Next Article in Journal
Fenofibrate Nanocrystal Composite Microparticles for Intestine-Specific Oral Drug Delivery System
Next Article in Special Issue
Special Issue “Carbohydrates 2018”
Previous Article in Journal
Long Term Effectiveness of Photodynamic Therapy for CIN Treatment
Previous Article in Special Issue
Synthesis and Biological Evaluation of Structurally Varied 5′-/6′-Isonucleosides and Theobromine-Containing N-Isonucleosidyl Derivatives
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Iminosugars: Effects of Stereochemistry, Ring Size, and N-Substituents on Glucosidase Activities

by
Luís O. B. Zamoner
,
Valquiria Aragão-Leoneti
and
Ivone Carvalho
*
School of Pharmaceutical Sciences of Ribeirão Preto, University of São Paulo, Av. do Café s/n, Monte Alegre, CEP14040-903 Ribeirão Preto, Brazil
*
Author to whom correspondence should be addressed.
Pharmaceuticals 2019, 12(3), 108; https://doi.org/10.3390/ph12030108
Submission received: 15 June 2019 / Revised: 8 July 2019 / Accepted: 10 July 2019 / Published: 12 July 2019
(This article belongs to the Special Issue Carbohydrates 2018)

Abstract

:
N-substituted iminosugar analogues are potent inhibitors of glucosidases and glycosyltransferases with broad therapeutic applications, such as treatment of diabetes and Gaucher disease, immunosuppressive activities, and antibacterial and antiviral effects against HIV, HPV, hepatitis C, bovine diarrhea (BVDV), Ebola (EBOV) and Marburg viruses (MARV), influenza, Zika, and dengue virus. Based on our previous work on functionalized isomeric 1,5-dideoxy-1,5-imino-D-gulitol (L-gulo-piperidines, with inverted configuration at C-2 and C-5 in respect to glucose or deoxynojirimycin (DNJ)) and 1,6-dideoxy-1,6-imino-D-mannitol (D-manno-azepane derivatives) cores N-linked to different sites of glucopyranose units, we continue our studies on these alternative iminosugars bearing simple N-alkyl chains instead of glucose to understand if these easily accessed scaffolds could preserve the inhibition profile of the corresponding glucose-based N-alkyl derivatives as DNJ cores found in miglustat and miglitol drugs. Thus, a small library of iminosugars (14 compounds) displaying different stereochemistry, ring size, and N-substitutions was successfully synthesized from a common precursor, D-mannitol, by utilizing an SN2 aminocyclization reaction via two isomeric bis-epoxides. The evaluation of the prospective inhibitors on glucosidases revealed that merely D-gluco-piperidine (miglitol, 41a) and L-ido-azepane (41b) DNJ-derivatives bearing the N-hydroxylethyl group showed inhibition towards α-glucosidase with IC50 41 µM and 138 µM, respectively, using DNJ as reference (IC50 134 µM). On the other hand, β-glucosidase inhibition was achieved for glucose-inverted configuration (C-2 and C-5) derivatives, as novel L-gulo-piperidine (27a) and D-manno-azepane (27b), preserving the N-butyl chain, with IC50 109 and 184 µM, respectively, comparable to miglustat with the same N-butyl substituent (40a, IC50 172 µM). Interestingly, the seven-membered ring L-ido-azepane (40b) displayed near twice the activity (IC50 80 µM) of the corresponding D-gluco-piperidine miglustat drug (40a). Furthermore, besides α-glucosidase inhibition, both miglitol (41a) and L-ido-azepane (41b) proved to be the strongest β-glucosidase inhibitors of the series with IC50 of 4 µM.

Graphical Abstract

1. Introduction

The major groups of glucosidase inhibitors that have been discovered are polyhydroxylated alkaloids containing piperidines, pyrrolidines, nor-tropanes, pyrrolizidines, and indolizidines as mono and bicyclic systems [1]. A great variety of these compounds, named iminosugars, have been isolated from natural sources, such as plants (Morus alba, Commelina communis), bacteria (Bacillus, Streptomyces), and fungi (Zygosaccharomyces rouxii for mulberry leaf fermentation) [2], and produced by synthetic strategies with potential inhibition properties not only over α- and β-glucosidases but also glycosyltransferases, glycogen phosphorylase [3,4,5], nucleoside phosphorylases [6], and sugar-nucleotide mutases (UDP-Galp mutase) [7]. High activity and specificity of iminosugars are associated with the ability of the nitrogen ring to mimic the transition state of pyranosidic or furanosidic units of natural glucosidase substrates that positively influence their shape and charge for enzyme binding.
Access to iminosugar analogues with N-substituted side chains has led to a variety of potent glucosidase and glycosyltransferases inhibitors with broad therapeutic applications, such as treatment of diabetes [8] and Gaucher disease [9], and even immunosuppressive activities [10] and antibacterial [11] and antiviral effects [12,13] against HIV [14], HPV [15], hepatitis C [16], bovine diarrhea (BVDV) [17], Ebola (EBOV) [18] and Marburg viruses (MARV) [19], influenza [20], Zika [21], and dengue virus [22,23]. Despite the α- and β-glucosidase inhibition promoted by nojirimycin itself (1), it was achieved a better profile for the corresponding 1-deoxynojirimycin (DNJ, 2) due to better stability and potency. Furthermore, a combination of these structural features can promote the cellular uptake of N-alkylated DNJ analogues, as shown by those containing long and linear alkyl chains, which displayed better activity in whole cells (human hepatoblastoma cells, HepG2) than purified pork glucosidase I [24]. Conversely, N-alkyl-less lipophilic N-alkyl groups, or even containing an oxygen atom, displayed lower cytotoxicity and significant activity against α-glucosidase, as described for N-methyl- (3) [25], N-butyl- (N-Bu-DNJ, miglustat, 4) [26], N-hydroxyethyl- (N-EtOH-DNJ, miglitol, 5), N-7-oxadecyl- (N-7-oxadecyl-DNJ, 6) [27,28], and N-glycyl-deoxynojirimycin (7) [29]. In fact, miglustat is particularly useful in the control of type I Gaucher disease [9] and Niemann–Pick type C (NPC) lysosomal storage diseases, via “substrate reduction therapy”, as well as miglitol in the treatment of non-insulin-dependent diabetes (type II) (Figure 1) to impair carbohydrate processing in the gut [30]. In addition, inhibition of the target human acid β-glucosidase (glucocerebrosidase, GCase) has been achieved by a set of derived iminosugars as new pharmacological chaperones for the treatment of Gaucher disease [31,32].
Besides the N-alkyl variations, several studies have pointed to the impact of the modification of DNJ hydroxyl groups involving C-2 to C-6 positions, assessing the influence of both stereochemistry and substituent variations on glycosidase activities. In general, the loss of α-glucosidase (I and II) and ceramide glycosyltransferase activities was evident by modifying C-2, C-3, and C-4, with the exception of N-butyl-1-deoxy-galactonojirimycin (migalastat, used for the treatment of Fabry disease) [30]. On the other hand, changes at C-1 and the ring nitrogen were allowed, based on the high inhibition revealed by DNJ and 1-azasugars (isofagomine (8), for instance), the latter obtained by replacing the anomeric carbon and the ring oxygen of glucose by nitrogen and carbon, respectively, with significant activity only against β-glucosidase. Interestingly, introduction of a hydroxyl group at the carbon (C-2) neighboring the nitrogen afforded a potent α-glucosidase inhibitor (noeuromycin, 9) that preserved the original β-glucosidase activity. Additionally, the extra hydroxymethylene group at the anomeric position of DNJ gave rise to α-homonojirimycin (10) and β-homonojirimycin (11) with the ability to inhibit α-glucosidase, which was even higher whilst bearing N-methyl or N-butyl substituents (Figure 1) [1]. Furthermore, seven-membered ring iminosugar have shown potential glucosidase inhibition [33,34,35]. Comprehensive studies on iminosugar derivatives can be found in literature reviews [36,37,38,39].
Inspired by a series of reported deoxynojirimycin disaccharides that were decorated with equal α- or β-glucopyranose units at C-2, C-3, and C-4 DNJ positions (12-17) [40,41], along with N-glycosylated deoxynojirimycin, MDL 73,945 (18) [42], we had reported an alternative approach, using functionalized isomeric 1,5-dideoxy-1,5-imino-D-gulitol (L-gulo-piperidines, with inverted configurations at C-2 and C-5 with respect to glucose or DNJ, 19-21) and 1,6-dideoxy-1,6-imino-D-mannitol (D-manno-azepane derivatives, 22-24) cores N-linked to different sites of glucopyranose units, such as C-1, C-3, and C-6 positions [43]. To reach this goal, we used a CuAAC reaction (copper azide alkyne cycloaddition reaction), as a click chemistry strategy, to connect six- and seven-membered iminosugars to glucose in different arrangements through triazole bridges to produce the most active α-glucosidase inhibitor (21) of the pseudo-disaccharide series, L-gulopiperidine attached to glucose C-6 position, with IC50 approximately three-fold lower than that of DNJ (Figure 1) [43].
Despite the reported loss of α-glucosidase activity under modification at the C-5 position of the iminosugar, such as displayed by 1-deoxy-L-ido-nojirimycin (with an inverted configuration at C-5 with respect to DNJ) [24] and for 1,5-dideoxy-1,5-iminoxylitol (lacking the C-5 hydroxymethyl group of DNJ) [27,28], we have been encouraged to continue our studies on L-gulo-piperidines based on the remarkable α-glucosidase inhibition previously obtained for pseudo-disaccharides with simultaneous inverted configurations at C-2 and C-5 positions in relation to glucose stereochemistry [43]. Thus, to understand the relative contribution of the ring size, stereochemistry, and N-alkyl substitution on glycosidase inhibition, we proceeded with the synthesis of a small library of N-substituted 1-deoxy-L-gulo-nojirimycin and D-manno-azepane derivatives and compared them with the corresponding classical N-substituted of 1-deoxy-D-gluco-piperidine (DNJ) and L-ido-azepane counterparts as glucose-type carbohydrate mimetics. To reach this goal, N-hydroxyethyl and N-butyl groups of miglitol and miglustat drugs, respectively, were investigated as highly important N-alkyl substitutions for glucosidase inhibition, besides N-phenethyl [44], and N-propynyl [43] as a less-active counterpart.

