Next Article in Journal
An Update on the Pivotal Roles of Probiotics, Their Components, and Metabolites in Preventing Colon Cancer
Next Article in Special Issue
Characterization of the Bacterial Composition of 47 Fermented Foods in Sweden
Previous Article in Journal
Possibility of Using Different Calcium Compounds for the Manufacture of Fresh Acid Rennet Cheese from Goat’s Milk
Previous Article in Special Issue
Temperature, Salinity and Garlic Additive Shape the Microbial Community during Traditional Beetroot Fermentation Process
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Latest Trends in Industrial Vinegar Production and the Role of Acetic Acid Bacteria: Classification, Metabolism, and Applications—A Comprehensive Review

by
Juan J. Román-Camacho
1,
Isidoro García-García
2,*,
Inés M. Santos-Dueñas
2,
Teresa García-Martínez
1 and
Juan C. Mauricio
1
1
Department of Agricultural Chemistry, Edaphology and Microbiology, Agrifood Campus of International Excellence ceiA3, University of Córdoba, 14014 Córdoba, Spain
2
Department of Inorganic Chemistry and Chemical Engineering, Agrifood Campus of International Excellence ceiA3, Nano Chemistry Institute (IUNAN), University of Córdoba, 14014 Córdoba, Spain
*
Author to whom correspondence should be addressed.
Foods 2023, 12(19), 3705; https://doi.org/10.3390/foods12193705
Submission received: 8 September 2023 / Revised: 3 October 2023 / Accepted: 6 October 2023 / Published: 9 October 2023
(This article belongs to the Special Issue The Microbial Community and Its Functions in Fermented Foods)

Abstract

:
Vinegar is one of the most appreciated fermented foods in European and Asian countries. In industry, its elaboration depends on numerous factors, including the nature of starter culture and raw material, as well as the production system and operational conditions. Furthermore, vinegar is obtained by the action of acetic acid bacteria (AAB) on an alcoholic medium in which ethanol is transformed into acetic acid. Besides the highlighted oxidative metabolism of AAB, their versatility and metabolic adaptability make them a taxonomic group with several biotechnological uses. Due to new and rapid advances in this field, this review attempts to approach the current state of knowledge by firstly discussing fundamental aspects related to industrial vinegar production and then exploring aspects related to AAB: classification, metabolism, and applications. Emphasis has been placed on an exhaustive taxonomic review considering the progressive increase in the number of new AAB species and genera, especially those with recognized biotechnological potential.

Graphical Abstract

1. Introduction

Since time immemorial, vinegar has been a fermented foodstuff widely used by mankind as a part of the diet and as a preservative, condiment, and flavoring additive. Because of its bio-healthy properties, vinegar was even traditionally used in ancient medicine [1,2]. Currently, there is a great variety of vinegars around the world, depending on the starter microbial composition, the raw material, and the technical methods used for production [3,4]. From an industrial point of view, the elaboration of vinegar is performed from an alcoholic medium in which a mixed culture of acetic acid bacteria (AAB) is used to carry out a biotransformation process of ethanol into acetic acid, which occurs in specific bioreactors (acetators) [5,6]. Despite the high quality of the final products obtained through traditional systems, including mainly solid-state fermentation and surface culture, their numerous disadvantages, including low efficiency, the slowness of the process, and a lack of control of the operating conditions, have led to the use of the submerged culture system [7,8]. The success of this method, widely implemented in Western and European countries, lies in the high yield and speed of the process under controlled stirring conditions due to the efficiency of mass transfer and continuous vigorous aeration [7,9]. In this sense, the choice of a working mode for the acetators also ensures a suitable environment for the development and activity of AAB, and the control and monitoring of fermentation conditions are some of the fundamental aspects to consider [5,7,10].
On the other hand, vinegar production would not be possible without the activity of the acetic acid bacteria (AAB). These microorganisms, strictly aerobic, can be found in a wide variety of natural and industrial environments. Their versatility and metabolic adaptability make them a taxonomic group of high interest for studying the optimization of obtaining multiple products, with acetic acid as one of the main components, as well as the essential mechanisms that allow them to grow under harsh conditions [11,12,13]. In this sense, the role of their membrane-bound and soluble dehydrogenase system may offer new opportunities in the development of innovative processes based on their capability to carry out the incomplete oxidation of several substrates, including alcohols, sugars, and sugar alcohols, for the production of organic acids [11,14]. Furthermore, the ability of AAB to produce exopolysaccharides is also of great interest for both research and industrial purposes, with some strains considered model organisms for understanding the mechanisms of cellulose synthesis. Moreover, at present, these are the most efficient microorganisms for producing them under controlled conditions [15,16,17]. Finally, the current state of omic technologies and efficient molecular modification methods may be applied to increase the understanding of physiological behavior and the characterization of new strains recovered from these complex media, as well as to exploit the full potential of AAB for producing vinegar and other related bioproducts [6,18,19,20,21,22].
Taking all the above into account, this review aims to provide an update containing some of the recent advances and challenges that have arisen around the main variables controlling industrial vinegar production as well as the characterization of the main microorganisms responsible for the process: AAB. Regarding the latter and considering recent research, which includes a progressive increase in the number of species and the description of new genera, especially those with recognized biotechnological potential, a taxonomic review has been carried out in this work.

2. The Historical Context of Vinegar: Origin and Uses

Vinegar has its origin in ancient civilizations. Concretely, the first testimony written on the use of vinegar comes from ancient Babylon, about 5000 years ago, when it was employed as a food preservative. Vinegar was “discovered” fortuitously when undisturbed wine stored in the open air turned spontaneously into vinegar [23]. Due to this phenomenon, known as “wine pitting”, and its sour taste, vinegar has been considered historically as a byproduct with poor commercial interest. However, the numerous applications and benefits of vinegar have been disclosed by mankind throughout history. Hippocrates (460–377 BCE) recommended using vinegar for cleaning ulcerations and for the treatment of sores [24]. Long afterward, in the 10th century, Sung Tse implemented the use of vinegar as a hand-washing agent to prevent infections, which led to an important development in the field of forensic medicine in China [1,25]. In the 18th century, American medical practitioners used vinegar to treat many ailments, including stomachache, high fever, and edema among others [1,24]. Nowadays, vinegar is widely consumed all over the world both directly and indirectly, as it is included in a great variety of dishes including sauces, ketchup, and mayonnaise [1].
Although vinegar has been traditionally used as a flavoring and food preservative, several studies evidence its nutritional potential effects, which can directly affect the health of consumers. This is due, in great part, to the healthy properties that acetic acid (the main constituent of vinegar) can exert on the human liver and gastrointestinal tract [26]. As a consequence, some of the benefits of vinegar may include appetite stimulation, recovery from exhaustion, antioxidant activity, lower lipid content in blood, and the regulation of blood pressure, which, in turn, have an effect on biomarkers for several diseases such as obesity, cancer, diabetes, and hypertension, among others (see Figure 1) [25,26,27,28,29,30,31,32].

3. Varieties of Vinegar

There is a great variety of vinegars around the world whose organoleptic properties are conferred by the starter microbial composition, raw material, and technical methods used for their production [3,4]. Vinegar has an alcoholic origin; it usually comes from the processing of vegetables or fruits [25]. In this way, several raw materials can be processed and used as acetification substrates, including wines, spirits, cereal grains (rice wines and malts), and fruit juices, among others. Other raw materials of animal origin can also be used, such as whey or honey [1,33,34]. This section will describe some of the most widely used varieties of vinegar in the world for which the raw materials are specific to particular regions and confer on the final product exceptional organoleptic properties and high quality.

3.1. Mediterranean Vinegar: Wine and Balsamic Vinegar

In Mediterranean countries, wine is the most used raw material due to the importance of grapevine cultivation in this region. Wine-producing countries are usually major vinegar-producing countries [19,35]. White and red wines allow for producing a large proportion of total wine vinegars [36]. In Spain, there are three Protected Designations of Origin (PDOs): “Sherry Vinegar”, “Vinegar of Condado de Huelva”, and “Vinegar of Montilla-Moriles”, all of them located in Andalusia [37]. Climate and soil factors of this region permit the growing of native varieties of grapes used for producing high-quality wines, which confer on the final vinegar exceptional organoleptic properties [4]. The aging process by the method known as “Criaderas y Soleras” is one of the singularities of these vinegars, which further enhances their uniqueness. In northern Italy, traditional balsamic vinegar has two PDOs: “L’Aceto Balsamico Tradizionale di Modena” and “L’Aceto Balsamico Tradizionale di Reggio Emilia” [38]. From the native varieties of grapes grown in this region near Modena, a must is obtained that is subsequently cooked. The cooked must is then subjected to simultaneous and spontaneous alcoholic and acetic acid fermentation, followed by a prolonged aging period of at least 12 years using barrels of different types of wood and sizes [38,39].

3.2. Spirit Vinegar

Spirit vinegar, also known as white vinegar, is obtained by the acetic acid fermentation of an alcohol medium produced from a previous alcoholic fermentation of an agricultural product containing fermentable sugars [40]. Spirit vinegar can reach the highest acidity levels [15–20% (w/v)], and although from a sensory point of view its organoleptic profile is usually very poor, in quantitative terms, it is one of the most produced vinegars worldwide, mainly in Great Britain, Germany, and the eastern USA [25,41]. For these reasons, spirit vinegar is used in studies that aim to achieve a high-yield acetification profile. Some of its main applications are as a cleaning product, seasoning, and food preservative [25].

3.3. Cereal Vinegar

Traditional cereal vinegar has a long history, with thousands of years of development and improvement that bind it to the Asian continent [42,43]. These vinegars differ according to several factors, such as the type of cereal (such as rice, sorghum, corn, barley and wheat) used as raw material, the microbial composition of starter cultures, elaboration procedures, and aging periods [42]. Among them, rice vinegar, obtained by the acetic acid fermentation of rice wine, “sake”, is popular in Asian countries. In Japan, vinegars are classified into polished rice vinegar (“Komesu”), unpolished rice vinegar (“Kurosu”), sake-less vinegar (“Kasuzu”), and other grain vinegars [44]. In China, different starchy substrates from each region are used for making some of the most famous Chinese vinegars, including Shanxi aged vinegar, Zhenjiang aromatic vinegar, Sichuan bran vinegar, and Fujian Monascus vinegar [45,46,47,48]. The particularities of processing the cereal vinegar include the use of solid-state fermentation (SSF), in which the starter culture is previously treated to allow the present dominant microbiota to carry out the saccharification and subsequent alcohol fermentation of the grains just before the acetic acid fermentation, thus producing a high-quality final product that is usually slightly milder and sweeter than Western vinegars [45,49].

3.4. Fruit Vinegar

Fruit vinegar is usually elaborated as an alternative for the exploitation of existing fruit surpluses, thus reducing the economic and environmental impact produced by the fruit industry [2]. Although Asian countries were the first to become interested in fruits as raw materials (other than grapes), their use and study in other parts of the world have been increasing in the last two decades [2]. The acidic nature of fruit vinegar and the high sensory impact of acetic acid on its organoleptic properties allow almost any type of fruit to be used for its elaboration. Among them, well-known cider vinegars are elaborated using apple juice through double alcoholic and acetic acid fermentation, especially in the United Kingdom, the United States, and Switzerland [50]. Many other fruits have been explored for the elaboration of vinegars, such as berry, persimmon, strawberry, pineapple, cherry, orange, mango, banana, and tomato, in the last few years [2]. Raw material processing is essential for the extraction of juice; crushing or pressing fruits are usually the most employed methods. Then, both the traditional surface culture and submerged culture can be employed as fermentation systems [2,8,27,50]. Depending on the fruit used, the final product will have a different nutritional composition; in the case of cider vinegar, the high polyphenol content of apples is responsible for its exclusive organoleptic properties, including a high astringency and viscosity as well as numerous health benefits [25,27,50,51].

4. Systems of Vinegar Production

Vinegar elaboration can be carried out by either solid-state fermentation (SSF) or liquid fermentation, with the latter including numerous techniques implemented in Western and European countries, mainly surface and submerged cultures [5,7]. The submerged culture is one of the main systems used to produce vinegar on an industrial scale [7,8,9,52].

4.1. Traditional Systems: Solid-State Fermentation and Surface Culture

SSF consists of a series of traditional techniques in which the microbiota responsible for the fermentation grow on substrates in the absence of free water [7]. These systems are often used in Asian countries to elaborate vinegar from grains (cereals) and to obtain high-quality vinegar after a period of aging [53,54]. SSF includes three main biological processes: (1) the liquefaction and saccharification of raw material starch, (2) alcohol fermentation, and (3) acetic acid fermentation. However, this method may be slower and less efficient than other techniques [7].
Among these traditional systems, the Orléans, Luxembourgish, and Schützenbach methods are the most used [4,55,56]. The Orléans or French method is the main surface culture system; it consists of an old procedure based on the use of wooden barrels to elaborate vinegar. The substrate used consists of a mixed culture of wine and vinegar, while acetic acid bacteria (AAB) are located on the surface, forming a biofilm known as “the mother of vinegar” [4,57]. During the production process, volumes of vinegar are harvested from the barrels every 8 to 10 days and replaced by wine. The acetification and aging processes occur simultaneously to obtain a high-quality product, but this method is too slow and involves high production costs [56,58]. The Luxembourgian and Schützenbach or German methods implement an immobilization system of AAB using supports made from wood shavings. These systems allow for increasing the contact surface between AAB and the acetification substrate, thus improving the oxygenation of the medium and the acetification yield [59]. Despite the high quality of vinegars obtained by these methods, they show some disadvantages, including the difficulties of controlling the system’s operational variables (such as temperature, oxygen supply, and volatile compounds) and ethanol losses through evaporation [4,56]. In addition, the bioprocess is normally slowed down, and the system does not allow for obtaining final products with acetic acid concentrations higher than 8–10% (w/v) [56,59].

4.2. Submerged Culture System

To industrialize vinegar production, the submerged culture was developed. Through this system, a submerged fermentation process takes place by which the ethanol content of the raw materials (such as spirits, wines, or fruit juice) is oxidized to acetic acid by AAB under controlled stirring conditions [7,8]. This biotransformation is carried out in short periods (24–48 h) and allows us to obtain high-acidity final products. This is mainly possible because of the efficiency of mass transfer and continuous vigorous aeration throughout the process [5,60]. Some aspects that contribute to the high efficiency of this method are as follows.

