Next Article in Journal
Comparative Efficacy of Botryocladia leptopoda Extracts in Scar Inhibition and Skin Regeneration: A Study on UV Protection, Collagen Synthesis, and Fibroblast Proliferation
Previous Article in Journal
The Effect of Lactiplantibacillus plantarum and Lacticaseiba-cillus Rhamnosus Strains on the Reduction of Hexachlorobenzene Residues in Fermented Goat Milk During Refrigerated Storage
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Bifunctional Azido(thio)ureas from an O-Protected 2-Amino-2-deoxy-d-glucopyranose: Synthesis and Structural Analyses

by
Concepción Sosa-Gil
1,
Esther Matamoros
1,2,3,*,
Pedro Cintas
1 and
Juan C. Palacios
1,*
1
Departamento de Química Orgánica e Inorgánica, Facultad de Ciencias, and Instituto del Agua, Cambio Climático y Sostenibilidad (IACYS)-Unidad de Química Verde y Desarrollo Sostenible, Universidad de Extremadura, 06006 Badajoz, Spain
2
Departamento de Química Orgánica, Universidad de Málaga, Campus Teatinos s/n, 29071 Málaga, Spain
3
Instituto de Investigación Biomédica de Málaga y Plataforma en Nanomedicina—IBIMA, Plataforma Bionand, Parque Tecnológico de Andalucía, 29590 Málaga, Spain
*
Authors to whom correspondence should be addressed.
Molecules 2024, 29(23), 5687; https://doi.org/10.3390/molecules29235687
Submission received: 10 November 2024 / Revised: 26 November 2024 / Accepted: 29 November 2024 / Published: 30 November 2024

Abstract

:
This publication reports a facile and convenient preparation of tri-O-acetyl-glucopyranoses, derived from the corresponding 2-deoxyaminosugar, where the vicinal anomeric and C2 positions are decorated by azido and (thio)ureido groups, respectively. This double functionalization leads to an inherently chiral core incorporating the versatile azido and (thio)ureido linkages prone to further manipulation. The latter also provides a structural element for hydrogen-bonded donor-acceptor (HB-DA) sites, which are of immense value in organocatalytic pursuits. A computation-aided conformational analysis unveils the landscape of available conformers and their relative stability. N-aryl (thio)ureas bearing substituents at ortho positions exist as mixtures of M- and P-atropisomeric conformers.

Graphical Abstract

1. Introduction

The preparation and importance of ureido derivatives dates back to the dawn of organic synthesis, and interest has grown at a fast pace over the past century, which can be exemplified by numerous applications. Thus, the urea function itself can form inclusion complexes with long chain alkanes and some functionalized molecules [1,2]. In such compounds, the urea molecules are bound together by intramolecular hydrogen bonds (HBs), leading to helical networks in which the host compounds can be oriented and accommodated. Urea structures also function as strong HB donors and acceptors [3] and have been widely employed for obtaining supramolecular architectures [4,5]. Substituted ureas and bis-ureas also form gels with organic liquids and water [6,7,8,9,10,11,12,13,14,15], and in some cases, the resulting organo/hydro-gels are thermally reversible and require very little concentration [16]. The scientific literature also describes pseudodisaccharides O-protected by acyl groups and linked by the thiourea function [17,18,19], while, surprisingly, the corresponding disaccharide-type derivatives linked by ureido groups are less known. Some ureas derived from carbohydrates exhibit interesting biological properties such as the SF-1993 [20] and CV-1 [21,22] antibiotics and the family of glycocinnamoyl espermidines [23,24]. Moreover, glycosylureas are inhibitors of α-glycosidase [25] and N-nitrosoureas, such as streptozotocin [26,27] and chlorozotocin [28], and possess antitumor activity [29]. These substances induce DNA crosslinking after its in vivo degradation. The isocyanate generated by this mechanism does not seem to be directly involved in the antitumor effect, but it does react with biogenic amines and proteins. Other side effects include the inhibition of RNA synthesis, DNA polymerase, and other enzymes involved in DNA repair, all accounting for the toxicity of nitrosoureas [30].
Methods aimed at synthesizing ureidosugars are usually limited to a few protocols, involving the condensation of monosaccharides with ureas to give symmetrical glycosylureas [31] and ureidodisaccharides [32,33], the reaction of glycosylamines [34,35], amino sugars [36,37,38,39,40] with isocyanates or their synthetic equivalents such as carbamates [41,42,43], the addition of water to diglycosylcarbodiimides [44,45], as well as by-products in various transformations of glycosyl isocyanates [43,46], each possessing inherent pros and cons.
It is likely that the synthetic protocol that still remains versatile enough to access a broad range of structurally diverse glycoconjugates and pseudo-oligosaccharides is the reaction of O-protected amino sugars with either isocyanates or isothiocyanates. Therefore, in this work, we describe the preparation of a series of ureas and thioureas linked to the sugar fragment through non-anomeric carbon atoms, starting from d-glucosamine as the available precursor. In order to increase the functionality relative to previous work, we thought of the labile azido group at the anomeric carbon as tri-O-acetyl-2-amino-2-deoxy-β-d-glucopyranosyl azide (1) can readily be obtained [47]. Organic azides can undergo a variety of synthetically useful transformations, particularly reducing into amine group [48,49,50] or the unique azide-alkyne cycloaddition [51,52,53,54]. On the other hand, sugar iso(thio)cyanates have long been known in carbohydrate chemistry, with the first example at the non-anomeric position being reported by Jochims and Seeliger [55] as early as 1965 using a phosgene stream. Clearly, the use of neat phosgene is to be highly discouraged in contemporary synthesis. A safer and less hazardous surrogate is a 1.93 M solution of phosgene in toluene, which was applied by Nowick et al. [56] to the preparation of isocyanates derived from amino acid esters. We have employed this in previous work in the synthesis of isocyanates from amino sugars (25) [57]. A third method, which is safe enough, involves triphosgene, Cl3C-OCOO-CCl3, a synthetic equivalent of phosgene, under heterogeneous conditions, similar to those of the Schotten–Baumann synthesis [58]. This way, isocyanate 6 has been obtained from 1 (Chart 1). In contrast, liquid thiophosgene, even if highly toxic, can easily be manipulated under heterogeneous biphasic conditions and removed by distillation. The method has been applied to the preparation of isothiocyanates 7–10 [17,18,19,55].

2. Results and Discussion

2.1. Syntheses and Structural Characterization

3,4,6-Tri-O-acetyl-2-amino-2-deoxy-β-d-glucopyranosyl azide (1) is an ideal starting material for the preparation of chiral derivatives functionalized at C1 and C2 pyranose positions. Its synthesis, according to an old protocol [47], involves the treatment of d-glucosamine hydrochloride (11) with acetyl bromide and the subsequent treatment of the resulting α-configured glucopyranosyl bromide hydrobromide (12) with silver azide, which follows an SN2 mechanism and proceeds with the reversal of the anomeric configuration, thus forming the β-glycosyl azide (Scheme 1).
The structure of 1 was confirmed by coincidence of its physical data with those described [47] and was supported by analytical and spectroscopic data. Thus, the IR spectrum shows both asymmetric and symmetric stretching vibrations of the NH2 group at 3380 and 3326 cm−1, respectively, together with an intense stretching band of the azido group at 2112 cm−1. The 1H NMR spectrum shows three signals for acetate methyl groups at ~2.1 ppm, while the anomeric proton (4.54 ppm) and the H-2 proton (2.83 ppm) lie in a shielding zone and are shifted upfield due to the low electronegativity of the vicinal azido and amino groups. The high coupling constants J1,2 = J2,3 = J3,4 = J4,5 (=9.5 Hz) are consistent with a d-gluco pyranoid ring in 4C1 conformation and β-anomeric configuration.
The reaction of 1 with equimolar amounts of an aryl isocyanate or aryl isothiocyanate (13) in benzene or CH2Cl2 at room temperature led to the ureido and thioureido derivatives 1422 (Scheme 2), which often crystallized spontaneously and could be isolated in moderate to high yields (45–97%), although their syntheses have not been optimized (Table S1). Again, all spectroscopic data fully support such structures. IR spectra show absorptions at ~3550–3250 cm−1, ~2100 cm−1, ~1750 cm−1, ~1250 cm−1, and ~1650 cm−1, accounting for stretching vibrations of the multiple functional groups present, namely NH, N=N=N, C=O (ester and amide), and C-O-C bonds [59]. Most protons and carbon atoms of compounds 1422 show similar chemical shifts. In the 1H NMR spectra, downfield resonances correspond in all cases to the NH protons of either ureido or thioureido linkages between 8.2 and 7.0 ppm and which are close to or overlapping the typical aromatic protons (~7.5–6.5 ppm). Signals at the upper field, in the interval between 5.2 and 3.7 ppm, belong to the aminosugar moiety and appear in the following order: H-3 > H-4 > H-1 > H-6 > H-6′ > H-2 > H-5 (Table 1). When comparing the signals of H-2 of 1422 with those of the starting product 1, a strong deshielding is generally observed (Δδ ~1.0 ppm for the ureido derivatives and Δδ ~2.1 ppm for the thioureido ones) as a consequence of the transformation of the NH2 group into the (thio)urea functionalities (Table 1). The remaining signals can easily be attributed to the aromatic protons of the starting isocyanate or isothiocyanate (Table S2). For meta- and para-OMe-substituted compounds, the methyl protons resonate at ~3.8 ppm. Unambiguous assignments could be established through double-resonance experiments, 2D correlations, and D2O isotopic exchange, with the latter proving that the singlet signal between 7.0 and 8.0 ppm corresponds to the NH group directly bound to the aromatic ring, whereas the doublet at ~6.0 ppm corresponds to the other NH linked to the sugar.
Likewise, 13C chemical shits (Table 2) could be assigned through polarization transfer (DEPT) and heteronuclear 1H-13C correlation spectra. As expected, the most deshielded signal at ~180 ppm can be attributed to the C=S bond of thioureas, while the carbonyl group of ureido derivatives appear at ~150 ppm. The carbonyls of the acetates groups are located between them at ~170 ppm, while the signals of their methyl groups invariably appear at ~21 ppm. The carbon signals of the pyranose ring generally follow a diagnostic sequence as follows: C-1 > C-5 > C-3 > C-4 > C-6 > C-2. Due to the poorly electron-withdrawing character of the azido group, the chemical shift in the anomeric carbon lies below 90 ppm. For substituted aryl rings, the methoxy groups of 1619 appear at ~55 ppm and the methyl signal of 22 at ~18 ppm.
A vigorously stirred biphasic reaction of thiophosgene and a chloroform solution of 1 in the presence of a saturated aqueous solution of CaCO3 made it possible to obtain 3,4,6-tri-O-acetyl-2-deoxy-2-isothiocyanato-β-d-glucopyranosyl azide (23) in a 58% yield (Scheme 3). The IR spectrum of the latter shows strong absorptions at 2120 and 2077 cm−1, typical of stretching vibrations of the N=N=N and N=C=S heterocumulenes, respectively. The isothiocyanato group causes a strong deshielding of the H-2 proton (~0.9 ppm) in the 1H NMR spectrum, while the carbon signal resonates at 142 ppm.
The possibility of condensing 23 with substituted anilines provides a straightforward route to thioureas. As shown in Scheme 3, the use of 2-chloro-6-methylaniline (24) led to the corresponding thiourea 25, precipitated from the reaction mixture as an orange solid after 72 h. The low isolated yield (22%) is most likely due to the steric hindrance caused by two substituents at the ortho positions. Although its spectroscopic data are similar to those outlined above for other thioureas, the interpretation of NMR spectra recorded at room temperature is challenging. The spectra are poorly resolved, and some signals are absent relative to similar structures, such as 22, for instance (Figure S44, Supplementary Materials). As we will see later, this behavior evidences the existence of a slow equilibrium between different conformers. Table 1 and Table 2 collect the most salient spectroscopic parameters of 25. The 13C NMR spectrum could be resolved at 223 K, the temperature at which the signals split into those of the individual conformers.

