Download as pdf or txt
Download as pdf or txt
You are on page 1of 34

Food Reviews International

ISSN: 8755-9129 (Print) 1525-6103 (Online) Journal homepage: https://www.tandfonline.com/loi/lfri20

A Review of Selection Criteria for Starter Culture


Development in the Food Fermentation Industry

Gilberto Vinicius De Melo Pereira, Dão P. De Carvalho Neto, Ana C. De O.


Junqueira, Susan G. Karp, Luiz A. J. Letti, Antonio I. Magalhães Júnior &
Carlos R. Soccol

To cite this article: Gilberto Vinicius De Melo Pereira, Dão P. De Carvalho Neto, Ana C. De O.
Junqueira, Susan G. Karp, Luiz A. J. Letti, Antonio I. Magalhães Júnior & Carlos R. Soccol (2019):
A Review of Selection Criteria for Starter Culture Development in the Food Fermentation Industry,
Food Reviews International, DOI: 10.1080/87559129.2019.1630636

To link to this article: https://doi.org/10.1080/87559129.2019.1630636

Published online: 20 Jun 2019.

Submit your article to this journal

Article views: 64

View Crossmark data

Full Terms & Conditions of access and use can be found at


https://www.tandfonline.com/action/journalInformation?journalCode=lfri20
FOOD REVIEWS INTERNATIONAL
https://doi.org/10.1080/87559129.2019.1630636

A Review of Selection Criteria for Starter Culture Development


in the Food Fermentation Industry
Gilberto Vinicius De Melo Pereira, Dão P. De Carvalho Neto, Ana C. De O. Junqueira,
Susan G. Karp, Luiz A. J. Letti, Antonio I. Magalhães Júnior, and Carlos R. Soccol
Bioprocess Engineering and Biotechnology Department, Federal University of Paraná (UFPR), Curitiba, Brazil

ABSTRACT KEYWORDS
Starter cultures are defined as selected microbial preparations used Alcoholic beverage; dairy;
to increase the efficiency of fermentation processes. In the food vinegar; coffee; soya; food
industry, numerous microbial cultures are used to ensure the produc- cultures
tion of safe and high-quality commodities. This review provides
a comprehensive theoretical guide for selecting microbial cultures
for use in the manufacturing of industrially important food commod-
ities, including dairy, meat, vegetables, alcoholic beverages, cocoa,
coffee, vinegar, and soy-based products. Related topics on fermented
food classification, microbial domestication, and regulatory require-
ments are addressed. The strategies reported in this review are useful
to help researchers choose methods and criteria selection.

Introduction
Fermentation is one of the oldest food-processing technologies. The classical fermentation
processes, such as winemaking, cheese making, and soya sauce production are documen-
ted as early as 600 B.C.[1] It was, however, with the penicillin production during World
War II that large-scale fermentations were first introduced.[2] Today a variety of fermented
foods is produced using this technology in large commercial enterprises. This includes
global-fermented foods, such as dairy, sausages, bread, alcoholic beverages, vegetables
(pickles, sauerkraut, and olives), and soy-based products.[3–5]
With the advancement of large-scale fermentations emerged the need to select well-
adapted microorganisms for industrial processing. Numerous microorganisms were selected
with specific characteristics for various food commodities, including Saccharomyces cerevisiae
used in the alcoholic beverage industry, Lactobacillus in lactic-acid-related products (fermen-
ted milks, vegetables, and meats), and molds in the production of soy-based products such as
miso, shoyu, and tempeh.[6–10] These microbial cultures are used to quickly start the fermen-
tation process and, hence, are known as starter cultures. The selective addition of starter
cultures to raw materials controls the overall fermentation process, reducing the risk of
fermentation failure while improving safety, stability, and end products’ sensorial value.[11]
The market for fermented food products is continuously in need of the implementation
and diversification of available products. For this purpose, an increasing number of
scientific studies have involved the selection of new strains with various and specific

CONTACT Carlos R. Soccol soccol@ufpr.br Bioprocess Engineering and Biotechnology Department, Federal
University of Paraná (UFPR), Curitiba 81531-970, Brazil
Color versions of one or more of the figures in the article can be found online at www.tandfonline.com/lfri.
© 2019 Taylor & Francis
2 G. VINICIUS DE MELO PEREIRA ET AL.

functional properties. The selection of food cultures involves a systematic approach


featuring multiple steps. It includes industry-specific characters for stress tolerance, key-
metabolite production, safety assessment, and technological parameter evaluation. In
addition, more recent studies have explored the complexity of microbial metabolism in
the production of specific molecules, such as enzymes, aromatic compounds, antimicro-
bials, and polysaccharides, which help to improve the texture, flavor and nutritional
aspects of fermented foods. In this review, a comprehensive theoretical guide is provided
for selecting microbial cultures for use in the manufacturing of industrially important
food commodities, including dairy, meat, vegetables, alcoholic beverages, cocoa, coffee,
vinegar, and soy-based products, to help researchers choose methods and criteria
selection.

Food fermentation classification


Foods produced by fermentative processes can be classified according to major biochem-
ical transformations that occur during processing (Fig. 1). In general, raw materials, such
as grapes, vegetables, milk, soy, and cocoa and coffee pulps, are rich substrates for
microbial growth. These substrates, consisting of water, fermentable sugars (e.g., glucose,
fructose, and sucrose), complex carbohydrates (cellulose and pectin), proteins, lipids,
minerals, and vitamins are converted into various end-metabolites responsible for con-
servation, nutritional enrichment, and sensorial improvement of industrially important
food commodities.
Globally consumed dairy foods, such as yogurt, cheeses, and fermented milks, are
produced by the metabolism of sugars into lactic acid by lactic acid bacteria (LAB).
Lactic acid fermentation is also used for the production of pickled vegetables, sauerkraut,
olives, and meats. These foods are produced by various LAB genera including,
Lactobacillus, Streptococcus, Lactococcus, Enterococcus, Pediococcus, and Leuconostoc.[3,12]
Among these, Lactobacillus species, such as Lb. acidophilus, Lb. pentosus, Lb. plantarum,
Lb. casei, Lb. paracasei, Lb. bulgaricus, Lb. rhamnosus, and Lb. lactis are the most
industrially exploited.[8,9,13–15] Metabolically, LAB areknown to produce high amounts
of lactic acid and other lower metabolites from diverse sources of carbon, including
lactose, fructose, glucose, and galactose. From glucose metabolism, LAB can be classified
as homofermentative bacteria, which produce lactic acid and carbon dioxide through the
Embden-Meyerhof-Parnas pathway or heterofermentative bacteria, which, in addition to
lactic acid, produce several other metabolites, including acetic acid, ethanol, and carbon
dioxide, via the pentose monophosphate pathway.[16,17] LAB also produce secondary
metabolites, including exopolysaccharides, enzymes, and bacteriocins, which are used to
increase the quality and shelf life of fermented foods.
Alcoholic fermentation is used in the food industry to produce beverages (wine, beer,
cachaça, sake, vodka, rum) and bread. The substrates include grape juice, diluted sugar-
cane juice, or hydrolyzed cereal grains, in which yeasts convert six-carbon molecules
(glucose and fructose) into nearly equal amounts of ethanol and carbon dioxide.[18,19]
During alcoholic beverage fermentation, suitable environmental conditions lead to a high
accumulation of ethanol, which is toxic to other microbial species. On the other hand,
ethanol is a minor product in bread because of its short fermentation time, which leads to
high rates of accumulated CO2.[20] Saccharomyces cerevisiae is the most commonly
FOOD REVIEWS INTERNATIONAL 3

Figure 1. Classification of industrial-fermented foods based on the major microbial metabolites.


4 G. VINICIUS DE MELO PEREIRA ET AL.

employed yeast in the food industry due to its high ethanol productivity and tolerance and
its ability to ferment a wide range of sugars.[21] Nonetheless, certain yeast strains, such as
Pichia sp., P. stipitis, and P. angophorae, have also been explored as effective ethanol
producers from various sugar sources.[22–24]
If the products of alcoholic fermentation are kept under aerobic conditions, acetic acid
bacteria (AAB) present in the environment oxidize some of the ethanol into acetic acid. This
process is useful in industrial production of vinegar with a final concentration of acetic acid
above 10%.[25] The industrial process of vinegar production involves three consecutive
phases: (i) conversion of sugars into ethanol by yeasts, (ii) oxidation of ethanol to acetic
acid by AAB, and (iii) an aging process ranging from 2 months (Balsamic vinegar) until 12
years (premium vinegar).[26] Saccharomyces and Zygosaccharomyces are the main yeast
species used in the first phase of alcoholic fermentation, and Gluconacetobacter (e.g.,
G. europaeus, G. oboediens, and G. intermedius) and Acetobacter (e.g., A. aceti and
A. pasteurianus) have been generally exploited for acetic acid production.[21,27–29]
Soy is a protein-rich substrate used for the production of oriental foods such as
shoyu, tempeh, miso, and natto. In certain soy-based products, the pH increases to
values as high as 8 due to the hydrolysis of proteins into peptides, amino acids, and
ammonia, known as alkaline fermentation process.[30] The essential proteolytic
microorganisms are filamentous fungi (e.g., Aspergillus and Rhizopus) in the produc-
tion of shoyu, tempeh, and miso, as well as Bacillus subtilis in the production of
natto.[7,10,31,32] The combination of high pH and free ammonia makes it very difficult
for other microorganisms that might spoil the product to grow. After complex
proteins and carbohydrates break down, lactic fermentation produces shoyu, tempeh,
and miso. For more of these fermentation processes’ biochemical details, the readers
are directed to review the work of Wang and Fung.[5]
Mixed fermentation is a classification used for food commodities, such as cocoa, coffee, and
black tea, carried out by different microbial groups, including yeasts, bacteria, and filamentous
fungi. These fermentative processes produce no major end-metabolites. Instead, ethanol,
acetic acid, lactic acid, and other metabolites are produced in similar proportions. In cocoa
fermentation, for example, the process begins with the dominance of yeast species that are
subsequently substituted by LAB which, in turn, decline after 48 h of fermentation and are
replaced by AAB.[33,34] The production of microbial aromatic molecules is considered essen-
tial for these processes in order to produce high-quality end products.[35–37]
Although the major above-mentioned biochemical events are used for food classifica-
tion, nearly all fermentation processes are the result of more than one microorganism. In
general, a microbial group that initiates fermentation will grow until the lack of substrate
or accumulation of by-products, inhibit further growth and activity.[38] However, during
this initial growth period, other organisms develop which are ready to take over when the
conditions become intolerable for the former ones. In wine production, final malolactic
fermentation (conversion of malic acid to lactic acid) by LAB is performed in red wines to
decrease astringency.[39,40] Vinegar fermentation is initiated by yeasts that work synergis-
tically, supplying ethanol for acetic acid bacteria growth.[41] In dairy products, several
other microbial groups are used together with LAB, including yeasts (Debaryomyces,
Candida, Yersinia, and Geotrichum species) and molds (Penicillum camemberti and
Penicillium roqueforti).[12]
FOOD REVIEWS INTERNATIONAL 5

Fermentation methods for food production


The production of fermented foods with superior quality depends on the presence,
growth, and metabolic activity of specific microorganisms. There are, essentially, three
manners of initiating food fermentation processes, viz., spontaneous fermentation, back-
slopping, and the use of starter culture. The earliest production of fermented foods was
based on spontaneous fermentation which relies on the indigenous microbiota present in
the raw material or from the environment. Milk and meat, for example, usually contain
the bacteria necessary to aid in the conversion of these materials into fermented cheese
and sausage, respectively, constituting a part of the ‘house flora’[42] Grapes and grape
crushing equipment contain yeasts responsible for the fermentation of sugars in ethanol
for wine production.[43,44] However, the presence of appropriate microorganisms does not
ensure a successful natural fermentation; the establishment of proper conditions and
bioavailability of nutrients for their optimal growth are also of vital importance.[45]
During vegetable fermentation for production of pickles and sauerkraut, low sodium
chloride concentration in brines stimulates the growth of LAB, whereas the elevated
concentrations of sugar present in grape promotes the growth of yeasts.[46–48]
A second strategy to initiate food fermentation processes is using part of a successful
fermentation as backslopping in order to ensure the dominance of the original
microbiota.[3,49] This procedure naturally selects well-adapted microorganisms reducing fer-
mentation time and increasing predictability. Backslopping inoculum has been previously
suggested for bread, beer, and cheese fermentations.[49,50] However, due to high microbial
load present in backslopping fermentation and, consequently, the difficulty to inoculate
a reproducible amount of cells, this method was gradually replaced by the use of selected, single
starter cultures. The primary function of a starter culture is to induce faster, more consistent
fermentation processes while controlling sensorial and nutritional aspects of the final product.
Today, a variety of microorganisms are used as starter cultures in the food industry. Figure 2
shows a Venn diagram constructed with the relationship of starter cultures and the different
food groups reported in Table 1. In general, Saccharomyces and Lactobacillus are the most
widely used microbial groups in the food industry, highlighting alcoholic beverages and dairy
products, respectively.[3,21] Interestingly, Venn diagram analysis pointed out that Candida is also
applied for the production of various food commodities since the use as an adjuvant in cheese
production or as starters for coffee, vegetables, alcoholic beverages, and vinegar fermentations.
Gluconobacter and Acetobacter species are mainly reported for vinegar production, filamentous
fungi (Aspergillus and Rhizopus) for soy-based foods, and several other microorganisms,
including yeast, bacteria, and filamentous fungi, in cocoa and coffee fermentations. Some
microbial species are explored to specific processes, such as Yarrowia in the coffee fermentation
process, Tolulospora in alcoholic beverages and dairy products, as well as a range of other non-
Saccharomyces species in the production of alcoholic beverages. The functions of each of these
species are reported in Table 1 and will be described later.

Microbial domestication
Many microorganisms have been domesticated by humans for use in food production.
The first microbial domestication clue occurred when fermentations began to be produced
from previous, well-succeeded processes (backslopping method) for the production of
6 G. VINICIUS DE MELO PEREIRA ET AL.

Figure 2. Venn diagram constructed from the relation starter culture and fermented foods according to
the data extracted from Table 1.

bread, beer, and cheese. This imposed a selective, man-made environment for the devel-
opment of specialized microbial strains. Next, with the advancement of microbiological
techniques and the development of starter cultures, several microbial phenotypes, highly
adapted to industrial processes (e.g., sugar-rich and oxygen-limited environments) were
generated.[165] A number of genetic events are reported from domestication processes,
such as genome decay, sugar metabolism, and stress tolerance enhancement, gene dupli-
cation, and regulatory rewiring of energy metabolism, which confers competitive advan-
tages during industrial fermentations. In the same sense, domestication is also
characterized by the potential loss of costly traits related to survival in nature, such as
ploidy alteration, loss of sexual reproduction ability, among others.[166]
Beer production is the most notable case of domestication since brewing yeasts are, in
general, propagated by continuous re-inoculation in a controlled environment without
contact for their ancestor. This led to the specialization of industrial species (viz.,
Saccharomyces cerevisiae and S. pastorianus) with enhanced ability to metabolize cereal-
derived carbohydrates, such as maltose and maltotriose, and tolerate high concentrations
of ethanol.[167] In winemaking, in contrast, yeasts are exposed to natural, uncontrolled
environments, with close contacts to wild yeasts. Thus, a range of hybridized yeasts with
a high tolerance for winemaking-associated stressful conditions are formed.[166] It is
Table 1. Developed starter cultures, selection criteria and inoculation effect for industrially important food commodities.
Product Starter culture Selection criteria Response to inoculation Reference
Lactic acid fermentation
[14,51–56]
Milk Lc. lactis; Stre. salivarius subsp. thermophilus; Lactic acid production; proteolytic and antifungal Improvement of fermentation process and sensory
Lb. delbrueckii ssp. bulgaricus activites; diacetyl, bacteriocin and exopolysaccharide attributes; pathogen growth inhibition
Lb. Harbinensis formation; phage resistance.
[8,15,57–71]
Cheese Lb. helveticus, Lb. delbrueckii subsp. lactis, Lb. Acid lactic production; volatile organic compounds, Accelerated cheese ripening; pathogen growth
paracasei subsp. paracasei, Lb. danicus, Lb. bacteriocin, and exopolysaccharides formation; inhibition; improvement of sensory attributes
rhamnosus, Lb. paraplantarum, Lb. acidophilus, Lb. proteolytic activity; salt tolerance.
plantarum, Lb. reuteri, Lb. brevis, Lb. plantarum, Stre.
thermophilus, Lc. lactis, Lc. cremosis, Lc. lactis subsp.
cremosis, Lc. lactis subsp. lactis, Lc. gasseri, W. cibaria,
Ent. faecium, Ent. durans, Leu. mesenteroides subsp.
cremosis, Leu. pseudomesenteroides
[13,72–75]
Olives Wick. anomalus, K. lactis, C. norvegica, Lb. pentosus, Acid lactic production; β-glucosidase activity; Reduction of the natural bitterness of raw olives.
Lb. plantarum, D. hansenii, T. delbrueckii capability to assimilate oleuropein and Accelerated fermentation; pathogen growth
polygalacturonic acid; antimicrobial activity. inhibition; high acetaldehyde formation.
[76–82]
Sauerkraut Lb. plantarum, Leu. mesenteroides, Leu. citreum, Lb. Acid lactic production; antibacterial activity; bile salt Faster fermentation process and dominance;
sakei, Lb. lactis, Lb. pentosus tolerance; nitrite depletion. improvement of sensory attributes; pathogen
growth inhibition; low nitrite formation.
[6,83–104]
Sausages Bif. animalis subsp. lactis, Lb. sakei, Lb. acidophilus, Ability to growth under different temperatures, pH Fermentation dominance; prevention of lipid
Lb. plantarum, Lb. delbrueckii, Lb. rhamnosus, Lb. and salt concentrations; negative for amino acid- oxidation; conservation of color; low content of
curvatus, Lb. casei ssp. casei, Lb. pentosus, Lb. decarboxylase; reduction of nitrite; low production biogenic amine; improvement of sensory attributes
farciminis, Lb. namurensis, Lb. casei/paracasei, Leu. of biogenic amines and carbon dioxide. Bacteriocin pathogen growth inhibition; efficient nitrite
mesenteroides, Lc. lactis ssp. lactis, Sta. equorum, Sta. production. reduction.
succinus, Sta. carnosus, Sta. xylosus, Sta.
saprophyticus, Sta. vitulus, Ped. pentosaceus, Ped.
acidilactici
Acetic acid fermentation
[27,29,41,105–
Vinegar Glu. xylinus, Gluconobacter sp., A. rancens, Glucose and ethanol tolerance; ethanol and acetate Fermentation 2 to 3 times faster than traditional
110]
A. lovanienis, A. xylinum, A. aceti, A. tropicalis, oxidation; cellulosic biopolymer formation; process, enriched aroma profile
A. pasteurianus, C. stellate, Sm. ludwigii, Z. bailii, thermotolerant; alcohol and aldehyde
Z. rouxii, Z. lentus, Z. sapae dehydrogenases formation; ability to grow in
different acetic acid concentration; aroma
compounds production.
Alcoholic fermentation
[111–119]
Beer S. cerevisiae, Sm. ludwigii, T. delbrueckii, P. kluyveri, Maltose and maltotriose fermentation; low H2 Improved fermentation process; beverage with high
H. vineae, Lachancea fermentati, L. thermotolerans, sensory quality; Improved fermentation at low
FOOD REVIEWS INTERNATIONAL

