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

Article Types

Countries / Regions

Search Results (84)

Search Parameters:
Keywords = polyNIPAM

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
15 pages, 2602 KiB  
Article
A Novel Approach for the Synthesis of Responsive Core–Shell Nanogels with a Poly(N-Isopropylacrylamide) Core and a Controlled Polyamine Shell
by Anna Harsányi, Attila Kardos, Pinchu Xavier, Richard A. Campbell and Imre Varga
Polymers 2024, 16(18), 2584; https://doi.org/10.3390/polym16182584 - 13 Sep 2024
Viewed by 590
Abstract
Microgel particles can play a key role, e.g., in drug delivery systems, tissue engineering, advanced (bio)sensors or (bio)catalysis. Amine-functionalized microgels are particularly interesting in many applications since they can provide pH responsiveness, chemical functionalities for, e.g., bioconjugation, unique binding characteristics for pollutants and [...] Read more.
Microgel particles can play a key role, e.g., in drug delivery systems, tissue engineering, advanced (bio)sensors or (bio)catalysis. Amine-functionalized microgels are particularly interesting in many applications since they can provide pH responsiveness, chemical functionalities for, e.g., bioconjugation, unique binding characteristics for pollutants and interactions with cell surfaces. Since the incorporation of amine functionalities in controlled amounts with predefined architectures is still a challenge, here, we present a novel method for the synthesis of responsive core–shell nanogels (dh < 100 nm) with a poly(N-isopropylacrylamide) (pNIPAm) core and a polyamine shell. To achieve this goal, a surface-functionalized pNIPAm nanogel was first prepared in a semi-batch precipitation polymerization reaction. Surface functionalization was achieved by adding acrylic acid to the reaction mixture in the final stage of the precipitation polymerization. Under these conditions, the carboxyl functionalities were confined to the outer shell of the nanogel particles, preserving the core’s temperature-responsive behavior and providing reactive functionalities on the nanogel surface. The polyamine shell was prepared by the chemical coupling of polyethyleneimine to the nanogel’s carboxyl functionalities using a water-soluble carbodiimide (EDC) to facilitate the coupling reaction. The efficiency of the coupling was assessed by varying the EDC concentration and reaction temperature. The molecular weight of PEI was also varied in a wide range (Mw = 0.6 to 750 kDa), and we found that it had a profound effect on how many polyamine repeat units could be immobilized in the nanogel shell. The swelling and the electrophoretic mobility of the prepared core–shell nanogels were also studied as a function of pH and temperature, demonstrating the successful formation of the polyamine shell on the nanogel core and its effect on the nanogel characteristics. This study provides a general framework for the controlled synthesis of core–shell nanogels with tunable surface properties, which can be applied in many potential applications. Full article
(This article belongs to the Special Issue Smart and Bio-Medical Polymers)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>The hydrodynamic size of (<b>a</b>) the carboxyl-functionalized pNIPAm nanogels and (<b>b</b>) a non-functionalized pNIPAm microgel in 10 mM HCl (pH = 2) and in 10 mM NaCl (pH = 7) as a function of temperature. Lines are only visual guides.</p>
Full article ">Figure 2
<p>(<b>a</b>) The electrophoretic mobility and (<b>b</b>) the hydrodynamic size of the carboxyl-functionalized pNIPAm nanogels as a function of pH at constant ionic strength (I = 10 mM). Lines are only visual guides.</p>
Full article ">Figure 3
<p>The variation in the microgel-bound PEI as a function of EDC excess used in the coupling reaction. The amount of microgel-bound PEI is given as a percentage of the total amount of PEI added to the reaction mixture. The lines are only visual guides.</p>
Full article ">Figure 4
<p>(<b>a</b>) The electrophoretic mobility and (<b>b</b>) the hydrodynamic size of the core–shell nanogels prepared with coupling low-molecular-weight PEI (<span class="html-italic">M<sub>w</sub></span> = 0.6 kDa) with a stoichiometric amount of EDC (in green) and using an EDC excess of ten times (in red). As a reference, data determined for the pNIPAm core are also plotted (in blue). Lines are only visual guides.</p>
Full article ">Figure 5
<p>The microgel-bound PEI as a function of the PEI molecular weight used in the coupling reaction. The amount of microgel-bound PEI is given as a percentage of the total amount of PEI added to the reaction mixture. The line is only a visual guide.</p>
Full article ">Figure 6
<p>The electrophoretic mobility of the core–shell nanogels prepared with coupling of PEIs with different molecular weights. As a reference, the electrophoretic mobility data determined for the pNIPAm core was also plotted (in blue). Lines are only visual guides.</p>
Full article ">Figure 7
<p>The hydrodynamic size of the core–shell pNIPAm nanogels prepared by coupling 10 kDa PEI as a function of temperature, measured in 10 mM HCl or in 10 mM NaOH. Lines are only visual guides.</p>
Full article ">Scheme 1
<p>A schematic representation of the synthesis protocol proposed to prepare pNIPAm nanogels with a polyamine shell. In the first step of the procedure, the precipitation polymerization of the nanogel beads is initiated at <span class="html-italic">t</span><sub>0</sub> with the addition of APS, and then after time <span class="html-italic">t</span><sub>1</sub> (when most of the monomers are already reacted), a second batch of monomers, including acrylic acid (AAc), is added to the reaction mixture to form the carboxyl-functionalized outer shell. These surface-functionalized nanogel beads are used to graft polyamine molecules to the carboxyl functionalities using EDC coupling to facilitate the formation of the polyamine shell in the second step of the protocol. (The yellow/green transition of the pNIPAm core indicates its decreasing crosslink density towards its surface; the blue shell depicts the acrylic-acid-functionalized outer shell of the nanogel core particles; the red shell represents the polyamine shell formed in the final coupling reaction.)</p>
Full article ">
16 pages, 8434 KiB  
Article
Keratin–PNIPAM Hybrid Microgels: Preparation, Morphology and Swelling Properties
by Elena Buratti, Maddalena Sguizzato, Giovanna Sotgiu, Roberto Zamboni and Monica Bertoldo
Gels 2024, 10(6), 411; https://doi.org/10.3390/gels10060411 - 20 Jun 2024
Viewed by 1120
Abstract
Combinations of synthetic polymers, such as poly(N-isopropylacrylamide) (PNIPAM), with natural biomolecules, such as keratin, show potential in the field of biomedicine, since these hybrids merge the thermoresponsive properties of PNIPAM with the bioactive characteristics of keratin. This synergy aims to produce hybrids that [...] Read more.
Combinations of synthetic polymers, such as poly(N-isopropylacrylamide) (PNIPAM), with natural biomolecules, such as keratin, show potential in the field of biomedicine, since these hybrids merge the thermoresponsive properties of PNIPAM with the bioactive characteristics of keratin. This synergy aims to produce hybrids that can respond to environmental stimuli while maintaining biocompatibility and functionality, making them suitable for various medical and biotechnological uses. In this study, we exploit keratin derived from wool waste in the textile industry, extracted via sulfitolysis, to synthesize hybrids with PNIPAM microgel. Utilizing two distinct methods—polymerization of NIPAM with keratin (HYB-P) and mixing preformed PNIPAM microgels with keratin (HYB-M)—resulted in hybrids with 20% and 25% keratin content, respectively. Dynamic light scattering (DLS) and transmission electron microscopic (TEM) analyses indicated the formation of colloidal systems with particle sizes of around 110 nm for HYB-P and 518 nm for HYB-M. The presence of keratin in both systems, 20% and 25%, respectively, was confirmed by spectroscopic (FTIR and NMR) and elemental analyses. Distinct structural differences were observed between HYB-P and HYB-M, suggesting a graft copolymer configuration for the former hybrid and a complexation for the latter one. Furthermore, these hybrids demonstrated temperature responsiveness akin to PNIPAM microgels and pH responsiveness, underscoring their potential for diverse biomedical applications. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Schemes of the synthesis of (<b>a</b>) pure PNIPAM, (<b>b</b>) PNIPAM–keratin hybrid microgel by polymerization of NIPAM in presence of keratin (KER) and (<b>c</b>) PNIPAM–keratin hybrid microgel by mixing a preformed PNIPAM microgel with keratin (KER).</p>
Full article ">Figure 2
<p>Microgels’ size distribution as determined by DLS analysis (<b>a</b>) and representative TEM micrographs of PNIPAM (<b>b</b>) (8500× mag), HYP-P (<b>c</b>) (28,000× mag) and HYB-M (<b>d</b>) (8500× mag) microgels.</p>
Full article ">Figure 3
<p>Comparison of the ATR spectra of PNIPAM, HYB-P, HYB-M and KER (<b>a</b>). Detail of the amide I spectral region (<b>b</b>). Second derivative of the amide I spectral region (<b>c</b>). Individual spectra are reported in the <a href="#app1-gels-10-00411" class="html-app">Supplementary Materials (Figures S1–S4)</a>.</p>
Full article ">Figure 4
<p>(<b>a</b>) Particle diameters, determined by DLS analysis at 25 °C, of HYB-P, HYB-M, KER and PNIPAM during dialysis into a membrane with a cutoff of 100 kDa for two weeks. (<b>b</b>) Particle size distribution, analyzed by DLS, of HYB-M after 1 day and 3 days of dialysis.</p>
Full article ">Figure 5
<p>Comparison of the <sup>1</sup>H NMR spectra of keratin (KER), PNIPAM, HYB-P and HYB-M in D<sub>2</sub>O. In the PNIPAM spectrum, the polymer structure with labels (a–d) for the assigned peaks are reported. In the keratin spectrum, the structure of the macromolecular chain is reported with labels (e and f), showing the oxidized cysteine (ox cys) repeating unit and a generic amino acid repeating unit. In the oxidized cysteine structure, S<sub>OX</sub> can be SOH (sulfenic acid), SOOH (sulfinic acid) or SOOOH (cysteic acid). Individual spectra are reported in the <a href="#app1-gels-10-00411" class="html-app">Supplementary Materials (Figures S5–S8)</a>.</p>
Full article ">Figure 6
<p>Data obtained by DLS analysis for PNIPAM, HYB-P and HYB-M: particle diameters as a function of temperature in the 20–40 °C range at pH 6 (<b>a</b>) and volume phase transition temperatures (VPTT) (<b>b</b>) and swelling ratios (α) (<b>c</b>) at pH 2.5, 6 and 8.5. Error bars for each point are standard deviations of three measurements; lines in (<b>a</b>) are fits of the data determined with a Boltzmann equation.</p>
Full article ">Figure 7
<p>Particle diameters as a function of temperature in the 20–40 °C range at three different pH levels (3.5, 6 and 8.5) for PNIPAM (<b>a</b>), HYB-P (<b>b</b>), HYB-M (<b>c</b>) and KER (<b>d</b>). Error bars for each point are standard deviations of three measurements; lines are fits of the data determined with a Boltzmann equation.</p>
Full article ">
13 pages, 9094 KiB  
Article
Preparation and Application of Responsive Nanocellulose Composites
by Yanhui Zhou, Lu Zhang and Yuan Li
Polymers 2024, 16(11), 1446; https://doi.org/10.3390/polym16111446 - 21 May 2024
Viewed by 710
Abstract
Cellulose nanofibrils/poly(N-Isopropylacrylamide) semi-interpenetrating networks (MMCNF-PNAs) were synthesized using an in situ fabrication (semi-IPN). The polymerization of N-isopropylacrylamide (NIPAM) (free radical) was conducted in the presence of magnetic modified cellulose nanofibrils (MMCNFs). The adsorption behaviors and surface morphology of the synthesized adsorbents were investigated [...] Read more.
Cellulose nanofibrils/poly(N-Isopropylacrylamide) semi-interpenetrating networks (MMCNF-PNAs) were synthesized using an in situ fabrication (semi-IPN). The polymerization of N-isopropylacrylamide (NIPAM) (free radical) was conducted in the presence of magnetic modified cellulose nanofibrils (MMCNFs). The adsorption behaviors and surface morphology of the synthesized adsorbents were investigated systematically. The adsorption behaviors of the as-prepared MMCNF-PNA towards methylene blue (MB, as the model contaminant) dye was studied, and the optimal adsorption conditions were also studied. The adsorption processes could be well fitted using pseudo-second-order and Elovich kinetic models. Meanwhile, Langmuir and Freundlich isotherm models were used to fit the adsorption which occurred at 25, 37 and 65 °C. The corresponding results showed that the Freundlich isotherm model fitted the adsorption process better, indicating that the dye’s adsorption happened via heterogeneous adsorptive energies on the prepared MMCNF-PNAs. Their desorption and reusability were also studied to verify magnetic responsivity. To sum up, MMCNF-PNAs are promising magnetic and thermal stimuli-responsive adsorbents, showing a controlled adsorption/desorption process. Full article
(This article belongs to the Special Issue Cellulose-Based Polymeric Materials)
Show Figures