2.2. Conformational Analysis of Urea/Thiourea Derivatives

All the new compounds show large coupling constants, J2,3 = J3,4 = J4,5 (>8.5 Hz), in agreement with a d-gluco-configured pyranoid ring in a 4C1 conformation (Table 3) and β-configuration at the anomeric position, the latter being witnessed by the large coupling constant J1,2 (~9.5 Hz).
Furthermore, the high value measured for JNH,H-2 > 7.5 Hz evidences an antiperiplanar relationship between the H-2 and NH protons. As a result, the half-plane containing the urea/thiourea group is arranged approximately perpendicular to the plane containing the pyranoid ring (Figure 1).
As mentioned above, some 1H NMR spectra of these (thio)ureido derivatives show broad signals with poor resolution. A preliminary surmise points to the average spectra resulting from rotamers that interconvert slowly at room temperature. The origin of such a restricted rotation should reasonably be ascribed to the presence of amide or thioamide bonds. In addition, compounds 22 and 25 could exhibit double configurational restrictions arising from both the urea/thiourea linkages and the bulky groups located at the ortho positions [60,61] (Scheme 4).
Variable-temperature experiments were carried out to verify this working hypothesis. Thus, when decreasing the temperature, the proton signals of compound 22 broadened and ultimately split into two rounded peaks, thus indicating the coexistence of two rotamers at least. Something similar happened when recording the spectra of thiourea 25 at different temperatures. As the temperature increased, the wide signals observed at room temperature evolved to sharp and well-defined peaks, while when decreasing the temperature to 223 K, double signal sets could be detected. The coalescence temperature (Tc) was determined to be 270 K. These experiments allowed us to quantify the energy barrier to rotation. At the coalescence temperature, the rate constant between two equilibrium states equally populated and uncoupled is given by the following expression [62,63]:
k = πΔν/(2)½
where Δν represents the frequency difference between the analogous signals, corresponding to both rotamers. This expression is not completely accurate when the equilibrium takes place between unequally populated states, yet it can be used to obtain an estimated value of the interconversion barrier because the error is usually small. On the other hand, according to the Theory of the Transition State, the variation in the rate constant of an elementary reaction step with temperature is given by Equation (2) [64]:
k = (kBTc/h)exp − (ΔG/RTc)
where R, kB, and h represent the ideal gas, Boltzmann, and Plank constants, respectively.
From Equations (1) and (2), the barrier (ΔG) can be deduced by equaling them in their logarithmic form as follows:
ln[πΔν/(2)½] = ln (kBTc/h) − (ΔG/RTc)
and rearranging the terms as follows:
ΔG/RTc = ln[kBTc(2)½/πhΔν]
resulting in an expression used for calculating the rotational barrier:
ΔG = RTcln[kBTc(2)½/πhΔν]
By substituting the constants in the latter, Equation (6), which has been employed in our calculations, results in
ΔG (cal/mol) = 1.987Tc [22.62 + ln(Tc/Δν)]
For compound 25, taking into account its coalescence temperature and the chemical shift difference Δν (in frequency units), the activation free energy (ΔG) corresponds to 11.2 kcal/mol. This rotational barrier is consistent with the existence of a slow equilibrium at room temperature, which probably corresponds to the interconversion of rotamers of the thioamide bond.

2.3. Theoretical Calculations

To clarify further and rationalize the above assumptions, a series of DFT-based calculations [65,66,67,68,69,70] have been performed on the stability of the different conformations adopted by ureas and thioureas in solution. Modeling was carried out with the Gaussian package [71] using the hybrid density functional M0-62X [72] in combination with the 6-311G(d,p) [73,74] and the def2-TZVP valence-triple-ζ [75] basis sets, both in the gas phase and when simulating the presence of chloroform (the SMD method) [76], the solvent in which NMR spectra were recorded.
Since the NH and the H-2 signals always show an antiperiplanar disposition (JNH,H-2 > 7.5 Hz), in line with other (thio)ureido derivatives from sugars [60,61], the possible conformers are reduced to four for 1421. Compounds 22 and 25, however, owing to the restricted rotation caused by ortho substituents, could also occur as a mixture of M and P atropisomers, thereby doubling the number of potential conformers (Figure 2).
Table 4 shows the energies calculated for the different conformations, (Z,Z), (Z,E), and (E,Z), of urea 14 and Scheme 5 for their optimized structures. The most stable rotamer, both in the gas phase and in CHCl3, is the (Z,Z) structure, which should be equilibrated in solution with the (Z,E) rotamer.
An analogous computation has been carried out for the conformational space of azido thiourea 15, whose data and optimized structures are shown in Table 4 and Scheme 6, respectively. Conversely, the (Z,E) conformer of 15 now possesses the greatest stability in both the gas phase and in solution, a fact that deviates from the stability order obtained for the conformers of 14. Clearly, the larger volume of the thiocarbonyl group C=S relative to that of C=O is behind this change.
Table 5 and Table 6 depict a selection of bond distances, virtual angles, and significant dihedral angles for the optimized structures of the (Z,Z), (Z,E), and (E,Z) conformers of 14 and 15. For the sake of clarity, Figure 3 and Figure 4 show the numbering codes to define the bond lengths and angles. In all cases, the pyranoid ring retains the 4C1 conformation, which minimizes the steric interactions of the ring substituents. Interestingly, the linear azide group adopts a quasi-axial arrangement and is parallel to the C=O/C=S groups in the (Z,Z) and (Z,E) conformers. The dihedral angles between N39-N18 and C37-O38/C37-S55 are close to 0°.
It is noteworthy that the distance between O38 and H48 at the aromatic ring of 14 is significantly lower than the sum of the van der Waals radii of both atoms (2.19–2.21 < rvwH + rvwO = 1.17 + 1.52 = 2.69 Å) [77], indicating a putative interaction that contributes to the coplanarity of the benzene ring with the flat urea system, as reflected by the low dihedral angle involving the C=O bond (C37-O38) and the C44-C45 bond of the aromatic ring (~5°). Thiourea 15, however, lacks such a coplanarity (~40°); the aromatic ring rotates to accommodate the bulky sulfur atom, and the distance between the latter and H47 at the aromatic ring is very close to the sum of the van der Waals radii (2.78–2.82~rvwH + rvwS = 1.17 + 1.8 = 2.97 Å).
On the other hand, the spatial orientation between the proton at C-2 of the pyranose and the NH group is close to an antiperiplanar arrangement. They form dihedral angles of ~140–160º in the (Z,Z), (Z,E), and (E,Z) conformers of 14 and 15. This geometry can also be inferred from experimental NMR data, i.e., the high values of JNH,H-2 (~8–9 Hz), which indicate an antiperiplanar relationship between both protons. Moreover, the downfield shifts shown by the H-2 of ureas (Δδ ~1 ppm) and thioureas (Δδ ~2 ppm) with respect to that observed for the H-2 signal of 1 agree with an arrangement in which the C=O or C=S bonds lie parallel to the C2-H bond. Overall, the conformational arrangement involving such bonds should generate, at least in part, a strong deshielding with anisotropic effects experienced by H-2 in particular (Figure 1).
A conformational analysis of the aryl moiety of 25 has been achieved, and the energy landscape of the arrangements generated around both N-C(S) bonds calculated, i.e., by rotating the dihedral angle θN-C(S)-N-Carom from 0° to 360° with a step size of 15° each. Again, the computational analysis was performed using the aforementioned 6-311G(d,p) and def2-TZVP valence-triple-ζ basis sets, with all geometries optimized in the gas phase at the M06-2X level of theory without any geometrical restriction. Bulk solvation has been reproduced with the SMD method as well. The M06-2X/def2-TZVP combination has been reported to provide suitable geometry optimization in terms of cost and accuracy for carbohydrate derivatives [78,79,80]. Such results are shown in Figure 5 and Table 7 and Table 8.
Results are similar at both levels of calculation. The most stable conformations found correspond to the ZZ and ZE arrangements (entries 1 and 2, respectively) and the calculated barrier to interconversion between them is ~10–12 kcal/mol (Table 9, Figure 6), a value very close to that found experimentally using dynamic NMR.
Furthermore, this value is similar to that reported for thioureido derivative 26 (~14 kcal/mol), having a close structural resemblance to 25, which was assigned to the interconversion between M and P rotamers [60]. However, the cyclic (thio)ureas 2731 showed higher barriers to this rotation, >22 kcal/mol, as measured by dynamic NMR and/or estimated by theoretical calculations (Chart 2). Accordingly, both rotamers are configurationally stable and could be isolated [60,61].
The gap found in such barriers could be reasonably ascribed to the greater flexibility of the acyclic thiourea system. To verify the point, we calculated the rotational profile around the angle θC2-C1-N-C(S). A simpler molecular model (33) was employed to avoid the conformational complications caused by the sugar moiety and to reduce the computational cost. The expected landscape is shown in Scheme 7, which was carried out from an angle of 280° with a step size of 20° until completing 360°, first clockwise and then counterclockwise. Data for all minima and maxima found are collected in Table 10 and depicted in Figure 7. The determination of the stabilities of the M and P rotamers and the corresponding transition states would allow us to estimate the barrier to rotation.
Both curves are practically coincidental and show that the rotation barrier takes a value of ~13.5 kcal/mol, similar to that determined for 25 and 26. Figure 8 shows the structures of the corresponding maxima and minima. It is worth pointing out that the initial ZZP-rotamer does not transform into the ZZM-rotamer, but into the ZEP; that is, the same transformation as for compound 25, as shown in Figure 5. Notably, the saddle points do not correspond to those represented in Scheme 7. It is evident that this transformation is energetically more favorable than atropisomerization, in which the half-planes defined by the thiourea linkage and the aromatic ring become coplanar, while either the chlorine atom or the methyl group interact with a bulky sulfur atom.

2.4. Unprotected Ureas

Selective O-deacetylation reactions were attempted using the conventional and mild conditions provided by a saturated solution of ammonia in methanol at room temperature. Compounds 14, 16, and 22 could be successfully deacetylated (Scheme 8) without affecting other functional groups, leading to the corresponding unprotected derivatives 3436, respectively, in high yields (Table S3). In stark contrast, the deacetylation protocol failed with thioureas, giving rise to messy reactions. As portrayed by 15, proton NMR monitoring shows a complex mixture, for which the multiple signals most likely result from the decomposition of the starting material.
This transformation can easily be monitored by TLC, and, after completion, the isolation protocol essentially involves filtration and further purification by recrystallization from ethanol. High-resolution mass spectra (HRMS, CI mode) allowed us to determine the molecular weight and compositions of 3436 (Figures S75–S77). The peak [M+Na]+ was detected in the case of 34 at m/z 346.1122, consistent with the formula C13H17O5N5Na and for 35 at m/z 376.1228, which agrees with C14H19O6N5Na. For urea 36, the [M+Na]+ peaks corresponding to the 35Cl and 37Cl isotopes could be detected, with an approximate intensity ratio 3:1, with the former being detected at m/z 394.0889 with composition C14H18O5N5ClNa.
All spectroscopic data support unequivocally the structures of such compounds as well. The absence of acetate absorptions in the IR spectra at ~1730 cm−1 evidence a complete deacetylation reaction. The typical broad bands of the OH groups and the NH peak appear between 3500 and 3100 cm−1, whereas the alcohol C-O bond lies at 1080 cm−1 (Figure S13). The intense signal of the azido group is preserved at ~2120 cm−1, and that of the urea carbonyl at ~1620 cm−1. Well-resolved 1H NMR spectra could be obtained in all cases, thus facilitating assignment (Table 11) and being corroborated by double-resonance experiments and D2O-induced isotopic exchange conducted on 34 (Figure S47). As expected, the most deshielded signal corresponds to the NH group linked directly to the aromatic ring (~8.50 ppm), while the NH bound to the pyranoid ring appears as a doublet at ~6.2 ppm. The anomeric proton resonates at ~4.5 ppm, and the remaining ring protons now follow the shift order H-6 > H-2, H-6′ > H-4, H-5 > H-3. The high coupling constants J2,3 = J3,4 = J4,5 (~9.5 Hz) indicate that the pyranoid ring maintains a 4C1 conformation with d-gluco configuration, and the high value of J1,2 (~9.0 Hz) evidences that the equatorial (β) orientation of the azido group remains unaffected. 13C NMR data are gathered in Table 12. Together with the absence of resonances for the acetate groups, the urea carbonyl signal is observed at ~160 ppm (Figure S72). The HMQC correlation spectrum enables the assignment of the pyranoid carbons following the order C-1 > C-4 > C-5 > C-3 > C-6 > C-2 (Table 11).

3. Conclusions

The present study repurposes the use and some applications of a per-O-acetylated-2-amino-2-deoxy-glucopyranosyl azide (1), described more than seven decades ago. This chiral synthon can be leveraged to obtain a series of new ureas and thioureas by reaction with aryl isocyanates and isothiocyanates, which can exist as mixtures of conformers in solution, while preserving the pyranoid conformation and the anomeric configuration. The new sugar isothiocyanate 3,4,6-tri-O-acetyl-2-deoxy-2-isothiocyanate-β-d-glucopyranosyl azide (23) has been obtained by a reaction of (1) with thiophosgene. Conformational analysis inferred from NMR data and theoretical calculations evidence a greater stability of the (Z,Z) conformer for ureas and the (Z,E) conformer for the corresponding thioureido derivatives, a fact attributed to the changes induced by carbonyl or thiocarbonyl groups. N-aryl (thio)ureas featuring ortho-disubstitution can also exist as atropisomeric conformers due to an additional bond rotation, for which the low energy barrier could be determined, as illustrated by the thioureido derivative 25 obtained by a reaction of 23 with 2-chloro-6-methylaniline. Azidoureas can easily be deacetylated with ammonia in methanol, affording derivatives with free OH groups and unaltered functionality, a process that failed for thioureas, nevertheless. In summary, the new family of carbohydrate (thio)ureas possessing the azido group at the anomeric position offers new synthetic avenues to be explored. Such manipulations range from expected click-type reactions, already developed for N-acetyl d-glucosamino azides, to the preparation of glycoconjugates and glycomimetics tailored for specific applications [49,81,82,83].