S formation; flocculation ability; low temperature


Schi. japonicas, Wick. anomalus. fermentation capacity; osmotolerant; production of temperatures; high production of flavor-active
volatile aroma compounds. compounds.
(Continued )
7
8

Table 1. (Continued).
Product Starter culture Selection criteria Response to inoculation Reference
[28,40,120–
Wine Lb. plantarum, O. oeni, Schi. pombe, Lb. plantarum, Malolactic fermentation capacity; volatile organic Improved fermentation process beverage with high
139]
Z. bailii, Lch. thermotolerans, S. cerevisiae, H. uvarum, compound production; glycerol formation; β- sensory quality (less acidity, color stability, and
D. pseudopolymorphus, P. guillermondii, Sm. ludwigii, glucosidase activity; low H2S formation; superior softness and aroma)
Bre. Bruxellensis, T. delbrueckii, M. pulcherrima, hydroxycinnamic acid decarboxylase activity;
C. zemplinina, H. vineae, Kz. gamospora, production of polysaccharides; low formation of
Z. kombuchaensis, P. kluyveri, H. guilliermondii. ethanol and acetic acid.
[140–145]
Cachaça S. cerevisiae, P. caribbica Resistance to high concentration of ethanol; β- Improvement of sensory attributes and fermentation
glucosidase activity; flocculation capacity; process
dominance and persistence in vats; non-H2
S production; resistance to antifungal components;
production of ethanol, glycerol, and volatile
compounds; low levels of acetaldehyde and
G. VINICIUS DE MELO PEREIRA ET AL.

methanol formation.
Soy-based products
[7,146]
Miso Asp. Oryzae Low acid phosphatase activity; antimutagenic Reductions of inosine monophosphate
activity. dephosphorylation activity; high antimutagenic
activity
[31,147,148]
Natto B. subtilis Enhanced production of extracellular cellulases and Reduction of hardness and development of sticky
proteases; ability to ferment soy protein texture; firmness, viscosity, ammonia content
improvement
[10,149]
Shoyu Asp. oryzae, Ped. halophilus, S. rouxii. Candida sp., Proteinase, peptidase and proteolytic activity; lactic Improvement of sensory attributes
R. oligosporus, Asp. oryzae acid, ethanol and flavor production; absence of
amino acid decarboxylase activity
[32,150,151]
Tempeh Rhizopus sp. Aromatic compounds production; proteolytic Improvement of sensory attributes; antioxidant
activity; free radical scavenging activity capacity improvement
Mixed fermentation
[35,152–156]
Coffee P. fermentans, S. cerevisiae, Lb. plantarum, Polygalacturonase, organic acids, and flavor-active Improvement of sensory attributes; accelerated
P. guilliermondii, C. parapsilosis, R. oligosporus, esters compounds production; pectinase, lipolytic fermentation process
Y. lipolytica and proteolytic activities
[157–164]
Cocoa S. cerevisiae, P. kluyveri, H. uvarum, Lb. plantarum, Lb. Ethanol, lactic acid, mannitol and acetic acid Improvement of sensory attributes; faster
fermentum, A. pasteurianus, A. tropicalis, A. aceti, production, stress tolerance, high aromatic fermentation process
K. marxianus production, pectinolytic activity.
The abbreviations of the microorganisms genera are as follows Streptococcus = Stre.; Lactobacillus = Lab.; Listeria = L.; Lactococcus = Lc.; Weissella = W.; Enterococcus = Ent.; Leuconostoc = Leu.;
Clostridium = Clos.; Bacillus = B.; Staphylococcus = Sta.; Wickerhamomyces = Wick.; Kluvyeromyces = K.; Candida = C.; Debaryomyces = D.; Torulaspora = T.; Escherichia = E.; Bifidobacterium =
Bif. = Pediococcus = Ped.; Gluconobacter = Glu.; Acetobacter = A.; Zygosaccharomyces = Z.; Saccharomycodes = Sm.; Saccharomyces = S.; Pichia = P.; Hanseniaspora = H.; Lachancea = Lch.;
Oeneococcus = O.; Schizosaccharomyces = Schi.; Brettanomyces = Bre.; Metschnikowia = M.; Kazachstania = Kz.; Aspergillus = Asp.; Rhizopus = R.; Yarrowia = Y.
FOOD REVIEWS INTERNATIONAL 9

estimated that more than 10 genetically distinct strains of S. cerevisiae have been found to
contribute to one wine fermentation.[168] A similar situation is also reported for other
fermented foods, such as bacteriocin-producing Streptococcus thermophilus strains used in
yogurt production,[169] thermotolerant Gluconobacter strains stable for vinegar
production,[170] and Lactobacillus species able to metabolize citric acid, the most abundant
organic acid present in cocoa pulp fermentation.[171]

Criteria for the selection of food-associated starter cultures


The selection of microorganisms in the food industry requires a systematic approach
similar to the one shown in Fig. 3. In general, the large number of microbial strains leads
the need to apply sequential tests to progressively reduce the number of candidates,
including (i) screening for stressful fermentation conditions, (ii) identification of key-
metabolite producers, and (iii) evaluation of technological parameters. At the end of this
procedure, the strains that present the highest number of functional properties and,
concomitantly, without any negative traits, are selected.
The primary screening is predominantly based on the assessment of the ability of
candidate strains to resist the stress conditions that fermentative processes impose. At the
beginning of the process, microbial cells are affected by osmotic stress due to the high
concentration of solutes in the fermentation medium, such as sugars in grapes, hydrolyzed
malt or sugar cane, proteins and lipids in dairy matrices, and sodium in vegetables.[172]
Cell exposition to such hypertonic conditions leads to an efflux of water from the cell,
decreasing both turgor pressure and water availability.[173] This causes the inactivation of

Figure 3. Schematic representation of the polyphasic approach required for food-associated starter
culture development.
10 G. VINICIUS DE MELO PEREIRA ET AL.

key enzymes of microbial metabolism, besides the increase in the concentration of toxic
ions and, ultimately, cell death. It is possible, however, to isolate strains that are well
adapted to these stressful environments by natural selection. These resistant strains
respond to osmotic stress by accumulating specific solutes, such as potassium ions,
amino acids (e.g., glutamine, glutamate, proline, y-aminobutyrate, and glycinebetaine)
and sugar (sucrose, trehalose, and glucosylglycerol), which decrease the activity of water
within the cytoplasm and increase both cell volume and turgor near their prestress
values.[174]
When the cell has adapted to the new environment and the fermentation begins, other
stressors become relevant, such as organic acids and alcohols accumulate, the temperature
changes and the environment acidifies.[172] Increase in temperature and/or decrease in pH
are probably the most frequent environmental factors that starter cultures must cope with
during fermentation processes. In vinegar and dairy products, for example, there is a high
rate of accumulation of acetic acid and lactic acid, respectively and, consequently, a sharp
decrease in pH takes place.[12,175] At concentrations as low as 5 g/L, organic acids begin to
affect cell growth and productivity.[172] The toxic effect results mostly from organic acid
dissociation inside microbial cells, causing a decrease of intracellular pH and metabolic
disturbance by the anion, among other deleterious effects.[176] Different studies demon-
strate the selection of lactic acid bacteria and acetic acid bacteria naturally resistant to
increased accumulations of organic acids, such as Acetobacter rancens, A. lovanienis,
A. xylinum, and Lactobacillus plantarum.[27,76,177] These resistant bacteria respond to
acid stress by altering membrane composition and/or increasing proton (H+) efflux and
amino acid catabolism as well as inducing the formation of DNA repair enzymes.[172]
During a fermentative process, exothermic reactions lead to a gradual increase in
temperature with concomitant effects on microbial growth and activity. In cocoa and
vinegar fermentation, for example, ethanol-to-acetic acid oxidation process is an extre-
mely exothermic reaction causing temperature rise as high as 50°C. The isolation of
naturally, thermo-resistant microbial strains allows a better performance of selected starter
culture to different fermentation processes, including dairy, cocoa, and vinegar
production.[27,105,178,179] Thermo-tolerant strains have the ability to increase the cytoplas-
mic membrane lipid-layer of acyl-chain length, saturation, branching (or altering anteiso
to iso branching), and/or cyclization of the fatty acids, which play an important role in
stabilizing the native conformation of the enzymes.[180] In addition, they increased the use
of sodium-ions rather than protons as coupling ion in energy transduction. Because of the
high H+ permeability, particularly at elevated temperatures, the use of a sodium motive
force can be of energetic advantage above a proton motive force.[181]
Most alcohols are natural metabolites of microorganisms, but very toxic to their hosts.
A typical example is ethanol whose toxicity is one of the most common stresses that
microbes encounter during the fermentation process. The primary effects of this molecule
appear to be increasing the fluidity of cell membrane and disruption of their phospholipid
components. Moreover, a number of physiological functions, including nutrient and ion
transport and cellular metabolism, may be altered by alcohol stress. Alcohol inhibition
occurs, normally, when its concentration reaches 12 g/L in the fermentation medium.[182]
After selecting best-adapted candidate strains, the second step is to search microbe cultures
with desirable metabolite production, which is specific for each fermentative process listed in
Fig. 1. This includes high production of ethanol (alcohol beverages), lactic acid (dairy, meat and
FOOD REVIEWS INTERNATIONAL 11

vegetable products), acetic acid (vinegar), and lytic enzymes (dairy and soy-associated products).
The production of aromatic molecules (such as acetaldehyde, isoamyl acetate, ethyl acetate,
phenylacetaldehyde, phenylethyl acetate, 1-octen-3-ol, and 2,3-butanediol) is of particular
interest in cocoa and coffee products, but, in general, specific metabolites are used to enhance
the sensorial quality of all food groups.
Finally, for commercial purpose, the starter culture must be amenable to large-scale
cultivation in relatively inexpensive substrates, besides being resistant for drying, packa-
ging, storage, and, finally, rehydration and reactivation processes.[183] These requirements
are necessary without the loss of essential and desirable fermentation properties.[168]

Starter culture development


Vinegar
The first step for the development of starter cultures for use in the manufacture of vinegar is
based on the selection of wild yeast and AAB with competitive activity during the full-
fermentation process. Initially, the high content of soluble solids of the raw material (grape,
rice, or sugar cane) leads to the need of selecting osmotolerant yeasts to perform the first stage
of ethanol production. In the same way, tolerance to high sugar concentration is applied to
selecting AAB, since after alcoholic fermentation, substantial quantities of sugars (20–40% of
the initial content) can still remain unfermented.[184,185] When the fermentation process
begins, oxidation reactions (manly sugar-to-ethanol and ethanol-to-acetic acid processes)
release a significant amount of energy and, consequently, a raise of the temperature to values
greater than 38°C takes place. The selection of thermotolerance microbial strains reduces
processing costs by eliminating the need for a cooling process.[105,186]
In the course of the fermentation process, ethanol and acetic acid accumulation are the
main factors that affect microbial growth and activity. Thus, microbial growth capacity at
10%-15% (v/v) of ethanol and 2–5% (v/v) of acetic acid are usually applied in vinegar
starter culture development.[29] Different AAB and yeast species have been selected for
vinegar production through stressful condition adaptation capacity (Table 1).
Zygosaccharomyces yeast and Gluconobacter species have been isolated from high sugar
content with the ability to grow over 30% of glucose,[41,187,188] while thermotolerant
Acetobacter species (Acetobacter tropicalis, A. pasteurianus) and Zygosaccharomyces lentus
have showed ability to grow up to 45°C.[106,107] In relation to acid tolerance, the high
alcohol dehydrogenase activity and stability in Gluconobater species enable this microbial
group to grow and stay metabolically active at high concentration of acetic acid.[189,190] On
the other hand, for Acetobacter species, in general, high acetic acid concentration causes
cell stress due to faster decrease of ADH activity.[191]
Polysaccharide production by AAB negatively affects the filterability of the product.[187]
Some Glucanoacetobacter species are known to synthesize high concentrations of dextrans,
levans, and cellulose and should be eliminated in selection programs.[192] Similarly, yeasts
with fructophilic activity should not be selected to avoid the formation of crystals in the
final product caused by residual glucose.[26] The yeast species Candida lactis-condensi and
C. stellata have been found to have the highest fructophilic phenotype, while
Hanseniaspora valbyensis, H. osmophila and S. cerevisiae show a glucophilic activity in
vinegar fermentations.[193]
12 G. VINICIUS DE MELO PEREIRA ET AL.

For the production of flavored vinegars, the selection of aroma-producing microbial


strains should be considered. A study conducted by Wang et al.[194] pointed out that, in
addition to the Gluconacetobacter and Acetobacter species, Lactobacillus, and Lactococcus
can be explored as potential vinegar aroma developers. In addition, Candida stellata,
H. osmophila and S. ludwigii have been found to produce high amount of flavor-active
compounds, such as succinic acid, ethyl acetate, and acetoin.[108–110] The role of yeast and
LAB in vinegar aroma formation has still not been well elucitated and a deeper knowledge
on this topic is required.

