Next Article in Journal
Therapeutic and Anti-Thrombotic Properties of Some Naturally Fermented Soybean Foods of the Eastern Himalayas
Next Article in Special Issue
Assessment of Growth Inhibition of Eugenol-Loaded Nano-Emulsions against Beneficial Bifidobacterium sp. along with Resistant Escherichia coli Using Flow Cytometry
Previous Article in Journal
Integration of Corn and Cane for Ethanol Production: Effects of Lactobacilli Contamination on Fermentative Parameters and Use of Ionizing Radiation Treatment for Disinfection
Previous Article in Special Issue
Probiotic Properties of Weissella confusa PP29 on Hibiscus sabdariffa L. Media
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Probiotics in the Sourdough Bread Fermentation: Current Status

by
Ingrid Teixeira Akamine
,
Felipe R. P. Mansoldo
and
Alane Beatriz Vermelho
*
Bioinovar, Institute of Microbiology Paulo de Góes, General Microbiology, Federal University of Rio de Janeiro—UFRJ, Rio de Janeiro 21941-902, Brazil
*
Author to whom correspondence should be addressed.
Fermentation 2023, 9(2), 90; https://doi.org/10.3390/fermentation9020090
Submission received: 21 November 2022 / Revised: 9 January 2023 / Accepted: 18 January 2023 / Published: 20 January 2023
(This article belongs to the Special Issue Postbiotics from Production to Their Health-Promoting Aspects)

Abstract

:
Sourdough fermentation is an ancient technique to ferment cereal flour that improves bread quality, bringing nutritional and health benefits. The fermented dough has a complex microbiome composed mainly of lactic acid bacteria and yeasts. During fermentation, the production of metabolites and chemical reactions occur, giving the product unique characteristics and a high sensory quality. Mastery of fermentation allows adjustment of gluten levels, delaying starch digestibility, and increasing the bio-accessibility of vitamins and minerals. This review focuses on the main steps of sourdough fermentation, the microorganisms involved, and advances in bread production with functional properties. The impact of probiotics on human health, the metabolites produced, and the main microbial enzymes used in the bakery industry are also discussed.

Graphical Abstract

1. Introduction

Sourdough fermentation is a technique that can use several types of flour, such as wheat, rye, or other cereals, and water. The oldest process of sourdough preparation is spontaneous fermentation and acidification due to the local microbiota in a complex interaction, mainly lactic acid bacteria (LAB) and yeasts [1,2,3]. It is a traditional fermentation process used for various foods, especially baked goods, improving nutritional and health benefits [4,5]. The best-known example is bread; however, other applications of this type of fermentation include biscuits, panettone, pasta, and beverages [6,7,8,9]. Several varieties of this traditional sourdough bread recipe have appeared, bringing innovations to the primary process to attend to the bakery industry and the consumer, which appreciates nutritious, healthy, and tasty food [10]. Sourdough can be used by the bakery industry in several applications through liquid, pasty, or dry formats with live microorganisms. Moreover, there are “ready-to-use” sourdoughs with microorganisms inactivated to promote shelf-stable products [11].
When sourdough is used, rather than Saccharomyces cerevisiae, a typical baker’s yeast, for the bread dough fermentation, superior properties in bread quality and technological value are achieved [12].
Probiotics are safe for human consumption and can produce metabolites that positively influence gastrointestinal [13] and bone health diseases [14]. Beneficial effects in eczema [15], allergies [16], respiratory tract infection [17], obesity [18], and cognitive and mental health are also described [19]. Probiotics act as gut microbiota, stimulating a complex communication with the body that is mediated by the metabolites generated [20]. Some LAB and yeasts present in sourdough fermentation are considered and presumed to be probiotics [21]. However, a complete analysis, such as a whole-genome analysis, is proposed for better probiotic characterization [22,23,24,25,26].
This study aims to review improvements in sourdough fermentation, its enzymes, metabolites, and technological advances to produce sourdough bread with desirable properties.

2. Sourdough Fermentation Types: Inoculum and Technology Processes

Sourdough fermentation can be classified according to the inoculum (type I, II, and III) and the technology process (type 0, I, II, and III). According to inoculum, type I uses the microorganisms present in the dough, which results in spontaneous fermentation and utilizes back-slopping to propagate the sourdough microbiota. This process is common in artisan bakeries and homes. Control parameters include temperature, number, and time of back-slopping in artisan bakeries. Type II inoculums are frequently utilized in industrial processes, where a starter culture is added to sourdough fermentation according to the objectives and results desirable in the final process. Type III is a hybrid and combines type II by using a starter culture with type I for propagating the sourdough with back-slopping. Type III is common in artisan bakeries and industrial ones [27].
According to the production and process, sourdough fermentation can also be classified into four major types. The simple type 0 fermentation starts with a flour–water mixture and is allowed for a limited time. LAB species that are naturally present in the flour, grow faster are more prolific than yeast. Aplevicz et al. 2013 [28] compared the growth of two Lactobacillus paracasei strains with two Saccharomyces cerevisiae strains. LAB had the highest counts over time and microbial growth after 10 h of fermentation. The highest value for LAB strains was 8.91 log CFU/g, and then 8.66 log CFU/g. One S. cerevisiae strain reached 8.03 log CFU/g, and the other, 0.21 log CFU/g. At 6 h of fermentation, LABs were 8.66 log CFU/g and 8.33 log CFU/g. Yeasts showed 7.18 log CFU/g and 7.08 log CFU/g.
The growth kinetics of six strains of LAB: Fructilactobacillus sanfranciscensis, were compared with five strains of yeast: Kazachstania humilis, from sourdough, by Altilia et al. 2021 [29]. Three predictive models were used to evaluate the behavior of co-cultivated microorganisms: The zwitwering model based on Gompertz’s equation, Baranyil and Roberts’ function, and Schiraldi’s function. The results showed that F. sanfranciscensis strains significantly steer the growth kinetics affecting the ratio of bacterial to yeast cells, where the yeast strains of K. humilis adapt to the bacterial strains. The authors discuss the possibility of metabolic interactions for stabilizing the sourdough consortium through communication, such as quorum sensing, to control population density, among other functionalities. In another co-culture article with LAB and yeast, the yeast population size decreased in the presence of LAB regardless of the strain, while the LAB’s population size was rarely influenced by the presence of yeast [30]. In type 0 fermentation, bioactive molecules and organic acids (lactic and acetic acids) are produced lowering the pH (pH~4), but it is controversial if this type of fermentation is a true sourdough. The time limit is insufficient to produce other characteristic products of sourdoughs and is more known as sponge dough [27].
Type I is a traditional method for preparing sourdough and consists of spontaneous fermentation in a flour–water mixture at an ambient temperature of less than 30 °C for 24 h or less and back-slopping frequently. This fermentation increases the quality of the final baked good [27,31].
Type II is when a starter culture initiates the fermentation process and is used to develop desirable characteristics for the baking industry. It is a sourdough in liquid form with controlled parameters [31,32].
Type III can be dried or lyophilized. It is preferable for use in an industrial bakery because it has a higher stability quality than fresh sourdough [31,33]. The fermentation temperature used is above 30 °C and in a single time stage between 24–72 h [34].
The fermentation process will change depending on the added materials, such as the addition of naturally rich products in microorganisms: fruits, yogurt, or another material [35]. It is observed that it is challenging to classify sourdough fermentation due to some nuances in the process production [34]. Figure 1 summarizes the major types of production and process of and their characteristics.

3. Sourdough Fermentation: Major Pathways

The sourdough fermentation process generates mainly acids, alcohols, aldehydes, esters, and ketones; it is the primary route of volatile organic compounds (VOCs) [36]. Sourdough bread has a complex profile, strongly influenced by the compounds generated during fermentation by the diverse array of microorganisms present, mainly yeasts and LAB.
Wheat flour is composed of starch (70–75%), water (14%), proteins (10–12%), non-starch polysaccharides (2–3%), lipids (2%), and soluble carbohydrates such as maltose, sucrose, and glucose (1.55–1.85%) The starch is degraded into glucose and maltose by flour amylase and glucoamylase by some sourdough LABs. Maltose is degraded into glucose by the enzyme maltose phosphorylase from LAB and by alpha-D-glucosidase, a maltase from Saccharomyces yeasts. In the dough, the available carbon source is thus maltose, followed by sucrose, glucose, and fructose, with some trisaccharides such as maltotriose and raffinose [37].
The glucose concentration increases during fermentation because other complex carbohydrates are metabolized by LAB and yeasts. The disaccharide lactose can be fermented by microbial enzymes liberating glucose and galactose after cleavage. Starting from glucose, homofermentative LAB produces lactic acid through glycolysis, while heterofermentative LAB generates, besides lactic acid, CO2, acetic acid, and/or ethanol [38]. Two major fermentation routes are found: lactic fermentation (LF) and alcoholic fermentation (AF). In lactic fermentation, the pyruvate molecules formed by glucose oxidation from the Embden–Meyerhof–Parnas or glycolytic pathway are reduced to lactic acid (homolactic fermentation). Streptococcus, Lactobacillus, and Enterococcus use this route (Figure 2A1). Another possibility is the pyruvate originating from a mixture of lactate, ethanol and/or acetic acid, through the oxidation of the coenzymes NADH + and H+ by the lactate dehydrogenase, and CO2 from a glucose molecule until ribulose 5 phosphate (heterolactic fermentation).
This fermentation is used by genera such as Leuconostoc, Bifidobacterium, Weissella, and some Lactobacillus. Some glycolytic enzymes are missing in these bacteria, and they use the pentose-P pathway to degrade the glucose (Figure 2A2). Classification according to the type of lactic fermentation defines three major Lactobacillus groups: group I: obligate homofermentative lactobacilli. Hexoses are exclusively fermented to lactic acid via the Embden–Meyerhof–Parnas pathway. The bacteria have the enzyme fructose-1,6-bisphosphate aldolase, but phosphoketolase is absent. Due to this, pentoses and gluconate are not fermented. This group includes the species L. acidophilus, L. delbrueckii, and L. salivarius; group II: facultative heterofermentative lactobacilli. Hexoses are fermented to lactic acid almost exclusively via the Embden–Meyerhof–Parnas pathway. The bacteria possess both aldolase and phosphoketolase enzymes, and ferment not only hexoses but also pentoses. In the presence of glucose, the enzymes of the phosphogluconate pathway are inhibited. This group includes L. casei, L. paracasei, L. plantarum, and L. curvatus. Organic acids, CO2, alcohol, and H2O2 could be produced; group III: composed by the obligate heterofermentative Lactobacilli. They have the enzyme phosphoketolase but not aldolase and ferment sugars in a heterofermentative mode. Hexoses are fermented via the phosphogluconate pathway, producing lactate, ethanol (acetic acid), and CO2 in equimolar amounts. Pentoses enter this pathway and can also be fermented [39,40,41,42].
After the enzymatic hydrolysis of pentosans and other complex carbohydrates, pentoses such as d-xylose, ribose, and l-arabinose are liberated in rye and wheat flour. Heterofermentative strains can ferment pentoses through part of the 6-phosphogluconate pathway [43]. Pentoses and hexoses are simultaneously rather than successively fermented by sourdough Lactobacilli. Lactobacilli are important during the fermentation of sourdoughs, and they can ferment pentose carbohydrates without producing CO2 because they have a constitutive phosphoketolase. Facultatively heterofermentative lactic acid bacteria produce the phosphoketolase enzyme in response to the presence of pentoses [42,44]. The contribution of LAB to the flavor of sourdough bread is associated with the production of lactic acid (fresh acidity) and acetic acid (sharp acidity). The conversion of amino acids such as phenylalanine (sweet), isoleucine (acidic), glycine, serine, and alanine (vinegar/sour) to aldehydes and ketones can form additional flavor compounds [36].
Alcoholic fermentation (AF) is characterized by the formation of ethanol and CO2 from pyruvate reduction (Figure 2B). AF is a metabolic pathway of yeasts. In this fermentation, yeast produces gas that promotes dough conditioning and increased volume and imparts desirable aromas and flavors in baked goods. A synergy occurs between the dough and the yeast, and in this context, the rate of CO2 production is determined by the activities of the glycolytic yeast enzymes. The retention of CO2 produced by yeast is a function of the wheat or cereal flour. It is essential to point out that glucose formed by LAB amylases from amide is a significant source of this substrate for yeasts. Other compounds are volatile organic compounds (VOCs), which arise from this interaction between yeasts and LABs. Diverse aldehydes, alcohols, esters, ketones, lactones, sulfur compounds, furan derivatives, and hydrocarbons are found, conferring flavor and other characteristics to sourdough bread [32]. Some VOCs are due to yeast fermentation, particularly of S. cerevisiae in mixed sourdough starters, including ethanol, 2,3-butanedione, 2-methyl-1-propanol, 3-methyl-1-butanol, and phenyl ethyl alcohol [45]. Other negatively correlated compounds have been associated with the fermentation of homofermentative and facultative heterofermentative LAB, including 2,3-butanedione and acetaldehyde [46].