4.2.1. The Bioreactor: Acetator Frings

The current success of vinegar-making industries is undoubtedly because of the use of the acetator developed by Hromatka and Ebner [61] and marketed by Heinrich Frings GmbH and Co., Bonn, Germany. These bioreactors have stainless steel tanks that can work with different volumes, from a pilot scale (8–10 L) to an industrial scale volume (20,000–100,000 L) [59]. They are equipped with coils as heat exchangers to maintain a constant temperature of 30–31 °C and an efficient volatile recovery system using exhaust gas condensers and scrubbers; as a result, the losses of volatile compounds because of stripping are considerably minimized [5,7,62]. Undoubtedly, the aeration system of these bioreactors confers on them a great part of their success; it consists of a turbine system that sucks air from the outside and releases it inside, resulting in very fine air bubbles, thus generating a homogeneous mixture with the culture medium [5,7,10,63]. Through this system, a higher oxygenation efficiency is reached, thus obtaining higher acetification yields than when using traditional methods.

4.2.2. Operating Modes

Another fundamental aspect lies in the operating modes used for these bioreactors, which mainly work in batch, semi-continuous, and continuous ways. Although this choice depends on the specific purpose, which may comprise many factors, in general, a suitable environment for the development and activity of acetic acid bacteria must be ensured [5,7]. According to several authors, AAB may show high sensitivity to different variables, including ethanol concentration, acetic acid concentration, the total strength of the medium (sum of the previous two), temperature, and available dissolved oxygen [64,65,66,67,68]. In this sense, numerous studies have demonstrated that using a continuous mode, a maximum acidity level of 8–10% (w/v) can be achieved because a higher concentration or even a low ethanol content can affect the specific growth rate of AAB [7]. On the other hand, the batch mode normally implies lower productivity and additional difficulty in the preparation and maintenance of starter cultures for each cycle [5]. Moreover, the semi-continuous working mode has been mainly imposed for the industrial production of vinegar [5,7,62,66]. In this method, each cycle is started by a loading phase that replenishes the reactor with a fresh medium to the working volume without exceeding a preset ethanol concentration. Then, an exhausting stage occurs, depleting the ethanol in the culture broth to a preset extent. Finally, a fraction of the volume content in the reactor is partially unloaded and the remaining volume is used as inoculum for the next cycle; Figure 2 shows an example of what might be a typical sequence of work in this process [19,52,67]. Working in this way, the operational variables are the initial concentration of ethanol in the culture medium, the concentration to which ethanol must be depleted for a cycle to be finished, the volume of the broth that is then unloaded, and the rate at which the bioreactor is loaded with fresh medium [64,65,67]. Because AAB are highly sensitive to both ethanol and acetic acid, cell concentration and viability can be strongly affected by fermentation conditions [5,60]. Therefore, an appropriate selection of the values of operational variables is essential for maintaining suitable ranges of both substrate and product concentrations, and in this way, the natural self-selection of the best-adapted AAB to the specific working medium is carried out [6,67]. This system also allows for obtaining high-strength vinegars that may reach high acidity levels [up to 15% (w/v)]. With the high demand for these products, it may be necessary to use dual-stage high-strength processes in which two fermentation tanks are operated in a synchronous mode, thus even achieving acetic acid concentrations above 20% (w/v) [9]. In this way, the stressful environment to which AAB are subjected detracts from the overall acetification rate [5].

4.2.3. Automation Systems

Each operating mode, particularly the semi-continuous mode, require control and monitoring because the fermentation conditions may induce variations in the development of cycles, even under identical conditions [5,69]. For this, the use of a monitoring system is necessary to obtain a constant recording of data of the main variables to be measured, including the volume of the medium, the concentration of ethanol and dissolved oxygen in it, and temperature [10,66]. Supervisory Control and Data Acquisition (SCADA) consists of monitoring software that provides a set of instructions for specific sensors controlled by signal acquisition modules [5]. This scheduling system also allows for sampling at critical moments of the cycle, such as at specific points in the loading and unloading periods, using different measuring devices including probes and transducers equipped with diverse sensors that continuously monitor and register all the values of each aforementioned main variable [67]. As an example, Figure 3 shows a pilot plant on a laboratory scale, consisting of a Frings acetator (8 L) working in a semi-continuous mode and equipped with a SCADA automation system that allows for the control of the main system variables.

5. General Characteristics of Acetic Acid Bacteria

The vinegar elaboration would not be possible without the activity of the acetic acid bacteria (AAB). These bacteria are Gram-negative or Gram-variable and their metabolism is strictly aerobic, using molecular oxygen (O2) as the last electron acceptor. Despite this, some strains of Acetobacter and Gluconobacter may tolerate low dissolved oxygen concentrations, for example, throughout the alcoholic fermentation in winemaking, when they may potentially be reactivated during the wine clarification [4,12]. AAB are catalase-positive and oxidase-negative, their optimum growing temperature usually ranges between 25 and 30 °C, and their optimum growing pH is between 5 and 6.5, although many AAB do not present difficulties when growing at much lower pH levels, such as between 3 and 4 [13,70,71]. The tolerance to low pH depends on parameters such as ethanol and acetic acid concentrations and oxygen availability [72]. Regarding their shape, most AAB are ellipsoidal or cylindrical, their size usually ranges between 0.4 and 1 μm wide and 0.8 and 4.5 μm long, and they can be observed under the microscope alone, in pairs, or as aggregates and chains (see Figure 4).
These microorganisms constitute a very heterogeneous bacterial group whose cells are normally mobile with peritrichous or polar flagellation. In nature, AAB are found on substrates containing sugars and/or alcohols such as fruit juice, wine, cider, beer, and vinegar [4,11]. On them, sugars and alcohols are incompletely oxidized, thus producing organic acids such as acetic acid from ethanol, mainly performed by the genera Acetobacter and Komagataeibacter, or gluconic acid from glucose, carried out by the genus Gluconobacter [41,73,74]. The ability of AAB to metabolize different carbon sources, such as alcohols and sugars, into organic acids is of great interest to the biotechnological industry; however, vinegar production is still the most extensively used industrial application.

6. Current Taxonomy of Acetic Acid Bacteria

Acetic acid bacteria are classified in the Acetobacteraceae family, which is included in the order Rhodospirillales of the class Alphaproteobacteria. Acetobacteraceae consists of two groups: an acetous group and an acidophilic group, based on ecological and phylogenic studies [75,76]. The former includes acetic acid bacteria, which share different a set of general features (see previous section) and include a wide diversity of genera: Acetobacter, Asaia, Gluconacetobacter, Gluconobacter, Granulibacter, and Komagataeibacter among others. The members of the acidophilic group have natures and origins that are physiologically and biochemically heterogeneous, and they include other acidophilic and neutrophilic genera like Acidiphilum and Roseomonas, among many others [76]. Recent research suggests that Acetobacteraceae family members are a potential source of many undiscovered bacterial metabolites that deserve further experimental exploration [75].
Regarding the acetic acid bacteria (AAB) group, Acetobacter was the first proposed genus [77]. In the 1960s, the taxonomy of AAB was significantly influenced by several studies based on the chemotaxonomy of the G + C content of DNA, quinone systems, cellular fatty acid composition, and DNA-DNA hybridization [78]. Throughout time, four main genera of AAB (Acetobacter, Gluconobacter, Gluconacetobacter, and Komagataeibacter) were established based on their membrane-bound dehydrogenases, which define their ethanol oxidation capabilities, and the type of respiratory coenzyme chain they contained [79,80]. Through the development of polyphasic classification techniques that integrate the analysis of several phenotypic, chemotactic, and genotypic data, new genera and species have been continuously reported [81]. Further, data from phylogenetic analysis based on 16S ribosomal RNA (rRNA) gene sequences have had a profound impact on the systematics of AAB, as well as the rest of the genera of the Acetobacteraceae family [76].
The activity of AAB is one of the main factors for the elaboration of different types of vinegar [4,6,7]. Species of Acetobacter and Komagataeibacter are usually among those mainly responsible for acetification because of their high oxidative abilities, although other minor fractions of microorganisms (Gluconacetobacter, Gluconobacter, among others) might coexist with the best-adapted ones [7,8,21]. Acetobacter spp. are widely found in wine, cereal, and balsamic vinegar, elaborated by traditional methods and early stages of those produced by submerged culture or low-acid vinegar because they may be damaged at acidity levels above 8–10% (w/v) [7,39]; Acetobacter pasteurianus is usually one of the most widely found species [53,73]. Komagataeibacter spp. (mainly relocated from Gluconacetobacter), which can resist acidity levels of up to 15–20% (w/v), are predominant in submerged cultures including spirit vinegar, late stages of the production of several wine and fruit vinegars, and even traditional vinegars, because of their tolerance to low acidity levels [8,41,63]; Komagataeibacter europaeus has been described as one of the main AAB for the industrial production of vinegar [19,57,82].
Currently, up to 47 genera and 207 species belonging to the Acetobacteraceae family have been identified according to Hördt et al. [83] and the List of Prokaryotic Names with Standing in Nomenclature (LPSN) database (Parte et al. 2020) [84]. Of them, 20 genera and 108 species currently belong to AAB, as can be seen in Table 1, which recompiles the most updated classification to date.

7. Metabolism of Acetic Acid Bacteria

The molecular and biochemical aspects that define the metabolism of acetic acid bacteria are increasingly becoming the target of much research. In this section, a general and updated overview of the main AAB metabolic pathways, especially those related to carbon source assimilation including alcohols, sugars, and sugar alcohols for the production of organic acids, has been provided. It is worth noting that many other related metabolic pathways, partially or completely unknown, are presently being studied by several authors [6,154,155,156,157,158].

7.1. Biotransformation of Ethanol to Acetic Acid

The overall oxidative biological reaction that defines the biotransformation of ethanol into acetic acid can be represented as follows:
C 2 H 5 O H + O 2 C H 3 C O O H + H 2 O    Δ H ° = 520   K J / m o l
AAB are chemoorganotrophs microorganisms that use ethanol from a medium of alcoholic origin as a carbon source. The genera Acetobacter and Komagataeibacter usually show a higher ethanol preference, although other AAB groups may show a preference for other carbon sources [7,41]. This biotransformation consists of an incomplete oxidation reaction of two steps. First, alcohol dehydrogenase (ADH) binds to pyrroloquinoline quinone (PQQ) to oxidize the ethanol into acetaldehyde. Next, acetaldehyde is oxidized to acetic acid by membrane-bound aldehyde dehydrogenase (ALDH); both enzymes are located on the periplasmic side of the inner cell membrane [159,160]. Oxidized nicotinamide adenine dinucleotide (NAD+) and nicotinamide adenine dinucleotide phosphate (NADP+), both located in the cytoplasm, may be used as coenzymes by NAD-ADH, NAD-ALDH, and NADP-ALDH [156,161]. The inner acetic acid can be completely oxidized by the acetyl-CoA synthase, which leads to the input of acetyl-CoA into the TCA cycle, and in this case, to CO2 and H2O providing energy (ATP) and detoxifying the cell [14]. Other organic acids such as lactic, pyruvic, malic, succinic, citric, and fumaric acids may be similarly metabolized [73]. Because of the strictly aerobic metabolism of AAB, the ADH-PQQ and ALDH complexes are closely linked to the respiratory chain, which transfers reducing equivalents from donor substrates to ubiquinone (UB). Then, electrons from the reduced UB, named ubiquinol (UBH2), are transferred to the final electron acceptor, oxygen (O2), by terminal ubiquinol oxidases (UOXs), producing H2O (see Figure 5) [14,161]. Some processes related to this central oxidative metabolism may include pathways that aim to obtain biosynthetic precursors of amino acids and nucleic acids, among others, in order to replenish cell material losses throughout the early stages of the acetification process. The TCA cycle also plays a crucial role in the assimilation of internal acetic acid; among the involved enzymes, succinyl-CoA: acetate CoA transferase (SCACT), encoded by aarC, New Orleans, LA, USA, is able to produce acetyl-CoA from inner acetate, being of significant importance in the tolerance to acetic acid [162,163]. Furthermore, membrane mechanisms dependent on proton motive force may be triggered for acetic acid release and cell detoxification at the final moments of the process; here, the importance of outer membrane proteins (OMPA) and efflux pumps (OPRM and ABC-transporters) in the control of the cellular output of acetic acid, as well as MLTA, participating in the maintenance of the peptidoglycan layer under these conditions, are highlighted; see Figure 6 [6].
The ADH complex of most AAB is composed of three subunits, although it may contain two subunits in some species [164]. Subunit I (72–78 kDa), encoded by the gene adhA, is a catalytic component containing a PQQ and a heme C moiety. Subunit II (44–45 kDa), encoded by the gene adhB, is a membrane-anchoring and ubiquinone-reducing component possessing three heme C moieties; both subunits participate in the intramolecular electron transport to the terminal UB. Subunit III (20 kDa), encoded by the gen adhS, which has no prosthetic group, facilitates the association of subunits I and II to the membrane and acts as a molecular chaperone for the folding and/or maturation of subunit I [73,161]. Several authors have related high ADH stability and activity with a high tolerance and production of acetic acid, mainly in species from the current genus Komagataeibacter [41,82]. The ALDH complex is composed of two or three subunits depending on the AAB species, and it acts as an operon. Although its optimum pH ranges between 4 and 5, the oxidation of acetaldehyde to acetate may be catalyzed at lower pH values. ALDH is highly sensitive to low oxygen concentrations and the presence of ethanol in the medium [73].

7.2. Carbohydrates Oxidation

AAB can metabolize different carbohydrates as carbon sources, mainly glucose, but also arabinose, fructose, galactose, mannose, ribose, sorbose, and xylose [73]. Most AAB have been characterized by non-functional glycolysis because of the absence of a phosphofructokinase enzyme; therefore, the pentose phosphate pathway (PPP) is the main metabolic route of AAB to oxidize the glucose available in the medium by the catalytic activity of the enzymes glucose-6-P dehydrogenase (G6PDH) and 6-phosphogluconate dehydrogenase (6PGD), providing metabolic precursors such as ribulose-5-phosphate and generating NADPH + H+ and energy [165,166,167]. Among AAB, several species from Gluconobacter have a glucose preference, and several Gluconobacter oxydans strains also exhibit the ability to oxidize glucose to gluconic acid via glucono-δ-lactone, forming D-gluconate. This oxidation reaction occurs in the periplasm using a membrane-bound pyrroloquinoline quinone-dependent glucose dehydrogenase (PQQ-GDH) located on the outer side of the cytoplasmic membrane. D-gluconate can be further oxidized rapidly to ketogluconates such as 2-ketogluconate (2-KGA), 5-ketogluconate (5-KGA), and 2,5-diketogluconic acid (2,5-DKGA), both in the periplasm and cytoplasm, by different oxidizing enzymes; see Figure 7 [166,168,169]. Glucose, gluconic acid, and ketogluconates can be assimilated by these bacteria, thus obtaining biomass and energy and acidifying the medium, possibly as part of their metabolic strategy to prevail over other glucose-like microorganisms [166]. Final products of PPP and the Entner Doudoroff pathway (EDP) may be completely oxidized to CO2 and H2O by Acetobacter, Gluconacetobacter, and Komagataeibacter spp. using the TCA cycle when the carbon source of the medium is exhausted; this is not the case for Gluconobacter spp., which show a non-functional TCA cycle [73,169]. Recent studies have proposed a molecular strategy in which K. europaeus, as the predominant species of a complex microbiota involved in the submerged production of vinegar from raw materials with high sugar content, might assimilate, firstly and before the ethanol, the glucose in the medium, draining biosynthetic precursors directly by using enzymes of PPP and the glycolysis to prevail over other species that exhibit high glucose preference [6]; see Figure 6.
AAB also exhibit the ability to oxidize several sugar alcohols such as glycerol, D-mannitol, and D-sorbitol, among others, with the use of glycerol as a carbon source being especially remarkable in winemaking, producing dihydroxyacetone (DHA) through the activity of some oxidizing enzymes, mainly glycerol dehydrogenase, and providing energy (ATP) via gluconeogenesis. Strains from Acetobacter pasteurianus, Gluconobacter oxydans, and Komagataeibacter xylinus are some of the most studied regarding this oxidative pathway [16,73].