Figure 1

Figure 1
<p>The FTIR spectra of CNF, MMCNF, MMCNF-PNA-1, MMCNF-PNA-2 and MMCNF-PNA-3.</p>
Full article ">Figure 2
<p>The SEM images of CNF (<b>a</b>), MMCNF (<b>b</b>), MMCNF-PNA-1 (<b>c</b>), MMCNF-PNA-2 (<b>d</b>) and MMCNF-PNA-3 (<b>e</b>), EDS-Fe element mapping for MMCNF-PNA-1, MMCNF-PNA-2 and MMCNF-PNA-3, the EDS images of Fe element on MMCNF-PNA-1 (<b>f</b>), MMCNF-PNA-2 (<b>g</b>), MMCNF-PNA-3 (<b>h</b>).</p>
Full article ">Figure 3
<p>The curves of TG (<b>a</b>) and DTG (<b>b</b>) (CNF, MMCNF, MMCNF-PNA-1, MMCNF-PNA-2 and MMCNF-PNA-3).</p>
Full article ">Figure 4
<p>Effect of contact time on adsorption (the area 1, 2, and 3 in the figures are divided to show the different adsorption stages).</p>
Full article ">Figure 5
<p>Kinetics of the MB adsorption using MMCNF-PNA-1 (<b>a</b>,<b>b</b>—pseudo-second-order), MMCNF-PNA-2 (<b>c</b>,<b>d</b>—pseudo-first-order) and MMCNF-PNA-3 (<b>e</b>,<b>f</b>—Elovich model).</p>
Full article ">Figure 6
<p>The adsorption isotherms for the adsorption of MB on MMCNF-PNA-1, MMCNF-PNA-2 and MMCNF-PNA-3: (<b>a</b>) Langmuir model and (<b>b</b>) Freundlich model.</p>
Full article ">Figure 7
<p>Desorption behavior of MB on adsorbents.</p>
Full article ">Figure 8
<p>The thermal-responsive behavior of the adsorbents.</p>
Full article ">
17 pages, 6415 KiB  
Article
The Effects of Incorporating Nanoclay in NVCL-NIPAm Hydrogels on Swelling Behaviours and Mechanical Properties
by Billy Shu Hieng Tie, Eyman Manaf, Elaine Halligan, Shuo Zhuo, Gavin Keane, Joseph Geever and Luke Geever
Nanomaterials 2024, 14(7), 597; https://doi.org/10.3390/nano14070597 - 28 Mar 2024
Cited by 2 | Viewed by 1065
Abstract
Following the formulation development from a previous study utilising N-vinylcaprolactam (NVCL) and N-isopropylacrylamide (NIPAm) as monomers, poly(ethylene glycol) dimethacrylate (PEGDMA) as a chemical crosslinker, and Irgacure 2959 as photoinitiator, nanoclay (NC) is now incorporated into the selected formulation for enhanced mechanical performance and [...] Read more.
Following the formulation development from a previous study utilising N-vinylcaprolactam (NVCL) and N-isopropylacrylamide (NIPAm) as monomers, poly(ethylene glycol) dimethacrylate (PEGDMA) as a chemical crosslinker, and Irgacure 2959 as photoinitiator, nanoclay (NC) is now incorporated into the selected formulation for enhanced mechanical performance and swelling ability. In this research, two types of NC, hydrophilic bentonite nanoclay (NCB) and surface-modified nanoclay (NCSM) of several percentages, were included in the formulation. The prepared mixtures were photopolymerised, and the fabricated gels were characterised through Fourier transform infrared spectroscopy (FTIR), cloud-point measurements, ultraviolet (UV) spectroscopy, pulsatile swelling, rheological analysis, and scanning electron microscopy (SEM). Furthermore, the effect of swelling temperature, NC types, and NC concentration on the hydrogels’ swelling ratio was studied through a full-factorial design of experiment (DOE). The successful photopolymerised NC-incorporated NVCL-NIPAm hydrogels retained the same lower critical solution temperature (LCST) as previously. Rheological analysis and SEM described the improved mechanical strength and polymer orientation of gels with any NCB percentage and low NCSM percentage. Finally, the temperature displayed the most significant effect on the hydrogels’ swelling ability, followed by the NC types and NC concentration. Introducing NC to hydrogels could potentially make them suitable for applications that require good mechanical performance. Full article
Show Figures

Figure 1

Figure 1
<p>Photopolymerised chemically crosslinked gels containing (<b>a</b>) 0.1% NCB, (<b>b</b>) 1.0% NCB, (<b>c</b>) 3.0% NCB, (<b>d</b>) 0.1% NCSM, (<b>e</b>) 1.0% NCSM, and (<b>f</b>) 3.0% NCSM.</p>
Full article ">Figure 2
<p>Interaction plot for the conducted DOE. Interaction terms are denoted by inserting an asterisk between the variables intended for interaction.</p>
Full article ">Figure 3
<p>FTIR spectra of FC34, 1.0NCB_FC34, and 1.0NCSM_FC34.</p>
Full article ">Figure 4
<p>Transparent aqueous solution of 1.0NCSM_FP34 (<b>left</b>) became cloudy (<b>right</b>).</p>
Full article ">Figure 5
<p>UV spectroscopy of the aqueous solutions applying a wavelength of 500 nm.</p>
Full article ">Figure 6
<p>Dried gels after pulsatile swelling: (<b>a</b>) FC34, (<b>b</b>) 0.1NCB_FC34, (<b>c</b>) 1.0NCB_FC34, (<b>d</b>) 3.0NCB_FC34, (<b>e</b>) 0.1NCSM_FC34, (<b>f</b>) 1.0NCSM_FC34, and (<b>g</b>) 3.0NCSM_FC34.</p>
Full article ">Figure 7
<p>Pulsatile swelling curves of FC34, 3.0NCB_FC34, and 3.0NCSM_FC34. Swelling at room temperature including 0–144 h, 168–192 h, 216–288 h, 312–336 h, and 360–384 h; swelling at 50 °C including 144–168 h, 192–216 h, 288–312 h, and 336–360 h.</p>
Full article ">Figure 8
<p>Comparison of NC-incorporated hydrogels storage modulus at 0.1% strain.</p>
Full article ">Figure 9
<p>SEM images under 400× magnification with identical scale.</p>
Full article ">Figure A1
<p>Pareto chart for the conducted DOE. The dashed line represents the reference line that is at 2.07, factor bars cross the reference line are statistically significant at the 0.05 level with the current model terms.</p>
Full article ">
18 pages, 6268 KiB  
Article
3D Polymer Gel Dosimeters with iCBCT 3D Reading and polyGeVero-CT Software Package for Quality Assurance in Radiotherapy
by Marek Kozicki, Piotr Maras and Malwina Jaszczak-Kuligowska
Materials 2024, 17(6), 1283; https://doi.org/10.3390/ma17061283 - 11 Mar 2024
Cited by 1 | Viewed by 916
Abstract
Dynamically evolving radiotherapy instruments require advancements in compatible 3D dosimetry systems. This paper reports on such tools for the coincidence test of the mechanical and radiation isocenter for a medical accelerator as part of the quality assurance in routine radiotherapy practice. Three-dimensional polymer [...] Read more.
Dynamically evolving radiotherapy instruments require advancements in compatible 3D dosimetry systems. This paper reports on such tools for the coincidence test of the mechanical and radiation isocenter for a medical accelerator as part of the quality assurance in routine radiotherapy practice. Three-dimensional polymer gel dosimeters were used in combination with 3D reading by iterative cone beam computed tomography and 3D data processing using the polyGeVero-CT software package. Different polymer gel dosimeters were used with the following acronyms: VIP, PAGAT, MAGIC, and NIPAM. The same scheme was used for each dosimeter: (i) irradiation sensitivity test for the iterative cone beam computed tomography reading to determine the appropriate monitor unit for irradiation, and (ii) verification of the chosen irradiation conditions by a star-shot 2D irradiation of each 3D dosimeter in the direction of performing the test. This work concludes with the optimum monitor unit per beam for each selected 3D dosimeter, delivers schemes for quick and easy determination of the radiation isocenter and performing the coincidence test. Full article
Show Figures

Figure 1

Figure 1
<p>A scheme illustrating the types of 3D dosimeters and their measurement methods.</p>
Full article ">Figure 2
<p>Photographs of non-irradiated and irradiated 3D polymer gel dosimeters. Irradiation was performed to study the impact of the number of monitor units (MU) on the reading of polymerized and crosslinked regions in 3D polymer gel dosimeters using iCBCT (mode pelvis, mean of 3 series) (TrueBeam, Varian, Palo Alto, CA, USA). The dosimeters are in 1 L containers (PH-5-DD1, GeVero Co., Lodz, Poland): VIP (<b>A</b>), MAGIC (<b>B</b>), and in ~0.6 L containers (PH-6-DD2, GeVero Co., Lodz, Poland): NIPAM (<b>C</b>), and PAGAT (<b>D</b>). The photographs of the dosimeters are before (left) and after (right) irradiation (500, 1000, and 1500 MU—bottom lines, and 2000, 5000, and 10,000 MU—top lines).</p>
Full article ">Figure 3
<p>Establishing the radiation field parameters for isocenter determination using 3D polymer gel dosimeters: VIP (<b>A</b>,<b>B</b>), MAGIC (<b>C</b>,<b>D</b>), NIPAM (<b>E</b>,<b>F</b>), PAGAT (<b>G</b>,<b>H</b>). Both iCBCT transversal images (TrueBeam, Varian, Palo Alto, CA, USA; fixed MLC gap of 2 mm) for two irradiated regions of 500, 1000, and 1500 (first region) and 2000, 5000, and 10,000 MU (second region) (<b>A</b>,<b>C</b>,<b>E</b>,<b>G</b>) and profiles (<b>B</b>,<b>D</b>,<b>F</b>,<b>H</b>) across these regions are shown. The position where the profiles were taken is indicated on the iCBCT images by blue (first region) and orange (second region) dotted lines. The profiles were smoothed using a mean filter: Kernel mode: 3D, kernel unit: mm, and kernel size 3. Data were processed using the polyGeVero-CT software package (GeVero Co., Lodz, Poland).</p>
Full article ">Figure 3 Cont.
<p>Establishing the radiation field parameters for isocenter determination using 3D polymer gel dosimeters: VIP (<b>A</b>,<b>B</b>), MAGIC (<b>C</b>,<b>D</b>), NIPAM (<b>E</b>,<b>F</b>), PAGAT (<b>G</b>,<b>H</b>). Both iCBCT transversal images (TrueBeam, Varian, Palo Alto, CA, USA; fixed MLC gap of 2 mm) for two irradiated regions of 500, 1000, and 1500 (first region) and 2000, 5000, and 10,000 MU (second region) (<b>A</b>,<b>C</b>,<b>E</b>,<b>G</b>) and profiles (<b>B</b>,<b>D</b>,<b>F</b>,<b>H</b>) across these regions are shown. The position where the profiles were taken is indicated on the iCBCT images by blue (first region) and orange (second region) dotted lines. The profiles were smoothed using a mean filter: Kernel mode: 3D, kernel unit: mm, and kernel size 3. Data were processed using the polyGeVero-CT software package (GeVero Co., Lodz, Poland).</p>
Full article ">Figure 4
<p>Determination of the radiation isocenter for the TrueBeam accelerator (Varian, Palo, Alto, CA, USA) using 3D polymer gel dosimeters: VIP (<b>A</b>,<b>B</b>), MAGIC (<b>C</b>,<b>D</b>), NIPAM (<b>E</b>,<b>F</b>), and PAGAT (<b>G</b>,<b>H</b>) and the polyGeVero-CT software package (GeVero Co., Lodz, Poland). Dosimeters in PH6-DD2 containers (~0.6 L, GeVero Co., Lodz, Poland) were irradiated in two regions with the photon beams crossed in a star-shot pattern to investigate the determination of the isocenter for the lower (bottom) and higher (top) monitor units per beam of 5000 and 10,000 (<b>A</b>,<b>B</b>), 5000 and 10,000 (<b>C</b>,<b>D</b>), 5000 and 10,000 (<b>E</b>,<b>F</b>), and 5000 and 10,000 MU (<b>G</b>,<b>H</b>). In (<b>A</b>,<b>C</b>,<b>E</b>,<b>G</b>) are photographs of the dosimeters after irradiation and in (<b>B</b>,<b>D</b>,<b>F</b>,<b>H</b>) are the results of data processing with the parameters of the radiation isocenter. Scale bars correspond to HU values.</p>
Full article ">Figure 4 Cont.
<p>Determination of the radiation isocenter for the TrueBeam accelerator (Varian, Palo, Alto, CA, USA) using 3D polymer gel dosimeters: VIP (<b>A</b>,<b>B</b>), MAGIC (<b>C</b>,<b>D</b>), NIPAM (<b>E</b>,<b>F</b>), and PAGAT (<b>G</b>,<b>H</b>) and the polyGeVero-CT software package (GeVero Co., Lodz, Poland). Dosimeters in PH6-DD2 containers (~0.6 L, GeVero Co., Lodz, Poland) were irradiated in two regions with the photon beams crossed in a star-shot pattern to investigate the determination of the isocenter for the lower (bottom) and higher (top) monitor units per beam of 5000 and 10,000 (<b>A</b>,<b>B</b>), 5000 and 10,000 (<b>C</b>,<b>D</b>), 5000 and 10,000 (<b>E</b>,<b>F</b>), and 5000 and 10,000 MU (<b>G</b>,<b>H</b>). In (<b>A</b>,<b>C</b>,<b>E</b>,<b>G</b>) are photographs of the dosimeters after irradiation and in (<b>B</b>,<b>D</b>,<b>F</b>,<b>H</b>) are the results of data processing with the parameters of the radiation isocenter. Scale bars correspond to HU values.</p>
Full article ">Figure 5
<p>Comparison of profiles for VIP, MAGIC, NIPAM, PAGAT, and PABIG<sup>nx</sup> from another study for comparison [<a href="#B18-materials-17-01283" class="html-bibr">18</a>]. Profiles drawn for each dosimeter in the position indicated in <a href="#materials-17-01283-f003" class="html-fig">Figure 3</a>H with yellow dashed line for the higher MU (10,000 MU) per beam used to irradiate the dosimeters (profiles filtered: mean filter, Kernel = 1, mode: 2D, unit: mm).</p>
Full article ">
14 pages, 3868 KiB  
Article
pNIPAm-Based pH and Thermoresponsive Copolymer Hydrogel for Hydrophobic and Hydrophilic Drug Delivery
by Anandhu Mohan, Madhappan Santhamoorthy, Thi Tuong Vy Phan and Seong-Cheol Kim
Gels 2024, 10(3), 184; https://doi.org/10.3390/gels10030184 - 7 Mar 2024
Cited by 8 | Viewed by 2576
Abstract
The regulated and targeted administration of hydrophobic and hydrophilic drugs is both promising and challenging in the field of drug delivery. Developing a hydrogel which is responsive to dual stimuli is considered a promising and exciting research area of study. In this work, [...] Read more.
The regulated and targeted administration of hydrophobic and hydrophilic drugs is both promising and challenging in the field of drug delivery. Developing a hydrogel which is responsive to dual stimuli is considered a promising and exciting research area of study. In this work, melamine functionalized poly-N-isopropyl acrylamide-co-glycidyl methacrylate copolymer has been developed by copolymerizing glycidyl methacrylate (GMA) monomer with N-isopropyl acrylamide (NIPAm) and further functionalized with melamine units (pNIPAm-co-pGMA-Mela). The prepared pNIPAm-co-pGMA-Mela copolymer hydrogel was characterized using various characterization techniques, including 1H NMR, FTIR, SEM, zeta potential, and particle size analysis. A hydrophobic drug (ibuprofen, Ibu) and hydrophilic drug (5-fluorouracil, 5-Fu) were selected as model drugs. Dual pH and temperature stimuli-responsive drug release behavior of the pNIPAm-co-pGMA-Mela hydrogel was evaluated under different pH (pH 7.4 and 4.0) and temperature (25 °C, 37 °C, and 45 °C) conditions. Furthermore, the in vitro biocompatibility of the developed pNIPAm-co-pGMA-Mela copolymer hydrogel was determined on MDA-MB-231 cells. The pH and temperature-responsive drug delivery study results reveal that the pNIPAm-co-pGMA-Mela hydrogel system is responsive to both pH and temperature stimuli and exhibits about ~100% of Ibu and 5-Fu, respectively, released at pH 4.0/45 °C. Moreover, the MTT assay and hemocompatibility analysis results proved that the pNIPAm-co-pGMA-Mela hydrogel system is biocompatible and hemocompatible, suggesting that that it could be used for drug delivery applications. The experimental results suggest that the proposed pNIPAm-co-pGMA-Mela hydrogel system is responsive to dual pH and temperature stimuli, and could be a promising drug carrier system for both hydrophilic and hydrophobic drug delivery applications. Full article
(This article belongs to the Special Issue Biopolymer Gels as Smart Drug Delivery and Theranostic Systems)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>(<b>A</b>) ¹H NMR spectra of pNIPAAm-co-pGMA copolymer and pNIPAAm-co-pGMA-Mela hydrogel; (<b>B</b>) FTIR analysis of (i) pNIPAAm-co-pGMA copolymer and (ii) pNIPAAm-co-pGMA-Mela hydrogel samples; (<b>C</b>) SEM images of (<b>i</b>) pNIPAAm-co-pGMA copolymer and (<b>ii</b>) pNIPAAm-co-pGMA-Mela hydrogel system.</p>
Full article ">Figure 2
<p>Zeta potential analysis of (<b>A</b>) pNIPAm-co-pGMA copolymer at (<b>i</b>) 25 °C and (<b>ii</b>) 45 °C and (<b>B</b>) pNIPAm-co-pGMA-Mela hydrogel at (<b>i</b>) 25 °C and (<b>ii</b>) 45 °C.</p>
Full article ">Figure 3
<p>Particle size analysis of (<b>A</b>) pNIPAm-co-pGMA copolymer and (<b>B</b>) pNIPAm-co-pGMA-Mela hydrogel at different temperature conditions; (<b>C</b>) particle size analysis of (i) pNIPAm-co-pGMA copolymer and (ii) pNIPAm-co-pGMA-Mela hydrogel at different pH conditions.</p>
Full article ">Figure 4
<p>Relative turbidity of (<b>A</b>) pNIPAm-co-pGMA copolymer and (<b>B</b>) pNIPAm-co-pGMA-Mela hydrogel samples at temperatures ranging from 25 °C to 55 °C; (<b>C</b>) schematic representation of the phase transition of the pNIPAm-co-pGMA-Mela hydrogel.</p>
Full article ">Figure 5
<p>Temperature-responsive swelling-deswelling behavior of (<b>A</b>) pNIPAm-co-pGMA copolymer and (<b>B</b>) pNIPAm-co-pGMA-Mela hydrogel at temperatures ranging from25 °C to 45 °C; (<b>C</b>) schematic representation of the sol–gel phase transition of pNIPAm-co-pGMA-Mela hydrogel.</p>
Full article ">Figure 6
<p>In vitro drug delivery efficiency of pNIPAAm-co-pGMA-Mela/Ibu hydrogel system: (<b>A</b>) Ibu release under various pH conditions, (<b>B</b>) Ibu release under various temperatures, and (<b>C</b>) Ibu release with different combinations of pH and temperature stimuli conditions.</p>
Full article ">Figure 7
<p>In vitro drug delivery efficiency of pNIPAAm-co-pGMA-Mela/5-Fu hydrogel system: (<b>A</b>) 5-Fu release at various pH conditions, (<b>B</b>) 5-Fu release at various temperatures, and (<b>C</b>) 5-Fu release with different combined pH and temperature conditions.</p>
Full article ">Figure 8
<p>(<b>A</b>) In vitro cytocompatibility of (i) pNIPAAm-co-pGMA-Mela hydrogel, (ii) 5-Fu drug-loaded pNIPAAm-co-pGMA-Mela/5-Fu hydrogel, and (iii) pure 5-Fu drug tested on MDA-MB-231 cells at different concentrations; (<b>B</b>) in vitro cytocompatibility of (i) pNIPAAm-co-pGMA-Mela hydrogel, (ii) Ibu drug-loaded pNIPAAm-co-pGMA-Mela/Ibu hydrogel, and (iii) pure Ibu drug tested on MDA-MB-231 cells at different concentrations; and (<b>C</b>) blood compatibility behavior of pNIPAAm-co-pGMA-Mela hydrogel at different sample concentrations. The inset image shows the in vitro blood compatibility of pNIPAAm-co-pGMA-Mela hydrogel by exposure of red blood cells to PBS, pNIPAAm-co-pGMA-Mela hydrogel sample, and Triton-X. Statistical significance refers to the cell toxicity with different samples (** significant <span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 9
<p>In vitro fluorescence cell images of (<b>i</b>) control without cells and (<b>ii</b>) MDA-MB-231 cells after treatment with pNIPAAm-co-pGMA-Mela hydrogel.</p>
Full article ">Scheme 1
<p>Illustration of the synthesis of pNIPAAm-co-pGMA copolymer (Step 1), melamine-functionalized pNIPAm-co-pGMA-Mela hydrogel (Step 2), and drug loading into the pNIPAAm-co-pGMA-Mela hydrogel system (Step 3).</p>
Full article ">
13 pages, 4423 KiB  
Article
Development of 3D Printed pNIPAM-Chitosan Scaffolds for Dentoalveolar Tissue Engineering
by Mehdi Salar Amoli, Resmi Anand, Mostafa EzEldeen, Liesbet Geris, Reinhilde Jacobs and Veerle Bloemen
Gels 2024, 10(2), 140; https://doi.org/10.3390/gels10020140 - 12 Feb 2024
Viewed by 1951
Abstract
While available treatments have addressed a variety of complications in the dentoalveolar region, associated challenges have resulted in exploration of tissue engineering techniques. Often, scaffold biomaterials with specific properties are required for such strategies to be successful, development of which is an active [...] Read more.
While available treatments have addressed a variety of complications in the dentoalveolar region, associated challenges have resulted in exploration of tissue engineering techniques. Often, scaffold biomaterials with specific properties are required for such strategies to be successful, development of which is an active area of research. This study focuses on the development of a copolymer of poly (N-isopropylacrylamide) (pNIPAM) and chitosan, used for 3D printing of scaffolds for dentoalveolar regeneration. The synthesized material was characterized by Fourier transform infrared spectroscopy, and the possibility of printing was evaluated through various printability tests. The rate of degradation and swelling was analyzed through gravimetry, and surface morphology was characterized by scanning electron microscopy. Viability of dental pulp stem cells seeded on the scaffolds was evaluated by live/dead analysis and DNA quantification. The results demonstrated successful copolymerization, and three formulations among various synthesized formulations were successfully 3D printed. Up to 35% degradability was confirmed within 7 days, and a maximum swelling of approximately 1200% was achieved. Furthermore, initial assessment of cell viability demonstrated biocompatibility of the developed scaffolds. While further studies are required to achieve the tissue engineering goals, the present results tend to indicate that the proposed hydrogel might be a valid candidate for scaffold fabrication serving dentoalveolar tissue engineering through 3D printing. Full article
(This article belongs to the Section Gel Analysis and Characterization)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>FTIR characterization of pNIPAM, chitosan and the copolymer with different ratios of constituents.</p>
Full article ">Figure 2
<p>(<b>A</b>) evaluation of print fidelity by measurement of perimeter ratio (<b>B</b>) Evaluation of print fidelity by calculating the pr value (<b>C</b>) evaluation of print resolution by measuring the area between the lines and (<b>D</b>) printability in 3D. Scale bar 10 mm. *: <span class="html-italic">p</span> ≤ 0.05 **: <span class="html-italic">p</span> ≤ 0.01 ***: <span class="html-italic">p</span> ≤ 0.001 ****: <span class="html-italic">p</span> ≤ 0.0001.</p>
Full article ">Figure 3
<p>(<b>A</b>) Rate of scaffold hydrolytic degradation in water at 37 °C (<b>B</b>) Rate of scaffold swelling at 37 °C.</p>
Full article ">Figure 4
<p>Surface morphology of the scaffolds obtained through SEM.</p>
Full article ">Figure 5
<p>(<b>A</b>) Live/dead viability assay of DPSCs seeded on the scaffolds and (<b>B</b>) DNA quantification of DPSCs. *: <span class="html-italic">p</span> ≤ 0.05 **: <span class="html-italic">p</span> ≤ 0.01 ***: <span class="html-italic">p</span> ≤ 0.001 ****: <span class="html-italic">p</span> ≤ 0.0001.</p>
Full article ">Scheme 1
<p>Schematic representation of the proposed reaction for synthesis of pNIPAM-chitosan copolymer.</p>
Full article ">
27 pages, 4694 KiB  
Article
Thermosensitive Polymeric Nanoparticles for Drug Co-Encapsulation and Breast Cancer Treatment
by Vanessa Franco Carvalho Dartora, Julia S. Passos, Leticia V. Costa-Lotufo, Luciana B. Lopes and Alyssa Panitch
Pharmaceutics 2024, 16(2), 231; https://doi.org/10.3390/pharmaceutics16020231 - 5 Feb 2024
Cited by 1 | Viewed by 2006
Abstract
Despite advances in breast cancer treatment, there remains a need for local management of noninvasive, low-grade ductal carcinoma in situ (DCIS). These focal lesions are well suited for local intraductal treatment. Intraductal administration supported target site drug retention, improved efficacy, and reduced systemic [...] Read more.
Despite advances in breast cancer treatment, there remains a need for local management of noninvasive, low-grade ductal carcinoma in situ (DCIS). These focal lesions are well suited for local intraductal treatment. Intraductal administration supported target site drug retention, improved efficacy, and reduced systemic exposure. Here, we used a poly(N-isopropyl acrylamide, pNIPAM) nanoparticle delivery system loaded with cytotoxic piplartine and an MAPKAP Kinase 2 inhibitor (YARA) for this purpose. For tumor environment targeting, a collagen-binding peptide SILY (RRANAALKAGELYKSILYGSG-hydrazide) was attached to pNIPAM nanoparticles, and the nanoparticle diameter, zeta potential, drug loading, and release were assessed. The system was evaluated for cytotoxicity in a 2D cell culture and 3D spheroids. In vivo efficacy was evaluated using a chemical carcinogenesis model in female Sprague–Dawley rats. Nanoparticle delivery significantly reduced the IC50 of piplartine (4.9 times) compared to the drug in solution. The combination of piplartine and YARA in nanoparticles further reduced the piplartine IC50 (~15 times). Treatment with these nanoparticles decreased the in vivo tumor incidence (5.2 times). Notably, the concentration of piplartine in mammary glands treated with nanoparticles (35.3 ± 22.4 μg/mL) was substantially higher than in plasma (0.7 ± 0.05 μg/mL), demonstrating targeted drug retention. These results indicate that our nanocarrier system effectively reduced tumor development with low systemic exposure. Full article
(This article belongs to the Special Issue Nanoparticles for Local Drug Delivery)
Show Figures

Figure 1

Figure 1
<p>Characterization, stability, and collagen binding of pNIPAM nanoparticles modified with SILY: (<b>A</b>) TEM micrograph of unloaded and drug-loaded pNIPAM nanoparticles modified with SILY, scale bar = 200 nm. The nanoparticles were lyophilized and resuspended in Milli-Q water for 4 h at 25 °C prior to TEM. Scale bar = 200 nm. (<b>B</b>) Dynamic light scattering for the hydrodynamic diameter temperature sweep from 17 °C to 42 °C of unloaded NPs and NPs and NP-SILY loaded with piplartine and YARA. (<b>C</b>) Degradation of fluorescently labeled nanoparticles over 7 days of dialysis against 1× PBS pH 7.4 and 3.5 at 544 nm of absorbance. * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.001, and *** <span class="html-italic">p</span> &lt; 0.0001 against pH 7.0. Data are the mean ± standard deviation of 6 replicates. (<b>D</b>) Collagen-binding assay demonstrating the ability of NP-SILY to bind to a collagen-I-coated surface. Particle binding increased with an increase in conjugated SILY, whereas blank nanoparticles did not show the ability to bind to the collagen plate.</p>
Full article ">Figure 2
<p>Release profiles of YARA (<b>A</b>) and piplartine (<b>B</b>) incorporated in NP-SILY in PBS at pH 3.5 and pH 7.4 over 120 h at 37 °C. Bars represent the average ± SEM (<span class="html-italic">n</span> = 12).</p>
Full article ">Figure 3
<p>MCF-7 (<b>A</b>) and T47-D (<b>B</b>) cell staining for Smad3 (pSpS423/425). Cells were treated with 10 ng/mL TGF-beta for 1 h before treatment with the nanoparticles containing YARA (NP-YARA). Nontreated cells were used as a control. As a secondary antibody, Alexa Fluor<sup>®</sup> 488 (Red) was used. Nuclei were stained with DAPI (Blue). The confocal images were captured at 60× magnification. Scale bar: 50 µm. (<b>C</b>) Quantification of the mean fluorescence intensity was by Fiji ImageJ2 software, and data are shown as the average ± standard deviation of 9 replicates in 3 independent experiments. Phosphorylation of HSP27 Serine 15 treated with YARA solution or NP-YARA for 48 h and compared with nontreated cells (positive control) in (<b>D</b>) MCF-7 and (<b>E</b>) T47-D cells. Error bars represent the standard deviation. * <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, and **** <span class="html-italic">p</span> &lt; 0.0001, ns = non-significant compared to the control.</p>
Full article ">Figure 4
<p>Viability of breast cancer cells after exposure to the unloaded nanoparticles and nanoparticles loaded with YARA, piplartine, or piplartine + YARA for 72 h: (<b>A</b>) comparison of the treatment of T47-D cells with unloaded (Unloaded NPs) or YARA-loaded nanoparticles (NP-SILY-YARA); (<b>B</b>) comparison of the treatment of MCF-7 cells with unloaded (Unloaded NPs) or YARA-loaded nanoparticles (NP-SILY-YARA); (<b>C</b>) comparison of the treatment of T47-D cells with piplartine solution and nanoparticles loaded with piplartine (NP-SILY-PIP) or piplartine+ YARA (NP-SILY-PIP_YARA); (<b>D</b>) comparison of the treatment of MCF-7 cells with piplartine solution and nanoparticles loaded with piplartine or piplartine + YARA. Data are shown as the average ± standard deviation of 10–15 replicates in 4–5 independent experiments. * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p &lt;</span> 0.01, *** <span class="html-italic">p &lt;</span> 0.001 and **** <span class="html-italic">p &lt;</span> 0.0001 compared to unloaded nanoparticles.</p>
Full article ">Figure 5
<p>Formation and viability of spheroids after treatment with the nanoparticles. Sizes of MCF-7 (<b>A</b>) and T47-D (<b>B</b>) spheroids after 5 days in culture. Bar = 400 µm. Viability of the MCF-7 (<b>C</b>) and T47-D (<b>D</b>) spheroids after exposure to the drug solution in DMSO or to the nanoparticles with encapsulated piplartine or piplartine + YARA for 72 h. Data are shown as the average ± standard deviation of 12 spheroids in 4 independent experiments. * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p &lt;</span> 0.01 compared to the piplartine solution (DMSO) and + <span class="html-italic">p</span> &lt; 0.05, ++ <span class="html-italic">p</span> &lt; 0.01 compared to NP-SILY-PIP.</p>
Full article ">Figure 6
<p>Uptake of rhodamine-labeled NP-SILY (50 μg/mL) and rhodamine solution (in DMSO) in cells as monolayers and spheroids after 48 h of incubation: (<b>A</b>) MCF-7 cells as monolayers; (<b>B</b>) T47-D cells as monolayers. The left images show nuclei highlighted in blue with Hoechst<sup>®</sup> 33342 dye, the center images display Rhodamine-B, and the right images present a composite of the two previous images. (<b>C</b>) Rhodamine penetration in T47-D spheroids following treatment with rhodamine solution. (<b>D</b>) Rhodamine penetration in T47-D spheroids following treatment with rhodamine-labeled NP-SILY at a 50 μg/mL nanoparticle concentration. Scale bar = 200 µm.</p>
Full article ">Figure 7
<p>In vivo mammary tissue retention of the fluorescent marker rhodamine administered in nanoparticles or as a solution: (<b>A</b>) whole animal images showing fluorescence staining after intraductal administration of NP-Rhodamine or rhodamine solution (control); (<b>B</b>) mammary tissue fluorescence intensity decay as a function of time, <span class="html-italic">n</span> = 3 animals/group. ** <span class="html-italic">p</span> &lt; 0.01 compared to NP-Rhodamine. (<b>C</b>) Histological sections of the mammary tissue of animals administered with blank NPs or NP-Rhodamine and compared with the untreated group. The pictures illustrate the structural integrity of ducts and the absence of inflammatory cell infiltration and edema. Scale bar = 100 μm.</p>
Full article ">Figure 8
<p>Effects of piplartine- and YARA-loaded NPs (NP-SILY-PIP-YARA) on the histological characteristics of breast tissue after MNU tumor induction compared to positive and negative control groups. A volume of 20 μL of the NP-SILY-PIP-YARA (1 mg/mL in 1× PBS) was administered per gland, and tissues were stained with H&amp;E. Scale bar = 100 μm.</p>
Full article ">
15 pages, 4963 KiB  
Article
Phase Transition Behaviors of Poly(N-isopropylacrylamide) Nanogels with Different Compositions Induced by (−)-Epigallocatechin-3-gallate and Ethyl Gallate
by Ke Deng, Yafei Wang, Lei Wang, Xianli Fan, Zhenyu Wu, Xue Wen, Wen Xie, Hong Wang, Zheng Zhou, Pengfei Chen and Xianggui Chen
Molecules 2023, 28(23), 7823; https://doi.org/10.3390/molecules28237823 - 28 Nov 2023
Cited by 2 | Viewed by 1249
Abstract
Phase transition behaviors of poly(N-isopropylacrylamide) nanogels with different compositions induced by (−)-epigallocatechin-3-gallate (EGCG) and ethyl gallate (EG) has been investigated systematically. Monodisperse poly(N-isopropylacrylamide-co-N-hydroxymethyl acrylamide) (P(NIPAM-co-NMAM)) and poly(N-isopropylacrylamide-co-2-hydroxyethyl methacrylate) (P(NIPAM- [...] Read more.
Phase transition behaviors of poly(N-isopropylacrylamide) nanogels with different compositions induced by (−)-epigallocatechin-3-gallate (EGCG) and ethyl gallate (EG) has been investigated systematically. Monodisperse poly(N-isopropylacrylamide-co-N-hydroxymethyl acrylamide) (P(NIPAM-co-NMAM)) and poly(N-isopropylacrylamide-co-2-hydroxyethyl methacrylate) (P(NIPAM-co-HEMA)) nanogels with different feeding monomer ratios were prepared by emulsion polymerization. P(NIPAM-co-NMAM) nanogels exhibit rapid isothermal phase transition behavior in EGCG solutions with low concentration (10−3 mol/L) in less than 10 minutes. The thermosensitive phase transition behaviors of nanogels are affected not only by the copolymerized monomers but also by the concentrations of EGCG and EG in aqueous solutions. Nanogels remain in a shrunken state and do not exhibit thermosensitive phase transition behaviors in EGCG solutions (≥5 mmol/L), whereas they display thermo-responsive phase transition behaviors in EG solutions. The volume phase transition temperature (VPTT) shifts to lower temperatures with increasing EG concentration. The diameters of P(NIPAM-co-NMAM) nanogels decrease with increasing EG concentration at temperatures between 29 and 33 °C. In contrast, the diameters of P(NIPAM-co-HEMA) nanogels increase with increasing EGCG concentration at temperatures between 37 and 45 °C. The results demonstrate the potential of nanogels for simple detection of EG and EGCG concentrations in aqueous solutions over a wide temperature range, and EGCG can serve as a signal for the burst-release of drugs from the P(NIPAM-co-NMAM)-based carriers at physiological temperature. Full article
(This article belongs to the Special Issue Synthesis and Application of Nanoparticles and Nanocomposites)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Synthesis of P(NPAM-<span class="html-italic">co</span>-HEMA) (<b>a</b>) and P(NIPAM-<span class="html-italic">co</span>-NMAM) (<b>b</b>) nanogels.</p>
Full article ">Figure 2
<p>SEM images of P(NPAM-<span class="html-italic">co</span>-HEMA) (<b>a</b>) and P(NIPAM-<span class="html-italic">co</span>-NMAM) (<b>b</b>) nanogels with different monomer molar ratios in the dry state.</p>
Full article ">Figure 3
<p>Good monodespersity of nanogels. (<b>a</b>–<b>c</b>) Diameter distribution of P(NIPAM-<span class="html-italic">co</span>-NMAM) with different monomer ratios of 10:1 (<b>a</b>), 9:1 (<b>b</b>), and 8:1 (<b>c</b>). (<b>d</b>–<b>g</b>) Diameter distribution of P(NIPAM-<span class="html-italic">co</span>-HEMA) with different monomer ratios of 99:1 (<b>d</b>), 95:5 (<b>e</b>), 9:1 (<b>f</b>), and 8:2 (<b>g</b>).</p>
Full article ">Figure 4
<p>FT-IR spectra of PNIPAM microgels, P(NPAM-<span class="html-italic">co</span>-HEMA) nanogels, and P(NIPAM-<span class="html-italic">co</span>-NMAM) nanogels.</p>
Full article ">Figure 5
<p>(<b>a</b>–<b>d</b>) Thermo-responsive hydrodynamic diameters of P(NIPAM-<span class="html-italic">co</span>-HEMA) nanogels with different compositions in water. (<b>e</b>) Absolute value of zeta potential of P(NIPAM-<span class="html-italic">co</span>-HEMA) nanogels with monomer ratio of 8:2 prepared with and without SDS. (<b>f</b>) Thermo-responsive factor of nanogels with different feeding monomers.</p>
Full article ">Figure 6
<p>(<b>a</b>–<b>c</b>) Thermo-responsive hydrodynamic diameters of P(NIPAM-<span class="html-italic">co</span>-NMAM) nanogels with different compositions in water. (<b>d</b>) Thermo-responsive factor of nanogels with different feeding monomer ratios. (<b>e</b>) Absolute value of zeta potential of P(NIPAM-<span class="html-italic">co</span>-NMAM) nanogels with monomer ratio of 8:1 prepared with and without SDS.</p>
Full article ">Figure 7
<p>DSC response under heating of powdered P(NIPAM-<span class="html-italic">co</span>-HEMA) nanogels in DI water. Successive curves are shifted vertically.</p>
Full article ">Figure 8
<p>Dynamic volume change (V/V<sub>0</sub>) of P(NIPAM-<span class="html-italic">co</span>-HEMA) (<b>a</b>) and P(NIPAM-<span class="html-italic">co</span>-NMAM) (<b>b</b>) nanogels in EGCG solutions with different concentrations (<b>a</b>,<b>b</b>) and EG solution (<b>c</b>) at 25 °C.</p>
Full article ">Figure 9
<p>Thermo-responsive hydrodynamic diameters of P(NIPAM-<span class="html-italic">co</span>-HEMA) (<b>a</b>–<b>c</b>) and P(NIPAM-<span class="html-italic">co</span>-NMAM) (<b>d</b>–<b>f</b>) nanogels in EG (<b>c</b>,<b>e</b>,<b>f</b>) and EGCG (<b>a</b>,<b>b</b>,<b>d</b>) aqueous solutions with 0, 5, 10, and 15 mmol/L concentration. HEMA95-5 and NMAM9-1 represent the NIPAM:HEMA = 95:5 and NIPAM:NMAM = 9:1, respectively. EGCG1, EGCG2, and EGCG3 indicate that the concentrations of EGCG are 5, 10, and 15 mmol/L. EG1, EG2, and EG3 indicate that the concentrations of EG are 5, 10, and 15 mmol/L.</p>
Full article ">Figure 10
<p>Effect of EG concentration in aqueous solution on the VPTT of P(NIPAM-<span class="html-italic">co</span>-NMAM) and P(NIPAM-<span class="html-italic">co</span>-HEMA) nanogels. The feeding monomer ratios of NIPAM:HEMA and NIPAM:NMAM are 95:5 and 9:1, respectively.</p>
Full article ">
18 pages, 6018 KiB  
Article
Preparation and Characterization of Thermoresponsive Polymer Scaffolds Based on Poly(N-isopropylacrylamide-co-N-tert-butylacrylamide) for Cell Culture
by Gilyana K. Kazakova, Victoria S. Presniakova, Yuri M. Efremov, Svetlana L. Kotova, Anastasia A. Frolova, Sergei V. Kostjuk, Yury A. Rochev and Peter S. Timashev
Technologies 2023, 11(5), 145; https://doi.org/10.3390/technologies11050145 - 18 Oct 2023
Cited by 2 | Viewed by 2325
Abstract
In the realm of scaffold-free cell therapies, there is a questto develop organotypic three-dimensional (3D) tissue surrogates in vitro, capitalizing on the inherent ability of cells to create tissues with an efficiency and sophistication that still remains unmatched by human-made devices. In this [...] Read more.
In the realm of scaffold-free cell therapies, there is a questto develop organotypic three-dimensional (3D) tissue surrogates in vitro, capitalizing on the inherent ability of cells to create tissues with an efficiency and sophistication that still remains unmatched by human-made devices. In this study, we explored the properties of scaffolds obtained by the electrospinning of a thermosensitive copolymer, poly(N-isopropylacrylamide-co-N-tert-butylacrylamide) (P(NIPAM-co-NtBA)), intended for use in such therapies. Two copolymers with molecular weights of 123 and 137 kDa and a content of N-tert-butylacrylamide of ca. 15 mol% were utilized to generate 3D scaffolds via electrospinning. We examined the morphology, solution viscosity, porosity, and thickness of the spun matrices as well as the mechanical properties and hydrophobic–hydrophilic characteristics of the scaffolds. Particular attention was paid to studying the influence of the thermosensitive polymer’s molecular weight and dispersity on the resultant scaffolds’ properties and the role of electroforming parameters on the morphology and mechanical characteristics of the scaffolds. The cytotoxicity of the copolymers and interaction of cells with the scaffolds were also studied. Our findings provide significant insight into approaches to optimizing scaffolds for specific cell cultures, thereby offering new opportunities for scaffold-free cell therapies. Full article
Show Figures

Figure 1

Figure 1
<p>Dependencies of the viscosity of P(NIPAM-<span class="html-italic">co</span>-NtBA) copolymers solutions on the shear rate.</p>
Full article ">Figure 2
<p>Electrospun P(NIPAM-co-NtBA) scaffolds from 15% solutions of Copolymer 1 and copolymer 2: (<b>A</b>) SEM images of scaffold surfaces capturing the nuances of the utilized copolymers; (<b>B</b>) histograms detailing the fiber diameter distributions; (<b>C</b>) SEM images emphasizing scaffold thickness variations (scaffold attached to the foil); (<b>D</b>) contact angle measurements showcasing the wettability properties. All scaffolds were electrospun with 1 mL of the solution using a dynamic collector rotating at 600 RPM and FAA.</p>
Full article ">Figure 3
<p>Electrospun P(NIPAM-<span class="html-italic">co</span>-NtBA) scaffolds from the Copolymer 2 solutions with different concentrations and collector speeds: (<b>A</b>) SEM images of scaffolds from the solution with 460 ± 50 mPa·s viscosity, showcasing the effects of the collector speeds (right to left: 2400, 1800, 600 RPM); (<b>B</b>) histograms detailing fiber diameter distributions corresponding to (<b>A</b>); (<b>C</b>) SEM images for scaffolds from the solution with 1460 ± 350 mPa·s viscosity, emphasizing the outcomes of varying collector speeds; (<b>D</b>) histograms elaborating the fiber diameter variations related to (<b>C</b>). All scaffolds were electrospun with 1 mL of a solution, a 15 cm nozzle-to-collector distance, and without FAA.</p>
Full article ">Figure 4
<p>Comparative analysis of electrospun P(NIPAM-<span class="html-italic">co</span>-NtBA) scaffolds from 15 wt% and 20 wt% solutions of Copolymer 2: (<b>A</b>) Contact angle measurements across the collector speeds of 600, 1800, and 2400 RPM; (<b>B</b>) SEM micrographs illustrating the thickness’s variation at the respective speeds. Electrospinning was conducted with 1 mL of the solution loading, 15 cm nozzle-to-collector distance, and without FAA.</p>
Full article ">Figure 5
<p>Comparative analysis of electrospun P(NIPAM-<span class="html-italic">co</span>-NtBA) scaffolds with and without FAA: (<b>A</b>) SEM images of the surface morphology; (<b>B</b>) cross-sectional SEM views of scaffolds, following their detachment from the foil, illustrating their inherent thickness; (<b>C</b>) variation in the fiber thickness distribution; (<b>D</b>) contact angle measurements. These scaffolds were produced using the dynamic collector rotating at 600 RPM from 1 mL of the 15% Copolymer 2 solution at a nozzle-to-collector distance of 15 cm.</p>
Full article ">Figure 6
<p>Electroformed P(NIPAM-<span class="html-italic">co</span>-NtBA) scaffolds at different nozzle-to-collector distances (15 and 23 cm): (<b>A</b>) Surface SEM images; (<b>B</b>) cross-sectional SEM views depicting the scaffold thickness on foil; (<b>C</b>) distribution of the fiber thickness; (<b>D</b>) measurements of contact angle. All scaffolds were fabricated using a dynamic collector rotating at 600 RPM and 1 mL of the Copolymer 1 solution with FAA.</p>
Full article ">Figure 7
<p>Comparative analysis of electrospun P(NIPAM-<span class="html-italic">co</span>-NtBA) scaffold characteristics prepared from different volumes (1 mL and 3 mL) of a Copolymer 1 solution. The tests were conducted at a maximum nozzle-to-collector distance of 23 cm with FAA enabled using a dynamic collector rotating at 600 RPM: (<b>A</b>) SEM images of the surface morphology; (<b>B</b>) SEM images of the scaffold thickness on foil; (<b>C</b>) distribution of the scaffold fiber thickness; (<b>D</b>) contact angle values for scaffolds prepared from different solution volumes; (<b>E</b>) mechanical data.</p>
Full article ">Figure 8
<p>Cytotoxicity of the pNIPAm-<span class="html-italic">co</span>-NtBA copolymer (<b>A</b>,<b>B</b>) and the cell behavior on the electrospun scaffolds (<b>C</b>). (<b>A</b>) Analysis of cell metabolic activity; Alamar Blue test. (<b>B</b>) Analysis of the proliferative activity of cells in the presence of the polymer; PicoGreen assay. (<b>C</b>) The REF52 cells’ morphology on the electrospun scaffolds, with low and high density, confocal microscopy of cells stained with Calcein AM and electrospun scaffolds fibers stained with Rhodamin C. (<b>D</b>) Analysis of the cell viability on the electrospun scaffolds using the Live/Dead assay. Calcein AM and propidium iodide staining of ARPE-19 cells grown on scaffolds for 3 weeks.</p>
Full article ">
14 pages, 4919 KiB  
Article
Thermo-Responsive Hydrogels Encapsulating Targeted Core–Shell Nanoparticles as Injectable Drug Delivery Systems
by Elif Gulin Ertugral-Samgar, Ali Murad Ozmen and Ozgul Gok
Pharmaceutics 2023, 15(9), 2358; https://doi.org/10.3390/pharmaceutics15092358 - 21 Sep 2023
Cited by 6 | Viewed by 2078
Abstract
As therapeutic agents that allow for minimally invasive administration, injectable biomaterials stand out as effective tools with tunable properties. Furthermore, hydrogels with responsive features present potential platforms for delivering therapeutics to desired sites in the body. Herein, temperature-responsive hydrogel scaffolds with embedded targeted [...] Read more.
As therapeutic agents that allow for minimally invasive administration, injectable biomaterials stand out as effective tools with tunable properties. Furthermore, hydrogels with responsive features present potential platforms for delivering therapeutics to desired sites in the body. Herein, temperature-responsive hydrogel scaffolds with embedded targeted nanoparticles were utilized to achieve controlled drug delivery via local drug administration. Poly(N-isopropylacrylamide) (pNIPAM) hydrogels, prepared with an ethylene-glycol-based cross-linker, demonstrated thermo-sensitive gelation ability upon injection into environments at body temperature. This hydrogel network was engineered to provide a slow and controlled drug release profile by being incorporated with curcumin-loaded nanoparticles bearing high encapsulation efficiency. A core (alginate)–shell (chitosan) nanoparticle design was preferred to ensure the stability of the drug molecules encapsulated in the core and to provide slower drug release. Nanoparticle-embedded hydrogels were shown to release curcumin at least four times slower compared to the free nanoparticle itself and to possess high water uptake capacity and more mechanically stable viscoelastic behavior. Moreover, this therapy has the potential to specifically address tumor tissues over-expressing folate receptors like ovaries, as the nanoparticles target the receptors by folic acid conjugation to the periphery. Together with its temperature-driven injectability, it can be concluded that this hydrogel scaffold with drug-loaded and embedded folate-targeting nanoparticles would provide effective therapy for tumor tissues accessible via minimally invasive routes and be beneficial for post-operative drug administration after tumor resection. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>FT-IR spectra of FA molecule (<b>A</b>), CS(AA NP) (<b>B</b>), and FA-CS(AA-Cur NP)(<b>C</b>).</p>
Full article ">Figure 2
<p>TEM images of prepared nanoparticles (<b>A</b>) and their stability behavior (<b>B</b>) (scale bar: 100 nm).</p>
Full article ">Figure 3
<p>Size distribution (<b>A</b>), stability (<b>B</b>), and morphological evaluation (<b>C</b>) of FA-CS(AA-Cur NP).</p>
Full article ">Figure 4
<p>Morphological and mechanical evaluations for blank pNIPAM-based hydrogel (<b>A</b>,<b>D</b>), FA-CS(AA-Cur NP)-incorporated hydrogel (<b>C</b>,<b>E</b>), and their comparative swelling profiles (<b>B</b>).</p>
Full article ">Figure 5
<p>Release profiles of curcumin from NP alone and NP-incorporated hydrogels at different pH conditions.</p>
Full article ">Figure 6
<p>Scratch before the injection (<b>A</b>) and after the injection (<b>B</b>) of NP-incorporated, thermo-responsive pNIPAM hydrogel and its degradation behavior (<b>C</b>,<b>D</b>) on chicken breast tissue incubated at 37 °C.</p>
Full article ">
16 pages, 1683 KiB  
Article
Preparation of Amino-Functionalized Poly(N-isopropylacrylamide)-Based Microgel Particles
by Anna Harsányi, Attila Kardos and Imre Varga
Gels 2023, 9(9), 692; https://doi.org/10.3390/gels9090692 - 28 Aug 2023
Cited by 1 | Viewed by 1517
Abstract
Responsive cationic microgels are a promising building block in several diagnostic and therapeutic applications, like transfection and RNA or enzyme packaging. Although the direct synthesis of cationic poly(N-isopropylacrylamide) (PNIPAm) microgel particles has a long history, these procedures typically resulted in low [...] Read more.
Responsive cationic microgels are a promising building block in several diagnostic and therapeutic applications, like transfection and RNA or enzyme packaging. Although the direct synthesis of cationic poly(N-isopropylacrylamide) (PNIPAm) microgel particles has a long history, these procedures typically resulted in low yield, low incorporation of the cationic comonomer, increased polydispersity, and pure size control. In this study, we investigated the possibility of the post-polymerization modification of P(NIPAm-co-acrylic acid) microgels to prepare primary amine functionalized microgels. To achieve this goal, we used 1-ethyl-3-(3-(dimethylamino)propyl)carbodiimide hydrochloride (EDC) mediated coupling of a diamine to the carboxyl groups. We found that by controlling the EDC excess in the reaction mixture, the amine functionalization of the carboxyl functionalized microgel could be varied and as much as 6–7 mol% amine content could be incorporated into the microgels. Importantly, the reaction was conducted at room temperature in an aqueous medium and it was found to be time efficient, making it a practical and convenient approach for synthesizing primary amine functionalized PNIPAm microgel particles. Full article
(This article belongs to the Special Issue Recent Advances in Microgels)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>(<b>a</b>) The relative monomer concentration in the function of the polymerization time for the P(NIPAm-co-10%AAc) synthesis. (<b>b</b>) The hydrodynamic diameter of the purified P(NIPAm-co-10%AAc) microgels in 10 mM HCl (pH = 2) and in 10 mM phosphate buffer (pH = 7) as a function of temperature. Each data point is the average of at least ten measurements and the standard errors of the datapoints are commensurate to the sizes of the symbols in the figure. The PDI of the samples were below 0.1.</p>
Full article ">Figure 2
<p>The relative concentration of EDC compared to its initial concentration as a function of reaction time in pure 50 mM MES buffer (black square). All other datasets were measured in EDC/P(NIPAm-co-10%AAc) solutions with a constant 5.0 mM acrylic acid content. Blue circles: stoichiometric EDC/carboxyl group ratio; Green up triangles: 3-fold, red diamonds: 6-fold, and magenta down triangles: 12-fold EDC excess. All samples contained 50 mM MES buffer and had a pH of 5.5.</p>
Full article ">Figure 3
<p>The electrophoretic mobility of the functionalized microgel particles as a function of solution pH. Different colors indicate microgels functionalized in different excesses of EDC as defined by the legends in the figure.</p>
Full article ">Figure 4
<p>The hydrodynamic diameter of the P(NIPAM-co-10%AAc) (blue circle) and two of the amine functionalized microgels as a function of solution pH. Red diamonds indicate the microgel functionalized in 6-fold EDC excess, while light blue squares represent the microgel functionalized in 24-fold EDC excess. Each data point is the average of at least ten measurements and the standard errors of the datapoints are commensurate to the sizes of the symbols in the figure. The PDI of the samples were below 0.1.</p>
Full article ">Figure 5
<p>The hydrodynamic diameter of the amine functionalized microgels measured at pH = 2 (positively charged microgels) and at pH = 11 (negatively charged microgels) as a function EDC excess used in the coupling reaction. Each data point is the average of at least ten measurements and the standard errors of the datapoints are commensurate to the sizes of the symbols in the figure.</p>
Full article ">Figure 6
<p>The hydrodynamic diameter of the amine functionalized microgel (made in 24-fold EDC excess) in the function of temperature at four different solution pH values. Each data point is the average of at least ten measurements and the standard errors of the datapoints are commensurate to the sizes of the symbols in the figure.</p>
Full article ">Scheme 1
<p>The main reactions taking place and the main products in an aqueous EDC solution in the presence of carboxyl functionalized microgels. <b>1</b>. The acid and base catalyzed hydrolysis of EDC. <b>2</b>. The EDC activation of the microgl carboxyl groups and three main reactions of the O-acylisourea intermediate ester. EDC activated reactive carbocations can undergo several other potential side reactions which are not depicted here. For additional information see Ref. [<a href="#B51-gels-09-00692" class="html-bibr">51</a>].</p>
Full article ">
13 pages, 3756 KiB  
Article
Dual pH- and Thermo-Sensitive Poly(N-isopropylacrylamide-co-allylamine) Nanogels for Curcumin Delivery: Swelling–Deswelling Behavior and Phase Transition Mechanism
by Madhappan Santhamoorthy and Seong-Cheol Kim
Gels 2023, 9(7), 536; https://doi.org/10.3390/gels9070536 - 1 Jul 2023
Cited by 5 | Viewed by 2293
Abstract
Curcumin (Cur) is a beneficial ingredient with numerous bioactivities. However, due to its low solubility and poor bioavailability, its therapeutic application is limited. In this work, we prepared poly-N-isopropylacrylamide p(NIPAm) and polyallylamine p(Am)-based nanogel (p(NIPAm-co-Am)) NG for a dual pH- and temperature-sensitive copolymer [...] Read more.
Curcumin (Cur) is a beneficial ingredient with numerous bioactivities. However, due to its low solubility and poor bioavailability, its therapeutic application is limited. In this work, we prepared poly-N-isopropylacrylamide p(NIPAm) and polyallylamine p(Am)-based nanogel (p(NIPAm-co-Am)) NG for a dual pH- and temperature-sensitive copolymer system for drug delivery application. In this copolymer system, the p(NIPAm) segment was incorporated to introduce thermoresponsive behavior and the p(Am) segment was incorporated to introduce drug binding sites (amine groups) in the resulting (p(NIPAm-co-Am)) NG system. Various instrumental characterizations including 1H nuclear magnetic resonance (1H NMR) spectroscopy, Fourier transform infrared (FT-IR) analysis, scanning electron microscopy (SEM), zeta potential, and particle size analysis were performed to confirm the copolymer synthesis. Curcumin (Cur), an anticancer bioactive substance, was employed to assess the in vitro drug loading and release performance of the resulting copolymer nanogels system at varied pH levels (pH 7.2, 6.5, and 4.0) and temperatures (25 °C, 37 °C, and 42 °C). The cytocompatibility of the p(NIPAm-co-Am) NG sample was also tested on MDA-MB-231 cells at various sample concentrations. All the study results indicate that the p(NIPAm-co-Am) NG produced might be effective for drug loading and release under pH and temperature dual-stimuli conditions. As a result, the p(NIPAm-co-Am) NG system has the potential to be beneficial in the use of drug delivery applications in cancer therapy. Full article
(This article belongs to the Special Issue Recent Advances in Gels Engineering for Drug Delivery)
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) ¹H NMR and (<b>b</b>) FT-IR spectra of p(NIPAm-co-Am) NG sample.</p>
Full article ">Figure 2
<p>(<b>a</b>,<b>b</b>) SEM images represent the surface morphology of the p(NIPAm-co-Am) NG sample.</p>
Full article ">Figure 3
<p>XPS spectra of (<b>a</b>) full scan spectrum; (<b>b</b>) C1s; (<b>c</b>) N1s; and (<b>d</b>) O1s elements, respectively.</p>
Full article ">Figure 4
<p>High-resolution XPS spectra of (<b>a</b>) C1s; (<b>b</b>) N1s; and (<b>c</b>) O1s elements, respectively.</p>
Full article ">Figure 5
<p>(<b>a</b>) Zeta potential of p(NIPAm-co-Am) NG system. (<b>b</b>) Schematic illustration of the phase transition of p(NIPAm-co-Am) NG under the pH and temperature stimuli conditions.</p>
Full article ">Figure 6
<p><b>Figure 6.</b> (<b>a</b>) DLS analysis of the p(NIPAm-co-Am) NG sample. (<b>b</b>) Relative turbidity of the p(NIPAm-co-Am) NG sample. (<b>c</b>) UV–vis absorption of the p(NIPAm-co-Am) NG sample at different solution temperatures.</p>
Full article ">Figure 7
<p>A proposed schematic illustration of the p(NIPAm-co-Am) NG system conformational change in each phase state.</p>
Full article ">Figure 8
<p>In vitro Cur delivery of p(NIPAm-co-Am) NG. (<b>a</b>) Cur release under pH stimuli, (<b>b</b>) Cur release under temperature stimuli, and (<b>c</b>) Cur release under pH and temperature stimuli conditions, respectively.</p>
Full article ">Figure 9
<p>In vitro biocompatibility of (a) p(NIPAm-co-Am) NG; (b) Cur loaded p(NIPAm co-Am) NG/Cur; and (c) pure Cur, respectively, at 37 °C. Statistical significance to the cell toxicity with different samples (*, significant <span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Scheme 1
<p>Schematic representation of drug loading and stimuli-responsive release behavior of the p(NIPAm-co-Am) NG system.</p>
Full article ">Scheme 2
<p>Illustration of the (a) synthesis of p(NIPAm-co-Am) copolymer NG system and (b) stimuli-responsive phase transition of p(NIPAm-co-Am) NG sample.</p>
Full article ">
20 pages, 12953 KiB  
Article
Modified Sulfanilamide Release from Intelligent Poly(N-isopropylacrylamide) Hydrogels
by Ana Dinić, Vesna Nikolić, Ljubiša Nikolić, Snežana Ilić-Stojanović, Stevo Najman, Maja Urošević and Ivana Gajić
Pharmaceutics 2023, 15(6), 1749; https://doi.org/10.3390/pharmaceutics15061749 - 16 Jun 2023
Viewed by 1427
Abstract
The aim of this study was to examine homopolymeric poly(N-isopropylacrylamide), p(NIPAM), hydrogels cross-linked with ethylene glycol dimethacrylate as carriers for sulfanilamide. Using FTIR, XRD and SEM methods, structural characterization of synthesized hydrogels before and after sulfanilamide incorporation was performed. The residual [...] Read more.
The aim of this study was to examine homopolymeric poly(N-isopropylacrylamide), p(NIPAM), hydrogels cross-linked with ethylene glycol dimethacrylate as carriers for sulfanilamide. Using FTIR, XRD and SEM methods, structural characterization of synthesized hydrogels before and after sulfanilamide incorporation was performed. The residual reactants content was analyzed using the HPLC method. The swelling behavior of p(NIPAM) hydrogels of different crosslinking degrees was monitored in relation to the temperature and pH values of the surrounding medium. The effect of temperature, pH, and crosslinker content on the sulfanilamide release from hydrogels was also examined. The results of the FTIR, XRD, and SEM analysis showed that sulfanilamide is incorporated into the p(NIPAM) hydrogels. The swelling of p(NIPAM) hydrogels depended on the temperature and crosslinker content while pH had no significant effect. The sulfanilamide loading efficiency increased with increasing hydrogel crosslinking degree, ranging from 87.36% to 95.29%. The sulfanilamide release from hydrogels was consistent with the swelling results—the increase of crosslinker content reduced the amount of released sulfanilamide. After 24 h, 73.3–93.5% of incorporated sulfanilamide was released from the hydrogels. Considering the thermosensitivity of hydrogels, volume phase transition temperature close to the physiological temperature, and the satisfactory results achieved for sulfanilamide incorporation and release, it can be concluded that p(NIPAM) based hydrogels are promising carriers for sulfanilamide. Full article
(This article belongs to the Section Nanomedicine and Nanotechnology)
Show Figures

Figure 1

Figure 1
<p>The chemical structure of sulfanilamide.</p>
Full article ">Figure 2
<p>The HPLC chromatograms and UV spectra of: (<b>a</b>) monomer NIPAM and (<b>b</b>) crosslinker EGDM.</p>
Full article ">Figure 3
<p>The swelling profiles of p(NIPAM) hydrogels at 20 ± 1 °C in the solution with the pH value of (<b>a</b>) 2.2 and (<b>b</b>) 7.4.</p>
Full article ">Figure 4
<p>The swelling profiles of p(NIPAM) hydrogels at 37 ± 1 °C in the solution with the pH value of (<b>a</b>) 2.2 and (<b>b</b>) 7.4.</p>
Full article ">Figure 5
<p>Dependence of the swelling degree of p(NIPAM) hydrogels, α, on the temperature.</p>
Full article ">Figure 6
<p><sup>1</sup>H-NMR spectrum of sulfanilamide.</p>
Full article ">Figure 7
<p>FTIR spectra of: (<b>a</b>) sulfanilamide, (<b>b</b>) p(NIPAM) hydrogel 2.5 mol%, and (<b>c</b>) p(NIPAM) hydrogel 2.5 mol% with incorporated sulfanilamide.</p>
Full article ">Figure 8
<p>Diffractograms of (<b>a</b>) sulfanilamide, (<b>b</b>) p(NIPAM) with 1.5 and 2.5 mol% of crosslinker, and p(NIPAM) with 1.5 mol% crosslinker and incorporated sulfanilamide.</p>
Full article ">Figure 9
<p>SEM micrographs of p(NIPAM) hydrogels: (<b>a</b>) 1.5 mol% (1000×), (<b>b</b>) 2.5 mol% (1000×), (<b>c</b>) 1.5 mol% with incorporated sulfanilamide (500×), and (<b>d</b>) 2.5 mol% with incorporated sulfanilamide (500×).</p>
Full article ">Figure 10
<p>HPLC chromatogram of sulfanilamide.</p>
Full article ">Figure 11
<p>Released sulfanilamide content at 37 ± 1 °C from p(NIPAM) hydrogels at: (<b>a</b>) pH = 2.2 and (<b>b</b>) pH = 7.4.</p>
Full article ">Figure 12
<p>Released sulfanilamide content at 20 ± 1 °C from p(NIPAM) hydrogels at: (<b>a</b>) pH = 2.2 and (<b>b</b>) pH = 7.4.</p>
Full article ">
18 pages, 4673 KiB  
Article
Injectable Hyaluronan-Based Thermoresponsive Hydrogels for Dermatological Applications
by Si Gou, Alexandre Porcello, Eric Allémann, Denis Salomon, Patrick Micheels, Olivier Jordan and Yogeshvar N. Kalia
Pharmaceutics 2023, 15(6), 1708; https://doi.org/10.3390/pharmaceutics15061708 - 11 Jun 2023
Cited by 10 | Viewed by 2253
Abstract
Most marketed HA-based dermal fillers use chemical cross-linking to improve mechanical properties and extend their lifetime in vivo; however, stiffer products with higher elasticity require an increased extrusion force for injection in clinical practice. To balance longevity and injectability, we propose a thermosensitive [...] Read more.
Most marketed HA-based dermal fillers use chemical cross-linking to improve mechanical properties and extend their lifetime in vivo; however, stiffer products with higher elasticity require an increased extrusion force for injection in clinical practice. To balance longevity and injectability, we propose a thermosensitive dermal filler, injectable as a low viscosity fluid that undergoes gelation in situ upon injection. To this end, HA was conjugated via a linker to poly(N-isopropylacrylamide) (pNIPAM), a thermosensitive polymer using “green chemistry”, with water as the solvent. HA-L-pNIPAM hydrogels showed a comparatively low viscosity (G′ was 105.1 and 233 for Candidate1 and Belotero Volume®, respectively) at room temperature and spontaneously formed a stiffer gel with submicron structure at body temperature. Hydrogel formulations exhibited superior resistance against enzymatic and oxidative degradation and could be administered using a comparatively lower injection force (49 N and >100 N for Candidate 1 and Belotero Volume®, respectively) with a 32G needle. Formulations were biocompatible (viability of L929 mouse fibroblasts was >100% and ~85% for HA-L-pNIPAM hydrogel aqueous extract and their degradation product, respectively), and offered an extended residence time (up to 72 h) at the injection site. This property could potentially be exploited to develop sustained release drug delivery systems for the management of dermatologic and systemic disorders. Full article
(This article belongs to the Special Issue Hyaluronic Acid for Medical Applications)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>(<b>A</b>) G′ storage modulus and (<b>B</b>) G″ loss modulus of HA-L-pNIPAM hydrogel candidates as a function of temperature (n = 3, Mean ± SD).</p>
Full article ">Figure 2
<p>Force in Newtons (N) required to extrude Belotero Volume<sup>®</sup>, Belotero Balance<sup>®</sup> and the three candidates as a function of the stroke distance of the piston in a syringe with a 30G, 32G and 34G needle (distance %) at 22 °C.</p>
Full article ">Figure 3
<p>Belotero Balance<sup>®</sup> and the three candidates storage modulus (G′), loss modulus (G″), and tangent delta (δ) normalized to initial values as function of time at 0.7 Hz and 37 °C after addition of 100 µL of hyaluronidase (100 U/mL) (<b>A</b>–<b>C</b>) (n = 3; Mean ± SD). After an addition of 100 µL H<sub>2</sub>O<sub>2</sub> (30% <span class="html-italic">w</span>/<span class="html-italic">w</span>) (<b>D</b>–<b>F</b>) (n = 3; ±sd). As a control condition, 100 µL of PBS buffer was added to 400 µL of the sample.</p>
Full article ">Figure 4
<p>Effects of raw HA and HA-L-pNIPAM polymers on L929 cells at 24 h of exposure. Data express the percentage of cell viability with respect to culture media control (n = 3, Mean ± SD).</p>
Full article ">Figure 5
<p>Comparison of subcutaneous injection of Candidates 1–3 in an ex vivo porcine skin model at T<sub>0</sub> (<b>A</b>–<b>C</b>) and T<sub>72</sub> (<b>D</b>–<b>F</b>). Light blue indicated the presence of the injected HA gel. Scale bar = 1 mm for (<b>A</b>,<b>E</b>) and 2 mm for (<b>B</b>–<b>D</b>,<b>F</b>).</p>
Full article ">Figure 6
<p>Distribution pattern and tissue integration of Candidate 1 at T<sub>0</sub> (<b>A</b>–<b>C</b>) and T<sub>72</sub> (<b>D</b>,<b>E</b>) in an ex vivo porcine skin model. Scale bar = 2 mm for (<b>A</b>), and 100 µm for (<b>B</b>–<b>E</b>).</p>
Full article ">
Back to TopTop