4. Experimental

4.1. General Methods

All reagents and solvents were obtained from commercial suppliers and used without further purification. Melting points were determined on Gallenkamp MPA (York, UK) and Electrothermal IA 9000 (York, UK) apparatuses and are uncorrected. Optical rotations were determined on a Perkin-Elmer 241 polarimeter (Waltham, MA, USA) at 22 ± 2 °C, with sodium (D line, λ = 589 nm) and mercury beams (λ = 578, 546, 436 nm). IR spectra were recorded in the range of 4000–600 cm−1 on an FT-IR Thermo spectrophotometer (Waltham, MA, USA). Solid samples were recorded on KBr (Merck (Darmstadt, Germany)) pellets. NMR spectra were measured on Bruker 400 and 500 AC/PC instruments (Karlsruhe, Germany) in DMSO-d6 or CDCl3. Structural elucidation was facilitated through (a) distortionless enhancement by polarization transfer (DEPT), (b) 2D correlation spectroscopy (COSY), (c) heteronuclear multiple-quantum correlation (HMQC), (d) heteronuclear multiple bond correlation (HMBC), (e) isotope exchange with deuterium oxide, and (f) variable-temperature experiments. All J values are given in hertz. Microanalyses were determined on a Leco® CHNS-932 analyzer (St. Joseph, MI, USA). High-resolution mass spectra (HRMS) were obtained using electrospray ionization (ESI) techniques with a 6520 Accurate-Mass Q-TOF LC/MS system from Agilent Technologies (Santa Clara, CA, USA) at the Servicio de Apoyo a la Investigación (SAIUEX) from the University of Extremadura.

4.2. Computational Details

The computational DFT study was carried out using the hybrid functional M06-2X [72] in conjunction with 6-311G(d,p) and def2-TZVP valence-triple-ζ basis sets [73,74,75], as implemented in the Gaussian09 package [71]. In all cases, frequency analyses were carried out to confirm the existence of true stationary points on the potential energy surface. All thermal corrections were calculated at the standard values of 1 atm at 298.15 K. Solvent effects were modeled through density-based, self-consistent reaction field (SCRF) theory of bulk electrostatics, i.e., the solvation model based on density (SMD) [76]. This solvation method accounts for long-range electrostatic polarization (bulk solvent) together with short-range effects due to cavitation, dispersion, and solvent structural effects.