8. Biotechnological Applications of Acetic Acid Bacteria

AAB are the main microorganisms responsible for vinegar production, but they are also used in different increasingly investigated biotechnological applications. Other foods can be produced as the result of the activity of AAB, as is the case of kombucha—a traditional beverage obtained by fermenting sugary tea with a symbiotic culture of acidophilic yeasts and bacteria including acetic acid bacteria (AAB) and lactic acid bacteria (LAB) immobilized in a microbial cellulose biofilm known as tea fungus [71,170]. First, yeasts transform sugars from tea into organic acids, ethanol, and CO2. Then, AAB may synthesize different compounds such as acetic acid (Acetobacter aceti, Acetobacter pasteurianus), gluconic acid (Gluconobacter oxydans), and bacterial cellulose (Komagataeibacter xylinus) due to the high biodiversity of AAB. This product is becoming more and more popular because of its probiotic characteristics as a treatment for gastrointestinal disorders and for improving general health and increasing longevity, with the benefits attributed to its acidic composition and high phenolic antioxidant content [71,170,171]. Another product, gluconic acid, is industrially obtained by the oxidation of glucose by several AAB, mainly Gluconobacter oxydans. Gluconic acid improves the sensory properties and increases the softness of other food products, including vinegar, and may also be used as an additive and preservative by the food industry. Due to its role in the aromatic profile of foods, gluconic acid has been proposed as a quality parameter of food products [71,172]. Gluconic acid is used in the pharmaceutical industry as gluconates of divalent metals, which function as mineral supplements to treat some diseases [173]. The high oxidative capability of numerous strains of Gluconobacter is also exploited to convert sugar alcohols, as in the case of D-sorbitol to L-sorbose—an important intermediate in the industrial production of L-ascorbic acid (vitamin C), an antioxidant often used in the food industry [73].
Among the biotechnological applications of AAB, the production of bacterial cellulose has attracted interest in recent years because of its extreme purity, unlike plant-derived cellulose, thus representing a promising alternative for many industries [71]. Among its multiple applications, bacterial cellulose is employed as a gelling, stabilizing, and thickening agent in foods, heart medicine, pharmacy, and skin repair in wound healing and burn treatments [174,175,176]. K. xylinus is the most commonly used species of AAB because of its capability to produce high amounts of bio-cellulose from different carbon and nitrogen sources, and this involves different enzymes such as glucose kinase, phosphoglucomutase, UDP-glucose pyrophosphorylase, and membrane-bound cellulose synthase [17]. Throughout strawberry vinegar fermentation, Ameyamaea, Gluconacetobacter, and Komagataeibacter were found, combining both culture-dependent and culture-independent methodologies, as the dominating bacterial genera in biofilms previously generated during the process [177]. Besides bio-cellulose, AAB may produce other microbial exopolysaccharides, such as levans, dextran, acetan, mannan, and gluconacetan, with important industrial applications [15,71].
Omics technologies have emerged in recent years as an alternative to solve many of the traditional hurdles for the isolation and characterization of acetic acid bacteria, especially those that limit the study of the richness and biodiversity of these microbiota inhabiting such selective media as vinegar. In fact, in the past few years, different omic and meta-omic approaches have been implemented, including (meta)genomics [8,21,49,177], transcriptomics [157,178], (meta)proteomics [6,19,41,158], and metabolomics [43,46,48]. These studies are allowing for the identification and characterization of most of the members of these communities involved in these production processes, describing their behavior both under different operating conditions and media, thus helping to elucidate the key role of the vinegar-making microbiota.
Using innovative shotgun metaproteomics (LC-MS/MS) and metagenomic techniques (16S rRNA amplicon sequencing), Román-Camacho et al. [6,21] approach the difficult problem of identifying and characterizing these complex microbial communities, as well as the influence of the use of various culture media for vinegar production; Komagataeibacter members, especially Komagataeibacter europaeus strains, were the main representatives. Likewise, these works explain the ability of AAB to adapt to different culture media through metabolic versatility. Shotgun metagenomics, which implements the use of sequencing data to infer potential metabolic functions encoded by the genomes of the community members, was applied to reveal the flavor metabolic network of the microbiota of a cereal vinegar [49].
Another aspect that also receives a lot of attention is related to the study of how AAB resist the aggressive environments in which they are normally found—for example, the high acidity values during the industrial production of vinegar. Using RNA-Seq transcriptomic analyses, Yang et al. [179] studied gene regulation changes to find possible relationships with the acidity of the medium, and Wang et al. [157] determined several mechanisms against high acid stress at different stages of acetic acid fermentation in a K. europaeus strain.
Metabolomics is another recent approach that, in the vinegar production area (especially Chinese cereal vinegars), is allowing us to determine different volatile aroma patterns throughout fermentation [43], as well as correlate numerous metabolites (volatile compounds, organic acids, and amino acids) with the main members of these microbiotas [46,48]; in these studies, head-space solid phase microextraction (HS-SPME) coupled with gas chromatography–mass spectrometry (GC–MS) was used for determining volatile compounds, and high-performance liquid chromatography (HPLC) was used for organic acids and amino acids.

9. Conclusions and Future Prospects

This review has attempted to summarize the current state of knowledge on vinegar production from the diversity of raw materials and starter cultures to system production and operating conditions used. Special emphasis has been placed on submerged cultures as a suitable method of industrial vinegar production and an updated taxonomic review of the main responsible microorganisms for the process: acetic acid bacteria. Because of the multiple variables that influence vinegar production, it is not easy for the industries to achieve a balance between them that is aimed at the optimization of this process. Considering current research and the evolution of the agri-food market, the improvement of the organoleptic properties of these unique products might be focused on the implementation of new operating conditions, the characterization of new raw materials, and the study of the microbial composition and behavior of microbiota inhabiting vinegar, especially those responsible for its elaboration. In recent years, omics tools have approached these strategies with high throughput without compromising the fitness of microbiota and the quality of the final product. In vinegar, they are allowed to solve many of the traditional hurdles for the isolation and characterization of microorganisms inhabiting these aggressive media. Despite this, the current omic analyses still have numerous limitations, such as the challenge of detecting low-abundant species, high levels of technical and biological noise, and the performance of few biological replicates because of the high cost. There is a pressing need to advance the field of sequencing, mass spectrometry technologies, and bioinformatics, implementing robust techniques for integrating, visualizing, and validating omic data. New challenges for industrial vinegar production might target obtaining innovative varieties of vinegar with sensorial profiles and healthy organoleptic properties; the employment of improved starter cultures from the selection of species or strains with a key role throughout acetification; advancements in obtaining vinegar isolates, their phenotypic characterization, and biotechnological enhancement; and the identification of marker genes, proteins, and metabolites throughout the process. This work might establish the first steps of a path toward the improvement of vinegar processing in the industry.

Author Contributions

Conceptualization, I.G.-G. and J.C.M.; methodology, J.J.R.-C., I.G.-G. and J.C.M.; validation, J.J.R.-C.; investigation, J.J.R.-C.; data curation, J.J.R.-C., I.M.S.-D. and T.G.-M.; writing—original draft preparation, J.J.R.-C., I.M.S.-D. and T.G.-M.; writing—review and editing, I.G.-G., I.M.S.-D., T.G.-M. and J.C.M.; visualization, J.J.R.-C.; funding acquisition, I.G.-G. and J.C.M. All authors have read and agreed to the published version of the manuscript.

Funding

The completion of this review has been made possible by the funding, over many years, of specific projects which can be found in the various publications of the authors cited in this article. In particular, as they are the most recent, the following two should be highlighted: “Junta de Andalucía” and the European Regional Development Fund, ERDF (FEDER), Ref. PY20_00590; and PID2021-127766OB-I00 from Spanish Ministry of Science and Innovation (MICINN/European Union FEDER). Likewise, reference should be made to the University of Cordoba’s own program for the promotion of research.

Data Availability Statement

The data used to support the findings of this study can be made available by the corresponding author upon request.

Conflicts of Interest

The authors are not aware of any affiliations, memberships, funding, or financial holdings that might be perceived as affecting the objectivity of this review.

