Nothing Special   »   [go: up one dir, main page]

You seem to have javascript disabled. Please note that many of the page functionalities won't work as expected without javascript enabled.
 
 
Sign in to use this feature.

Years

Between: -

Subjects

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Journals

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Article Types

Countries / Regions

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Search Results (1,087)

Search Parameters:
Keywords = 2D-NMR spectroscopy

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
20 pages, 3603 KiB  
Article
Water-Soluble Intracellular Polysaccharides (IPSW-2 to 4) from Phellinus igniarius Mycelia: Fractionation, Structural Elucidation, and Antioxidant Activity
by Isaac Duah Boateng and Xiaoming Yang
Foods 2024, 13(22), 3581; https://doi.org/10.3390/foods13223581 - 9 Nov 2024
Viewed by 526
Abstract
Phellinus igniarius is a medicinal fungus. Nonetheless, research on its water-soluble intracellular polysaccharides (IPSW-2 to 4) fractionation, structural elucidation, and antioxidant activity is limited. In this study, water-soluble intracellular polysaccharides (IPSW-2 to 4) were extracted and fractionated from P. igniarius mycelia, and their [...] Read more.
Phellinus igniarius is a medicinal fungus. Nonetheless, research on its water-soluble intracellular polysaccharides (IPSW-2 to 4) fractionation, structural elucidation, and antioxidant activity is limited. In this study, water-soluble intracellular polysaccharides (IPSW-2 to 4) were extracted and fractionated from P. igniarius mycelia, and their antioxidant and structural properties were assessed using GC-FID, GC-MS, FTIR, and NMR spectroscopy (1H and 13C). In the water-eluted P. igniarius polysaccharide fractions (IPS30W, IPS60W, and IPS80W) of anion-exchange chromatography, the polysaccharide content was 79.05%, 68.25%, and 62.06%, with higher yields of 25.07%, 21.38%, and 20.34%, respectively. In contrast, the salt (NaCl) elution fractions (IPS30S1, IPS60S1, IPS60S2, and IPS80S1) of anion-exchange chromatography had lower polysaccharide content and yield. Hence, water elution fractions (IPS30W, IPS60W, and IPS80W) were selected for further purification. After repeated purification using size-exclusion chromatography, IPSW-2 to 4 were obtained with a yield of 8% to 15.83%. The IPSW-2 to IPSW-4 structures were elucidated, and they showed no triple helical conformation. Based on periodate oxidation, Smith degradation, methylation analysis, and 1H and 13C NMR spectroscopy, the primary structures of IPSW-2, IPSW-3, and IPSW-4 were all glucan, with the main chain consisting of (1→6)-α-D-Glcp, (1→3,4)-α-D-Glcp, and (1→3, 6)-α-D-Glcp, with α-D-Glcp as a side chain. Finally, antioxidant analysis showed that IPS30W, IPS60W, and IPS80W were all more capable of scavenging superoxide anions than the polysaccharides of Phyllostachys (13.8%) and floribunda (15.1%) at the same concentration (0.40 mg/mL). This will serve as a guide for the development of functional foods. Full article
(This article belongs to the Section Food Nutrition)
Show Figures

Figure 1

Figure 1
<p>Flow diagram of the extraction, fractionation, and purification of water-soluble intracellular polysaccharides from <span class="html-italic">P. igniarius</span> mycelia.</p>
Full article ">Figure 2
<p>(<b>A</b>) Amounts of fractionated products. (<b>B</b>–<b>D</b>) The yield and content of carbohydrate and protein of (<b>B</b>) crude polysaccharides, (<b>C</b>) polysaccharides by DEAE-52 cellulose column chromatography, (<b>D</b>) polysaccharides by SephacrylTM S-400 gel-filtration column chromatography. Note: Different letters within groups indicate significant differences in the same column, <span class="html-italic">p</span> &lt; 0.05. For (<b>C</b>), the yield (%) calculation is based on dry ethanol fractional polysaccharides IPS30, IPS60, and IPS80, respectively. For (<b>D</b>), the yield (%) calculation is based on dry separation products by DEAE-52 cellulose column chromatography. IPSW-1 is from IPS30W, IPSW-2 and IPSW-3 from IPS60W, and IPSW-4 from IPS80.</p>
Full article ">Figure 3
<p>FT-IR spectrum of IPSW-2 (<b>a</b>), IPSW-3 (<b>b</b>), and IPSW-4 (<b>c</b>) before methylation and after methylation.</p>
Full article ">Figure 3 Cont.
<p>FT-IR spectrum of IPSW-2 (<b>a</b>), IPSW-3 (<b>b</b>), and IPSW-4 (<b>c</b>) before methylation and after methylation.</p>
Full article ">Figure 4
<p>a–b: NMR spectra of IPSW-2 using (<b>a</b>) <sup>1</sup>H NMR and (<b>b</b>) <sup>13</sup>C NMR. c–d: NMR spectra of IPSW-3 using (<b>c</b>) <sup>1</sup>H NMR and (<b>d</b>) <sup>13</sup>C NMR. e–f: NMR spectra of IPSW-4 using (<b>e</b>) <sup>1</sup>H NMR and (<b>f</b>) <sup>13</sup>C NMR. Note: For (<b>b</b>): A, B, C, and D are 99.70, 99.62, 95.84, and 98.55 ppm, respectively; for (<b>d</b>): A, B, and C are 99.71, 99.59, and 95.77 ppm, respectively; and for (<b>f</b>): A, B, C, and D are 99.72, 99.60, 95.81, and 98.54 ppm, respectively.</p>
Full article ">Figure 5
<p>Dependence of the λmax of Congo red analysis of various polysaccharide complexes at various concentrations of sodium hydroxide.</p>
Full article ">Figure 6
<p>a–b: Scavenging activity (% inhibition) of (<b>a</b>) IPS30, IPS60, and IPS80 and (<b>b</b>) vitamin C on superoxide anion radicals. c–d: Scavenging activity of (<b>c</b>) IPS30, IPS60, and IPS80 and (<b>d</b>) vitamin C on hydroxyl radicals. e–f: The reduction power of (<b>e</b>) IPS30, IPS60, and IPS80 and (<b>f</b>) vitamin C radical. g–h: Scavenging activity of (<b>g</b>) IPS30, IPS60, and IPS80 and (<b>h</b>) Vitamin C on DPPH.</p>
Full article ">
17 pages, 1405 KiB  
Article
Phytochemical Analysis and Biological Evaluation of Carob Leaf (Ceratonia siliqua L.) Crude Extracts Using NMR and Mass Spectroscopic Techniques
by Themistoklis Venianakis, Nikolaos Parisis, Atalanti Christou, Vlasios Goulas, Nikolaos Nikoloudakis, George Botsaris, Tjaša Goričan, Simona Golič Grdadolnik, Andreas G. Tzakos and Ioannis P. Gerothanassis
Molecules 2024, 29(22), 5273; https://doi.org/10.3390/molecules29225273 - 7 Nov 2024
Viewed by 465
Abstract
Carob leaves have gained attention for their bioactive properties and traditional medicinal uses, including as treatment for diabetes, digestive disorders, and microbial infections. The aim of this study was to explore the phytochemical composition of carob leaf acetone extracts using advanced spectroscopic techniques. [...] Read more.
Carob leaves have gained attention for their bioactive properties and traditional medicinal uses, including as treatment for diabetes, digestive disorders, and microbial infections. The aim of this study was to explore the phytochemical composition of carob leaf acetone extracts using advanced spectroscopic techniques. The combined use of heteronuclear nuclear magnetic resonance (NMR) experiments with 1D selective nuclear Overhauser effect spectroscopy (NOESY) offers detailed structural insights and enables the direct identification and quantification of key bioactive constituents in carob leaf extract. In particular, the NMR and mass spectrometry techniques revealed the presence of myricitrin as a predominant flavonoid, as well as a variety of glycosylated derivatives of myricetin and quercetin, in acetone extract. Furthermore, siliquapyranone and related gallotannins are essential constituents of the extract. The potent inhibitory effects of the carob leaf extract on Staphylococcus aureus (MIC = 50 μg mL−1) and a-glucosidase enzyme (IC50 = 67.5 ± 2.4 μg mL−1) were also evaluated. Finally, the antibacterial potency of carob leaf constituents were calculated in silico; digalloyl-parasorboside and gallic acid 4-O-glucoside exert a stronger bactericidal activity than the well-known myricitrin and related flavonoids. In summary, our findings provide valuable insights into the bioactive composition and health-promoting properties of carob leaves and highlight their potential for pharmaceutical and nutraceutical applications. Full article
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) 500 MHz <sup>1</sup>H NMR spectrum of the phenol-OH region of 20.14 mg carob leaf acetone extract in 500 μL DMSO-d<sub>6</sub>. (<b>b</b>) The same spectrum as in (<b>a</b>) after titration with trifluoroacetic acid solution in DMSO-d<sub>6</sub>. NMR spectra were acquired at 295 K and with 64 scans, 1.9 s acquisition time, and a relaxation delay of 4.0 s.</p>
Full article ">Figure 2
<p>800 MHz <sup>1</sup>H NMR spectrum of the expanded OH(5) region of flavonoids and their integrals relative to that of the major analyte myricetin-3-<span class="html-italic">O</span>-α- rhamnopyranoside (12.700 ppm).</p>
Full article ">Figure 3
<p>800 MHz <sup>1</sup>H-<sup>13</sup>C HMBC spectrum of the selected region of 20.14 mg carob leaf acetone extract in 500 μL DMSO-d<sub>6</sub>. Selected region of 800 MHz <sup>1</sup>H-<sup>13</sup>C HMBC spectrum of 20.14 mg carob leaf acetone extract in 500 μL DMSO-d<sub>6</sub>. The double arrows denote the connectivities of -OH(5) and -OH(7) with characteristic carbon atoms of myricetin-3-<span class="html-italic">O</span>-α-<span class="html-small-caps">l</span>-rhamnopyranoside. The spectrum was acquired at 298 K and with 56 scans and 1024 increments. The total experimental time was 14 h and 30 min.</p>
Full article ">Figure 4
<p>800 MHz HSQC-TOCSY spectrum of selected region of 20.14 mg carob leaf acetone extract in 500 μL DMSO-d<sub>6</sub>. Characteristic connectivities between the methyl carbon and all protons of the α-rhamnopyranoside ring are shown. The spectrum was acquired at 298 K and with 64 scans, 1024 increments, and a mixing time of 100 ms. The total experimental time was 1 day and 7 h.</p>
Full article ">Figure 5
<p>800 MHz 1D selective NOESY NMR spectrum of 20.14 mg carob leaf acetone extract in 500 μL DMSO-d<sub>6</sub>. The irradiated proton of myricetin-3-<span class="html-italic">O</span>-α-<span class="html-small-caps">l</span>-rhamnopyranoside is denoted with thunder and the protons showing connectivities are circled. The spectrum was acquired at 298 K and with 256 scans and a mixing time of 600 ms. The total experimental time was 30 min.</p>
Full article ">
15 pages, 4680 KiB  
Article
Recyclable Thermoplastic Elastomer from Furan Functionalized Hairy Nanoparticles with Polystyrene Core and Polydimethylsiloxane Hairs
by Md Hanif Uddin, Sultan Alshali, Esam Alqurashi, Saber Alyoubi, Natalia Walters and Ishrat M. Khan
Polymers 2024, 16(22), 3117; https://doi.org/10.3390/polym16223117 - 7 Nov 2024
Viewed by 473
Abstract
Polymers synthesized with end-of-life consideration allow for recovery and reprocessing. “Living-anionic polymerization (LAP)” and hydrosilylation reaction were utilized to synthesize hair-end furan functionalized hairy nanoparticles (HNPs) with a hard polystyrene (PS) core and soft polydimethylsiloxane (PDMS) hairs via a one-pot approach. The synthesis [...] Read more.
Polymers synthesized with end-of-life consideration allow for recovery and reprocessing. “Living-anionic polymerization (LAP)” and hydrosilylation reaction were utilized to synthesize hair-end furan functionalized hairy nanoparticles (HNPs) with a hard polystyrene (PS) core and soft polydimethylsiloxane (PDMS) hairs via a one-pot approach. The synthesis was carried out by first preparing the living core through crosslinking styrene with divinylbenzene using sec-butyl lithium, followed by the addition of the hexamethylcyclotrisiloxane (D3) monomer to the living core. The living polymer was terminated by dimethylchlorosilane to obtain the HNPs with Si-H functional end groups. The furan functionalization was carried out by the hydrosilylation reaction between the Si-H of the functionalized HNP and 2-vinyl furan. Additionally, furan functionalized polystyrene (PS) and polydimethylsiloxane (PDMS) were also synthesized by LAP. 1H NMR and ATR-IR spectra confirmed the successful synthesis of the target polymers. Differential scanning calorimetry showed two glass transition temperatures indicative of a polydimethylsiloxane soft phase and a polystyrene hard phase, suggesting that the HNPs are microphase separated. The furan functionalized HNPs form thermo-reversible networks upon crosslinking with bismaleimide (BMI) via a Diels−Alder coupling reaction. The kinetics of the forward Diels–Alder reaction between the functionalized polymer and BMI were studied at three different temperatures: 50 °C, 60 °C, and 70 °C by UV–Vis spectroscopy. The activation energy for the furan functionalized HNPs reaction with the bismaleimide was lower compared to the furan functionalized polystyrene and polydimethylsiloxane linear polymers. The crosslinked polymer network formed from the Diels−Alder forward reaction dissociates at around 140–154 °C, and the HNPs are recovered. The recovered HNPs can be re-crosslinked at 50 °C. The results suggest that furan functionalized HNPs are promising building blocks for preparing thermo-reversible elastomeric networks. Full article
Show Figures

Figure 1

Figure 1
<p>Schematic illustration of crosslinking and decrosslinking of furan functionalized HNPs with bismaleimide.</p>
Full article ">Figure 2
<p>ATR-FTIR spectrum of (a) HNP-SiH (black), (b) HNP-F (red).</p>
Full article ">Figure 3
<p>The 500 MHz <sup>1</sup>H NMR spectra of (<b>a</b>) Silylhydride functionalized HNP(HNP-SiH) and (<b>b</b>) Furan functionalized HNP(HNP-F).</p>
Full article ">Figure 4
<p>DLS data of HNP before (<b>a</b>) and after (<b>b</b>) functionalization.</p>
Full article ">Figure 5
<p>(<b>a</b>) TGA first-derivative curve and (<b>b</b>) DSC thermograms of HNP-F (red), PS-F (blue), PDMS-F (pink).</p>
Full article ">Figure 6
<p>(<b>a</b>) 10% (<span class="html-italic">w</span>/<span class="html-italic">w</span>) HNP-F and BMI dissolved in DMF and THF (1:1) before crosslinking; (<b>b</b>) crosslinked gel after heating for 2 h at 50 °C; (<b>c</b>,<b>d</b>) decrosslinking after heating at 150 °C for 1 min and quench cooling; (<b>e</b>,<b>f</b>) crosslinking again at 50 °C for 2 h.</p>
Full article ">Figure 7
<p>ATR-FTIR spectra of (a) crosslinked HNP-F with BMI (black) and (b) HNP-F (red).</p>
Full article ">Figure 8
<p>(<b>A</b>) UV absorbance of BMI reaction with HNP-F with time; (<b>B</b>) first-order kinetics of HNP-F with BMI; (<b>C</b>) second-order reaction kinetics of HNP-F; and (<b>D</b>) activation energies of HNP-F, PS-F, and PDMS-F reactions with BMI.</p>
Full article ">Figure 9
<p>(<b>a</b>) Heating curve of DSC analysis of crosslinked HNP-F with BMI; (<b>b</b>) heating curve of the DSC analysis of PS-F with BMI. The retro-DA peak, which indicates the cleavage of PS-BMI, is observed at 143 °C. (<b>c</b>) PDMS shows cleavage of PDMS-BMI at around 150 °C.</p>
Full article ">Figure 10
<p>SEM image of (<b>a</b>) uncrosslinked HNP-F with BMI, (<b>b</b>) crosslinked HNP-F with BMI, and (<b>c</b>) decrosslinked HNP-F after heating at 150 °C. AFM images of crosslinked HNP-F with BMI at different magnifications (<b>d</b>,<b>e</b>).</p>
Full article ">Scheme 1
<p>Synthesis of furan functionalized HNP-F.</p>
Full article ">Scheme 2
<p>The crosslinking of HNP-F with bismaleimide (BMI).</p>
Full article ">
22 pages, 3760 KiB  
Article
Synthesis and Docking Studies of Novel Spiro[5,8-methanoquinazoline-2,3′-indoline]-2′,4-dione Derivatives
by Tünde Faragó, Rebeka Mészáros, Edit Wéber and Márta Palkó
Molecules 2024, 29(21), 5112; https://doi.org/10.3390/molecules29215112 - 29 Oct 2024
Viewed by 469
Abstract
In this study, a set of spiro[5,8-methanoquinazoline-2,3′-indoline]-2′,4-dione derivatives 3ap were synthesized starting from unsubstituted and N-methyl-substituted diendo- and diexo-2-aminonorbornene carboxamides, as well as various substituted isatins. The typical method involves a condensation reaction of alicyclic aminocarboxamide and isatin [...] Read more.
In this study, a set of spiro[5,8-methanoquinazoline-2,3′-indoline]-2′,4-dione derivatives 3ap were synthesized starting from unsubstituted and N-methyl-substituted diendo- and diexo-2-aminonorbornene carboxamides, as well as various substituted isatins. The typical method involves a condensation reaction of alicyclic aminocarboxamide and isatin in the presence of a catalyst, using a solvent and an acceptable temperature. We developed a cost-effective and ecologically benign high-speed ball milling (HSBM), microwave irradiation (MW), and continuous flow (CF) technique to synthesize spiroquinazolinone molecule 3a. The structures of the synthesized compounds 3ap were determined using 1D and 2D NMR spectroscopies. Furthermore, docking studies and absorption, distribution, metabolism, and toxicity (ADMET) predictions were used in this work. In agreement with the corresponding features found in the case of both the SARS-CoV-2 main protease (RCSB Protein Data Bank: 6LU7) and human mast cell tryptase (RCSB Protein Data Bank: 2ZA5) based on the estimated total energy and binding affinity, H bonds, and hydrophobicity in silico, compound 3d among our 3ag, 3ik, and 3m derivatives was found to be our top-rated compound. Full article
Show Figures

