[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
remove_circle_outline

Journals

Article Types

Countries / Regions

Search Results (77)

Search Parameters:
Keywords = piroxicam

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
17 pages, 1977 KiB  
Article
Pyrimidine Derivatives as Selective COX-2 Inhibitors with Anti-Inflammatory and Antioxidant Properties
by Beata Tylińska, Anna Janicka-Kłos, Tomasz Gębarowski, Paulina Nowotarska, Stanisława Plińska and Benita Wiatrak
Int. J. Mol. Sci. 2024, 25(20), 11011; https://doi.org/10.3390/ijms252011011 - 13 Oct 2024
Viewed by 410
Abstract
Pyrimidine derivatives exhibit a wide range of biological activities, including anti-inflammatory properties. The aim of this study was to investigate the effects of tested pyrimidine derivatives on the activity of cyclooxygenase isoenzymes (COX-1 and COX-2), antioxidant properties, and their ability to inhibit the [...] Read more.
Pyrimidine derivatives exhibit a wide range of biological activities, including anti-inflammatory properties. The aim of this study was to investigate the effects of tested pyrimidine derivatives on the activity of cyclooxygenase isoenzymes (COX-1 and COX-2), antioxidant properties, and their ability to inhibit the growth of inflammatory cells. In vitro tests were conducted to assess the ability of pyrimidine derivatives L1–L4 to inhibit COX-1 and COX-2 activity using the TMPD oxidation assay (N,N,N',N'-tetramethyl-p-phenylenediamine). The compounds’ ability to inhibit the growth of lipopolysaccharide (LPS)-stimulated THP-1 (human leukemia monocytic) monocyte cells and their impact on reactive oxygen species (ROS) levels in an inflammatory model were also evaluated. The binding properties of human serum albumin (HSA) were assessed using UV–Vis spectroscopy, circular dichroism (CD), and isothermal titration calorimetry (ITC). Among the tested pyrimidine derivatives, L1 and L2 showed high selectivity towards COX-2, outperforming piroxicam and achieving results comparable to meloxicam. In the sulforhodamine B (SRB) assay, L1 and L2 demonstrated dose-dependent inhibition of LPS-stimulated THP-1 cell growth. Additionally, ROS assays indicated that these compounds reduced free radical levels, confirming their antioxidant properties. Binding studies with albumin revealed that L1 and L2 formed stable complexes with HSA. These results suggest that these compounds could serve as a basis for further research into anti-inflammatory and anticancer drugs with reduced toxicity. Full article
(This article belongs to the Section Biochemistry)
17 pages, 2659 KiB  
Systematic Review
Non-Steroidal Anti-Inflammatory Drugs Administered Intra-Articularly in Temporomandibular Joint Disorders: A Systematic Review and Meta-Analysis
by Filip Bliźniak, Maciej Chęciński, Kamila Chęcińska, Karolina Lubecka, Monika Kamińska, Mariusz Szuta, Dariusz Chlubek and Maciej Sikora
J. Clin. Med. 2024, 13(14), 4056; https://doi.org/10.3390/jcm13144056 - 11 Jul 2024
Cited by 2 | Viewed by 969
Abstract
Objectives: This systematic review was designed to summarize randomized controlled trials of intra-articular administration of non-steroidal anti-inflammatory drugs (NSAIDs) for temporomandibular disorders. Methods: Randomized controlled trials regarding intra-articular injections of non-steroidal anti-inflammatory drugs for temporomandibular disorders were included in the review. [...] Read more.
Objectives: This systematic review was designed to summarize randomized controlled trials of intra-articular administration of non-steroidal anti-inflammatory drugs (NSAIDs) for temporomandibular disorders. Methods: Randomized controlled trials regarding intra-articular injections of non-steroidal anti-inflammatory drugs for temporomandibular disorders were included in the review. The final search was conducted on 16 June 2024 in the Bielefeld Academic Search Engine, PubMed, and Scopus databases. Results: Of the 173 identified studies, 6 were eligible for review. In trials comparing arthrocentesis alone to arthrocentesis with NSAIDs, slight differences in joint pain were noted. For tenoxicam, differences were under 1 point on a 0–10 scale after 4 weeks, with inconsistent results. Piroxicam showed no significant difference, and pain levels were minimal in both groups. For maximum mouth opening (MMO), tenoxicam showed no significant difference. Piroxicam increased MMO by nearly 5 mm, based on one small trial with bias concerns. Conclusions: Currently, there is no strong scientific evidence supporting the injection of NSAIDs into the temporomandibular joint to relieve pain or increase jaw movement. Preliminary reports on piroxicam with arthrocentesis and tenoxicam or diclofenac without rinsing justify further research. Full article
(This article belongs to the Section Dentistry, Oral Surgery and Oral Medicine)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Piroxicam—chemical structure. Author: Fuse809, Public domain, via Wikimedia Commons.</p>
Full article ">Figure 2
<p>Tenoxicam—chemical structure. Author: Jü, Public domain, via Wikimedia Commons.</p>
Full article ">Figure 3
<p>Diclofenac—chemical structure. Author: Harbin, Public domain, via Wikimedia Commons.</p>
Full article ">Figure 4
<p>PRISMA flow diagram.</p>
Full article ">Figure 5
<p>Risk of bias assessment—part 1.</p>
Full article ">Figure 6
<p>Risk of bias assessment—part 2.</p>
Full article ">Figure 7
<p>Pain change.</p>
Full article ">Figure 8
<p>MMO change.</p>
Full article ">
12 pages, 662 KiB  
Review
Update on Evidence and Directions in Temporomandibular Joint Injection Techniques: A Rapid Review of Primary Research
by Karolina Lubecka, Kamila Chęcińska, Filip Bliźniak, Maciej Chęciński, Natalia Turosz, Iwona Rąpalska, Adam Michcik, Dariusz Chlubek and Maciej Sikora
J. Clin. Med. 2024, 13(14), 4022; https://doi.org/10.3390/jcm13144022 - 10 Jul 2024
Cited by 2 | Viewed by 1633
Abstract
This rapid review summarizes the latest primary research in temporomandibular joint (TMJ) injection treatment. The final literature searches were conducted on 4 January 2024. Selection was performed systematically following predefined eligibility criteria. Randomized control trials concerning the treatment of TMJ disorders with intra-articular [...] Read more.
This rapid review summarizes the latest primary research in temporomandibular joint (TMJ) injection treatment. The final literature searches were conducted on 4 January 2024. Selection was performed systematically following predefined eligibility criteria. Randomized control trials concerning the treatment of TMJ disorders with intra-articular injections were included. Studies on more invasive interventions were excluded. Quality of life, joint pain and range of mandibular mobility were assessed. Ultimately, 12 studies covering a total of 603 patients qualified. They concerned: (1) arthrocentesis (AC) and the administration of, (2) injectable platelet-rich fibrin (I-PRF), (3) platelet-rich plasma (PRP), (4) hyaluronic acid (HA), (5) non-steroidal anti-inflammatory drugs (NSAIDs), and (6) hypertonic dextrose (HD) with a local anesthetic. The dominant approach was to perform arthrocentesis before administering the appropriate injection substance (I-PRF, PRP, HA, or NSAID). Two current studies on the intra-articular administration of NSAIDs, specifically tenoxicam and piroxicam, are noteworthy. A mixture of PRP and HA was injected in another two trials. These two innovative approaches may prove to be significant directions for further research on injection treatment of TMJs. Full article
(This article belongs to the Section Dentistry, Oral Surgery and Oral Medicine)
Show Figures

Figure 1

Figure 1
<p>Flow diagram.</p>
Full article ">
22 pages, 5195 KiB  
Article
Xanthan–Polyurethane Conjugates: An Efficient Approach for Drug Delivery
by Narcis Anghel, Iuliana Spiridon, Maria-Valentina Dinu, Stelian Vlad and Mihaela Pertea
Polymers 2024, 16(12), 1734; https://doi.org/10.3390/polym16121734 - 19 Jun 2024
Viewed by 680
Abstract
The antifungal agent, ketoconazole, and the anti-inflammatory drug, piroxicam, were incorporated into matrices of xanthan or oleic acid-esterified xanthan (Xn) and polyurethane (PU), to develop topical drug delivery systems. Compared to matrices without bioactive compounds, which only showed a nominal compressive stress of [...] Read more.
The antifungal agent, ketoconazole, and the anti-inflammatory drug, piroxicam, were incorporated into matrices of xanthan or oleic acid-esterified xanthan (Xn) and polyurethane (PU), to develop topical drug delivery systems. Compared to matrices without bioactive compounds, which only showed a nominal compressive stress of 32.18 kPa (sample xanthan–polyurethane) at a strain of 71.26%, the compressive resilience of the biomaterials increased to nearly 50.04 kPa (sample xanthan–polyurethane–ketoconazole) at a strain of 71.34%. The compressive strength decreased to around 30.67 kPa upon encapsulating a second drug within the xanthan–polyurethane framework (sample xanthan–polyurethane–piroxicam/ketoconazole), while the peak sustainable strain increased to 87.21%. The Weibull model provided the most suitable fit for the drug release kinetics. Unlike the materials based on xanthan–polyurethane, those made with oleic acid-esterified xanthan–polyurethane released the active ingredients more slowly (the release rate constant showed lower values). All the materials demonstrated antimicrobial effectiveness. Furthermore, a higher volume of piroxicam was released from oleic acid-esterified xanthan–polyurethane–piroxicam (64%) as compared to xanthan–polyurethane–piroxicam (44%). Considering these results, materials that include polyurethane and either modified or unmodified xanthan showed promise as topical drug delivery systems for releasing piroxicam and ketoconazole. Full article
(This article belongs to the Special Issue Progress in Polymer Networks)
Show Figures

Figure 1

Figure 1
<p>FTIR spectra of xanthan (a) and xanthan esterified with oleic acid (b).</p>
Full article ">Figure 2
<p><sup>1</sup>H–NMR spectra for Xa (<b>a</b>) and XaAO (<b>b</b>).</p>
Full article ">Figure 3
<p>FTIR spectra for materials containing unmodified xanthan: (a) Xn-PU; (b) Xn-PU-P; (c) Xn-PU-K; (d) Xn-PU-P/K.</p>
Full article ">Figure 4
<p>FTIR spectra for materials containing xanthan esterified with oleic acid: (a) XnOA-PU; (b) XnOA-PU-P; (c) XnOA-PU-K; (d) XnOA-PU-P/K.</p>
Full article ">Figure 5
<p>Mechanical properties of swollen Xn-PU-based formulations under compression: (<b>A</b>) Stress–strain profiles of Xn-PU-based formulations; (<b>B</b>) linear dependence of stress–strain curves; (<b>C</b>) elastic modulus calculated according to the standard procedure; (<b>D</b>) maximum sustained compression (dark blue color) and compressive strength (light blue color). The standard deviation is presented as error bars.</p>
Full article ">Figure 6
<p>Mechanical properties of swollen XnOA-PU-based formulations under compression: (<b>A</b>) Stress–strain profiles of XnOA-PU-based formulations; (<b>B</b>) linear dependence of stress–strain curves (<b>C</b>) elastic modulus (dark blue color) calculated according to the standard procedure; (<b>D</b>) maximum sustained compression (dark blue color) and compressive strength (light blue color). The standard deviation is presented as error bars.</p>
Full article ">Figure 7
<p>SEM micrographs of biomaterials.</p>
Full article ">Figure 8
<p>Cumulative drug release over time for biomaterials containing only one bioactive principle.</p>
Full article ">Figure 9
<p>Cumulative drug release over time for biomaterials containing both bioactive principles.</p>
Full article ">Figure 10
<p>Drug loading capacity of xanthan–polyurethane matrices for piroxicam, ketoconazole, and a combination of both.</p>
Full article ">Figure 11
<p>Drug loading efficiency of xanthan–polyurethane matrices for piroxicam, ketoconazole, and a combination of both.</p>
Full article ">Figure 12
<p>In vitro anti-inflammatory activity of the tested materials: (a) Xn-PU; (b) Xn-PU-P; (c) Xn-PU-K; (d) Xn-PU-P/K; (e) XnOA-PU; (f) XnOA-PU-P; (g) XnOA-PU-K; (h) XnOA-PU-P/K.</p>
Full article ">Scheme 1
<p>Synthesis of xanthan oleate.</p>
Full article ">Scheme 2
<p>Synthesis pathways of polyurethane water dispersion.</p>
Full article ">
12 pages, 6694 KiB  
Article
Surface-Enhanced Raman Scattering for Probe Detection via Gold Nanorods and AuNRs@SiO2 Composites
by Huiqin Li, Yanyu Tian, Shaotian Yan, Lijun Ren, Rong Ma, Weiwei Zhao, Hongge Zhang and Shumei Dou
Coatings 2024, 14(5), 530; https://doi.org/10.3390/coatings14050530 - 24 Apr 2024
Viewed by 922
Abstract
In this paper, a self-assembly method was used to prepare gold nanorod composites, and a seed-growth method was used to adjust the amount of AgNO3 solution, enabling the preparation of gold nanorods with different aspect ratios. AuNRs@SiO2 nanocomposite particles were then [...] Read more.
In this paper, a self-assembly method was used to prepare gold nanorod composites, and a seed-growth method was used to adjust the amount of AgNO3 solution, enabling the preparation of gold nanorods with different aspect ratios. AuNRs@SiO2 nanocomposite particles were then prepared by using the Stöber method to coat the gold nanorod surface with silica. Transmission electron microscopy showed that the maximum aspect ratio of the gold nanorods was 4.53, which was achieved using 2 mL of 10 mM AgNO3 solution. The Raman-scattering intensity of the gold nanorods was studied using rhodamine 6G, thiram, melamine, and piroxicam, and detection limits of 10−8 M, 10−5 M, and 10−3 M were, respectively, achieved. As a substrate, these gold nanorods showed good repeatability and reproducibility, and trace detection was successfully achieved. A transmission electron microscopy analysis shows that the SiO2 shell became thicker with increasing tetraethyl orthosilicate addition. Using AuNRs@SiO2 as the base and R6G, thiram, and piroxicam as the probes, measurable detection limits of 10−9 M, 10−6 M, and 10−5 M were achieved, and this composite also showed excellent repeatability and reproducibility. Full article
(This article belongs to the Section Plasma Coatings, Surfaces & Interfaces)
Show Figures

Figure 1

Figure 1
<p>TEM images of AuNRs prepared with different volumes of 10 mM AgNO<sub>3</sub> solution: (<b>a</b>) 1.0 mL, (<b>b</b>) 1.25 mL, (<b>c</b>) 1.5 mL, (<b>d</b>) 1.75 mL, (<b>e</b>) 2.0 mL, and (<b>f</b>) 2.25 mL.</p>
Full article ">Figure 2
<p>TEM images of AuNRs@SiO<sub>2</sub> composites prepared with different volumes of TEOS solution (20% in methanol): (<b>a</b>) 150 μL, (<b>b</b>) 270 μL, (<b>c</b>) 360 μL, and (<b>d</b>) 540 μL.</p>
Full article ">Figure 3
<p>SERS-detection profiles of 10<sup>−5</sup> M R6G probe molecules enhanced by AuNR substrates prepared with different volumes of AgNO<sub>3</sub> solution.</p>
Full article ">Figure 4
<p>SERS-detection profiles of R6G solutions of different concentrations enhanced by AuNR substrates prepared with 2 mL of 10 mM AgNO<sub>3</sub> solution.</p>
Full article ">Figure 5
<p>(<b>a</b>) SERS-detection profiles of R6G probe molecules enhanced by AuNR substrates stored for different amounts of time; (<b>b</b>) intensity histogram of the peak at 1361 cm<sup>−1</sup> obtained using 10 randomly acquired signals.</p>
Full article ">Figure 6
<p>SERS-detection profiles of thiram solution probe molecules with a AuNR substrate.</p>
Full article ">Figure 7
<p>SERS-detection profiles of piroxicam solutions enhanced by a AuNR substrate.</p>
Full article ">Figure 8
<p>SERS-detection profiles of 10<sup>−5</sup> M R6G probe molecules enhanced by AuNRs@SiO<sub>2</sub> substrates prepared with different volumes of TEOS solution (20% in methanol).</p>
Full article ">Figure 9
<p>SERS-detection profiles of different R6G concentrations enhanced by using AuNRs@SiO<sub>2</sub> as a substrate.</p>
Full article ">Figure 10
<p>(<b>a</b>) SERS-detection profiles of the R6G probe molecule enhanced by AuNRs@SiO<sub>2</sub> substrates after storage for different amounts of time; (<b>b</b>) intensity histogram of the peak at 1361 cm<sup>−1</sup> obtained using 10 randomly acquired signals.</p>
Full article ">Figure 11
<p>SERS-detection profiles of thiram solutions enhanced by a AuNR@SiO<sub>2</sub> substrate.</p>
Full article ">Figure 12
<p>SERS-detection profiles of piroxicam solutions enhanced by a AuNR@SiO<sub>2</sub> substrate.</p>
Full article ">
13 pages, 3032 KiB  
Article
Preparation and Investigation of a Nanosized Piroxicam Containing Orodispersible Lyophilizate
by Petra Party, Sándor Soma Sümegi and Rita Ambrus
Micromachines 2024, 15(4), 532; https://doi.org/10.3390/mi15040532 - 15 Apr 2024
Viewed by 1218
Abstract
Non-steroidal anti-inflammatory piroxicam (PRX) is a poorly water-soluble drug that provides relief in different arthritides. Reducing the particle size of PRX increases its bioavailability. For pediatric, geriatric, and dysphagic patients, oral dispersible systems ease administration. Moreover, fast disintegration followed by drug release and [...] Read more.
Non-steroidal anti-inflammatory piroxicam (PRX) is a poorly water-soluble drug that provides relief in different arthritides. Reducing the particle size of PRX increases its bioavailability. For pediatric, geriatric, and dysphagic patients, oral dispersible systems ease administration. Moreover, fast disintegration followed by drug release and absorption through the oral mucosa can induce rapid systemic effects. We aimed to produce an orodispersible lyophilizate (OL) consisting of nanosized PRX. PRX was solved in ethyl acetate and then sonicated into a poloxamer-188 solution to perform spray-ultrasound-assisted solvent diffusion-based nanoprecipitation. The solid form was formulated via freeze drying in blister sockets. Mannitol and sodium alginate were applied as excipients. Dynamic light scattering (DLS) and nanoparticle tracking analysis (NTA) were used to determine the particle size. The morphology was characterized by scanning electron microscopy (SEM). To establish the crystallinity, X-ray powder diffraction (XRPD) and differential scanning calorimetry (DSC) were used. A disintegration and in vitro dissolution test were performed. DLS and NTA presented a nanosized PRX diameter. The SEM pictures showed a porous structure. PRX became amorphous according to the XRPD and DSC curves. The disintegration time was less than 1 min and the dissolution profile improved. The final product was an innovative anti-inflammatory drug delivery system. Full article
Show Figures

Figure 1

Figure 1
<p>Schematic figure of the preparation method (created using <a href="http://BioRender.com" target="_blank">BioRender.com</a>, accessed on 12 February 2024.).</p>
Full article ">Figure 2
<p>NTA results of the NS (red color shows the distribution of the particles, blue color shows the size of the particles at exact points.</p>
Full article ">Figure 3
<p>SEM images of the NS (<b>a</b>) and the OL (<b>b</b>).</p>
Full article ">Figure 4
<p>DSC results of the initial materials, the NS, and the OL.</p>
Full article ">Figure 5
<p>XRPD results of the initial materials, the NS, and the OL.</p>
Full article ">Figure 6
<p>FTIR results of two initial materials: the PM and the OL.</p>
Full article ">Figure 7
<p>Dissolution of the initial drug and the OL. Data are means ± SD (<span class="html-italic">n</span> = 3 independent measurements).</p>
Full article ">
19 pages, 1407 KiB  
Article
Transport of Non-Steroidal Anti-Inflammatory Drugs across an Oral Mucosa Epithelium In Vitro Model
by Grace C. Lin, Heinz-Peter Friedl, Sarah Grabner, Anna Gerhartl and Winfried Neuhaus
Pharmaceutics 2024, 16(4), 543; https://doi.org/10.3390/pharmaceutics16040543 - 15 Apr 2024
Viewed by 1326
Abstract
Non-steroidal anti-inflammatory drugs (NSAIDs) are one of the most prescribed drugs to treat pain or fever. However, oral administration of NSAIDs is frequently associated with adverse effects due to their inhibitory effect on the constitutively expressed cyclooxygenase enzyme 1 (COX-1) in, for instance, [...] Read more.
Non-steroidal anti-inflammatory drugs (NSAIDs) are one of the most prescribed drugs to treat pain or fever. However, oral administration of NSAIDs is frequently associated with adverse effects due to their inhibitory effect on the constitutively expressed cyclooxygenase enzyme 1 (COX-1) in, for instance, the gastrointestinal tract. A systemic delivery, such as a buccal delivery, of NSAIDs would be beneficial and additionally has the advantage of a non-invasive administration route, especially favourable for children or the elderly. To investigate the transport of NSAIDs across the buccal mucosa and determine their potential for buccal therapeutic usage, celecoxib, diclofenac, ibuprofen and piroxicam were tested using an established oral mucosa Transwell® model based on human cell line TR146. Carboxyfluorescein and diazepam were applied as internal paracellular and transcellular marker molecule, respectively. Calculated permeability coefficients revealed a transport ranking of ibuprofen > piroxicam > diclofenac > celecoxib. Transporter protein inhibitor verapamil increased the permeability for ibuprofen, piroxicam and celecoxib, whereas probenecid increased the permeability for all tested NSAIDs. Furthermore, influence of local inflammation of the buccal mucosa on the transport of NSAIDs was mimicked by treating cells with a cytokine mixture of TNF-α, IL-1ß and IFN-γ followed by transport studies with ibuprofen (+ probenecid). Cellular response to pro-inflammatory stimuli was confirmed by upregulation of cytokine targets at the mRNA level, increased secreted cytokine levels and a significant decrease in the paracellular barrier. Permeability of ibuprofen was increased across cell layers treated with cytokines, while addition of probenecid increased permeability of ibuprofen in controls, but not across cell layers treated with cytokines. In summary, the suitability of the in vitro oral mucosa model to measure NSAID transport rankings was demonstrated, and the involvement of transporter proteins was confirmed; an inflammation model was established, and increased NSAID transport upon inflammation was measured. Full article
(This article belongs to the Special Issue Transport of Drugs through Biological Barriers—an Asset or Risk)
Show Figures

Figure 1

Figure 1
<p>Measured cleared volume [µL] from blanks and cells over time [min] of (<b>A</b>) celecoxib, (<b>B</b>) diclofenac, (<b>C</b>) ibuprofen and (<b>D</b>) piroxicam with corresponding diazepam values from experiments. Representative values for carboxyfluorescein (CF) measured in the piroxicam study shown in (<b>E</b>). Results shown as mean ± SD from three independent experiments (<span class="html-italic">n</span> = 4–8). Permeability coefficient of carboxyfluroescein, diazepam and NSAIDs without inhibitors and upon treatment with verapamil and probenecid shown as µm/min (<b>F</b>) and normalised to the respective permeability coefficient of CF (<b>G</b>). Results shown as mean ± SEM from three independent experiments (<span class="html-italic">n</span> = 4–28). Statistical analysis performed as two-way ANOVA with post hoc Holm–Sidak test, * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01, *** <span class="html-italic">p</span> &lt; 0.001 and α = 0.05.</p>
Full article ">Figure 2
<p>Samples from 48 h inflammation studies (control: media, INF: 100 ng/mL IL-1ß, IFN-γ and TNF-α, INF+Ibu: 100 ng/mL IL-1ß, IFN-γ, TNF-α and 100 µM ibuprofen, Ibu: 100 µM ibuprofen) analysed for mRNA expression of JAK1, TRAF2 and RELA. (<b>A</b>) Expression values referred to housekeeping gene 18SrRNA and shown normalised to samples treated with INF as mean ± SD from three independent experiments. Statistical analysis performed as two-way ANOVA with post hoc Holm–Sidak test, * <span class="html-italic">p</span> &lt; 0.05, *** <span class="html-italic">p</span> &lt; 0.001, α = 0.05. (<b>B</b>) Representative Western blot for COX-1 (70 kDa), COX-2 (70–72 kDa) and ß-actin (42 kDa) after inflammation studies. (<b>C</b>) Measured concentration of cytokines in media from inflammation studies on the apical (saliva) and basolateral (blood) compartment. Results shown as mean ± SD from three independent experiments. Statistical analysis performed as two-way ANOVA with post hoc Holm–Sidak test, ** <span class="html-italic">p</span> &lt; 0.01 and α = 0.05.</p>
Full article ">Figure 3
<p>Measured TEER values [Ω × cm<sup>2</sup>] at start (0 h) and end (48 h) of experiment from control cells, cells treated with cytokines (INF), cytokines and 100 µM ibuprofen (INF + Ibu) and 100 µM Ibuprofen (Ibu) (<b>A</b>). Cells showing TEER values lower than 100 Ω*cm<sup>2</sup> at beginning of experiment were excluded from further data analysis. Results shown as mean ± SD from three independent experiments (<span class="html-italic">n</span> = 13–17). Statistical analysis performed as one-way ANOVA following post hoc Holm–Sidak test with *** <span class="html-italic">p</span> &lt; 0.001 and α = 0.05. (<b>B</b>) Corresponding permeability coefficient values of permeability assays with carboxyfluorescein performed at end of experiment (48 h). Results shown as mean ± SD from three independent experiments (<span class="html-italic">n</span> = 13–17). Statistical analysis performed as one-way ANOVA with α = 0.05. (<b>C</b>) Measured permeability coefficient from transport studies with diazepam, ibuprofen and carboxyfluorescein (CF) with and without probenecid (Prob) performed after 48 h. Results shown as mean ± SD from three independent experiments (<span class="html-italic">n</span> = 3–6). Statistical analysis performed as two-way ANOVA following post hoc Holm–Sidak test with *** <span class="html-italic">p</span> &lt; 0.001 and α = 0.05.</p>
Full article ">
25 pages, 4065 KiB  
Article
Material-Sparing Feasibility Screening for Hot Melt Extrusion
by Amanda Pluntze, Scott Beecher, Maria Anderson, Dillon Wright and Deanna Mudie
Pharmaceutics 2024, 16(1), 76; https://doi.org/10.3390/pharmaceutics16010076 - 5 Jan 2024
Viewed by 1901
Abstract
Hot melt extrusion (HME) offers a high-throughput process to manufacture amorphous solid dispersions. A variety of experimental and model-based approaches exist to predict API solubility in polymer melts, but these methods are typically aimed at determining the thermodynamic solubility and do not take [...] Read more.
Hot melt extrusion (HME) offers a high-throughput process to manufacture amorphous solid dispersions. A variety of experimental and model-based approaches exist to predict API solubility in polymer melts, but these methods are typically aimed at determining the thermodynamic solubility and do not take into account kinetics of dissolution or the associated degradation of the API during thermal processing, both of which are critical considerations in generating a successful amorphous solid dispersion by HME. This work aims to develop a material-sparing approach for screening manufacturability of a given pharmaceutical API by HME using physically relevant time, temperature, and shear. Piroxicam, ritonavir, and phenytoin were used as model APIs with PVP VA64 as the dispersion polymer. We present a screening flowchart, aided by a simple custom device, that allows rapid formulation screening to predict both achievable API loadings and expected degradation from an HME process. This method has good correlation to processing with a micro compounder, a common HME screening industry standard, but only requires 200 mg of API or less. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Illustration of the cross section of the MiniMixer.</p>
Full article ">Figure 2
<p>Illustration of the micro compounder (<b>left</b>), showing the sample path during the recirculation (<b>middle</b>) and extrusion modes (<b>right</b>).</p>
Full article ">Figure 3
<p>Degradation prescreening results for the three API systems, with the limiting temperatures listed for each test. (<b>Left</b>) TGA results showing the wt% mass loss after 15 min at each temperature, with the dashed lines showing the 0.5–1 wt% thresholds. The colored data highlight those that led to the determined limiting temperature for each system. (<b>Right</b>) Visual assessment results showing images of the samples after being held isothermally for 15 min at each temperature. The circled samples show the data that led to the determined limiting temperature for each system.</p>
Full article ">Figure 4
<p>Diffractograms of the samples processed on the micro compounder at various API loadings, with the corresponding API diffractogram shown in gray for each system.</p>
Full article ">Figure 5
<p>Dissolution of PXCM (<b>left</b>) and PHY (<b>right</b>) into PVP-VA64 as determined by PXRD at 150 °C and 215 °C, respectively, comparing an unmixed sample and MiniMixer processing.</p>
Full article ">Figure 6
<p>Degradation comparisons for PXCM (20 wt%), RTV (50 wt%), and PHY (25 wt%) MiniMixer (lighter, open circles), and micro compounder processed extrudates (darker, filled circles) as a function of temperature with 3 min of mixing. The colored bar highlights the data collected at the maximum processing temperature determined from degradation prescreening.</p>
Full article ">Figure 7
<p>Degradation over 4 h for an unmixed sample (open circles) compared to the 3 min processed MiniMixer (lighter filled diamond) and micro compounder (darker filled diamond) samples for the three systems studied. The inlay shows a close-up of the first 30 min.</p>
Full article ">Figure 8
<p>HME feasibility screening flowchart for a given API–polymer formulation.</p>
Full article ">
16 pages, 2278 KiB  
Article
Computational Amendment of Parenteral In Situ Forming Particulates’ Characteristics: Design of Experiment and PBPK Physiological Modeling
by Nada M. El Hoffy, Ahmed S. Yacoub, Amira M. Ghoneim, Magdy Ibrahim, Hussein O. Ammar and Nermin Eissa
Pharmaceutics 2023, 15(10), 2513; https://doi.org/10.3390/pharmaceutics15102513 - 23 Oct 2023
Viewed by 1478
Abstract
Lipid and/or polymer-based drug conjugates can potentially minimize side effects by increasing drug accumulation at target sites and thus augment patient compliance. Formulation factors can present a potent influence on the characteristics of the obtained systems. The selection of an appropriate solvent with [...] Read more.
Lipid and/or polymer-based drug conjugates can potentially minimize side effects by increasing drug accumulation at target sites and thus augment patient compliance. Formulation factors can present a potent influence on the characteristics of the obtained systems. The selection of an appropriate solvent with satisfactory rheological properties, miscibility, and biocompatibility is essential to optimize drug release. This work presents a computational study of the effect of the basic formulation factors on the characteristics of the obtained in situ-forming particulates (IFPs) encapsulating a model drug using a 21.31 full factorial experimental design. The emulsion method was employed for the preparation of lipid and/or polymer-based IFPs. The IFP release profiles and parameters were computed. Additionally, a desirability study was carried out to choose the optimum formulation for further morphological examination, rheological study, and PBPK physiological modeling. Results revealed that the type of particulate forming agent (lipid/polymer) and the incorporation of structure additives like Brij 52 and Eudragit RL can effectively augment the release profile as well as the burst of the drug. The optimized formulation exhibited a pseudoplastic rheological behavior and yielded uniformly spherical-shaped dense particulates with a PS of 573.92 ± 23.5 nm upon injection. Physiological modeling simulation revealed the pioneer pharmacokinetic properties of the optimized formulation compared to the observed data. These results assure the importance of controlling the formulation factors during drug development, the potentiality of the optimized IFPs for the intramuscular delivery of piroxicam, and the reliability of PBPK physiological modeling in predicting the biological performance of new formulations with effective cost management. Full article
(This article belongs to the Special Issue Dosage Form Formulation Technologies for Improving Bioavailability)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Example of 3D-response surface plots for the effect of formulation factors on (<b>A</b>) PS, (<b>B</b>) Box–Cox transformation for PS, (<b>C</b>) PDI, (<b>D</b>) Q0.5, (<b>E</b>) MDT, and (<b>F</b>) Box–Cox transformation for MDT.</p>
Full article ">Figure 2
<p>(<b>A</b>) Release profile of the prepared IFPs and (<b>B</b>) the desirability study marked with the chosen formulation.</p>
Full article ">Figure 3
<p>Release profiles of the chosen formulation IFP3 against modified formulations with (<b>A</b>) structural additives and (<b>B</b>) different solvents.</p>
Full article ">Figure 4
<p>(<b>A</b>) Transmission electron micrograph (TEM) of the optimized formulations (IFP3-EBD). (<b>B</b>) PBPK simulated PX plasma concentration–time curves following IM application of IFP3-EBD.</p>
Full article ">
14 pages, 2519 KiB  
Article
Vaccination against Extracellular Vimentin for Treatment of Urothelial Cancer of the Bladder in Client-Owned Dogs
by Diederik J. M. Engbersen, Judy R. van Beijnum, Arno Roos, Marit van Beelen, Jan David de Haan, Guy C. M. Grinwis, Jack A. Schalken, J. Alfred Witjes, Arjan W. Griffioen and Elisabeth J. M. Huijbers
Cancers 2023, 15(15), 3958; https://doi.org/10.3390/cancers15153958 - 3 Aug 2023
Cited by 2 | Viewed by 1757
Abstract
It was recently shown that targeting extracellular vimentin (eVim) is safe and effective in preclinical models. Here, we report the safety and efficacy in client-owned dogs with spontaneous bladder cancer of CVx1, an iBoost technology-based vaccine targeting eVim in combination with COX-2 inhibition. [...] Read more.
It was recently shown that targeting extracellular vimentin (eVim) is safe and effective in preclinical models. Here, we report the safety and efficacy in client-owned dogs with spontaneous bladder cancer of CVx1, an iBoost technology-based vaccine targeting eVim in combination with COX-2 inhibition. This was a single-arm prospective phase 1/2 study with CVx1 in 20 client-owned dogs with spontaneous UC which involved four subcutaneous vaccinations with CVx1 at 2-week intervals for induction of antibody titers, followed by maintenance vaccinations at 2-month intervals. Additionally, daily cyclooxygenase (COX)-2 inhibition with meloxicam was given. The response was assessed by antibody titers, physical condition, abdominal ultrasound and thorax X-ray. The primary endpoints were the development of antibody titers, as well as overall survival compared to a historical control group receiving carboplatin and COX-2 inhibition with piroxicam. Kaplan–Meier survival analysis was performed. All dogs developed antibodies against eVim. Titers were adequately maintained for the duration of this study. A median overall survival of 374 days was observed, which was 196 days for the historical control group (p < 0.01). Short-term grade 1–2 toxicity at the injection site and some related systemic symptoms peri-vaccination were observed. No toxicity was observed related to the induced antibody response. A limitation of this study is the single-arm prospective setting. CVx1 plus meloxicam consistently induced efficient antibody titers, was well tolerated and showed prolonged survival. The results obtained merit further development for human clinical care. Full article
(This article belongs to the Section Tumor Microenvironment)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Antibody response in dogs vaccinated with CVx1. (<b>A</b>) Schematic overview of the study set-up. An ultrasound of the bladder and X-ray of the thorax were performed at diagnosis. Induction vaccinations were given in 2-week intervals with maintenance vaccinations every 2 months. During the induction phase, blood samples were taken every two weeks (S0–S3), at the time point of vaccination and during maintenance monthly blood samples (S4–S∞) were taken. (<b>B</b>) Anti-eVim antibody titers at start of treatment (S0) and after each induction vaccination (S1–S4). (<b>C</b>) Anti-eVim antibody titers divided by gender (male (M) <span class="html-italic">n</span> = 8; female (F) <span class="html-italic">n</span> = 12) at time point S3. (<b>D</b>) Anti-eVim antibody titers divided by body weight (&gt;25 kg (green bar) <span class="html-italic">n</span> = 4; 10–25 kg (red bar) <span class="html-italic">n</span> = 7; &lt;10 kg (blue bar) <span class="html-italic">n</span> = 9). Data are depicted as mean ± SEM.</p>
Full article ">Figure 2
<p>Vimentin is expressed in the tumor vasculature of UC. Hematoxylin/eosin (HE) staining of three different UC tissue samples (<b>upper row</b>). The same samples were stained with mouse monoclonal anti-vimentin antibody (<b>lower row</b>) and show a vascular staining pattern for vimentin. Representative images of a panel of <span class="html-italic">n</span> = 10 are shown. Scale bar 35 μm.</p>
Full article ">Figure 3
<p>Clinical response in dogs vaccinated with CVx1. Kaplan–Meier curve of probability (%) of survival of the CVx1 vaccine (red line) compared to the historical control group (Ctrl; black line) of [<a href="#B21-cancers-15-03958" class="html-bibr">21</a>]. The dotted line marks the median overall survival.</p>
Full article ">Figure 4
<p>Sera of dogs vaccinated with CVx1 react with both murine and human vimentin. (<b>A</b>) Anti-eVim ELISA showing reactivity of the serum of CVx1-vaccinated dog (dbc1 and dbc2) blood samples S1, S2 with dVim (dog vimentin; dark blue bars) and mVim (mouse vimentin; light blue bars). (<b>B</b>) Immunofluorescence of human (HMEC-1) and mouse (SVEC) endothelial cells, stained with pooled dog serum (green) and mouse monoclonal anti-vimentin antibody (red). Single color channels are shown in black/white. (<b>C</b>) Western blot analysis of mouse and human endothelial cell lysates (<b>top</b>) and recombinant vimentin protein (~55 kDa) (<b>bottom</b>) with pooled dog serum (green) and mouse monoclonal anti-vimentin antibody (red). Original blots are shown in <a href="#app1-cancers-15-03958" class="html-app">Figure S4B,C</a>. Overlapping protein detection is visible as yellow at the expected molecular weight ~55 kD.</p>
Full article ">
15 pages, 3277 KiB  
Article
Upturn Strategies for Arachidonic Acid-Induced MC3T3-E1—625 nm Irradiation in Combination with NSAIDs: Dissipating Inflammation and Promoting Healing
by Danyang Liu, Byunggook Kim, Wenqi Fu, Siyu Zhu, Jaeseok Kang, Oksu Kim and Okjoon Kim
Photonics 2023, 10(5), 535; https://doi.org/10.3390/photonics10050535 - 6 May 2023
Cited by 1 | Viewed by 1486
Abstract
Oral surgery, such as tooth extractions and dental implantations, can cause inflammation in the surrounding tissue, especially in bones. Anti-inflammatory drugs are crucial for pain relief and wound healing. Nonsteroidal anti-inflammatory drugs (NSAIDs) and light-emitting diode irradiation (LEDI) at 625 nm have been [...] Read more.
Oral surgery, such as tooth extractions and dental implantations, can cause inflammation in the surrounding tissue, especially in bones. Anti-inflammatory drugs are crucial for pain relief and wound healing. Nonsteroidal anti-inflammatory drugs (NSAIDs) and light-emitting diode irradiation (LEDI) at 625 nm have been used as therapies to reduce inflammation, which ultimately promotes wound healing. The mechanism of these two methods, however, is different, which possibly makes the combined use of the two approaches effective. Therefore, the efficacy of 625 nm LEDI, NSAIDs, or a combination of both on anti-inflammatory and wound healing effects were analyzed in MC3T3-E1. In this study, piroxicam, ibuprofen, indomethacin, and celecoxib were selected as the NSAIDs. The effect of LEDI at 625 nm was investigated by cell viability, prostaglandin E2 (PGE2) release, and the expression of inflammation-related proteins and cell migration-related proteins were evaluated. Additionally, alkaline phosphatase staining with activity, cell migration assay and BrdU cell proliferation assays were performed. Both LEDI and NSAIDs reduced cyclooxygenase-2 (COX-2) and PGE2. Additionally, LEDI promoted cell migration, proliferation, and bone formation as well, but not by NSAIDs. Thus, a combination of LEDI and NSAIDs can benefits the cells in inflammation, which provides upturn strategies for bone healing after tooth extraction. Full article
Show Figures

Figure 1

Figure 1
<p>Schematic of the irradiation system. (<b>A</b>) LEDI machine schematic; (<b>B</b>) schematic of irradiation; (<b>C</b>) irradiation physical diagram; (<b>D</b>) parameters of irradiation instrument.</p>
Full article ">Figure 2
<p>Effects of AA on MC3T3-E1. (<b>A</b>) MC3T3-E1 were treated with 0–400 μM AA for 6 h for cell viability. (<b>B</b>) The expressions of COX-1 and COX-2 were analyzed by Western blotting in a dose-dependent manner of AA. (<b>C</b>) The relative expression levels of the proteins (protein/β-actin) were shown using Image J software. (<b>D</b>) After AA treatment, the release of PGE<sub>2</sub> was detected using an ELISA kit. The PGE<sub>2</sub> release increased after AA induction, and the levels of PGE<sub>2</sub> were positively correlated with the AA concentrations. *** <span class="html-italic">p</span> &lt; 0.001.</p>
Full article ">Figure 3
<p>Effects of 625 nm LEDI on cell viability and PGE<sub>2</sub> in AA-induced MC3T3-E1. (<b>A</b>) MC3T3-E1 were treated with irradiation before AA treating (PreIR+AA), AA treating and irradiation simultaneously (SimulIR+AA), and irradiation after AA treating (PostIR+AA) for cell viability and (<b>B</b>) PGE<sub>2</sub> assay. PreIR+AA group had the most decreased PGE<sub>2</sub> release. (<b>C</b>) Effects of irradiation on COX-1 and COX-2 expression in AA-induced MC3T3-E1 analyzed by Western blotting. (<b>D</b>) The relative expression levels of proteins (protein/β-actin) were shown using Image J software. Compared to the AA group, PreIR+AA group significantly decreased COX-2, which was consistent with PGE<sub>2</sub> release. *** <span class="html-italic">p</span> &lt; 0.001.</p>
Full article ">Figure 4
<p>Effects of NSAIDs on MC3T3-E1 cell viability and PGE<sub>2</sub> release. (<b>A</b>–<b>D</b>) MC3T3-E1 were treated with four kinds of NSAIDs: Piroxicam, ibuprofen, indomethacin, and celecoxib at different concentrations for 1 h, and the cell viability was evaluated. (<b>E</b>–<b>H</b>) Among them, after treating with 10 nM–1 µM of piroxicam, 0.1–10 µM of ibuprofen, 50 nM–0.5 µM of indomethacin, and 10 nM–1 µM of celecoxib, PGE<sub>2</sub> was detected using an ELISA kit. PIRO: piroxicam; IBP: ibuprofen; INDO: indomethacin; CELE: celecoxib; * <span class="html-italic">p</span> &lt; 0.1, *** <span class="html-italic">p</span> &lt; 0.001.</p>
Full article ">Figure 5
<p>Effects of 625 nm LEDI and NSAIDs on cell migration in AA-induced MC3T3-E1. (<b>A</b>) Pictures of the cell migration process. (<b>B</b>) Migration area was measured by Image J software. (<b>C</b>) The expressions of E-cadherin, Vimentin and N-cadherin by Western blotting. (<b>D</b>) The relative expression levels of proteins (protein/β-actin) were shown using Image J software. * <span class="html-italic">p</span> &lt; 0.1, ** <span class="html-italic">p</span> &lt; 0.01, *** <span class="html-italic">p</span> &lt; 0.001. Compared with the presence or absence of AA.</p>
Full article ">Figure 6
<p>Effects of combined 625 nm LEDI and NSAIDs on cell migration and proliferation in AA-Induced MC3T3-E1. (<b>A</b>) The expressions of E-cadherin, N-cadherin, Vimentin and PCNA were analyzed by Western blotting. (<b>B</b>,<b>C</b>) The relative expression levels of proteins (protein/β-actin) were shown using Image J software. (<b>D</b>) Cell proliferation was evaluated with the BrdU Cell Proliferation Assay Kit. (<b>E</b>) After treatment, the release of PGE<sub>2</sub> from MC3T3-E1 was detected by an ELISA kit. * <span class="html-italic">p</span> &lt; 0.1, *** <span class="html-italic">p</span> &lt; 0.001.</p>
Full article ">Figure 7
<p>Effects of combined use of 625 nm LEDI and NSAIDs on bone calcification in AA-induced MC3T3-E1 (<b>A</b>,<b>B</b>) ALP staining and their relative expression was measured by absorbance. (<b>C</b>) ALP activity was detected by an ELISA kit. Cele: celecoxib * <span class="html-italic">p</span> &lt; 0.1, ** <span class="html-italic">p</span> &lt; 0.01, *** <span class="html-italic">p</span> &lt; 0.001.</p>
Full article ">
14 pages, 3664 KiB  
Article
Chrysin Is Immunomodulatory and Anti-Inflammatory against Complete Freund’s Adjuvant-Induced Arthritis in a Pre-Clinical Rodent Model
by Muhammad Asif Faheem, Tasleem Akhtar, Nadia Naseem, Usman Aftab, Muhammad Shoaib Zafar, Safdar Hussain, Muhammad Shahzad and Glenda Carolyn Gobe
Pharmaceutics 2023, 15(4), 1225; https://doi.org/10.3390/pharmaceutics15041225 - 12 Apr 2023
Cited by 3 | Viewed by 2293
Abstract
Chrysin (5,7-dihydroxyflavone) has many pharmacological properties including anti-inflammatory actions. The objective of this study was to evaluate the anti-arthritic activity of chrysin and to compare its effect with the non-steroidal anti-inflammatory agent, piroxicam, against complete Freund’s adjuvant (CFA)-induced arthritis in a pre-clinical model [...] Read more.
Chrysin (5,7-dihydroxyflavone) has many pharmacological properties including anti-inflammatory actions. The objective of this study was to evaluate the anti-arthritic activity of chrysin and to compare its effect with the non-steroidal anti-inflammatory agent, piroxicam, against complete Freund’s adjuvant (CFA)-induced arthritis in a pre-clinical model in rats. Rheumatoid arthritis was induced by injecting CFA intra-dermally in the sub-plantar region of the left hind paw of rats. Chrysin (50 and 100 mg/kg) and piroxicam (10 mg/kg) were given to rats with established arthritis. The model of arthritis was characterized using an index of arthritis, with hematological, biological, molecular, and histopathological parameters. Treatment with chrysin significantly reduced the arthritis score, inflammatory cells, erythrocyte sedimentation rate, and rheumatoid factor. Chrysin also reduced the mRNA levels of tumor necrosis factor, nuclear factor kappa-B, and toll-like recepter-2 and increased anti-inflammatory cytokines interleukin-4 and -10, as well as the hemoglobin levels. Using histopathology and microscopy, chrysin reduced the severity of arthritis in joints, infiltration of inflammatory cells, subcutaneous inflammation, cartilage erosion, bone erosion, and pannus formation. Chrysin showed comparable effects to piroxicam, which is used for the treatment of rheumatoid arthritis. The results showed that chrysin possesses anti-inflammatory and immunomodulatory effects that make it a potential drug for the treatment of arthritis. Full article
Show Figures

Figure 1

Figure 1
<p>Comparison of body weight (<b>a</b>) and arthritis score (<b>b</b>) in all experimental groups (n = 10 per group). ### indicates <span class="html-italic">p</span>-value ≤ 0.001 and shows significant difference from the untreated controls. *** indicates <span class="html-italic">p</span>-value ≤ 0.001 and shows a significant difference from the CFA arthritic group.</p>
Full article ">Figure 2
<p>Comparison of hemoglobin (<b>a</b>) and total leucocyte count (TLC) (<b>b</b>) in all experimental groups (n = 10 per group). ### indicates <span class="html-italic">p</span>-value ≤ 0.001 and shows significant difference from the untreated controls. ** indicates <span class="html-italic">p</span>-value ≤ 0.01 and *** indicates <span class="html-italic">p</span>-value ≤ 0.001. These show a significant difference from the CFA arthritic group.</p>
Full article ">Figure 3
<p>Comparison of erythrocyte sedimentation rate (ESR) (<b>a</b>) and rheumatoid factor (<b>b</b>) in all experimental groups (n = 10 per group). ## indicates <span class="html-italic">p</span>-value ≤ 0.01, and ### indicates <span class="html-italic">p</span>-value ≤ 0.001, showing a significant difference from control. * indicates <span class="html-italic">p</span>-value ≤ 0.05, ** indicates <span class="html-italic">p</span>-value ≤ 0.01 and *** indicates <span class="html-italic">p</span>-value ≤0.001 and they indicate a significant difference from the CFA arthritic group.</p>
Full article ">Figure 4
<p>Comparison of relative mRNA expression level of TNF (<b>a</b>), NF-κB (<b>b</b>), TLR-2 (<b>c</b>), IL-4 (<b>d</b>), and IL-10 (<b>e</b>) in all of the experimental groups (n = 10 per group). ## indicates <span class="html-italic">p</span>-value ≤ 0.01, ### indicates <span class="html-italic">p</span>-value ≤ 0.001, and they show significant differences from the control group. * indicates <span class="html-italic">p</span>-value ≤ 0.05, ** indicates <span class="html-italic">p</span>-value ≤ 0.01, and *** indicates <span class="html-italic">p</span>-value ≤ 0.001, and they show a significant difference from the CFA arthritic group.</p>
Full article ">Figure 5
<p>Histological images (H&amp;E stain, 100X) of examples of ankle joints of different groups including the control group (<b>a</b>), arthritic group (<b>b</b>), chrysin 50 mg/kg (<b>c</b>), chrysin 100 mg/kg (<b>d</b>), and piroxicam 10 mg/kg groups (<b>e</b>) (n = 10 per group). The control group shows no inflammation. Using the extent of inflammatory cell infiltration as an indicator of effect, the CFA arthritic group shows severe inflammation/extensive inflammatory cell infiltration; while (<b>c</b>–<b>e</b>) show an improvement in the extent of inflammatory cell infiltration after treatment with chrysin (50 mg/kg or 100 mg/kg, respectively) or piroxicam in CFA arthritic rats. (<b>a</b>): <span class="html-fig-inline" id="pharmaceutics-15-01225-i001"><img alt="Pharmaceutics 15 01225 i001" src="/pharmaceutics/pharmaceutics-15-01225/article_deploy/html/images/pharmaceutics-15-01225-i001.png"/></span> shows stratified squamous keratinized epithelium. (<b>b</b>): ▲ Shows chronic granulomatous inflammation, <span class="html-fig-inline" id="pharmaceutics-15-01225-i001"><img alt="Pharmaceutics 15 01225 i001" src="/pharmaceutics/pharmaceutics-15-01225/article_deploy/html/images/pharmaceutics-15-01225-i001.png"/></span> Shows pannus formation, <span class="html-fig-inline" id="pharmaceutics-15-01225-i002"><img alt="Pharmaceutics 15 01225 i002" src="/pharmaceutics/pharmaceutics-15-01225/article_deploy/html/images/pharmaceutics-15-01225-i002.png"/></span> Shows peri-articular inflammation. (<b>c</b>): ▲ Showing synovial hyperplasia, <span class="html-fig-inline" id="pharmaceutics-15-01225-i001"><img alt="Pharmaceutics 15 01225 i001" src="/pharmaceutics/pharmaceutics-15-01225/article_deploy/html/images/pharmaceutics-15-01225-i001.png"/></span> Shows articular cartilage with no inflammation. (<b>d</b>): <span class="html-fig-inline" id="pharmaceutics-15-01225-i001"><img alt="Pharmaceutics 15 01225 i001" src="/pharmaceutics/pharmaceutics-15-01225/article_deploy/html/images/pharmaceutics-15-01225-i001.png"/></span> = mild synovial hyperplasia, ▲ = mild peri articular inflammation, <span class="html-fig-inline" id="pharmaceutics-15-01225-i002"><img alt="Pharmaceutics 15 01225 i002" src="/pharmaceutics/pharmaceutics-15-01225/article_deploy/html/images/pharmaceutics-15-01225-i002.png"/></span> = intact articular cartilage. (<b>e</b>): <span class="html-fig-inline" id="pharmaceutics-15-01225-i001"><img alt="Pharmaceutics 15 01225 i001" src="/pharmaceutics/pharmaceutics-15-01225/article_deploy/html/images/pharmaceutics-15-01225-i001.png"/></span> shows mild synovial inflammation and hyperplasia, ▲ shows intact articular cartilage.</p>
Full article ">Figure 6
<p>Comparison of histological scoring while assessing severity of inflammation (<b>a</b>), subcutaneous inflammation (<b>b</b>), synovial inflammation (<b>c</b>), cartilage erosion (<b>d</b>), bone erosion (<b>e</b>), and pannus formation (<b>f</b>) in rats of all experimental groups (n = 10 per group). ## indicates <span class="html-italic">p</span>-value ≤ 0.01, ### indicates <span class="html-italic">p</span>-value ≤ 0.001 showing significant difference from control. ** indicates <span class="html-italic">p</span>-value ≤ 0.01 and *** indicates <span class="html-italic">p</span>-value ≤ 0.001 showing a significant difference from the CFA arthritic group.</p>
Full article ">
16 pages, 2622 KiB  
Article
Study of Formulation and Process Variables for Optimization of Piroxicam Nanosuspension Using 32 Factorial Design to Improve Solubility and In Vitro Bioavailability
by Yahya Alhamhoom, Sandip M. Honmane, Umme Hani, Riyaz Ali M. Osmani, Geetha Kandasamy, Rajalakshimi Vasudevan, Sharanya Paramshetti, Ravindra R. Dudhal, Namrata K. Kengar and Manoj S. Charde
Polymers 2023, 15(3), 483; https://doi.org/10.3390/polym15030483 - 17 Jan 2023
Cited by 7 | Viewed by 2503
Abstract
Piroxicam is a Biopharmaceutical Classification System (BCS) Class II drug having poor aqueous solubility and a short half-life. The rationale behind the present research was to develop a Piroxicam nanosuspension to enhance the solubility and thereby the in vitro bioavailability of the drug. [...] Read more.
Piroxicam is a Biopharmaceutical Classification System (BCS) Class II drug having poor aqueous solubility and a short half-life. The rationale behind the present research was to develop a Piroxicam nanosuspension to enhance the solubility and thereby the in vitro bioavailability of the drug. Piroxicam nanosuspension (PRX NS) was prepared by an anti-solvent precipitation technique and optimized using a full-factorial design. Herein, the nanosuspension was prepared using polymer polyvinylpyrrolidone (PVP) K30® and Poloxamer 188® as a stabilizer to improve the solubility and in vitro bioavailability of the drug. Nine formulations were prepared based on 32 full-factorial experimental designs to study the effect of the formulation variables such as concentration of poloxamer 188 (%) (X1) and stirring speed (rpm) (X2) as a process variable on the response of particle size (nm) and solubility (µg/mL). The prepared NS was characterized by phase solubility, Fourier-transform infrared (FT-IR), differential scanning calorimetry (DSC), X-ray powder diffraction (XRPD), transmission electron microscopy (TEM), particle size, zeta potential, entrapment efficiency, and percent drug release. DSC and XRPD analysis of freeze-dried NS formulation showed conversion of PRX into a less crystalline form. NS formulations showed a reduction in the size from 443 nm to 228 nm with −22.5 to −30.5 mV zeta potential and % drug entrapment of 89.76 ± 0.76. TEM analysis confirmed the size reduction at the nano level. The solubility was increased from 44 μg/mL to 87 μg/mL by altering the independent variables. The solubility of PRX NS in water was augmented by 14- to 15-fold (87.28 μg/mL) than pure PRX (6.6 μg/mL). The optimized formulation (NS9) at drug-to-stabilizer concentration exhibited a greater drug release of approximately 96.07% after 120 min as compared to the other NS formulations and pure PRX (36.78%). Thus, all these results revealed that the prepared NS formulations have improved the solubility and in vitro dissolution compared to the pure drug. Furthermore, an increase in the drug release was observed from the NS than that of the pure PRX. All these outcomes signified that the prepared PRX NS showed an increase in solubility and in vitro dissolution behavior; which subsequently would aid in attainment of enhanced bioavailability. Full article
(This article belongs to the Special Issue Polymers Synthesis and Characterization)
Show Figures

Figure 1

Figure 1
<p>UV spectra depicting λ<sub>max</sub> of PRX in methanol (<b>A</b>), methanolic HCl (<b>B</b>), PBS pH 6.8 (<b>C</b>) and PBS pH 7.4 (<b>D</b>).</p>
Full article ">Figure 2
<p>Overlain FT-IR spectra of PRX (A), physical mixture of PRX and PVP K30<sup>®</sup> (B), physical mixture of PRX and Poloxamer 188<sup>®</sup> (C), physical mixture of PRX and mannitol (D), physical mixture of PRX, PVP K30<sup>®</sup>, Poloxamer 188<sup>®</sup> and mannitol (E), and freeze-dried optimized formulation (NS9) (F).</p>
Full article ">Figure 3
<p>Linear and 3D response surface plot for independent variable’s effects on the response particle size (<b>A</b>,<b>B</b>), respectively, and on response solubility (<b>C</b>,<b>D</b>), respectively.</p>
Full article ">Figure 4
<p>Particle size distribution curve (<b>A</b>) and zeta potential peak (<b>B</b>) of the optimized formulation NS9.</p>
Full article ">Figure 5
<p>TEM micrograph of optimized PRX NS formulation (NS9).</p>
Full article ">Figure 6
<p>Overlain DSC thermograms of the pure drug (PRX) (A) and freeze-dried formulation (NS9) (B).</p>
Full article ">Figure 7
<p>XRPD diffractograms of (<b>A</b>) pure drug PRX and (<b>B</b>) freeze-dried optimized formulation NS9.</p>
Full article ">Figure 8
<p>In vitro drug release profiles of prepared PRX NS formulations.</p>
Full article ">
21 pages, 2412 KiB  
Article
Biomedical Applications of Thermosensitive Hydrogels for Controlled/Modulated Piroxicam Delivery
by Snežana Ilić-Stojanović, Ljubiša Nikolić, Vesna Nikolić, Ivan Ristić, Suzana Cakić and Slobodan D. Petrović
Gels 2023, 9(1), 70; https://doi.org/10.3390/gels9010070 - 15 Jan 2023
Cited by 5 | Viewed by 2091
Abstract
The objectives of this study are the synthesis of thermosensitive poly(N-isopropylacrylamide-co-2-hydroxypropyl methacrylate), p(NiPAm-HPMet), hydrogels and the analysis of a drug-delivery system based on piroxicam, as a model drug, and synthesized hydrogels. A high pressure liquid chromatography method has been [...] Read more.
The objectives of this study are the synthesis of thermosensitive poly(N-isopropylacrylamide-co-2-hydroxypropyl methacrylate), p(NiPAm-HPMet), hydrogels and the analysis of a drug-delivery system based on piroxicam, as a model drug, and synthesized hydrogels. A high pressure liquid chromatography method has been used in order to determine both qualitative and quantitative amounts of unreacted monomers and crosslinkers from polymerized hydrogels. Swelling kinetics and the order of a swelling process of the hydrogels have been analyzed at 10 and 40 °C. The copolymers’ thermal properties have been monitored by the differential scanning calorimetry (DSC) method. DSC termograms have shown that melting occurs in two temperature intervals (142.36–150.72 °C and 153.14–156.49 °C). A matrix system with incorporated piroxicam has been analyzed by using FTIR and SEM methods. Structural analysis has demonstrated that intermolecular non-covalent interactions have been built between side-groups of copolymer and loaded piroxicam. Morphology of p(NiPAm-HPMet) after drug incorporation indicates the piroxicam presence into the copolymer pores. Kinetic parameters of the piroxicam release from hydrogels at 37 °C and pH 7.4 indicate that the fluid transport mechanism corresponds to Fickian diffusion. As a result, formulation of thermosensitive p(NiPAm-HPMet) hydrogels with incorporated piroxicam could be of interest for further testing as a drug carrier for modulated and prolonged release, especially for topical administration. Full article
(This article belongs to the Section Gel Applications)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Differential scanning calorimetry (DSC) curves (first derivative of heat flow) of the synthesized poly(<span class="html-italic">N</span>-isopropylacrylamide-<span class="html-italic">co</span>-2-hydroxypropyl methacrylate) xerogels.</p>
Full article ">Figure 2
<p>Swelling reversibility of p(NiPAm-HPMet) hydrogel at alternating swelling/contraction for three cycles under the temperature change from 10 °C to 50 °C. Error bars represent the standard deviation of the means of three replicates.</p>
Full article ">Figure 3
<p>FTIR spectra of: pure p(NiPAm-HPMet) hydrogel, p(NiPAm-HPMet) hydrogel with incorporated piroxicam and piroxicam with the structural formula.</p>
Full article ">Figure 4
<p>The piroxicam incorporation into the poly(<span class="html-italic">N</span>-isopropylacrylamide-<span class="html-italic">co</span>-2-hydroxypropyl methacrylate) hydrogels (the potential intramolecular interactions is indicated).</p>
Full article ">Figure 5
<p>SEM micrographs of poly(<span class="html-italic">N</span>-isopropylacrylamide-<span class="html-italic">co</span>-2-hydroxypropyl methacrylate) with incorporated piroxicam in the equilibrium swollen state at: (<b>a</b>) magnification 200×, scale bar 100 μm; (<b>b</b>) magnification 2000×, scale bar 10 μm.</p>
Full article ">Figure 6
<p>(<b>a</b>) HPLC chromatogram of piroxicam, (<b>b</b>) UV spectrum of piroxicam from DAD detector.</p>
Full article ">Figure 7
<p>In vitro cumulative release of piroxicam from series of p(NiPAm-HPMet) hydrogels at 37 °C and pH 7.4: (<b>a</b>) in mg/g<sub>xerocel</sub> during 24 h; (<b>b</b>) in percentages during first 4 h. Error bars represent the standard deviation of the means of three replicates.</p>
Full article ">
14 pages, 1438 KiB  
Systematic Review
Phonophoresis through Nonsteroidal Anti-Inflammatory Drugs for Knee Osteoarthritis Treatment: Systematic Review and Meta-Analysis
by Francisco Javier Martin-Vega, David Lucena-Anton, Alejandro Galán-Mercant, Veronica Perez-Cabezas, Carlos Luque-Moreno, Maria Jesus Vinolo-Gil and Gloria Gonzalez-Medina
Biomedicines 2022, 10(12), 3254; https://doi.org/10.3390/biomedicines10123254 - 14 Dec 2022
Cited by 1 | Viewed by 3293
Abstract
Knee osteoarthritis (OA) is the most common joint disease. The administration of nonsteroidal anti-inflammatory drugs (NSAIDs) by phonophoresis is a therapeutic alternative to relieve pain in inflammatory pathologies. The main aim was to analyze the efficacy of the application of NSAIDs by phonophoresis [...] Read more.
Knee osteoarthritis (OA) is the most common joint disease. The administration of nonsteroidal anti-inflammatory drugs (NSAIDs) by phonophoresis is a therapeutic alternative to relieve pain in inflammatory pathologies. The main aim was to analyze the efficacy of the application of NSAIDs by phonophoresis in knee OA. A systematic review and meta-analysis of controlled clinical trials were performed between January and March 2021 in the following databases: Web of Science, Scopus, PubMed, Cinahl, SciELO, and PEDro. The PEDro scale was used to evaluate the level of evidence of the selected studies. The RevMan 5.4 statistical software was used to obtain the meta-analysis. Eight studies were included, of which five were included in the meta-analysis, involving 195 participants. The NSAIDs used through phonophoresis were ibuprofen, piroxicam, diclofenac sodium, diclofenac diethylammonium, ketoprofen, and methyl salicylate. The overall result for pain showed not-conclusive results, but a trend toward significance was found in favor of the phonophoresis group compared to the control group (standardized mean difference (SMD) = −0.92; 95% confidence interval: −1.87–0.02). Favorable results were obtained for physical function (SMD = −1.34; 95% CI: −2.00–0.68). Based on the selected studies, the application of NSAIDs by phonophoresis is effective in relieving the symptoms of knee OA. Future long-term studies are recommended. Full article
(This article belongs to the Section Cell Biology and Pathology)
Show Figures

Figure 1

Figure 1
<p>Flow diagram of the systematic review and meta-analysis.</p>
Full article ">Figure 2
<p>Forest plot for visual analogue scale. The red squares indicate the weight assigned to the study. The horizontal lines depict the confidence interval. The black rhombuses show the overall result.</p>
Full article ">Figure 3
<p>Forest plot for Western Ontario and McMaster Universities Osteoarthritis Index. The red squares indicate the weight assigned to the study. The horizontal lines depict the confidence interval. The black rhombuses show the overall result.</p>
Full article ">Figure 4
<p>Funnel plot showing publication bias for visual analogue scale.</p>
Full article ">Figure 5
<p>Funnel plot showing publication bias for Western Ontario and McMaster Universities Osteoarthritis Index.</p>
Full article ">
Back to TopTop