4.3. Synthetic Procedures

3,4,6-Tri-O-acetyl-2-amino-2-deoxy-β-d-glucopyranosyl azide (1) [47]. To a solution of 12 (2.25 g, 6.1 mmol) in chloroform (20 mL) a suspension of silver azide in chloroform (from 1.0 g of sodium azide) was added. The mixture was vigorously refluxed, filtered, and washed with chloroform. The filtrate was evaporated to dryness and the crude product (1.6 g, 95%) was crystallized from ethanol, affording yellowish needles (1.1 g, 60%), which had m.p. 120–121 °C. IR (KBr) νmax 3380, 3326 (NH2), 2112 (N3), 1748 (acetate), 1237 (C-O-C), 1040 (C-O), 905 cm−1. 1H NMR (400 MHz, CDCl3): δ (ppm) 5.03 (t, 3H, H-3, J 15.2 Hz, J 7.6 Hz), 4.97 (t, 3H, H-4), 4.54 (d, 1H, H-1, J 7.2 Hz), 4.30 (dd, 1H, H-6, J 10.0 Hz, J 4.0 Hz), 4.14 (dd, 1H, H-6a, J 9.6 Hz, J 1.6 Hz), 3.80–3.77 (m, 1H, H-5), 2.83 (t, 1H, H-2, J 14.8 Hz, J 7.2 Hz), 2.10 (s, 3H, OAc), 2.08 (s, 3H, OAc), 2.03 (s,3H, OAc), 1.50 (bs, 2H, NH). 13C{1H} NMR (100 MHz, CDCl3): δ (ppm) 170.6 (2CO), 169.6 (CO), 91.8 (C-1), 75.1 (C-5), 73.9 (C-3), 68.2 (C-4), 61.9 (C-6), 55.8 (C-2), 20.7, 20.6, and 20.5 (OAc).
3,4,6-Tri-O-acetyl-2-deoxy-2-(3-phenylureido)-β-d-glucopyranosyl azide (14). Phenyl isocyanate (0.16 mL, 1.51 mmol) was added under constant stirring to a cooled (ice bath) solution of 1 (0.50 g, 1.51 mmol) in methylene chloride (4 mL). Immediately, a white crystalline solid separated (0.24 g, 97%), which had m.p. 170–171 °C. [α]D −11.9°; [α]578 −11.9°; [α]546 −12.6°; [α]436 −21.9° (c 0.5, chloroform). IR (KBr) ν ¯ max 3349, 3285, 2117, 1752, 1652, 1598, 1566, 1225, 1062, 1039, 735, 666 cm−1. 1H NMR (500 MHz, CDCl3): δ (ppm) 7.44 (s, 1H, NH-Ph), 7.27 (m, 4H, arom.), 7.06 (t, 1H, arom., 13.5 Hz, J 6.5 Hz), 5.70 (d, 1H, NH, J 8.5 Hz), 5.29 (t, 1H, H-3, J 9.5 Hz), 5.09 (t, 1H, H-4), 4.78 (d, 1H, H-1, J 9.5 Hz), 4.26 (dd, 1H, H-6, J 12.0 Hz, J 5.0 Hz), 4.16 (dd, 1H, H-6′, J 2.0 Hz), 3.82 (c, 1H, H-2, J 19.0 Hz), 3.77 (m, 1H, H-5), 2.10 (s, 3H, OAc), 2.04 (s, 3H, OAc), 2.02 (s, 3H, OAc). 13C{1H} NMR (125 MHz, CDCl3): δ (ppm) 170.1, 170.7, 169.3 (CO, acetate), 155.6 (CO, urea), 138.0, 129.3 (2C), 124.2, 121.1 (2C) (arom.), 89.0 (C1), 73.8 (C5), 72.5 (C3), 68.3 (C4), 62.0 (C6), 54.9 (C2), 20.7, (OAc), 20.6 (OAc). Anal. Calculated for C19H23N5O8: C, 50.78; H, 5.16; N, 15.58. Found: C, 50.71; H, 5.08; N 15.74.
3,4,6-Tri-O-acetyl-2-deoxy-2-(3-phenylthioureido)-β-d-glucopyranosyl azide (15). Phenyl isothiocyanate (0.180 mL, 1.51 mmol) was added under constant stirring to a cooled (ice bath) solution of 1 (0.50 g, 1.51 mmol) in methylene chloride (4 mL). Immediately, a white crystalline solid separated (0.20 g, 45%), which showed m.p. 149–150 °C. [α]D −44.9°; [α]578 −46.8°; [α]546 −51.7°; [α]436 −83.0° (c 0.5, chloroform). IR (KBr) ν ¯ max 3337, 3297, 3064, 2112, 1750, 1729, 1541, 1238, 1100, 1086, 1040, 951 cm−1. 1H NMR (500 MHz, CDCl3): δ (ppm) 8.15 (s, 1H, NH), 7.45 (t, 2H, arom, J 7.5 Hz), 7.35 (m, 1H, arom.), 7.21 (d, 2H, arom.), 5.99 (d, 1H, NH, J 9.0 Hz), 5.18 (m, 2H, H-3, and H-4), 4.93 (bs, 1H, H-2), 4.60 (bs, 1H, H-1), 4.25 (dd, 1H, H-6, J 12.5 Hz, 5.0 Hz), 4.16 (dd, 1H, H-6′, J 2.5 Hz), 3.73 (m, 1H, H-5), 2.10 (s, 3H, OAc), 2.09 (s, 3H, OAc), 2.01 (s, 3H, OAc). 13C{1H} NMR (125 MHz, CDCl3): δ (ppm) 181.8 (CS), 171.1 (CO), 170.6 (CO), 169.1 (CO), 135.5, 130.2 (2C), 127.8, 125.6 (2C) (arom), 88.5 (C1), 74.1 (C5), 72.6 (C3), 67.7 (C4), 61.7 (C6), 58.6 (C2), 20.7 (OAc), 20.6 (OAc), 20.5 (OAc). 1H NMR (400 MHz, DMSO-d6): δ (ppm) 9.82 (s, 1H, NH), 7.66 (d, 1H, NH, J 7.6 Hz), 7.36 (t, 2H, arom., J 7.6 Hz), 7.27 (d, 2H, arom., J 8.0 Hz), 7.17 (t, 1H, arom.), 5.30 (bs, 1H, H-3), 4.95 (t, 3H, H-1, H-2, H-4), 4.21 (dd, 1H, H-6, J 12.4 Hz, J 4.8 Hz), 4.10 (dd, 1H, H-6′, J 2.0 Hz), 3.97 (m, 1H, H-5), 2.04 (s, 3H, OAc), 1.99 (s, 3H, OAc), 1.97 (s, 3H, OAc). 13C{1H} NMR (100 MHz, DMSO-d6): δ (ppm) 181.1 (CS), 170.0 (CO), 169.8 (CO), 169.2 (CO), 138.1, 128.8 (2C), 128.2, 123.7 (2C) (arom.), 87.4 (C1), 72.7 (C5), 72.6 (C3), 67.9 (C4), 61.5 (C6), 57.6 (C2), 20.4 (OAc), 20.4 (OAc), 20.3 (OAc). Anal. Calculated for C19H23N5O7S: C, 49.03; H, 4.98; N, 15.05; S, 6.89. Found: C, 49.32; H, 5.02; N, 15.15; S, 7.04.
3,4,6-Tri-O-acetyl-2-deoxy-2-[3-(3-methoxyphenyl)ureido]-β-d-glucopyranosyl azide (16). 3-Methoxyphenyl isocyanate (0.14 mL, 1.1 mmol) was added under constant stirring to a cooled (ice bath) solution of 1 (0.34 g, 1.0 mmol) in methylene chloride (4 mL). The mixture was evaporated, and the residue was treated with ethyl ether and crystallized from benzene, affording a white solid (0.26 g, 52%) that had m.p. 159–160 °C. [α]D −30.5°; [α]578 −31.5°; [α]546 −35.4°; [α]436 −52.4° (c 0.5, chloroform). IR (KBr) ν ¯ max 3322, 3203, 3093, 2121, 1754, 1649, 1610, 1585, 1288, 1223, 1090, 1032 cm−1. 1H NMR (500 MHz, CDCl3): δ (ppm) 7.30 (s, 1H, arom.), 7.21 (t, 1H, arom., J 8.5 Hz), 6.98 (t, 1H, NH, J 2.5 Hz), 6.83 (dd, 1H, arom., J 7.5 Hz, J 1.0 Hz), 6.65 (dd, 1H, arom.), 5.54 (d, 1H, NH, J 9.0 Hz), 5.30 (t, 1H, H-3, J 9.5 Hz), 5.10 (t, 1H, H-4), 4.81 (d, 1H, H-1, J 9.5 Hz), 4.28 (dd, 1H, H-6, J 12.5 Hz, J 5.5 Hz), 4.18 (dd, 1H, H-6′, J 2.5 Hz), 3.85 (t, 1H, H-2, J 9.0 Hz), 3.81 (m, 1H, H-5), 3.78 (s, 3H, CH3), 2.12 (s, 3H, OAc), 2.06 (s, 3H, OAc), 2.04 (s, 3H, OAc). 13C{1H} NMR (125 MHz, CDCl3): δ (ppm) 171.1 (CO), 170.7 (CO), 169.3 (CO), 160.4 (CO), 155.3 (1C), 139.3 (1C), 130.0 (1C), 113.2 (1C), 109.7 (1C), 107.0 (1C) (arom.), 89.0 (C1), 73.8 (C3), 72.6 (C5), 68.3 (C4), 62.0 (C6), 55.2 (C2), 54.8 (CH3), 20.7 (OAc), 20.7 (OAc), 20.5 (OAc). Anal. Calculated for C20H25N5 O9: C, 50.10; H, 5.26; N, 14.61. Found: C, 49.90; H, 5.25; N, 14.57.
3,4,6-Tri-O-acetyl-2-deoxy-2-[3-(3-methoxyphenyl)thioureido]-β-d-glucopyranosyl azide (17). 3-Methoxyphenyl isothiocyanate (0.17 mL, 1.2 mmol) was added under constant stirring to a cooled (ice bath) solution of 1 (0.34 g, 1.0 mmol) in methylene chloride (4 mL). The mixture was evaporated, and the residue was treated with ethyl ether and crystallized from benzene, yielding a white solid (0.32 g, 64%) that showed m.p. 108–109 °C. [α]D −30.5°; [α]578 −31.5°; [α]546 −35.4°; [α]436 −45.5° (c 0.5, chloroform). IR (KBr) ν ¯ max 3325, 3154, 3019, 1752, 1732, 1601, 1549, 1233, 1037 cm−1. 1H NMR (400 MHz, CDCl3): δ (ppm) 8.09 (s, 1H, NH), 7.34 (t, 1H, arom., J 8.4 Hz), 6.87 (d, 1H, arom., J 7.2 Hz), 6.77 (s, 2H, arom.), 6.05 (d, 1H, NH, J 9.2 Hz), 5.19 (m, 2H, H-3, and H-4), 4.98 (bs, 1H, H-2), 4.60 (bs, 1H, H-1), 4.26 (dd, 1H, H-6, J 12.4 Hz, J 4.8 Hz), 4.16 (dd, 1H, H-6′, J 2.4 Hz), 3.83 (s, 3H, CH3), 3.74 (m, 1H, H-5), 2.11 (s, 3H, OAc), 2.10 (s, 3H, OAc), 2.02 (s, 3H, OAc). 13C{1H} NMR (100 MHz, CDCl3): δ (ppm) 181.5 (C=S), 171.0 (CO), 170.7 (CO), 169.1 (CO), 160.9 (1C), 131.0 (1C), 128.3 (1C), 117.3 (2C), 113.6 (1C), 111.0 (C1) (arom.), 74.0 (C5), 72.6 (C3), 67.7 (C4), 61.7 (C6), 58.6 (C2), 55.5 (CH3), 20.7 (OAc), 20.70 (OAc), 20.51 (OAc). Anal. Calculated for C20H25N5 O8S: C, 48.48; H, 5.09; N, 14.13; S, 6.47. Found: C, 48.24; H, 5.21; N, 14.11; S, 6.23.
3,4,6-Tri-O-acetyl-2-deoxy-2-[3-(4-methoxyphenyl)ureido]-β-d-glucopyranosyl azide (18). 4-Methoxyphenyl isocyanate (0.14 mL, 1.2 mmol) was added under constant stirring to a cooled (ice bath) solution of 1 (0.34 g, 1.0 mmol) in methylene chloride (4 mL). A solid precipitated, which was filtered and crystallized subsequently from benzene, affording a white solid (0.22 g, 64%) that showed m.p. 139–140 °C. [α]D −42.7°; [α]578 −43.4°; [α]546 −49.6°; [α]436 −80.3° (c 0.6, chloroform);. IR (KBr) ν ¯ max 3636, 3570, 3343, 3184, 2119, 1732, 1549, 1515, 1227, 1103, 1086 cm−1. 1H NMR (500 MHz, CDCl3): δ (ppm) 7.87 (s, 1H, NH), 7.13 (d, 2H, arom., J 9.0 Hz), 6.96 (d, 2H, arom.), 5.74 (d, 1H, NH, J 7.5 Hz), 5.18 (t, 1H, H-4, J 9.0 Hz), 5.10 (bs, 1H, H-3), 4.93 (bs, 1H, H-2), 4.54 (bs, 1H, H-1), 4.25 (dd, 1H, H-6, J 12.5 Hz, J 4.5 Hz), 4.16 (dd, 1H, H-6′, J 2.5 Hz), 3.83 (s, 3H, CH3), 3.72 (m, 1H, H-5), 2.10 (s, 6H, OAc), 2.01 (s, 3H, OAc). 13C{1H} NMR (125 MHz, CDCl3): δ (ppm) 182.3 (CO), 171.0 (CO), 170.6 (CO), 169.0 (CO), 159.4 (1C), 128.0 (2C), 115.3 (2C) (arom.), 88.5 (C1), 74.1 (C5), 72.6 (C3), 67.7 (C4), 61.7 (C6), 58.6 (C2), 55.5 (CH3), 20.7 (OAc), 20.7 (OAc), 20.5 (OAc). Anal. Calculated for C20H25N5 O9: C, 50.10; H, 5.26; N, 14.61. Found: C, 49.89; H, 5.36; N, 14.57.
3,4,6-Tri-O-acetyl-2-deoxy-2-[3-(4-methoxyphenyl)tioureido]-β-d-glucopyranosyl azide (19). 4-Methoxyphenyl isothiocyanate (0.14 mL, 1.0 mmol) was added under vigorous stirring to a cooled (ice bath) solution of 1 (0.34 g, 1.0 mmol) in methylene chloride (4 mL). The mixture was evaporated, and the residue was treated with ethyl ether. A solid (0.32 g, 64%) was obtained, which crystallized from benzene and had m.p. 142–143 °C. [α]D −45.4°; [α]578 −43.4°; [α]546 −49.6°; [α]436 −80.3° (c 0.5, chloroform). IR (KBr) ν ¯ max 3636, 3380, 3344, 3186, 2112, 1748, 1548, 1238, 1104, 1040 cm−1. 1H NMR (400 MHz, CDCl3): δ (ppm) 7.90 (s, 1H, NH), 7.13 (d, 2H, arom., J 8.8 Hz), 6.95 (d, 2H, arom.), 5.75 (d, 1H, NH, J 7.6 Hz), 5.18 (t, 1H, H-3, J 9.6 Hz), 5.10 (bs, 1H, H-4), 4.93 (bs, 1H, H-2), 4.55 (bs, 1H, H-1), 4.25 (dd, 1H, H-6, J 12.4 Hz, J 4.8 Hz), 4.16 (dd, 1H, H-6′, J 2.4 Hz), 3.83 (s, 3H, CH3), 3.71 (m, 1H, H-5), 2.10 (s, 6H, OAc), 2.01 (s, 3H, OAc). 13C{1H} NMR (100 MHz, CDCl3): δ (ppm) 182.2 (CS), 171.1 (CO), 170.7 (CO), 169.1 (CO), 159.4 (1C), 128.3 (1C), 128.1 (2C), 115.3 (2C) (arom.), 88.5 (C1), 74.1 (C5), 72.6 (C3), 67.6 (C4), 61.7 (C6), 58.6 (C2), 55.5 (CH3), 20.8 (OAc), 20.7 (OAc), 20.5 (OAc). Anal. Calculated for C20H25N5 O8S: C, 48.48; H, 5.09; N, 14.13; S, 6.47. Found: C, 48.13; H, 5.15; N, 14.11; S, 6.27.
3,4,6-Tri-O-acetyl-2-deoxy-2-[3-(2-fluorophenyl)tioureido]-β-d-glucopyranosyl azide (20). 2-Fluorophenyl isothiocyanate (0.13 mL, 1.1 mmol) was added under constant stirring to a cooled (ice bath) solution of 1 (0.50 g, 1.5 mmol) in methylene chloride (4 mL). Immediately, a white crystalline solid precipitated (0.31 g, 69%), which showed m.p. 132–133 °C. [α]D −53.1°; [α]578 −55.4°; [α]546 −63.5°; [α]436 −114.6° (c 0.5, chloroform). IR (KBr) ν ¯ max 3331, 3068, 2962, 2929, 2112, 1751, 1729, 1544, 1241, 1100, 1085, 1040 cm−1. 1H NMR (500 MHz, CDCl3): δ (ppm) 7.96 (s, 1H, NH), 7.50 (bs, 1H, arom.), 7.31 (m, 1H, arom.), 7.19 (m, 2H, arom.), 6.18 (bs, 1H, NH), 5.17 (m, 2H, H-3, and H-4), 4.86 (bs, 1H, H-2), 4.66 (bs, 1H, H-1), 4.26 (dd, 1H, H-6, J 12.5 Hz, J 5.0 Hz), 4.17 (dd, 1H, H-6′, J 2.0 Hz), 3.77 (m, 1H, H-5), 2.10 (s, 3H, OAc), 2.07 (s, 3H, OAc), 2.01 (s, 3H, OAc). 13C{1H} NMR (125 MHz, CDCl3): δ (ppm) 181.6 (CS), 171.6 (CO), 170.7 (CO), 169.2 (CO), 144.6 (1C), 143.2 (1C), 125.2 (2C), 123.1 (2C) (arom.), 88.6 (C1), 74.1 (C5), 72.8 (C3), 67.6 (C4), 61.7 (C6), 58.7 (C2), 20.8 (OAc), 20.7 (OAc), 20.5 (OAc). Anal. Calculated for C19H22N5FO7S: C, 47.20; H, 4.59; N, 14.49; S, 6.63. Found: C, 47.29; H, 4.73; N, 14.42; S, 6.76.
3,4,6-Tri-O-acetyl-2-deoxy-2-[3-(4-nitrophenyl)thioureido]-β-d-glucopyranosyl azide (21). 4-Nitrophenyl isothiocyanate (0.18 g, 1.0 mmol) was added under vigorous stirring to a cooled (ice bath) solution of 1 (0.33 g, 1.0 mmol) in methylene chloride (4 mL). The mixture was evaporated and the residue was treated with ethyl ether and crystallized from benzene, yielding a yellow solid (0.32 g, 64%) that had m.p. 100–101 °C. [α]578 −14.7°; [α]546 −12.8°; [α]436 −47.6° (c 0.5, chloroform). IR (KBr) ν ¯ max 3331, 3073, 2119, 1750, 1598, 1539, 1234, 1111, 1046 cm−1; 1H NMR (500 MHz, CDCl3): δ (ppm) 8.61 (s, 1H, NH), 8.21 (d, 2H, arom., J 9.5 Hz), 7.57 (d, 2H, arom.), 6.64 (d, 1H, NH, J 8.5 Hz), 5.22 (m, 2H, H-3, and H-4), 4.75 (bs, 1H, H-2), 4.31 (dd, 1H, H-6, J 12.5 Hz, J 4.5 Hz), 4.19 (dd, 1H, H-6′, J 2.5 Hz), 3.82 (m, 1H, H-5), 2.12 (s, 3H, OAc), 2.10 (s, 3H, OAc), 2.06 (s, 3H, OAc). 13C{1H} NMR (125 MHz, CDCl3): δ (ppm) 182.5 (CS), 171.4 (CO), 170.7 (CO), 169.1 (CO), 127.7 (1C), 125.0 (1C), 116.8 (2C), 1116.7 (2C) (arom.), 88.6 (C1), 74.0 (C5), 72.5 (C3), 67.8 (C4), 61.7 (C6), 58.7 (C2), 20.7 (OAc), 20.6 (OAc), 20.5 (OAc). HRMS (m/z) [M+H]+: calculated for C19H22N6O9S, 510.1050, found 510.1067.
3,4,6-Tri-O-acetyl-2-deoxy-2-[3-(2-chloro-6-methylphenyl)ureido]-β-d-glucopyranosyl azide (22). 2-Chloro-6-methylphenyl isocyanate (0.07 mL, 0.5 mmol) was added under constant stirring to a cooled (ice bath) solution of 1 (0.16 g, 0.5 mmol) in methylene chloride (4 mL). The mixture was stirred for 24 h, after which a solid (0.13 g, 54%) was obtained, which crystallized from benzene and showed m.p. 220–221 °C. [α]589 −21.8°; [α]578 −23.0°; [α]546 −26.5°; [α]436 −44.3° (c 0.5, chloroform). IR (KBr) ν ¯ max 3301, 2120, 1753, 1646, 1597, 1233, 1070, 1036 cm−1. 1H NMR (500 MHz, DMSO-d6): δ (ppm) 7.97 (s, 1H, NH), 7.31 (d, 1H, arom., J 8.0 Hz), 7.20 (d, 1H, arom., J 6.5 Hz), 7.15 (t, 1H, arom.), 6.38 (bs, 1H, NH), 5.23 (t, 1H, H-3, J 9.5 Hz), 4.94 (d, 1H, H1, J 9.0 Hz), 4.89 (t, 1H, H-4), 4.18 (dd, 1H, H-6, J 12.5 Hz, J 5.0 Hz), 4.08 (dd, 1H, H-6′, J 2.0 Hz), 4.00 (m, 1H, H-5), 3.80 (q, 1H, H-2, J 19.5 Hz, J 10.0 Hz), 2.18 (s, 3H, CH3), 2.03 (s, 3H, OAc), 1.98 (s, 3H, OAc), 1.96 (s, 3H, OAc). 1H NMR (400 MHz, CDCl3): δ (ppm) 7.31 (m, 1H, arom.), 7.17 (m, 1H, arom.), 6.24 (bs, 1H, NH), 5.26 (t, 1H, H-3, J 9.6 Hz), 5.07 (t, 1H, H-4), 4.81 (d, 1H, NH, J 8.8 Hz), 4.75 (bs, 1H, H-1), 4.26 (dd, 1H, H-6, J 12.4 Hz, J 4.8 Hz), 4.15 (dd, 1H, H-6′, J 2.0 Hz), 3.75 (m, 2H, H-2, and H-5), 2.31 (s, 3H, CH3), 2.10 (s, 3H, OAc), 2.04 (s, 3H, OAc), 2.01 (s, 3H, OAc). 13C{1H} NMR (100 MHz, DMSO-d6): δ (ppm) 170.0 (CO), 169.5 (CO), 169.2 (CO), 154.7 (CO), 138.7 (1C), 137.2 (1C), 134.1, 128.9 (2C), 126.7 (2C), 88.0 (C1), 72.6 (C5), 72.3 (C3), 68.3 (C4), 61.7 (C6), 53.4 (C2), 20.4 (OAc), 20.3 (2OAc), 18.1 (CH3). Anal. Calculated for C20H24ClN5 O8: C, 48.25; H, 4.86; N, 14.07; Cl, 7.12. Found: C, 48.10; H, 4.66; N, 14.17.
3,4,6-Tri-O-acetyl-2-deoxy-2-isothiocianato-β-d-glucopyranosyl azide (23). To a vigorously stirred mixture of 1 (0.22 g, 0.67 mmol) in chloroform (5 mL), calcium carbonate (0.17 g, 0.68 mmol) and water (2 mL) was added thiophosgene (0.06 mL, 0.78 mmol). Stirring was maintained at room temperature for 48 h. After this time, the mixture was filtered and extracted with chloroform. The filtrate was washed with water. The organic phase was dried over calcium chloride and evaporated to dryness, yielding an orange residue from which a solid separates when treating with ethyl ether (0.15 g, 58%). Crystallized from benzene it showed m.p. 88–89 °C. [α]589 14.3°; [α]578 15.0°; [α]546 17.9°; [α]436 33.9° (c 0.5, chloroform). IR (KBr) ν ¯ max 2120, 2077, 1750, 1739, 1239, 1218, 1062, 1033 cm−1. 1H NMR (400 MHz, CDCl3): δ (ppm) 5.20 (t, 1H, H-3, J 10.0 Hz), 4.99 (t, 1H, H-4), 4.75 (d, 1H, H-1, J 8.8 Hz), 4.28 (dd,1H, H-6, J 12.8 Hz, J 4.8 Hz), 4.13 (dd, 1H, H-6′, J 2.4 Hz), 3.80 (m, 1H, H-5), 3.74 (dd, 1H, H-2, J 10.4 Hz, J 9.2 Hz), 2.10 (s, 3H, OAc), 2.09 (s, 3H, OAc), 2.03 (s, 3H, OAc). 13C{1H} NMR (100 MHz, CDCl3): δ (ppm) 170.5 (CO), 169.7 (CO), 169.4 (CO), 141.9 (CS), 88.5 (C1), 74.1 (C5), 72.6 (C3), 67.4 (C4), 61.4 (C6), 60.2 (C2), 20.7 (OAc), 20.6 (OAc), 20.5 (OAc). Anal. Calculated for C13H16N4 O7S: C, 41.93; H, 4.33; N, 15.05; S, 8.61. Found: C, 42.01; H, 4.01; N, 15.00; S, 8.41.
3,4,6-Tri-O-acetyl-2-deoxy-2-[3-(2-chloro-6-methylphenyl)thioureido]-β-d-glucopyranosyl azide (25). 2-Chloro-6-methylaniline (0.03 mL, 0.27 mmol) was added under constant stirring to a cooled (ice bath) solution of 23 (0.10 g, 0.27 mmol) in methylene chloride (4 mL). The mixture was kept at room temperature for 72 h, after which it was evaporated to dryness. When the resulting orange residue was treated with ethyl ether/petroleum ether, a solid (0.031 g, 22%) was obtained, which crystallized from benzene and had m.p. 164–165 °C. IR (KBr) ν ¯ max 3331, 2118, 1750, 1730, 1241, 1104, 1088, 1053 cm−1. 1H NMR (500 MHz, CDCl3, 330 K): δ (ppm) 7.51 (bs, 1H, NH), 7.35 (d, 1H, arom., J 7.5 Hz), 7.23 (m, 2H, arom.), 5.51 (bs, 1H, NH), 5.12 (d, 1H, H-1, J 4.0 Hz), 4.76 (bs, 1H, H-3), 4.59 (bs, 1H, H-4), 4.24 (dd, 1H, H-6, J 12.5 Hz, J 5.0 Hz), 4.17 (dd, 1H, H-6′, J 2.5 Hz), 3.73 (bs, 1H, H-5), 2.31 (s, 3H, CH3), 2.08 (s, 3H, OAc), 2.05 (s, 3H, OAc), 1.98 (s, 3H, OAc). 13C{1H} NMR (100 MHz, CDCl3, 223 K): δ (ppm) 180.7 (CS), 171.3 (CO), 171.2 (CO), 169.3 (CO), 139.4 (1C), 133.6 (1C), 130.3 (1C), 130.0 (1C), 129.9 (1C), 127.9 (1C) (arom.), 88.2 (C1), 73.3 (C5), 71.3 (C3), 66.2 (C6), 70.0 (C4), 57.9 (C2), 21.0 (OAc), 20.9 (OAc), 18.3 (OAc). HRMS (m/z) [M+Na]+: calculated for C20H24O7N5ClNaS 536.0965, found 536.0977.
Deacetylation reaction. General procedure. To a solution of the per-O-acetyl ureido derivative (0.25 mmol) in methanol (4 mL) was added a saturated solution of ammonia in methanol. The reaction was monitored by TLC (silica gel plate) using benzene/methanol (9:1) as eluent.
2-Deoxy-2-(3-phenylureido)-β-d-glucopyranosyl azide (34). Starting from 14 and according to the general procedure, the reaction mixture was evaporated to dryness and the crude product (81%) was crystallized from ethanol and had m.p. 205–206 °C. IR (KBr) ν ¯ max 3445, 3351, 3276, 2120, 1619, 1596, 1562, 1237, 1084, 1029 cm−1. 1H NMR (500 MHz, DMSO-d6): δ (ppm) 8.52 (s, 1H, NH), 7.40 (d, 2H, arom., J 8.0 Hz), 7.22 (t, 2H, arom., J 7.5 Hz), 6.90 (t, 1H, arom.), 6.12 (d, 1H, NH, J 9.0 Hz), 5.15 (d, 2H, OH-3, and OH-4, J 5.5 Hz), 4.67 (t, 1H, OH-6, J 6.0 Hz), 4.51 (d, 1H, H-1, J 9.0 Hz), 3.72 (ddd, 1H, H-6, J 12.5 Hz, J 5.5 Hz), 3.50 (ddd, 1H, H-6′), 3.45 (c, 1H, H-2), 3.30 (m, 2H, H-4, and H-5), 3.15 (td, 1H, H-3, J 9.5 Hz, J 5.5 Hz). 13C{1H} NMR (125 MHz, DMSO-d6): δ (ppm) 155.0 (CO), 140.3, 128.6 (2C), 121.0, 117.6 (2C) (arom.), 88.9 (C1), 79.1 and 74.1 (C4, C5), 70.2 (C3), 60.7 (C6), 55.3 (C2). HRMS (m/z) [M+Na]+: calculated for C13H17O5N5Na 346.1114, found 346.1122.
2-Deoxy-2-[3-(4-methoxyphenyl)ureido]-β-d-glucopyranosyl azide (35). Starting from 15 and in agreement with the general procedure, the title compound crystallized from the reaction mixture. It was filtered, washed with ethyl ether (23%), recrystallized from ethanol, and showed m.p. 214–215 °C. IR (KBr) ν ¯ max 3470, 3305, 3108, 2110, 1641, 1597, 1561, 1488, 1240, 1090, 1054 cm−1. 1H NMR (400 MHz, DMSO-d6): δ (ppm) 8.55 (s, 1H, NH), 7.13 (m, 2H, arom.), 6.88 (d, 1H, arom., J 7.6 Hz), 6.48 (dd, 1H, arom., J 8.0 Hz, J 2.4 Hz), 6.11 (d, 1H, NH, J 8.8 Hz), 5.16 (m, 2H, OH-3, and OH-4), 4.69 (t, 1H, OH-6, J 5.6 Hz), 4.51 (d, 1H, H-1, J 9.0 Hz), 3.72 (m, 1H, H-6), 3.70 (s, 3H, CH3), 3.47 (m, 2H, H-2, and H-6′), 3.30 (m, 2H, H-4, and H-5), 3.14 (td, 1H, H-3, J 8.8 Hz, J 5.2 Hz). 13C{1H} NMR (100 MHz, DMSO-d6): δ (ppm) 159.5 (CO), 154.9 (1C), 141.5 (1C), 129.3 (1C), 110.0 (1C), 106.4 (1C), 103.5 (1C) (arom.), 88.8 (C1), 79.1 (C4 ó C5), 74.1 (C4 ó C5), 70.2 (C3), 60.7 (C6), 55.3 (C2), 54.8 (CH3). HRMS (m/z) [M+Na]+: calculated for C14H19O6N5Na 373.1217, found 373.1228.
2-Deoxy-2-[3-(2-chloro-6-methylphenyl)ureido]-β-d-glucopyranosyl azide (36). Starting from 22 and in agreement with the general procedure, the title compound crystallized from the reaction mixture. It was filtered, washed with ethyl ether (79%), recrystallized from ethanol, and showed m.p. 206–207 °C. IR (KBr) ν ¯ max 3300, 2227, 2121, 1638, 1567, 1468, 1455, 1248, 1073, 1035 cm−1. 1H NMR (400 MHz, DMSO-d6): δ (ppm) 7.81 (s, 1H, NH), 7.29 (d, 1H, arom., J 8.0 Hz), 7.18 (d, 1H, arom., J 7.2 Hz), 7.13 (t, 1H, arom.), 6.31 (d, 1H, NH, J 8.4 Hz), 5.13 (d, 1H, OH, J 5.6 Hz), 5.09 (d, 1H, OH), 4.67 (t, 1H, OH-6, J 5.6 Hz), 4.46 (d, 1H, H-1, J 8.8 Hz), 3.72 (dd, 1H, H-6, J 10.4 Hz, J 5.6 Hz), 3.46 (m, 2H, H-6′, and H-2), 3.30 (m, 2H, H-4, and H-5), 3.13 (td, 1H, H-3, J 8.8 Hz, J 5.2 Hz). 13C{1H} NMR (100 MHz, DMSO-d6): δ (ppm) 155.4 (CO), 138.5 (1C), 134.6 (1C), 131.5 (1C), 128.9 (1C), 126.7 (1C), 126.6 (1C) (arom.), 89.2 (C1), 79.2 and 74.1 (C4, C5), 70.4 (C3), 60.8 (C6), 55.8 (C2), 18.4 (CH3). HRMS (m/z) [M+Na]+: calculated for C14H18O5N5ClNa 394.0885, found 394.0889.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/molecules29235687/s1, FTIR and NMR spectra, and computational data mentioned through the main text.

