[go: up one dir, main page]

 
 
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

Journals

Article Types

Countries / Regions

Search Results (22)

Search Parameters:
Keywords = ferrogels

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
20 pages, 7987 KiB  
Article
Remote Positioning of Spherical Alginate Ferrogels in a Fluid Flow by a Magnetic Field: Experimental and Computer Simulation
by Felix Blyakhman, Alexander Safronov, Ilya Starodumov, Darya Kuznetsova and Galina Kurlyandskaya
Gels 2023, 9(9), 711; https://doi.org/10.3390/gels9090711 - 1 Sep 2023
Cited by 2 | Viewed by 1194
Abstract
This work belongs to the development of mechanical force-responsive drug delivery systems based on remote stimulation by an external magnetic field at the first stage, assisting the positioning of a ferrogel-based targeted delivery platform in a fluid flow. Magnetically active biopolymer beads were [...] Read more.
This work belongs to the development of mechanical force-responsive drug delivery systems based on remote stimulation by an external magnetic field at the first stage, assisting the positioning of a ferrogel-based targeted delivery platform in a fluid flow. Magnetically active biopolymer beads were considered a prototype implant for the needs of replacement therapy and regenerative medicine. Spherical calcium alginate ferrogels (FGs)~2.4 mm in diameter, filled with a 12.6% weight fraction of magnetite particles of 200–300 nm in diameter, were synthesized. A detailed characterization of the physicochemical and magnetic properties of FGs was carried out, as were direct measurements of the field dependence of the attractive force for FG-beads. The hydrodynamic effects of the positioning of FG-beads in a fluid flow by a magnetic field were studied experimentally in a model vessel with a fluid stream. Experimental results were compared with the results of mathematical and computer modeling, showing reasonable agreement. The contributions of the hydrodynamic and magnetic forces acting on the FG-bead in a fluid flow were discussed. Obtained forces for a single ferrogel implant were as high as 0 to 10−4 N for the external field range of 0 to 35 kA/m, perfectly in the range of mechanical force stimuli in biological systems. Full article
(This article belongs to the Special Issue Innovative Biopolymer-Based Hydrogels)
Show Figures

Figure 1

Figure 1
<p>Top view of the ferrogel beads under the study (<b>a</b>). General arrangement of identical superparamagnetic spheres self-organized as one-layer bead arrays placed onto flat surfaces: 1—line for indication of nearest neighbors corresponding to central red bead; 2—the second level of the neighbors corresponding to central red bead. Red dashed line indicates the bead cluster corresponding to experimental one (<b>b</b>). Description of the positioning of ferrogel spherical sample for observation of the surface properties by optical microscopy (<b>c</b>). Optical microscopy image of the surface of FG bead (<b>d</b>).</p>
Full article ">Figure 2
<p>XRD pattern for magnetite particles used in the preparation of ferrogel beads. Numbers correspond to Miller indexes.</p>
Full article ">Figure 3
<p>TEM images of magnetite particles. The inset gives numerical particle size distribution (PSD) calculated by the analysis of 632 images.</p>
Full article ">Figure 4
<p>Magnetic hysteresis loops for magnetite particles. M—magnetization; H—magnetic field strength. Insert reflects part of the hysteresis loops close to zero fields, representing the low external magnetic field range at which the primary magnetization curve becomes better visible.</p>
Full article ">Figure 5
<p>Thermal analysis plots for CaAlg dried hydrogel (<b>a</b>) and ferrogel (<b>b</b>). The green line (TG) shows the mass loss of the sample. The blue line (DSC) corresponds to the thermal effect of decomposition. The red (H<sub>2</sub>O) and black (CO<sub>2</sub>) lines correspond to the mass spectral signal proportional to the water and carbon dioxide content released in decomposition.</p>
Full article ">Figure 6
<p>(<b>a</b>) Dependence of the attractive force (<span class="html-italic">F</span>) in ferrogels on the magnetic field strength (<span class="html-italic">H</span>) for 36 magnetic beads. The confidence interval for <span class="html-italic">F</span> is given at <span class="html-italic">p</span> = 0.05 (n = 9): y = 0.00002x<sup>2</sup> + 0.0044x for the fit of the experimental curve. 1—calculations in accordance with Equation (6) (case of non-interacting beads); 2—calculations in accordance with Equation (6) but corrected for dipole-dipole interactions taking into account the first group of nearest neighbors; 3—correction of case 2 for possible change of primary magnetization curve taking into account interactions with the first group of nearest neighbors; 4—calculations in accordance with Equation (6) but corrected for dipole-dipole interactions taking into account the second group of nearest neighbors (see also <a href="#gels-09-00711-f001" class="html-fig">Figure 1</a>b). (<b>b</b>) Magnetic hysteresis loop of gel and ferrogel beads; inset shows the hysteresis loop of the blank gel bead in the scale, making the slope of the magnetization curve visible.</p>
Full article ">Figure 7
<p>Geometry of magnetic fields for a central line of the beads in beads array (central line, <a href="#gels-09-00711-f001" class="html-fig">Figure 1</a>b); 1—the first type neighbors, 2—the second type neighbors. Red arrows indicate magnetic fields created by the central bead for the positions of the other beads.</p>
Full article ">Figure 8
<p>Side view of a fragment of the experimental setup for positioning the ferrogel bead in a magnetic field. The diameter of the bead is about 2.4 mm. See explanation in the text.</p>
Full article ">Figure 9
<p>Effect of the magnetic field strength on the spatial position of the ferrogel bead in the tube at a fluid flow rate of 400 mL per sec. Tracks 1, 2, and 3 correspond to the magnetic field strengths at the surface of the EM core of 40, 32, and 24 kA/m, respectively. The origin of the coordinates corresponds to the beginning of the observation. The thick horizontal line at the X-axis represents the position of the electromagnet core. See explanation in the text.</p>
Full article ">Figure 10
<p>Effect of the magnetic field strength on the linear velocity of the ferrogel bead in the tube at fluid flow of 400 mL per sec. Tracks 1, 2, and 3 correspond to the magnetic field strength at the surface of the EM core of 40, 32, and 24 kA/m, respectively. The origin of the coordinates corresponds to the beginning of the observation. The thick horizontal line at the X-axis represents the position of the electromagnet core. See explanation in the text.</p>
Full article ">Figure 11
<p>Color plot of pressure distribution at the surface and around the ferrogel sample in contact with the wall (<b>a1</b>,<b>b1</b>) and in the presence of a gap between the sample and the tube wall (<b>a2</b>,<b>b2</b>). White lines represent the streamlines of the liquid.</p>
Full article ">Figure 12
<p>Computer simulation of the spatial position of the ferrogel bead in the tube at fluid flow using the model (the model description is presented in <a href="#sec4dot5-gels-09-00711" class="html-sec">Section 4.5</a>; see also <a href="#app1-gels-09-00711" class="html-app">Supplementary Videos S3 and S4</a>). Tracks 1 and 2 correspond to the magnetic field strengths at the surface of the EM core of 40 and 32 kA/m, respectively. The origin of the coordinates corresponds to the beginning of the observation. The thick horizontal line at the X-axes represents the position of the electromagnet core.</p>
Full article ">Figure 13
<p>Computer simulation of the linear velocity of the ferrogel bead in the tube at fluid flow using the model (the model description is presented in <a href="#sec4dot5-gels-09-00711" class="html-sec">Section 4.5</a>; see also <a href="#app1-gels-09-00711" class="html-app">Supplementary Videos S3 and S4</a>). Tracks 1 and 2 correspond to the magnetic field strengths at the surface of the EM core of 40 and 32 kA/m, respectively. The origin of the coordinates corresponds to the beginning of the observation. The thick horizontal line at the X-axis represents the position of the electromagnet core.</p>
Full article ">Figure 14
<p>Schematic sketch of the experimental setup for the measurement of the attractive force of ferrogel beads on a magnet. 1—cuvette filled with water and beads; 2—EM; 3—DC power supply; 4—FT. See explanation in the text.</p>
Full article ">Figure 15
<p>Schematic of experimental hydrodynamic setup for positioning ferrogel beads in a fluid flow. 1—main reservoir mounted on a tripod with adjustable height, 2—straight section of the tube fixed on a horizontal platform, 3—electromagnet connected to a power supply (not present), 4—buffer reservoir, 5—peristaltic pump, 6—video camera, 7—light source. See explanation in the text.</p>
Full article ">Figure 16
<p>Scheme of the computational domain and the used boundary conditions. Inflow boundary “1” defines a fluid flow with a fixed velocity, flowing in a cylindrical domain with rigid walls “3”. The flow interacts with the ball and exits through outflow boundary “2” defined by zero pressure gradient.</p>
Full article ">Figure 17
<p>Scheme of forces acting on the gel sample and its initial position in the computational domain.</p>
Full article ">
20 pages, 3659 KiB  
Article
Using SMART Magnetic Fluids and Gels for Prevention and Destruction of Bacterial Biofilms
by Jarosƚaw E. Król and Garth D. Ehrlich
Microorganisms 2023, 11(6), 1515; https://doi.org/10.3390/microorganisms11061515 - 7 Jun 2023
Viewed by 1705
Abstract
Biofouling is a major problem in all natural and artificial settings where solid surfaces meet liquids in the presence of living microorganisms. Microbes attach to the surface and form a multidimensional slime that protects them from unfavorable environments. These structures, known as biofilms, [...] Read more.
Biofouling is a major problem in all natural and artificial settings where solid surfaces meet liquids in the presence of living microorganisms. Microbes attach to the surface and form a multidimensional slime that protects them from unfavorable environments. These structures, known as biofilms, are detrimental and very hard to remove. Here, we used SMART magnetic fluids [ferrofluids (FFs), magnetorheological fluids (MRFs), and ferrogels (FGs) containing iron oxide nano/microparticles] and magnetic fields to remove bacterial biofilms from culture tubes, glass slides, multiwell plates, flow cells, and catheters. We compared the ability of different SMART fluids to remove biofilms and found that commercially available, as well as homemade, FFs, MRFs, and FGs can successfully remove biofilm more efficiently than traditional mechanical methods, especially from textured surfaces. In tested conditions, SMARTFs reduced bacterial biofilms by five orders of magnitude. The ability to remove biofilm increased with the amount of magnetic particles; therefore, MRFs, FG, and homemade FFs with high amounts of iron oxide were the most efficient. We showed also that SMART fluid deposition can protect a surface from bacterial attachment and biofilm formation. Possible applications of these technologies are discussed. Full article
(This article belongs to the Special Issue Bacterial Biofilm Formation and Eradication)
Show Figures