Lactic-acid-related products
In general, the development of starter culture for lactic fermentations is based on selecting
stress resistant and lactic acid-producing LAB strains.[195] The lactic fermentation-
associated stressful conditions include temperature changes, osmotic stress (such as high
concentrations of lipids and proteins in milk and meat and sodium chloride employed in
cheese and vegetable processing), and lactic acid accumulation.[12] These features are
required for fast microbial adaptation and growth, ensuring the development of physical
properties (texture, viscosity, and body) and taste of manufactured products.[12,196,197]
Based on these criteria, different LAB strains belonging to species involved in yogurt and
cheese production (Lactococcus lactis, S. thermophilus, and Lb. delbrueckii), olives (Lb.
plantarum and Lb. pentosus), fermented vegetables (Leuc. mesenteroides and Lb. plan-
tarum), and meats (Lb. pentosus and Ped. acidilactici) have been selected.[12,72,83,197–200]
The wide diversity of food products generated from lactic fermentation requires the
adoption of specific selection criteria for starter culture development. LAB starter cultures
in meat processing, for example, should be able to modulate the flavor of the finished
product without the generation of biogenic amines, since these compounds, resulting from
the decarboxylation of free amino acids, have cytotoxicity effect to human intestinal
cells.[201] Studies revealed that the use of homofermentative LAB, mainly Lactobacillus
spp. and Pediococcus spp., in association with proteolytic bacteria such as Staphylococcus
xylosus and S. carnosus, allows the reduction of biogenic amines in fermented
sausages.[94,98,99,202–205] Bacteriocin production, such as aureocin, enterocin, nisin, pedio-
cin, plantaricin, and lacticin is another characteristic used for meat starter culture devel-
opment in order to inhibit the growth of pathogenic bacteria and increase product shelf
life. Strains of Lactobacillus sakei, Lb. curvatus, and Lb. plantarum have been selected
based on the production of nisin, plantaricin, and lacticin active against Listeria mono-
cytogenes, Clostridium perfringens, Bacillus cereus, and Staphylococcus aureus.[100,101,206,207]
The ability to produce bacteriocins is also searched for the development of starter cultures
for the production of fermented milk, cheeses, and vegetables.[14,51,64–66,77–79] For vege-
table production, starter cultures can be selected based on the low or high production of
carbon dioxide. In pickles and olives, the high carbon dioxide content can form “bloaters”
and reduce the yield of the product. On the other hand, in cabbage fermentation, the
production of carbon dioxide is desirable in order to assist in the formation of an
anaerobic environment during sauerkraut production.[208] Daeschel et al.[209] isolated
and selected the mutant strain Lb. plantarum 965, which is used nowadays as
a commercial starter culture for pickles production due to low carbon dioxide formation.
FOOD REVIEWS INTERNATIONAL 13

In dairy products, the selection of exopolysaccharide-producing LAB has been performed


to improve the texture and functional properties of milk products.[52,67,68]
Lactic acid is a nonvolatile, odorless compound with neutral contribution to aroma
development in lactic fermentations.[12] Thus, the production of other functional, aro-
matic molecules has been used as a criterion for lactic starter culture selection. The major
aromatic molecules used for LAB selection include aldehydes (acetaldehyde, methional,
2-methylpropanal, phenylacetaldehyde) organic acids (acetic, butyric, propionic, and
succinic acids), esters (ethyl acetate, ethyl isobutanoate, phenylethyl acetate, ethyl hex-
anoate, ethylbutyrate), higher alcohols (1-octen-3-ol, 2,3-butanediol, methanethiol,
3-methyl-1-butanol; butan-2-ol, 3-methylbutanediol, phenylethanol), carboxylic acids
(2-methyl butyric acid, phenylacetic acid, benzoic acid, 2-methyl butyric acid), and
ketones (2-hexanone, acetoin, diacetyl, 2,3-pentanedione).[8,152,210–213] Indirectly, aromatic
microorganisms can be selected through the production of extracellular enzymes that
convert precursor molecules into the actual aroma compounds. In protein-rich environ-
ments, proteolytic enzymes play a major role in the formation of aromatic molecules
derived from the release of amino acids from complex proteins.[53,69,70,102] The milk
proteins, in particular caseins, fulfill the aroma precursor function, for instance for
Lactobacillus used in cheese and yogurt manufacturing.[214–216] Also, lipid degradation
plays a role in aroma formation of fermented meat and dairy products.[57,71,217] In yogurt,
despite a long list of volatile components that are found in the product, acetaldehyde,
ethanol, and diketones have a significant impact on the development of desirable sensory
characteristics.[218,219] The detailed review of Smid and Kleerebezem[212] described the
various precursor molecules for the formation of aroma compounds in connection with
the metabolic pathways involved in lactic acid fermentation.
The choice of microorganisms that present probiotic activities is becoming a new focus
for lactic starter culture development. This fact relies on the ability of these microorgan-
isms to confer, besides sensorial modifications, several functional characteristics such as
antimicrobial, anticarcinogenic, antidepressive, antioxidant, and cholesterol-lowering
activities.[220–222] Among the most-desired probiotic selection criteria are host-associated
stress resistance, epithelium adhesion ability, and antimicrobial activity.[223] The use of
probiotics is widespread in the production of fermented milks. However, other substrates
such as cheese and meat are being recognized as probiotic carriers and useful matrices to
protect the probiotic bacteria in the harsh conditions of the gastrointestinal tract.[224,225]

Alcoholic fermentation
The search for yeasts with rapid capacity of fermenting sugars into ethanol is the main
selection criteria for starter culture development in the alcoholic beverage industry. The
efficiency of ethanol production is, in general, increased by selecting yeast strains capable
of rapid growth in sugary starting materials, such as cereal starches in beers, whiskies, and
sake production, sucrose-rich plants (molasses or sugar juice from sugarcane) in rums,
cachaça, and tequila, and fructose-rich fruits in wine. The evaluation of sugar metabolism
should be carried out under fermentation-related stressful conditions, such as high
temperatures (35–45°C) and ethanol concentration (over 20%). In addition, yeast resis-
tance to low pH is an important attribute to make the process less susceptible to bacterial
infection. Other common selection criteria include: (i) low production of hydrogen sulfide
14 G. VINICIUS DE MELO PEREIRA ET AL.

(H2S), a byproduct of yeasts’ metabolism that is responsible for conferring off-flavors; (ii)
low production of acetic acid and other organic acids that may affect sensory quality and
ethanol yield; (iii) flocculation capacity, which facilitates downstream processing and
reduces cost of cells recovery; and (iv) volatile aroma compounds production, such as
higher alcohols, esters, organic acids, sulfur, and aldehydes.[130,133,134,226,227] Yeasts that
have been selected for alcoholic beverage production are summarized in Table 1.
Saccharomyces cerevisiae is the common specie used in the different processes due to its
high ethanol tolerance and productivity as well as the ability of fermenting a wide range of
sugars. However, other non-Saccharomyces species have also been investigated mainly due
to their high aromatic metabolites formation.
The selection of yeasts with reduced ethanol production is a practice performed by several
authors for wine starter culture development.[135,228] This is because ethanol compromises the
perception of wine aromatic complexity as well as rejection by health-conscious consumers.
One strategy adopted for lowering alcohol levels in wine is selecting non-Saccharomyces yeasts
with respiratory metabolism, such as Metschnikowia pulcherrima, Starmerella bacillaris,
Hanseniaspora uvarum, H. opuntiae, H. vinae, Pichia kudriavzevii, Candida flavescens,
Zygosaccharomyces bailii, Z. sapae, and Z. bisporus.[229–232] The combination of these uncon-
ventional yeasts with S. cerevisiae has been suggested to improve various attributes in wines,
such as aroma profile, sensory complexity, and color stability. In addition, non-Saccharomyces
yeasts usually produce functional enzymes, including glucosidase, pectinase, and
lichenase.[136,233] Pectinase activity reduces turbidity during vinification, lichenase can aid
wine filterability, while beta-glucosidase activity is related to the ability to hydrolyze terpe-
noids, norisoprenoids, benzene derivatives, and aliphatic components leading to an enriched
wine flavor. Finally, non-Saccharomyces yeasts with high glycerol production capacity have
been selected to enhance the softness and mouthfeel of wine.[133]
For distilled beverages production (e.g., cachaça, sake, and tequila), the selection of high
ethanol-producing yeasts is of extreme relevance to increase the yield during distillation.
Thus, strains of S. cerevisiae able to tolerate an ethanol accumulation greater than 15% are
usually selected.[234] However, other specific selection parameters include low production
of methanol and high production of ethyl lactate and ethyl acetate.[140–142]
Although fermentation of cereal extracts by Saccharomyces for beer production is
a well-standardized industrial process, recent studies have selected new yeast strains and
established different selection criteria. Araújo et al.[111] isolated different stress-tolerant
S. cerevisiae strains from cachaça fermentation able to produce high-quality beers. The
criteria used include efficiency on maltose and maltotriose fermentation, ethanol produc-
tion, and low production of acids, glycerol, and hydrogen sulfide. Canonico et al.[112]
selected a strain of Torulaspora delbrueckii on the basis of glucose, maltose, and sucrose
fermentation efficiency into ethanol. Although the T. delbrueckii strain did not present an
efficient fermentation capacity compared to a commercial Saccharomyces cerevisae, the co-
culturing of both resulted in a low alcohol beer with pleasant sensory analysis and a lighter
color. In the same sense, Holt et al.[113] selected different non-Saccharomyces yeasts (Pichia
kluyverii, Brettanomyces ssp., T. delbrueckii) with ability to support high osmolarity levels
(40% glucose and 40% fructose), and able to enrich beer aroma through the production of
flavor-active compounds, such as esters, higher alcohol, and phenolic compounds. The co-
culturing of these yeasts with an S. cerevisiae strain resulted in different-flavored beers.
Pichia kluyverii strains can add a fruity aroma in response to the production of isoamyl
FOOD REVIEWS INTERNATIONAL 15

acetate, Brettanomyces species conferred spicy notes by the release of ethyl phenolic
compounds, while T. delbrueckii enhanced the 4-vinyl guaiacol concentration attributing
clove-like aroma for the final product.

Soy-based products
In the manufacturing of miso and shoyu, soybeans are initially inoculated with mycelium of
Aspergillus oryzae previously grown on rice. The mold rice, or koji serves as a source of
enzymes. Thus, the main mold selection criterion is based on the production of macerating
enzymes (pectinase, cellulose, and hemicellulase), proteinase, peptidase, and amylases. The
production of such enzymes by A. oryzae enables the release of simple sugars and amino acids
for the second phase of lactic fermentation. Other common selection criteria include (i)
glutaminase activity for the improvement of flavor components, (ii) low phosphatase activity
to avoid the degradation of purinic ribonucleotides that enhance the undesired umami taste of
glutamate, and (iii) antimutagenic properties to increase biological properties of the
product.[7,146,235] For the second step in shoyu production (i.e., lactic acid fermentation),
the ability to produce lactic acid in high salt concentration is the main selection criterion.
Strains of Pediococcus halophilus are most commonly selected for this stage.[236,237] Finally, the
production of ethanol and flavor substances in high salt concentration by Saccharomyces
rouxii and Candida spp. are important characteristics to improve final product quality.[10,149]
In the production of miso, the second-step fermentation is performed by yeasts (S. rouxii and
Zygosaccharomyces spp.) and lactic acid bacteria that usually come from a previous fermenta-
tion or can be added as pure cultures. Hesseltine and Shibasaki[238] produced good quality
miso with a single culture of S. rouxii added after koji fermentation.
In the case of natto, the fermentation is exclusively performed by Bacillus subtilis. Selection of
starters has been based on enhanced production of extracellular cellulases and proteases to
promote soybean digestion, as well as the production of extracellular polyglutamic acid that
confers the sticky texture of the product.[31,147] In addition, different B. subtilis strains have been
compared in improving firmness, viscosity, and ammonia content of the final product.[148]
For tempeh production, the selection of proteases-producing Rhizopus sp. has been
proposed to increase protein bioavailability.[150] Other parameters include the production
of aromatic compounds to provide a pleasant and slightly cheese-like aroma,[151] produc-
tion of niacin, potassium, iron, and vitamins B1 and B2,[239] and free radical scavenging
activity related to product antioxidant acitivity.[32] Starzyńska-Janiszewska et al.[240] eval-
uated the addition of a Lactobacillus plantarum culture in tempeh fermentation and
obtained increased protein bioavailability, antioxidant capacity, and nutritional quality.
Keuth and Bisping[241] studied the addition of vitamin B12-producing bacteria (non-
pathogenic Citrobacter freundii and Klebsiella pneumoniae, isolated from Indonesian
tempeh and soaking water samples) to the mold fermentation and produced a bio-
enriched tempeh containing physiologically-active B12 vitamin.

Mixed fermentation
Cocoa and coffee are the main food commodities produced on a global scale by mixed
fermentation. Although a large number of microorganisms are responsible for these
fermentation processes, such as yeast, LAB, and AAB, it is possible to delineate common
16 G. VINICIUS DE MELO PEREIRA ET AL.

selection criteria. They include the ability to grow and dominate indigenous microorgan-
isms, produce pectinolytic enzymes in order to facilitate the dry process, and synthesize
aromatic molecules to aid in product quality.[36,37,157,242]
The complexity of cocoa fermentation requires the presence of a mixed starter culture
composed of functional microorganisms. In this case, LAB have been selected by the ability to
produce lactic acid, while efficient ethanol-to-acetic acid oxidation ability is searched for AAB.
Some species selected through these characteristics include Lactobacillis plantarum, Lb.
fermentum, Acetobacter pasteurianus, and A. tropicalis.[157–159] Yeasts with high capacity to
produce ethanol metabolize citric acid and produce high amounts of pectinolytic enzymes are
widely studied. These include S. cerevisiae, Pichia kluivery, Kluvyeromyces marxianus, and
Hanseniaspora uvarum.[160–162] Yeasts, LAB, and AAB are also selected by aroma compound
production, such as benzyl alcohol, phenethyl alcohol, ethyl hexanoate, ethyl octanoate, benzyl
acetate, and 2-phenethyl acetate.[160,161,178] The concomitant or sequential inoculation of
yeasts, LAB, and/or AAB has been suggested as an efficient strategy to achieve all these
functional properties.[33,157–159,161–163]
The selection of yeast with low ethanol and high aromatic compounds production (e.g.,
acetaldehyde, ethyl acetate, isoamyl acetate, and phenylethyl alcohol) is the main focus in
coffee starter studies. Pichia fermentans, P. guilliermondi, and Yarrowia lipolytica are the main
species selected.[35,153,154,243,244] LAB are also being exploited as a promising coffee starter
culture. In this case, rapid lactic acid production is used to accelerate the fermentative process.
Thus, homofermentive species, such as Lactobacillus plantarum, show attractive potential.[152]

Regulatory aspects and quality control


Similarly for all other ingredients used in foods, the risk of introducing new microbial
cultures into the market should be assessed. Some initiatives by the European community
(European Food Safety Agency (EFSA)) and United States (Food and Drugs Administration
(FDA)) have been dedicated to establishing criteria for the safety assessment of microbial
cultures used in foods. In 2007, EFSA introduced, after public hearing, the Qualified
Presumption of Safety QPS program, which includes a list of 93 microbial species consid-
ered safe for human use.[245] The QPS list takes into account aspects related to taxonomy
definition, body of knowledge (literature report and historical use), and safety concerns
(pathogenicity, virulence, and plasmid-associated antibiotic resistance spreading).[246,247] In
the US, the FDA uses a similar classification system known as GRAS (Generally Recognized
as Safe).[248] A microbial species that is not present in QPS or GRAS lists should undergo
a previous evaluation before it can be placed on the market.
Microbial cultures used in the food industry are usually referred to as working
culture. Genetic variations of working culture may result in lack of safety status and
productivity. A strict focus on quality control at every stage in the value chain should
be applied to ensure microbial stability. This includes maintenance of reference stock at
−80ºC, system for traceability, and use of newly stocked culture in each batch of
fermentation.[245] Additionally, molecular fingerprinting tools are indispensable for
molecular typing and detection of genetic variation. Chromosomal stability in yeast
can be monitored by pulsed-field gel electrophoresis (PFGE), while plasmid stability in
bacteria is usually evaluated by fluorescent amplified fragment length polymorphism
(FAFLP).[249,250]
FOOD REVIEWS INTERNATIONAL 17

Concluding remarks and future prospects


Starter cultures can be explored in a range of ways to improve processes and products
within the food industry. The selection of underutilized microbes should be based on
industrial fermentation fitness (stressful resistance), efficient production of primary meta-
bolites (ethanol, lactic acid, acetic acid, enzymes, secondary aromatic molecules, antimi-
crobial compounds, and polysaccharides), and technological properties (drying,
packaging, and storage capacity). Industrial fermentation processes require the maximiza-
tion of physiological behavior from food cultures. In this sense, genetic manipulation
techniques, such as recombinant DNA technology and Clustered Regularly Interspaced
Short Palindromic Repeats (CRISPR) can be used to generate improved cultures that meet
the industrial requirement.[215] Nonetheless, the use of genetically modified organism
(GMO) has conflicts with government regulations and consumer acceptability.[251,252] It
is possible to induce genetic modifications using naturally occurring genetic events, such
as hybridization, induced mutations, conjugation, and protoplast fusion, which are not
categorized as GMO.[253]
The process of starter culture selection has recently been improved using “omics”
technologies – i.e., genomic, transcriptomic, proteomic, and metabolomic. Genomics
can be used as a predictive tool for screening microbial cultures carrying genes related
to aroma molecules formation, enzymes production, and safety concerns (antibiotic
resistance and toxins production).[254–258] Accordingly, transcriptome, proteome, meta-
bolomics profiling can be used to evaluate the expression of these genes under different
industrial-related stress conditions, as well as identify posttranslational modifications that
affect protein functionality.[259]
In conclusion, the wide diversity of raw materials, processing methods, and end
products necessitate the application of specific selection tests for each food group. In
this article, strategies and methods are recommended for the selection of functional
microbes for the production of dairy, meat, vegetables, alcoholic beverages, cocoa, coffee,
vinegar, and soy-based products as useful support for upcoming starter culture selection
and validation process.

Acknowledgments
The authors thank Conselho Nacional de Desenvolvimento Científico e Tecnológico do Brasil
(CNPq) for the research scholarship.

Funding
This work was supported by the Conselho Nacional de Desenvolvimento Científico e Tecnológico
[303254/2017-3].

References
[1] Salque, M.; Bogucki, P. I.; Pyzel, J.; Sobkowiak-Tabaka, I.; Grygiel, R.; Szmyt, M.;
Evershed, R. P. Earliest Evidence for Cheese Making in the Sixth Millennium Bc in
Northern Europe. Nature. 2012, 493(7433), 522–525. DOI: 10.1038/nature11698.
18 G. VINICIUS DE MELO PEREIRA ET AL.