Author Contributions

C.S.-G. and E.M. performed most experiments, spectroscopic studies and computational studies. J.C.P. conceived the conceptual ideas and manuscript outline. J.C.P. and P.C. drafted the manuscript. C.S.-G., E.M., P.C. and J.C.P. reviewed and critically edited the final content. All authors have read and agreed to the published version of the manuscript.

Funding

We thank the Junta de Extremadura and Fondo Europeo de Desarrollo Regional for financial support (Grant GR21039).

Data Availability Statement

All data can be obtained from the authors upon reasonable request. This manuscript is part of Ph.D. theses, by one of us (C.S.-G.) under a Creative Commons Licence, which are available at https://dehesa.unex.es/handle/10662/6467, from institutional repositories. (Accessed on 4 November 2024).

Acknowledgments

We gratefully acknowledge the Servicio de Apoyo a la Investigación (SAIUEX) at the University of Extremadura for analytical and spectroscopic resources, and the computational facilities at the LUSITANIA Supercomputing Centre supported by Cenits and Computaex Foundation. This manuscript is dedicated to the memory of María Dolores Méndez-Cordero.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Brown, M.E.; Hollingsworth, M.D. Stress-induced Domain Reorientation in Urea Inclusion Compounds. Nature 1995, 376, 323–327. [Google Scholar] [CrossRef]
  2. Harris, K.D.M. Meldola Lecture: Understanding the Properties of Urea and Thiourea Inclusion Compounds. Chem. Soc. Rev. 1997, 26, 279–289. [Google Scholar] [CrossRef]
  3. Choi, K.H.; Hamilton, A.D. Macrocyclic Anion Receptors Based on Directed Hydrogen Bonding Interactions. Coord. Chem. Rev. 2003, 240, 101–110. [Google Scholar] [CrossRef]
  4. Castellano, R.K.; Nuckolls, C.; Rebek, J., Jr. Transfer of Chiral Information through Molecular Assembly. J. Am. Chem. Soc. 1999, 121, 11156–11163. [Google Scholar] [CrossRef]
  5. Cho, Y.L.; Rudkevich, D.M.; Shivanyuk, A.; Rissanen, K.; Rebek, J., Jr. Hydrogen Bonding Effects in Calix[4]arene Capsules. Chem. Eur. J. 2000, 6, 3788–3796. [Google Scholar] [CrossRef] [PubMed]
  6. Terech, P.; Weiss, R.G. Low Molecular Mass Gelators of Organic Liquids and the Properties of Their Gels. Chem. Rev. 1997, 97, 3133–3160. [Google Scholar] [CrossRef] [PubMed]
  7. Van Esch, J.; De Feyter, S.; Kellogg, R.M.; De Schryver, S.; Feringa, B.L. Self-Assembly of Bisurea Compounds in Organic Solvents and on Solid Substrates. Chem. Eur. J. 1997, 3, 1238–1243. [Google Scholar] [CrossRef]
  8. Estroff, L.A.; Hamilton, A.D. Effective Gelation of Water Using a Series of Bis-urea Dicarboxylic Acids. Angew. Chem. Int. Ed. Engl. 2000, 39, 3447–3450. [Google Scholar] [CrossRef]
  9. De Loos, M.; Van Esch, J.; Kellogg, R.M.; Feringa, B.L. Chiral Recognition in Bis-Urea-Based Aggregates and Organogels through Cooperative Interactions. Angew. Chem. Int. Ed. 2001, 40, 613–616. [Google Scholar] [CrossRef]
  10. Kölbel, M.; Menger, F.M. Molecular Recognition among Structurally Similar Components of a Self-Assembling Soft Material. Langmuir 2001, 17, 4490–4492. [Google Scholar] [CrossRef]
  11. Beginn, U.; Tartsch, B. Bis[(alkoxy)benzoylsemicarbazides]–A New Class of Powerful Organogelators. Chem. Commun. 2001, 1924–1925. [Google Scholar] [CrossRef] [PubMed]
  12. Tamaru, S.-I.; Uchino, S.-Y.; Takeuchi, M.; Ikeda, M.; Hatano, T.; Shinkai, S. A Porphyrin-Based Gelator Assembly Which Is Reinforced by Peripheral Urea Groups and Chirally Twisted by Chiral Urea Additives. Tetrahedron Lett. 2002, 43, 3751–3755. [Google Scholar] [CrossRef]
  13. Wang, G.; Hamilton, A.D. Low Molecular Weight Organogelators for Water. Chem. Commun. 2003, 310–311. [Google Scholar] [CrossRef] [PubMed]
  14. Babu, P.; Sangeetha, N.M.; Vijayakumar, P.; Mitra, U.; Rissanen, K.; Raju, A.R. Pyrene-Derived Novel One- and Two-Component Organogelators. Chem. Eur. J. 2003, 9, 1922–1932. [Google Scholar] [CrossRef] [PubMed]
  15. George, S.J.; Ajayaghosh, A. Self-Assembled Nanotapes of Oligo(p-phenylene vinylene)s: Sol-Gel-Controlled Optical Properties in Fluorescent π-Electronic Gels. Chem. Eur. J. 2005, 11, 3217–3227. [Google Scholar] [CrossRef]
  16. George, M.; Tan, G.; John, V.T.; Weiss, R.G. Urea and Thiourea Derivatives as Low Molecular-Mass Organogelators. Chem. Eur. J. 2005, 11, 3243–3254. [Google Scholar] [CrossRef]
  17. Ávalos, M.; Babiano, R.; Cintas, P.; Jiménez, J.L.; Palacios, J.C.; Fuentes, J. Synthesis of Acylated Thioureylenedisaccharides. J. Chem. Soc. Perkin Trans. 1 1990, 495–501. [Google Scholar] [CrossRef]
  18. Ávalos, M.; Fuentes, J.; Gómez, I.M.; Jiménez, J.L.; Palacios, J.C.; Ortiz, M.C. Synthesis of 1,3,4,6-Tetra-O-acetyl-2-[3-alkyl-(aryl)-thioreido]-2-deoxy-α-d-glucopyranoses and their Transformation Into 2-Alkyl(aryl)amino-(1,2-dideoxy-α-d-glucopyrano)[2,1-d]-2-thiazolines. Carbohydr. Res. 1986, 154, 49–62. [Google Scholar] [CrossRef]
  19. Spanu, P.; Ulgheri, F. Synthetic Approaches to Carbohydrate-Based Ureas. Curr. Org. Chem. 2008, 12, 1071–1092. [Google Scholar] [CrossRef]
  20. Omoto, S.; Shomura, T.; Suzuki, H.; Inouye, S. Studies on Actinomycetales Producing Antibiotics Only on Agar Culture. J. Antibiot. 1979, 32, 436–441. [Google Scholar] [CrossRef]
  21. Ichimura, M.; Koguchi, T.; Yasuzawa, T.; Tomita, F. CV-1, A New Antibiotic Produced by a Strain of Streptomyces Sp. I. Fermentation, Isolation and Biological Properties of the Antibiotic. J. Antibiot. 1987, 40, 723–726. [Google Scholar] [CrossRef] [PubMed]
  22. Yasuzawa, T.; Yoshida, M.; Ichimura, M.; Shirahata, K.; Sano, H. CV-1, A New Antibiotic Produced by a Strain of Streptomyces Sp. II. Structure Determination. J. Antibiot. 1987, 40, 727–731. [Google Scholar] [CrossRef] [PubMed]
  23. Dobashi, K.; Nagaoka, K.; Watanabe, Y.; Nishida, M.; Hamada, M.; Naganawa, H.; Takita, T.; Takeuchi, T.; Umezawa, H. Glycocinnasperimicin D, a New Member of Glycocinnamoylspermidine Antibiotic. J. Antibiot. 1985, 38, 1166–1170. [Google Scholar] [CrossRef] [PubMed]
  24. McKay, M.J.; Nguyen, H.M. Recent Developments in Glycosyl Urea Synthesis. Carbohydr. Res. 2014, 385, 18–44. [Google Scholar] [CrossRef]
  25. Tewari, N.; Tiwari, V.K.; Misrha, R.C.; Tripathi, R.P.; Sristava, A.K.; Ahmad, R.; Srivastava, R.; Srivastava, B.S. Synthesis and Bioevaluation of Glycosyl Ureas as A-Glucosidase Inhibitors and Their Effect on Mycobacterium. Biorg. Med. Chem. 2003, 11, 2911–2922. [Google Scholar] [CrossRef]
  26. Herr, R.R.; Jahnke, H.K.; Argoudelis, A.D. The Structure of Streptozotocin. J. Am. Chem. Soc. 1967, 89, 4808–4809. [Google Scholar] [CrossRef]
  27. Wiley, P.F.; McMichael, D.L.; Koert, J.M.; Wiley, V.H. Analogs and Conversion Products of Streptozotocin. J. Antibiot. 1976, 29, 1218–1225. [Google Scholar] [CrossRef]
  28. Hammer, C.F.; Loranger, R.A.; Schein, P.S. Structures of the Decomposition Products of Chlorozotocin: New Intramolecular Carbamates of 2-Amino-2-deoxyhexoses. J. Org. Chem. 1981, 46, 1521–1531. [Google Scholar] [CrossRef]
  29. Gnewuch, C.T.; Sosnovsky, G. A Critical Appraisal of the Evolution of N-Nitrosoureas as Anticancer Drugs. Chem. Rev. 1997, 97, 829–1014. [Google Scholar] [CrossRef]
  30. Silverman, R.B. The Organic Chemistry of Drug Design and Drug Action; Academic Press: San Diego, CA, USA, 1992; pp. 251–254. [Google Scholar]
  31. Schoorl, M.N. Les uréides (carbamides) des sucres. Recl. Trav. Chim. Pays-Bas 1903, 22, 31–77. [Google Scholar] [CrossRef]
  32. Benn, M.H.; Jones, A.S. 761. Glycosylureas. Part I. Preparation and Some Reactions of d-Glucosylureas and d-Ribosylureas. J. Chem. Soc. 1960, 3837–3841. [Google Scholar] [CrossRef]
  33. Jones, A.S.; Ross, G.W. The Structure of d-Glucosylureas. Tetrahedron 1962, 18, 189–193. [Google Scholar] [CrossRef]
  34. Plusquellec, D.; Roulleau, F.; Brown, E. Chimie des Sucres sans Groupements Protecteurs: Synthese de Carbamates, D’urees Et de Thiourees en Position 1 du Lactose. Tetrahedron Lett. 1984, 25, 1901–1904. [Google Scholar] [CrossRef]
  35. Goodman, I. Glycosyl Ureides. Adv. Carbohydr. Chem. 1958, 13, 215–236. [Google Scholar]
  36. Ichikawa, Y.; Matsukawa, Y.; Nishiyama, T.; Isobe, M. Synthesis, Characterization, and Reaction of Crystalline Fischer’s Glucopyranosyl Isocyanate. Eur. J. Org. Chem. 2004, 2004, 586–591. [Google Scholar] [CrossRef]
  37. Ávalos, M.; Babiano, R.; Cintas, P.; Jiménez, J.L.; Palacios, J.C.; Valencia, C. The Reaction of 2-Amino-2-deoxyhexopyranoses with Isocyanates. Synthesis of Ureas and Their Transformation into Heterocyclic Derivatives. Tetrahedron 1993, 49, 2655–2675. [Google Scholar] [CrossRef]
  38. Ávalos, M.; Babiano, R.; Cintas, P.; Jiménez, J.L.; Palacios, J.C.; Valencia, C. On the Mechanism of Formation of Glycofurano[2,1-D]-imidazolidin-2-ones. Reaction of 2-Amino-2-deoxyheptopyranoses with isocyanates. Tetrahedron 1993, 49, 2676–2690. [Google Scholar] [CrossRef]
  39. Maya, I.; López, O.; Maza, S.; Fernández-Bolaños, J.G.; Fuentes, J. A Facile Access to Ureido Sugars. Synthesis of Urea-Bridged Β-Cyclodextrins. Tetrahedron Lett. 2003, 44, 8539–8543. [Google Scholar] [CrossRef]
  40. Knapp, S.; Jaramillo, C.; Freeman, B. An Ezomycin Model Glycosylation. J. Org. Chem. 1994, 59, 4800–4804. [Google Scholar] [CrossRef]
  41. Liu, Q.; Luedtke, N.W.; Tor, Y. A Simple Conversion of Amines into Monosubstituted Ureas in Organic and Aqueous Solvents. Tetrahedron Lett. 2001, 42, 1445–1447. [Google Scholar] [CrossRef]
  42. Piekarska-Bartoszewicz, B.; Temeriusz, A. A New Method for the Syntesis of Ureido Sugars. Carbohydr. Res. 1990, 203, 302–307. [Google Scholar] [CrossRef]
  43. Ichikawa, Y.; Nishiyama, T.; Isobe, M. Stereospecific Synthesis of Urea-Tethered Neoglycoconjugates Starting from Glucopyranosyl Carbamates. Tetrahedron 2004, 60, 2621–2627. [Google Scholar] [CrossRef]
  44. Kóvacs, J.; Pintér, I.; Messmer, A.; Töth, G.; Duddeck, H. A New Route to Cyclic Urea Derivatives of Sugars Via Phosphinimines. Carbohydr. Res. 1987, 166, 101–111. [Google Scholar] [CrossRef]
  45. Pérez, V.M.D.; Mellet, C.O.; Fuentes, J.; Fernández, J.M.G. Synthesis of Glycosyl(Thio)Ureido Sugars Via Carbodiimides and their Conformational Behaviour in Water. Carbohydr. Res. 2000, 326, 161–175. [Google Scholar] [CrossRef] [PubMed]
  46. Ukita, C.; Hamada, A.; Yoshida, M. Synthesis of 1-(β-d-Ribofuranosyl) Urea and Related Compounds. Chem. Pharm. Bull. 1964, 12, 454–459. [Google Scholar] [CrossRef]
  47. Bertho, A.; Révész, A. Über 1-β,2-Diamino-3,4,6-triacetyl-d-glucose. Justus Liebigs Ann. Chem. 1953, 581, 161–167. [Google Scholar] [CrossRef]
  48. Bolton, C.H.; Hough, L.; Khan, M.Y. Some Further Studies on the Synthesis of Glycopeptide Derivatives: 2-Acetamido-2-deoxy-Β-d-glucopyranosylamine Derivatives. Biochem. J. 1966, 101, 184–190. [Google Scholar] [CrossRef]
  49. Ogunsina, M.; Pan, H.; Samadder, P.; Arthur, G.; Schweizer, F. Structure Activity Relationships of N-Linked and Diglycosylated Glucosamine-Based Antitumor Glycerolipids. Molecules 2013, 18, 15288–15304. [Google Scholar] [CrossRef]
  50. Ghirardello, M.; Ledru, M.H.; Sau, A.; Galán, M.C. Chemo-Selective Rh-Catalysed Hydrogenation of Azides into Amines. Carbohydr. Res. 2020, 489, 107948. [Google Scholar] [CrossRef]
  51. Meldal, M.; Tornøe, C.V. Cu-Catalyzed Azide-Alkyne Cycloaddition. Chem. Rev. 2008, 108, 2952–3015. [Google Scholar] [CrossRef]
  52. He, X.-P.; Zeng, Y.-L.; Zang, Y.; Li, J.; Field, R.A.; Chen, G.-R. Carbohydrate CuAAC Click Chemistry for Therapy and Diagnosis. Carbohydr. Res. 2016, 429, 1–22. [Google Scholar] [CrossRef] [PubMed]
  53. Thakur, K.; Khare, N.K. Synthesis of Glycoconjugate Mimics by ‘Click Chemistry’. Carbohydr. Res. 2019, 484, 107775. [Google Scholar] [CrossRef]
  54. Sutcharitruk, W.; Sirion, U.; Saeeng, R. One-Pot Synthesis of Substituted-Amino Triazole-Glycosides. Carbohydr. Res. 2019, 484, 107780. [Google Scholar] [CrossRef] [PubMed]
  55. Jochims, J.C.; Seeliger, A. Isocyanato- und Isothiocyanato-Derivate des d-Glucosamins. Tetrahedron 1965, 21, 2611–2616. [Google Scholar] [CrossRef]
  56. Nowick, J.S.; Powell, N.A.; Nguyen, T.M.; Noronha, G. An Improved Method for the Synthesis of Enantiomerically Pure Amino Acid Ester Isocyanates. J. Org. Chem. 1992, 57, 7364–7366. [Google Scholar] [CrossRef]
  57. Ávalos, M.; Babiano, R.; Cintas, P.; Hursthouse, M.B.; Jiménez, J.L.; Light, M.E.; Palacios, J.C.; Pérez, E.M.S. Synthesis of Sugar Isocyanates and Their Application to the Formation of Ureido-Linked Disaccharides. Eur. J. Org. Chem. 2006, 2006, 657–671. [Google Scholar] [CrossRef]
  58. Smith, M.B.; March, J. Advanced Organic Chemistry, 5th ed.; John Wiley & Sons, Inc.: New York, NY, USA, 2001. [Google Scholar]
  59. Nakanishi, K.; Solomon, P. Infrared Absorption Spectroscopy; Holden-Day: San Francisco, CA, USA, 1977; pp. 23, 40–43. [Google Scholar]
  60. Ávalos, M.; Babiano, R.; Cintas, P.; Hursthouse, M.B.; Jiménez, J.L.; Light, M.E.; Palacios, J.C.; Silvero, G. Non-Biaryl Atropisomers Derived from Carbohydrates. Part 3: Rotational Isomerism of Sterically Hindered Heteroaryl Imidazolidine-2-ones and 2-Thiones. Tetrahedron 2005, 61, 7931–7944. [Google Scholar] [CrossRef]
  61. Ávalos, M.; Babiano, R.; Cintas, P.; Hursthouse, M.B.; Jiménez, J.L.; Light, M.E.; Palacios, J.C.; Silvero, G. Non-Biaryl Atropisomers Derived from Carbohydrates. Part 4: Absolute Stereochemistry of Carbohydrate-Based Imidazolidine-2-ones and 2-Thiones with Axial and Central Chirality. Tetrahedron 2005, 61, 7945–7959. [Google Scholar] [CrossRef]
  62. Eliel, E.L.; Wilen, S.H. Stereochemistry of Organic Compounds; John Wiley & Sons, Inc.: New York, NY, USA, 1994; p. 503. [Google Scholar]
  63. Abraham, R.J.; Loftus, P. Proton and Carbon-13 NMR Spectroscopy; Heyden: London, UK, 1981; p. 165. [Google Scholar]
  64. Williams, D.H.; Fleming, I. Spectroscopic Methods in Organic Chemistry; McGraw-Hill: New York, NY, USA, 1984; p. 103. [Google Scholar]
  65. Parr, R.G.; Yang, W. Density-Functional Theory of Atoms and Molecules; Oxford University Press: Oxford, UK, 1989. [Google Scholar]
  66. Andzelm, J.; Wimmer, E. Density Functional Gaussian-Type-Orbital Approach to Molecular Geometries, Vibrations, and Reaction Energies. J. Chem. Phys. 1992, 96, 1280–1303. [Google Scholar] [CrossRef]
  67. Becke, A.D. Density-Functional Thermochemistry. I. The Effect of the Exchange-Only Gradient Correction. J. Chem. Phys. 1992, 96, 2155–2160. [Google Scholar] [CrossRef]
  68. Gill, P.M.W.; Johnson, B.G.; Pople, J.A.; Frisch, M.J. The Performance of the Becke-Lee-Yang-Parr (B-LYP) Density Functional Theory with Various Basis Sets. J. Chem. Phys. Lett. 1992, 197, 499–505. [Google Scholar] [CrossRef]
  69. Sosa, C.; Lee, C. Density Functional Description of Transition Structures Using Nonlocal Corrections. Silylene Insertion Reactions into the Hydrogen Molecule. J. Chem. Phys. 1993, 98, 8004–8011. [Google Scholar] [CrossRef]
  70. Stephens, P.J.; Devlin, F.J.; Frisch, M.J.; Chabalowski, C.F. Ab Initio Calculation of Vibrational Absorption and Circular Dichroism Spectra Using Density Functional Force Fields. J. Phys. Chem. 1994, 98, 11623–11627. [Google Scholar] [CrossRef]
  71. 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.; et al. Gaussian 09; Revision A.1; Gaussian, Inc.: Wallingford, CT, USA, 2009. [Google Scholar]
  72. Zhao, Y.; Truhlar, D.G. The M06 Suite of Density Functionals for Main Group Thermochemistry, Thermochemical Kinetics, Noncovalent Interactions, Excited States, and Transition Elements: Two New Functionals and Systematic Testing of Four M06-Class Functionals and 12 Other Functionals. Theor. Chem. Acc. 2008, 120, 215–241. [Google Scholar]
  73. McLean, A.D.; Chandler, G.S. Contracted Gaussian Basis Sets for Molecular Calculations. I. Second Row Atoms, Z = 11–18. J. Chem. Phys. 1980, 72, 5639–5648. [Google Scholar] [CrossRef]
  74. Krishnan, R.; Binkley, J.S.; Seeger, R.; Pople, J.A. Self-Consistent Molecular Orbital Methods. XX. A Basis Set for Correlated Wave Functions. J. Chem. Phys. 1980, 72, 650–654. [Google Scholar] [CrossRef]
  75. Weigend, F.; Ahlrichs, R. Balanced Basis Sets of Split Valence, Triple Zeta Valence and Quadruple Zeta Valence Quality for H to Rn: Design and Assessment of Accuracy. Phys. Chem. Chem. Phys. 2005, 7, 3297–3305. [Google Scholar] [CrossRef]
  76. Marenich, A.V.; Cramer, C.J.; Truhlar, D.G. Universal Solvation Model Based on Solute Electron Density and on a Continuum Model of the Solvent Defined by the Bulk Dielectric Constant and Atomic Surface Tensions. J. Phys. Chem. B 2009, 113, 6378–6396. [Google Scholar] [CrossRef]
  77. Bondi, A. Van Der Waals Volumes And Radii. J. Phys. Chem. 1964, 68, 441–451. [Google Scholar] [CrossRef]
  78. Csonka, G.I.; French, A.D.; Johnson, G.P.; Stortz, C.A. Evaluation of Density Functionals and Basis Sets for Carbohydrates. J. Chem. Theory Comput. 2009, 5, 679–692. [Google Scholar] [CrossRef]
  79. St. John, P.C.; Kim, Y.; Kim, S.; Paton, R.S. Prediction of Organic Homolytic Bond Dissociation Enthalpies and Near Chemical Accuracy with Sub-Second Computational Cost. Nat. Commun. 2020, 11, 2328. [Google Scholar] [CrossRef] [PubMed]
  80. Turner, J.A.; Adrianov, T.; Zakaria, M.A.; Taylor, M.S. Effects of Configuration and Substitution on C−H Bond Dissociation Enthalpies in Carbohydrate Derivatives: A Systematic Computational Study. J. Org. Chem. 2022, 87, 1421–1433. [Google Scholar] [CrossRef] [PubMed]
  81. Garifullin, B.F.; Khabibulina, L.R.; Belenok, M.G.; Saifina, L.F.; Zarubaev, V.V.; Slita, A.V.; Volobueva, A.S.; Semenov, V.E.; Kataev, V.E. Synthesis and Antiviral Activity of 1,2,3-Triazolyl Nucleoside Analogues with N-Acetyl-d-Glucosamine Residue. Nucleosides Nucleotides Nucleic Acids 2023, 42, 743–765. [Google Scholar] [CrossRef] [PubMed]
  82. Yule, L.R.; Bower, R.L.; Kaur, H.; Kowalczyk, R.; Hay, D.L.; Brimble, M.A. Synthesis and Amylin Receptor Activity of Glycomimetics of Pramlintide Using Click Chemistry. Org. Biomol. Chem. 2016, 14, 5238–5245. [Google Scholar] [CrossRef]
  83. Ortega-Muñoz, M.; Perez-Balderas, F.; Morales-Sanfrutos, J.; Hernandez-Mateo, F.; Isac-García, J.; Santoyo-Gonzalez, F. Click Multivalent Heterogeneous Neoglycoconjugates-Modular Synthesis and Evaluation of Their Binding Affinities. Eur. J. Org. Chem. 2009, 2009, 2454–2473. [Google Scholar] [CrossRef]
Chart 1. Structures of azido precursor 1 and iso(thio)cyanates 210 used throughout this work.
Chart 1. Structures of azido precursor 1 and iso(thio)cyanates 210 used throughout this work.
Molecules 29 05687 ch001
Scheme 1. Two-step preparation of 3,4,6-Tri-O-acetyl-2-amino-2-deoxy-β-d-glucopyranosyl azide.
Scheme 1. Two-step preparation of 3,4,6-Tri-O-acetyl-2-amino-2-deoxy-β-d-glucopyranosyl azide.
Molecules 29 05687 sch001
Scheme 2. Direct preparation of ureido or thioureido glucopyranosyl azides.
Scheme 2. Direct preparation of ureido or thioureido glucopyranosyl azides.
Molecules 29 05687 sch002
Scheme 3. General route towards 2-isothiocyanato-β-d-glucopyranosyl azide (23) and coupling with anilines.
Scheme 3. General route towards 2-isothiocyanato-β-d-glucopyranosyl azide (23) and coupling with anilines.
Molecules 29 05687 sch003
Figure 1. Experimental relative disposition in solution of urea/tiourea and sugar ring fragments.
Figure 1. Experimental relative disposition in solution of urea/tiourea and sugar ring fragments.
Molecules 29 05687 g001
Scheme 4. Atropisomeric conformers derived from o,o’-disubstituted N-aryl ureas.
Scheme 4. Atropisomeric conformers derived from o,o’-disubstituted N-aryl ureas.
Molecules 29 05687 sch004
Figure 2. Conformational space for (thio)ureido derivatives showing their potential interconversions.
Figure 2. Conformational space for (thio)ureido derivatives showing their potential interconversions.
Molecules 29 05687 g002
Scheme 5. Optimized structures for the conformational rotamers of azido urea 14 (color code: carbon, gray; hydrogen, white; oxygen, red and nitrogen, blue).
Scheme 5. Optimized structures for the conformational rotamers of azido urea 14 (color code: carbon, gray; hydrogen, white; oxygen, red and nitrogen, blue).
Molecules 29 05687 sch005
Scheme 6. Optimized structures for the conformational rotamers of azido thiourea 15 (color code: carbon, gray; hydrogen, white; oxygen, red; nitrogen, blue and sulfur, yellow).
Scheme 6. Optimized structures for the conformational rotamers of azido thiourea 15 (color code: carbon, gray; hydrogen, white; oxygen, red; nitrogen, blue and sulfur, yellow).
Molecules 29 05687 sch006
Figure 3. Atom numbering for the calculated structure of 14 (Z,Z conformer) (color code: carbon, gray; hydrogen, white; oxygen, red and nitrogen, blue).
Figure 3. Atom numbering for the calculated structure of 14 (Z,Z conformer) (color code: carbon, gray; hydrogen, white; oxygen, red and nitrogen, blue).
Molecules 29 05687 g003
Figure 4. Atom numbering for the calculated structure of 15 (Z,E conformer) (color code: carbon, gray; hydrogen, white; oxygen, red; nitrogen, blue and sulfur, yellow).
Figure 4. Atom numbering for the calculated structure of 15 (Z,E conformer) (color code: carbon, gray; hydrogen, white; oxygen, red; nitrogen, blue and sulfur, yellow).
Molecules 29 05687 g004
Figure 5. Conformational profiles of the aryl moiety of 25 at the M06-2X/6-311G(d,p) level in the gas phase.
Figure 5. Conformational profiles of the aryl moiety of 25 at the M06-2X/6-311G(d,p) level in the gas phase.
Molecules 29 05687 g005
Figure 6. Optimized structures for the two conformational minima of 25 and the transition structure (‡) at the M06-2X/6-311G(d,p) level.
Figure 6. Optimized structures for the two conformational minima of 25 and the transition structure (‡) at the M06-2X/6-311G(d,p) level.
Molecules 29 05687 g006
Chart 2. Acyclic and cyclic atropisomeric (thio)ureido derivatives.
Chart 2. Acyclic and cyclic atropisomeric (thio)ureido derivatives.
Molecules 29 05687 ch002
Scheme 7. Conformational profile of the M/P interconversion for thiourea 33.
Scheme 7. Conformational profile of the M/P interconversion for thiourea 33.
Molecules 29 05687 sch007
Figure 7. Conformational profiles for the M/P interconversion of 33 at the M06-2X/6-311G(d,p) level in the gas phase.
Figure 7. Conformational profiles for the M/P interconversion of 33 at the M06-2X/6-311G(d,p) level in the gas phase.
Molecules 29 05687 g007
Figure 8. Optimized structures for the stationary points related to the rotational evolution in compound 33 and the transition structure (‡).
Figure 8. Optimized structures for the stationary points related to the rotational evolution in compound 33 and the transition structure (‡).
Molecules 29 05687 g008
Scheme 8. Deacetylation of glucosyl azidoureas.
Scheme 8. Deacetylation of glucosyl azidoureas.
Molecules 29 05687 sch008
Table 1. 1H NMR data (δ, ppm) for the aminosugar moiety of 1, 1423, and 25.
Table 1. 1H NMR data (δ, ppm) for the aminosugar moiety of 1, 1423, and 25.
CompoundNHArNHH-1H-2H-3H-4H-5H-6H-6′
1 b 1.50 s4.54 d2.83 t5.03 t4.97 t3.78 ddd4.30 dd4.14 dd
14 a7.44 s5.70 d4.78 d3.82 q5.29 t5.09 t3.76 m4.26 dd4.15 dd
15 a8.15 s5.99 d4.60 bs4.93 bs5.18 m3.73 m4.25 dd4.16 dd
16 a6.98 t5.54 d4.81 d3.85 t5.30 t5.10 t3.81 m4.28 dd4.18 dd
17 b8.09 s6.05 d4.60 bs4.98 m5.19 m3.74 m4.26 dd4.16 dd
18 a7.87 s5.74 d4.54 bs4.93 bs5.10 bs5.18 t3.72 m4.25 dd4.15 dd
19 b7.90 s5.75 d4.55 bs4.93 bs5.18 t5.10 bs3.71 m4.25 dd4.16 dd
20 a7.96 s6.18 bs4.66 bs4.86 bs5.17 m3.77 m4.26 dd4.17 dd
21 a8.61 s6.64 dd4.75 bs5.22 m3.82 m4.31 dd4.19 dd
22 c7.97 s6.38 bs4.94 d3.80 q5.23 t4.89 t4.00 m4.18 dd4.08 dd
23 b 4.75 d3.74 d5.20 t5.00 t3.80 m4.28 dd4.13 dd
25 a,d7.51 s5.51 bs5.12 d---4.76 bs4.59 bs3.73 bs4.24 dd4.17 dd
a In CDCl3 at 500 MHz. b In CDCl3 at 400 MHz. c In DMSO-d6 at 500 MHz. d At 330 K.
Table 2. 13C NMR data (δ, ppm) of 1, 1423 and 25.
Table 2. 13C NMR data (δ, ppm) of 1, 1423 and 25.
CompoundCO dCSC-1C-2C-3C-4C-5C-6
1 b170.6---91.855.873.968.275.161.9
14 a155.6---89.054.972.668.373.862.0
15 a---181.288.558.672.667.774.161.7
16 a155.3---89.054.872.668.373.862.0
17 b---181.588.958.672.667.774.062.0
18 a159.4---88.558.672.767.774.161.7
19 b---182.288.558.672.667.774.161.7
20 a---181.688.658.872.967.674.161.7
21 a---182.588.658.772.567.874.061.7
22 c154.7---88.053.472.368.372.662.0
23 b---141.988.560.272.667.474.161.4
25 b,e---180.7088.257.971.361.073.367.2
a In CDCl3 at 125 MHz. b In CDCl3 at 100 MHz. c In DMSO-d6 at 100 MHz. d Urea. e At 223 K.
Table 3. NMR coupling constants (Hz) for 1 and 1422.
Table 3. NMR coupling constants (Hz) for 1 and 1422.
CompoundJNH,H-2J1,2J2,3J3,4J4,5J5,6J5,6′J6,6′
1 a---9.09.59.59.55.02.012.0
14 a8.59.59.59.510.05.01.512.0
15 a9.0---------9.05.02.512.5
16 a9.09.59.59.510.05.52.512.5
17 b11.5---------9.06.03.015.5
18 a7.5------9.010.04.52.512.5
19 b7.6------9.69.64.82.412.4
20 a------------9.55.02.012.5
21 a------------9.54.52.512.5
22 c---9.510.09.510.05.02.512.5
a In CDCl3 at 500 MHz. b In CDCl3 at 400 MHz. c In DMSO-d6 at 500 MHz.
Table 4. Relative electronic and Gibbs free energies calculated for conformers of compounds 14 and 15 a.
Table 4. Relative electronic and Gibbs free energies calculated for conformers of compounds 14 and 15 a.
1415
ConformerΔE dΔG dΔE eΔG eΔE dΔG dΔE eΔG e
Z,Z b0.00.00.00.03.93.23.02.8
Z,E b0.60.11.23.20.00.00.00.0
E,Z b3.84.73.35.95.65.43.53.3
Z,Z c0.00.00.00.03.92.52.32.1
Z,E c0.61.01.10.90.00.00.00.0
E,Z c4.34.83.43.76.45.14.43.9
a In kcal/mol. b M062X/6-311G(d,p). c M062X/def2-TZVP. d In the gas phase. e In CHCl3.
Table 5. Selected bond distances (Å) and dihedral angles (°) for conformers of 14.
Table 5. Selected bond distances (Å) and dihedral angles (°) for conformers of 14.
(Z,Z) a(Z,E) a(E,Z) a(Z,Z) b(Z,E) b(E,Z) b
O38—H482.19---2.202.21---2.20
N39—N18—C37—O38−8.59−7.67----8.23−7.58---
O38—C37—C44—C454.70---9.365.25---9.36
O38—C37—N43—C44−0.74−179.556.43−1.40173.76−2.75
H42—N41—C2—H10133.32137.16164.24136.67139.35162.07
a At the M062X/6-311G(d,p) level. b At the M062X/def2-TZVP level.
Table 6. Selected bond distances (Å) and dihedral angles (°) for conformers of 15.
Table 6. Selected bond distances (Å) and dihedral angles (°) for conformers of 15.
(Z,Z) a(Z,E) a(E,Z) a(Z,Z) b(Z,E) b(E,Z) b
S55—H472.82---2.962.78---2.86
N38—N18—C37—S55−8.14−1.81----7.08-1.06---
S55—C37—C43—C4440.64---53.4739.98---50.74
S55—C37—N42—C43−1.17−159.9513.860.31−164.0216.66
H41—N40—C2—H10145.29159.38160.72147.66162.65160.44
a At the M062X/6-311G(d,p) level. b At the M062X/def2-TZVP level.
Table 7. Relative energy minima (in kcal/mol) and their dihedral angles (in °) found for 25 a.
Table 7. Relative energy minima (in kcal/mol) and their dihedral angles (in °) found for 25 a.
Gas PhaseChloroform
Entry ΔE bΔG bθN-C(S)-N-CaromθN-C(S)-N-C2ΔE cΔG cθN-C(S)-N-CaromθN-C(S)-N-C2
1min 12.10.0178.4161.02.91.9187.7176.5
2min 20.0−0.1357.1176.20.00.0355.8176.9
a At the M062X/6-311G(d,p). b In the gas phase. c In CHCl3.
Table 8. Relative energy minima (in kcal/mol) and their dihedral angles (in °) found for 25 a.
Table 8. Relative energy minima (in kcal/mol) and their dihedral angles (in °) found for 25 a.
Gas PhaseChloroform
Entry ΔE bΔG bθN-C(S)-N-CaromθN-C(S)-N-C2ΔE cΔG cθN-C(S)-N-CaromθN-C(S)-N-C2
1min 10.6−1.0180.8164.21.51.6187.0176.3
2min 20.00.0358.1176.70.00.0356.7177.7
a At the M062X/def2-TZVP level. b In the gas phase. c In CHCl3.
Table 9. Relative energy maxima (in kcal/mol) and their dihedral angles (in °) found for 25 a.
Table 9. Relative energy maxima (in kcal/mol) and their dihedral angles (in °) found for 25 a.
∆E∆GθN-C(S)-N-CaromθN-C(S)-N-C2‡ b
max11.212.2296.9165.2−32.49
a At the M062X/6-311G(d,p) level in the gas phase. b Imaginary frequencies for the transition structure (‡) in cm−1.
Table 10. Stationary points for the rotational interconversion of thiourea 33 a.
Table 10. Stationary points for the rotational interconversion of thiourea 33 a.
Gas PhaseChloroform
Entry ΔE bΔG bθC2-C1-N-C(S) c‡ dΔE bΔG bθC2-C1-N-C(S) c‡ d
1min 10.000.0070.90 0.000.0074.95
2min 23.053.85279.66 2.381.98273.72
3max 111.6513.64206.91−66.1912.3813.78205.32−69.18
4max 210.6113.46331.35−73.5311.8814.29331.15−70.93
a At the M062X/6-311G(d,p) in the gas phase. b In kcal/mol. c In degrees. d Imaginary frequencies for the transition structure (‡) in cm−1.
Table 11. 1H NMR data (δ, ppm) of 3436 a.
Table 11. 1H NMR data (δ, ppm) of 3436 a.
CompoundNHArNHH-1H-2H-3H-4/H-5H-6H-6′
34 b8.52 s6.12 d4.51 d3.45 c3.15 td3.30 m3.72 dd3.50 dd
35 a8.55 s6.11 d4.51 d3.47 m3.14 td3.30 m3.72 m3.47 m
36 a7.81 s6.31 d4.46 d3.46 m3.13 td3.30 m3.72 dd3.46 m
a In DMSO-d6 at 400 MHz. b In DMSO-d6 at 500 MHz.
Table 12. 13C NMR data (δ, ppm) of 3436 a.
Table 12. 13C NMR data (δ, ppm) of 3436 a.
CompoundCOC1C2C3C4C5C6CH3
34 b155.088.955.370.279.274.160.8---
35 a159.688.855.370.279.274.160.754.8
36 a155.489.255.870.479.374.160.818.4
a In DMSO-d6 at 100 MHz. b In DMSO-d6 at 125 MHz.
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Sosa-Gil, C.; Matamoros, E.; Cintas, P.; Palacios, J.C. Bifunctional Azido(thio)ureas from an O-Protected 2-Amino-2-deoxy-d-glucopyranose: Synthesis and Structural Analyses. Molecules 2024, 29, 5687. https://doi.org/10.3390/molecules29235687

AMA Style

Sosa-Gil C, Matamoros E, Cintas P, Palacios JC. Bifunctional Azido(thio)ureas from an O-Protected 2-Amino-2-deoxy-d-glucopyranose: Synthesis and Structural Analyses. Molecules. 2024; 29(23):5687. https://doi.org/10.3390/molecules29235687

Chicago/Turabian Style

Sosa-Gil, Concepción, Esther Matamoros, Pedro Cintas, and Juan C. Palacios. 2024. "Bifunctional Azido(thio)ureas from an O-Protected 2-Amino-2-deoxy-d-glucopyranose: Synthesis and Structural Analyses" Molecules 29, no. 23: 5687. https://doi.org/10.3390/molecules29235687

APA Style

Sosa-Gil, C., Matamoros, E., Cintas, P., & Palacios, J. C. (2024). Bifunctional Azido(thio)ureas from an O-Protected 2-Amino-2-deoxy-d-glucopyranose: Synthesis and Structural Analyses. Molecules, 29(23), 5687. https://doi.org/10.3390/molecules29235687

Article Metrics

Back to TopTop