Figure 1

Figure 1
<p>Structure of novel spiro[5,8-methanoquinazoline-2,3′-indoline]-2′,4-dione derivatives <b>3a</b>–<b>p</b>.</p>
Full article ">Figure 2
<p>NOE interactions proving the relative configuration of <b>3a</b> and <b>3e</b>. Red arrows show the detected NOESY cross-peaks; black crossed arrows represent missing NOE contacts.</p>
Full article ">Figure 3
<p>(<b>a</b>) Molecular interactions and binding pose of compound <b>3d</b> at the interface of SARS-CoV-2 main protease (PDB: 6LU7); H bonds between the macromolecule and compound <b>3d</b> are shown as yellow dashes and distances are in Å units; (<b>b</b>) Connelly surface of docking pose of 6LU7 with <b>3d</b> shown as stick model; (<b>c</b>) receptor surface: H-bond donor vs. acceptor ability of surrounding amino acids; (<b>d</b>) 2D interaction map between 6LU7 and <b>3d</b>.</p>
Full article ">Figure 4
<p>(<b>a</b>) Molecular interactions and binding pose of compound <b>3e</b> at the interface of SARS-CoV-2 main protease (PDB: 6LU7); H bonds between the macromolecule and compound <b>3e</b> are shown as yellow dashes and distances are in Å units; (<b>b</b>) Connelly surface of docking pose of 6LU7 with <b>3e</b> shown as stick model; (<b>c</b>) receptor surface: H-bond donor vs. acceptor ability of surrounding amino acids; (<b>d</b>) 2D interaction map between 6LU7 and <b>3e</b>.</p>
Full article ">Figure 5
<p>(<b>a</b>) Molecular interactions and binding pose of compound <b>3d</b> at the interface of human tryptase with potent non-peptide inhibitor (PDB: 2ZA5); H bonds between the macromolecule and compound <b>3d</b> are shown as yellow dashes and distances are in Å units; (<b>b</b>) Connelly surface of docking pose of 2ZA5 with <b>3d</b> shown as stick model; (<b>c</b>) receptor surface: H-bond donor vs. acceptor ability of surrounding amino acids; (<b>d</b>) 2D interaction map between 2ZA5 and <b>3d</b>.</p>
Full article ">Scheme 1
<p>Spirocondensation reaction of <span class="html-italic">diexo</span>- and <span class="html-italic">diendo</span>-2-aminonorbornene carboxamides <b>1a</b>–<b>d</b> with unsubstituted, 5-methyl, 5-iodo- and 7-chloro-substituted isatins <b>2a</b>–<b>d</b>.</p>
Full article ">Scheme 2
<p>The spirocondensation reaction pathway.</p>
Full article ">
14 pages, 1133 KiB  
Article
Potential Antioxidant Compounds from the Spores of Dicranopteris linearis and the Branches of Averrhoa bilimbi
by Thuc-Huy Duong, Thi-Minh-Dinh Tran, Phuong-Mai To, Nguyen-Hong-Nhi Phan, Thi-Phuong Nguyen, Huong Thuy Le and Jirapast Sichaem
Antioxidants 2024, 13(11), 1319; https://doi.org/10.3390/antiox13111319 - 29 Oct 2024
Viewed by 439
Abstract
This study focused on bio-guided isolation based on antioxidant activities from Dicranopteris linearis spores and Averrhoa bilimbi branches. The total phenolic content (TPC), total flavonoid content (TFC), and antioxidant activities of the extracts were determined. For D. linearis spores, the ethyl acetate (EA) [...] Read more.
This study focused on bio-guided isolation based on antioxidant activities from Dicranopteris linearis spores and Averrhoa bilimbi branches. The total phenolic content (TPC), total flavonoid content (TFC), and antioxidant activities of the extracts were determined. For D. linearis spores, the ethyl acetate (EA) extract exhibited the highest TPC (120.13 ± 0.04 mg GAE/g) and TFC (21.94 ± 0.30 mg QE/g), along with strong DPPH antioxidant activity (96.3 ± 0.3% inhibition, IC50 of 39.4 ± 0.3 µg/mL). For A. bilimbi branches, the n-hexane–ethyl acetate (HEA) extract showed the highest TPC (165.21 ± 0.24 mg GAE/g) and TFC (26.20 ± 0.01 mg QE/g), with significant DPPH antioxidant activity (89.6 ± 0.7% inhibition, IC50 of 39.7 ± 1.9 µg/mL). Phytochemical investigation led to the identification of ten compounds (D1D10) from D. linearis spores and twelve compounds (A1A12) from A. bilimbi branches. Notably, compound A1 was identified as a new natural compound. The chemical structures were elucidated through NMR spectroscopy and comparison with existing literature. The antioxidant activities of selected compounds (D3D5, D8D10, and A1A11) were evaluated using DPPH and ABTS free radical scavenging assays. Among them, compound A3 exhibited the strongest antioxidant activities (IC50 of 7.1 ± 0.1 µg/mL for DPPH and 14.8 ± 0.1 for ABTS, respectively). The results of this study highlight the potential of D. linearis and A. bilimbi for use in natural product-based antioxidant applications. Full article
(This article belongs to the Section Natural and Synthetic Antioxidants)
Show Figures

Figure 1

Figure 1
<p>Chemical structures of compounds <b>D1</b>–<b>D10</b> and <b>A1</b>–<b>A12</b>.</p>
Full article ">Figure 2
<p>HPLC chromatograms of the <span class="html-italic">A. bilimbi</span> HEA extract and the <span class="html-italic">D. linearis</span> EA extract, showing the presence of compounds <b>D1</b>–<b>D10</b>, <b>A3</b>, <b>A4</b>, <b>A6</b>, <b>A9</b>, and the new compound <b>A1</b>.</p>
Full article ">Figure 3
<p>Selected heteronuclear multiple bond correlations of compounds <b>A1</b> and <b>A2</b>.</p>
Full article ">
32 pages, 10320 KiB  
Article
Synthesis and Structure of Novel Hybrid Compounds Containing Phthalazin-1(2H)-imine and 4,5-Dihydro-1H-imidazole Cores and Their Sulfonyl Derivatives with Potential Biological Activities
by Łukasz Balewski, Maria Gdaniec, Anna Hering, Christophe Furman, Alina Ghinet, Jakub Kokoszka, Anna Ordyszewska and Anita Kornicka
Int. J. Mol. Sci. 2024, 25(21), 11495; https://doi.org/10.3390/ijms252111495 - 26 Oct 2024
Viewed by 587
Abstract
A novel hybrid compound—2-(4,5-dihydro-1H-imidazol-2-yl)phthalazin-1(2H)-imine (5) was synthesized and converted into di-substituted sulfonamide derivatives 6ao and phthalazine ring opening products—hydrazonomethylbenzonitriles 7am. The newly prepared compounds were characterized using elemental analyses, IR and NMR [...] Read more.
A novel hybrid compound—2-(4,5-dihydro-1H-imidazol-2-yl)phthalazin-1(2H)-imine (5) was synthesized and converted into di-substituted sulfonamide derivatives 6ao and phthalazine ring opening products—hydrazonomethylbenzonitriles 7am. The newly prepared compounds were characterized using elemental analyses, IR and NMR spectroscopy, as well as mass spectrometry. Single crystal X-ray diffraction data were collected for the representative compounds 5, 6c, 6e, 7g, and 7k. The antiproliferative activity of compound 5, sulfonyl derivatives 6ao and benzonitriles 7am was evaluated on approximately sixty cell lines within nine tumor-type subpanels, including leukemia, lung, colon, CNS, melanoma, ovarian, renal, prostate, and breast. None of the tested compounds showed any activity against the cancer cell lines used. The antioxidant properties of all compounds were assessed using the DPPH, ABTS, and FRAP radical scavenging methods, as well as the β-carotene bleaching test. Antiradical tests revealed that among the investigated compounds, a moderate ABTS antiradical effect was observed for sulfonamide 6j (IC50 = 52.77 µg/mL). Benzonitrile 7i bearing two chlorine atoms on a phenyl ring system showed activity in a β-carotene bleaching test (IC50 = 86.21 µg/mL). Finally, the interaction AGE/RAGE in the presence of the selected phthalazinimines 6a, 6b, 6g, 6m, and hydrazonomethylbenzonitriles 7a, 7cg, and 7ik was determined by ELISA assay. A moderate inhibitory potency toward RAGE was found for hydrazonomethylbenzonitriles—7d with an electron-donating methoxy group (R = 3-CH3O-C6H4) and 7f, 7k with an electron-withdrawing substituent (7f, R = 2-Cl-C6H4; 7k, R = 4-NO2-C6H4). Full article
(This article belongs to the Section Bioactives and Nutraceuticals)
Show Figures