Figure 1

Figure 1
<p>Effect of SMART ferrofluids on bacterial viability. SMARTF abbreviations are in <a href="#microorganisms-11-01515-t001" class="html-table">Table 1</a>. SMART fluids at various concentrations (0.1%, 1%, and 10%) were added to the exponentially-grown <span class="html-italic">E. coli</span> C cultures (OD600 = 0.6) and the cell density (CFU/mL) was measured after 2 h. Star symbols mark the SMART fluids under which the cell reductions showed a dose-dependent manner. Compared to the control (i.e., no exposure to SMART fluids), the star-labeled SMART fluids all showed significant cell reduction (<span class="html-italic">p</span> &lt; 0.05, <span class="html-italic">t</span>-test) at a concentration of 10%.</p>
Full article ">Figure 2
<p>Efficiency of biofilm removal by SMART fluids in test tube experiment. The remaining cell number (CFU/mL) was measured and compared to the control (i.e., no exposure to the SMARTFs). The resulting percentage of biofilm on each condition was shown. SMARTF abbreviations and details are in <a href="#microorganisms-11-01515-t001" class="html-table">Table 1</a>. The “negative” remaining biofilm under the MRFs was because the MRFs removed the biofilm so thoroughly that the surface was much cleaner (i.e., contained less biofilm) than in the “no treatment” case.</p>
Full article ">Figure 3
<p>Effect of iron oxide concentration on biofilm removal. Iron oxide FFs at various concentrations (increasing from K1 to K7) were added to a biofilm of <span class="html-italic">E. coli</span> C grown on a culture tube for 8 days. The remaining cell number (CFU/mL) was measured and compared to the control (i.e., no exposure). The resulting percentage of biofilm on each condition was shown. See <a href="#microorganisms-11-01515-t001" class="html-table">Table 1</a> for more details about the SMART fluids.</p>
Full article ">Figure 4
<p>Biofilm removal from microscope slides. (<b>A</b>) Stages of biofilm removal treatment. From the left: slide biofilm, CV stained for better visualization, FM sliding, cleaned slide. (<b>B</b>) Efficiency of biofilm removal with SMART fluids presented as the number of remaining cells in CFU/mL. C—control without SMARTF treatment. See <a href="#app1-microorganisms-11-01515" class="html-app">Figure S3</a> for experiment details.</p>
Full article ">Figure 5
<p>Microscopic analyses of <span class="html-italic">E. coli</span> slides. (<b>A</b>) Control, (<b>B</b>) EHF3, (<b>C</b>) K5, (<b>D</b>) MRF-122EG, (<b>E</b>) AMT-Dampro, (<b>F</b>) AMT-Magnaflo, (<b>G</b>) AMT-Smartec, (<b>H</b>) AMT-Rheotec, (<b>I</b>) FM. Red bar scales represent 20 µm.</p>
Full article ">Figure 6
<p>Microscope analyses of removal of submerged biofilm in a 24-well plate. (<b>A</b>) Control, (<b>B</b>) EHF3, (<b>C</b>) K5, (<b>D</b>) MRF-122EG, (<b>E</b>) AMT-Dampro, (<b>F</b>) AMT-Magnaflo, (<b>G</b>) AMT-Smartec, (<b>H</b>) AMT-Rheotec, (<b>I</b>) FM, (<b>J</b>) quantitative analysis of microscope pictures. Mag. 20×, Red bar scales represent 100 µm.</p>
Full article ">Figure 7
<p>Twelve-day-old <span class="html-italic">E. coli</span> C biofilm in a urinary catheter before (<b>A</b>,<b>C</b>) and after (<b>B</b>,<b>D</b>) treatment with FM.</p>
Full article ">Figure 8
<p>Removing biofilm from vertical flow cells using EHF3 (<b>A</b>) and FM (<b>B</b>). From the left: <span class="html-italic">E. coli</span> C biofilm on both sides of the flow cell; flow cell with one wall cleaned; removing biofilm from the second wall; flow cell with both sides cleaned.</p>
Full article ">Figure 9
<p>Efficiency of biofilm removal from glass slide (<b>A</b>) and ABS textured plastic (<b>B</b>) by cell scraping (SC), vortexing (VOX), FM, and EHF3 (FF).</p>
Full article ">Figure 10
<p>Biofilm formation inhibition by ferrofluid. Upper panel—schematic of slide construction, ferrofluid location, and experimental setup; center panel—biofilm growth on microscope slides (obverse and reverse) not stained (left) and stained with crystal violet; bottom panel—microscopic analyses of <span class="html-italic">E. coli</span> slides. (<b>A</b>) Control, (<b>B</b>) EHF3, (<b>C</b>) K5, (<b>D</b>) MRF-122EG, (<b>E</b>) AMT-Dampro, (<b>F</b>) AMT-Magnaflo, (<b>G</b>) AMT-Smartec, (<b>H</b>) AMT-Rheotec, (<b>I</b>) FM. Blue bar scales represent 20 µm.</p>
Full article ">
16 pages, 3079 KiB  
Article
Design of Spherical Gel-Based Magnetic Composites: Synthesis and Characterization
by Pavel A. Shabadrov, Alexander P. Safronov, Nadezhda M. Kurilova and Felix A. Blyakhman
J. Compos. Sci. 2023, 7(5), 177; https://doi.org/10.3390/jcs7050177 - 1 May 2023
Cited by 4 | Viewed by 1522
Abstract
The purpose of the study was the synthesis and the physicochemical characterization of spherical beads of magnetically active composite ferrogels (FGs) with diameters of 2–3 mm for further application to the needs of targeted drug delivery and/or replacement therapy. Spherical FGs based on [...] Read more.
The purpose of the study was the synthesis and the physicochemical characterization of spherical beads of magnetically active composite ferrogels (FGs) with diameters of 2–3 mm for further application to the needs of targeted drug delivery and/or replacement therapy. Spherical FGs based on a physical network of calcium alginate (CaAlg), a chemical network of polyacrylamide (PAAm), and a combined network of calcium alginate and polyacrylamide (PAAm/CaAlg) were analyzed. FGs were filled with γ-Fe2O3 magnetic nanoparticles (MNPs) obtained by using the electrical explosion of wire method. A comparative study of the swelling behavior and of the structural features of the polymeric network in CaAlg, PAAm/CaAlg, and PAAm spherical beads was performed. It was shown that the densest network was provided by a combination of chemical and physical networking in PAAm/CaAlg FGs. If the physical network were removed from FGs it resulted in a substantial increase in the average diameter and the swelling ratio of spherical beads and a decrease in the MNPs concentration in the swollen FGs by approximately two times. It was shown that irrespective of the gel composition, the embedding of maghemite nanoparticles led to an increase in the swelling ratio of the polymeric network. This indicated the absence of strong intermolecular interactions between the polymer and the filler. The results obtained might be useful for the design of magnetically active spherical FG beads of a given size and controlled physicochemical properties. Full article
(This article belongs to the Section Polymer Composites)
Show Figures

Figure 1

Figure 1
<p>TEM micrograph (JEOL JEM2100) of maghemite γ-Fe<sub>2</sub>O<sub>3</sub> EEW MNPs (<b>A</b>) used for the synthesis of ferrogels. Particle size distribution (<b>B</b>) was obtained by a graphical analysis of 2473 TEM images of MNPs. X-ray diffractogram (<b>C</b>) of MNPs with Miller indexes (Bruker D8 Discover).</p>
Full article ">Figure 2
<p>Photos of ferrogel (<b>a</b>,<b>c</b>) and hydrogel (<b>b</b>,<b>d</b>) beads. (<b>a</b>)—CaAlg/PAAm ferrogel; (<b>b</b>)—CaAlg/PAAm blank hydrogel; (<b>c</b>)—PAAm ferrogel; (<b>d</b>)—CaAlg blank hydrogel. The actual sizes of presented spherical beads are shown in the table right below.</p>
Full article ">Figure 3
<p>The effect of acrylamide on the surface tension of Na alginate solutions.</p>
Full article ">Figure 4
<p>The swelling ratio of gel beads with different polymer networks.</p>
Full article ">Figure 5
<p>The scheme of PAAm ferrogel bead formation from the combined PAAm/CaAlg interpenetrating network.</p>
Full article ">Figure 6
<p>TG/DSC curves for CaAlg dried hydrogels (<b>a</b>) and ferrogels (<b>b</b>). The green line (TG) shows the mass loss of the sample. The blue line (DSC) corresponds to the thermal effect of decomposition. The red (H<sub>2</sub>O) and black (CO<sub>2</sub>) lines correspond to the mass spectral signal proportional to the water and carbon dioxide content released in decomposition.</p>
Full article ">Figure 7
<p>The swelling ratio of the polymer network for hydrogel and ferrogel beads depending on the composition.</p>
Full article ">Figure 8
<p>The dependence of the swelling ratio (<b>a</b>) and the weight fraction of PAAm in the gel (<b>b</b>) on the molar concentration of the monomer. Curves created in SigmaPlot 11.0 software by nonlinear approximation regression (<b>a</b>) and using the quadratic function (<b>b</b>) according to Equation (7).</p>
Full article ">
16 pages, 5612 KiB  
Article
Gelation in Alginate-Based Magnetic Suspensions Favored by Poor Interaction among Sodium Alginate and Embedded Particles
by Alexander P. Safronov, Elena V. Rusinova, Tatiana V. Terziyan, Yulia S. Zemova, Nadezhda M. Kurilova, Igor. V. Beketov and Andrey Yu. Zubarev
Appl. Sci. 2023, 13(7), 4619; https://doi.org/10.3390/app13074619 - 6 Apr 2023
Cited by 3 | Viewed by 1649
Abstract
Alginate gels are extensively tested in biomedical applications for tissue regeneration and engineering. In this regard, the modification of alginate gels and solutions with dispersed magnetic particles gives extra options to control the rheo-elastic properties both for the fluidic and gel forms of [...] Read more.
Alginate gels are extensively tested in biomedical applications for tissue regeneration and engineering. In this regard, the modification of alginate gels and solutions with dispersed magnetic particles gives extra options to control the rheo-elastic properties both for the fluidic and gel forms of alginate. Rheological properties of magnetic suspensions based on Na-alginate water solution with embedded magnetic particles were studied with respect to the interfacial adhesion of alginate polymer to the surface of particles. Particles of magnetite (Fe3O4), metallic iron (Fe), metallic nickel (Ni), and metallic nickel with a deposited carbon layer (Ni@C) were taken into consideration. Storage modulus, loss modulus, and the shift angle between the stress and the strain were characterized by the dynamic mechanical analysis in the oscillatory mode. The intensity of molecular interactions between alginate and the surface of the particles was characterized by the enthalpy of adhesion which was determined from calorimetric measurements using a thermodynamic cycle. Strong interaction at the surface of the particles resulted in the dominance of the “fluidic” rheological properties: the prevalence of the loss modulus over the storage modulus and the high value of the shift angle. Meanwhile, poor interaction of alginate polymer with the surface of the embedded particles favored the “elastic” gel-like properties with the dominance of the storage modulus over the loss modulus and low values of the shift angle. Full article
(This article belongs to the Special Issue Novel Nanomaterials and Nanostructures)
Show Figures

Figure 1

Figure 1
<p>TEM microphotographs of magnetic particles: (<b>a</b>)—magnetite (Fe<sub>3</sub>O<sub>4</sub>); (<b>b</b>)—metallic iron (Fe); (<b>c</b>)—metallic nickel (Ni); (<b>d</b>)—nickel particles coated with carbon shell (Ni@C); (<b>e</b>)—HRTEM image of the carbon layer on the surface of Ni particle.</p>
Full article ">Figure 2
<p>Shear stress dependence of the dynamic mechanical parameters for Na-alginate water solutions. Frequency of shear oscillations 1Hz, with a temperature of 25 °C. Concentration 5% (wt.): 1—storage modulus, <span class="html-italic">G′</span>; 2—loss modulus, <span class="html-italic">G</span>″; 3—shift angle, <span class="html-italic">δ</span>. Concentration 10% (wt.): 4—storage modulus, <span class="html-italic">G′</span>; 5—loss modulus, <span class="html-italic">G</span>″; 6—shift angle, <span class="html-italic">δ</span>.</p>
Full article ">Figure 3
<p>Concentration dependences of the dynamic moduli and shift angle at 1 Hz frequency and shear stress 1 Pa.</p>
Full article ">Figure 4
<p>Shear stress dependence of the storage modulus (<span class="html-italic">G</span>′), the loss modulus (<span class="html-italic">G</span>″), and the shift angle (<span class="html-italic">δ</span>) for the magneto-rheological suspension of 2.8% Fe<sub>3</sub>O<sub>4</sub> MPs dispersed in 5% Na-alginate solution. Frequency of shear oscillations 1 Hz, temperature 25 °C.</p>
Full article ">Figure 5
<p>Concentration dependences of storage modulus <span class="html-italic">G</span>′ and loss modulus <span class="html-italic">G</span>″ for the magneto-rheological suspensions with different magnetic particles. (<b>a</b>)—Fe and Fe<sub>3</sub>O<sub>4</sub> MPs, (<b>b</b>)—Ni and Ni@C MPs. <span class="html-italic">G</span>′ and <span class="html-italic">G</span>″ are averaged over the plateau in 0.1–1 Pa range of shear stress. Frequency 1 Hz. Temperature 25 °C.</p>
Full article ">Figure 6
<p>Concentration dependences of shift angle <span class="html-italic">δ</span> for magneto-rheological suspensions with different magnetic particles. <span class="html-italic">δ</span> is averaged over the plateau at 0.1–1 Pa range of shear stress. Frequency 1 Hz. Temperature 25 °C.</p>
Full article ">Figure 7
<p>Enthalpy of dissolution of Na-alginate + MPs composites in water (Δ<span class="html-italic">H</span><sub>4</sub>) at 25 °C. Values at the ordinate axes correspond to the enthalpy of dissolution of Na-alginate (left axis—Δ<span class="html-italic">H</span><sub>1</sub>) and to the enthalpy of wetting of MNPs (right axis—Δ<span class="html-italic">H</span><sub>2</sub>). Experimental data for Ni and Ni@C are presented.</p>
Full article ">Figure 8
<p>Concentration dependences of the enthalpy of adhesion of Na-alginate polymer to the surface of MNPs. Temperature 25 °C. Lines are drawn for the eye-guide only.</p>
Full article ">Figure 9
<p>Schematic presentation of the two types of structuring in Na-alginate MRS with MPs depended on the interfacial adhesion among the particles and the polymer. (<b>a</b>)—strong adhesion between Na-alginate and MPs leads to the polymeric layers at the surface which favor fluid-type behavior. It is the case of Fe<sub>3</sub>O<sub>4</sub> and Ni@C MNPs. (<b>b</b>)—weak interaction at the interface leads to the networking of MPs due to the magnetic forces which promote elastic deformation and the gelation of MRS. It is the case of Fe and Ni MPs.</p>
Full article ">
12 pages, 2216 KiB  
Article
Starch-g-Acrylic Acid/Magnetic Nanochitin Self-Healing Ferrogels as Flexible Soft Strain Sensors
by Pejman Heidarian and Abbas Z. Kouzani
Sensors 2023, 23(3), 1138; https://doi.org/10.3390/s23031138 - 19 Jan 2023
Cited by 6 | Viewed by 1847
Abstract
Mechanically robust ferrogels with high self-healing ability might change the design of soft materials used in strain sensing. Herein, a robust, stretchable, magneto-responsive, notch insensitive, ionic conductive nanochitin ferrogel was fabricated with both autonomous self-healing and needed resilience for strain sensing application without [...] Read more.
Mechanically robust ferrogels with high self-healing ability might change the design of soft materials used in strain sensing. Herein, a robust, stretchable, magneto-responsive, notch insensitive, ionic conductive nanochitin ferrogel was fabricated with both autonomous self-healing and needed resilience for strain sensing application without the need for additional irreversible static chemical crosslinks. For this purpose, ferric (III) chloride hexahydrate and ferrous (II) chloride as the iron source were initially co-precipitated to create magnetic nanochitin and the co-precipitation was confirmed by FTIR and microscopic images. After that, the ferrogels were fabricated by graft copolymerisation of acrylic acid-g-starch with a monomer/starch weight ratio of 1.5. Ammonium persulfate and magnetic nanochitin were employed as the initiator and crosslinking/nano-reinforcing agents, respectively. The ensuing magnetic nanochitin ferrogel provided not only the ability to measure strain in real-time under external magnetic actuation but also the ability to heal itself without any external stimulus. The ferrogel may also be used as a stylus for a touch-screen device. Based on our findings, our research has promising implications for the rational design of multifunctional hydrogels, which might be used in applications such as flexible and soft strain sensors, health monitoring, and soft robotics. Full article
(This article belongs to the Special Issue Nanostructured Materials Systems for Optical Sensing)
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) Synthesis of magnetic nanochitin at a stoichiometric aqueous quantity of Fe(II)/Fe(III) salt solution and an alkaline solution in a non-oxidizing environment; (<b>b</b>) SEM photograph of nanochitin; (<b>c</b>) TEM photograph of magnetic nanochitin; (<b>d</b>) magnetic hysteresis loop of magnetic nanochitin at the temperature of 300 K; (<b>e</b>) FTIR spectra of (I) magnetic nanochitin, (II) nanochitin, and (III) acrylic acid-g-starch/magnetic nanochitin ferrogel; (<b>f</b>) the possibility for interaction between iron oxide nanoparticles and polymeric chains; (<b>g</b>) SEM photographs of ferrogel containing 0.5 wt.% magnetic nanochitin; (<b>h</b>) SEM photographs of ferrogel containing 1 wt.% magnetic nanochitin; (<b>i</b>) SEM photographs of ferrogel containing 1.5 wt.% magnetic nanochitin; and (<b>j</b>) SEM photographs of ferrogel containing 2 wt.% magnetic nanochitin.</p>
Full article ">Figure 2
<p>(<b>a</b>) Ultimate tensile strength for ferrogels at different magnetic nanochitin concentrations; (<b>b</b>) Young’s modulus for ferrogels at different magnetic nanochitin concentrations; (<b>c</b>) elongation at break of ferrogels at different magnetic nanochitin concentrations; (<b>d</b>) toughness of ferrogels at different magnetic nanochitin concentrations; (<b>e</b>) tensile strength healing efficiency of ferrogels at different magnetic nanochitin concentrations; and (<b>f</b>) notch sensitivity analysis of ferrogels at varying concentrations.</p>
Full article ">Figure 3
<p>(<b>a</b>) Resistance response to different strains for original and healed ferrogels; (<b>b</b>) electrical recovery of ferrogel after (i) disconnection, (ii) connection, (iii) disconnection, (iv) connection, (v) disconnection, and (vi) connection from the cutline; (<b>c</b>) the real-time resistance change detection of a ferrogel connected to an artificial wooden elbow at various angles; (<b>d</b>) gauge factor curve of the ferrogel strain sensor containing 1.5 wt.% magnetic nanochitin; (<b>e</b>) using the ferrogel as a magnetic strain sensor by exposing the ferrogel to the external magnet; (<b>f</b>) photographs of magnetic sensitivity of the ferrogel (i and iii) before exposing to a permanent magnet and (ii and iv) after exposing to a permanent magnet.</p>
Full article ">
25 pages, 6831 KiB  
Article
Hyperthermia of Magnetically Soft-Soft Core-Shell Ferrite Nanoparticles
by Venkatesha Narayanaswamy, Jayalakshmi Jagal, Hafsa Khurshid, Imaddin A. Al-Omari, Mohamed Haider, Alexander S. Kamzin, Ihab M. Obaidat and Bashar Issa
Int. J. Mol. Sci. 2022, 23(23), 14825; https://doi.org/10.3390/ijms232314825 - 26 Nov 2022
Cited by 6 | Viewed by 2448
Abstract
Magnetically soft-soft MnFe2O4-Fe3O4 core-shell nanoparticles were synthesized through a seed-mediated method using the organometallic decomposition of metal acetyl acetonates. Two sets of core-shell nanoparticles (S1 and S2) of similar core sizes of 5.0 nm and different [...] Read more.
Magnetically soft-soft MnFe2O4-Fe3O4 core-shell nanoparticles were synthesized through a seed-mediated method using the organometallic decomposition of metal acetyl acetonates. Two sets of core-shell nanoparticles (S1 and S2) of similar core sizes of 5.0 nm and different shell thicknesses (4.1 nm for S1 and 5.7 nm for S2) were obtained by changing the number of nucleating sites. Magnetic measurements were conducted on the nanoparticles at low and room temperatures to study the shell thickness and temperature dependence of the magnetic properties. Interestingly, both core-shell nanoparticles showed similar saturation magnetization, revealing the ineffective role of the shell thickness. In addition, the coercivity in both samples displayed similar temperature dependencies and magnitudes. Signatures of spin glass (SG) like behavior were observed from the field-cooled temperature-dependent magnetization measurements. It was suggested to be due to interface spin freezing. We observed a slight and non-monotonic temperature-dependent exchange bias in both samples with slightly higher values for S2. The effective magnetic anisotropy constant was calculated to be slightly larger in S2 than that in S1. The magnetothermal efficiency of the chitosan-coated nanoparticles was determined by measuring the specific absorption rate (SAR) under an alternating magnetic field (AMF) at 200–350 G field strengths and frequencies (495.25–167.30 kHz). The S2 nanoparticles displayed larger SAR values than the S1 nanoparticles at all field parameters. A maximum SAR value of 356.5 W/g was obtained for S2 at 495.25 kHz and 350 G for the 1 mg/mL nanoparticle concentration of ferrogel. We attributed this behavior to the larger interface SG regions in S2, which mediated the interaction between the core and shell and thus provided indirect exchange coupling between the core and shell phases. The SAR values of the core-shell nanoparticles roughly agreed with the predictions of the linear response theory. The concentration of the nanoparticles was found to affect heat conversion to a great extent. The in vitro treatment of the MDA-MB-231 human breast cancer cell line and HT-29 human colorectal cancer cell was conducted at selected frequencies and field strengths to evaluate the efficiency of the nanoparticles in killing cancer cells. The cellular cytotoxicity was estimated using flow cytometry and an MTT assay at 0 and 24 h after treatment with the AMF. The cells subjected to a 45 min treatment of the AMF (384.50 kHz and 350 G) showed a remarkable decrease in cell viability. The enhanced SAR values of the core-shell nanoparticles compared to the seeds with the most enhancement in S2 is an indication of the potential for tailoring nanoparticle structures and hence their magnetic properties for effective heat generation. Full article
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) XRD pattern, (<b>b</b>) field-dependent magnetization hysteresis loops at 5 and 300 K, (<b>c</b>) ZFC and FC temperature-dependent magnetization plots, (<b>d</b>) bright-field TEM image, (<b>e</b>) SAED pattern, and (<b>f</b>) size distribution of the MnFe<sub>2</sub>O<sub>4</sub> nanoparticles.</p>
Full article ">Figure 2
<p>(<b>a</b>) XRD of the core-shell nanoparticles S1 and S2. (<b>b</b>) Shift in the highest intensity peak (311) of the S1 and S2 nanoparticles.</p>
Full article ">Figure 3
<p>TEM bright-field images, SAED patterns, and size-distribution histograms of the S1 (<b>a</b>–<b>c</b>) and S2 (<b>d</b>–<b>f</b>) core-shell nanoparticles.</p>
Full article ">Figure 3 Cont.
<p>TEM bright-field images, SAED patterns, and size-distribution histograms of the S1 (<b>a</b>–<b>c</b>) and S2 (<b>d</b>–<b>f</b>) core-shell nanoparticles.</p>
Full article ">Figure 4
<p>The (311) peak positions obtained from the Gaussian fitting of the XRD diffraction peak (Black-Experimental data and Red-Fitted data).</p>
Full article ">Figure 5
<p>MH plots of the S1 and S2 core-shell nanoparticles at 300 K.</p>
Full article ">Figure 6
<p>Coercivity as a function of temperature and field cooling for the 0, 1, 2, and 3 T field-cooling values for the (<b>a</b>) S1 and (<b>b</b>) S2 nanoparticles. Saturation magnetization as a function of temperature and field cooling for the 0, 1, 2, and 3 T field-cooling values for the (<b>c</b>) S1 and (<b>d</b>) S2 nanoparticles.</p>
Full article ">Figure 7
<p>The M-T plots under the ZFC and FC conditions for the (<b>a</b>) S1 and (<b>b</b>) S2 samples.</p>
Full article ">Figure 8
<p>Zoomed portion of the MH plots used for the exchange bias calculations (<b>a</b>) S1 at 25 K under 1 T and (<b>b</b>) S2 at 200 K under 0 T.</p>
Full article ">Figure 9
<p>M vs. 1/H<sup>2</sup> plots obtained at high field strengths for the (<b>a</b>) S1 and (<b>b</b>) S2 nanoparticles at room temperature (experimental data are marked by symbols, and the solid lines represent a linear fit of the experimental data using Equation (4)). (<b>c</b>) Temperature-dependent K<sub>eff</sub> values of the S1 and S2 nanoparticles.</p>
Full article ">Figure 10
<p>(<b>a</b>) Heating profiles of the seed, S1, and S2 core-shell nanoparticles at 495.25 kHz and a 350 G field strength, (<b>b</b>) frequency-dependent heating profiles of the S1 (1 mg/mL) nanoparticles under 350 G, (<b>c</b>) frequency-dependent heating profiles of the S2 (1 mg/mL) nanoparticles under 350 G, and (<b>d</b>) frequency-dependent SAR values of the S1 and S2 nanoparticles under a 350 G AC field in the frequency range of 495.25–167.30 kHz.</p>
Full article ">Figure 11
<p>Heating profiles of the S2 core-shell nanoparticles; (<b>a</b>) frequency-dependent heating profiles of the S2 (2 mg/mL) nanoparticles under 350 G, (<b>b</b>) field-dependent heating profiles of the S2 (2 mg/mL) nanoparticles at a frequency of 495.25 kHz, (<b>c</b>) frequency-dependent heating profiles of the S2 (3 mg/mL) nanoparticles under 350 G, (<b>d</b>) concentration vs. frequency-dependent SAR under a field amplitude of 350 G, and (<b>e</b>) field-dependent SAR for the 3 mg/mL concentration at frequencies of 495.25, 330.6, and 167.35 kHz.</p>
Full article ">Figure 12
<p>Concentration-dependent SAR values vs. the field at frequencies of (<b>a</b>) 495.10 kHz and (<b>b</b>) 330.6 kHz.</p>
Full article ">Figure 13
<p>MTT assay after the AMF treatment; (<b>a</b>) MDA-MB-231 0 h, (<b>b</b>) HT-29 0 h, (<b>c</b>) MDA-MB-231 24 h, and (<b>d</b>) HT-29 24 h (**** indicates that P value is less than 0.0001).</p>
Full article ">Figure 14
<p>(<b>a</b>) Apoptosis assay of the MDA-MB-231 cell line 0 h after the AMF treatment. (<b>b</b>) Apoptosis assay of the MDA-MB-231 cell line 24 h after the AMF treatment. (<b>c</b>) Apoptosis assay of the HT29 cell line 0 h after the AMF treatment. (<b>d</b>) Apoptosis assay of the HT29 cell line 24 h after the AMF treatment.</p>
Full article ">Figure 14 Cont.
<p>(<b>a</b>) Apoptosis assay of the MDA-MB-231 cell line 0 h after the AMF treatment. (<b>b</b>) Apoptosis assay of the MDA-MB-231 cell line 24 h after the AMF treatment. (<b>c</b>) Apoptosis assay of the HT29 cell line 0 h after the AMF treatment. (<b>d</b>) Apoptosis assay of the HT29 cell line 24 h after the AMF treatment.</p>
Full article ">Figure 14 Cont.
<p>(<b>a</b>) Apoptosis assay of the MDA-MB-231 cell line 0 h after the AMF treatment. (<b>b</b>) Apoptosis assay of the MDA-MB-231 cell line 24 h after the AMF treatment. (<b>c</b>) Apoptosis assay of the HT29 cell line 0 h after the AMF treatment. (<b>d</b>) Apoptosis assay of the HT29 cell line 24 h after the AMF treatment.</p>
Full article ">
16 pages, 3409 KiB  
Article
Mechanical Force Acting on Ferrogel in a Non-Uniform Magnetic Field: Measurements and Modeling
by Felix A. Blyakhman, Alexander P. Safronov, Andrey Yu. Zubarev, Grigory Yu. Melnikov, Sergey Yu. Sokolov, Aitor Larrañaga Varga and Galina V. Kurlyandskaya
Micromachines 2022, 13(8), 1165; https://doi.org/10.3390/mi13081165 - 23 Jul 2022
Cited by 4 | Viewed by 1671
Abstract
The development of magnetoactive microsystems for targeted drug delivery, magnetic biodetection, and replacement therapy is an important task of present day biomedical research. In this work, we experimentally studied the mechanical force acting in cylindrical ferrogel samples due to the application of a [...] Read more.
The development of magnetoactive microsystems for targeted drug delivery, magnetic biodetection, and replacement therapy is an important task of present day biomedical research. In this work, we experimentally studied the mechanical force acting in cylindrical ferrogel samples due to the application of a non-uniform magnetic field. A commercial microsystem is not available for this type of experimental study. Therefore, the original experimental setup for measuring the mechanical force on ferrogel in a non-uniform magnetic field was designed, calibrated, and tested. An external magnetic field was provided by an electromagnet. The maximum intensity at the surface of the electromagnet was 39.8 kA/m and it linearly decreased within 10 mm distance from the magnet. The Ferrogel samples were based on a double networking polymeric structure which included a chemical network of polyacrylamide and a physical network of natural polysaccharide guar. Magnetite particles, 0.25 micron in diameter, were embedded in the hydrogel structure, up to 24% by weight. The forces of attraction between an electromagnet and cylindrical ferrogel samples, 9 mm in height and 13 mm in diameter, increased with field intensity and the concentration of magnetic particles, and varied within 0.1–30 mN. The model provided a fair evaluation of the mechanical forces that emerged in ferrogel samples placed in a non-uniform magnetic field and proved to be useful for predicting the deformation of ferrogels in practical bioengineering applications. Full article
(This article belongs to the Special Issue Nanoparticles in Biomedical Sciences)
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) XRD spectrum of magnetic particles with Miller indexes corresponding to each reflection; (<b>b</b>) transmission electron microscopy images; (<b>c</b>) scanning electron microscopy images, of magnetite particles used as a magnetic filler for ferrogels and model epoxy composite; (<b>d</b>) the average PSD for 632 particles, the line corresponds to the normal distribution probability function with a median of 248 nm and a dispersion of 57 nm.</p>
Full article ">Figure 2
<p>Schematic sketch of the experimental setup for measuring the mechanical force on ferrogel in a non-uniform magnetic field. (<b>1</b>) Cuvette with a ferrogel sample; (<b>2</b>) electromagnet; (<b>3</b>) stabilized power supply; (<b>4</b>) force transducer. See explanation in the text.</p>
Full article ">Figure 3
<p>Magnetic hysteresis loops of magnetite particles in the VSM: (<b>a</b>) Magnetite particles (Particles 100%), ferrogels with different concentration of the magnetite particles used as a filler (FG 23.8%, FG 16.7%, FG 6.5%, FG 4.3%, Gel 0.0%—the numbers correspond to the weight percent of the particles), and epoxy composite with 30.1 wt.% of the particles. Inset shows the concentration dependence of the magnetic moment of the samples of different types mentioned above for the magnetization measured in the external field of 1450 kA/m, being close to saturation magnetization. Points are experimental data, a line is the linear fit, the arrow indicates the point corresponding to model epoxy composite; (<b>b</b>) the low-field dependence of the initial magnetization (primary magnetization curve) of the model epoxy sample (sphere) with 30% (wt.) of magnetite particles, measurement performed starting with the demagnetized state. The solid lines are guides for observations.</p>
Full article ">Figure 4
<p>General sketch of the geometry of the considered measuring system: <span class="html-italic">1</span>—ferrogel sample; <span class="html-italic">2</span>—electromagnet. Small blue arrows illustrate the distribution of the magnetic field lines and the big magenta arrow shows the direction of the force F orientation.</p>
Full article ">Figure 5
<p>Characteristic plots of the field strength of the electromagnet: (<b>a</b>) Dependence of the field in the center of the surface of the magnet on the electric current value. H(I) dependence can be fitted by the equation y = 76.6X + 1.5; (<b>b</b>) dependences of the field strength on the distance from the surface. Dependences can be fitted by the following equations: 1 − y = 0.6X + 10.0, 2 − y = 0.3X + 5.8, 3 − y = 1.2X + 18.9, 4 − y = 0.5X + 11.0, 5 − y = 1.6X + 24.8, and −y = 0.7X + 14.9.</p>
Full article ">Figure 6
<p>Mechanical forces emerged in ferrogel samples in a non-uniform magnetic field as a function of the field intensity at the surface of the electromagnet. Symbols correspond to the experimental values measured for the ferrogel sampes with different weight fraction of magnetite particles given in the legend. Lines are plots calculated using Equation (6).</p>
Full article ">Figure 7
<p>(<b>a</b>) Acting force as a function of magnetite weight fraction in ferrogel at selected levels (23.9, 31.8, and 39.8 kA/m) of magnetic field intensity at the surface of the electromagnet. Symbols correspond to the experimental values. Lines give the model evaluation for the dependence according to Equation (6); (<b>b</b>) parts of the magnetic hysteresis loops measured by VSM for ferrogel samples with 23.8, 16.7, 6.5, and 4.3 wt.% of magnetic filler. Dashed vertical lines indicate the same fields as used for part (<b>a</b>): 1—23.9 kA/m, 2—31.8 kA/m, and 3—39.8 kA/m. It is important to mention that the field strengths at the part (<b>a</b>) correspond to the gradient field intensity at the magnet surface. However, the field strengths at the part (<b>b</b>) correspond to the uniform external magnetic field used for VSM measurements.</p>
Full article ">
7 pages, 3503 KiB  
Proceeding Paper
Precipitation of Iron Oxide in Hydrogel with Superparamagnetic and Stimuli-Responsive Properties
by Alice Mieting, Sitao Wang, Mia Schliephake, Daniela Franke, Margarita Guenther, Stefan Odenbach and Gerald Gerlach
Chem. Proc. 2021, 5(1), 49; https://doi.org/10.3390/chemproc2021005049 - 17 Dec 2021
Viewed by 1467
Abstract
In this work, we present a template-based preparation of iron oxide-containing hydrogels (ferrogels) with ionic sensitive and superparamagnetic properties. The influence of the cross-linked template polyacrylamide and the concentration of the iron salts and sodium hydroxide on the precipitation of the iron oxide [...] Read more.
In this work, we present a template-based preparation of iron oxide-containing hydrogels (ferrogels) with ionic sensitive and superparamagnetic properties. The influence of the cross-linked template polyacrylamide and the concentration of the iron salts and sodium hydroxide on the precipitation of the iron oxide particles is investigated with respect to the stability of the ferrogels. Scanning electron microscope images show cubic particles, which can be semiquantitatively classified in three groups of particle size with respect to the dilution level. Magnetic hysteresis curves reveal a sigmoidal shape without remanence and coercivity for all samples. The higher cross-linked ferrogels, in comparison with the lower cross-linked ferrogels, possess a steady-state degree of swelling in ultrapure water and a stimuli-sensitive deswelling over a wide range of varying ionic strengths. Thus, they are suitable candidates for applications in sensing and microfluidics. Full article
Show Figures

Figure 1

Figure 1
<p>Scheme of the precipitation of iron oxide in hydrogel using iron chloride and sodium hydroxide solutions (<b>A</b>). Overview of the transformation of a sensor hydrogel to a ferrogel by soaking the samples in iron salt solution with orange-yellow coloring and forming black-brown-colored ferrogels in sodium hydroxide at different dilution levels (<b>B</b>).</p>
Full article ">Figure 2
<p>Representative SEM images of sensor ferrogels (<b>a</b>–<b>c</b>) and actuator ferrogels (<b>d</b>–<b>f</b>) of the different dilution levels. Scale bar: 1 µm.</p>
Full article ">Figure 3
<p>Magnetic hysteresis loop of the sensor ferrogels (<b>a</b>) and the actuator ferrogels (<b>b</b>) of the different dilution levels. Inset in the diagrams shows the curves at a low magnetic field.</p>
Full article ">Figure 4
<p>Swelling cycles showing the repeatability of the deswelling in 1 M NaCl solution and swelling in ultrapure water (0 M NaCl) of the sensor ferrogels (<b>a</b>) and actuator ferrogels (<b>b</b>).</p>
Full article ">Figure 5
<p>Swelling sensitivity of sensor ferrogels in NaCl solutions versus ionic strength.</p>
Full article ">
17 pages, 4497 KiB  
Article
Echogenic Advantages of Ferrogels Filled with Magnetic Sub-Microparticles
by Olga A. Dinislamova, Antonina V. Bugayova, Tatyana F. Shklyar, Alexander P. Safronov and Felix A. Blyakhman
Bioengineering 2021, 8(10), 140; https://doi.org/10.3390/bioengineering8100140 - 11 Oct 2021
Cited by 3 | Viewed by 2497
Abstract
Ultrasonic imaging of ferrogels (FGs) filled with magnetic nanoparticles does not reflect the inner structure of FGs due to the small size of particles. To determine whether larger particle size would improve the acoustic properties of FGs, biocompatible hydrogels filled with 100–400 nm [...] Read more.
Ultrasonic imaging of ferrogels (FGs) filled with magnetic nanoparticles does not reflect the inner structure of FGs due to the small size of particles. To determine whether larger particle size would improve the acoustic properties of FGs, biocompatible hydrogels filled with 100–400 nm iron oxide magnetic sub-microparticles with weight fraction up to 23.3% were synthesized and studied. Polymeric networks of synthesized FGs were comprised of chemically cross-linked polyacrylamide with interpenetrating physical network of natural polysaccharide—Guar or Xanthan. Cylindrical samples approximately 10 mm in height and 13 mm in diameter were immersed in a water bath and examined using medical ultrasound (8.5 MHz). The acoustic properties of FGs were characterized by the intensity of reflected echo signal. It was found that the echogenicity of sub-microparticles provides visualization not only of the outer geometry of the gel sample but of its inner structure as well. In particular, the echogenicity of FGs interior depended on the concentration of magnetic particles in the FGs network. The ultrasound monitoring of the shape, dimensions, and inner structure of FGs in the applied external magnetic field is demonstrated. It is especially valuable for the application of FGs in tissue engineering and regenerative medicine. Full article
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) SEM image of magnetite particles; (<b>b</b>) particle size distribution given by graphical image analysis of 632 particles in spherical approximation.</p>
Full article ">Figure 2
<p>General view of PAAm/Xanthan (<b>a</b>) and PAAm/Guar (<b>b</b>) FGs samples filled with MPs at various weight fractions. (<b>a</b>) Weight fraction of MPs: 0.0%, 1.9%, 4.3%, 5.7%, 12.9%, 22.3%; (<b>b</b>) 0.0%, 2.9%, 4.6%, 10.3%, 16.0%, 23.3% (from the left to the right). Diameter of samples is about 13 mm.</p>
Full article ">Figure 3
<p>Example of ferrogel visualization. (PAAm/Xanthan, 12.9% of MPs). Squares with marks 1–4 illustrate locations where specific parameters of brightness were calculated. See also explanation in the text.</p>
Full article ">Figure 4
<p>Optical microphotograph of FG with PAAm/Xanthan network and iron oxide content 1.9% (wt.). Image obtained by “Micromed-I” microscope (Micromed Ltd., St. Petersburg, Russia).</p>
Full article ">Figure 5
<p>Dependence of swelling ratio of FG in PAAm/Guar and PAAm/Xanthan series on the content of iron oxide MPs. Lines are given for an eye-guide only.</p>
Full article ">Figure 6
<p>Dependences of storage modulus (G’) and loss modulus (G’’) on the MPs concentration for samples PAAm/Xanthan (<b>a</b>) and PAAm/Guar (<b>b</b>). Vertical bars reflect the confidence interval with <span class="html-italic">p</span> = 0.01 (<span class="html-italic">n</span> = 5). Equations of linear regression are: (<b>a</b>) y = 0.15x + 1.60 (R<sup>2</sup> = 0.973) for G’ and y = 0.01x + 0.11 (R<sup>2</sup> = 0.85) for G’’; (<b>b</b>) y = 0.11x + 2.81 (R<sup>2</sup> = 0.98) for G’ and y = 0.01x + 0.26 (R<sup>2</sup> = 0.95) for G’’.</p>
Full article ">Figure 7
<p>Examples of ultrasound imaging of gels with various content of fillers. (<b>a</b>) PAAm gel; (<b>b</b>) PAAm gel filled with iron oxide nanoparticles (~14 nm in diameter), weight fraction of 2%; (<b>c</b>) PAAm/Xanthan blank gel; (<b>d</b>) PAAm/Guar blank gel; (<b>e</b>) PAAm/Xanthan gel filled with iron oxide microparticles, weight fraction 1.9%; (<b>f</b>) PAAm/Xanthan gel filled with iron oxide microparticles, weight fraction 22.3%; (<b>g</b>) PAAm/Guar gel filled with iron oxide microparticles, weight fraction 2.9%; (<b>h</b>) PAAm/Guar gel filled with iron oxide microparticles, weight fraction 23.3%.</p>
Full article ">Figure 7 Cont.
<p>Examples of ultrasound imaging of gels with various content of fillers. (<b>a</b>) PAAm gel; (<b>b</b>) PAAm gel filled with iron oxide nanoparticles (~14 nm in diameter), weight fraction of 2%; (<b>c</b>) PAAm/Xanthan blank gel; (<b>d</b>) PAAm/Guar blank gel; (<b>e</b>) PAAm/Xanthan gel filled with iron oxide microparticles, weight fraction 1.9%; (<b>f</b>) PAAm/Xanthan gel filled with iron oxide microparticles, weight fraction 22.3%; (<b>g</b>) PAAm/Guar gel filled with iron oxide microparticles, weight fraction 2.9%; (<b>h</b>) PAAm/Guar gel filled with iron oxide microparticles, weight fraction 23.3%.</p>
Full article ">Figure 8
<p>Dependences of maximal brightness inside the FGs on the distance from the upper gel/water boundary in PAAm/Xanthan sample (<b>a</b>) and PAAm/Guar sample (<b>b</b>). Vertical bars reflect the confidence interval with <span class="html-italic">p</span> = 0.01 (<span class="html-italic">n</span> = 13). See also explanation in the text.</p>
Full article ">Figure 9
<p>Dependences of maximal (1) and average brightness (2) at gel/water boundary on the MPs concentration for samples PAAm/Xanthan (<b>a</b>) and PAAm/Guar (<b>b</b>). Vertical bars reflect the confidence interval with <span class="html-italic">p</span> = 0.01 (<span class="html-italic">n</span> = 60). Linear regressions with <span class="html-italic">x</span> being the concentration of iron oxide MPs in % (wt.): (<b>a</b>) <span class="html-italic">I<sub>max</sub></span> = 6.5<span class="html-italic">x</span> + 129.6 (R<sup>2</sup> = 0.953), <span class="html-italic">I<sub>av</sub></span> = 4.9<span class="html-italic">x</span> + 110 (R<sup>2</sup> = 0.968); (<b>b</b>) <span class="html-italic">I<sub>max</sub></span> = 5.9<span class="html-italic">x</span> + 122.4 (R<sup>2</sup> = 0.969), <span class="html-italic">I<sub>av</sub></span> = 5.0<span class="html-italic">x</span> + 99.5 (R<sup>2</sup> = 0.962).</p>
Full article ">Figure 10
<p>Dependences of maximal (1) and average brightness (2) at gel bulk on the MPs concentration for samples PAAm/Xanthan (<b>a</b>) and PAAm/Guar (<b>b</b>). Vertical bars reflect the confidence interval with <span class="html-italic">p</span> = 0.01.</p>
Full article ">Figure 11
<p>Dependences of the maximum (1) and average (2) brightness on the storage modulus for PAAm/Xanthan (<b>a</b>) and PAAm/Guar (<b>b</b>). Linear regressions: (<b>a</b>) <span class="html-italic">I<sub>max</sub></span> = 31.7<span class="html-italic">G</span>’ + 68.1 (R<sup>2</sup> = 0.946), <span class="html-italic">I<sub>av</sub></span> = 27.8<span class="html-italic">G</span>’ + 49.2 (R<sup>2</sup> = 0.930); (<b>b</b>) <span class="html-italic">I<sub>max</sub></span> = 48.5<span class="html-italic">G</span>’ − 17.1 (R<sup>2</sup> = 0.950), <span class="html-italic">I<sub>av</sub></span> = 41.7<span class="html-italic">G</span>’ − 25 (R<sup>2</sup> = 0.921).</p>
Full article ">Figure 12
<p>Ultrasound visualization of ferrogel deformation in applied magnetic field. FG with PAAm/Guar network and 23.3% content of iron oxide. (<b>a</b>) no field; (<b>b</b>) 400 Oe gradient field applied.</p>
Full article ">
34 pages, 3386 KiB  
Review
Tuning the Cell and Biological Tissue Environment through Magneto-Active Materials
by Jorge Gonzalez-Rico, Emanuel Nunez-Sardinha, Leticia Valencia, Angel Arias, Arrate Muñoz-Barrutia, Diego Velasco and Daniel Garcia-Gonzalez
Appl. Sci. 2021, 11(18), 8746; https://doi.org/10.3390/app11188746 - 19 Sep 2021
Cited by 7 | Viewed by 3175
Abstract
This review focuses on novel applications based on multifunctional materials to actuate biological processes. The first section of the work revisits the current knowledge on mechanically dependent biological processes across several scales from subcellular and cellular level to the cell-collective scale (continuum approaches). [...] Read more.
This review focuses on novel applications based on multifunctional materials to actuate biological processes. The first section of the work revisits the current knowledge on mechanically dependent biological processes across several scales from subcellular and cellular level to the cell-collective scale (continuum approaches). This analysis presents a wide variety of mechanically dependent biological processes on nervous system behaviour; bone development and healing; collective cell migration. In the second section, this review presents recent advances in smart materials suitable for use as cell substrates or scaffolds, with a special focus on magneto-active polymers (MAPs). Throughout the manuscript, both experimental and computational methodologies applied to the different treated topics are reviewed. Finally, the use of smart polymeric materials in bioengineering applications is discussed. Full article
(This article belongs to the Section Applied Biosciences and Bioengineering)
Show Figures

Figure 1

Figure 1
<p>Examples of biological cells and tissue, whose functional response greatly depends on mechanical effects: (<b>A</b>) mechano-electrophysiological response in brain tissue, experimental (adapted with permission from [<a href="#B1-applsci-11-08746" class="html-bibr">1</a>]) (left) and modelling (adapted with permission from [<a href="#B2-applsci-11-08746" class="html-bibr">2</a>]) (right) insights; (<b>B</b>) bone adaptation to mechanical loading, experimental (adapted with permission from [<a href="#B3-applsci-11-08746" class="html-bibr">3</a>]) (left) and modelling (adapted with permission from [<a href="#B4-applsci-11-08746" class="html-bibr">4</a>]) (right) insights; (<b>C</b>) cancer cell proliferation and migration depending on mechanical properties of the substrate, experimental (adapted with permission from [<a href="#B5-applsci-11-08746" class="html-bibr">5</a>]) (left) and modelling (adapted with permission from [<a href="#B6-applsci-11-08746" class="html-bibr">6</a>]) (right) insights; (<b>D</b>) durotaxis in epithelial cell migration depending on mechanical properties of the substrate, experimental (adapted with permission from [<a href="#B7-applsci-11-08746" class="html-bibr">7</a>]) (left) and modelling (adapted with permission from [<a href="#B8-applsci-11-08746" class="html-bibr">8</a>]) (right) insights.</p>
Full article ">Figure 2
<p>Scheme of different manufacturing techniques for magneto-active hydrogels or ferrogels: (<b>A</b>) blending method; (<b>B</b>) in situ precipitation method; (<b>C</b>) grafting-onto method.</p>
Full article ">Figure 3
<p>Scheme of microstructural-based approach, accounting for both magnetic particles and polymer matrix components, and how these models can be scaled up to the micromechanical level. Figure adapted from Garcia-Gonzalez and Hossain [<a href="#B154-applsci-11-08746" class="html-bibr">154</a>].</p>
Full article ">Figure 4
<p>Stages of the deformation process of an MAH surrounded by a solvent with null chemical potential (<math display="inline"><semantics> <mrow> <msub> <mi mathvariant="sans-serif">μ</mi> <mi mathvariant="normal">s</mi> </msub> <mo>=</mo> <mn>0</mn> </mrow> </semantics></math>) and exposed to an external magnetic field. The MAH experiences, in a timeline: (i) initial swelling ratio <math display="inline"><semantics> <mrow> <msub> <mi mathvariant="sans-serif">λ</mi> <mi mathvariant="normal">o</mi> </msub> </mrow> </semantics></math>; (ii) solvent diffusion until reaching a stable swelling ratio <math display="inline"><semantics> <mrow> <msub> <mi mathvariant="sans-serif">λ</mi> <mrow> <mi>sta</mi> </mrow> </msub> </mrow> </semantics></math>; (iii) application of an external magnetic field <span class="html-italic">h</span>; solvent diffusion until reaching a new stable state. Figure adapted from Garcia-Gonzalez and Landis [<a href="#B157-applsci-11-08746" class="html-bibr">157</a>].</p>
Full article ">Figure 5
<p>Bioengineering applications of magneto-active polymers in: (<b>A</b>) bone tissue engineering: mechanical stimulation based on the deformation of a magnetic composite leads to enhanced osteogenic differentiation (adapted with permission from [<a href="#B193-applsci-11-08746" class="html-bibr">193</a>]); (<b>B</b>) drug delivery: alginate ferrogel deformation by application of a magnetic field leads to controlled drug release (adapted with permission from [<a href="#B191-applsci-11-08746" class="html-bibr">191</a>]); (<b>C</b>) hyperthermia treatment: application of a magnetic ferrogel (PNIPAm-Fe<sub>3</sub>O<sub>4</sub>) for tumour encasing and hyperthermia cancer therapy (adapted with permission from [<a href="#B144-applsci-11-08746" class="html-bibr">144</a>]); (<b>D</b>) soft robotics (adapted with permission from [<a href="#B190-applsci-11-08746" class="html-bibr">190</a>]); (<b>E</b>) neuronal regeneration: remote magnetic orientation of collagen fibres in a hydrogel for directed neuronal regeneration (adapted with permission from [<a href="#B198-applsci-11-08746" class="html-bibr">198</a>]).</p>
Full article ">
13 pages, 1497 KiB  
Article
A Model for the Magnetoimpedance Effect in Non-Symmetric Nanostructured Multilayered Films with Ferrogel Coverings
by Nikita A. Buznikov and Galina V. Kurlyandskaya
Sensors 2021, 21(15), 5151; https://doi.org/10.3390/s21155151 - 29 Jul 2021
Cited by 8 | Viewed by 2233
Abstract
Magnetoimpedance (MI) biosensors for the detection of in-tissue incorporated magnetic nanoparticles are a subject of special interest. The possibility of the detection of the ferrogel samples mimicking the natural tissues with nanoparticles was proven previously for symmetric MI thin-film multilayers. In this work, [...] Read more.
Magnetoimpedance (MI) biosensors for the detection of in-tissue incorporated magnetic nanoparticles are a subject of special interest. The possibility of the detection of the ferrogel samples mimicking the natural tissues with nanoparticles was proven previously for symmetric MI thin-film multilayers. In this work, in order to describe the MI effect in non-symmetric multilayered elements covered by ferrogel layer we propose an electromagnetic model based on a solution of the 4Maxwell equations. The approach is based on the previous calculations of the distribution of electromagnetic fields in the non-symmetric multilayers further developed for the case of the ferrogel covering. The role of the asymmetry of the film on the MI response of the multilayer–ferrogel structure is analyzed in the details. The MI field and frequency dependences, the concentration dependences of the MI for fixed frequencies and the frequency dependence of the concentration sensitivities are obtained for the detection process by both symmetric and non-symmetric MI structures. Full article
(This article belongs to the Special Issue Sensors and Biosensors Related to Magnetic Nanoparticles)
Show Figures

Figure 1

Figure 1
<p>Schematic representation of the MI multilayered structures. Symmetric structures of “sandwich” type with equal thicknesses of ferromagnetic and conductive layers (<b>a</b>). Non-symmetric structure of “sandwich” type with different thicknesses of ferromagnetic top and bottom layers (<b>b</b>). Symmetric MI structure with equal thicknesses of top and bottom identical multilayers consisting of ferromagnetic soft magnetic sub-layers separated by non-ferromagnetic spacers and conductive central layer (<b>c</b>). Non-symmetric MI structures with different thicknesses of top and bottom multilayers consisting of ferromagnetic soft magnetic sub-layers separated by non-ferromagnetic spacers and conductive central layer (<b>d</b>). Studied non-symmetric multilayered films: the multilayer with ferrogel placed onto the surface of thin layer (<b>e</b>) and onto the surface of thick layer (<b>f</b>).</p>
Full article ">Figure 2
<p>MI ratio Δ<span class="html-italic">Z</span>/<span class="html-italic">Z</span> as a function of the external field <span class="html-italic">H</span><span class="html-italic"><sub>e</sub></span> at <span class="html-italic">f</span> = <span class="html-italic">ω</span>/2<span class="html-italic">π</span> = 100 MHz for the multilayered element with ferrogel placed onto the surface of thin layer (<span class="html-italic">m</span> = 3, <span class="html-italic">n</span> = 4) (<b>a</b>) and onto the surface of thick layer (<span class="html-italic">m</span> = 4, <span class="html-italic">n</span> = 3) (<b>b</b>) (see also <a href="#sensors-21-05151-f001" class="html-fig">Figure 1</a>e,f). Curves 1, multilayered film without gel; curves 2, film with pure gel (hydrogel) (<span class="html-italic">H</span><span class="html-italic"><sub>p</sub></span> = 0); curves 3, <span class="html-italic">H</span><span class="html-italic"><sub>p</sub></span> = 0.25 Oe; curves 4, <span class="html-italic">H</span><span class="html-italic"><sub>p</sub></span> = 0.5 Oe; curves 5, <span class="html-italic">H</span><span class="html-italic"><sub>p</sub></span> = 0.75 Oe. Parameters used for calculations are <span class="html-italic">l</span> = 1 cm, <span class="html-italic">w</span> = 0.2 mm, 2<span class="html-italic">d</span><sub>0</sub> = 500 nm, <span class="html-italic">d</span><sub>1</sub> = 3 nm, <span class="html-italic">d</span><sub>2</sub> = 100 nm, <span class="html-italic">M</span> = 750 G, <span class="html-italic">H</span><span class="html-italic"><sub>a</sub></span> = 6 Oe, <span class="html-italic">ψ</span> = −0.1π, <span class="html-italic">σ</span><sub>0</sub> = 5 × 10<sup>17</sup> s<sup>−1</sup>, <span class="html-italic">σ</span><sub>1</sub> = 5 × 10<sup>16</sup> s<sup>−1</sup>, <span class="html-italic">σ</span><sub>2</sub> = 3 × 10<sup>16</sup> s<sup>−1</sup>, <span class="html-italic">κ</span> = 0.02, <span class="html-italic">d</span><sub>3</sub> = 1 mm, <span class="html-italic">H</span><span class="html-italic"><sub>c</sub></span> = 6.5 Oe, <span class="html-italic">H</span><sub>1</sub> = 750 Oe and <span class="html-italic">ε</span> = 80.</p>
Full article ">Figure 3
<p>Frequency dependence of maximum MI ratio (Δ<span class="html-italic">Z</span>/<span class="html-italic">Z</span>)<sub>max</sub>: curve 1, multilayered film without gel; curve 2, multilayer with pure gel (hydrogel) placed onto the surface of thin layer (<span class="html-italic">m</span> = 3, <span class="html-italic">n</span> = 4); curve 3, multilayer with pure gel placed onto the surface of thick layer (<span class="html-italic">m</span> = 4, <span class="html-italic">n</span> = 3) (see also <a href="#sensors-21-05151-f001" class="html-fig">Figure 1</a>e,f). Other parameters used for calculations are the same as in <a href="#sensors-21-05151-f002" class="html-fig">Figure 2</a>.</p>
Full article ">Figure 4
<p>Frequency dependence of maximum MI ratio (Δ<span class="html-italic">Z</span>/<span class="html-italic">Z</span>)<sub>max</sub> for the multilayered sensitive element with ferrogel placed onto the surface of thin layer (<span class="html-italic">m</span> = 3, <span class="html-italic">n</span> = 4) (<b>a</b>) or onto the surface of thick layer (<span class="html-italic">m</span> = 4, <span class="html-italic">n</span> = 3) (<b>b</b>) (see also <a href="#sensors-21-05151-f001" class="html-fig">Figure 1</a>e,f). Curves 1 correspond to the multilayered film without gel; curves 2, film with pure gel (hydrogel) (<span class="html-italic">H</span><span class="html-italic"><sub>p</sub></span> = 0); curves 3 <span class="html-italic">H</span><span class="html-italic"><sub>p</sub></span> = 0.25 Oe; curves 4, <span class="html-italic">H</span><span class="html-italic"><sub>p</sub></span> = 0.5 Oe; curves 5, <span class="html-italic">H</span><span class="html-italic"><sub>p</sub></span> = 0.75 Oe. Other parameters used for calculations are the same as in <a href="#sensors-21-05151-f002" class="html-fig">Figure 2</a>.</p>
Full article ">Figure 5
<p>Maximum MI ratio (Δ<span class="html-italic">Z</span>/<span class="html-italic">Z</span>)<sub>max</sub> as a function of the field <span class="html-italic">H</span><span class="html-italic"><sub>p</sub></span> at different frequencies for the multilayer with ferrogel placed onto the surface of thin layer (<span class="html-italic">m</span> = 3, <span class="html-italic">n</span> = 4) (<b>a</b>) and onto the surface of thick layer (<span class="html-italic">m</span> = 4, <span class="html-italic">n</span> = 3) (<b>b</b>) (see also <a href="#sensors-21-05151-f001" class="html-fig">Figure 1</a>e,f). Other parameters used for calculations are the same as in <a href="#sensors-21-05151-f002" class="html-fig">Figure 2</a>.</p>
Full article ">Figure 6
<p>Frequency dependence of the concentration sensitivity <span class="html-italic">S</span><span class="html-italic"><sub>c</sub></span>: curve 1, symmetric multilayered film (<span class="html-italic">m</span> = <span class="html-italic">n</span> = 4); curve 2, non-symmetric multilayer with ferrogel placed onto the surface of thin layer (<span class="html-italic">m</span> = 3, <span class="html-italic">n</span> = 4); curve 3, non-symmetric multilayer with ferrogel placed onto the surface of thick layer (<span class="html-italic">m</span> = 4, <span class="html-italic">n</span> = 3). Other parameters used for calculations are the same as in <a href="#sensors-21-05151-f002" class="html-fig">Figure 2</a>.</p>
Full article ">
21 pages, 6351 KiB  
Article
Specific Absorption Rate Dependency on the Co2+ Distribution and Magnetic Properties in CoxMn1-xFe2O4 Nanoparticles
by Venkatesha Narayanaswamy, Imaddin A. Al-Omari, Aleksandr S. Kamzin, Bashar Issa, Huseyin O. Tekin, Hafsa Khourshid, Hemant Kumar, Ambresh Mallya, Sangaraju Sambasivam and Ihab M. Obaidat
Nanomaterials 2021, 11(5), 1231; https://doi.org/10.3390/nano11051231 - 7 May 2021
Cited by 11 | Viewed by 2502
Abstract
Mixed ferrite nanoparticles with compositions CoxMn1-xFe2O4 (x = 0, 0.2, 0.4, 0.6, 0.8, and 1.0) were synthesized by a simple chemical co-precipitation method. The structure and morphology of the nanoparticles were obtained by X-ray diffraction [...] Read more.
Mixed ferrite nanoparticles with compositions CoxMn1-xFe2O4 (x = 0, 0.2, 0.4, 0.6, 0.8, and 1.0) were synthesized by a simple chemical co-precipitation method. The structure and morphology of the nanoparticles were obtained by X-ray diffraction (XRD), transmission electron microscope (TEM), Raman spectroscopy, and Mössbauer spectroscopy. The average crystallite sizes decreased with increasing x, starting with 34.9 ± 0.6 nm for MnFe2O4 (x = 0) and ending with 15.0 ± 0.3 nm for CoFe2O4 (x = 1.0). TEM images show an edge morphology with the majority of the particles having cubic geometry and wide size distributions. The mixed ferrite and CoFe2O4 nanoparticles have an inverse spinel structure indicated by the splitting of A1g peak at around 620 cm−1 in Raman spectra. The intensity ratios of the A1g(1) and A1g(2) peaks indicate significant redistribution of Co2+ and Fe3+ cations among tetrahedral and octahedral sites in the mixed ferrite nanoparticles. Magnetic hysterics loops show that all the particles possess significant remnant magnetization and coercivity at room temperature. The mass-normalized saturation magnetization is highest for the composition with x = 0.8 (67.63 emu/g), while CoFe2O4 has a value of 65.19 emu/g. The nanoparticles were PEG (poly ethylene glycol) coated and examined for the magneto thermic heating ability using alternating magnetic field. Heating profiles with frequencies of 333.45, 349.20, 390.15, 491.10, 634.45, and 765.95 kHz and 200, 250, 300, and 350 G field amplitudes were obtained. The composition with x = 0.2 (Co0.2Mn0.8Fe2O4) with saturation magnetization 57.41 emu/g shows the highest specific absorption rate (SAR) value of 190.61 W/g for 10 mg/mL water dispersions at a frequency of 765.95 kHz and 350 G field strength. The SAR values for the mixed ferrite and CoFe2O4 nanoparticles increase with increasing concentration of particle dispersions, whereas for MnFe2O4, nanoparticles decrease with increasing the concentration of particle dispersions. SARs obtained for Co0.2Mn0.8Fe2O4 and CoFe2O4 nanoparticles fixed in agar ferrogel dispersions at frequency of 765.95 kHz and 350 G field strength are 140.35 and 67.60 W/g, respectively. This study shows the importance of optimizing the occupancy of Co2+ among tetrahedral and octahedral sites of the spinel system, concentration of the magnetic nanoparticle dispersions, and viscosity of the surrounding medium on the magnetic properties and heating efficiencies. Full article
(This article belongs to the Special Issue Nanoparticles for Bio-Medical Applications)
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) X-ray diffraction patterns of Co<sub>x</sub>Mn<sub>1-x</sub>Fe<sub>2</sub>O<sub>4</sub> (<span class="html-italic">x</span> = 0, 0.2, 0.4, 0.6, 0.8, and 1.0) nanoparticles. (<b>b</b>) Highest intensity peak (311) position shift with respect to cobalt concentration.</p>
Full article ">Figure 2
<p>TEM bright field images, HRTEM, selected area electron diffraction patterns, and size distribution of (<b>a</b>–<b>d</b>) CoFe<sub>2</sub>O<sub>4</sub> and (<b>e</b>–<b>h</b>) Co<sub>0.2</sub>Mn<sub>0.8</sub>Fe<sub>2</sub>O<sub>4</sub> nanoparticles.</p>
Full article ">Figure 2 Cont.
<p>TEM bright field images, HRTEM, selected area electron diffraction patterns, and size distribution of (<b>a</b>–<b>d</b>) CoFe<sub>2</sub>O<sub>4</sub> and (<b>e</b>–<b>h</b>) Co<sub>0.2</sub>Mn<sub>0.8</sub>Fe<sub>2</sub>O<sub>4</sub> nanoparticles.</p>
Full article ">Figure 3
<p>Raman spectra of Co<sub>x</sub>Mn<sub>1-x</sub>Fe<sub>2</sub>O<sub>4</sub> (<span class="html-italic">x</span> = 0, 0.2, 0.4, 0.6, 0.8, and 1.0) nanoparticles (the composition and fitting of the peaks are marked in the figures; the green line indicates the peak fitting).</p>
Full article ">Figure 4
<p>Room temperature Mössbauer spectrum of Co<sub>0.2</sub>Mn<sub>0.8</sub>Fe<sub>2</sub>O<sub>4</sub>. The dots represent the experimental data and the red solid line represents the fitting.</p>
Full article ">Figure 5
<p>(<b>a</b>) Magnetic hysteresis loops obtained at room temperature in the field range of −2.0 T to +2.0 T for Co<sub>x</sub>Mn<sub>1-x</sub>Fe<sub>2</sub>O<sub>4</sub> (<span class="html-italic">x</span> = 0.0, 0.2, 0.4, 0.6, 0.8, and 1.0) nanoparticles. (<b>b</b>) Mass normalized saturation magnetization (M<sub>s</sub>) value vs. Co<sup>2+</sup> concentration of mixed ferrite (line in (<b>b</b>) is just a guide for the eye).</p>
Full article ">Figure 6
<p>The magnetic hysteresis loops obtained at temperature 5 K (<b>a</b>) under zero field cooled (ZFC) condition and (<b>b</b>) under 1 T field cooled condition. (<b>c</b>) The exchange bias field as a function of composition at temperature 5 K under several field cooled values 0 and 1 T. (<b>d</b>) The vertical hysteresis loop shifts as a function of temperature at ZFC and 1 T field cooled conditions (lines in (<b>c</b>,<b>d</b>) are just guides for the eye).</p>
Full article ">Figure 7
<p>Heating profiles of PEG-coated CoFe<sub>2</sub>O<sub>4</sub> nanoparticles: (<b>a</b>) at 350 G for 10 mg/mL concentration, (<b>b</b>) at 765.95 kHz for 10 mg/mL concentration, (<b>c</b>) at 765.95 kHz for 7 mg/mL concentration, and (<b>d</b>) PEG-coated Co<sub>0.2</sub>Mn<sub>0.8</sub>Fe<sub>2</sub>O<sub>4</sub> at 350 G for 10 mg/mL concentration.</p>
Full article ">Figure 8
<p>(<b>a</b>) SAR values as a function of the compositions Co<sub>x</sub>Mn<sub>1-x</sub>Fe<sub>2</sub>O<sub>4</sub> (<span class="html-italic">x</span> = 0, 0.2, 0.4, 0.6, 0.8, and 1.0) nanoparticle with concentrations of 3, 5, 7, and 10 mg/mL at 765.95 kHz frequency and 350 G field strength. (<b>b</b>) Concentration dependent SAR of MnFe<sub>2</sub>O<sub>4</sub>, CoFe<sub>2</sub>O<sub>4</sub>, and Co<sub>0.2</sub>Mn<sub>0.8</sub>Fe<sub>2</sub>O<sub>4</sub> nanoparticles at 765.95 kHz frequency and 350 G field strength (lines are just guides for the eye).</p>
Full article ">Figure 9
<p>(<b>a</b>) SAR values as a function of alternating magnetic field (AMF) frequency at field strength of 350 G for 10 mg/mL of CoFe<sub>2</sub>O<sub>4</sub> and Co<sub>0.2</sub>Mn<sub>0.8</sub>Fe<sub>2</sub>O<sub>4</sub> nanoparticles and (<b>b</b>) SAR values as a function of field strength at fixed AMF frequency of 765.95 kHz for 10 mg/mL of CoFe<sub>2</sub>O<sub>4</sub> and Co<sub>0.2</sub>Mn<sub>0.8</sub>Fe<sub>2</sub>O<sub>4</sub> nanoparticles (lines are just guides for the eye).</p>
Full article ">Figure 10
<p>(<b>a</b>) Images of agar gel and ferrogel of the nanoparticles used for hyperthermia measurements. (<b>b</b>) Heating profiles of 1 mL of distilled water and agar gel.</p>
Full article ">Figure 11
<p>Heating profiles of PEG-coated 10 mg/mL CoFe<sub>2</sub>O<sub>4</sub> nanoparticles and agar ferrogel: (<b>a</b>) at 350 G as a function of frequency and (<b>b</b>) at 765.95 kHz as a function of field strength. Heating profiles of PEG-coated 10 mg/mL Co<sub>0.2</sub>Mn<sub>0.8</sub>Fe<sub>2</sub>O<sub>4</sub> nanoparticles and agar ferrogel: (<b>c</b>) at 350 G fixed field strength and as a function of frequency and (<b>d</b>) at 765.95 kHz fixed frequency and as a function of field strength.</p>
Full article ">Figure 11 Cont.
<p>Heating profiles of PEG-coated 10 mg/mL CoFe<sub>2</sub>O<sub>4</sub> nanoparticles and agar ferrogel: (<b>a</b>) at 350 G as a function of frequency and (<b>b</b>) at 765.95 kHz as a function of field strength. Heating profiles of PEG-coated 10 mg/mL Co<sub>0.2</sub>Mn<sub>0.8</sub>Fe<sub>2</sub>O<sub>4</sub> nanoparticles and agar ferrogel: (<b>c</b>) at 350 G fixed field strength and as a function of frequency and (<b>d</b>) at 765.95 kHz fixed frequency and as a function of field strength.</p>
Full article ">Figure 12
<p>(<b>a</b>) SAR values as a function of AMF frequency at field strength of 350 G for 10 mg/mL of CoFe<sub>2</sub>O<sub>4</sub> and Co<sub>0.2</sub>Mn<sub>0.8</sub>Fe<sub>2</sub>O<sub>4</sub> nanoparticles and agar ferrogel. (<b>b</b>) SAR values as a function of field strength at fixed AMF frequency of 756.95 kHz for 10 mg/mL of CoFe<sub>2</sub>O<sub>4</sub> and Co<sub>0.2</sub>Mn<sub>0.8</sub>Fe<sub>2</sub>O<sub>4</sub> nanoparticles and agar ferrogel (lines are just guides for the eye).</p>
Full article ">
20 pages, 2794 KiB  
Article
Magnetic Properties of Iron Oxide Nanoparticles Do Not Essentially Contribute to Ferrogel Biocompatibility
by Felix A. Blyakhman, Alexander P. Safronov, Emilia B. Makarova, Fedor A. Fadeyev, Tatyana F. Shklyar, Pavel A. Shabadrov, Sergio Fernandez Armas and Galina V. Kurlyandskaya
Nanomaterials 2021, 11(4), 1041; https://doi.org/10.3390/nano11041041 - 19 Apr 2021
Cited by 11 | Viewed by 2649
Abstract
Two series of composite polyacrylamide (PAAm) gels with embedded superparamagnetic Fe2O3 or diamagnetic Al2O3 nanoparticles were synthesized, aiming to study the direct contribution of the magnetic interactions to the ferrogel biocompatibility. The proliferative activity was estimated for [...] Read more.
Two series of composite polyacrylamide (PAAm) gels with embedded superparamagnetic Fe2O3 or diamagnetic Al2O3 nanoparticles were synthesized, aiming to study the direct contribution of the magnetic interactions to the ferrogel biocompatibility. The proliferative activity was estimated for the case of human dermal fibroblast culture grown onto the surfaces of these types of substrates. Spherical non-agglomerated nanoparticles (NPs) of 20–40 nm in diameter were prepared by laser target evaporation (LTE) electrophysical technique. The concentration of the NPs in gel was fixed at 0.0, 0.3, 0.6, or 1.2 wt.%. Mechanical, electrical, and magnetic properties of composite gels were characterized by the dependence of Young’s modulus, electrical potential, magnetization measurements on the content of embedded NPs. The fibroblast monolayer density grown onto the surface of composite substrates was considered as an indicator of the material biocompatibility after 96 h of incubation. Regardless of the superparamagnetic or diamagnetic nature of nanoparticles, the increase in their concentration in the PAAm composite provided a parallel increase in the cell culture proliferation when grown onto the surface of composite substrates. The effects of cell interaction with the nanostructured surface of composites are discussed in order to explain the results. Full article
Show Figures

Figure 1

Figure 1
<p>TEM images of LTE MNPs (<b>a</b>) and ANPs (<b>b</b>). Inserts: histograms of PSDs; lines give fitting of PSD with lognormal distribution function.</p>
Full article ">Figure 2
<p>PSD of MNPs and ANPs in water suspensions measured by DLS. The probability function corresponds to the intensity of dynamic light scattering related to the fraction of nanoparticles with a certain diameter.</p>
Full article ">Figure 3
<p>Magnetic hysteresis loops of Fe<sub>2</sub>O<sub>3</sub> and Al<sub>2</sub>O<sub>3</sub> LTE as-prepared nanoparticles. Inset shows the same responses in low magnetization scale.</p>
Full article ">Figure 4
<p>Magnetic hysteresis loops of gels and gel-based composites: MNP-based (<b>a</b>) and ANP-based (<b>b</b>) materials. G: blank gel; FG: Fe<sub>2</sub>O<sub>3</sub>-based ferrogels; AG: Al<sub>2</sub>O<sub>3</sub>-based gel composites.</p>
Full article ">Figure 5
<p>General view of gel-based samples: (<b>a</b>) ANPs-based composites, (<b>b</b>) MNPs-based composites, and the completely transparent sample is a blank gel; the diameter of the gel samples was equal to 13 mm. The values of Young’s modulus for PAAm hydrogels filed with different concentrations of MNPs or ANPs. Data presented as mean value and standard deviation bar (<span class="html-italic">n</span> = 5).</p>
Full article ">Figure 6
<p>Dependence of electrical potential in gels on the concentration of Fe<sub>2</sub>O<sub>3</sub> or Al<sub>2</sub>O<sub>3</sub> nanoparticles. Data presented as mean value and SD bar (<span class="html-italic">n</span> = 7).</p>
Full article ">Figure 7
<p>Fibroblasts on the surface of gel-based substrates. Cells were cultured for 96 h and fixed for fluorescent microscopy. Cell nuclei were stained with DAPI (4,6-diamidino-2-phenylindole), the cytoplasm was stained with pyrazolone yellow. (<b>a</b>) blank gel substrate; (<b>b</b>) gel-based substrate with 1.2% Al<sub>2</sub>O<sub>3</sub> nanoparticles; (<b>c</b>) gel-based substrate with 1.2% Fe<sub>2</sub>O<sub>3</sub> nanoparticles.</p>
Full article ">Figure 8
<p>General view of the surface of dry of PAAm gel-base composites with Fe<sub>2</sub>O<sub>3</sub> (<b>a</b>) or Al<sub>2</sub>O<sub>3</sub> (<b>b</b>) nanoparticles (1.2 wt.%). Scanning electron microscopy.</p>
Full article ">
23 pages, 4256 KiB  
Article
Magnetorheological Effect of Magnetoactive Elastomer with a Permalloy Filler
by Dmitry Borin, Gennady Stepanov, Anton Musikhin, Andrey Zubarev, Anton Bakhtiiarov and Pavel Storozhenko
Polymers 2020, 12(10), 2371; https://doi.org/10.3390/polym12102371 - 15 Oct 2020
Cited by 18 | Viewed by 2751 | Correction
Abstract
Within the frames of this study, the synthesis of a permalloy to be used as a filler for magnetoactive and magnetorheological elastomers (MAEs and MREs) was carried out. By means of the mechanochemical method, an alloy with the composition 75 wt.% of Fe [...] Read more.
Within the frames of this study, the synthesis of a permalloy to be used as a filler for magnetoactive and magnetorheological elastomers (MAEs and MREs) was carried out. By means of the mechanochemical method, an alloy with the composition 75 wt.% of Fe and 25 wt.% of Ni was obtained. The powder of the product was utilized in the synthesis of MAEs. Study of the magnetorheological (MR) properties of the elastomer showed that in a ~400 mT magnetic field the shear modulus of the MAE increased by a factor of ~200, exhibiting an absolute value of ~8 MPa. Furthermore, we obtained experimentally a relative high loss factor for the studied composite; this relates to the size and morphology of the synthesized powder. The composite with such properties is a very perspective material for magnetocontrollable damping devices. Under the action of an external magnetic field, chain-like structures are formed inside the elastomeric matrix, which is the main determining factor for obtaining a high MR effect. The effect of chain-like structures formation is most pronounced in the region of small strains, since structures are partially destroyed at large strains. A proposed theoretical model based on chain formation sufficiently well describes the experimentally observed MR effect. The peculiarity of the model is that chains of aggregates of particles, instead of individual particles, are considered. Full article
(This article belongs to the Special Issue Magnetic Polymer Composites: Design and Application)
Show Figures

Figure 1

Figure 1
<p>Carbonyl iron microparticles reduced in hydrogen.</p>
Full article ">Figure 2
<p>Carbonyl nickel microparticles.</p>
Full article ">Figure 3
<p>Fe/Ni alloy microparticles.</p>
Full article ">Figure 4
<p>Magnetization curves of Fe–Ni systems with xFe = 0, 25, 50, 75, and 100 wt.%.</p>
Full article ">Figure 5
<p>The size distribution for 1—Fe, 2—Ni and 3—Fe/Ni alloy.</p>
Full article ">Figure 6
<p>Shear modulus as a function of deformation in various magnetic fields.</p>
Full article ">Figure 7
<p>Shear storage modulus of the Fe/Ni alloy and carbonyl iron-based samples as a func-tion of magnetic field (at a strain of 0.00014).</p>
Full article ">Figure 8
<p>Dependence of the loss factor of the material on deformation (strain) and magnetic field for the Fe–Ni alloy-based sample.</p>
Full article ">Figure 9
<p>Dependence of the loss factor of the material on deformation (strain) and magnetic field for the reference carbonyl iron-based sample.</p>
Full article ">Figure 10
<p>Sketch of the model system: (<b>a</b>) isotropic distributed spherical-shaped agglomerates; (<b>b</b>) agglomerates united into a linear chain aligned along an externally applied field; (<b>c</b>) deviation of the chain from the field direction under macroscopic shear deformation.</p>
Full article ">Figure 11
<p>Sketch of the cubic lattice. The centers of each agglomerate can be located at any point within its segment <span class="html-italic">S<sub>1</sub></span>.</p>
Full article ">Figure 12
<p>Sketch of three first stages (<span class="html-italic">k</span> = 0, 1, 2) of the agglomerates aggregation. The horizontal arrows illustrate evolution of the agglomerates in time. The segments of possible positions of the chains are shown. The segment’s boundaries are for the poles of the agglomerates at the chains’ extremities. The single agglomerates and the chains are shown in the centers of the segments of their possible positions.</p>
Full article ">Figure 13
<p>Illustration of the chains’ displacement towards each other. Left—the relative position of the chains immediately after their formation; right—after the displacement. Horizontal arrow—evolution in time.</p>
Full article ">Figure 14
<p>Illustration of the distance between centers of <span class="html-italic">k</span>-th and <span class="html-italic">j</span>-th agglomerates located in neighboring two <span class="html-italic">n</span>-agglomerate chains.</p>
Full article ">Figure 15
<p>Illustration of modeling the chain-like aggregates as ellipsoids of revolution along the axis z. The chain parallel to the applied field corresponds to the non-deformed sample, and the declined chain to the sheared sample.</p>
Full article ">Figure 16
<p>The mean number <math display="inline"> <semantics> <mrow> <mfenced close="&gt;" open="&lt;"> <mi>n</mi> </mfenced> </mrow> </semantics> </math> of agglomerates in the chains vs. applied magnetic field induction <span class="html-italic">B</span>. System parameters: <math display="inline"> <semantics> <mrow> <msub> <mi>χ</mi> <mn>0</mn> </msub> <mo>=</mo> <mn>300</mn> <mo>,</mo> <mo> </mo> <msub> <mi>G</mi> <mn>0</mn> </msub> <mo>=</mo> <mn>10</mn> <mo> </mo> <mi>kPa</mi> <mo>,</mo> <mo> </mo> <msub> <mi>M</mi> <mi>s</mi> </msub> <mo>=</mo> <mn>820</mn> <mo> </mo> <mi mathvariant="normal">k</mi> <mfrac> <mi mathvariant="normal">A</mi> <mi mathvariant="normal">m</mi> </mfrac> <mo>,</mo> <mo> </mo> <msub> <mi>d</mi> <mi>a</mi> </msub> <mo>=</mo> <mn>20</mn> <mo> </mo> <mo>µ</mo> <mi mathvariant="normal">m</mi> <mo>,</mo> <mi mathvariant="sans-serif">Φ</mi> <mo>=</mo> <mn>0.3</mn> </mrow> </semantics> </math>.</p>
Full article ">Figure 17
<p>Shear modulus vs. the applied magnetic field. System parameters: <math display="inline"> <semantics> <mrow> <msub> <mi>χ</mi> <mn>0</mn> </msub> <mo>=</mo> <mn>300</mn> <mo>,</mo> <mo> </mo> <msub> <mi>G</mi> <mn>0</mn> </msub> <mo>=</mo> <mn>10</mn> <mo> </mo> <mi>kPa</mi> <mo>,</mo> <mo> </mo> <msub> <mi>M</mi> <mi>s</mi> </msub> <mo>=</mo> <mn>820</mn> <mo> </mo> <mi mathvariant="normal">k</mi> <mfrac> <mi mathvariant="normal">A</mi> <mi mathvariant="normal">m</mi> </mfrac> <mo>,</mo> <mo> </mo> <msub> <mi>d</mi> <mi>a</mi> </msub> <mo>=</mo> <mn>20</mn> <mo> </mo> <mo>µ</mo> <mi mathvariant="normal">m</mi> <mo>,</mo> <mi mathvariant="sans-serif">Φ</mi> <mo>=</mo> <mn>0.437</mn> <mo>,</mo> <msub> <mi mathvariant="sans-serif">Φ</mi> <mi>m</mi> </msub> <mo>=</mo> <mn>0.63</mn> </mrow> </semantics> </math>. Line—theory, squares—experimental data shown in <a href="#polymers-12-02371-f007" class="html-fig">Figure 7</a>. Volume concentration of the embedded particles is 30%.</p>
Full article ">
20 pages, 4781 KiB  
Article
Effects of Constant Magnetic Field to the Proliferation Rate of Human Fibroblasts Grown onto Different Substrates: Tissue Culture Polystyrene, Polyacrylamide Hydrogel and Ferrogels γ-Fe2O3 Magnetic Nanoparticles
by Felix A. Blyakhman, Grigory Yu. Melnikov, Emilia B. Makarova, Fedor A. Fadeyev, Daiana V. Sedneva-Lugovets, Pavel A. Shabadrov, Stanislav O. Volchkov, Kamiliya R. Mekhdieva, Alexander P. Safronov, Sergio Fernández Armas and Galina V. Kurlyandskaya
Nanomaterials 2020, 10(9), 1697; https://doi.org/10.3390/nano10091697 - 28 Aug 2020
Cited by 15 | Viewed by 2833
Abstract
The static magnetic field was shown to affect the proliferation, adhesion and differentiation of various types of cells, making it a helpful tool for regenerative medicine, though the mechanism of its impact on cells is not completely understood. In this work, we have [...] Read more.
The static magnetic field was shown to affect the proliferation, adhesion and differentiation of various types of cells, making it a helpful tool for regenerative medicine, though the mechanism of its impact on cells is not completely understood. In this work, we have designed and tested a magnetic system consisting of an equidistant set of the similar commercial permanent magnets (6 × 4 assay) in order to get insight on the potential of its experimental usage in the biological studies with cells culturing in a magnetic field. Human dermal fibroblasts, which are widely applied in regenerative medicine, were used for the comparative study of their proliferation rate on tissue culture polystyrene (TCPS) and on the polyacrylamide ferrogels with 0.00, 0.63 and 1.19 wt % concentrations of γ-Fe2O3 magnetic nanoparticles obtained by the well-established technique of laser target evaporation. We used either the same batch as in previously performed but different biological experiments or the same fabrication conditions for fabrication of the nanoparticles. This adds special value to the understanding of the mechanisms of nanoparticles contributions to the processes occurring in the living systems in their presence. The magnetic field increased human dermal fibroblast cell proliferation rate on TCPS, but, at the same time, it suppressed the growth of fibroblasts on blank gel and on polyacrylamide ferrogels. However, the proliferation rate of cells on ferrogels positively correlated with the concentration of nanoparticles. Such a dependence was observed both for cell proliferation without the application of the magnetic field and under the exposure to the constant magnetic field. Full article
(This article belongs to the Special Issue Nanomaterials and Nanotechnology for Regenerative Medicine)
Show Figures

Figure 1

Figure 1
<p>The geometry of the experiment and the model of the magnetic matrix (<b>a</b>). Segmented view of the mesh in finite-element modeling (<b>b</b>).</p>
Full article ">Figure 2
<p>Magnetic field distribution created by the magnetic system in XY plane at Z = 0 cm.</p>
Full article ">Figure 3
<p>Magnetic field distribution created by the magnetic system in XY plane at Z = 2 cm: (<b>a</b>) results of measurements; (<b>b</b>) results of finite-element computer simulation.</p>
Full article ">Figure 4
<p>Arrangement of the cellular experiments with the use of plates’ stack. See explanation in text.</p>
Full article ">Figure 5
<p>(<b>a</b>) TEM image (JEOL JEM2100) of iron oxide magnetic nanoparticles (MNPs) used in the synthesis of ferrogels; (<b>b</b>) particle size distribution obtained by the graphical processing of 2570 images. The line corresponds to the lognormal fitting of the histogram.</p>
Full article ">Figure 6
<p>(<b>a</b>) XRD pattern of iron oxide MNPs with Miller indexes for the inverse spinel lattice of magnetite (or maghemite). (<b>b</b>) Gel samples used for cells culturing (from left to right): BG, FG1, FG2. Diameter of gel samples is 13 mm.</p>
Full article ">Figure 7
<p>(<b>a</b>) Zero-field cooled and field cooled (ZFC-FC) thermomagnetic curves of air-dry laser target evaporation (LTE) iron oxide MNPs; (<b>b</b>) magnetic hysteresis loops of LTE iron oxide MNPs measured for fixed selected temperatures; (<b>c</b>) magnetic hysteresis loops of blank gel (BG) and ferrogels FG1 and FG2; (<b>d</b>) magnetic hysteresis loops of blank gel and ferrogels in the low-fields region. Dashed line indicates external magnetic field strength of 400 Oe, i.e., typical field strength created by magnetic system in the experiments with cell cultures. Blue arrow indicates maximum value of saturation magnetization for FG2 and red arrow indicates maximum value of saturation magnetization for FG1.</p>
Full article ">Figure 8
<p>SEM image of dried FG1 sample (7.7 wt. % of MNPs in dried state, 0.63 wt. % in swollen state) covered with 20 nm conductive carbon layer in order to enhance conductivity. Brighter agglomerates remind the aggregations of MNPs.</p>
Full article ">Figure 9
<p>Human dermal fibroblasts on tissue culture polystyrene (TCPS) without magnetic field application (<b>a</b>) and in magnetic field (<b>b</b>) after 4 days of growth. Magnification ×100, staining of the cell cytoplasm with pyrazolone yellow.</p>
Full article ">Figure 10
<p>The effects of a magnetic field on the fibroblasts’ proliferation rate in different scaffolds. Data are presented as <span class="html-italic">X ± m</span>, n = 54 (9 fields of view per well ×6 wells). Label TCPS (1) corresponds to the proliferation of fibroblasts from donor #1, the numerical values are given in <a href="#nanomaterials-10-01697-t002" class="html-table">Table 2</a> (all 24 wells). All other labels correspond to the proliferation of fibroblasts from donor #2. A level of significance <span class="html-italic">p</span> &lt; 0.05 was determined for FG1 (2), FG2 (2), and TCPS (2).</p>
Full article ">
Back to TopTop