[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
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
remove_circle_outline
remove_circle_outline

Article Types

Countries / Regions

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Search Results (933)

Search Parameters:
Keywords = cellulase

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
15 pages, 5509 KiB  
Article
Effect of Alkaline and Hydrothermal Pretreatments in Sugars and Ethanol Production from Rice Husk Waste
by José Ayala-Armijos, Byron Lapo, Carolina Beltrán, Joaquín Sigüenza, Braulio Madrid, Estefanía Chérrez, Verónica Bravo and Diana Sanmartín
Resources 2024, 13(9), 128; https://doi.org/10.3390/resources13090128 - 16 Sep 2024
Abstract
This study investigates the effectiveness of hydrothermal and alkaline pretreatment methods in enhancing the concentration of fermentable sugars derived from rice husk waste. After the pretreatments, enzymatic hydrolysis and fermentation processes were executed to evaluate the ethanol production from each pretreatment. Rice husk [...] Read more.
This study investigates the effectiveness of hydrothermal and alkaline pretreatment methods in enhancing the concentration of fermentable sugars derived from rice husk waste. After the pretreatments, enzymatic hydrolysis and fermentation processes were executed to evaluate the ethanol production from each pretreatment. Rice husk powder measuring ≤250 µm was used. For the alkaline pretreatment, sodium hydroxide (NaOH) was used at concentrations ranging from 0.5, 1 to 1.5% w/v. The efficacy of the hydrothermal pretreatment method was evaluated after 15, 30 and 45 min at 120 °C. The enzymatic hydrolysis process was performed over 144 h at 50 °C, pH 4.8 with an enzyme loading of 30 FPU (filter paper units). Fermentation was carried out at 37 °C using a strain of Saccharomyces cerevisiae Hansen 1883 (NCYC 366). Results indicated that the optimal conditions for alkaline pretreatment were observed at a 1.5% NaOH, while the best hydrothermal procedure was achieved at 120 °C and 45 min. The impact of these pretreatments was assessed based on the efficiency of enzymatic hydrolysis. The alkaline pretreatment resulted in 81.70% conversion of cellulose to glucose and 96.30% conversion of hemicellulose to xylose. In contrast, the hydrothermal pretreatment achieved 93% cellulose-to-glucose conversion and 83.35% hemicellulose-to-xylose conversion. The ethanol production registered ranged from 13 to 13.23 g·L−1, corresponding to a conversion factor of 0.43 for ethanol from fermentable sugars. Full article
(This article belongs to the Special Issue Alternative Use of Biological Resources)
Show Figures

Figure 1

Figure 1
<p>Relative crystallinity of raw and milled rice husk (≤250 µm). (<b>A</b>) Raw rice husk. (<b>B</b>) Milled rice husk.</p>
Full article ">Figure 2
<p>FTIR spectra of the rice husks treated with NaOH (0.5, 1 and 1.5%).</p>
Full article ">Figure 3
<p>FTIR spectra of rice husks with hydrothermal pretreatment at 120 °C (15, 30 and 45 min of exposition).</p>
Full article ">Figure 4
<p>Klason lignin content after hydrothermal and alkaline pretreatments.</p>
Full article ">Figure 5
<p>Cellobiose formation during the progress of the enzymatic hydrolysis of rice husk with: (<b>a</b>) alkaline pretreatment and (<b>b</b>) hydrothermal pretreatment.</p>
Full article ">Figure 6
<p>Glucose formation during the progress of the enzymatic hydrolysis of rice husk with: (<b>a</b>) alkaline pretreatment and (<b>b</b>) hydrothermal pretreatment.</p>
Full article ">Figure 7
<p>Formation of xylose during the progress of the enzymatic hydrolysis of rice husk with: (<b>a</b>) alkaline pretreatment and (<b>b</b>) hydrothermal pretreatment.</p>
Full article ">Figure 8
<p>Substrate uptake by <span class="html-italic">S. cerevisiae</span> Hansen 1883 (NCYC 366) during fermentation from the hydrolysates pretreated with: (<b>a</b>) alkaline pretreatment and (<b>b</b>) hydrothermal pretreatment.</p>
Full article ">Figure 9
<p>Ethanol concentration produced from hydrolysates pretreated with (<b>a</b>) alkaline pretreatment and (<b>b</b>) hydrothermal.</p>
Full article ">Figure 10
<p>Mass balance of the two process applied: (<b>a</b>) with alkaline pretreatment and (<b>b</b>) with hydrothermal pretreatment.</p>
Full article ">
17 pages, 26558 KiB  
Article
Synergistically Enhanced Enzymatic Hydrolysis of Sugarcane Bagasse Mediated by a Recombinant Endo-Xylanase from Streptomyces ipomoeae
by Zhong Li, Youqing Dong, Junli Liu, Liang Xian, Aixing Tang, Qingyun Li, Qunliang Li and Youyan Liu
Processes 2024, 12(9), 1997; https://doi.org/10.3390/pr12091997 - 16 Sep 2024
Abstract
Xylanase is commonly thought to effectively cooperate with cellulase to promote the bioconversion of lignocellulose. In this study, a novel xylanase, SipoEnXyn10A (Xyn10A), previously identified from Streptomyces ipomoeae, was employed to investigate its synergetic effects on sugarcane bagasse (SCB) transformation. It was [...] Read more.
Xylanase is commonly thought to effectively cooperate with cellulase to promote the bioconversion of lignocellulose. In this study, a novel xylanase, SipoEnXyn10A (Xyn10A), previously identified from Streptomyces ipomoeae, was employed to investigate its synergetic effects on sugarcane bagasse (SCB) transformation. It was shown that the relative increase in reducing sugars reached up to 65%, with enhanced yields of glucose and xylose by 78% and 50%, respectively, in the case of the replacement of cellulase with an equivalent amount of Xyn10A at an enzyme loading of 12.5%. The highest degrees of synergy (DS) for glucose and xylose could reach 2.57 and 1.84. Moreover, the hydrolysis rate increased evidently, and the reaction time to reach the same yield of glucose and xylose was shortened by 72 h and 96 h, respectively. This study on synergistic mechanisms demonstrated that the addition of Xyn10A could cause the destruction of substrates’ morphology and the dissolution of lignin components but could not change the accessibility and crystallinity of substrate cellulose. The joint effect of cellulase and xylanase during the hydrolysis process was thought to result in a synergistic mechanism. Full article
(This article belongs to the Section Chemical Processes and Systems)
Show Figures

Figure 1

Figure 1
<p>Degradation of five different pretreated SCBs by synergizing of Xyn10A with C2730 and S10048. Values are the mean ± SD of three parallel experiments.</p>
Full article ">Figure 2
<p>Effects of (<b>a</b>) pH and (<b>b</b>) temperature on reducing sugar yield and DS of each system. (■) mixed enzyme group, (◆) single cellulase group, (▲) single Xyn10A group, (●) degree of synergy. Values are the mean ± SD of three parallel experiments.</p>
Full article ">Figure 3
<p>Hydrolysis of pretreated SCB using different ratios of cellulase and xylanase. Effect of Xyn10A ratio on (<b>a</b>) glucose yield and (<b>b</b>) DS<sub>glu</sub> of each system were determined under different total enzyme loadings: (hollow) 6 mg/g substrates and (solid) 12 mg/g substrates. For each system, (square) mixed enzyme group, (diamond) single cellulase group, and (round) DS<sub>glu</sub>. Values are the mean ± SD of two parallel experiments.</p>
Full article ">Figure 4
<p>Hydrolysis of pretreated SCB using different ratios of cellulase and xylanase. Effects of Xyn10A ratio on (<b>a</b>) xylose yield and (<b>b</b>) DS<sub>xyl</sub> of each system were determined under different total enzyme loadings: 6 mg/g substrates (hollow) and 12 mg/g substrates (solid). For each system, (square) mixed enzyme group, (diamond) single cellulase group, (triangle) single Xyn10A group, and (round) DS<sub>xyl</sub>. Values are the mean ± SD of two parallel experiments.</p>
Full article ">Figure 5
<p>Monosaccharide yield and its DS during the synergistic degradation process, (<b>a</b>) glucose and (<b>b</b>) xylose. For each system: (■) mixed enzyme group, (◆) single cellulase group, (▲) single Xyn10A group, (●) DS. Values are the mean ± SD of two parallel experiments.</p>
Full article ">Figure 6
<p>Effect of Xyn10A pre-adsorption on (<b>a</b>) glucose and (<b>b</b>) xylose yield. For each system:(square) mixed enzyme group, (diamond) single cellulase group, (triangle) single Xyn10A group. P Xyn10A represents the system with Xyn10A pre-adsorption, while Xyn10A represents the system without Xyn10A pre-adsorption. Values are the mean ± SD of two parallel experiments.</p>
Full article ">Figure 7
<p>X-ray diffraction patterns of different SCB samples. (Red) mixed enzyme group, (blue) commercial cellulase group, (green) single Xyn10A group, (black) control.</p>
Full article ">Figure 8
<p>FTIR spectra of different SCB samples. (Red) mixed enzyme group, (blue) commercial cellulase group, (green) single Xyn10A group, (black) control.</p>
Full article ">
15 pages, 2534 KiB  
Article
Construction of Microbial Consortium to Enhance Cellulose Degradation in Corn Straw during Composting
by Jie Li, Juan Li, Ruopeng Yang, Ping Yang, Hongbo Fu, Yongchao Yang and Chaowei Liu
Agronomy 2024, 14(9), 2107; https://doi.org/10.3390/agronomy14092107 - 16 Sep 2024
Abstract
The improper treatment of crop straw not only leads to resource wastage but also adversely impacts the ecological environment. However, the application of microorganisms can accelerate the decomposition of crop straw and improve its utilization. In this study, cellulose-degrading microbial strains were isolated [...] Read more.
The improper treatment of crop straw not only leads to resource wastage but also adversely impacts the ecological environment. However, the application of microorganisms can accelerate the decomposition of crop straw and improve its utilization. In this study, cellulose-degrading microbial strains were isolated from naturally decayed corn straw and screened using Congo red staining, along with assessing variations in carboxymethyl cellulase (CMCase) activity, filter paper enzyme (FPase) activity and β-glucosidase (β-Gase) activity, as well as the degradation rate. The eight strains, namely Neurospora intermedia isolate 29 (A1), Streptomyces isolate FFJC33 (A2), Gibberella moniliformis isolate FKCB-009 (A3), Fusarium fujikuroi isolate EFS3(2) (A4), Fusarium Fujikuroi isolate FZ04 (A5), Lysine bacillus macroides strain LNHL43 (B1), Bacillus subtilis strain MPF30 (B2) and Paenibacilli lautus strain ALEB-P1 (C), were identified and selected for microbial strain consortium design based on their high activities of CMCase, FPase and β-Gase. The fungi, bacteria and actinomycete strains were combined without antagonistic effects on corn straw decomposition. The results showed the A2B2 combination had a significantly higher FPase at 55.44 U/mL and β-Gase at 25.73 U/mL than the other two strain combinations (p < 0.05). Additionally, the degradation rate of this combination was 40.33%, which was considerably higher than that of the other strains/consortia. The strain combination A4B2C also had superior enzyme activity, including CMCase with a value of 35.03 U/mL, FPase with a value of 63.59 U/mL and β-Gase with a value of 26.15 U/mL, which were significantly different to those of the other three strain combinations (p < 0.05). Furthermore, seven single microbial strains with high cellulase activities were selected to construct various microbial consortiums for in situ composting in order to evaluate their potential. Taken as a whole, the results of composting, including temperature, moisture content, pH, E4/E6 value and seed germination index, indicated that the microbial strain consortium consisting of Neurospora intermediate isolate 29, Fusarium fujikuroi isolate EFS3(2), Fusarium fujikuroi isolate FZ04, Lysinibacillus macrolides, Lysinibacillus sphaericus, Bacillus subtilis and Paenibacillus lautus was advantageous for corn straw decomposition and yielded high-quality compost. The screened flora was able to effectively degrade corn straw. This study provides a novel solution for the construction of a microbial consortium for the composting of corn straw. Full article
(This article belongs to the Section Agricultural Biosystem and Biological Engineering)
Show Figures

Figure 1

Figure 1
<p>(<b>A</b>) Enzyme activity of single strains. (<b>B</b>) Degradation rate of corn straw using single strain. (<b>C</b>) Enzyme activity of combination of two microbial strains. (<b>D</b>) Enzyme activity of combination of three microbial strains. Different lowercase letters indicate significant differences between different treatments within the same enzyme.</p>
Full article ">Figure 2
<p>Hydrolytic circles of fungi, bacteria and actinomycetes.</p>
Full article ">Figure 3
<p>(<b>A</b>) Degradation rate of corn straw using two-microbial-strain combinations. (<b>B</b>) Degradation rate of corn straw using three-microbial-strain combinations. (<b>C</b>) pH changes of corn straw using two-microbial-strain combinations. (<b>D</b>) pH changes of corn straw using three-microbial-strain combinations.</p>
Full article ">Figure 4
<p>(<b>A</b>) Variation of pile and ambient temperature. (<b>B</b>) Moisture content variation of the pile. (<b>C</b>) pH value variation of the pile. (<b>D</b>) E4/E6 value variation of the pile. (<b>E</b>) Seed germination index variation of the pile.</p>
Full article ">
13 pages, 5151 KiB  
Article
Direct Conversion of Minimally Pretreated Corncob by Enzyme-Intensified Microbial Consortia
by Alei Geng, Nana Li, Anaiza Zayas-Garriga, Rongrong Xie, Daochen Zhu and Jianzhong Sun
Agriculture 2024, 14(9), 1610; https://doi.org/10.3390/agriculture14091610 - 14 Sep 2024
Viewed by 212
Abstract
The presence of diverse carbohydrate-active enzymes (CAZymes) is crucial for the direct bioconversion of lignocellulose. In this study, various anaerobic microbial consortia were employed for the degradation of 10 g/L of minimally pretreated corncob. The involvement of lactic acid bacteria (LAB) and a [...] Read more.
The presence of diverse carbohydrate-active enzymes (CAZymes) is crucial for the direct bioconversion of lignocellulose. In this study, various anaerobic microbial consortia were employed for the degradation of 10 g/L of minimally pretreated corncob. The involvement of lactic acid bacteria (LAB) and a CAZyme-rich bacterium (Bacteroides cellulosilyticus or Paenibacillus lautus) significantly enhanced the lactic acid production by Ruminiclostridium cellulolyticum from 0.74 to 2.67 g/L (p < 0.01), with a polysaccharide conversion of 67.6%. The supplement of a commercial cellulase cocktail, CTec 2, into the microbial consortia continuously promoted the lactic acid production to up to 3.35 g/L, with a polysaccharide conversion of 80.6%. Enzymatic assays, scanning electron microscopy, and Fourier transform infrared spectroscopy revealed the substantial functions of these CAZyme-rich consortia in partially increasing enzyme activities, altering the surface structure of biomass, and facilitating substrate decomposition. These results suggested that CAZyme-intensified consortia could significantly improve the levels of bioconversion of lignocellulose. Our work might shed new light on the construction of intensified microbial consortia for direct conversion of lignocellulose. Full article
(This article belongs to the Section Agricultural Technology)
Show Figures

Figure 1

Figure 1
<p>Chemical productions on 10 g/L minimally pretreated corncob by a series of microbial consortia after 3 days of incubation. For the species composition of the consortia, refer to <a href="#agriculture-14-01610-t001" class="html-table">Table 1</a>. Control indicates the chemical composition after inoculation but without incubation. Values are the mean of three replicates. Error bars indicate standard deviation.</p>
Full article ">Figure 2
<p>Heatmap of the extracellular enzyme activities of different microbial consortia. <sup>a</sup> Crude enzymes from the culture grown on wheat bran and <sup>b</sup> crude enzymes from the culture grown on corncob. Data were normalized in a row. For the species composition of the consortia, refer to <a href="#agriculture-14-01610-t001" class="html-table">Table 1</a>. Values are the mean of three replicates.</p>
Full article ">Figure 3
<p>SEM of the corncob before and after treatment by a series of microbial consortia. (<b>A</b>) Blank control; (<b>B</b>) consortium A (<span class="html-italic">R. cellulolyticum</span> alone); (<b>C</b>) consortium C; (<b>D</b>) consortium D; (<b>E</b>) consortium E; and (<b>F</b>) consortium F. For the species composition of the consortia, refer to <a href="#agriculture-14-01610-t001" class="html-table">Table 1</a>. The images were taken at 2.0 kv, the magnification was set at 2000 times, and the scale bars in the insertion are 4 μm.</p>
Full article ">Figure 4
<p>FTIR analysis of the corncob before and after treatment by the microbial consortia. Letters on the graph refer to the treatment by the specific consortia. The assignment of the bands is shown in <a href="#app1-agriculture-14-01610" class="html-app">Table S2</a>.</p>
Full article ">Figure 5
<p>Microbial population structures of different consortia during various time spans. (<b>A</b>) At the start point; (<b>B</b>) after 1 day; (<b>C</b>) after 2 days; and (<b>D</b>) after 3 days of incubation (end of fermentation). Letters on the horizontal axis refer to the specific consortia, and different species are distinguished by color.</p>
Full article ">Figure 6
<p>Chemical productions on 10 g/L minimally pretreated corncob by enzyme-enhanced microbial consortia after 3 days of incubation. For the species composition of the consortia, refer to <a href="#agriculture-14-01610-t001" class="html-table">Table 1</a>. Values are the mean of three replicates. Error bars indicate standard deviation.</p>
Full article ">
13 pages, 12746 KiB  
Article
Characterization and Pathogenicity of Colletotrichum truncatum Causing Hylocereus undatus Anthracnose through the Changes of Cell Wall-Degrading Enzymes and Components in Fruits
by Shuwu Zhang, Yun Liu, Jia Liu, Enchen Li and Bingliang Xu
J. Fungi 2024, 10(9), 652; https://doi.org/10.3390/jof10090652 - 13 Sep 2024
Viewed by 311
Abstract
Anthracnose is one of the destructive diseases of pitaya that seriously affects the plant growth and fruit quality and causes significant yield and economic losses worldwide. However, information regarding the species of pathogens that cause anthracnose in pitaya (Hylocereus undatus) fruits [...] Read more.
Anthracnose is one of the destructive diseases of pitaya that seriously affects the plant growth and fruit quality and causes significant yield and economic losses worldwide. However, information regarding the species of pathogens that cause anthracnose in pitaya (Hylocereus undatus) fruits in Gansu Province, China, and its pathogenic mechanism is unknown. Thus, the purposes of our present study were to identify the species of pathogens causing H. undatus fruits anthracnose based on the morphological and molecular characteristics and determine its pathogenic mechanism by physiological and biochemical methods. In our present study, forty-six isolates were isolated from the collected samples of diseased H. undatus fruits and classified as three types (named as H-1, H-2, and H-3), according to the colony and conidium morphological characteristics. The isolation frequencies of H-1, H-2, and H-3 types were 63.04%, 21.74%, and 15.22%, respectively. The representative single-spore isolate of HLGTJ-1 in H-1 type has significant pathogenicity, and finally we identified Colletotrichum truncatum as the pathogen based on the morphological characteristics as well as multi-locus sequence analysis. Moreover, the H. undatus fruits inoculated with C. truncatum had a significantly increased activity of cell wall-degrading enzymes (CWDEs) cellulase (Cx), β-glucosidase (β-Glu), polygalacturonase (PG), and pectin methylgalacturonase (PMG), while having a decreased level of cell wall components of original pectin and cellulose in comparison to control. The average increased activities of Cx, β-Glu, PG, and PMG were 30.73%, 40.40%, 51.55%, and 32.23% from day 0 to 6 after inoculation, respectively. In contrast, the average decreased contents of original pectin and cellulose were 1.82% and 16.47%, respectively, whereas the average increased soluble pectin content was 38.31% in comparison to control. Our results indicate that C. truncatum infection increased the activities of CWDEs in H. undatus fruits to disassemble their cell wall components, finally leading to the fruits’ decay and deterioration. Thus, our findings will provide significant evidence in the controlling of pitaya anthracnose in the future. Full article
(This article belongs to the Special Issue Control of Postharvest Fungal Diseases)
Show Figures

Figure 1

Figure 1
<p>Symptoms of <span class="html-italic">Hylocereus undatus</span> fruit anthracnose at different time periods in Wuwei city, China. (<b>A</b>) and (<b>B</b>): the symptoms at the initial stage; (<b>C</b>–<b>F</b>): the symptoms at a later stage.</p>
Full article ">Figure 2
<p>The pathogenicity test of the representative isolate of HLGTJ-1 on <span class="html-italic">Hylocereus undatus</span> fruit after inoculation. (<b>A</b>) Fruit inoculation with the PDA discs without the HLGTJ-1 isolate (control); (<b>B</b>,<b>C</b>) fruits inoculation with the mycelial discs of the HLGTJ-1 isolate at 3 and 7 days after inoculation, respectively.</p>
Full article ">Figure 3
<p>Morphological characteristics of the representative isolate of HLGTJ-1. (<b>A</b>,<b>B</b>) The front and reverse views of the colony, respectively; (<b>C</b>) the conidial mass produced on the front of the colony; (<b>D</b>) the conidial mass observed under a stereoscope; (<b>E</b>) the acervuli and setae observed under a stereoscope; (<b>F</b>) the acervulus and setae observed under a microscope using the hand-sliced method; (<b>G</b>,<b>H</b>) conidia; (<b>I</b>,<b>J</b>) appressorium.</p>
Full article ">Figure 4
<p>Multi-locus phylogenetic tree of the single-spore isolate of HLGTJ-1 based on the combined sequences (HJ-ITS region and <span class="html-italic">HJ-GAPDH</span> and <span class="html-italic">HJ-HIS3</span> genes) by the maximum likelihood (ML) method. Bootstraps supporting values higher than 50% from the 1000 replicates are presented at the nodes.</p>
Full article ">Figure 5
<p>Changes in activities of Cx (<b>A</b>), β-Glu (<b>B</b>), PG (<b>C</b>), and PMG (<b>D</b>) in <span class="html-italic">Hylocereus undatus</span> fruits at different days after inoculation with the isolate. Different letters in Figure are significantly different at <span class="html-italic">p</span> &lt; 0.05.</p>
Full article ">Figure 6
<p>Changes in original pectin (<b>A</b>), soluble pectin (<b>B</b>), and cellulose (<b>C</b>) contents in <span class="html-italic">Hylocereus undatus</span> fruits at different days after inoculation with the isolate. Different letters in Figure are significantly different at <span class="html-italic">p</span> &lt; 0.05.</p>
Full article ">
13 pages, 3636 KiB  
Article
Improving the Catalytic Efficiency of an AA9 Lytic Polysaccharide Monooxygenase MtLPMO9G by Consensus Mutagenesis
by Yao Meng, Wa Gao, Xiaohua Liu, Tang Li, Kuikui Li and Heng Yin
Catalysts 2024, 14(9), 614; https://doi.org/10.3390/catal14090614 - 12 Sep 2024
Viewed by 233
Abstract
Cellulose is one of the most abundant renewable resources in nature. However, its recalcitrant crystalline structure hinders efficient enzymatic depolymerization. Unlike cellulases, lytic polysaccharide monooxygenases (LPMOs) can oxidatively cleave glycosidic bonds in the crystalline regions of cellulose, playing a crucial role in its [...] Read more.
Cellulose is one of the most abundant renewable resources in nature. However, its recalcitrant crystalline structure hinders efficient enzymatic depolymerization. Unlike cellulases, lytic polysaccharide monooxygenases (LPMOs) can oxidatively cleave glycosidic bonds in the crystalline regions of cellulose, playing a crucial role in its enzymatic depolymerization. An AA9 LPMO from Myceliophthora thermophila was previously identified and shown to exhibit a highly efficient catalytic performance. To further enhance its catalytic efficiency, consensus mutagenesis was applied. Compared with the wild-type enzyme, the oxidative activities of mutants A165S and P167N increased by 1.8-fold and 1.4-fold, respectively, and their catalytic efficiencies (kcat/Km) improved by 1.6-fold and 1.2-fold, respectively. The mutants also showed significantly enhanced activity in the synergistic degradation of cellulose with cellobiohydrolase. Additionally, the P167N mutant exhibited better H2O2 tolerance. A molecular dynamics analysis revealed that the increased activity of mutants A165S and P167N was due to the closer proximity of the active center to the substrate post-mutation. This study demonstrates that selecting appropriate mutation sites via a semi-rational design can significantly improve LPMO activity, providing valuable insights for the protein engineering of similar enzymes. Full article
(This article belongs to the Section Biocatalysis)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>(<b>A</b>) PCR identification of the transformants of <span class="html-italic">Mt</span>LPMO9G mutants. As the intact linearized plasmid destroyed the original AOX1 gene during the insertion of the gene and, at the same time, formed a new AOX1 gene, a fragment of the target gene (1500 bp) and a fragment of the AOX1 gene of <span class="html-italic">Pichia pastoris</span> (2000 bp) appeared at the same time in the electrophoretic detection. (<b>B</b>) SDS-PAGE identification of the purified <span class="html-italic">Mt</span>LPMO9G mutants. M: protein marker. Lane 1, mutant A165S; lane 2, mutant N166G; lane 3, mutant P167N. The presence of the single band at 43 kDa corresponds with the target protein, indicating that purified proteins were obtained.</p>
Full article ">Figure 2
<p>The activity analysis of <span class="html-italic">Mt</span>LPMO9G and its mutants. (<b>A</b>) HPLC analysis of the product profile of <span class="html-italic">Mt</span>LPMO9G and its mutants on PASC; (<b>B</b>) quantification of cellobionic acid production following the further treatment of soluble LPMO products by CBH I. Compared with the WT (black), the activity of mutant A165S (blue) increased by about 78%, the activity of mutant P167N (red) increased by about 43%, and there was no significant change in mutant N166G (green). The data are presented as the mean ± standard deviation. * <span class="html-italic">p</span> &lt; 0.05 and ** <span class="html-italic">p</span> &lt; 0.01.</p>
Full article ">Figure 3
<p>Comparison of the thermal stability of <span class="html-italic">Mt</span>LPMO9G and its mutants. As shown in the table in the figure, the <span class="html-italic">Tm</span> value of mutant A165S (blue) was reduced by 0.37 °C and the <span class="html-italic">Tm</span> value of mutant P167N (red) was reduced by 7.19 °C compared with the WT (black).</p>
Full article ">Figure 4
<p>Comparison of the synergy between <span class="html-italic">Mt</span>LPMO9G and its mutants and CBH II. The synergistic effect of mutants A165S (blue) and P167N (red) with CBH was significantly increased when the ratio of LPMO to CBH II was 1:10 and 10:1 compared with the WT (black). The data are presented as the mean ± standard deviation. ** <span class="html-italic">p</span> &lt; 0.01, and *** <span class="html-italic">p</span> &lt; 0.001.</p>
Full article ">Figure 5
<p>Comparison of the substrate-binding ability of <span class="html-italic">Mt</span>LPMO9G and its mutants. The <span class="html-italic">B<sub>max</sub></span> was calculated to be increased in mutant P167N (red) compared with the WT (black).</p>
Full article ">Figure 6
<p>Comparison of H<sub>2</sub>O<sub>2</sub> tolerance ability of <span class="html-italic">Mt</span>LPMO9G and its mutants. (<b>A</b>) Reaction after 5 h; (<b>B</b>) reaction after 12 h; (<b>C</b>) reaction after 24 h. At 24 h after the reaction began, the H<sub>2</sub>O<sub>2</sub> tolerance of mutant P167N (red) showed a significant increase compared with the WT (black) under the addition of 100 μM H<sub>2</sub>O<sub>2</sub>. The data are presented as the mean ± standard deviation. ** <span class="html-italic">p</span> &lt; 0.01.</p>
Full article ">Figure 7
<p>Michaelis–Menten kinetics of <span class="html-italic">Mt</span>LPMO9G and its mutants.</p>
Full article ">Figure 8
<p>Interactions of <span class="html-italic">Mt</span>LPMO9G and its mutants with cellulose probed by MD simulations. (<b>A</b>) Overview of <span class="html-italic">Mt</span>LPMO9G (green) interacting with cellulose (pink) at t = 20 ns, where the addition of sugar chains to LPMOs was better able to measure the interaction of LPMOs with cellulose; (<b>B</b>) distance between Cu(II) in the active center and the substrate plane during 20 ns of the simulation.</p>
Full article ">
19 pages, 3230 KiB  
Article
The Phylogenomic Characterization of Planotetraspora Species and Their Cellulases for Biotechnological Applications
by Noureddine Bouras, Mahfoud Bakli, Guendouz Dif, Slim Smaoui, Laura Șmuleac, Raul Paşcalău, Esther Menendez and Imen Nouioui
Genes 2024, 15(9), 1202; https://doi.org/10.3390/genes15091202 - 12 Sep 2024
Viewed by 279
Abstract
This study aims to evaluate the in silico genomic characteristics of five species of the genus Planotetraspora: P. kaengkrachanensis, P. mira, P. phitsanulokensis, P. silvatica, and P. thailandica, with a view to their application in therapeutic research. [...] Read more.
This study aims to evaluate the in silico genomic characteristics of five species of the genus Planotetraspora: P. kaengkrachanensis, P. mira, P. phitsanulokensis, P. silvatica, and P. thailandica, with a view to their application in therapeutic research. The 16S rRNA comparison indicated that these species were phylogenetically distinct. Pairwise comparisons of digital DNA-DNA hybridization (dDDH) and OrthoANI values between these studied type strains indicated that dDDH values were below 62.5%, while OrthoANI values were lower than 95.3%, suggesting that the five species represent distinct genomospecies. These results were consistent with the phylogenomic study based on core genes and the pangenome analysis of these five species within the genus Planotetraspora. However, the genome annotation showed some differences between these species, such as variations in the number of subsystem category distributions across whole genomes (ranging between 1979 and 2024). Additionally, the number of CAZYme (Carbohydrate-Active enZYme) genes ranged between 298 and 325, highlighting the potential of these bacteria for therapeutic research applications. The in silico physico-chemical characteristics of cellulases from Planotetraspora species were analyzed. Their 3D structure was modeled, refined, and validated. A molecular docking analysis of this cellulase protein structural model was conducted with cellobiose, cellotetraose, laminaribiose, carboxymethyl cellulose, glucose, and xylose ligand. Our study revealed significant interaction between the Planotetraspora cellulase and cellotetraose substrate, evidenced by stable binding energies. This suggests that this bacterial enzyme holds great potential for utilizing cellotetraose as a substrate in various applications. This study enriches our understanding of the potential applications of Planotetraspora species in therapeutic research. Full article
(This article belongs to the Section Microbial Genetics and Genomics)
Show Figures

Figure 1

Figure 1
<p>Phylogenomic tree based on genome sequences in the TYGS tree inferred with FastME 2.1.6.1 [<a href="#B16-genes-15-01202" class="html-bibr">16</a>] from the Genome BLAST Distance Phylogeny approach (GBDP); distances calculated from genome sequences. The branch lengths are scaled in terms of GBDP distance formula d5. The numbers above branches are GBDP pseudo-bootstrap support values &gt;70% from 100 replications. The tree was rooted at the midpoint [<a href="#B13-genes-15-01202" class="html-bibr">13</a>]. The different color indicates the different species cluster.</p>
Full article ">Figure 2
<p>Maximum likelihood core gene phylogenomic tree. The core genes were identified using the Roary program [<a href="#B20-genes-15-01202" class="html-bibr">20</a>] and MEGA software version 11 [<a href="#B21-genes-15-01202" class="html-bibr">21</a>] with 1000 bootstrap replications to assess statistical support. This illustrates the evolutionary relationship between the species of <span class="html-italic">Planotetraspora</span>. <span class="html-italic">Nocardiopsis algeriensis</span> CECT 8712<sup>T</sup> was used as an outgroup. Bar 0.02 nucleotide substitution per site.</p>
Full article ">Figure 3
<p>Pangenome analysis of five species of genomes of <span class="html-italic">Planotetraspora</span> was conducted using Roary [<a href="#B20-genes-15-01202" class="html-bibr">20</a>]. Matrix destitution of genes across the pangenome of all <span class="html-italic">Planotetraspora</span> species (from Rotary), and <span class="html-italic">Nocardiopsis algeriensis</span> CECT 8712<sup>T</sup> was used as an outgroup. Pangenome visualization is displayed as presence (blue) and absence (white) output using Phandango [<a href="#B24-genes-15-01202" class="html-bibr">24</a>].</p>
Full article ">Figure 4
<p>The SWISS−MODEL generated the 3D and refined structure of <span class="html-italic">Planotetraspora</span> cellulase enzymes and its validation. (<b>A</b>) This structural model was refined by ModRefiner and visualized by PyMOL. α-helix (chocolate), β-sheet (cyan), and loop (yellow). (<b>B</b>) Ramachandran plot for the refined cellulase obtained by PROCHECK. The residues found in favored (A, B, and L), additional allowed (a, b, l, and p), generously allowed (~a, ~b, ~l, and ~p), and disallowed regions are delineated with red, yellow, beige, and white color coding, respectively. All non-glycine and non-proline residues are depicted as filled black squares, while glycines (non−terminal) are represented as filled black triangles.</p>
Full article ">Figure 5
<p>Molecular docking analysis of the cellotetraose ligand with the <span class="html-italic">Planotetraspora</span> cellulase enzyme receptor. (<b>A</b>) The cartoon representation of the protein receptor and its ligand. (<b>B</b>) Cavity illustration showed the region on the protein receptor surface where bonding occurs with the ligand. (<b>C</b>) Vina score and cavity size models table displayed the Vina score of this receptor and the ligand docking with the best highlighted score. The enzyme–ligand interactions were visualized, employing the ball and stick style for the ligand and the surface style for the receptor. Ligand and receptor colors were configured based on elements and B-factor, respectively. (<b>D</b>) Interaction residues depicted the interacting residues of the receptor with the ligand as visualized in BIOVIA Discovery Studio Visualizer. (<b>E</b>) 2D diagram of residue interaction types between ligand and receptor as visualized in BIOVIA Discovery Studio Visualizer.</p>
Full article ">
12 pages, 2976 KiB  
Article
Improved Sugar Recovery from Mandarin Peel under Optimal Enzymatic Hydrolysis Conditions and Application to Bioethanol Production
by Hyerim Son, Jeongho Lee and Hah Young Yoo
Processes 2024, 12(9), 1960; https://doi.org/10.3390/pr12091960 - 12 Sep 2024
Viewed by 284
Abstract
Mandarin peel (MP) has gained attention as a feedstock for flavonoid recovery via the extraction process based on the biorefinery concept, but residues remain after the extraction. Toward an integrated biorefinery concept, this study aimed to valorize extracted MP (eMP) by using it [...] Read more.
Mandarin peel (MP) has gained attention as a feedstock for flavonoid recovery via the extraction process based on the biorefinery concept, but residues remain after the extraction. Toward an integrated biorefinery concept, this study aimed to valorize extracted MP (eMP) by using it in bioethanol production. For efficient fermentable sugar production, the effect of enzymatic hydrolysis conditions on sugar conversion from eMP was investigated, and the results showed that combining cellulase and cellobiase resulted in a higher enzymatic glucose conversion (78.2%) than the use of the individual enzymes (37.5% and 45.6%). Pectinase played an essential role in enhancing enzymatic arabinose conversion, and the optimal conditions were determined to be pH 4 and 90 units of the three enzymes. Under optimal conditions, the sugar yield was 199 g glucose and 47 g arabinose/kg eMP, and the hydrolysate was used in bioethanol fermentation. The results showed that the bioethanol production was 3.78 g/L (73.9% yield), similar to the control medium (3.79 g/L; 74.2% yield), although the cell growth of the yeast was slightly delayed in the eMP hydrolysate medium. This study highlights the potential of eMP as a low-cost feedstock for sugar and bioethanol production. Full article
(This article belongs to the Special Issue Platform Chemicals and Novel Materials from Biomass)
Show Figures

Figure 1

Figure 1
<p>The effects of enzyme types and their combination on enzymatic glucose and arabinose conversion from extracted mandarin peel (CL, cellulase; CB, cellobiase; P, pectinase). The data are presented as mean values. Error bars indicate standard deviations (<span class="html-italic">n</span> = 3).</p>
Full article ">Figure 2
<p>Effect of pH on enzymatic glucose and arabinose conversion from extracted mandarin peel. The data are presented as mean values. Error bars indicate standard deviations (<span class="html-italic">n</span> = 3).</p>
Full article ">Figure 3
<p>The effect of enzyme loading on glucose and arabinose conversion from extracted mandarin peel. The data are presented as mean values. Error bars indicate standard deviations (<span class="html-italic">n</span> = 3).</p>
Full article ">Figure 4
<p>Bioethanol fermentation by <span class="html-italic">Saccharomyces cerevisiae</span> using extracted mandarin peel hydrolysates. The data are presented as mean values. Error bars indicate standard deviations (<span class="html-italic">n</span> = 3).</p>
Full article ">Figure 5
<p>Mass balance of extracted mandarin peel to produce sugars and bioethanol through saccharification and fermentation processes.</p>
Full article ">Figure A1
<p>Effect of pH on cellulase (<b>a</b>), cellobiase (<b>b</b>), and pectinase (<b>c</b>) activity.</p>
Full article ">
15 pages, 2556 KiB  
Article
Optimization of Consolidated Bioprocessing Fermentation of Uncooked Sweet Potato Residue for Bioethanol Production by Using a Recombinant Amylolytic Saccharomyces cerevisiae Strain via the Orthogonal Experimental Design Method
by Xin Wang, Chenchen Gou, Haobo Zheng, Na Guo, Yanling Li, Aimei Liao, Na Liu, Hailong Tian and Jihong Huang
Fermentation 2024, 10(9), 471; https://doi.org/10.3390/fermentation10090471 - 12 Sep 2024
Viewed by 346
Abstract
An amylolytic industrial yeast strain named 1974-GA-temA, developed previously by our research team by coexpressing the α-amylase and glucoamylase genes, combines enzyme production, sweet potato residue (SPR) hydrolysis, and glucose fermentation into ethanol in a one-step process. This consolidated bioprocessing (CBP) method has [...] Read more.
An amylolytic industrial yeast strain named 1974-GA-temA, developed previously by our research team by coexpressing the α-amylase and glucoamylase genes, combines enzyme production, sweet potato residue (SPR) hydrolysis, and glucose fermentation into ethanol in a one-step process. This consolidated bioprocessing (CBP) method has great application potential in the commercial production of bioethanol from SPR, but important fermentation parameters should be optimized to further increase the ethanol concentration and yield. In this study, the effects of the initial fermentation pH, solid-to-liquid ratio, inoculation volume, addition of exogenous enzyme, and supplementation with metal ions were systemically investigated. Single-factor experiments revealed that the optimal pH was 4.0. In the solid-to-liquid ratio test, an increase in the solid-to-liquid ratio corresponded with a gradual increase in the ethanol concentration, peaking at 1:5. However, the ethanol yield gradually decreased, with the optimal solid-to-liquid ratio identified as 1:5. The ethanol concentration and yield reached 9.73 g/L and 5.84%, respectively. Additionally, an increase in the inoculum size resulted in increased ethanol concentration and yield, with the optimal inoculum level determined to be 10%. An ethanol concentration of 7.87 g/L was attained under these specified conditions, equating to an ethanol yield of 4.72%. Further analysis was conducted to assess the effects of exogenous cellulase, hemicellulase, and pectinase, both individually and in combination, on ethanol concentration and yield. The results indicated that pectinase had a particularly significant effect. The highest ethanol concentration was observed when all three enzymes were administered concurrently, yielding 27.27 g/L ethanol. Then, the role of metal ions in SPR fermentation was evaluated. The metal ions did not significantly affect the process, with the exception of copper ions. The addition of copper ions at a specific concentration of 0.2 g/100 g SPR increased the ethanol concentration. However, concentrations exceeding 0.2 g/100 g SPR inhibited yeast cell growth. Finally, orthogonal optimization was employed to determine the optimal combination of factors: pH, 4.0; solid-to-liquid ratio, 1:6; inoculation volume, 10%; cellulase and pectinase addition; and the absence of Cu2+ addition. Under these conditions, strain 1974-GA-temA produced 34.83 ± 0.62 g/L ethanol after 8 days of fermentation, corresponding to a 20.90% ± 0.37% ethanol yield. This value markedly exceeds the outcomes of all the conducted orthogonal experiments. The fermentation optimization experiments in this study are expected to increase ethanol production during the CBP fermentation of SPR. Full article
(This article belongs to the Section Fermentation Process Design)
Show Figures

Figure 1

Figure 1
<p>Effect of initial pH on ethanol production and yield from SPR fermentation. Error bars represent the standard deviation from the mean of three replicates. Letters a–c: Significant difference in relation to ethanol concentration or ethanol yield in pH 4.0 (<span class="html-italic">p</span> &lt; 0.01).</p>
Full article ">Figure 2
<p>Effects of the solid-to-liquid ratio on ethanol production and yield from SPR fermentation. Error bars represent the standard deviation from the mean of three replicates. Letters a–e: Significant difference in relation to ethanol concentration in solid-to-liquid ratio of 1:3 or ethanol yield in solid-to-liquid ratio of 1:8 (<span class="html-italic">p</span> &lt; 0.01).</p>
Full article ">Figure 3
<p>Effects of inoculation volume on ethanol production and yield from SPR fermentation. Error bars represent the standard deviation from the mean of three replicates. Letters a–c: Significant difference in relation to ethanol concentration or ethanol yield in inoculum size of 10% (<span class="html-italic">p</span> &lt; 0.01).</p>
Full article ">Figure 4
<p>Impact of both combined and individual additions of cellulase, hemicellulase, and pectinase on ethanol production and yield during the CBP fermentation of SPR by strain 1974-GA-temA. Error bars represent the standard deviation from the mean of three replicates. Letters a–c: Significant difference in relation to ethanol concentration or ethanol yield in condition of C + H + P (<span class="html-italic">p</span> &lt; 0.01). ctl: without enzyme addition; C: cellulase; H: hemicellulase; P: pectinase.</p>
Full article ">Figure 5
<p>Impact of metal ions on the production of ethanol from the CBP fermentation of SPR by strain 1974-GA-temA. The influence of CuSO<sub>4</sub> (<b>a</b>), FeSO<sub>4</sub> (<b>b</b>), ZnSO<sub>4</sub> (<b>c</b>), MgSO<sub>4</sub> (<b>d</b>), NaCl (<b>e</b>), KCl (<b>f</b>), CaCl<sub>2</sub> (<b>g</b>), and MnCl<sub>2</sub> (<b>h</b>) were shown. Error bars represent the standard deviation from the mean of three replicates. The letters on the bar graphs represent the results of significance analyses (<span class="html-italic">p</span> &lt; 0.01).</p>
Full article ">Figure 6
<p>Ethanol production from optimized SPR CBP fermentation by strain 1974-GA-temA. The data represent the means of three repeats with standard deviations.</p>
Full article ">
14 pages, 3381 KiB  
Article
Mass Transfer Resistance and Reaction Rate Kinetics for Carbohydrate Digestion with Cell Wall Degradation by Cellulase
by Yongmei Sun, Shu Cheng, Jingying Cheng and Timothy A. G. Langrish
Foods 2024, 13(18), 2881; https://doi.org/10.3390/foods13182881 - 11 Sep 2024
Viewed by 425
Abstract
This paper introduces an enzymatic approach to estimate internal mass-transfer resistances during food digestion studies. Cellulase has been used to degrade starch cell walls (where cellulose is a significant component) and reduce the internal mass-transfer resistance, so that the starch granules are released [...] Read more.
This paper introduces an enzymatic approach to estimate internal mass-transfer resistances during food digestion studies. Cellulase has been used to degrade starch cell walls (where cellulose is a significant component) and reduce the internal mass-transfer resistance, so that the starch granules are released and hydrolysed by amylase, increasing the starch hydrolysis rates, as a technique for measuring the internal mass-transfer resistance of cell walls. The estimated internal mass-transfer resistances for granular starch hydrolysis in a beaker and stirrer system for simulating the food digestion range from 2.2 × 107 m−1 s at a stirrer speed of 100 rpm to 6.6 × 107 m−1 s at 200 rpm. The reaction rate constants for cellulase-treated starch are about three to eight times as great as those for starch powder. The beaker and stirrer system provides an in vitro model to quantitatively understand external mass-transfer resistance and compare mass-transfer and reaction rate kinetics in starch hydrolysis during food digestion. Particle size analysis indicates that starch cell wall degradation reduces starch granule adhesion (compared with soaked starch samples), though the primary particle sizes are similar, and increases the interfacial surface area, reducing internal mass-transfer resistance and overall mass-transfer resistance. Dimensional analysis (such as the Damköhler numbers, Da, 0.3–0.5) from this in vitro system shows that mass-transfer rates are greater than reaction rates. At the same time, SEM (scanning electron microscopy) images of starch particles indicate significant morphology changes due to the cell wall degradation. Full article
(This article belongs to the Special Issue Enzymes' Chemistry in Food)
Show Figures

Figure 1

Figure 1
<p>Diagram of mass-transfer resistance change for starch hydrolysis. (<b>A</b>) Raw starch; (<b>B</b>) Cellulase and raw starch.The plant cell wall affects the internal mass-transfer resistance. Degraded plant cell walls (cellulose) in sample (<b>B</b>) reduce the internal mass-transfer resistance.</p>
Full article ">Figure 2
<p>SEM images for the morphological analysis of (<b>a</b>) starch powder vs. (<b>b</b>) cellulase-treated starch (cellulolysis).</p>
Full article ">Figure 3
<p>Evidence of starch cell wall degradation. Yellowish starch after one hour of cell wall degradation, but the yellow colour was almost gone after four hours of degradation.</p>
Full article ">Figure 4
<p>Starch hydrolysis before and after cell wall degradation (cellulose hydrolysis).</p>
Full article ">Figure 5
<p>Cumulative particle size distributions for starch samples (soaking and degradation).</p>
Full article ">Figure 6
<p>Particle size analysis for starch samples (soaking and degradation).</p>
Full article ">Figure 7
<p>Morphology analysis of the changes in starch from soaking and cell-wall degradation. Samples were dried in a laboratory drying oven at 50 °C for 24 h.</p>
Full article ">Figure A1
<p>The slope of the glucose concentration–time curve for cellulase-treated starch hydrolysis according to <a href="#foods-13-02881-f004" class="html-fig">Figure 4</a>. The <span class="html-italic">X</span>-axis is the reaction time (h), and the <span class="html-italic">Y</span>-axis is the glucose concentration (mg/mL). The stirrer speed is 100 rpm.</p>
Full article ">
15 pages, 5040 KiB  
Article
Transcriptome Analysis Reveals the Effect of Oyster Mushroom Spherical Virus Infection in Pleurotus ostreatus
by Yifan Wang, Junjie Yan, Guoyue Song, Zhizhong Song, Matthew Shi, Haijing Hu, Lunhe You, Lu Zhang, Jianrui Wang, Yu Liu, Xianhao Cheng and Xiaoyan Zhang
Int. J. Mol. Sci. 2024, 25(17), 9749; https://doi.org/10.3390/ijms25179749 - 9 Sep 2024
Viewed by 305
Abstract
Oyster mushroom spherical virus (OMSV) is a mycovirus that inhibits mycelial growth, induces malformation symptoms, and decreases the yield of fruiting bodies in Pleurotus ostreatus. However, the pathogenic mechanism of OMSV infection in P. ostreatus is poorly understood. In this study, RNA [...] Read more.
Oyster mushroom spherical virus (OMSV) is a mycovirus that inhibits mycelial growth, induces malformation symptoms, and decreases the yield of fruiting bodies in Pleurotus ostreatus. However, the pathogenic mechanism of OMSV infection in P. ostreatus is poorly understood. In this study, RNA sequencing (RNA-seq) was conducted, identifying 354 differentially expressed genes (DEGs) in the mycelium of P. ostreatus during OMSV infection. Verifying the RNA-seq data through quantitative real-time polymerase chain reaction on 15 DEGs confirmed the consistency of gene expression trends. Both Gene Ontology and Kyoto Encyclopedia of Genes and Genomes analyses highlighted the pivotal role of primary metabolic pathways in OMSV infection. Additionally, significant changes were noted in the gene expression levels of carbohydrate-active enzymes (CAZymes), which are crucial for providing the carbohydrates needed for fungal growth, development, and reproduction by degrading renewable lignocellulose. The activities of carboxymethyl cellulase, laccase, and amylase decreased, whereas chitinase activity increased, suggesting a potential mechanism by which OMSV influenced mycelial growth through modulating CAZyme activities. Therefore, this study provided insights into the pathogenic mechanisms triggered by OMSV in P. ostreatus. Full article
(This article belongs to the Special Issue Advances in Plant Virus Diseases and Virus-Induced Resistance)
Show Figures

Figure 1

Figure 1
<p>The mycelium growth of OMSV-free (Mock) and OMSV−infected (OMSV) <span class="html-italic">P. ostreatus</span> strains. (<b>a</b>) The mycelium growth on PDA plates after seven days of cultivation. (<b>b</b>) RT-PCR detection of OMSV. Lane M, DNA Marker2000. Numbers 1–3 represent three biological replicates.</p>
Full article ">Figure 2
<p>Gene expression analysis of OMSV-free (Mock) and OMSV-infected (OMSV) strains of <span class="html-italic">P. ostreatus</span>. (<b>a</b>) Principal component analysis (PCA) of each sample. The FPKM values of each sample were used to perform PCA. (<b>b</b>) The FPKM box plots of each sample. The horizontal axis represents the sample names, while the vertical axis displays the log10 (FPKM) values. Different colors denote distinct samples.</p>
Full article ">Figure 3
<p>Differential gene analysis of OMSV-free (Mock) and OMSV-infected (OMSV) <span class="html-italic">P. ostreatus</span> strains. (<b>a</b>) Volcano plot of all identified genes. Genes with upregulated expression are indicated by red dots, those with downregulated expression by blue dots, and genes not differentially expressed by gray dots. (<b>b</b>) The statistical map of DEGs. The horizontal axis indicates the comparison names, and the vertical axis indicates the number of DEGs. (<b>c</b>) Heatmap displaying the clustering analysis of all DEGs for the OMSV-free and OMSV-infected strains of <span class="html-italic">P. ostreatus</span>. The expression levels of DEGs were normalized using the log10 FPKM method. Each row represents a single gene, and each column corresponds to a sample group. The color gradient from blue to red indicates that the FPKM value ranges from low to high.</p>
Full article ">Figure 4
<p>KOG analysis of the DEGs of OMSV-free and OMSV-infected <span class="html-italic">P. ostreatus</span> strains. The vertical axis indicates the number of DEGs within a specific functional cluster, while the horizontal axis represents the functional classes.</p>
Full article ">Figure 5
<p>GO and KEGG function classification of DEGs of OMSV-free and OMSV-infected <span class="html-italic">P. ostreatus</span> strains. (<b>a</b>) GO pathway-annotated genes. The vertical axis represented the name of the enriched GO term, while the horizontal axis indicated the number of DEGs within the corresponding term. Different colors represent three different categories. (<b>b</b>) KEGG pathway-annotated genes. Different colors represent four different categories. (<b>c</b>) Statistics of KEGG enrichment.</p>
Full article ">Figure 5 Cont.
<p>GO and KEGG function classification of DEGs of OMSV-free and OMSV-infected <span class="html-italic">P. ostreatus</span> strains. (<b>a</b>) GO pathway-annotated genes. The vertical axis represented the name of the enriched GO term, while the horizontal axis indicated the number of DEGs within the corresponding term. Different colors represent three different categories. (<b>b</b>) KEGG pathway-annotated genes. Different colors represent four different categories. (<b>c</b>) Statistics of KEGG enrichment.</p>
Full article ">Figure 6
<p>Validation of selected DEGs of OMSV-free (Mock) and OMSV−infected (OMSV) strains of <span class="html-italic">P. ostreatus</span> using qRT-PCR. The statistical analysis included a one-way analysis of variance (ANOVA) and <span class="html-italic">t</span>-tests, * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01, and *** <span class="html-italic">p</span> &lt; 0.001.</p>
Full article ">Figure 7
<p>The enzyme activity of OMSV-free (Mock) and OMSV−infected (OMSV) <span class="html-italic">P. ostreatus</span> strains during mycelial growth. (<b>a</b>) Laccase activity; (<b>b</b>) amylase activity; (<b>c</b>) CMCase activity; (<b>d</b>) chitinase activity. Numbers 5–8 denote the days of mycelial growth in liquid medium. The statistical analysis included a one-way analysis of variance (ANOVA) and <span class="html-italic">t</span>-tests, * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01, and *** <span class="html-italic">p</span> &lt; 0.001.</p>
Full article ">
18 pages, 3024 KiB  
Article
Millet Bran Dietary Fibers Modified by Heating and Enzymolysis Combined with Carboxymethylation, Acetylation, or Crosslinking: Influences on Properties of Heat-Induced Egg White Protein Gel
by Yan Li, Chen Feng, Xueying Wang, Yajun Zheng, Xinling Song, Nan Wang and Danhong Liu
Foods 2024, 13(17), 2827; https://doi.org/10.3390/foods13172827 - 5 Sep 2024
Viewed by 334
Abstract
Applications of millet bran dietary fiber (MBDF) in the food industry are limited by its poor hydration properties. Herein, MBDF was modified by heating, xylanase and cellulase treatment separately combined with carboxymethylation, acetylation, and phosphate crosslinking, and the effects of the modified MBDFs [...] Read more.
Applications of millet bran dietary fiber (MBDF) in the food industry are limited by its poor hydration properties. Herein, MBDF was modified by heating, xylanase and cellulase treatment separately combined with carboxymethylation, acetylation, and phosphate crosslinking, and the effects of the modified MBDFs on heat-induced egg white protein gel (H-EWG) were studied. The results showed that three composite modifications, especially heating and dual enzymolysis combined with carboxymethylation, increased the surface area, soluble fiber content, and hydration properties of MBDF (p < 0.05). MBDF and the modified MBDFs all made the microstructure of H-EWG denser and decreased its α-helix content. Three composite modifications, especially heating and dual enzymolysis combined with carboxymethylation, enhanced the improving effect of MBDF on the WRA (from 24.89 to 35.53 g/g), pH, hardness (from 139.93 to 323.20 g), chewiness, and gumminess of H-EWPG, and enhanced the gastric stability at 3–5 g/100 g. MBDFs modified with heating and dual enzymolysis combined with acetylation or crosslinking were more effective in increasing the antioxidant activity of the gastrointestinal hydrolysates of H-EWG than MBDF (p < 0.05). Overall, heating, xylanase and cellulase treatment separately combined with carboxymethylation, acetylation and crosslinking can enhance the hydration properties and the improving effect of millet bran fibers on H-EWG properties. Full article
(This article belongs to the Section Food Biotechnology)
Show Figures

Figure 1

Figure 1
<p>Scanning electron micrographs of MBDF (<b>A</b>), MBDF-HDEC (<b>B</b>), MBDF-HDEPC (<b>C</b>), and MBDF-HDEA (<b>D</b>) with a magnification of 5000×, at 1 μm. MBDF, millet bran dietary fiber; MBDF-HDEC, MBDF modified by heating and dual enzymes hydrolysis combined with carboxymethylation; MBDF-HDEPC, MBDF modified by heating and dual enzymes hydrolysis combined with phosphate crosslinking; MBDF-HDEA, MBDF modified by heating and dual enzymes hydrolysis combined with acetylation.</p>
Full article ">Figure 2
<p>Fourier-transformed infrared spectroscopy of MBDF, MBDF-HDEC, MBDF-HDEPC, and MBDF-HDEA.</p>
Full article ">Figure 3
<p>Scanning electron micrographs of native egg white gel (<b>A</b>), egg white gel containing 5 g/100 g MBDF (<b>B</b>), egg white gel containing 5 g/100 g MBDF-HDEPC (<b>C</b>), egg white gel containing 5 g/100 g MBDF-HDEA (<b>D</b>); egg white gel containing 5 g/100 g MBDF-HDEC (<b>E</b>), egg white gel containing 4 g/100 g MBDF-HDEC (<b>F</b>), egg white gel containing 3 g/100 g MBDF-HDEC (<b>G</b>), egg white gel containing 2 g/100 g MBDF-HDEC (<b>H</b>), and egg white gel containing 1 g/100 g MBDF-HDEC (<b>I</b>) with a magnification of 5000×, at 1 μm.</p>
Full article ">Figure 4
<p>Relative content of protein secondary structure of heat-induced egg white gels fortified with MBDFs at an addition amount of 5 g/100 g. H-EWP, heat-induced egg white protein gel; H-EWP/MBDF, heat-induced egg white gel with MBDF; H-EWP/HDEC, heat-induced egg white gel with MBDF-HDEC; H-EWP/HDEPC, heat-induced egg white gel with MBDF-HDEPC; and H-EWP/HDEA, heat-induced egg white gel with MBDF-HDEA.</p>
Full article ">Figure 5
<p>Effects of different addition amount of MBDFs on the water-holding ability (<b>A</b>), pH (<b>B</b>), dehydration rate in the freeze-thaw cycle (<b>C</b>), and optical transparency (<b>D</b>) of heat-induced egg white gel (H-EWG). Different lower letters (a–e) near the lines or on the bars mean significant difference (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 6
<p>Effects of MBDF, MBDF-HDEC, MBDF-HDEPC, and MBDF-HDEA (at addition amount of 5/100 g) on the free amino acid production (<b>A</b>) and ABTS<sup>+</sup> scavenging activity (<b>B</b>) of heat-induced egg white protein gel under gastrointestinal digestion. Different lowercase letters (a–f) in the same column indicate significant difference (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">
16 pages, 2948 KiB  
Article
High-Level Expression of β-Glucosidase in Aspergillus niger ATCC 20611 Using the Trichoderma reesei Promoter Pcdna1 to Enhance Cellulose Degradation
by Jingjing Chang, Juan Wang, Zhihong Li, Lu Wang, Peng Lu, Yaohua Zhong and Hong Liu
Fermentation 2024, 10(9), 461; https://doi.org/10.3390/fermentation10090461 - 5 Sep 2024
Viewed by 423
Abstract
β-glucosidase is a key component of cellulase for its function in hydrolyzing cellobiose to glucose in the final step of cellulose degradation. The high-level expression of β-glucosidase is essential for cellulose conversion. Aspergillus niger ATCC 20611 has the potential for efficient protein expression [...] Read more.
β-glucosidase is a key component of cellulase for its function in hydrolyzing cellobiose to glucose in the final step of cellulose degradation. The high-level expression of β-glucosidase is essential for cellulose conversion. Aspergillus niger ATCC 20611 has the potential for efficient protein expression because of its ability to secret enzymes for the industrial production of fructooligosaccharides, but it lacks robust promoters for high-level protein expression. Here, the development of A. niger 20611 as a powerful protein expression system exploited the conserved constitutive promoter Pgpd1 of the glyceraldehyde-3-phosphate dehydrogenase-encoding gene from Trichoerma reesei to drive the expression of the enhanced green fluorescent protein in A. niger ATCC 20611. The mycelium of the transformant AGE9 exhibited intense fluorescence. Then, the promotor Pgpd1 was used to drive the expression of β-glucosidase and the enzyme activity of transformants AGB1 and AGB33 were 1.02 and 0.51 U/mL, respectively. These results demonstrate that the promotor Pgpd1 from T. reesei was applicable for A. niger ATCC 20611. Furthermore, the T. reesei-specific robust promoter Pcdna1 was used to drive the expression of β-glucosidase. The β-glucosidase exhibited a high-level expression with a yield of 15.2 U/mL, which was over 13.9 times higher than that driven by the promoter Pgpd1. The β-glucosidase was thermally stable and accounted for 85% of the total extracellular proteins. Subsequently, the fermentation broth including β-glucosidase was directly added to the cellulase mixture of T. reesei for saccharification of the acid-treated corncob residues and the delignified corncob residues, which increased the saccharification efficiency by 26.21% and 29.51%, respectively. Thus, β-glucosidase exhibited a high level of expression in A. niger ATCC 20611 and enhanced cellulose degradation by addition in vitro. In addition, the robust promoter Pcdna1 of T. reesei could drive the high-level expression of protein in A. niger ATCC 20611. These results demonstrate that the promoters in filamentous fungi could be employed across species in A. niger ATCC 20611 and further facilitated the efficient expression of β-glucosidase to optimize cellulases for efficient cellulose transformation. Full article
(This article belongs to the Special Issue Biorefinery of Lignocellulosic Biomass, 2nd Edition)
Show Figures

Figure 1

Figure 1
<p>Construction of the <span class="html-italic">gpd1</span> promoter-driven EGFP expression strains of <span class="html-italic">A. niger</span> ATCC 20611: (<b>A</b>) Schematic representation of the <span class="html-italic">egfp</span> expression under the control of <span class="html-italic">gpd1</span> promoter. The resistance gene <span class="html-italic">ptrA</span> was used as the reporter gene. (<b>B</b>) The transformant strains expressing EGFP under the control of <span class="html-italic">gpd1</span> promoter were selected on an MM plate with 0.3% PT (the left). The plate diagram on the right showed the growth of <span class="html-italic">A. niger</span> ATCC 20611 on a MM plate with 0.3% PT as the control. (<b>C</b>) The transcript levels of <span class="html-italic">egfp</span> in the transformant AGE9 and the parental strain <span class="html-italic">A. niger</span> ATCC 20611 were detected by RT–qPCR at 24 h. The relative expression was the level of transcripts normalized to that of <span class="html-italic">actin</span> gene. (<b>D</b>) Detection of the fluorescence in the mycelium of <span class="html-italic">A. niger</span> AGF9-1 and the parental strain ATCC 20611 at 48 h by light and fluorescent microscopy. Error bars indicate the standard deviation (SD) of three biological replicates. Student’s <span class="html-italic">t</span>-test: ** <span class="html-italic">p</span> &lt; 0.01. ND, not detected.</p>
Full article ">Figure 2
<p>Construction of the <span class="html-italic">gpd1</span> promoter-driven β-glucosidase (BGLA) expression strains of <span class="html-italic">A. niger</span> ATCC 20611: (<b>A</b>) Schematic representation of the <span class="html-italic">bglA</span> expression cassette, which contained the promotor of <span class="html-italic">gpd1</span> (P<span class="html-italic">gpd1</span>), the signal peptides of <span class="html-italic">cbh1</span> (SP), the β-glucosidase-encoding gene (<span class="html-italic">bglA</span>) and the terminator of <span class="html-italic">cbh1</span> (T<span class="html-italic">cbh1</span>). (<b>B</b>) Detection of the β-glucosidase activities of transformants on the CMC-esculin plates. The parental strain ATCC 20611 was used as control. (<b>C</b>) The ratios of halo diameter to colony diameter of the transformants in (<b>B</b>). (<b>D</b>) The transcript levels of <span class="html-italic">bglA</span> in transformants AGB1, AGB33 and the parental strain ATCC 20611 at 24 h by RT–qPCR. The relative expression was the level of transcripts normalized to that of <span class="html-italic">actin</span> gene. (<b>E</b>) Detection of β-glucosidase activity using p-Nitrophenyl-β-d-glucopyranoside (pNPG) as the substrate of the <span class="html-italic">cdna1</span> promoter-driven β-glucosidase (BGLA) expression strains and the parental strain ATCC 20611 after 4-day fermentation. Error bars indicate the standard deviation (SD) of three biological replicates. Student’s <span class="html-italic">t</span>-test: ** <span class="html-italic">p</span> &lt; 0.01. ND, not detected.</p>
Full article ">Figure 3
<p>Transcript analysis of endoplasmic reticulum-associated genes (<span class="html-italic">bip1</span>, <span class="html-italic">pdi1 yos9</span> and <span class="html-italic">der1</span>) by RT-qPCR. The transcript levels of UPR—(<b>A</b>) and ERAD—(<b>B</b>) related genes of transformants AGB1, AGB33 and the parental strain ATCC 20611. The relative expression was the level of transcripts normalized to that of the <span class="html-italic">actin</span> gene. Error bars indicate the standard deviation (SD) of three biological replicates. Student’s <span class="html-italic">t</span>-test: ** <span class="html-italic">p</span> &lt; 0.01, * <span class="html-italic">p</span> &lt; 0.05. n.s., no significant differences.</p>
Full article ">Figure 4
<p>Construction of the <span class="html-italic">cdna1</span> promoter-driven β-glucosidase (BGLA) expression strains of <span class="html-italic">A. niger</span> ATCC 20611: (<b>A</b>) Schematic representation of the <span class="html-italic">bglA</span> expression under the control of <span class="html-italic">gpd1</span> promoter, which contained the promotor <span class="html-italic">gpd1</span> (P<span class="html-italic">cdna1</span>), the signal peptides of <span class="html-italic">cbh1</span> (SP), the β-glucosidase-encoding gene (<span class="html-italic">bglA</span>) and the terminator of <span class="html-italic">cbh1</span> (T<span class="html-italic">cbh1</span>). (<b>B</b>) Detection of β-glucosidase activities of transformant on the CMC-esculin plate. The parental strain ATCC 20611 as control. (<b>C</b>) The ratios of halo diameter to colony diameter. The data assessed the β-glucosidase production capacity. Analysis of the expression level of <span class="html-italic">bglA</span>. (<b>D</b>) The transcript levels of <span class="html-italic">bglA</span> in transformants ACB8, ACB11 and the parental strain ATCC 20611 at 24 h by RT–qPCR. The relative expression was the level of transcripts normalized to that of the <span class="html-italic">actin</span> gene. Error bars indicate the standard deviation (SD) of three biological replicates. Student’s <span class="html-italic">t</span>-test: ** <span class="html-italic">p</span> &lt; 0.01. ND, not detected.</p>
Full article ">Figure 5
<p>Analysis of the expression level of <span class="html-italic">bglA</span>: (<b>A</b>) Detection of β-glucosidase activity using p-Nitrophenyl-β-d-glucopyranoside (pNPG) as the substrate of the <span class="html-italic">cdna1</span> promoter-driven β-glucosidase (BGLA) expression strains and the parental strain ATCC 20611 after 4-day fermentation. (<b>B</b>) SDS PAGE showing secreted BGLA expression. (<b>C</b>) Detection of extracellular total protein concentration by Bradford’s method. Error bars indicate the standard deviation (SD) of three biological replicates. Student’s <span class="html-italic">t</span>-test: ** <span class="html-italic">p</span> &lt; 0.01, * <span class="html-italic">p</span> &lt; 0.05.</p>
Full article ">Figure 6
<p>Transcript analysis of endoplasmic reticulum-associated genes (<span class="html-italic">bip1</span>, <span class="html-italic">pdi1 yos9</span> and <span class="html-italic">der1</span>) by RT-qPCR: The transcript levels of UPR—(<b>A</b>) and ERAD—(<b>B</b>) related genes of transformants ACB8, ACB11 and the parental strain ATCC 20611. The relative expression was the level of transcripts normalized to that of actin. Error bars indicate the standard deviation (SD) of three biological replicates. Student’s <span class="html-italic">t</span>-test: ** <span class="html-italic">p</span> &lt; 0.01, * <span class="html-italic">p</span> &lt; 0.05.</p>
Full article ">Figure 7
<p>Enzymatic characterization of BGLA in <span class="html-italic">A. niger</span> ACB 8. The BGLA in <span class="html-italic">T. reesei</span> QVB-1 was the control: (<b>A</b>) Optimal temperature. The activity of BGLA was detected in the temperature range of 30–80 °C at pH 4.8. (<b>B</b>) and (C) Thermostability. After preincubation of the fermentation broths of the ACB8 and QVB-1 at 50, 60 and 70 °C for 0.5 h to 10 h, respectively, the residual activities were determined at pH 4.8 and 50 °C and the initial activity at pH 4.8 and 50 °C was defined as 100%. (<b>D</b>) Optimal pH. The activity of BGLA was detected at 50 °C in the pH range 3 to 8. (<b>E</b>) pH stability. After preincubation of the fermentation broths of the ACB8 and QVB-1 in citric acid buffer of different pH for 24 h in 4 °C, respectively, the residual activities were determined at pH 4.8 and 50 °C and the initial activity at pH 4.8 and 50 °C was defined as 100%.</p>
Full article ">Figure 8
<p>Saccharification of different pretreated corncob: (<b>A</b>) Glucose released from saccharification of acid-pretreated corncob residues by <span class="html-italic">T. reesei</span> SN1 with the addition of the fermentation broth of <span class="html-italic">A. niger</span> ACB8 for 48 h. (<b>B</b>) Glucose released from the saccharification of delignified corncob residues by <span class="html-italic">T. reesei</span> SN1 adding the fermentation broth of ACB8 for 48 h. Saccharification by adding the fermentation broth of the parental strain <span class="html-italic">A. niger</span> ATCC 20611 was used as the control. Error bars indicate the standard deviation (SD) of three biological replicates. Student’s <span class="html-italic">t</span>-test: * <span class="html-italic">p</span> &lt; 0.05. n.s., no significant differences.</p>
Full article ">
17 pages, 2814 KiB  
Article
Statistical Optimization and Purification of Cellulase Enzyme Production from Trichosporon insectorum
by Hanane Touijer, Najoua Benchemsi, Muhammad Irfan, Annabella Tramice, Meryem Slighoua, Ramzi A. Mothana, Abdullah R. Alanzi, Bousta Dalila and Hicham Bekkari
Fermentation 2024, 10(9), 453; https://doi.org/10.3390/fermentation10090453 - 1 Sep 2024
Viewed by 605
Abstract
Enzymatic degradation of cellulosic biomass represents the most sustainable and environmentally friendly method for producing liquid biofuel, widely utilized in various commercial processes. While cellulases are predominantly produced by bacteria and fungi, the enzymatic potential of cellulase-producing yeasts remains significantly less explored. In [...] Read more.
Enzymatic degradation of cellulosic biomass represents the most sustainable and environmentally friendly method for producing liquid biofuel, widely utilized in various commercial processes. While cellulases are predominantly produced by bacteria and fungi, the enzymatic potential of cellulase-producing yeasts remains significantly less explored. In this study, the yeast strain Trichosporon insectorum, isolated from the gut of the coprophagous beetle Gymnopleurus sturmii, was utilized for cellulase production in submerged fermentation. A central composite design was employed to optimize cellulase production, with substrate concentration, temperature, and pH as dependent variables. The highest CMCase activity of 0.71 IU/mL was obtained at 1% substrate concentration, pH 5, and an incubation temperature of 40 °C for 72 h of fermentation using cellulose as a carbon source. For FPase production, the high value was 0.23 IU/mL at 0.5% CMC, pH 6, and an incubation temperature of 40 °C for 72 h. After purification, the enzymes produced by T. insectorum represent 39% of the total proteins. The results of this study offer an alternative strategy for utilizing various carbon sources, both soluble (CMC, carboxymethylcellulose) and insoluble (cellulose), to efficiently produce cellulase for the degradation of lignocellulosic materials. This approach holds promising benefits for sustainable waste management. Full article
(This article belongs to the Section Industrial Fermentation)
Show Figures

Figure 1

Figure 1
<p>Observed and predicted values of dependent variables for CMCase production in media with CMC or FC.</p>
Full article ">Figure 2
<p>Observed and predicted values of dependent variables for FPase production in media with CMC or FC.</p>
Full article ">Figure 3
<p>Contour plots of the different variables for CMCase production in media with CMC or FC.</p>
Full article ">Figure 3 Cont.
<p>Contour plots of the different variables for CMCase production in media with CMC or FC.</p>
Full article ">Figure 4
<p>Contour plots of the different variables for FPase production in media with CMC or FC.</p>
Full article ">Figure 4 Cont.
<p>Contour plots of the different variables for FPase production in media with CMC or FC.</p>
Full article ">Figure 5
<p>Optical density of purified protein fractions.</p>
Full article ">Figure 6
<p>Thin layer chromatography of the purified fractions (demonstration of residual CMC by the α-naphthol test).</p>
Full article ">Figure 7
<p>Thin layer chromatography of the purified fractions (demonstration of proteins by the ninhydrin test).</p>
Full article ">
25 pages, 2773 KiB  
Article
Application of Conventional and Hybrid Thermal-Enzymatic Modified Wheat Flours as Clean Label Bread Improvers
by Piotr Lewko, Agnieszka Wójtowicz and Marek Gancarz
Appl. Sci. 2024, 14(17), 7659; https://doi.org/10.3390/app14177659 - 29 Aug 2024
Viewed by 436
Abstract
A new flour blend (F) composed of selected milling and leaving passages with a high content of non-starch polysaccharides underwent thermal (T), hydrothermal (H) or hybrid processing and was used along with cellulase (C) and cellulase-xylanase complex (CX) to produce bread. This modified [...] Read more.
A new flour blend (F) composed of selected milling and leaving passages with a high content of non-starch polysaccharides underwent thermal (T), hydrothermal (H) or hybrid processing and was used along with cellulase (C) and cellulase-xylanase complex (CX) to produce bread. This modified flour can be considered a clean label product. In this study, blends of common and treated flours were tested for dough properties and rheology. The modified flours were added at 10 and 20% to the base wheat flour. A pan bread was then prepared to test their suitability for bread baking. Dough and bread properties were subsequently assessed. Accordingly, dough with added thermally, hydrothermally, and hybrid modified flours revealed differences in rheology. Addition of hybrid enzymatic-hydrothermal treated flour increased dough tenacity by 23% and baking strength by 26%, but decreased dough extensibility by 19%, whereas hybrid enzymatic-thermal modification increased water absorption by 6% and bread yield from 146.77% to 150.02% when modified flour was added at 20%. Breads with added modified flours demonstrated a 16% increase in bread volume, 8% lower baking loss, and 14% greater density, with no negative effect on color and texture. Thus, hybrid thermal-enzymatic treatment of the developed flours can be recommended as a suitable method for enhancing the utilization of waste flour fractions and increasing their value by enabling them to be considered as clean label bread improvers. Full article
Show Figures

Figure 1

Figure 1
<p>Rheological features of raw materials composition with added modified flours as compared to control common bread flour: (<b>a</b>) protein weakening (C2); (<b>b</b>) starch gelatinization (C3); (<b>c</b>) amylase activity (C4); (<b>d</b>) starch retrogradation (C5); K—control; F—flour; T—dry thermal treatment; H—hydrothermal treatment; C—cellulase enzyme; CX—cellulase-xylanase enzyme complex; 10 and 20—% of modified flour in bread recipe; dash line—level for control sample; <sup>a–l</sup>—means indicated with similar letters in columns do not differ significantly at α = 0.05.</p>
Full article ">Figure 2
<p>Bread samples with the addition of modified flours: K—control; F—flour; T—dry thermal treatment; H—hydrothermal treatment; C—cellulase enzyme; CX—cellulase-xylanase enzyme complex; 10 and 20—% of modified flour in bread recipe.</p>
Full article ">Figure 3
<p>PCA analysis: (<b>a</b>) projection of variables all parameters on the PC1 and PC2 loadings plot; (<b>b</b>) projection of modified wheat flours on the PC1 and PC2 scores plot.</p>
Full article ">Figure 4
<p>PCA analysis: (<b>a</b>) projection of variables parameters on the PC1 and PC3 loadings plot; (<b>b</b>) projection of modified wheat flours on the PC1 and PC3 scores plot.</p>
Full article ">
Back to TopTop