[2] Carvalho Neto, D. P.; Pereira, G. V.; de, M.; Finco, A. M. O.; Letti, L. A. J.; Silva, J. G.;
Vandenberghe, L. P. S.; Soccol, R. Efficient Coffee Beans Mucilage Layer Removal Using
Lactic Acid Fermentation in a Stirred-Tank Bioreactor: Kinetic, Metabolic and Sensorial
Studies. Food Biosci. 2018, 26, 80–87. DOI: 10.1016/j.fbio.2018.10.005.
[3] de Vuyst, L.;. Lactic Acid Bacteria as Functional Starter Cultures for the Food Fermentation
Industry. Trends Food Sci. Technol. 2004, 15(2), 67–78. DOI: 10.1016/j.tifs.2003.09.004.
[4] Di Cagno, R.; Coda, R.; De Angelis, M.; Gobbetti, M. Exploitation of Vegetables and Fruits
through Lactic Acid Fermentation. Food Microbiol. 2013, 33(1), 1–10. DOI: 10.1016/J.
FM.2012.09.003.
[5] Wang, J.; Fung, D. Y. C. Alkaline-Fermented Foods: A Review with Emphasis on Pidan
Fermentation. Crit. Rev. Microbiol. 1996, 22(2), 101–138. DOI: 10.3109/10408419609106457.
[6] 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. DOI: 10.1016/J.MEATSCI.2016.06.007.
[7] Marui, J.; Tada, S.; Fukuoka, M.; Wagu, Y.; Shiraishi, Y.; Kitamoto, N.; Sugimoto, T.;
Hattori, R.; Suzuki, S.; Kusumoto, K.-I. Reduction of the Degradation Activity of
Umami-Enhancing Purinic Ribonucleotide Supplement in Miso by the Targeted
Suppression of Acid Phosphatases in the Aspergillus Oryzae Starter Culture. Int. J. Food
Microbiol. 2013, 166(2), 238–243. DOI: 10.1016/j.ijfoodmicro.2013.07.006.
[8] Randazzo, C. L.; Pitino, I.; De, L. S.; Scifò, G. O.; Caggia, C. Effect of Wild Strains Used as
Starter Cultures and Adjunct Cultures on the Volatile Compounds of the Pecorino Siciliano
Cheese. Int. J. Food Microbiol. 2008, 122, 269–278. DOI: 10.1016/j.ijfoodmicro.2007.12.005.
[9] Xiong, T.; Li, X.; Guan, Q.; Peng, F.; Xie, M. Starter Culture Fermentation of Chinese
Sauerkraut: Growth, Acidification and Metabolic Analyses. Food Control. 2014, 41,
122–127. DOI: 10.1016/J.FOODCONT.2013.12.033.
[10] Yulifianti, R.; Ginting, E. Proteolytic Activity of Selected Moulds in the First Fermentation of
Black-Seeded Soysauce. Ser. Earth Environ. Sci. 2018, 102, 012097. DOI: 10.1088/1755-1315/
102/1/012097.
[11] Tamime, A. Y.;. Microbiology of Starter Cultures. In Dairy Microbiology Handbook: The
Microbiology of Milk and Milk Products; Robinson, R.K., Ed.; Wiley: New York, 2003; pp
261–366.
[12] Johnson, M. E.; Steele, J. L. Fermented Dairy Products. In Food Microbiology: Fundamentals
and Frontiers; Doyle, M.P., Buchanan, R.L., Eds.; ASM Press: Washington, 2013; pp
825–840.
[13] Blana, V. A.; Grounta, A.; Tassou, C. C.; Nychas, G.-J. E.; Panagou, E. Z. Inoculated
Fermentation of Green Olives with Potential Probiotic Lactobacillus Pentosus and
Lactobacillus Plantarum Starter Cultures Isolated from Industrially Fermented Olives.
Food Microbiol. 2014, 38, 208–218. DOI: 10.1016/J.FM.2013.09.007.
[14] Simova, E. D.; Beshkova, D. M.; Angelov, M. P.; Dimitrov, Z. P. Bacteriocin Production by
Strain Lactobacillus Delbrueckii Ssp. Bulgaricus BB18 during Continuous Prefermentation of
Yogurt Starter Culture and Subsequent Batch Coagulation of Milk. J. Ind. Microbiol.
Biotechnol. 2008, 35(6), 559–567. DOI: 10.1007/s10295-008-0317-x.
[15] de Souza, C. H. B.; Buriti, F. C. A.; Behrens, J. H.; Saad, S. M. I. Sensory Evaluation of
Probiotic Minas Fresh Cheese with Lactobacillus Acidophilus Added Solely or in Co-Culture
with a Thermophilic Starter Culture. Int. J. Food Sci. Technol. 2008, 43(5), 871–877. DOI:
10.1111/j.1365-2621.2007.01534.x.
[16] Carr, F. J.; Chill, D.; Maida, N. The Lactic Acid Bacteria: A Literature Survey. Crit. Rev.
Microbiol. 2002, 28(4), 281–370. DOI: 10.1080/1040-840291046759.
[17] Endo, A.; Dicks, L. M. T. Pgysiology of the LAB. In Lactic Acid Bacteria: Biodiversity and
Taxonomy; Holzapfel, W.H., Wood, B.J.B., Eds.; Wiley Blackwell: Chichester, 2014; pp
13–30.
FOOD REVIEWS INTERNATIONAL 19

[18] Marques, W. L.; Raghavendran, V.; Stambuk, B. U.; Gombert, A. K. Sucrose and
Saccharomyces Cerevisiae : A Relationship Most Sweet. FEMS Yeast Res. 2016, 16(1),
fov107. DOI: 10.1093/femsyr/fov107.
[19] Walker, G.; Stewart, G. Saccharomyces Cerevisiae in the Production of Fermented
Beverages. Beverages. 2016, 2(4), 30. DOI: 10.3390/beverages2040030.
[20] Chiotellis, E.; Campbell, G. M. Proving of Bread Dough II: Measurement of Gas Production
and Retention. Food Bioprod. Process. 2003, 81(3), 207–216. DOI: 10.1205/
096030803322437974.
[21] Piškur, J.; Rozpedowska, E.; Polakova, S.; Merico, A.; Compagno, C. How Did
Saccharomyces Evolve to Become a Good Brewer? Trends Genet. 2006, 22(4), 183–186.
DOI: 10.1016/j.tig.2006.02.002.
[22] Agbogbo, F. K.; Coward-Kelly, G. Cellulosic Ethanol Production Using the Naturally
Occurring Xylose-Fermenting Yeast, Pichia Stipitis. Biotechnol. Lett. 2008, 30(9),
1515–1524. DOI: 10.1007/s10529-008-9728-z.
[23] Horn, S. J.; Aasen, I. M.; Ostgaard, K. Ethanol Production from Seaweed Extract. J. Ind.
Microbiol. Biotechnol. 2000, 25(5), 249–254. DOI: 10.1038/sj.jim.7000065.
[24] Yu, Z.; Zhang, H. Pretreatments of Cellulose Pyrolysate for Ethanol Production by
Saccharomyces Cerevisiae, Pichia Sp. YZ-1 and Zymomonas Mobilis. Biomass Bioenergy.
2003, 24(3), 257–262. DOI: 10.1016/S0961-9534(02)00147-2.
[25] Gullo, M.; Verzelloni, E.; Canonico, M. Aerobic Submerged Fermentation by Acetic Acid
Bacteria for Vinegar Production: Process and Biotechnological Aspects. Process Biochem.
2014, 49(10), 1571–1579. DOI: 10.1016/J.PROCBIO.2014.07.003.
[26] Solieri, L.; Giudici, P. Yeasts Associated to Traditional Balsamic Vinegar: Ecological and
Technological Features. Int. J. Food Microbiol. 2008, 125(1), 36–45. DOI: 10.1016/j.
ijfoodmicro.2007.06.022.
[27] Saeki, A.; Theeragool, G.; Matsushita, K.; Toyama, H.; Lotong, N.; Adachi, O. Development
of Thermotolerant Acetic Acid Bacteria Useful for Vinegar Fermentation at Higher
Temperatures. Biosci. Biotechnol. Biochem. 1997, 61(1), 138–145. DOI: 10.1271/bbb.61.138.
[28] Garavaglia, J.; Schneider, R. C. S.; Camargo Mendes, S. D.; Welke, J. E.; Zini, C. A.;
Caramão, E. B.; Valente, P. Evaluation of Zygosaccharomyces Bailii BCV 08 as a
Co-Starter in Wine Fermentation for the Improvement of Ethyl Esters Production.
Microbiol. Res. 2015, 173, 59–65. DOI: 10.1016/J.MICRES.2015.02.002.
[29] Gullo, M.; Caggia, C.; De Vero, L.; Giudici, P. Characterization of Acetic Acid Bacteria in
“Traditional Balsamic Vinegar.”. Int. J. Food Microbiol. 2006, 106(2), 209–212. DOI:
10.1016/j.ijfoodmicro.2005.06.024.
[30] Parkouda, C.; Nielsen, D. S.; Azokpota, P.; Ouba, L. I. I.; Amoa-Awua, W. K.; Thorsen, L.;
Hounhouigan, J. D.; Jensen, J. S.; Tano-Debrah, K.; Diawara, B.;; et al. The Microbiology of
Alkaline-Fermentation of Indigenous Seeds Used as Food Condiments in Africa and Asia.
Crit. Rev. Microbiol. 2009, 35(2), 139–156.
[31] Kubo, Y.; Inaoka, T.; Hachiya, T.; Miyake, M.; Hase, S.; Nakagawa, R.; Hasegawa, H.;
Funane, K.; Sakakibara, Y.; Kimura, K. Development of a Rifampicin-Resistant Bacillus
Subtilis Strain for Natto-Fermentation Showing Enhanced Exoenzyme Production.
J. Biosci. Bioeng. 2013, 115(6), 654–657. DOI: 10.1016/J.JBIOSC.2012.12.012.
[32] Sheih, I. C.;. Preparation of High Free Radical Scavenging Tempeh by a Newly Isolated
Rhizopus Sp. R-69 from Indonesia. Food Sci. Agric. Chem. 2000, 2(1), 35–40.
[33] Schwan, R. F.; Wheals, A. E. The Microbiology of Cocoa Fermentation and Its Role in
Chocolate Quality. Crit. Rev. Food Sci. Nutr. 2004, 44(4), 205–221. DOI: 10.1080/
10408690490464104.
[34] Schwan, R. F.; Pereira, G. V. M.; Fleet, G. H. Microbial Activities during Cocoa
Fermentation. In Cocoa and Coffee Fermentation; Schwan, R.F., Fleet, G.H., Eds.; CRC
Press: Boca Raton, 2014; pp 129–192.
[35] Pereira, G. V. M.; Neto, E.; Soccol, V. T.; Medeiros, A. B. P.; Woiciechowski, A. L.;
Soccol, C. R. Conducting Starter Culture-Controlled Fermentations of Coffee Beans during
20 G. VINICIUS DE MELO PEREIRA ET AL.

on-Farm Wet Processing: Growth, Metabolic Analyses and Sensorial Effects. Food Res. Int.
2015, 75, 348–356. DOI: 10.1016/j.foodres.2015.06.027.
[36] Pereira, G. V.; M., D.; Soccol, V. T.; Brar, S. K.; Neto, E.; Soccol, C. R. Microbial Ecology and
Starter Culture Technology in Coffee Processing. Crit. Rev. Food Sci. Nutr. 2017, 57(13),
2775–2788. DOI: 10.1080/10408398.2015.1067759.
[37] Pereira, G. V.; de, M.; Soccol, V. T.; Soccol, C. R. Current State of Research on Cocoa and
Coffee Fermentations. Curr. Opin. Food Sci. 2016, 7, 50–57. DOI: 10.1016/J.
COFS.2015.11.001.
[38] Kovárová-Kovar, K.; Egli, T. Growth Kinetics of Suspended Microbial Cells: From
Single-Substrate-Controlled Growth to Mixed-Substrate Kinetics. Microbiol. Mol. Biol. Rev.
1998, 62(3), 646–666.
[39] Liu, S.-Q.;. Malolactic Fermentation in Wine - beyond Deacidification. J. Appl. Microbiol.
2002, 92(4), 589–601. DOI: 10.1046/j.1365-2672.2002.01589.x.
[40] Malherbe, S.; Menichelli, E.; Du Toit, M.; Tredoux, A.; Muller, N.; Naes, T.; Nieuwoudt, H.
The Relationships between Consumer Liking, Sensory and Chemical Attributes of Vitis
Vinifera L. Cv. Pinotage Wines Elaborated with Different Oenococcus Oeni Starter
Cultures. J. Sci. Food Agric. 2013, 93(11), 2829–2840. DOI: 10.1002/jsfa.6115.
[41] Turtura, G. C.; Benfenati, L. Caratteristiche Microbiologiche E Chimiche Dell’Aceto
Balsamico Naturale. Ann. Microbiol. 1988, 38, 51–74.
[42] Santos, E. M.; González-Fernández, C.; Jaime, I.; Rovira, J. Comparative Study of Lactic Acid
Bacteria House Flora Isolated in Different Varieties of `Chorizo’. Int. J. Food Microbiol.
1998, 39(1–2), 123–128. DOI: 10.1016/S0168-1605(97)00128-1.
[43] Renouf, V.; Claisse, O.; Lonvaud-Funel, A. Understanding the Microbial Ecosystem on the
Grape Berry Surface through Numeration and Identification of Yeast and Bacteria. Aust.
J. Grape Wine Res. 2005, 11(3), 316–327. DOI: 10.1111/j.1755-0238.2005.tb00031.x.
[44] Rosini, G.; Federici, F.; Martini, A. Yeast Flora of Grape Berries during Ripening. Microb.
Ecol. 1982, 8(1), 83–89. DOI: 10.1007/BF02011464.
[45] Steinkraus, K. H.;. Industrialization of Indigenous Fermented Foods, 2nd ed.; Marcel Dekker:
New York, 2004.
[46] Hurtado, A.; Reguant, C.; Bordons, A.; Rozès, N. Lactic Acid Bacteria from Fermented
Olives. Food Microbiol. 2012, 31, 1–8. DOI: 10.1016/j.fm.2012.02.003.
[47] Fleet, G. H.;. Yeast Interactions and Wine Flavour. Int. J. Food Microbiol. 2003, 86(1–2),
11–22. DOI: 10.1016/S0168-1605(03)00245-9.
[48] Rainieri, S.; Pretorius, I. S. Selection and Improvement of Wine Yeasts. Ann. Microbiol.
2000, 50, 15–31.
[49] Holzapfel, W. H.;. Appropriate Starter Culture Technologies for Small-Scale Fermentation in
Developing Countries. Int. J. Food Microbiol. 2002, 75(3), 197–212. DOI: 10.1016/S0168-
1605(01)00707-3.
[50] Holzapfel, W.;. Use of Starter Cultures in Fermentation on a Household Scale. Food Control.
1997, 8(5–6), 241–258. DOI: 10.1016/S0956-7135(97)00017-0.
[51] Benkerroum, N.; Oubel, H.; Mimoun, L. B. Behavior of Listeria Monocytogenes and
Staphylococcus Aureus in Yogurt Fermented with a Bacteriocin-Producing Thermophilic
Starter. J. Food Prot. 2002, 65(5), 799–805. DOI: 10.4315/0362-028X-65.5.799.
[52] Han, X.; Yang, Z.; Jing, X.; Yu, P.; Zhang, Y.; Yi, H.; Zhang, L. Improvement of the Texture
of Yogurt by Use of Exopolysaccharide Producing Lactic Acid Bacteria. Biomed Res. Int.
2016, 2016, 1–6. DOI: 10.1155/2016/7945675.
[53] Rahman, I. E. A.; Dirar, H. A.; Osman, M. A. Microbiological and Biochemical Changes and
Sensory Evaluation of Camel Milk Fermented by Selected Bacterial Starter Cultures. African
J. Food Sci. 2009, 3(12), 398–405.
[54] Madera, C.; García, P.; Janzen, T.; Rodríguez, A.; Suárez, J. E. Characterisation of
Technologically Proficient Wild Lactococcus Lactis Strains Resistant to Phage Infection.
Int. J. Food Microbiol. 2003, 86(3), 213–222. DOI: 10.1016/S0168-1605(03)00042-4.
[55] Delavenne, E.; Cliquet, S.; Trunet, C.; Barbier, G.; Mounier, J.; Le Blay, G. Characterization
of the Antifungal Activity of Lactobacillus Harbinensis K.V9.3.1Np and Lactobacillus
FOOD REVIEWS INTERNATIONAL 21

