UNIVERSITY OF NIGERIA, NSUKKA Donatus.pdfAccording to Oyeku et al. (2009), it consists mainly of...

97
1 Digitally Signed by: Content manager’s Name DN : CN = Weabmaster’s name O= University of Nigeria, Nsukka OU = Innovation Centre Nwamarah Uche UNIVERSITY OF NIGERIA, NSUKKA ALCOHOL DEHYDROGENASE Onah Donatus

Transcript of UNIVERSITY OF NIGERIA, NSUKKA Donatus.pdfAccording to Oyeku et al. (2009), it consists mainly of...

  • 1

    Digitally Signed by: Content manager’s Name

    DN : CN = Weabmaster’s name

    O= University of Nigeria, Nsukka

    OU = Innovation Centre

    Nwamarah Uche

    UNIVERSITY OF NIGERIA, NSUKKA

    ALCOHOL DEHYDROGENASE

    Onah Donatus

  • 2

    CHAPTER ONE

    INTRODUCTION AND LITERATURE REVIEW

    1.0. Introduction

    Palm wine is an important alcoholic beverage resulting from the spontaneous fermentation of the

    sap of the palm, which has been attributed to yeast and bacteria (Onwuka, 2011; Opara et al.,

    2013). It is the fermentad sap of certain varieties of palm trees including raphia palm (Raphia

    hookery or R. vinifera) (Ali, 2008). Fresh palm wine is sweat, clear, neutral, colourless juice

    containing minimal sugar (less than 0.5%) small amount of protein, gums and minerals (Opara et

    al., 2013). According to Oyeku et al. (2009), it consists mainly of water, sugar, vitamins and

    many aroma and flavour components in very small amounts. In traditional African societies, the

    palm wine play a significant role in customary practices, especially the distilled product from the

    palm wine, a potent gin called by various names in West Africa (Amoa-Awua et al., 2006).

    Over ten million people consume palm wine in West Africa (Onwuka, 2011)

    Traditionally, it is believed that when taken by nursing mothers; palm wine stimulates lactations

    and also has diuretic effect. Palm wine has also been used to enhance men’s potency due to yeast

    cell concentration. It could also be used for leavening of dough ad used in African medicine

    particularly in the treatment of measles and malaria (Onwuka, 2011).

    Despite all these good qualities of palm wine, it is a highly perishable sap due to fermentation

    which starts soon after the sap is collected and within an hour or two becomes reasonably high

    in alcohol (up to 4%). If palm wine is allowed to continue to ferment for more than 24hrs, it

    starts to turn into vinegar. This makes it unacceptable to consumers and creates losses to the food

    service industries. Fermentation in palm wine is possible because it constitutes a good growth

    medium for numerous microorganisms especially for yeast, lactic acid and acetic acid bacteria

    (Bechem et al., 2007). Saccharomyces cerevisae constitutes about 70% of the total yeast of palm

    wine and the activities of these microbes are believed to be responsible for conversion of sugar

    in palm wine to alcohol after a short time while bacteria induces the conversion of alcohol into

    vinegar (Onwuka, 2011).

  • 3

    Authorities who have studied the succession of micro flora in palm wine consistently reported

    the emergence of Acetobacter after about 24hrs of fermentation, at which time, alcohol was

    present in reasonable quantities (Opara et al., 2013). Earlier researches on the microbiology of

    palm wine had isolated Acetobacter from palm wine and these bacteria are believed to be

    responsible for souring of palm wine which is not acceptable by many.

    Acetic acid bacteria, Acetobacter and Gluconobacter, as well known as vinegar producers are

    able to oxidize ethanol to acetic acid by two sequential catalytic reactions of alcohol

    dehydrogenase and aldehyde dehydrogenase which are located on the periplasmic side of their

    cytoplasmic membrane (Abolhassan et al., 2007). Though these enzymes are important in

    industrial production of acetic acid, they are nevertheless spoilage molecules for many types of

    food and juices including palm wine (Ameh et al., 2011).

    Many attempts have been made to control palm wine spoilage at microbial level (Enwefa et al.,

    2004). Locally, the rural people use special leaves such as bitter leave to cover the wine

    container which they believe kills or disallows the influx of microorganism into the wine.

    Unfortunately, this method do not take care of the organisms already in the wine itself, hence

    deterioration continues (Onwuka, 2011). Also, the bark of a tree S. gabonensis was often added

    to the fresh palm wine. This impacts an amber colour and bitter taste to the wine. Although it

    delays souring of the wine and also lowers the titrable acidity (Ojimelukwe, 2002), the extract

    could not inhibit several yeast and bacteria present in the wine. With increasing availability of

    modern methods, efforts were directed towards the use of chemicals and pasteurization. Attempt

    to preserve palm wine using sulfite failed because at this pH, the concentration of sulfite required

    to suppress microfloral activities would be excessive for human consumption. Moreover, the use

    of chemical preservatives are discouraged due to the belief of cancer promotion. Convectional

    heating methods have been employed to delay spoilage, but the attractive flavor of palm wine is

    destroyed, giving room for arguments between wine drinkers and service men on the freshness of

    the beverage.

    Currently, palm wine is bottled on commercial scale with 37.5mg/l of metabisulphite and

    pasteurized at 65oC for 35mins, but the search for a more convenient, safer and effective method

    must continue.