4. Probiotics and Postbiotics in Sourdough and the Impact on Human Health

Probiotics are live microorganisms that confer a health benefit when administered in adequate amounts. In addition, new terms have been used to name microbial metabolites from non-viable cells, including paraprobiotics, parapsychobiotics, ghost probiotics, matabiotics, tyndallized probiotics, bacterial lysates, and postbiotics. However, the last term is being mightily used for the vital concept of promoting health. Postbiotics means non-living microorganisms and is a preparation with inanimate microbial cells or their components that promote host health. Metabolites and cellular structures from microorganisms from sourdough are potential postbiotics [47,48].
Sourdough microbiota are involved in a complex interaction that engenders beneficial compounds for human consumption. With the current molecular methods, a better characterization of sourdough microbiota has been described [49,50]. Moreover, whole-genome sequencing has permitted more accuracy in confirming the requirements for characterizing probiotic microorganisms [25].
The probiotics and postbiotics of sourdough provide health benefits in addition to the essential nutritional components of the bread itself. A fermented food passes through a process with microorganisms and enzymes converting macronutrients into bioactive products. LAB is one major microorganism group with several probiotic strains applicable to cereal fermentation. LAB are classified as Gram-positive bacteria which include low guanine + cytosine (G + C) content as well as being acid-tolerant, non-motile, non-spore-forming, and are rod- or cocci-shaped. The main function of LAB is to produce lactic acid, that is, the acidification [40]. The most common genera found in sourdough were Lactobacillus, Leuconostoc, Weissella, and Pediococcus. However, the genera Streptococcus, Enterococcus, and Lactococcus are found, although less frequently [51,52]. The second group of microorganisms are the yeast, and more than 20 species dominated by the genera Saccharomyces and Candida are found. LABs and yeasts are naturally found in the microbiota of the cereal flour used or can be added as a starter. Some native microorganisms found in sourdough are listed in Table 1.
No strong relation between the microbiota and regional origin has been proven until now [34,53]. However, in Brazil, Lactobacillus farciminis was the prevalent species rather than Lactobacillus sanfranciscensis. Other bacteria could be present. At a temperature of 30 ± 1 °C, unusual bacterial groups such as Pseudomonas and Enterobacteriaceae can proliferate. At a temperature of 21 ± 1 °C, there may be a prevalence of LAB, which are safer for human consumption [54].
Other factors that have been identified that could influence the microbiota. Insects could be a reservoir of Lactobacillus sanfranciscensis [55], commonly found in type I sourdough. Lactobacillus delbrueckii and Lactobacillus reuteri groups from vertebrate–host interactions are typical occurrences in type II sourdough [33]. In Italy, bakery LAB species such as Weissela spp. and Streptococcus spp. are in the local environment [50]. In addition, the water used in sourdough type I could influence its microbiota [56]. It has been demonstrated that Lactobacillus, Streptococcus, Enterococcus, and Lactococcus are endophytic components of wheat and, as such, a source of sourdough microbiota [52]. The sourdough microbiota changes with decreasing pH, whereas at the beginning, there is abundant Proteobacteria phylum in mature sourdoughs, which at the end is replaced by Firmicutes [54,57]. De Angelis et al. [58] demonstrated that the extraction rate of wheat flour from cultivars Triticum aestivum and Triticum durum modified the microbiota community in sourdough type I. The study showed that during fermentations, both controlled and uncontrolled pH conditions up to 4.0 yield similar LAB and yeast species [59].
A meta-analysis with 583 sourdough-related literature articles in the period 1999–2017 was done by Kerrebroeck et al. 2017 [34] based on the study of species diversity (LAB and yeasts) and process condition (type I and II). The authors concluded that the species number of yeasts is lower than LABs and the microbial diversity of type I sourdoughs is lower than type II. Through the principal component (PC) analysis, three clusters were detected. The first group had a high relative abundance of Pichia kudriavzevii and/or Torulaspora delbrueckii. In the second group, a high relative abundance of Wickerhamomyces anomalus and/or Candida glabrata was established, and the third Saccharomyces cerevisiae and/or C. humilis species were detected.
A systematic analysis from 1990 to 2020 using “sourdough” as a keyword showed the following number of articles by geographic location: North America (13), South America (6), Africa (31), Europe (175), Asia (54), China (20), and Oceania (1—New Zealand). In Latin America and America, very few studies have been published. Lactobacillus plantarium and Saccharomyces cerevisiae appeared in several sourdough microbiota characterizations [1]. It also showed that some bacteria not belonging to the LAB group could significantly function in sourdough and need to be explored more. Some species are listed in Table 1. On the other hand, in a study that characterized 500 samples from sourdough microbiota, almost all from the United States of America (429 samples) showed evidence of different bacterial groups of the Rhodospirillales order despite the Lactobacillales being the dominant order. Lactobacillus sanfranciscensis was the prevalent species. It was demonstrated that the acetic acid bacteria group substantially impacted aroma and leavening properties in sourdough fermentation [53].
A sourdough starter is in essence, a culture of probiotic yeast and lactic bacteria to be added to the fermentation process, besides the native microbiome of the cereals and external contaminations.
Starters also can be used to propagate the sourdough as a live fermentation. In this context, it is interesting to explore the technological and biological properties of the selected starter associations of bacteria/yeast to improve the quality of fermentation and bread properties. Acidification and proteolysis analysis (including free amino acids), and volatile compound profiles are examples. In a recent review of Landis et al. [53], working with 500 sourdough starters from four continents, it was suggested that geographical location has little influence on the microbial diversity of bacteria and yeast species. This result contrasts sharply with widespread assumptions relative to the biogeographic distribution of starters from sourdough [34]. In Brazil, for instance, Lactobacillus farciminis was the prevalent species rather than Lactobacillus sanfranciscensis. Table 2 lists some starters used in sourdough and produced compounds.
Table 1. Native microbiota found in sourdough.
Table 1. Native microbiota found in sourdough.
MicrobiotaReference
BacteriaAcetobacter lovaniensis spp.
Acetobacter malorum ssp.
Acetobacter pasteurianus/papaya
Acetobacter tropicalis
Enterecoccus durans
Enterobacter hormaechei/cloacae
Enterococcus faecalis
Enterococcus faecium
Enterococcus gilvus
Enterococcus hirae
Gluconobacter frateurii spp.
Gluconobacter sphaericus spp.
Komagataeibacter cluster 4
Lactobacillus brevis
Lactobacillus coryniformis
Lactobacillus curvatus
Lactobacillus diolivorans
Lactobacillus farciminis
Lactobacillus fermentum
Lactobacillus gallinarum
Lactobacillus kimchii
Lactobacillus otakiensis
Lactobacillus parabrevis
Lactobacillus paracasei
Lactobacillus paralimentarius
Lactobacillus plantarium
Lactobacillus sakei
Lactobacillus sanfranciscensis
Lactobacillus xiangfangensis
Lactococcus lactis
Leuconostoc
Leuconostoc citreum
Pediococcus
Pediococcus parvulus
Pediococcus pentosaceus
Psychrobacter
Streptococcus
Weissella
[1,34,53,54,60,61]
YeastsCida glabrata
Cida humilis
Hanseniaspora uvarum
Kazachstania humilis (synonym Cida humilis)
Kazachstania servazzii
Kazachstania unispora
Kluyveromyces aestuarii
Kluyveromyces lactis
Kluyveromyces marxianus
Pichia fermentans (synonym Cida lambica)
Pichia kudriazevii
Saccharomyces cerevisiae
Saccharomyces uvarum
Saccharomycestales sp.
Torulaspora delbrueckii
Wickerhamomyces anomalus
Yarrowia keelungensis
[1,34,53,61,62,63]
Table 2. Starters used in sourdough bread production with identified compounds.
Table 2. Starters used in sourdough bread production with identified compounds.
InoculumCompoundPropertyReference
Lactobacillus sanfranciscensis and Lactobacillus reuteriGlutaminase activitySignificantly influences wheat bread flavor[64]
Lactobacillus diolivorans and Lactobacillus buchneriPropionateIncrease antifungal property on bread[65]
Weissella cibaria MG1DextransImproved volume, crumb softness, and shelf-life[66]
Lactobacillus hammessiMonohydroxy fatty acidAntifungal property improving the bread shelf-life[67]
Lactobacillus paracasei RN5, Lactobacillus plantarum X2, Lactobacillus brevis LBRZ7, Lactobacillus fermentum LBRH10, Lactobacillus buchneri LBRZ6, and Propionibacterium frendenreichii ssp. Shermanii NBIMCC 327AntimicrobialPrevent bacterial and mold spoilage[68]
Lactobacillus curvatus 750(13), Pediococcus acidilactici EKO26, Pediococcus pentosaceus 1850(3), Lactobacillus coryniformis pA, Weissella cibaria EKO31, Pediococcus pentosaceus EKO23, Lactobacillus plantarum KKp 593/p, Lactobacillus helveticus 10, Lactobacillus plantarum W37/54, Lactobacillus sakei 750(20), and Lactobacillus rhamnosus Lr (23)Proteolytic activityReducing allergenic proteins and improving the quality of bakery products[69]
Lactobacillus curvatus MA2, Pediococcus pentosaceus OA2, and Pediococcus acidilactici O1A1Phytase and antioxidant activitiesImprove textural and sensory features of bread[70]
Weissella cibaria PON10030, Weissella cibaria PON10032, Lactobacillus citreum PON 10079, and Lactobacillus citreum PON10080Volatile organic compoundsImprove the taste of bread[71]
Saccharomyces bayanusAromatic compoundsImprove the sensory profile of bread[72]
Propionibacterium freudenreichiiB12 Vitamin Potential improvement in nutritional and health value of bread[73]
Torulaspora delbrueckiiAromatic compoundsImprove the sensory profile of bread[72]
Lactobacillus reuteri LTH5448 and Lactobacillus reuteri 100-23γ-glutamyl dipeptidesInfluence in the salty taste[74]
Lactobacillus sanfranciscensis, Candida milleri, and transglutaminaseIsodipeptide bonds, ketones, medium-chain fat acids, and alcoholsPositive effects on bread rheological features, shelf-life, and aroma profile[75]
Lactobacillus acidophilus ATCC20552 and Bifidobacterium lactis Bb 12AntimicrobialInhibit rope-forming B. subtilus[76]
Kluyveromyces marxianus and Saccharomyces cerevisiaeInulinaseReduction of fructans, consequently FODMAPs in dough prepared with whole wheat flour[77]
Kluyveromyces marxianusInulinase, phytaseReduction of FODMAPs
Highest porosity and lowest hardness
[62,78]
Saccharomyces cerevisiae and Pediococcus pentosaceusPhytase activityPhytic acid decrease[79]
Gluconobacter oxydans IMDO A845Higher amount of lactic acidPositive aroma profile of sourdough bread[80]
Leuconostoc citreum FDR241glycosyltransferase dextran concentration in sourdough[81]
Enterococcus mundtii QAUSD01 Proteolytic activityGluten-degrading[60]
Wickerhamomyces anomalus QAUWA03Proteolytic activityGluten-degrading[60]
Lactobacillus plantarumPhenolic acid esterase, decarboxylases, reductase, and wide range of glycosil hydrolasesIts influence in bread quality needs study[33]
Weissella cibaria VTT E-153485PeptidaseIncreased proteolysis in faba bean dough [82]
Weissella confusa VTT E-143403DextranIncreased viscosity in faba bean dough[82]
Pediococcus pentosaceus VTT E-153483 and Leuconostoc kimchi VTT E-153484PhytaseReduction of phytic acid in faba bean dough[82]
Lactobacillus amylovorus DSM19280 6% and Weisella cibaria MG1 18%Organic acid and exopolysaccharideLow-salt bread with desirable shelf-life, and high sensory quality (volume and crumb texture)[83]
Lactobacillus brevis and Lactobacillus plantarum at 35 °CVolatile compoundsImprove texture and aromatic properties of sourdough bread[84]
Lactobacillus reuteriOrganic acid, saturated fatty acid, hydroxy fatty acidAnti-aflatoxigenic capability and antifungal activity[85]
Lactobacillus plantarum 29DAN and L. plantarum 98APolyphenolAntioxidant activity[86]
Lactobacillus plantarum NOS7315, Lactobacillus rossiae NOS7307, Lactobacillus brevis NOS7311, and Saccharomyces cerevisiae PS7314Synergistic fermentationImproved bread sensory characteristic [87]
Bacillus licheniformisExopolysaccharides (EPS)
Immunomodulatory potential[88]
Lactobacillus paracasei K5Organic acid, higher the complexity of volatile compoundsDecrease spoilage, increase shelf-life, and improve sensory properties in sourdough bread[89]
Enterococcus mundtii QAUSD01 and Wickerhamomyces anomalus QAUWA03Proteolytic activityToxic gliadin degraded in the sourdough fermentation[90]
Lactobacillus plantarum CH1Antifungal compoundsDo not interfere in the sensory quality of bread[91]
Streptococcus thermophilusGlucosyltransferase BBread with lowly digestible starch and textural improvements[92]
Pediococcus pentosaceus SP2Organic acid contentReduce mold and rope spoilage[93]
Enterococcus faecium and Kluyveromyces aestuariiHigh phenolic and antioxidant capacity, respectivelyImprove bread quality[62]
Lactobacillus reuteri TMW1.656ReutericyclinInhibition of growth of rope-forming bacilli in bread[94]
Wickerhamomyces anomalus P4PhytaseDecrease phytate and increase mineral solubilization in sourdough bread[95]
Levilactobacillus brevis TMW 1.211, Pediococcus claussenii TMW 2.340 from breweriesO2-substituted (1,3)-β-D-glucan Improving water binding capacity[96]
Weissella confusa/cibaria 3MI3 from sourdoughDextran Technological properties of dough and bread, such as water absorption, rheology, stability in cold storage, bread staling, and syneresis of starch gels/avoided the resistant starch formation[97]
Pediococcus lolii B72 and Lactiplantibacillus plantarum E75 from mature sourdoughVolatiles compoundsImproving sensorial acceptability[98]
The microbial production of bioactive peptides, enzymes, organic acids, exopolysaccharides (EPS), and vitamins are considered the primary metabolites responsible for antioxidant, antimicrobial, and probiotic activities [99]. Relative to fermentative processes, many bioproducts are present in sourdough fermentation. Koistinen et al. [100] demonstrated 118 bioactive compounds in sourdough fermentation. Some of these health benefits and others are summarized in Table 3.
Table 3. Microorganisms and the compounds they produce, and the health benefits.
Table 3. Microorganisms and the compounds they produce, and the health benefits.
MicroorganismsCompoundBenefitReference
Probiotics Streptococcus thermophilus, Lactobacillus plantarum, L. acidophilus, L. casei, L. delbrueckii spp. bulgaricus, Bifidobacterium breve, B. longum, and B. infantisPeptidaseAlfa-gliadin degradation, reduced wheat allergenic[101]
LAB from sourdoughGamma-aminobutyric acid (GABA)ACE-inhibitory activity[102]
LAB from sourdoughMultifactors Low-glycemic index in the white wheat bread[103]
Lactobacillus reuteriExopolysaccharideAntiadhesive properties, inhibition enterotoxigenic Escherichia coli[104]
Lactobacillus brevis with S. cerevisiae var. Chevalieri; L. Fermentum; L. Fermentum with phytaseHigher total phenolic and a lower molar ratio of lactic to acetic acidReduce glycemic index[105]
L. curvatus SAL33 and L. Brevis AM7Peptide lunasinCancer preventive[106]
Bifidobacterium strainsPhytaseIncrease iron bioavailability in bread[107]
Weissella ciabaria MG1; L. reuteri VIP, L. reuteri Y2OligosaccharidesImproved nutritional quality of sorghum bread[108]
L. brevisPhytaseDecrease phytate levels, improve mineral bio-accessibility[109]
L. Sakei KTU05-6Organic acids, bacteriocinsBio-preservative [110]
Weissella confusa LBAE C39-2Dextransucrase (glycoside hydrolase)Alfa-glucans/ oligosaccharides or glycoconjugates[111]
L. rossiae DSM 15814 from sourdoughVitamin B12, folate, and riboflavinNutritional improvement[112]
Lactobacillus amylovorus DSM 19280 and Weisella cibaria MG1Glutamate accumulationNaCl reduction in bread [83]
Traditional sourdough LAB starter cultureEssential and non-essential amino acids, flavonoids, antioxidant peptidesNutritional improvement protects against oxidative stress and degenerative disease through phenolic compounds[113]
LAB from traditional Austrian sourdoughsFructose metabolized/antifungal and anti-bacillus propertiesDecrease FODMAPS/ control molds[114]
Lactobacillus plantarum ZJUFT17 from Chinese sourdoughIn mice, decreased: the profile, insulin resistance, lipopolysaccharide, cytokines interleukin (IL)-1β, tumor necrosis factor (TNF)-αManaging gut microbiota, decreasing pathogenic and pro-inflammatory microbes, and stimulating anti-obesity ones[115]
Levilactobacillus brevis TMW 1.211, Pediococcus claussenii TMW 2.340 from breweriesO2-substituted (1,3)-β-D-glucan Prebiotic effect in bread, improving water binding capacity[96]
Lactobacillus plantarum ZJUFB2 from Chinese sourdoughProbiotic effect on gut microbiotaPrevent insulin resistance and modulate gut microbiota.[116]
Levilactobacillus brevis TMW 1.2112, Pediococcus claussenii TMW 2.340Dietary fiber, short acid fat chain SCFA, butyrate, propionate,β-glucanHealthy environment in the colon, chemopreventive[117]
Pediococcus pentosaceus F01, Levilactobacillus brevis MRS4, Lactiplantibacillus plantarum H64, and C48Γ-aminobutyric acid (GABA)Bread from surplus bread with high nutritional value[118]
Weissella cibaria PDER21α-D-glucanAntioxidant properties[119]
One important point is that most probiotics die when the bread is baked at high temperatures. However, most health benefits continue, mostly, not with probiotics at intestinal epithelial cell colonization but in sourdough dough. It is essential to note that active biomolecules could be present in the baked bread. Different conditions, varying temperatures, baking time, and fermentation parameters could influence the heat resistance of some bioproducts, and they could be present as nutrients and fibers such as β-glucan and resistant starch [120]. Increasingly, the literature discusses this important point. If the baking time is shortened by increasing the baking temperature or reducing the bread size, higher residual viability of LAB species may be obtained after baking [121]. Studies with L. plantarum at three different baking temperatures showed a reduction from 8.8 log CFU/g to 4–5 log CFU/G [122]. Concerning yeast, the count decreased from an initial value of 9 log CFU/g to 4 log CFU/g with baking at 200 °C for 13 min [121].
During the baking process, cell lysis of microorganisms delivers cellular debris that may also have beneficial properties functioning as postbiotics. Examples of postbiotic compounds that are present in sourdough are short-chain fatty acids (SCFAs), secreted proteins/peptides, bacteriocins, secreted biosurfactants, amino acids, flavonoids, EPS, vitamins, organic acids, and other molecules discussed in this review. Remains of microbial cell structures such as peptidoglycan from bacterial cell walls, pili, fimbriae, flagella, cell-surface-associated proteins, cell-wall-bound biosurfactants, and cell supernatants are also postbiotic components that have potential health benefits in the host [47,123].
There is both a growing demand from consumers for additive-free, safe, and nutritious foods and for bread with a longer shelf-life and less staling due to microbial spoilage, and sourdough offers some advantages [124].
Several microorganisms, mainly LAB, have antimicrobial properties, which favors their use as probiotics and as a bioprotective culture in fermented products. Koistinen et al. [125] suggest that metabolites with this effect produced by LAB may synergistically modulate the local microbial ecology, such as in the gut. Some strains of LAB can produce a variety of other antimicrobial compounds, such as organic acids (lactic, acetic, formic, propionic, and butyric acids). Organic acids have been described with antifungal actions against Aspergillus, Fusarium, Monilia, and Penicillium [126]. Ethanol, fatty acids, enzymes, acetoin, hydrogen peroxide, diacetyl, antifungal compounds (propionate, phenyl-lactate, hydroxyphenyl-lactate, cyclic dipeptides, and 3-hydroxy fatty acids), bacteriocins (nisin, reuterin, reutericyclin, pediocin, lacticin, enterocin, and others), and bacteriocin-like inhibitory substances (BLIS) are also described in the literature [127].
Bacteriocins are a class of antimicrobial proteins or peptides considered safe for the body and do not induce resistance. This antimicrobial can be applied against harmful food pathogens and used as a food preservative (preserving foods such as dairy products, canned foods, and meats) and a natural antiseptic with bactericidal or bacteriostatic effects. Lactococcus lactis is the microorganism producing nisin, the first fully identified bacteriocin [128]. Among lactic acid bacteria, one of the most versatile and adaptable is Lactiplantibacillus plantarum, with a long history of use in foods and claims of benefits as a probiotic. This species has shown promising results in protecting post-harvest fruits and controlling mold in foods such as beverages and bread [129]. Lactiplantibacillus plantarum strain ITM21B has been used to prolong bread shelf-life due to the production of antimicrobial substances, such as lactic, acetic, phenyl lactic (PLA), and hydroxy-phenyl lactic (OH-PLA) acids [130]. Varsha et al. 2014 [131] demonstrated that the antifungal compounds produced during fermentation resist sterilization temperatures and remain active adding value for the baking industry.
The antimicrobial effect reduces the salt content in bread. Helping reduce sodium consumption to the recommended level of 5–6 g/day will benefit health, particularly regarding blood pressure and cardiovascular disease. Taking into consideration that bread and cereals represent the major sources of salt in the human diet, microbial bioprocessing has been proposed to reduce salt content while keeping the flavor [132].
In addition, the hydrolysis of flour proteins leads to the production of amino acids, organic acids, and other metabolites, improving the nutritional and functional values of sourdough bread [133].
Dietary sourdough bread increases short-chain fatty acid, isovaleric and 2-methyl butyric acids, and γ-aminobutyric acid (GABA) content, among other bioactive molecules [134]. The γ-aminobutyric acid GABA is a non-protein amino acid synthesized from L-glutamate by the glutamate decarboxylase [135]. It is a mammalian neurotransmitter involved in critical regulatory functions, hypotensive properties, anti-depressive effects, diuretic, and antioxidant effects [136]. It is present in many medicines and supplements [137].
Several strains of LAB can carry out activities that promote human health, such as modulating the immune response, preventing cancer, reducing chronic intestinal inflammation, and cholesterol levels. Other functions include improving the intestinal barrier, inhibiting pathogenic organisms, and beneficial interactions with the endogenous intestinal microbiota [138]. These bacteria are present in sourdough and liberate bioproducts in bread dough, such as vitamin B12, folate, and riboflavin [112].
Although probiotics from sourdough are not present in the intestine environment, they provide gastrointestinal benefits, including the pre-digestion of non-nutritional molecules. Sourdough bacteria and yeast decrease the phytic acid present in wheat. The phytate (myo-inositol hexaphosphate) content of whole wheat products is a concern because it inhibits the absorption of minerals, has a negative effect on the nutritional properties, and causes gastrointestinal disorders, digestive discomfort, and flatulence. On the other hand, removing wheat bran and germ reduces the nutritional value of bread. The acidification during sourdough fermentation stimulates endogenous grain phytase, but LAB and yeast phytase activity significantly reduces the phytate content of the bread [124]. Another advantage is the reduction of fermentable oligo-, di-, and monosaccharides and polyols such as sorbitol and mannitol (FODMAPS) and amylase–trypsin inhibitors (ATIs) from wheat. Both are related to trigger intestinal symptoms in irritable bowel syndrome and non-celiac wheat sensitivity. A study by Boakye et al. 2022 [139], described that the longer sourdough fermentation time (12 h) caused up to 69%, 69%, and 41% reductions in fructans, raffinose, and ATIs, respectively. Low FODMAPs may be valuable for people with gastrointestinal disorders.
The reduction of gluten content is important for people with gluten-related disorders. The long fermentation process of sourdough also helps break down gluten proteins in wheat. Moreover, certain bacteria produce peptidases. Some Lactobacillus strains, including L. sanfranciscensis, L.rossiae, L. plantarum, L. brevis, L. pentosus, L. alimentarius, L. fermentum, L. paracasei, L. casei subsp. casei, and P. pentosaceus degrade gluten. Other gluten-degrading bacteria have been isolated form sourdough, such as Bacillus spp. [140], and Enterococcus mundtii and the yeast Wickerhamomyces anomaly from locally fermented sourdoughs (Khamir) [60]. Gluten has baking properties, but its protein components, gliadin, and glutenins, could trigger undesirable immune responses, causing inflammation and damage [141].
Antioxidant and anti-inflammatory activities of peptides from cooked sourdough breads were described [142], where longer fermentation (72–96 h) increases the production of aromatic compounds with antioxidant activity [143].
LAB and yeast strains were isolated and molecularly identified from traditional Iranian sourdough. Based on total phenol production and antioxidant capacity assessments, all the identified traditional sourdough microorganisms significantly produced phenolic compounds. They showed significant antioxidant capacity improving bread’s health benefits and quality [62].
EPS such as β-glucan, dextran, and inulin, are metabolites from LAB identified in sourdough fermentation. The β-glucan, for instance, is a prebiotic homopolysaccharide formed by glucose, presenting substantial health benefits, such as stabilized cholesterol levels, anti-inflammatory effects, and benefits for probiotic microorganisms [96]. The conclusions reached by Schlörmann et al. [117] provide significant insights into the general chemopreventive and prebiotic effects of LAB-generated sourdoughs and bread. In addition, exopolysaccharides are produced by LAB in sourdough, favoring great water retention. β-glucans also contribute to the viscoelastic properties of dough [96]. In the case of thermal treatment at 60 °C and 80 °C for up to 60 min, the molecular mass of the EPS was not significantly affected [144].
The glycemic index (GI) is the speed at which sugar enters our body’s bloodstream and this parameter is affected by probiotics from sourdough. It is reported that sourdough bread has a lower GI than yeast bread. Bread with carbohydrates and starches, which are rapidly assimilated, increase the GI. An explanation of the decrease in the glycemic index is the higher level of lactic acid produced under these fermentation conditions. Lactic acid effectively diminishes postprandial glycemic and insulinemic responses, whereas acetic acid brings out a delayed gastric emptying rate. Moreover, the starch availability is reduced under baking heat, thus allowing the breakdown of sucrose to form EPS such as glucans which contribute to the rise in dietary fiber content [145,146]. Mutlu et al. [147] discussed that LAB could contribute to lowering the glycemic index by limiting the digestion rate of starch and converting glucose from digested starch into several bioproducts such as organic acids, fatty acids, sugar alcohols, bioactive peptides, and indigestible exopolysaccharides. Similar results were found by Demirkesen-Bicak et al. [148], where the authors demonstrated results supporting the consumption of wheat and whole wheat bread produced by type 2 fermentation due to higher resistant starch and slowly digestible starch, and lower rapidly digestible starch and glycemic index values showing beneficial effects. However, some controversy exists in the literature. A systematic review selected 18 studies published up to June 2021 in the EMBASE, MEDLINE, Scopus, and Web of Science databases. The authors concluded that the consumption of sourdough bread has a lower influence on blood glucose compared to that of industrial bread or glucose [149].
In addition to the health effects, sourdough brings flavor, aroma, and better texture of bread due to LAB enzymatic hydrolysis processes. Flavor from volatile compounds in sourdough are due to peptidases conversion of glutamine released from cereal proteins to glutamate, and conversion of arginine to ornithine by LAB during baking. Microbial and enzymatic reactions, such as lipid oxidation by cereal enzymes, improve flavor and texture. LAB and yeast in sourdough produce aroma precursors such as free amino acids, which lead to the generation of aldehydes or corresponding alcohols [126,150]. Table 3 summarizes the functionality of several microorganisms studied in this fermentation. It was demonstrated that sourdough bread fermentation conditions considerably influence bioactive compound viability [100]. Furthermore, many studies have described unique characteristics and molecules formed from the selective use of inoculum as a starter in bread making.
Using a starter culture and varying the temperature and fermentation time changes the flavor and texture in sourdough bread [84]. Moreover, selected LAB and manipulated temperature of sourdough fermentation could increase the β-glucan yield [96]. Methods for maintaining viable starter cells for use in the baking industry have been studied widely, such as immobilization techniques [93]. Studies showed that using the spray drying technique to preserve the viability of lactic acid bacteria and yeasts in the dry mass was much more effective than other drying techniques [5]. However, this area needs further studies since they can vary with the cell type or microbial consortia [151]. Studies have been realized to develop better kinetic control in the spray drying process of LAB, which is fundamental for their growth and viability [152]. Understanding the stress response mechanisms used for LAB can help industries develop better processes to produce LAB at a large scale [153]. Encapsulation of LAB has been demonstrated as a method to improve the viability of LAB during drying processes [154]. For instance, microencapsulation of cells by spray drying with cellulose acetate phthalate was more efficient than with maltodextrin [155]. Using probiotics in bread is a challenge due to the higher temperature of baking. However, a recent study showed that the strain of Lactobacillus plantarum P8 survived baking and increased the number of its cells during storage [122].