Rhamnosus K.C8.3.1I In Yogurt. Food Microbiol. 2015, 45(Pt A), 10–17. DOI: 10.1016/j.
fm.2014.04.017.
[56] Kuhl, G. C.; Gusso, A. P.; Porto, B. L. S.; Müller, C. M. O.; Mazzon, R. R.; Oliveira, M. A. L.;
Richards, N. S. P.; dos, S.; Lindner, J. D. D. Selection of Lactic Acid Bacteria for the
Optimized Production of Sheep’s Milk Yogurt with a High Conjugated Linoleic Acid
Content. J. Food Res. 2017, 6(4), 44. DOI: 10.5539/jfr.v6n4p44.
[57] Nieto-Arribas, P.; Poveda, J. M.; Seseña, S.; Palop, L.; Cabezas, L. Technological
Characterization of Lactobacillus Isolates from Traditional Manchego Cheese for Potential
Use as Adjunct Starter Cultures. Food Control. 2009, 20(12), 1092–1098. DOI: 10.1016/j.
foodcont.2009.03.001.
[58] De Angelis, M.; de Candia, S.; Calasso, M. P.; Faccia, M.; Guinee, T. P.; Simonetti, M. C.;
Selection, G. M. Use of Autochthonous Multiple Strain Cultures for the Manufacture of
High-Moisture Traditional Mozzarella Cheese. Int. J. Food Microbiol. 2008, 125(2), 123–132.
DOI: 10.1016/j.ijfoodmicro.2008.03.043.
[59] Ayad, E. H. E.;. Starter Culture Development for Improving Safety and Quality of Domiati
Cheese. Food Microbiol. 2009, 26(5), 533–541. DOI: 10.1016/J.FM.2009.03.007.
[60] Mills, S.; Serrano, L. M.; Griffin, C.; Connor, P. M. O.; Schaad, G.; Bruining, C.; Hill, C.;
Ross, R. P.; Meijer, W. C. Inhibitory Activity of Lactobacillus Plantarum LMG P-26358
against Listeria Innocua When Used as an Adjunct Starter in the Manufacture of Cheese.
Microbial Cell Factories. 2011, 10(Suppl 1), 1–11.
[61] Nascimento, M. S.; Moreno, I.; Kuaye, A. Y. Applicability of Bacteriocin-Producing
Lactobacillus Plantarum, Enterococcus Faecium and Lactococcus Lactis Ssp. Lactis as
Adjunct Starter in Minas Frescal Cheesemaking. Int. J. Dairy Technol. 2008, 61(4),
352–357. DOI: 10.1111/j.1471-0307.2008.00426.x.
[62] Ho, V. T. T.; Lo, R.; Bansal, N.; Turner, M. S. Characterisation of Lactococcus Lactis Isolates
from Herbs, Fruits and Vegetables for Use as Biopreservatives against Listeria
Monocytogenes in Cheese. Food Control. 2018, 85, 472–483. DOI: 10.1016/J.
FOODCONT.2017.09.036.
[63] Pedersen, T. B.; Ristagno, D.; Mcsweeney, P. L. H.; Vogensen, F. K.; Ardö, Y. Potential
Impact on Cheese Fl Avour of Heterofermentative Bacteria from Starter Cultures. Int. Dairy
J. 2013, 33(2), 112–119. DOI: 10.1016/j.idairyj.2013.03.003.
[64] Matijašić, B. B.; Rajšp, M. K.; Perko, B.; Rogelj, I. Inhibition of Clostridium Tyrobutyricum
in Cheese by Lactobacillus Gasseri. Int. Dairy J. 2007, 17(2), 157–166. DOI: 10.1016/J.
IDAIRYJ.2006.01.011.
[65] Rilla, N.; Martínez, B.; Delgado, T.; Rodríguez, A. Inhibition of Clostridium Tyrobutyricum
in Vidiago Cheese by Lactococcus Lactis Ssp. Lactis IPLA 729, a Nisin Z Producer. Int.
J. Food Microbiol. 2003, 85(1–2), 23–33. DOI: 10.1016/S0168-1605(02)00478-6.
[66] Martı́nez-Cuesta, M. C.; Requena, T.; Peláez, C. Use of a Bacteriocin-Producing
Transconjugant as Starter in Acceleration of Cheese Ripening. Int. J. Food Microbiol.
2001, 70(1–2), 79–88. DOI: 10.1016/S0168-1605(01)00516-5.
[67] Gómez-Ruiz, J. Á.; Cabezas, L.; Martínez-Castro, I.; González-Viñas, M. Á.; Poveda, J. M.
Influence of a Defined-Strain Starter and Lactobacillus Plantarum as Adjunct Culture on
Volatile Compounds and Sensory Characteristics of Manchego Cheese. Eur. Food Res.
Technol. 2008, 227(1), 181–190. DOI: 10.1007/s00217-007-0708-7.
[68] Lynch, K. M.; Mcsweeney, P. L. H.; Arendt, E. K.; Uniacke-lowe, T.; Galle, S.; Isolation, C. A.
And Characterisation of Exopolysaccharide-Producing Weissella and Lactobacillus and
Their Application as Adjunct Cultures in Cheddar Cheese. Int. Dairy J. 2014, 34(1),
125–134. DOI: 10.1016/j.idairyj.2013.07.013.
[69] Garde, S.; Tomillo, J.; Gaya, P.; Medina, M.; Nuñez, M. Proteolysis in Hispánico Cheese
Manufactured Using a Mesophilic Starter, a Thermophilic Starter, and Bacteriocin-
Producing Lactococcus Lactis Subsp. Lactis INIA 415 Adjunct Culture. J. Agric. Food
Chem. 2002, 50(12), 3479–3485. DOI: 10.1021/jf011291d.
[70] Terzić-Vidojević, A.; Tonković, K.; Leboš Pavunc, A.; Beganović, J.; Strahinić, I.; Kojić, M.;
Veljović, K.; Golić, N.; Kos, B.; Čadež, N.;, et al. Evaluation of Autochthonous Lactic Acid
22 G. VINICIUS DE MELO PEREIRA ET AL.

Bacteria as Starter Cultures for Production of White Pickled and Fresh Soft Cheeses. LWT -
Food Sci. Technol. 2015, 63(1), 298–306. DOI: 10.1016/J.LWT.2015.03.050.
[71] Milesi, M. M.; Mcsweeney, P. L. H.; Hynes, E. R. Viability and Contribution to Proteolysis of
an Adjunct Culture of Lactobacillus Plantarum in Two Model Cheese Systems : Cheddar
Cheese-Type and Soft-Cheese Type. J. Appl. Microbiol. 2008, 105, 884–892. doi:10.1111/
j.1365-2672.2008.03813.x
[72] Psani, M.; Kotzekidou, P. Technological Characteristics of Yeast Strains and Their Potential
as Starter Adjuncts in Greek-Style Black Olive Fermentation. World J. Microbiol. Biotechnol.
2006, 22(12), 1329–1336. DOI: 10.1007/s11274-006-9180-y.
[73] Bevilacqua, A.; Beneduce, L.; Sinigaglia, M.; Corbo, M. R. Selection of Yeasts as Starter
Cultures for Table Olives. J. Food Sci. 2013, 78(5), M742–M751. DOI: 10.1111/1750-
3841.12117.
[74] Ruiz-Barba, J. L.; Jiménez-Díaz, R. A Novel Lactobacillus Pentosus-Paired Starter Culture for
Spanish-Style Green Olive Fermentation. Food Microbiol. 2012, 30(1), 253–259. DOI:
10.1016/J.FM.2011.11.004.
[75] Panagou, E. Z.; Schillinger, U.; Franz, C. M. A. P.; Nychas, G. E. Microbiological and
Biochemical Profile of Cv. Conservolea Naturally Black Olives during Controlled
Fermentation with Selected Strains of Lactic Acid Bacteria. Food Microbiol. 2008, 25(2),
348–358. DOI: 10.1016/j.fm.2007.10.005.
[76] Xiong, T.; Song, S.; Huang, X.; Feng, C.; Liu, G.; Huang, J.; Screening, X. M. Identification of
Functional Lactobacillus Specific for Vegetable Fermentation. J. Food Sci. 2013, 78(1), M84–
M89. DOI: 10.1111/j.1750-3841.2012.03003.x.
[77] Tolonen, M.; Rajaniemi, S.; Pihlava, J.; Johansson, T. Formation of Nisin, Plant-Derived
Biomolecules and Antimicrobial Activity in Starter Culture Fermentations of Sauerkraut.
Food Microbiol. 2004, 21(2), 167–179. DOI: 10.1016/S0740-0020(03)00058-3.
[78] Chang, J. Y.; Chang, H. C. Growth Inhibition of Foodborne Pathogens by Kimchi Prepared
with Bacteriocin-Producing Starter Culture. J. Food Sci. 2011, 76(1), M72–M78. DOI:
10.1111/j.1750-3841.2010.01965.x.
[79] Rao, Y.; Chang, W.; Xiang, W.; Li, M.; Che, Z.; Tang, J. Screening and Performance of
L Actobacillus Plantarum E11 with Bacteriocin-Like Substance Secretion as Fermentation
Starter of Sichuan Pickle. J. Food Saf. 2013, 33(4), 445–452. DOI: 10.1111/jfs.12075.
[80] Beganović, J.; Pavunc, A. L.; Gjuračić, K.; Špoljarec, M.; Šušković, J.; Kos, B. Improved
Sauerkraut Production with Probiotic Strain Lactobacillus Plantarum L4 and Leuconostoc
Mesenteroides LMG 7954. J. Food Sci. 2011, 76(2), M124–M129. DOI: 10.1111/j.1750-
3841.2010.02030.x.
[81] Jung, J. Y.; Lee, S. H.; Lee, H. J.; Seo, H.-Y.; Park, W.-S.; Jeon, C. O. Effects of Leuconostoc
Mesenteroides Starter Cultures on Microbial Communities and Metabolites during Kimchi
Fermentation. Int. J. Food Microbiol. 2012, 153(3), 378–387. DOI: 10.1016/j.
ijfoodmicro.2011.11.030.
[82] Yan, P.-M.; Xue, W.-T.; Tan, -S.-S.; Zhang, H.; Chang, X.-H. Effect of Inoculating Lactic
Acid Bacteria Starter Cultures on the Nitrite Concentration of Fermenting Chinese Paocai.
Food Control. 2008, 19(1), 50–55. DOI: 10.1016/J.FOODCONT.2007.02.008.
[83] Baka, A. M.; Papavergou, E. J.; Pragalaki, T.; Bloukas, J. G.; Kotzekidou, P. Effect of Selected
Autochthonous Starter Cultures on Processing and Quality Characteristics of Greek
Fermented Sausages. LWT - Food Sci. Technol. 2011, 44(1), 54–61. DOI: 10.1016/J.
LWT.2010.05.019.
[84] Bedia, M.; Méndez, L.; Bañón, S. Evaluation of Different Starter Cultures (Staphylococci Plus
Lactic Acid Bacteria) in Semi-Ripened Salami Stuffed in Swine Gut. Meat Sci. 2011, 87(4),
381–386. DOI: 10.1016/J.MEATSCI.2010.11.015.
[85] 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(7), 1562–1571. DOI: 10.1016/J.LWT.2011.01.028.
FOOD REVIEWS INTERNATIONAL 23

[86] Cenci-Goga, B. T.; Rossitto, P. V.; Sechi, P.; Parmegiani, S.; Cambiotti, V.; Cullor, J. S. Effect
of Selected Dairy Starter Cultures on Microbiological, Chemical and Sensory Characteristics
of Swine and Venison (Dama Dama) Nitrite-Free Dry-Cured Sausages. Meat Sci. 2012, 90
(3), 599–606. DOI: 10.1016/j.meatsci.2011.09.022.
[87] Klingberg, T. D.; Axelsson, L.; Naterstad, K.; Elsser, D.; Budde, B. B. Identification of
Potential Probiotic Starter Cultures for Scandinavian-Type Fermented Sausages. Int.
J. Food Microbiol. 2005, 105, 419–431. DOI: 10.1016/j.ijfoodmicro.2005.03.020.
[88] Lee, J.; Kim, C.; Kunz, B. Identification of Lactic Acid Bacteria Isolated from Kimchi and
Studies on Their Suitability for Application as Starter Culture in the Production of
Fermented Sausages. Meat Sci. 2006, 72, 437–445. DOI: 10.1016/j.meatsci.2005.08.013.
[89] Neffe-Skocińska, K.; Okoń, A.; Kołożyn-Krajewska, D.; Dolatowski, Z. Amino Acid Profile
and Sensory Characteristics of Dry Fermented Pork Loins Produced with a Mixture of
Probiotic Starter Cultures. J. Sci. Food Agric. 2017, 97(9), 2953–2960. DOI: 10.1002/
jsfa.8133.
[90] Ratanaburee, A.; Kantachote, D.; Charernjiratrakul, W.; Sukhoom, A. Enhancement of γ-
Aminobutyric Acid (GABA) in Nham (Thai Fermented Pork Sausage) Using Starter
Cultures of Lactobacillus Namurensis NH2 and Pediococcus Pentosaceus HN8. Int.
J. Food Microbiol. 2013, 167(2), 170–176. DOI: 10.1016/J.IJFOODMICRO.2013.09.014.
[91] Ravyts, F.; Barbuti, S.; Frustoli, M. A.; Parolari, G.; Saccani, G.; De Vuyst, L.;
Competitiveness, L. F. Antibacterial Potential of Bacteriocin-Producing Starter Cultures in
Different Types of Fermented Sausages. J. Food Prot. 2008, 71(9), 1817–1827.
[92] Rubio, R.; Aymerich, T.; Bover-cid, S.; Guàrdia, M. D.; Arnau, J.; Garriga, M. Probiotic
Strains Lactobacillus Plantarum 299V and Lactobacillus Rhamnosus GG as Starter Cultures
for Fermented Sausages. LWT - Food Sci. Technol. 2013, 54(1), 51–56. DOI: 10.1016/j.
lwt.2013.05.014.
[93] Rubio, R.; Jofré, A.; Martín, B.; Aymerich, T.; Garriga, M. Characterization of Lactic Acid
Bacteria Isolated from Infant Faeces as Potential Probiotic Starter Cultures for Fermented
Sausages. Food Microbiol. 2014, 38, 303–311. DOI: 10.1016/j.fm.2013.07.015.
[94] Bover-Cid, S.; Izquierdo-Pulido, M.; Vidal-Carou, M. C. Mixed Starter Cultures to Control
Biogenic Amine Production in Dry Fermented Sausages. J. Food Prot. 2000, 63(11),
1556–1562. DOI: 10.4315/0362-028X-63.11.1556.
[95] Maria, A.; Simion, C.; 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(1), 123–131. DOI:
10.1016/j.foodcont.2013.06.047.
[96] Talon, R.; Leroy, S.; Lebert, I.; Giammarinaro, P.; Chacornac, J.; Latorre-moratalla, M.;
Vidal-carou, C.; Zanardi, E.; Conter, M.; Safety Improvement, L. A. Preservation of
Typical Sensory Qualities of Traditional Dry Fermented Sausages Using Autochthonous
Starter Cultures. Int. J. Food Microbiol. 2008, 126, 227–234. DOI: 10.1016/j.
ijfoodmicro.2008.05.031.
[97] Wang, X. H.; Ren, H. Y.; Liu, D. Y.; Zhu, W. Y.; Wang, W. Effects of Inoculating
Lactobacillus Sakei Starter Cultures on the Microbiological Quality and Nitrite Depletion
of Chinese Fermented Sausages. Food Control. 2013, 32(2), 591–596. DOI: 10.1016/j.
foodcont.2013.01.050.
[98] Gardini, F.; Martuscelli, M.; Crudele, M. A.; Paparella, A.; Suzzi, G. Use of Staphylococcus
Xylosus as a Starter Culture in Dried Sausages: Effect on the Biogenic Amine Content. Meat
Sci. 2002, 61(3), 275–283. DOI: 10.1016/S0309-1740(01)00193-0.
[99] Lu, S.; Xu, X.; Zhou, G.; Zhu, Z.; Meng, Y.; Sun, Y. Effect of Starter Cultures on Microbial
Ecosystem and Biogenic Amines in Fermented Sausage. Food Control. 2010, 21(4), 444–449.
DOI: 10.1016/j.foodcont.2009.07.008.
[100] Benkerroum, N.; Daoudi, A.; Hamraoui, T.; Ghalfi, H.; Thiry, C.; Duroy, M.; Evrart, P.;
Roblain, D.; Thonart, P. Lyophilized Preparations of Bacteriocinogenic Lactobacillus
Curvatus and Lactococcus Lactis Subsp. Lactis as Potential Protective Adjuncts to Control
24 G. VINICIUS DE MELO PEREIRA ET AL.

Listeria Monocytogenes in Dry-Fermented Sausages. J. Appl. Microbiol. 2005, 98(1), 56–63.


DOI: 10.1111/j.1365-2672.2004.02419.x.
[101] Gao, Y.; Li, D.; Liu, X. Bacteriocin-Producing Lactobacillus Sakei C2 as Starter Culture in
Fermented Sausages. Food Control. 2014, 35(1), 1–6. DOI: 10.1016/J.
FOODCONT.2013.06.055.
[102] Casaburi, A.; Di Monaco, R.; Cavella, S.; Toldrá, F.; Ercolini, D.; Villani, F. Proteolytic and
Lipolytic Starter Cultures and Their Effect on Traditional Fermented Sausages Ripening and
Sensory Traits. Food Microbiol. 2008, 25(2), 335–347. DOI: 10.1016/j.fm.2007.10.006.
[103] Ammor, S.; Dufour, E.; Zagorec, M.; Chaillou, S.; Characterization, C. I. Selection of
Lactobacillus Sakei Strains Isolated from Traditional Dry Sausage for Their Potential Use
as Starter Cultures. Food Microbiol. 2005, 22(6), 529–538. DOI: 10.1016/J.FM.2004.11.016.
[104] Babić, I.; Markov, K.; Kovačević, D.; Trontel, A.; Slavica, A.; Đugum, J.; Čvek, D.;
Svetec, I. K.; Posavec, S.; Frece, J. Identification and Characterization of Potential
Autochthonous Starter Cultures from a Croatian “Brand” Product “Slavonski Kulen.”.
Meat Sci. 2011, 88(3), 517–524. DOI: 10.1016/J.MEATSCI.2011.02.003.
[105] Moonmangmee, D.; Adachi, O.; Ano, Y.; Shinagawa, E.; Toyama, H.; Theeragool, G.;
Lotong, N.; Isolation, M. K. Characterization of Thermotolerant Gluconobacter Strains
Catalyzing Oxidative Fermentation at Higher Temperatures. Biosci. Biotechnol. Biochem.
2000, 64(11), 2306–2315. DOI: 10.1271/bbb.64.2306.
[106] Ndoye, B.; Lebecque, S.; Dubois-Dauphin, R.; Tounkara, L.; Guiro, A.-T.; Kere, C.;
Diawara, B.; Thonart, P. Thermoresistant Properties of Acetic Acids Bacteria Isolated from
Tropical Products of Sub-Saharan Africa and Destined to Industrial Vinegar. Enzyme
Microb. Technol. 2006, 39(4), 916–923. DOI: 10.1016/J.ENZMICTEC.2006.01.020.
[107] Steels, H.; Bond, C. J.; Collins, M. D.; Roberts, I. N.; Stratford, M.; James, S. A.
Zygosaccharomyces Lentus Sp. Nov., A New Member of the Yeast Genus
Zygosaccharomyces Barker. Int. J. Syst. Bacteriol. 1999, 49(1), 319–327. DOI: 10.1099/
00207713-49-1-319.
[108] Ciani, M.;. Wine Vinegar Production Using Base Wines Made with Different Yeast Species.
J. Sci. Food Agric. 1998, 78(2), 290–294. DOI: 10.1002/(SICI)1097-0010(199810)78:2<290::
AID-JSFA120>3.0.CO;2-A.
[109] Saeki, A.;. Studies on Acetic Acid Fermentation. III Continuous Production of Vinegar with
Immobilized Saccharomycodes Ludwigii Cells and Immobilized Acetobacter Aceti Cells
Entrapped in Calcium Alginate Gel Beads. J. Jpn. Soc. Food Sci. 1990, 37, 722–725.
[110] Granchi, L.; Ganucci, D.; Messini, A.; Vincenzini, M. Oenological Properties of
Hanseniaspora Osmophila and Kloeckera Corticis from Wines Produced by Spontaneous
Fermentations of Normal and Dried Grapes. FEMS Yeast Res. 2002, 2(3), 403–407.
[111] Araújo, T. M.; Souza, M. T.; Diniz, R. H. S.; Yamakawa, C. K.; Soares, L. B.; Lenczak, J. L.; de
Castro Oliveira, J. V.; Goldman, G. H.; Barbosa, E. A.; Campos, A. C. S.;, et al. Cachaça Yeast
Strains: Alternative Starters to Produce Beer and Bioethanol. Antonie Van
Leeuwenhoek.2018, 111(10), 1749–1766. DOI: 10.1007/s10482-018-1063-3.
[112] Canonico, L.; Comitini, F.; Ciani, M. Torulaspora Delbrueckii Contribution in Mixed
Brewing Fermentations with Different Saccharomyces Cerevisiae Strains. Int. J. Food
Microbiol. 2017, 259, 7–13. DOI: 10.1016/J.IJFOODMICRO.2017.07.017.
[113] Holt, S.; Mukherjee, V.; Lievens, B.; Verstrepen, K. J.; Thevelein, J. M. Bioflavoring by
Non-Conventional Yeasts in Sequential Beer Fermentations. Food Microbiol. 2018, 72,
55–66. DOI: 10.1016/J.FM.2017.11.008.
[114] Figueiredo, B. I. C.; Saraiva, M. A. F.; Pimenta, P. P. S.; Testasicca, M. C. S.;
Sampaio, G. M. S.; Da Cunha, A. C.; Afonso, L. C. C.; de Queiroz, M. V.; Castro, I. M.;
Brandão, R. L. New Lager Brewery Strains Obtained by Crossing Techniques Using Cachaça
(Brazilian Spirit) Yeasts. Appl. Environ. Microbiol. 2017, 83(20), e01582–17. DOI: 10.1128/
AEM.01582-17.
[115] De Francesco, G.; Turchetti, B.; Sileoni, V.; Marconi, O.; Perretti, G. Screening of New
Strains of Saccharomycodes Ludwigii and Zygosaccharomyces Rouxii to Produce
Low-Alcohol Beer. J. Inst. Brew. 2015, 121(1), 113–121. DOI: 10.1002/jib.185.
FOOD REVIEWS INTERNATIONAL 25

[116] Tronchoni, J.; Curiel, J. A.; Morales, P.; Torres-Pérez, R.; Gonzalez, R. Early Transcriptional
Response to Biotic Stress in Mixed Starter Fermentations Involving Saccharomyces
Cerevisiae and Torulaspora Delbrueckii. Int. J. Food Microbiol. 2017, 241, 60–68. DOI:
10.1016/J.IJFOODMICRO.2016.10.017.
[117] Toh, D. W. K.; Chua, J. Y.; Liu, S. Q. Impact of Simultaneous Fermentation with
Saccharomyces Cerevisiae and Torulaspora Delbrueckii on Volatile and Non-Volatile
Constituents in Beer. LWT. 2018, 91, 26–33. DOI: 10.1016/J.LWT.2018.01.025.
[118] Saerens, S.; Swiegers, J. Production of Low-Alcohol or Alcohol-Free Beer with Pichia
Kluyveri Yeast Strains, March 2014.
[119] Osburn, K.; Amaral, J.; Metcalf, S. R.; Nickens, D. M.; Rogers, C. M.; Sausen, C.; Caputo, R.;
Miller, J.; Li, H.; Tennessen, J. M.;, et al. Primary Souring: A Novel Bacteria-Free Method for
Sour Beer Production. Food Microbiol. 2018, 70, 76–84. DOI: 10.1016/J.FM.2017.09.007.
[120] Lucio, O.; Pardo, I.; Krieger-Weber, S.; Heras, J. M.; Ferrer, S. Selection of Lactobacillus
Strains to Induce Biological Acidification in Low Acidity Wines. LWT - Food Sci. Technol.
2016, 73, 334–341. DOI: 10.1016/J.LWT.2016.06.031.
[121] Wang, S.; Li, S.; Zhao, H.; Gu, P.; Chen, Y.; Zhang, B.; Zhu, B. Acetaldehyde Released by
Lactobacillus Plantarum Enhances Accumulation of Pyranoanthocyanins in Wine during
Malolactic Fermentation. Food Res. Int. 2018, 108, 254–263. DOI: 10.1016/j.
foodres.2018.03.032.
[122] Brizuela, N. S.; Bravo-Ferrada, B. M.; Pozo-Bayón, M. Á.; Semorile, L.; Elizabeth
Tymczyszyn, E. Changes in the Volatile Profile of Pinot Noir Wines Caused by
Patagonian Lactobacillus Plantarum and Oenococcus Oeni Strains. Food Res. Int. 2018,
106, 22–28. DOI: 10.1016/J.FOODRES.2017.12.032.
[123] Gobbi, M.; Comitini, F.; Domizio, P.; Romani, C.; Lencioni, L.; Mannazzu, I.; Lachancea
Thermotolerans, C. M. And Saccharomyces Cerevisiae in Simultaneous and Sequential
Co-Fermentation: A Strategy to Enhance Acidity and Improve the Overall Quality of
Wine. Food Microbiol. 2013, 33(2), 271–281. DOI: 10.1016/j.fm.2012.10.004.
[124] Villena, M. A.; Iranzo, J. F. Ú.; Pérez, A. I. B. β-Glucosidase Activity in Wine Yeasts:
Application in Enology. Enzyme Microb. Technol. 2007, 40(3), 420–425. DOI: 10.1016/J.
ENZMICTEC.2006.07.013.
[125] Benito, S.; Morata, A.; Palomero, F.; González, M. C.; Suárez-Lepe, J. A. Formation of
Vinylphenolic Pyranoanthocyanins by Saccharomyces Cerevisiae and Pichia Guillermondii
in Red Wines Produced following Different Fermentation Strategies. Food Chem. 2011, 124
(1), 15–23. DOI: 10.1016/J.FOODCHEM.2010.05.096.
[126] Giovani, G.; Rosi, I.; Quantification, B. M. Characterization of Cell Wall Polysaccharides
Released by Non-Saccharomyces Yeast Strains during Alcoholic Fermentation. Int. J. Food
Microbiol. 2012, 160(2), 113–118. DOI: 10.1016/j.ijfoodmicro.2012.10.007.
[127] Viana, F.; Belloch, C.; Vallés, S.; Manzanares, P. Monitoring a Mixed Starter of
Hanseniaspora Vineae–Saccharomyces Cerevisiae in Natural Must: Impact on
2-Phenylethyl Acetate Production. Int. J. Food Microbiol. 2011, 151(2), 235–240. DOI:
10.1016/j.ijfoodmicro.2011.09.005.
[128] Anfang, N.; Brajkovich, M.; Goddard, M. R. Co-Fermentation with Pichia Kluyveri Increases
Varietal Thiol Concentrations in Sauvignon Blanc. Aust. J. Grape Wine Res. 2009, 15(1), 1–8.
DOI: 10.1111/j.1755-0238.2008.00031.x.
[129] Moreira, N.; Mendes, F.; Guedes de Pinho, P.; Hogg, T.; Vasconcelos, I. Heavy Sulphur
Compounds, Higher Alcohols and Esters Production Profile of Hanseniaspora Uvarum and
Hanseniaspora Guilliermondii Grown as Pure and Mixed Cultures in Grape Must. Int.
J. Food Microbiol. 2008, 124(3), 231–238. DOI: 10.1016/j.ijfoodmicro.2008.03.025.
[130] Beckner Whitener, M. E.; Carlin, S.; Jacobson, D.; Weighill, D.; Divol, B.; Conterno, L.; Du
Toit, M.; Vrhovsek, U. Early Fermentation Volatile Metabolite Profile of
Non-Saccharomyces Yeasts in Red and White Grape Must: A Targeted Approach. LWT -
Food Sci. Technol. 2015, 64(1), 412–422. DOI: 10.1016/j.lwt.2015.05.018.
[131] Sadineni, V.; Kondapalli, N.; Obulam, V. S. R. Effect of Co-Fermentation with Saccharomyces
Cerevisiae and Torulaspora Delbrueckii or Metschnikowia Pulcherrima on the Aroma and
26 G. VINICIUS DE MELO PEREIRA ET AL.

Sensory Properties of Mango Wine. Ann. Microbiol. 2012, 62(4), 1353–1360. DOI: 10.1007/
s13213-011-0383-6.
[132] Comitini, F.; Gobbi, M.; Domizio, P.; Romani, C.; Lencioni, L.; Mannazzu, I.; Ciani, M.
Selected Non-Saccharomyces Wine Yeasts in Controlled Multistarter Fermentations with
Saccharomyces Cerevisiae. Food Microbiol. 2011, 28(5), 873–882. DOI: 10.1016/j.
fm.2010.12.001.
[133] De Benedictis, M.; Bleve, G.; Grieco, F.; Tristezza, M.; Tufariello, M.; Grieco, F. An
Optimized Procedure for the Enological Selection of Non-Saccharomyces Starter Cultures.
Antonie Van Leeuwenhoek. 2011, 99(2), 189–200. DOI: 10.1007/s10482-010-9475-8.
[134] Rantsiou, K.; Dolci, P.; Giacosa, S.; Torchio, F.; Tofalo, R.; Torriani, S.; Suzzi, G.; Rolle, L.;
Cocolin, L. Candida Zemplinina Can Reduce Acetic Acid Produced by Saccharomyces
Cerevisiae in Sweet Wine Fermentations. Appl. Environ. Microbiol. 2012, 78(6),
1987–1994. DOI: 10.1128/AEM.06768-11.
[135] Tofalo, R.; Schirone, M.; Torriani, S.; Rantsiou, K.; Cocolin, L.; Perpetuini, G.; Suzzi, G.
Diversity of Candida Zemplinina Strains from Grapes and Italian Wines. Food Microbiol.
2012, 29(1), 18–26. DOI: 10.1016/J.FM.2011.08.014.
[136] Hu, K.; Jin, G.-J.; Xu, Y.-H.; Tao, Y.-S. Wine Aroma Response to Different Participation of
Selected Hanseniaspora Uvarum in Mixed Fermentation with Saccharomyces Cerevisiae.
Food Res. Int. 2018, 108, 119–127. DOI: 10.1016/j.foodres.2018.03.037.
[137] Sun, S. Y.; Chen, Z. X.; Jin, C. W. Combined Influence of Lactic Acid Bacteria Starter and
Final PH on the Induction of Malolactic Fermentation and Quality of Cherry Wines. LWT.
2018, 89, 449–456. DOI: 10.1016/J.LWT.2017.11.023.
[138] Jiang, J.; Sumby, K. M.; Sundstrom, J. F.; Grbin, P. R.; Jiranek, V. Directed Evolution of
Oenococcus Oeni Strains for More Efficient Malolactic Fermentation in a Multi-Stressor
Wine Environment. Food Microbiol. 2018, 73, 150–159. DOI: 10.1016/J.FM.2018.01.005.
[139] Benito, S.; Palomero, F.; Morata, A.; Calderón, F.; Suárez-Lepe, J. A. New Applications for
Schizosaccharomyces Pombe in the Alcoholic Fermentation of Red Wines. Int. J. Food Sci.
Technol. 2012, 47(10), 2101–2108. DOI: 10.1111/j.1365-2621.2012.03076.x.
[140] Barbosa, E. A.; Souza, M. T.; Diniz, R. H. S.; Godoy-Santos, F.; Faria-Oliveira, F.;
Correa, L. F. M.; Alvarez, F.; Coutrim, M. X.; Afonso, R. J. C. F.; Castro, I. M.;, et al.
Quality Improvement and Geographical Indication of Cachaça (Brazilian Spirit) by Using
Locally Selected Yeast Strains. J. Appl. Microbiol. 2016, 121(4), 1038–1051. DOI: 10.1111/
jam.13216.
[141] Gomes, F. C. O.; Silva, C. L. C.; Marini, M. M.; Oliveira, E. S.; Rosa, C. A. Use of Selected
Indigenous Saccharomyces Cerevisiae Strains for the Production of the Traditional Cachaça in
Brazil. J. Appl. Microbiol. 2007, 103(6), 2438–2447. DOI: 10.1111/j.1365-2672.2007.03486.x.
[142] Silva, C. L. C.; Vianna, C. R.; Cadete, R. M.; Santos, R. O.; Gomes, F. C. O.; Oliveira, E. S.;
Rosa, C. A.; Selection, G. Chemo-Sensory Evaluation of Flocculent Starter Culture Strains of
Saccharomyces Cerevisiae in the Large-Scale Production of Traditional Brazilian Cachaça.
Int. J. Food Microbiol. 2009, 131(2–3), 203–210. DOI: 10.1016/j.ijfoodmicro.2009.02.027.
[143] Duarte, W. F.; Amorim, J. C.; Schwan, R. F. The Effects of Co-Culturing Non-Saccharomyces
Yeasts with S. Cerevisiae on the Sugar Cane Spirit (Cachaça) Fermentation Process. Antonie Van
Leeuwenhoek. 2013, 103(1), 175–194. DOI: 10.1007/s10482-012-9798-8.
[144] de Souza, A. P. G.; Vicente, M.; de, A.; Klein, R. C.; Fietto, L. G.; Coutrim, M. X.;
Afonso, R. J.; de, C. F.; Araújo, L. D.; Da Silva, P. H. A.;, et al. Strategies to Select Yeast
Starters Cultures for Production of Flavor Compounds in Cachaça Fermentations. Antonie
Van Leeuwenhoek. 2012, 101(2), 379–392. DOI: 10.1007/s10482-011-9643-5.
[145] Campos, C. R.; Silva, C. F.; Dias, D. R.; Basso, L. C.; Amorim, H. V.; Schwan, R. F. Features
of Saccharomyces Cerevisiae as a Culture Starter for the Production of the Distilled Sugar
Cane Beverage, Cachaça in Brazil. J. Appl. Microbiol. 2009, 108(6), 1871–1879. DOI:
10.1111/j.1365-2672.2009.04587.x.
[146] Watanabe, T.; Owari, K.; Hori, K.; Takahashi, K. Selection of Koji Mold Strain for Making
Functional Miso as Rich Antimutagenic Activity. Nippon Shokuhin Kagaku Kogaku Kaishi.
2004, 51(12), 698–702. DOI: 10.3136/nskkk.51.698.
FOOD REVIEWS INTERNATIONAL 27

[147] Kubo, Y.; Rooney, A. P.; Tsukakoshi, Y.; Nakagawa, R.; Hasegawa, H.; Kimura, K. Phylogenetic
Analysis of Bacillus Subtilis Strains Applicable to Natto (Fermented Soybean) Production. Appl.
Environ. Microbiol. 2011, 77(18), 6463–6469. DOI: 10.1128/AEM.00448-11.
[148] Wei, Q.; Wolf-Hall, C.; Chang, K. C. Natto Characteristics as Affected by Steaming Time,
Bacillus Strain, and Fermentation Time. J. Food Sci. 2001, 66(1), 167–173. DOI: 10.1111/
j.1365-2621.2001.tb15601.x.
[149] Sugiyama, S.;. Selection of Micro-Organisms for Use in the Fermentation of Soy Sauce. Food
Microbiol. 1984, 1(4), 339–347. DOI: 10.1016/0740-0020(84)90067-4.
[150] Heskamp, M.-L.; Barz, W. Expression of Proteases by Rhizopus Species during Tempeh
Fermentation of Soybeans. Nahrung/Food. 1998, 42(01), 23–28. DOI: 10.1002/(SICI)1521-
3803(199802)42:01<23::AID-FOOD23>3.0.CO;2-3.
[151] Supriyanto,; Fujio, Y.; Hayakawa, I. Statistical Characterization of Tempeh Starter from the
Aroma Components of Soybean Tempeh. In Developments in Food Engineering; Yano, T.,
Matsuno, R., Nakamura, K. Eds.; Springer US: Boston, MA, 1994; pp 522–524. DOI: 10.1007/
978-1-4615-2674-2_167.
[152] Pereira, G. V. M.; Carvalho Neto, D. P.; Medeiros, A. B. P.; Soccol, V. T.; Neto, E.;
Woiciechowski, A. L.; Soccol, C. R. Potential of Lactic Acid Bacteria to Improve the
Fermentation and Quality of Coffee during On-Farm Processing. Int. J. Food Sci. Technol.
2016, 51(7), 1689–1695. DOI: 10.1111/ijfs.13142.
[153] Pereira, G. V.; de, M.; Soccol, V. T.; Pandey, A.; Medeiros, A. B. P.; Lara, J. M. R. A.;
Gollo, A. L.; Soccol, C. R. Isolation, Selection and Evaluation of Yeasts for Use in
Fermentation of Coffee Beans by the Wet Process. Int. J. Food Microbiol. 2014, 188,
60–66. DOI: 10.1016/j.ijfoodmicro.2014.07.008.
[154] Lee, L. W.; Tay, G. Y.; Cheong, M. W.; Curran, P.; Yu, B.; Liu, S. Q. Modulation of the
Volatile and Non-Volatile Profiles of Coffee Fermented with Yarrowia Lipolytica: II. Roasted
Coffee. LWT. 2017, 80, 32–42. DOI: 10.1016/J.LWT.2017.01.070.
[155] Silva, C. F.; Vilela, D. M.; de Souza Cordeiro, C.; Duarte, W. F.; Dias, D. R.; Schwan, R. F.
Evaluation of a Potential Starter Culture for Enhance Quality of Coffee Fermentation. World
J. Microbiol. Biotechnol. 2013, 29(2), 235–247. DOI: 10.1007/s11274-012-1175-2.
[156] Lee, L. W.; Cheong, M. W.; Curran, P.; Yu, B.; Liu, S. Q. Modulation of Coffee Aroma via
the Fermentation of Green Coffee Beans with Rhizopus Oligosporus: II. Effects of Different
Roast Levels. Food Chem. 2016, 211, 925–936. DOI: 10.1016/j.foodchem.2016.05.073.
[157] Pereira, G. V.; de, M.; Miguel, M. G.; da, C. P.; Ramos, C. L.; Schwan, R. F. Microbiological
and Physicochemical Characterization of Small-Scale Cocoa Fermentations and Screening of
Yeast and Bacterial Strains to Develop a Defined Starter Culture. Appl. Environ. Microbiol.
2012, 78(15), 5395–5405. DOI: 10.1128/AEM.01144-12.
[158] Lefeber, T.; Janssens, M.; Camu, N.; De Vuyst, L. Kinetic Analysis of Strains of Lactic Acid
Bacteria and Acetic Acid Bacteria in Cocoa Pulp Simulation Media toward Development of
a Starter Culture for Cocoa Bean Fermentation. Appl. Environ. Microbiol. 2010, 76(23),
7708–7716. DOI: 10.1128/AEM.01206-10.
[159] Lefeber, T.; Papalexandratou, Z.; Gobert, W.; Camu, N.; De Vuyst, L. On-Farm
Implementation of a Starter Culture for Improved Cocoa Bean Fermentation and Its
Influence on the Flavour of Chocolates Produced Thereof. Food Microbiol. 2012, 30(2),
379–392. DOI: 10.1016/j.fm.2011.12.021.
[160] Batista, N. N.; Ramos, C. L.; Dias, D. R.; Pinheiro, A. C. M.; Schwan, R. F. The Impact of
Yeast Starter Cultures on the Microbial Communities and Volatile Compounds in Cocoa
Fermentation and the Resulting Sensory Attributes of Chocolate. J. Food Sci. Technol. 2016,
53(2), 1101–1110. DOI: 10.1007/s13197-015-2132-5.
[161] Crafack, M.; Keul, H.; Eskildsen, C. E.; Petersen, M. A.; Saerens, S.; Blennow, A.; Skovmand-
Larsen, M.; Swiegers, J. H.; Petersen, G. B.; Heimdal, H.;, et al. Impact of Starter Cultures
and Fermentation Techniques on the Volatile Aroma and Sensory Profile of Chocolate. Food
Res. Int. 2014, 63, 306–316. DOI: 10.1016/J.FOODRES.2014.04.032.
28 G. VINICIUS DE MELO PEREIRA ET AL.

[162] Jespersen, L.; Nielsen, D.; Honholt, S.; Occurrence, J. M. Diversity of Yeasts Involved in
Fermentation of West African Cocoa Beans. FEMS Yeast Res. 2005, 5(4–5), 441–453. DOI:
10.1016/j.femsyr.2004.11.002.
[163] Sandhya, M. V. S.; Yallappa, B. S.; Varadaraj, M. C.; Puranaik, J.; Rao, L. J.; Janardhan, P.;
Murthy, P. S. Inoculum of the Starter Consortia and Interactive Metabolic Process in
Enhancing Quality of Cocoa Bean (Theobroma Cacao) Fermentation. LWT - Food Sci.
Technol. 2016, 65, 731–738. DOI: 10.1016/J.LWT.2015.09.002.
[164] Batista, N. N.; Ramos, C. L.; Ribeiro, D. D.; Pinheiro, A. C. M.; Schwan, R. F. Dynamic
Behavior of Saccharomyces Cerevisiae, Pichia Kluyveri and Hanseniaspora Uvarum during
Spontaneous and Inoculated Cocoa Fermentations and Their Effect on Sensory
Characteristics of Chocolate. LWT - Food Sci. Technol. 2015, 63(1), 221–227. DOI:
10.1016/J.LWT.2015.03.051.
[165] Gibbons, J. G.; Rinker, D. C. The Genomics of Microbial Domestication in the Fermented
Food Environment. Curr. Opin. Genet. Dev. 2015, 35, 1–8. DOI: 10.1016/J.GDE.2015.07.003.
[166] Gallone, B.; Mertens, S.; Gordon, J. L.; Maere, S.; Verstrepen, K. J.; Origins, S. J.;
Evolution, D. Diversity of Saccharomyces Beer Yeasts. Curr. Opin. Biotechnol. 2018, 49,
148–155. DOI: 10.1016/j.copbio.2017.08.005.
[167] Gallone, B.; Steensels, J.; Prahl, T.; Soriaga, L.; Saels, V.; Herrera-Malaver, B.; Merlevede, A.;
Roncoroni, M.; Voordeckers, K.; Miraglia, L.;, et al. Domestication and Divergence of
Saccharomyces Cerevisiae Beer Yeasts. Cell.2016, 166(6), 1397–1410.e16. DOI: 10.1016/j.
cell.2016.08.020.
[168] Fleet, G. H.;. Wine Yeasts for the Future. FEMS Yeast Res. 2008, 8(7), 979–995. DOI:
10.1111/j.1567-1364.2008.00427.x.
[169] Vendramin, V.; Treu, L.; Campanaro, S.; Lombardi, A.; Corich, V.; Genome
Comparison, G. A. Physiological Characterization of Eight Streptococcus Thermophilus
Strains Isolated from Italian Dairy Products. Food Microbiol. 2017, 63, 47–57. DOI:
10.1016/J.FM.2016.11.002.
[170] Saichana, N.; Matsushita, K.; Adachi, O.; Frébort, I.; Frebortova, J. Acetic Acid Bacteria:
A Group of Bacteria with Versatile Biotechnological Applications. Biotechnol. Adv. 2015, 33
(6), 1260–1271. DOI: 10.1016/j.biotechadv.2014.12.001.
[171] Ouattara, H. D.; Ouattara, H. G.; Droux, M.; Reverchon, S.; Nasser, W.; Niamke, S. L. Lactic
Acid Bacteria Involved in Cocoa Beans Fermentation from Ivory Coast: Species Diversity
and Citrate Lyase Production. Int. J. Food Microbiol. 2017, 256, 11–19. DOI: 10.1016/j.
ijfoodmicro.2017.05.008.
[172] Yousef, A. E.; Courtney, P. D. Basics of Stress Adaptation and Implications in New-
Generation Foods. In Microbial Stress Adaptation and Food Safety; Yousef, A.E., Juneja, V.
K., Eds.; CRC Press: Boca Raton, 2003, 2–8.
[173] Bauer, F. F.; Pretorius, I. S. Yeast Stress Response and Fermentation Efficiency: How to
Survive the Making of Wine -A Review. S. Afr. J. Enol. Vitic. 2000, 21, 27–46.
[174] Csonka, L. N.;. Physiological and Genetic Responses of Bacteria to Osmotic Stress. Microbiol.
Rev. 1989, 53(1), 121–147.
[175] Mas, A.; Torija, M. J.; García-Parrilla, M.; del, C.; Troncoso, A. M. Acetic Acid Bacteria and
the Production and Quality of Wine Vinegar. Sci. World J. 2014, 2014, 1–6. DOI: 10.1155/
2014/394671.
[176] Trček, J.; Mira, N. P.; Jarboe, L. R. Adaptation and Tolerance of Bacteria against Acetic Acid.
Appl. Microbiol. Biotechnol. 2015, 99(15), 6215–6229. DOI: 10.1007/s00253-015-6762-3.
[177] Trček, J.; Ramuš, J.; Raspor, P. Phenotypic Characterization and RAPD-PCR Profiling of
Acetobacter Sp. Isolated from Spirit Vinegar Production. Food Technol. Biotechnol. 1997, 35
(1), 63–67.
[178] Meersman, E.; Steensels, J.; Struyf, N.; Paulus, T.; Saels, V.; Mathawan, M.; Allegaert, L.;
Vrancken, G.; Verstrepen, K. J. Tuning Chocolate Flavor through Development of
Thermotolerant Saccharomyces Cerevisiae Starter Cultures with Increased Acetate Ester
Production. Appl. Environ. Microbiol. 2016, 82(2), 732–746. DOI: 10.1128/AEM.02556-15.
FOOD REVIEWS INTERNATIONAL 29

[179] Sieuwerts, S.; Molenaar, D.; van Hijum, S. A. F. T.; Beerthuyzen, M.; Stevens, M. J. A.;
Janssen, P. W. M.; Ingham, C. J.; de Bok, F. A. M.; de Vos, W. M.; van Hylckama
Vlieg, J. E. T. Mixed-Culture Transcriptome Analysis Reveals the Molecular Basis of
Mixed-Culture Growth in Streptococcus Thermophilus and Lactobacillus Bulgaricus. Appl.
Environ. Microbiol. 2010, 76(23), 7775–7784. DOI: 10.1128/AEM.01122-10.
[180] Parsons, J. B.; Rock, C. O. Bacterial Lipids: Metabolism and Membrane Homeostasis. Prog.
Lipid Res. 2013, 52(3), 249–276. DOI: 10.1016/J.PLIPRES.2013.02.002.
[181] Tolner, B.; Poolman, B.; Konings, W. N. Adaptation of Microorganisms and Their Transport
Systems to High Temperatures. Comp. Biochem. Physiol. A. Physiol. 1997, 118(3), 423–428.
DOI: 10.1016/S0300-9629(97)00003-0.
[182] Bai, F. W.; Anderson, W. A.; Moo-Young, M. Ethanol Fermentation Technologies from
Sugar and Starch Feedstocks. Biotechnol. Adv. 2008, 26(1), 89–105. DOI: 10.1016/J.
BIOTECHADV.2007.09.002.
[183] Soubeyrand, V.; Julien, A.; Sablayrolles, J.-M. Rehydration Protocols for Active Dry Wine
Yeasts and the Search for Early Indicators of Yeast Activity. Am. J. Enol. Vitic. 2006, 57, 4.
[184] Berthels, N.; Corderootero, R.; Bauer, F.; Thevelein, J.; Pretorius, I. S. Discrepancy in
Glucose and Fructose Utilisation during Fermentation by Wine Yeast Strains. FEMS Yeast
Res. 2004, 4(7), 683–689. DOI: 10.1016/j.femsyr.2004.02.005.
[185] Chatonnet, P.; Dubourdieu, D.; Boidron, J. N. The Influence of Brettanomyces/Dekkera Sp.
Yeasts and Lactic Acid Bacteria on the Ethylphenol Content of Red Wines. Am. J. Enol.
Vitic. 1995, 46(4), 463–468.
[186] Adachi, O.; Moonmangmee, D.; Toyama, H.; Yamada, M.; Shinagawa, E.; Matsushita, K.
New Developments in Oxidative Fermentation. Appl. Microbiol. Biotechnol. 2003, 60(6),
643–653. DOI: 10.1007/s00253-002-1155-9.
[187] Gullo, M.; Giudici, P. Acetic Acid Bacteria in Traditional Balsamic Vinegar: Phenotypic
Traits Relevant for Starter Cultures Selection. Int. J. Food Microbiol. 2008, 125(1), 46–53.
DOI: 10.1016/j.ijfoodmicro.2007.11.076.
[188] Solieri, L.; Chand Dakal, T.; Giudici, P. Zygosaccharomyces Sapae Sp. Nov., Isolated from
Italian Traditional Balsamic Vinegar. Int. J. Syst. Evol. Microbiol. 2013, 63(Pt 1), 364–371.
DOI: 10.1099/ijs.0.043323-0.
[189] Sievers, M.; Teuber, M. The Microbiology and Taxonomy of Acetobacter Europaeus in
Commercial Vinegar Production. J. Appl. Bacteriol. 1995, 79, 84–95.
[190] Trcek, J.; Toyama, H.; Czuba, J.; Misiewicz, A.; Matsushita, K. Correlation between Acetic
Acid Resistance and Characteristics of PQQ-Dependent ADH in Acetic Acid Bacteria. Appl.
Microbiol. Biotechnol. 2006, 70(3), 366–373. DOI: 10.1007/s00253-005-0073-z.
[191] Steiner, P.; Sauer, U. Long-Term Continuous Evolution of Acetate ResistantAcetobacter
Aceti. Biotechnol. Bioeng. 2003, 84(1), 40–44. DOI: 10.1002/bit.10741.
[192] Mamlouk, D.; Gullo, M. Acetic Acid Bacteria: Physiology and Carbon Sources Oxidation.
Indian J. Microbiol. 2013, 53(4), 377–384. DOI: 10.1007/s12088-013-0414-z.
[193] Solieri, L.; Giudici, P. Yeast Starter Selection for Traditional Balsamic Vinegar. Ind. Delle
Bevande. 2005, 34, 526–531.
[194] Wang, Z.-M.; Lu, Z.-M.; Shi, J.-S.; Xu, Z.-H. Exploring Flavour-Producing Core Microbiota
in Multispecies Solid-State Fermentation of Traditional Chinese Vinegar. Sci. Rep. 2016, 6
(1), 26818. DOI: 10.1038/srep26818.
[195] Buckenhüskes, H. J.;. Selection Criteria for Lactic Acid Bacteria to Be Used as Starter
Cultures for Various Food Commodities. FEMS Microbiol. Rev. 1993, 12(1–3), 253–271.
DOI: 10.1016/0168-6445(93)90067-J.
[196] Bacus, J. N.;. Fermented Meat and Poultry Products. Adv. Meat Res. 1986, 2, 123–164.
[197] Breidt, F.; McFeeters, R. F.; Perez-Diaz, I.; Lee, C. H. Fermented Vegetables. In Food
Microbiology: Fundamentals and Frontiers; Doyle, M.P., Buchanan, R.L., Eds.; ASM Press:
Washington, 2013; pp 841–856.
[198] Kargozari, M.; Moini, S.; Akhondzadeh Basti, A.; Emam-Djomeh, Z.; Gandomi, H.; Revilla
Martin, I.; Ghasemlou, M.; Carbonell-Barrachina, Á. A. Effect of Autochthonous Starter
Cultures Isolated from Siahmazgi Cheese on Physicochemical, Microbiological and Volatile
30 G. VINICIUS DE MELO PEREIRA ET AL.

Compound Profiles and Sensorial Attributes of Sucuk, a Turkish Dry-Fermented Sausage.


Meat Sci. 2014, 97(1), 104–114. DOI: 10.1016/J.MEATSCI.2014.01.013.
[199] Erkkilä, S.; Petäjä, E. Screening of Commercial Meat Starter Cultures at Low PH and in the
Presence of Bile Salts for Potential Probiotic Use. Meat Sci. 2000, 55(3), 297–300. DOI:
10.1016/S0309-1740(99)00156-4.
[200] Villani, F.; Casaburi, A.; Pennacchia, C.; Filosa, L.; Russo, F.; Ercolini, D. Microbial Ecology
of the Soppressata of Vallo Di Diano, a Traditional Dry Fermented Sausage from Southern
Italy, and in Vitro and in Situ Selection of Autochthonous Starter Cultures. Appl. Environ.
Microbiol. 2007, 73(17), 5453–5463. DOI: 10.1128/AEM.01072-07.
[201] Lorenzo, J. M.; Munekata, P. E. S.; Domínguez, R. Role of Autochthonous Starter Cultures in
the Reduction of Biogenic Amines in Traditional Meat Products. Curr. Opin. Food Sci. 2017,
14, 61–65. DOI: 10.1016/J.COFS.2017.01.009.
[202] Bover-Cid, S.; Hugas, M.; Izquierdo-Pulido, M.; Vidal-Carou, M. C. Reduction of Biogenic
Amine Formation Using a Negative Amino Acid-Decarboxylase Starter Culture for
Fermentation of Fuet Sausages. J. Food Prot. 2000, 63(2), 237–243. DOI: 10.4315/0362-
028X-63.2.237.
[203] Komprda, T.; Smělá, D.; Pechová, P.; Kalhotka, L.; Štencl, J.; Klejdus, B. Effect of Starter
Culture, Spice Mix and Storage Time and Temperature on Biogenic Amine Content of Dry
Fermented Sausages. Meat Sci. 2004, 67, 607–616. DOI: 10.1016/j.meatsci.2004.01.003.
[204] 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(3),
846–853. DOI: 10.1016/j.foodchem.2011.05.033.
[205] Yongjin, H.; Wenshui, X.; Xiaoyong, L. Changes in Biogenic Amines in Fermented Silver
Carp Sausages Inoculated with Mixed Starter Cultures. Food Chem. 2007, 104, 188–195.
DOI: 10.1016/j.foodchem.2006.11.023.
[206] Dicks, L. M. T.; Mellett, F. D.; Hoffman, L. C. Use of Bacteriocin-Producing Starter Cultures
of Lactobacillus Plantarum and Lactobacillus Curvatus in Production of Ostrich Meat
Salami. Meat Sci. 2004, 66(3), 703–708. DOI: 10.1016/J.MEATSCI.2003.07.002.
[207] Noonpakdee, W.; Santivarangkna, C.; Jumriangrit, P.; Sonomoto, K.; Panyim, S. Isolation of
Nisin-Producing Lactococcus Lactis WNC 20 Strain from Nham, a Traditional Thai
Fermented Sausage. Int. J. Food Microbiol. 2003, 81, 137–145.
[208] Steinkraus, K. H.;. Lactic Acid Fermentation in the Production of Foods from Vegetables,
Cereals and Legumes. Antonie van Leeuwenhoek. 1983, 49, 337–348.
[209] Daeschel, M. A.; McFeeters, R. F.; Fleming, H. P.; Klaenhammer, T. R.; Sanozky, R. B.
Mutation and Selection of Lactobacillus Plantarum Strains that Do Not Produce Carbon
Dioxide from Malate. Appl. Environ. Microbiol. 1984, 47(2), 419–420.
[210] Kondyli, E.; Katsiari, M. C.; Masouras, T.; Voutsinas, L. P. Free Fatty Acids and Volatile
Compounds of Low-Fat Feta-Type Cheese Made with a Commercial Adjunct Culture. Food
Chem. 2002, 79, 199–205. DOI: 10.1016/S0308-8146(02)00132-2.
[211] McSweeney, P. L. H.; Sousa, M. J. Biochemical Pathways for the Production of Flavour
Compounds in Cheeses during Ripening: A Review. Lait. 2000, 80(3), 293–324. DOI:
10.1051/lait:2000127.
[212] Smid, E. J.; Kleerebezem, M. Production of Aroma Compounds in Lactic Fermentations.
Annu. Rev. Food Sci. Technol. 2014, 5(1), 313–326. DOI: 10.1146/annurev-food-030713-
092339.
[213] Smit, G.; Smit, B.; Engels, W. Flavour Formation by Lactic Acid Bacteria and Biochemical
Flavour Profiling of Cheese Products. FEMS Microbiol. Rev. 2005, 29(3), 591–610. DOI:
10.1016/j.femsre.2005.04.002.
[214] Cheng, H.;. Volatile Flavor Compounds in Yogurt: A Review. Crit. Rev. Food Sci. Nutr. 2010,
50(10), 938–950. DOI: 10.1080/10408390903044081.
[215] Guldfeldt, L. U.; Sorensen, K. I.; Stroman, P.; Behrndt, H.; Williams, D.; Johansen, E. Effect
of Starter Cultures with a Genetically Modified Peptidolytic or Lytic System on Cheddar
Cheese Ripening. Int. Dairy J. 2001, 11, 373–382. DOI: 10.1016/S0958-6946(01)00066-8.
FOOD REVIEWS INTERNATIONAL 31

[216] Menéndez, S.; Centeno, J. A.; Godínez, R.; Rodríguez-Otero, J. L. Effects of Lactobacillus
Strains on the Ripening and Organoleptic Characteristics of Arzúa-Ulloa Cheese. Int. J. Food
Microbiol. 2000, 59(1–2), 37–46.
[217] Nieto-Arribas, P.; Seseña, S.; Poveda, J. M.; Palop, L.; Genotypic, C. L. Technological
Characterization of Leuconostoc Isolates to Be Used as Adjunct Starters in Manchego
Cheese Manufacture. Food Microbiol. 2010, 27, 85–93. DOI: 10.1016/j.fm.2009.08.006.
[218] Ott, A.; Fay, L. B.; Determination, C. A. Origin of the Aroma Impact Compounds of Yogurt
Flavor. J. Agric. Food Chem. 1997, 45(3), 850–858. DOI: 10.1021/JF960508E.
[219] Tamime, A. Y.; Robinson, R. K. In Yoghurt: Science and Technology; CRC Press: Boca Raton,
1985.
[220] Liong, M.-T.; Lee, B.-H.; Choi, S.-B.; Lew, L.-C.; Lau, A.-S.-Y.; Daliri, E. B.-M. Cholesterol-
Lowering Effects of Probiotics and Prebiotics. In Probiotics and Prebiotics: Current Research
and Future Trends; Venema, K., Do Carmo, A.P. Eds.; Caister Academic Press: Norfolk,
2015; pp 429–446. DOI: 10.21775/9781910190098.29.
[221] Marchesi, J. R.; Adams, D. H.; Fava, F.; Hermes, G. D. A.; Hirschfield, G. M.; Hold, G.;
Quraishi, M. N.; Kinross, J.; Smidt, H.; Tuohy, K. M.;, et al. The Gut Microbiota and Host
Health: A New Clinical Frontier. Gut.2016, 65(2), 330–339. DOI: 10.1136/gutjnl-2015-
309990.
[222] Zoumpopoulou, G.; Pot, B.; Tsakalidou, E.; Dairy Probiotics:, P. K. Beyond the Role of
Promoting Gut and Immune Health. Int. Dairy J. 2017, 67, 46–60. DOI: 10.1016/J.
IDAIRYJ.2016.09.010.
[223] Pereira, G.; V. de, M.; Coelho, B. D. O.; 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, in press. DOI: 10.1016/j.biotechadv.2018.09.003.
[224] Kołożyn-Krajewska, D.; Dolatowski, Z. J. Probiotic Meat Products and Human Nutrition.
Process Biochem. 1761–1772, 2012(47). DOI: 10.1016/j.procbio.2012.09.017.
[225] Ross, R. P.; Desmond, C.; Fitzgerald, G. F.; Stanton, C. Overcoming the Technological
Hurdles in the Development of Probiotic Foods. J. Appl. Microbiol. 2005, 98(6),
1410–1417. DOI: 10.1111/j.1365-2672.2005.02654.x.
[226] Da Conceição, L. E. F. R.; Saraiva, M. A. F.; Diniz, R. H. S.; Oliveira, J.; Barbosa, G. D.;
Alvarez, F.; Da Mata Correa, L. F.; Mezadri, H.; Coutrim, M. X.; Afonso, R. J.;, et al.
Biotechnological Potential of Yeast Isolates from Cachaça: The Brazilian Spirit. J. Ind.
Microbiol. Biotechnol.2015, 42(2), 237–246. DOI: 10.1007/s10295-014-1528-y.
[227] Torriani, S.; Felis, G. E.; Selection Criteria, F. F. Tools for Malolactic Starters Development:
An Update. Ann. Microbiol. 2011, 61(1), 33–39. DOI: 10.1007/s13213-010-0072-x.
[228] Quirós, M.; Rojas, V.; Gonzalez, R.; Morales, P. Selection of Non-Saccharomyces Yeast
Strains for Reducing Alcohol Levels in Wine by Sugar Respiration. Int. J. Food Microbiol.
2014, 181, 85–91. DOI: 10.1016/J.IJFOODMICRO.2014.04.024.
[229] Contreras, A.; Hidalgo, C.; Henschke, P. A.; Chambers, P. J.; Curtin, C.; Varela, C.
Evaluation of Non-Saccharomyces Yeasts for the Reduction of Alcohol Content in Wine.
Appl. Environ. Microbiol. 2014, 80(5), 1670–1678. DOI: 10.1128/AEM.03780-13.
[230] Rossouw, D.; Bauer, F. F. Exploring the Phenotypic Space of Non-Saccharomyces Wine
Yeast Biodiversity. Food Microbiol. 2016, 55, 32–46. DOI: 10.1016/J.FM.2015.11.017.
[231] Englezos, V.; Rantsiou, K.; Torchio, F.; Rolle, L.; Gerbi, V.; Cocolin, L. Exploitation of the
Non-Saccharomyces Yeast Starmerella Bacillaris (Synonym Candida Zemplinina) in Wine
Fermentation: Physiological and Molecular Characterizations. Int. J. Food Microbiol. 2015,
199, 33–40. DOI: 10.1016/j.ijfoodmicro.2015.01.009.
[232] Gobbi, M.; De Vero, L.; Solieri, L.; Comitini, F.; Oro, L.; Giudici, P.; Ciani, M. Fermentative
Aptitude of Non-Saccharomyces Wine Yeast for Reduction in the Ethanol Content in Wine.
Eur. Food Res. Technol. 2014, 239(1), 41–48. DOI: 10.1007/s00217-014-2187-y.
[233] López, M. C.; Mateo, J. J.; Maicas, S. Screening of β-Glucosidase and β-Xylosidase Activities
in Four Non- Saccharomyces Yeast Isolates. J. Food Sci. 2015, 80(8), C1696–C1704. DOI:
10.1111/1750-3841.12954.
32 G. VINICIUS DE MELO PEREIRA ET AL.

[234] Vicente, M. D. A.; Fietto, L. G.; Castro, I.; de, M.; Gonçalves Dos Santos, A. N.;
Coutrim, M. X.; Brandão, R. L. Isolation of Saccharomyces Cerevisiae Strains Producing
Higher Levels of Flavoring Compounds for Production of “Cachaça” the Brazilian Sugarcane
Spirit. Int. J. Food Microbiol. 2006, 108(1), 51–59. DOI: 10.1016/J.
IJFOODMICRO.2005.10.018.
[235] Kaewkrod, A.; Niamsiri, N.; Likitwattanasade, T.; Lertsiri, S. Activities of Macerating
Enzymes are Useful for Selection of Soy Sauce Koji. LWT. 2018, 89, 735–739. DOI:
10.1016/J.LWT.2017.11.020.
[236] Luh, B. S.;. Industrial Production of Soy Sauce. J. Ind. Microbiol. 1995, 14(6), 467–471. DOI:
10.1007/BF01573959.
[237] Kanbe, C.; Uchida, K. Citrate Metabolism by Pediococcus Halophilus. Appl. Environ.
Microbiol. 1987, 53(6), 1257–1262.
[238] Hesseltine, C. W.; Miso, S. K., III. Pure Culture Fermentation with Saccharomyces Rouxii.
Appl. Microbiol. 1961, 9(6), 515–518.
[239] Omolola, M. O.; Otunola, E. T. Preliminary Studies on Tempeh Flour Produced from Three
Different Rhizopus Species. Int. J. Biotechnol. Food Sci. 2013, 1(5), 90–96.
[240] Starzyńska-Janiszewska, A.; Stodolak, B.; Mickowska, B. Effect of Controlled Lactic Acid
Fermentation on Selected Bioactive and Nutritional Parameters of Tempeh Obtained from
Unhulled Common Bean (Phaseolus Vulgaris) Seeds. J. Sci. Food Agric. 2014, 94(2), 359–366.
DOI: 10.1002/jsfa.6385.
[241] Keuth, S.; Bisping, B. Formation of Vitamins by Pure Cultures of Tempe Moulds and
Bacteria during the Tempe Solid Substrate Fermentation. J. Appl. Bacteriol. 1993, 75(5),
427–434. DOI: 10.1111/j.1365-2672.1993.tb02798.x.
[242] De Vuyst, L.; Weckx, S. The Cocoa Bean Fermentation Process: From Ecosystem Analysis to
Starter Culture Development. J. Appl. Microbiol. 2016, 121(1), 5–17. DOI: 10.1111/
jam.13045.
[243] Evangelista, S. R.; Miguel, M. G. C. P.; Cordeiro, C. S.; Silva, C. F.; Pinheiro, A. C. M.;
Schwan, R. F. Inoculation of Starter Cultures in a Semi-Dry Coffee (Coffea Arabica)
Fermentation Process. Food Microbiol. 2014, 44, 87–95. DOI: 10.1016/j.fm.2014.05.013.
[244] Evangelista, S. R.; Silva, C. F.; Miguel, M. G. P.; da, C.; Cordeiro, C.; de, S.;
Pinheiro, A. C. M.; Duarte, W. F.; Schwan, R. F. Improvement of Coffee Beverage Quality
by Using Selected Yeasts Strains during the Fermentation in Dry Process. Food Res. Int.
2014, 61, 183–195. DOI: 10.1016/j.foodres.2013.11.033.
[245] Laulund, S.; Wind, A.; Derkx, P. M. F.; Regulatory, Z. V. Safety Requirements for Food
Cultures. Microorganisms. 2017, 5, 28. DOI: 10.3390/microorganisms5020028.
[246] Ricci, A.; Allende, A.; Bolton, D.; Chemaly, M.; Davies, R.; Girones, R.; Koutsoumanis, K.;
Salvador, P.; Lindqvist, R.; Nørrung, B.;, et al. Update of the List of QPS-Recommended
Biological Agents Intentionally Added to Food or Feed as Notified to EFSA 8: Suitability of
Taxonomic Units Notified to EFSA until March 2018. Efsa J.2018, 16(7), 5315. DOI:
10.2903/j.efsa.2018.5315.
[247] Barlow, S.; Chesson, A.; Collins, J.; Fernandes, T.; Flynn, A.; Hardy, T.; Jansson, B.;
Knaap, A.; Kuiper, H.; Le, N. P.;; et al. Opinion of the Scientific Committee on a Request
from EFSA Related to a Generic Approach to the Safety Assessment by EFSA of
Microorganisms Used in Food/Feed and the Production of Food/Feed Additives. Efsa J.
2005, 226, 1–12.
[248] FDA. Microorganisms & Microbial-Derived Ingredients Used in Food (Partial List).
[249] Ohi, H.; Okazaki, N.; Uno, S.; Miura, M.; Hiramatsu, R. Chromosomal DNA Patterns and
Gene Stability of Pichia Pastoris. Yeast. 1998, 14(10), 895–903. DOI: 10.1002/(SICI)1097-
0061(199807)14:10<895::AID-YEA288>3.0.CO;2-9.
[250] Vancanneyt, M.; Huys, G.; Lefebvre, K.; Vankerckhoven, V.; Goossens, H.; Swings, J.
Intraspecific Genotypic Characterization of Lactobacillus Rhamnosus Strains Intended for
Probiotic Use and Isolates of Human Origin. Appl. Environ. Microbiol. 2006, 72(8),
5376–5383. DOI: 10.1128/AEM.00091-06.
FOOD REVIEWS INTERNATIONAL 33

[251] Grunert, K. G.; Bredahl, L.; Scholderer, J. Four Questions on European Consumers ’
Attitudes toward the Use of Genetic Modification in Food Production. Innov. Food Sci.
Emerg. Technol. 2003, 4, 435–445. DOI: 10.1016/S1466-8564.
[252] Mills, S.; Sullivan, O. O.; Hill, C.; Fitzgerald, G.; Ross, R. P. The Changing Face of Dairy
Starter Culture Research : From Genomics to Economics. Int. J. Dairy Technol. 2010, 63(2),
149–170. DOI: 10.1111/j.1471-0307.2010.00563.x.
[253] de Vos, W. M.;. Safe and Sustainable Systems for Food-Grade Fermentations by Genetically
Modified Lactic Acid Bacteria. Int. Dairy J. 1999, 9, 3–10. DOI: 10.1016/S0958-6946(99)
00038-2.
[254] Howell, K. S.; Klein, M.; Swiegers, J. H.; Hayasaka, Y.; Elsey, G. M.; Fleet, G. H.; Høj, P. B.;
Pretorius, I. S.; Lopes, M. A. B. Genetic Determinants of Volatile-Thiol Release by
Saccharomyces Cerevisiae during Wine Fermentation. Appl. Environ. Microbiol. 2005, 71
(9), 5420–5426. DOI: 10.1128/AEM.71.9.5420.
[255] Wu, Q.; Tun, H. M.; Leung, F. C.; Shah, N. P. Genomic Insights into High
Exopolysaccharide-Producing Dairy Starter Bacterium Streptococcus Thermophilus ASCC
1275. Sci. Rep. 2014, 4, 4974. DOI: 10.1038/srep04974.
[256] Kastner, S.; Perreten, V.; Bleuler, H.; Hugenschmidt, G.; Lacroix, C.; Meile, L. Antibiotic
Susceptibility Patterns and Resistance Genes of Starter Cultures and Probiotic Bacteria Used
in Food. Syst. Appl. Microbiol. 2006, 29, 145–155. DOI: 10.1016/j.syapm.2005.07.009.
[257] Broadbent, J. R.; Hughes, J. E.; Welker, D. L.; Tompkins, T. A.; Steele, L. Complete Genome
Sequence for Lactobacillus Helveticus CNRZ 32, an Industrial Cheese Starter and Cheese
Flavor Adjunct. Genome Announc. 2013, 1(4), 1–2. DOI: 10.1128/genomeA.00590-13.
Copyright.
[258] Rosenstein, R.; Nerz, C.; Biswas, L.; Resch, A.; Raddatz, G.; Schuster, S. C.; Götz, F. Genome
Analysis of the Meat Starter Culture Bacterium Staphylococcus Carnosus TM300. Appl.
Environ. Microbiol. 2009, 75(3), 811–822. DOI: 10.1128/AEM.01982-08.
[259] Ferrocino, I.; Cocolin, L. Current Perspectives in Food-Based Studies Exploiting
Multi-Omics Approaches. Curr. Opin. Food Sci. 2017, 13, 10–15. DOI: 10.1016/j.
cofs.2017.01.002.

You might also like