References

  1. Ho, C.W.; Lazim, A.M.; Fazry, S.; Zaki, U.K.H.H.; Lim, S.J. Varieties, Production, Composition and Health Benefits of Vinegars: A Review. Food Chem. 2017, 221, 1621–1630. [Google Scholar] [CrossRef]
  2. Luzón-Quintana, L.M.; Castro, R.; Durán-Guerrero, E. Biotechnological Processes in Fruit Vinegar Production. Foods 2021, 10, 945. [Google Scholar] [CrossRef] [PubMed]
  3. Li, S.; Li, P.; Feng, F.; Luo, L.X. Microbial Diversity and Their Roles in the Vinegar Fermentation Process. Appl. Microbiol. Biotechnol. 2015, 99, 4997–5024. [Google Scholar] [CrossRef] [PubMed]
  4. Mas, A.; Torija, M.J.; García-Parrilla, M.D.C.; Troncoso, A.M. Acetic Acid Bacteria and the Production and Quality of Wine Vinegar. Sci. World J. 2014, 2014, 394671. [Google Scholar] [CrossRef] [PubMed]
  5. García-García, I.; Santos-Dueñas, I.M.; Jiménez-Ot, C.; Jiménez-Hornero, J.E.; Bonilla-Venceslada, J.L. Vinegar Engineering. In Vinegars of the World; Solieri, L., Giudici, P., Eds.; Springer: Milan, Italy, 2009; pp. 97–120. ISBN 978-88-470-0865-6. [Google Scholar]
  6. Román-Camacho, J.J.; Mauricio, J.C.; Santos-Dueñas, I.M.; García-Martínez, T.; García-García, I. Unraveling the Role of Acetic Acid Bacteria Comparing Two Acetification Profiles From Natural Raw Materials: A Quantitative Approach in Komagataeibacter Europaeus. Front. Microbiol. 2022, 13, 840119. [Google Scholar] [CrossRef]
  7. Gullo, M.; Verzelloni, E.; Canonico, M. Aerobic Submerged Fermentation by Acetic Acid Bacteria for Vinegar Production: Process and Biotechnological Aspects. Process Biochem. 2014, 49, 1571–1579. [Google Scholar] [CrossRef]
  8. Trček, J.; Mahnič, A.; Rupnik, M. Diversity of the Microbiota Involved in Wine and Organic Apple Cider Submerged Vinegar Production as Revealed by DHPLC Analysis and Next-Generation Sequencing. Int. J. Food Microbiol. 2016, 223, 57–62. [Google Scholar] [CrossRef]
  9. Álvarez-Cáliz, C.M.; Santos-Dueñas, I.M.; Jiménez-Hornero, J.E.; García-García, I. Optimization of the Acetification Stage in the Production of Wine Vinegar by Use of Two Serial Bioreactors. Appl. Sci. 2021, 11, 1217. [Google Scholar] [CrossRef]
  10. Qi, Z.; Yang, H.; Xia, X.; Quan, W.; Wang, W.; Yu, X. Achieving High Strength Vinegar Fermentation via Regulating Cellular Growth Status and Aeration Strategy. Process Biochem. 2014, 49, 1063–1070. [Google Scholar] [CrossRef]
  11. Garcia-Garcia, I.; Gullo, M.; Chen, F.; Garcia-Martinez, T. Editorial: Acetic Acid Bacteria. Front. Microbiol. 2023, 14, 1142659. [Google Scholar] [CrossRef]
  12. 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, 1260–1271. [Google Scholar] [CrossRef] [PubMed]
  13. Trček, J.; Mira, N.P.; Jarboe, L.R. Adaptation and Tolerance of Bacteria against Acetic Acid. Appl. Microbiol. Biotechnol. 2015, 99, 6215–6229. [Google Scholar] [CrossRef] [PubMed]
  14. He, Y.; Xie, Z.; Zhang, H.; Liebl, W.; Toyama, H.; Chen, F. Oxidative Fermentation of Acetic Acid Bacteria and Its Products. Front. Microbiol. 2022, 13, 879246. [Google Scholar] [CrossRef]
  15. Anguluri, K.; La China, S.; Brugnoli, M.; De Vero, L.; Pulvirenti, A.; Cassanelli, S.; Gullo, M. Candidate Acetic Acid Bacteria Strains for Levan Production. Polymers 2022, 14, 2000. [Google Scholar] [CrossRef]
  16. La China, S.; Zanichelli, G.; De Vero, L.; Gullo, M. Oxidative Fermentations and Exopolysaccharides Production by Acetic Acid Bacteria: A Mini Review. Biotechnol. Lett. 2018, 40, 1289–1302. [Google Scholar] [CrossRef]
  17. Kuo, C.H.; Chen, J.H.; Liou, B.K.; Lee, C.K. Utilization of Acetate Buffer to Improve Bacterial Cellulose Production by Gluconacetobacter xylinus. Food Hydrocoll. 2016, 53, 98–103. [Google Scholar] [CrossRef]
  18. Rizo, J.; Guillén, D.; Farrés, A.; Díaz-Ruiz, G.; Sánchez, S.; Wacher, C.; Rodríguez-Sanoja, R. Omics in Traditional Vegetable Fermented Foods and Beverages. Crit. Rev. Food Sci. Nutr. 2020, 60, 791–809. [Google Scholar] [CrossRef]
  19. Román-Camacho, J.J.; Santos-Dueñas, I.M.; García-García, I.; Moreno-García, J.; García-Martínez, T.; Mauricio, J.C. Metaproteomics of Microbiota Involved in Submerged Culture Production of Alcohol Wine Vinegar: A First Approach. Int. J. Food Microbiol. 2020, 333, 108797. [Google Scholar] [CrossRef]
  20. Román-Camacho, J.J.; Mauricio, J.C.; Santos-Dueñas, I.M.; García-Martínez, T.; García-García, I. Functional Metaproteomic Analysis of Alcohol Vinegar Microbiota during an Acetification Process: A Quantitative Proteomic Approach. Food Microbiol. 2021, 98, 103799. [Google Scholar] [CrossRef]
  21. Román-Camacho, J.J.; García-García, I.; Santos-Dueñas, I.M.; Ehrenreich, A.; Liebl, W.; García-Martínez, T.; Mauricio, J.C. Combining Omics Tools for the Characterization of the Microbiota of Diverse Vinegars Obtained by Submerged Culture: 16S RRNA Amplicon Sequencing and MALDI-TOF MS. Front. Microbiol. 2022, 13, 1055010. [Google Scholar] [CrossRef]
  22. Walsh, A.M.; Crispie, F.; Claesson, M.J.; Cotter, P.D. Translating Omics to Food Microbiology. Annu. Rev. Food Sci. Technol. 2017, 8, 113–134. [Google Scholar] [CrossRef]
  23. Andrés-Barrao, C.; Barja, F. Acetic Acid Bacteria Strategies Contributing to Acetic Acid Resistance during Oxidative Fermentation. In Acetic Acid Bacteria: Fundamentals and Food Applications; Sengun, I.Y., Ed.; Food Biology Series “A Science Publishers Book”; CRC Press: Boca Raton, FL, USA, 2017; pp. 92–119. ISBN 978-1-315-15349-0. [Google Scholar]
  24. Johnston, C.S.; Gaas, C.A. Vinegar: Medicinal Uses and Antiglycemic Effect. Med. Gen. Med. 2006, 8, 61. [Google Scholar]
  25. Budak, N.H.; Aykin, E.; Seydim, A.C.; Greene, A.K.; Guzel-Seydim, Z.B. Functional Properties of Vinegar. J. Food Sci. 2014, 79, R757–R764. [Google Scholar] [CrossRef] [PubMed]
  26. Ali, Z.; Wang, Z.; Amir, R.M.; Younas, S.; Wali, A.; Adowa, N.; Ayim, I. Potential Uses of Vinegar as a Medicine and Related in Vivo Mechanisms. Int. J. Vitam. Nutr. Res. 2016, 86, 140–151. [Google Scholar] [CrossRef]
  27. Budak, N.H.; Kumbul Doguc, D.; Savas, C.M.; Seydim, A.C.; Kok Tas, T.; Ciris, M.I.; Guzel-Seydim, Z.B. Effects of Apple Cider Vinegars Produced with Different Techniques on Blood Lipids in High-Cholesterol-Fed Rats. J. Agric. Food Chem. 2011, 59, 6638–6644. [Google Scholar] [CrossRef] [PubMed]
  28. Budak, H.N.; Guzel-Seydim, Z.B. Antioxidant Activity and Phenolic Content of Wine Vinegars Produced by Two Different Techniques. J. Sci. Food Agric. 2010, 90, 2021–2026. [Google Scholar] [CrossRef] [PubMed]
  29. Chang, J.; Fang, T. Survival of Escherichia coli O157:H7 and Salmonella enterica Serovars Typhimurium in Iceberg Lettuce and the Antimicrobial Effect of Rice Vinegar against E. coli O157:H7. Food Microbiol. 2007, 24, 745–751. [Google Scholar] [CrossRef]
  30. Chou, C.H.; Liu, C.W.; Yang, D.J.; Wu, Y.H.S.; Chen, Y.C. Amino Acid, Mineral, and Polyphenolic Profiles of Black Vinegar, and Its Lipid Lowering and Antioxidant Effects in Vivo. Food Chem. 2015, 168, 63–69. [Google Scholar] [CrossRef]
  31. Honsho, S.; Sugiyama, A.; Takahara, A.; Satoh, Y.; Nakamura, Y.; Hashimoto, K. A Red Wine Vinegar Beverage Can Inhibit the Renin-Angiotensin System: Experimental Evidence in Vivo. Biol. Pharm. Bull. 2005, 28, 1208–1210. [Google Scholar] [CrossRef]
  32. Yamashita, H.; Fujisawa, K.; Ito, E.; Idei, S.; Kawaguchi, N.; Kimoto, M.; Hiemori, M.; Tsuji, H. Improvement of Obesity and Glucose Tolerance by Acetate in Type 2 Diabetic Otsuka Long-Evans Tokushima Fatty (OLETF) Rats. Biosci. Biotechnol. Biochem. 2007, 71, 1236–1243. [Google Scholar] [CrossRef]
  33. Lazim, A.M.; Lim, S.J.; Ho, C.W.; Fazry, S. Types of Vinegars. In Advances in Vinegar Production; Bekatorou, A., Ed.; CRC Press: Boca Raton, FL, USA, 2019; pp. 19–28. ISBN 978-1-351-20847-5. [Google Scholar]
  34. Parrondo, J.; Garcia, L.A.; Diaz, M. Whey Vinegar. In Vinegars of the World; Solieri, L., Giudici, P., Eds.; Springer: Milan, Italy, 2009; pp. 273–288. ISBN 978-88-470-0865-6. [Google Scholar]
  35. Maestre, O.; Santos-Dueñas, I.M.; Peinado, R.; Jiménez-Ot, C.; García-García, I.; Mauricio, J.C. Changes in Amino Acid Composition during Wine Vinegar Production in a Fully Automatic Pilot Acetator. Process Biochem. 2008, 43, 803–807. [Google Scholar] [CrossRef]
  36. Sellmer-Wilsberg, S. Wine and Grape Vinegars. In Vinegars of the World; Solieri, L., Giudici, P., Eds.; Springer: Milan, Italy, 2009; pp. 145–156. ISBN 978-88-470-0865-6. [Google Scholar]
  37. Durán-Guerrero, E.; Castro, R.; García-Moreno, M.D.V.; Rodríguez-Dodero, M.D.C.; Schwarz, M.; Guillén-Sánchez, D. Aroma of Sherry Products: A Review. Foods 2021, 10, 753. [Google Scholar] [CrossRef] [PubMed]
  38. Giudici, P.; Gullo, M.; Solieri, L. Traditional Balsamic Vinegar. In Vinegars of the World; Solieri, L., Giudici, P., Eds.; Springer: Milan, Italy, 2009; pp. 157–177. ISBN 978-88-470-0865-6. [Google Scholar]
  39. Gullo, M.; Giudici, P. Acetic Acid Bacteria in Traditional Balsamic Vinegar: Phenotypic Traits Relevant for Starter Cultures Selection. Int. J. Food Microbiol. 2008, 125, 46–53. [Google Scholar] [CrossRef] [PubMed]
  40. Grierson, B. Malt and Distilled Malt Vinegar. In Vinegars of the World; Solieri, L., Giudici, P., Eds.; Springer: Milan, Italy, 2009; pp. 135–143. ISBN 978-88-470-0865-6. [Google Scholar]
  41. Andrés-Barrao, C.; Saad, M.M.; Cabello Ferrete, E.; Bravo, D.; Chappuis, M.-L.; Ortega Pérez, R.; Junier, P.; Perret, X.; Barja, F. Metaproteomics and Ultrastructure Characterization of Komagataeibacter Spp. Involved in High-Acid Spirit Vinegar Production. Food Microbiol. 2016, 55, 112–122. [Google Scholar] [CrossRef]
  42. Giudici, P.; Corradini, G.; Bonciani, T.; Wu, J.; Chen, F.; Lemmetti, F. The Flavor and Taste of Cereal Chinese Vinegars. Acetic Acid Bact. 2017, 6, 6370. [Google Scholar] [CrossRef]
  43. Zhang, X.; Wang, P.; Xu, D.; Wang, W.; Zhao, Y. Aroma Patterns of Beijing Rice Vinegar and Their Potential Biomarker for Traditional Chinese Cereal Vinegars. Food Res. Int. 2019, 119, 398–410. [Google Scholar] [CrossRef]
  44. Murooka, Y.; Nanda, K.; Yamashita, M. Rice Vinegars. In Vinegars of the World; Solieri, L., Giudici, P., Eds.; Springer: Milan, Italy, 2009; pp. 121–133. ISBN 978-88-470-0865-6. [Google Scholar]
  45. Chen, F.; Li, L.; Qu, J.; Chen, C. Cereal Vinegars Made by Solid-State Fermentation in China. In Vinegars of the World; Solieri, L., Giudici, P., Eds.; Springer: Milan, Italy, 2009; pp. 243–259. ISBN 978-88-470-0865-6. [Google Scholar]
  46. Jiang, Y.; Lv, X.; Zhang, C.; Zheng, Y.; Zheng, B.; Duan, X.; Tian, Y. Microbial Dynamics and Flavor Formation during the Traditional Brewing of Monascus Vinegar. Food Res. Int. 2019, 125, 108531. [Google Scholar] [CrossRef]
  47. Xu, W.; Huang, Z.; Zhang, X.; Li, Q.; Lu, Z.; Shi, J.; Xu, Z.; Ma, Y. Monitoring the Microbial Community during Solid-State Acetic Acid Fermentation of Zhenjiang Aromatic Vinegar. Food Microbiol. 2011, 28, 1175–1181. [Google Scholar] [CrossRef]
  48. Zhu, Y.; Zhang, F.; Zhang, C.; Yang, L.; Fan, G.; Xu, Y.; Sun, B.; Li, X. Dynamic Microbial Succession of Shanxi Aged Vinegar and Its Correlation with Flavor Metabolites during Different Stages of Acetic Acid Fermentation. Sci. Rep. 2018, 8, 8612. [Google Scholar] [CrossRef]
  49. Wu, L.-H.; Lu, Z.-M.; Zhang, X.-J.; Wang, Z.-M.; Yu, Y.-J.; Shi, J.-S.; Xu, Z.-H. Metagenomics Reveals Flavour Metabolic Network of Cereal Vinegar Microbiota. Food Microbiol. 2017, 62, 23–31. [Google Scholar] [CrossRef]
  50. Joshi, V.K.; Sharma, S. Cider Vinegar: Microbiology, Technology and Quality. In Vinegars of the World; Solieri, L., Giudici, P., Eds.; Springer: Milan, Italy, 2009; pp. 197–207. ISBN 978-88-470-0865-6. [Google Scholar]
  51. Ousaaid, D.; Mechchate, H.; Laaroussi, H.; Hano, C.; Bakour, M.; El Ghouizi, A.; Conte, R.; Lyoussi, B.; El Arabi, I. Fruits Vinegar: Quality Characteristics, Phytochemistry, and Functionality. Molecules 2021, 27, 222. [Google Scholar] [CrossRef] [PubMed]