5. Enzymes in Sourdough

Enzymes are responsible for several biochemical events in sourdough fermentation. Although some enzymes are present in the cereal, most of them are produced by microorganisms. From this perspective, the selection of native yeasts in sourdough for use in the baking industry has increased, and they are better adapted to improve their stability [156,157]. In the same way, bacteria, especially lactic acid bacteria, are widely selected for use in industrial bread making [71].
Many valuable properties observed in the select starters of bacteria and yeasts are related to their enzyme activities or involvement in the enzyme activities in the sourdough system. The wheat grain has enzymes that can be activated or inhibited by the low pH of the sourdough environment, and the microbiota present in sourdough contributes to its enzymes in the sourdough metabolic process [3]. The mechanisms of LAB and yeast enzymatic action during sourdough fermentation are becoming well-known [54]. The hydrolases demonstrate an essential role in sourdough bread. However, other enzyme classes, including transferases and oxidoreductase, are involved in the properties of sourdough fermentation [81,158]. Figure 3 shows the sourdough fermentation process and the major enzymes involved.

5.1. Transferases

The more expressive metabolic activities of microbial communities during fermentation are acidification, leavening, and flavor formation, all related to the metabolism of carbohydrates [159]. Transferases are enzymes that catalyze the transfer of specific functional groups from one molecule to another. Glycosyltransferases (EC 2.4) are involved in the biosynthesis of exopolysaccharides in sourdough. It is an enzymatic group produced by the LAB from sourdough microbiota, such as Lactobacillus, Leuconostoc spp., Leuconostoc citreum, and Weissella species. Glucansucrases are glycosyltransferases from LAB that split sucrose. The resulting glucose builds EPS-type α-glucan polymers such as dextran, mutan, alternan, and reuteran. Fructansucrase, found in Gram-negative and Gram-positive bacteria, is another glycosyltransferase that synthesizes levan (levansucrase) or inulin (inulosucrase). The inulosucrase is observed only in LAB [160,161].
EPS formed during sourdough fermentation influences the dough’s viscoelastic properties and improves the texture and shelf-life. In this context, EPS can replace hydrocolloids used as bread improvers and meet consumer demands for reduced use of food additives [162].
Adding a glucosyltransferase B, made by Streptococcus thermophilus with the activity of 4,6-α-glucanotransferase in the preparation of wheat flour dough, results in bread with better health benefits. This enzyme increased the number of short branches and percentage of (α1→6) linkages in starch, resulting in slow starch digestibility and low retrogradation properties. It also reduced hardness, gumminess, and chewiness, indicating improvements in texture [92].
Transglutaminase (TGase, EC 2.3.2.13) catalyzes the cross-linking of proteins between the ε-amino group of a lysine residue and the γ-carboxamide (acyl) group of a glutamine residue. Which results in an intra- or intermolecular bridge highly resistant to proteolysis, which leads to an increase in the structure and texture of protein substrates [163,164]. It is an important enzyme for the food industry, where studies report that its application improves dough handling properties, increases bread stability, and volume. The protein polymers resulting from the action of TGase can modify the rheological properties of gluten, transforming a weak gluten into a strong one [75].

5.2. Oxidoreductases

Oxidative enzymes are applied to treat wheat flour to restore the gluten network. Those enzymes can be divided into exogenous oxidative enzymes (laccase, glucose oxidase, hexose oxidase) and endogenous oxidative enzymes (tyrosinase, peroxidase, catalase, sulfhydryl oxidase, lipoxygenase, protein disulfide isomerase, NAD(P)H-dependent dehydrogenase, thioredoxin reductase, and glutathione reductase) [165].
Oxidoreductase (EC 1) catalyzes reactions of electron transfer. Glutathione reductase (EC 1.8.1.7) is the most known enzyme in this group related to sourdough bread properties. It acts on a sulfur group of donors, with NAD+ or NADP+ as the acceptor [166].
Gluten protein presents intra- and intermolecular chains of disulfide bonds between amino acids [167]. The glutathione reductase activity, identified in the microbiota of natural fermentation, interferes with these thiol bonds [158]. The glutathione reductase deficient mutant strain (Lactobacillus sanfranciscensis) produced bread with a softer texture and higher specific volume than bread made with traditional biological yeast and non-mutant Lactobacillus sanfranciscensis activity [167].
Glucose oxidase (EC 1.1.3.4) and pyranose oxidase (EC 1.1.3.10) are oxidoreductases produced by several fungi, mostly from Aspergillus niger in the case of glucose oxidase. One molecule of glucose oxidase catalyzes the conversion of β-D- glucose into hydrogen peroxide (H2O2) and gluconic acid [168]. According to Xu et al. [169] the resulting H2O2 can further oxidize the free sulfhydryl units in the gluten protein and promote the formation of disulfide bonds in the gluten network, thereby strengthening the gluten network. In addition, glucose oxidase has antimicrobial activity against fungi and foodborne bacteria [170]. These effects enhance the rheological properties, volume and height, and stability of the bread [169,171,172].
Laccases (benzenediol: oxidoreductase oxygen, EC 1.10.3.2) are multi-copper enzymes and they belong to a group of so-called polyphenol oxidases. They catalyze the oxidation of a wide range of substrates while simultaneously reducing molecular oxygen to water [173,174,175]. Laccase activity has been found to enhance the viscoelastic dough qualities and in recent years it has been used as a dough and bread improver, primarily utilized in commercial preparations [175,176]. Pentosans are polymerized by laccase and favor the aggregation of glutenins, which lowers the protein content of glutenin macropolymers. While not all researchers agree, some authors report that laccases are also important in baking because they can cross-link the arabinoxylans found in whole flour, improving the gluten index and inducing hetero-crosslinking between arabinoxylans and the gluten matrix [165].

5.3. Lyases

Lyases (EC 4) are a group of enzymes that cleave carbon bonds with oxygen, carbon, nitrogen, and others and eventually form double bonds or a new ring [177].
The glutamate decarboxylase (EC 4.1.1.15) is commonly found in the sourdough environment and acts in L-glutamate’s decarboxylation to form γ-aminobutyric acid (GABA), a non-protein amino acid with various health benefits.
Amino acid catabolism by LAB is related to sourdough properties such as taste and nutrition. However, these reactions can accumulate toxic biogenic amines, especially tyramine, cadaverine, and putrescine. It is well-known that a high GABA content in foods is highly desirable as it is a non-protein amino acid with many health benefits, and some LAB, such as Levilactobacillus brevis, are excellent producers [178,179]. LAB and yeast are the most important GABA producers because they are commercially valuable as starters in fermented foods. GABA has several physiological functions, such as inhibiting neurotransmission in the central nervous system. Other related functions include relaxation, sleeplessness, enhanced immunity under stress, increasing the concentration of growth hormone in plasma, preventing diabetes, inhibiting the invasion and metastasis of various types of cancer, anti-inflammatory effects, blood pressure regulation, and antioxidant effects. It is also a hormonal regulator (for review, see Diana et al., 2014 [178] and Polak et al. 2021 [179].
Indeed, fermented foods represent an excellent source of dietary GABA. In a recent study, Lactobacillus brevis A7 and Lactobacillus farciminis A11, demonstrated high GABA-producing capability when used to prepare bread containing 20% amaranth flour (amaranth has a high concentration of essential amino acids) [180].
Calabrese et al. [181] analyzed spontaneous sourdough metagenomes and transcriptomes to identify microorganisms and their role in the fermentation process. Based on meta-omics data collected and their identified microbiota, a reconstructed synthetic microbial community was established. The results revealed a complex network with dominant, subdominant, and satellite communities engaged in different functional pathways. Aspartate ammonia-lyase (EC 4.3.1.1), the enzyme that converts aspartate into fumarate, was uniquely transcribed by Lactiplantibacillus plantarum. This approach was effective and significant for studying these complex metabolic frameworks bringing new knowledge into this complex fermentation.

5.4. Hydrolases

Hydrolases (EC 3) are enzymes that use water to break various chemical bond types.
Amylase is the most popular hydrolase in baked good processes. However, peptidases have received more attention recently due to investigations into their role with gluten. Other hydrolases have essential roles in the baking processes, such as cellulase, xylanase, and pentosanase [182]. Table 4 presents the main enzymes used in the bakery and the microorganisms that produce them, according to Dahiya et al. [183].
Table 4. List of microorganisms that produce the main enzymes used in the bakery.
Table 4. List of microorganisms that produce the main enzymes used in the bakery.
EnzymeMicroorganismsReference
XylanaseSporotrichum thermophile BJAMDU5[184]
Pichia pastoris[185]
Bacillus subtilis[186]
Myceliophthora thermophila BJTLRMDU3[187]
PhytaseLactobacillus casei[188]
Enterobacter sp. ACSS[189]
Sporotrichum thermophile[190]
Aspergillus niger NCIM 563[191]
AmylaseRhizopus oryzae[192]
Bacillus subtilis US586[193]
Bacillus subtilis M13[194]
Streptomyces badiun DB-1[195]
Glucose OxidaseAspergillus niger[196]
Penicillium notatum[197]
Aspergillus niger[198]
Aspergillus niger[199]
PeptidaseRhizopus oryzae[192]
Bacillus subtilis PF1[200]
Bacillus subtilis[201]
Bacillus pumilus SG2[202]
LipaseAspergillus niger MTCC 872[203]
Pseudomonas fluorescens (NRLL B-2641)[204]
Bacillus subtilis I-4[205]
Bacillus sp. MPTK 912[206]
CellulaseSporotrichum thermophile BJAMDU5[203]
Streptomyces strain C188[207]
Cellulomonas uda[208]
Trichoderma reesei NCIM 1186[209]

5.4.1. Amylase, Inulinase, and Their Impact on Bread Structure and Properties

Several types of amylases act on starch, mainly liberating glucose for the fermentation and improvement of the bread’s color, flavor, and shelf-life [210]. The main groups are endoamylases, exoamylases, and debranching amylases. Among the endoamylases, α-amylase (EC 3.2.1.1) is the main one, acting on α-1,4 bonds in the amylose and amylopectin chains. The group of exoamylases includes glucoamylase (EC 3.2.1.3), α-glucosidase (EC 3.2.1.20), and β-amylase (EC 3.2.1.2) promoting the hydrolysis of terminal glycosidic units. Debranching amylases act on α-1,6 branch bonds, and pullulanase (EC 3.2.1.41) is the leading representative [211].
Starch/amide is present in around 70–75% of wheat flour. It comprises a linear chain of α-1,4 glycopyranoside (amylose) and a linear chain of α-1,4 glycopyranoside with α-1,6 branch points (amylopectin). During the baking process, the bread undergoes several changes in starch retrogradation relative to the crystalline structure of amylose and amylopectin. These changes occur not only in baking but also during the cooling and storage processes [212]. Starch retrogradation is a process in which disaggregated amylose and amylopectin chains in a gelatinized starch paste reassociate to form a more ordered structure [213]. It is a pivotal issue in bread aging and is the cause of bread hardening.
The amylases produced by LAB strains reduce the aging process during bread storage [214]. Amylase-producing microorganisms in the sourdough environment are essential in converting starch into fermentable carbohydrates such as maltodextrins, maltose, sucrose, and glucose. LAB are sources of maltose phosphorylase, which generate D-glucose, β-D-glucose, 1-phosphate, and 1,6-α-glucosidase, which hydrolyzes α-(1–6)-glucooligosaccharides [3]. In addition, wheat has α-amylase, β-amylase, and glucoamylase activity, but at pH < 4.5, only glucoamylase maintains its activity. Microorganisms in the sourdough microbiota produce inulinase (β-2, 1-D-fructan-fructan-hydrolase—EC 3.2.1.7)—a glycosylase that hydrolyzes the β-2.1 bonds of fructose of the fructose polymer, inulin. Inulinase can reduce the number of oligosaccharides, disaccharides, monosaccharides, and fermentable polyols (FODMAPs) in bread [78]. FODMAPs are associated with irritable bowel syndrome. However, it is not appropriate to eliminate these polyols but only to reduce them to an adequate level, as these compounds are considered prebiotics to maintain the healthy intestinal microbiota [215]. Kluyveromyces marxianus yeast strains have reduced the concentration of FODMAPs in whole wheat bread [78].

5.4.2. Cellulase, Phytase, and Xylanase for Mineral Bio-Accessibility Improvement in Bread

Cellulases are classified into endoglucanases (act in the middle of cellulose) and exoglucanases (cleave the extremities of the polymer). The endoglucanase is the endo-β-1,4-glucanase (EC 3.2.1.4), cleaving intramolecular bonds of β-1,4-glycosidic. The exoglucanases are the β-glucosidase (EC 3.2.1.21) and exo-β-1,4-glucan cellobiohydrolase (EC 3.2.1.91), which cleave ends of the cellulose and glycosidic terminals (liberating Cellobiose), respectively [216].
Endocellulase acts on cellulose and β-glucan substrates, removing insoluble arabinoxylans, contributing to gluten network formation [217]. Subsequently, there is a decrease in the hardness of the bread and, consequently, an improvement in the sensory evaluations. Cellulase action may also provide an anti-staling effect, which may be related to alterations in water distribution between the starch–protein matrix [182]. In the study carried out by Li et al. (2014) [218], the authors added β-glucanase to a mixture of barley flour (30%) and wheat flour (70%) with approximately 1.5% glucan as the starting material. As a result, they noticed an improvement in the maneuverability of the dough, an increase in softness and elasticity, an increase in the specific volume of the bread, and a reduction in hardness.
The enzyme phytase or myoinositol-hexaphosphate phosphohydrolase is present in plants, bacteria, yeasts, and filamentous fungi. It hydrolyzes phytate molecules and releases inorganic phosphate, dephosphorylation-dependent inositol esters, and minerals linked in the phytate structure [219]. Two classes of phytase are categorized depending on where the hydrolysis of phytate begins, -3-phytase (EC 3.1.3.8) or 6-phytase (EC 3.1.3.26); the 3-phytase being the one that removes the orthophosphate group from the position C3 characteristic of microorganisms. Two classes of phytase are found depending on the position where the hydrolysis of phytase -3-phytase (EC 3.1.3.8) or -6-phytase (EC 3.1.3.26). The 3-phytase removes the orthophosphate group from the position C3 characteristic of microorganisms [220].
Xylan is the main hemicellulosic component of hardwoods and accounts for approximately 30% of the woody cell wall. Xylanases (EC 3.2.1.8) are enzymes that hydrolyze this polysaccharide. They are of two types: β-1,4-endoxylanase cleaving internal glycosidic linkages, and α-D-xylosidase which cleaves xylo-oligosaccharides, and xylobiose forming xylose, a disaccharide of two xylose molecules [211]. It is important to highlight that xylanase from various sources has different mechanisms. Two mechanisms can summarize the improvement in bread quality due to the addition of xylanase: (i) the removal of arabinoxylans from gluten alters the distribution of water between gluten proteins and arabinoxylans; (ii) pentosan destruction and viscosity reduction effect [182].
Aslam et al. [221] reviewed that wheat is a reservoir of minerals, mainly in the aleurone tissue between the endosperm and the seed coat. However, these minerals are complexed with phytate molecules that reduce their bio-accessibility. Xylanase and cellulase showed satisfactory results to facilitate access into the aleurone tissue for the action of phytase, improving iron bio-accessibility [222].
Although some aspects of mineral assimilation by the human body need to be considered, such as diet, blood flow, gut epithelium, and intestinal microbiota [223], some LAB and yeasts may have phytase activity that improves mineral bio-accessibility in sourdough bread [79,224,225,226]. Bifidobacterium strains have demonstrated the ability to increase the accessibility of iron through its phytase activity during bread fermentation [107].
Pentosanase is a hemicellulase that acts on pentosans, particularly hydrolyzing water-unextractable arabinoxylan (WUA) into water-extractable arabinoxylan (WEA). The WUA represents 85% of the total arabinoxylan in whole wheat flour. The WEA affects the bread making process by reducing viscosity and making the dough more malleable. However, WUA promotes lower bread quality for interfering with the gluten network because it competes with gluten for water [169]. Adding pentosanase has shown a good outcome for bread making, with higher specific volume and crumb texture improvement [227]. Moreover, Martínez-Anaya et al. [228] showed that sourdough with starter modified the results of pentosanase used in the bread making process, with the fraction of soluble pentosan increased considerably, and a lower xylose/arabinose ratio.

