for distinguishing between cyclic and acyclic activated complexes and among cyclic activated complexes in addition reactions. Chem. Rev. 1975, 75, 439–489. [Google Scholar] [CrossRef]
  17. Mehta, G.; Chandrasekhar, J. Electronic control of facial selection in additions to sterically unbiased ketones and olefins. Chem. Rev. 1999, 99, 1437–1467. [Google Scholar] [CrossRef]
  18. Huisgen, R.; Ooms, P.H.J.; Mingin, M.; Allinger, N.L. Exceptional reactivity of the bicyclo[2.2.1]heptene double bond. J. Am. Chem. Soc. 1980, 102, 3951–3953. [Google Scholar] [CrossRef]
  19. Rondan, N.G.; Paddon-Row, M.N.; Caramella, P.; Houk, K.N. Nonpanar alkenes and carbonyls: A molecular distortion which parallels addition stereoselectivity. J. Am. Chem. Soc. 1981, 103, 2436–2438. [Google Scholar] [CrossRef]
  20. Houk, K.N.; Rondan, N.G.; Brown, F.K.; Jorgensen, W.L.; Madura, J.D.; Spellmeyer, D.C. Electronic origins and consequences of pyramidalization of asymmetric alkenes in ground and triplet excited states. J. Am. Chem. Soc. 1983, 105, 5980–5988. [Google Scholar]
  21. Coutouli-Argyropoulou, E.; Pilanidou, P. An entry to new isoxazoline analogues of dideoxynucleosides by bromonitrile oxide 1,3-dipolar cycloaddition. Tetrahedron Lett. 2003, 44, 3755–3758. [Google Scholar] [CrossRef]
  22. Quadrelli, P.; Bovio, B.; Piccanello, A. Structure determination of 8-benzyl-5-phenyl-3-oxa-4,8-diazatricyclo[5.2.1.02,6]dec-4-ene and 1-(9-Ethoxy-5-phenyl-3-oxa-4,8-diaza-tricyclo[5.2.1.02,6]dec-4-en-8-yl)-ethanone: Their synthesis, chemical relationship and comparison with similar compounds. J. Chem. Crystallogr. 2012, 42, 43–66. [Google Scholar] [CrossRef]
  23. Cremer, D.; Pople, J. General definition of ring puckering coordinates. J. Am. Chem. Soc. 1975, 97, 1354–1358. [Google Scholar] [CrossRef]
  24. Bernasconi, C.F.; Rappoport, Z. Recent advances in our mechanistic understanding of SNV reactions. Acc. Chem. Res. 2009, 42, 993–1003. [Google Scholar] [CrossRef]
  25. Fernandez, I.; Bickelhaupt, F.M.; Uggerud, E. Reactivity in nucleophilic vinylic substitution (SNV):SNversus SNVσ mechanistic dichotomy. J. Org. Chem. 2013, 78, 8574–8484. [Google Scholar] [CrossRef]
  26. Frisch, M.J.; Trucks, G.W.; Schlegel, H.B.; Scuseria, G.E.; Robb, M.A.; Cheeseman, J.R.; Scalmani, G.; Barone, V.; Mennucci, B.; Petersson, G.A. Calculated Structure of 4a at the B3LYP 6–31G* Level by Means of Gaussian 09; Revision D.01; Gaussian, Inc.: Wallingford, CT, USA, 2009. [Google Scholar]
  27. Sheldrick, G.M. SHELXL-93. Program for the Refinement of Crystal Structure; University of Göttingen: Göttingen, Germany, 1993. [Google Scholar]
  28. Johnson, C.K. ORTEP, Report ORNL-3793; Oak Ridge National Laboratory: Oak Ridge TN, USA, 1966. [Google Scholar]
  29. Wilson, A.J.C. Determination of absolute from relative X-ray intensity data. Nature 1942, 150, 152. [Google Scholar] [CrossRef]
  30. André, C.; Legrand, B.; Deng, C.; Didierjean, C.; Pickaert, G.; Martinez, J.; Averlant-Petit, M.C.; Amblard, M.; Calmes, M. (S)-ABOC: A rigid bicyclic β-amino acid as turn inducer. Org. Lett. 2012, 14, 960–963. [Google Scholar] [CrossRef]
  31. Gorska, K; Winssinger, N. Reactions templated by nucleic acids: More ways to translate oligonucleotide-based instructions into emerging function. Angew. Chem. Int. Ed. 2013, 52, 6820–6843. [Google Scholar] [CrossRef]
  • Sample Availability: Not available.

Share and Cite

MDPI and ACS Style

Memeo, M.G.; Lapolla, F.; Bovio, B.; Quadrelli, P. Three-Dimensional Heterocycles: New Uracil-Based Structures Obtained by Nucleophilic Substitution at the sp2 Carbon of Bromoisoxazoline. Molecules 2014, 19, 8661-8678. https://doi.org/10.3390/molecules19068661

AMA Style

Memeo MG, Lapolla F, Bovio B, Quadrelli P. Three-Dimensional Heterocycles: New Uracil-Based Structures Obtained by Nucleophilic Substitution at the sp2 Carbon of Bromoisoxazoline. Molecules. 2014; 19(6):8661-8678. https://doi.org/10.3390/molecules19068661

Chicago/Turabian Style

Memeo, Misal Giuseppe, Francesco Lapolla, Bruna Bovio, and Paolo Quadrelli. 2014. "Three-Dimensional Heterocycles: New Uracil-Based Structures Obtained by Nucleophilic Substitution at the sp2 Carbon of Bromoisoxazoline" Molecules 19, no. 6: 8661-8678. https://doi.org/10.3390/molecules19068661

Article Metrics

Back to TopTop