Figure 1

Figure 1
<p>Structures of clinically approved phthalazines (AMG-900, Vatalanib) and imidazoline-fused derivative (Copanlisib) with anti-cancer activities.</p>
Full article ">Figure 2
<p>General structure of designed hybrid compound: 2-(4,5-dihydro-1<span class="html-italic">H</span>-imidazol-2-yl)phthalazin-1(2<span class="html-italic">H</span>)-imine (<b>I</b>) and its derivatives: <span class="html-italic">N</span>-(2-(1-(aryl(alkyl)sulfonyl)-4,5-dihydro-1<span class="html-italic">H</span>-imidazol-2-yl)phthalazin-1(2<span class="html-italic">H</span>)-ylidene)aryl(alkyl)sulfonamides (<b>II</b>) and 2-(((1-(aryl(alkyl)sulfonyl)imidazolidin-2-ylidene)hydrazono)methyl)benzonitriles (<b>III</b>).</p>
Full article ">Figure 3
<p><sup>1</sup>H-<sup>13</sup>C HMBC spectrum of compound <b>5</b>.</p>
Full article ">Figure 4
<p><sup>1</sup>H-<sup>1</sup>H ROESY spectrum of compound <b>5</b>.</p>
Full article ">Figure 5
<p>Molecular structure of <b>5</b>. Displacement ellipsoids are shown at the 50% probability level.</p>
Full article ">Figure 6
<p>Arrangement of molecules via π–π stacking and C-H···π interactions in <b>5</b>.</p>
Full article ">Figure 7
<p>Molecular structure of <b>6e</b> (<b>left</b>) and <b>6c</b> (<b>right</b>). Displacement ellipsoids are shown at the 50% probability level.</p>
Full article ">Figure 8
<p>Molecular structure of <b>7g</b> (<b>left</b>) and <b>7k</b> (<b>right</b>). Displacement ellipsoids are shown at the 50% probability level.</p>
Full article ">Figure 9
<p>Structures of two possible tautomers (<b>A</b>) and (<b>B</b>) of compound <b>7k</b> and relative energy (ΔE, kcal/mol) calculated in vacuum at B3LYP/6.31G** level of theory.</p>
Full article ">Figure 10
<p>Orbital diagrams of HOMO and LUMO, and Mulliken atomic charge distribution (right) for the optimized structure of tautomer B of compound <b>7k</b>.</p>
Full article ">Scheme 1
<p>Synthesis of 2-(4,5-dihydro-1<span class="html-italic">H</span>-imidazol-2-yl)phthalazin-1(2<span class="html-italic">H</span>)-imine (<b>5</b>) (isolated yields in paretheses).</p>
Full article ">Scheme 2
<p>Synthesis of <span class="html-italic">N</span>-(2-(1-(aryl(alkyl)sulfonyl)-4,5-dihydro-1<span class="html-italic">H</span>-imidazol-2-yl)phthalazin-1(2<span class="html-italic">H</span>)-ylidene)aryl(alkyl)sulfonamides <b>6a</b>–<b>o</b> and 2-(((1-(aryl(alkyl)sulfonyl)imidazolidin-2-ylidene)hydrazono)methyl)benzonitriles <b>7a</b>–<b>m</b> (isolated yields in parentheses).</p>
Full article ">Scheme 3
<p>Proposed mechanism for the formation of benzonitriles <b>7</b>.</p>
Full article ">
27 pages, 8072 KiB  
Article
Preparation of Ibuprofen-Loaded Inhalable γCD-MOFs by Freeze-Drying Using the QbD Approach
by Anett Motzwickler-Németh, Petra Party, Péter Simon, Milena Sorrenti, Rita Ambrus and Ildikó Csóka
Pharmaceutics 2024, 16(11), 1361; https://doi.org/10.3390/pharmaceutics16111361 - 24 Oct 2024
Viewed by 491
Abstract
Background/Objectives: Research on cyclodextrin-based metal-organic frameworks (CD-MOFs) is still in its infancy, but their potential for use in drug delivery—expressly in the lung—seems promising. We aimed to use the freeze-drying method to create a novel approach for preparing CD-MOFs. MOFs consisting of γ-cyclodextrin [...] Read more.
Background/Objectives: Research on cyclodextrin-based metal-organic frameworks (CD-MOFs) is still in its infancy, but their potential for use in drug delivery—expressly in the lung—seems promising. We aimed to use the freeze-drying method to create a novel approach for preparing CD-MOFs. MOFs consisting of γ-cyclodextrin (γCD) and potassium cations (K+) were employed to encapsulate the poorly water-soluble model drug Ibuprofen (IBU) for the treatment of cystic fibrosis (CF). Methods: Using the LeanQbD® software (v2022), we designed the experiments based on the Quality by Design (QbD) concept. According to QbD, we identified the three most critical factors, which were the molar ratio of the IBU to the γCD, incubation time, and the percentage of the organic solvent. light-, scanning electron microscope (SEM) and laser diffraction were utilized to observe the morphology and particle size of the samples. In addition, the products were characterized by Differential Scanning Calorimetry (DSC), X-ray Powder Diffraction (XRPD), Fourier Transform Infrared Spectroscopy (FT-IR) and nuclear magnetic resonance spectroscopy (NMR). Results: Based on characterizations, we concluded that a γCD-MOF/IBU complex was also formed using the freeze-drying method. Using formulations with optimal aerodynamic properties, we achieved 38.10 ± 5.06 and 47.18 ± 4.18 Fine Particle Fraction% (FPF%) based on the Andersen Cascade Impactor measurement. With these formulations, we achieved a fast dissolution profile and increased IBU solubility. Conclusions: This research successfully demonstrates the innovative use of freeze-drying to produce γCD-MOFs for inhalable IBU delivery. The method enabled to modify the particle size, which was crucial for successful pulmonary intake, emphasizing the need for further investigation of these formulations as effective delivery systems. Full article
(This article belongs to the Section Nanomedicine and Nanotechnology)
Show Figures

Figure 1

Figure 1
<p>The chemical structure of IBU.</p>
Full article ">Figure 2
<p>The chemical compositions of α-, β-, and γ-cyclodextrins (<b>a</b>), along with a schematic representation of CDs (<b>b</b>) and CD-MOFs (<b>c</b>).</p>
Full article ">Figure 3
<p>Ishikawa cause-effect diagram for IBU-loaded CD-MOFs.</p>
Full article ">Figure 4
<p>Schematic representation of the working principle of Andersen Cascade Impactor.</p>
Full article ">Figure 5
<p>Evaluation of correlation between CQAs-CPPs/CMAs (<b>a</b>) based on the preliminary RA, and the relative occurrence and relative severity of the factors (<b>b</b>).</p>
Full article ">Figure 6
<p>Light microscopic pictures of samples prepared by vapor diffusion method: γCD-MOFs without (<b>a</b>) and in the presence of IBU (<b>b</b>).</p>
Full article ">Figure 7
<p>SEM micrographs of freeze-dried Sample 10 (1:1 mol; 0 h; 20%) (<b>a</b>), 13 (1:1 mol; 24 h; 10%) (<b>b</b>), 11 (1:1 mol; 48 h; 0 h) (<b>c</b>), and 4 (1:2 mol; 48 h; 10%) (<b>d</b>).</p>
Full article ">Figure 8
<p>Three-dimensional surface plots of the effect of independent variables on the particle size (<b>a</b>) and Span values (<b>b</b>–<b>d</b>).</p>
Full article ">Figure 8 Cont.
<p>Three-dimensional surface plots of the effect of independent variables on the particle size (<b>a</b>) and Span values (<b>b</b>–<b>d</b>).</p>
Full article ">Figure 9
<p>XRPD patterns of raw materials and γCD-MOFs with IBU prepared by vapor diffusion and freeze-dried method.</p>
Full article ">Figure 10
<p>Three-dimensional surface plots of the effect of independent variables on the crystallinity indexes (<b>a</b>–<b>c</b>).</p>
Full article ">Figure 11
<p>DSC curves of raw IBU and γCD-MOFs containing IBU.</p>
Full article ">Figure 12
<p>FT-IR spectra of control γCD-MOFs and freeze-dried γCD-MOFs containing IBU.</p>
Full article ">Figure 13
<p>In vitro aerodynamic distribution of the freeze-dried samples.</p>
Full article ">Figure 14
<p>Release profile of raw IBU and CD-MOFs with the best MMAD and FPF% values in simulated lung media at 37 °C.</p>
Full article ">Figure 15
<p><sup>1</sup>H NMR spectrum of Sample 3 in D<sub>2</sub>O.</p>
Full article ">
19 pages, 11956 KiB  
Article
Synthesis of New Pyrazolo[3,4-d]pyrimidine Derivatives: NMR Spectroscopic Characterization, X-Ray, Hirshfeld Surface Analysis, DFT, Molecular Docking, and Antiproliferative Activity Investigations
by Mohamed El Hafi, El Hassane Anouar, Sanae Lahmidi, Mohammed Boulhaoua, Mohammed Loubidi, Ashwag S. Alanazi, Insaf Filali, Mohamed Hefnawy, Lhoussaine El Ghayati, Joel T. Mague and El Mokhtar Essassi
Molecules 2024, 29(21), 5020; https://doi.org/10.3390/molecules29215020 - 24 Oct 2024
Viewed by 804
Abstract
Four new pyrazolo[3,4-d]pyrimidines (P1P4) were successfully synthesized in good relative yields by reacting 3-methyl-1-phenyl-1H-pyrazolo[3,4-d]pyrimidin-4-ol with various alkylating agents (methyl iodide, propargyl bromide, and phenacyl bromide) at room temperature in DMF solvent, employing liquid–solid phase transfer [...] Read more.
Four new pyrazolo[3,4-d]pyrimidines (P1P4) were successfully synthesized in good relative yields by reacting 3-methyl-1-phenyl-1H-pyrazolo[3,4-d]pyrimidin-4-ol with various alkylating agents (methyl iodide, propargyl bromide, and phenacyl bromide) at room temperature in DMF solvent, employing liquid–solid phase transfer catalysis. The P1P4 structures were elucidated using NMR spectroscopy and X-ray diffraction. Intermolecular interactions in P1P4 were analyzed via Hirshfeld surface analysis and 2D fingerprint plots. Geometrical parameters were accurately modeled by DFT calculations using the B3LYP hybrid functional combined with a 6–311++G(d,p) basis set. The antiproliferative activity of P1P4 towards colorectal carcinoma (HCT 116), human hepatocellular carcinoma (HepG2), and human breast cancer (MCF-7) cell lines, along with one normal cell line (WI38) was investigated using the MTT assay and sunitinib as a reference. Compounds P1 and P2 exhibited antiproliferative activities comparable to the reference drug towards all tested cells, with an IC50 range of 22.7–40.75 µM. Both compounds also showed high selectivity indices and minimal cytotoxic effects on the normal cell line. Molecular docking revealed that the significant antiproliferative activity may attributed to the number and type of intermolecular hydrogen bonding established between pyrazolo[3,4-d]pyrimidines and DNA topoisomerase, a common target for various anticancer agents. Full article
(This article belongs to the Section Organic Chemistry)
Show Figures

Figure 1

Figure 1
<p>Chemical structures of reported pyrazolo[3,4-<span class="html-italic">d</span>]pyrimidine derivatives.</p>
Full article ">Figure 2
<p>2D structures (<b>left</b>) and perspective views of <b>P1</b>–<b>P4</b> (<b>right</b>). The dashed line in <b>P3</b> represents the intramolecular hydrogen bond.</p>
Full article ">Figure 3
<p>A portion of one layer in <b>P1</b> projected onto <math display="inline"><semantics> <mrow> <mfenced separators="|"> <mrow> <mn>10</mn> <mover accent="true"> <mrow> <mn>3</mn> </mrow> <mo>¯</mo> </mover> </mrow> </mfenced> <mo>,</mo> </mrow> </semantics></math> with the b-axis horizontal and running from left to right. N—H···O hydrogen bonds and C—H···π(ring) interactions are depicted, respectively, by violet and green dashed lines. Non-interacting hydrogen atoms are omitted for clarity.</p>
Full article ">Figure 4
<p>Elevation view of the packing in <b>P2</b> seen parallel to (<math display="inline"><semantics> <mrow> <mn>10</mn> <mover accent="true"> <mrow> <mn>1</mn> </mrow> <mo>¯</mo> </mover> </mrow> </semantics></math>) with C—H···N hydrogen bonds and C—H···π(ring) interactions depicted, respectively, by light blue and green dashed lines.</p>
Full article ">Figure 5
<p>A portion of two chains in <b>P3</b> viewed along the <span class="html-italic">a</span>-axis direction with C—H···O and C—H···N hydrogen bonds depicted, respectively, by black and light blue dashed lines. The π-stacking interactions are depicted by orange dashed lines.</p>
Full article ">Figure 6
<p>A portion of one chain of <b>P4</b> view along the <span class="html-italic">b</span>-axis. Dashed lines C—H···O represent hydrogen intermolecular bonding.</p>
Full article ">Figure 7
<p>The optimized geometries of <b>P1</b>–<b>P4</b> (<b>left</b>) and their superposition with X-ray-generated ones (<b>right</b>, the green color corresponds to the optimized geometry, and the red color corresponds to the X-ray-generated structure).</p>
Full article ">Figure 8
<p>The d<sub>norm</sub> surfaces for viewing hydrogen bonding interactions in <b>P1</b>–<b>P4</b> crystals.</p>
Full article ">Figure 8 Cont.
<p>The d<sub>norm</sub> surfaces for viewing hydrogen bonding interactions in <b>P1</b>–<b>P4</b> crystals.</p>
Full article ">Figure 9
<p>2D fingerprints of the highest intercontacts in <b>P1</b>–<b>P4</b>.</p>
Full article ">Figure 9 Cont.
<p>2D fingerprints of the highest intercontacts in <b>P1</b>–<b>P4</b>.</p>
Full article ">Figure 10
<p>Bar representation of the in vitro antiproliferative activity of <b>P1</b>–<b>P4</b> and sunitinib against HCT 116, HepG2, and MCF-7 cancer cell lines, and one normal cell line WI38.</p>
Full article ">Figure 11
<p>2D binding affinities of <b>P1</b>–<b>P4</b> into the DNA topoisomerase binding site.</p>
Full article ">Figure 11 Cont.
<p>2D binding affinities of <b>P1</b>–<b>P4</b> into the DNA topoisomerase binding site.</p>
Full article ">Figure 12
<p>3D binding affinities of <b>P1</b>–<b>P4</b> into the DNA topoisomerase binding site.</p>
Full article ">Scheme 1
<p>Synthesis path of <b>P1</b>–<b>P4</b>.</p>
Full article ">Scheme 2
<p>Tautomerism of compound <b>P1</b>.</p>
Full article ">
11 pages, 3016 KiB  
Article
Mapping Natural Sugars Metabolism in Acute Myeloid Leukaemia Using 2D Nuclear Magnetic Resonance Spectroscopy
by Christina Muhs, Islam Alshamleh, Christian Richter, Hubert Serve and Harald Schwalbe
Cancers 2024, 16(21), 3576; https://doi.org/10.3390/cancers16213576 - 23 Oct 2024
Viewed by 532
Abstract
Metabolism plays a central role in cancer progression. Rewiring glucose metabolism is essential for fulfilling the high energy and biosynthetic demands as well as for the development of drug resistance. Nevertheless, the role of other diet-abundant natural sugars is not fully understood. In [...] Read more.
Metabolism plays a central role in cancer progression. Rewiring glucose metabolism is essential for fulfilling the high energy and biosynthetic demands as well as for the development of drug resistance. Nevertheless, the role of other diet-abundant natural sugars is not fully understood. In this study, we performed a comprehensive 2D NMR spectroscopy tracer-based assay with a panel of 13C-labelled sugars (glucose, fructose, galactose, mannose and xylose). We assigned over 100 NMR signals from metabolites derived from each sugar and mapped them to metabolic pathways, uncovering two novel findings. First, we demonstrated that mannose has a semi-identical metabolic profile to that of glucose with similar label incorporation patterns. Second, next to the known role of fructose in driving one-carbon metabolism, we explained the equally important contribution of galactose to this pathway. Interestingly, we demonstrated that cells growing with either fructose or galactose became less sensitive to certain one-carbon metabolism inhibitors such as 5-Flurouracil and SHIN1. In summary, this study presents the differential metabolism of natural sugars, demonstrating that mannose has a comparable profile to that of glucose. Conversely, galactose and fructose contribute to a greater extent to one-carbon metabolism, which makes them important modulators for inhibitors targeting this pathway. To our knowledge, this is the first NMR study to comprehensively investigate the metabolism of key natural sugars in AML and cancer. Full article
(This article belongs to the Section Methods and Technologies Development)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Proliferation and metabolic adaptability of AML cells to natural sugars. (<b>A</b>) Comparisons of growth rates of 4 AML cell lines growing with different sugars over a period of 8 days. (<b>B</b>) Viability of 4 different AML cell lines on day 8 after growing in different sugars relative to day 0. (<b>C</b>) Sugar consumption in 4 different AML cell lines after 24 h of sugar switch (<span class="html-italic">n</span> = 2–3 technical replicates).</p>
Full article ">Figure 2
<p>NMR metabolic profiling of natural sugars metabolism in AML. (<b>A</b>) <sup>1</sup>H-<sup>13</sup>C 2D-HSQC example NMR spectrum of <sup>13</sup>C galactose labelling and spectral assignment in Molm13 cells. Full <sup>1</sup>H-<sup>13</sup>C 2D-HSQC of all sugars for Molm 13 is shown in <a href="#app1-cancers-16-03576" class="html-app">Supplementary Figure S4</a>. (<b>B</b>) Heat maps displaying fold change in <sup>13</sup>C intensity of selected metabolites labelled with different <sup>13</sup>C sugars (relative to glucose) in 5 AML cell lines. Full heat map is shown in <a href="#app1-cancers-16-03576" class="html-app">Supplementary Figure S1</a>. G3P (glycerol-3-phosphate), Ga1P (galactose-1-phosphate), UDP-Glc (UDP-glucose), alpha-KG (alpha-ketoglutaric acid), NAcAsp (N-acetyl aspartate), ATP (Adenosine triphosphate), NAD (nicotinamide adenine dinucleotide), UDP-Ga (uridine diphosphate galactose), GTP (guanosine triphosphate).</p>
Full article ">Figure 3
<p>Galactose and fructose support one-carbon metabolism. (<b>A</b>) Heat map of the fold change of <sup>13</sup>C label incorporation into a selection of metabolites depicting one-carbon and redox metabolism. (<b>B</b>) Overview of fructose and galactose metabolic fates based on the <sup>13</sup>C label incorporation patterns. (<b>C</b>) <sup>13</sup>C label incorporation ratio from fructose or galactose into serine and glycine in 5 different AML cell lines. (<b>D</b>) Schematic illustration of the two different settings for the contrast; (i) <sup>12</sup>C galactose + <sup>13</sup>C fructose or (ii) <sup>13</sup>C galactose + <sup>12</sup>C fructose. (<b>E</b>) Fructose and galactose consumption in a contrasting metabolic assay. (<b>F</b>) Fold change of <sup>13</sup>C label enrichment from the contrasting metabolic assay of galactose and fructose. Full bar plot is shown in <a href="#app1-cancers-16-03576" class="html-app">Supplementary Figure S3</a>.</p>
Full article ">Figure 4
<p>Galactose reduces sensitivity to one-carbon metabolism inhibitors. (<b>A</b>) IC50 curves of Molm13 and THP1 cells treated with SHIN1 and 5-FU. Cells were previously adapted for 6 days’ growth in media containing either glucose, fructose or galactose. Relative survival of Molm13 and THP1 treated with 5-FU or SHIN1 at approximately IC50 concentrations. (<b>B</b>) Schematic displaying the one-carbon metabolism inhibitors [<a href="#B19-cancers-16-03576" class="html-bibr">19</a>,<a href="#B20-cancers-16-03576" class="html-bibr">20</a>].</p>
Full article ">
12 pages, 1666 KiB  
Article
Distinct Solubilization Mechanisms of Medroxyprogesterone in Gemini Surfactant Micelles: A Comparative Study with Progesterone
by Hiromichi Nakahara, Kazutaka Koga and Keisuke Matsuoka
Molecules 2024, 29(20), 4945; https://doi.org/10.3390/molecules29204945 - 19 Oct 2024
Viewed by 562
Abstract
The solubilization behavior of medroxyprogesterone (MP) within gemini surfactant micelles (14-6-14,2Br) was investigated and compared with that of progesterone to uncover distinct solubilization mechanisms. We employed 1H-NMR and 2D ROESY spectroscopy to elucidate the spatial positioning of MP within the [...] Read more.
The solubilization behavior of medroxyprogesterone (MP) within gemini surfactant micelles (14-6-14,2Br) was investigated and compared with that of progesterone to uncover distinct solubilization mechanisms. We employed 1H-NMR and 2D ROESY spectroscopy to elucidate the spatial positioning of MP within the micelle, revealing that MP integrates more deeply into the micellar core. This behavior is linked to the unique structural features of MP, particularly its 17β-acetyl group, which promotes enhanced interactions with the hydrophobic regions of the micelle, while the 6α-methyl group interacts with the hydrophilic regions of the micelle. The 2D ROESY correlations specifically highlighted interactions between the hydrophobic chains of the surfactant and two protons of MP, H22 and H19. Complementary machine learning and electron density analyses supported these spectroscopic findings, underscoring the pivotal role of the molecular characteristics of MP in its solubilization behavior. These insights into the solubilization dynamics of MP not only advance our understanding of hydrophobic compound incorporation in gemini surfactant micelles but also indicate the potential of 14-6-14,2Br micelles for diverse drug delivery applications. Full article
(This article belongs to the Special Issue Surfactants at the Soft Interfacial Layer)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Chemical structures of (<b>a</b>) 14-6-14,2Br<sup>−</sup> and (<b>b</b>) MP. The protons of each molecule are labeled as Ha−Hf for the surfactant and H1−H22 for MP. In (<b>b</b>), MP is shown with its steroidal backbone, 17α-hydroxy group (indicated by red arrows), and 17β-acetyl group using the chair conformation notations (right) to represent its structure.</p>
Full article ">Figure 2
<p>A 2D ROESY spectrum of a 5 mM 14-6-14,2Br<sup>−</sup> solution with MP maximally solubilized in D<sub>2</sub>O at 298.2 K. Blue dashed regions show correlation signals indicating proximity and interactions between surfactant protons and MP in the micellar system.</p>
Full article ">Figure 3
<p>The ROE spectra of 2 mM (<b>bottom</b>) and 5 mM 14-6-14,2Br<sup>−</sup>/MP (<b>top</b>) of obtained sliced data using one-dimensional processing at diagonal peak H18 and diagonal peaks of 2D ROESY indicating the irradiation positions used to observe ROE correlations. For comparison, the <sup>1</sup>H-NMR spectrum of 5 mM 14-6-14,2Br<sup>−</sup>/MP is shown in the middle. Purple dotted arrows indicate the proton H18 of MP, and red solid arrows indicate the proton Ha of the surfactant.</p>
Full article ">Figure 4
<p>Structural fingerprint of MP highlighting the contribution of different structural elements to its solubility (log<span class="html-italic">S</span>) in water. The elements contributing positively to the solubility are highlighted in red, while those contributing negatively are in blue. The intensity of the color corresponds with the magnitude of the contribution. This fingerprint provides a visual representation of how different parts of the progesterone molecule influence its solubility behavior.</p>
Full article ">
16 pages, 3212 KiB  
Article
Synthesis, Characterization and Structural Study of the Two Ionic Hydrogen-Bonded Organic Frameworks Based on Sterically Crowded Bifunctional Moieties
by Kira E. Vostrikova, Vladimir P. Kirin and Denis G. Samsonenko
Chemistry 2024, 6(5), 1271-1286; https://doi.org/10.3390/chemistry6050073 - 16 Oct 2024
Viewed by 515
Abstract
Small bifunctional molecules are attractive for use as models in different areas of knowledge. How can their functional groups interact in solids? This is important to know for the prediction of the physical and chemical properties of the materials based on them. In [...] Read more.
Small bifunctional molecules are attractive for use as models in different areas of knowledge. How can their functional groups interact in solids? This is important to know for the prediction of the physical and chemical properties of the materials based on them. In this study, two new hydrogen-bonded organic frameworks (HOFs) based on sterically demanding molecular compounds, bis(1-hydroxy-2-methylpropane-2-aminium) sulfate (1) and 2-methyl-4-oxopentan-2-aminium hydrogen ethanedioate hydrate (2), were synthesized and fully characterized by means of FTIR and NMR spectroscopies, as well as by X-ray powder diffraction and thermogravimetric analyses. Their molecular and crystal structures were established through single-crystal X-ray diffraction analysis. It was shown that both compounds have a layered structure due to the formation of a 2D hydrogen-bonding network, the layers being linked by systematically arranged Van der Waals contacts between the methyl groups of organic cations. To unveil some dependencies between the chemical nature of bifunctional molecules and their solid structure, Hirschfeld surface (HS) analysis was carried out for HOFs 1, 2, and their known congeners 1-hydroxy-2-methylpropan-2-aminium hemicarbonate (3) and 1-hydroxy-2-methylpropan-2-aminium (1-hydroxy-2-methylpropan-2-yl) carbamate (4). HS was performed to quantify and visualize the close intermolecular atomic contacts in the crystal structures. It is clearly seen that H–H contacts make the highest contributions to the amino alcohol based compounds 1, 3 and 4, with a maximal value of 65.2% for compound 3 having CO32− as a counterion. A slightly lower contribution of H–H contacts (64.4%) was found for compound 4, in which the anionic part is represented by 1-hydroxy-2-methylpropan-2-yl carbamate. The significant contribution of the H–H contacts in the bifunctional moieties is due to the presence of a quaternary carbon atom with a short three-carbon chain. Full article
(This article belongs to the Section Supramolecular Chemistry)
Show Figures

Figure 1

Figure 1
<p>Asymmetric unit of the compounds: (<b>a</b>) bis(1-hydroxy-2-methylpropane-2-ammonium) sulfate (<b>1</b>); (<b>b</b>) 2-methyl-4-oxopentan-2-aminium hydrogen ethanedioate hydrate (<b>2</b>).</p>
Full article ">Figure 2
<p>Formation of a layer in the crystal of compound <b>1</b> due to hydrogen bonding (view along <span class="html-italic">b</span> axis).</p>
Full article ">Figure 3
<p>Eight types of hydrogen bonds in compound <b>1</b>. The distances of the hydrogen bonds are given between the atoms O and N or O and O in Å.</p>
Full article ">Figure 4
<p>The 2D network organization and the interlayer Van der Waals interactions, which are realized through contacts between the methyl groups of 2-methyl-4-oxopentan-2-aminium moieties in the crystal of compound <b>2</b> (view along <span class="html-italic">b</span> axis).</p>
Full article ">Figure 5
<p>Hirschfeld surface for the compounds (view along <span class="html-italic">a</span>-axis): (<b>a</b>) <b>1</b>; (<b>b</b>) <b>2</b>; (<b>c</b>) <b>3</b> [<a href="#B11-chemistry-06-00073" class="html-bibr">11</a>]; (<b>d</b>) <b>4</b> [<a href="#B12-chemistry-06-00073" class="html-bibr">12</a>].</p>
Full article ">Scheme 1
<p>Bifunctional molecules amino alcohol and amino ketone, as well as their derivatives.</p>
Full article ">Scheme 2
<p>Reaction of addition of ammonia to mesityl oxide (<b>top</b>); titration of the reaction mixture with an ethanol solution of oxalic acid (<b>bottom</b>).</p>
Full article ">
23 pages, 6558 KiB  
Article
Unravelling Different Water Management Strategies in Three Olive Cultivars: The Role of Osmoprotectants, Proteins, and Wood Properties
by Sara Parri, Claudia Faleri, Marco Romi, José C. del Río, Jorge Rencoret, Maria Celeste Pereira Dias, Sara Anichini, Claudio Cantini and Giampiero Cai
Int. J. Mol. Sci. 2024, 25(20), 11059; https://doi.org/10.3390/ijms252011059 - 15 Oct 2024
Viewed by 632
Abstract
Understanding the responses of olive trees to drought stress is crucial for improving cultivation and developing drought-tolerant varieties. Water transport and storage within the plant is a key factor in drought-tolerance strategies. Water management can be based on a variety of factors such [...] Read more.
Understanding the responses of olive trees to drought stress is crucial for improving cultivation and developing drought-tolerant varieties. Water transport and storage within the plant is a key factor in drought-tolerance strategies. Water management can be based on a variety of factors such as stomatal control, osmoprotectant molecules, proteins and wood properties. The aim of the study was to evaluate the water management strategy under drought stress from an anatomical and biochemical point of view in three young Italian olive cultivars (Giarraffa, Leccino and Maurino) previously distinguished for their physiological and metabolomic responses. For each cultivar, 15 individuals in pots were exposed or not to 28 days of water withholding. Every 7 days, the content of sugars (including mannitol), proline, aquaporins, osmotins, and dehydrins, in leaves and stems, as well as the chemical and anatomical characteristics of the wood of the three cultivars, were analyzed. ‘Giarraffa’ reduced glucose levels and increased mannitol production, while ‘Leccino’ accumulated more proline. Both ‘Leccino’ and ‘Maurino’ increased sucrose and aquaporin levels, possibly due to their ability to remove embolisms. ‘Maurino’ and ‘Leccino’ accumulated more dehydrins and osmotins. While neither genotype nor stress affected wood chemistry, ‘Maurino’ had a higher vessel-to-xylem area ratio and a larger hydraulic diameter, which allows it to maintain a high transpiration rate but may make it more susceptible to cavitation. The results emphasized the need for an integrated approach, highlighting the importance of the relative timing and sequence of each parameter analyzed, allowing, overall, to define a “strategy” rather than a “response” to drought of each cultivar. Full article
(This article belongs to the Special Issue Molecular Advances in Olive and Its Derivatives)
Show Figures

Figure 1

Figure 1
<p>Sugar levels identified by HPLC in Giarraffa (GIA), Leccino (LEC) and Maurino (MAU) under control (CTRL, black) and drought stress (DS, orange). (<b>A</b>) Glucose in leaf; (<b>B</b>) glucose in stem; (<b>C</b>) fructose in leaf; (<b>D</b>) fructose in stem; (<b>E</b>) sucrose in leaf; (<b>F</b>) sucrose in stem; (<b>G</b>) mannitol in leaf; (<b>H</b>) mannitol in stem, all expressed in mg g<sup>−1</sup> tissue dry weight (DW). Data in each column are presented as mean ± standard error. Within each time point, different letters denote statistical significance (<span class="html-italic">p</span>-value &lt; 0.05) according to Tukey’s multiple post hoc tests.</p>
Full article ">Figure 1 Cont.
<p>Sugar levels identified by HPLC in Giarraffa (GIA), Leccino (LEC) and Maurino (MAU) under control (CTRL, black) and drought stress (DS, orange). (<b>A</b>) Glucose in leaf; (<b>B</b>) glucose in stem; (<b>C</b>) fructose in leaf; (<b>D</b>) fructose in stem; (<b>E</b>) sucrose in leaf; (<b>F</b>) sucrose in stem; (<b>G</b>) mannitol in leaf; (<b>H</b>) mannitol in stem, all expressed in mg g<sup>−1</sup> tissue dry weight (DW). Data in each column are presented as mean ± standard error. Within each time point, different letters denote statistical significance (<span class="html-italic">p</span>-value &lt; 0.05) according to Tukey’s multiple post hoc tests.</p>
Full article ">Figure 2
<p>Proline content in leaves (<b>A</b>) and stems (<b>B</b>) of Giarraffa (GIA), Leccino (LEC), and Maurino (MAU) cultivars under control (CTRL, black) and drought stress (DS, orange). Contents are expressed as μg g<sup>−1</sup> tissue dry weight (DW). Values in each column are presented as mean ± standard error. Within each time point, different letters denote statistical significance (<span class="html-italic">p</span>-value &lt; 0.05) according to Tukey’s multiple post hoc tests.</p>
Full article ">Figure 3
<p>PIP1 aquaporin levels in stems of Giarraffa (GIA), Leccino (LEC) and Maurino (MAU) cultivars under control (CTRL) and drought-stress (DS) conditions, at the beginning of stress (t0), two weeks later (t2) and four weeks later (t4). (<b>A</b>) Membranes immunoblotted with anti-aquaporin antibodies from the above experimental groups; (<b>B</b>) relative blot quantification expressed as integrated density (i.d.).</p>
Full article ">Figure 4
<p>Dehydrin levels in leaves of Giarraffa (GIA), Leccino (LEC) and Maurino (MAU) cultivars after two (t2) and four (t4) weeks of stress. (<b>A</b>) Membranes immunoblotted with anti-dehydrin antibodies from the above experimental groups; (<b>B</b>) relative quantification of the blots expressed as integrated density (i.d.).</p>
Full article ">Figure 5
<p>Dehydrin levels in stems of Giarraffa (GIA), Leccino (LEC) and Maurino (MAU) cultivars under control (CTRL) and drought-stress (DS) conditions, at the beginning of stress (t0), two weeks later (t2) and four weeks later (t4). (<b>A</b>) Membranes immunoblotted with anti-dehydrin antibodies from the above experimental groups; (<b>B</b>) relative blotting quantification expressed as integrated density (i.d.).</p>
Full article ">Figure 6
<p>Osmotin levels in leaves of Giarraffa (GIA), Leccino (LEC) and Maurino (MAU) cultivars after two (t2) and four (t4) weeks of stress. (<b>A</b>) Membranes immunoblotted with anti-osmotin antibodies from the above experimental groups; (<b>B</b>) relative quantification of blotting expressed as integrated density (i.d.).</p>
Full article ">Figure 7
<p>Osmotin levels in stems of Giarraffa (GIA), Leccino (LEC) and Maurino (MAU) cultivars under control (CTRL) and drought-stress (DS) conditions, at the beginning of stress (t0), two weeks (t2) and four weeks (t4). (<b>A</b>) Membranes immunoblotted with anti-osmotin antibodies from the above experimental groups; (<b>B</b>) relative blotting quantification expressed as integrated density (i.d.).</p>
Full article ">Figure 8
<p>2D-HSQC NMR spectra of stems from three olive cultivars (Giarraffa, Leccino, and Maurino) subjected to drought stress (DS) (bottom) and their corresponding stem controls (top). The primary lignin structures identified are also shown. A: β-<span class="html-italic">O</span>-4′ alkyl-aryl ethers; B: β-5′ phenylcoumarans; C: β-β′ resinols; F: β-1′-spirodienones Cinnamyl alcohol end-groups (I), cinnamaldehyde end-groups (J), <span class="html-italic">p</span>-hydroxyphenyl units (H), guaiacyl units (G), syringyl units (S), and Cα-oxidized syringyl units (Sʹ). The yellow boxes reflect semi-quantitative estimates of lignin units and compounds. Composition is expressed in molar percent (H + G + S = 100%), and end-groups are expressed as a fraction of the total lignin inter-unit linkage types A–F.</p>
Full article ">Figure 9
<p>Stem sections of <span class="html-italic">Olea europaea</span> cultivars Giarraffa (<b>A</b>), Leccino (<b>B</b>), and Maurino (<b>C</b>). ph: phloem, x: xylem vessels; cz: cambial zone; r: parenchyma ray; f: fibers; p: paratracheal parenchyma; bar corresponds to 20 µm. (<b>D</b>) Frequency distributions (number of vessels by 5 µm diameter) of vessel lumen diameters in the three olive cultivars.</p>
Full article ">
17 pages, 2241 KiB  
Article
Antibacterial and Antioxidant Activities of Flavonoids, Phenolic and Flavonoid Glycosides from Gouania longispicata Leaves
by Hannington Gumisiriza, Eunice Apio Olet, Lydia Mwikali, Racheal Akatuhebwa, Timothy Omara, Julius Bunny Lejju and Duncan Crispin Sesaazi
Microbiol. Res. 2024, 15(4), 2085-2101; https://doi.org/10.3390/microbiolres15040140 - 11 Oct 2024
Viewed by 1004
Abstract
The leaves of Gouania longispicata Engl. (GLE) have been traditionally used to treat more than forty ailments in Uganda, including stomachache, lung and skin cancers, syphilis, toothache, and allergies. In this study, pure compounds were isolated from the methanolic extract of GLE leaves [...] Read more.
The leaves of Gouania longispicata Engl. (GLE) have been traditionally used to treat more than forty ailments in Uganda, including stomachache, lung and skin cancers, syphilis, toothache, and allergies. In this study, pure compounds were isolated from the methanolic extract of GLE leaves and their structures elucidated using ultraviolet visible spectroscopy, liquid chromatography–tandem mass spectrometry, high performance liquid chromatography, and 1D and 2D NMR techniques. The antibacterial and antioxidant activities of the compounds were assessed using the broth dilution and DPPH assays, respectively. Two known flavonoid glycosides (kaempferol-3-O-α-rhamnopyranoside and rutin), a phenolic glycoside (4,6-dihydroxy-3-methylacetophenone-2-O-β-D-glucopyranoside), and flavonoids (kaempferol and quercetin) were characterized. This is the first time that the kaempferol derivative, the acetophenone as well as free forms of quercetin, kaempferol, and rutin, are being reported in GLE and the Gouania genus. The compounds exhibited antibacterial activity against Streptococcus pneumoniae and Escherichia coli with minimum inhibitory concentrations between 16 µg/mL and 125 µg/mL. The radical scavenging activities recorded half-minimum inhibitory concentrations (IC50) ranging from 18.6 ± 1.30 µg/mL to 28.1 ± 0.09 µg/mL. The IC50 of kaempferol and quercetin were not significantly different from that of ascorbic acid (p > 0.05), highlighting their potential as natural antioxidant agents. These results lend credence to the use of GLE leaves in herbal treatment of microbial infections and oxidative stress-mediated ailments. Full article
Show Figures

Figure 1

Figure 1
<p><span class="html-italic">Gouania longispicata</span> shoot sampled from Rukungiri District, Western Uganda. Photo taken by Hannington Gumisiriza.</p>
Full article ">Figure 2
<p>Structures of compounds characterized in methanolic extract of GLE leaves.</p>
Full article ">Figure 3
<p>Visualized HMBC correlations in compound <b>1</b> isolated from methanolic extract of GLE leaves. The linkages between the sugar to the aglycone and the B-ring to the C-ring are shown.</p>
Full article ">Figure 4
<p>Minimum inhibitory concentration of isolated compounds from methanolic extract of GLE against <span class="html-italic">S. pneumoniae</span> and <span class="html-italic">E. coli</span>. Values are means of replicates (<span class="html-italic">n</span> = 3). The numbers <b>1</b>–<b>5</b> refer to the isolated compounds: kaempferol-3-<span class="html-italic">O</span>-α-rhamnopyranoside (compound <b>1</b>), 4,6-dihydroxy-3-methylacetophenone-2-<span class="html-italic">O</span>-β-D-glucopyranoside (compound <b>2</b>), kaempferol (compound <b>3</b>), quercetin (compound <b>4</b>), and rutin (compound <b>5</b>), respectively, while CIP = Ciprofloxacin.</p>
Full article ">
27 pages, 4908 KiB  
Article
Potent Biological Activity of Fluorinated Derivatives of 2-Deoxy-d-Glucose in a Glioblastoma Model
by Maja Sołtyka-Krajewska, Marcin Ziemniak, Anna Zawadzka-Kazimierczuk, Paulina Skrzypczyk, Ewelina Siwiak-Niedbalska, Anna Jaśkiewicz, Rafał Zieliński, Izabela Fokt, Stanisław Skóra, Wiktor Koźmiński, Krzysztof Woźniak, Waldemar Priebe and Beata Pająk-Tarnacka
Biomedicines 2024, 12(10), 2240; https://doi.org/10.3390/biomedicines12102240 - 1 Oct 2024
Viewed by 1375
Abstract
Background: One defining feature of various aggressive cancers, including glioblastoma multiforme (GBM), is glycolysis upregulation, making its inhibition a promising therapeutic approach. One promising compound is 2-deoxy-d-glucose (2-DG), a d-glucose analog with high clinical potential due to its ability to [...] Read more.
Background: One defining feature of various aggressive cancers, including glioblastoma multiforme (GBM), is glycolysis upregulation, making its inhibition a promising therapeutic approach. One promising compound is 2-deoxy-d-glucose (2-DG), a d-glucose analog with high clinical potential due to its ability to inhibit glycolysis. Upon uptake, 2-DG is phosphorylated by hexokinase to 2-DG-6-phosphate, which inhibits hexokinase and downstream glycolytic enzymes. Unfortunately, therapeutic use of 2-DG is limited by poor pharmacokinetics, suppressing its efficacy. Methods: To address these issues, we synthesized novel halogenated 2-DG analogs (2-FG, 2,2-diFG, 2-CG, and 2-BG) and evaluated their glycolytic inhibition in GBM cells. Our in vitro and computational studies suggest that these derivatives modulate hexokinase activity differently. Results: Fluorinated compounds show the most potent cytotoxic effects, indicated by the lowest IC50 values. These effects were more pronounced in hypoxic conditions. 19F NMR experiments and molecular docking confirmed that fluorinated derivatives bind hexokinase comparably to glucose. Enzymatic assays demonstrated that all halogenated derivatives are more effective HKII inhibitors than 2-DG, particularly through their 6-phosphates. By modifying the C-2 position with halogens, these compounds may overcome the poor pharmacokinetics of 2-DG. The modifications seem to enhance the stability and uptake of the compounds, making them effective at lower doses and over prolonged periods. Conclusions: This research has the potential to reshape the treatment landscape for GBM and possibly other cancers by offering a more targeted, effective, and metabolically focused therapeutic approach. The application of halogenated 2-DG analogs represents a promising advancement in cancer metabolism-targeted therapies, with the potential to overcome current treatment limitations. Full article
(This article belongs to the Section Cancer Biology and Oncology)
Show Figures

Figure 1

Figure 1
<p>Chemical structures of <span class="html-small-caps">d</span>-glucose, 2-DG, and its halogenated derivatives: 2-FG, 2,2-diFG, 2-BG, 2-CG.</p>
Full article ">Figure 2
<p>Viability of U-87 and U-251 cells after 72 h treatment with various concentrations of (<b>A</b>) 2-DG [0.5–20 mM] and its halogen derivatives: (<b>B</b>) 2-FG [1–10 mM], (<b>C</b>) 2,2-diFG [0.5–15 mM]. Protein synthesis inhibitor CHX [20 μM] was used as a positive cytotoxic control. Significant differences between the treatment and control means are indicated by *** <span class="html-italic">p</span> &lt; 0.001.</p>
Full article ">Figure 3
<p>Proliferation of U-87 and U-251 cells after 72 h treatment with various concentrations of (<b>A</b>) 2-DG [2.5–20 mM] and its halogen derivatives: (<b>B</b>) 2-FG [1–5 mM], (<b>C</b>) 2,2-diFG [1–10 mM]. Protein synthesis inhibitor CHX [20 μM] was used as a positive cytotoxic control. Significant differences between the treatment and control means are indicated by *** <span class="html-italic">p</span> &lt; 0.001.</p>
Full article ">Figure 4
<p>Protein synthesis of U-87 and U-251 cells after 72 h treatment with various concentrations of (<b>A</b>) 2-DG [2.5–10 mM] and its halogen derivatives: (<b>B</b>) 2-FG [1–5 mM], (<b>C</b>) 2,2-diFG [1–10 mM]. Protein synthesis inhibitor CHX [20 μM] was used as a positive cytotoxic control. Significant differences between the treatment and control means are indicated by *** <span class="html-italic">p</span> &lt; 0.001.</p>
Full article ">Figure 5
<p>Viability of U-87 and U-251 cells after 72 h treatment with various concentrations of (<b>A</b>) 2-DG [0.5–20 mM] and its halogen derivatives: (<b>B</b>) 2-FG [1–10 mM], (<b>C</b>) 2,2-diFG [0.5–15 mM] in normoxia and hypoxia-like (DMOG + Rho) conditions. Significant differences between the treatment and control means are indicated by * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01, *** <span class="html-italic">p</span> &lt; 0.001, ns—no statistical significance.</p>
Full article ">Figure 6
<p>Intracellular (cells) and extracellular (medium) lactate production of U-251 and U-87 cells after 72 h treatment with various concentrations of (<b>A</b>) 2-DG [2.5–10 mM] and its halogen derivatives: (<b>B</b>) 2-FG [1–5 mM], (<b>C</b>) 2,2-diFG [0.5–5 mM]. Significant differences between the treatment and control means are indicated by *** <span class="html-italic">p</span> &lt; 0.001.</p>
Full article ">Figure 7
<p>Viability of U-87 cells after 72 h combined treatment of CQ [10 μM] with IC<sub>50</sub> concentrations of 2-DG [5 mM], 2-FG [3 mM], and 2,2-diFG [5 mM]. Significant differences between the treatment and control means are indicated by ** <span class="html-italic">p</span> &lt; 0.01, *** <span class="html-italic">p</span> &lt; 0.001, ns—no statistical significance.</p>
Full article ">Figure 8
<p>Affinity of halogenated analogs of 2-DG to hexokinase determined by <sup>19</sup>F NMR relaxation experiment. In each case, the <span class="html-italic">K<sub>d</sub></span> was determined separately for α and β anomers, and the corresponding fitting curves are depicted as blue and orange, respectively.</p>
Full article ">Figure 9
<p>Molecular docking of 2-DG and its derivatives to crystal structure HKII (PDB entry 2NTZ). In each panel, selected hydrogen atoms and amino acid residues of HKII were omitted for clarity, making binding site visible. Typical H···O hydrogen bonds are depicted as pink dotted lines. H-bonds with less stringed geometrical constraints and close contacts involving position 2 in the pyranose ring are depicted as green dotted lines. If not mentioned otherwise only the α anomers are shown. Panels A-F show the docking of following compounds: (<b>A</b>) Glc (experimental data from literature, PDB entry 2NTZ); (<b>B</b>) 2-DG; (<b>C</b>) 2-FG, superposition of both anomers, H···O hydrogen bonds formed by the β anomer are depicted as orange dotted lines, H···F hydrogen bonds are depicted as green dotted lines; (<b>D</b>) 2,2-diFG, superposition of both anomers, H···O hydrogen coloring scheme is as in the panel C, H···F hydrogen bonds are depicted as dotted lines colored green and black for anomers α and β, respectively; (<b>E</b>) 2-CG; (<b>F</b>) 2-BG.</p>
Full article ">Figure 10
<p>Hirshfeld surface analysis of ligand-protein interactions for selected compounds (<b>A</b>) 2-DG (red—strong O···H H bonds, orange—H···H contacts), (<b>B</b>) 2-FG, anomers α and β are shown on top and bottom, respectively (dark pink—possible weak F···O halogen bond, lime—medium strong F···H H-bonds, (<b>C</b>) 2,2′-diFG, anomers α and β are shown at top and bottom, respectively (red—atomic clash with water molecules (docking artifact), dark pink—possible weak F···O halogen bond), (<b>D</b>) 2-CG (violet—possible weak Cl···O halogen bond, green—weak Cl···H H-bonds, lime—medium strong F···H H-bonds, dark pink—possible weak F···O halogen bond).</p>
Full article ">Figure 11
<p>Inhibition of HKII by 2-DG and its fluorinated-derivatives (2-FG, 2,2-diFG). The HKII activity is normalized to 1.0 in the absence of any inhibitors.</p>
Full article ">
22 pages, 3749 KiB  
Article
Isolation and Structural Characterization of Natural Deep Eutectic Solvent Lignin from Brewer’s Spent Grains
by Karina Antoun, Malak Tabib, Sarah Joe Salameh, Mohamed Koubaa, Isabelle Ziegler-Devin, Nicolas Brosse and Anissa Khelfa
Polymers 2024, 16(19), 2791; https://doi.org/10.3390/polym16192791 - 1 Oct 2024
Viewed by 751
Abstract
Brewer’s spent grains (BSG) offer valuable opportunities for valorization beyond its conventional use as animal feed. Among its components, lignin—a natural polymer with inherent antioxidant properties—holds significant industrial potential. This work investigates the use of microwave-assisted extraction combined with acidic natural deep eutectic [...] Read more.
Brewer’s spent grains (BSG) offer valuable opportunities for valorization beyond its conventional use as animal feed. Among its components, lignin—a natural polymer with inherent antioxidant properties—holds significant industrial potential. This work investigates the use of microwave-assisted extraction combined with acidic natural deep eutectic solvents (NaDESs) for efficient lignin recovery, evaluating three different NaDES formulations. The results indicate that choline chloride–lactic acid (ChCl-LA), a NaDES with superior thermal stability as confirmed via thermogravimetric analysis (TGA), is an ideal solvent for lignin extraction at 150 °C and 15 min, achieving a balance of high yield and quality. ChCl-LA also demonstrated good solubility and cell disruption capabilities, while microwaves significantly reduced processing time and severity. Under optimal conditions, i.e., 150 °C, 15 min, in the presence of ChCl-LA NaDES, the extracted lignin achieved a purity of up to 79% and demonstrated an IC50 (inhibitory concentration 50%) of approximately 0.022 mg/L, indicating a relatively strong antioxidant activity. Fourier transform infrared (FTIR) and 2D-HSQC NMR (heteronuclear single quantum coherence nuclear magnetic resonance) spectroscopy confirmed the successful isolation and preservation of its structural integrity. This study highlights the potential of BSG as a valuable lignocellulosic resource and underscores the effectiveness of acidic NaDESs combined with microwave extraction for lignin recovery. Full article
(This article belongs to the Section Biobased and Biodegradable Polymers)
Show Figures

Figure 1

Figure 1
<p>TGA curves of NaDESs used. ChCl-choline chloride, FA-formic acid, OX-oxalic acid, and LA-lactic acid.</p>
Full article ">Figure 2
<p>Effect of temperature variation on (<b>A</b>) lignin yield and (<b>B</b>) lignin purity (Klason lignin).</p>
Full article ">Figure 3
<p>Scavenging activity for extracted lignin for DPPH radical. (<b>A</b>) variation of fractionation temperature; (<b>B</b>) variation of fractionation time.</p>
Full article ">Figure 4
<p>FTIR analysis of extracted lignin. (<b>A</b>) variation in fractionation temperature. (<b>B</b>) variation in fractionation duration.</p>
Full article ">Figure 4 Cont.
<p>FTIR analysis of extracted lignin. (<b>A</b>) variation in fractionation temperature. (<b>B</b>) variation in fractionation duration.</p>
Full article ">Figure 5
<p>2D-HSQC NMR spectrum of extracted lignin with variation in the temperature; 130 °C, 150 °C, and 170 °C (15 min). (<b>A</b>–<b>C</b>): side-chain regions: (<b>A′</b>–<b>C′</b>): aromatic regions.</p>
Full article ">Figure 6
<p>2D-HSQC NMR spectrum of lignin with variation of the extraction duration from 10 to 30 min (150 °C). (<b>A</b>–<b>E</b>): Side-chain regions: (<b>A′</b>–<b>E′</b>): Aromatic regions.</p>
Full article ">Figure 6 Cont.
<p>2D-HSQC NMR spectrum of lignin with variation of the extraction duration from 10 to 30 min (150 °C). (<b>A</b>–<b>E</b>): Side-chain regions: (<b>A′</b>–<b>E′</b>): Aromatic regions.</p>
Full article ">Figure 7
<p>Main lignin substructures cited in the spectra of <a href="#polymers-16-02791-f005" class="html-fig">Figure 5</a> and <a href="#polymers-16-02791-f006" class="html-fig">Figure 6</a>: (A) β-O-4′ alkyl-aryl ethers; (A′) acetylated β-O-4′ substructures; (B) phenylcoumarans; (C) resinols; (I) p-hydroxycinnamyl alcohol end groups; (S) syringyl units; (H) p-hydroxyphenyl units; (G) guaiacyl units; (FA) formic acid; (PCA) p-coumarates.</p>
Full article ">
Back to TopTop