5.4.3. Lipase and the Baking Technology

Lipase and esterase (EC 3.1) catalyze the reaction that forms ester bonds. The sub-class triacylglycerol acyl hydrolase (EC 3.1.1.3) acts on triglycerides releasing fatty acids and glycerol.
Both phospholipase A1 (EC 3.1.1.32) and phospholipase A2 (EC 3.1.1.4) liberated fatty acids on phospholipids and other lipophilic compounds. These groups are the most important in the bread making process [210,211,229].
In baking, using lipase to replace emulsifiers, such as monoglycerides, proved to be an efficient alternative to replace chemical additives [230]. Triacylglycerol acyl hydrolase can break down triglycerides, which interferes negatively with the gas retention gluten properties [210]. Sourdough fermentation inhibits the lipase activity due to decreased pH in the system, which is an efficient alternative for neutralizing the lipase activity in wheat germ [231]. Wheat germ, rich in vitamins, amino acids, fiber, and minerals, is nutritionally more important in wheat. However, it causes oxidation in wheat flour during storage because this needs to be removed before producing wheat flour [232]. A study showed that wheat germ could produce bread with a superior nutritional, textural, and sensory quality after the sourdough process to obtain sourdough-fermented wheat germ free of oxidation potential [233].

5.4.4. Peptidase and Implications for the Gluten Network

Peptidases hydrolyze peptide bonds of proteins and peptides. They are still known as proteases, proteinases, and proteolytic enzymes [234].
The International Union of Biochemistry and Molecular Biology (IUBMB) classifies peptidases with criteria based on reaction and catalytic type. Endopeptidases (EC 3.4.21-99) hydrolyze internal peptide bonds, while exopeptidases (EC 3.4.11-19) remove amino acids from the ends of peptide chains. These enzymes have a complex classification involving the reaction and catalytic type and the evolutive relations. The MEROPS platform is a complete site about peptidases, including their complete classification [234].
Gluten is a complex protein responsible for wheat bread’s most peculiar characteristics, such as expandability and retention of gas produced in fermentation. Its quality is more important than quantity [235]. Gluten structure varies with the cultivar and cultivation conditions [236]. It is composed of gliadin (α, β, γ, ω), glutenin, and polypeptide fractions that combine with gliadin or glutenin [237]. Figure 4 seeks to demonstrate the complexity of gluten formation and the water effect. Gluten forms a network retaining carbon dioxide during baking, making dough expand. It is responsible for dough’s rheological characteristics due to gliadin, which ensures the dough’s viscosity, and glutenin, with the property of elasticity. During sourdough fermentation, the partial hydrolysis of gliadin and glutenin proteins occurs because of the acidification and activation of cereal peptidases. In addition, endogenous flour peptidases became activated at a low pH reducing the gluten disulfide bonds [38]. Cysteine proteinases in wheat grain are capable of hydrolyzing both gliadins and glutenin. The hydrolysis of glutenins during sourdough fermentation results in depolymerization and subsequent solubilization [238]. Endopeptidases may be applied to improve the rheological properties, and one example is the endo- and exopeptidases from A. oryzae. The fungi peptidases are employed to modify wheat gluten by limited proteolysis resulting in better rheology and shorter dough mixing for a higher loaf [239]. Lactic acid bacteria also exhibit proteolytic activity, decreasing the content of native gluten [231,240,241]. Several studies have focused on LAB peptidases and demonstrated that in different ways, that LAB in sourdough fermentation has a proteolytic activity for gluten hydrolysis [60]. Celiac disease or gluten enteropathy is an autoimmune disease caused by gluten-containing food and results in those with it to adhere to a strict gluten-free diet. One characteristic of gluten is that it is resistant to proteolysis by human digestive peptidases [242]. There is considerable interest in producing gluten-free bread that is specific for celiac consumers. The gliadin fraction, especially α2-gliadin, has been considered the most allergenic part of gluten [243]. Another possible treatment is enzyme therapy, in which peptidases are ingested with food to increase the hydrolysis of resistant immunogenic peptides containing proline and glutamine amino acids [244]. The hydrolysis of some parts of gluten could favor the appearance of allergen terminals and cause more harm than benefit to a celiac person [243]. On the other hand, hydrolyzing gluten to reduce the allergenic components potentially benefits non-celiac people [69]. A group of LAB hydrolyzed gluten, but not the most allergenic part in the central region of alpha-gliadin. However, other allergenic proteins were degraded, improving the wheat digestibility [245]. A bioprospection analysis of gliadin-cleaving proteolytic activity among 20 Lactobacillus strains showed that the most active strain, L. casei, was able to hydrolyze the 33-mer immunogenic peptide of α-gliadin by 82% in 8 h and completely in 12 h [246]. Bread digestibility might be positively affected by long sourdough fermentation [148]. L. rhamnosus, Pediococcus pentosaceus, and Lactobacillus curvatus used gluten as the only source of nitrogen, significantly reducing the allergenic composition of wheat and improving digestibility [69]. Strains of Lactobacillus plantarum ES137 and Pediococcus acidilactici ES22, isolated from sourdough fermentation, showed high proteolytic capacity [247]. Lactobacilli in sourdough have proteolytic capacity to produce γ-glutamyl dipeptides, other peptides, and amino acids as flavor precursor compounds in the bread produced [248]. Other bacteria, such as Bacillus spp. isolated from sourdough, are capable of gluten hydrolysis [140].

6. General Regulation for Microbes Used in Sourdough Bread

Searching for rules that maintain food safety is a critical concern due to the growing use of microorganisms as foods or probiotics. At this moment, there is an effort at Codex Alimentarius to harmonize rules for probiotics and the use of microorganisms in food [251].
In the United States, there are rules defined by the Food and Drug Administration (FDA) for the use of microorganisms in food production. It is necessary to qualify as a Generally Recognized as Safe for Use (GRAS) microorganism [252]. In Brazil, the biological fermenting mixtures and their microbial content are classified as a supporting and processing agent and are exempted from registration by ANVISA [253]. However, food with nutritional allegations or probiotic use needs to be registered [253], and microorganisms isolated from national raw materials, or related territory, are recognized as national genetic patrimony [254], and they need to be registered too.
The requirements for a microorganism to be considered GRAS include taxonomic identification, absence of virulence, enterotoxins, hemolytic activity, and transferable antibiotic resistance genes. A probiotic microorganism still needs to prove stress tolerance (low pH, gastric enzymes, bile salts) up to levels found in the intestine, as well as adhesion capacity and antipathogenic activity, and to be approved in clinical trials [255]. Moreover, FAO/WHO advises in its guidelines that the evaluation of probiotics should be done using molecular information from strains, because the probiotic properties are related to this level of identification [256].
In France, since 1993, a law specifies the characteristics of sourdough bread, such as pH 4.3, acetic acid 900 ppm, at least, with sourdough consisting of lactic acid bacteria and yeasts. It is possible to add Saccharomyces cerevisiae to a maximum of 0.2% on the flour weight. If sourdough is dry, it needs at least one billion live food bacteria and one to ten million live yeasts per gram [257].
In Europe, since 1997, with the novel food regulation, all food that started to be produced after 15 May 1997 is considered a novel food and has specific rules to be followed [258]. For microorganisms usually found in sourdough and used as a starter, it is challenging to define it as a novel food because sourdough predates 1997 [259]. Therefore, if the starter used in the sourdough fermentation is unusual or has some additional novel properties, it could be considered a novel food. For instance, UV-treated baker’s yeast (Saccharomyces cerevisiae) has been considered a novel food since 2017 because it is alleged to increase vitamin D2 [260]. Moreover, the EU1151/2012 recognizes that baker products could have a geographical denomination [11].
Regarding microorganism patents, the United States, Japan, Europe, India, and participants of The Trade-related Aspects of Intellectual Property Rights (TRIPS) agreement consider that microorganisms should be patented [261]. However, Brazil, for instance, allows patenting only for genetically modified microorganisms [262].

7. Conclusions

The role of microorganisms in sourdough bread has been extensively demonstrated in the literature. Increasing information is arising about the fermentation process and other metabolic routes promoting improvements in quality and adding interesting nutritional, health, and sensory characteristics. Currently, several benefits of probiotic and postbiotics elements are well established in terms of the bread’s properties, such as rheological and organoleptic properties, as well as human health and dough quality. However, there are challenges to overcome to produce a microbial starter or consortium with high technological properties on an industrial scale. Better digestibility, satiety, and antioxidants, among other properties, could be obtained. The sourdough ecosystem needs to be explored more concerning the dynamic of the interaction among the microorganisms, including the quorum sensing process and the metabolic regulation of sourdough. Molecular and omics tools are bringing more information about the sourdough microbiome and the complex metabolic network, allowing the complete identification of the non-cultivable microbial population and establishing probiotic strains. A better characterization of the postbiotics present in the dough is also essential. Another promising area of study is relative to the addition of other microbial products that can improve the final quality of bread. Sourdough bread and the microorganisms involved in its production are an area for increased research activity. The breaking and sharing of bread is of biblical and global importance. Further studies will elucidate the mechanisms of action of these bioproducts, metabolic routes, and their effects on bread, and the production of safe food leading to technological advances.

Author Contributions

Conceptualization and writing, I.T.A.; writing, formal analysis, original draft preparation, F.R.P.M.; conceptualization, writing, review, editing, project administration, and funding acquisition, A.B.V. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded in part by the postgraduate program of the Paulo de Góes Institute of Microbiology, Federal University of Rio de Janeiro (UFRJ), through the Coordenação de Aperfeiçoamento Pessoal de Nível Superior (CAPES) [grant number 001], Conselho Nacional de Desenvolvimento Científico e Tecnológico (MCTI-CNPq) grant code [309461/2019-7], and Fundação de Amparo à Pesquisa do Estado do Rio de Janeiro (FAPERJ), grant code [“CIENTISTA DO NOSSO ESTADO” 26/202.630/2019 247088].

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

We would like to thank Andrew Macrae for the English review and SlideModel.com for the diagram in Figure 1. The images in the graphical abstract are from TogoTV (©2016 DBCLS TogoTV) and Servier Medical Art (smart.servier.com).

Conflicts of Interest

The funders had no role in the design of the study; in the collection, analyses, or interpretation of data; in the writing of the manuscript; or in the decision to publish the results.

References

  1. Arora, K.; Ameur, H.; Polo, A.; Di Cagno, R.; Rizzello, C.G.; Gobbetti, M. Thirty Years of Knowledge on Sourdough Fermentation: A Systematic Review. Trends Food Sci. Technol. 2021, 108, 71–83. [Google Scholar] [CrossRef]
  2. De Vuyst, L.; Neysens, P. The Sourdough Microflora: Biodiversity and Metabolic Interactions. Trends Food Sci. Technol. 2005, 16, 43–56. [Google Scholar] [CrossRef]
  3. Gänzle, M.G. Enzymatic and Bacterial Conversions during Sourdough Fermentation. Food Microbiol. 2014, 37, 2–10. [Google Scholar] [CrossRef]
  4. Bartkiene, E.; Özogul, F.; Rocha, J.M. Bread Sourdough Lactic Acid Bacteria—Technological, Antimicrobial, Toxin-Degrading, Immune System-, and Faecal Microbiota-Modelling Biological Agents for the Preparation of Food, Nutraceuticals and Feed. Foods 2022, 11, 452. [Google Scholar] [CrossRef] [PubMed]
  5. Reale, A.; Di Renzo, T.; Preziuso, M.; Panfili, G.; Cipriano, L.; Messia, M.C. Stabilization of Sourdough Starter by Spray Drying Technique: New Breadmaking Perspective. LWT 2019, 99, 468–475. [Google Scholar] [CrossRef]
  6. Erten, H.; Tanguler, H.; Canbaş, A. A Traditional Turkish Lactic Acid Fermented Beverage: Shalgam (Salgam). Food Rev. Int. 2008, 24, 352–359. [Google Scholar] [CrossRef]
  7. Fois, S.; Piu, P.P.; Sanna, M.; Roggio, T.; Catzeddu, P. In Vivo and In Vitro Starch Digestibility of Fresh Pasta Produced Using Semolina-Based or Wholemeal Semolina-Based Liquid Sourdough. Foods 2021, 10, 2507. [Google Scholar] [CrossRef]
  8. Sahin, A.W.; Rice, T.; Zannini, E.; Lynch, K.M.; Coffey, A.; Arendt, E.K. The Incorporation of Sourdough in Sugar-Reduced Biscuits: A Promising Strategy to Improve Techno-Functional and Sensory Properties. Eur. Food Res. Technol. 2019, 245, 1841–1854. [Google Scholar] [CrossRef]
  9. Venturi, M.; Guerrini, S.; Vincenzini, M. Stable and Non-Competitive Association of Saccharomyces cerevisiae, Candida Milleri and Lactobacillus Sanfranciscensis during Manufacture of Two Traditional Sourdough Baked Goods. Food Microbiol. 2012, 31, 107–115. [Google Scholar] [CrossRef]
  10. Coda, R.; Cagno, R.D.; Gobbetti, M.; Rizzello, C.G. Sourdough Lactic Acid Bacteria: Exploration of Non-Wheat Cereal-Based Fermentation. Food Microbiol. 2014, 37, 51–58. [Google Scholar] [CrossRef]
  11. Brandt, M.J. Industrial Production of Sourdoughs for the Baking Branch—An Overview. Int. J. Food Microbiol. 2019, 302, 3–7. [Google Scholar] [CrossRef] [PubMed]
  12. Kezer, G. Functional Perspective on Sourdough Bread. Turk. J. Agric. Food Sci. Technol. 2022, 10, 1410–1414. [Google Scholar] [CrossRef]
  13. Hanson, L.; VandeVusse, L.; Forgie, M.; Malloy, E.; Singh, M.; Scherer, M.; Kleber, D.; Dixon, J.; Hryckowian, A.J.; Safdar, N. A Randomized Controlled Trial of an Oral Probiotic to Reduce Antepartum Group B Streptococcus Colonization and Gastrointestinal Symptoms. Am. J. Obstet. Gynecol. MFM 2023, 5, 100748. [Google Scholar] [CrossRef]
  14. Lee, Y.; Oh, H.; Jo, M.; Cho, H.; Park, Y. Synergistic Effect of N-3 PUFA and Probiotic Supplementation on Bone Loss Induced by Chronic Mild Stress through the Brain–Gut–Bone Axis. J. Funct. Foods 2023, 100, 105363. [Google Scholar] [CrossRef]
  15. Dotterud, C.K.; Storrø, O.; Johnsen, R.; Øien, T. Probiotics in Pregnant Women to Prevent Allergic Disease: A Randomized, Double-Blind Trial: Probiotics in Pregnant Women to Prevent Allergic Disease. Br. J. Dermatol. 2010, 163, 616–623. [Google Scholar] [CrossRef] [PubMed]
  16. Rahim, M.H.A.; Hazrin-Chong, N.H.; Harith, H.H.; Wan-Mohtar, W.A.A.Q.I.; Sukor, R. Roles of Fermented Plant-, Dairy- and Meat-Based Foods in the Modulation of Allergic Responses. Food Sci. Hum. Wellness 2023, 12, 691–701. [Google Scholar] [CrossRef]
  17. Dekker, J.; Quilter, M.; Qian, H. Comparison of Two Probiotics in Follow-on Formula: Bifidobacterium animalis Subsp. Lactis HN019 Reduced Upper Respiratory Tract Infections in Chinese Infants. Benef. Microbes 2022, 13, 341–353. [Google Scholar] [CrossRef]
  18. He, F.; Shen, X.; Cheng, R.; Marotta, F. Screening Potential Probiotics Against Obesity and Metabolism Abnormalities in the Elderly. In Gut Microbiota in Aging and Chronic Diseases; Marotta, F., Ed.; Healthy Ageing and Longevity; Springer International Publishing: Cham, Switzerland, 2023; Volume 17, pp. 375–386. ISBN 978-3-031-14022-8. [Google Scholar]
  19. Shi, S.; Zhang, Q.; Sang, Y.; Ge, S.; Wang, Q.; Wang, R.; He, J. Probiotic Bifidobacterium longum BB68S Improves Cognitive Functions in Healthy Older Adults: A Randomized, Double-Blind, Placebo-Controlled Trial. Nutrients 2022, 15, 51. [Google Scholar] [CrossRef]
  20. Jenkins, G.; Mason, P. The Role of Prebiotics and Probiotics in Human Health: A Systematic Review with a Focus on Gut and Immune Health. Food Nutr. J. 2022, 6, 245. [Google Scholar]
  21. Staniszewski, A.; Kordowska-Wiater, M. Probiotic and Potentially Probiotic Yeasts—Characteristics and Food Application. Foods 2021, 10, 1306. [Google Scholar] [CrossRef]
  22. Belicová, A.; Mikulášová, M.; Dušinský, R. Probiotic Potential and Safety Properties of Lactobacillus plantarum from Slovak Bryndza Cheese. BioMed Res. Int. 2013, 2013, e760298. [Google Scholar] [CrossRef] [PubMed]
  23. Llopis, S.; Hernández-Haro, C.; Monteoliva, L.; Querol, A.; Molina, M.; Fernández-Espinar, M.T. Pathogenic Potential of Saccharomyces Strains Isolated from Dietary Supplements. PLoS ONE 2014, 9, e98094. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  24. Perricone, M.; Bevilacqua, A.; Corbo, M.R.; Sinigaglia, M. Technological Characterization and Probiotic Traits of Yeasts Isolated from Altamura Sourdough to Select Promising Microorganisms as Functional Starter Cultures for Cereal-Based Products. Food Microbiol. 2014, 38, 26–35. [Google Scholar] [CrossRef] [PubMed]
  25. Wang, Y.; Liang, Q.; Lu, B.; Shen, H.; Liu, S.; Shi, Y.; Leptihn, S.; Li, H.; Wei, J.; Liu, C.; et al. Whole-Genome Analysis of Probiotic Product Isolates Reveals the Presence of Genes Related to Antimicrobial Resistance, Virulence Factors, and Toxic Metabolites, Posing Potential Health Risks. BMC Genom. 2021, 22, 210. [Google Scholar] [CrossRef]
  26. Jiang, S.; Cai, L.; Lv, L.; Li, L. Pediococcus Pentosaceus, a Future Additive or Probiotic Candidate. Microb. Cell Factories 2021, 20, 45. [Google Scholar] [CrossRef]
  27. De Vuyst, L.; Comasio, A.; Kerrebroeck, S.V. Sourdough Production: Fermentation Strategies, Microbial Ecology, and Use of Non-Flour Ingredients. Crit. Rev. Food Sci. Nutr. 2022. [Google Scholar] [CrossRef]
  28. Aplevicz, K.S.; Ogliari, P.J.; Sant’Anna, E.S. Influence of Fermentation Time on Characteristics of Sourdough Bread. Braz. J. Pharm. Sci. 2013, 49, 233–239. [Google Scholar] [CrossRef] [Green Version]
  29. Altilia, S.; Foschino, R.; Grassi, S.; Antoniani, D.; Dal Bello, F.; Vigentini, I. Investigating the Growth Kinetics in Sourdough Microbial Associations. Food Microbiol. 2021, 99, 103837. [Google Scholar] [CrossRef]
  30. Carbonetto, B.; Nidelet, T.; Guezenec, S.; Perez, M.; Segond, D.; Sicard, D. Interactions between Kazachstania Humilis Yeast Species and Lactic Acid Bacteria in Sourdough. Microorganisms 2020, 8, 240. [Google Scholar] [CrossRef] [Green Version]
  31. Chavan, R.S.; Chavan, S.R. Sourdough Technology-A Traditional Way for Wholesome Foods: A Review. Compr. Rev. Food Sci. Food Saf. 2011, 10, 169–182. [Google Scholar] [CrossRef]
  32. Warburton, A.; Silcock, P.; Eyres, G.T. Impact of Sourdough Culture on the Volatile Compounds in Wholemeal Sourdough Bread. Food Res. Int. 2022, 161, 111885. [Google Scholar] [CrossRef] [PubMed]
  33. Gänzle, M.G.; Zheng, J. Lifestyles of Sourdough Lactobacilli–Do They Matter for Microbial Ecology and Bread Quality? Int. J. Food Microbiol. 2019, 302, 15–23. [Google Scholar] [CrossRef]
  34. Van Kerrebroeck, S.; Maes, D.; De Vuyst, L. Sourdoughs as a Function of Their Species Diversity and Process Conditions, a Meta-Analysis. Trends Food Sci. Technol. 2017, 68, 152–159. [Google Scholar] [CrossRef]
  35. Siepmann, F.B.; Ripari, V.; Waszczynskyj, N.; Spier, M.R. Overview of Sourdough Technology: From Production to Marketing. Food Bioprocess Technol. 2018, 11, 242–270. [Google Scholar] [CrossRef]
  36. Limbad, M.; Gutierrez Maddox, N.; Hamid, N.; Kantono, K. Sensory and Physicochemical Characterization of Sourdough Bread Prepared with a Coconut Water Kefir Starter. Foods 2020, 9, 1165. [Google Scholar] [CrossRef] [PubMed]
  37. Mietton, L.; Samson, M.-F.; Marlin, T.; Godet, T.; Nolleau, V.; Guezenec, S.; Segond, D.; Nidelet, T.; Desclaux, D.; Sicard, D. Impact of Leavening Agent and Wheat Variety on Bread Organoleptic and Nutritional Quality. Microorganisms 2022, 10, 1416. [Google Scholar] [CrossRef] [PubMed]
  38. De Vuyst, L.; Van Kerrebroeck, S.; Leroy, F. Microbial Ecology and Process Technology of Sourdough Fermentation. In Advances in Applied Microbiology; Elsevier: Amsterdam, The Netherlands, 2017; Volume 100, pp. 49–160. ISBN 978-0-12-812048-4. [Google Scholar]
  39. Coelho, M.C.; Malcata, F.X.; Silva, C.C.G. Lactic Acid Bacteria in Raw-Milk Cheeses: From Starter Cultures to Probiotic Functions. Foods 2022, 11, 2276. [Google Scholar] [CrossRef]
  40. Bintsis, T. Lactic acid bacteria as starter cultures: An update in their metabolism and genetics. AIMS Microbiol. 2018, 4, 665–684. [Google Scholar] [CrossRef]
  41. Castillo Martinez, F.A.; Balciunas, E.M.; Salgado, J.M.; Domínguez González, J.M.; Converti, A.; de Souza Oliveira, R.P. Lactic Acid Properties, Applications and Production: A Review. Trends Food Sci. Technol. 2013, 30, 70–83. [Google Scholar] [CrossRef]
  42. Gaglio, R.; Alfonzo, A.; Polizzotto, N.; Corona, O.; Francesca, N.; Russo, G.; Moschetti, G.; Settanni, L. Performances of Different Metabolic Lactobacillus Groups during the Fermentation of Pizza Doughs Processed from Semolina. Fermentation 2018, 4, 61. [Google Scholar] [CrossRef] [Green Version]
  43. Di Cagno, R.; De Angelis, M.; Gallo, G.; Settanni, L.; Berloco, M.G.; Siragusa, S.; Parente, E.; Corsetti, A.; Gobbetti, M. Genotypic and Phenotypic Diversity of Lactobacillus rossiae Strains Isolated from Sourdough. J. Appl. Microbiol. 2007, 103, 821–835. [Google Scholar] [CrossRef] [PubMed]
  44. Gänzle, M.; Gobbetti, M. Physiology and Biochemistry of Lactic Acid Bacteria. In Handbook on Sourdough Biotechnology; Springer: New York, NY, USA, 2013; pp. 183–216. ISBN 978-1-4614-5424-3. [Google Scholar]
  45. Yang, H.; Liu, T.; Zhang, G.; He, G. Intraspecific Diversity and Fermentative Properties of Saccharomyces cerevisiae from Chinese Traditional Sourdough. LWT 2020, 124, 109195. [Google Scholar] [CrossRef]
  46. Kaseleht, K.; Paalme, T.; Mihhalevski, A.; Sarand, I. Analysis of Volatile Compounds Produced by Different Species of Lactobacilli in Rye Sourdough Using Multiple Headspace Extraction: Volatiles Produced by LAB in Sourdough. Int. J. Food Sci. Technol. 2011, 46, 1940–1946. [Google Scholar] [CrossRef]
  47. Nataraj, B.H.; Ali, S.A.; Behare, P.V.; Yadav, H. Postbiotics-Parabiotics: The New Horizons in Microbial Biotherapy and Functional Foods. Microb. Cell Factories 2020, 19, 168. [Google Scholar] [CrossRef]
  48. Salminen, S.; Collado, M.C.; Endo, A.; Hill, C.; Lebeer, S.; Quigley, E.M.M.; Sanders, M.E.; Shamir, R.; Swann, J.R.; Szajewska, H.; et al. Publisher Correction: The International Scientific Association of Probiotics and Prebiotics (ISAPP) Consensus Statement on the Definition and Scope of Postbiotics. Nat. Rev. Gastroenterol. Hepatol. 2021, 18, 671. [Google Scholar] [CrossRef]
  49. El Sheikha, A.F.; Hu, D.-M.M. Molecular Techniques Reveal More Secrets of Fermented Foods. Crit. Rev. Food Sci. Nutr. 2020, 60, 11–32. [Google Scholar] [CrossRef] [PubMed]
  50. Weckx, S.; Van Kerrebroeck, S.; De Vuyst, L. Omics Approaches to Understand Sourdough Fermentation Processes. Int. J. Food Microbiol. 2019, 302, 90–102. [Google Scholar] [CrossRef]
  51. Damián, M.R.; Cortes-Perez, N.G.; Quintana, E.T.; Ortiz-Moreno, A.; Garfias Noguez, C.; Cruceño-Casarrubias, C.E.; Sánchez Pardo, M.E.; Bermúdez-Humarán, L.G. Functional Foods, Nutraceuticals and Probiotics: A Focus on Human Health. Microorganisms 2022, 10, 1065. [Google Scholar] [CrossRef]
  52. Minervini, F.; Celano, G.; Lattanzi, A.; Tedone, L.; de Mastro, G.; Gobbetti, M.; de Angelis, M. Lactic Acid Bacteria in Durum Wheat Flour Are Endophytic Components of the Plant during Its Entire Life Cycle. Appl. Environ. Microbiol. 2015, 81, 6736–6748. [Google Scholar] [CrossRef] [Green Version]
  53. Landis, E.A.; Oliverio, A.M.; McKenney, E.A.; Nichols, L.M.; Kfoury, N.; Biango-Daniels, M.; Shell, L.K.; Madden, A.A.; Shapiro, L.; Sakunala, S.; et al. The Diversity and Function of Sourdough Starter Microbiomes. eLife 2021, 10, e61644. [Google Scholar] [CrossRef]
  54. Menezes, L.A.A.; Sardaro, M.L.S.; Duarte, R.T.D.; Mazzon, R.R.; Neviani, E.; Gatti, M.; De Dea Lindner, J. Sourdough Bacterial Dynamics Revealed by Metagenomic Analysis in Brazil. Food Microbiol. 2020, 85, 103302. [Google Scholar] [CrossRef] [PubMed]
  55. Groenewald, W.H.; Reenen, C.A.V.; Todorov, S.D.; Toit, M.D.; Witthuhn, R.C.; Holzapfel, W.H.; Dicks, L.M.T. Identification of Lactic Acid Bacteria from Vinegar Flies Based on Phenotypic and Genotypic Characteristics. Am. J. Enol. Vitic. 2006, 57, 519–525. [Google Scholar] [CrossRef]
  56. Minervini, F.; Dinardo, F.R.; De Angelis, M.; Gobbetti, M. Tap Water Is One of the Drivers That Establish and Assembly the Lactic Acid Bacterium Biota during Sourdough Preparation. Sci. Rep. 2019, 9, 570. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  57. Dinardo, F.R.; Minervini, F.; De Angelis, M.; Gobbetti, M.; Gänzle, M.G. Dynamics of Enterobacteriaceae and Lactobacilli in Model Sourdoughs Are Driven by PH and Concentrations of Sucrose and Ferulic Acid. LWT 2019, 114, 108394. [Google Scholar] [CrossRef]
  58. De Angelis, M.; Minervini, F.; Siragusa, S.; Rizzello, C.G.; Gobbetti, M. Wholemeal Wheat Flours Drive the Microbiome and Functional Features of Wheat Sourdoughs. Int. J. Food Microbiol. 2019, 302, 35–46. [Google Scholar] [CrossRef]
  59. Van Kerrebroeck, S.; Bastos, F.C.C.; Harth, H.; De Vuyst, L. A Low PH Does Not Determine the Community Dynamics of Spontaneously Developed Backslopped Liquid Wheat Sourdoughs but Does Influence Their Metabolite Kinetics. Int. J. Food Microbiol. 2016, 239, 54–64. [Google Scholar] [CrossRef]
  60. Sakandar, H.A.; Usman, K.; Imran, M. Isolation and Characterization of Gluten-Degrading Enterococcus mundtii and Wickerhamomyces anomalus, Potential Probiotic Strains from Indigenously Fermented Sourdough (Khamir). LWT Food Sci. Technol. 2018, 91, 271–277. [Google Scholar] [CrossRef]
  61. Fraberger, V.; Unger, C.; Kummer, C.; Domig, K.J. Insights into Microbial Diversity of Traditional Austrian Sourdough. LWT 2020, 127, 109358. [Google Scholar] [CrossRef]
  62. Fekri, A.; Torbati, M.; Yari Khosrowshahi, A.; Bagherpour Shamloo, H.; Azadmard-Damirchi, S. Functional Effects of Phytate-Degrading, Probiotic Lactic Acid Bacteria and Yeast Strains Isolated from Iranian Traditional Sourdough on the Technological and Nutritional Properties of Whole Wheat Bread. Food Chem. 2020, 306, 125620. [Google Scholar] [CrossRef]
  63. Arici, M.; Ozulku, G.; Yildirim, R.M.; Sagdic, O.; Durak, M.Z. Biodiversity and Technological Properties of Yeasts from Turkish Sourdough. Food Sci. Biotechnol. 2018, 27, 499–508. [Google Scholar] [CrossRef]
  64. Vermeulen, N.; Gänzle, M.G.; Vogel, R.F. Glutamine Deamidation by Cereal-Associated Lactic Acid Bacteria. J. Appl. Microbiol. 2007, 103, 1197–1205. [Google Scholar] [CrossRef]
  65. Zhang, C.; Brandt, M.J.; Schwab, C.; Gänzle, M.G. Propionic Acid Production by Cofermentation of Lactobacillus buchneri and Lactobacillus diolivorans in Sourdough. Food Microbiol. 2010, 27, 390–395. [Google Scholar] [CrossRef] [PubMed]
  66. Sandra, G.; Schwab, C.; Bello, F.D.; Coffey, A.; Gänzle, M.; Arendt, E. Comparison of the Impact of Dextran and Reuteran on the Quality of Wheat Sourdough Bread. J. Cereal Sci. 2012, 56, 531–537. [Google Scholar] [CrossRef]
  67. Black, B.A.; Zannini, E.; Curtis, J.M.; Gänzle, M.G. Antifungal Hydroxy Fatty Acids Produced during Sourdough Fermentation: Microbial and Enzymatic Pathways, and Antifungal Activity in Bread. Appl. Environ. Microbiol. 2013, 79, 1866–1873. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  68. Denkova, R.; Ilieva, S.; Denkova, Z.; Georgieva, L.; Yordanova, M.; Nikolova, D.; Evstatieva, Y. Production of Wheat Bread without Preservatives Using Sourdough Starters. Biotechnol. Biotechnol. Equip. 2014, 28, 889–898. [Google Scholar] [CrossRef]
  69. Stefańska, I.; Piasecka-Jóźwiak, K.; Kotyrba, D.; Kolenda, M.; Stecka, K.M. Selection of Lactic Acid Bacteria Strains for the Hydrolysis of Allergenic Proteins of Wheat Flour. J. Sci. Food Agric. 2016, 96, 3897–3905. [Google Scholar] [CrossRef]
  70. Mamhoud, A.; Nionelli, L.; Bouzaine, T.; Hamdi, M.; Gobbetti, M.; Rizzello, C.G. Selection of Lactic Acid Bacteria Isolated from Tunisian Cereals and Exploitation of the Use as Starters for Sourdough Fermentation. Int. J. Food Microbiol. 2016, 225, 9–19. [Google Scholar] [CrossRef]
  71. Corona, O.; Alfonzo, A.; Ventimiglia, G.; Nasca, A.; Francesca, N.; Martorana, A.; Moschetti, G.; Settanni, L. Industrial Application of Selected Lactic Acid Bacteria Isolated from Local Semolinas for Typical Sourdough Bread Production. Food Microbiol. 2016, 59, 43–56. [Google Scholar] [CrossRef] [Green Version]
  72. Aslankoohi, E.; Herrera-Malaver, B.; Rezaei, M.N.; Steensels, J.; Courtin, C.M.; Verstrepen, K.J. Non-Conventional Yeast Strains Increase the Aroma Complexity of Bread. PLoS ONE 2016, 11, e0165126. [Google Scholar] [CrossRef] [Green Version]
  73. Edelmann, M.; Chamlagain, B.; Santin, M.; Kariluoto, S.; Piironen, V. Stability of Added and in Situ-Produced Vitamin B12 in Breadmaking. Food Chem. 2016, 204, 21–28. [Google Scholar] [CrossRef] [Green Version]
  74. Zhao, C.J.; Gänzle, M.G. Synthesis of Taste-Active γ-Glutamyl Dipeptides during Sourdough Fermentation by Lactobacillus Reuteri. J. Agric. Food Chem. 2016, 64, 7561–7568. [Google Scholar] [CrossRef] [PubMed]
  75. Scarnato, L.; Montanari, C.; Serrazanetti, D.I.; Aloisi, I.; Balestra, F.; Del Duca, S.; Lanciotti, R. New Bread Formulation with Improved Rheological Properties and Longer Shelf-Life by the Combined Use of Transglutaminase and Sourdough. LWT Food Sci. Technol. 2017, 81, 101–110. [Google Scholar] [CrossRef]
  76. Elsanhoty, R.M.; Ghonamy, A.G.; El-Adly, N.A.; Fawzy Ramadan, M. Impact of Lactic Acid Bacteria and Bifidobacterium on the Survival of Bacillus subtilus During Fermentation of Wheat Sourdough. J. Food Process. Preserv. 2017, 41, e13086. [Google Scholar] [CrossRef]
  77. Struyf, N.; Laurent, J.; Verspreet, J.; Verstrepen, K.J.; Courtin, C.M. Saccharomyces cerevisiae and Kluyveromyces marxianus Cocultures Allow Reduction of Fermentable Oligo-, Di-, and Monosaccharides and Polyols Levels in Whole Wheat Bread. J. Agric. Food Chem. 2017, 65, 8704–8713. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  78. Struyf, N.; Vandewiele, H.; Herrera-Malaver, B.; Verspreet, J.; Verstrepen, K.J.; Courtin, C.M. Kluyveromyces marxianus Yeast Enables the Production of Low FODMAP Whole Wheat Breads. Food Microbiol. 2018, 76, 134–145. [Google Scholar] [CrossRef] [PubMed]
  79. Karaman, K.; Sagdic, O.; Durak, M.Z. Use of Phytase Active Yeasts and Lactic Acid Bacteria Isolated from Sourdough in the Production of Whole Wheat Bread. LWT Food Sci. Technol. 2018, 91, 557–567. [Google Scholar] [CrossRef]
  80. Van Kerrebroeck, S.; Comasio, A.; Harth, H.; De Vuyst, L. Impact of Starter Culture, Ingredients, and Flour Type on Sourdough Bread Volatiles as Monitored by Selected Ion Flow Tube-Mass Spectrometry. Food Res. Int. 2018, 106, 254–262. [Google Scholar] [CrossRef]
  81. Coda, R.; Xu, Y.; Moreno, D.S.; Mojzita, D.; Nionelli, L.; Rizzello, C.G.; Katina, K. Performance of Leuconostoc citreum FDR241 during Wheat Flour Sourdough Type I Propagation and Transcriptional Analysis of Exopolysaccharides Biosynthesis Genes. Food Microbiol. 2018, 76, 164–172. [Google Scholar] [CrossRef]
  82. Rizzello, C.G.; Coda, R.; Wang, Y.; Verni, M.; Kajala, I.; Katina, K.; Laitila, A. Characterization of Indigenous Pediococcus pentosaceus, Leuconostoc Kimchii, Weissella Cibaria and Weissella Confusa for Faba Bean Bioprocessing. Int. J. Food Microbiol. 2019, 302, 24–34. [Google Scholar] [CrossRef]
  83. Belz, M.C.E.; Axel, C.; Arendt, E.K.; Lynch, K.M.; Brosnan, B.; Sheehan, E.M.; Coffey, A.; Zannini, E. Improvement of Taste and Shelf Life of Yeasted Low-Salt Bread Containing Functional Sourdoughs Using Lactobacillus amylovorus DSM 19280 and Weisella Cibaria MG1. Int. J. Food Microbiol. 2019, 302, 69–79. [Google Scholar] [CrossRef]
  84. Siepmann, F.B.; Sousa de Almeida, B.; Waszczynskyj, N.; Spier, M.R. Influence of Temperature and of Starter Culture on Biochemical Characteristics and the Aromatic Compounds Evolution on Type II Sourdough and Wheat Bread. Lwt 2019, 108, 199–206. [Google Scholar] [CrossRef]
  85. Sadeghi, A.; Ebrahimi, M.; Mortazavi, S.A.; Abedfar, A. Application of the Selected Antifungal LAB Isolate as a Protective Starter Culture in Pan Whole-Wheat Sourdough Bread. Food Control 2019, 95, 297–307. [Google Scholar] [CrossRef]
  86. Antognoni, F.; Mandrioli, R.; Potente, G.; Taneyo Saa, D.L.; Gianotti, A. Changes in Carotenoids, Phenolic Acids and Antioxidant Capacity in Bread Wheat Doughs Fermented with Different Lactic Acid Bacteria Strains. Food Chem. 2019, 292, 211–216. [Google Scholar] [CrossRef] [PubMed]
  87. Yu, Y.; Wang, L.; Qian, H.; Zhang, H.; Li, Y.; Wu, G.; Qi, X.; Xu, M.; Rao, Z. Effect of Selected Strains on Physical and Organoleptic Properties of Breads. Food Chem. 2019, 276, 547–553. [Google Scholar] [CrossRef] [PubMed]
  88. Kook, S.-Y.; Lee, Y.; Jeong, E.-C.; Kim, S. Immunomodulatory Effects of Exopolysaccharides Produced by Bacillus licheniformis and Leuconostoc mesenteroides Isolated from Korean Kimchi. J. Funct. Foods 2019, 54, 211–219. [Google Scholar] [CrossRef]
  89. Mantzourani, I.; Plessas, S.; Odatzidou, M.; Alexopoulos, A.; Galanis, A.; Bezirtzoglou, E.; Bekatorou, A. Effect of a Novel Lactobacillus paracasei Starter on Sourdough Bread Quality. Food Chem. 2019, 271, 259–265. [Google Scholar] [CrossRef]
  90. Sakandar, H.A.; Kubow, S.; Azadi, B.; Faryal, R.; Ali, B.; Ghazanfar, S.; Quraishi, U.M.; Imran, M. Wheat Fermentation With Enterococcus mundtii QAUSD01 and Wickerhamomyces anomalus QAUWA03 Consortia Induces Concurrent Gliadin and Phytic Acid Degradation and Inhibits Gliadin Toxicity in Caco-2 Monolayers. Front. Microbiol. 2019, 9, 3312. [Google Scholar] [CrossRef]
  91. Ouiddir, M.; Bettache, G.; Leyva Salas, M.; Pawtowski, A.; Donot, C.; Brahimi, S.; Mabrouk, K.; Coton, E.; Mounier, J. Selection of Algerian Lactic Acid Bacteria for Use as Antifungal Bioprotective Cultures and Application in Dairy and Bakery Products. Food Microbiol. 2019, 82, 160–170. [Google Scholar] [CrossRef]
  92. Li, D.; Zhao, Y.; Fei, T.; Wang, Y.; Lee, B.H.; Shim, J.H.; Xu, B.; Li, Z.; Li, X. Effects of Streptococcus thermophilus GtfB Enzyme on Dough Rheology, Bread Quality and Starch Digestibility. Food Hydrocoll. 2019, 96, 134–139. [Google Scholar] [CrossRef]
  93. Plessas, S.; Mantzourani, I.; Bekatorou, A. Evaluation of Pediococcus pentosaceus SP2 as Starter Culture on Sourdough Bread Making. Foods 2020, 9, 77. [Google Scholar] [CrossRef] [Green Version]
  94. Li, Z.; Siepmann, F.B.; Rojas Tovar, L.E.; Chen, X.; Gänzle, M.G. Effect of Copy Number of the SpoVA2mob Operon, Sourdough and Reutericyclin on Ropy Bread Spoilage Caused by Bacillus spp. Food Microbiol. 2020, 91, 103507. [Google Scholar] [CrossRef] [PubMed]
  95. Li, Z.; Zhou, M.; Cui, M.; Wang, Y.; Li, H. Improvement of Whole Wheat Dough Fermentation for Steamed Bread Making Using Selected Phytate-Degrading Wickerhamomyces anomalus P4. J. Cereal Sci. 2021, 100, 103261. [Google Scholar] [CrossRef]
  96. Bockwoldt, J.A.; Fellermeier, J.; Steffens, E.; Vogel, R.F.; Ehrmann, M.A. β-Glucan Production by Levilactobacillus brevis and Pediococcus claussenii for In Situ Enriched Rye and Wheat Sourdough Breads. Foods 2021, 10, 547. [Google Scholar] [CrossRef]
  97. Buksa, K.; Kowalczyk, M.; Boreczek, J. Extraction, Purification and Characterisation of Exopolysaccharides Produced by Newly Isolated Lactic Acid Bacteria Strains and the Examination of Their Influence on Resistant Starch Formation. Food Chem. 2021, 362, 130221. [Google Scholar] [CrossRef]
  98. Blaiotta, G.; Romano, R.; Trifuoggi, M.; Aponte, M.; Miro, A. Development of a Wet-Granulated Sourdough Multiple Starter for Direct Use. Foods 2022, 11, 1278. [Google Scholar] [CrossRef] [PubMed]
  99. Muhammad, Z.; Ramzan, R.; Abdelazez, A.; Amjad, A.; Afzaal, M.; Zhang, S.; Pan, S. Assessment of the Antimicrobial Potentiality and Functionality of Lactobacillus plantarum Strains Isolated from the Conventional Inner Mongolian Fermented Cheese Against Foodborne Pathogens. Pathogens 2019, 8, 71. [Google Scholar] [CrossRef] [Green Version]
  100. Koistinen, V.M.; Mattila, O.; Katina, K.; Poutanen, K.; Aura, A.M.; Hanhineva, K. Metabolic Profiling of Sourdough Fermented Wheat and Rye Bread. Sci. Rep. 2018, 8, 5684. [Google Scholar] [CrossRef] [Green Version]
  101. Angelis, M.D.; Rizzello, C.G.; Fasano, A.; Clemente, M.G.; Simone, C.D.; Silano, M.; Vincenzi, M.D.; Losito, I.; Gobbetti, M. VSL#3 Probiotic Preparation Has the Capacity to Hydrolyze Gliadin Polypeptides Responsible for Celiac Sprue Probiotics and Gluten Intolerance. Biochim. Biophys. Acta BBA Mol. Basis Dis. 2006, 1762, 80–93. [Google Scholar] [CrossRef] [Green Version]
  102. Rizzello, C.G.; Cassone, A.; Cagno, R.D.; Gobbetti, A.M. Synthesis of Angiotensin I-Converting Enzyme (ACE)-Inhibitory Peptides and γ-Aminobutyric Acid (GABA) during Sourdough Fermentation by Selected Lactic Acid Bacteria. J. Agric. Food Chem. 2008, 56, 6936–6943. [Google Scholar] [CrossRef]
  103. De Angelis, M.; Damiano, N.; Rizzello, C.G.; Cassone, A.; Di Cagno, R.; Gobbetti, M. Sourdough Fermentation as a Tool for the Manufacture of Low-Glycemic Index White Wheat Bread Enriched in Dietary Fibre. Eur. Food Res. Technol. 2009, 229, 593–601. [Google Scholar] [CrossRef]
  104. Wang, Y.; Gänzle, M.G.; Schwab, C. Exopolysaccharide Synthesized by Lactobacillus Reuteri Decreases the Ability of Enterotoxigenic Escherichia Coli To Bind to Porcine Erythrocytes. Appl. Environ. Microbiol. 2010, 76, 4863–4866. [Google Scholar] [CrossRef] [PubMed]
  105. Novotni, D.; Curić, D.; Bituh, M.; Colić Barić, I.; Skevin, D.; Cukelj, N. Glycemic Index and Phenolics of Partially-Baked Frozen Bread with Sourdough. Int. J. Food Sci. Nutr. 2011, 62, 26–33. [Google Scholar] [CrossRef] [PubMed]
  106. Rizzello, C.G.; Nionelli, L.; Coda, R.; Gobbetti, M. Synthesis of the Cancer Preventive Peptide Lunasin by Lactic Acid Bacteria during Sourdough Fermentation. Nutr. Cancer 2012, 64, 111–120. [Google Scholar] [CrossRef]
  107. Sanz-Penella, J.M.; Laparra, J.M.; Sanz, Y.; Haros, M. Assessment of Iron Bioavailability in Whole Wheat Bread by Addition of Phytase-Producing Bifidobacteria. J. Agric. Food Chem. 2012, 60, 3190–3195. [Google Scholar] [CrossRef] [PubMed]
  108. Galle, S.; Schwab, C.; Dal Bello, F.; Coffey, A.; Gänzle, M.G.; Arendt, E.K. Influence of In-Situ Synthesized Exopolysaccharides on the Quality of Gluten-Free Sorghum Sourdough Bread. Int. J. Food Microbiol. 2012, 155, 105–112. [Google Scholar] [CrossRef] [PubMed]
  109. Sümengen, M.; Dinçer, S.; Kaya, A. Phytase Production from Lactobacillus brevis. Turk. J. Biol. 2012, 36, 533–541. [Google Scholar] [CrossRef]
  110. Cizeikiene, D.; Juodeikiene, G.; Paskevicius, A.; Bartkiene, E. Antimicrobial Activity of Lactic Acid Bacteria against Pathogenic and Spoilage Microorganism Isolated from Food and Their Control in Wheat Bread. Food Control 2013, 31, 539–545. [Google Scholar] [CrossRef]
  111. Amari, M.; Arango, L.F.G.; Gabriel, V.; Robert, H.; Morel, S.; Moulis, C.; Gabriel, B.; Remaud-Siméon, M.; Fontagné-Faucher, C. Characterization of a Novel Dextransucrase from Weissella confusa Isolated from Sourdough. Appl. Microbiol. Biotechnol. 2013, 97, 5413–5422. [Google Scholar] [CrossRef]
  112. De Angelis, M.; Bottacini, F.; Fosso, B.; Kelleher, P.; Calasso, M.; Di Cagno, R.; Ventura, M.; Picardi, E.; van Sinderen, D.; Gobbetti, M. Lactobacillus Rossiae, a Vitamin B12 Producer, Represents a Metabolically Versatile Species within the Genus Lactobacillus. PLoS ONE 2014, 9, e107232. [Google Scholar] [CrossRef] [Green Version]
  113. Omoba, O.S.; Isah, L.R. Influence of Sourdough Fermentation on Amino Acids Composition, Phenolic Profile, and Antioxidant Properties of Sorghum Biscuits. Prev. Nutr. Food Sci. 2018, 23, 220–227. [Google Scholar] [CrossRef]
  114. Fraberger, V.; Ammer, C.; Domig, K.J. Functional Properties and Sustainability Improvement of Sourdough Bread by Lactic Acid Bacteria. Microorganisms 2020, 8, 1895. [Google Scholar] [CrossRef] [PubMed]
  115. Liu, T.; Li, Y.; Zhao, M.; Mo, Q.; Feng, F. Weight-Reducing Effect of Lactobacillus plantarum ZJUFT17 Isolated from Sourdough Ecosystem. Nutrients 2020, 12, 977. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  116. Zhong, H.; Wang, J.; Abdullah; Hafeez, M.A.; Guan, R.; Feng, F. Lactobacillus plantarum ZJUFB2 Prevents High Fat Diet-Induced Insulin Resistance in Association With Modulation of the Gut Microbiota. Front. Nutr. 2021, 8, 754222. [Google Scholar] [CrossRef] [PubMed]
  117. Schlörmann, W.; Bockwoldt, J.A.; Hübner, S.M.; Wittwer, E.; Reiners, S.; Lorkowski, S.; Dawczynski, C.; Ehrmann, M.A.; Glei, M. Use of the β-Glucan-Producing Lactic Acid Bacteria Strains Levilactobacillus brevis and Pediococcus claussenii for Sourdough Fermentation—Chemical Characterization and Chemopreventive Potential of In Situ-Enriched Wheat and Rye Sourdoughs and Breads. Nutrients 2022, 14, 1510. [Google Scholar] [CrossRef] [PubMed]
  118. Verni, M.; Vekka, A.; Immonen, M.; Katina, K.; Rizzello, C.G.; Coda, R. Biosynthesis of Γ-aminobutyric Acid by Lactic Acid Bacteria in Surplus Bread and Its Use in Bread Making. J. Appl. Microbiol. 2022, 133, 76–90. [Google Scholar] [CrossRef] [PubMed]
  119. Yilmaz, M.T.; Ispirli, H.; Taylan, O.; Alamoudi, M.; Dertli, E. Bioactive and Technological Properties of an α-D-Glucan Synthesized by Weissella cibaria PDER21. Carbohydr. Polym. 2022, 285, 119227. [Google Scholar] [CrossRef]
  120. Saa, D.T.; Di Silvestro, R.; Nissen, L.; Dinelli, G.; Gianotti, A. Effect of Sourdough Fermentation and Baking Process Severity on Bioactive Fiber Compounds in Immature and Ripe Wheat Flour Bread. Leb. Wiss Technol 2018, 89, 322–328. [Google Scholar] [CrossRef]
  121. Debonne, E.; Van Bockstaele, F.; Van Driessche, M.; De Leyn, I.; Eeckhout, M.; Devlieghere, F. Impact of Par-Baking and Packaging on the Microbial Quality of Par-Baked Wheat and Sourdough Bread. Food Control 2018, 91, 12–19. [Google Scholar] [CrossRef]
  122. Zhang, L.; Taal, M.A.; Boom, R.M.; Chen, X.D.; Schutyser, M.A.I. Effect of Baking Conditions and Storage on the Viability of Lactobacillus plantarum Supplemented to Bread. Leb. Wiss Technol 2018, 87, 318–325. [Google Scholar] [CrossRef]
  123. Salminen, S.; Collado, M.C.; Endo, A.; Hill, C.; Lebeer, S.; Quigley, E.M.M.; Sanders, M.E.; Shamir, R.; Swann, J.R.; Szajewska, H.; et al. The International Scientific Association of Probiotics and Prebiotics (ISAPP) Consensus Statement on the Definition and Scope of Postbiotics. Nat. Rev. Gastroenterol. Hepatol. 2021, 18, 649–667. [Google Scholar] [CrossRef]
  124. Kazakos, S.; Mantzourani, I.; Plessas, S. Quality Characteristics of Novel Sourdough Breads Made with Functional Lacticaseibacillus paracasei SP5 and Prebiotic Food Matrices. Foods 2022, 11, 3226. [Google Scholar] [CrossRef]
  125. Koistinen, V.M.; Hedberg, M.; Shi, L.; Johansson, A.; Savolainen, O.; Lehtonen, M.; Aura, A.; Hanhineva, K.; Landberg, R. Metabolite Pattern Derived from Lactiplantibacillus plantarum—Fermented Rye Foods and In Vitro Gut Fermentation Synergistically Inhibits Bacterial Growth. Mol. Nutr. Food Res. 2022, 66, 2101096. [Google Scholar] [CrossRef] [PubMed]
  126. Lau, S.W.; Chong, A.Q.; Chin, N.L.; Talib, R.A.; Basha, R.K. Sourdough Microbiome Comparison and Benefits. Microorganisms 2021, 9, 1355. [Google Scholar] [CrossRef]
  127. Reis, J.A.; Paula, A.T.; Casarotti, S.N.; Penna, A.L.B. Lactic Acid Bacteria Antimicrobial Compounds: Characteristics and Applications. Food Eng. Rev. 2012, 4, 124–140. [Google Scholar] [CrossRef]
  128. Chen, J.; Pang, H.; Wang, L.; Ma, C.; Wu, G.; Liu, Y.; Guan, Y.; Zhang, M.; Qin, G.; Tan, Z. Bacteriocin-Producing Lactic Acid Bacteria Strains with Antimicrobial Activity Screened from Bamei Pig Feces. Foods 2022, 11, 709. [Google Scholar] [CrossRef]
  129. De Simone, N.; Rocchetti, M.T.; la Gatta, B.; Spano, G.; Drider, D.; Capozzi, V.; Russo, P.; Fiocco, D. Antimicrobial Properties, Functional Characterisation and Application of Fructobacillus fructosus and Lactiplantibacillus plantarum Isolated from Artisanal Honey. Probiotics Antimicrob. Proteins 2022. [Google Scholar] [CrossRef]
  130. Lavermicocca, P.; Valerio, F.; Evidente, A.; Lazzaroni, S.; Corsetti, A.; Gobbetti, M. Purification and Characterization of Novel Antifungal Compounds from the Sourdough Lactobacillus plantarum Strain 21B. Appl. Environ. Microbiol. 2000, 66, 4084–4090. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  131. Varsha, K.K.; Priya, S.; Devendra, L.; Nampoothiri, K.M. Control of Spoilage Fungi by Protective Lactic Acid Bacteria Displaying Probiotic Properties. Appl. Biochem. Biotechnol. 2014, 172, 3402–3413. [Google Scholar] [CrossRef]
  132. Valerio, F.; Conte, A.; Di Biase, M.; Lattanzio, V.M.T.; Lonigro, S.L.; Padalino, L.; Pontonio, E.; Lavermicocca, P. Formulation of Yeast-Leavened Bread with Reduced Salt Content by Using a Lactobacillus plantarum Fermentation Product. Food Chem. 2017, 221, 582–589. [Google Scholar] [CrossRef]
  133. Hernández-Figueroa, R.H.; Mani-López, E.; López-Malo, A. Antifungal Capacity of Poolish-Type Sourdough Supplemented with Lactiplantibacillus plantarum and Its Aqueous Extracts in Vitro and Bread. Antibiotics 2022, 11, 1813. [Google Scholar] [CrossRef]
  134. Da Ros, A.; Polo, A.; Rizzello, C.G.; Acin-Albiac, M.; Montemurro, M.; Di Cagno, R.; Gobbetti, M. Feeding with Sustainably Sourdough Bread Has the Potential to Promote the Healthy Microbiota Metabolism at the Colon Level. Microbiol. Spectr. 2021, 9, 3. [Google Scholar] [CrossRef] [PubMed]
  135. Siragusa, S.; De Angelis, M.; Di Cagno, R.; Rizzello, C.G.; Coda, R.; Gobbetti, M. Synthesis of γ-Aminobutyric Acid by Lactic Acid Bacteria Isolated from a Variety of Italian Cheeses. Appl. Environ. Microbiol. 2007, 73, 7283–7290. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  136. Sarasa, S.B.; Mahendran, R.; Muthusamy, G.; Thankappan, B.; Selta, D.R.F.; Angayarkanni, J. A Brief Review on the Non-Protein Amino Acid, Gamma-Amino Butyric Acid (GABA): Its Production and Role in Microbes. Curr. Microbiol. 2020, 77, 534–544. [Google Scholar] [CrossRef] [PubMed]
  137. Khanlari, Z.; Moayedi, A.; Ebrahimi, P.; Khomeiri, M.; Sadeghi, A. Enhancement of Γ-aminobutyric Acid (GABA) Content in Fermented Milk by Using Enterococcus faecium and Weissella confusa Isolated from Sourdough. J. Food Process. Preserv. 2021, 45, e15869. [Google Scholar] [CrossRef]
  138. Arena, M.P.; Capozzi, V.; Russo, P.; Drider, D.; Spano, G.; Fiocco, D. Immunobiosis and Probiosis: Antimicrobial Activity of Lactic Acid Bacteria with a Focus on Their Antiviral and Antifungal Properties. Appl. Microbiol. Biotechnol. 2018, 102, 9949–9958. [Google Scholar] [CrossRef]
  139. Boakye, P.G.; Kougblenou, I.; Murai, T.; Okyere, A.Y.; Anderson, J.; Bajgain, P.; Philipp, B.; LaPlante, B.; Schlecht, S.; Vogel, C.; et al. Impact of Sourdough Fermentation on FODMAPs and Amylase-Trypsin Inhibitor Levels in Wheat Dough. J. Cereal Sci. 2022, 108, 103574. [Google Scholar] [CrossRef]
  140. Rashmi, B.S.; Gayathri, D.; Vasudha, M.; Prashantkumar, C.S.; Swamy, C.T.; Sunil, K.S.; Somaraja, P.K.; Prakash, P. Gluten Hydrolyzing Activity of Bacillus spp Isolated from Sourdough. Microb. Cell Factories 2020, 19, 130. [Google Scholar] [CrossRef]
  141. Gobbetti, M.; Giuseppe Rizzello, C.; Di Cagno, R.; De Angelis, M. Sourdough Lactobacilli and Celiac Disease. Food Microbiol. 2007, 24, 187–196. [Google Scholar] [CrossRef]
  142. Luti, S.; Mazzoli, L.; Ramazzotti, M.; Galli, V.; Venturi, M.; Marino, G.; Lehmann, M.; Guerrini, S.; Granchi, L.; Paoli, P.; et al. Antioxidant and Anti-Inflammatory Properties of Sourdoughs Containing Selected Lactobacilli Strains Are Retained in Breads. Food Chem. 2020, 322, 126710. [Google Scholar] [CrossRef]
  143. Colosimo, R.; Gabriele, M.; Cifelli, M.; Longo, V.; Domenici, V.; Pucci, L. The Effect of Sourdough Fermentation on Triticum Dicoccum from Garfagnana: 1H NMR Characterization and Analysis of the Antioxidant Activity. Food Chem. 2020, 305, 125510. [Google Scholar] [CrossRef]
  144. Nachtigall, C.; Rohm, H.; Jaros, D. Degradation of Exopolysaccharides from Lactic Acid Bacteria by Thermal, Chemical, Enzymatic and Ultrasound Stresses. Foods 2021, 10, 396. [Google Scholar] [CrossRef] [PubMed]
  145. Gobbetti, M.; De Angelis, M.; Corsetti, A.; Di Cagno, R. Biochemistry and Physiology of Sourdough Lactic Acid Bacteria. Trends Food Sci. Technol. 2005, 16, 57–69. [Google Scholar] [CrossRef]
  146. Polese, B.; Nicolai, E.; Genovese, D.; Verlezza, V.; La Sala, C.N.; Aiello, M.; Inglese, M.; Incoronato, M.; Sarnelli, G.; De Rosa, T.; et al. Postprandial Gastrointestinal Function Differs after Acute Administration of Sourdough Compared with Brewer’s Yeast Bakery Products in Healthy Adults. J. Nutr. 2018, 148, 202–208. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  147. Mutlu, C.; Candal-Uslu, C.; Özhanlı, H.; Arslan-Tontul, S.; Erbas, M. Modulating of Food Glycemic Response by Lactic Acid Bacteria. Food Biosci. 2022, 47, 101685. [Google Scholar] [CrossRef]
  148. Demirkesen-Bicak, H.; Arici, M.; Yaman, M.; Karasu, S.; Sagdic, O. Effect of Different Fermentation Condition on Estimated Glycemic Index, In Vitro Starch Digestibility, and Textural and Sensory Properties of Sourdough Bread. Foods 2021, 10, 514. [Google Scholar] [CrossRef]
  149. Rolim, M.E.; Fortes, M.I.; Von Frankenberg, A.; Duarte, C.K. Consumption of Sourdough Bread and Changes in the Glycemic Control and Satiety: A Systematic Review. Crit. Rev. Food Sci. Nutr. 2022. [Google Scholar] [CrossRef]
  150. Hansen, A.; Schieberle, P. Generation of Aroma Compounds during Sourdough Fermentation: Applied and Fundamental Aspects. Trends Food Sci. Technol. 2005, 16, 85–94. [Google Scholar] [CrossRef]
  151. Albagli, G.; do Monte Schwartz, I.; Amaral, P.F.F.; Ferreira, T.F.; Finotelli, P.V. How Dried Sourdough Starter Can Enable and Spread the Use of Sourdough Bread. LWT 2021, 149, 111888. [Google Scholar] [CrossRef]
  152. Mao, H.; Chen, X.D.; Fu, N. Exploring the Integrity of Cellular Membrane and Resistance to Digestive Juices of Dehydrated Lactic Acid Bacteria as Influenced by Drying Kinetics. Food Res. Int. 2022, 157, 111395. [Google Scholar] [CrossRef]
  153. Gao, X.; Kong, J.; Zhu, H.; Mao, B.; Cui, S.; Zhao, J. Lactobacillus, Bifidobacterium and Lactococcus Response to Environmental Stress: Mechanisms and Application of Cross-Protection to Improve Resistance against Freeze-Drying. J. Appl. Microbiol. 2022, 132, 802–821. [Google Scholar] [CrossRef]
  154. Rajam, R.; Subramanian, P. Encapsulation of Probiotics: Past, Present and Future. Beni-Suef Univ. J. Basic Appl. Sci. 2022, 11, 46. [Google Scholar] [CrossRef]
  155. Costa, K.; Silva, L.S.; Kobori, C.N.; da Silva, A.M.; Nicoli, J.R. Microencapsulation of Bifidobacterium longum 51A Cells by Spray Drying and Its Incorporation in Acerola (Malpighia Emarginata) Pulp Powder. Int. J. Food Sci. Technol. 2022, 57, 323–329. [Google Scholar] [CrossRef]
  156. Katina, K.; Salmenkallio-Marttila, M.; Partanen, R.; Forssell, P.; Autio, K. Effects of Sourdough and Enzymes on Staling of High-Fibre Wheat Bread. LWT Food Sci. Technol. 2006, 39, 479–491. [Google Scholar] [CrossRef]
  157. Palla, M.; Agnolucci, M.; Calzone, A.; Giovannetti, M.; Di Cagno, R.; Gobbetti, M.; Rizzello, C.G.; Pontonio, E. Exploitation of Autochthonous Tuscan Sourdough Yeasts as Potential Starters. Int. J. Food Microbiol. 2019, 302, 59–68. [Google Scholar] [CrossRef] [PubMed]
  158. Tang, K.X.; Zhao, C.J.; Gänzle, M.G. Effect of Glutathione on the Taste and Texture of Type I Sourdough Bread. J. Agric. Food Chem. 2017, 65, 4321–4328. [Google Scholar] [CrossRef] [PubMed]
  159. Menezes, L.A.A.; Molognoni, L.; de Sá Ploêncio, L.A.; Costa, F.B.M.; Daguer, H.; Dea Lindner, J.D. Use of Sourdough Fermentation to Reducing FODMAPs in Breads. Eur. Food Res. Technol. 2019, 245, 1183–1195. [Google Scholar] [CrossRef]
  160. van Hijum, S.A.F.T.; Kralj, S.; Ozimek, L.K.; Dijkhuizen, L.; van Geel-Schutten, I.G.H. Structure-Function Relationships of Glucansucrase and Fructansucrase Enzymes from Lactic Acid Bacteria. Microbiol. Mol. Biol. Rev. 2006, 70, 157–176. [Google Scholar] [CrossRef] [Green Version]
  161. Yu, L.; Qian, Z.; Ge, J.; Du, R. Glucansucrase Produced by Lactic Acid Bacteria: Structure, Properties, and Applications. Fermentation 2022, 8, 629. [Google Scholar] [CrossRef]
  162. Galle, S.; Arendt, E.K. Exopolysaccharides from Sourdough Lactic Acid Bacteria. Crit. Rev. Food Sci. Nutr. 2014, 54, 891–901. [Google Scholar] [CrossRef]
  163. Cozzolino, A.; Di Pierro, P.; Mariniello, L.; Sorrentino, A.; Masi, P.; Porta, R. Incorporation of Whey Proteins into Cheese Curd by Using Transglutaminase. Biotechnol. Appl. Biochem. 2003, 38, 289–295. [Google Scholar] [CrossRef]
  164. Scarnato, L.; Serrazanetti, D.I.; Aloisi, I.; Montanari, C.; Del Duca, S.; Lanciotti, R. Combination of Transglutaminase and Sourdough on Gluten-Free Flours to Improve Dough Structure. Amino Acids 2016, 48, 2453–2465. [Google Scholar] [CrossRef] [PubMed]
  165. Pourmohammadi, K.; Abedi, E. Enzymatic Modifications of Gluten Protein: Oxidative Enzymes. Food Chem. 2021, 356, 129679. [Google Scholar] [CrossRef]
  166. BRENDA Information on EC 1.8.1.7—Glutathione-Disulfide Reductase. Available online: https://www.brenda-enzymes.org/enzyme.php?ecno=1.8.1.7#reactschemes (accessed on 20 November 2022).
  167. Xu, D.; Tang, K.; Hu, Y.; Xu, X.; Gänzle, M.G. Effect of Glutathione Dehydrogenase of Lactobacillus sanfranciscensis on Gluten Properties and Bread Volume in Type i Wheat Sourdough Bread. J. Agric. Food Chem. 2018, 66, 9770–9776. [Google Scholar] [CrossRef] [PubMed]
  168. Khan, J.; Khurshid, S.; Sarwar, A.; Aziz, T.; Naveed, M.; Ali, U.; Makhdoom, S.I.; Nadeem, A.A.; Khan, A.A.; Sameeh, M.Y.; et al. Enhancing Bread Quality and Shelf Life via Glucose Oxidase Immobilized on Zinc Oxide Nanoparticles—A Sustainable Approach towards Food Safety. Sustainability 2022, 14, 14255. [Google Scholar] [CrossRef]
  169. Xu, X.; Yang, Q.; Luo, Z.; Xiao, Z. Effects of Sourdough Fermentation and an Innovative Compound Improver on the Baking Performance, Nutritional Quality, and Antistaling Property of Whole Wheat Bread. ACS Food Sci. Technol. 2022, 2, 825–835. [Google Scholar] [CrossRef]
  170. Visvanathan, R.; Jayathilake, C.; Liyanage, R.; Sivakanesan, R. Applicability and Reliability of the Glucose Oxidase Method in Assessing α-Amylase Activity. Food Chem. 2019, 275, 265–272. [Google Scholar] [CrossRef]
  171. Decamps, K.; Joye, I.J.; Courtin, C.M.; Delcour, J.A. Glucose and Pyranose Oxidase Improve Bread Dough Stability. J. Cereal Sci. 2012, 55, 380–384. [Google Scholar] [CrossRef]
  172. Decamps, K.; Joye, I.J.; Rakotozafy, L.; Nicolas, J.; Courtin, C.M.; Delcour, J.A. The Bread Dough Stability Improving Effect of Pyranose Oxidase from Trametes multicolor and Glucose Oxidase from Aspergillus niger: Unraveling the Molecular Mechanism. J. Agric. Food Chem. 2013, 61, 7848–7854. [Google Scholar] [CrossRef]
  173. Manhivi, V.E.; Amonsou, E.O.; Kudanga, T. Laccase-Mediated Crosslinking of Gluten-Free Amadumbe Flour Improves Rheological Properties. Food Chem. 2018, 264, 157–163. [Google Scholar] [CrossRef]
  174. Mayolo-Deloisa, K.; González-González, M.; Rito-Palomares, M. Laccases in Food Industry: Bioprocessing, Potential Industrial and Biotechnological Applications. Front. Bioeng. Biotechnol. 2020, 8, 222. [Google Scholar] [CrossRef] [Green Version]
  175. Niño-Medina, G.; Gutiérrez-Soto, G.; Urías-Orona, V.; Hernández-Luna, C.E. Effect of Laccase from Trametes maxima CU1 on Physicochemical Quality of Bread. Cogent Food Agric. 2017, 3, 1328762. [Google Scholar] [CrossRef]
  176. Papadimitriou, K.; Zoumpopoulou, G.; Georgalaki, M.; Alexandraki, V.; Kazou, M.; Anastasiou, R.; Tsakalidou, E. Sourdough Bread. In Innovations in Traditional Foods; Elsevier: Amsterdam, The Netherlands, 2019; pp. 127–158. ISBN 978-0-12-814887-7. [Google Scholar]
  177. IUBMB International Union of Biochemistry and Molecular Biology. Available online: https://iubmb.qmul.ac.uk/ (accessed on 20 November 2022).
  178. Diana, M.; Rafecas, M.; Quílez, J. Free Amino Acids, Acrylamide and Biogenic Amines in Gamma-Aminobutyric Acid Enriched Sourdough and Commercial Breads. J. Cereal Sci. 2014, 60, 639–644. [Google Scholar] [CrossRef]
  179. Polak, T.; Mejaš, R.; Jamnik, P.; Kralj Cigić, I.; Poklar Ulrih, N.; Cigić, B. Accumulation and Transformation of Biogenic Amines and Gamma-Aminobutyric Acid (GABA) in Chickpea Sourdough. Foods 2021, 10, 2840. [Google Scholar] [CrossRef] [PubMed]
  180. Venturi, M.; Galli, V.; Pini, N.; Guerrini, S.; Granchi, L. Use of Selected Lactobacilli to Increase γ-Aminobutyric Acid (GABA) Content in Sourdough Bread Enriched with Amaranth Flour. Foods 2019, 8, 218. [Google Scholar] [CrossRef] [Green Version]
  181. Calabrese, F.M.; Ameur, H.; Nikoloudaki, O.; Celano, G.; Vacca, M.; JFLemos, W., Jr.; Manzari, C.; Vertè, F.; Di Cagno, R.; Pesole, G.; et al. Metabolic Framework of Spontaneous and Synthetic Sourdough Metacommunities to Reveal Microbial Players Responsible for Resilience and Performance. Microbiome 2022, 10, 148. [Google Scholar] [CrossRef] [PubMed]
  182. Pourmohammadi, K.; Abedi, E. Hydrolytic Enzymes and Their Directly and Indirectly Effects on Gluten and Dough Properties: An Extensive Review. Food Sci. Nutr. 2021, 9, 3988–4006. [Google Scholar] [CrossRef] [PubMed]
  183. Dahiya, S.; Bajaj, B.K.; Kumar, A.; Tiwari, S.K.; Singh, B. A Review on Biotechnological Potential of Multifarious Enzymes in Bread Making. Process Biochem. 2020, 99, 290–306. [Google Scholar] [CrossRef]
  184. Bala, A.; Singh, B. Concomitant Production of Cellulase and Xylanase by Thermophilic Mould Sporotrichum thermophile in Solid State Fermentation and Their Applicability in Bread Making. World J. Microbiol. Biotechnol. 2017, 33, 109. [Google Scholar] [CrossRef] [PubMed]
  185. Cunha, L.; Martarello, R.; de Souza, P.M.; de Freitas, M.M.; Barros, K.V.G.; Ferreira Filho, E.X.; Homem-de-Mello, M.; Magalhães, P.O. Optimization of Xylanase Production from Aspergillus foetidus in Soybean Residue. Enzyme Res. 2018, 2018, 6597017. [Google Scholar] [CrossRef] [Green Version]
  186. Guo, Y.; Gao, Z.; Xu, J.; Chang, S.; Wu, B.; He, B. A Family 30 Glucurono-Xylanase from Bacillus subtilis LC9: Expression, Characterization and Its Application in Chinese Bread Making. Int. J. Biol. Macromol. 2018, 117, 377–384. [Google Scholar] [CrossRef]
  187. Adigüzel, A.O.; Tunçer, M. Production, Characterization and Application of a Xylanase from Streptomyces sp. AOA40 in Fruit Juice and Bakery Industries. Food Biotechnol. 2016, 30, 189–218. [Google Scholar] [CrossRef]
  188. García-Mantrana, I.; Monedero, V.; Haros, M. Myo-Inositol Hexakisphosphate Degradation by Bifidobacterium pseudocatenulatum ATCC 27919 Improves Mineral Availability of High Fibre Rye-Wheat Sour Bread. Food Chem. 2015, 178, 267–275. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  189. Chanderman, A.; Puri, A.K.; Permaul, K.; Singh, S. Production, Characteristics and Applications of Phytase from a Rhizosphere Isolated Enterobacter sp. ACSS. Bioprocess Biosyst. Eng. 2016, 39, 1577–1587. [Google Scholar] [CrossRef] [PubMed]
  190. Ranjan, B.; Satyanarayana, T. Recombinant HAP Phytase of the Thermophilic Mold Sporotrichum Thermophile: Expression of the Codon-Optimized Phytase Gene in Pichia pastoris and Applications. Mol. Biotechnol. 2016, 58, 137–147. [Google Scholar] [CrossRef] [PubMed]
  191. Buddhiwant, P.; Bhavsar, K.; Kumar, V.R.; Khire, J.M. Phytase Production by Solid-State Fermentation of Groundnut Oil Cake by Aspergillus niger: A Bioprocess Optimization Study for Animal Feedstock Applications. Prep. Biochem. Biotechnol. 2016, 46, 531–538. [Google Scholar] [CrossRef]
  192. Benabda, O.; M’hir, S.; Kasmi, M.; Mnif, W.; Hamdi, M. Optimization of Protease and Amylase Production by Rhizopus oryzae Cultivated on Bread Waste Using Solid-State Fermentation. J. Chem. 2019, 2019, e3738181. [Google Scholar] [CrossRef] [Green Version]
  193. Trabelsi, S.; Ben Mabrouk, S.; Kriaa, M.; Ameri, R.; Sahnoun, M.; Mezghani, M.; Bejar, S. The Optimized Production, Purification, Characterization, and Application in the Bread Making Industry of Three Acid-Stable Alpha-Amylases Isoforms from a New Isolated Bacillus subtilis Strain US586. J. Food Biochem. 2019, 43, e12826. [Google Scholar] [CrossRef]
  194. Rana, N.; Verma, N.; Vaidya, D.; Dipta, B. Application of Bacterial Amylase in Clarification of Juices and Bun Making. J. Pharm. Phytochem. 2017, 6, 859–864. [Google Scholar]
  195. Shivlata, L.; Satyanarayana, T. Characteristics of Raw Starch-Digesting α-Amylase of Streptomyces badius DB-1 with Transglycosylation Activity and Its Applications. Appl. Biochem. Biotechnol. 2017, 181, 1283–1303. [Google Scholar] [CrossRef]
  196. Kiesenhofer, D.P.; Mach, R.L.; Mach-Aigner, A.R. Glucose Oxidase Production from Sustainable Substrates. Curr. Biotechnol. 2017, 6, 238–244. [Google Scholar] [CrossRef]
  197. Hassan, B.A.; Jebor, M.A.; Ali, Z.M. Purification and Characterization of the Glucose Oxidase from Penicillium notatum. Int. J. Pharm. Qual. Assur. 2018, 9, 55–63. [Google Scholar] [CrossRef]
  198. Yogananth, N.; Chanthru, A.; Sivanesan, D. Optimization of Various Parameters for the Production of Glucose Oxidase Using Aspergillus niger. Int. J. Curr. Microbiol. App. Sci. 2012, 1, 23–28. [Google Scholar]
  199. Hamid, H.M.; Khalil-ur-Rehman; Zia, M.A.; Asgher, M. Optimization of Various Parameters for the Production of Glucose Oxidase from Rice Polishing Using Aspergillus niger. Biotechnol. Pak. 2003, 2, 1–7. [Google Scholar]
  200. Bhange, K.; Chaturvedi, V.; Bhatt, R. Simultaneous Production of Detergent Stable Keratinolytic Protease, Amylase and Biosurfactant by Bacillus subtilis PF1 Using Agro Industrial Waste. Biotechnol. Rep. 2016, 10, 94–104. [Google Scholar] [CrossRef]
  201. Ozdenefe, M.S.; Dincer, S.; Unal, M.U.; Kayis, F.B.; Takci, H.A.M.; Arkut, A. Optimization of Culture Conditions for Alkaline Protease Production from Waste Breads Using Bacillus subtilis. Romanian Biotechnol. Lett. 2017, 22, 12597–12610. [Google Scholar]
  202. Sangeetha, R.; Geetha, A.; Arulpandi, I. Optimization of Protease and Lipase Production by Bacillus pumilus SG 2 Isolated from an Industrial Effluent. Internet J Microbiol 2008, 5, 2. [Google Scholar]
  203. Nema, A.; Patnala, S.H.; Mandari, V.; Kota, S.; Devarai, S.K. Production and Optimization of Lipase Using Aspergillus niger MTCC 872 by Solid-State Fermentation. Bull. Natl. Res. Cent. 2019, 43, 82. [Google Scholar] [CrossRef] [Green Version]
  204. Tanyol, M.; Uslu, G.; Yönten, V. Optimization of Lipase Production on Agro-Industrial Residue Medium by Pseudomonas fluorescens (NRLL B-2641) Using Response Surface Methodology. Biotechnol. Biotechnol. Equip. 2015, 29, 64–71. [Google Scholar] [CrossRef]
  205. Iqbal, S.A.; Rehman, A. Characterization of Lipase from Bacillus subtilis I-4 and Its Potential Use in Oil Contaminated Wastewater. Braz. Arch. Biol. Technol. 2015, 58, 789–797. [Google Scholar] [CrossRef] [Green Version]
  206. Kumar, D.J.M.; Rejitha, R.; Devika, S.; Balakumaran, M.D.; Rebecca, A.I.N.; Kalaichelvan, P.T. Production, Optimization and Purification of Lipase from Bacillus sp. MPTK 912 Isolated from Oil Mill Effluent. Adv. Appl. Sci. Res. 2012, 3, 930–938. [Google Scholar]
  207. Mahmoud, A.-M.Y.; Asif, A.M.J.-F.; Hani, Z.A. Production, Purification and Characterization of Cellulase from Streptomyces sp. Afr. J. Microbiol. Res. 2014, 8, 348–354. [Google Scholar] [CrossRef] [Green Version]
  208. Swathy, R.; Rambabu, K.; Banat, F.; Ho, S.-H.; Chu, D.-T.; Show, P.L. Production and Optimization of High Grade Cellulase from Waste Date Seeds by Cellulomonas Uda NCIM 2353 for Biohydrogen Production. Int. J. Hydrog. Energy 2020, 45, 22260–22270. [Google Scholar] [CrossRef]
  209. Verma, N.; Kumar, V.; Bansal, M.C. Utility of Luffa cylindrica and Litchi chinensis Peel, an Agricultural Waste Biomass in Cellulase Production by Trichoderma reesei under Solid State Cultivation. Biocatal. Agric. Biotechnol. 2018, 16, 483–492. [Google Scholar] [CrossRef]
  210. Oort, M.V.; Whitehurst, R.J. Enzymes in Food Technology; John Wiley & Sons: Hoboken, NJ, USA, 2009; ISBN 978-1-4443-0994-2. [Google Scholar]
  211. Dalmaso, G.Z.L.; Ferreira, D.; Vermelho, A.B. Marine Extremophiles: A Source of Hydrolases for Biotechnological Applications. Mar. Drugs 2015, 13, 1925–1965. [Google Scholar] [CrossRef] [Green Version]
  212. Kanazawa, S.; Sanabria, M.; Monteiro, M. Influence of the Fermentation Methods on the Resistant Starch Formation by X-Ray Diffraction. SN Appl. Sci. 2021, 3, 191. [Google Scholar] [CrossRef]
  213. Wang, S.; Li, C.; Copeland, L.; Niu, Q.; Wang, S. Starch Retrogradation: A Comprehensive Review. Compr. Rev. Food Sci. Food Saf. 2015, 14, 568–585. [Google Scholar] [CrossRef]
  214. Corsetti, A.; Gobbetti, M.; De Marco, B.; Balestrieri, F.; Paoletti, F.; Russi, L.; Rossi, J. Combined Effect of Sourdough Lactic Acid Bacteria and Additives on Bread Firmness and Staling. J Agric Food Chem 2000, 48, 3044–3051. [Google Scholar] [CrossRef]
  215. Muir, J.G.G.; Varney, J.E.E.; Ajamian, M.; Gibson, P.R.R. Gluten-Free and Low-FODMAP Sourdoughs for Patients with Coeliac Disease and Irritable Bowel Syndrome: A Clinical Perspective. Int. J. Food Microbiol. 2019, 290, 237–246. [Google Scholar] [CrossRef]
  216. Vermelho, A.B.; Couri, S. Methods to Determine Enzymatic Activity; Bentham Science Publishers: Saif Zone, Sharjah, United Arab Emirates, 2013. [Google Scholar]
  217. Melim Miguel, A.S.; Souza, T.; da Costa Figueiredo, E.V.; Paulo Lobo, B.W.; Maria, G. Enzymes in Bakery: Current and Future Trends. In Food Industry; Muzzalupo, I., Ed.; InTech: London, UK, 2013; ISBN 978-953-51-0911-2. [Google Scholar]
  218. Li, Z.; Dong, Y.; Zhou, X.; Xiao, X.; Zhao, Y.; Yu, L. Dough Properties and Bread Quality of Wheat–Barley Composite Flour as Affected by β-Glucanase. Cereal Chem. J. 2014, 91, 631–638. [Google Scholar] [CrossRef]
  219. Lei, X.G.; Porres, J.M. Phytase Enzymology, Applications, and Biotechnology. Biotechnol. Lett. 2003, 25, 1787–1794. [Google Scholar] [CrossRef]
  220. Ries, E.F. Estudo Da Produção, Caracterização e Aplicação de Nova Fitase de Saccharomyces cerevisiae. Ph.D. Thesis, Universidade Estadual de Campinas, Campinas, Brazil, 2010; 130p. [Google Scholar]
  221. Aslam, M.F.; Ellis, P.R.; Berry, S.E.; Latunde-Dada, G.O.; Sharp, P.A. Enhancing Mineral Bioavailability from Cereals: Current Strategies and Future Perspectives. Nutr. Bull. 2018, 43, 184–188. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  222. Baye, K.; Guyot, J.-P.; Icard-Vernière, C.; Rochette, I.; Mouquet-Rivier, C. Enzymatic Degradation of Phytate, Polyphenols and Dietary Fibers in Ethiopian Injera Flours: Effect on Iron Bioaccessibility. Food Chem. 2015, 174, 60–67. [Google Scholar] [CrossRef] [PubMed]
  223. Brouns, F. Phytic Acid and Whole Grains for Health Controversy. Nutrients 2022, 14, 25. [Google Scholar] [CrossRef] [PubMed]
  224. Iglesias-Puig, E.; Monedero, V.; Haros, M. Bread with Whole Quinoa Flour and Bifidobacterial Phytases Increases Dietary Mineral Intake and Bioavailability. LWT Food Sci. Technol. 2015, 60, 71–77. [Google Scholar] [CrossRef] [Green Version]
  225. Nuobariene, L.; Hansen, Å.S.; Arneborg, N. Isolation and Identification of Phytase-Active Yeasts from Sourdoughs. LWT Food Sci. Technol. 2012, 48, 190–196. [Google Scholar] [CrossRef]
  226. Nuobariene, L.; Arneborg, N.; Hansen, Å.S. Phytase Active Yeasts Isolated From Bakery Sourdoughs. In Proceedings of the 9th Baltic Conference on Food Science and Technology “Food for Consumer Well-Being”, Jelgava, Latvia, 8–9 May 2014; pp. 223–227. [Google Scholar]
  227. Mohammadi, M.; Zoghi, A.; Azizi, M.H. Effect of Xylanase and Pentosanase Enzymes on Dough Rheological Properties and Quality of Baguette Bread. J. Food Qual. 2022, 2022, 2910821. [Google Scholar] [CrossRef]
  228. Martínez-Anaya, M.A.; Devesa, A. Influence of Enzymes in Sourdough Wheat Breadmaking. Changes in Pentosans / Influencia de Las Enzimas y Los Iniciadores Microbianos En Panificación. Cambios En Las Pentosanas. Food Sci. Technol. Int. 2000, 6, 109–116. [Google Scholar] [CrossRef]
  229. Leahu, A.; Codină, G.G.; Mironeasa, S.; Roşu, A.-I. Effects of A2 Phospholipase on Dough Rheological Properties and Bread Characteristics. Food Environ. Saf. J. 2011, 10, 1. [Google Scholar]
  230. Gandra, K.M.; Del Bianchi, M.; Godoy, V.P.; Queiroz, F.P.C.; Steel, C.J. Aplicação de Lipase e Monoglicerídeo Em Pão de Forma Enriquecido Com Fibras. Cienc. E Tecnol. Aliment. 2008, 28, 182–192. [Google Scholar] [CrossRef] [Green Version]
  231. Gobbetti, M.; De Angelis, M.; Di Cagno, R.; Calasso, M.; Archetti, G.; Rizzello, C.G. Novel Insights on the Functional/Nutritional Features of the Sourdough Fermentation. Int. J. Food Microbiol. 2018, 302, 103–113. [Google Scholar] [CrossRef]
  232. Zhu, K.-X.; Zhou, H.-M.; Qian, H.-F. Proteins Extracted from Defatted Wheat Germ: Nutritional and Structural Properties. Cereal Chem. 2006, 83, 69–75. [Google Scholar] [CrossRef]
  233. Rizzello, C.G.; Nionelli, L.; Coda, R.; De Angelis, M.; Gobbetti, M. Effect of Sourdough Fermentation on Stabilisation, and Chemical and Nutritional Characteristics of Wheat Germ. Food Chem. 2010, 119, 1079–1089. [Google Scholar] [CrossRef]
  234. Rawlings, N.D.; Morton, F.R.; Kok, C.Y.; Kong, J.; Barrett, A.J. The MEROPS Database of Proteolytic Enzymes, Their Substrates and Inhibitors in 2017 and a Comparison with Peptidases in the PANTHER Database. Nucleic Acids Res. 2018, 36, D320–D325. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  235. Vermeulen, N.; Kretzer, J.; Machalitza, H.; Vogel, R.F.; Gänzle, M.G. Influence of Redox-Reactions Catalysed by Homo- and Hetero-Fermentative Lactobacilli on Gluten in Wheat Sourdoughs. J. Cereal Sci. 2006, 43, 137–143. [Google Scholar] [CrossRef]
  236. Wieser, H. Chemistry of Gluten Proteins. Food Microbiol. 2007, 24, 115–119. [Google Scholar] [CrossRef]
  237. Bietz, J.; Wall, J. Identity of High Molecular Weight Gliadin and Ethanol-Soluble Glutenin Subunits of Wheat: Relation to Gluten Structure. Cereal Chem. 1980, 57, 415–421. [Google Scholar]
  238. Gänzle, M.G.; Loponen, J.; Gobbetti, M. Proteolysis in Sourdough Fermentations: Mechanisms and Potential for Improved Bread Quality. Trends Food Sci. Technol. 2008, 19, 513–521. [Google Scholar] [CrossRef]
  239. De Souza, T.S.P.; de Andrade, C.J.; Koblitz, M.G.B.; Fai, A.E.C. Microbial Peptidase in Food Processing: Current State of the Art and Future Trends. Catal. Lett. 2023, 153, 114–137. [Google Scholar] [CrossRef]
  240. Di Cagno, R.; Rizzello, C.G.; De Angelis, M.; Cassone, A.; Giuliani, G.; Benedusi, A.; Limitone, A.; Surico, R.F.; Gobbetti, M. Use of Selected Sourdough Strains of Lactobacillus for Removing Gluten and Enhancing the Nutritional Properties of Gluten-Free Bread. J. Food Prot. 2008, 71, 1491–1495. [Google Scholar] [CrossRef]
  241. Thiele, C.; Grassl, S.; Gänzle, M. Gluten Hydrolysis and Depolymerization during Sourdough Fermentation. J. Agric. Food Chem. 2004, 52, 1307–1314. [Google Scholar] [CrossRef]
  242. Shan, L.; Qiao, S.-W.; Arentz-Hansen, H.; Molberg, Ø.; Gray, G.M.; Sollid, L.M.; Khosla, C. Identification and Analysis of Multivalent Proteolytically Resistant Peptides from Gluten: Implications for Celiac Sprue. J. Proteome Res. 2005, 4, 1732–1741. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  243. Engström, N.; Sandberg, A.S.; Scheers, N. Sourdough Fermentation of Wheat Flour Does Not Prevent the Interaction of Transglutaminase 2 with A2-Gliadin or Gluten. Nutrients 2015, 7, 2134. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  244. Dunaevsky, Y.E.; Tereshchenkova, V.F.; Belozersky, M.A.; Filippova, I.Y.; Oppert, B.; Elpidina, E.N. Effective Degradation of Gluten and Its Fragments by Gluten-Specific Peptidases: A Review on Application for the Treatment of Patients with Gluten Sensitivity. Pharmaceutics 2021, 13, 1603. [Google Scholar] [CrossRef] [PubMed]
  245. Reale, A.; Di Stasio, L.; Di Renzo, T.; De Caro, S.; Ferranti, P.; Picariello, G.; Addeo, F.; Mamone, G. Bacteria Do It Better! Proteomics Suggests the Molecular Basis for Improved Digestibility of Sourdough Products. Food Chem. 2021, 359, 129955. [Google Scholar] [CrossRef]
  246. Alvarez-Sieiro, P.; Redruello, B.; Ladero, V.; Martín, M.C.; Fernández, M.; Alvarez, M.A. Screening Sourdough Samples for Gliadin-Degrading Activity Revealed Lactobacillus casei Strains Able to Individually Metabolize the Coeliac-Disease-Related 33-Mer Peptide. Can. J. Microbiol. 2016, 62, 422–430. [Google Scholar] [CrossRef] [Green Version]
  247. Lancetti, R.; Sciarini, L.; Pérez, G.T.; Salvucci, E. Technological Performance and Selection of Lactic Acid Bacteria Isolated from Argentinian Grains as Starters for Wheat Sourdough. Curr. Microbiol. 2021, 78, 255–264. [Google Scholar] [CrossRef]
  248. Zhao, C.J.; Schieber, A.; Gänzle, M.G. Formation of Taste-Active Amino Acids, Amino Acid Derivatives and Peptides in Food Fermentations—A Review. Food Res. Int. 2016, 89, 39–47. [Google Scholar] [CrossRef]
  249. Feng, W.; Ma, S.; Wang, X. Quality Deterioration and Improvement of Wheat Gluten Protein in Frozen Dough. Grain Oil Sci. Technol. 2020, 3, 29–37. [Google Scholar] [CrossRef]
  250. Schopf, M.; Scherf, K.A. Water Absorption Capacity Determines the Functionality of Vital Gluten Related to Specific Bread Volume. Foods 2021, 10, 228. [Google Scholar] [CrossRef]
  251. Codex Alimentarius. Agenda Item 11 Cx/Nfsdu 19/41/11. Joint FAO/WHO Food Standards Programme. In Proceedings of the Codex Committee on Nutrition and Foods for Special Dietary Uses Forty-First Session, Dusseldorf, Germany, 24–29 November 2019. [Google Scholar]
  252. FDA. Generally Recognized as Safe (GRAS). Available online: https://www.fda.gov/food/food-ingredients-packaging/generally-recognized-safe-gras (accessed on 10 October 2022).
  253. Anvisa Resolução Da Diretoria Colegiada—RDC No 241, DE 26 DE JULHO DE 2018; Imprensa Nacional: Brasília, Brazil, 2018.
  254. Planalto. LEI Nº 13.123. 2015. Available online: https://www.planalto.gov.br/ccivil_03/_ato2015-2018/2015/lei/l13123.htm (accessed on 10 October 2022).
  255. De Melo Pereira, G.V.; de Oliveira Coelho, B.; Magalhães Júnior, A.I.; Thomaz-Soccol, V.; Soccol, C.R. How to Select a Probiotic? A Review and Update of Methods and Criteria. Biotechnol. Adv. 2018, 36, 2060–2076. [Google Scholar] [CrossRef]
  256. FAO; WHO. Probiotics in Food: Health and Nutritional Properties and Guidelines for Evaluation; FAO Food and Nutrition Paper; Food and Agriculture Organization of the United Nations: Rome, Italy, 2006; ISBN 978-92-5-105513-7. [Google Scholar]
  257. Legifrance. France Décret N°93-1074. 1993. Available online: https://www.legifrance.gouv.fr/loda/id/JORFTEXT000000727617/ (accessed on 10 October 2022).
  258. Eur Regulamento (UE) 2015/2283 do Parlamento Europeu e do Conselho. Available online: https://eur-lex.europa.eu/legal-content/PT/TXT/PDF/?uri=CELEX:32015R2283&from=EN (accessed on 6 October 2022).
  259. Europa Safety Assessment and Regulatory Aspects of Micro-Organisms in Feed and Food Applications. Available online: https://food.ec.europa.eu/system/files/2020-12/sci-com_scan-old_report_out85.pdf (accessed on 4 October 2022).
  260. Eur-Lex Establishing the Union List of Novel Foods in Accordance with Regulation (EU) 2015/2283 of the European Parliament and of the Council on Novel Foods. Available online: https://eur-lex.europa.eu/legal-content/EN/TXT/PDF/?uri=CELEX:32017R2470&from=EN (accessed on 2 October 2022).
  261. Balachandra Nair, R.; Ramachandranna, P.C. Patenting of Microorganisms: Systems and Concerns. J. Commer. Biotechnol. 2010, 16, 337–347. [Google Scholar] [CrossRef]
  262. Planalto. LEI Nº 9.279. 1996. Available online: https://www.planalto.gov.br/ccivil_03/leis/l9279.htm (accessed on 10 October 2022).
Figure 1. Sourdough production of the four types of processes. Type II and III can be scaled to an industrial level.
Figure 1. Sourdough production of the four types of processes. Type II and III can be scaled to an industrial level.
Fermentation 09 00090 g001
Figure 2. Major sourdough fermentation pathways. (A) Lactic fermentation. (B) Alcoholic fermentation. (A1), homolactic fermentation; (A2), heterolactic fermentation. (B3), alcoholic fermentation from yeast. The formation of ethanol and CO2 from the reduction of pyruvate characterizes alcoholic fermentation.
Figure 2. Major sourdough fermentation pathways. (A) Lactic fermentation. (B) Alcoholic fermentation. (A1), homolactic fermentation; (A2), heterolactic fermentation. (B3), alcoholic fermentation from yeast. The formation of ethanol and CO2 from the reduction of pyruvate characterizes alcoholic fermentation.
Fermentation 09 00090 g002
Figure 3. General scheme of bread sourdough fermentation and the major enzymes involved. EPS: Exopolysaccharides.
Figure 3. General scheme of bread sourdough fermentation and the major enzymes involved. EPS: Exopolysaccharides.
Fermentation 09 00090 g003
Figure 4. Some possible interactions of wheat proteins. (A)-low water content. (B)-baking mixture with optimal water content, based on the work of Feng et al. and Schopf and Scherf [249,250].
Figure 4. Some possible interactions of wheat proteins. (A)-low water content. (B)-baking mixture with optimal water content, based on the work of Feng et al. and Schopf and Scherf [249,250].
Fermentation 09 00090 g004
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Akamine, I.T.; Mansoldo, F.R.P.; Vermelho, A.B. Probiotics in the Sourdough Bread Fermentation: Current Status. Fermentation 2023, 9, 90. https://doi.org/10.3390/fermentation9020090

AMA Style

Akamine IT, Mansoldo FRP, Vermelho AB. Probiotics in the Sourdough Bread Fermentation: Current Status. Fermentation. 2023; 9(2):90. https://doi.org/10.3390/fermentation9020090

Chicago/Turabian Style

Akamine, Ingrid Teixeira, Felipe R. P. Mansoldo, and Alane Beatriz Vermelho. 2023. "Probiotics in the Sourdough Bread Fermentation: Current Status" Fermentation 9, no. 2: 90. https://doi.org/10.3390/fermentation9020090

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop