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

Journals

Article Types

Countries / Regions

Search Results (40)

Search Parameters:
Keywords = host–guest polymer complex

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
18 pages, 4409 KiB  
Review
Supramolecular Chemistry of Polymer-Based Molecular Tweezers: A Minireview
by Bahareh Vafakish and Lee D. Wilson
Surfaces 2024, 7(3), 752-769; https://doi.org/10.3390/surfaces7030049 - 14 Sep 2024
Viewed by 701
Abstract
Polymer-based molecular tweezers have emerged as a prominent research area due to their enhanced ability to form host–guest complexes, driven by advancements in their design and synthesis. The impact of the spacer structure on the tweezers is predominant. They can be rigid, flexible, [...] Read more.
Polymer-based molecular tweezers have emerged as a prominent research area due to their enhanced ability to form host–guest complexes, driven by advancements in their design and synthesis. The impact of the spacer structure on the tweezers is predominant. They can be rigid, flexible, and stimuli-responsive. Herein, a new generation of molecular tweezers is introduced as polymer-based molecular tweezers. The integration of molecular tweezers onto biopolymers has significantly expanded their potential applications, making them promising candidates, especially in drug delivery, owing to their biocompatibility, adaptive structural features, and versatile interaction capabilities. The unique structure of polymer-based molecular tweezers, particularly when integrated with biopolymers, creates a unique nano-environment that enhances their interaction with guest molecules. This minireview focuses on the synthesis and applications of polymer-based molecular tweezers and examines how the incorporation of various spacers affects their binding affinity and specificity. These features highlight the advancement of these polymer-based systems, emphasizing their potential applications, particularly in drug delivery, water treatment technology, and future research opportunities. Full article
(This article belongs to the Collection Featured Articles for Surfaces)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>The molecular structure of some well-known molecular tweezers synthesized by different research groups. Redrawn from references [<a href="#B7-surfaces-07-00049" class="html-bibr">7</a>,<a href="#B13-surfaces-07-00049" class="html-bibr">13</a>,<a href="#B14-surfaces-07-00049" class="html-bibr">14</a>,<a href="#B17-surfaces-07-00049" class="html-bibr">17</a>,<a href="#B19-surfaces-07-00049" class="html-bibr">19</a>,<a href="#B20-surfaces-07-00049" class="html-bibr">20</a>], respectively.</p>
Full article ">Figure 2
<p>Chemical structure of water-soluble Zimmerman molecular tweezers on the dextran scaffold. Redrawn from reference [<a href="#B47-surfaces-07-00049" class="html-bibr">47</a>].</p>
Full article ">Figure 3
<p>Maitra’s molecular tweezers immobilized on the polystyrene. The image illustrates the attachment of the tweezers to a polystyrene bead. Redrawn from reference [<a href="#B19-surfaces-07-00049" class="html-bibr">19</a>].</p>
Full article ">Figure 4
<p>(<b>a</b>) Bis-pyrenyl tweezers on a polyamide scaffold appropriate for aromatic guests. Redrawn from reference [<a href="#B49-surfaces-07-00049" class="html-bibr">49</a>]. (<b>b</b>) The UV-visible spectra of host–guest blend reveals a prominent absorption band at 526 nm, indicating charge transfer between them. The blue line is a pure diimide guest, the red line is a polymer-based host, and the black line is a host–guest blend. Reprinted with permission from reference [<a href="#B49-surfaces-07-00049" class="html-bibr">49</a>].</p>
Full article ">Figure 5
<p>Structure of zinc porphyrin tweezers immobilized on TentaGel polymer (cross-linked polystyrene network with grafted polyethylene glycol units). Redrawn from reference [<a href="#B51-surfaces-07-00049" class="html-bibr">51</a>].</p>
Full article ">Figure 6
<p>(<b>a</b>) U- and W-shape of PEG-bound molecular tweezers in acidic and neutral pH. Adapted with permission from reference [<a href="#B52-surfaces-07-00049" class="html-bibr">52</a>]. (<b>b</b>) Chemical structure of mitoxantrone which was used as the guest. The arrow shows decrease of fluorescence intensity. (<b>c</b>) Emission fluorescence of polymer based tweezers upon addition of mitoxantrone at neutral (left) and acidic pH (right). Reprinted with permission from reference [<a href="#B52-surfaces-07-00049" class="html-bibr">52</a>].</p>
Full article ">Figure 7
<p>Water-soluble metal-bis-porphyrin end-capped with PEG. Adapted with permission from reference [<a href="#B56-surfaces-07-00049" class="html-bibr">56</a>].</p>
Full article ">Figure 8
<p>(<b>a</b>) Structure of three different metalloporphyrin–peptoid conjugates (MPPCs) and (<b>b</b>) host–guest complex formation between MPPCs guests as follows: achiral host–chiral guest, chiral host–chiral guest, and chiral host–chiral guest. Adapted with permission from reference [<a href="#B58-surfaces-07-00049" class="html-bibr">58</a>].</p>
Full article ">Figure 9
<p>Schematic proposed structure of the stimuli-responsive catalyst. Formation of hybrid metal nanomaterial, PAAgCHI/Fe<sub>3</sub>O<sub>4</sub>/Ag (L, H), where L and H refer to low and high levels of Ag nanoparticle loadings. The following color scheme defines the various components: red sphere (Fe<sub>3</sub>O<sub>4</sub>), gray sphere (Ag NPs), blue line (chitosan), and green line (grafted PAA; PAAg). Copied with permission from Ref. [<a href="#B65-surfaces-07-00049" class="html-bibr">65</a>].</p>
Full article ">Figure 10
<p>Schematic illustration of the dual-function adsorption properties on surface-patterned flax with variable chitosan loading. The surface properties vary from a predominantly positive charge due to chitosan surface coating effects to a predominantly negative surface charge for pristine flax fiber (FFR; without the chitosan coating). The arrows in the lower panel highlight the incremental adsorption of RB (right to left), whereas MB increases from left to right. Copied with permission from ref. [<a href="#B77-surfaces-07-00049" class="html-bibr">77</a>].</p>
Full article ">Scheme 1
<p>Schematic representation of three different molecular tweezers based on the type of the spacer, including (<b>a</b>) flexible, (<b>b</b>) rigid, and (<b>c</b>) stimuli-responsive tweezers activated by external triggers, such as light or other forms of stimulation denoted by the zig-zag line. Adapted with permission from reference [<a href="#B4-surfaces-07-00049" class="html-bibr">4</a>].</p>
Full article ">Scheme 2
<p>The schematic arrangement of tweezers (red) and guest (green) parts to show the supramolecular polymers connected by noncovalent interactions.</p>
Full article ">Scheme 3
<p>Schematic illustration of the stepwise synthesis process of SPS. Copied with permission from ref. [<a href="#B78-surfaces-07-00049" class="html-bibr">78</a>].</p>
Full article ">
33 pages, 17015 KiB  
Review
The Many Faces of Cyclodextrins within Self-Assembling Polymer Nanovehicles: From Inclusion Complexes to Valuable Structural and Functional Elements
by Ivana Jarak, Sara Ramos, Beatriz Caldeira, Cátia Domingues, Francisco Veiga and Ana Figueiras
Int. J. Mol. Sci. 2024, 25(17), 9516; https://doi.org/10.3390/ijms25179516 - 1 Sep 2024
Viewed by 1055
Abstract
Most chemotherapeutic agents are poorly soluble in water, have low selectivity, and cannot reach the tumor in the desired therapeutic concentration. On the other hand, sensitive hydrophilic therapeutics like nucleic acids and proteins suffer from poor bioavailability and cell internalization. To solve this [...] Read more.
Most chemotherapeutic agents are poorly soluble in water, have low selectivity, and cannot reach the tumor in the desired therapeutic concentration. On the other hand, sensitive hydrophilic therapeutics like nucleic acids and proteins suffer from poor bioavailability and cell internalization. To solve this problem, new types of controlled release systems based on nano-sized self-assemblies of cyclodextrins able to control the speed, timing, and location of therapeutic release are being developed. Cyclodextrins are macrocyclic oligosaccharides characterized by a high synthetic plasticity and potential for derivatization. Introduction of new hydrophobic and/or hydrophilic domains and/or formation of nano-assemblies with therapeutic load extends the use of CDs beyond the tried-and-tested CD-drug host–guest inclusion complexes. The recent advances in nano drug delivery have indicated the benefits of the hybrid amphiphilic CD nanosystems over individual CD and polymer components. This review provides a comprehensive overview of the most recent advances in the design of CDs self-assemblies and their use for delivery of a wide range of therapeutic molecules. It aims to offer a valuable insight into the many roles of CDs within this class of drug nanocarriers as well as current challenges and future perspectives. Full article
(This article belongs to the Special Issue Cyclodextrins for Drug/iRNA Co-delivery)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Representation of βCD.</p>
Full article ">Figure 2
<p>Influence of structural parameters on nanoassembly morphology illustrated by example of acylated cyclodextrins (β-CD-C10): (<b>A</b>) influence of critical packing parameters on morphology; (<b>B</b>) cryo TEM images of non-PEGylated (<b>left</b>) and PEGylated (<b>right</b>) β-CD-C10. Reproduced with permission from [<a href="#B21-ijms-25-09516" class="html-bibr">21</a>].</p>
Full article ">Figure 3
<p>Amphiphilic βCD-g-PMA-co-PLGA for co-delivery of drug and adjuvant. (<b>A</b>) Schematic representation of polymer synthetic sequence; (<b>B</b>) schematic representation of drug/adjuvant activity and delivery; (<b>C</b>) quantitative diagram of therapeutic outcome (cell population distribution) after treatment. PB blank micelles; B2D coloaded micelles; BD and BC singly loaded micelles; Dox and Con free drugs; 2D free drug combination. Reproduced with permission from [<a href="#B57-ijms-25-09516" class="html-bibr">57</a>].</p>
Full article ">Figure 4
<p>Schematic representation of dandelion-like CD-based micelles for dual chemo and phototherapy: (<b>A</b>) βCD-NHC<sub>12</sub>H<sub>25</sub>/HSPC self-assembly and dandelion shell growth; (<b>B</b>) stimuli-sensitive degradation of camptothecin-CD/camtothecin-cRGDfk prodrugs and drug release; (<b>C</b>) active targeting of cells with high integrin αvβ3 expression is followed by GSH-induced vesicle degradation and release of active ingredients. Photodynamic therapy complements chemotherapeutic activity upon irradiation. Reproduced with permission from [<a href="#B69-ijms-25-09516" class="html-bibr">69</a>].</p>
Full article ">Figure 5
<p>Schematic representation of multifunctional mPEG-PLGA poly(pseudo)rotaxane for delivery of Oxa/anti-inflammatory prodrug: (<b>A</b>) PPR self-assembly; (<b>B</b>) stimuli-sensitive release of Oxa and DN, and GSH trapping by αCD-itaconate. Reproduced with permission from [<a href="#B52-ijms-25-09516" class="html-bibr">52</a>].</p>
Full article ">Figure 6