  52. Lee, S.; Lee, J.-A.; Park, G.-G.; Jang, J.-K.; Park, Y.-S. Semi-Continuous Fermentation of Onion Vinegar and Its Functional Properties. Molecules 2017, 22, 1313. [Google Scholar] [CrossRef] [PubMed]
  53. Wu, J.; Gullo, M.; Chen, F.; Giudici, P. Diversity of Acetobacter pasteurianus Strains Isolated From Solid-State Fermentation of Cereal Vinegars. Curr. Microbiol. 2010, 60, 280–286. [Google Scholar] [CrossRef] [PubMed]
  54. Xia, M.; Zheng, Y.; Zhang, X.; Xie, S.; Wu, Y.; Xia, T.; Wang, M. Solid-State Fermentation Systems for Vinegar Production. In Advances in Vinegar Production; Bekatorou, A., Ed.; CRC Press: Boca Raton, FL, USA, 2019; pp. 141–154. ISBN 978-1-351-20847-5. [Google Scholar]
  55. Bekatorou, A. Advances in Immobilized Biocatalyst Technologies for Vinegar Production. In Advances in Vinegar Production; Bekatorou, A., Ed.; CRC Press: Boca Raton, FL, USA, 2019; pp. 117–139. ISBN 978-1-351-20847-5. [Google Scholar]
  56. Tesfaye, W.; Morales, M.L.; García-Parrilla, M.C.; Troncoso, A.M. Wine Vinegar: Technology, Authenticity and Quality Evaluation. Trends Food Sci. 2002, 13, 12–21. [Google Scholar] [CrossRef]
  57. Andres-Barrao, C.; Weber, A.; Chappuis, M.L. Acetic Acid Bacteria Population Dynamics and Natural Imposition of Gluconacetobacter europaeus during Submerged Vinegar Production. Arch. Sci. 2011, 64, 99–114. [Google Scholar] [CrossRef]
  58. Raspor, P.; Goranovič, D. Biotechnological Applications of Acetic Acid Bacteria. Crit. Rev. Biotechnol. 2008, 28, 101–124. [Google Scholar] [CrossRef]
  59. El Vinagre de Vino; Llaguno Marchena, C.; Polo, M.C. (Eds.) Consejo Superior de Investigaciones Científicas: Madrid, Spain, 1991; ISBN 978-84-00-07205-6. [Google Scholar]
  60. García-García, I.; Jiménez-Hornero, J.E.; Santos-Dueñas, I.M.; González-Granados, Z.; Cañete-Rodríguez, A.M. Modeling and Optimization of Acetic Acid Fermentation. In Advances in Vinegar Production; Bekatorou, A., Ed.; CRC Press: Boca Raton, FL, USA, 2019; pp. 299–325. ISBN 978-1-351-20847-5. [Google Scholar]
  61. Hromatka, O.; Ebner, H. Vinegar by Submerged Oxidative Fermentation. Ind. Eng. Chem. 1959, 51, 1279–1280. [Google Scholar] [CrossRef]
  62. De Ory, I.; Romero, L.E.; Cantero, D. Operation in Semi-Continuous with a Closed Pilot Plant Scale Acetifier for Vinegar Production. J. Food Eng. 2004, 63, 39–45. [Google Scholar] [CrossRef]
  63. Fernández-Pérez, R.; Torres, C.; Sanz, S.; Ruiz-Larrea, F. Strain Typing of Acetic Acid Bacteria Responsible for Vinegar Production by the Submerged Elaboration Method. Food Microbiol. 2010, 27, 973–978. [Google Scholar] [CrossRef]
  64. Baena-Ruano, S.; Santos-Dueñas, I.M.; Mauricio, J.C.; García-García, I. Relationship between Changes in the Total Concentration of Acetic Acid Bacteria and Major Volatile Compounds during the Acetic Acid Fermentation of White Wine. J. Sci. Food Agric. 2010, 90, 2675–2681. [Google Scholar] [CrossRef]
  65. Baena-Ruano, S.; Jiménez-Ot, C.; Santos-Dueñas, I.M.; Jiménez-Hornero, J.E.; Bonilla-Venceslada, J.L.; Álvarez-Cáliz, C.; García-García, I. Influence of the Final Ethanol Concentration on the Acetification and Production Rate in the Wine Vinegar Process. J. Chem. Technol. Biotechnol. 2010, 85, 908–912. [Google Scholar] [CrossRef]
  66. García-García, I.; Cantero-Moreno, D.; Jiménez-Ot, C.; Baena-Ruano, S.; Jiménez-Hornero, J.; Santos-Dueñas, I.; Bonilla-Venceslada, J.; Barja, F. Estimating the Mean Acetification Rate via On-Line Monitored Changes in Ethanol during a Semi-Continuous Vinegar Production Cycle. J. Food Eng. 2007, 80, 460–464. [Google Scholar] [CrossRef]
  67. Jiménez-Hornero, J.E.; Santos-Dueñas, I.M.; García-García, I. Modelling Acetification with Artificial Neural Networks and Comparison with Alternative Procedures. Processes 2020, 8, 749. [Google Scholar] [CrossRef]
  68. Santos-Dueñas, I.M.; Jimenez-Hornero, J.E.; Cañete-Rodriguez, A.M.; Garcia-Garcia, I. Modeling and Optimization of Acetic Acid Fermentation: A Polynomial-Based Approach. Biochem. Eng. J. 2015, 99, 35–43. [Google Scholar] [CrossRef]
  69. Kalogianni, D. Rapid Detection Methods for Online Monitoring of Vinegar Fermentations. In Advances in Vinegar Production; Bekatorou, A., Ed.; CRC Press: Boca Raton, FL, USA, 2019; pp. 421–440. ISBN 978-1-351-20847-5. [Google Scholar]
  70. Wang, B.; Shao, Y.; Chen, F. Overview on Mechanisms of Acetic Acid Resistance in Acetic Acid Bacteria. World J. Microbiol. Biotechnol. 2015, 31, 255–263. [Google Scholar] [CrossRef]
  71. Gomes, R.J.; Borges, M.D.F.; Rosa, M.D.F.; Castro-Gómez, R.J.H.; Spinosa, W.A. Acetic Acid Bacteria in the Food Industry: Systematics, Characteristics and Applications. Food Technol. Biotechnol. 2018, 56, 139. [Google Scholar] [CrossRef]
  72. Malimas, T.; Thi Lan Vu, H.; Muramatsu, Y.; Yukphan, P.; Tanasupawat, S.; Yamada, Y. Systematics of Acetic Acid Bacteria. In Acetic Acid Bacteria; Sengun, I.Y., Ed.; Food Biology Series “A Science Publishers Book”; CRC Press: Boca Raton, FL, USA, 2017; pp. 3–43. ISBN 978-1-315-15349-0. [Google Scholar]
  73. Mamlouk, D.; Gullo, M. Acetic Acid Bacteria: Physiology and Carbon Sources Oxidation. Indian J. Microbiol. 2013, 53, 377–384. [Google Scholar] [CrossRef]
  74. Qiu, X.; Zhang, Y.; Hong, H. Classification of Acetic Acid Bacteria and Their Acid Resistant Mechanism. AMB Expr. 2021, 11, 29. [Google Scholar] [CrossRef]
  75. Guzman, J.; Vilcinskas, A. Genome Analysis Suggests the Bacterial Family Acetobacteraceae Is a Source of Undiscovered Specialized Metabolites. Antonie Leeuwenhoek 2022, 115, 41–58. [Google Scholar] [CrossRef]
  76. Komagata, K.; Iino, T.; Yamada, Y. The Family Acetobacteraceae. In The Prokaryotes; Rosenberg, E., DeLong, E.F., Lory, S., Stackebrandt, E., Thompson, F., Eds.; Springer: Berlin/Heidelberg, Germany, 2014; pp. 3–78. ISBN 978-3-642-30196-4. [Google Scholar]
  77. Skerman, V.B.D.; McGowan, V.; Sneath, P.H.A. Approved Lists of Bacterial Names. Int. J. Syst. Bacteriol. 1980, 30, 225–420. [Google Scholar] [CrossRef]
  78. Gillis, M.; De Ley, J. Intra- and Intergeneric Similarities of the Ribosomal Ribonucleic Acid Cistrons of Acetobacter and Gluconobacter. Int. J. Syst. Bacteriol. 1980, 30, 7–27. [Google Scholar] [CrossRef]
  79. Yamada, Y. Acetobacter xylinus Sp. nov., nom. rev., for the Cellulose-Forming and Cellulose-Less, Acetate-Oxidizing Acetic Acid Bacteria with the Q-10 System. J. Gen. Appl. Microbiol. 1983, 29, 417–420. [Google Scholar] [CrossRef]
  80. Yamada, Y.; Yukphan, P.; Lan Vu, H.T.; Muramatsu, Y.; Ochaikul, D.; Tanasupawat, S.; Nakagawa, Y. Description of Komagataeibacter gen. nov., with Proposals of New Combinations (Acetobacteraceae). J. Gen. Appl. Microbiol. 2012, 58, 397–404. [Google Scholar] [CrossRef] [PubMed]
  81. Cleenwerck, I.; De Vos, P. Polyphasic Taxonomy of Acetic Acid Bacteria: An Overview of the Currently Applied Methodology. Int. J. Food Microbiol. 2008, 125, 2–14. [Google Scholar] [CrossRef]
  82. Trček, J.; Jernejc, K.; Matsushita, K. The Highly Tolerant Acetic Acid Bacterium Gluconacetobacter europaeus Adapts to the Presence of Acetic Acid by Changes in Lipid Composition, Morphological Properties and PQQ-Dependent ADH Expression. Extremophiles 2007, 11, 627–635. [Google Scholar] [CrossRef]
  83. Hördt, A.; López, M.G.; Meier-Kolthoff, J.P.; Schleuning, M.; Weinhold, L.-M.; Tindall, B.J.; Gronow, S.; Kyrpides, N.C.; Woyke, T.; Göker, M. Analysis of 1000+ Type-Strain Genomes Substantially Improves Taxonomic Classification of Alphaproteobacteria. Front. Microbiol. 2020, 11, 468. [Google Scholar] [CrossRef]
  84. Parte, A.C.; Sardà Carbasse, J.; Meier-Kolthoff, J.P.; Reimer, L.C.; Göker, M. List of Prokaryotic Names with Standing in Nomenclature (LPSN) Moves to the DSMZ. Int. J. Syst. Bacteriol. 2020, 70, 5607–5612. [Google Scholar] [CrossRef]
  85. Cleenwerck, I.; Vandemeulebroecke, K.; Janssens, D.; Swings, J. Re-Examination of the Genus Acetobacter, with Descriptions of Acetobacter cerevisiae Sp. nov. and Acetobacter malorum Sp. nov. Int. J. Syst. Bacteriol. 2002, 52, 1551–1558. [Google Scholar] [CrossRef]
  86. Lisdiyanti, P.; Kawasaki, H.; Seki, T.; Yamada, Y.; Uchimura, T.; Komagata, K. Identification of Acetobacter Strains Isolated from Indonesian Sources, and Proposals of Acetobacter syzygii Sp. nov., Acetobacter cibinongensis Sp. nov., and Acetobacter orientalis Sp. nov. J. Gen. Appl. Microbiol. 2001, 47, 119–131. [Google Scholar] [CrossRef]
  87. Sombolestani, A.S.; Cleenwerck, I.; Cnockaert, M.; Borremans, W.; Wieme, A.D.; De Vuyst, L.; Vandamme, P. novel Acetic Acid Bacteria from Cider Fermentations: Acetobacter conturbans Sp. nov. and Acetobacter fallax Sp. nov. Int. J. Syst. Bacteriol. 2020, 70, 6163–6171. [Google Scholar] [CrossRef]
  88. Lisdiyanti, P.; Kawasaki, H.; Seki, T.; Yamada, Y.; Uchimura, T.; Komagata, K. Systematic Study of the Genus Acetobacter with Descriptions of Acetobacter indonesiensis Sp. nov., Acetobacter tropicalis Sp. nov., Acetobacter orleanensis (Henneberg 1906) Comb. nov., Acetobacter lovaniensis (Frateur 1950) Comb. nov., and Acetobacter estunensis (Carr 1958) Comb. nov. J. Gen. Appl. Microbiol. 2000, 46, 147–165. [Google Scholar] [CrossRef] [PubMed]
  89. Cleenwerck, I.; Gonzalez, A.; Camu, N.; Engelbeen, K.; De Vos, P.; De Vuyst, L. Acetobacter fabarum Sp. nov., an Acetic Acid Bacterium from a Ghanaian Cocoa Bean Heap Fermentation. Int. J. Syst. Bacteriol. 2008, 58, 2180–2185. [Google Scholar] [CrossRef] [PubMed]
  90. Tanasupawat, S.; Kommanee, J.; Yukphan, P.; Muramatsu, Y.; Nakagawa, Y.; Yamada, Y. Acetobacter Farinalis Sp. nov., an Acetic Acid Bacterium in the α-Proteobacteria. J. Gen. Appl. Microbiol. 2011, 57, 159–167. [Google Scholar] [CrossRef] [PubMed]
  91. Yukphan, P.; Charoenyingcharoen, P.; Kingcha, Y.; Likhitrattanapisal, S.; Muangham, S.; Tanasupawat, S.; Yamada, Y. Acetobacter garciniae Sp. nov., an Acetic Acid Bacterium Isolated from Fermented Mangosteen Peel in Thailand. Int. J. Syst. Bacteriol. 2021, 71, 005052. [Google Scholar] [CrossRef] [PubMed]
  92. Cleenwerck, I.; Camu, N.; Engelbeen, K.; De Winter, T.; Vandemeulebroecke, K.; De Vos, P.; De Vuyst, L. Acetobacter ghanensis Sp. nov., a novel Acetic Acid Bacterium Isolated from Traditional Heap Fermentations of Ghanaian Cocoa Beans. Int. J. Syst. Bacteriol. 2007, 57, 1647–1652. [Google Scholar] [CrossRef] [PubMed]
  93. Spitaels, F.; Li, L.; Wieme, A.; Balzarini, T.; Cleenwerck, I.; Van Landschoot, A.; De Vuyst, L.; Vandamme, P. Acetobacter lambici Sp. nov., Isolated from Fermenting Lambic Beer. Int. J. Syst. Bacteriol. 2014, 64, 1083–1089. [Google Scholar] [CrossRef]
  94. Ferrer, S.; Mañes-Lázaro, R.; Benavent-Gil, Y.; Yépez, A.; Pardo, I. Acetobacter musti Sp. nov., Isolated from Bobal Grape Must. Int. J. Syst. Bacteriol. 2016, 66, 957–961. [Google Scholar] [CrossRef]
  95. Dutta, D.; Gachhui, R. novel Nitrogen-Fixing Acetobacter nitrogenifigens Sp. nov., Isolated from Kombucha Tea. Int. J. Syst. Bacteriol. 2006, 56, 1899–1903. [Google Scholar] [CrossRef]
  96. Silva, L.R. Acetobacter oeni Sp. nov., Isolated from Spoiled Red Wine. Iny. J. Syst. Bacteriol. 2006, 56, 21–24. [Google Scholar] [CrossRef]
  97. Lino, T.; Suzuki, R.; Kosako, Y.; Ohkuma, M.; Komagata, K.; Uchimura, T. Acetobacter okinawensis Sp. nov., Acetobacter papayae Sp. nov., and Acetobacter persicus Sp. nov.; novel Acetic Acid Bacteria Isolated from Stems of Sugarcane, Fruits, and a Flower in Japan. J. Gen. Appl. Microbiol. 2012, 58, 235–243. [Google Scholar] [CrossRef]
  98. Baek, J.H.; Kim, K.H.; Moon, J.Y.; Yeo, S.-H.; Jeon, C.O. Acetobacter oryzoeni Sp. nov., Isolated from Korean Rice Wine Vinegar. Int. J. Syst. Bacteriol. 2020, 70, 2026–2033. [Google Scholar] [CrossRef] [PubMed]
  99. Kim, K.H.; Cho, G.Y.; Chun, B.H.; Weckx, S.; Moon, J.Y.; Yeo, S.H.; Jeon, C.O. Acetobacter oryzifermentans sp. nov., Isolated from Korean Traditional Vinegar and Reclassification of the Type Strains of Acetobacter pasteurianus subsp. ascendens (Henneberg 1898) and Acetobacter pasteurianus subsp. paradoxus (Frateur 1950) as Acetobacter ascendens sp. nov., comb. nov. Syst. Appl. Microbiol. 2018, 41, 324–332. [Google Scholar] [CrossRef] [PubMed]