2. Results and Discussion

2.1. Chemistry

Initially, target 1-deoxy-L-gulo-nojirimycin (26-29a) and D-manno-azepane derivatives (26-29b) were synthesized by a regiospecific C2-symmetric unprotected bis-epoxide opening strategy in the presence of primary amines, followed by an SN2 aminocyclization reaction to give a mixture of both six- and seven-membered iminosugar isomers [45,46,47]. As previously reported, the synthesis of unprotected bis-epoxide 25 was promptly achieved from the simple and commercially available starting material, D-mannitol, in two steps, by tosylation of both D-mannitol primary alcohols (71%), followed by a base-promoted intramolecular SN2 reaction to give 1,2:5,6-dianhydro-D-mannitol 25 (29%), Scheme 1A [43]. Opening of the homochiral C2-symmetric bis-epoxide 25 by alkylamines (at the less-hindered position of one epoxy function) led to the formation of secondary amines, which promoted an SN2 aminocyclization reaction to give a mixture of polyhydroxy-piperidine and -azepane by 6-exo-tet or 7-endo-tet processes, respectively. In this series, azepane was isolated in a slightly higher proportion than DNJ, indicating the free 3,4 diol in bis-epoxide 25 did not affect the regioselectivity. Conversely, higher yields of DNJ than azepane derivatives were previously reported using benzyl protecting groups at C-3 and C-4 of bis-epoxide in the presence of several amines (benzylamine, for instance) [46]. Interestingly, this ratio can be reversed in the presence of a Lewis acid (perchloric acid), which catalyzes epoxide opening to mainly give azepane derivatives. Furthermore, the exclusive formation of seven-membered azasugars using a more rigid trans-acetonide protecting group, as observed in 1,2:5,6-dianhydro-3,4-O-isopropylidene-D-mannitol or L-iditol, led to the conclusion that formation of both polyhydroxy-piperidine and -azepane regioisomers can be achieved by an aminocyclization of a flexible bis-epoxide bearing a free or acyclic hydroxyl protecting groups at C-3 and C-4, and the ratio varies according to the experimental conditions [46].
Therefore, the microwave-assisted aminocyclization reaction of bis-epoxide 25 was carried out with four different primary amines (propargylamine, butylamine, ethanolamine, or phenethylamine), which resulted in a mixture of N-alkyl substituted polyhydroxypiperidine 26-29a and azepane 26b-29b regioisomers, being iminosugars 27a,b and 28a,b novel compounds. To prevent laborious separation of a and b regioisomers, some mixtures were treated with acetic anhydride and pyridine for prompt separation of the per-O-acetylated 1-deoxy-L-gulo-nojirimycin and D-manno-azepane by column chromatography, isolated in different ratio and yields over two steps, as depicted in Table 1. However, the separation of protected regioisomers 31a,b was more demanding and required HPLC purification, mainly because the deprotected products 28a,b were inseparable by chromatographic column or even HPLC under test conditions. Lastly, deprotection of the regioisomers in the presence of sodium methoxide gave products 26-29a and 26-29b in quantitative yields. Eventually, derivatives 27a,b and 29a,b bearing a more lipophilic side chain (butyl and phenethyl, respectively) were separated without the need of previous protection by using chromatography column eluted with DCM/MeOH (4:1). However, the yields of pure polyhydroxy-piperidines and -azepanes were much lower (approximately 10%, 0.8:1 ratio, respectively) than using protection/deprotection strategies (Table 1).
In order to keep the stereo-control during the reaction and obtain the iminosugars with the same stereochemistry of glucose, we pursued the classical procedure based on the protection of 1,2- and 5,6- positions of D-mannitol to produce the diisopropylidene intermediate 32 [48], which was benzylated at 3,4- positions and then deprotected under acid catalysis to give compound 33 (Scheme 1B) [47]. Briefly, selective protection of primary hydroxyl functions with bulk groups, followed by activation of the O-2 and O-5 with mesyl chloride and treatment of 34 in MeOH with concentrated HCl allowed the preparation of bis-epoxide 35 because of the intramolecular attack of the released primary hydroxyl functions that displaces the leaving mesyl groups. Then, bis-epoxide 35, comprising inverted configurations at C-2 and C-5 comparatively to 25, was converted to the corresponding mixture of N-substituted 1-deoxy-D-gluco-nojirimycin (36-38a) and 1,6-dideoxy-1,6-imino-L-ido-azepane derivatives (36-38b) in approximately 1:1 ratio under treatment with propargylamine, butylamine, ethanolamine, or phenethylamine for the aminocyclization reaction, as described for bis-epoxide 25. Attempts to generate the N-phenethyl derivative of this series were unsuccessful since D-glucitol was isolated as a major product, possibly because phenethylamine promoted a regioselective opening of partially protected 1,2-epoxide (35) and then an O-cyclization leading to glucitol, as reported using ammonium formate [49].
After chromatographic separation of regioisomers, removal of the benzyl groups was better achieved under treatment with trimethylsilyl iodine [50] rather than hydrogenation conditions [47] to give final products 39-41a-b in moderate to quantitative yields (45–100%).

2.2. Biological Assays

Initially, the small library of iminosugar derivatives (26-29a,b and 39-41a,b) was screened for α-glucosidase inhibition (from Saccharomyces cerevisiae) activities using p-nitrophenyl α-D-glucopyranoside as substrate and prospective inhibitors at 1.0 mM concentration. To broaden the scope of the analysis, β-glucosidase (almond) activity of the same set of compounds was conducted using the corresponding p-nitrophenyl β-D-glucopyranoside.

2.2.1. Yeast α-glucosidase Activities

Based on the IC50 values using α-glucosidase, the greatest inhibition was verified for both piperidine and azepane DNJ derivatives bearing the N-hydroxylethyl group, D-gluco-piperidine (miglitol, 41a, IC50 41 µM) and L-ido-azepane (41b, IC50 138 µM), using the DNJ as the reference (IC50 134 µM) (Table 2). The α-glucosidase inhibition promoted by L-ido-azepane 41b was significant and related to DNJ, although with a three-fold lower activity than D-gluco-piperidine (41a). In spite of finding a patent for azepane 38b, the data were inaccessible [51], and mixed results were found for nonsubstituted L-ido-azepane with inhibition properties (Ki 4.8 µM) lower than the corresponding D-gluco-piperidine (DNJ, Ki 0.44 µM) assayed on Bacillus stearothermophilus α-glucosidase [46] and high (Ki 29.4 µM) [52] to weak activity (IC50 772 µM [33] or 35% inhibition at 1 mM [53]) using yeast α-glucosidase. In addition, weak or no α-glucosidase inhibition was observed for N-propynyl (39a,b) and N-butyl (40a,b) DNJ and azepane derivatives in these assays, confirming reported data for 39a and miglustat (40a) [54] and 40b (14% inhibition at 1 mM) [33] both using yeast α-glucosidase.
In respect to L-gulo-piperidine and D-manno-azepane series, with inverted configurations at C-2 and C-5, derivatives (26-29a,b) bearing N-hydroxyethyl, N-butyl, N-propynyl [43], or N-phenethyl chains on the endocyclic nitrogen proved to be inactive against yeast α-glucosidase at the tested concentration (15–2000 µM), leading to loss of activity even for the N-hydroxyethyl derivatives (28a,b) when compared to 41a,b. Reported α-glucosidase inhibition data for nonsubstituted L-gulo-piperidine and D-manno-azepane were found as weak as 30% and 55%, respectively, tested at 1 mM in Bacillus stearothermophilus [46] or 21% in yeast α-glucosidase at 240 µM [52]. Thus, it was evident that α-glucosidase activities were considerably affected by iminosugar stereochemistry, ring size, and N-substitutions, and inversion of configuration was detrimental for activity regardless of the N-substituents here described, suggesting the wrong orientation of at least two hydroxyl groups attached at C-2 and C-5, which led to reduced binding affinity at yeast α-glucosidase active sites.

