Next Article in Journal
Biological Demalication and Deacetification of Musts and Wines: Can Wine Yeasts Make the Wine Taste Better?
Previous Article in Journal
The Application of State-of-the-Art Analytic Tools (Biosensors and Spectroscopy) in Beverage and Food Fermentation Process Monitoring
Previous Article in Special Issue
Production and Quality Analysis of Wine from Honey and Coconut Milk Blend Using Saccharomyces cerevisiae
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Spontaneous Food Fermentations and Potential Risks for Human Health

by
Vittorio Capozzi
1,
Mariagiovanna Fragasso
2,
Rossana Romaniello
1,
Carmen Berbegal
1,3,
Pasquale Russo
4 and
Giuseppe Spano
1,*
1
Department of Agriculture, Food and Environment Sciences, University of Foggia, Via Napoli 25, 71122 Foggia, Italy
2
Cereal Research Centre, Council for Agricultural Research and Economics, 71122 Foggia, Italy
3
EnolabERI BioTecMed, Universitat de València, 46100 Valencia, Spain
4
Promis Biotech srl, Via Napoli 25, 71122 Foggia, Italy
*
Author to whom correspondence should be addressed.
Fermentation 2017, 3(4), 49; https://doi.org/10.3390/fermentation3040049
Submission received: 31 August 2017 / Revised: 19 September 2017 / Accepted: 25 September 2017 / Published: 28 September 2017
(This article belongs to the Special Issue Microbiology and Food Hygiene)

Abstract

:
Fermented foods and beverages are a heterogeneous class of products with a relevant worldwide significance for human economy, nutrition and health for millennia. A huge diversity of microorganisms is associated with the enormous variety in terms of raw materials, fermentative behavior and obtained products. In this wide microbiodiversity it is possible that the presence of microbial pathogens and toxic by-products of microbial origin, including mycotoxins, ethyl carbamate and biogenic amines, are aspects liable to reduce the safety of the consumed product. Together with other approaches (e.g., use of preservatives, respect of specific physico-chemical parameters), starter cultures technology has been conceived to successfully dominate indigenous microflora and to drive fermentation to foresee the desired attributes of the matrix, assuring quality and safety. Recent trends indicate a general return to spontaneous food fermentation. In this review, we point out the potential risks for human health associated with uncontrolled (uninoculated) food fermentation and we discuss biotechnological approaches susceptible to conciliate fermented food safety, with instances of an enhanced contribution of microbes associated to spontaneous fermentation.

Graphical Abstract

1. Fermented Foods and Beverages: Scientific Dimension, Social Relevance, and Economic Significance

A large basket of food and beverages is obtained from a microbial-based transformation of food raw materials. Different classes of microorganisms can be involved, mainly yeasts and bacteria, with a certain role of molds. The obtained fermented foods and beverages have ben staple foods for millennia, with a considerable importance in the human diet for reasons of generally enhanced shelf-life, palatability, safety and nutritional quality [1]. In fact, the desired fermentation process consists of protechnological microbial development in the given matrix, with direct and indirect effects of primary and secondary microbial metabolism. Protechnological microorganisms, in order to obtain energy and to sustain their anabolic processes, reduce the content of carbohydrates and other macromolecules available in the raw matrix, accumulating catabolic products (e.g., lactic acid, ethanol). These biological dynamics, together with the possible release of antimicrobial compounds [2], reduce the risks of undesired microbial developments (thus increasing product shelf-life and safety level). On the other hand, both primary and secondary metabolites strongly influence palatability and sensory attributes of the fermented matrices [3]. Finally, these modifications performed by microorganisms radically change the nutritional quality of the food/beverage, also because of the production of biomolecules of nutritional interest synthetized by fermenting cells [4]. Certainly, this is just a partial brief overview on the impact of fermenting microbes on the global (e.g., hygienic, sensory, nutritional, functional) quality of fermented foods/beverages, a topic that will be partially deepened in the next section. What is important to remark is the importance of the so-called “omics” approaches (metagenomics, transcriptomics, proteomics, metabolomics, volatomics) to offer new perspectives in the understanding of microbial contributions to the global quality of fermented foods [5,6,7,8,9].
From a geographical perspective, several scientific studies elucidate the microbiological and nutritional significance of fermented foods and beverages in the different continents: Asia (e.g., [10,11]), Africa (e.g., [11,12]), America (e.g., [11,13]), Australia (e.g., [11,14]), and Europe (e.g., [11,15]). On the other hand, as a function of the nature of the matrices (raw material) subjected to fermentation, recent literature suggests categorization of worldwide fermented foods and beverages into nine principal classes: (a) fermented cereals; (b) fermented vegetables and bamboo shoots; (c) fermented legumes; (d) fermented roots/tubers; (e) fermented milk products; (f) fermented and preserved meat products; (g) fermented, dried and smoked fish products; (h) miscellaneous fermented products; and (i) alcoholic beverages ([16]). Several productions can be considered worldwide diffused or with a national/continental presence, while considerable diversity in terms of matrices and of kind of fermentations has a regional diffusion. The widespread geographical diffusion ([11,16]) and the relevance in terms of consumption across populations make fermented foods and beverage a global sector, with an increasing relevance in human nutrition and economy. The significance in actual human nutrition framework is also well testified by the direct and indirect inclusion in the dietary guidelines of several countries (e.g., the Indian pyramid explicitly promotes the consumption of fermented foods) [17,18]. In order to underline the economic/social importance of fermentations, we have to remember that the so-called “ancient biotechnology” was empirically discovered and replicated as a food preservation technique [19].

2. The Microbiology of Fermented Foods and Beverages: “Microbiodiversity”, Impact on Food Qualities and on Human Health

A wide diversity of microorganisms is associated with the huge diversity in terms of raw materials, fermentative behavior and obtained products [16]. Table 1 reports just a few exemplificative cases of fermented products belonging to the nine categories proposed by Tamang et al. [16]. This brief overview also helps to provide an idea of the dimension of the microbial diversity associate to food fermentations: fermentative processes associated with the production of these twenty fermented products involved more than fifty microbial genera and hundreds of different species of bacteria, yeasts, and filamentous fungi. In addition, (i) if we consider that behind each species there is a consistent intraspecific diversity and that (ii) often desired and undesired microbial features associated with food fermentations are often strain-dependent characters, the potentially articulate impact of this diversity on global food quality appears clear.
The microbiota associated with the fermenting matrix is in a strong relationship, in a sort of continuous dichotomy, with the acceptance of the final products: it is liable to positively and negatively affect the main quality and safety attributes. In fact, microorganisms may improve or depreciate (i) the safety of the foodstuffs, modulating the content of biological and chemical contaminants (e.g., [22]); (ii) the palatability and the aroma, releasing volatile organic compounds and influencing the taste and the texture (e.g., [23]); (iii) the nutritional quality, modifying the quantity of macro- and micro-nutrients and their digestibility and bioavailability (e.g., [24]); (iv) the presence (in quality and quantity) of biological and/or chemical entities susceptible to maximize the consumer health (out of the benefits associated to the nutritional contribute) (e.g., [1,25]). Other case studies are reported in Table 2 to better exemplify how microbes can affect the global quality of fermented foods and beverages, encompassing applications such as toxic compound degradation, texturizing properties, bio-fortifications, and addition of functional ingredients [26,27,28,29].
In particular, the benefits for human health are of outstanding relevance: transformation of food constituents, biosynthesis of compounds with nutritional and/or functional importance, delivery of commensal microbes to the human gastro-intestinal tract, and delivery of probiotic strains [1]. Naturally, the presence of several possible benefits led us to focus on minimizing the risks for products safety in order to assure consumers health.

3. Risks for Human Health Associated with Fermentations

Just as for fresh or alternatively processed foods, there is a certain risk of contaminants that pose hazard to human health associated with fermented foods. We have to separate (micro) biological risks from chemical risks of microbiological origin. Cases of microbial pathogens have been reported in association with several fermented foods, such as cheese, sausages, fermented fish and fermented cereals [41]. On the other hand, we have to consider toxic by-products of microbial origin, including mycotoxins, ethyl carbamate and biogenic amines [41] (Table 3). As it is possible to notice, in only a few examples, we report different relevant pathogens (Bacillus cereus, Escherichia coli, Salmonella sp., Escherichia coli O157:H7, Staphylococcus aureus, Vibrio cholera, Listeria monocytogenes, Aeromonas, Klebsiella, Campylobacter and Shigella sp.), potent neurotoxin (Ochratoxin A), and several molecules belonging to the class of biogenic amines [42,43,44,45]. All cases that testify well the relevance of risks of microbial origin associated with this considerable class of foodstuffs.
Naturally, we have to remember the risk of the presence of biological, chemical and physical contaminants that do not deal with the microbiological dimension (e.g., insects, pesticides, glass sliver). However, the treatment of these categories of contaminants is out of the scope of this review, especially if we consider that their presence can be generally considered as unaffected in case of both spontaneous fermentation and inoculation with starter cultures.
Coming back to microbial associated hazards for human health, it is essential to separate (i) risks associated with microbial genera/species usually not found in association with fermented matrices, and (ii) genera/species detected in the monitoring of spontaneous fermentation. The first is the case of pathogens, while to the second class often belong to producers of mycotoxins, ethyl carbamate and biogenic amines. Hence, the second class is more insidious, considering that also in the same species we can find strains of protechnological interest and strains liable to produce toxic compounds. With this concern, in Table 4, it is possible to find a (non-exhaustive) list of microbial species associated with fermented matrices (in accordance with data reported in Table 1), for which the selection of strains liable to produce compounds toxic (to different extents) for human health has been reported in literature.
Moreover, we have to consider that not all the species reported by Tamang et al. [16] (e.g., those reported in Table 1) in association with worldwide food fermentations are (i) recognized as safe for human use in the framework of the principal national legislative frameworks (e.g., Generally Recognized as Safe (GRAS) by U.S. Food and Drug Administration; Qualified Presumption of Safety (QPS) by the European Food Safety Authority) [64], and/or (ii) included in the “inventory of microbial species used in food fermentations” proposed by the European Food and Feed Cultures Association ([65]). Moreover, from a biological point of view, we have to stress two other crucial concerns that deal with the safety of strains associated with food/beverages fermentations: (i) the occurrence of virulence traits; and (ii) the transfer of antibiotic resistance determinants [64,66,67].
All the mentioned aspects underline the presence of a situation that is particularly complex considering that in the case of fermented food a microbial development is desired, and thus is more difficult to limit the multiplication of undesired microbes. In addition, we have to consider recent tendencies such as the coming back to “natural” processing that in some cases increase food safety risks of microbial origin. It was the case of E. coli O157:H7 organisms found to survive for 20 days in unpreserved refrigerated apple cider artisanally produced (apples were not washed, cider was not pasteurized, and no preservatives were added) [48]. Finally, we have to underline that in specific clinical cases, also food-delivering microorganisms generally recognized as safe, such as lactic acid bacteria of protechnological interest and/or probiotic strains, can cause illness [68].

4. Spontaneous Versus Induced Fermentation: Starter Cultures, Scientific Evidence and Actual Trends

From an historical perspective, the management of microbial resources performed for millennia, without any awareness of existence of microscopic organisms, can be summarized as the inoculation of the raw material with a small amount of matrix from a previous successful fermentation [69]. In other terms, this practice foresees the use the microbiota that had demonstrated efficiency the day before as inoculum for the new production. This management has been declined in the different production domains and using different languages, for example: “back-slopping” (sourdough preparation), “sieroinnesto” and “lattoinnesto” (dairy production), “pied de cuve” (enological productions), “inoculum enrichment”. The shifting from this empirical regimen to the modern microbial cultures framework got through the discovery of microbes and the advent of microbiological disciplines, that in several cases started with studies and applications in the field of food microbiology (e.g., Pasteur′s “Études sur le vin”; Emil Christian Hansen′s pure-culturing techniques applied to yeasts for beer production) [69]. The use of starter cultures coincides with the industrialization of fermented productions, satisfying the exigencies of standardization associated with modern large-scale fermentations. In Table 5, we propose some recent definitions reported in the literature.
Various scientific evidence on the comparison of spontaneous versus inoculated fermentation processes (e.g., those reported in Table 6) testified the crucial importance of the starter culture technology to assure food safety worldwide, at all levels of fermented food production: household, traditional, and industrial. This desired affect is achieved by means of different biological activities such as faster acidification activity, domination of the indigenous microflora, reduction of fermentation time and reduction of undesired microbial strains/species and toxic compounds [72,73,74,75] (Table 6).
Certainly, it is mandatory to underline that starter culture technology is not effective per se. The effectiveness of each tailored starter culture on a specific aspect of the inoculated food matrices is a function of the quality of biotechnological solutions conceived to cope with a given specific real problem. In general, considering the safety assurance, it is crucial the assessment of the safety of species/strains and of the quality/purity of biomass preparation [64,85]. More in depth, the scientific demonstration of efficacy regarding the specific safety target is important. For example, Van Ba et al. [86], evaluating several starter cultures and an inoculated control in sausage fermentation, found that not all the starter cultures were effective in lowering the content of different biogenic amines (in comparison with uninoculated control). We also have to remember that starter cultures have not always been found to improve a specific safety target. That was the case of Torrea et al., who found concentrations of nonvolatile amines and phenethylamine in wines from inoculated must were superior to those of the control wine (uninoculated must) [87]. On the other hand, with the concern of strain safety, is interesting to remark on the advances allowed by integrated genome-based assessments of the safety of specific microbial biotypes (e.g., [88,89]).
Generally, the importance of the adoption of a starter culture regimen in order to minimize the risk of food-borne diseases [43] has also been confirmed by means of artificially inoculated pathogens [90].
Starter cultures exploit changes as a function of peculiarities associated with the degree of development of the different countries. In developing countries, the significance of starter cultures is strongly related to the importance of food preservation, yield increasing, and food security [91]. By contrast, in the “Western world”, starter cultures are also designed (i) to pursue personalized nutrition; (ii) to reach new health targets; and (iii) to sustainably increase shelf life, particularly of artisanal, traditional, typical, organic and biodynamic productions [64,92].
Recent social, economic and environmental trends imply a progressive return to the reliance of spontaneous fermentations in the sector of traditional, typical, and artisanal fermented foods (including Geographical Indications), but also considering organic and biodynamic productions. Increasing evidence from scientists and stakeholders from different disciplines/fields counterpoise the use of commercial starter cultures with the exploitation of spontaneous fermentation, preferring this second approach in the management of food fermentations [93]. From this point of view, the return to spontaneous fermentations represents a strategy to restore a content of tradition, typicality and artisanality already loose. The other leading idea is that the use of commercial starter cultures corresponds to a drift from the “natural” manufacture of fermented foods, with a proportion of “synthetic fertilizers” that are connected to “agricultural production” as well as “commercial microbial starter cultures” that are connected to “food fermentations” [92]. For example, winemakers are constantly searching for new techniques to modulate wine style, and the exploitation of indigenous yeasts present in grape must is re-emerging as a commercial option in several wine regions (wines made with indigenous or “wild” yeasts are perceived to be more complex by showing a greater diversity of flavors) [94,95]. It is a concept, shifting to cheese-making practice, well summarized by Piero Sardo, President of the Slow Food Foundation for Biodiversity Onlus: “With industrial starters, cheese made in the mountains will be only slightly different from cheese made in the plains […]. With the industrial packets the music is already written and the cheese-maker just plays it” [64]. All these points of view are indirectly sustained by recent scientific evidence indicating the presence of a so-called “microbial terroir” dimension in association with specific fermented productions liable to influence perceived quality [96,97,98,99,100]. The recent debate on the proposal of microorganisms as a driver of sensory innovation in the gastronomy/artisanal fermented food production fields testify well to the scientific point of view about spontaneous fermentation by autochthonous microorganisms [101,102]. Relying on spontaneous uncontrolled fermentations poses serious challenges for the safety and the quality of fermented foods. In fact, spontaneous fermentation generally increases the risks of the implantation/domination of microbial strains dangerous for the human health [64]. In addition, the risk of spoiling microbial communities in food matrices is considerable.

5. How to Conciliate Fermented Food Safety with Instances of an Enhanced Contribution of Microbes Associated to Spontaneous Fermentation

In this section, we delve into the two main biotechnological solutions that could conjugate (a) safety/quality of fermented foods and (b) the adoption of a microbial regimen in food fermentations compatible with enhanced contribution of microbes associated with spontaneous fermentation: (i) the design of multi-strains starter cultures based on the selection of ecotypes from spontaneous fermentations [15,92,103] (“top-down” solution); and (ii) the application of innovative biotechnologies and microbiological methods to monitor the safety of spontaneous fermentations (“bottom-up” solution).
The design of tailored starter cultures for specific productions in such a way to mimic protechnological microbial diversity associated to spontaneous fermentation is a solution experienced worldwide (few examples in Table 7).
To better understand the dimension of the existing efforts in the design of “tailored” starter cultures for traditional, typical, and artisanal fermented foods (including geographical indications), you can find in Table 8 a non-exhaustive list of scientific works reported in the recent literature on the characterization of microbes associated with Apulian (Southern Italian region) spontaneous fermentations (carried out also to select strains of protechnological interest). Taken together, information reported in Table 7 (international interest in the topic) and Table 8 (regional interest in the topic) testifies well the dimension of a biotechnological latent potential globally characterized as a function of local productions, and that might be exploited to pursue unique sensory quality without compromising the safety of the production.
On the other hand, the opportunity exists to use an integrated approach of combined molecular and microbiological methods to assess the safety of the microbiota associated with spontaneous fermentation. The main example is the development and application of sequence-based molecular technologies (phylobiomics, metagenomics and metatranscriptomics) for examining the diversity and safety of indigenous microbiota associated with traditional fermented foods and beverages [134]. In addition, the combination of culture-independent and culture-dependent analysis might be used to verify the (legislative) safety standard compliance of dominant strains associated with spontaneous fermentation (e.g., adoption of a QPS approach to dominant LAB associated with Grana Padano cheese whey starters, as a proposed approach susceptible to be extended to other types of undefined-strain cultures [135]). In general, all these molecular and physiological approaches might help determine the presence of strains associated with the indigenous microbiota that could pose risks to human health (e.g., presence of genes involved in biogenic amines production and verification of the corresponding phenotype in vivo). Naturally, it is needed to take account of this importance to periodically perform these monitoring activities, considering the possible variability of microbial consortia associated with “inoculum enrichment” practices.

6. Conclusions

Fermented foods and beverages represent a worldwide category of edible products with a prominent significance for human economy, nutrition and health for millennia. A huge diversity of microorganisms has been detected in association with spontaneous fermentations all around the world. In the framework of this heterogeneous microbiota, it is possible to find microbial pathogens and/or strains liable to synthetize toxic by-products such as mycotoxins, ethyl carbamate and biogenic amines. These microbial contaminants can reduce the safety of the corresponding fermented product. Starter culture technology represents a cornerstone in the assurance of quality and the safety of fermentation. However, recent economic, productive and social trends have led to the rediscovery of the potential of spontaneous fermentation in improving the unique quality of fermented products. With this review, we remember the potential risks for human health associated with uncontrolled (uninoculated) food fermentations and we point out how modern microbial biotechnologies offer solutions to conciliate fermented food safety with instances of an enhanced contribution of microbes associated with spontaneous fermentation.

Acknowledgments

This research was supported by the Apulia Region in the framework of the Projects “Sviluppo di approcci microbiologici innovativi per il miglioramento della qualità di vini tipici Regionali—NEWINE (Bando “Ricerca e sperimentazione in Agricoltura”; Project code PRS_042), “Biotecnologie degli alimenti per l’innovazione e la competitività delle principali filiere regionali: estensione della conservabilità e aspetti funzionali—BIOTECA” (Bando “Aiuti a Sostegno Cluster Tecnologici Regionali”; Project code QCBRAJ6) and “Innovazioni di processo e di prodotto nel comparto dei vini spumanti da vitigni autoctoni pugliesi”—IPROVISP (Bando “Aiuti a Sostegno Cluster Tecnologici Regionali”; Project code VJBKVF4). Vittorio Capozzi was supported by Fondo di Sviluppo e Coesione 2007-2013—APQ Ricerca Regione Puglia “Programma regionale a sostegno della specializzazione intelligente e della sostenibilità sociale ed ambientale—FutureInResearch”. Pasquale Russo was supported by a grant of the Apulian Region in the framework of “Peform Tech (Puglia Emerging Food Technology)” project (practice code LPIJ9P2). Carmen Berbegal was supported by “Programa Atracció de Talent VLC-Campus 2015 de la Universitat de València”.

Author Contributions

Vittorio Capozzi, Carmen Berbegal, Pasquale Russo and Giuseppe Spano conceived and designed the different chapters; Vittorio Capozzi, Mariagiovanna Fragasso, Carmen Berbegal, Pasquale Russo and Rossana Romaniello wrote the paper. Giuseppe Spano critically read the paper.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Marco, M.L.; Heeney, D.; Binda, S.; Cifelli, C.J.; Cotter, P.D.; Foligné, B.; Gänzle, M.; Kort, R.; Pasin, G.; Pihlanto, A.; et al. Health benefits of fermented foods: Microbiota and beyond. Curr. Opin. Biotechnol. 2017, 44, 94–102. [Google Scholar] [CrossRef] [PubMed]
  2. Russo, P.; Arena, M.P.; Fiocco, D.; Capozzi, V.; Drider, D.; Spano, G. Lactobacillus plantarum with broad antifungal activity: A promising approach to increase safety and shelf-life of cereal-based products. Int. J. Food Microbiol. 2017, 247, 48–54. [Google Scholar] [CrossRef] [PubMed]
  3. Romano, A.; Capozzi, V.; Spano, G.; Biasioli, F. Proton transfer reaction-mass spectrometry: Online and rapid determination of volatile organic compounds of microbial origin. Appl. Microbiol. Biotechnol. 2015, 99, 3787–3795. [Google Scholar] [CrossRef] [PubMed]
  4. Capozzi, V.; Russo, P.; Dueñas, M.T.; López, P.; Spano, G. Lactic acid bacteria producing B-group vitamins: A great potential for functional cereals products. Appl. Microbiol. Biotechnol. 2012, 96, 1383–1394. [Google Scholar] [CrossRef] [PubMed]
  5. Ercolini, D. High-Throughput Sequencing and Metagenomics: Moving Forward in the Culture-Independent Analysis of Food Microbial Ecology. Appl. Environ. Microbiol. 2013, 79, 3148–3155. [Google Scholar] [CrossRef] [PubMed]
  6. Xu, Y.-J. Foodomics: A novel approach for food microbiology. TrAC Trends Anal. Chem. 2017. [Google Scholar] [CrossRef]
  7. Mozzi, F.; Ortiz, M.E.; Bleckwedel, J.; De Vuyst, L.; Pescuma, M. Metabolomics as a tool for the comprehensive understanding of fermented and functional foods with lactic acid bacteria. Food Res. Int. 2013, 54, 1152–1161. [Google Scholar] [CrossRef]
  8. Pinu, F.R.; Villas-boas, S.G. Rapid Quantification of Major Volatile Metabolites in Fermented Food and Beverages Using Gas Chromatography-Mass Spectrometry. Metabolites 2017, 7, 37. [Google Scholar] [CrossRef] [PubMed]
  9. Capozzi, V.; Yener, S.; Khomenko, I.; Farneti, B.; Cappellin, L.; Gasperi, F.; Scampicchio, M.; Biasioli, F. PTR-ToF-MS Coupled with an Automated Sampling System and Tailored Data Analysis for Food Studies: Bioprocess Monitoring, Screening and Nose-space Analysis. J. Vis. Exp. JoVE 2017. [Google Scholar] [CrossRef] [PubMed]
  10. Rhee, S.J.; Lee, J.-E.; Lee, C.-H. Importance of lactic acid bacteria in Asian fermented foods. Microb. Cell Fact. 2011, 10, S5. [Google Scholar] [CrossRef] [PubMed]
  11. Tamang, J.P.; Kailasapathy, K. Fermented Foods and Beverages of the World; CRC Press: Boca Raton, FL, USA, 2010; ISBN 978-1-4200-9496-1. [Google Scholar]
  12. Odunfa, S.A.; Oyewole, O.B. African fermented foods. In Microbiology of Fermented Foods; Springer: Boston, MA, USA, 1998; pp. 713–752. ISBN 978-1-4613-7990-4. [Google Scholar]
  13. Penna, A.L.B.; Nero, L.A.; Todorov, S.D. Fermented Foods of Latin America: From Traditional Knowledge to Innovative Applications; CRC Press: Boca Raton, FL, USA, 2017; ISBN 978-1-315-35243-5. [Google Scholar]
  14. Campbell-Platt, G. Fermented foods—A world perspective. Food Res. Int. 1994, 27, 253–257. [Google Scholar] [CrossRef]
  15. Capozzi, V.; Russo, P.; Spano, G. Microbial information regimen in EU geographical indications. World Pat. Inf. 2012, 34, 229–231. [Google Scholar] [CrossRef]
  16. Tamang, J.P.; Watanabe, K.; Holzapfel, W.H. Review: Diversity of Microorganisms in Global Fermented Foods and Beverages. Front. Microbiol. 2016, 7. [Google Scholar] [CrossRef] [PubMed]
  17. Chilton, S.N.; Burton, J.P.; Reid, G. Inclusion of Fermented Foods in Food Guides around the World. Nutrients 2015, 7, 390–404. [Google Scholar] [CrossRef] [PubMed]
  18. Bell, V.; Ferrão, J.; Fernandes, T. Nutritional Guidelines and Fermented Food Frameworks. Foods 2017, 6, 65. [Google Scholar] [CrossRef] [PubMed]
  19. Paul Ross, R.; Morgan, S.; Hill, C. Preservation and fermentation: Past, present and future. Int. J. Food Microbiol. 2002, 79, 3–16. [Google Scholar] [CrossRef]
  20. Petruzzi, L.; Capozzi, V.; Berbegal, C.; Corbo, M.R.; Bevilacqua, A.; Spano, G.; Sinigaglia, M. Microbial Resources and Enological Significance: Opportunities and Benefits. Front. Microbiol. 2017, 8. [Google Scholar] [CrossRef] [PubMed]
  21. Greppi, A.; Rantsiou, K.; Padonou, W.; Hounhouigan, J.; Jespersen, L.; Jakobsen, M.; Cocolin, L. Determination of yeast diversity in ogi, mawè, gowé and tchoukoutou by using culture-dependent and -independent methods. Int. J. Food Microbiol. 2013, 165, 84–88. [Google Scholar] [CrossRef] [PubMed]
  22. Russo, P.; Capozzi, V.; Spano, G.; Corbo, M.R.; Sinigaglia, M.; Bevilacqua, A. Metabolites of Microbial Origin with an Impact on Health: Ochratoxin A and Biogenic Amines. Front. Microbiol. 2016, 7. [Google Scholar] [CrossRef] [PubMed]
  23. Benozzi, E.; Romano, A.; Capozzi, V.; Makhoul, S.; Cappellin, L.; Khomenko, I.; Aprea, E.; Scampicchio, M.; Spano, G.; Märk, T.D.; et al. Monitoring of lactic fermentation driven by different starter cultures via direct injection mass spectrometric analysis of flavour-related volatile compounds. Food Res. Int. 2015, 76, 682–688. [Google Scholar] [CrossRef] [PubMed]
  24. Capozzi, V.; Menga, V.; Digesu, A.M.; De Vita, P.; van Sinderen, D.; Cattivelli, L.; Fares, C.; Spano, G. Biotechnological production of vitamin B2-enriched bread and pasta. J. Agric. Food Chem. 2011, 59, 8013–8020. [Google Scholar] [CrossRef] [PubMed]
  25. Bove, P.; Russo, P.; Capozzi, V.; Gallone, A.; Spano, G.; Fiocco, D. Lactobacillus plantarum passage through an oro-gastro-intestinal tract simulator: Carrier matrix effect and transcriptional analysis of genes associated to stress and probiosis. Microbiol. Res. 2013, 168, 351–359. [Google Scholar] [CrossRef] [PubMed]
  26. De Bellis, P.; Tristezza, M.; Haidukowski, M.; Fanelli, F.; Sisto, A.; Mulè, G.; Grieco, F. Biodegradation of Ochratoxin A by Bacterial Strains Isolated from Vineyard Soils. Toxins 2015, 7, 5079–5093. [Google Scholar] [CrossRef] [PubMed]
  27. Capozzi, V.; Russo, P.; Fragasso, M.; De Vita, P.; Fiocco, D.; Spano, G. Biotechnology and Pasta-Making: Lactic Acid Bacteria as a New Driver of Innovation. Front. Microbiol. 2012, 3. [Google Scholar] [CrossRef] [PubMed]
  28. Klayraung, S.; Viernstein, H.; Sirithunyalug, J.; Okonogi, S. Probiotic Properties of Lactobacilli Isolated from Thai Traditional Food. Sci. Pharm. 2008, 76, 485–504. [Google Scholar] [CrossRef]
  29. Russo, P.; de Chiara, M.L.V.; Capozzi, V.; Arena, M.P.; Amodio, M.L.; Rascón, A.; Dueñas, M.T.; López, P.; Spano, G. Lactobacillus plantarum strains for multifunctional oat-based foods. LWT Food Sci. Technol. 2016, 68, 288–294. [Google Scholar] [CrossRef]
  30. Müller, R.; Rappert, S. Pyrazines: Occurrence, formation and biodegradation. Appl. Microbiol. Biotechnol. 2010, 85, 1315–1320. [Google Scholar] [CrossRef] [PubMed]
  31. Anraku, K.; Nonaka, K.; Yamaga, T.; Yamamoto, T.; Shin, M.-C.; Wakita, M.; Hamamoto, A.; Akaike, N. Removal of Toxin (Tetrodotoxin) from Puffer Ovary by Traditional Fermentation. Toxins 2013, 5, 193–202. [Google Scholar] [CrossRef] [PubMed]
  32. Capozzi, V.; Russo, P.; Ladero, V.; Fernandez, M.; Fiocco, D.; Alvarez, M.A.; Grieco, F.; Spano, G. Biogenic Amines Degradation by Lactobacillus plantarum: Toward a Potential Application in Wine. Front. Microbiol. 2012, 3. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  33. Magalhães da Veiga Moreira, I.; de Figueiredo Vilela, L.; da Cruz Pedroso Miguel, M.G.; Santos, C.; Lima, N.; Freitas Schwan, R. Impact of a Microbial Cocktail Used as a Starter Culture on Cocoa Fermentation and Chocolate Flavor. Molecules 2017, 22, 766. [Google Scholar] [CrossRef] [PubMed]
  34. Capozzi, V.; Makhoul, S.; Aprea, E.; Romano, A.; Cappellin, L.; Sanchez Jimena, A.; Spano, G.; Gasperi, F.; Scampicchio, M.; Biasioli, F. PTR-MS Characterization of VOCs Associated with Commercial Aromatic Bakery Yeasts of Wine and Beer Origin. Molecules 2016, 21, 483. [Google Scholar] [CrossRef] [PubMed]
  35. Fessard, A.; Kapoor, A.; Patche, J.; Assemat, S.; Hoarau, M.; Bourdon, E.; Bahorun, T.; Remize, F. Lactic Fermentation as an Efficient Tool to Enhance the Antioxidant Activity of Tropical Fruit Juices and Teas. Microorganisms 2017, 5, 23. [Google Scholar] [CrossRef] [PubMed]
  36. Hemalatha, S.; Platel, K.; Srinivasan, K. Influence of germination and fermentation on bioaccessibility of zinc and iron from food grains. Eur. J. Clin. Nutr. 2007, 61, 342–348. [Google Scholar] [CrossRef] [PubMed]
  37. Fawale, O.S.; Gbadamosi, S.O.; Ige, M.M.; Kadiri, O. Effects of cooking and fermentation on the chemical composition, functional, and antinutritional properties of kariya (Hildergardia barteri) seeds. Food Sci. Nutr. 2017. [Google Scholar] [CrossRef]
  38. Morelli, L. Yogurt, living cultures, and gut health. Am. J. Clin. Nutr. 2014, 99, 1248S–1250S. [Google Scholar] [CrossRef] [PubMed]
  39. Arena, M.P.; Caggianiello, G.; Russo, P.; Albenzio, M.; Massa, S.; Fiocco, D.; Capozzi, V.; Spano, G. Functional Starters for Functional Yogurt. Foods 2015, 4, 15–33. [Google Scholar] [CrossRef] [PubMed]
  40. Pérez-Ramos, A.; Mohedano, M.L.; López, P.; Spano, G.; Fiocco, D.; Russo, P.; Capozzi, V. In Situ β-Glucan Fortification of Cereal-Based Matrices by Pediococcus parvulus 2.6: Technological Aspects and Prebiotic Potential. Int. J. Mol. Sci. 2017, 18, 1588. [Google Scholar] [CrossRef] [PubMed]
  41. Nout, M.J.R. Fermented foods and food safety. Food Res. Int. 1994, 27, 291–298. [Google Scholar] [CrossRef]
  42. Mena, C.; Almeida, G.; Carneiro, L.; Teixeira, P.; Hogg, T.; Gibbs, P.A. Incidence of Listeria monocytogenes in different food products commercialized in Portugal. Food Microbiol. 2004, 21, 213–216. [Google Scholar] [CrossRef]
  43. Gadaga, T.H.; Nyanga, L.K.; Mutukumira, A.N. The Occurrence, Growth and Control of Pathogens in African Fermented Foods. Afr. J. Food Agric. Nutr. Dev. 2007, 4. [Google Scholar] [CrossRef]
  44. Ostry, V.; Malir, F.; Ruprich, J. Producers and Important Dietary Sources of Ochratoxin A and Citrinin. Toxins 2013, 5, 1574–1586. [Google Scholar] [CrossRef] [PubMed]
  45. Suzzi, G.; Gardini, F. Biogenic amines in dry fermented sausages: A review. Int. J. Food Microbiol. 2003, 88, 41–54. [Google Scholar] [CrossRef]
  46. Park, K.M.; Kim, H.J.; Jeong, M.C.; Koo, M. Occurrence of Toxigenic Bacillus cereus and Bacillus thuringiensis in Doenjang, a Korean Fermented Soybean Paste. J. Food Prot. 2016, 79, 605–612. [Google Scholar] [CrossRef] [PubMed]
  47. Centers for Disease Control and Prevention (CDC). Escherichia coli O157:H7 outbreak linked to commercially distributed dry-cured salami—Washington and California, 1994. Morb. Mortal. Wkly. Rep. 1995, 44, 157–160. [Google Scholar]
  48. Besser, R.E.; Lett, S.M.; Weber, J.T.; Doyle, M.P.; Barrett, T.J.; Wells, J.G.; Griffin, P.M. An Outbreak of Diarrhea and Hemolytic Uremic Syndrome from Escherichia coli O157:H7 in Fresh-Pressed Apple Cider. JAMA 1993, 269, 2217–2220. [Google Scholar] [CrossRef] [PubMed]
  49. Adedeji, B.S.; Ezeokoli, O.T.; Ezekiel, C.N.; Obadina, A.O.; Somorin, Y.M.; Sulyok, M.; Adeleke, R.A.; Warth, B.; Nwangburuka, C.C.; Omemu, A.M.; et al. Bacterial species and mycotoxin contamination associated with locust bean, melon and their fermented products in south-western Nigeria. Int. J. Food Microbiol. 2017, 258, 73–80. [Google Scholar] [CrossRef] [PubMed]
  50. Gilsdorf, A.; Jansen, A.; Alpers, K.; Dieckmann, H.; van Treeck, U.; Hauri, A.M.; Fell, G.; Littmann, M.; Rautenberg, P.; Prager, R.; et al. A nationwide outbreak of Salmonella Bovismorbificans PT24, Germany, December 2004-March 2005. Eur. Surveill. 2005, 10, E050324.1. [Google Scholar]
  51. Dano, S.D.; Manda, P.; Dembélé, A.; Kouassi Abla, A.M.-J.; Bibaud, J.H.; Gouet, J.Z.; Ze Maria Sika, C.B. Influence of Fermentation and Drying Materials on the Contamination of Cocoa Beans by Ochratoxin A. Toxins 2013, 5, 2310–2323. [Google Scholar] [CrossRef] [PubMed]
  52. Beneduce, L.; Romano, A.; Capozzi, V.; Lucas, P.; Barnavon, L.; Bach, B.; Vuchot, P.; Grieco, F.; Spano, G. Biogenic amine in wines. Ann. Microbiol. 2010, 60, 573–578. [Google Scholar] [CrossRef]
  53. Rabie, M.A.; Elsaidy, S.; el-Badawy, A.-A.; Siliha, H.; Malcata, F.X. Biogenic amine contents in selected Egyptian fermented foods as determined by ion-exchange chromatography. J. Food Prot. 2011, 74, 681–685. [Google Scholar] [CrossRef] [PubMed]
  54. Uthurry, C.A.; Lepe, J.A.S.; Lombardero, J.; García Del Hierro, J.R. Ethyl carbamate production by selected yeasts and lactic acid bacteria in red wine. Food Chem. 2006, 94, 262–270. [Google Scholar] [CrossRef]
  55. Da Rocha, M.E.B.; Freire, F.; Maia, F.; Guedes, M.I.F.; Rondina, D. Mycotoxins and their effects on human and animal health. Food Control 2014, 36, 159–165. [Google Scholar] [CrossRef]
  56. Milićević, D.R.; Škrinjar, M.; Baltić, T. Real and Perceived Risks for Mycotoxin Contamination in Foods and Feeds: Challenges for Food Safety Control. Toxins 2010, 2, 572–592. [Google Scholar] [CrossRef] [PubMed]
  57. Ladero, V.; Fernández, M.; Calles-Enríquez, M.; Sánchez-Llana, E.; Cañedo, E.; Martín, M.C.; Alvarez, M.A. Is the production of the biogenic amines tyramine and putrescine a species-level trait in enterococci? Food Microbiol. 2012, 30, 132–138. [Google Scholar] [CrossRef] [PubMed]
  58. Capozzi, V.; Ladero, V.; Beneduce, L.; Fernández, M.; Alvarez, M.A.; Benoit, B.; Laurent, B.; Grieco, F.; Spano, G. Isolation and characterization of tyramine-producing Enterococcus faecium strains from red wine. Food Microbiol. 2011, 28, 434–439. [Google Scholar] [CrossRef] [PubMed]
  59. Tristezza, M.; Vetrano, C.; Bleve, G.; Spano, G.; Capozzi, V.; Logrieco, A.; Mita, G.; Grieco, F. Biodiversity and safety aspects of yeast strains characterized from vineyards and spontaneous fermentations in the Apulia Region, Italy. Food Microbiol. 2013, 36, 335–342. [Google Scholar] [CrossRef] [PubMed]
  60. Spano, G.; Russo, P.; Lonvaud-Funel, A.; Lucas, P.; Alexandre, H.; Grandvalet, C.; Coton, E.; Coton, M.; Barnavon, L.; Bach, B.; et al. Biogenic amines in fermented foods. Eur. J. Clin. Nutr. 2010, 64 (Suppl. S3), S95–S100. [Google Scholar] [CrossRef] [PubMed]
  61. Straub, B.W.; Kicherer, M.; Schilcher, S.M.; Hammes, W.P. The formation of biogenic amines by fermentation organisms. Z. Lebensm. Unters. Forsch. 1995, 201, 79–82. [Google Scholar] [CrossRef] [PubMed]
  62. Azevedo, Z.; Couto, J.A.; Hogg, T. Citrulline as the main precursor of ethyl carbamate in model fortified wines inoculated with Lactobacillus hilgardii: A marker of the levels in a spoiled fortified wine. Lett. Appl. Microbiol. 2002, 34, 32–36. [Google Scholar] [CrossRef] [PubMed]
  63. Arena, M.E.; Manca de Nadra, M.C. Biogenic amine production by Lactobacillus. J. Appl. Microbiol. 2001, 90, 158–162. [Google Scholar] [CrossRef] [PubMed]
  64. Russo, P.; Spano, G.; Capozzi, V. Safety evaluation of starter cultures. In Starter Cultures in Food Production; Speranza, B., Bevilacqua, A., Corbo, M.R., Sinigaglia, M., Eds.; John Wiley & Sons, Ltd.: Hoboken, NJ, USA, 2017; pp. 101–128. ISBN 978-1-118-93379-4. [Google Scholar]
  65. Bourdichon, F.; Casaregola, S.; Farrokh, C.; Frisvad, J.C.; Gerds, M.L.; Hammes, W.P.; Harnett, J.; Huys, G.; Laulund, S.; Ouwehand, A.; et al. Food fermentations: Microorganisms with technological beneficial use. Int. J. Food Microbiol. 2012, 154, 87–97. [Google Scholar] [CrossRef] [PubMed]
  66. Rossi, F.; Rizzotti, L.; Felis, G.E.; Torriani, S. Horizontal gene transfer among microorganisms in food: Current knowledge and future perspectives. Food Microbiol. 2014, 42, 232–243. [Google Scholar] [CrossRef] [PubMed]
  67. Gueimonde, M.; Sánchez, B.; G de Los Reyes-Gavilán, C.; Margolles, A. Antibiotic resistance in probiotic bacteria. Front. Microbiol. 2013, 4, 202. [Google Scholar] [CrossRef] [PubMed]
  68. Bernardeau, M.; Vernoux, J.P.; Henri-Dubernet, S.; Guéguen, M. Safety assessment of dairy microorganisms: The Lactobacillus genus. Int. J. Food Microbiol. 2008, 126, 278–285. [Google Scholar] [CrossRef] [PubMed]
  69. Brandt, M.J. Starter cultures for cereal based foods. Food Microbiol. 2014, 37, 41–43. [Google Scholar] [CrossRef] [PubMed]
  70. Vogel, R.F.; Hammes, W.P.; Habermeyer, M.; Engel, K.-H.; Knorr, D.; Eisenbrand, G. Microbial food cultures—Opinion of the Senate Commission on Food Safety (SKLM) of the German Research Foundation (DFG). Mol. Nutr. Food Res. 2011, 55, 654–662. [Google Scholar] [CrossRef] [PubMed]
  71. Hansen, E.B. STARTER CULTURES/Uses in the Food Industry. In Encyclopedia of Food Microbiology (Second Edition); Batt, C.A., Tortorello, M.L., Eds.; Academic Press: Oxford, UK, 2014; pp. 529–534. ISBN 978-0-12-384733-1. [Google Scholar]
  72. De Angelis, M.; Campanella, D.; Cosmai, L.; Summo, C.; Rizzello, C.G.; Caponio, F. Microbiota and metabolome of un-started and started Greek-type fermentation of Bella di Cerignola table olives. Food Microbiol. 2015, 52, 18–30. [Google Scholar] [CrossRef] [PubMed]
  73. Shukla, S.; Park, H.-K.; Lee, J.-S.; Kim, J.-K.; Kim, M. Reduction of biogenic amines and aflatoxins in Doenjang samples fermented with various Meju as starter cultures. Food Control 2014, 42, 181–187. [Google Scholar] [CrossRef]
  74. Chen, X.; Li, J.; Zhou, T.; Li, J.; Yang, J.; Chen, W.; Xiong, Y.L. Two efficient nitrite-reducing Lactobacillus strains isolated from traditional fermented pork (Nanx Wudl) as competitive starter cultures for Chinese fermented dry sausage. Meat Sci. 2016, 121, 302–309. [Google Scholar] [CrossRef] [PubMed]
  75. Sanchart, C.; Rattanaporn, O.; Haltrich, D.; Phukpattaranont, P.; Maneerat, S. Enhancement of gamma-aminobutyric acid (GABA) levels using an autochthonous Lactobacillus futsaii CS3 as starter culture in Thai fermented shrimp (Kung-Som). World J. Microbiol. Biotechnol. 2017, 33, 152. [Google Scholar] [CrossRef] [PubMed]
  76. Sabatini, N.; Mucciarella, M.R.; Marsilio, V. Volatile compounds in uninoculated and inoculated table olives with Lactobacillus plantarum (Olea europaea L., cv. Moresca and Kalamata). LWT Food Sci. Technol. 2008, 41, 2017–2022. [Google Scholar] [CrossRef]
  77. Comunian, R.; Ferrocino, I.; Paba, A.; Daga, E.; Campus, M.; Di Salvo, R.; Cauli, E.; Piras, F.; Zurru, R.; Cocolin, L. Evolution of microbiota during spontaneous and inoculated Tonda di Cagliari table olives fermentation and impact on sensory characteristics. LWT Food Sci. Technol. 2017, 84, 64–72. [Google Scholar] [CrossRef]
  78. Yang, J.; Ji, Y.; Park, H.; Lee, J.; Park, S.; Yeo, S.; Shin, H.; Holzapfel, W.H. Selection of functional lactic acid bacteria as starter cultures for the fermentation of Korean leek (Allium tuberosum Rottler ex Sprengel.). Int. J. Food Microbiol. 2014, 191, 164–171. [Google Scholar] [CrossRef] [PubMed]
  79. Lorenzo, J.M.; Gómez, M.; Fonseca, S. Effect of commercial starter cultures on physicochemical characteristics, microbial counts and free fatty acid composition of dry-cured foal sausage. Food Control 2014, 46, 382–389. [Google Scholar] [CrossRef]
  80. Ciuciu Simion, A.M.; Vizireanu, C.; Alexe, P.; Franco, I.; Carballo, J. Effect of the use of selected starter cultures on some quality, safety and sensorial properties of Dacia sausage, a traditional Romanian dry-sausage variety. Food Control 2014, 35, 123–131. [Google Scholar] [CrossRef]
  81. Xie, C.; Wang, H.-H.; Nie, X.-K.; Chen, L.; Deng, S.-L.; Xu, X.-L. Reduction of biogenic amine concentration in fermented sausage by selected starter cultures. CyTA J. Food 2015, 13, 491–497. [Google Scholar] [CrossRef]
  82. Tosukhowong, A.; Visessanguan, W.; Pumpuang, L.; Tepkasikul, P.; Panya, A.; Valyasevi, R. Biogenic amine formation in Nham, a Thai fermented sausage, and the reduction by commercial starter culture, Lactobacillus plantarum BCC 9546. Food Chem. 2011, 129, 846–853. [Google Scholar] [CrossRef] [PubMed]
  83. Lonvaud-Funel, A. Biogenic amines in wines: Role of lactic acid bacteria. FEMS Microbiol. Lett. 2001, 199, 9–13. [Google Scholar] [CrossRef] [PubMed]
  84. Patrignani, F.; Ndagijimana, M.; Belletti, N.; Gardini, F.; Vernocchi, P.; Lanciotti, R. Biogenic Amines and Ethyl Carbamate in Primitivo Wine: Survey of Their Concentrations in Commercial Products and Relationship with the Use of Malolactic Starter. J. Food Prot. 2012, 75, 591–596. [Google Scholar] [CrossRef] [PubMed]
  85. Costantini, A.; Vaudano, E.; Del Prete, V.; Danei, M.; Garcia-Moruno, E. Biogenic amine production by contaminating bacteria found in starter preparations used in winemaking. J. Agric. Food Chem. 2009, 57, 10664–10669. [Google Scholar] [CrossRef] [PubMed]
  86. Van Ba, H.; Seo, H.-W.; Kim, J.-H.; Cho, S.-H.; Kim, Y.-S.; Ham, J.-S.; Park, B.-Y.; Kim, H.-W.; Kim, T.-B.; Seong, P.-N. The effects of starter culture types on the technological quality, lipid oxidation and biogenic amines in fermented sausages. LWT Food Sci. Technol. 2016, 74, 191–198. [Google Scholar] [CrossRef]
  87. Torrea, D.; Ancín, C. Content of Biogenic Amines in a Chardonnay Wine Obtained through Spontaneous and Inoculated Fermentations. J. Agric. Food Chem. 2002, 50, 4895–4899. [Google Scholar] [CrossRef] [PubMed]
  88. Salvetti, E.; Orrù, L.; Capozzi, V.; Martina, A.; Lamontanara, A.; Keller, D.; Cash, H.; Felis, G.E.; Cattivelli, L.; Torriani, S.; et al. Integrate genome-based assessment of safety for probiotic strains: Bacillus coagulans GBI-30, 6086 as a case study. Appl. Microbiol. Biotechnol. 2016, 100, 4595–4605. [Google Scholar] [CrossRef] [PubMed]
  89. Orrù, L.; Salvetti, E.; Cattivelli, L.; Lamontanara, A.; Michelotti, V.; Capozzi, V.; Spano, G.; Keller, D.; Cash, H.; Martina, A.; et al. Draft Genome Sequence of Bacillus coagulans GBI-30, 6086, a Widely Used Spore-Forming Probiotic Strain. Genome Announc. 2014, 2. [Google Scholar] [CrossRef] [PubMed]
  90. Panagou, E.Z.; Tassou, C.C.; Vamvakoula, P.; Saravanos, E.K.A.; Nychas, G.-J.E. Survival of Bacillus cereus vegetative cells during Spanish-style fermentation of conservolea green olives. J. Food Prot. 2008, 71, 1393–1400. [Google Scholar] [CrossRef] [PubMed]
  91. Holzapfel, W.H. Appropriate starter culture technologies for small-scale fermentation in developing countries. Int. J. Food Microbiol. 2002, 75, 197–212. [Google Scholar] [CrossRef]
  92. Spano, G.; Capozzi, V. Food Microbial Biodiversity and “Microbes of Protected Origin”. Front. Microbiol. 2011, 2. [Google Scholar] [CrossRef]
  93. Capozzi, V.; Spano, G.; Fiocco, D. Transdisciplinarity and Microbiology Education. J. Microbiol. Biol. Educ. JMBE 2012, 13, 70–73. [Google Scholar] [CrossRef] [PubMed]
  94. Varela, C.; Siebert, T.; Cozzolino, D.; Rose, L.; Mclean, H.; Henschke, P.A. Discovering a chemical basis for differentiating wines made by fermentation with “wild” indigenous and inoculated yeasts: Role of yeast volatile compounds. Aust. J. Grape Wine Res. 2009, 15, 238–248. [Google Scholar] [CrossRef]
  95. Liu, P.-T.; Lu, L.; Duan, C.-Q.; Yan, G.-L. The contribution of indigenous non-Saccharomyces wine yeast to improved aromatic quality of Cabernet Sauvignon wines by spontaneous fermentation. LWT Food Sci. Technol. 2016, 71, 356–363. [Google Scholar] [CrossRef]
  96. Bokulich, N.A.; Thorngate, J.H.; Richardson, P.M.; Mills, D.A. Microbial biogeography of wine grapes is conditioned by cultivar, vintage, and climate. Proc. Natl. Acad. Sci. USA 2014, 111, E139–E148. [Google Scholar] [CrossRef] [PubMed]
  97. Capozzi, V.; Garofalo, C.; Chiriatti, M.A.; Grieco, F.; Spano, G. Microbial terroir and food innovation: The case of yeast biodiversity in wine. Microbiol. Res. 2015, 181, 75–83. [Google Scholar] [CrossRef] [PubMed]
  98. Knight, S.; Klaere, S.; Fedrizzi, B.; Goddard, M.R. Regional microbial signatures positively correlate with differential wine phenotypes: Evidence for a microbial aspect to terroir. Sci. Rep. 2015, 5. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  99. Bokulich, N.A.; Collins, T.S.; Masarweh, C.; Allen, G.; Heymann, H.; Ebeler, S.E.; Mills, D.A. Associations among Wine Grape Microbiome, Metabolome, and Fermentation Behavior Suggest Microbial Contribution to Regional Wine Characteristics. mBio 2016, 7, e00631-16. [Google Scholar] [CrossRef] [PubMed]
  100. Garofalo, C.; Arena, M.P.; Laddomada, B.; Cappello, M.S.; Bleve, G.; Grieco, F.; Beneduce, L.; Berbegal, C.; Spano, G.; Capozzi, V. Starter Cultures for Sparkling Wine. Fermentation 2016, 2, 21. [Google Scholar] [CrossRef]
  101. Johnson, A.J. Artisanal food microbiology. Nat. Microbiol. 2016, 1, nmicrobiol201639. [Google Scholar] [CrossRef] [PubMed]
  102. Cocolin, L.; Gobbetti, M.; Neviani, E.; Daffonchio, D. Ensuring safety in artisanal food microbiology. Nat. Microbiol. 2016, 1, nmicrobiol2016171. [Google Scholar] [CrossRef] [PubMed]
  103. Maqueda, M.; Pérez-Nevado, F.; Regodón, J.A.; Zamora, E.; Alvarez, M.L.; Rebollo, J.E.; Ramírez, M. A low-cost procedure for production of fresh autochthonous wine yeast. J. Ind. Microbiol. Biotechnol. 2011, 38, 459–469. [Google Scholar] [CrossRef] [PubMed]
  104. Okeke, C.A.; Ezekiel, C.N.; Nwangburuka, C.C.; Sulyok, M.; Ezeamagu, C.O.; Adeleke, R.A.; Dike, S.K.; Krska, R. Bacterial Diversity and Mycotoxin Reduction During Maize Fermentation (Steeping) for Ogi Production. Front. Microbiol. 2015, 6. [Google Scholar] [CrossRef] [PubMed]
  105. Romi, W.; Ahmed, G.; Jeyaram, K. Three-phase succession of autochthonous lactic acid bacteria to reach a stable ecosystem within 7 days of natural bamboo shoot fermentation as revealed by different molecular approaches. Mol. Ecol. 2015, 24, 3372–3389. [Google Scholar] [CrossRef] [PubMed]
  106. Feng, Z.; Gao, W.; Ren, D.; Chen, X.; Li, J. Evaluation of bacterial flora during the ripening of Kedong sufu, a typical Chinese traditional bacteria-fermented soybean product. J. Sci. Food Agric. 2013, 93, 1471–1478. [Google Scholar] [CrossRef] [PubMed]
  107. Huch Née Kostinek, M.; Hanak, A.; Specht, I.; Dortu, C.M.; Thonart, P.; Mbugua, S.; Holzapfel, W.H.; Hertel, C.; Franz, C.M.A.P. Use of Lactobacillus strains to start cassava fermentations for Gari production. Int. J. Food Microbiol. 2008, 128, 258–267. [Google Scholar] [CrossRef] [PubMed]
  108. Motato, K.E.; Milani, C.; Ventura, M.; Valencia, F.E.; Ruas-Madiedo, P.; Delgado, S. Bacterial diversity of the Colombian fermented milk “Suero Costeño” assessed by culturing and high-throughput sequencing and DGGE analysis of 16S rRNA gene amplicons. Food Microbiol. 2017, 68, 129–136. [Google Scholar] [CrossRef] [PubMed]
  109. Casquete, R.; Benito, M.J.; Martín, A.; Ruiz-Moyano, S.; Hernández, A.; Córdoba, M.G. Effect of autochthonous starter cultures in the production of “salchichón”, a traditional Iberian dry-fermented sausage, with different ripening processes. LWT Food Sci. Technol. 2011, 44, 1562–1571. [Google Scholar] [CrossRef]
  110. Guan, L.; Cho, K.H.; Lee, J.-H. Analysis of the cultivable bacterial community in jeotgal, a Korean salted and fermented seafood, and identification of its dominant bacteria. Food Microbiol. 2011, 28, 101–113. [Google Scholar] [CrossRef] [PubMed]
  111. Gullo, M.; Giudici, P. Acetic acid bacteria in traditional balsamic vinegar: Phenotypic traits relevant for starter cultures selection. Int. J. Food Microbiol. 2008, 125, 46–53. [Google Scholar] [CrossRef] [PubMed]
  112. Cordero-Bueso, G.; Esteve-Zarzoso, B.; Gil-Díaz, M.; García, M.; Cabellos, J.M.; Arroyo, T. Improvement of Malvar Wine Quality by Use of Locally-Selected Saccharomyces cerevisiae Strains. Fermentation 2016, 2, 7. [Google Scholar] [CrossRef]
  113. Ricciardi, A.; Parente, E.; Piraino, P.; Paraggio, M.; Romano, P. Phenotypic characterization of lactic acid bacteria from sourdoughs for Altamura bread produced in Apulia (Southern Italy). Int. J. Food Microbiol. 2005, 98, 63–72. [Google Scholar] [CrossRef] [PubMed]
  114. Aquilanti, L.; Zannini, E.; Zocchetti, A.; Osimani, A.; Clementi, F. Polyphasic characterization of indigenous lactobacilli and lactococci from PDO Canestrato Pugliese cheese. LWT Food Sci. Technol. 2007, 40, 1146–1155. [Google Scholar] [CrossRef]
  115. De Candia, S.; De Angelis, M.; Dunlea, E.; Minervini, F.; McSweeney, P.L.H.; Faccia, M.; Gobbetti, M. Molecular identification and typing of natural whey starter cultures and microbiological and compositional properties of related traditional Mozzarella cheeses. Int. J. Food Microbiol. 2007, 119, 182–191. [Google Scholar] [CrossRef] [PubMed]
  116. De Angelis, M.; de Candia, S.; Calasso, M.P.; Faccia, M.; Guinee, T.P.; Simonetti, M.C.; Gobbetti, M. Selection and use of autochthonous multiple strain cultures for the manufacture of high-moisture traditional Mozzarella cheese. Int. J. Food Microbiol. 2008, 125, 123–132. [Google Scholar] [CrossRef] [PubMed]
  117. Bevilacqua, A.; Perricone, M.; Cannarsi, M.; Corbo, M.R.; Sinigaglia, M. Technological and spoiling characteristics of the yeast microflora isolated from Bella Di Cerignola table olives. Int. J. Food Sci. Technol. 2009, 44, 2198–2207. [Google Scholar] [CrossRef]
  118. 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]
  119. Bevilacqua, A.; Altieri, C.; Corbo, M.R.; Sinigaglia, M.; Ouoba, L.I.I. Characterization of Lactic Acid Bacteria Isolated from Italian Bella di Cerignola Table Olives: Selection of Potential Multifunctional Starter Cultures. J. Food Sci. 2010, 75, M536–M544. [Google Scholar] [CrossRef] [PubMed]
  120. Speranza, B.; Bevilacqua, A.; Corbo, M.R.; Altieri, C.; Sinigaglia, M. Selection of autochthonous strains as promising starter cultures for Fior di Latte, a traditional cheese of southern Italy. J. Sci. Food Agric. 2015, 95, 88–97. [Google Scholar] [CrossRef] [PubMed]
  121. Bleve, G.; Tufariello, M.; Durante, M.; Perbellini, E.; Ramires, F.A.; Grieco, F.; Cappello, M.S.; De Domenico, S.; Mita, G.; Tasioula-Margari, M.; et al. Physico-chemical and microbiological characterization of spontaneous fermentation of Cellina di Nardò and Leccino table olives. Front. Microbiol. 2014, 5. [Google Scholar] [CrossRef] [PubMed]
  122. Garofalo, C.; El Khoury, M.; Lucas, P.; Bely, M.; Russo, P.; Spano, G.; Capozzi, V. Autochthonous starter cultures and indigenous grape variety for regional wine production. J. Appl. Microbiol. 2015, 118, 1395–1408. [Google Scholar] [CrossRef] [PubMed]
  123. Tristezza, M.; Tufariello, M.; Capozzi, V.; Spano, G.; Mita, G.; Grieco, F. The Oenological Potential of Hanseniaspora uvarum in Simultaneous and Sequential Co-fermentation with Saccharomyces cerevisiae for Industrial Wine Production. Front. Microbiol. 2016, 7. [Google Scholar] [CrossRef] [PubMed]
  124. Lamontanara, A.; Orrù, L.; Cattivelli, L.; Russo, P.; Spano, G.; Capozzi, V. Genome Sequence of Oenococcus oeni OM27, the First Fully Assembled Genome of a Strain Isolated from an Italian Wine. Genome Announc. 2014, 2. [Google Scholar] [CrossRef] [PubMed]
  125. Capozzi, V.; Russo, P.; Lamontanara, A.; Orrù, L.; Cattivelli, L.; Spano, G. Genome Sequences of Five Oenococcus oeni Strains Isolated from Nero Di Troia Wine from the Same Terroir in Apulia, Southern Italy. Genome Announc. 2014, 2. [Google Scholar] [CrossRef] [PubMed]
  126. Lamontanara, A.; Caggianiello, G.; Orrù, L.; Capozzi, V.; Michelotti, V.; Bayjanov, J.R.; Renckens, B.; van Hijum, S.A.F.T.; Cattivelli, L.; Spano, G. Draft Genome Sequence of Lactobacillus plantarum Lp90 Isolated from Wine. Genome Announc. 2015, 3, e00097-15. [Google Scholar] [CrossRef] [PubMed]
  127. Garofalo, C.; Russo, P.; Beneduce, L.; Massa, S.; Spano, G.; Capozzi, V. Non-Saccharomyces biodiversity in wine and the ‘microbial terroir’: A survey on Nero di Troia wine from the Apulian region, Italy. Ann. Microbiol. 2016, 66, 143–150. [Google Scholar] [CrossRef]
  128. Tristezza, M.; di Feo, L.; Tufariello, M.; Grieco, F.; Capozzi, V.; Spano, G.; Mita, G.; Grieco, F. Simultaneous inoculation of yeasts and lactic acid bacteria: Effects on fermentation dynamics and chemical composition of Negroamaro wine. LWT Food Sci. Technol. 2016, 66, 406–412. [Google Scholar] [CrossRef]
  129. Garofalo, C.; Tristezza, M.; Grieco, F.; Spano, G.; Capozzi, V. From grape berries to wine: Population dynamics of cultivable yeasts associated to “Nero di Troia” autochthonous grape cultivar. World J. Microbiol. Biotechnol. 2016, 32, 59. [Google Scholar] [CrossRef] [PubMed]
  130. Capozzi, V.; Di Toro, M.R.; Grieco, F.; Michelotti, V.; Salma, M.; Lamontanara, A.; Russo, P.; Orrù, L.; Alexandre, H.; Spano, G. Viable But Not Culturable (VBNC) state of Brettanomyces bruxellensis in wine: New insights on molecular basis of VBNC behaviour using a transcriptomic approach. Food Microbiol. 2016, 59, 196–204. [Google Scholar] [CrossRef] [PubMed]
  131. Di Toro, M.R.; Capozzi, V.; Beneduce, L.; Alexandre, H.; Tristezza, M.; Durante, M.; Tufariello, M.; Grieco, F.; Spano, G. Intraspecific biodiversity and “spoilage potential” of Brettanomyces bruxellensis in Apulian wines. LWT Food Sci. Technol. 2015, 60, 102–108. [Google Scholar] [CrossRef]
  132. Berbegal, C.; Peña, N.; Russo, P.; Grieco, F.; Pardo, I.; Ferrer, S.; Spano, G.; Capozzi, V. Technological properties of Lactobacillus plantarum strains isolated from grape must fermentation. Food Microbiol. 2016, 57, 187–194. [Google Scholar] [CrossRef] [PubMed]
  133. Petruzzi, L.; Bevilacqua, A.; Corbo, M.R.; Speranza, B.; Capozzi, V.; Sinigaglia, M. A Focus on Quality and Safety Traits of Saccharomyces cerevisiae Isolated from Uva di Troia Grape Variety. J. Food Sci. 2017, 82, 124–133. [Google Scholar] [CrossRef] [PubMed]
  134. Van Hijum, S.A.; Vaughan, E.E.; Vogel, R.F. Application of state-of-art sequencing technologies to indigenous food fermentations. Curr. Opin. Biotechnol. 2013, 24, 178–186. [Google Scholar] [CrossRef] [PubMed]
  135. Rossetti, L.; Carminati, D.; Zago, M.; Giraffa, G. A qualified presumption of safety approach for the safety assessment of Grana Padano whey starters. Int. J. Food Microbiol. 2009, 130, 70–73. [Google Scholar] [CrossRef] [PubMed]
Table 1. Non-exhaustive list of fermented foods belonging to the main categories of global fermented foods. Information reported in accordance to Tamang et al. [16] (to Petruzzi et al. [20] for wine and Greppi et al. [21] for ogi and for tchoukoutou).
Table 1. Non-exhaustive list of fermented foods belonging to the main categories of global fermented foods. Information reported in accordance to Tamang et al. [16] (to Petruzzi et al. [20] for wine and Greppi et al. [21] for ogi and for tchoukoutou).
Major Groups Global Fermented FoodsFermented Product (Raw Material, Geographical Diffusion): Microorganisms Involved in the Fermentation Process
(a) Fermented cerealsSourdough (Rye, wheat; America, Europe, Australia): Lb. sanfranciscensis, Lb. alimentarius, Lb. buchneri, Lb. casei, Lb. delbrueckii, Lb. fructivorans, Lb. plantarum, Lb. reuteri, Lb. johnsonii, Cand. humili, Issatchenkia orientalis
Ogi (Maize, sorghum, millet; Nigeria): Lb. plantarum, Lb. pantheris, Lb. vaccinostercus, Corynebacterium sp., Aerobacter sp., Candida krusei, Clavispora lusitaniae, Sacch. cerevisiae, Rhodotorula sp., Cephalosporium sp., Fusarium sp., Aspergillus sp., Penicillium sp.
Idli (Rice, black gram or other dehusked pulses; India, Sri Lanka, Malaysia, Singapore): Leuc. mesenteroides, Lb. delbrueckii, Lb. fermenti, Lb. coryniformis, Ped. acidilactis, Ped. cerevisae, Streptococcus sp., Ent. faecalis, Lact. lactis, B. amyloliquefaciens, Cand. cacaoi, Cand. fragicola, Cand. glabrata, Cand. kefyr, Cand. pseudotropicalis, Cand. sake, Cand. tropicalis, Deb. hansenii, Deb. tamarii, Issatchenkia terricola, Rhiz. graminis, Sacch. cerevisiae, Tor. candida, Tor. holmii
(b) Fermented vegetables and bamboo shootsTable Olives (Olive; USA, Spain, Portugal, Italy, Greece, Peru, Chile): Leuc. mesenteroides, Ped. pentosaceus; Lb. plantarum Lb. pentosus/Lb. plantarum, Lb. paracollinoides, Lb. vaccinostercus/Lb. suebicus and Pediococcus sp. non-lactics (Gordonia sp./Pseudomonas sp., Halorubrum orientalis, Halosarcina pallid, Sphingomonas sp./Sphingobium sp./Sphingopyxis sp., Thalassomonas agarivorans) and yeasts (Candida cf. apicola, Pichia sp., Pic. manshurica/Pic. galeiformis, Sacch. cerevisiae)
Kimchi (Cabbage, green onion, hot pepper, ginger; Korea): Leuc. mesenteroides, Leuc. citreum, Leuc. gasicomitatum, Leuc. kimchii, Leuc. inhae, W. koreensis, W. kimchii, W. cibaria, Lb. plantarum, Lb. sakei, Lb. delbrueckii, Lb. buchneri, Lb. brevis, Lb. fermentum, Ped. acidilactici, Ped. pentosaceus, Lc. lactis, yeasts species of Candida, Halococcus, Haloterrigena, Kluyveromyces, Lodderomyces, Natrialba, Natronococcus, Pichia, Saccharomyces, Sporisorium, Trichosporon
Soibum (Bamboo shoot; India): Lb. plantarum, Lb. brevis, Lb. coryniformis, Lb. delbrueckii, Leuc. fallax, Leuc. Lact. lactis, Leuc. mesenteroides, Ent. durans, Strep. lactis, B. subtilis, B. lichniformis, B. coagulans, B. cereus, B. pumilus, Pseudomonas fluorescens, Saccharomyces sp., Torulopsis sp.
(c) Fermented legumesTempeh (Soybean; Indonesia, The Netherlands, Japan, USA): Rhiz. oligisporus, Rhiz. arrhizus, Rhiz. oryzae, Rhiz. stolonifer, Asp. niger, Citrobacter freundii, Enterobacter cloacae, K. pneumoniae, K. pneumoniae subsp. ozaenae, Pseudomas fluorescens, Lb. fermentum, Lb. lactis, Lb. plantarum, Lb. reuteri
Dawadawa (Locust bean; Ghana, Nigeria): B. pumilus, B. licheniformis, B. subtilis, B. firmus, B. atrophaeus, B. amyloliquefaciens, B. mojavensis, Lysininbacillus sphaericus.
(d) Fermented roots/tubersFufu (Cassava; West Africa): Bacillus sp., Lb. plantarum, Leuc. mesenteroides, Lb. cellobiosus, Lb. brevis; Lb. coprophilus, Lc. lactis; Leuc. lactis, Lb. bulgaricus, Klebsiella sp., Leuconostoc sp., Corynebacterium sp., Candida sp.
Tapé (Cassava; Indonesia): Streptococcus sp., Rhizopus sp., Saccharomycopsis fibuligera
(e) Fermented milk productsCheese (Animal milk; Worldwide): Lc. lactis subsp. cremoris, Lc. lactis subsp. lactis, Lb. delbrueckii subsp. delbrueckii, Lb. delbrueckii subsp. lactis, Lb. helveticus, Lb. casei, Lb. plantarum, Lb. salivarius, Leuconostoc spp., Strep. thermophilus, Ent. durans, Ent. faecium, and Staphylococcus spp., Brevibacterium linens, Propionibacterium freudenreichii, Debaryomyces hansenii, Geotrichum candidum, Penicillium camemberti, P. roqueforti
Kefir (Goat, sheep, cow; Russia): Lb. brevis, Lb. caucasicus, Lb. kefiri, Strep. thermophilus, Lb. bulgaricus, Lb. plantarum, Lb. casei, Lb. brevis, Tor. holmii, Tor. delbruechii
(f) Fermented and preserved meat productsChorizo (Pork; Spain): Lb. sake, Lb. curvatus, Lb. plantarum
Nem-chua (Pork, salt, cooked rice; Vietnam): Lb. pentosus, Lb. plantarum, Lb. brevis, Lb. paracasei, Lb. fermentum, Lb. acidipiscis, Lb. farciminis, Lb. rossiae, Lb. fuchuensis, Lb. namurensis, Lc. lactis, Leuc. citreum, Leuc. fallax, Ped. acidilactici, Ped. pentosaceus, Ped. stilesii, Weissella cibaria, W. paramesenteroides
(g) Fermented, dried and smoked fish productsNgari (Fish, salt; India): Lact. lactis, Lb. plantarum, Lb. pobuzihii, Lb. fructosus, Lb. amylophilus, Lb. coryniformis, Ent. faecium, B. subtilis, B. pumilus, B indicus, Micrococcus sp., Staphy. cohnii subsp. cohnii, Staphy. carnosus, Tetragenococcus halophilus subsp. flandriensis, Clostridium irregular, Azorhizobium caulinodans, Candida sp., Saccharomycopsis sp.
Surströmming (Fish; Sweden): Haloanaerobium praevalens
(h) Miscellaneous fermented productsBalsamic Vinegar (Grape must; Italy): Acetobacter aceti subsp. aceti, Acetobacter pasteurianus, Acetobacter polyxygenes, Acetobacter xylinum, Acetobacter malorum, Acetobacter pomorum, Candida lactis-condensi, Candida stellata, Hanseniaspora valbyensis, Hanseniaspora osmophila, Saccharomycodes ludwigii, Sacch. cerevisiae, Zygosaccharomyces bailii, Zygosaccharomyces bisporus, Zygosaccharomyces lentus, Zygosaccharomyces mellis, Zygosaccharomyces Pseudorouxii, Zygosaccharomyces Rouxii
Pidan (duck eggs; Chinese): B. cereus, B. macerans, Staph. cohnii, Staph. epidermidis, Staph. Haemolyticus, Staph. warneri
(i) Alcoholic beveragesWine (Grape must; Worldwide): Saccharomyces and non-Saccharomyces (so-called “wild”) yeasts (e.g., Candida colliculosa, C. stellata, Hanseniaspora uvarum, Kloeckera apiculata, Kl. thermotolerans, Torulaspora delbrueckii, Metschnikowia pulcherrima, Pichia fermentans, Schizosaccharomyces pombe, Hanseniaspora uvarum); bacteria (Oenococcus oeni, Lactobacillus plantatum)
Pulque (cactus (Agave) plant of Mexico): LAB (Lc. lactis subsp. lactis, Lb. acetotolerans, Lb. acidophilus, Lb. hilgardii, Lb. kefir, Lb. plantarum, Leuc. citreum, Leuc. kimchi, Leuc. mesenteroides, Leuc. pseudomesenteroides), the γ-Proteobacteria (Erwinia rhapontici, Enterobacter spp., and Acinetobacter radioresistens, several α-Proteobacteria), Zymomonas mobilis, Acetobacter malorum, A. pomorium, Microbacterium arborescens, Flavobacterium johnsoniae, Gluconobacter oxydans, Hafnia alvei
Tchoukoutou (spontaneously fermented beer) (sorghum, Benin): S. cerevisiae, Candida krusei, Clavispora lusitaniae, Candida rugosa
Table 2. Some examples about the impact of microorganisms associated with fermentations on the main aspect of global quality of fermented products.
Table 2. Some examples about the impact of microorganisms associated with fermentations on the main aspect of global quality of fermented products.
Global QualityPositive Effect of FermentationsReferences
hygienic qualityPyrazines biodegradation[30]
Toxins biodegradation[26,31]
Biogenic amines biodegradation[32]
sensory qualityFlavor improvement[33,34]
Texturizing properties[29,35]
nutritional qualityVitamin bio-fortification[4,27]
Increased bioaccessibility of minerals[36]
Reduction of antinutritional properties[37]
Reduction of lactose[38]
functional qualityAntioxidant activity enhancement[35]
Probiotic properties of selected strains[28,39]
Bio-fortification in microbial β-glucans[40]
Gluten degradation[41]
Table 3. Examples of the presence of contaminants of microbial origin hazardous for human health in fermented matrices.
Table 3. Examples of the presence of contaminants of microbial origin hazardous for human health in fermented matrices.
Safety IssueEvidenceReference
PathogensSome fermented foods (ready-to-eat) were found positive for the presence of Listeria monocytogenes in a global survey on several types of food products on sale in Portugal[42]
The most commonly encountered pathogens in African fermented foods include Bacillus cereus, Escherichia coli, Salmonella sp., Staphylococcus aureus, Vibrio cholera, Aeromonas, Klebsiella, Campylobacter and Shigella sp.[43]
51 doenjang samples have been found broadly contaminated with Bacillus cereus; while, only one sample was positive for Bacillus thuringiensis. All B. cereus isolates from doenjang were positive for diarrheal toxin genes.[46]
Epidemiologic investigation linked Escherichia coli O157:H7 infection with consumption of a commercial dry-cured salami product distributed in several western U.S. states.[47]
An outbreak of diarrhea and hemolytic uremic syndrome from Escherichia coli O157:H7 in fresh-pressed apple cider.[48]
Two commonly consumed traditional condiments (iru and ogiri) and their respective raw seeds (locust bean and melon) were found contaminated with potentially pathogenic species such as Alcaligenes faecalis, Bacillus anthracis, Proteus mirabilis and Staphylococcus sciuri subsp. sciuri occurred in the samples.[49]
Consumption of fermented raw pork sausage was associated with infection Salmonella enterica serovar Bovismorbificans.[50]
MycotoxinDietary sources of Ochratoxin A including fermented foods.[44]
Fermentation influences contamination of cocoa beans by Ochratoxin A.[51]
Biogenic aminesBioproduction of putrescine, histamine, tyramine and cadaverine in wine is a bacterial strain-dependent character.[52]
Bioproduction of tyramine in cheese up to considerable levels (e.g., 2010 mg/kg in Egyptian blue cheese).[53]
In fermented sausages, biogenic amines are mainly produced by fermentative microbial population.[45]
Ethyl carbamateEthyl carbamate produced by selected yeasts and lactic acid bacteria in red wine.[54]
Table 4. Presence of at least one strain belonging to the genera/species reported in Table 1 (thus associated with fermented matrices) and for which has been reported in literature a concern of safety nature.
Table 4. Presence of at least one strain belonging to the genera/species reported in Table 1 (thus associated with fermented matrices) and for which has been reported in literature a concern of safety nature.
Safety IssueStrainReference
MIC 1Aspergillus sp.[55]
MICCephalosporium sp.[56]
BA 2Enterococcus durans[57]
BAEnterococcus faecalis[57]
BAEnterococcus faecium[58]
MICFusarium sp.[55]
BAIssatchenkia terricola[59]
BALactobacillus buchneri[60]
BALactobacillus brevis[60]
BALactobacillus curvatus[61]
BALactobacillus hilgardii[60]
EC 3Lactobacillus hilgardii[62]
BALactobacillus mali[60]
BALactobacillus plantarum[63]
BALactobacillus reuteri[61]
BALeuconostoc mesenteroides[60]
BAMetschnikowia pulcherrima[59]
BAMicrococcus spp.[61]
BAOenococcus oeni[60]
MICPenicillium sp.[55]
BAPichia manshurica[59]
BAStaphylococcus carnosus[61]
1 mycotoxins, 2 biogenic amines, 3 ethyl carbamate precursors.
Table 5. A selection of definitions for “starter cultures” reported in recent scientific publications.
Table 5. A selection of definitions for “starter cultures” reported in recent scientific publications.
DefinitionReference
Starter cultures“Starter cultures” are preparations of live microorganisms or their resting forms, whose metabolic activity has desired effects in the fermentation substrate, the food. The preparations may contain unavoidable residues from the culture substrate and additives that support the vitality and technological functionality of the microorganisms (such as antifreeze or antioxidant compounds).[70]
Microbial food cultures“Microbial food cultures” (MFC) are live bacteria, yeasts or molds used in food production. MFC preparations are formulations, consisting of one or more microbial species and/or strains, including media components carried over from the fermentation and components which are necessary for their survival, storage, standardization, and to facilitate their application in the food production process.[65]
Commercial starter cultures“Commercial starter cultures” are standardized inoculum to be used for the production of fermented foods. Starter cultures are produced by specialized manufactures. Rigorous quality assurance and quality control are conducted to ensure performance, composition, and safety of the culture.[71]
Table 6. Selected scientific evidences concerning the effect of starter cultures inoculation in comparison with spontaneous fermentation.
Table 6. Selected scientific evidences concerning the effect of starter cultures inoculation in comparison with spontaneous fermentation.
MatrixEvidencesReference
Table olivesSelected lactobacilli and yeast showed a fast acidification of brine. Olives inoculated with lactobacilli and yeast showed the lowest biodiversity and the highest appreciation for both texture profile analysis and sensory evaluation[72]
Inoculation of brine medium with lactic acid bacteria starters significantly influenced aroma profiles [76]
Autochthonous starter produced same sensory quality as natural traditional table olives in a shorter time[77]
Fermented leek kimchi.Leeks fermented with Weissella confusa LK4 showed the highest radical scavenging effects and reducing ability. Total flavonoid and poly-phenolic contents changed during fermentation and showed correlation with anti-oxidant effects[78]
Soybean fermented productA reduction of biogenic amines and aflatoxins has been reported in Doenjang samples fermented with various Meju as starter cultures[73]
Dry-cured foal sausageThe inclusion of starter cultures contributes to improve the hygienic quality (stronger acidification and reduced count of Enterobacteriaceae) of foal sausages without significant effect on lipolysis, texture and appearance[79]
Chinese fermented dry sausagesNitrite content of all inoculated sausages declined rapidly during ripening compared to non-inoculated[74]
Fermented sausageSelected starter cultures improve quality, safety and sensorial properties of Dacia sausage, a traditional Romanian dry-sausage variety[80]
Evidences indicated that the selected Lactobacillus plantarum strain had a strong effect on inhibiting the production of biogenic amines[81]
Starter cultures minimize the formation of biogenic amines during the process of Nham fermentation[82]
Thai fermented shrimp (Kung-Som)Starter culture enhance GABA content, improved microbiological safety (dominated the total microflora) as well as reduced fermentation time[75]
WineIn cellars where biogenic amines are usually high, repeated experiments showed that in inoculated wines, biogenic amine concentrations were very low, while uninoculated control wines contained all the usual amines[83]
The use of a selected malolactic starter resulted in reductions in biogenic amines concentrations in wine produced by started malolactic fermentation compared with wine produced by spontaneous malolactic fermentation[84]
Table 7. Exemplificative cases of characterization of indigenous microorganisms selected from spontaneous fermentation in different Continents.
Table 7. Exemplificative cases of characterization of indigenous microorganisms selected from spontaneous fermentation in different Continents.
Major Groups Global Fermented FoodsProduct NameCountryReference
(a) Fermented cerealsOgiNigeria[104]
(b) Fermented vegetables and bamboo shootsSoidonIndia[105]
(c) Fermented legumesKedong sufuChina[106]
(d) Fermented roots/tubersGariKenya[107]
(e) Fermented milk productsSuero CosteñoColombia[108]
(f) Fermented and preserved meat productsSalchichónSpain[109]
(g) Fermented, dried and smoked fish productsJeotgalKorea[110]
(h) Miscellaneous fermented productsVinegarItaly[111]
(i) Alcoholic beveragesMalvar WinePhilippines.[112]
Table 8. Exemplificative cases of characterization of indigenous microorganisms selected from spontaneous fermentation in Apulian region (Southern Italy) in the last 15 years.
Table 8. Exemplificative cases of characterization of indigenous microorganisms selected from spontaneous fermentation in Apulian region (Southern Italy) in the last 15 years.
Fermented Foods/BeveragesAutochthonous Variety/Typical ProductStudied MicroorganismsMain Potential Impact on Food QualityReferences
BreadSourdough for Altamura bread (bread)Lactic acid bacteriaSensory quality[113]
CheeseCanestrato Pugliese (cheese)Lactobacilli and lactococciSensory quality[114]
CheeseMozzarella cheese (cheese)Lactobacilli, lactococci, streptococci and enterococciSensory quality[115]
CheeseMozzarella cheese (cheese)Lactobacillus plantarum, Lactobacillus helveticus, Lactobacillus delbrueckii subsp. lactis, Streptococcus thermophilus, Enterococcus faecalis, Enterococcus duransSensory quality[116]
Table olivesBella di Cerignola (Olives)YeastsSensory quality[117]
BreadSourdough for Altamura bread (bread)YeastsFunctional quality[118]
Table olivesBella di Cerignola (Olives)Lactic acid bacteriaFunctional quality[119]
Wine Different yeast speciesHygienic quality[59]
CheeseFior di Latte (cheese)Lactic acid bacteriaSensory quality[120]
Table olivesCellina di Nardò and Leccino (olives)Yeasts and lactic acid bacteriaSensory quality[121]
WineUva di Troia (grape)Oenococcus oeni, Saccharomyces cerevisiaeSensory quality[122]
WineNegroamaro (grape)Hanseniaspora uvarum, Saccharomyces cerevisiaeSensory quality[123]
WineUva di Troia (grape)Oenococcus oeni, Lactobacillus plantarumSensory quality[124,125,126]
WineUva di Troia (grape)Non-Saccharomyces yeastsSensory quality[127]
WineNegroamaro (grape)Yeasts and lactic acid bacteriaSensory quality[128]
WineUva di Troia (grape)YeastsSensory quality[129]
WineDifferent autochthonous grape varietiesBrettanomyces bruxellensisSensory quality[130,131]
WineUva di Troia (grape)Lactobacillus plantarumSensory quality[132]
WineUva di Troia (grape)Saccharomyces cerevisiaeHygienic quality[133]

Share and Cite

MDPI and ACS Style

Capozzi, V.; Fragasso, M.; Romaniello, R.; Berbegal, C.; Russo, P.; Spano, G. Spontaneous Food Fermentations and Potential Risks for Human Health. Fermentation 2017, 3, 49. https://doi.org/10.3390/fermentation3040049

AMA Style

Capozzi V, Fragasso M, Romaniello R, Berbegal C, Russo P, Spano G. Spontaneous Food Fermentations and Potential Risks for Human Health. Fermentation. 2017; 3(4):49. https://doi.org/10.3390/fermentation3040049

Chicago/Turabian Style

Capozzi, Vittorio, Mariagiovanna Fragasso, Rossana Romaniello, Carmen Berbegal, Pasquale Russo, and Giuseppe Spano. 2017. "Spontaneous Food Fermentations and Potential Risks for Human Health" Fermentation 3, no. 4: 49. https://doi.org/10.3390/fermentation3040049

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