<p>Self-assembled CDs for protein delivery: (<b>A</b>) scheme of a simple synthetic workflow for protein derivatization and functionalization via host–guest insertion. The structure of derivatized CD is presented; (<b>B</b>) confocal micrographs and flow cytometry histograms of cellular uptake of proteins (β-galactosidase and IgG) modified with Ad-NHS and R8-CDOH. Reproduced with permission from [<a href="#B80-ijms-25-09516" class="html-bibr">80</a>]. Copyright 2020 American Chemical Society.</p>
Full article ">Figure 7
<p>Polyion-based CD nano-assembly for intracellular protein delivery: (<b>A</b>) schematic representation of CD derivatization with cationic ligands; (<b>B</b>) formation of protein-CD nanoassembly; (<b>C</b>) intracellular distribution of protein-CD complexes; (<b>D</b>) molecular modelling of interactions between nuclear transported importin and CD ligands; (<b>E</b>) scheme of intracellular uptake and transport of CD-protein complex. Reproduced with permission from [<a href="#B83-ijms-25-09516" class="html-bibr">83</a>]. Copyright 2022 Elsevier.</p>
Full article ">Figure 8
<p>Vesicle-shaped self-assembly of amphiphilic NH<sub>2</sub>Pr-CD-C12 for intracellular siRNA delivery: (<b>A</b>) Structure of amphiphilic CD and derivatization of formed vesicles; (<b>B</b>–<b>D</b>) in vivo biodistribution of siRNA-loaded NH<sub>2</sub>Pr-CD-C12. Reproduced with permission from [<a href="#B93-ijms-25-09516" class="html-bibr">93</a>] under the CC-BY 4.0 license. ** <span class="html-italic">p</span> &lt; 0.01.</p>
Full article ">
18 pages, 7519 KiB  
Review
Recent Progress in Solid-State Room Temperature Afterglow Based on Pure Organic Small Molecules
by Xin Shen, Wanhua Wu and Cheng Yang
Molecules 2024, 29(13), 3236; https://doi.org/10.3390/molecules29133236 - 8 Jul 2024
Cited by 1 | Viewed by 1130
Abstract
Organic room temperature afterglow (ORTA) can be categorized into two key mechanisms: continuous thermally activated delayed fluorescence (TADF) and room-temperature phosphorescence (RTP), both of which involve a triplet excited state. However, triplet excited states are easily quenched by non-radiative transitions due to oxygen [...] Read more.
Organic room temperature afterglow (ORTA) can be categorized into two key mechanisms: continuous thermally activated delayed fluorescence (TADF) and room-temperature phosphorescence (RTP), both of which involve a triplet excited state. However, triplet excited states are easily quenched by non-radiative transitions due to oxygen and molecular vibrations. Solid-phase systems provide a conducive environment for triplet excitons due to constrained molecular motion and limited oxygen permeation within closely packed molecules. The stimulated triplet state tends to release energy through radiative transitions. Despite numerous reports on RTP in solid-phase systems in recent years, the complexity of these systems precludes the formulation of a universal theory to elucidate the underlying principles. Several strategies for achieving ORTA luminescence in the solid phase have been developed, encompassing crystallization, polymer host-guest doping, and small molecule host-guest doping. Many of these systems exhibit luminescent responses to various physical stimuli, including light stimulation, mechanical stimuli, and solvent vapor exposure. The appearance of these intriguing luminescent phenomena in solid-phase systems underscores their significant potential applications in areas such as light sensing, biological imaging, and information security. Full article
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) Simplified Jablonski diagram for host-guest ORTA multiple luminescence; (<b>b</b>) intermolecular electron cloud overlap diagram and host-guest doped material interaction model.</p>
Full article ">Figure 2
<p>(<b>a</b>) Structures of BP and its derivatives; (<b>b</b>) PL spectra of BP in THF solution and crystalline state at room temperature and 77 K (left) and phosphorescence decay curve of BP crystals at room temperature. (<b>c</b>) Photographs of crystals of DCBP, DBBP, BBP, and ABP taken under normal laboratory lighting and 365-nm UV light illumination at room temperature. (<b>d</b>) the corresponding emission spectra of prompt fluorescence (dotted line) and phosphorescence (solid lines). Reproduced with permission from ref. [<a href="#B20-molecules-29-03236" class="html-bibr">20</a>] Copyright © 2024, American Chemical Society.</p>
Full article ">Figure 3
<p>(<b>a</b>) Chemical structures of Br6A (1) and Br6 (2). (<b>b</b>) Diagram of Br6 crystal packing.</p>
Full article ">Figure 4
<p>The chemical structures of the luminous small molecule guest used for polymers host doping system and in situ polymerization with MMA.</p>
Full article ">Figure 5
<p>(<b>a</b>) Molecule design of NIC-DMAC; (<b>b</b>) natural transition orbitals of S<sub>1</sub>, T<sub>1,,</sub> and T<sub>2</sub> states for NIC-DMAC calculated in the gas phase by the TD-M062X/6–31g(d) method, blue for hole and orange for particle; SOCME (ζ) were calculated in the gas phase at the M062X/def2-tzvp level and the proposed mechanism for the ISC and RIC processes of NIC-DMAC. Reproduced with permission from ref. [<a href="#B23-molecules-29-03236" class="html-bibr">23</a>] Copyright © 2024 Wiley-VCH GmbH.</p>
Full article ">Figure 6
<p>Construction strategy for efficient dark blue room-temperature phosphorescence with tunable lifetime. Reproduced with permission from ref. [<a href="#B33-molecules-29-03236" class="html-bibr">33</a>] Copyright © 2024 Wiley-VCH GmbH.</p>
Full article ">Figure 7
<p>(<b>a</b>) Temperature-dependent afterglow of the ICZ-p1-0.5 wt.%-PMMA film and the corresponding temperature-dependent color chart for comparison temperature sensing. (<b>b</b>) Afterglow spectra at different UV light irradiation times of this film at 298 K in air. Reproduced with permission from ref. [<a href="#B32-molecules-29-03236" class="html-bibr">32</a>] Copyright © 2024 Wiley-VCH GmbH.</p>
Full article ">Figure 8
<p>(<b>a</b>) Molecular and single-crystal structures of mono-DMACDPS and Me-DMACDPS; Normalized PL spectra of mono-DMACDPS and Me-DMACDPS in the crystal at 300 K. Transient photoluminescence spectra of mono-DMACDPS and Me-DMACDPS in the crystal: image of mono-DMACDPS crystal with 365 nm lamp on and off. Reproduced with permission from ref. [<a href="#B21-molecules-29-03236" class="html-bibr">21</a>] Copyright © 2024 Wiley-VCH Verlag GmbH &amp; Co. KGaA, Weinheim (<b>b</b>) Exciplex-induced TADF, persistent RTP, and ML in a doping system with OPTZ-CN as the host and PTZ-CN as the guest. (<b>c</b>) Molecular structures and temperature-dependent steady-state PL spectra of bTEoCN, bTENCo, and bTEpCN. Reproduced with permission from ref. [<a href="#B41-molecules-29-03236" class="html-bibr">41</a>] Copyright © 2024 Wiley-VCH GmbH.</p>
Full article ">Figure 9
<p>(<b>a</b>) Molecular structure and single crystals of the G-crystal Y-crystal, and R-crystal of DMAC-PSBF2. (<b>b</b>) The molecular structure and two possible conformations of CzS-ph-3F and a proposed diagram for the stimulus-responsive RTP effect based on conformational changes and photographs of Czs-ph-3F taken before and after turning off 365 nm UV irradiation under ambient conditions. Reproduced with permission from ref. [<a href="#B79-molecules-29-03236" class="html-bibr">79</a>] Copyright © 2024 Wiley-VCH GmbH. (<b>c</b>) Molecular structures of TA analogs and images of TA analogs:PMMA films under 254 nm hand lamp, before and after elimination of oxygen by nitrogen purging, respectively, the numbers 1 and 2 in molecule TA means the different reactive sites). Reproduced with permission from ref. [<a href="#B74-molecules-29-03236" class="html-bibr">74</a>] Copyright © 2024 Wiley-VCH GmbH. (<b>d</b>) Rational design of molecular rotors, molecular packing modes in crystals, and proposed mechanism for dynamic ultralong organic phosphorescence (step 1: intersystem crossing; step 2: photo-activation, and step 3: thermally deactivation). Reproduced with permission from ref. [<a href="#B71-molecules-29-03236" class="html-bibr">71</a>] Copyright © 2024 Wiley-VCH Verlag GmbH &amp; Co. KGaA, Weinheim.</p>
Full article ">Figure 10
<p>The adjacent dimers in single crystals of DMAC-PSBF2 (G-crystal, Y-crystal, and R-crystal).</p>
Full article ">Figure 11
<p>Calculated SOC constants (ξ) between the S<sub>0</sub>/S<sub>1</sub> and T<sub>n</sub> (Δ<span class="html-italic">E</span><sub>S1Tn</sub> &lt; 0.3 eV) for Czs-ph-3F(ax.) and Czs-ph-3F(eq.). Reproduced with permission from ref. [<a href="#B79-molecules-29-03236" class="html-bibr">79</a>] Copyright © 2024 Wiley-VCH GmbH.</p>
Full article ">Figure 12
<p>Intermolecular interactions and twisting angles of BCzT in a single crystal before and after photoactivation (BCzT-a) at 100 K. Reproduced with permission from ref. [<a href="#B71-molecules-29-03236" class="html-bibr">71</a>] Copyright © 2024 Wiley-VCH Verlag GmbH &amp; Co. KGaA, Weinheim.</p>
Full article ">
12 pages, 2872 KiB  
Article
Nanogels Based on N,N-Dimethylacrylamide and β-Cyclodextrin Triacrylate for Enhanced Solubility and Therapeutic Efficacy of Aripiprazole
by Siyka Stoilova, Dilyana Georgieva, Rositsa Mihaylova, Petar D. Petrov and Bistra Kostova
Gels 2024, 10(4), 217; https://doi.org/10.3390/gels10040217 - 22 Mar 2024
Cited by 1 | Viewed by 1712
Abstract
Aripiprazole (ARZ) is a medication used for the treatment of various diseases such as schizophrenia, bipolar disorder, major depressive disorder, autism, and Tourette’s syndrome. Despite its therapeutic benefits, ARZ is characterized by a poor water solubility which provoked the development of various delivery [...] Read more.
Aripiprazole (ARZ) is a medication used for the treatment of various diseases such as schizophrenia, bipolar disorder, major depressive disorder, autism, and Tourette’s syndrome. Despite its therapeutic benefits, ARZ is characterized by a poor water solubility which provoked the development of various delivery systems in order to enhance its solubility. In the present work, a nanoscale drug delivery system based on N,N-dimethylacrylamide (DMAA) and β-cyclodextrin triacrylate (β-CD-Ac3) as potential aripiprazole delivery vehicles was developed. The nanogels were synthesized by free radical polymerization of DMAA in the presence of β-CD-Ac3 as a crosslinking agent and then loaded with ARZ via host-guest inclusion complexation. The blank- and drug-loaded nanogels were evaluated using different methods. Fourier transform infrared (FTIR) spectroscopy was employed to confirm the incorporation of β-CD moieties into the polymer network. Dynamic light scattering (DLS) was used to study the size of the developed systems. The samples exhibited a monomodal particle size distribution and a relatively narrow dispersity index. The hydrodynamic diameter (Dh) of the gels varied between 107 and 129 nm, with a tendency for slightly larger particles as the β-CD-Ac3 fraction increased. Loading the drug into the nanocarrier resulted in slightly larger particles than the blank gels, but their size was still in the nanoscopic range (166 to 169 nm). The release profiles in PBS were studied and a sustained release pattern with no significant burst effect was observed. A cytotoxicity assessment was also conducted to demonstrate the non-toxicity and biocompatibility of the studied polymers. Full article
(This article belongs to the Special Issue Gel-Based Materials: Preparations and Characterization (2nd Edition))
Show Figures

Figure 1

Figure 1
<p>Synthetic scheme of preparation of the poly (N,N-dimethylacrylamide)/β-cyclodextrin triacrylate nanogel.</p>
Full article ">Figure 2
<p>FTIR spectra of the two reagents, DMAA and β-CD-Ac<sub>3</sub>, and PDMAA-β-CD-Ac<sub>3</sub> nanogel.</p>
Full article ">Figure 3
<p>(<b>a</b>) Digital images of nanogels synthesized from DMAA and β-CD-Ac<sub>3</sub> (1, NG2) and BAA (2, NG4), and (<b>b</b>) UV–Vis spectra of nanogels synthesized from DMAA and β-CD-Ac<sub>3</sub> (NG2) and BAA (NG4).</p>
Full article ">Figure 4
<p>Hydrodynamic diameter distribution plots of blank nanogels comprising different number of β-CD units.</p>
Full article ">Figure 5
<p>Representative AFM 2D (<b>a</b>) and 3D (<b>b</b>) height images of nanogel synthesized from DMAA and β-CD-Ac<sub>3</sub> at 1:1 feed ratio (NG1).</p>
Full article ">Figure 6
<p>In vitro release of aripiprazole from the two nanogel samples.</p>
Full article ">Figure 7
<p>Effect of filtered and unfiltered NG2 nanogel on the cell viability in a normal murine fibroblast cell line (CCL-1) and a cutaneous T-cell lymphoma in vitro model (HUT-78) after continuous 72 h exposure to two different concentrations.</p>
Full article ">
12 pages, 5815 KiB  
Article
Co-Crystallization between Aliphatic Polyesters through Co-Inclusion Complexation with Small Molecule
by Jia-Yao Chen, Xue-Wen Zhang, Tian-Yu Wu and Hai-Mu Ye
Molecules 2023, 28(10), 4091; https://doi.org/10.3390/molecules28104091 - 15 May 2023
Cited by 1 | Viewed by 1408
Abstract
Crystalline/crystalline blends of polymer have shown advantages in the preparation of new polymeric materials. However, the regulation of co-crystallization in a blend is still full of challenges due to the preferential self-crystallization driven by thermodynamics. Here, an inclusion complex approach is proposed to [...] Read more.
Crystalline/crystalline blends of polymer have shown advantages in the preparation of new polymeric materials. However, the regulation of co-crystallization in a blend is still full of challenges due to the preferential self-crystallization driven by thermodynamics. Here, an inclusion complex approach is proposed to facilitate the co-crystallization between crystalline polymers, because the crystallization process displays a prominent kinetics advantage when polymer chains are released from the inclusion complex. Poly(butylene succinate) (PBS), poly(butylene adipate) (PBA) and urea are chosen to form co-inclusion complexes, where PBS and PBA chains play as isolated guest molecules and urea molecules construct the host channel framework. The coalesced PBS/PBA blends are obtained by fast removing the urea framework and systematically investigated by differential scanning calorimetry, X-ray diffraction, proton nuclear magnetic resonance and Fourier transformation infrared spectrometry. It is demonstrated that PBA chains are co-crystallized into PBS extended-chain crystals in the coalesced blends, while such a phenomenon has not been detected in simply co-solution-blended samples. Though PBA chains could not be totally accommodated in the PBS extended-chain crystals, their co-crystallized content increases with the initial feeding ratio of PBA. Consequently, the melting point of the PBS extended-chain crystal gradually declines from 134.3 °C to 124.2 °C with an increasing PBA content. The PBA chains playing as defects mainly induce lattice expansion along the a-axis. In addition, when the co-crystals are soaked in tetrahydrofuran, some of the PBA chains are extracted out, leading to damage to the correlative PBS extended-chain crystals. This study shows that co-inclusion complexation with small molecules could be an effective way to promote co-crystallization behavior in polymer blends. Full article
(This article belongs to the Special Issue Exclusive Feature Papers in Macromolecular Chemistry)
Show Figures

Figure 1

Figure 1
<p>DSC scans of PBS, PBA and their simply-blended samples during the melt-cooling (<b>A</b>) and subsequent heating (<b>B</b>) processes at a constant rate of 10 °C/min, and (i)–(vi) indicate PBS, PBS/PBA-90/10, PBS/PBA-80/20, PBS/PBA-70/30, PBS/PBA-60/40 and PBA, respectively.</p>
Full article ">Figure 2
<p>DSC heating curves of coalesced PBS/PBA samples at a rate of 10 °C/min, and (i)–(v) indicate PBS, PBS/PBA-90/10, PBS/PBA-80/20, PBS/PBA-70/30 and PBS/PBA-60/40, respectively.</p>
Full article ">Figure 3
<p>Melting temperature <span class="html-italic">T</span><sub>m</sub> and melting enthalpy Δ<span class="html-italic">H</span><sub>m</sub> (normalized by the mass fraction of PBS) for the indicated samples as a function of PBS content in coalesced PBS/PBA blends.</p>
Full article ">Figure 4
<p>Wide angle X-ray diffractograms of simply-blended PBS/PBA samples at 26 °C (<b>A</b>) and 70 °C (<b>B</b>), and (i)–(v) indicate PBS, PBS/PBA-90/10, PBS/PBA-80/20, PBS/PBA-70/30 and PBS/PBA-60/40, respectively.</p>
Full article ">Figure 5
<p>Wide angle X-ray diffractograms of coalesced PBS/PBA blend at 26 °C (<b>A</b>) and 70 °C (<b>B</b>), and (i)–(v) indicate PBS, PBS/PBA-90/10, PBS/PBA-80/20, PBS/PBA-70/30 and PBS/PBA-60/40, respectively.</p>
Full article ">Figure 6
<p>The 2nd derivatives of Wide angle X-ray diffractograms of coalesced PBS/PBA samples at 70 °C. (i)–(v) indicate PBS, PBS/PBA-90/10, PBS/PBA-80/20, PBS/PBA-70/30 and PBS/PBA-60/40, respectively.</p>
Full article ">Figure 7
<p>The chemical structures of PBA and PBS (<b>A</b>) and the <sup>1</sup>H-NMR spectra of coalesced PBS/PBA-90/10 (<b>B</b>), PBS/PBA-80/20 (<b>C</b>), PBS/PBA-70/30 (<b>D</b>) and PBS/PBA-60/40 (<b>E</b>).</p>
Full article ">Figure 8
<p>DSC heating curves of coalesced PBS/PBA samples at a rate of 10 °C/min after being soaked in THF. (i)–(iv) indicate PBS/PBA-90/10, PBS/PBA-80/20, PBS/PBA-70/30 and PBS/PBA-60/40, respectively.</p>
Full article ">Figure 9
<p>FTIR spectra of the coalesced PBS and PBS/PBA samples in the wavenumber range of 1780–1680 cm<sup>–1</sup> after being etched by methylamine vapor for 48 h and corresponding differential spectra (grey dash line) between etched PBS (black dash dot lines) and PBS/PBA samples (blue solid lines). The blue solid lines in (<b>A</b>–<b>D</b>) indicate PBS/PBA-90/10, PBS/PBA-80/20, PBS/PBA-70/30 and PBS/PBA-60/40 after etching, respectively. The spectrum of amorphous PBA (green dot line) is inserted in (<b>D</b>) for comparison.</p>
Full article ">
11 pages, 4286 KiB  
Article
β-Cyclodextrin Polymer-Based Fluorescence Enhancement Strategy via Host–Guest Interaction for Sensitive Assay of SARS-CoV-2
by Shanshan Gao, Gege Yang, Xiaohui Zhang, Rui Shi, Rongrong Chen, Xin Zhang, Yuancheng Peng, Hua Yang, Ying Lu and Chunxia Song
Int. J. Mol. Sci. 2023, 24(8), 7174; https://doi.org/10.3390/ijms24087174 - 12 Apr 2023
Viewed by 1782
Abstract
Nucleocapsid protein (N protein) is an appropriate target for early determination of viral antigen-based severe acute respiratory syndrome coronavirus 2 (SARS-CoV-2). We have found that β-cyclodextrin polymer (β-CDP) has shown a significant fluorescence enhancement effect for fluorophore pyrene via host–guest interaction. Herein, we [...] Read more.
Nucleocapsid protein (N protein) is an appropriate target for early determination of viral antigen-based severe acute respiratory syndrome coronavirus 2 (SARS-CoV-2). We have found that β-cyclodextrin polymer (β-CDP) has shown a significant fluorescence enhancement effect for fluorophore pyrene via host–guest interaction. Herein, we developed a sensitive and selective N protein-sensing method that combined the host–guest interaction fluorescence enhancement strategy with high recognition of aptamer. The DNA aptamer of N protein modified with pyrene at its 3′ terminal was designed as the sensing probe. The added exonuclease I (Exo I) could digest the probe, and the obtained free pyrene as a guest could easily enter into the hydrophobic cavity of host β-CDP, thus inducing outstanding luminescent enhancement. While in the presence of N protein, the probe could combine with it to form a complex owing to the high affinity between the aptamer and the target, which prevented the digestion of Exo I. The steric hindrance of the complex prevented pyrene from entering the cavity of β-CDP, resulting in a tiny fluorescence change. N protein has been selectively analyzed with a low detection limit (11.27 nM) through the detection of the fluorescence intensity. Moreover, the sensing of spiked N protein from human serum and throat swabs samples of three volunteers has been achieved. These results indicated that our proposed method has broad application prospects for early diagnosis of coronavirus disease 2019. Full article
(This article belongs to the Special Issue Cyclodextrins: Properties and Applications)
Show Figures

Figure 1

Figure 1
<p>Schematic design of aptamer-based sensitive assay of SARS-CoV-2 N protein using β-cyclodextrin polymer to enhance pyrene fluorescence.</p>
Full article ">Figure 2
<p>(<b>A</b>) Fluorescence emission spectra in addition to β-CDP or β-CD monomer. (<b>B</b>) Fluorescence intensity of pyrene in addition to β-CDP vs. time. The concentration of β-CDP was 1.5 mg/mL. The concentration of pyrene labeled aptamer was 400 nM. The maximum emission wavelength was set at 376 nm.</p>
Full article ">Figure 3
<p>Fluorescence emission spectra in addition to pyrene-labeled scrambled DNA sequence and N protein aptamer. The concentration of β-CDP was 1.5 mg/mL. The concentration of N protein was 600 nM. The concentration of pyrene-labeled aptamer or pyrene-labeled scrambled DNA sequence was set at 400 nM. The maximum emission wavelength was set at 376 nm.</p>
Full article ">Figure 4
<p>(<b>A</b>) CD spectra of aptamer after incubation with N protein at 37 °C for 30 min (black curve), aptamer (red curve), and N protein alone (blue curve). The concentration of N protein and aptamer was equal to 1 μM. (<b>B</b>) Agarose gel electrophoresis. The concentration of N protein, aptamer, and Exo I was 600 nM, 400 nM, and 240 U/mL, respectively.</p>
Full article ">Figure 5
<p>(<b>A</b>) Optimization of pH value. The activity of Exo I was 240 U/mL. The concentration of N protein, aptamer, and β-CDP was 200 nM, 400 nM, and 1.5 mg/mL, respectively. (<b>B</b>) (F<sub>0</sub>–F)/F<sub>0</sub> vs. the centration of β-CDP. The activity of Exo I was 240 U/mL. The concentration of N protein and aptamer was 200 nM and 400 nM, respectively. (<b>C</b>) (F<sub>0</sub>–F)/F<sub>0</sub> vs. the activity of Exo I. The concentration of aptamer was 400 nM. The concentration of N protein was 200 nM. The concentration of β-CDP was 1.5 mg/mL. (<b>D</b>) (F<sub>0</sub>–F)/F<sub>0</sub> vs. the concentration of aptamer. The concentration of Exo I and N protein were 240 U/mL and 200 nM, respectively. The concentration of β-CDP was 1.5 mg/mL. Error bars indicated the standard deviations of three experiments.</p>
Full article ">Figure 6
<p>(<b>A</b>) Fluorescence emission spectra of the system under the concentration of N protein from 0 to 600 nM. (<b>B</b>) The linear relationship between the concentration of N protein (the concentration from left to right was 0, 40, 100, 140, 200, and 300 nM, respectively) and (F<sub>0</sub>–F)/F<sub>0</sub> (S/N = 3). The concentration of aptamer, Exo I, and β-CDP was 400 nM, 240 U/mL, and 1.5 mg/mL, respectively. The excitation wavelength was set at 345 nm, and the maximum emission wavelength was set at 376 nm.</p>
Full article ">Figure 7
<p>(F<sub>0</sub>–F)/F<sub>0</sub> value of potential interferences in human blood samples (<b>A</b>) and throat swabs (<b>B</b>). The concentration of potential interferences or N protein was all set at 600 nM.</p>
Full article ">Figure 8
<p>(<b>A</b>) (F<sub>0</sub>–F)/F<sub>0</sub> value at various concentrations of N protein spiked to the human serum sample. (<b>B</b>) (F<sub>0</sub>–F)/F<sub>0</sub> value at various concentrations of N protein spiked to the throat swabs sample.</p>
Full article ">
20 pages, 3723 KiB  
Article
Injectable Hydrogels Based on Cyclodextrin/Cholesterol Inclusion Complexation and Loaded with 5-Fluorouracil/Methotrexate for Breast Cancer Treatment
by Saud Almawash, Ahmed M. Mohammed, Mohamed A. El Hamd and Shaaban K. Osman
Gels 2023, 9(4), 326; https://doi.org/10.3390/gels9040326 - 12 Apr 2023
Cited by 3 | Viewed by 2043
Abstract
Breast cancer is the second most common cancer in women worldwide. Long-term treatment with conventional chemotherapy may result in severe systemic side effects. Therefore, the localized delivery of chemotherapy helps to overcome such a problem. In this article, self-assembling hydrogels were constructed via [...] Read more.
Breast cancer is the second most common cancer in women worldwide. Long-term treatment with conventional chemotherapy may result in severe systemic side effects. Therefore, the localized delivery of chemotherapy helps to overcome such a problem. In this article, self-assembling hydrogels were constructed via inclusion complexation between host β-cyclodextrin polymers (8armPEG20k-CD and pβ-CD) and the guest polymers 8-armed poly(ethylene glycol) capped either with cholesterol (8armPEG20k-chol) or adamantane (8armPEG20k-Ad) and were loaded with 5-fluorouracil (5-FU) and methotrexate (MTX). The prepared hydrogels were characterized by SEM and rheological behaviors. The in vitro release of 5-FU and MTX was studied. The cytotoxicity of our modified systems was investigated against breast tumor cells (MCF-7) using an MTT assay. Additionally, the histopathological changes in breast tissues were monitored before and after their intratumor injection. The results of rheological characterization indicated the viscoelastic behavior in all cases except for 8armPEG-Ad. In vitro release results showed a variable range of release profiles from 6 to 21 days, depending on the hydrogel composition. MTT findings indicated the inhibition ability of our systems against the viability of cancer cells depending on the kind and concentration of the hydrogel and the incubation period. Moreover, the results of histopathology showed the improvement of cancer manifestation (swelling and inflammation) after intratumor injection of loaded hydrogel systems. In conclusion, the obtained results indicated the applicability of the modified hydrogels as injectable vehicles for both loading and controlled release of anticancer therapies. Full article
(This article belongs to the Special Issue Cancer Cell Biology in Biological Hydrogel)
Show Figures

Figure 1

Figure 1
<p>Photos of self-assembling hydrogel based on CD-chol/Ad inclusion complexation: (<b>A</b>) 8armPEG20k-chol/8armPEG20k-CD, (<b>B</b>) 8armPEG20k-Ad/pβ-CD, (<b>C</b>) 8armPEG20k-chol/pβ-CD, (<b>D</b>) 8armPEG-OH/pβ-CD, and (<b>E</b>) 8armPEG20k-chol/native β-CD.</p>
Full article ">Figure 2
<p>SEM micrographs showing the channels and cross sections of different lyophilized hydrogel systems at a magnification power (×100). (<b>a</b>–<b>c</b>) The graphs display the hydrogel formulas: (<b>a</b>) formula A, (<b>b</b>) formula B, and (<b>c</b>) formula C. (<b>d</b>–<b>f</b>) The graphs display the physical mixtures of the hydrogel components: (<b>d</b>) physical mixture of formula A, (<b>e</b>) physical mixture of formula B, and (<b>f</b>) physical mixture of formula C.</p>
Full article ">Figure 3
<p>In vitro release profiles of both 5-FU and MTX from (<b>A</b>) hydrogel formula A, composed of 30%, <span class="html-italic">w</span>/<span class="html-italic">v</span> 8armPEG-chol/8rmPEG-CD (1:1%, <span class="html-italic">w</span>/<span class="html-italic">w</span> ratio), and from (<b>B</b>) hydrogel formula C, composed of 10%, <span class="html-italic">w</span>/<span class="html-italic">v</span> 8armPEG20k-chol/pβ-CD (1:1%, <span class="html-italic">w</span>/<span class="html-italic">w</span> ratio) at 37 °C in PBS.</p>
Full article ">Figure 4
<p>Cytotoxicity assay (% cell viability), as a function of drug concentration, loaded into the modified gel systems; G4 (8armPEG20k-CD/8armPEG20k-chol), G5 (pβ-CD/8armPEG20k-Ad), and G6 (pβ-CD/8armPEG20k-chol) compared with G2 (5-FU free saline solution) and G3 (5-FU/MTX free saline solution) against MCF-7 breast cancer cell line. The results are presented as the average of three independent measurements ± SD.</p>
Full article ">Figure 5
<p>Cytotoxicity assay (% of cell viability), as a function of the incubation period, of the modified gel systems; G4 (8armPEG20k-CD/8armPEG20k-chol), G5 (pβ-CD/8armPEG20k-Ad), and G6 (pβ-CD/8armPEG20k-chol) compared with G2 (5-FU free saline solution) and G3 (5-FU/MTX free saline solution) against MCF-7 breast cancer cell line. The results are presented as the average of three independent measurements ± SD.</p>
Full article ">Figure 6
<p>Photographs of the treated rats, including normal rat (<b>A</b>), untreated tumor-induced rat (<b>B</b>), and the various treated groups (<b>C</b>–<b>E</b>), including 5-FU/MTX saline solution injected group (<b>C</b>), the drug-loaded hydrogel system A (<b>D</b>), and the drug-loaded gel system C (<b>E</b>).</p>
Full article ">Figure 7
<p>Photomicrographs display some histopathological appraisals (arrows), which were detected in the mammary gland tissues from six treatment groups: (<b>A</b>) the unmedicated group showing numerous clusters of neoplastic cells with marked edema; (<b>B</b>) the group receiving (5-FU/MTX) saline solution showing proliferating neoplastic cells with inflammatory reactions; (<b>C</b>) the group receiving hydrogel formula A showing few proliferating mammary ducts and inflammatory cells infiltration; and (<b>D</b>) the group receiving hydrogel formula C showing few neoplastic cells forming ducts with mild inflammatory edema; 100× (H&amp;E).</p>
Full article ">Figure 8
<p>The percentage of relative tumor volume (%RTV) as an indication of antitumor efficacy of the modified hydrogel systems (formulas A and C) in comparison with the free drugs and untreated groups after their local injection into the breast tumor (<span class="html-italic">n</span> = 8).</p>
Full article ">Figure 9
<p>The effect of the modified hydrogel systems (A and C) loaded with dual anticancer (5-FU/MTX) on the body weight of rats in comparison with untreated animals.</p>
Full article ">
16 pages, 3798 KiB  
Article
Preparation, Characterization, and Antioxidant Activity of L-Ascorbic Acid/HP-β-Cyclodextrin Inclusion Complex-Incorporated Electrospun Nanofibers
by Nabab Khan, Amit Kumar Singh and Ankit Saneja
Foods 2023, 12(7), 1363; https://doi.org/10.3390/foods12071363 - 23 Mar 2023
Cited by 12 | Viewed by 2768
Abstract
L-Ascorbic acid (LAA) is a key vitamin, implicated in a variety of physiological processes in humans. Due to its free radical scavenging activity, it is extensively employed as an excipient in pharmaceutical products and food supplements. However, its application is greatly impeded by [...] Read more.
L-Ascorbic acid (LAA) is a key vitamin, implicated in a variety of physiological processes in humans. Due to its free radical scavenging activity, it is extensively employed as an excipient in pharmaceutical products and food supplements. However, its application is greatly impeded by poor thermal and aqueous stability. Herein, to improve the stability and inhibit oxidative degradation, we prepared LAA-cyclodextrin inclusion complex-incorporated nanofibers (NFs). The continuous variation method (Job plot) demonstrated that LAA forms inclusions with hydroxypropyl-β-cyclodextrin (HP-β-CD) at a 2:1 molar stoichiometric ratio. The NFs were prepared via the single step electrospinning technique, without using any polymer matrix. The solid-state characterizations of LAA/HP-β-CD-NF via powder x-ray diffractometry (PXRD), Fourier-transform infrared (FT-IR) analysis, differential scanning calorimetry (DSC), thermal gravimetric analysis (TGA), and nuclear magnetic resonance (1H NMR and 2D-NOESY) spectroscopy, reveal the effective encapsulation of the LAA (guest molecule) inside the HP-β-CD (host) cavity. The SEM micrograph reveals an average fiber diameter of ~339 nm. The outcomes of the thermal investigations demonstrated that encapsulation of LAA within HP-β-CD cavities provides improved thermal stability of LAA (by increasing the thermal degradation temperature). The radical scavenging assay demonstrated the enhanced antioxidant potential of LAA/HP-β-CD-NF, as compared to native LAA. Overall, the study shows that cyclodextrin inclusion complex-incorporated NFs, are an effective approach for improving the limitations associated with LAA, and provide promising avenues in its therapeutic and food applications. Full article
(This article belongs to the Special Issue Health Foods: Molecular Nutrition Mechanisms and Product Development)
Show Figures

Figure 1

Figure 1
<p>Schematic illustration of molecular structure of: (<b>a</b>) L-ascorbic acid, (<b>b</b>) hydroxypropyl-<span class="html-italic">β</span>-cyclodextrin, and (<b>c</b>) formation of LAA/HP-<span class="html-italic">β</span>-CD-NFs, via the electrospinning technique.</p>
Full article ">Figure 2
<p>Graphical representation of continuous variation plot (Job plot) for the complexation of LAA with HP-<span class="html-italic">β</span>-CD, from absorbance measurements at 25 °C. R = [(LAA)/ (LAA) + (HP-<span class="html-italic">β</span>-CD)], ∆A = absorbance difference of LAA with and without HP-<span class="html-italic">β</span>-CD.</p>
Full article ">Figure 3
<p>Solid state characterization of LAA/HP-<span class="html-italic">β</span>-CD-NFs. (<b>a</b>) Surface micrograph of LAA/HP-<span class="html-italic">β</span>-CD-NFs captured by scanning electron microscope, (<b>b</b>,<b>c</b>) photographs of the obtained LAA/HP-<span class="html-italic">β</span>-CD-NFs and electrospinning solution, respectively, (<b>d</b>) histogram representing average diameter of LAA/HP-<span class="html-italic">β</span>-CD-NFs, determined using the ImageJ 1.53t software, based on at least 50 counts, (<b>e</b>) FT-IR spectrum and (<b>f</b>) XRD diffractogram of LAA (blue, stars in red color indicates sharp diffraction peaks), HP-<span class="html-italic">β</span>-CD (green), and LAA/HP-<span class="html-italic">β</span>-CD-NFs (orange).</p>
Full article ">Figure 4
<p>Thermal analysis representing overlay thermogram: (<b>a</b>) DSC, (<b>b</b>) TGA, and (<b>c</b>) DTG analyses of LAA (blue), HP-<span class="html-italic">β</span>-CD (green), and LAA/HP-<span class="html-italic">β</span>-CD-NFs (orange).</p>
Full article ">Figure 5
<p>Nuclear magnetic resonance spectroscopy characterization. (<b>a</b>) <sup>1</sup>H NMR of LAA, HP-<span class="html-italic">β</span>-CD in DMSO-d<sub>6</sub>, and LAA/HP-<span class="html-italic">β</span>-CD-NF in D<sub>2</sub>O, (<b>b</b>) 2-D NOESY spectrum of LAA/HP-<span class="html-italic">β</span>-CD-NF in D<sub>2</sub>O, demonstrating contour region (zoomed in region), revealing interaction of LAA protons with protons of HP-<span class="html-italic">β</span>-CD.</p>
Full article ">Figure 6
<p>(<b>a</b>) Graphical representation of concentration-dependent antioxidant performance (% DPPH radical scavenging) of LAA and LAA/HP-<span class="html-italic">β</span>-CD-NF. Visual representation of series of different concentrations of obtained solutions of (<b>b</b>) LAA and (<b>c</b>) LAA/HP-<span class="html-italic">β</span>-CD-NF, demonstrating change in color of sample solutions (due to scavenging of DPPH radical), from purple to yellow at highest concentration of LAA in LAA/HP-<span class="html-italic">β</span>-CD-NF (30 µM), indicating its antioxidant potential. The values are expressed as mean ± standard deviation (n = 3), and the statistical analysis at the significance level of * <span class="html-italic">p</span> &lt; 0.05 was achieved by using one-way analysis of variance (ANOVA), followed by Tukey’s post hoc test, using GraphPad Prism 9 software.</p>
Full article ">
12 pages, 3309 KiB  
Article
Stimuli-Responsive Designer Supramolecular Polymer Gel
by M. Douzapau, Srayoshi Roy Chowdhury, Surajit Singh, Olamilekan Joseph Ibukun and Debasish Haldar
Chemistry 2023, 5(1), 691-702; https://doi.org/10.3390/chemistry5010048 - 22 Mar 2023
Cited by 3 | Viewed by 2534
Abstract
This paper reports a stimuli-responsive designer supramolecular polymer gel in dimethylsulphoxide (DMSO)/water (1:2) based on a dipeptide amphiphile and β-cyclodextrin (β-CD) The dipeptide amphiphile contains caproic acid at the N terminus and methyl ester at the C terminus. From X-ray single crystal diffraction, [...] Read more.
This paper reports a stimuli-responsive designer supramolecular polymer gel in dimethylsulphoxide (DMSO)/water (1:2) based on a dipeptide amphiphile and β-cyclodextrin (β-CD) The dipeptide amphiphile contains caproic acid at the N terminus and methyl ester at the C terminus. From X-ray single crystal diffraction, the amphiphile adopts a kink-like conformation. The amphiphile self-assembled to form a parallel sheet-like structure stabilized by multiple intermolecular hydrogen bonds. Moreover, the parallel sheet-like structure is also stabilized by edge-to-edge ππ stacking interactions. In higher-order packing, it forms a corrugated sheet-like structure stabilized by hydrophobic interactions. The dipeptide amphiphile interacts with β-cyclodextrin and forms gel through supramolecular polymer formation in (DMSO)/water (1:2) by a simple heating-cooling cycle. The sol-to-gel transformation is because of a host–guest complex between compound 1 and β-CD and the formation of supramolecular polymer accompanied by microstructure changes from nanofibers to microrods. The gel is temperature responsive with a Tgel of 70 °C. The supramolecular polymer gel is also responsive to stimuli such aspicric acid and HCl. The extensive spectroscopic studies show that the aromatic hydrophobic side chain of compound 1 forms a host–guest complex with β-CD. These results will be helpful for the design of advanced programable eco-friendly functional materials. Full article
(This article belongs to the Section Supramolecular Chemistry)
Show Figures

Figure 1

Figure 1
<p>Schematic presentation of compound <b>1</b> and α, β, andγ-cyclodextrin.</p>
Full article ">Figure 2
<p>The ORTEP diagram with atom numbering scheme of compound <b>1</b>. 50% Probability. [C in grey, O in red, N in blue].</p>
Full article ">Figure 3
<p>(<b>a</b>) The solid-state structure of compound <b>1</b> showing a parallel sheet-like structure stabilized by multiple intermolecular hydrogen bonds. Intermolecular hydrogen bonds are shown as black dotted lines. (<b>b</b>) Solid state structure of compound <b>1</b> showing edge-to-edge π–π stacking interaction. π–π stacking interactions are shown as a space fill model; [C in grey, O in red, N in blue].</p>
Full article ">Figure 4
<p>The solid-state packing diagram of compound <b>1</b> showing a supramolecular corrugated sheet-like structure stabilized by intermolecular hydrophobic interactions. Intermolecular hydrogen hydrophobic interactions are shown in red circles [C in grey, O in red, N in blue].</p>
Full article ">Figure 5
<p>(<b>a</b>) The image showing compound <b>1</b> and β-CD supramolecular gel in DMSO-H<sub>2</sub>O. (<b>b</b>) Effect of strain on supramolecular gel in DMSO-H<sub>2</sub>O; (<b>c</b>) Effect of angular frequency on supramolecular gel in DMSO-H<sub>2</sub>O.</p>
Full article ">Figure 6
<p>(<b>a</b>) FE-SEM image of compound <b>1</b> in DMSO and (<b>b</b>) FE-SEM image of xerogel form by compound <b>1</b>, β-CD in DMSO and water mixture.</p>
Full article ">Figure 7
<p>(<b>a</b>) The emission spectra of compound <b>1</b> show bands at 348 and 408 nm. The emission intensities of the bandsgradually increase with the gradual addition of β-CD (excitation at 280 nm). (<b>b</b>) The binding stoichiometry of compound 1 and β-CD in DMSO by fitting emission data at 348, 327, and 408 nm using the bindfit methods, <a href="http://supramolecular.org" target="_blank">http://supramolecular.org</a> (accessed on 21 February 2023). (<b>c</b>) Solid state FT-IR spectra of compound <b>1</b> (black), β-CD (blue), and xerogel obtained from compound <b>1</b> and β-CD supramolecular gel in DMSO-H<sub>2</sub>O (red). (<b>d</b>) Part of the <sup>1</sup>H-NMR spectra of compound <b>1</b> (red); compound <b>1</b> and β-CD Gel in DMSO-D6 (blue), showing shifting of compound <b>1</b> aromatic peaks with the addition of β-CD.</p>
Full article ">Figure 8
<p>The images showing compound <b>1</b> and β-CD supramolecular gel in DMSO-H<sub>2</sub>O is (<b>a</b>,<b>b</b>) thermo-responsive; (<b>c</b>) HCl responsive; and (<b>d</b>) picric acid-responsive.</p>
Full article ">Figure 9
<p>The fluorescence spectroscopic studies of compound <b>1</b> (black), picric acid (blue), compound 1-picric acid charge transfer complex (red), and compound <b>1</b> and β-CD supramolecular gel in DMSO-H<sub>2</sub>O degraded by picric acid (green). (Excitation at 280 nm).</p>
Full article ">Figure 10
<p>The proposed model forsupramolecular polymer formation due tothe inclusion complex between compound <b>1</b> and β-CD (<b>left</b>) and the decomposition of the supramolecular polymer by addition of picric acid (<b>right</b>).</p>
Full article ">Scheme 1
<p>Synthesis of compound <b>1</b>.</p>
Full article ">
34 pages, 6949 KiB  
Review
Mass Spectrometry of Esterified Cyclodextrins
by Diana-Andreea Blaj, Marek Kowalczuk and Cristian Peptu
Molecules 2023, 28(5), 2001; https://doi.org/10.3390/molecules28052001 - 21 Feb 2023
Cited by 3 | Viewed by 2970
Abstract
Cyclodextrins are cyclic oligosaccharides that have received special attention due to their cavity-based structural architecture that imbues them with outstanding properties, primarily related to their capacity to host various guest molecules, from low-molecular-mass compounds to polymers. Cyclodextrin derivatization has been always accompanied by [...] Read more.
Cyclodextrins are cyclic oligosaccharides that have received special attention due to their cavity-based structural architecture that imbues them with outstanding properties, primarily related to their capacity to host various guest molecules, from low-molecular-mass compounds to polymers. Cyclodextrin derivatization has been always accompanied by the development of characterization methods, able to unfold complicated structures with increasing precision. One of the important leaps forward is represented by mass spectrometry techniques with soft ionization, mainly matrix-assisted laser desorption/ionization (MALDI) and electrospray ionization (ESI). In this context, esterified cyclodextrins (ECDs) benefited also from the formidable input of structural knowledge, thus allowing the understanding of the structural impact of reaction parameters on the obtained products, especially for the ring-opening oligomerization of cyclic esters. The current review envisages the common mass spectrometry approaches such as direct MALDI MS or ESI MS analysis, hyphenated liquid chromatography-mass spectrometry, and tandem mass spectrometry, employed for unraveling the structural features and particular processes associated with ECDs. Thus, the accurate description of complex architectures, advances in the gas phase fragmentation processes, assessment of secondary reactions, and reaction kinetics are discussed in addition to typical molecular mass measurements. Full article
(This article belongs to the Special Issue Identification of Biomolecules by Mass Spectrometry)
Show Figures

Figure 1

Figure 1
<p>MALDI MS spectrum of succinic-anhydride-modified γ-cyclodextrin (reprinted with permission from [<a href="#B22-molecules-28-02001" class="html-bibr">22</a>]).</p>
Full article ">Figure 2
<p>MALDI-MS spectra of β-CD decanoate esters (<b>A</b>) and β-CD butyrate esters (<b>B</b>). Masses of sodium adducts (first series) and potassium adducts (second series) are shown with numbers indicating the degree of substitution (reprinted with permission from [<a href="#B5-molecules-28-02001" class="html-bibr">5</a>]).</p>
Full article ">Figure 3
<p>The mass spectra of ECD: (<b>a</b>) mono-2-O-poly(δ-VL)-β-CD (<span class="html-italic">m/z = 1134 (β-CD) + n</span> × <span class="html-italic">100 (δ-VL) + 23 (Na<sup>+</sup>))</span>, (<b>b</b>) mono-2-O-poly(ε-CL)-β-CD (<span class="html-italic">m/z = 1134 (β-CD) + n</span> × <span class="html-italic">114 (ε-CL) + 23 (Na<sup>+</sup>))</span>, and (<b>c</b>) mono-2-O-(5-benzyloxypentanoyl)-poly(δ-VL)-β-CD (<span class="html-italic">m/z = 1324 (mono-2-O-(6-benzoxy-pentanoyl)-β-CD) + n</span> × <span class="html-italic">100 (δ-VL) + 23 (Na<sup>+</sup>))</span> (reprinted with permission from [<a href="#B38-molecules-28-02001" class="html-bibr">38</a>]).</p>
Full article ">Figure 4
<p>Total ion chromatogram (<b>A</b>) and LC-MS spectrum of PHB-CD derivatives (<b>B</b>) (reprinted with permission from [<a href="#B41-molecules-28-02001" class="html-bibr">41</a>]).</p>
Full article ">Figure 5
<p>Mn (number-average molecular mass) evolution of β-CDLA product determined by <sup>1</sup>H NMR and MALDI MS (DHB—dried droplet, CHCA—thin-layer) (<b>A</b>) and the agreement between the <span class="html-italic">Mn</span> evolution determined by <sup>1</sup>H NMR and CHCA (<b>B</b>) (reprinted with permission from [<a href="#B51-molecules-28-02001" class="html-bibr">51</a>]).</p>
Full article ">Figure 6
<p>MALDI MS spectrum of CDLA derivative showing the MS peak series detected in the CDLA sample (reprinted with permission from [<a href="#B51-molecules-28-02001" class="html-bibr">51</a>]).</p>
Full article ">Figure 7
<p>ELSD chromatogram of crude β-CDLA mixture and MALDI MS spectrum (reprinted with permission from [<a href="#B49-molecules-28-02001" class="html-bibr">49</a>]).</p>
Full article ">Figure 8
<p>The substitution pattern evolution of CDCL derivatives with Mn increase (reprinted with permission from [<a href="#B46-molecules-28-02001" class="html-bibr">46</a>]).</p>
Full article ">Figure 9
<p>Organo-activator influence on CDCL <span class="html-italic">Mn</span> evolution (reprinted with permission from [<a href="#B46-molecules-28-02001" class="html-bibr">46</a>]).</p>
Full article ">Figure 10
<p>MALDI MS spectrum of CDCL derivatives obtained in DMF or DMSO and DMF degradation reaction in the presence of CDCL (reprinted with permission from [<a href="#B46-molecules-28-02001" class="html-bibr">46</a>]).</p>
Full article ">Figure 11
<p>MALDI LID (TOF/TOF) of β-CDLA precursor ions: sodium- (<b>A</b>), potassium- (<b>B</b>), and lithium-charged (<b>C</b>) (reprinted with permission from [<a href="#B50-molecules-28-02001" class="html-bibr">50</a>]).</p>
Full article ">Scheme 1
<p>The structure of native CDs.</p>
Full article ">Scheme 2
<p>Structures of ECDs and OECD.</p>
Full article ">Scheme 3
<p><span class="html-italic">Thermolysin</span>-catalyzed transesterification of vinyl esters in the presence of CDs (adapted from [<a href="#B31-molecules-28-02001" class="html-bibr">31</a>]).</p>
Full article ">Scheme 4
<p>Ring-opening reaction of cyclic esters in the presence of cyclodextrins.</p>
Full article ">Scheme 5
<p>Charge-induced ring opening and linearization of proton-charged β-CD, resulting in the formation of a new non-reducing end (reprinted with permission from [<a href="#B112-molecules-28-02001" class="html-bibr">112</a>]).</p>
Full article ">Scheme 6
<p>Proposed fragmentation mechanism for the formation of the fragment with 2 glycoside units (reprinted with permission from [<a href="#B113-molecules-28-02001" class="html-bibr">113</a>]).</p>
Full article ">Scheme 7
<p>Fragmentation of polyesters on the acyl or alkyl part of the ester bond.</p>
Full article ">Scheme 8
<p>Fragmentation pathways of PHB-CD (adapted from [<a href="#B114-molecules-28-02001" class="html-bibr">114</a>]).</p>
Full article ">Scheme 9
<p>Fragmentation of CDCL derivatives at the level of oligocaprolactone chain and of the substituted β-CD (reprinted with permission from [<a href="#B46-molecules-28-02001" class="html-bibr">46</a>]).</p>
Full article ">
15 pages, 4329 KiB  
Article
A Supramolecular Nanoassembly of Lenvatinib and a Green Light-Activatable NO Releaser for Combined Chemo-Phototherapy
by Francesca Laneri, Nadia Licciardello, Yota Suzuki, Adriana C. E. Graziano, Federica Sodano, Aurore Fraix and Salvatore Sortino
Pharmaceutics 2023, 15(1), 96; https://doi.org/10.3390/pharmaceutics15010096 - 28 Dec 2022
Cited by 4 | Viewed by 2051
Abstract
The chemotherapeutic Lenvatinib (LVB) and a nitric oxide (NO) photodonor based on a rhodamine antenna (RD-NO) activatable by the highly compatible green light are supramolecularly assembled by a β-cyclodextrin branched polymer (PolyCD). The poorly water-soluble LVB and RD-NO solubilize very well within the [...] Read more.
The chemotherapeutic Lenvatinib (LVB) and a nitric oxide (NO) photodonor based on a rhodamine antenna (RD-NO) activatable by the highly compatible green light are supramolecularly assembled by a β-cyclodextrin branched polymer (PolyCD). The poorly water-soluble LVB and RD-NO solubilize very well within the polymeric host leading to a ternary supramolecular nanoassembly with a diameter of ~55 nm. The efficiency of the NO photorelease and the typical red fluorescence of RD-NO significantly enhance within the polymer due to its active role in the photochemical and photophysical deactivation pathways. The co-presence of LVB within the same host does not affect either the nature or the efficiency of the photoinduced processes of RD-NO. Besides, irradiation of RD-NO does not lead to the decomposition of LVB, ruling out any intermolecular photoinduced process between the two guests despite sharing the same host. Ad-hoc devised Förster Resonance Energy Transfer experiments demonstrate this to be the result of the not close proximity of the two guests, which are confined in different compartments of the same polymeric host. The supramolecular complex is stable in a culture medium, and its biological activity has been evaluated against HEP-G2 hepatocarcinoma cell lines in the dark and under irradiation with visible green light, using LVB at a concentration well below the IC50. Comparative experiments performed using the polymeric host encapsulating the individual LVB and RD-NO components under the same experimental conditions show that the moderate cell mortality induced by the ternary complex in the dark increases significantly upon irradiation with visible green light, more likely as the result of synergism between the NO photogenerated and the chemotherapeutic. Full article
(This article belongs to the Section Pharmaceutical Technology, Manufacturing and Devices)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>(<b>A</b>) Absorption spectrum of LVB in PBS (pH 7.4) at 4.1 µM (a) and in the presence of PolyCD (2 mg mL<sup>−1</sup>) at 25 µM (b), 40 µM (c), 50 µM (d) and 60 µM I (e). The spectrum of a 60 µM LVB solution in methanol is shown for the sake of comparison (f). (<b>B</b>) Absorbance value at 245 nm of LVB at different concentrations in the presence of PolyCD. T = 25 °C; cell path = 0.1 cm.</p>
Full article ">Figure 2
<p>(<b>A</b>) Absorption spectra of PolyCD (2 mg mL<sup>−1</sup>) in PBS (pH 7.4) co-loaded with LVB (25 µM) and RD-NO (6 µM) (a) and, for the sake of comparison, loaded only with LVB (b) and RD-NO (c). T = 25 °C; cell path = 1 cm. (<b>B</b>) Hydrodynamic diameter of sample a in <a href="#pharmaceutics-15-00096-f002" class="html-fig">Figure 2</a>A.</p>
Full article ">Figure 3
<p>(<b>A</b>) Absorption spectral changes observed upon 532 nm light irradiation of a PBS (pH 7.4) solution of PolyCD (2 mg mL<sup>−1</sup>) co-loaded with LVB (25 µM) and RD-NO (6 µM). The inset shows the absorbance changes at 397 nm of PolyCD (2 mg mL<sup>−1</sup><b>)</b> co-loaded with LVB (25 µM) and RD-NO (6 µM) (λ) and, for comparison, the same host in the absence of LVB (ν). (<b>B</b>) NO release profile observed for PolyCD (2 mg mL<sup>−1</sup>) co-loaded with LVB (25 µM) and RD-NO (6 µM) (a) and, for comparison, the same host in the absence of LVB (b). T = 25 °C.</p>
Full article ">Figure 4
<p>(<b>A</b>) Fluorescence emission spectrum observed at λ<sub>exc</sub> = 520 nm light excitation for a PBS (pH 7.4) solution of PolyCD (2 mg mL<sup>−1</sup>) co-loaded with LVB (25 µM) and RD-NO (6 µM). (<b>B</b>) Fluorescence decay and the related fitting of the same solution recorded at λ<sub>exc</sub> = 455 nm and λ<sub>em</sub> = 590 nm. T = 25 °C.</p>
Full article ">Figure 5
<p>Normalized fluorescence spectra of LVB (a) (λ<sub>exc</sub> = 240 nm), NBF (d) (λ<sub>exc</sub> = 440 nm), and absorption spectra of NBF (c) and RD-NO (b). All guests are loaded in PolyCD (2 mg mL<sup>−1</sup>) dissolved in PBS (pH 7.4). T = 25 °C.</p>
Full article ">Figure 6
<p>(<b>A</b>) Absorption spectrum of PolyCD (2 mg mL<sup>−1</sup>) in PBS (pH 7.4) co-loaded with LVB, RD-NO, and NBF (5 µM). (<b>B</b>) Fluorescence emission spectra of the sample as in (<b>A</b>) (a) and PolyCD (2 mg mL<sup>−1</sup>) in PBS (pH 7.4) loaded only with LVB at the same concentration (b) recorded at λ<sub>exc</sub> = 240 nm. (<b>C</b>) Fluorescence emission spectra of the sample as in (<b>A</b>) (a) and PolyCD (2 mg mL<sup>−1</sup>) in PBS (pH 7.4) loaded only with NBF (b) and RD-NO (c) recorded at λ<sub>exc</sub> = 445 nm.</p>
Full article ">Figure 7
<p>Fluorescence microscopy analysis of HEP-G2 hepatocarcinoma cell lines treated with PolyCD (2 mg mL<sup>−1</sup>) in PBS (pH 7.4) loaded with LVB (25 µM), RD-NO (6 µM) and both components and stained with DAPI. The cells were analyzed with a DAPI emission filter (<b>A</b>), a rhodamine emission filter (<b>B</b>), or by merging images (<b>A</b>,<b>B</b>) (<b>C</b>). Scale bar = 50 µM.</p>
Full article ">Figure 8
<p>Viability of HEP-G2 hepatocarcinoma cells as a function of the concentration of free LVB.</p>
Full article ">Figure 9
<p>Cell viability was observed 24 h after incubating HEP-G2 hepatocarcinoma cells with free LVB, PolyCD, and the supramolecular complexes PolyCD/LVB, PolyCD/RD-NO, and PolyCD/LVB/RD-NO in the dark and upon different irradiation times (occurred after the first 4 h of incubation) at λ<sub>exc</sub> &gt; 500 nm. [LVB] = 50 µM; [RD-NO] = 6 µM; [PolyCD] = 2 mg mL<sup>−1</sup>.</p>
Full article ">Scheme 1
<p>Molecular structures of the guests LVB and RD-NO, which are supramolecularly assembled in the branched polymeric host PolyCD.</p>
Full article ">
21 pages, 1634 KiB  
Review
Cyclodextrin-Based Polymeric Materials Bound to Corona Protein for Theranostic Applications
by Donya Esmaeilpour, Jens Albert Broscheit and Sergey Shityakov
Int. J. Mol. Sci. 2022, 23(21), 13505; https://doi.org/10.3390/ijms232113505 - 4 Nov 2022
Cited by 4 | Viewed by 2866
Abstract
Cyclodextrins (CDs) are cyclic oligosaccharide structures that could be used for theranostic applications in personalized medicine. These compounds have been widely utilized not only for enhancing drug solubility, stability, and bioavailability but also for controlled and targeted delivery of small molecules. These compounds [...] Read more.
Cyclodextrins (CDs) are cyclic oligosaccharide structures that could be used for theranostic applications in personalized medicine. These compounds have been widely utilized not only for enhancing drug solubility, stability, and bioavailability but also for controlled and targeted delivery of small molecules. These compounds can be complexed with various biomolecules, such as peptides or proteins, via host-guest interactions. CDs are amphiphilic compounds with water-hating holes and water-absorbing surfaces. Architectures of CDs allow the drawing and preparation of CD-based polymers (CDbPs) with optimal pharmacokinetic and pharmacodynamic properties. These polymers can be cloaked with protein corona consisting of adsorbed plasma or extracellular proteins to improve nanoparticle biodistribution and half-life. Besides, CDs have become famous in applications ranging from biomedicine to environmental sciences. In this review, we emphasize ongoing research in biomedical fields using CD-based centered, pendant, and terminated polymers and their interactions with protein corona for theranostic applications. Overall, a perusal of information concerning this novel approach in biomedicine will help to implement this methodology based on host-guest interaction to improve therapeutic and diagnostic strategies. Full article
(This article belongs to the Special Issue Recent Insights in Chemistry and Technology of Cyclodextrins)
Show Figures

Figure 1

Figure 1
<p>Structure and conformation of natural CDs.</p>
Full article ">Figure 2
<p>Diagram describes the production process of drug-loaded single-chain polymer nanoparticles (SCNPs) for dual-responsive drug release [<a href="#B38-ijms-23-13505" class="html-bibr">38</a>]. The abbreviations are PDI*: perylenediimide, PCL: polycaprolactone, PEG: polyethyleneglycol, and GSH: glutathione.</p>
Full article ">Figure 3
<p>The interaction between protein corona, NP, and cellular receptors (physiological reactions). After NP injection into the blood, protein corona starts to aggregate at the NP surface as a part of biological identity. After that, cellular response can be triggered by specific protein receptors at the cell surface [<a href="#B44-ijms-23-13505" class="html-bibr">44</a>,<a href="#B45-ijms-23-13505" class="html-bibr">45</a>].</p>
Full article ">Figure 4
<p>Classification of polymers by molecular topology.</p>
Full article ">Figure 5
<p>The kinetic proceeding of protein amyloid fibrillation comprising (Lag phase) aggregation of misfolded monomers into tiny intermediate oligomers; (Growth phase) re-arrangement of these oligomers into organized fibrils containing the cross-beta structure; (Saturation phase) association of beta structured oligomers into proto-fibrils [<a href="#B120-ijms-23-13505" class="html-bibr">120</a>].</p>
Full article ">
15 pages, 4874 KiB  
Article
Stimulus-Responsive, Gelatin-Containing Supramolecular Nanofibers as Switchable 3D Microenvironments for Cells
by Kentaro Hayashi, Mami Matsuda, Masaki Nakahata, Yoshinori Takashima and Motomu Tanaka
Polymers 2022, 14(20), 4407; https://doi.org/10.3390/polym14204407 - 19 Oct 2022
Cited by 3 | Viewed by 2508
Abstract
Polymer- and/or protein-based nanofibers that promote stable cell adhesion have drawn increasing attention as well-defined models of the extracellular matrix. In this study, we fabricated two classes of stimulus-responsive fibers containing gelatin and supramolecular crosslinks to emulate the dynamic cellular microenvironment in vivo. [...] Read more.
Polymer- and/or protein-based nanofibers that promote stable cell adhesion have drawn increasing attention as well-defined models of the extracellular matrix. In this study, we fabricated two classes of stimulus-responsive fibers containing gelatin and supramolecular crosslinks to emulate the dynamic cellular microenvironment in vivo. Gelatin enabled cells to adhere without additional surface functionalization, while supramolecular crosslinks allowed for the reversible switching of the Young’s modulus through changes in the concentration of guest molecules in culture media. The first class of nanofibers was prepared by coupling the host–guest inclusion complex to gelatin before electrospinning (pre-conjugation), while the second class of nanofibers was fabricated by coupling gelatin to polyacrylamide functionalized with host or guest moieties, followed by conjugation in the electrospinning solution (post-conjugation). In situ AFM nano-indentation demonstrated the reversible switching of the Young’s modulus between 2–3 kPa and 0.2–0.3 kPa under physiological conditions by adding/removing soluble guest molecules. As the concentration of additives does not affect cell viability, the supramolecular fibers established in this study are a promising candidate for various biomedical applications, such as standardized three-dimensional culture matrices for somatic cells and the regulation of stem cell differentiation. Full article
(This article belongs to the Special Issue Fabrication and Application of Electrospun Nanofibers)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p><sup>1</sup>H NMR spectra of (<b>a</b>) gelatin modified with host-guest pair and (<b>b</b>) unmodified gelatin in D<sub>2</sub>O, acquired on a 500 MHz JEOL ECA-500 NMR spectrometer at 25 °C. Chemical shifts were referenced to maleic acid as the standard (δ = 6.2 ppm). Insets show positions of protons in chemical structures corresponding to the peaks in <sup>1</sup>H NMR spectra: 1-6 for CD unit, a-c for Ad unit, and * for maleic acid.</p>
Full article ">Figure 2
<p><sup>1</sup>H NMR spectra of (<b>a</b>) βCD-gelatin and (<b>b</b>) Ad-gelatin in D<sub>2</sub>O containing trifluoroacetic acid (12.5 <span class="html-italic">v</span>/<span class="html-italic">v</span>%), acquired on a 500 MHz JEOL ECA-500 NMR spectrometer at 30 °C. Chemical shifts were referenced to -C<span class="html-italic">H</span>- of the main chain as the standard (b position, δ = 2.2 ppm). Peaks a-c correspond to the main-chain protons, whereas peaks 1–6 correspond to CD and Ad protons.</p>
Full article ">Figure 3
<p>Optical microscopy images of gelatin-βCD-Ad fibers (<b>a</b>) before chemical crosslinking (in air) and (<b>b</b>) after crosslinking with [EDC] = [NHS] =12.5 mM, (<b>c</b>) 25.0 mM, and (<b>d</b>) 37.5 mM in PBS.</p>
Full article ">Figure 4
<p>Topographic profiles of a single gelatin-βCD-Ad fiber using particle-assisted AFM prior to nano-indentation. (<b>a</b>) Schematic illustration of particle-assisted AFM. A topographic scan is essential to find the center of each fiber. (<b>b</b>) Height and (<b>c</b>) line profiles of a single gelatin-βCD-Ad fiber prepared in [EDC] = [NHS] = 12.5 mM, 25.0 mM, and 37.5 mM.</p>
Full article ">Figure 5
<p>Characteristic force–distance curves of a single gelatin-βCD-Ad fiber crosslinked at different concentrations of EDC and NHS: (<b>a</b>) [EDC] = [NHS] =12.5 mM, (<b>b</b>) 25.0 mM, and (<b>c</b>) 37.5 mM. The best fit results obtained using the Hertz model for a spherical indenter is shown as solid lines, yielding the following Young’s moduli: <span class="html-italic">E</span><sub>12.5</sub> = 16 kPa, <span class="html-italic">E</span><sub>25.0</sub> = 32 kPa, and <span class="html-italic">E</span><sub>37.5</sub> = 42 kPa. (<b>d</b>) Young’s modulus of gelatin-βCD-Ad fibers in the absence and presence of 10 mM Ad-COONa (<span class="html-italic">N</span> = 4).</p>
Full article ">Figure 6
<p>AFM topographic profiles of βCD-gelatin/Ad-gelatin fibers (<b>a</b>) before chemical crosslinking (in air) and after crosslinking with (<b>b</b>) [EDC] = [NHS] = 100 mM, (<b>c</b>) 200 mM, (<b>d</b>) 400 mM, and (<b>e</b>) 2 M (in PBS). (<b>f</b>) Force–distance curves of a βCD-gelatin/Ad-gelatin fiber crosslinked in [EDC] = [NHS] = 2 M.</p>
Full article ">Figure 7
<p>Young’s modulus (<b>a</b>,<b>c</b>) and thickness (<b>b</b>,<b>d</b>) of βCD-Gelatin/Ad-Gelatin fibers in the absence and presence of 5 mM Ad-COONa monitored by in situ AFM nano-indentation (<span class="html-italic">N</span> = 3). Fibers were crosslinked in (<b>a</b>,<b>b</b>) 400 mM and (<b>c</b>,<b>d</b>) [EDC] = [NHS] = 2 M for 4 h.</p>
Full article ">Scheme 1
<p>Overview of the fabrication of stimulus-responsive nanofibers.</p>
Full article ">Scheme 2
<p>Preparation of pre-conjugated gelatin-βCD-Ad (Method 1).</p>
Full article ">Scheme 3
<p>Synthesis of (<b>a</b>) βCD gelatin and (<b>b</b>) Ad-gelatin for post-conjugation (Method 2).</p>
Full article ">
20 pages, 1960 KiB  
Review
A Promising Review on Cyclodextrin Conjugated Paclitaxel Nanoparticles for Cancer Treatment
by Kamini Velhal, Sagar Barage, Arpita Roy, Jaya Lakkakula, Ramesh Yamgar, Mohammed S. Alqahtani, Krishna Kumar Yadav, Yongtae Ahn and Byong-Hun Jeon
Polymers 2022, 14(15), 3162; https://doi.org/10.3390/polym14153162 - 3 Aug 2022
Cited by 14 | Viewed by 2897
Abstract
This review presented the unique characteristics of different types of cyclodextrin polymers by non-covalent host–guest interactions to synthesize an inclusion complex. Various cancers are treated with different types of modified cyclodextrins, along with the anticancer drug paclitaxel. PTX acts as a mitotic inhibitor, [...] Read more.
This review presented the unique characteristics of different types of cyclodextrin polymers by non-covalent host–guest interactions to synthesize an inclusion complex. Various cancers are treated with different types of modified cyclodextrins, along with the anticancer drug paclitaxel. PTX acts as a mitotic inhibitor, but due to its low dissolution and permeability in aqueous solutions, it causes considerable challenges for drug delivery system (DDS) designs. To enhance the solubility, it is reformulated with derivatives of cyclodextrins using freeze-drying and co-solvent lyophilization methods. The present supramolecular assemblies involve cyclodextrin as a key mediator, which is encapsulated with paclitaxel and their controlled release at the targeted area is highlighted using different DDS. In addition, the application of cyclodextrins in cancer treatment, which reduces the off-target effects, is briefly demonstrated using various types of cancer cell lines. A new nano-formulation of PTX is used to improve the antitumor activity compared to normal PTX DDS in lungs and breast cancer is well defined in the present review. Full article
(This article belongs to the Section Biobased and Biodegradable Polymers)
Show Figures

Figure 1

Figure 1
<p>Structure of three parent cyclodextrin. (Adapted with permission from Ref. [<a href="#B14-polymers-14-03162" class="html-bibr">14</a>]. Copyright © 2009 American Chemical Society).</p>
Full article ">Figure 2
<p>Schematic illustration of the chemical structures and construction of the HACD-AuNPs and the drug @HACD-AuNPs. (Adapted with permission from Ref. [<a href="#B25-polymers-14-03162" class="html-bibr">25</a>]).</p>
Full article ">Figure 3
<p>Scanning electron microscopy (SEM) photomicrographs of PCX (<b>a</b>); PCX:6OCaproβCD inclusion complex (<b>b</b>) and PCX:PC βCDC6 inclusion complex (<b>c</b>). (Adapted with permission from Ref. [<a href="#B30-polymers-14-03162" class="html-bibr">30</a>]).</p>
Full article ">Figure 4
<p>Schematic illustration of the construction of pH-sensitive PTX nanoformulation based on acetylated a-CD (Ac-aCD). (Adapted with permission from Ref. [<a href="#B40-polymers-14-03162" class="html-bibr">40</a>]. Copyright © 2013 Elsevier Ltd).</p>
Full article ">
28 pages, 14096 KiB  
Review
Lead(II)-Azido Metal–Organic Coordination Polymers: Synthesis, Structure and Application in PbO Nanomaterials Preparation
by Jaber Dadashi, Mohammad Khaleghian, Younes Hanifehpour, Babak Mirtamizdoust and Sang Woo Joo
Nanomaterials 2022, 12(13), 2257; https://doi.org/10.3390/nano12132257 - 30 Jun 2022
Cited by 5 | Viewed by 2851
Abstract
The current study aims to explain recent developments in the synthesis of Pb(II)-azido metal-organic coordination polymers. Coordination polymers are defined as hybrid materials encompassing metal-ion-based, organic linkers, vertices, and ligands, serving to link the vertices to 1D, 2D, or 3D periodic configurations. The [...] Read more.
The current study aims to explain recent developments in the synthesis of Pb(II)-azido metal-organic coordination polymers. Coordination polymers are defined as hybrid materials encompassing metal-ion-based, organic linkers, vertices, and ligands, serving to link the vertices to 1D, 2D, or 3D periodic configurations. The coordination polymers have many applications and potential properties in many research fields, primarily dependent on particular host–guest interactions. Metal coordination polymers (CPs) and complexes have fascinating structural topologies. Therefore, they have found numerous applications in different areas over the past two decades. Azido-bridged complexes are inorganic coordination ligands with higher fascination that have been the subject of intense research because of their coordination adaptability and magnetic diversity. Several sonochemical methods have been developed to synthesize nanostructures. Researchers have recently been interested in using ultrasound in organic chemistry synthetics, since ultrasonic waves in liquids accelerate chemical reactions in heterogeneous and homogeneous systems. The sonochemical synthesis of lead–azide coordination compounds resulted from very fantastic morphologies, and some of these compounds are used as precursors for preparing nano lead oxide. The ultrasonic sonochemistry approach has been extensively applied in different research fields, such as medical imaging, biological cell disruption, thermoplastic welding, food processing, and waste treatment. CPs serve as appropriate precursors for preparing favorable materials at the nanoscale. Using these polymers as precursors is beneficial for preparing inorganic nanomaterials such as metal oxides. Full article
(This article belongs to the Special Issue Latest Advances in Inorganic Nanomaterials)
Show Figures

Figure 1

Figure 1
<p>Progress in interest in synthesizing coordination polymers (CPs) (Source: Scopus).</p>
Full article ">Figure 2
<p>From 1D CP in [Pb(baea)(N<sub>3</sub>).(N<sub>3</sub>)]<sub>n</sub> to 2D CP with supramolecular interactions. Reprinted with permission from [<a href="#B83-nanomaterials-12-02257" class="html-bibr">83</a>]. Copyright 2012, Springer Nature.</p>
Full article ">Figure 3
<p>ORTEP diagram and representation of Pb(II) space in [Pb<sub>2</sub>(µ-N<sub>3</sub>)(µ-NO<sub>3</sub>)L<sub>2</sub>]<sub>n</sub>. Reprinted with permission from [<a href="#B79-nanomaterials-12-02257" class="html-bibr">79</a>]. Copyright 2013, Elsevier Ltd., Amsterdam, The Netherlands.</p>
Full article ">Figure 4
<p>ORTEP view of [Pb<sub>2</sub>(pbap)(N<sub>3</sub>)<sub>4</sub>] (1) with atom-numbering scheme and 20% probability ellipsoids for all non-H atoms. Reprinted with permission from [<a href="#B84-nanomaterials-12-02257" class="html-bibr">84</a>]. Copyright 2010, Elsevier Ltd., Amsterdam, The Netherlands.</p>
Full article ">Figure 5
<p>(<b>a</b>) A piece of the CP indicating the one-dimensional (1D) zig-zag polymeric chain and (<b>b</b>) the relation between chains linked with dual symmetry through Pb–N3 relations. Reprinted with permission from [<a href="#B86-nanomaterials-12-02257" class="html-bibr">86</a>]. Copyright 2015, Springer Nature.</p>
Full article ">Figure 6
<p>A piece of the CP indicating a 1D polymer. Reprinted with permission from [<a href="#B87-nanomaterials-12-02257" class="html-bibr">87</a>]. Copyright 2009, WILEY-VCH Verlag GmbH &amp; Co. KGaA, Weinheim.</p>
Full article ">Figure 7
<p>SEM images of the nano-hexagonal rods. Reprinted with permission from [<a href="#B88-nanomaterials-12-02257" class="html-bibr">88</a>]. Copyright 2012, Springer Nature.</p>
Full article ">Figure 8
<p>Packing of 1D chains for the formation of 3D supramolecular layers through pi…pi stacking relations. Reprinted with permission from [<a href="#B88-nanomaterials-12-02257" class="html-bibr">88</a>]. Copyright 2012, Springer Nature.</p>
Full article ">Figure 9
<p>Piece of CP indicating a 1D zig-zag polymer. Reprinted with permission from [<a href="#B81-nanomaterials-12-02257" class="html-bibr">81</a>]. Copyright 2013, Springer Nature.</p>
Full article ">Figure 10
<p>SEM images of PbO nanopowders created by [Pb(pcih)N<sub>3</sub>MeOH]<sub>n</sub> calculations. Reprinted with permission from [<a href="#B89-nanomaterials-12-02257" class="html-bibr">89</a>]. Copyright 2013, Springer Nature.</p>
Full article ">Figure 11
<p>From 3 nuclear building blocks (monomer) to 1D CP. Reprinted with permission from [<a href="#B90-nanomaterials-12-02257" class="html-bibr">90</a>]. Copyright 2015, Springer Nature.</p>
Full article ">Figure 12
<p>Piece of the CP indicating the 1D polymer. Reprinted with permission from [<a href="#B91-nanomaterials-12-02257" class="html-bibr">91</a>]. Copyright 2011, WILEY-VCH Verlag GmbH &amp; Co. KGaA, Weinheim.</p>
Full article ">Figure 13
<p>C axis indicates the lead atoms’ coordination. Reprinted with permission from [<a href="#B92-nanomaterials-12-02257" class="html-bibr">92</a>]. Copyright 2010, Elsevier Ltd., Amsterdam, The Netherlands.</p>
Full article ">Figure 14
<p>SEM images of (<b>a</b>) [Pb(3-pyc)(N<sub>3</sub>)(H<sub>2</sub>O)]<sub>n</sub> (1) nanostructure balls in aqua solution, (<b>b</b>) [Pb(3-pyc)(N<sub>3</sub>)(H<sub>2</sub>O)]<sub>n</sub> (1) nanoplates inethanolic solution, (<b>c</b>) [Pb(3-pyc)I]<sub>n</sub> (2) NPs, and (<b>d</b>) [Pb(3-pyc)Br]<sub>n</sub> (3) nanofibers. Reprinted with permission from [<a href="#B93-nanomaterials-12-02257" class="html-bibr">93</a>]. Copyright 2010, Elsevier Ltd., Amsterdam, The Netherlands.</p>
Full article ">Figure 15
<p>ORTEP image of complex [Pb(2-pyc)(N<sub>3</sub>)(H<sub>2</sub>O)]<sub>n</sub>. i: −x, y + 1/2, −z + 1/2; ii: −x, −y, −z; iii: −x, −y + 1/2, z + 1/2. Reprinted with permission from [<a href="#B81-nanomaterials-12-02257" class="html-bibr">81</a>]. Copyright 2010, Springer Nature.</p>
Full article ">Figure 16
<p>Molecular configuration of [Pb<sub>2</sub>(tmph)<sub>2</sub>(μ-N<sub>3</sub>)<sub>2</sub>(CH<sub>3</sub>COO)<sub>2</sub>]. Reprinted with permission from [<a href="#B94-nanomaterials-12-02257" class="html-bibr">94</a>]. Copyright 2015, Elsevier Ltd., Amsterdam, The Netherlands.</p>
Full article ">Figure 17
<p>Schematic view of Pb(II) space. Reprinted with permission from [<a href="#B94-nanomaterials-12-02257" class="html-bibr">94</a>]. Copyright 2015, Elsevier Ltd., Amsterdam, The Netherlands.</p>
Full article ">Figure 18
<p>[Pb<sub>3</sub>(tmph)<sub>4</sub>(µ-N<sub>3</sub>)<sub>5</sub>(µ-NO<sub>3</sub>)] molecular structure (A = i = −x, −y, −z). Reprinted with permission from [<a href="#B95-nanomaterials-12-02257" class="html-bibr">95</a>]. Copyright 2015, Elsevier Ltd., Amsterdam, The Netherlands.</p>
Full article ">Figure 19
<p>SEM photographs of [Pb(phen)(μ-N<sub>3</sub>)(μ-NO<sub>3</sub>)]<sub>n</sub> nanorods obtained from ultrasonic radiation. Reprinted with permission from [<a href="#B96-nanomaterials-12-02257" class="html-bibr">96</a>]. Copyright 2012, Springer Nature.</p>
Full article ">Figure 20
<p>A piece of the CP indicating a 1D polymer. Reprinted with permission from [<a href="#B97-nanomaterials-12-02257" class="html-bibr">97</a>]. Copyright 2006, WILEY-VCH Verlag GmbH &amp; Co. KGaA, Weinheim.</p>
Full article ">Figure 21
<p>XRD pattern of PbO prepared after thermolysis of compound [Pb2(tmph)2(µ-N3)2(CH3COO)2]. Reprinted with permission from [<a href="#B94-nanomaterials-12-02257" class="html-bibr">94</a>]. Copyright 2015, Elsevier Ltd., Amsterdam, The Netherlands.</p>
Full article ">Figure 22
<p>SEM photographs of PbO nanopowders (produced by thermolysis of nanorods). Reprinted with permission from [<a href="#B94-nanomaterials-12-02257" class="html-bibr">94</a>]. Copyright 2015, Elsevier Ltd., Amsterdam, The Netherlands.</p>
Full article ">Figure 23
<p>TEM photographs of PbO nanopowders (produced by thermolysis of nanorods). Reprinted with permission from [<a href="#B95-nanomaterials-12-02257" class="html-bibr">95</a>]. Copyright 2015, Elsevier Ltd., Amsterdam, The Netherlands.</p>
Full article ">Figure 24
<p>XRD pattern of PbO prepared after thermolysis of compound [Pb3(tmph)4(µ-N3)5(µ-NO3)]n. Reprinted with permission from [<a href="#B95-nanomaterials-12-02257" class="html-bibr">95</a>]. Copyright 2015, Elsevier Ltd., Amsterdam, The Netherlands.</p>
Full article ">Scheme 1
<p>Some of the applications and potential properties of coordination polymers.</p>
Full article ">Scheme 2
<p>The stereochemical effect of the 6s pair on the coordination sphere (a ligand’s D donor atom).</p>
Full article ">Scheme 3
<p>Usual bridging states of azido ligand.</p>
Full article ">Scheme 4
<p>Application of ultrasonic sonochemistry method in different research fields.</p>
Full article ">Scheme 5
<p>Schematic graphs of synthetic methods.</p>
Full article ">Scheme 6
<p>Materials produced and synthetic manners.</p>
Full article ">Scheme 7
<p>Produced materials and synthetic approaches.</p>
Full article ">Scheme 8
<p>Produced materials and synthetic approaches.</p>
Full article ">
Back to TopTop