  100. Sokollek, S.J.; Hertel, C.; Hammes, W.P. Description of Acetobacter oboediens Sp. nov. and Acetobacter pomorum Sp. nov., Two New Species Isolated from Industrial Vinegar Fermentations. Int. J. Syst. Bacteriol. 1998, 48, 935–940. [Google Scholar] [CrossRef] [PubMed]
  101. Vu, H.T.L.; Yukphan, P.; Bui, V.T.T.; Charoenyingcharoen, P.; Malimas, S.; Nguyen, L.K.; Muramatsu, Y.; Tanaka, N.; Tanasupawat, S.; Le, B.T.; et al. Acetobacter sacchari Sp. nov., for a Plant Growth-Promoting Acetic Acid Bacterium Isolated in Vietnam. Ann. Microbiol. 2019, 69, 1155–1163. [Google Scholar] [CrossRef]
  102. Ndoye, B.; Cleenwerck, I.; Engelbeen, K.; Dubois-Dauphin, R.; Guiro, A.T.; Van Trappen, S.; Willems, A.; Thonart, P. Acetobacter senegalensis Sp. nov., a Thermotolerant Acetic Acid Bacterium Isolated in Senegal (Sub-Saharan Africa) from Mango Fruit (Mangifera indica L.). Int. J. Syst. Bacteriol. 2007, 57, 1576–1581. [Google Scholar] [CrossRef] [PubMed]
  103. Li, L.; Wieme, A.; Spitaels, F.; Balzarini, T.; Nunes, O.C.; Manaia, C.M.; Van Landschoot, A.; De Vuyst, L.; Cleenwerck, I.; Vandamme, P. Acetobacter sicerae Sp. nov., Isolated from Cider and Kefir, and Identification of Species of the Genus Acetobacter by DnaK, GroEL and RpoB Sequence Analysis. Int. J. Syst. Bacteriol. 2014, 64, 2407–2415. [Google Scholar] [CrossRef]
  104. Pitiwittayakul, N.; Theeragool, G.; Yukphan, P.; Chaipitakchonlatarn, W.; Malimas, T.; Muramatsu, Y.; Tanasupawat, S.; Nakagawa, Y.; Yamada, Y. Acetobacter suratthanensis Sp. nov., an Acetic Acid Bacterium Isolated in Thailand. Ann. Microbiol. 2016, 66, 1157–1166. [Google Scholar] [CrossRef]
  105. Pitiwittayakul, N.; Yukphan, P.; Chaipitakchonlatarn, W.; Yamada, Y.; Theeragool, G. Acetobacter thailandicus Sp. nov., for a Strain Isolated in Thailand. Ann. Microbiol. 2015, 65, 1855–1863. [Google Scholar] [CrossRef]
  106. Urakami, T.; Tamaoka, J.; Suzuki, K.-I.; Komagata, K. Acidomonas Gen. nov., Incorporating Acetobacter methanolicus as Acidomonas methanolica comb. nov. Int. J. Syst. Bacteriol. 1989, 39, 50–55. [Google Scholar] [CrossRef]
  107. Yukphan, P.; Malimas, T.; Muramatsu, Y.; Takahashi, M.; Kaneyasu, M.; Potacharoen, W.; Tanasupawat, S.; Nakagawa, Y.; Hamana, K.; Tahara, Y.; et al. Ameyamaea chiangmaiensis gen. nov., Sp. nov., an Acetic Acid Bacterium in the α-Proteobacteria. Biosci. Biotechnol. Biochem. 2009, 73, 2156–2162. [Google Scholar] [CrossRef]
  108. Suzuki, R.; Zhang, Y.; Iino, T.; Kosako, Y.; Komagata, K.; Uchimura, T. Asaia astilbes Sp. nov., Asaia platycodi Sp. nov., and Asaia prunellae Sp. nov., novel Acetic Acid Bacteria Isolated from Flowers in Japan. J. Gen. Appl. Microbiol. 2010, 56, 339–346. [Google Scholar] [CrossRef]
  109. Yamada, Y.; Katsura, K.; Kawasaki, H.; Widyastuti, Y.; Saono, S.; Seki, T.; Uchimura, T.; Komagata, K. Asaia bogorensis gen. nov., Sp. nov., an Unusual Acetic Acid Bacterium in the α-Proteobacteria. Int. J. Syst. Bacteriol. 2000, 50, 823–829. [Google Scholar] [CrossRef] [PubMed]
  110. Yukphan, P.; Potacharoen, W.; Tanasupawat, S.; Tanticharoen, M.; Yamada, Y. Asaia krungthepensis Sp. nov., an Acetic Acid Bacterium in the α-Proteobacteria. Int. J. Syst. Bacteriol. 2004, 54, 313–316. [Google Scholar] [CrossRef] [PubMed]
  111. Malimas, T.; Yukphan, P.; Takahashi, M.; Kaneyasu, M.; Potacharoen, W.; Tanasupawat, S.; Nakagawa, Y.; Tanticharoen, M.; Yamada, Y. Asaia lannaensis Sp. nov., a New Acetic Acid Bacterium in the Alphaproteobacteria. Biosci. Biotechnol. Biochem. 2008, 72, 666–671. [Google Scholar] [CrossRef] [PubMed]
  112. Katsura, K.; Kawasaki, H.; Potacharoen, W.; Saono, S.; Seki, T.; Yamada, Y.; Uchimura, T.; Komagata, K. Asaia siamensis Sp. nov., an Acetic Acid Bacterium in the α-Proteobacteria. Int. J. Syst. Bacteriol. 2001, 51, 559–563. [Google Scholar] [CrossRef] [PubMed]
  113. Kommanee, J.; Tanasupawat, S.; Yukphan, P.; Malimas, T.; Muramatsu, Y.; Nakagawa, Y.; Yamada, Y. Asaia spathodeae Sp. nov., an Acetic Acid Bacterium in the α-Proteobacteria. J. Gen. Appl. Microbiol. 2010, 56, 81–87. [Google Scholar] [CrossRef]
  114. Yun, J.-H.; Lee, J.-Y.; Hyun, D.-W.; Jung, M.-J.; Bae, J.-W. Bombella apis Sp. nov., an Acetic Acid Bacterium Isolated from the Midgut of a Honey Bee. Int. J. Syst. Bacteriol. 2017, 67, 2184–2188. [Google Scholar] [CrossRef]
  115. Hilgarth, M.; Redwitz, J.; Ehrmann, M.A.; Vogel, R.F.; Jakob, F. Bombella favorum Sp. nov. and Bombella mellum Sp. nov., Two novel Species Isolated from the Honeycombs of Apis mellifera. Int. J. Syst. Bacteriol. 2021, 71, 004633. [Google Scholar] [CrossRef]
  116. Li, L.; Praet, J.; Borremans, W.; Nunes, O.C.; Manaia, C.M.; Cleenwerck, I.; Meeus, I.; Smagghe, G.; De Vuyst, L.; Vandamme, P. Bombella intestini Gen. nov., Sp. nov., an Acetic Acid Bacterium Isolated from Bumble Bee Crop. Int. J. Syst. Bacteriol. 2015, 65, 267–273. [Google Scholar] [CrossRef]
  117. Roh, S.W.; Nam, Y.-D.; Chang, H.-W.; Kim, K.-H.; Kim, M.-S.; Ryu, J.-H.; Kim, S.-H.; Lee, W.-J.; Bae, J.-W. Phylogenetic Characterization of Two novel Commensal Bacteria Involved with Innate Immune Homeostasis in Drosophila melanogaster. Appl. Environ. Microbiol. 2008, 74, 6171–6177. [Google Scholar] [CrossRef]
  118. Ramírez-Bahena, M.H.; Tejedor, C.; Martín, I.; Velázquez, E.; Peix, A. Endobacter medicaginis gen. nov., Sp. nov., Isolated from Alfalfa Nodules in an Acidic Soil. Int. J. Syst. Bacteriol. 2013, 63, 1760–1765. [Google Scholar] [CrossRef] [PubMed]
  119. Nishijima, M.; Tazato, N.; Handa, Y.; Tomita, J.; Kigawa, R.; Sano, C.; Sugiyama, J. Gluconacetobacter tumulisoli Sp. nov., Gluconacetobacter takamatsuzukensis Sp. nov. and Gluconacetobacter aggeris Sp. nov., Isolated from Takamatsuzuka Tumulus Samples before and during the Dismantling Work in 2007. Int. J. Syst. Bacteriol. 2013, 63, 3981–3988. [Google Scholar] [CrossRef] [PubMed]
  120. Tazato, N.; Nishijima, M.; Handa, Y.; Kigawa, R.; Sano, C.; Sugiyama, J. Gluconacetobacter tumulicola Sp. nov. and Gluconacetobacter asukensis Sp. nov., Isolated from the Stone Chamber Interior of the Kitora Tumulus. Int. J. Syst. Bacteriol. 2012, 62, 2032–2038. [Google Scholar] [CrossRef] [PubMed]
  121. Fuentes-Ramírez, L.E.; Bustillos-Cristales, R.; Tapia-Hernández, A.; Jiménez-Salgado, T.; Wang, E.T.; Martínez-Romero, E.; Caballero-Mellado, J. novel Nitrogen-Fixing Acetic Acid Bacteria, Gluconacetobacter johannae Sp. nov. and Gluconacetobacter azotocaptans Sp. nov., Associated with Coffee Plants. Int. J. Syst. Bacteriol. 2001, 51, 1305–1314. [Google Scholar] [CrossRef]
  122. Yamada, Y.; Hoshino, K.; Ishikawa, T. The Phylogeny of Acetic Acid Bacteria Based on the Partial Sequences of 16S Ribosomal RNA: The Elevation of the Subgenus Gluconoacetobacter to the Generic Level. Biosci. Biotechnol. Biochem. 1997, 61, 1244–1251. [Google Scholar] [CrossRef]
  123. Sombolestani, A.S.; Cleenwerck, I.; Cnockaert, M.; Borremans, W.; Wieme, A.D.; Moutia, Y.; Spaepen, S.; De Vuyst, L.; Vandamme, P. Gluconacetobacter dulcium Sp. nov., a novel Gluconacetobacter Species from Sugar-Rich Environments. Int. J. Syst. Bacteriol. 2021, 71, 004569. [Google Scholar] [CrossRef]
  124. Schüller, G.; Hertel, C.; Hammes, W.P. Gluconacetobacter entanii Sp. nov., Isolated from Submerged High-Acid Industrial Vinegar Fermentations. Int. J. Syst. Bacteriol. 2000, 50, 2013–2020. [Google Scholar] [CrossRef]
  125. Franke, I.H.; Fegan, M.; Hayward, C.; Leonard, G.; Stackebrandt, E.; Sly, L.I. Description of Gluconacetobacter sacchari Sp. nov., a New Species of Acetic Acid Bacterium Isolated from the Leaf Sheath of Sugar Cane and from the Pink Sugar-Cane Mealy Bug. Int. J. Syst. Bacteriol. 1999, 49, 1681–1693. [Google Scholar] [CrossRef]
  126. Yukphan, P.; Charoenyingcharoen, P.; Malimas, S.; Muramatsu, Y.; Nakagawa, Y.; Tanasupawat, S.; Yamada, Y. Gluconobacter aidae Sp. nov., an Acetic Acid Bacteria Isolated from Tropical Fruits in Thailand. Int. J. Syst. Bacteriol. 2020, 70, 4351–4357. [Google Scholar] [CrossRef]
  127. Yukphan, P.; Takahashi, M.; Potacharoen, W.; Tanasupawat, S.; Nakagawa, Y.; Tanticharoen, M.; Yamada, Y. Gluconobacter albidus (Ex Kondo and Ameyama 1958) Sp. nov., nom. rev., an Acetic Acid Bacterium in the α-Proteobacteria. J. Gen. Appl. Microbiol. 2004, 50, 235–242. [Google Scholar] [CrossRef]
  128. Sombolestani, A.S.; Cleenwerck, I.; Cnockaert, M.; Borremans, W.; Wieme, A.D.; De Vuyst, L.; Vandamme, P. Characterization of novel Gluconobacter Species from Fruits and Fermented Food Products: Gluconobacter cadivus Sp. nov., Gluconobacter vitians Sp. nov. and Gluconobacter potus Sp. nov. Int. J. Syst. Bacteriol. 2019, 71, 004751. [Google Scholar] [CrossRef] [PubMed]
  129. Spitaels, F.; Wieme, A.; Balzarini, T.; Cleenwerck, I.; Van Landschoot, A.; De Vuyst, L.; Vandamme, P. Gluconobacter cerevisiae Sp. nov., Isolated from the Brewery Environment. Int. J. Syst. Bacteriol. 2014, 64, 1134–1141. [Google Scholar] [CrossRef] [PubMed]
  130. Yamada, Y.; Akita, M. An Electrophoretic Comparison of Enzymes in Strains of Gluconobacter Species. J. Gen. Appl. Microbiol. 1984, 30, 115–126. [Google Scholar] [CrossRef]
  131. Mason, L.M.; Claus, G.W. Phenotypic Characteristics Correlated with Deoxyribonucleic Acid Sequence Similarities for Three Species of Gluconobacter: G. Oxydans (Henneberg 1897) De Ley 1961, G. frateurii Sp. nov., and G. asaii Sp. nov. Int. J. Syst. Bacteriol. 1989, 39, 174–184. [Google Scholar] [CrossRef]
  132. Malimas, T.; Yukphan, P.; Takahashi, M.; Muramatsu, Y.; Kaneyasu, M.; Potacharoen, W.; Tanasupawat, S.; Nakagawa, Y.; Tanticharoen, M.; Yamada, Y. Gluconobacter japonicus Sp. nov., an Acetic Acid Bacterium in the Alphaproteobacteria. Int. J. Syst. Bacteriol. 2009, 59, 466–471. [Google Scholar] [CrossRef] [PubMed]
  133. Malimas, T.; Yukphan, P.; Lundaa, T.; Muramatsu, Y.; Takahashi, M.; Kaneyasu, M.; Potacharoen, W.; Tanasupawat, S.; Nakagawa, Y.; Suzuki, K.; et al. Gluconobacter kanchanaburiensis Sp. nov., a Brown Pigment-Producing Acetic Acid Bacterium for Thai Isolates in the Alphaproteobacteria. J. Gen. Appl. Microbiol. 2009, 55, 247–254. [Google Scholar] [CrossRef]
  134. Malimas, T.; Yukphan, P.; Takahashi, M.; Kaneyasu, M.; Potacharoen, W.; Tanasupawat, S.; Nakagawa, Y.; Tanticharoen, M.; Yamada, Y. Gluconobacter kondonii Sp. nov., an Acetic Acid Bacterium in the α-Proteobacteria. J. Gen. Appl. Microbiol. 2007, 53, 301–307. [Google Scholar] [CrossRef]
  135. De Ley, J. Comparative Carbohydrate Metabolism and a Proposal for a Phylogenetic Relationship of the Acetic Acid Bacteria. J. Gen. Microbiol. 1961, 24, 31–50. [Google Scholar] [CrossRef]
  136. Malimas, T.; Yukphan, P.; Takahashi, M.; Muramatsu, Y.; Kaneyasu, M.; Potacharoen, W.; Tanasupawat, S.; Nakagawa, Y.; Tanticharoen, M.; Yamada, Y. Gluconobacter roseus (Ex Asai 1935) Sp. nov., nom. rev., a Pink-Colored Acetic Acid Bacterium in the Alphaproteobacteria. J. Gen. Appl. Microbiol. 2008, 54, 119–125. [Google Scholar] [CrossRef]
  137. Malimas, T.; Yukphan, P.; Takahashi, M.; Muramatsu, Y.; Kaneyasu, M.; Potacharoen, W.; Tanasupawat, S.; Nakagawa, Y.; Tanticharoen, M.; Yamada, Y. Gluconobacter sphaericus (Ameyama 1975) Comb. nov., a Brown Pigment-Producing Acetic Acid Bacterium in the Alphaproteobacteria. J. Gen. Appl. Microbiol. 2008, 54, 211–220. [Google Scholar] [CrossRef]
  138. Tanasupawat, S.; Thawai, C.; Yukphan, P.; Moonmangmee, D.; Itoh, T.; Adachi, O.; Yamada, Y. Gluconobacter thailandicus Sp. nov., an Acetic Acid Bacterium in the α-Proteobacteria. J. Gen. Appl. Microbiol. 2004, 50, 159–167. [Google Scholar] [CrossRef] [PubMed]
  139. Yukphan, P.; Malimas, T.; Lundaa, T.; Muramatsu, Y.; Takahashi, M.; Kaneyasu, M.; Tanasupawat, S.; Nakagawa, Y.; Suzuki, K.; Tanticharoen, M.; et al. Gluconobacter wancherniae Sp. nov., an Acetic Acid Bacterium from Thai Isolates in the α-Proteobacteria. J. Gen. Appl. Microbiol. 2010, 56, 67–73. [Google Scholar] [CrossRef] [PubMed]
  140. Greenberg, D.E.; Porcella, S.F.; Stock, F.; Wong, A.; Conville, P.S.; Murray, P.R.; Holland, S.M.; Zelazny, A.M. Granulibacter bethesdensis gen. nov., Sp. nov., a Distinctive Pathogenic Acetic Acid Bacterium in the Family Acetobacteraceae. Int. J. Syst. Evol. Microbiol. 2006, 56, 2609–2616. [Google Scholar] [CrossRef]
  141. Naloka, K.; Yukphan, P.; Matsutani, M.; Matsushita, K.; Theeragool, G. Komagataeibacter diospyri Sp. nov., a novel Species of Thermotolerant Bacterial Nanocellulose-Producing Bacterium. Int. J. Syst. Bacteriol. 2020, 70, 251–258. [Google Scholar] [CrossRef]
  142. Yamada, Y. Transfer of Gluconacetobacter kakiaceti, Gluconacetobacter medellinensis and Gluconacetobacter maltaceti to the Genus Komagataeibacter as Komagataeibacter kakiaceti comb. nov., Komagataeibacter medellinensis comb. nov. and Komagataeibacter maltaceti comb. nov. Int. J. Syst. Bacteriol. 2014, 64, 1670–1672. [Google Scholar] [CrossRef]
  143. Marič, L.; Cleenwerck, I.; Accetto, T.; Vandamme, P.; Trček, J. Description of Komagataeibacter melaceti Sp. nov. and Komagataeibacter melomenusus Sp. nov. Isolated from Apple Cider Vinegar. Microorganisms 2020, 8, 1178. [Google Scholar] [CrossRef] [PubMed]
  144. Lisdiyanti, P.; Kawasaki, H.; Widyastuti, Y.; Saono, S.; Seki, T.; Yamada, Y.; Uchimura, T.; Komagata, K. Kozakia baliensis Gen. nov., Sp. nov., a novel Acetic Acid Bacterium in the α-Proteobacteria. Int. J. Syst. Bacteriol. 2002, 52, 813–818. [Google Scholar] [CrossRef]
  145. Yukphan, P.; Malimas, T.; Potacharoen, W.; Tanasupawat, S.; Tanticharoen, M.; Yamada, Y. Neoasaia chiangmaiensis gen. nov., Sp. nov., a novel Osmotolerant Acetic Acid Bacterium in the α-Proteobacteria. J. Gen. Appl. Microbiol. 2005, 51, 301–311. [Google Scholar] [CrossRef]
  146. Yukphan, P.; Malimas, T.; Muramatsu, Y.; Potacharoen, W.; Tanasupawat, S.; Nakagawa, Y.; Tanticharoen, M.; Yamada, Y. Neokomagataea Gen. nov., with Descriptions of Neokomagataea thailandica Sp. nov. and Neokomagataea tanensis Sp. nov., Osmotolerant Acetic Acid Bacteria of the α-Proteobacteria. Biosci. Biotechnol. Biochem. 2011, 75, 419–426. [Google Scholar] [CrossRef]
  147. Thi Lan Vu, H.; Yukphan, P.; Chaipitakchonlatarn, W.; Malimas, T.; Muramatsu, Y.; Thi Tu Bui, U.; Tanasupawat, S.; Cong Duong, K.; Nakagawa, Y.; Thanh Pham, H.; et al. Nguyenibacter vanlangensis gen. nov., Sp. nov., an Unusual Acetic Acid Bacterium in the α-Proteobacteria. J. Gen. Appl. Microbiol. 2013, 59, 153–166. [Google Scholar] [CrossRef]
  148. Brandão, P.R.; Crespo, M.T.B.; Nascimento, F.X. Phylogenomic and Comparative Analyses Support the Reclassification of Several Komagataeibacter Species as novel Members of the Novacetimonas gen. nov. and Bring New Insights into the Evolution of Cellulose Synthase Genes. Int. J. Syst. Bacteriol. 2022, 72, 005252. [Google Scholar] [CrossRef] [PubMed]
  149. Jojima, Y.; Mihara, Y.; Suzuki, S.; Yokozeki, K.; Yamanaka, S.; Fudou, R. Saccharibacter floricola gen. nov., Sp. nov., a novel Osmophilic Acetic Acid Bacterium Isolated from Pollen. Int. J. Syst. Bacteriol. 2004, 54, 2263–2267. [Google Scholar] [CrossRef] [PubMed]
  150. Loganathan, P.; Nair, S. Swaminathania salitolerans gen. nov., Sp. nov., a Salt-Tolerant, Nitrogen-Fixing and Phosphate-Solubilizing Bacterium from Wild Rice (Porteresia Coarctata Tateoka). Int. J. Syst. Bacteriol. 2004, 54, 1185–1190. [Google Scholar] [CrossRef] [PubMed]
  151. Malimas, T.; Chaipitakchonlatarn, W.; Thi Lan Vu, H.; Yukphan, P.; Muramatsu, Y.; Tanasupawat, S.; Potacharoen, W.; Nakagawa, Y.; Tanticharoen, M.; Yamada, Y. Swingsia samuiensis gen. nov., Sp. nov., an Osmotolerant Acetic Acid Bacterium in the α-Proteobacteria. J. Gen. Appl. Microbiol. 2013, 59, 375–384. [Google Scholar] [CrossRef] [PubMed]
  152. Vu, H.T.L.; Malimas, T.; Chaipitakchonlatarn, W.; Bui, V.T.T.; Yukphan, P.; Bui, U.T.T.; Muramatsu, Y.; Sitdhipol, J.; Tanasupawat, S.; Duong, K.C.; et al. Tanticharoenia aidae Sp. nov., for Acetic Acid Bacteria Isolated in Vietnam. Ann. Microbiol. 2016, 66, 417–423. [Google Scholar] [CrossRef]
  153. Yukphan, P.; Malimas, T.; Muramatsu, Y.; Takahashi, M.; Kaneyasu, M.; Tanasupawat, S.; Nakagawa, Y.; Suzuki, K.; Potacharoen, W.; Yamada, Y. Tanticharoenia sakaeratensis Gen. nov., Sp. nov., a New Osmotolerant Acetic Acid Bacterium in the α-Proteobacteria. Biosci. Biotechnol. Biochem. 2008, 72, 672–676. [Google Scholar] [CrossRef] [PubMed]
  154. Li, Y.; Yan, P.; Lei, Q.; Li, B.; Sun, Y.; Li, S.; Lei, H.; Xie, N. Metabolic Adaptability Shifts of Cell Membrane Fatty Acids of Komagataeibacter hansenii HDM1-3 Improve Acid Stress Resistance and Survival in Acidic Environments. J. Ind. Microbiol. Biotechnol. 2019, 46, 1491–1503. [Google Scholar] [CrossRef]
  155. Sankuan, X.; Cuimei, Z.; Bingqian, F.; Yu, Z.; Menglei, X.; Linna, T.; Jia, S.; Xinyi, Z.; Min, W. Metabolic Network of Ammonium in Cereal Vinegar Solid-State Fermentation and Its Response to Acid Stress. Food Microbiol. 2021, 95, 103684. [Google Scholar] [CrossRef]
  156. Sriherfyna, F.H.; Matsutani, M.; Hirano, K.; Koike, H.; Kataoka, N.; Yamashita, T.; Nakamaru-Ogiso, E.; Matsushita, K.; Yakushi, T. The Auxiliary NADH Dehydrogenase Plays a Crucial Role in Redox Homeostasis of Nicotinamide Cofactors in the Absence of the Periplasmic Oxidation System in Gluconobacter oxydans NBRC3293. Appl. Environ. Microbiol. 2021, 87, e02155-20. [Google Scholar] [CrossRef]
  157. Wang, L.; Hong, H.; Zhang, C.; Huang, Z.; Guo, H. Transcriptome Analysis of Komagataeibacter europaeus CGMCC 20445 Responses to Different Acidity Levels During Acetic Acid Fermentation. Pol. J. Microbiol. 2021, 70, 305–313. [Google Scholar] [CrossRef]
  158. Zheng, Y.; Zhang, R.; Yin, H.; Bai, X.; Chang, Y.; Xia, M.; Wang, M. Acetobacter pasteurianus Metabolic Change Induced by Initial Acetic Acid to Adapt to Acetic Acid Fermentation Conditions. Appl. Microbiol. Biotechnol. 2017, 101, 7007–7016. [Google Scholar] [CrossRef] [PubMed]
  159. Adachi, O.; Tayama, K.; Shinagawa, E.; Matsushita, K.; Ameyama, M. Purification and Characterization of Membrane-Bound Aldehyde Dehydrogenase from Gluconobacter suboxydans. Agric. Biol. Chem. 1980, 44, 503–515. [Google Scholar] [CrossRef]
  160. Ameyama, M.; Adachi, O. Alcohol Dehydrogenase from Acetic Acid Bacteria, Membrane-Bound. Meth. Enzymol. 1982, 89, 450–457. [Google Scholar]
  161. Qin, Z.; Yu, S.; Chen, J.; Zhou, J. Dehydrogenases of Acetic Acid Bacteria. Biotechnol. Adv. 2022, 54, 107863. [Google Scholar] [CrossRef]
  162. Mullins, E.A.; Francois, J.A.; Kappock, T.J. A specialized citric acid cycle requiring succinyl-coenzyme A (CoA): Acetate CoA-transferase (AarC) confers acetic acid resistance on the acidophile Acetobacter aceti. J. Bacteriol. 2008, 190, 4933–4940. [Google Scholar] [CrossRef]
  163. Arai, H.; Sakurai, K.; Ishii, M. Metabolic features of Acetobacter aceti. In Acetic Acid Bacteria; Matsushita, K., Toyama, H., Tonouchi, N., Okamoto-Kainuma, A., Eds.; Springer: Tokyo, Japan, 2016; pp. 255–271. ISBN 978-4-431-55931-3. [Google Scholar]
  164. Yakushi, T.; Matsushita, K. Alcohol Dehydrogenase of Acetic Acid Bacteria: Structure, Mode of Action, and Applications in Biotechnology. Appl. Microbiol. Biotechnol. 2010, 86, 1257–1265. [Google Scholar] [CrossRef]
  165. Adler, P.; Frey, L.J.; Berger, A.; Bolten, C.J.; Hansen, C.E.; Wittmann, C. The Key to Acetate: Metabolic Fluxes of Acetic Acid Bacteria under Cocoa Pulp Fermentation-Simulating Conditions. Appl. Environ. Microbiol. 2014, 80, 4702–4716. [Google Scholar] [CrossRef]
  166. García-García, I.; Cañete-Rodríguez, A.M.; Santos-Dueñas, I.M.; Jiménez-Hornero, J.E.; Ehrenreich, A.; Liebl, W.; García-Martínez, T.; Mauricio, J.C. Biotechnologically Relevant Features of Gluconic Acid Production by Acetic Acid Bacteria. Acetic Acid Bact. 2017, 6, 6458. [Google Scholar] [CrossRef]
  167. Yin, H.; Zhang, R.; Xia, M.; Bai, X.; Mou, J.; Zheng, Y.; Wang, M. Effect of Aspartic Acid and Glutamate on Metabolism and Acid Stress Resistance of Acetobacter pasteurianus. Microb Cell. Fact. 2017, 16, 109. [Google Scholar] [CrossRef]
  168. Bringer, S.; Bott, M. Central Carbon Metabolism and Respiration in Gluconobacter oxydans. In Acetic Acid Bacteria; Matsushita, K., Toyama, H., Tonouchi, N., Okamoto-Kainuma, A., Eds.; Springer: Tokyo, Japan, 2016; pp. 235–253. ISBN 978-4-431-55931-3. [Google Scholar]
  169. Kiefler, I.; Bringer, S.; Bott, M. Metabolic Engineering of Gluconobacter oxydans 621H for Increased Biomass Yield. Appl. Microbiol. Biotechnol. 2017, 101, 5453–5467. [Google Scholar] [CrossRef]
  170. Villarreal-Soto, S.A.; Beaufort, S.; Bouajila, J.; Souchard, J.-P.; Taillandier, P. Understanding Kombucha Tea Fermentation: A Review. J. Food Sci. 2018, 83, 580–588. [Google Scholar] [CrossRef] [PubMed]
  171. Ayed, L.; Ben Abid, S.; Hamdi, M. Development of a Beverage from Red Grape Juice Fermented with the Kombucha Consortium. Ann. Microbiol. 2017, 67, 111–121. [Google Scholar] [CrossRef]
  172. Mounir, M.; Shafiei, R.; Zarmehrkhorshid, R.; Hamouda, A.; Ismaili Alaoui, M.; Thonart, P. Simultaneous Production of Acetic and Gluconic Acids by a Thermotolerant Acetobacter Strain during Acetous Fermentation in a Bioreactor. J. Biosci. Bioeng. 2016, 121, 166–171. [Google Scholar] [CrossRef]
  173. Cañete-Rodríguez, A.M.; Santos-Dueñas, I.M.; Jiménez-Hornero, J.E.; Ehrenreich, A.; Liebl, W.; García-García, I. Gluconic Acid: Properties, Production Methods and Applications—An Excellent Opportunity for Agro-Industrial by-Products and Waste Bio-Valorization. Process Biochem. 2016, 51, 1891–1903. [Google Scholar] [CrossRef]
  174. Mohammadkazemi, F.; Azin, M.; Ashori, A. Production of Bacterial Cellulose Using Different Carbon Sources and Culture Media. Carbohydr. Polym. 2015, 117, 518–523. [Google Scholar] [CrossRef]
  175. Shi, Z.; Zhang, Y.; Phillips, G.O.; Yang, G. Utilization of Bacterial Cellulose in Food. Food Hydrocoll. 2014, 35, 539–545. [Google Scholar] [CrossRef]
  176. Ullah, H.; Wahid, F.; Santos, H.A.; Khan, T. Advances in Biomedical and Pharmaceutical Applications of Functional Bacterial Cellulose-Based Nanocomposites. Carbohydr. Polym. 2016, 150, 330–352. [Google Scholar] [CrossRef]
  177. Alvarez-Ordóñez, A.; Coughlan, L.M.; Briandet, R.; Cotter, P.D. Biofilms in Food Processing Environments: Challenges and Opportunities. Annu. Rev. Food Sci. Technol. 2019, 10, 173–195. [Google Scholar] [CrossRef]
  178. Peng, M.Y.; Zhang, X.J.; Huang, T.; Zhong, X.Z.; Chai, L.J.; Lu, Z.M.; Shi, J.S.; Xu, Z.H. Komagataeibacter europaeus improves community stability and function in solid-state cereal vinegar fermentation ecosystem: Non-abundant species plays important role. Food Res. Int. 2021, 150, 110815. [Google Scholar] [CrossRef]
  179. Yang, H.; He, Y.; Liao, J.; Li, X.; Zhang, J.; Liebl, W.; Chen, F. RNA-Seq Transcriptomic Analysis Reveals Gene Expression Profiles of Acetic Acid Bacteria under High-Acidity Submerged Industrial Fermentation Process. Front. Microbiol. 2022, 13, 956729. [Google Scholar] [CrossRef]
Figure 1. Functional properties and health benefits of vinegar on human metabolism.
Figure 1. Functional properties and health benefits of vinegar on human metabolism.
Foods 12 03705 g001
Figure 2. Submerged culture for vinegar production working in a semi-continuous mode. Each cycle of acetification starts by loading the tank to its working volume (8 L) without exceeding a preset ethanol concentration [5% (v/v)]. When ethanol concentration is depleted to 1.0–1.5% (v/v), 50% of the reactor content (4 L) is unloaded. This system is maintained for the following production cycles.
Figure 2. Submerged culture for vinegar production working in a semi-continuous mode. Each cycle of acetification starts by loading the tank to its working volume (8 L) without exceeding a preset ethanol concentration [5% (v/v)]. When ethanol concentration is depleted to 1.0–1.5% (v/v), 50% of the reactor content (4 L) is unloaded. This system is maintained for the following production cycles.
Foods 12 03705 g002