2.2.2. Almond β-glucosidase Activities:

Conversely, the assessment of the series on β-glucosidase revealed that novel L-gulo-piperidine (27a) and D-manno-azepane (27b) derivatives (inverted configurations at C-2 and C-5 related to glucose) preserving the N-butyl chain showed significant activity, IC50 109 and 184 µM, respectively, comparable to miglustat (40a, IC50 172 µM), although three- to five-fold lower than DNJ (IC50 33 µM) (Table 2). In this particular case, the N-butyl chain seems to play an important role in β-glucosidase inhibition since the reported data for nonsubstituted was low for both L-gulo-piperidine (13% at 1 mM) and D-manno-azepane (1% at 1 mM) or no inhibition at 240 µM on the same enzyme [46,52]. Interestingly, for the same set that preserve the N-butyl chain but display glucose stereochemistry, the seven-membered ring derivative L-ido-azepane 40b displayed nearly twice the activity (IC50 80 µM) of the corresponding D-gluco-piperidine drug (40a), which resembled the stronger β-glucosidase inhibition achieved for nonsubstituted L-ido-azepane (Ki 17 µM [46], 12.8 µM [52], or IC50 38 µM [53]) than nonsubstituted D-gluco-piperidine (Ki 1700 µM) [46]. Furthermore, both derivatives bearing the N-hydroxyethyl chain, as occurs in miglitol (41a) and L-ido-azepane 41b, proved to be the strongest β-glucosidase inhibitors with IC50 of 4 µM. Based on all these findings, it was possible to infer that almond β-glucosidase active sites can accept and interact with a wider range of iminosugars than yeast α-glucosidase.
See dose-response curves obtained from Yeast α-Glucosidase and Almond β-Glucosidase assays in Supplementary Materials.

3. Conclusions

In summary, a series of iminosugars were successfully synthesized from a common precursor, D-mannitol, to produce two alternative bis-epoxides, further modified by an SN2 aminocyclization reaction to give a mixture of both N-substituted six- and seven-membered iminosugar isomers. Besides the ring size, two additional structural variations were also pursued to broaden the scope of reported strategies, as stereochemistry (maintenance of glucose stereochemistry or inversion of configuration at C-2 and C-5 positions) and N-chain of the endocyclic nitrogen (N-propynyl, -butyl, -hydroxyethyl, and -phenethyl). Classical polyhydroxypiperidines, miglustat and miglitol drugs that maintain glucose configuration (D-gluco-nojirimycin, DNJ) and bear N-butyl and N-hydroxyethyl chains, respectively, were synthesized and used as reference for evaluation of the series towards α- and β-glucosidases. Assessment of α-glucosidase activity of iminosugars revealed solely miglitol as the most active of the series, followed by the corresponding L-ido-azepane isomer. All other iminosugars proved to not be inhibitors of yeast α-glucosidase. On the other hand, all N-butyl iminosugars having either glucose stereochemistry, D-gluco-piperidine miglustat drug and L-ido-azepane, or inverted configurations at C-2 and C-5 related to glucose, L-gulo-piperidine and D-manno-azepane derivatives, displayed significant inhibition of almond β-glucosidase. In spite of that, the strongest inhibition was achieved for D-gluco-piperidine miglitol drug and the corresponding L-ido-azepane iminosugars containing N-hydroxyethyl chains, but, in these tests, no activity was accomplished for inverted configuration counterparts. Thus, we observed that glucosidase inhibition promoted by some polyhydroxypiperidines was accompanied by proportional inhibition of the corresponding polyhydroxyazepane isomers bearing the same N-chain regardless of the ring stereochemistry. In addition, the findings of this study on β-glucosidase inhibition by L-gulo-piperidine and D-manno-azepane series are relevant considering their straightforward synthesis compared to DNJ series.

4. Material and Methods

1H (300 MHz) and 13C (75 MHz) NMR spectra were recorded on a Bruker® Ultrashield 300 NMR spectrometer. 1H (400 MHz) and 13C (100 MHz) NMR spectra were recorded on a Bruker® Avance 400 MHz NMR spectrometer. 1H (500 MHz) and 13C (125 MHz) NMR spectra were recorded on a Bruker® Avance 500 MHz NMR spectrometer. All spectra were recorded at room temperature (~20 °C) in Sigma Aldrich® deuterated solvents. Chemical shifts (δ) were expressed in parts per million (ppm) relative to the reference peak. Coupling constants (J) were expressed in Hertz (Hz). Splitting patterns in 1H NMR spectra were designated as s (singlet), br s (broad singlet), d (doublet), br d (broad doublet), t (triplet), q (quartet), dd (doublet of doublets), dt (doublet of triplets), ddd (doublet of doublet of doublets), and m (multiplet). Optical rotations were measured on a Jasco P-2000 polarimeter at 22.5 nd using a sodium lamp and wavelength of 589 nm at 22.5 °C. HPLC purifications were designed in a Shimadzu® SCL-10A HPLC system, Diode Array Detector Shimadzu® SPD-M10A, and processed on Class-VP software. Purification of the compounds 31a and 31b was performed in HPLC using a Macherey–Nagel CLC-ODS semiprep column, Methanol 40%, flowrate 4.0 mL/min, and 200 nm. High-resolution mass spectra (HRMS) were obtained on a Bruker Daltonics MicrOTOF-Q II ESI-TOF mass spectrometer, and an Exactive Plus Orbitrap mass spectrometer (Thermo Scientific, Germany) was equipped with an electrospray (H-ESI-II) probe and operated in negative ionization mode. The system was controlled by Xcalibur and Tune (Thermo Scientific). For biological assays, absorbance at 405 nm was measured using SpectraMax M2 Molecular Devices®.
1,2:5,6-di-anhydro-D-mannitol (25) [43]: D-Mannitol (5.05 g, 27.4 mmol) was solubilized in pyridine (25.0 mL) and heated to 120 °C for 15 min. After that, the solution was refrigerated to 0 °C, treated dropwise with tosyl chloride solution (13.10 g, 68.7 mmol/10 mL pyridine) for 1 h, and stirred at 0 °C for 3 h then at r.t. for 1 h. The mixture was coevaporated with toluene, and the solid was diluted in dichloromethane and washed with HCl 1 mol·L−1 and saturated NaHCO3. The organic layer was dried over MgSO4 and concentrated. The crude mixture of 1,6-di-O-tosyl-D-mannitol was solubilized in a mixture of acetonitrile and water (38.0 mL, 2:1, v/v), and a small portion of phenolphthalein was added to it. The mixture was stirred at 35–40 °C and titrated with NaOH 5 mol·L−1 until the solution remained pink. After that, a mixture of Na2CO3 (87.4 g) in ethyl acetate (324 mL) was added and stirred vigorously. The solid was filtrated and washed with ethyl acetate. The organic solution was dried over MgSO4, filtered, and concentrated. The resulting mixture was purified by flash chromatography (hexane and ethyl acetate, 7:3, v/v) to yield 1 (1.12 g, 7.70 mmol, 28%). 1H NMR (400 MHz, CD3OD): δ 2.77 (2H, dd, J 2.7 Hz, 5.3 Hz, H-1a, H-6a); 2.82 (2H, dd, J 4.0 Hz, 5.3 Hz, H-1b, H-6b); 3.11–3.17 (2H, m, H-2, H-5); 3.46–3.52 (2H, m, H-3, H-4). 13C NMR (100 MHz, CDCl3): δ 46.1 (C-1, C-6); 52.8 (C-2, C-5); 73.3 (C-3, C-4).
1,2:5,6-Di-O-isopropilidene D-mannitol (32) [48]
1,2:5,6-Dianhydro-3,4-di-O-benzyl-L-iditol (35): prepared as described by Wilkinson et al. [47]
General procedure for the synthesis ofL-gulo-piperidine and D-manno-azepane derivatives(26a,b-31a,b): 1,2:5,6-dianhydro-D-glucitol was solubilized in methanol and 2.5 eq of primary amine (propargylamine, ethanolamine, butylamine, or phenethylamine) was added to the solution. The mixture was heated to 90 °C in a microwave for 5 min (150 W) in a sealed vessel. The solvents were evaporated in vacuum. Compounds 27a,b and 29a,b were separated in a flash column (dichloromethane and methanol, 8:2 v/v). Compounds 26a,b and 28a,b were acetylated by the addition of pyridine (8.0 mL) and acetic anhydride (22.8 mL) and stirred at r.t. After 3 h, ice was added to the mixture, the product was extracted with ethyl ether, and the organic layer was evaporated and dried with MgSO4. Compounds 30a,b were separated by column chromatography (toluene and ethyl acetate, 3:7, v/v). Compounds 31a,b were purified in a HPLC C-18 semiprep column in methanol/water 40% and flow rate 4.0 mL/min.
See 1H, 13C and bidimensional NMR and ESI HRMS spectra of compounds 26-31a,b and 36-41a,b in Supplementary Materials.
General procedure for removal of Acetyl groups: Compounds 30a,b and 31a,b were dissolved in methanol (1.0 mL), and sodium methoxide 1 M was added dropwise to the solution until pH 9.0 and checked with Tornassol. After half an hour, TLC showed total consumption of starting material, then it was neutralized with ion exchange resin DOWEX® 50WX4-50. After that, the mixture was filtered through a Celite® pad and concentrated in vacuo.
General procedure for removal o Benzyl groups [50]: The corresponding product was solubilized in dichloromethane, and Me3SI (4 eq) was added slowly. The reaction was allowed to stir at r.t. for 15 min. TLC showed total consumption of starting material, then the reaction was quenched with methanol. The product was purified in a SPE-C18 silica pad with methanol/water 1:1 v/v.
N-Propynyl-1,5-dideoxy-1,5-imino-L-gulitol (26a) [43]: 80% (0.0082 g, 0.041 mmol): 1H NMR (500 MHz, D2O): δ 2.71–2.77 (2H, m, H-1a, ≡CH); 2.85 (1H, t, J 10.9 Hz, H-1b); 2.88–2.93 (1H, m, H-5); 3.46 (1H, d, J 17.6 Hz, H-7a); 3.68 (1H, d, J 17.6 Hz, H-7b); 3.83 (1H, dd, J 5.4 Hz, 11.7 Hz, H-6a); 3.89 (1H, dd, J 3.9 Hz, 11.7 Hz, H-6b); 3.91–3.94 (1H, m, H-3); 4.05–4.14 (2H, m, H-2, H-4). ESI HRMS: [M+H]+ calculated for C9H15NO4 202.1074; found 202.1076.
N-Propynyl-1,6-dideoxy-1,6-imino-D-mannitol (26b) [43]: 86% (0.0088 g, 0.044 mmol): 1H NMR (500 MHz, D2O): δ 2.66 (1H, t, J 2.2 Hz, ≡CH); 2.85 (4H, m, H-1a; H-1b; H-6a; H-6b); 3.37 (1H, d, J 17.0 Hz, H-7a); 3.41 (1H, d, J 17.0 Hz, H-7b); 3.89–3.91 (2H, m, H-3, H-4); 4.12 (2H, t, J 4.8 Hz, H-2, H-5). ESI HRMS: [M+H]+ calculated for C9H15NO4 202.1074; found 202.1073.
N-Butyl-1,5-dideoxy-1,5-imino-L-gulitol (27a): [ α ] D 22.5 29.5 (c 1.8, MeOH), 1H NMR (500 MHz, D2O): δ 0.83 (3H, t, J 7.4 Hz, H-10); 1.17–1.28 (2H, m, H-9); 1.37–1.50 (2H, m, H-8); 2.50–2.92 (5H, m, 2xH-1, H-5, 2xH-7); 3.75 (1H, dd, J 5.6. 11.8, H-6a); 3.78–3.85 (2H, m, H-3, H-6b); 3.94–4.00 (1H, m, H-2); 4.05–4.11 (1H, m, H-4). 13C NMR (125 MHz, D2O): δ 13.3 (C-10); 20.0 (C-9); 27.4 (C-8); 53.1 (C-1 or C-7); 55.1 (C-1 or C-7); 58.5 (C-6); 68.2 (C-4); 70.1 (C-3); 70.4 (C-3); 72.8 (C-2). ESI HRMS: [M+H]+ calculated for C10H22NO4 220.1544; found 220.1538.
N-Butyl-1,6-dideoxy-1,6-imino-D-mannitol (27b): [ α ] D 22.5 −62.0 (c 2.2, MeOH), 1H NMR (500 MHz, D2O): δ 0.76 (3H, t, J 7.3 Hz, 3xH-10); 1.07–1.25 (2H, m, 2xH-9); 1.30–1.44 (2H, m, 2xH-8); 2.47–2.57 (2H, m, 2xH-7); 2.65–2.85 (4H, m, 2xH-1, 2xH-6); 3.73–3.79 (2H, m, H-3, H-4); 3.94–4.04 (2H, m, H-2, H-5). 13C NMR (125 MHz, D2O): δ 13.2 (C-10); 20.0 (C-9); 27.5 (C-8); 55.2 (C-1, C-6); 58.4 (C-7); 68.3 (C-2; C-5); 72.8 (C-4; C-3). ESI HRMS: [M+H]+ calculated for C10H22NO4 220.1544; found 220.1543.
N-Hydroxyethyl-1,5-dideoxy-1,5-imino-L-gulitol (28a): [ α ] D 22.5 16.1 (c 0.6, MeOH), 1H NMR (400 MHz, CD3OD): δ 2.53–2.71 (2H, m, H-1a, H-7a); 2.78–2.89 (2H, m, H-5, H-7b); 2.97 (1H, ddd, J 4.9; 7.0; 13.5 Hz, H-1b); 3.55–3.98 (7H, m, H-2, H-3, H-4, H-6a, H-6b, H-8a, H-8b). 13C NMR (100 MHz, CD3OD): δ 51.9 (C-7); 55.1 (C-1); 58.6 (C-8); 60.0 (C-6); 61.0 (C-5); 66.5, 70.9, 71.3 (C-2, C-3, C-4). ESI HRMS: [M+H]+ calculated for C8H18NO5 208.1180; found 208.1177
N-Hydroxyethyl-1,6-imino-D-mannitol (28b): [ α ] D 22.5 −15.3 (c 0.5, MeOH), 1H NMR (400 MHz, CD3OD): δ 2.58–2.80 (4H, m, H-1a; H-6ª; 2xH-7); 2.89 (2H, dd, J 4.3; 13.2 Hz, H-1b, H-6b); 3.59 (2H, dd, J 6.0; 12.5 Hz, 2xH-8); 3.86–3.92 (2H, m, H-3, H-4); 4.00–4.08 (2H, m, H-2, H-5). 13C NMR (100 MHz, CD3OD): δ 58.4 (C-1; C-6); 60.4 (C-8); 61.8 (C-7); 70.9 (C-2; C-5); 74.0 (C-3; C-4). ESI HRMS: [M+H]+ calculated for C8H18NO5 208.1180; found 208.1182.
N-Phenethyl-1,5-dideoxy-1,5-imino-L-gulitol (29a) [46]
N-Phenethyl-1,6-dideoxy-1,6-imino-D-mannitol (29b) [46]
N-Propynyl-2,3,4,6-tetra-O-acetyl-1,5-dideoxy-1,5-imino-L-gulitol (30a) [43]: 1H NMR (400 MHz, CDCl3): δ 2.03, 2.09, 2.13, 2.14 (12H, 4s, 4×CH3); 2.33 (1H, t, J 2.3 Hz, ≡CH); 2.80 (1H, dd, J 4.7 Hz, 11.2 Hz, H-1a); 2.97 (1H, dd, J 9.6 Hz, 11.2 Hz, H-1b); 3.28 (1H, ddd, J 3.3 Hz, 5.8 Hz, 10.0 Hz, H-5); 3.44 (1H, dd, J 2.3 Hz, 17.8 Hz, H-7a); 3.67 (1H, dd, J 2.3 Hz, 17.8 Hz, H-7b); 4.20 (1H, dd, J 5.8 Hz, 11.6 Hz, H-6b); 4.24 (1H, dd, J 3.3 Hz, 11.6 Hz, H-6a); 5.20–5.27 (3H, m, H-2, H-3, H-4). 13C NMR (100 MHz, CDCl3): δ 20.8; 20.9 (CH3); 43.9 (C-1); 49.6 (C-7); 55.5 (C-5); 61.4 (C-6); 66.5 (C-3, C-4); 68.7 (C-2); 77.2 (≡CH); 169.3; 169.7; 170.5 (C=O). ESI HRMS: [M+H]+ calculated for C17H24NO8 370.1496; found 370.1496.
N-Propynyl-2,3,4,6-tetra-O-acetyl-1,6-dideoxy-1,6-imino-D-mannitol (30b) [43]: 1H NMR (400 MHz, CDCl3): δ 2.05, 2.12 (12H, 2s, 4×CH3); 2.25 (t, J 2.3 Hz, ≡CH); 2.91 (2H, dd, J 5.6 Hz, 13.8 Hz, H-1a, H-6a); 2.99 (2H, dd, J 4.4 Hz, 13.8 Hz, H-1b, H-6b); 3.39 (1H, dd, J 2.3 Hz, 17.3 Hz, H-7a); 3.45 (1H, dd, J 2.3 Hz, 17.3 Hz, H-7b); 5.40 (2H, dt, J 1.2 Hz, 4.6 Hz, H-2, H-5); 5.48 (2H, t, J 1.2 Hz, H-3, H-4). 13C NMR (100 MHz, CDCl3): δ 20.8; 21.0 (CH3); 47.9 (C-7); 54.1 (C-1, C-6); 69.9 (C-2, C-5); 70.9 (C-3 e C-4); 77.2 (≡CH); 169.9; 170.1 (C=O). ESI HRMS: [M+H]+ calculated for C17H24NO8 370.1496; found 370.1535.
N-Acetoxyethyl-2,3,4,6-tetra-O-acetyl-1,5-dideoxy-1,5-imino-L-gulitol (31a):1H NMR (400 MHz, CDCl3): δ 2.06; 2.08; 2.10 (18H, 3s, 5xCH3); 2.89 (1H, dd, J 5.0; 13.7 Hz, H-1a); 2.96 (2H, t, H-7, J 5.9 Hz, H-7); 3.04 (1H, dd, J 2.6; 13.7 Hz, H-1b); 3.45 (1H, dd, J 4.8; 11.4 Hz, H-5); 4.05 (1H, dd, J 5.6; 11.3 Hz, H-6a); 4.10–4.21 (2H, m, 2xH-8); 4.37 (1H, dd, J 7.0; 11.8 Hz, H-6b); 5.15 (1H, dd, J 3.3; 9.0 Hz, H-3); 5.20–5.25 (1H, m, H-2); 5.20–5.25 (1H, m, H-2); 5.28 (1H, dd, J 4.9; 9.0 Hz, H-4). 13C NMR (100 MHz, CDCl3): δ 20.8, 20.9, 21.0 (5xCH3); 48.9 (C-1); 52.7 (C-7); 58.2 (C-5); 59.9 (C-6); 68.0 (C-3); 68.1, 68.3 (C-2, C-4).
N-Acetoxyethyl-2,3,4,6-tetra-O-acetyl-1,6-dideoxy-1,6-imino-D-mannitol (31b):1H NMR (400 MHz, CDCl3): δ 2.03; 2.07; 2.10 (18H, 3s, 5xCH3); 2.85 (2H, t, J 5.8 Hz, H-4, H-4); 2.91 (2H, dd, J 5.6; 14.0 Hz, H-1a, H-1a); 3.04 (2H, dd, J 4.3; 14.0 Hz, H-1b, H-1b); 4.11 (2H, dd, J 5.7; 9.1 Hz, H-5, H-5); 5.29-5.37 (2H, m, H-2, H-2); 5.43-5.47 (2H, m, H-3, H-3). 13C NMR (100 MHz, CDCl3): δ 20.8; 20.9; 21.0 (5xCH3); 54.8 (C-1, C-1); 56.4 (C-4); 62.2 (C-5); 70.2 (C-2, C-2); 70.6 (C-3, C-3).
N-Propynyl-2,3,4,6-tetra-O-acetyl-1,5-dideoxy-1,5-imino-D-glucitol (36a):1H NMR (400 MHz, CDCl3): δ 2.27 (1H, t, J 2.2 Hz, ≡CH); 2.50 (1H, d, J 9.5 Hz, H-5); 2.60 (1H, t, J 10.7 Hz, H-1a); 2.92 (1H, dd, J 4.9; 10.9 Hz, H-1b); 3.31–3.44 (2H, m, 2xH-7); 3.62–3.91 (5H, m, H-2, H-3, H-4, 2xH-6); 4.77 (2H, dd, J 3.8; 11.2 Hz, CH2Ph); 4.98 (2H, dd, J 11.2; 13.4 Hz, CH2Ph); 7.29–7.45 (10H, m, H-Ph). 13C RMN (100 MHz, CDCl3): δ 42.1 (C-1); 56.1 (C-7); 57.3 (C-6); 63.1 (C-2 or C-5); 69.4 (C-2 or C-5); 74.6 (C≡CH); 75.2 (CH2Ph); 87.3 (C-3; C-4); 127.9, 128.6, 128.7 (Ar); 138.1 (Cq). ESI HRMS: [M+H]+ calculated for C23H28NO4 382.2013; found 382.2002.
N-Propynyl-2,3,4,6-tetra-O-acetyl-1,6-dideoxy-1,6-imino-L-iditol (36b):1H NMR (400 MHz, CDCl3): δ 2.31 (1H, t, J 2.3 Hz, ≡CH); 2.77 (2H, dd, J 8.2; 12.3 Hz, H-1a, H-6a); 2.98 (2H, dd, J 1.1; 12.7, H-1b, H6b); 3.40–3.56 (2H, m, 2xH-7); 3.63–3.70 (2H, m, H-3, H-4); 3.80–3.93 (2H, m, H-2, H-5); 4.67 (2H, d, J 11.2 Hz, CH2Ph); 4.81 (2H, d, J 11.2 Hz, CH2Ph); 7.29–7.42 (10H, m, H-Ph). 13C RMN (100 MHz, CDCl3): δ 48.8 (C-7); 56.9 (C-1, C-6); 68.0 (C-2, C-5); 73.8 (CH2Ph); 78.2 (C≡CH); 86.5 (C-3, C-4); 127.9, 128.0, 128.6 (Ar); 137.9 (Cq). ESI HRMS: [M+H]+ calculated for C23H28NO4 382.2013; found 382.2001.
N-Butyl-2,3,4,6-tetra-O-acetyl-1,5-dideoxy-1,5-imino-D-glucitol (37a):1H NMR (400 MHz, CDCl3): δ 0.94 (3H, t, J 6.9 Hz, 3xH-10); 1.21–1.38 (2H, m, 2xH-9); 1.39–1.56 (2H, m, 2xH-8); 2.26 (1H, t, J 10.6Hz, H-5); 2.32–2.54 (3H, m, H-1a, H-1b, H-7a); 2.70–2.82 (1H, m, H-7b); 3.13 (1H, dd, J 4.6; 11.2 Hz, H-6a); 3.38 (1H, t, J 8.8 Hz, H-3); 3.60–3.73 (2H, m, H-4, H-6b); 3.77–3.93 (2H, m, H-2, H-XX); 4.75 (2H, dd, J 5.6; 11.2 Hz, CH2Ph); 4.96 (2H, t ap., J 11.0 Hz, CH2Ph); 7.29–7.48 (10H, m, Ar). 13C RMN (100 MHz, CDCl3): δ 14.0 (C-10); 20.6 (C-9); 27.3 (C-8); 52.1 (C-1 or C-7); 55.1 (C-1 or C-7); 57.5 (C-6); 64.9 (C-2 or C-5); 69.3(C-2 or C-5); 74.9 (CH2Ph); 75.1 (CH2Ph); 78.1 (C-3 or C-4); 86.8 (C-3 or C-4); 127.8, 127.9, 128.0, 128.5, 128.7 (C-Ar); 138.1 (Cq); 138.5 (Cq). ESI HRMS: [M+H]+ calculated for C24H34NO4 400.2483; found 400.2477.
N-Butyl-2,3,4,6-tetra-O-acetyl-1,6-dideoxy-1,6-imino-L-iditol (37b):1H NMR (400 MHz, CDCl3): δ 0.93 (3H, t, J 7.2 Hz, 3xH-10); 1.25–1.40 (2H, m, 2xH-9); 1.42–1.58 (2H, m, 2xH-8); 2.52–2.69 (4H, m, H-1a, H-6a, 2xH-7); 2.92 (2H, d, J 12.5 Hz, H-1b, H-6b); 3.60–3.70 (2H, m, H-3, H-4); 3.78–3.89 (2H, m, H-2, H-5); 4.66 (2H, d, J 11.2 Hz, CH2Ph); 4.79 (2H, d, J 11.2 Hz, CH2Ph); 7.29–7.43 (10H, m, H-Ph). 13C RMN (100 MHz, CDCl3): δ 13.9 (C-10); 20.4 (C-9); 29.3 (C-8); 57.6 (C-7); 59.0 (C-1; C-6); 67.7 (C-2, C-5); 73.6 (CH2Ph); 87.0 (C-3, C-4); 127.9, 128.6 (Ar); 137.0 (Cq). ESI HRMS: [M+H]+ calculated for C24H34NO4 400.2483; found 400.2475.
N-Hydroxyethyl-2,3,4,6-tetra-O-acetyl-1,5-dideoxy-1,5-imino-D-glucitol (38a):1H NMR (400 MHz, CDCl3): δ 2.29 (1H, dd, J 9.7, 11.3 Hz, H-1a); 2.37–2.51 (2H, m, H-5, H-7a); 2.94–3.08 (1H, m, H-7b); 3.16 (1H, dd, J 4.4; 11.5 Hz, H-1b); 3.43 (1H, t, J 8.3 Hz, H-3); 3.54–3.77 (4H, m, H-2, H-4, H-8a, H-8b); 3.88 (2H, qd, J 2.8; 12.3 Hz, H-6a, H-6b); 4.67–4.81 (2H, m, CH2Ph); 4.84–4.97 (2H, m, CH2Ph); 7.29–7.45 (10H, m, Ph-H). 13C RMN (100 MHz, CDCl3): δ 53.2 (C-7); 55.3 (C-1); 57.8 (C-6); 59.6 (C-8); 65.7 (C-5); 69.0 (C-2 or C-4); 74.8 (CH2Ph); 78.0 (C-2 or C-4); 85.9 (C-3); 128.0, 128.6 (Ar); 138.4 (Cq). ESI HRMS: [M+H]+ calculated for C22H30NO5 388.2119; found 388.2120.
N-Hydroxyethyl-2,3,4,6-tetra-O-acetyl-1,6-dideoxy-1,6-imino-L-iditol (38b):1H NMR (400 MHz, CDCl3): δ 2.66–2.80 (4H, m, 2xH-1, 2xH-2); 2.99 (2H, dd, J 2.8; 13.1 Hz, 2xH-7); 3.62–3.72 (4H, m, H-2, H-5, 2xH-8); 3.80–3.90 (2H, m, H-3, H-4); 4.65 (2H, d, J 11.3 Hz, CH2Ph); 4.84 (2H, d, J 11.3 Hz, CH2Ph); 7.28–7.42 (10H, m, H-Ph). 13C RMN (100 MHz, CDCl3): δ 59.0 (C-7); 59.7 (C-8); 60.7 (C-1; C-6); 70.0 (C-3; C-4); 74.2 (CH2Ph); 85.4 (C-2; C-5); 127.9, 128.0, 128.6, 137.9 (Ar). ESI HRMS: [M+H]+ calculated for C22H30NO5 388.2119; found 388.2102.
N-Propynyl-1,5-dideoxy-1,5-imino-D-glucitol (39a) [47]
N-Propynyl-1,6-dideoxy-1,6-imino-L-iditol (39b) [47]
N-Butyl-1,5-dideoxy-1,5-imino-D-glucitol (40a) [55]
N-Butyl-1,6-dideoxy-1,6-imino-L-iditol (40b): [ α ] D 22.5 1.3 (c 1.1, MeOH), 1H NMR (400 MHz, D2O): δ 0.86 (3H, t, J 8.0 Hz, 3xH-10); 1.24–1.36 (2H, m, 2xH-9); 1.56–1.74 (2H, m, 2xH-8); 3.14–3.23 (2H, m, H-7); 3.26–3.37 (4H, m, 2xH-1, 2xH-6); 3.58–3.67 (2H, m, H-2, H-5); 4.01–4.08 (2H, m, H-3, H-4). 13C RMN (100 MHz, D2O): δ 12.9 (C-10); 19.3 (C-9); 25.6 (C-8); 58.6 (C-1; C-6; C-7) 67.1 (C-3 or C-4); 67.7 (C-3 or C-4); (C-2; C-5). ESI HRMS: [M+H]+ calculated for C10H22NO4 220.1544; found 220.1545.
N-Hydroxyethyl-1,5-dideoxy-1,5-imino-D-glucitol (41a) [55]
N-Hydroxyethyl-1,6-dideoxy-1,6-imino-L-iditol (41b): [ α ] D 22.5 −9.1 (c 0.6, MeOH), 1H NMR (400 MHz, D2O): δ 2.62–2.77 (4H, m, H-1a; H-1b; H-6a; H-6b); 2.91 (2H, dd, J 3.8;13.7 Hz, H-7a, H-7b); 3.40–3.48 (2H, m, H-2, H-3); 3.69 (4H, m, H-4, H-5, H-8a, H-8b,). 13C RMN (100 MHz, D2O): δ 58.4 (C-7), 59.3 (C-1, C-6, C-8), 70.8 (C-4, C-5), 75.6 (C-2, C-3). ESI HRMS: [M+H]+ calculated for C8H18NO5 208.1185; found 208.1176.
Biological assays [43]: Yeast α-glucosidase (EC 3.2.1.20) and almond β-glucosidase (EC 3.2.1.21) activity was assessed using a 96-well plate assay. Assays contained 20 mM NaOAc at pH 6.8 (α-glucosidase) and pH 6.2 (β-glucosidase), 10 mM PIPES (piperazine-N,N′-bis(2-ethanesulfonic acid), 0.1 mM EDTA, α-glucosidase (5 μg/mL), β-glucosidase (6 μg/mL), and inhibitor (0.1–2 mM). Enzyme and inhibitor were equilibrated at 37 °C for 30 min. The reaction was initiated by the addition of p-nitrophenyl α-D-glucopyranoside (200 μM) or p-nitrophenyl β-D-glucopyranoside (200 μM), and then it was quenched with 100 μL of sodium carbonate 3.0 M after 25 min incubation at 37 °C. Assays were repeated in duplicate and data averaged.

Supplementary Materials

The following are available online at https://0-www-mdpi-com.brum.beds.ac.uk/1424-8247/12/3/108/s1, 1H, 13C and bidimensional NMR and ESI HRMS spectra of compounds 26-31a,b and 36-41a,b; Table: IC50 of final compounds and Dose-response curves obtained from Yeast α-Glucosidase and Almond β-Glucosidase assays.

Author Contributions

Conceptualization, I.C.; methodology, L.O.B.Z., V.A.-L.; software, L.O.B.Z.; validation, L.O.B.Z. and I.C.; formal analysis, L.O.B.Z. and I.C.; investigation, L.O.B.Z., Valquiria Aragão-Leoneti and I.C.; resources, L.O.B.Z., Valquiria Aragão-Leoneti and I.C.; data curation, L.O.B.Z. and I.C.; writing—original draft preparation, I.C.; writing—review and editing, L.O.B.Z. and I.C.; visualization, L.O.B.Z. and I.C.; supervision, I.C.; project administration, I.C.; funding acquisition, I.C.

Funding

This research was funded by Fundação de Amparo à Pesquisa do Estado de São Paulo (FAPESP), grant number 2007/00910-6, Conselho Nacional de Desenvolvimento Científico e Tecnológico (CNPq), ), grant number 503709/2011-5, and Coordenação de Aperfeiçoamento de Pessoal de Nível Superior (CAPES).

Acknowledgments

We acknowledge financial support from Fundação de Amparo à Pesquisa do Estado de São Paulo (FAPESP, Proc. n. 2007/00910-6), Conselho Nacional de Desenvolvimento Científico e Tecnológico (CNPq, Proc. n. 503709/2011-5), and Coordenação de Aperfeiçoamento de Pessoal de Nível Superior (CAPES).

Conflicts of Interest

The authors declare no conflict of interest. The funders had no role in the design of the study; in the collection, analyses, or interpretation of data; in the writing of the manuscript, or in the decision to publish the results.

References

  1. Melo, E.B.; Gomes, A.S.; Carvalho, I. α-and β-Glucosidase inhibitors: Chemical structure and biological activity. Tetrahedron 2006, 62, 10277–10302. [Google Scholar]
  2. Gao, K.; Zheng, C.; Wang, T.; Zhao, H.; Wang, J.; Wang, Z.; Zhai, X.; Jia, Z.; Chen, J.; Zhou, Y.; et al. Review 1-Deoxynojirimycin: Occurrence, Extraction, Chemistry, Oral Pharmacokinetics, Biological Activities and In Silico Target Fishing. Molecules 2016, 21, 1600. [Google Scholar] [CrossRef] [PubMed]
  3. Schuster, M.; Blechert, S. Inhibition of fucosyltransferase V by a GDP-azasugar. Bioorg. Med. Chem. Lett. 2001, 11, 1809–1811. [Google Scholar] [CrossRef]
  4. Jakobsen, P.; Lundbeck, J.M.; Kristiansen, M.; Breinholt, J.; Demuth, H.; Pawlas, J.; Candela, M.P.T.; Andersen, B.; Westergaard, N.; Lundgren, K.; et al. Imino sugars: Potential inhibitors of liver glycogen phosphorylase. Bioorg. Med. Chem. 2001, 9, 733–744. [Google Scholar] [CrossRef]
  5. Compain, P.; Martin, O.R. Carbohydrate mimetics-based glycosyltransferase inhibitors. Bioorg. Med. Chem. 2001, 9, 3077–3092. [Google Scholar] [CrossRef]
  6. Fedorov, A.; Shi, W.; Kicska, G.; Fedorov, E.; Tyler, P.C.; Furneaux, R.H.; Hanso, J.C.; Gainsford, G.J.; Larese, J.Z.; Schramm, V.L.; et al. Transition state structure of purine nucleoside phosphorylase and principles of atomic motion in enzymatic catalysis. Biochemistry 2001, 40, 853–860. [Google Scholar] [CrossRef] [PubMed]
  7. Lee, R.E.; Smith, M.D.; Pickering, L.; Fleet, G.W.J. An approach to combinatorial library generation of galactofuranose mimics as potential inhibitors of mycobacterial cell wall biosynthesis: Synthesis of a peptidomimetic of uridine 5′-diphosphogalactofuranose (UDP-galf). Tetrahedron Lett. 1999, 40, 8689–8692. [Google Scholar] [CrossRef]
  8. Campo, V.L.; Aragão-Leoneti, V.; Carvalho, I. Glycosidases and diabetes: Metabolic changes, mode of action and therapeutic perspectives. In Carbohydrate Chemistry; Amélia Pilar, R., Thisbe, L., Eds.; Royal Society of Chemistry: Cambridge, UK, 2013; Volume 9, pp. 181–203. [Google Scholar]
  9. Rosenbloom, B.E.; Weinreb, N.J. Gaucher disease: A comprehensive review. Crit. Rev. Oncog. 2013, 18, 163–175. [Google Scholar] [CrossRef] [PubMed]
  10. Yang, X.; Xiong, D.; Song, C.; Tai, G.; Ye, X. Synthesis of N-dialkylphosphoryl iminosugar derivatives and their immunosuppressive activities. Org. Biomol. Chem. 2015, 13, 9364–9368. [Google Scholar] [CrossRef] [PubMed]
  11. Warfield, K.; Ramstedt, U. Iminosugars as Antibacterial Compounds and Uses Thereof for Treating Bacterial Infections. PCT Int. App. WO 2014143999 A1, 18 September 2014. [Google Scholar]
  12. Alonzi, D.S.; Scott, K.A.; Dwek, R.A.; Zitzmann, N. Iminosugar antivirals: The therapeutic sweet spot. Biochem. Soc. Trans. 2017, 45, 571–582. [Google Scholar] [CrossRef]
  13. Chang, J.; Block, T.M.; Guo, J. Antiviral therapies targeting host ER alpha-glucosidases: Current status and future directions. Antivir. Res. 2013, 99, 251–260. [Google Scholar] [CrossRef] [PubMed]
  14. Fowler, P.A.; Haines, A.H.; Taylor, R.J.K.; Chrystal, E.J.T.; Gravestock, M.B. Synthesis and biological activity of acyclic analogues of nojirimycin. J. Chem. Soc. Perkin Trans. 1994, 1, 2229–2235. [Google Scholar] [CrossRef]
  15. Wetherilla, L.F.; Wassona, C.W.; Swinscoea, G.; Kealya, D.; Fosterb, R.; Griffinc, S.; Macdonald, A. Alkyl-imino sugars inhibit the pro-oncogenic ion channel function of humanpapillomavirus (HPV) E5. Antivir. Res. 2018, 158, 113–121. [Google Scholar] [CrossRef] [PubMed]
  16. Jacob, J.R.; Mansfield, K.; You, J.E.; Tennant, B.C.; Kim, Y.H. Natural iminosugar derivatives of 1-deoxynojirimycin inhibit glycosylation of hepatitis viral envelope proteins. J. Microbiol. 2007, 45, 431–440. [Google Scholar] [PubMed]
  17. Ouzounov, S.; Mehta, A.; Dwek, R.A.; Block, T.M.; Jordan, R. The combination of interferon α-2b and n-butyl deoxynojirimycin has a greater than additive antiviral effect upon production of infectious bovine viral diarrhea virus (BVDV) in vitro: Implications for hepatitis C virus (HCV) therapy. Antivir. Res. 2002, 55, 425–435. [Google Scholar] [CrossRef]
  18. Miller, J.L.; Spiro, S.G.; Dowall, S.D.; Taylor, I.; Rule, A.; Alonzi, D.S.; Sayce, A.C.; Wright, E.; Bentley, E.M.; Thom, R.; et al. In Vivo Efficacy of Iminosugars in a Lethal Ebola Virus Guinea Pig Model. PLoS ONE 2016, 11, 1–18. [Google Scholar] [CrossRef] [PubMed]
  19. Warfield, K.L.; Warren, T.K.; Qiu, X.; Wells, J.; Mire, C.; Geisbert, J.B.; Stuthman, K.S.; Garza, N.L.; Tongeren, S.A.V.; Shurtleff, A.C.; et al. Assessment of the potential for host-targeted iminosugars UV-4 and UV-5 activity against filovirus infections in vitro and in vivo. Antivir. Res. 2017, 138, 22–31. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  20. Tyrrell, B.E.; Sayce, A.C.; Warfield, K.L.; Miller, J.L.; Zitzmann, N. Iminosugars: Promising therapeutics for influenza infection. Crit. Rev. Microbiol. 2017, 43, 521–545. [Google Scholar] [CrossRef] [PubMed]
  21. Treston, A.M.; Warfield, K.L. Methods of Treating Zika Virus Infection. PCT Int. App. WO 2017201052 A1, 23 November 2017. [Google Scholar]
  22. Miller, J.L.; Tyrrell, B.E.; Zitzmann, N. Mechanisms of Antiviral Activity of Iminosugars Against Dengue Virus. In Dengue and Zika: Control and Antiviral Treatment Strategies; Rolf Hilgenfeld, R., Vasudevan, S.G., Eds.; Springer: Berlin/Heidelberg, Germany, 2018; Volume 1062, pp. 277–301. [Google Scholar]
  23. Sayce, A.C.; Alonzi, D.S.; Killingbeck, S.S.; Tyrrell, B.E.; Hill, M.L.; Caputo, A.T.; Iwaki, R.; Kinami, K.; Ide, D.; Kiappes, J.L.; et al. Iminosugars Inhibit Dengue Virus Production viaInhibition of ER Alpha-Glucosidases Not Glycolipid Processing Enzymes. PLoS Negl. Trop. Dis. 2010, 3, e0004524. [Google Scholar] [CrossRef]
  24. Tan, A.; Broek, L.V.D.; Boeckel, S.V.; Ploegh, H.; Bolscher, J.J. Chemical modification of the glucosidase inhibitor 1-deoxynojirimycin. Structure-activity relationships. Biol. Chem. 1991, 266, 14504–14510. [Google Scholar]
  25. Collins, P.; Ferrier, R. Monosaccharides: Their Chemistry and Their Roles in Natural Products; Wiley: New York, NY, USA, 1995; pp. 37–38. [Google Scholar]
  26. Sorbera, L.A.; Castaner, J.; Bayes, M. Miglustat. Drugs Fut. 2003, 28, 229–236. [Google Scholar] [CrossRef]
  27. Asano, N.; Oseki, k.; Kizu, H.; Matsui, K. Nitrogen-in-the-Ring Pyranoses and Furanoses: Structural Basis of Inhibition of Mammalian Glycosidases. J. Med. Chem. 1994, 37, 3701–3706. [Google Scholar] [CrossRef] [PubMed]
  28. Van den Broek, L.A.G.M.; Vermaas, D.J.; van Kemenade, F.J.; Tan, M.C.C.A.; Rotteveel, F.T.M.; Zandberg, P.; Butters, T.D.; Miedema, F.; Ploegh, H.L.; van Boeckel, C.A.A. Synthesis of oxygen-substituted N-alkyl 1-deoxynojirimycin derivatives: Aza sugar α-glucosidase inhibitors showing antiviral (HIV-1) and immunosuppressive activity. Recl. Trav. Chim. Pays-Bas 1994, 113, 507–516. [Google Scholar] [CrossRef]
  29. Hines, J.; Chang, H.; Gerdeman, M.S.; Warn, D.E. Isotope edited NMR studies of glycosidases: Design and synthesis of a novel glycosidase inhibitor. Bioorg. Med. Chem. Lett. 1999, 9, 1255–1260. [Google Scholar] [CrossRef]
  30. Horne, G.; Wilson, F.X.; Tinsley, J.; Williams, D.H.; Storer, R. Iminosugars past, present and future: Medicines for tomorrow. Drug Discov. Today 2011, 16, 107–118. [Google Scholar] [CrossRef]
  31. Parmeggiani, C.; Catarzi, S.; Matassini, C.; D’Adamio, G. Human Acid β-Glucosidase Inhibition by Carbohydrate Derived Iminosugars: Towards New Pharmacological Chaperones for Gaucher Disease. ChemBioChem 2015, 16, 2054–2064. [Google Scholar] [CrossRef]
  32. Trapero, A.; Llebaria, A. Glucocerebrosidase inhibitors for the treatment of Gaucher disease. Future Med. Chem. 2013, 5, 573–590. [Google Scholar] [CrossRef] [Green Version]
  33. Désiré, J.; Mondon, M.; Fontelle, N.; Nakagawa, S.; Hirokami, Y.; Adachi, I.; Iwaki, R.; Fleet, G.W.J.; Alonzi, D.S.; Twigg, G.; et al. N-and C-alkylation of seven-membered iminosugars generates potent glucocerebrosidase inhibitors and F508del-CFTR correctors. Org. Biomol. Chem. 2014, 12, 8977–8996. [Google Scholar] [CrossRef]
  34. Shih, T.L.; Liang, M.T.; Wu, K.D.; Lin, C.H. Synthesis of polyhydroxy 7-and N-alkyl-azepanes as potent glycosidase inhibitors. Carbohydr. Res. 2011, 346, 183–190. [Google Scholar] [CrossRef]
  35. Taghzouti, H.; Goumain, S.; Harakat, D.; Portella, C.; Behr, J.B.; Plantier-Royon, R. Synthesis of 2-carboxymethyl polyhydroxyazepanes and their evaluation as glycosidase inhibitors. Bioorg. Chem. 2015, 58, 11–17. [Google Scholar] [CrossRef]
  36. Nash, R.J.; Kato, A.; Yu, C.Y.; Fleet, G.W. Iminosugars as therapeutic agents: Recent advances and promising trends. Future Med. Chem. 2011, 2011 3, 513–521. [Google Scholar] [CrossRef]
  37. Brás, N.F.; Cerqueira, N.M.; Ramos, M.J.; Fernandes, P.A. Glycosidase inhibitors: A patent review (2008–2013). Expert Opin. Ther. Pat. 2014, 24, 857–874. [Google Scholar] [CrossRef] [PubMed]
  38. Asano, N. Iminosugars: The Potential of Carbohydrate Analogs. In Carbohydrate Chemistry: State of the Art and Challenges for Drug Development; Cipolla, L., Ed.; University of Milano-Bicocca: Milano, Italy, 2015; Chapter 11; pp. 279–301. [Google Scholar] [CrossRef]
