[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 (40)

Search Parameters:
Keywords = 10-hydroxystearic acid

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
20 pages, 3254 KiB  
Article
Acai Oil-Based Organogel Containing Hyaluronic Acid for Topical Cosmetic: In Vitro and Ex Vivo Assessment
by Suellen Christtine da Costa Sanches, Lindalva Maria de Meneses Costa Ferreira, Rayanne Rocha Pereira, Desireé Gyles Lynch, Ingryd Nayara de Farias Ramos, André Salim Khayat, José Otávio Carrera Silva-Júnior, Alessandra Rossi and Roseane Maria Ribeiro-Costa
Pharmaceutics 2024, 16(9), 1195; https://doi.org/10.3390/pharmaceutics16091195 - 11 Sep 2024
Viewed by 775
Abstract
Organogels are semi-solid pharmaceutical forms whose dispersing phase is an organic liquid, for example, an oil, such as acai oil, immobilized by a three-dimensional network formed by the gelling agent. Organogels are being highlighted as innovative release systems for cosmetic active ingredients such [...] Read more.
Organogels are semi-solid pharmaceutical forms whose dispersing phase is an organic liquid, for example, an oil, such as acai oil, immobilized by a three-dimensional network formed by the gelling agent. Organogels are being highlighted as innovative release systems for cosmetic active ingredients such as hyaluronic acid for topical applications. Acai oil was evaluated for its physicochemical parameters, fatty acid composition, lipid quality index, spectroscopic pattern (Attenuated total reflectance Fourier Transform Infrared Spectroscopy), thermal behavior, total phenolic, total flavonoids, and total carotenoids and β-carotene content. The effectiveness of the organogel incorporated with hyaluronic acid (OG + HA) was evaluated through ex vivo permeation and skin retention tests, in vitro tests by Attenuated total reflectance Fourier Transform Infrared Spectroscopy and Differential Scanning Calorimetry. The physicochemical analyses highlighted that the acai oil exhibited quality standards in agreement with the regulatory bodies. Acai oil also showed high antioxidant capacity, which was correlated with the identified bioactive compounds. The cytotoxicity tests demonstrated that the formulation OG + HA does not release toxic substances into the biological environment that could impede cell growth, adhesion, and efficacy. In vitro and ex vivo analyses demonstrated that after 6 h of application, OG + HA presented a high level of hydration, thermal protection and release of HA. Thus, it can be concluded that the OG + HA formulation has the potential for physical–chemical applications, antioxidant quality, and potentially promising efficacy for application in the cosmetic areas. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>ATR-FTIR spectrum of acai oil.</p>
Full article ">Figure 2
<p>Thermogravimetric and derived thermogravimetry curves of acai oil.</p>
Full article ">Figure 3
<p>Thermogram of the acai oil.</p>
Full article ">Figure 4
<p>Effect of different concentrations of OG + HA on the viability of fibroblast cells of the MRC5 lineage incubated for 48 h at 37 °C in DMEM. Data were expressed as mean ± standard deviation (n = 3). ***: significative values for <span class="html-italic">p</span> &lt; 0.0001 in relation to the control and other concentrations of OG + HA. NS: not significative values among themselves or compared to the control.</p>
Full article ">Figure 5
<p>Effect of OG + HA on pig skin fragments after different retention times in stratum corneum and epidermis/dermis.</p>
Full article ">Figure 6
<p>ATR-FTIR spectra of pig membranes after different retention times of OG + HA.</p>
Full article ">Figure 7
<p>Effect of OG + HA on pig skin fragments after different permeation/retention times. Data were expressed as mean ± standard deviation (n = 3). NS: not significant values among themselves or compared to the control.</p>
Full article ">Figure A1
<p>Ethics Committee for the use of Animals.</p>
Full article ">
12 pages, 6191 KiB  
Article
Chemiluminescent Reaction Induced by Mixing of Fluorescent-Dye-Containing Molecular Organogels with Aqueous Oxidant Solutions
by Yutaka Ohsedo and Kiho Miyata
Gels 2024, 10(8), 492; https://doi.org/10.3390/gels10080492 - 25 Jul 2024
Viewed by 650
Abstract
Chemiluminescence in solution-based systems has been extensively studied for the chemical analysis of biomolecules. However, investigations into the control of chemiluminescence reactions in gel-based systems, which offer flexibility in reaction conditions (such as the softness of the reaction environment), have only recently begun [...] Read more.
Chemiluminescence in solution-based systems has been extensively studied for the chemical analysis of biomolecules. However, investigations into the control of chemiluminescence reactions in gel-based systems, which offer flexibility in reaction conditions (such as the softness of the reaction environment), have only recently begun in polymer materials, with limited exploration in low-molecular-weight gelator (LMWG) systems. In this study, we investigated the chemiluminescence behaviors in the gel states using LMWG systems and evaluated their applicability to fluorescent-dye-containing molecular organogel systems/oxidant-containing aqueous systems. Using diethyl succinate organogels composed of 12-hydroxystearic acid as a molecular organogelator, we examined the fluorescent properties of various fluorescent dyes mixed with oxidant aqueous solutions. As the reaction medium transitioned from the solution to the gel state, the emission color and chemiluminescence duration changed significantly, and distinct characteristics were observed, for each dye. This result indicates that the chemiluminescence behavior differs significantly between the solution and gel states. Additionally, visual inspection and dynamic viscoelastic measurements of the mixed fluorescent dye-containing molecular gels and oxidant-containing aqueous solutions confirmed that the chemiluminescence induced by the mixing occurred within the gel phase. Furthermore, the transition from the solution to the gel state may allow for the modulation of the mixing degree, thereby enabling control over the progression of the chemiluminescence reaction. Full article
(This article belongs to the Special Issue Gel Formation and Processing Technologies for Material Applications)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>(<b>a</b>) Schematic illustration of the research concept of mixing-induced chemiluminescent gel. (<b>b</b>) Chemical structures of the organogelator and fluorescent dyes in this study.</p>
Full article ">Figure 2
<p>Vial inversion method for the gelation test and thixotropy test: (<b>a</b>) Gelation test of 12-HS ethyl phthalate organogels; the solution was heated to dissolve and allowed to cool before inverting the vial. (<b>b</b>) Thixotropy test (the vial inversion method for thixotropy) of 12-HS ethyl phthalate organogels; shear force was applied for 10 s in a vortex mixer to bring the vial to a sol state; then the vial was allowed to stand for 1 min before it was inverted.</p>
Full article ">Figure 3
<p>Dynamic rheological properties with respect to (<b>a</b>) strain sweep, (<b>b</b>) frequency sweep, and (<b>c</b>) thixotropic behavior for the ethyl phthalate organogels containing BD, GD, and RD (denoted such as Gel-BD).</p>
Full article ">Figure 4
<p>Electronic absorption spectra (<b>a</b>) and fluorescent spectra (<b>b</b>) of the fluorescent dyes in ethyl phthalate solutions.</p>
Full article ">Figure 5
<p>Chemiluminescent spectra after mixing in the solution and gel states.</p>
Full article ">Figure 6
<p>Chemiluminescent spectra after mixing in the solution and gel states.</p>
Full article ">Figure 7
<p>Dynamic rheological properties with respect to (<b>a</b>) strain sweep, (<b>b</b>) frequency sweep, and (<b>c</b>) thixotropic behavior the ethyl phthalate organogels containing BD, GD, and RD with aqueous oxidant solutions (denoted such as Mixed Gel-BD).</p>
Full article ">Figure 8
<p>The time-dependent changes in the chemiluminescence intensity. (<b>a</b>) BD systems, (<b>b</b>) GD systems, and (<b>c</b>) RD systems.</p>
Full article ">Figure 9
<p>Measured fluorescence lifetime decay in the solution state (<b>a</b>) and gel state (<b>b</b>). Blue curve: excitation light, red curve: example of fitting curve.</p>
Full article ">
11 pages, 4545 KiB  
Article
Energy Recovery from Municipal Sewage Sludge: An Environmentally Friendly Source for the Production of Biochemicals
by Luigi di Bitonto, Antonella Angelini and Carlo Pastore
Appl. Sci. 2024, 14(12), 4974; https://doi.org/10.3390/app14124974 - 7 Jun 2024
Viewed by 707
Abstract
In this work, a detailed analysis of the lipid component in primary sludge and sewage scum up-taken from several wastewater treatment plants located in southern Italy was carried out. Lipids in the primary sludge accounted for 200–250 mg/g of the total solids (TS), [...] Read more.
In this work, a detailed analysis of the lipid component in primary sludge and sewage scum up-taken from several wastewater treatment plants located in southern Italy was carried out. Lipids in the primary sludge accounted for 200–250 mg/g of the total solids (TS), with calcium soaps as a main component (70–82%), while total lipids made up about 350–500 mg/gTS in the sewage scum and consisted mainly of FFAs (45–60%) and calcium soaps (27–35%). In addition, estolides and 10-hydroxystearic acid were also quantified. A specific valorization process was then developed and tested for either primary sludge or sewage scum. In detail, lipids were first recovered, chemically activated by the addition of acids (calcium soaps were converted to free fatty acids) and finally reacted with methanol to obtain methyl esters. The lipid recovery from primary sludge and sewage scum was particularly efficient (recoverability of 92–99%). The conversion of the starting acids into FAMEs (yield > 98%) was achieved under very mild conditions (70 °C, 2 h) with AlCl3·6H2O as a catalyst. Biodiesel (according to EN14214), methyl 10-hydroxystearate and methyl estolides were efficiently isolated by distillation under vacuum. Finally, a feasibility study of the proposed processes was carried out to evaluate their possible integration into a wastewater treatment plant, critically analyzing both the positive aspects and the relative limitations. Full article
(This article belongs to the Special Issue Waste Valorization, Green Technologies and Circular Economy)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Geolocation of WWTPs located in the Apulia region, southern Italy.</p>
Full article ">Figure 2
<p>Chemical composition of FFAs extracted from primary sludge and sewage scum and determination of Average Molecular Weight (AMW).</p>
Full article ">Figure 3
<p>Gas chromatogram of crude products obtained from the esterification of lipids extracted from primary sludge and sewage scum.</p>
Full article ">Figure 4
<p>Scheme of processes for the valorization of primary sludge and sewage scum.</p>
Full article ">
11 pages, 1353 KiB  
Article
Sulfonic Resins as Catalysts for the Oxidation of Alcohols with H2O2/KBr
by Vicente Dorado, Clara I. Herrerías and José M. Fraile
Catalysts 2024, 14(1), 74; https://doi.org/10.3390/catal14010074 - 17 Jan 2024
Viewed by 1390
Abstract
Sulfonic resins can replace homogeneous sulfonic acids in the oxidation of alcohols with the H2O2/KBr system. The performance of different resins was tested with methyl 9(10)-hydroxystearate, a secondary fatty alcohol. The structural features of the resin were more important [...] Read more.
Sulfonic resins can replace homogeneous sulfonic acids in the oxidation of alcohols with the H2O2/KBr system. The performance of different resins was tested with methyl 9(10)-hydroxystearate, a secondary fatty alcohol. The structural features of the resin were more important than the acid strength for the catalytic performance of this reaction. The optimization of the reaction conditions allows the recovery of the resin, although regeneration is required due to the active role of KBr, and a significant loss of sulfonic groups can be detected after nine runs. In the case of primary fatty alcohols, the oxidation leads to carboxylic acids, which are esterified with the starting alcohol under the acidic conditions. For cyclic secondary alcohols, the steric hindrance around the hydroxyl group seems to be important for the efficiency of the oxidation to ketone. Full article
(This article belongs to the Special Issue Advances in the Catalytic Behavior of Ion-Exchange Resins)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Recovery of Dowex 50W×8 in the oxidation of methyl 9(10)-hydroxystearate (<b>1a</b> + <b>1b</b>) with H<sub>2</sub>O<sub>2</sub>/KBr. Reaction conditions: 0.5 mmol of methyl 9(10)-hydroxystearate, 0.1 mmol of KBr, 0.75 mmol of 60% H<sub>2</sub>O<sub>2</sub>, 0.2 mmol of sulfonic groups, 1 mL of CH<sub>2</sub>Cl<sub>2</sub>/H<sub>2</sub>O (9:1 <span class="html-italic">v</span>/<span class="html-italic">v</span>), rt, 24 h. Blue bars indicate the result after the reactivation of the resin with HCl.</p>
Full article ">Scheme 1
<p>Oxidation of methyl 9(10)-hydroxystearate (<b>1a</b> + <b>1b</b>).</p>
Full article ">Scheme 2
<p>Proposed mechanism for the oxidation of alcohols with the KBr/H<sub>2</sub>O<sub>2</sub>/sulfonic resin method (adapted from [<a href="#B24-catalysts-14-00074" class="html-bibr">24</a>]).</p>
Full article ">Scheme 3
<p>Partial deactivation of sulfonic resins and reactivation with HCl.</p>
Full article ">Scheme 4
<p>Oxidation of different alcohols with KBr/H<sub>2</sub>O<sub>2</sub>/Dowex 50W×8.</p>
Full article ">
25 pages, 2345 KiB  
Article
Asymmetric Synthesis of Saturated and Unsaturated Hydroxy Fatty Acids (HFAs) and Study of Their Antiproliferative Activity
by Olga G. Mountanea, Christiana Mantzourani, Dimitrios Gkikas, Panagiotis K. Politis and George Kokotos
Biomolecules 2024, 14(1), 110; https://doi.org/10.3390/biom14010110 - 15 Jan 2024
Viewed by 1302
Abstract
Hydroxy fatty acids (HFAs) constitute a class of lipids, distinguished by the presence of a hydroxyl on a long aliphatic chain. This study aims to expand our insights into HFA bioactivities, while also introducing new methods for asymmetrically synthesizing unsaturated and saturated HFAs. [...] Read more.
Hydroxy fatty acids (HFAs) constitute a class of lipids, distinguished by the presence of a hydroxyl on a long aliphatic chain. This study aims to expand our insights into HFA bioactivities, while also introducing new methods for asymmetrically synthesizing unsaturated and saturated HFAs. Simultaneously, a procedure previously established by us was adapted to generate new HFA regioisomers. An organocatalytic step was employed for the synthesis of chiral terminal epoxides, which either by alkynylation or by Grignard reagents resulted in unsaturated or saturated chiral secondary alcohols and, ultimately, HFAs. 7-(S)-Hydroxyoleic acid (7SHOA), 7-(S)-hydroxypalmitoleic acid (7SHPOA) and 7-(R)- and (S)-hydroxymargaric acids (7HMAs) were synthesized for the first time and, together with regioisomers of (R)- and (S)-hydroxypalmitic acids (HPAs) and hydroxystearic acids (HSAs), whose biological activity has not been tested so far, were studied for their antiproliferative activities. The unsaturation of the long chain, as well as an odd-numbered (C17) fatty acid chain, led to reduced activity, while the new 6-(S)-HPA regioisomer was identified as exhibiting potent antiproliferative activity in A549 cells. 6SHPA induced acetylation of histone 3 in A549 cells, without affecting acetylated α-tubulin levels, suggesting the selective inhibition of histone deacetylase (HDAC) class I enzymes, and was found to inhibit signal transducer and activator of transcription 3 (STAT3) expression. Full article
(This article belongs to the Collection Bioactive Lipids in Inflammation, Diabetes and Cancer)
Show Figures

Figure 1

Figure 1
<p>Saturated hydroxy fatty acids (SHFAs) and their roles.</p>
Full article ">Figure 2
<p>Antiproliferative activity of various HFAs (human lung cancer cell line A549). Cells were treated with increasing concentrations (10 µM, 25 µM, 35 µM, 50 µM, 75 µM, 100 µM) of the test compounds for 72 h and cell viability was determined by the MTT assay for a minimum of three experiments (AVG ± SEM).</p>
Full article ">Figure 3
<p>Inhibition curves of enantiomers 6SHPA, 6RHSA and 6SHSA in A549 cells. IC<sub>50</sub> values were derived from the dose–response relationship for a minimum of six experiments (95% CI log IC<sub>50</sub> −4.53 to −4.38, −4.71 to −3.71 and −4.57 to −4.28, respectively).</p>
Full article ">Figure 4
<p>Antiproliferative activity of HFAs (human astrocytoma cell line SF268). Cells were treated with increasing concentrations (10 µM, 25 µM, 35 µM, 50 µM, 75 µM, 100 µM) of the test compounds for 72 h and cell viability was determined by the MTT assay for a minimum of three experiments (AVG ± SEM).</p>
Full article ">Figure 5
<p>Inhibition curves of enantiomers 6SHSA and 6SHPA in SF268 cells. IC<sub>50</sub> values were derived from the dose–response relationship for a minimum of six experiments (95% CI log IC<sub>50</sub> −4.15 to −4.06 and −4.23 to −4.17, respectively).</p>
Full article ">Figure 6
<p>6SHPA induces the acetylation of histone 3 and inhibits STAT3 expression in A549 cells. Cells treated with 6SHPA (50 µM) versus DMSO and assessed after 48 h. (<b>A</b>,<b>B</b>) Western blot analysis for acetylated α-tubulin upon 6SHPA treatment. (<b>A</b>,<b>C</b>) Western blot analysis for acetylated histone 3 upon 6SHPA treatment. (<b>A</b>,<b>D</b>) Western blot analysis for STAT3 expression upon 6SHPA treatment. Protein expression levels quantitated by using the ImageJ software. All experiments were repeated three times (<span class="html-italic">n</span> = 3). For all cases, *** <span class="html-italic">p</span> &lt; 0.001. Original images can be found in <a href="#app1-biomolecules-14-00110" class="html-app">Supplementary File S1</a>.</p>
Full article ">Scheme 1
<p>Asymmetric synthesis of unsaturated HFAs. (a) PCC, dry CH<sub>2</sub>Cl<sub>2</sub>; (b) i. (2<span class="html-italic">S</span>,5<span class="html-italic">R</span>)-2-(<span class="html-italic">tert</span>-butyl)-3,5-dimethylimidazolidin-4-one trifluoroacetate (20%), 2,3,4,5,6,6-hexachlorocyclohexa-2,4-dien-1-one, THF; ii. NaBH<sub>4</sub>, EtOH, iii. KOH/EtOH/H<sub>2</sub>O; (c) 1-decyne or 1-octyne, n-BuLi, BF<sub>3</sub>·OEt<sub>2</sub>, dry THF; (d) Lindlar’s cat. (5%), H<sub>2</sub>, quinoline, CH<sub>3</sub>OH; (e) AcCl, dry pyridine, dry CH<sub>2</sub>Cl<sub>2</sub>; (f) TBAF, dry THF; (g) Jones reagent, acetone; (h) LiOH·H<sub>2</sub>O, THF:H<sub>2</sub>O.</p>
Full article ">Scheme 2
<p>Asymmetric synthesis of (<span class="html-italic">R</span>)-6HPA, (<span class="html-italic">R</span>)-8HPA, (<span class="html-italic">R</span>)-11HPA, (<span class="html-italic">R</span>)-7HMA, (<span class="html-italic">R</span>)-8HSA and (<span class="html-italic">R</span>)-11HSA. (a) i. (2<span class="html-italic">S</span>,5<span class="html-italic">R</span>)-2-(<span class="html-italic">tert</span>-butyl)-3,5-dimethylimidazolidin-4-one trifluoroacetate (20%), 2,3,4,5,6,6-hexachlorocyclohexa-2,4-dien-1-one, THF; ii. NaBH<sub>4</sub>, EtOH; iii. KOH/EtOH/H<sub>2</sub>O; (b) RMgX, CuI, dry THF; (c) AcCl, dry pyridine, dry CH<sub>2</sub>Cl<sub>2</sub>; (d) H<sub>2</sub>, 10% Pd/C, EtOH; (e) Jones reagent, acetone; (f) LiOH·H<sub>2</sub>O, THF:H<sub>2</sub>O.</p>
Full article ">Scheme 3
<p>Asymmetric synthesis of (<span class="html-italic">S</span>)- and (<span class="html-italic">R</span>)-6HSAs. (a) i. (2<span class="html-italic">R</span>,5<span class="html-italic">S</span>)-2-(<span class="html-italic">tert</span>-butyl)-3,5-dimethylimidazolidin-4-one trifluoroacetate (20%), 2,3,4,5,6,6-hexachlorocyclohexa-2,4-dien-1-one, THF; ii. NaBH<sub>4</sub>, EtOH; iii. KOH/EtOH/H<sub>2</sub>O; (b) i. (2<span class="html-italic">S</span>,5<span class="html-italic">R</span>)-2-(<span class="html-italic">tert</span>-butyl)-3,5-dimethylimidazolidin-4-one trifluoroacetate (20%), 2,3,4,5,6,6-hexachlorocyclohexa-2,4-dien-1-one, THF; ii. NaBH<sub>4</sub>, EtOH; iii. KOH/EtOH/H<sub>2</sub>O; (c) 1-undecyne, n-BuLi, BF<sub>3</sub>·OEt<sub>2</sub>, dry THF; (d) 10% Pd/BaSO<sub>4</sub>, H<sub>2</sub>, EtOAc; (e) AcCl, dry pyridine, dry CH<sub>2</sub>Cl<sub>2</sub>; (f) TBAF, dry THF; (g) Jones reagent, acetone; (h) LiOH·H<sub>2</sub>O, THF:H<sub>2</sub>O.</p>
Full article ">
12 pages, 3331 KiB  
Article
Synthesis of a Grease Thickener from Cashew Nut Shell Liquor
by Son A. Hoang, Khanh D. Pham, Nhung H. Nguyen, Ha T. Tran, Ngoc Hoang and Chi M. Phan
Molecules 2023, 28(22), 7624; https://doi.org/10.3390/molecules28227624 - 16 Nov 2023
Viewed by 1271
Abstract
Thickener, also known as a gelling agent, is a critical component of lubricating greases. The most critical property of thickener, temperature resistance, is determined by the molecular structure of the compounds. Currently, all high-temperature-resistant thickeners are based on 12-hydroxystearic acid, which is exclusively [...] Read more.
Thickener, also known as a gelling agent, is a critical component of lubricating greases. The most critical property of thickener, temperature resistance, is determined by the molecular structure of the compounds. Currently, all high-temperature-resistant thickeners are based on 12-hydroxystearic acid, which is exclusively produced from castor oil. Since castor oil is also an important reagent for other processes, finding a sustainable alternative to 12-hydroxystearic acid has significant economic implications. This study synthesises an alternative thickener from abundant agricultural waste, cashew nut shell liquor (CNSL). The synthesis and separation procedure contains three steps: (i) forming and separating calcium anacardate by precipitation, (ii) forming and separating anacardic acid (iii) forming lithium anacardate. The obtained lithium anacardate can be used as a thickener for lubricating grease. It was found that the recovery of anacardic acid was around 80%. The optimal reaction temperature and time conditions for lithium anacardate were 100 °C and 1 h, respectively. The method provides an economical alternative to castor and other vegetable oils. The procedure presents a simple pathway to produce the precursor for the lubricating grease from agricultural waste. The first reaction step can be combined with the existing distillation of cashew nut shell processing. An effective application can promote CNSL to a sustainable feedstock for green chemistry. The process can also be combined with recycled lithium from the spent batteries to improve the sustainability of the battery industry. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Structure of the natural-based fatty acids. Natural anacardic acids include four compounds with different saturation levels in the hydrocarbon chain.</p>
Full article ">Figure 2
<p>Effect of temperature on anacardic acid recovery efficiency.</p>
Full article ">Figure 3
<p>Effect of reaction time on anacardic acid recovery efficiency.</p>
Full article ">Figure 4
<p>FT-IR spectrum of anacardic acid.</p>
Full article ">Figure 5
<p><sup>13</sup>C NMR spectrum of anacardic acid.</p>
Full article ">Figure 6
<p><sup>1</sup>H NMR spectrum of anacardic acid.</p>
Full article ">Figure 7
<p>FTIR of the product from LiOH-anacardic acid reactions at different temperatures.</p>
Full article ">Figure 8
<p>FTIR of the product from LiOH-anacardic acid reactions at different Li:anacardic molar ratios 0.5:1; 1:1; 1.5:1).</p>
Full article ">Figure 9
<p>FTIR of the product from LiOH-anacardic acid reactions at 100 °C, lithium anacardate molar ratio 1:1 and different reaction times (0.5 h; 1 h; 1.5 h).</p>
Full article ">Figure 10
<p>Grease obtained from lithium anacardate and paraffin oil.</p>
Full article ">
13 pages, 2641 KiB  
Article
The Metabolomic Characteristics and Dysregulation of Fatty Acid Esters of Hydroxy Fatty Acids in Breast Cancer
by Linlin Qin, Na An, Bifeng Yuan, Quanfei Zhu and Yuqi Feng
Metabolites 2023, 13(11), 1108; https://doi.org/10.3390/metabo13111108 - 24 Oct 2023
Cited by 1 | Viewed by 1542
Abstract
Lipid reprogramming metabolism is crucial for supporting tumor growth in breast cancer and investigating potential tumor biomarkers. Fatty acid esters of hydroxy fatty acids (FAHFAs) are a class of endogenous lipid metabolites with anti-diabetic and anti-inflammatory properties that have been discovered in recent [...] Read more.
Lipid reprogramming metabolism is crucial for supporting tumor growth in breast cancer and investigating potential tumor biomarkers. Fatty acid esters of hydroxy fatty acids (FAHFAs) are a class of endogenous lipid metabolites with anti-diabetic and anti-inflammatory properties that have been discovered in recent years. Our previous targeted analysis of sera from breast cancer patients revealed a significant down-regulation of several FAHFAs. In this study, we aimed to further explore the relationship between FAHFAs and breast cancer by employing chemical isotope labeling combined with liquid chromatography−mass spectrometry (CIL-LC-MS) for profiling of FAHFAs in tumors and adjacent normal tissues from breast cancer patients. Statistical analysis identified 13 altered isomers in breast cancer. These isomers showed the potential to distinguish breast cancer tissues with an area under the curve (AUC) value above 0.9 in a multivariate receiver operating curve model. Furthermore, the observation of up-regulated 9-oleic acid ester of hydroxy stearic acid (9-OAHSA) and down-regulated 9-hydroxystearic acid (9-HSA) in tumors suggests that breast cancer shares similarities with colorectal cancer, and their potential mechanism is to attenuate the effects of pro-apoptotic 9-HSA by enhancing the synthesis of FAHFAs, thereby promoting tumor survival and progression through this buffering system. Full article
(This article belongs to the Section Endocrinology and Clinical Metabolic Research)
Show Figures

Figure 1

Figure 1
<p>Schematic flow diagram of CIL-LC-MS screening of FAHFAs in breast tumor and adjacent normal tissues.</p>
Full article ">Figure 2
<p>Identification of OAHSA isomers using the retention index ester bond position rule. The upper panel shows the extracted ion chromatograms of OAHSA in (<b>A</b>) tumor tissues, (<b>B</b>) adjacent normal tissues, and (<b>C</b>) standards (Peak 2, 13−OAHSA; Peak 3, 12−OAHSA; Peak 4, 10−OAHSA; Peak 5, 9−OAHSA; and Peak 6, 5−OAHSA), respectively. Regression curves of log<sub>10</sub>(RI) (RI, retention index) versus the ester bond positions of OAHSA (<b>D</b>) were constructed with OAHSA standards (dark blue dots), and <b>1</b> (red dot) represents the predicted 11-OAHSA.</p>
Full article ">Figure 3
<p>Statistical analysis of screening FAHFA isomers for breast cancer characteristics. (<b>A</b>) OPLS−DA score plots of FAHFAs show a clear separation between tumor tissue and normal tissue adjacent to tumor tissues. (<b>B</b>) Volcano plots display that 6 isomers are up−regulated and 7 isomers are down−regulated in tumor tissues, and fold change was calculated by the ratio of relative abundance of FAHFAs in tumor versus adjacent normal tissues. (<b>C</b>) Venn diagram shows that FAHFAs characteristic of breast cancer were obtained by screening for thresholds that were met both in OPLS−DA analysis (VIP &gt; 1) and volcano plot analysis (<span class="html-italic">p</span> &lt; 0.05, FC &gt; 1.5). (<b>D</b>) Hierarchical clustering heatmaps show the relative abundance of 13 significantly differentiated FAHFAs in breast tissues, visualizing characteristic alteration of FAHFAs in tumor tissues and contributing to a precise classification of tissues.</p>
Full article ">Figure 4
<p>The multivariate ROC models were constructed using 13 differentiated FAHFAs, using three different machine learning methods, Random Forest (RF), Support Vector Machine (SVM), and Partial Least Squares Discriminant Analysis (PLS-DA), respectively.</p>
Full article ">Figure 5
<p>Violin scatter plots of characteristic FAHFAs and their related HFAs that can be used to diagnose breast cancer. *, <span class="html-italic">p</span> &lt; 0.05; ****, <span class="html-italic">p</span> &lt; 0.0001; ns, not significant.</p>
Full article ">
20 pages, 3433 KiB  
Article
Aqueous Binary Mixtures of Stearic Acid and Its Hydroxylated Counterpart 12-Hydroxystearic Acid: Fine Tuning of the Lamellar/Micelle Threshold Temperature Transition and of the Micelle Shape
by Maëva Almeida, Daniel Dudzinski, Bastien Rousseau, Catherine Amiel, Sylvain Prévost, Fabrice Cousin and Clémence Le Coeur
Molecules 2023, 28(17), 6317; https://doi.org/10.3390/molecules28176317 - 29 Aug 2023
Cited by 3 | Viewed by 1116
Abstract
This study examines the structures of soft surfactant-based biomaterials which can be tuned by temperature. More precisely, investigated here is the behavior of stearic acid (SA) and 12-hydroxystearic acid (12-HSA) aqueous mixtures as a function of temperature and the 12-HSA/SA molar ratio (R). [...] Read more.
This study examines the structures of soft surfactant-based biomaterials which can be tuned by temperature. More precisely, investigated here is the behavior of stearic acid (SA) and 12-hydroxystearic acid (12-HSA) aqueous mixtures as a function of temperature and the 12-HSA/SA molar ratio (R). Whatever R is, the system exhibits a morphological transition at a given threshold temperature, from multilamellar self-assemblies at low temperature to small micelles at high temperature, as shown by a combination of transmittance measurements, Wide Angle X-ray diffraction (WAXS), small angle neutron scattering (SANS), and differential scanning calorimetry (DSC) experiments. The precise determination of the threshold temperature, which ranges between 20 °C and 50 °C depending on R, allows for the construction of the whole phase diagram of the system as a function of R. At high temperature, the micelles that are formed are oblate for pure SA solutions (R = 0) and prolate for pure 12-HSA solutions (R = 1). In the case of mixtures, there is a progressive continuous transition from oblate to prolate shapes when increasing R, with micelles that are almost purely spherical for R = 0.33. Full article
(This article belongs to the Special Issue Responsive Soft Materials Based on Biomolecules)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>(<b>A</b>) Transmittance as a function of temperature for HSA/SA mixtures from R = 0 to R = 1 upon cooling at a cooling rate of 0.2 °C/min. (<b>B</b>) Temperatures of transition determined by turbidimetry and by DSC as a function of ratio R, where error bars are estimated from the width of the peak.</p>
Full article ">Figure 2
<p>Enthalpograms obtained for 2 wt% mixture in fatty acid at various R ratios upon cooling at a cooling rate of 1 °C/min. Data are shifted for clarity.</p>
Full article ">Figure 3
<p>WAXS diffractograms for the different HSA/SA samples ratio, from pure SA (R = 0) to pure HSA (R = 1), at three different temperatures (20 °C from reference [<a href="#B37-molecules-28-06317" class="html-bibr">37</a>], 36 °C, and 50 °C). The spectra were successively shifted in intensity by an offset of 0.0033 cm<sup>−1</sup> for clarity. Data at 20 °C with permission from reference [<a href="#B37-molecules-28-06317" class="html-bibr">37</a>].</p>
Full article ">Figure 4
<p>SANS intensity profiles at four temperatures ((<b>A</b>): 30 °C, (<b>B</b>): 37 °C, (<b>C</b>): 45 °C, and (<b>D</b>): 60 °C) for the different samples in D<sub>2</sub>O for different HSA/SA ratios, from pure SA (R = 0) to pure HSA (R = 1). The spectra are successively shifted by a factor of 10 in intensity for clarity (data for R = 0 in absolute scale). The black and red continuous lines correspond to the best fit of the data either by a lamellar model or by a model of an elliptical micelles in interactions (see description in <a href="#app1-molecules-28-06317" class="html-app">Supplementary Materials</a>). Some data showing a combination of scattering of lamellar objects and micelles are not fitted (see explanation in text).</p>
Full article ">Figure 5
<p>Value of lamellar spacing. (<b>A</b>) Caillé Parameter (<b>B</b>) and thickness (<b>C</b>) obtained from fitting of lamellar phase at different temperature. Data at 20 °C are already published in [<a href="#B37-molecules-28-06317" class="html-bibr">37</a>] and adapted with permission from Ref. [<a href="#B37-molecules-28-06317" class="html-bibr">37</a>] but are represented for the sake of comparison.</p>
Full article ">Figure 6
<p>Evolution (<b>A</b>) of polar and equatorial radiuses obtained from micelles’ fit at 45 and 60 °C, (<b>B</b>) of the ellipticity (<b>C</b>) of the number of aggregations <span class="html-italic">N<sub>volume</sub></span> and <span class="html-italic">N<sub>weight</sub></span> as a function of the ratio R. (<b>D</b>) <span class="html-italic">S</span>(<span class="html-italic">q</span>)<sub>(q→0)</sub> obtained from division on the scattered intensity by the fitted form factor. (<b>E</b>) Schematic representation of the evolution of micelles’ shape as a function of the ratio R from prolate micelles to oblate ones. The SA molecules are represented in orange and the 12-HSA molecules in red.</p>
Full article ">Figure 7
<p>Schemes of the structures of the self-assembled aggregates as a function of R from pure SA (R = 0) to pure 12-HSA (R = 1). The SA molecules are represented in orange and the 12-HSA molecules in red. The total amount of fatty acid is constant and fixed at 2 wt%.</p>
Full article ">
24 pages, 3977 KiB  
Article
Aqueous Binary Mixtures of Stearic Acid and Its Hydroxylated Counterpart 12-Hydroxystearic Acid: Cascade of Morphological Transitions at Room Temperature
by Maëva Almeida, Daniel Dudzinski, Catherine Amiel, Jean-Michel Guigner, Sylvain Prévost, Clémence Le Coeur and Fabrice Cousin
Molecules 2023, 28(11), 4336; https://doi.org/10.3390/molecules28114336 - 25 May 2023
Cited by 4 | Viewed by 2209
Abstract
Here, we describe the behavior of mixtures of stearic acid (SA) and its hydroxylated counterpart 12-hydroxystearic acid (12-HSA) in aqueous mixtures at room temperature as a function of the 12-HSA/SA mole ratio R. The morphologies of the self-assembled aggregates are obtained through a [...] Read more.
Here, we describe the behavior of mixtures of stearic acid (SA) and its hydroxylated counterpart 12-hydroxystearic acid (12-HSA) in aqueous mixtures at room temperature as a function of the 12-HSA/SA mole ratio R. The morphologies of the self-assembled aggregates are obtained through a multi-structural approach that combines confocal and cryo-TEM microscopies with small-angle neutron scattering (SANS) and wide-angle X-ray scattering (WAXS) measurements, coupled with rheology measurements. Fatty acids are solubilized by an excess of ethanolamine counterions, so that their heads are negatively charged. A clear trend towards partitioning between the two types of fatty acids is observed, presumably driven by the favorable formation of a H-bond network between hydroxyl OH function on the 12th carbon. For all R, the self-assembled structures are locally lamellar, with bilayers composed of crystallized and strongly interdigitated fatty acids. At high R, multilamellar tubes are formed. The doping via a low amount of SA molecules slightly modifies the dimensions of the tubes and decreases the bilayer rigidity. The solutions have a gel-like behavior. At intermediate R, tubes coexist in solution with helical ribbons. At low R, local partitioning also occurs, and the architecture of the self-assemblies associates the two morphologies of the pure fatty acids systems: they are faceted objects with planar domains enriched in SA molecules, capped with curved domains enriched in 12-HSA molecules. The rigidity of the bilayers is strongly increased, as well their storage modulus. The solutions remain, however, viscous fluids in this regime. Full article
(This article belongs to the Special Issue Responsive Soft Materials Based on Biomolecules)
Show Figures

Figure 1

Figure 1
<p>Representation of (<b>A</b>) 12-hydroxy stearic acid (12-HAS) molecule and (<b>B</b>) stearic acid (SA) molecules.</p>
Full article ">Figure 2
<p>Photographs of the samples at all ratios R examined in the paper. The inset at R = 0.4 shows the heterogeneous macroscopic aspect.</p>
Full article ">Figure 3
<p>Confocal microscope images with an ×100 magnification obtained for different 12-has/SA samples at 2 wt% in the presence of Nile Red with different ratios to pure SA (R = 0) in figure (<b>A</b>) to pure 12-HSA (R = 1) in figure (<b>I</b>). Figures (<b>B</b>–<b>H</b>) correspond, respectively, to samples with R ratios equal to 0.2, 0.3, 0.4, 0.5, 0.6, 0.75 and 0.9.</p>
Full article ">Figure 4
<p>Cryo-TEM images obtained for different samples ratios from pure 12-HSA (R = 1, images <b>A</b>–<b>C</b>) to pure SA (R = 0, images <b>L</b>–<b>N</b>) and intermediate surfactant concentrations R = 0.4 (images <b>D</b>–<b>G</b>) and R = 0.1 (images <b>H</b>–<b>K</b>).</p>
Full article ">Figure 5
<p>Right: (<b>A</b>) SANS intensity profiles at 20 °C for the different samples in D<sub>2</sub>O for all 12-HSA/SA ratios, from pure SA (R = 0) to pure 12-HSA (R = 1). The spectra are successively shifted by a factor of 10 in intensity for clarity (data for R = 0 on an absolute scale). The green lines correspond to the characteristic decays in the low q region. The black and red continuous lines correspond to the best fit of the data (see description of the model in <a href="#app1-molecules-28-04336" class="html-app">Supplementary Materials</a>). The dotted part of the fitted model for R = 1, R = 0.9 and R = 0.75 correspond to the q-range where there is multiple scattering. The sample at R = 0.4 is not fitted as it is macroscopically heterogeneous (see main text). Left (<b>B</b>–<b>D</b>): Respectively, the Caillé parameters, interlamellar distances and lamella thicknesses as functions of R, obtained via SANS data fitting.</p>
Full article ">Figure 6
<p>WAXS diffractograms for the different 12-HSA/SA sample ratios, from pure SA (R = 0) to pure 12-HSA (R = 1), as well as a pure 12-HSA sample with a fatty acid/ethanolamine ratio of r = 0.5 for comparison. For clarity, the spectra were shifted in intensity from each other by 0.0025 cm<sup>−1</sup>.</p>
Full article ">Figure 7
<p>(<b>A</b>) Plateau value of storage G<sub>0</sub>′ and loss G<sub>0</sub>″ modulus measured at 1 Hz for the different sample ratios from pure SA (R = 0) to pure 12-HSA (R = 1) at 20 °C (loss factor in inset). (<b>B</b>) Shear stress as a function of the strain amplitude at a constant frequency of 1 Hz for selected samples (R = 0.4; 0.6 and 1). Asymptotic curves (dashed lines) depict the elastic regime followed by the solution yields. The associated yield stress is shown in the inset for R ranging from 0.4 to 1).</p>
Full article ">Figure 8
<p>Schemes of the structures of the self-assembled aggregates as a function of R from pure SA (R = 0) to pure 12-HSA (R = 1) at 20 °C. The SA molecules are shown in orange and the 12-HSA molecules in red. The scale is not kept constant from one structure to another. The dashed part of diagram corresponds to the range of R where several structures coexist in the solution.</p>
Full article ">
14 pages, 2198 KiB  
Article
Ultrastable and Responsive Foams Based on 10-Hydroxystearic Acid Soap for Spore Decontamination
by Carolina Dari, Fabrice Cousin, Clemence Le Coeur, Thomas Dubois, Thierry Benezech, Arnaud Saint-Jalmes and Anne-Laure Fameau
Molecules 2023, 28(11), 4295; https://doi.org/10.3390/molecules28114295 - 24 May 2023
Cited by 3 | Viewed by 1644
Abstract
Currently, there is renewed interest in using fatty acid soaps as surfactants. Hydroxylated fatty acids are specific fatty acids with a hydroxyl group in the alkyl chain, giving rise to chirality and specific surfactant properties. The most famous hydroxylated fatty acid is 12-hydroxystearic [...] Read more.
Currently, there is renewed interest in using fatty acid soaps as surfactants. Hydroxylated fatty acids are specific fatty acids with a hydroxyl group in the alkyl chain, giving rise to chirality and specific surfactant properties. The most famous hydroxylated fatty acid is 12-hydroxystearic acid (12-HSA), which is widely used in industry and comes from castor oil. A very similar and new hydroxylated fatty acid, 10-hydroxystearic acid (10-HSA), can be easily obtained from oleic acid by using microorganisms. Here, we studied for the first time the self-assembly and foaming properties of R-10-HSA soap in an aqueous solution. A multiscale approach was used by combining microscopy techniques, small-angle neutron scattering, wide-angle X-ray scattering, rheology experiments, and surface tension measurements as a function of temperature. The behavior of R-10-HSA was systematically compared with that of 12-HSA soap. Although multilamellar micron-sized tubes were observed for both R-10-HSA and 12-HSA, the structure of the self-assemblies at the nanoscale was different, which is probably due to the fact that the 12-HSA solutions were racemic mixtures, while the 10-HSA solutions were obtained from a pure R enantiomer. We also demonstrated that stable foams based on R-10-HSA soap can be used for cleaning applications, by studying spore removal on model surfaces in static conditions via foam imbibition. Full article
(This article belongs to the Special Issue Responsive Soft Materials Based on Biomolecules)
Show Figures

Figure 1

Figure 1
<p>The surface tension of aqueous solutions of 10-HSA/MEA as a function of concentration at T = 20 ± 1 °C.</p>
Full article ">Figure 2
<p>(<b>a</b>) Phase contrast microscopy picture of 10-HSA at 25 °C. (<b>b</b>) SANS intensity profile of 10-HSA at 25 °C. The black line corresponds to the best fit of the data described in the <a href="#app1-molecules-28-04295" class="html-app">Supplementary Materials</a>. (<b>c</b>) Oscillatory measurements of 10-HSA, elastic <span class="html-italic">G</span>′ (●), and viscous <span class="html-italic">G</span>″ (○) moduli plotted as a function of the strain amplitude, where <span class="html-italic">γ</span> is a constant <span class="html-italic">f</span> = 1 Hz at 25 °C.</p>
Full article ">Figure 3
<p>(<b>a</b>) Phase contrast microscopy picture of 10-HSA at 65 °C. (<b>b</b>) Variations of elastic G′ (●) and viscous G″ (○) moduli upon heating. The moduli were measured at γ = 0.1% and <span class="html-italic">f</span> = 1 Hz. (<b>c</b>) SANS intensity profile of 10-HSA at 75 °C. The line corresponds to the best fit of the spectrum described in the <a href="#app1-molecules-28-04295" class="html-app">Supplementary Materials</a>.</p>
Full article ">Figure 4
<p>(<b>a</b>) Evolution of the foam volume (purple) and of the liquid volume in the foam (green) over time at 25 °C. (<b>b</b>) Evolution of the foam stability by measuring the evolution of foam volume over time at different temperatures: at 25 °C when multilamellar tubes were present and at 65 °C when spherical micelles were present.</p>
Full article ">Figure 5
<p>(<b>a</b>) Spore log reduction on the model surface after 30 min for water, 10-HSA dispersion, hand-shaken (HS) foam, and double-syringe (DS) foam. The small letters, a and b, indicate groups of statistical differences according to Tukey’s test (<span class="html-italic">p</span> &lt; 0.05). (<b>b</b>) Epifluorescence microscopy pictures of the model surface contaminated by fluorescent spores before cleaning, and (<b>c</b>) after 30 min of cleaning with DS foam. The scale bar represents 50 µm in all the pictures.</p>
Full article ">Figure 6
<p>Epifluorescence microscopy pictures of the penetration of the aqueous dispersion of spores into a double-syringe foam over time. Here, t = 0 corresponds to the image taken just after contact with the foam. The scale bar represents 1 mm in all the pictures.</p>
Full article ">Figure 7
<p>Schematic showing the methodology for spore removal on stainless-steel surfaces using foam. (1) Five drops (1 µL) of the spore suspension were placed on the surface of the model plate with a micropipette; (2) the plates were dried in an oven for 1 h at 30 °C; (3) the soiled plates were then placed in the tubes containing the foams and were kept in a horizontal position for 30 min; (4) each plate was sampled with a dry cotton swab; (5) the swab was put in a tube with 5 mL of sterile Mill-Q water and vortexed for 1 min at 2400 rpm; (6) serial dilutions were made in sterile Milli-Q water for each tube and were then placed in tryptic soy agar and incubated for 24 h at 30 °C; (7) the number of colony-forming units (CFU) was counted manually.</p>
Full article ">
14 pages, 3024 KiB  
Article
Correlation between Physical Properties of 12-Hydroxystearic Acid Organogels and Hansen Solubility Parameters
by Yuya Murakami, Taisei Uchiyama and Atsushi Shono
Gels 2023, 9(4), 314; https://doi.org/10.3390/gels9040314 - 7 Apr 2023
Cited by 1 | Viewed by 1722
Abstract
The Hansen solubility parameter (HSP) is a useful index for reasoning the gelation behavior of low-molecular-weight gelators (LMWGs). However, the conventional HSP-based methods only “classify” solvents that can and cannot form gels and require many trials to achieve this. For engineering purposes, quantitative [...] Read more.
The Hansen solubility parameter (HSP) is a useful index for reasoning the gelation behavior of low-molecular-weight gelators (LMWGs). However, the conventional HSP-based methods only “classify” solvents that can and cannot form gels and require many trials to achieve this. For engineering purposes, quantitative estimation of gel properties using the HSP is highly desired. In this study, we measured critical gelation concentrations based on three distinct definitions, mechanical strength, and light transmittance of organogels prepared with 12-hydroxystearic acid (12HSA) and correlated them with the HSP of solvents. The results demonstrated that the mechanical strength, in particular, strongly correlated with the distance of 12HSA and solvent in the HSP space. Additionally, the results indicated that the constant volume-based concentration should be used when comparing the properties of organogels to a different solvent. These findings are helpful in efficiently determining the gelation sphere of new LMWGs in HSP space and contribute to designing organogels with tunable physical properties. Full article
(This article belongs to the Special Issue Advances in Polymer Rheology)
Show Figures

Figure 1

Figure 1
<p>The degree of gelation with a weight fraction of 12HSA for hexane (Al06), cyclohexene (Cy06), hexachloroacetone (XCa), and ethyl chloroformate (Xec).</p>
Full article ">Figure 2
<p>The result of oscillation stress sweep measurements of a 12HSA organogel prepared with styrene (<span class="html-italic">w</span> = 1.10%): (<b>a</b>) Storage and loss moduli with various stress; (<b>b</b>) Stress/strain curve.</p>
Full article ">Figure 3
<p>The obtained rheological property of 12HSA organogels prepared with hexane (Al06), octane (Al08), and nonane (Al09): (<b>a</b>) Yield stress (<math display="inline"><semantics> <mrow> <msub> <mi>τ</mi> <mi mathvariant="normal">y</mi> </msub> </mrow> </semantics></math>) with a weight fraction of 12HSA; (<b>b</b>) Yield strain (<math display="inline"><semantics> <mrow> <msub> <mi>γ</mi> <mi mathvariant="normal">y</mi> </msub> </mrow> </semantics></math>) with a weight fraction of 12HSA.</p>
Full article ">Figure 4
<p>The degree of gelation with a weight fraction of 12HSA for hexane (Al06), octane (Al08), and nonane (Al09).</p>
Full article ">Figure 5
<p>3D plots of solvent used for 12HSA [<a href="#B30-gels-09-00314" class="html-bibr">30</a>] organogel preparation with their <math display="inline"><semantics> <mrow> <msub> <mi>w</mi> <mrow> <mi>CGC</mi> </mrow> </msub> </mrow> </semantics></math>. Note that black plots indicate that 12HSA was insoluble in the solvent.</p>
Full article ">Figure 6
<p>Correlation between yield stress (<math display="inline"><semantics> <mrow> <msub> <mi>τ</mi> <mi mathvariant="normal">y</mi> </msub> </mrow> </semantics></math>) and distance in HSP space: (<b>a</b>) Fixed concentration at <math display="inline"><semantics> <mrow> <mi>w</mi> <mo>=</mo> <mn>1.0</mn> <mo>%</mo> </mrow> </semantics></math>; (<b>b</b>) Fixed concentration at <math display="inline"><semantics> <mrow> <mi>v</mi> <mo>=</mo> <mn>80</mn> <mtext> </mtext> <mrow> <mi>mmol</mi> <mo>/</mo> <mi mathvariant="normal">L</mi> <mo>-</mo> <mi>solvent</mi> </mrow> </mrow> </semantics></math>; (<b>c</b>) Fixed concentration at <math display="inline"><semantics> <mrow> <mi>x</mi> <mo>=</mo> <mn>2.8</mn> <mtext> </mtext> <mrow> <mi>mol</mi> <mo>%</mo> </mrow> </mrow> </semantics></math>.</p>
Full article ">Figure 7
<p>Correlation between light transmittance (<math display="inline"><semantics> <mrow> <msub> <mi>α</mi> <mrow> <mi>gel</mi> </mrow> </msub> </mrow> </semantics></math>) with the distance in HSP space: (<b>a</b>) Fixed concentration at <math display="inline"><semantics> <mrow> <mi>w</mi> <mo>=</mo> <mn>1.0</mn> <mo>%</mo> </mrow> </semantics></math>; (<b>b</b>) Fixed concentration at <math display="inline"><semantics> <mrow> <mi>v</mi> <mo>=</mo> <mn>80</mn> <mtext> </mtext> <mrow> <mi>mmol</mi> <mo>/</mo> <mi mathvariant="normal">L</mi> <mo>-</mo> <mi>solvent</mi> </mrow> </mrow> </semantics></math>; (<b>c</b>) Fixed concentration at <math display="inline"><semantics> <mrow> <mi>x</mi> <mo>=</mo> <mn>2.8</mn> <mtext> </mtext> <mrow> <mi>mol</mi> <mo>%</mo> </mrow> </mrow> </semantics></math>.</p>
Full article ">
12 pages, 3126 KiB  
Communication
Targeted and Suspect Fatty Acid Profiling of Royal Jelly by Liquid Chromatography—High Resolution Mass Spectrometry
by Christiana Mantzourani and Maroula G. Kokotou
Biomolecules 2023, 13(3), 424; https://doi.org/10.3390/biom13030424 - 23 Feb 2023
Cited by 2 | Viewed by 1642
Abstract
Royal jelly (RJ) is a bee product produced by the mandibular and hypopharyngeal glands of worker honeybees which has attracted special attention because of its numerous pharmacological activities and its applications to dermatology and cosmetics. In 2020, we demonstrated a liquid chromatography–high resolution [...] Read more.
Royal jelly (RJ) is a bee product produced by the mandibular and hypopharyngeal glands of worker honeybees which has attracted special attention because of its numerous pharmacological activities and its applications to dermatology and cosmetics. In 2020, we demonstrated a liquid chromatography–high resolution mass spectrometry (LC–HRMS) method for the determination of seven medium-chain FFAs in RJ samples. The aim of the present work was to extend our studies on FA profiling of RJ, exploring the presence of common long-chain saturated, mono-unsaturated and poly-unsaturated free FAs in RJ samples using this LC–HRMS method. Among twenty common FAs studied by a targeted approach, palmitic acid, stearic acid and oleic acid were found at concentrations higher than the rest of the FAs (the concentrations of these three acids ranged from 37.4 to 48.0, from 17.7 to 24.0 and from 9.4 to 11.1 mg/100 g of fresh RJ, respectively). The high mass accuracy of LC–HRMS allowed the application of a suspect approach, which enabled the exploration of various C9 and C11 FAs, as well as hydroxylated C12 FAs. Nonenoic acid was indicated as the most abundant among these acids. In addition, for the first time, the presence of a variety of regio-isomers of hydroxymyristic, hydroxypalmitic and hydroxystearic acids was demonstrated in RJ samples. Full article
Show Figures

Figure 1

Figure 1
<p>Extracted ion chromatograms (EICs) of the analytes in a representative RJ sample.</p>
Full article ">Figure 2
<p>Relative average contents of RJ common long-chain FAs.</p>
Full article ">Figure 3
<p>EICs of nonanoic, nonenoic, hydroxynonanoic and hydroxynonenoic acids in a representative RJ sample.</p>
Full article ">Figure 4
<p>EICs of undecanoic, undecenoic, hydroxyundecanoic and hydroxyundecenoic acids in a representative RJ sample.</p>
Full article ">Figure 5
<p>EICs of hydroxydodecanoic, hydroxydodecenoic and dihydroxydodecanoic acids (<b>A</b>) and zoom (<b>B</b>) in a representative RJ sample.</p>
Full article ">Figure 6
<p>EICs of hydroxytetradecanoic acid (<b>A</b>) and zoom (<b>B</b>) in a representative RJ sample.</p>
Full article ">Figure 7
<p>EICs of hydroxypalmitic and hydroxystearic acids in a representative RJ sample.</p>
Full article ">
17 pages, 1893 KiB  
Article
Organogel of Acai Oil in Cosmetics: Microstructure, Stability, Rheology and Mechanical Properties
by Suellen Christtine da Costa Sanches, Maria Inês Ré, José Otávio Carréra Silva-Júnior and Roseane Maria Ribeiro-Costa
Gels 2023, 9(2), 150; https://doi.org/10.3390/gels9020150 - 10 Feb 2023
Cited by 8 | Viewed by 2266
Abstract
Organogel (OG) is a semi-solid material composed of gelling molecules organized in the presence of an appropriate organic solvent, through physical or chemical interactions, in a continuous net. This investigation aimed at preparing and characterizing an organogel from acai oil with hyaluronic acid [...] Read more.
Organogel (OG) is a semi-solid material composed of gelling molecules organized in the presence of an appropriate organic solvent, through physical or chemical interactions, in a continuous net. This investigation aimed at preparing and characterizing an organogel from acai oil with hyaluronic acid (HA) structured by 12-hydroxystearic acid (12-HSA), aiming at topical anti-aging application. Organogels containing or not containing HA were analyzed by Fourier-transform Infrared Spectroscopy, polarized light optical microscopy, thermal analysis, texture analysis, rheology, HA quantification and oxidative stability. The organogel containing hyaluronic acid (OG + HA) has a spherulitic texture morphology with a net-like structure and absorption bands that evidenced the presence of HA in the three-dimensional net of organogel. The thermal analysis confirmed the gelation and the insertion of HA, as well as a good thermal stability, which is also confirmed by the study of oxidative stability carried out under different temperature conditions for 90 days. The texture and rheology studies indicated a viscoelastic behavior. HA quantification shows the efficiency of the HA cross-linking process in the three-dimensional net of organogel with 11.22 µg/mL for cross-linked HA. Thus, it is concluded that OG + HA shows potentially promising physicochemical characteristics for the development of a cosmetic system. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>FTIR spectrum of 12HSA, OG, AH, and OG + HA submitted to a resolution of 2 cm<sup>−1</sup> and a range of 4000 to 400 cm<sup>−1</sup>.</p>
Full article ">Figure 2
<p>Microphotographs of OG (<b>A</b>) and OG + HA (<b>B</b>). Magnification at 50×.</p>
Full article ">Figure 3
<p>TG (<b>A</b>) DTG (<b>B</b>) curve of HA, 12HT, OG and OG + HA in N<sub>2</sub> atmosphere (50 mL/min) and 10 °C/min heating, in a temperature range from 25 to 600 °C.</p>
Full article ">Figure 4
<p>Differential exploratory calorimetry curve of OG and OG + HA in N<sub>2</sub> atmosphere (50 mL/min) and 10 °C/min heating, over a temperature range from 25 to 550 °C.</p>
Full article ">Figure 5
<p>Strength over time in a double compression test of OG (<b>A</b>) and OG + HA (<b>B</b>).</p>
Full article ">Figure 6
<p>Apparent viscosity of OG and OG + HA.</p>
Full article ">Figure 7
<p>Complex module of OG (<b>A</b>) and OG + HA (<b>B</b>).</p>
Full article ">
18 pages, 3631 KiB  
Article
Investigating the Synthesis and Characteristics of UV-Cured Bio-Based Epoxy Vegetable Oil-Lignin Composites Mediated by Structure-Directing Agents
by Brindusa Balanuca, Raluca Sanda Komartin, Madalina Ioana Necolau, Celina Maria Damian and Raluca Stan
Polymers 2023, 15(2), 439; https://doi.org/10.3390/polym15020439 - 13 Jan 2023
Cited by 4 | Viewed by 1511
Abstract
Bio-based composites were developed from the epoxy derivatives of Lallemantia iberica oil and kraft lignin (ELALO and EpLnK), using UV radiation as a low energy consumption tool for the oxiranes reaction. To avoid the filler sedimentation or its inhomogeneous distribution in the oil [...] Read more.
Bio-based composites were developed from the epoxy derivatives of Lallemantia iberica oil and kraft lignin (ELALO and EpLnK), using UV radiation as a low energy consumption tool for the oxiranes reaction. To avoid the filler sedimentation or its inhomogeneous distribution in the oil matrix, different structure-directing agents (SDA) were employed: 1,3:2,4-dibenzylidene-D-sorbitol (DBS), 12-hydroxystearic acid (HSA) and sorbitan monostearate (Span 60). The SDA and EpLnK effect upon the ELALO-based formulations, their curing reaction and the performance of the resulting materials were investigated. Fourier-transform Infrared Spectrometry (FTIR) indicates different modes of molecular arrangement through H bonds for the initial ELALO-SDA or ELALO-SDA-EpLnK systems, also confirming the epoxy group’s reaction through the cationic mechanism for the final composites. Gel fraction measurements validate the significant conversion of the epoxides for those materials containing SDAs or 1% EpLnK; an increased EpLnK amount (5%), with or without SDA addition, conduced to an inefficient polymerization process, with the UV radiation being partially absorbed by the filler. Thermo-gravimetric and dynamic-mechanical analyses (TGA and DMA) revealed good properties for the ELALO-based materials. By loading 1% EpLnK, the thermal stability was improved to with 10 °C (for Td3%) and the addition of each SDA differently influenced the Tg values but also gave differences in the glassy and rubbery states when the storage moduli were interrogated, depending on their chemical structures. Water affinity and morphological studies were also carried out. Full article
(This article belongs to the Section Polymer Composites and Nanocomposites)
Show Figures

Figure 1

Figure 1
<p>Chemical structure of the used SDA: (<b>a</b>) DBS, (<b>b</b>) Span 60, (<b>c</b>) HAS.</p>
Full article ">Figure 2
<p><sup>1</sup>H-NMR spectra of (<b>a</b>) crude LALO and (<b>b</b>) epoxidized LALO.</p>
Full article ">Figure 3
<p>FTIR spectra registered for (<b>a</b>) LALO and (<b>b</b>) ELALO.</p>
Full article ">Figure 4
<p>FTIR spectra of the initial (unpolymerized) ELALO and ELALO-SDA systems. Detail—cropped and zoomed spectral region from 1000–800 cm<sup>−1</sup>.</p>
Full article ">Figure 5
<p>FTIR spectra of the final F1–F4 materials (after UV treatment).</p>
Full article ">Figure 6
<p>FTIR spectra of the initial and final F7–F9 composites. (Before curing—not. B; after curing—not. A).</p>
Full article ">Figure 7
<p>DMA curves (Tan delta vs. temperature) recorder for selected ELALO-based materials. Influence of the SDA—left graph and influence of both additives—right graph.</p>
Full article ">Figure 8
<p>SEM images for different ELALO-based materials (ELALO (<b>F1</b>), ELALO-EpLnK1 (<b>F5</b>), ELALO-DBS-EpLnK1 (<b>F7</b>), ELALO-Span-EpLnK1 (<b>F8</b>), ELALO-HSA-EpLnK1 (<b>F9</b>); 2000× magnification).</p>
Full article ">
21 pages, 3905 KiB  
Article
Development and Characterization of Novel In-Situ-Forming Oleogels
by Anne Dümichen, Henrike Lucas, Marie-Luise Trutschel and Karsten Mäder
Pharmaceutics 2023, 15(1), 254; https://doi.org/10.3390/pharmaceutics15010254 - 11 Jan 2023
Cited by 4 | Viewed by 2435
Abstract
PLGA-based in situ forming implants (ISFI) often require a high amount of potentially toxic solvents such as N methyl-Pyrrolidone (NMP). The aim of the present study was to develop lipid in-situ-forming oleogels (ISFOs) as alternative delivery systems. 12-Hydroxystearic acid (12-HSA) was selected as [...] Read more.
PLGA-based in situ forming implants (ISFI) often require a high amount of potentially toxic solvents such as N methyl-Pyrrolidone (NMP). The aim of the present study was to develop lipid in-situ-forming oleogels (ISFOs) as alternative delivery systems. 12-Hydroxystearic acid (12-HSA) was selected as the oleogelling agent and three different oleoformulations were investigated: (a) 12-HSA, peanut oil (PO), NMP; (b) 12-HSA, medium-chain triglycerides (MCT), ethanol; (c) 12-HSA, isopropyl myristate (IPM), ethanol. The effects of the 12-HSA concentration, preparation method, and composition on the mechanical stability were examined using a texture analysis and oscillating rheology. The texture analysis was used to obtain information on the compression strength. The amplitude sweeps were analyzed to provide information on the gel strength and the risk of brittle fractures. The frequency sweeps allowed insights into the long-term stability and risk of syneresis. The syringeability of the ISFOs was tested, along with their acute and long-term cytotoxicity in vitro. The developed ISFOs have the following advantages: (1) the avoidance of highly acidic degradation products; (2) low amounts of organic solvents required; (3) low toxicity; (4) low injection forces, even with small needle sizes. Therefore, ISFOs are promising alternatives to the existing polymer/NMP-based ISFIs. Full article
(This article belongs to the Special Issue Advanced Pharmaceutical Science and Technology in Germany)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>The compression force required to create a 1 mm indentation. The force required to indent the gel wafer by 1 mm was extracted from the force–displacement-curves. The data are categorized by the manufacturing process and 12-HSA concentration. The compositions of the gels are coded by the liquid lipid used (PO <span class="html-fig-inline" id="pharmaceutics-15-00254-i001"><img alt="Pharmaceutics 15 00254 i001" src="/pharmaceutics/pharmaceutics-15-00254/article_deploy/html/images/pharmaceutics-15-00254-i001.png"/></span>, MCT <span class="html-fig-inline" id="pharmaceutics-15-00254-i002"><img alt="Pharmaceutics 15 00254 i002" src="/pharmaceutics/pharmaceutics-15-00254/article_deploy/html/images/pharmaceutics-15-00254-i002.png"/></span>, IPM <span class="html-fig-inline" id="pharmaceutics-15-00254-i003"><img alt="Pharmaceutics 15 00254 i003" src="/pharmaceutics/pharmaceutics-15-00254/article_deploy/html/images/pharmaceutics-15-00254-i003.png"/></span>).</p>
Full article ">Figure 2
<p><b>Left</b>: Elastic (<span class="html-italic">G</span>′: <span class="html-fig-inline" id="pharmaceutics-15-00254-i004"><img alt="Pharmaceutics 15 00254 i004" src="/pharmaceutics/pharmaceutics-15-00254/article_deploy/html/images/pharmaceutics-15-00254-i004.png"/></span>) and viscous component values (<span class="html-italic">G</span>″: <span class="html-fig-inline" id="pharmaceutics-15-00254-i005"><img alt="Pharmaceutics 15 00254 i005" src="/pharmaceutics/pharmaceutics-15-00254/article_deploy/html/images/pharmaceutics-15-00254-i005.png"/></span>). <b>Right</b>: Ratios of the flow point (<math display="inline"><semantics> <mrow> <msub> <mi>τ</mi> <mi>F</mi> </msub> </mrow> </semantics></math>) and the yield point (<math display="inline"><semantics> <mrow> <msub> <mi>τ</mi> <mi>γ</mi> </msub> </mrow> </semantics></math> ). From amplitude sweeps acquired via oscillating rheometry, the shear moduli of <span class="html-italic">G</span>′ and <span class="html-italic">G</span>″ within the LVER and flow (<math display="inline"><semantics> <mrow> <msub> <mi>τ</mi> <mi>F</mi> </msub> </mrow> </semantics></math> ) and yield point (<math display="inline"><semantics> <mrow> <msub> <mi>τ</mi> <mi>γ</mi> </msub> </mrow> </semantics></math> ) were extracted. The data are categorized by the manufacturing technique and 12-HSA concentration. The compositions of the gels are coded based on the liquid lipid (PO <span class="html-fig-inline" id="pharmaceutics-15-00254-i001"><img alt="Pharmaceutics 15 00254 i001" src="/pharmaceutics/pharmaceutics-15-00254/article_deploy/html/images/pharmaceutics-15-00254-i001.png"/></span>, MCT <span class="html-fig-inline" id="pharmaceutics-15-00254-i002"><img alt="Pharmaceutics 15 00254 i002" src="/pharmaceutics/pharmaceutics-15-00254/article_deploy/html/images/pharmaceutics-15-00254-i002.png"/></span>, IPM <span class="html-fig-inline" id="pharmaceutics-15-00254-i003"><img alt="Pharmaceutics 15 00254 i003" src="/pharmaceutics/pharmaceutics-15-00254/article_deploy/html/images/pharmaceutics-15-00254-i003.png"/></span>).</p>
Full article ">Figure 3
<p><b>Left</b>: Linearity limits. <b>Right</b>: Tangent of the phase angle (<span class="html-italic">tanδ</span>). The respective linearity limits and <span class="html-italic">tanδ</span> values within the linear range were extracted from the frequency sweeps obtained using oscillatory rheometry. The data are categorized according to the manufacturing technique and 12-HSA concentration. The compositions of the gels are coded based on the liquid lipid (PO <span class="html-fig-inline" id="pharmaceutics-15-00254-i001"><img alt="Pharmaceutics 15 00254 i001" src="/pharmaceutics/pharmaceutics-15-00254/article_deploy/html/images/pharmaceutics-15-00254-i001.png"/></span>, MCT <span class="html-fig-inline" id="pharmaceutics-15-00254-i002"><img alt="Pharmaceutics 15 00254 i002" src="/pharmaceutics/pharmaceutics-15-00254/article_deploy/html/images/pharmaceutics-15-00254-i002.png"/></span>, IPM <span class="html-fig-inline" id="pharmaceutics-15-00254-i003"><img alt="Pharmaceutics 15 00254 i003" src="/pharmaceutics/pharmaceutics-15-00254/article_deploy/html/images/pharmaceutics-15-00254-i003.png"/></span>).</p>
Full article ">Figure 4
<p>The syringeability of the different ISFOs with a 26 G cannula. The maximum force required to compress the syringe and eject the sample was extracted from the force–displacement curves obtained via the texture analysis. The data are presented as arithmetic means ± the standard deviation (SD); <span class="html-italic">n</span> = 3. The data are categorized by manufacturing method and 12-HSA concentration. The compositions of the gels are coded based on the liquid lipid used (PO <span class="html-fig-inline" id="pharmaceutics-15-00254-i001"><img alt="Pharmaceutics 15 00254 i001" src="/pharmaceutics/pharmaceutics-15-00254/article_deploy/html/images/pharmaceutics-15-00254-i001.png"/></span>, MCT <span class="html-fig-inline" id="pharmaceutics-15-00254-i002"><img alt="Pharmaceutics 15 00254 i002" src="/pharmaceutics/pharmaceutics-15-00254/article_deploy/html/images/pharmaceutics-15-00254-i002.png"/></span>, IPM <span class="html-fig-inline" id="pharmaceutics-15-00254-i003"><img alt="Pharmaceutics 15 00254 i003" src="/pharmaceutics/pharmaceutics-15-00254/article_deploy/html/images/pharmaceutics-15-00254-i003.png"/></span>). The pure liquid lipid and solvents were also ejected for comparison.</p>
Full article ">Figure 5
<p>Acute toxicity–cell viability levels after exposure to solvents: (<b>a</b>) 3T3 cells after 24 h; (<b>b</b>) NHDF cells after 24 h; (<b>c</b>) 3T3 cells after 96 h; (<b>d</b>) NHDF cells after 96 h. Positive and negative controls for each assay are provided in the gray sections to demonstrate the performance of the experimental design. The cells were treated with solutions of NMP (PO ISFOs) <span class="html-fig-inline" id="pharmaceutics-15-00254-i006"><img alt="Pharmaceutics 15 00254 i006" src="/pharmaceutics/pharmaceutics-15-00254/article_deploy/html/images/pharmaceutics-15-00254-i006.png"/></span> and ethanol (MCT and IPM ISFOs) <span class="html-fig-inline" id="pharmaceutics-15-00254-i007"><img alt="Pharmaceutics 15 00254 i007" src="/pharmaceutics/pharmaceutics-15-00254/article_deploy/html/images/pharmaceutics-15-00254-i007.png"/></span>. The viability values are given as a function of the solvent concentration and the corresponding 12-HSA concentration, and as arithmetic mean ± SD; <span class="html-italic">n</span> = 3. Since no higher 12-HSA concentrations than 15% were tested, the amount of ISFO injected into the extraction medium was increased for the last 4 datapoints (250, 300, 350, and 400 µL into 800 µL of medium; for all other datapoints 200 µL was injected into 800 µL).</p>
Full article ">Figure 6
<p>Long-term toxicity–cell viability levels after exposure to ISFO extracts: (<b>a</b>) 3T3 cells after 24 h; (<b>b</b>) NHDF cells after 24 h; (<b>c</b>) 3T3 cells after 96 h of exposure; (<b>d</b>) NHDF cells after 96 h. Positive and negative controls for each assay are provided in the gray sections to demonstrate the performance of the experimental design. The cells were treated with PO ISFO extracts (<span class="html-fig-inline" id="pharmaceutics-15-00254-i008"><img alt="Pharmaceutics 15 00254 i008" src="/pharmaceutics/pharmaceutics-15-00254/article_deploy/html/images/pharmaceutics-15-00254-i008.png"/></span>), MCT-ISFO extracts (<span class="html-fig-inline" id="pharmaceutics-15-00254-i009"><img alt="Pharmaceutics 15 00254 i009" src="/pharmaceutics/pharmaceutics-15-00254/article_deploy/html/images/pharmaceutics-15-00254-i009.png"/></span>), and IPM ISFO extracts (<span class="html-fig-inline" id="pharmaceutics-15-00254-i010"><img alt="Pharmaceutics 15 00254 i010" src="/pharmaceutics/pharmaceutics-15-00254/article_deploy/html/images/pharmaceutics-15-00254-i010.png"/></span>). The viability values are given as a function of the 12-HSA concentrations as arithmetic means ± SD; <span class="html-italic">n</span> = 3. Since no higher 12-HSA concentrations than 15% were tested, the amount of ISFO injected into the extraction medium was increased for the last 4 datapoints (250, 300, 350, and 400 µL into 800 µL of medium; for all other datapoints 200 µL was injected into 800 µL).</p>
Full article ">Figure A1
<p>The setup for the gel wafer preparation. Upside down glass funnels were covered with dialysis membranes, filled with 2.5 g of sample, and solidified in a desiccator used as a water bath. The water bath was filled with PBS at pH 7.4 with at least 10x the volume of the gel being solidified. The buffer was heated to 37 °C. The gelling process took 48 h and the buffer was exchanged after 24 h.</p>
Full article ">Figure A2
<p>The cell toxicity testing setup. On a 96 well plate, each column is dedicated to a different test solution. Column 1 is filled with pure cell medium as a blank, column 2 contains cells in medium as a negative control, and column 3 contains cells in medium stressed by a 0.05% Triton<sup>TM</sup> X 100 solution. Columns 4–8 contain cells in medium treated with test solutions corresponding to increasing 12-HSA-concentrations (3–15%). Columns 9–12 are treated with a corresponding 12-HSA-concentration to column 8 but at higher <span class="html-italic">v</span>/<span class="html-italic">v</span> ratios.</p>
Full article ">Figure A3
<p>A comparison of all force–distance curves acquired via the texture analysis of the gel wafers. All force–distance curves are categorized based on their manufacturing methods in the columns and their lipid components in the rows. All triplicate measurements are shown in the same color.</p>
Full article ">Figure A4
<p>An illustration of an analysis of an amplitude sweep. As an example, the strain-controlled amplitude sweep of a 10-IPM-organogel is shown, measured at 1 Hz within a shear range of 0.001% to 1% at 37 °C. The analyzed parameters <math display="inline"><semantics> <mrow> <msub> <mi>γ</mi> <mi>L</mi> </msub> <mo>,</mo> <mo> </mo> <msub> <mi>τ</mi> <mi>Y</mi> </msub> <mo>,</mo> <mo> </mo> <msub> <mi>τ</mi> <mi>F</mi> </msub> </mrow> </semantics></math>, and LVER are marked. The end of the LVER is called the linearity limit. It is determined by <span class="html-italic">G</span>′, the elastic component of the sample, as it shows a plateau within the LVER. Following DIN 1810-2, the last value of <span class="html-italic">G</span>′, which is within a 10% range of the plateau <span class="html-italic">G</span>′, is the end of the LVER. The related shear strain is the linearity limit <math display="inline"><semantics> <mrow> <msub> <mi>γ</mi> <mi>L</mi> </msub> </mrow> </semantics></math> and the related shear stress is called the yield stress or yield point <math display="inline"><semantics> <mrow> <msub> <mi>τ</mi> <mi>Y</mi> </msub> </mrow> </semantics></math>. When the sample starts to flow, <span class="html-italic">G</span>″, the viscous component, outweighs the elastic component, which is why it is called the flow point <math display="inline"><semantics> <mrow> <msub> <mi>τ</mi> <mi>F</mi> </msub> </mrow> </semantics></math>. By definition, this is the intersection point of <span class="html-italic">G</span>′ and <span class="html-italic">G</span>″, and the required stress for the sol–gel transition is the flow stress <math display="inline"><semantics> <mrow> <msub> <mi>τ</mi> <mi>F</mi> </msub> </mrow> </semantics></math>.</p>
Full article ">Figure A5
<p>An illustration of an analysis of a frequency sweep. As an example, the frequency sweep of a melted 10-IPM-ISFO is shown, measured at a 0.003% shear strain over frequencies between 0.03 and 100 Hz at 37 °C. The analyzed parameter <math display="inline"><semantics> <mrow> <msub> <mi>f</mi> <mi>L</mi> </msub> </mrow> </semantics></math> is marked and the linear region is shown as a gray shaded area. To compare the data, the end of the linearity was determined by fitting the <span class="html-italic">G</span>′-data using a least-squares linear fit. The frequency of the first data point that exceeds an error of 10% is referred to as the frequency of the linearity limit <math display="inline"><semantics> <mrow> <msub> <mi>f</mi> <mi>L</mi> </msub> </mrow> </semantics></math>. For emulsions and gels, the tendencies for syneresis can be estimated by the value of the <span class="html-italic">tanδ</span>; hence, the average of <span class="html-italic">δ</span> in the linearity range was determined and <span class="html-italic">tanδ</span> calculated.</p>
Full article ">Figure A6
<p>The syringeability levels using 20, 22, 24, and 26 G cannulas. The maximum forces required to compress the syringe and eject the sample were extracted from the force–distance curves acquired from the texture analysis. The data are presented as arithmetic means ± SD; <span class="html-italic">n</span> = 3. The data are categorized by the manufacturing technique and 12-HSA concentration. The compositions of the gels are coded based on the liquid lipids that are used (PO <span class="html-fig-inline" id="pharmaceutics-15-00254-i001"><img alt="Pharmaceutics 15 00254 i001" src="/pharmaceutics/pharmaceutics-15-00254/article_deploy/html/images/pharmaceutics-15-00254-i001.png"/></span>, MCT <span class="html-fig-inline" id="pharmaceutics-15-00254-i002"><img alt="Pharmaceutics 15 00254 i002" src="/pharmaceutics/pharmaceutics-15-00254/article_deploy/html/images/pharmaceutics-15-00254-i002.png"/></span>, IPM <span class="html-fig-inline" id="pharmaceutics-15-00254-i003"><img alt="Pharmaceutics 15 00254 i003" src="/pharmaceutics/pharmaceutics-15-00254/article_deploy/html/images/pharmaceutics-15-00254-i003.png"/></span>).</p>
Full article ">
Back to TopTop