Figure 3. Pilot plant, on a laboratory scale, equipped with a Frings Acetator (8 L) working in a semi-continuous mode and a SCADA automation system that controls the main variables of the system. Biochemical Engineering Laboratory, Chemical Engineering Section, University of Córdoba, Spain.
Figure 3. Pilot plant, on a laboratory scale, equipped with a Frings Acetator (8 L) working in a semi-continuous mode and a SCADA automation system that controls the main variables of the system. Biochemical Engineering Laboratory, Chemical Engineering Section, University of Córdoba, Spain.
Foods 12 03705 g003
Figure 4. Images taken from a sample (two different planes) of vinegar, which is normally comprised of a complex microbiota of acetic acid bacteria, working inside the bioreactor used (Frings, 8 L) by scanning electron microscopy (SEM).
Figure 4. Images taken from a sample (two different planes) of vinegar, which is normally comprised of a complex microbiota of acetic acid bacteria, working inside the bioreactor used (Frings, 8 L) by scanning electron microscopy (SEM).
Foods 12 03705 g004
Figure 5. Incomplete oxidation reaction of ethanol into acetic acid both (1) at the cell membrane level (green box): PQQ-ADH, PQQ-dependent alcohol dehydrogenase; ALDH, membrane-bound aldehyde dehydrogenase; UB, ubiquinone; UBH2, ubiquinol; UOX, ubiquinol oxidase; and (2) the cytoplasm level (red box): NAD-ADH, NAD-dependent alcohol dehydrogenase; NAD-ALDH, NAD-dependent aldehyde dehydrogenase; NADP-ALDH, NADP-dependent aldehyde dehydrogenase. ATP, energy; TCA cycle, Tricarboxylic Acid Cycle. Adapted by the authors from He et al. [14].
Figure 5. Incomplete oxidation reaction of ethanol into acetic acid both (1) at the cell membrane level (green box): PQQ-ADH, PQQ-dependent alcohol dehydrogenase; ALDH, membrane-bound aldehyde dehydrogenase; UB, ubiquinone; UBH2, ubiquinol; UOX, ubiquinol oxidase; and (2) the cytoplasm level (red box): NAD-ADH, NAD-dependent alcohol dehydrogenase; NAD-ALDH, NAD-dependent aldehyde dehydrogenase; NADP-ALDH, NADP-dependent aldehyde dehydrogenase. ATP, energy; TCA cycle, Tricarboxylic Acid Cycle. Adapted by the authors from He et al. [14].
Foods 12 03705 g005
Figure 6. Molecular strategy of K. europaeus proposed to prevail throughout acetification process. The colors of the proteins represent the phase in which the protein had the highest quantification value. Fine wine (FW), beer (B), end of loading (EL), just before unloading (UL). A, oxidation of ethanol into acetic acid; B, TCA cycle; C, glycolysis; D, pentose phosphate pathway; E, amino acids and purines formation from biosynthetic precursors; F, regulation of protein synthesis; G, membrane mechanisms for acetic acid release. GAP, glyceraldehyde 3-phosphate; DHAP, dihydroxyacetone phosphate; 3PG, 3-phosphoglycerate; PEP, phosphoenol-pyruvate; FGAR, formylglycinamide ribonucleotide; FGAM, formylglycinamidine ribonucleotide; IMP, inosine monophosphate. Adapted by the authors from Román-Camacho et al. [6].
Figure 6. Molecular strategy of K. europaeus proposed to prevail throughout acetification process. The colors of the proteins represent the phase in which the protein had the highest quantification value. Fine wine (FW), beer (B), end of loading (EL), just before unloading (UL). A, oxidation of ethanol into acetic acid; B, TCA cycle; C, glycolysis; D, pentose phosphate pathway; E, amino acids and purines formation from biosynthetic precursors; F, regulation of protein synthesis; G, membrane mechanisms for acetic acid release. GAP, glyceraldehyde 3-phosphate; DHAP, dihydroxyacetone phosphate; 3PG, 3-phosphoglycerate; PEP, phosphoenol-pyruvate; FGAR, formylglycinamide ribonucleotide; FGAM, formylglycinamidine ribonucleotide; IMP, inosine monophosphate. Adapted by the authors from Román-Camacho et al. [6].
Foods 12 03705 g006
Figure 7. Glucose metabolism in Gluconobacter. Cell membrane enzymes: PQQ-GDH, PQQ-dependent D-glucose dehydrogenase; GADH, FAD-dependent D-gluconate 2-dehydrogenase; 2KGADH, FAD-dependent 2-keto-D-gluconate dehydrogenase; GA5DH, PQQ-dependent D-gluconate 5-dehydrogenase. Cytoplasm enzymes: GDH-NAD, NADP-dependent D-glucose dehydrogenase; GA5DH-NAD, NADP-dependent D-gluconate 5-dehydrogenase; 2KGR, 2-keto-D-gluconate reductase; 5KGR, 5-keto-D-gluconate reductase; G6PDH, glucose-6-phosphate dehydrogenase; 6PGD, 6-phosphogluconate dehydrogenase. Compounds: G3P, glyceraldehyde 3-phosphate. Pathways: PPP, pentose phosphate pathway; EDP, Entner Doudoroff pathway. Adapted by the authors from García-García et al. [166].
Figure 7. Glucose metabolism in Gluconobacter. Cell membrane enzymes: PQQ-GDH, PQQ-dependent D-glucose dehydrogenase; GADH, FAD-dependent D-gluconate 2-dehydrogenase; 2KGADH, FAD-dependent 2-keto-D-gluconate dehydrogenase; GA5DH, PQQ-dependent D-gluconate 5-dehydrogenase. Cytoplasm enzymes: GDH-NAD, NADP-dependent D-glucose dehydrogenase; GA5DH-NAD, NADP-dependent D-gluconate 5-dehydrogenase; 2KGR, 2-keto-D-gluconate reductase; 5KGR, 5-keto-D-gluconate reductase; G6PDH, glucose-6-phosphate dehydrogenase; 6PGD, 6-phosphogluconate dehydrogenase. Compounds: G3P, glyceraldehyde 3-phosphate. Pathways: PPP, pentose phosphate pathway; EDP, Entner Doudoroff pathway. Adapted by the authors from García-García et al. [166].
Foods 12 03705 g007
Table 1. Current classification of the acetic acid bacteria group. A total of 20 genera and 108 species are shown. Source: LPSN database (https://www.bacterio.net (accessed on 10 June 2022)).
Table 1. Current classification of the acetic acid bacteria group. A total of 20 genera and 108 species are shown. Source: LPSN database (https://www.bacterio.net (accessed on 10 June 2022)).
GeneraSpeciesReferences
Acetobacter1Acetobacter acetiSkerman et al. [77]
2Acetobacter cerevisiaeCleenwerck et al. [85]
3Acetobacter cibinongensisLisdiyanti et al. [86]
4Acetobacter conturbansSombolestani et al. [87]
5Acetobacter estunensisLisdiyanti et al. [88]
6Acetobacter fabarumCleenwerck et al. [89]
7Acetobacter fallaxSombolestani et al. [87]
8Acetobacter farinalisTanasupawat et al. [90]
9Acetobacter garciniaeYukphan et al. [91]
10Acetobacter ghanensisCleenwerck et al. [92]
11Acetobacter indonesiensisLisdiyanti et al. [88]
12Acetobacter lambiciSpitaels et al. [93]
13Acetobacter lovaniensisLisdiyanti et al. [88]
14Acetobacter malorumCleenwerck et al. [85]
15Acetobacter mustiFerrer et al. [94]
16Acetobacter nitrogenifiensDutta and Gachhui [95]
17Acetobacter oeniSilva et al. [96]
18Acetobacter okinawensisLino et al. [97]
19Acetobacter orientalisLisdiyanti et al. [86]
20Acetobacter orleanensisLisdiyanti et al. [88]
21Acetobacter oryzoeniBaek et al. [98]
22Acetobacter oryzifermentansKim et al. [99]
23Acetobacter papayaeLino et al. [97]
24Acetobacter pasteurianusSkerman et al. [77]
25Acetobacter persiciLino et al. [97]
26Acetobacter pomorumSokollek et al. [100]
27Acetobacter sacchariVu et al. [101]
28Acetobacter senegalensisNdoye et al. [102]
29Acetobacter siceraeLi et al. [103]
30Acetobacter suratthaniensisPitiwittayakul et al. [104]
31Acetobacter syzygiiLisdiyanti et al. [86]
32Acetobacter thailandicusPitiwittayakul et al. [105]
33Acetobacter tropicalisLisdiyanti et al. [88]
Acidomonas34Acidomonas methanolicaUrakami et al. [106]
Ameyamaea35Ameyamaea chiangmaiensisYukphan et al. [107]
Asaia36Asaia astilbisSuzuki et al. [108]
37Asaia bogorensisYamada et al. [109]
38Asaia krungthepensisYukphan et al. [110]
39Asaia lannensisMalimas et al. [111]
40Asaia platycodiSuzuki et al. [108]
41Asaia prunellaeSuzuki et al. [108]
42Asaia siamensisKatsura et al. [112]
43Asaia spathodeaeKommanee et al. [113]
Bombella44Bombella apisYun et al. [114]
45Bombella favorumHilgarth et al. [115]
46Bombella intestiniLi et al. [116]
47Bombella mellumHilgarth et al. [115]
Commensalibacter48Commensalibacter intestiniRoh et al. [117]
Endobacter49Endobacter medicaginisRamírez-Bahena et al. [118]
Gluconacetobacter50Gluconacetobacter aggerisNishijima et al. [119]
51Gluconacetobacter asukensisTazato et al. [120]
52Gluconacetobacter azotocaptansFuentes-Ramírez et al. [121]
53Gluconacetobacter diazotrophicusYamada et al. [122]
54Gluconacetobacter dulciumSombolestani et al. [123]
55Gluconacetobacter entaniiSchüller et al. [124]
56Gluconacetobacter johannaeFuentes-Ramírez et al. [121]
57Gluconacetobacter liquefaciensYamada et al. [122]
58Gluconacetobacter sacchariFranke et al. [125]
59Gluconacetobacter takamatsuzukensisNishijima et al. [119]
60Gluconacetobacter tumulicolaTazato et al. [120]
61Gluconacetobacter tumulisoliNishijima et al. [119]
Gluconobacter62Gluconobacter aidaeYukphan et al. [126]
63Gluconobacter albidusYukphan et al. [127]
64Gluconobacter cadivusSombolestani et al. [128]
65Gluconobacter cerevisiaeSpitaels et al. [129]
66Gluconobacter cerinusYamada and Akita [130]
67Gluconobacter frateuriiMason and Claus [131]
68Gluconobacter japonicusMalimas et al. [132]
69Gluconobacter kanchanaburiensisMalimas et al. [133]
70Gluconobacter kondoniiMalimas et al. [134]
71Gluconobacter morbiferRoh et al. [117]
72Gluconobacter oxydansDe Ley [135]
73Gluconobacter potusSombolestani et al. [128]
74Gluconobacter roseusMalimas et al. [136]
75Gluconobacter sphaericusMalimas et al. [137]
76Gluconobacter thailandicusTanasupawat et al. [138]
77Gluconobacter vitiansSombolestani et al. [126]
78Gluconobacter wancherniaeYukphan et al. [139]
Granulibacter79Granulibacter bethesdensisGreenberg et al. [140]
Komagataeibacter80Komagataeibacter diospyriNaloka et al. [141]
81Komagataeibacter europaeusYamada et al. [80]
82Komagataeibacter intermediusYamada et al. [80]
83Komagataeibacter kakiacetiYamada [142]
84Komagataeibacter kombuchaeYamada et al. [80]
85Komagataeibacter medellinensisYamada [142]
86Komagataeibacter melacetiMarič et al. [143]
87Komagataeibacter melomenususMarič et al. [143]
88Komagataeibacter nataicolaYamada et al. [80]
89Komagataeibacter oboediensYamada et al. [80]
90Komagataeibacter rhaeticusYamada et al. [80]
91Komagataeibacter saccharivoransYamada et al. [80]
92Komagataeibacter sucrofermentansYamada et al. [80]
93Komagataeibacter swingsiiYamada et al. [80]
94Komagataeibacter xylinusYamada et al. [80]
Kozakia95Kozakia baliensisLisdiyanti et al. [144]
Neoasaia96Neoasaia chiangmaiensisYukphan et al. [145]
Neokomagataea97Neokomagataea tanensisYukphan et al. [146]
98Neokomagataea thailandicaYukphan et al. [146]
Nguyenibacter99Nguyenibacter vanlangensisVu et al. [147]
Novacetimonas100
101
Novacetimonas cocois
Novacetimonas hansenii
Brandao et al. [148]
Brandao et al. [148]
102Novacetimonas maltacetiBrandao et al. [148]
103Novacetimonas pomacetiBrandao et al. [148]
Saccharibacter104Saccharibacter floricolaJojima et al. [149]
Swaminathania105Swaminathania saitoleransLoganathan and Nair [150]
Swingsia106Swingsia samuiensisMalimas et al. [151]
Tanticharoenia107Tanticharoenia aidaeVu et al. [152]
108Tanticharoenia sakaeratensisYukphan et al. [153]
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Román-Camacho, J.J.; García-García, I.; Santos-Dueñas, I.M.; García-Martínez, T.; Mauricio, J.C. Latest Trends in Industrial Vinegar Production and the Role of Acetic Acid Bacteria: Classification, Metabolism, and Applications—A Comprehensive Review. Foods 2023, 12, 3705. https://doi.org/10.3390/foods12193705

AMA Style

Román-Camacho JJ, García-García I, Santos-Dueñas IM, García-Martínez T, Mauricio JC. Latest Trends in Industrial Vinegar Production and the Role of Acetic Acid Bacteria: Classification, Metabolism, and Applications—A Comprehensive Review. Foods. 2023; 12(19):3705. https://doi.org/10.3390/foods12193705

Chicago/Turabian Style

Román-Camacho, Juan J., Isidoro García-García, Inés M. Santos-Dueñas, Teresa García-Martínez, and Juan C. Mauricio. 2023. "Latest Trends in Industrial Vinegar Production and the Role of Acetic Acid Bacteria: Classification, Metabolism, and Applications—A Comprehensive Review" Foods 12, no. 19: 3705. https://doi.org/10.3390/foods12193705

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

Article Metrics

Back to TopTop