  39. Wadood, A.; Ghufran, M.; Khan, A.; Azam, S.S.; Jelani, M.; Uddin, R. Selective glycosidase inhibitors: A patent review (2012–present). Int. J. Biol. Macromol. 2018, 111, 82–91. [Google Scholar] [CrossRef] [PubMed]
  40. Asano, N.; Oseki, K.; Kaneko, E.; Matsui, K. Enzymic synthesis of alpha- and beta-D-glucosides of 1-deoxynojirimycin and their glycosidase inhibitory activities. Carbohydr. Res. 1994, 258, 255–266. [Google Scholar] [CrossRef]
  41. Yoshikuni, Y.; Ezure, Y.; Seto, T.; Mori, K.; Watanabe, M.; Enomoto, H. Synthesis and alpha-glucosidase-inhibiting activity of a new alpha-glucosidase inhibitor, 4-O-alpha-D-glucopyranosylmoranoline and its N-substituted derivatives. Chem. Pharm. Bull. 1989, 37, 106–109. [Google Scholar] [CrossRef] [PubMed]
  42. Robinson, K.M.; Begovic, M.E.; Rhinehart, B.L.; Heineke, E.W.; Ducep, J.B.; Kastner, P.R.; Marshall, F.N.; Danzin, C. New potent α-glucohydrolase inhibitor MDL 73945 with long duration of action in rats. Diabetes 1991, 40, 825–830. [Google Scholar] [CrossRef]
  43. Zamoner, L.O.B.; Aragão-Leoneti, V.; Mantoani, S.P.; Rugen, M.D.; Nepogodiev, S.A.; Field, R.A.; Carvalho, I. CuAAC click chemistry with N-propargyl 1,5-dideoxy-1,5-imino-Dgulitol and N-propargyl 1,6-dideoxy-1,6-imino-D-mannitol provides access to triazole-linked piperidine and azepane pseudo-disaccharide iminosugars displaying glycosidase inhibitory properties. Carbohydr. Res. 2016, 429, 29–37. [Google Scholar] [CrossRef] [PubMed]
  44. Alonzi, D.S.; Dwek, R.A.; Butters, T.D. Improved cellular inhibitors for glycoprotein processinga-glucosidases:biological characterisation of alkyl-and arylalkyl-N-substituted deoxynojirimycins. Tetrahedron Asymmetry 2009, 20, 897–901. [Google Scholar] [CrossRef]
  45. Poitout, L.; Le Merrer, Y.; Depezay, J.C.L. Polyhydroxylated piperidines and azepanes from D-mannitol synthesis of 1-deoxynojirimycin and analogues. Tetrahedron Lett. 1994, 35, 3293–3296. [Google Scholar] [CrossRef]
  46. Le Merrer, Y.; Poitout, L.; Depezay, J.C.; Dosbaa, I.; Geoffroy, S.; Foglietti, M. Synthesis of azasugars as potent inhibitors of glycosidases. Bioorg. Med. Chem. 1997, 5, 519–533. [Google Scholar] [CrossRef]
  47. Wilkinson, B.L.; Bornaghi, L.F.; Lopez, M.; Healy, P.C.; Poulsen, S.; Houston, T.A. Synthesis of N-Propargyl Imino-Sugar Scaffolds for Compound Library Generation using Click Chemistry. Aust. J. Chem. 2010, 63, 821–829. [Google Scholar] [CrossRef]
  48. Jurczak, J.; Bauer, T.; Chmielewski, M.A. general approach to the synthesis of 2,3-di-O-protected derivatives of D-glyceraldehyde. Carbohydr. Res. 1987, 164, 493–498. [Google Scholar] [CrossRef]
  49. Aragão-Leoneti, V.; Carvalho, I. Simple and efficient synthesis of 2,5-anhydro-D-glucitol. Tetrahedron Lett. 2013, 54, 1087–1089. [Google Scholar] [CrossRef]
  50. Jung, M.E.; Lyster, M.A. Quantitative dealkylation of alkyl ethers via treatment with trimethylsilyl iodide. A new method for ether hydrolysis. J. Org. Chem. 1977, 42, 3761–3764. [Google Scholar] [CrossRef]
  51. Kasai, K.; Okada, K.; Saito, S.; Tokutake, M.; Tobe, K. Preparation of N-substituted-hexahydro-3,4,5,6-tetrahydroxyazepine as Glycosidase Inhibitors. Jpn. Kokai Tokkyo Koho JP 2001002648 A, 9 January 2001. [Google Scholar]
  52. Qian, X.; Morís-Varas, F.; Fitzgerald, M.C.; Wong, C.-H. C2-Symmetrical Tetrahydroxyazepanes as Inhibitors of Glycosidases and HIV/FIV Proteases. Bioorg. Med. Chem. 1996, 4, 2055–2069. [Google Scholar] [CrossRef]
  53. Li, H.; Liu, T.; Zhang, Y.; Favre, S.; Pierre, C.B.; Vogel, T.D.B.; Oikonomakos, N.G.; Marrot, J.; Blériot, Y. New Synthetic Seven-Membered 1-Azasugars Displaying Potent Inhibition Towards Glycosidases and Glucosylceramide Transferase. ChemBioChem 2008, 9, 253–260. [Google Scholar] [CrossRef] [PubMed]
  54. Cendret, V.; Legigan, T.; Mingot, A.; Thibaudeau, S.; Adachi, I.; Forcella, M.; Parenti, P.; Bertrand, J.; Becq, F.; Norez, C.; et al. Synthetic deoxynojirimycin derivatives bearing a thiolated, fluorinated or unsaturated N-alkyl chain: Identification of potent α-glucosidase and trehalase inhibitors as well as F508del-CFTR correctors. Org. Biomol. Chem. 2015, 13, 10734–10744. [Google Scholar] [CrossRef]
  55. Zhang, Z.X.; Wu, B.; Wang, B.; Li, T.H.; Zhang, P.F.; Guo, L.N.; Wang, W.J.; Zhao, W.; Wang, P.G. Facile and stereo-controlled synthesis of 2-deoxynojirimycin, Miglustat and Miglitol. Tetrahedron Lett. 2011, 52, 3802–3804. [Google Scholar] [CrossRef]
Figure 1. Examples of polyhydroxylated piperidine and azepane iminosugars reported, some of them displaying glucosidase inhibition.
Figure 1. Examples of polyhydroxylated piperidine and azepane iminosugars reported, some of them displaying glucosidase inhibition.
Pharmaceuticals 12 00108 g001
Scheme 1. (A) Synthesis of iminosugars 26-29a and 26-29b from D-mannitol, via bis-epoxide 25. Reagents and conditions: (i) TsCl, py, 71%; (ii) NaOH, CH3CN:H2O, 40 °C, 29%; (iii) Primary amine: propargylamine, butylamine, ethanolamine, or phenethylamine, MeOH, MW, 90 °C; (iv) Ac2O, Py; for yields over two steps see Table 1; (v) NaOMe, MeOH (quant). (B) Synthesis of iminosugars 39-41a and 39-41b from D-mannitol, via bis-epoxide 35; (vi) 2,2-dimethoxypropane, TsOH, 96%; (vii) NaH, BnBr, n-Bu4NI, THF, 93%; (viii) HCl, MeOH, 0 °C, (quant); (ix) TBDMS chloride, imidazole, DMF, 0 °C, 88%; (x) MsCl, NEt3, DCM, 92%; (xi) HCl, MeOH, then NaOH, H2O, 70%; and (xii) TMSI, DCM, rt, then MeOH, 45–100%.
Scheme 1. (A) Synthesis of iminosugars 26-29a and 26-29b from D-mannitol, via bis-epoxide 25. Reagents and conditions: (i) TsCl, py, 71%; (ii) NaOH, CH3CN:H2O, 40 °C, 29%; (iii) Primary amine: propargylamine, butylamine, ethanolamine, or phenethylamine, MeOH, MW, 90 °C; (iv) Ac2O, Py; for yields over two steps see Table 1; (v) NaOMe, MeOH (quant). (B) Synthesis of iminosugars 39-41a and 39-41b from D-mannitol, via bis-epoxide 35; (vi) 2,2-dimethoxypropane, TsOH, 96%; (vii) NaH, BnBr, n-Bu4NI, THF, 93%; (viii) HCl, MeOH, 0 °C, (quant); (ix) TBDMS chloride, imidazole, DMF, 0 °C, 88%; (x) MsCl, NEt3, DCM, 92%; (xi) HCl, MeOH, then NaOH, H2O, 70%; and (xii) TMSI, DCM, rt, then MeOH, 45–100%.
Pharmaceuticals 12 00108 sch001
Table 1. Yields obtained from microwave-assisted aminocyclization reaction of bis-epoxides 25 and 35.
Table 1. Yields obtained from microwave-assisted aminocyclization reaction of bis-epoxides 25 and 35.
Primary Amine for Aminocyclization ReactionYield (%)
Polyhydroxy-Piperidine Polyhydroxy-Azepane
1-deoxy-L-gulo-nojirimycin
26-29a
1-deoxy-D-gluco-nojirimycin (DNJ)
39-41a
D-manno-azepane
26-29b
L-ido-azepane
39-41b
Propargylamine32403537
Butylamine20332438
Ethanolamine17222128
Phenethylamine*4-5-
* Low yields obtained when the reaction mixture was purified directly by chromatographic column, without previous acetylation.
Table 2. α- and β-Glucosidase activities of synthesized iminosugars having alternative stereochemistry, ring size, and N-alkyl and N-arylalkyl chains on the endocyclic nitrogen.
Table 2. α- and β-Glucosidase activities of synthesized iminosugars having alternative stereochemistry, ring size, and N-alkyl and N-arylalkyl chains on the endocyclic nitrogen.
Iminosugars with Inverted Configuration at C-2 and C-5 with Respect to GlucoseIminosugars Preserving Glucose Stereochemistry
Inhibition (µM)Inhibition (µM)
α-Glucosidaseβ-Glucosidase α-Glucosidaseβ-Glucosidase
----DNJ Pharmaceuticals 12 00108 i001134.4 ± 2.133.1 ± 3.1
26a Pharmaceuticals 12 00108 i002NI1716 ± 12.839a Pharmaceuticals 12 00108 i0032527 ± 82.2635.7 ± 8.5
26b Pharmaceuticals 12 00108 i004NINI39b Pharmaceuticals 12 00108 i005NI3437 ± 70.6
27a Pharmaceuticals 12 00108 i006NI109.7 ± 9.340a Pharmaceuticals 12 00108 i007NI172.8 ± 1.7
27b Pharmaceuticals 12 00108 i0082031 ± 17.1184.6 ± 2.640b Pharmaceuticals 12 00108 i009NI80.0 ± 4.9
28a Pharmaceuticals 12 00108 i010NINI41a Pharmaceuticals 12 00108 i01141.3 ± 10.14.0 ± 1.5
28b Pharmaceuticals 12 00108 i012NINI41b Pharmaceuticals 12 00108 i013138.8 ± 1.24.0 ± 1.4
29a Pharmaceuticals 12 00108 i014NINI----
29b Pharmaceuticals 12 00108 i015NINI----
Enzyme inhibition: IC50 in µM, α-Glucosidase from Saccharomyces cerevisiae and β-Glucosidase from almonds. NI: no inhibition. DNJ; deoxynojirimycin

Share and Cite

MDPI and ACS Style

Zamoner, L.O.B.; Aragão-Leoneti, V.; Carvalho, I. Iminosugars: Effects of Stereochemistry, Ring Size, and N-Substituents on Glucosidase Activities. Pharmaceuticals 2019, 12, 108. https://0-doi-org.brum.beds.ac.uk/10.3390/ph12030108

AMA Style

Zamoner LOB, Aragão-Leoneti V, Carvalho I. Iminosugars: Effects of Stereochemistry, Ring Size, and N-Substituents on Glucosidase Activities. Pharmaceuticals. 2019; 12(3):108. https://0-doi-org.brum.beds.ac.uk/10.3390/ph12030108

Chicago/Turabian Style

Zamoner, Luís O. B., Valquiria Aragão-Leoneti, and Ivone Carvalho. 2019. "Iminosugars: Effects of Stereochemistry, Ring Size, and N-Substituents on Glucosidase Activities" Pharmaceuticals 12, no. 3: 108. https://0-doi-org.brum.beds.ac.uk/10.3390/ph12030108

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop