[go: up one dir, main page]
More Web Proxy on the site http://driver.im/
You seem to have javascript disabled. Please note that many of the page functionalities won't work as expected without javascript enabled.
 
 
Sign in to use this feature.

Years

Between: -

Subjects

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Journals

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Article Types

Countries / Regions

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Search Results (366)

Search Parameters:
Keywords = skim milk

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
17 pages, 2900 KiB  
Article
Optimization of the Cryoprotectants for Direct Vat Set Starters in Sichuan Paocai Using Response Surface Methodology
by Lianqun Wu, Zhenying Yang, Ying Zhang, Ling Li, Chunli Tan, Lixia Pan, Yanping Wu, Kai Zhong and Hong Gao
Foods 2025, 14(2), 157; https://doi.org/10.3390/foods14020157 - 7 Jan 2025
Viewed by 468
Abstract
The quality of Sichuan paocai in natural fermentation is often inconsistent due to the complexity of its microbial community and environmental influences. To address this, dominant microbial strains were selectively inoculated to improve the product’s quality and safety. However, vacuum freeze-drying, commonly used [...] Read more.
The quality of Sichuan paocai in natural fermentation is often inconsistent due to the complexity of its microbial community and environmental influences. To address this, dominant microbial strains were selectively inoculated to improve the product’s quality and safety. However, vacuum freeze-drying, commonly used to prepare direct vat set (DVS) starters, can significantly damage strains due to freezing stress. This study aimed to optimize a freeze-drying protection system for Lactiplantibacillus plantarum and Bacillus subtilis to enhance their survival. Using response surface methodology, combinations of cryoprotectants were evaluated. A formulation comprising skim milk powder, glycerol, sucrose, and L-proline significantly improved strain viability after lyophilization, outperforming single cryoprotectants. Further investigation into storage conditions revealed that low temperatures (−20 °C) provided the best preservation for DVS starters. Furthermore, the optimized DVS starters demonstrated excellent fermentation performance in Sichuan paocai, enhancing its color, flavor, and sensory quality compared to natural fermentation. These findings offer a reliable freeze-drying protection strategy for survival and viability of L. plantarum and B. subtilis. Full article
Show Figures

Figure 1

Figure 1
<p>Effects of different kinds and different gradient concentrations of protectants on the survival rate of cryoprotectants of <span class="html-italic">L. plantarum</span> (<b>A</b>) and <span class="html-italic">B. subtilis</span> (<b>B</b>). For each gradient concentration of the same kind of cryoprotectant, the values corresponding to different letters are significantly different (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 2
<p>Response surface plots (<b>A<sub>1</sub></b>–<b>C<sub>1</sub></b>) and contour plots (<b>A<sub>2</sub></b>–<b>C<sub>2</sub></b>) of the two-way interaction of skim milk powder, sucrose, and L-proline for the response surface model <span class="html-italic">L. plantarum</span>.</p>
Full article ">Figure 3
<p>Response surface plots (<b>A<sub>1</sub></b>–<b>C<sub>1</sub></b>) and contour plots (<b>A<sub>2</sub></b>–<b>C<span class="html-italic"><sub>2</sub></span></b>) of the two-way interactions between skim milk powder, sucrose, and glycerol for the response surface model <span class="html-italic">B. subtilis</span>.</p>
Full article ">Figure 4
<p>Survival rates of <span class="html-italic">L. plantarum</span> (<b>A</b>) powder and <span class="html-italic">B. subtilis</span> (<b>B</b>) powder at different storage temperatures. For the same storage time, different letters indicate that the corresponding values at different temperatures were significantly different (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 5
<p>Physicochemical indexes and sensory evaluation of Sichuan paocai made using three different fermentation methods. Changes in pH (<b>A</b>), changes in nitrite content (<b>B</b>), sensory scores on the fourth and seventh day (<b>C</b>), differences in texture and structure on the seventh day of fermentation (<b>D</b>), and chromaticity color differences between the three groups of paocai and radish raw material (RM) on the seventh day of fermentation (<b>E</b>). Values corresponding to different letters in the determination of different indicators were significantly different (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">
14 pages, 1109 KiB  
Article
Isolation of Actinobacteria from Date Palm Rhizosphere with Enzymatic, Antimicrobial, Antioxidant, and Protein Denaturation Inhibitory Activities
by Maria Smati, Amina Bramki, Fatima Zohra Makhlouf, Rihab Djebaili, Beatrice Farda, Fatima Zohra Abdelhadi, Nahla Abdelli, Mahmoud Kitouni and Marika Pellegrini
Biomolecules 2025, 15(1), 65; https://doi.org/10.3390/biom15010065 - 5 Jan 2025
Viewed by 456
Abstract
Arid ecosystems constitute a promising source of actinobacteria producing new bioactive molecules. This study aimed to explore different biological activities of actinomycetes isolated from the rhizosphere of Phoenix dactylifera L. in the Ghardaia region, Algeria. A total of 18 actinobacteria were isolated and [...] Read more.
Arid ecosystems constitute a promising source of actinobacteria producing new bioactive molecules. This study aimed to explore different biological activities of actinomycetes isolated from the rhizosphere of Phoenix dactylifera L. in the Ghardaia region, Algeria. A total of 18 actinobacteria were isolated and studied for their enzymatic and antimicrobial activities. All isolates shared cellulase and catalase activity; most of them produced amylase (94%), esterase (84%), lecithinase and lipoproteins (78%), caseinase (94%), and gelatinase (72%). The isolates could coagulate (56%) or peptonize (28%) skim milk. Overall, 72% of the isolates exhibited significant antibacterial activity against at least one test bacteria, while 56% demonstrated antifungal activity against at least one test fungi. Based on enzyme production and antimicrobial activity, isolate SGI16 was selected for secondary metabolite extraction by ethyl acetate. The crude extract of SGI16 was analyzed using DPPH and BSA denaturation inhibition tests, revealing significant antioxidant power (IC50 = 7.24 ± 0.21 μg mL−1) and protein denaturation inhibitory capacity (IC50 = 492.41 ± 0.47 μg mL−1). Molecular identification based on 16S rDNA analysis showed that SGI16 belonged to the genus Streptomyces. The findings highlight that date palms’ rhizosphere actinobacteria are a valuable source of biomolecules of biotechnological interest. Full article
(This article belongs to the Section Natural and Bio-derived Molecules)
Show Figures

Figure 1

Figure 1
<p>Antibacterial activity of actinobacteria. Vertical bars represent standard error (<span class="html-italic">n</span> = 3). Different letters indicate significant differences at <span class="html-italic">p</span> ≤ 0.05 according to one-way ANOVA followed by Tukey’s HSD test.</p>
Full article ">Figure 2
<p>Antifungal activity of actinobacteria. Vertical bars represent standard error (<span class="html-italic">n</span> = 3). Different letters indicate significant differences at <span class="html-italic">p</span> ≤ 0.05 according to one-way ANOVA followed by Tukey’s HSD test.</p>
Full article ">Figure 3
<p>Phylogenetic tree constructed using the neighbor-joining method showing the relationship of isolate SGI16 with the closest species of the genus <span class="html-italic">Streptomyces</span>. Bootstrap values above 50% (for 1000 replicates) are indicated. Accession numbers for each sequence are shown in parentheses. The scale bar indicates 0.02 substitutions per nucleotide position. <span class="html-italic">Escherichia coli</span> NBRC(T) 102,203 was used as an outgroup.</p>
Full article ">
12 pages, 2650 KiB  
Article
A Sensitive and Selective Electrochemical Aptasensor for Carbendazim Detection
by Suthira Pushparajah, Mahnaz Shafiei and Aimin Yu
Biosensors 2025, 15(1), 15; https://doi.org/10.3390/bios15010015 - 3 Jan 2025
Viewed by 374
Abstract
Carbendazim (CBZ) is used to prevent fungal infections in agricultural crops. Given its high persistence and potential for long-term health effects, it is crucial to quickly identify pesticide residues in food and the environment in order to mitigate excessive exposure. Aptamer-based sensors offer [...] Read more.
Carbendazim (CBZ) is used to prevent fungal infections in agricultural crops. Given its high persistence and potential for long-term health effects, it is crucial to quickly identify pesticide residues in food and the environment in order to mitigate excessive exposure. Aptamer-based sensors offer a promising solution for pesticide detection due to their exceptional selectivity, design versatility, ease of use, and affordability. Herein, we report the development of an electrochemical aptasensor for CBZ detection. The sensor was fabricated through a one-step electrodeposition of platinum nanoparticles (Pt NPs) and reduced graphene oxide (rGO) on a glassy carbon electrode (GCE). Then, a CBZ-specific aptamer was attached via Pt-sulfur bonds. Upon combining CBZ with the aptamer on the electrode surface, the redox reaction of the electrochemical probe K4[Fe(CN)6] is hindered, resulting in a current drop. Under optimized conditions (pH of 7.5 and 25 min of incubation time), the proposed aptasensor showed a linear current reduction to CBZ concentrations between 0.5 and 15 nM. The limit of detection (LOD) for this proposed aptasensor is 0.41 nM. Along with its repeatable character, the aptasensor demonstrated better selectivity for CBZ compared to other potential compounds. The recovery rates for detecting CBZ in skim milk and tap water using the standard addition method were 98% and 96%, respectively. The proposed aptasensor demonstrated simplicity, sensitivity, and selectivity for detecting CBZ with satisfactory repeatability. It establishes a strong foundation for environmental monitoring of CBZ. Full article
Show Figures

Figure 1

Figure 1
<p>Schematic illustration of the preparation of the electrochemical aptasensor for CBZ detection.</p>
Full article ">Figure 2
<p>SEM images of (<b>A</b>) Pt-rGO/GCE and (<b>B</b>) Apt-Pt-rGO/GCE. XPS spectra of (<b>C</b>) wide scan of Apt-Pt-rGO/GCE, and (<b>D</b>) Peak binding energy shift of Pt 4f (a) before and (b) after aptamer immobilization.</p>
Full article ">Figure 3
<p>(<b>A</b>) CV plots and (<b>B</b>) Nyquist diagrams of EIS of (a) bare GCE, (b) Pt-rGO/GCE, and (c) Apt-Pt-rGO/GCE in a 0.1 M KCl solution containing 1.0 mM K<sub>4</sub>[Fe(CN)<sub>6</sub>].</p>
Full article ">Figure 4
<p>(<b>A</b>) DPVs of 1.0 mM K<sub>4</sub>[Fe(CN)<sub>6</sub>] at the aptasensor before and after adding 4 nM and 10 nM of CBZ in pH 7.0 PBS containing 0.1 M KCl. The effects of (<b>B</b>) incubation time (pH fixed at 7.0) and (<b>C</b>) pH (incubation time fixed at 25 min) on the CBZ current response.</p>
Full article ">Figure 5
<p>(<b>A</b>) DPV responses of the aptasensor toward CBZ with different concentrations (0, 0.5, 1, 2, 4, 6, 8, 10, and 15 nM) in pH 7.5 PBS containing 1.0 mM K<sub>4</sub>[Fe(CN)<sub>6</sub>] and 0.1 M KCl. (<b>B</b>) Linear curve of ΔI vs. CBZ concentration (nM).</p>
Full article ">Figure 6
<p>(<b>A</b>) Selectivity performance of the aptasensor in 10 nM of ciprofloxacin, acetaminophen, ascorbic acid, glucose, NaCl, KI, KNO<sub>3</sub>, and (NH<sub>4</sub>)<sub>2</sub>SO<sub>4</sub> in pH 7.5 PBS containing 1.0 mM K<sub>4</sub>[Fe(CN)<sub>6</sub>]. (<b>B</b>) Repeatability of the aptasensor in five samples containing 15 nM CBZ. (<b>C</b>) Current response of the aptasensor to 2 nM of CBZ when kept at 4 °C for 0, 7, 14, and 21 days.</p>
Full article ">
17 pages, 1523 KiB  
Article
Effect of Fortification with High-Milk-Protein Preparations on Yogurt Quality
by Justyna Żulewska, Maria Baranowska, Marika Magdalena Bielecka, Aneta Zofia Dąbrowska, Justyna Tarapata, Katarzyna Kiełczewska and Adriana Łobacz
Foods 2025, 14(1), 80; https://doi.org/10.3390/foods14010080 - 1 Jan 2025
Viewed by 735
Abstract
Protein-enriched yogurts have become increasingly popular among consumers seeking to boost their daily protein intake. The incorporation of milk proteins and protein preparations in yogurt production not only enhances nutritional value but also improves texture, viscosity, and overall sensory properties—key factors that influence [...] Read more.
Protein-enriched yogurts have become increasingly popular among consumers seeking to boost their daily protein intake. The incorporation of milk proteins and protein preparations in yogurt production not only enhances nutritional value but also improves texture, viscosity, and overall sensory properties—key factors that influence consumer acceptance. The main objective of this study was to evaluate the influence of casein and whey protein preparations on the physicochemical properties, viability of lactic acid bacteria, and sensory attributes of yogurts. Yogurts were enriched with 2% (w/w) protein preparations, including micellar casein preparation (CN85), whey protein isolate (WPI), whey protein concentrate (WPC60), and protein preparations obtained from skim milk by membrane filtration: micellar casein concentrate (CN75) and serum protein concentrate (SPC). The yogurts were produced using the thermostatic method, and their chemical composition, rheological properties, syneresis, firmness, lactic acid bacteria population, and sensory attributes were evaluated. The effects of high-protein preparations derived from skim milk through laboratory-scale membrane filtration processes (SPC, CN75) were compared with those of commercially available protein preparations (SMP, CN85, WPI, and WPC). Obtained results demonstrated that the membrane filtration-derived preparations (SPC and CN75) exhibited advantageous physicochemical properties and supported robust viability of yogurt and probiotic bacteria. However, their sensory quality was marginally inferior compared to the commercial preparations (SMP, CN85, WPI, and WPC). These findings indicate the potential applicability of membrane filtration-derived protein preparations in yogurt production while underscoring the necessity for further investigation to enhance and optimize their sensory characteristics. Full article
(This article belongs to the Special Issue Comprehensive Coverage of the Latest Research in the Dairy Industry)
Show Figures

Figure 1

Figure 1
<p>The production scheme for yogurts enriched with high-protein preparations. SMP—skim milk powder with 35% protein, WPI—whey protein isolate with 91% protein content, WPC—whey protein concentrate with 60% protein content, CN85—micellar casein preparation with 85% protein content, CN75—micellar casein concentrate with 75% protein content, SPC—serum protein concentrate with 67% protein content, M/YSMP—milk/yogurt with skim milk powder with 35% protein content, control sample, M/YWPI—milk/yogurt with whey protein isolate, M/YWPC—milk/yogurt with whey protein concentrate with 60% protein content, M/YCN85—milk/yogurt with micellar casein concentrate with 85% protein content, M/YCN75—milk/yogurt with micellar casein concentrate with 75% protein content, M/YSPC—milk/yogurt with serum protein with 67% protein content.</p>
Full article ">Figure 2
<p>(<b>a</b>) SDS-PAGE electrophoretogram of yogurt samples. (<b>b</b>). Average content of casein and whey proteins in a densitometric analysis of electrophoretograms (n = 2). CN—casein, WP—whey protein, TN—total nitrogen, YSMP—yogurt with skim milk powder with 35% protein content, control sample, YCN75—yogurt with micellar casein concentrate with 75% protein content, YCN85—yogurt with micellar casein concentrate with 85% protein content, YSPC—yogurt with serum protein with 67% protein content, YWPI—yogurt with whey protein isolate, and YWPC—yogurt with whey protein concentrate with 60% protein content.</p>
Full article ">Figure 3
<p>Flow curves of experimental yogurts after 3 days of storage.</p>
Full article ">Figure 4
<p>Flow curves of experimental yogurts after 21 days of storage.</p>
Full article ">Figure 5
<p>Firmness of yogurts enriched with 2% <span class="html-italic">w</span>/<span class="html-italic">w</span> of high-protein preparations after 3rd and 21st days of storage. Values are means ± SD (n = 6). YSMP—yogurt with skim milk powder with 35% protein content, control sample, YCN75—milk/yogurt with micellar casein concentrate with 75% protein content, YCN85—yogurt with micellar casein concentrate with 85% protein content, YSPC—yogurt with serum protein with 67% protein content, YWPI—yogurt with whey protein isolate, and YWPC—yogurt with whey protein concentrate with 60% protein content.</p>
Full article ">Figure 6
<p>Syneresis of yogurts enriched with 2% <span class="html-italic">w</span>/<span class="html-italic">w</span> of high-protein preparations after 3rd and 21st days of storage. Values are means ± SD (n = 6). YSMP—yogurt with skim milk powder with 35% protein content, control sample, YCN75—yogurt with micellar casein concentrate with 75% protein content, YCN85—yogurt with micellar casein concentrate with 85% protein content, YSPC—yogurt with serum protein with 67% protein content, YWPI—yogurt with whey protein isolate, and YWPC—yogurt with whey protein concentrate with 60% protein content.</p>
Full article ">
24 pages, 2390 KiB  
Article
Goat’s Milk Powder Enriched with Red (Lycium barbarum L.) and Black (Lycium ruthenicum Murray) Goji Berry Extracts: Chemical Characterization, Antioxidant Properties, and Prebiotic Activity
by Danijel D. Milinčić, Aleksandar Ž. Kostić, Steva Lević, Uroš M. Gašić, Dragana D. Božić, Relja Suručić, Tijana D. Ilić, Viktor A. Nedović, Bojana B. Vidović and Mirjana B. Pešić
Foods 2025, 14(1), 62; https://doi.org/10.3390/foods14010062 - 29 Dec 2024
Viewed by 773
Abstract
The current trend in food innovations includes developing products containing plant ingredients or extracts rich in bioactive compounds. This study aimed to prepare and characterize skimmed thermally treated goat’s milk powders enriched with lyophilized fruit extracts of Lycium ruthenicum Murray (GMLR) and Lycium [...] Read more.
The current trend in food innovations includes developing products containing plant ingredients or extracts rich in bioactive compounds. This study aimed to prepare and characterize skimmed thermally treated goat’s milk powders enriched with lyophilized fruit extracts of Lycium ruthenicum Murray (GMLR) and Lycium barbarum L. (GMLB). Proximate analysis, ultra-high performance liquid chromatography with quadrupole time-of-flight mass spectrometry (UPLC-Q-TOF-MS), Fourier transform infrared spectroscopy using attenuated total reflection (FTIR-ATR), and electrophoretic analysis were assessed. Total phenolic content (TPC), total protein content, and antioxidant properties of enriched goat milk powders were determined spectrophotometrically, and prebiotic potential was evaluated by the broth microdilution method. A total of 25 phenolic compounds and 18 phenylamides were detected in the enriched goat milk powders. Electrophoretic analysis showed the absence of proteolysis in the prepared powders. The GMLR showed the highest TPC and displayed a ferric ion-reducing power, probably contributed by anthocyanins and some phenylamides. GMLR and GMLB had higher ABTS radical scavenging activity but lower ferrous ion-chelating capacity than control goat′s milk powder. GMLB and GMLR in a dose-dependent manner (0.3–5 mg/mL) showed a growth-promoting effect on probiotic strains. In summary, prepared goji/goat milk powders, primarily GMLR, might be used as prebiotic supplements or functional food additives. Full article
(This article belongs to the Section Nutraceuticals, Functional Foods, and Novel Foods)
Show Figures

Figure 1

Figure 1
<p>Images of spray-dried powders: (<b>a</b>) goat’s milk powder without extract (control sample) (GM); (<b>b</b>) goat’s milk powder enriched with <span class="html-italic">L. ruthenicum</span> extract (GMLR); and (<b>c</b>) goat’s milk powder enriched with <span class="html-italic">L. barbarum</span> extract (GMLB).</p>
Full article ">Figure 2
<p>FTIR-ATR spectra of goat′s milk (GM) and goat′s milk/goji extracts (GMLR and GMLB) powders. Abbreviations: “GMLR”—goat’s milk powder enriched with <span class="html-italic">L. ruthenicum</span> extract; “GMLB”—goat’s milk powder enriched with <span class="html-italic">L. barbarum</span> extract; and “GM”—goat’s milk powder without extract (control sample).</p>
Full article ">Figure 3
<p>Characteristic MS/MS fragmentation patterns and predicted structures of non-glycosylated spermidine derivatives: (<b>a</b>) <span class="html-italic">N</span>-dihydrocaffeoyl-<span class="html-italic">N</span>′-coumaroyl spermidine (<span class="html-italic">m</span>/<span class="html-italic">z</span> 456); (<b>b</b>) <span class="html-italic">N</span>-caffeoyl-<span class="html-italic">N</span>′-dihydrocaffeoyl spermidine (<span class="html-italic">m</span>/<span class="html-italic">z</span> 472); and (<b>c</b>) <span class="html-italic">N</span>-feruloyl-<span class="html-italic">N</span>′-dihydrocaffeoyl spermidine (<span class="html-italic">m</span>/<span class="html-italic">z</span> 486).</p>
Full article ">Figure 4
<p>Electrophoretic patterns of goat′s milk (GM) and goat′s milk/goji extract (GMLR and GMLB) powders, analyzed by SDS-PAGE in reducing conditions. Lines: “GMLR”—goat’s milk powder enriched with <span class="html-italic">L. ruthenicum</span> extract; “GMLB”—goat’s milk powder enriched with <span class="html-italic">L. barbarum</span> extract; “GM”—goat’s milk powder without extract (control sample); molecular weight standard (LMW); and bovine milk protein standard (SP). Abbreviations: bovine serum albumin (BSA); immunoglobulin hard chain (Ighc); αs2-casein (αs2-CN); αs1-casein (αs1-CN); β-casein (β-CN); κ-casein (κ-CN); β-lactoglobulins (β-LG); and α-lactalbumin (α-LA).</p>
Full article ">Figure 5
<p>The prebiotic potential of the goat′s milk and goat′s milk/goji extract powders on various probiotic strains of microorganisms. The results are presented as minimum and maximum percent of growth stimulation of prepared powders in the range of 0.312–5 mg/mL. Abbreviations: “GMLR”—goat’s milk powder enriched with <span class="html-italic">L. ruthenicum</span> extract; “GMLB”—goat’s milk powder enriched with <span class="html-italic">L. barbarum</span> extract; and “GM”—goat’s milk powder without extract (control sample).</p>
Full article ">
14 pages, 2795 KiB  
Article
The Refractive Index of Human Milk Serum: Natural Variations and Dependency on Macronutrient Concentrations
by Johanna R. de Wolf, Kawthar Ali, Chris G. Legtenberg, Wietske Verveld and Nienke Bosschaart
Foods 2024, 13(24), 4124; https://doi.org/10.3390/foods13244124 - 20 Dec 2024
Viewed by 439
Abstract
The refractive index (RI) of human milk serum (also known as whey, milk soluble fraction or milk plasma) depends on the individual molecular species dissolved in the serum and their concentrations. Although the human milk serum RI is known to influence milk analysis [...] Read more.
The refractive index (RI) of human milk serum (also known as whey, milk soluble fraction or milk plasma) depends on the individual molecular species dissolved in the serum and their concentrations. Although the human milk serum RI is known to influence milk analysis methods based on light scattering, the RI dependency on human milk serum composition is currently unknown. Therefore, we systematically evaluate how the RI depends on natural variations in macronutrient concentrations in the soluble fraction. We measure RI variations in serum simulating samples with controlled macronutrient concentrations, as well as skimmed and whole fore-, bulk, and hindmilk from 19 donors. For both types of samples, we relate the measured RI to the macronutrient composition. From the serum simulating samples, we observe that the RI depends more on variations in whey protein, than carbohydrate concentrations, while minerals have negligible influence. For all donated samples, the average RI was 1.3470 (range 1.3466–1.3474). Per donor, no significant differences were observed in RI between fore-, bulk, and hindmilk. We conclude that protein and solids-not-fat (i.e., the total contribution of carbohydrates, proteins and minerals present in milk) concentrations are most predictive for human milk serum RI. Full article
(This article belongs to the Section Dairy)
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>–<b>d</b>) the measured refractive index (o) of the serum simulating samples as a function of the carbohydrate concentration, for different mineral concentrations (0.090, 0.130, 0.170, and 0.210 g/dL). Between the four different subplots, the protein concentration is varied (0.45, 0.75, 1.05, and 1.35 g/dL). The lines are linear fits on the refractive index against the carbohydrate concentration per constant mineral and protein concentration, with the slope <span class="html-italic">a</span> and offset <span class="html-italic">b</span> displayed in the tables below each graph. The error bars represent the standard deviation of the measurements and may fall behind data points. (<b>e</b>) The parameter <span class="html-italic">b</span> from the tables below graphs a-d as a function of the protein concentration. The lines are linear fits on the parameter <span class="html-italic">b</span> against the protein concentration with slope <span class="html-italic">c</span> and offset <span class="html-italic">d</span>, as displayed in the table on the right. The error bars represent the 95% confidence interval of the linear fit on the refractive index against carbohydrate concentration.</p>
Full article ">Figure 2
<p>The refractive index of the donated human milk samples within (<b>a</b>) the measured refractive index per sample and (<b>b</b>) a histogram of the refractive index of all skimmed mature bulk milk samples.</p>
Full article ">Figure 3
<p>The measured refractive index of whole milk samples as a function of the measured refractive index of skimmed milk samples for (<b>a</b>) foremilk, (<b>b</b>) bulk milk, and (<b>c</b>) hindmilk. (<b>d</b>) The difference in measured refractive index between the whole milk and skimmed milk samples as a function of fat concentration of the whole milk sample.</p>
Full article ">Figure 4
<p>The refractive index of skimmed bulk milk samples as a function of (<b>a</b>) fat concentration, (<b>b</b>) carbohydrate concentration, (<b>c</b>) protein concentration, (<b>d</b>) and the solids-not-fat concentration.</p>
Full article ">Figure 5
<p>(<b>a</b>) Refractive index, (<b>b</b>) protein concentration, and (<b>c</b>) the solids-not-fat as a function of lactation period. The legend in figure c applies to a and b as well.</p>
Full article ">
15 pages, 5388 KiB  
Article
Electrostatic Spray Drying of a Milk Protein Matrix—Impact on Maillard Reactions
by Doll Chutani, Todor Vasiljevic, Thom Huppertz and Eoin Murphy
Molecules 2024, 29(24), 5994; https://doi.org/10.3390/molecules29245994 - 19 Dec 2024
Viewed by 669
Abstract
Electrostatic spray drying (ESD) of a milk protein matrix comprising whey protein isolate (WPI), skim milk powder (SMP), and lactose was compared to conventional spray drying (CSD) and freeze-drying (FD). ESD and CSD were used to produce powders at low (0.12–0.14), medium (0.16–0.17), [...] Read more.
Electrostatic spray drying (ESD) of a milk protein matrix comprising whey protein isolate (WPI), skim milk powder (SMP), and lactose was compared to conventional spray drying (CSD) and freeze-drying (FD). ESD and CSD were used to produce powders at low (0.12–0.14), medium (0.16–0.17), and high (0.31–0.36) levels of water activity (aw), while FD powders targeted low aw (0.12). Maillard reaction indicators were studied after drying and during storage for up to 28 days at 20, 40, or 60 °C by measuring free -NH2 groups, as an indicator of available lysine, and 5-hydroxymethylfurfural (HMF). After drying, levels of residual free -NH2 groups were ~15% higher in ESD and FD powders than in their CSD counterparts. CSD powders also had ~14% higher HMF concentrations compared to their ESD and FD counterparts. Storage led to reductions in free -NH2 groups and increases in HMF content in all powders, the extent of which increased with increasing storage temperature. Reductions in free -NH2 groups followed first-order reaction kinetics at 20 and 40 °C but second-order reaction kinetics at 60 °C. Lactose crystallization was detected in high-aw CSD powders after 14 d at 40 °C and in both CSD and ESD powders after 7 d at 60 °C. Overall, we found that ESD is a gentle drying technology which enables production of powders with lower Maillard reaction markers. Full article
(This article belongs to the Special Issue Feature Papers in Food Chemistry—3rd Edition)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Impact of electrostatic spray drying (<span class="html-fig-inline" id="molecules-29-05994-i001"><img alt="Molecules 29 05994 i001" src="/molecules/molecules-29-05994/article_deploy/html/images/molecules-29-05994-i001.png"/></span>), conventional spray drying (<span class="html-fig-inline" id="molecules-29-05994-i002"><img alt="Molecules 29 05994 i002" src="/molecules/molecules-29-05994/article_deploy/html/images/molecules-29-05994-i002.png"/></span>), and freeze-drying (<span class="html-fig-inline" id="molecules-29-05994-i003"><img alt="Molecules 29 05994 i003" src="/molecules/molecules-29-05994/article_deploy/html/images/molecules-29-05994-i003.png"/></span>) on concentrations of free -NH<sub>2</sub> groups when powders were produced at different water activity levels. The error bars represent the standard deviations from two experimental trials of each drying condition.</p>
Full article ">Figure 2
<p>Impact of electrostatic spray drying (<span class="html-fig-inline" id="molecules-29-05994-i001"><img alt="Molecules 29 05994 i001" src="/molecules/molecules-29-05994/article_deploy/html/images/molecules-29-05994-i001.png"/></span>), conventional spray drying (<span class="html-fig-inline" id="molecules-29-05994-i002"><img alt="Molecules 29 05994 i002" src="/molecules/molecules-29-05994/article_deploy/html/images/molecules-29-05994-i002.png"/></span>), and freeze-drying (<span class="html-fig-inline" id="molecules-29-05994-i003"><img alt="Molecules 29 05994 i003" src="/molecules/molecules-29-05994/article_deploy/html/images/molecules-29-05994-i003.png"/></span>) on HMF (5-hydroxymethylfurfural) concentrations when powders were produced at different water activity levels. The error bars represent the standard deviations from two experimental trials of each drying condition.</p>
Full article ">Figure 3
<p>Levels of free -NH<sub>2</sub> groups in low- (<b>a</b>,<b>d</b>,<b>g</b>), medium- (<b>b</b>,<b>e</b>,<b>h</b>), or high- (<b>c</b>,<b>f</b>,<b>i</b>) water-activity powders produced by electrostatic spray drying (<span class="html-fig-inline" id="molecules-29-05994-i004"><img alt="Molecules 29 05994 i004" src="/molecules/molecules-29-05994/article_deploy/html/images/molecules-29-05994-i004.png"/></span>), conventional spray drying (<span class="html-fig-inline" id="molecules-29-05994-i005"><img alt="Molecules 29 05994 i005" src="/molecules/molecules-29-05994/article_deploy/html/images/molecules-29-05994-i005.png"/></span>), and freeze-drying (<span class="html-fig-inline" id="molecules-29-05994-i006"><img alt="Molecules 29 05994 i006" src="/molecules/molecules-29-05994/article_deploy/html/images/molecules-29-05994-i006.png"/></span>) as a function of storage time at 20 (<b>a</b>–<b>c</b>), 40 (<b>d</b>–<b>f</b>), or 60 (<b>g</b>–<b>i</b>) °C. The arrows (<span class="html-fig-inline" id="molecules-29-05994-i007"><img alt="Molecules 29 05994 i007" src="/molecules/molecules-29-05994/article_deploy/html/images/molecules-29-05994-i007.png"/></span>: ESD and <span class="html-fig-inline" id="molecules-29-05994-i008"><img alt="Molecules 29 05994 i008" src="/molecules/molecules-29-05994/article_deploy/html/images/molecules-29-05994-i008.png"/></span>: CSD) indicate the time point at which crystallization was detected in the sample via FTIR. FD samples exclusively contained samples with low a<sub>w</sub>. The error bars represent the standard deviations from two experimental trials of each drying condition.</p>
Full article ">Figure 4
<p>Levels of HMF in low- (<b>a</b>,<b>d</b>,<b>g</b>), medium- (<b>b</b>,<b>e</b>,<b>h</b>) or high- (<b>c</b>,<b>f</b>,<b>i</b>) water-activity powders produced by electrostatic spray drying (<span class="html-fig-inline" id="molecules-29-05994-i004"><img alt="Molecules 29 05994 i004" src="/molecules/molecules-29-05994/article_deploy/html/images/molecules-29-05994-i004.png"/></span>), conventional spray drying (<span class="html-fig-inline" id="molecules-29-05994-i005"><img alt="Molecules 29 05994 i005" src="/molecules/molecules-29-05994/article_deploy/html/images/molecules-29-05994-i005.png"/></span>), and freeze-drying (<span class="html-fig-inline" id="molecules-29-05994-i006"><img alt="Molecules 29 05994 i006" src="/molecules/molecules-29-05994/article_deploy/html/images/molecules-29-05994-i006.png"/></span>) as a function of storage time at 20 (<b>a</b>–<b>c</b>), 40 (<b>d</b>–<b>f</b>), or 60 (<b>g</b>–<b>i</b>) °C. The arrows (<span class="html-fig-inline" id="molecules-29-05994-i007"><img alt="Molecules 29 05994 i007" src="/molecules/molecules-29-05994/article_deploy/html/images/molecules-29-05994-i007.png"/></span>: ESD and <span class="html-fig-inline" id="molecules-29-05994-i008"><img alt="Molecules 29 05994 i008" src="/molecules/molecules-29-05994/article_deploy/html/images/molecules-29-05994-i008.png"/></span>: CSD) indicate the time point at which crystallization was detected in the sample via FTIR. FD samples exclusively contained samples with low a<sub>w</sub>. The error bars represent the standard deviations from two experimental trials of each drying condition.</p>
Full article ">Figure 5
<p>Graphical representation of the experimental design of this study.</p>
Full article ">Figure 6
<p>FTIR spectra of (ESD) powder at time t = 0 of storage in the range 4000–400 cm<sup>−1</sup> with the annotated region of interest indicated between the red dashed lines. FTIR—Fourier Transform Infrared and ESD—electrostatic spray drying.</p>
Full article ">
12 pages, 1653 KiB  
Article
Amino Acid Composition of Dried Bovine Dairy Powders from a Range of Product Streams
by Simon R. Gilmour, Stephen E. Holroyd, Maher D. Fuad, Dave Elgar and Aaron C. Fanning
Foods 2024, 13(23), 3901; https://doi.org/10.3390/foods13233901 - 3 Dec 2024
Viewed by 654
Abstract
The amino acid (AA) content of multiple samples of various dairy powders was determined, providing a comprehensive evaluation of the differences in AA profiles attributable to distinct manufacturing processes. Products examined included whole milk powder (WMP), skim milk powder (SMP), cheese whey protein [...] Read more.
The amino acid (AA) content of multiple samples of various dairy powders was determined, providing a comprehensive evaluation of the differences in AA profiles attributable to distinct manufacturing processes. Products examined included whole milk powder (WMP), skim milk powder (SMP), cheese whey protein concentrate (WPC-C), lactic acid casein whey protein concentrate (WPC-L), high-fat whey protein concentrate (WPC-HF), hydrolyzed whey protein concentrate (WPH), whey protein isolate (WPI), and demineralized whey protein (D90). WMP and SMP exhibited broadly similar AA profiles, with minor differences likely due to the minimal milk fat protein content, which is nearly absent from SMP. Comparative analysis of WPC-C and WPC-L indicated higher levels of threonine, serine, glutamic acid, and proline in WPC-C but lower levels of tyrosine, phenylalanine, and tryptophan, attributed to the different methods of separation from casein proteins. WPI and WPC-HF originate from similar sweet whey streams but follow divergent processing methods; consequent on this were variations in the levels of all AAs except histidine. The nanofiltration step in D90 production retains its non-protein nitrogen content and affects its AA profile; consequently, D90 consistently exhibited lower AA levels than WPC-C. These findings underscore the significant impact of manufacturing processes on dairy powder AA composition. Full article
(This article belongs to the Section Dairy)
Show Figures

Figure 1

Figure 1
<p>Amino acid profile of hydrolyzed whey protein concentrate (WPH).</p>
Full article ">Figure 2
<p>Amino acid profiles of whole milk powder (WMP) (<span style="color:#D0CECE">■</span>) and skim milk powder (SMP) (<span style="color:gray">■</span>). * Indicates statistical difference (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 3
<p>Amino acid profiles of cheese whey protein concentrate (WPC-C) (<span style="color:#D0CECE">■</span>) and lactic acid casein whey protein concentrate (WPC-L) (<span style="color:gray">■</span>). * Indicates statistical difference (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 4
<p>Amino acid profiles of whey protein isolate (WPI) (<span style="color:#D0CECE">■</span>) and high-fat whey protein concentrate (WPC-HF) (<span style="color:gray">■</span>). * Indicates statistical difference (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 5
<p>Amino acid profiles of demineralized whey (D90) (<span style="color:#D0CECE">■</span>) and cheese whey protein concentrate (WPC-C) (<span style="color:gray">■</span>). * Indicates statistical difference (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 6
<p>Amino acid profiles of demineralized whey (D90) (<span style="color:#D0CECE">■</span>) and cheese whey protein concentrate (WPC-C) (<span style="color:gray">■</span>) as a percentage of total amino acid content. * Indicates statistical difference (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">
10 pages, 1008 KiB  
Article
Determination of Protein Interaction in Milk Protein Concentrate Powders Manufactured from pH-Adjusted and Heat-Treated Skim Milk
by Kavya Dileep, Hari Meletharayil and Jayendra K. Amamcharla
Foods 2024, 13(23), 3832; https://doi.org/10.3390/foods13233832 - 28 Nov 2024
Viewed by 739
Abstract
The influence of heating as a pretreatment on the structural and functional attributes of milk protein concentrate (MPC) powders derived from ultrafiltered/diafiltered (UF/DF) skim milk is under-reported. This research delves into the impact of pH and heat treatment on skim milk’s properties before [...] Read more.
The influence of heating as a pretreatment on the structural and functional attributes of milk protein concentrate (MPC) powders derived from ultrafiltered/diafiltered (UF/DF) skim milk is under-reported. This research delves into the impact of pH and heat treatment on skim milk’s properties before UF/DF and how these changes affect the resulting MPC powders. By adjusting the pH of skim milk to 6.5, 6.8, or 7.1 and applying thermal treatment at 90 °C for 15 min to one of two divided lots (with the other serving as a control), we studied the protein interactions in MPC. Post-heat treatment, the skim milk’s pH was adjusted back to 6.8, followed by ultrafiltration and spray drying to produce MPC powders with protein content of 82.38 ± 2.72% on a dry matter basis. MPC dispersions from these powders at 5% protein (w/w) were also evaluated for particle size, viscosity, and heat coagulation time (HCT) to further understand how the protein interactions in skim milk influence the properties of MPC dispersions. Capillary electrophoresis was used to assess the casein and whey protein distribution in both the soluble and colloidal phases. Findings revealed that preheating skim milk at pH 7.1 increased serum phase interactions, while heating skim milk preadjusted to a pH of 6.5 promoted whey protein–casein interactions at the micellar interface. Notably, the D (4,3) of casein micelles was larger for dispersions from milk with a preheated pH of 6.5 compared to other pH levels, correlating positively with enhanced dispersion viscosity due to increased volume fraction. These results support the potential for tailoring MPC powder functionality in various food applications through the precise control of the milk’s pre-treatment conditions. Full article
Show Figures

Figure 1

Figure 1
<p>Experimental design for the manufacture of MPC powders heated and unheated at different pH values.</p>
Full article ">Figure 2
<p>Typical capillary gel electropherogram of serum phase of (<b>A</b>) MPC dispersions of MPC powders manufactured from unheated skim milk at pH 6.8, (<b>B</b>) MPC dispersions of MPC powders manufactured from heated skim milk at pH 6.8.</p>
Full article ">
20 pages, 3951 KiB  
Article
Lactiplantibacillus plantarum for the Preparation of Fermented Low-Bitter Enzymatic Skim Milk with Antioxidant Ability
by Yi Jiang, Longfei Zhang, Yushi Jin, Haiyan Xu, Yating Liang, Zihan Xia, Chenchen Zhang, Chengran Guan, Hengxian Qu, Yunchao Wa, Wenqiong Wang, Yujun Huang, Ruixia Gu and Dawei Chen
Foods 2024, 13(23), 3828; https://doi.org/10.3390/foods13233828 - 27 Nov 2024
Viewed by 647
Abstract
A high degree of hydrolysis can reduce the allergenicity of milk, while lactic acid bacteria (LAB) fermentation can further enhance the antioxidant ability of enzymatic milk. LAB with a strong antioxidant ability was screened, and the effects of LAB on the bitterness, taste [...] Read more.
A high degree of hydrolysis can reduce the allergenicity of milk, while lactic acid bacteria (LAB) fermentation can further enhance the antioxidant ability of enzymatic milk. LAB with a strong antioxidant ability was screened, and the effects of LAB on the bitterness, taste and flavor of enzymatic skim milk (ESM) with a high degree of hydrolysis were investigated in this paper, in addition to the response surface methodology optimized the conditions of the LAB fermentation of ESM. The results indicate that the skim milk hydrolyzed by Protamex has a higher degree of hydrolysis and lower bitterness. The scavenging rate of 2,2-Diphenyl-1-picrylhydrazyl (DPPH) free radical, the inhibition rate of hydroxyl radical (·OH) and the superoxide dismutase (SOD) activity of Lactiplantibacillus plantarum 16 and Lactococcus lactis subsp. lactis m16 are significantly higher than those of other strains (p < 0.05), while the improvement effect of L. plantarum 16 on the bitterness and flavor of ESM is better than that of L. lactis subsp. lactis m16. The fermented ESM has a strong antioxidant ability and low bitterness when the inoculum quantity of L. plantarum 16 is 5%, fermentation at 37 °C for 18 h and the pH of the ESM is 6.5, for which the DPPH free radical scavenging rate is 61.32%, the ·OH inhibition rate is 83.35%, the SOD activity rate is 14.58 and the sensory evaluation is 4.25. The contents of amino acids related to bitterness and antioxidants were reduced and increased, respectively. The ESM fermented by L. plantarum 16 has a good flavor, antioxidant ability and low bitterness. Full article
Show Figures

Figure 1

Figure 1
<p>Hydrolysis degree of ESM (<span class="html-italic">n</span> = 3, x ± sd). Different letters indicate significant differences (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 2
<p>Bitter taste of ESM by three kinds of proteases (<span class="html-italic">n</span> = 10, x ± sd). Different lowercase letters indicate significant differences (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 3
<p>The effect of quinine concentration on bitterness response value.</p>
Full article ">Figure 4
<p>Electronic tongue flavor analysis of ESM.</p>
Full article ">Figure 5
<p>The antioxidant ability of ESM fermented by LAB (<span class="html-italic">n</span> = 3, x ± sd). Different lowercase letters of the same indicator indicate significant difference (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 6
<p>Bitterness evaluation value of ESM fermented by LAB (<span class="html-italic">n</span> = 3, x ± sd). Different lowercase letters indicate significant differences (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 7
<p>Electronic tongue taste profiles of ESM fermented by LAB.</p>
Full article ">Figure 8
<p>The antioxidant ability of ESM fermented by <span class="html-italic">L. plantarum</span> 16 at different fermentation temperatures. (<b>A</b>), DPPH free radical scavenging rate; (<b>B</b>), ·OH inhibition rate; (<b>C</b>), SOD activity. Different lowercase letters indicate significant differences (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 9
<p>The antioxidant ability of ESM fermented by <span class="html-italic">L. plantarum</span> 16 at different fermentation times. (<b>A</b>), DPPH free radical scavenging rate; (<b>B</b>), ·OH inhibition rate; (<b>C</b>), SOD activity. Different lowercase letters indicate significant differences (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 10
<p>The antioxidant ability of ESM fermented by <span class="html-italic">L. plantarum</span> 16 under different inoculum quantity. (<b>A</b>), DPPH free radical scavenging rate; (<b>B</b>), ·OH inhibition rate; (<b>C</b>), SOD activity. Different lowercase letters indicate significant differences (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 11
<p>The antioxidant ability of <span class="html-italic">L. plantarum</span> 16-fermented ESM under different enzyme release pH conditions. (<b>A</b>), DPPH free radical scavenging rate; (<b>B</b>), ·OH inhibition rate; (<b>C</b>), SOD activity. Different lowercase letters indicate significant differences (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 12
<p>Response surface and contour plots of the impact of interaction on comprehensive evaluation values. The lighter the color, the less the influence of the experimental factors on the fuzzy comprehensive evaluation in this response surface plot.</p>
Full article ">
11 pages, 2116 KiB  
Article
Enzymatic Oxidation of Aflatoxin M1 in Milk Using CotA Laccase
by Yongpeng Guo, Hao Lv, Zhiyong Rao, Zhixiang Wang, Wei Zhang, Yu Tang and Lihong Zhao
Foods 2024, 13(22), 3702; https://doi.org/10.3390/foods13223702 - 20 Nov 2024
Viewed by 632
Abstract
Aflatoxin M1 (AFM1) in milk poses a significant threat to human health. This study examined the capacity of Bacillus licheniformis CotA laccase to oxidize AFM1. The optimal conditions for the CotA laccase-catalyzed AFM1 oxidation were observed at [...] Read more.
Aflatoxin M1 (AFM1) in milk poses a significant threat to human health. This study examined the capacity of Bacillus licheniformis CotA laccase to oxidize AFM1. The optimal conditions for the CotA laccase-catalyzed AFM1 oxidation were observed at pH 8.0 and 70 °C, achieving an AFM1 oxidation rate of 86% in 30 min. The Km and Vmax values for CotA laccase with respect to AFM1 were 18.91 μg mL−1 and 9.968 μg min−1 mg−1, respectively. Computational analysis suggested that AFM1 interacted with CotA laccase via hydrogen bonding and van der Waals interactions. Moreover, the oxidation products of AFM1 mediated by CotA laccase were identified as the C3-hydroxy derivatives of AFM1 by HPLC-FLD and UPLC-TOF/MS. Toxicological assessment revealed that the hepatotoxicity of AFM1 was substantially reduced following oxidation by CotA laccase. The efficacy of CotA laccase in removing AFM1 in milk was further tested, and the result showed that the enzyme agent achieved an AFM1 removal rate of 83.5% in skim milk and 65.1% in whole milk. These findings suggested that CotA laccase was a novel AFM1 oxidase capable of eliminating AFM1 in milk. More effort is still needed to improve the AFM1 oxidase activity of CotA laccase in order to shorten the processing time when applying the enzyme in the milk industry. Full article
Show Figures

Figure 1

Figure 1
<p>Enzymatic properties and kinetics of CotA laccase-mediated AFM<sub>1</sub> oxidation. Effect of pH (<b>A</b>), temperature (<b>B</b>), and metal ions (<b>C</b>) on CotA laccase-mediated AFM<sub>1</sub> oxidation. (<b>D</b>) Michaelis–Menten plot of CotA laccase-catalyzed AFM<sub>1</sub> oxidation.</p>
Full article ">Figure 2
<p>Identification of CotA laccase-mediated AFM<sub>1</sub> oxidation products. (<b>A</b>) HPLC chromatograms of AFM<sub>1</sub> and CotA laccase-mediated AFM<sub>1</sub> oxidation products. (<b>B</b>) Mass spectra analysis of AFM<sub>1</sub> and CotA laccase-mediated AFM<sub>1</sub> oxidation products. (<b>C</b>) The reaction scheme for AFM<sub>1</sub> oxidation by CotA laccase.</p>
Full article ">Figure 3
<p>Molecular docking analysis of AFM<sub>1</sub> with CotA laccase. (<b>A</b>) The two-dimensional interaction model of AFM<sub>1</sub> with CotA laccase. (<b>B</b>) The three-dimensional interaction model of AFM<sub>1</sub> with CotA laccase.</p>
Full article ">Figure 4
<p>Evaluation of the cytotoxic effects of AFM<sub>1</sub> and its oxidation products. (<b>A</b>) Viability of L-02 cells following exposure to 100 μΜ of AFM<sub>1</sub> and CotA laccase-catalyzed AFM<sub>1</sub> oxidation products. (<b>B</b>) LDH activity. (<b>C</b>,<b>D</b>) Apoptosis rate of L-02 cells. Different letters denote statistically significant differences between groups (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 5
<p>Elimination of AFM<sub>1</sub> in milk by CotA laccase. Effect of CotA laccase amount (<b>A</b>) and incubation time (<b>B</b>) on AFM<sub>1</sub> degradation rate in skim milk and whole milk.</p>
Full article ">
22 pages, 6438 KiB  
Article
Evaluation of Hazelnut Cake Flour for Use as a Milk Powder Replacer in Ice Cream
by Mirela Lučan Čolić, Antun Jozinović, Jasmina Lukinac, Marko Jukić and Martina Antunović
Appl. Sci. 2024, 14(22), 10303; https://doi.org/10.3390/app142210303 - 9 Nov 2024
Viewed by 658
Abstract
Hazelnut oil cake, a by-product in the cold-pressing of hazelnut oil, is a rich in valuable nutrients, which makes it a promising option for supplementation or as a raw material in the development of functional products. The aim of this work was to [...] Read more.
Hazelnut oil cake, a by-product in the cold-pressing of hazelnut oil, is a rich in valuable nutrients, which makes it a promising option for supplementation or as a raw material in the development of functional products. The aim of this work was to study the influence of partial or complete replacing of skim milk powder (SMP) with hazelnut press cake flour (HPCF) in varying ratios (0%, 25%, 50%, 75%, and 100%) on the physicochemical properties and sensory attributes of milk ice cream. The replacement modified the chemical composition of the ice cream mixture, resulting in a reduction (p < 0.05) of milk solids non-fat (MSNF), protein, and carbohydrates content, while simultaneously elevating the hazelnut content, and total fat content. This modification influenced the rheological characteristics of the ice cream mixtures, leading to an increase in the consistency coefficient from 1.32 to 7.66 Pa sn. Furthermore, a decline in overrun values (from 26.99% to 15.85%), an increase in hardness (from 6881.71 to 23,829.30 g), retarded melting properties, and variations in colour attributes were observed with higher concentrations of HPCF. In the sensory evaluation test, it was found that consumer acceptance was enhanced for the samples with partial substitution of SMP when compared to standard milk ice cream. The findings suggest that a replacement of milk powder with hazelnut cake by up to 75% is achievable, in order to obtain functional ice cream with adequate physicochemical and sensorial qualities. Full article
(This article belongs to the Section Food Science and Technology)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Ice cream samples presented during the sensory analysis.</p>
Full article ">Figure 2
<p>Rheology properties of ice cream mixtures: (<b>a</b>) apparent viscosity (Pa s) and (<b>b</b>) flow curves–shear stress (Pa) at various shear rates (1/s). C—control ice cream; HN1, HN2, HN3, HN4—samples with 25%, 50%, 75%, and 100% replacements of SMP with HPCF.</p>
Full article ">Figure 3
<p>Texture profile of ice cream samples: (<b>a</b>) hardness (g), (<b>b</b>) adhesiveness (g∙s), (<b>c</b>) cohesiveness, and (<b>d</b>) gumminess. The results are expressed as mean ± SD. Means in the same column with different superscripts are significantly different according to the Tukey HSD test (<span class="html-italic">p</span> &lt; 0.05). C—control ice cream; HN1, HN2, HN3, HN4—samples with 25%, 50%, 75%, and 100% replacements of SMP with HPCF.</p>
Full article ">Figure 4
<p>Melting behaviour of ice cream samples: (<b>a</b>) pictures taken during the melting and (<b>b</b>) melting curves. The results are expressed as mean ± SD. C—control ice cream; HN1, HN2, HN3, HN4—samples with 25%, 50%, 75%, and 100% replacements of SMP with HPCF.</p>
Full article ">Figure 5
<p>Sensory scores of ice cream samples: (<b>a</b>) overall acceptability and (<b>b</b>) specific sensory attributes. The results are expressed as mean ± SD. Means in the same column with different superscripts are significantly different according to the Tukey HSD test (<span class="html-italic">p</span> &lt; 0.05). C—control ice cream; HN1, HN2, HN3, HN4—samples with 25%, 50%, 75%, and 100% replacements of SMP with HPCF.</p>
Full article ">Figure 6
<p>Principal component analysis (PCA): (<b>a</b>) bootstrap ellipses and (<b>b</b>) biplot graph for ice cream with hazelnut press cake flour. C—control ice cream; HN1, HN2, HN3, HN4—samples with 25%, 50%, 75%, and 100% replacements of SMP with HPCF; CH—carbohydrate, TS—total solid, MSNF—milk solid non-fat, TA—titratable acidity, OR—overrun, FDI—fat destabilisation index.</p>
Full article ">Figure A1
<p>Schematic diagram of the study protocol.</p>
Full article ">
12 pages, 1966 KiB  
Article
Purification and Identification of the Nematicidal Activity of S1 Family Trypsin-Like Serine Protease (PRA1) from Trichoderma longibrachiatum T6 Through Prokaryotic Expression and Biological Function Assays
by Nan Ma, Hang Lv, Solomon Boamah, Shuwu Zhang and Bingliang Xu
Genes 2024, 15(11), 1437; https://doi.org/10.3390/genes15111437 - 6 Nov 2024
Viewed by 723
Abstract
Background/Objectives: Heterodera avenae is a highly significant plant-parasitic nematode, causing severe economic losses to global crop production each year. Trichoderma species have been found to parasitize nematodes and control them by producing enzymes that degrade eggshells. The T. longibrachiatum T6 (T6) strain has [...] Read more.
Background/Objectives: Heterodera avenae is a highly significant plant-parasitic nematode, causing severe economic losses to global crop production each year. Trichoderma species have been found to parasitize nematodes and control them by producing enzymes that degrade eggshells. The T. longibrachiatum T6 (T6) strain has been demonstrated the parasitic and lethal effects on H. avenae cysts and eggs, associated with the increased serine protease activity and trypsin-like serine protease gene (PRA1) expression. Methods: Our present study aimed to purify the recombinant PRA1 protease through a prokaryotic expression system and identify its nematicidal activity. Results: The recombinant PRA1 protease was identified as S1 family trypsin-like serine protease, with a molecular weight of 43.16 kDa. The purified soluble protease exhibited the optimal activity at 35 °C and pH 8.0, and also demonstrating higher hydrolytic ability toward casein and skimmed milk. Meanwhile, the Ca2+ and Mg2+ enhanced its activity, while the inhibitor PMSF significantly reduced it. The contents of H. avenae eggs leaked out after treatment with the recombinant PRA1 protease, with egg hatching inhibition and relative hatching inhibition rates at 70.60% and 66.58%, respectively. In contrast, there was no sign of content dissolution, and embryos developed normally in the control group. Conclusions: Our present study revealed that the PRA1 protease of T6 strain has a lethal effect on H. avenae eggs, which providing a theoretical basis for developing biocontrol agents to control nematodes. Full article
(This article belongs to the Section Plant Genetics and Genomics)
Show Figures

Figure 1

Figure 1
<p>Construction of the expression vector with 1.2% agarose gel electrophoresis detection: (<b>A</b>) The <span class="html-italic">E. coli</span> DH5α detection using PCR, where M represents the 2000 DNA Marker, lanes 1–8 are pET-32a-PRA1 recombinant plasmid that has been transformed to <span class="html-italic">E. coli</span> DH5α, lane 9 is the pET-32a empty plasmid amplification (positive control), and lane 10 is the sterile water instead of template amplification (negative control). (<b>B</b>) Dual-enzyme digestion verification of the recombinant plasmid pET-32a-PRA1, where M represents the 10,000 DNA Marker, and lane 1 represent the double-enzyme digestion products of the recombinant plasmid pET-32a-PRA1.</p>
Full article ">Figure 2
<p>Determination of recombinant PRA1 protease expression by SDS-PAGE: (<b>A</b>) Induction of recombinant PRA1 protease in <span class="html-italic">E. coli</span> BL21 (DE3) using IPTG, where M represents the protein marker; the first lane is a non-induced pET-32a empty vector at 28 °C for 4 h; the second lane represents a positive control, which is an <span class="html-italic">E. coli</span> BL21 (DE3) solution of pET-32a empty vector that has been transformed into competent cells and induced at 28 °C for 4 h; the third lane displays an <span class="html-italic">E. coli</span> BL21 (DE3) solution of recombinant protease that has not been induced; the fourth lane represents a supernatant of recombinant protease that has been induced at 28 °C for 4 h; and the fifth lane represents precipitation of recombinant protease that has been induced at 28 °C for 4 h. (<b>B</b>) Analysis of the isolated recombinant protease using SDS-PAGE, where lane 1 is the purified recombinant protease and M is the protein marker.</p>
Full article ">Figure 3
<p>Effects of temperature and pH on the recombinant PRA1 protease’s enzyme activity: (<b>A</b>) Recombinant PRA1 protease at its optimal temperature. (<b>B</b>) The recombinant PRA1 protease’s thermal stability. (<b>C</b>) The recombinant PRA1 protease at its optimal pH. (<b>D</b>) The pH stability of the recombinant PRA1 protease. The means ± standard errors are displayed for the data, and columns labeled with various letters denote significant differences at <span class="html-italic">p</span> &lt; 0.05.</p>
Full article ">Figure 4
<p>Morphological analysis of <span class="html-italic">H. avenae</span> eggs subjected to recombinant PRA1 protease treatment. Eggs after treatment with recombinant PRA1 protease are represented in images (<b>A1</b>–<b>A5</b>), taken at 2, 4, 6, 8, and 10 days under 400× magnification, respectively. Samples (<b>B1</b>–<b>B5</b>) served as positive controls, representing eggs at 2, 4, 6, 8, and 10 days post-treatment with the <span class="html-italic">E. coli</span> BL21 (DE3) solution containing the pET-32a empty vector, which was transformed into competent <span class="html-italic">E. coli</span> BL21 (DE3) cells, respectively. Observations for (<b>B1</b>–<b>B3</b>) were conducted at a magnification of 400×, while (<b>B4</b>,<b>B5</b>) were observed at 200× magnification. (<b>C1</b>–<b>C5</b>) served as negative controls, representing eggs at 2, 4, 6, 8, and 10 days post-treatment with sterile water, respectively. Observations for (<b>C1</b>–<b>C3</b>) were conducted at 400× magnification, while (<b>C4</b>,<b>C5</b>) were examined at 200× magnification.</p>
Full article ">
17 pages, 8438 KiB  
Article
Bacterial Cellulose–Silk Hydrogel Biosynthesized by Using Coconut Skim Milk as Culture Medium for Biomedical Applications
by Junchanok Chaikhunsaeng, Phasuwit P. Phatchayawat, Suchata Kirdponpattara and Muenduen Phisalaphong
Gels 2024, 10(11), 714; https://doi.org/10.3390/gels10110714 - 6 Nov 2024
Viewed by 846
Abstract
In this study, hydrogel films of biocomposite comprising bacterial cellulose (BC) and silk (S) were successfully fabricated through a simple, facile, and cost-effective method via biosynthesis by Acetobacter xylinum in a culture medium of coconut skim milk/mature coconut water supplemented with the powders [...] Read more.
In this study, hydrogel films of biocomposite comprising bacterial cellulose (BC) and silk (S) were successfully fabricated through a simple, facile, and cost-effective method via biosynthesis by Acetobacter xylinum in a culture medium of coconut skim milk/mature coconut water supplemented with the powders of thin-shell silk cocoon (SC). Coconut skim milk/mature coconut water and SC are the main byproducts of coconut oil and silk textile industries, respectively. The S/BC films contain protein, carbohydrate, fat, and minerals and possess a number of properties beneficial to wound healing and tissue engineering, including nontoxicity, biocompatibility, appropriate mechanical properties, flexibility, and high water absorption capacity. It was demonstrated that silk could fill into a porous structure and cover fibers of the BC matrix with very good integration. In addition, components (fat, protein, etc.) in coconut skim milk could be well incorporated into the hydrogel, resulting in a more elastic structure and higher tensile strength of films. The tensile strength and the elongation at break of BC film from coconut skim milk (BCM) were 212.4 MPa and 2.54%, respectively, which were significantly higher than BC film from mature coconut water (BCW). A more elastic structure and relatively higher tensile strength of S/BCM compared with S/BCW were observed. The films of S/BCM and S/BCW showed very high water uptake ability in the range of 400–500%. The presence of silk in the films also significantly enhanced the adhesion, proliferation, and cell-to-cell interaction of Vero and HaCat cells. According to multiple improved properties, S/BC hydrogel films are high-potential candidates for application as biomaterials for wound dressing and tissue engineering. Full article
(This article belongs to the Special Issue Designing Gels for Wound Dressing)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Thin-shell silk cocoon (<b>A</b>) and silk cocoon powder (<b>B</b>).</p>
Full article ">Figure 2
<p>FESEM images of surface morphologies (<b>top</b>) and cross-section (<b>bottom</b>) at magnification of 20,000× (<b>A</b>) BCW, (<b>B</b>) 0.10S/BCW, (<b>C</b>) 0.15S/BCW, (<b>D</b>) 0.20S/BCW, (<b>E</b>) BCM, (<b>F</b>) 0.10S/BCM, (<b>G</b>) 0.15S/BCM, and (<b>H</b>) 0.20S/BCM.</p>
Full article ">Figure 3
<p>Mechanical properties of dried BC films prepared from mature coconut water (CW) and coconut skim milk (CM) culture medium supplemented with silk at 0.00–0.25%.</p>
Full article ">Figure 4
<p>FTIR spectra of SC, BCW film, and S/BCW films.</p>
Full article ">Figure 5
<p>FTIR spectra of SC, BCM film, and S/BCM films.</p>
Full article ">Figure 6
<p>XRD patterns: BCW and S/BCW films (<b>A</b>); BCM and S/BCM films (<b>B</b>).</p>
Full article ">Figure 7
<p>Toxicity test against Vero, L929, and HaCat cell lines on BCW, S/BCW, BCM, and S/BCM films for (<b>A</b>) 24 h and (<b>B</b>) 48 h.</p>
Full article ">Figure 8
<p>SEM images of Vero (<b>up</b>) and HaCat (<b>down</b>) cell adhesion on the cover glass (control) for 48 h at magnification of (<b>A</b>,<b>C</b>) 100×and (<b>B</b>,<b>D</b>) 1000×.</p>
Full article ">Figure 9
<p>Vero cell adhesion on (<b>A</b>) BCW, (<b>B</b>) 0.10S/BCW, (<b>C</b>) 0.15S/BCW, (<b>D</b>) 0.20S/BCW, (<b>E</b>) BCM, (<b>F</b>) 0.10S/BCM, (<b>G</b>) 0.15S/BCM, and (<b>H</b>) 0.20S/BCM after cultivation for 48 h.</p>
Full article ">Figure 10
<p>HaCat cell adhesion on (<b>A</b>) BCW, (<b>B</b>) 0.10S/BCW, (<b>C</b>) 0.15S/BCW, (<b>D</b>) 0.20S/BCW, (<b>E</b>) BCM, (<b>F</b>) 0.10S/BCM, (<b>G</b>) 0.15S/BCM, and (<b>H</b>) 0.20S/BCM after cultivation for 48 h.</p>
Full article ">Figure 11
<p>Schematic diagram of the fabrication process of BC and S/BC films.</p>
Full article ">
33 pages, 4683 KiB  
Article
Component Distribution, Shear-Flow Behavior, and Sol–Gel Transition in Mixed Dispersions of Casein Micelles and Serum Proteins
by Hossein Gholamian, Maksym Loginov, Marie-Hélène Famelart, Florence Rousseau, Fabienne Garnier-Lambrouin and Geneviève Gésan-Guiziou
Foods 2024, 13(21), 3480; https://doi.org/10.3390/foods13213480 - 30 Oct 2024
Viewed by 970
Abstract
The shear flow and solid–liquid transition of mixed milk protein dispersions with varying concentrations of casein micelles (CMs) and serum proteins (SPs) are integral to key dairy processing operations, including microfiltration, ultrafiltration, diafiltration, and concentration–evaporation. However, the rheological behavior of these dispersions has [...] Read more.
The shear flow and solid–liquid transition of mixed milk protein dispersions with varying concentrations of casein micelles (CMs) and serum proteins (SPs) are integral to key dairy processing operations, including microfiltration, ultrafiltration, diafiltration, and concentration–evaporation. However, the rheological behavior of these dispersions has not been sufficiently studied. In the present work, dispersions of CMs and SPs with total protein weight fractions (ωPR) of 0.021–0.28 and SP to total protein weight ratios (RSP) of 0.066–0.214 and 1 were prepared by dispersing the respective protein isolates in the permeate from skim milk ultrafiltration and then further concentrated via osmotic compression. The partition of SPs between the CMs and the dispersion medium was assessed by measuring the dry matter content and viscosity of the dispersion medium after separating it from the CMs via ultracentrifugation. The rheological properties were studied at 20 °C via shear rheometry, and the sol–gel transition was characterized via oscillatory measurements. No absorption of SPs by CMs was observed in dispersions with ωPR = 0.083–0.126, regardless of the RSP. For dispersions of SPs with ωPR ≤ 0.21, as well as the dispersion medium of mixed dispersions with ωPR = 0.083–0.126, the high shear- rate-limiting viscosity was described using Lee’s equation with an SP voluminosity (vSP) of 2.09 mL·g−1. For the mixed dispersions with a CM volume fraction of φCM ≤ 0.37, the relative high shear-rate-limiting viscosity was described using Lee’s equation with a CM voluminosity (vCM) of 4.15 mL·g−1 and a vSP of 2.09 mL·g−1, regardless of the RSP. For the mixed dispersions with φCM > 0.55, the relative viscosity increased significantly with an increasing RSP (this was explained by an increase in repulsion between CMs). However, the sol–gel transition was independent of the RSP and was observed at φCM ≈ 0.65. Full article
(This article belongs to the Section Dairy)
Show Figures

Figure 1

Figure 1
<p>Theoretical DM content of the dispersion medium (<math display="inline"><semantics> <mrow> <msubsup> <mrow> <mi>ω</mi> </mrow> <mrow> <mi>D</mi> <mi>M</mi> <mo>,</mo> <mi>t</mi> <mi>h</mi> </mrow> <mrow> <mi>d</mi> <mi>m</mi> </mrow> </msubsup> </mrow> </semantics></math>) (calculated under the assumption set C (<a href="#foods-13-03480-t002" class="html-table">Table 2</a>) vs. measured total DM content of supernatants (<math display="inline"><semantics> <mrow> <msubsup> <mrow> <mi>ω</mi> </mrow> <mrow> <mi>D</mi> <mi>M</mi> <mo>,</mo> <mi>e</mi> <mi>x</mi> <mi>p</mi> </mrow> <mrow> <mi>d</mi> <mi>m</mi> </mrow> </msubsup> </mrow> </semantics></math>) of mixed milk protein dispersions). Data for suspensions with <span class="html-italic">R<sub>SP</sub></span> = 0.066 and <math display="inline"><semantics> <mrow> <msubsup> <mrow> <mi>ω</mi> </mrow> <mrow> <mi>P</mi> <mi>R</mi> </mrow> <mrow> <mi>D</mi> </mrow> </msubsup> </mrow> </semantics></math> = 0.083 and 0.125 (squares), <span class="html-italic">R<sub>SP</sub></span> = 0.158 and <math display="inline"><semantics> <mrow> <msubsup> <mrow> <mi>ω</mi> </mrow> <mrow> <mi>P</mi> <mi>R</mi> </mrow> <mrow> <mi>D</mi> </mrow> </msubsup> </mrow> </semantics></math> = 0.105 and 0.126 (triangles), and <span class="html-italic">R<sub>SP</sub></span> = 0.214 and <math display="inline"><semantics> <mrow> <msubsup> <mrow> <mi>ω</mi> </mrow> <mrow> <mi>P</mi> <mi>R</mi> </mrow> <mrow> <mi>D</mi> </mrow> </msubsup> </mrow> </semantics></math> = 0.126 (inverted triangles). The dashed line corresponds to <math display="inline"><semantics> <mrow> <msubsup> <mrow> <mi>ω</mi> </mrow> <mrow> <mi>D</mi> <mi>M</mi> <mo>,</mo> <mi>e</mi> <mi>x</mi> <mi>p</mi> </mrow> <mrow> <mi>d</mi> <mi>m</mi> </mrow> </msubsup> </mrow> </semantics></math> = <math display="inline"><semantics> <mrow> <msubsup> <mrow> <mi>ω</mi> </mrow> <mrow> <mi>D</mi> <mi>M</mi> <mo>,</mo> <mi>t</mi> <mi>h</mi> </mrow> <mrow> <mi>d</mi> <mi>m</mi> </mrow> </msubsup> </mrow> </semantics></math>. The determination coefficient for the data fitting by this equation (<span class="html-italic">r</span><sup>2</sup>) is shown near the dashed line. The DM content at the origin is that of PUF (0.0521).</p>
Full article ">Figure 2
<p>(<b>a</b>) Steady-state apparent viscosity (<span class="html-italic">η</span>) as a function of shear rate (<math display="inline"><semantics> <mrow> <mover accent="true"> <mrow> <mi>γ</mi> </mrow> <mo>˙</mo> </mover> </mrow> </semantics></math>) for WPI dispersions (<span class="html-italic">R<sub>SP</sub></span> = 1) prepared by mixing WPI powder with PUF (solid symbols) and following osmotic compression (open symbols); total protein concentrations (<math display="inline"><semantics> <mrow> <msubsup> <mrow> <mi>ω</mi> </mrow> <mrow> <mi>P</mi> <mi>R</mi> </mrow> <mrow> <mi>D</mi> </mrow> </msubsup> </mrow> </semantics></math>) are shown next to the curves. (<b>b</b>) <math display="inline"><semantics> <mrow> <mi>η</mi> <mo>(</mo> <mrow> <mrow> <mi>γ</mi> </mrow> </mrow> <mo>|</mo> <mo>)</mo> </mrow> </semantics></math> for supernatants produced via ultracentrifugation of mother dispersions; the <span class="html-italic">R<sub>SP</sub></span> in dispersions used for ultracentrifugation are shown next to the curves; the dashed line indicates experimental data for PUF.</p>
Full article ">Figure 3
<p>(<b>a</b>) Relative viscosity (<math display="inline"><semantics> <mrow> <msub> <mrow> <mi>η</mi> </mrow> <mrow> <mi>r</mi> <mi>P</mi> <mi>U</mi> <mi>F</mi> </mrow> </msub> </mrow> </semantics></math>) at a shear rate of 100 s<sup>−1</sup> as a function of SP concentration (<math display="inline"><semantics> <mrow> <msubsup> <mrow> <mi>ω</mi> </mrow> <mrow> <mi>P</mi> <mi>R</mi> </mrow> <mrow> <mi>D</mi> </mrow> </msubsup> <msub> <mrow> <mi>ρ</mi> </mrow> <mrow> <mi>D</mi> </mrow> </msub> </mrow> </semantics></math>) in liquid (solid symbols) and compressed (open symbols) dispersions of WPI; solid curve corresponds to data fitting by Lee’s equation, dashed curves correspond to lower and upper limits of the 95% prediction interval; left and bottom axes correspond to results over the entire range of <math display="inline"><semantics> <mrow> <msubsup> <mrow> <mi>ω</mi> </mrow> <mrow> <mi>P</mi> <mi>R</mi> </mrow> <mrow> <mi>D</mi> </mrow> </msubsup> <msub> <mrow> <mi>ρ</mi> </mrow> <mrow> <mi>D</mi> </mrow> </msub> </mrow> </semantics></math> studied (circles), while the right and top axes correspond to results for the lowest <math display="inline"><semantics> <mrow> <msubsup> <mrow> <mi>ω</mi> </mrow> <mrow> <mi>P</mi> <mi>R</mi> </mrow> <mrow> <mi>D</mi> </mrow> </msubsup> <msub> <mrow> <mi>ρ</mi> </mrow> <mrow> <mi>D</mi> </mrow> </msub> </mrow> </semantics></math> (diamonds). (<b>b</b>) Relative viscosity (<math display="inline"><semantics> <mrow> <msub> <mrow> <mi>η</mi> </mrow> <mrow> <mi>r</mi> <mi>P</mi> <mi>U</mi> <mi>F</mi> </mrow> </msub> </mrow> </semantics></math>) at a shear rate of 100 s<sup>−1</sup> as a function of SP concentration (<math display="inline"><semantics> <mrow> <msubsup> <mrow> <mi>ω</mi> </mrow> <mrow> <mi>S</mi> <mi>P</mi> </mrow> <mrow> <mi>d</mi> <mi>m</mi> </mrow> </msubsup> <msub> <mrow> <mi>ρ</mi> </mrow> <mrow> <mi>D</mi> </mrow> </msub> </mrow> </semantics></math>) in supernatants produced via ultracentrifugation of dispersions with different <span class="html-italic">R<sub>SP</sub></span> and <math display="inline"><semantics> <mrow> <msubsup> <mrow> <mi>ω</mi> </mrow> <mrow> <mi>P</mi> <mi>R</mi> </mrow> <mrow> <mi>D</mi> </mrow> </msubsup> </mrow> </semantics></math>: <span class="html-italic">R<sub>SP</sub></span> = 0.066 and <math display="inline"><semantics> <mrow> <msubsup> <mrow> <mi>ω</mi> </mrow> <mrow> <mi>P</mi> <mi>R</mi> </mrow> <mrow> <mi>D</mi> </mrow> </msubsup> </mrow> </semantics></math> = 0.083 and 0.125 (squares), <span class="html-italic">R<sub>SP</sub></span> = 0.158 and <math display="inline"><semantics> <mrow> <msubsup> <mrow> <mi>ω</mi> </mrow> <mrow> <mi>P</mi> <mi>R</mi> </mrow> <mrow> <mi>D</mi> </mrow> </msubsup> </mrow> </semantics></math> = 0.105 and 0.126 (triangles), and <span class="html-italic">R<sub>SP</sub></span> = 0.214 and <math display="inline"><semantics> <mrow> <msubsup> <mrow> <mi>ω</mi> </mrow> <mrow> <mi>P</mi> <mi>R</mi> </mrow> <mrow> <mi>D</mi> </mrow> </msubsup> </mrow> </semantics></math> = 0.126 (inverted triangles); solid and dashed curves are the same as in (<b>a</b>).</p>
Full article ">Figure 4
<p>Steady-state apparent viscosity (<span class="html-italic">η</span>) as a function of shear rate (<math display="inline"><semantics> <mrow> <mover accent="true"> <mrow> <mi>γ</mi> </mrow> <mo>˙</mo> </mover> </mrow> </semantics></math>) for dispersions with <span class="html-italic">R<sub>SP</sub></span> = 0.066 (<b>a</b>), 0.158 (<b>b</b>), and 0.214 (<b>c</b>): solid symbols—liquid dispersions prepared by mixing and diluting CNI and WPI powders with PUF; open symbols—compressed dispersions prepared via osmotic compression; values of <math display="inline"><semantics> <mrow> <msubsup> <mrow> <mi>ω</mi> </mrow> <mrow> <mi>P</mi> <mi>R</mi> </mrow> <mrow> <mi>D</mi> </mrow> </msubsup> </mrow> </semantics></math> are shown next to the curves.</p>
Full article ">Figure 5
<p>Apparent (<span class="html-italic">η</span>) (<b>a</b>) and relative (<math display="inline"><semantics> <mrow> <msubsup> <mrow> <mi>η</mi> </mrow> <mrow> <mi>r</mi> <mi>d</mi> <mi>m</mi> </mrow> <mrow> <mi>D</mi> </mrow> </msubsup> </mrow> </semantics></math>) (<b>b</b>) viscosity of dispersions with near-Newtonian behavior at <math display="inline"><semantics> <mrow> <mover accent="true"> <mrow> <mi>γ</mi> </mrow> <mo>˙</mo> </mover> <mo>=</mo> <mn>100</mn> <msup> <mrow> <mi>s</mi> </mrow> <mrow> <mo>−</mo> <mn>1</mn> </mrow> </msup> </mrow> </semantics></math> as a function of <math display="inline"><semantics> <mrow> <msubsup> <mrow> <mi>φ</mi> </mrow> <mrow> <mi>C</mi> <mi>M</mi> </mrow> <mrow> <mi>D</mi> </mrow> </msubsup> </mrow> </semantics></math> calculated for <span class="html-italic">v<sub>CM</sub></span> = 4.15 mL·g<sup>−1</sup>: <span class="html-italic">R<sub>SP</sub></span> = 0.066 (squares), 0.158 (triangles), and 0.214 (inverted triangles); the solid curve in <a href="#foods-13-03480-f005" class="html-fig">Figure 5</a>b was calculated using the modified Lee’s equation (Equation (18)).</p>
Full article ">Figure 6
<p>Apparent viscosity (<span class="html-italic">η</span>) measured at <math display="inline"><semantics> <mrow> <mover accent="true"> <mrow> <mi>γ</mi> </mrow> <mo>˙</mo> </mover> </mrow> </semantics></math> = 10 s<sup>−1</sup> (<b>a</b>) and 100 s<sup>−1</sup> (<b>b</b>) as a function of total protein concentration (<math display="inline"><semantics> <mrow> <msubsup> <mrow> <mi>ω</mi> </mrow> <mrow> <mi>P</mi> <mi>R</mi> </mrow> <mrow> <mi>D</mi> </mrow> </msubsup> </mrow> </semantics></math>) for dispersions with <span class="html-italic">R<sub>SP</sub></span> = 0.066 (CNI, squares), 0.158 (triangles), 0.214 (inverted triangles), and 1 (WPI, circles).</p>
Full article ">Figure 7
<p>Relative viscosity of dispersion (<span class="html-italic">η<sub>r</sub></span>) at <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>η</mi> </mrow> <mrow> <mi>d</mi> <mi>m</mi> </mrow> </msub> <mover accent="true"> <mrow> <mi>γ</mi> </mrow> <mo>˙</mo> </mover> </mrow> </semantics></math> of 0.005 Pa (<b>a</b>) and 0.15 Pa (<b>b</b>) as a function of the volume fraction of casein micelles (CMs) in dispersions with <span class="html-italic">R<sub>SP</sub></span> = 0.066 (squares), 0.158 (triangles), and 0.214 (inverted triangles); dashed vertical lines correspond to <math display="inline"><semantics> <mrow> <msubsup> <mrow> <mi>φ</mi> </mrow> <mrow> <mi>C</mi> <mi>M</mi> </mrow> <mrow> <mi>D</mi> </mrow> </msubsup> </mrow> </semantics></math> = 0.55 and 0.65.</p>
Full article ">Figure 8
<p>Elastic (<span class="html-italic">G</span>′) (solid symbols) and loss (<span class="html-italic">G</span>″) (open symbols) moduli as a function of frequency (<span class="html-italic">θ</span>) for mixed milk protein dispersions with <span class="html-italic">R<sub>SP</sub></span> = 0.006 (CNI) (<b>a</b>), 0.158 (<b>b</b>), 0.214 (<b>c</b>), and 1 (WPI) (<b>d</b>). Symbols correspond to experimental data; total protein concentrations (<math display="inline"><semantics> <mrow> <msubsup> <mrow> <mi>ω</mi> </mrow> <mrow> <mi>P</mi> <mi>R</mi> </mrow> <mrow> <mi>D</mi> </mrow> </msubsup> </mrow> </semantics></math>) are shown next to the curves.</p>
Full article ">Figure 9
<p>Elastic modulus <span class="html-italic">G</span>′ (measured at a frequency (<math display="inline"><semantics> <mrow> <mi>θ</mi> </mrow> </semantics></math>) of 0.1 Hz) and exponent <span class="html-italic">m</span> (estimated using Equation (20)) for mixed milk protein dispersions prepared by osmotic compression for <span class="html-italic">R<sub>SP</sub></span> = 0.066 (CNI) (<b>a</b>), 0.158 (<b>b</b>), and 0.214 (<b>c</b>). The left and right edges of shaded zones correspond to dispersions with liquid-like (<span class="html-italic">G</span>′ &lt; <span class="html-italic">G</span>″) or solid-like (<span class="html-italic">G</span>′ &gt; <span class="html-italic">G</span>″) behavior, respectively, in the range of <math display="inline"><semantics> <mrow> <mi>θ</mi> </mrow> </semantics></math> of 0.1–10.0 Hz.</p>
Full article ">Figure 10
<p>Composition of dispersions in the sol–gel transition region as a function of the relative concentration of serum proteins (<span class="html-italic">R<sub>SP</sub></span>): particle volume fraction of casein micelles (CMs) in the dispersion (<math display="inline"><semantics> <mrow> <msubsup> <mrow> <mi>φ</mi> </mrow> <mrow> <mi>C</mi> <mi>M</mi> </mrow> <mrow> <mi>D</mi> <mo>,</mo> <mi>s</mi> <mi>g</mi> </mrow> </msubsup> </mrow> </semantics></math>) (squares), the volume fraction of proteins in the dispersion (<math display="inline"><semantics> <mrow> <msubsup> <mrow> <mi>φ</mi> </mrow> <mrow> <mi>P</mi> <mi>R</mi> </mrow> <mrow> <mi>D</mi> <mo>,</mo> <mi>s</mi> <mi>g</mi> </mrow> </msubsup> </mrow> </semantics></math>) (circles), and volume fraction of serum proteins in the dm (<math display="inline"><semantics> <mrow> <msubsup> <mrow> <mi>φ</mi> </mrow> <mrow> <mi>S</mi> <mi>P</mi> </mrow> <mrow> <mi>d</mi> <mi>m</mi> <mo>,</mo> <mi>s</mi> <mi>g</mi> </mrow> </msubsup> </mrow> </semantics></math>) (diamonds). Symbols correspond to the middle of the sol–gel transition region, while error bars indicate its upper and lower limits (shaded zones in <a href="#foods-13-03480-f009" class="html-fig">Figure 9</a>). The dashed horizontal line indicates the mean <math display="inline"><semantics> <mrow> <msubsup> <mrow> <mi>φ</mi> </mrow> <mrow> <mi>C</mi> <mi>M</mi> </mrow> <mrow> <mi>D</mi> <mo>,</mo> <mi>s</mi> <mi>g</mi> </mrow> </msubsup> </mrow> </semantics></math> (0.65).</p>
Full article ">Figure 11
<p>(<b>a</b>) Theoretical dry matter (DM) content of the dispersion medium (<math display="inline"><semantics> <mrow> <msubsup> <mrow> <mi>ω</mi> </mrow> <mrow> <mi>D</mi> <mi>M</mi> <mo>,</mo> <mi>t</mi> <mi>h</mi> </mrow> <mrow> <mi>d</mi> <mi>m</mi> </mrow> </msubsup> </mrow> </semantics></math>) (calculated for the assumed values of <span class="html-italic">K<sub>SP</sub></span>, <span class="html-italic">f<sub>o</sub></span>, and <span class="html-italic">f<sub>eMN</sub></span> listed) vs. measured total DM content of supernatants (<math display="inline"><semantics> <mrow> <msubsup> <mrow> <mi>ω</mi> </mrow> <mrow> <mi>D</mi> <mi>M</mi> <mo>,</mo> <mi>e</mi> <mi>x</mi> <mi>p</mi> </mrow> <mrow> <mi>d</mi> <mi>m</mi> </mrow> </msubsup> </mrow> </semantics></math>) of mixed milk protein dispersions. Data for suspensions with <span class="html-italic">R<sub>SP</sub></span> = 0.066 and <math display="inline"><semantics> <mrow> <msubsup> <mrow> <mi>ω</mi> </mrow> <mrow> <mi>P</mi> <mi>R</mi> </mrow> <mrow> <mi>D</mi> </mrow> </msubsup> </mrow> </semantics></math> = 0.083 and 0.125 (squares), <span class="html-italic">R<sub>PS</sub></span> = 0.158 and <math display="inline"><semantics> <mrow> <msubsup> <mrow> <mi>ω</mi> </mrow> <mrow> <mi>P</mi> <mi>R</mi> </mrow> <mrow> <mi>D</mi> </mrow> </msubsup> </mrow> </semantics></math> = 0.105 and 0.126 (triangles), and <span class="html-italic">R<sub>SP</sub></span> = 0.214 and <math display="inline"><semantics> <mrow> <msubsup> <mrow> <mi>ω</mi> </mrow> <mrow> <mi>P</mi> <mi>R</mi> </mrow> <mrow> <mi>D</mi> </mrow> </msubsup> </mrow> </semantics></math> = 0.126 (inverted triangles). The dashed line is the 1:1 line. The DM content at the origin is that of PUF (0.0521). (<b>b</b>) Relative viscosity (<math display="inline"><semantics> <mrow> <msub> <mrow> <mi>η</mi> </mrow> <mrow> <mi>r</mi> <mi>P</mi> <mi>U</mi> <mi>F</mi> </mrow> </msub> </mrow> </semantics></math>) at a shear rate of 100 s<sup>−1</sup> as a function of the concentration of serum protein in supernatants (<math display="inline"><semantics> <mrow> <msubsup> <mrow> <mi>ω</mi> </mrow> <mrow> <mi>S</mi> <mi>P</mi> </mrow> <mrow> <mi>d</mi> <mi>m</mi> </mrow> </msubsup> <msub> <mrow> <mi>ρ</mi> </mrow> <mrow> <mi>D</mi> </mrow> </msub> </mrow> </semantics></math>) (calculated for the assumed values of <span class="html-italic">K<sub>SP</sub></span>, <span class="html-italic">f<sub>o</sub></span>, and <span class="html-italic">f<sub>eMN</sub></span> listed); data for dispersions with <span class="html-italic">R<sub>SP</sub></span> = 0.066 and <math display="inline"><semantics> <mrow> <msubsup> <mrow> <mi>ω</mi> </mrow> <mrow> <mi>P</mi> <mi>R</mi> </mrow> <mrow> <mi>D</mi> </mrow> </msubsup> </mrow> </semantics></math> = 0.083 and 0.125 (squares), <span class="html-italic">R<sub>SP</sub></span> = 0.158 and <math display="inline"><semantics> <mrow> <msubsup> <mrow> <mi>ω</mi> </mrow> <mrow> <mi>P</mi> <mi>R</mi> </mrow> <mrow> <mi>D</mi> </mrow> </msubsup> </mrow> </semantics></math> = 0.105 and 0.126 (triangles), and <span class="html-italic">R<sub>SP</sub></span> = 0.214 and <math display="inline"><semantics> <mrow> <msubsup> <mrow> <mi>ω</mi> </mrow> <mrow> <mi>P</mi> <mi>R</mi> </mrow> <mrow> <mi>D</mi> </mrow> </msubsup> </mrow> </semantics></math> = 0.126 (inverted triangles); solid curve corresponds to data described by Lee’s equation, while dashed curves correspond to lower and upper limits of the 95% prediction interval.</p>
Full article ">
Back to TopTop