  • 4

    1.1. Acetobacter

    Scientist have advocated the control of biological activities at molecular level because of its

    safety and the purity of the products. This draws our attention to alcohol dehydrogenase, one of

    the enzymes in Acetobacter responsible for deterioration of palm wine by converting alcohol, the

    most wanted component of palm wine, into acetic acid, with a view to investigating the

    possibility of controlling palm wine spoilage at the enzyme level. This entails that the enzyme is

    isolated, purified and characterized and that the effect of such parameter as pH, temperature and

    ethanol concentration on the activity of alcohol dehydrogenase are investigated.

    The knowledge of the effect of these parameters on the activity of alcohol dehydrogenase will be

    indispensible in regulating the activity of alcohol dehydrogenase. In this study, alcohol

    dehydrogenase was extracted from Acetobacter which was isolated from palm wine, partially

    purified, characterized and its thermal and pH stability investigated.

    Since their first discovery and reporting as a unique group, the acetic acid bacteria (bacteria that

    produce acetic acid) have been labeled with numerous genetic names, which have been the

    subject of extensive discussion and revision. The eighth edition of Berger’s Manual of

    Determinative Bacteriology (Buchanan and Gibbons, 1974) recognized only two genera,

    Acetobacter (motile by petrichous flagella or non-motile) and Gluconobacter (motile by polar

    flagella or nonmotile), and placed the genus Gluconobacter with the family Pseudomonadaceae;

    however, the genus Acetobacter was not assigned to any particular family and was grouped

    within the genera of uncertain affiliation. The Approved List of Bacterial Names, (Skerman et

    al., 1980) acknowledged both the genera Acetobacter and Gluconobacter. The ninth edition of

    Bergey’s Manual of Systematic Bacteriology (Buchanan and Gibbons, 1984) recognized the fact

    that the genera Gluconobacter and Acetobacter were closely related; hence they were placed

    within the family Acetobacteraceae. Members of the family are united by their unique ability to

    oxidize ethanol to acetic acid. Under this family we have genera Acetobacter, Gluconobacter and

    Frateuria. Today, acetic acid bacteria have been classified into 24 different genera. The major

    genera involved in vinegar production include: Acetobacter, Gluconobacter, Gluconacetobacter,

    Asaia, Neoasaia, Saccharibacter, Frateuria and Kozakia (De Vero and Giudici, 2008).

    The microorganisms present in wine-making processes are mainly yeasts, lactic acid

    bacteria and acetic acid bacteria, because of the extreme conditions in grape must (juices before

  • 5

    or during fermentation) such as the low pH (between 3 and 4) or high sugar concentration.

    Saccharomyces species (mainly Saccharomyces cerevisiae) are responsible for converting the

    sugars in grape must into ethanol and CO2 (Drysdale and Fleet, 1988).

    Lactic acid bacteria decrease the acidity of the wine and convert malic acid into lactic

    acid and CO2. This is a one-step reaction known as malolactic fermentation, which usually takes

    place once the alcoholic fermentation is over (Ribereu-Gayon et al., 2002).

    Acetic acid bacteria (AAB) play a negative role in the wine-making process because they

    alter the organoleptic characteristics of the wine and, in some cases, can also lead to stuck and

    sluggish fermentation. AAB modify wine, mainly because they produce acetic acid, acetaldehyde

    and ethyl acetate. They are also involved in other industrial processes of considerable interest for

    biotechnology such as the production of cellulose, sorbose and vinegar (Du Toit and Pretorius,

    2002).

    Acetic acid bacteria can be found in different stages of the wine-making process: for

    example, grape ripening, must, alcoholic fermentation, and bottled and stored wine. Although it

    has been known that wine can be altered by acetic acid bacteria ever since Pasteur, and they have

    a highly undesirable impact on the alcoholic fermentation processes, relatively little is

    understood about how they behave. Other microorganisms such as yeasts and lactic acid bacteria

    are also present during alcoholic fermentations and have been studied in much greater depth.

    1.2. General Characteristics of acetic acid bacteria

    Acetic acid bacteria (AAB) are gram negative, ellipsoidal (regular oval) to rod-shaped,

    and can occur singly, in pairs or in chains. They are motile due to the presence of flagella which

  • 6

    can be both peritrichous (having flagella uniformly distributed over the body surface, as certain

    bacteria) or polar (ie when the flagellum is located at one end of the cell). They do not form

    endospores as a defensive resistance. They have obligate aerobic metabolism, with oxygen as the

    terminal electron acceptor. The optimum pH for the growth of AAB is 5-6.5 (Holt et al., 1994).

    However, these bacteria can grow at lower pH values of between 3 to 4. They vary between 0.4-

    1µm long. They are catalase positive and oxidase negative. AAB can present pigmentation in

    solid cultures and can produce different kinds of polysaccharides (De Ley et al., 1984)

    AAB occur in sugar and alcoholised, slightly acid niches such as flowers, fruits, beer,

    wine, cider, vinegar, souring fruit juices and honey. On these substrates, they oxidize the sugars

    and alcohols, resulting to an accumulation of organic acids as final products. When the substrate

    is ethanol, acetic acid is produced, and this is where the name of the bacterial group comes from.

    However, these bacteria also oxidize glucose to gluconic acid, galactose to galactonic acid,

    arabinose to arabinoic acid. Some of these transformations carried out by AAB are considered

    interest for the biotechnological industry. The best known industrial application of AAB is

    vinegar production but they are also used to produce sorbose, from sorbitol, and cellulose.

    Fig 1. Electron microscope photography of Acetobacter (Gonzalex et al., 2004)

  • 7

    1.3. Physiological Role of Acetic Acid Bacteria

    One of the main characteristics of AAB is their ability to oxidize a wide variety of

    substrates and to accumulate the products of their metabolism in the media without toxicity for

    the bacteria. This ability is basically due to the dehydrogenase activity in the cell membrane.

    These dehydrogenases are closely related to the cytochrome chain (Matsushita et al., 1985).

    1.3.1. Ethanol Metabolism

    The oxidation of ethanol to acetic acid is the best known characteristics of acetic acid

    bacteria. Ethanol oxidation by AAB takes place in two steps. In the first one, ethanol is oxidized

    to acetaldehyde and in the second step acetaldehyde is oxidized into acetate. In both reactions,

    electrons are transferred and these are later accepted by oxygen.

    Two enzymes play a critical role in this oxidation process, both of which are bound to the

    cytoplasmic membrane: they are alcohol dehydrogenase and aldehyde dehydrogenase. Both

    enzymes have their active site on the outer surface of the cytoplasmic membrane (Adachi et al.,

    1978; Saeki et al., 1997).

    The bacteria can produce high concentration of acetic acid, up to 150g/l (Sievers et al.,

    1996; Lu et al., 1999), which makes them very important to the vinegar industry. Their

    resistance is strain dependent (Namba et al., 1984). The enzyme citrate synthase plays a key role

    in this resistance, because it detoxicates acetic acid by incorporation into the tricarboxylic or

    glyoxylate cycles, but only when ethanol is not present in the media. According to the report of

    Menzel and Gottschalk (1985), Acetobacter strain decrease their internal pH in response to a

  • 8

    lower external pH. However, an adaptation to high acetate concentration seems to be a

    prerequisite for high tolerance (Lasko et al., 2000).

    1.3.2. Primary and Polyalcohol Metabolism

    A considerable number of AAB can oxidize alcohols into sugars; mannitol into fructose;

    sorbitol into sorbose or eritritol into eritrulose. An important ability in oenology is to use

    glycerol as a carbon source (De Ley et al., 1984), which is converted into dihydroxyacetone, a

    small amount of which is used for energy synthesis.

    The enzymes that catalyse all these reactions are located in the cell membrane and induce a high

    accumulation of substrates in the media, which make AAB suitable microorganisms for the

    biotechnological industry (Deppenmeirer et al., 2002)

    1.3.3. Carbohydrate Metabolism

    AAB can metabolise different carbohydrates as carbon sources. Acetobacter species can

    use sugars through the hexose monophosphate pathway (Warburg-Dickens pathway) (De Ley et

    al., 1984; Drysdale and Fleet, 1988). And also through the EMbden-Meyerhof-Parnas and

    Entner-Doudoroff pathways (Attwood et al., 1991), although such authors as Drysdale and Fleet

    (1988) say that this last pathway is not used by AAB to metabolise glucose. From here they are

    further metabolized to CO2 and water via the tricarboxylic acid pathway, which is not functional

    in Gluconobacter species, although the complete oxidation is only functional when there is no

    carbon source in the media.

    Glycerol dihydroxyactetone

  • 9

    Sugar is more preferred as a carbon source by Gluconobacter than by Acetobacter because the

    species of this genus can obtain energy more efficiently by the metabolisation of the sugar via

    pentose phosphate pathway (De Ley et al., 1984).

    Glucose metabolism by these species produces a considerable number of industrially important

    metabolites (Olijve and Kok, 1979; Weenk et al., 1984; Qazi et al., 1991; Qazi et al., 1993;

    Velizarov and Beschkov, 1994). Some of these metabolites are 2-ketogluconic, 5-ketogluconic

    and 2,5-diketogluconic acids.

    The most characteristic reaction is the direct oxidation of glucose into Glucono-δ-lactone, which

    is oxidized into gluconic acid. This last reaction is particularly active in Gluconobacter species in

    media with high concentration of sugars such as grapes and must. This metabolite can be used as

    an indicator of the presence of these bacteria.

    Acetic acid bacteria can also use other carbohydrate, such as arabinose, fructose, galactose,

    mannitol, mannose, ribose, sorbitol and oxylose (De Ley et al., 1984) (Fig. 1)

    1.3.4. Organic Acid Metabolism

    AAB are able to metabolise a variety of organic acids. They do so through the tricarboxylic acid

    cycle which oxidizes these acids to CO2 and water. Gloconobacer, which lacks a functional

    tricarboxylic acid cycle, is unable to oxidize most organic acids (Holt et al., 1994). Acetic, citric,

    fumaric, lactic, malic, pyruvic, and succinic acids are completely oxidized to CO2 and water.

    These changes are very important in wine making, because they mean that the quality of the

    wines decrease.

    Another important by product of lactate metabolism is acetoin (important in the world of

    oenology (the scientific study of all aspect of wine and wine making) (De Ley, 1959). The

    buttery aroma of this compound is considered to be an unwanted flavor in wine, in which its

    detection limit is 150g/l (Romano and Suzzi, 1996; Du Toit and Pretorius, 2000).

    1.3.5. Nitrogen Metabolism

    Although some AAB species (Gluconacetobacter diazotrophicus) (Gillis et al., 1989) can fix

    atmospheric nitrogen, most of them use ammonium as a carbon source (De Ley et al., 1984). So

  • 10

    these bacteria can synthesize all the amino acids and nitrogenated compounds from ammonium.

    Depending on the amino acid in the media, their growth can either be stimulated or inhibited. So,

    glutamate, glutamine, proline, and histidine stimulate the growth of AAB, whereas valine for

    Gluconobacter oxidans and threonine and homoserine for Acetobacter aceti seem to have an

    inhibitory effect (Belly and Claus, 1972). However, no studies have been made on the nutritional

    needs of AAB nitrogen in wine. It has been observed that AAB selectively prefer some amino

    acids during vinegar production (Valero et al., 2003), and leave significant amount of

    ammonium in the media.

    1.4. History of Acetobacter

    The first taxonomist of AAB is the French scientist, Pasteur. Studying the Orleans method of

    vinegar production, he demonstrated that the acetic acid came from ethanol oxidation and that

    long-term oxidation of acetic acid converted it into CO2 and water. His results led him to

    formulate the involvement of the microorganisms in the process of transforming alcohol into

    vinegar, and confirmed the existence of Mycoderma aceti which Persoon had already described

    in 1882. Subsequently, in the year 1879 Hansen observed that the microbial flora which

    converted alcohol into acetic acid was not pure and consisted of various bacterial species. It was

    through the work of Beijerinck (1899) that the genera Acetobacter was proposed.

    1.5. Taxonomy of Acetobacter and other acetic acid bacteria

    The first classification was proposed by Hansen in 1894, based on the occurrence of a film in

    the liquid media, and its reaction with iodine. Asai (1934) formulated the proposal of classifying

    AAB into two genera: Acetobacter and Gluconobacter. The main differences between these two

    genera were both cytological (based on the cell bacterial cell structure, function and formation)

    and physiological (the scientific study of an organism’s vital functions, including growth and

    development, the absorption and processing of nutrient, the synthesis and distribution of proteins

    and organic molecules, and the functioning of different tissues, organs and other anatomic

    structures). The main physiological difference was that Acetobacter oxidized ethanol into acetic

    acid and, subsequently, completed the oxidation of acetic acid into water and CO2.

    Gluconobacter species, on the other hand, were unable to complete this oxidation of acetic acid.

    It was Frateur (1950) who formulated a classification based mainly on five physiological criteria:

  • 11

    1. Presence of catalase

    2. Gluconic acid production from glucose

    3. Oxidation of acetic acid into CO2 and water

    4. The oxidation of lactate into CO2 and water and

    5. The oxidation of glycerol into hydroxyacetone

    On the bases of these criteria, he divided Acetobacter genera into four groups

    1. Peroxydans

    2. Oxydans

    3. Mesoxydans and

    4. Suboxydans

    Those AAB that had peritrichous flagella and were able to completely oxidize ethanol into

    CO2 and water, are grouped into the genus Acetobacter while those that had polar flagella and

    unable to perform the complete oxidation are grouped into the genera Gluconobacter. The

    taxonomical keys for bacteria taxonomy have been historically collected in Bergey’s Manual of

    Systematic Bacteriology. In its last edition (De Ley et al., 1984), some molecular techniques

    were included as fatty acid composition, soluble protein electrophoresis, percentage of G + C

    content, and DNA-DNA hybridization. Gluconobacter and Acetobacter genera were included in

    the family of Acetobacteraceae. Acetobacter genus was composed by 4 species: A. aceti, A.

    pasteurianus, A. liquefaciens and A. hansenii. The Gluconobacter genus only consisted of G.

    oxydans.

    1.5.1. Taxonomy Based on Molecular Techniques

    Classification of AAB based initially on morphological and physiological criteria has been

    submitted to continuous variation and reorientations. These variations are due, basically, to the

    application of molecular techniques to the taxonomic study. DNA-DNA hybridization,

  • 12

    percentage base ratio determination, and 16S rDNA sequence analysis are the most common

    techniques used for this purpose.

    1.5.2. DNA-DNA Hybridization

    From taxonomical point of view, this is the most widely used for describing new

    species within bacterial groups. The technique measures the degree of similarity between the

    genomes of different species. When several species are compared in this way, the similarity

    values make it possible to arrange the species in a phylogenetic tree, which shows the degree of

    intraspecific and interspecific similarity.

    1.5.3. Percentage Base Ratio Determination

    This was one of the first molecular tools to be used in bacterial taxonomy. It calculates

    the percentage of G + C in a bacterial genome (G for guanine and C for cytosine. Guanine and

    adenine are nitrogenous bases in DNA). Although this percentage must be taken into

    consideration, by itself it cannot identify a given microorganism. In AAB, the % value of G + C

    vary between 55.5 and 64.5%.

    1.5.4. 16S rDNS Sequence Analysis

    The 16S rDNA gene is a highly preserved region with small changes that can be characteristic

    of different species. Ribosomal genes are compared in most taxonomical studies of bacteria.

    Acetobacteraceae family is no exception in this reorganization of species and genera. Six

    new AAB genera have been added to both the Acetobacter and Gluconobacter genera mentioned

    above. These include

    1. Acidomonas (Urakami et al., 1989)

    2. Gluconacetobacter (Yamada et al., 1997)

    3. Asaia (Yamada et al., 2000)

    4. Saccharibacter (Jojima et al., 2004)

    5. Swminathania (Loganathan and Nair, 2004) and

  • 13

    6. Kozakia (Lisdiyanti et al., 2002)

    At present, the Acetobacteriaceae family consists of 8 genera and 38 species. It has been

    proposed that the following species should be added to what was previously established by

    Bergey’s (De Ley et al., 1984). These are Acetobacter cerevisiae, A. malorum (Ceenwerck et al.,

    2002), A. tropicalis, A. orleaniensis, A. lovaniensis, and A. estuniensis, A. syzgii, A.

    cibinongenesis and A. oreintalis (Lisdiyanti et al., 2001), A. pomorum and A. oboediens

    (Sokollek et al., 1998), A. intermedians (Boesch et al., 1998), Kozakia baliensis (Lisdiyanti et

    al., 2002), Gluconobacter johannae and Ga. azotocaptuans (Fuentes-Ramirez et al., 2001), Ga

    swigsii and Ga. rhaeticus (Cleenwerck et al., 2005) and Ga. sacchari (Franke et al., 1999), Asaia

    krungthepensis (Yukuphan et al., 2004), As. siamensis (Katsura et al., 2001), As. bogorensis

    (Yamada et al., 2000), Saccaribacter floricola (Jojima et al., 2001), Swaminathania salitolerans

    (Loganathan and Nair, 2004). A. oboediens and A. intermedius were subsequently reclassified as

    Glucoacetobacter by Yamada (2000).

    These are summarized in table 1.

  • 14

    Table 1. Species of acetic acid bacteria

    Acetobacter Gluconacetobacter Gluconobacter

    A. cerevisiae,

    A. malorum

    A. tropicalis,

    A. orleaniensis,

    A. lovaniensis

    A. estuniensis,

    A. syzgii,

    A. cibinongenesis

    A. oreintalis

    A. pomorum

    A.aceti

    A. pasteurianus

    A.indonosiensis

    A. peroxydans

    Ga. johannae

    Ga. azotocaptans

    Ga swigsii

    Ga. rhaeticus

    Ga. Sacchari

    Ga. hansenii

    Ga. entanii

    Ga. xylinus

    Ga. liquefaciens

    Ga. diazotrophicus

    Ga. europaeus

    Ga. Oboediens

    Ga. intermedius

    G. oxydans

    G. frateurii

    G. assaii

    Asaia

    As. Bogorensis

    As. Siamensis

    As. Indonesiensis

    As. rugthepensis

    Swaminathania

    S. salitolerans

    Acidomonas Kozakia

    Ac. methanolica K. baliensis

    Saccharibacter

    Sa. floricola

    Gonzalex et al., 2004

    1.6. Isolation of Acetobacter

    These physiological difference among genera made it possible to develop differential

    culture media. Various media have been reported for isolating AAB whose carbon source is

  • 15

    glucose, mannitol, ethanol etc. some of these media can also incorporate CaCO3 or bromocresol-

    green as acid indicators (Swings and De Ley, 1981; De Ley et al., 1984; Drysdale and Fleet,

    1988). Culture media are usually supplemented with pimaricin in the agar plates to prevent the

    yeast from growing and with penicillin to eliminate lactic acid bacteria.

    Most of the widely used culture media are GYC (5% D-glucose, 1% yeast extract, 0.5%

    CaCO3 and 2% agar (w/v), described by Carr and Passmore (1979), and, YPM ( 2.5% mannitol,

    0.5% yeast extract, 0.3% peptone and 2% agar (w/v)). Plates must be incubated for between 2 to

    4 days at 28oC under aerobic conditions. These culture media are suitable for wine samples (Du

    Toit and Lamberchts, 2002; Bartowsky et al., 2003), and no problems have been detected

    culturing and isolating AAB from wine samples.

    Nevertheless, some works indicate the difficulty of culturing this bacterial groups from

    vinegar samples (Sokollek et al., 1998). This problem has been partially solved by introducing a

    double agar layer (0.5% agar in the lower layer and 1% agar in the upper layer (w/v) into the

    cultures and media containing ethanol and acetic acid in an attempt to stimulate the atmosphere

    of the acetification tanks, such as AE medium (Entani et al., 1985).

    1.7. Identification of acetic acid bacteria

    Identification of acetic acid bacteria is done using either classical method or molecular

    techniques.

    1.7.1. Classical Methods

    Classical microbiological taxonomy has traditionally used morphological and

    physiological differences among the species to discriminate between them. The tests could

    only discriminate at the species level, although the physiological methods would not be able

    to distinguish the currently described species. At the genus level, several characteristics can

    contribute to the differentiation. The Gluconobacter genus cannot completely oxidize acetic

    acid into CO2 and water. The main characteristic of Acidomonas is that it can grow in

    methanol, and Asaia is characterized by its inability to grow in a media with an acetic acid

    concentration higher than 0.35%. The other two genera, Gluconacetobacter and Acetobacter,

    can be differentiated on the bases of their ubiquinone content. Ubiquinone Q9 is present in

  • 16

    Acetobacter, and ubiquinone Q10 in Gluconacetobacter (Trcek and Teuber, 2002). Kozakia

    have low similarity values of % G + C content among the other genera (7 – 25% lower than

    the other species), the major ubiquinone is Q10 and have a weak activity in oxidation of

    lactate and acetate into carbon dioxide and water. The genus Saccharibacter has a negligible

    or very weak productivity of acetic acid from ethanol and the osmophilic growth properties

    (its adaptation to environment with high osmotic pressure, such as high sugar concentration)

    distinguished this genus from other AAB. Swaminathania genus is able to fix nitrogen and

    solubilized phosphate in the presence of NaCl. Some of the phenotypic characteristics of the

    former species described in Bergey’s Manual are shown in Table 2.

    Table 2. Phenotypic characteristics of the species belonging to the Acetobacter and

    Gluconobacter

    Characteristics A. aceti A. liquefaciens A. pasteurianus A. hansenii G. oxydans

    Ethanol overoxidation + + + + _

    Growth in:

    ethanol + + V _ _

    Sodium acetate + V V _ _

    Dulcitol _ _ _ _ _

    Glycerol Cetogenesis + + _ V V

    Lactate oxidation + + + + _

    Pigment production _ + _ _ +

    Source: Gonzalex et al., 2004

    1.7.2. Molecular Techniques

    There are many molecular techniques of identifying AAB both on species level and on

    strain level. One of them is PCR-RFLP of the rDNA 16S method. This technique was used by

    Ruiz et al. (2000) to identify AAB and is appropriate for differentiating and characterizing

  • 17

    microorganisms on the basis of their phylogenetic relationships (phylogenetic analysis exploits

    the changes in DNA sequence that arise through mutations during evolution to reconstruct the

    evolutionary history of different groups of organisms) (Carlotti and Funke, 1994). In eubacterial

    DNA, the RNA loci include 16S, 23S and 5S rRNA gene, which are separated by internally

    transcribed spacer (ITS) regions. The techniques consist on the amplification of the 16S rDNA

    regions followed by the digestion of the amplified fragment with a restriction enzyme. The DNA

    fragments obtained are separated by electrophoresis. The resulting patterns are characteristic of

    every species and make it possible to characterize almost all the AAB species.

    One of the techniques used to identify AAB on strain level is Random amplified

    polymorphic DNA-PCR (RAPD-PC). RAPD fingerprint based on the amplification of the

    genomic DNA with a single primer of arbitrary sequence, of 9 or 10 bases of length, which

    hybridise with sufficient affinity to chromosomal DNA sequences at low annealing temperatures

    so that they can be used to initiate the amplification of bacterial genome regions. The

    amplification is followed by agarose gel electrophoresis, which yields a band pattern that should

    be characteristic of the particular bacterial strain (Caetano-Anolles et al., 1991; Meunier and

    Grimont, 1993). The technique has already been used to characterize rice vinegar AAB. They

    managed to discriminate among AAB strains and the patterns yielded between 7 and 8 DNA

    fragments).

    1.8. Ecology of Acetobacter

    Ecology is the science of the study of the relationship between living organism and its

    environment. AAB can grow in different environment and the components of the environment

    affect its growth and activities

    1.8.1. Acetobacter in palm wine

    The sap of the oil palm tree (Elaeis guinneesis) serves as a rich substrate for various types

    of micro-organisms to grow. However, it is as a source for producing palm wine that the

    substrate is mainly known for (Amoa-Awua et al., 2006). In various African countries and

    beyond, the sap of the palm tree is tapped and allowed to undergo spontaneous fermentation,

    which allows the proliferation of yeasts species to convert the sweet substrate into an alcoholic

    beverage. Fresh palm wine is sweat, clear, neutral, colourless juice containing minimal sugar

  • 18

    (less than 0.5%) small amount of protein, gums and minerals (Opara et al., 2013). According to

    Oyeku et al. (2009), it consists mainly of water, sugar, vitamins and many aroma and flavour

    components in very small amounts.

    In various traditional African societies, the palm wine plays a significant role in

    customary practices, especially the distilled product from the palm wine, a potent gin called by

    various names in West Africa.

    Traditionally, it is believed that when drank by nursing mothers; palm wine stimulates

    lactations and also has diuretic effect. Palm wine has also been used to enhance men’s potency

    due to yeast cell concentration. It could also be used for leavening of dough and used in African

    medicine particularly in the treatment of measles and malaria (Onwuka, 2011).

    Despite all these good qualities of palm wine, it is a highly perishable sap due to

    fermentation which starts soon after the sap is collected and within an hour or two becomes

    reasonably high in alcohol (up to 4%). If palm wine is allowed to continue to ferment for more

    than 24 hours, it starts to turn into vinegar. This makes it unacceptable to consumers and creates

    losses to the food service industries. Fermentation in palm wine is possible because it constitutes

    a good growth medium for numerous microorganisms especially for yeast, lactic acid and acetic

    acid bacteria (Bechem et al., 2007). According to the research of Amoa-Awua et al. (2006),

    concurrent alcoholic, lactic acid and acetic acid fermentation occurs during the tapping of palm

    wine from oil palm trees. Yeast growth dominated by S. cerevisiae starts immediately after

    tapping begins and alcohol concentrations become substantial in the product after the third day.

    Lactic acid bacteria dominated by L. plantarum and L. mesenteriodes are responsible for a rapid

    acidification of the product during the first 24 h of tapping whilst the growth of acetic acid

    bacteria involving both Acetobacter and Gluconobacter species become pronounced after the

    buildup in alcohol concentrations on the third day. Increases in the alcohol level of palm wine

    are faster in the container into which the palm wine accumulates during the tapping than in the

    receptacle cut out in the tree trunk, and samples which accumulate overnight have alcohol

    contents of over 3%. During the holding/marketing of palm wine, the concentration of alcohol

    increases from 3% to over 7% in 24 h, remains high for the next 3 days and begins to drop. The

    concentration of acetic acid increases slowly from a concentration of about 0.42–0.48% and after

    4 days had exceeded the acceptable level of 0.6% (Amoa-Awua et al., 2006). Saccharomyces

    cerevisae constitutes about 70% of the total yeast of palm wine and the activities of these

  • 19

    microbes are believed to be responsible for conversion of sugar in palm wine to alcohol after a

    short time while bacteria induces the conversion of alcohol into vinegar (Onwuka, 2011).

    Because of the central role that the alcoholic beverage has played in the traditional society, it is

    important that the microbiology and biochemistry of the fermentation process are well

    understood.

    Authorities who have studied the succession of micro flora in palm wine consistently reported

    the emergence of Acetobacter after about 24 hours of fermentation, at which time, alcohol was

    present in reasonable quantities (Opara et al., 2013). Earlier researches on the microbiology of

    palm wine had isolated Acetobacter from palm wine and these bacteria are believed to be

    responsible for souring of palm wine which is not acceptable by many.

    Acetic acid bacteria, Acetobacter and Gluconobacter, as well known as vinegar producers are

    able to oxidize ethanol to acetic acid by two sequential catalytic reactions of alcohol

    dehydrogenase and aldehyde dehydrogenase which are located on the periplasmic side of their

    cytoplasmic membrane (Abolhassan et al., 2007). Though these enzymes are important in

    industrial production of acetic acid, they are nevertheless spoilage molecules for many types of

    food and juices including palm wine (Ameh et al., 2011).

    Previous studies on the microbiology of oil palm tree (E. guineensis) and R. hookeri have

    incriminated several bacterial and yeast flora to be involved in the fermentation process (Okafor,

    1975). Acetobacter species were earlier isolated from oil palm wine (Faparusi, 1973; Okafar,

    1975).

    1.8.2. Acetobacter in other wines

    Alcohol fermentations are carried out by yeast (mainly Saccharomyces cervisiae), which

    are responsible for transforming the sugars present in the musts (glucose and fructose) into

    ethanol. The second group of microorganisms involved in the wine production is lactic acid

    bacteria. These bacteria are responsible for the malolactic fermentation, the process by which the

    malic acid is transformed into lactic acid, thus deacidifying and softening the wine. The third

    group of wine microorganisms are the acetic acid bacteria. Unlike the other microorganisms

    involved in fermentation processes, they have received very little attention, and little is known

    about their behavior and dynamics in wine making processes or their contribution to the spoilage

    of must and wines. According to Margalith (1981), acetic acid in wine becomes objectionable at

  • 20

    concentration exceeding 0.7–1.2g/l. Acetic acid is the main volatile acid in wines and its

    presence is frequently described as volatile acidity (Margalith 1981). An excess of acetic acid in

    wine is the main problems found nowadays in wineries. Another consequences of high volatile

    acidity in wines is the presence of ethyl acetate, which also gives the wines a vinegary taint

    (traces of undesirable quality) and makes the wine smell like glue.

    The wine making process begins in the vineyard. The grapes acquire and harbor the right sugar

    and physiological composition of their juice so that, once they have been crushed, it can be

    transformed into wine by yeast. The growth of AAB has been reported during various steps of

    the wine-making process, including some conditions in which they would not be expected to

    grow.

    1.8.3. Acetobacter in Grapes and Musts

    As the grapes become mature, the amount of sugars (glucose and fructose) increases.

    Those sugars are an optimum growing media for AAB, and in particular for G. oxydans, because

    this species clearly prefer ethanol as the carbon source. In these conditions the predominant

    species in grapes is usually G. oxydans, and the most common populations are around 102-

    105cfu/g (Joyeux et al., 1984a; Du Toit and Lambercht, 2002) (cfu stands for colony forming

    unit, a measure of the number of viable cells capable of producing new colonies when seeded,

    that are contained in a culture). Because of G. oxydans’ low tolerance of ethanol, it disappears in

    the first stages of alcoholic fermentation. Acetobacter and Gluconacetobacter species have also

    been isolated from unspoiled grapes, although in very low amounts (Du Toit and Lamberchts,

    2002).

    Damaged, rotten or Botrytis-infected grapes can be infected by yeasts and acetic acid

    bacteria. Yeasts can start metabolizing the sugars in grapes into ethanol, which are then oxidized

    into acetic acid by acetic acid bacteria. Damaged grapes contain AAB population, mainly

    belonging to Acetobacter species (A. aceti and A. pasteurianus) up to 106cfu/g (Joyeux et al.,

    1984b; Grossman and Becker, 1984). These grapes contain high concentrations of acetic acid,

    ethanol and glycerol, and small amounts of ethyl acetate (Sponholz and Dietrich, 1985; Drysdale

    and Fleet, 1989b). Both ethanol and glycerol are the products of yeast metabolism. The glycerol

    produced can be metabolized by AAB into dihydroxyacetone, which affects the sensory quality

  • 21

    of the wine and can bind to SO2 thus decreasing its antimicrobial properties. Gluconic acid arises

    from the metabolism of glucose by AAB (Drysdale and Fleet, 1988), and can be further oxidized

    to produce 5-keto and 2-ketogluconic acid.

    Thus, grape juice composition can be significantly altered if the berries are infected with

    acetic acid bacteria. The changes not only have an adverse effect on the sensory quality of wine

    but also on the growth of yeasts during alcoholic fermentation (Drysdale and Fleet, 1989a) and

    the possible growth of lactic acid bacteria (Joyeux et al., 1984b)

    Adding SO2 to the musts is common practice in cellars (a wine cellar is a storage room

    for wine in bottles or barrels, or more rarely in carboys, amphorae or plastic containers, because

    it inhibits the microorganism and hinders the development of undesirable organisms such as

    AAB. So the presence and growth of AAB in must will depend on the concentration of SO2

    whether it is present in the free or the bonded form. The free form consists of molecular sulphur

    dioxide, bisulphate ands sulphite ions. Only the molecular SO2 has anti-microbial effects. The

    proportion of molecular SO2 represents from 1% to 10% of the free form depending on the pH of

    the wine, therefore, the lower the pH is, the higher proportion of molecular SO2 will exist, and

    the higher anti-bacterial effect (Ribereau-Gayon et al., 2000). In this process the must may also

    be contaminated by AAB resident in the cellar because of such processes as grapes juice racking

    and pumping.

    1.8.4. Acetobacter during Fermentation

    During alcoholic fermentation both Saccharomyces and non-Saccharomyces yeasts

    develop enormously and can reach populations up to 107-108 cfu/ml. During this process, sugars

    from must are transformed into ethanol by yeasts, which make this new media more suitable for

    Acetobacter and Gluconacetobacter species. In this process, a considerable amount of CO2 is

    produced because of the yeast metabolism, and this creates anaerobic conditions that are

    theoretically unsuitable for AAB growth. Recent studies by Du Toit et al. (2005), however,

    suggest that some AAB strains can survive for a long period under relatively anaerobic

    conditions in wine. The pH is usually around 3.5, and the optimum pH for AAB development is

    5.5-6.3 (Holt et al., 1994), although some AAB have been isolated at pH 3.0. The pH is also

    important for the state in which we can find SO2 in wine. Low concentration of SO2 does not

  • 22

    affect the culturability of some AAB strains, and sulphur dioxide does not completely eliminate

    the presence of AAB (Du Toit et al., 2005). AAB are able to grow in wines containing 20mg/l of

    free SO2 (Joyeux et al., 1984a), which means that the common levels of SO2 in wines are not

    enough to inhibit AAB growth. Watanabe and Iino (1984) found that 100mg/l of total SO2 were

    needed to inhibit the growth of Acetobacter species in grape must.

    The temperature at which alcoholic fermentation takes place depends on the type of

    vinification. Red wine fermentations take place between 25 and 38oC, which is the same as the

    optimum temperature for AAB growth (Holt et al., 1994), and therefore does not seem to prevent

    AAB development. The temperature of white and rose fermentations ranges from 18-19oC and

    the effect of low-temperature fermentations on the AAB population has not been studied yet.

    Growth of these bacteria during alcoholilc fermentation may also be linked to the number of

    bacteria and yeast in the must at the start of the fermentation (Watanabe and Iino, 1984). The

    predominant species during alcoholic fermentation are commonly A. aceti, A. pasteurianus, Ga.

    Liquefaciens and Ga. Hansenii (Joyeux et al., 1984b; Du Toit and Lamberchts, 2002), although

    G. oxydans have also been isolated as the only species during the fermentation.

    In spite of these adverse conditions during alcoholic fermentation, some authors (Du Toit

    et al., 2005) have detected that AAB can survive and even grow during this process. If the

    quality of the wines is to be good, it is of vital importance to keep the numbers of AAB low. This

    can be done by using healthy grapes, inoculating a high quality of yeast, adding SO2, clarifying

    the must and lowering the pH by adding acid (Du Toit and Pretorius, 2002).

    If AAB grow a lot in the first stages of alcoholic fermentation, fermentation may become stuck

    or sluggish and there may be renewed growth of AAB and the reduction in the quality of the

    wines during their storage (Joyeux et al., 1984b)

    1.8.5. Acetobacter during ageing and wine maturation

    During storage, the major species found belong to Acetobacter (A. aceti and A.

    pasteurianus). These bacteria have been isolated from the top, middle and bottom of the tanks

    and barrels, suggesting that AAB can actually survive under the semi-anaerobic conditions

    occurring in wine containers. This can be explained by the ability of AAB to use compounds,

  • 23

    such as quinines and educible dyes, as electron acceptors (Du Toit and Pretorius, 2002). The

    main product obtained from the presence of AAB at this point is acetic acid, although

    considerable amounts of acetaldehyde and ethyl acetate are produced (Drysdale and Fleet,

    1989b) and glycerol metabolises to dihydroxyacetone. The pumping over and racking of wine

    may stimulate the growth of AAB and lead to populations up to 108cel/ml (Joyeux et al., 1984b;

    Drysdale and Fleet, 1989b), because of the intake of oxygen during these operations. The

    number of bacteria usually decreases drastically after bottling, because of the relatively

    anaerobic conditions in a bottle. However, the excessive addition of oxygen during bottling can

    increase the number of AAB. The ethanol concentration of wine is around 10-15% (v/v). As

    mentioned above, ethanol is a good carbon source for AAB, but it can also inhibit AAB growth

    at high concentrations. However, it is well known that these bacteria can grow in wine

    containing between 10-14% (v/v) (Joyeux et al., 1984a; Drysdale and Fleet, 1989a; Koselbalan

    and Ozlingen, 1992; Du Toit and Pretorius, 2002). It has been reported by Saeki et al. (1997) that

    AAB can overcome the inhibitory effect, and become tolerant to ethanol. In this respect AAB

    have been isolated from sake and tequila (beverages with a much higher ethanol concentration

    than wine) (Joyeux et al., 1984a), although Drysdale and Fleet (1989b) observed a weak growth

    of AAB even at 10oC.

    1.8.6. Acetobacter in vinegar production

    Vinegar is a precious food additive and complement as well as effective preservative against

    food spoilage that is produced by Acetic acid bacteria and contains essential nutrients such as

    amino acids regarding its fruit source (Kocher et al., 2006). Food and Drug

    Administration(FDA), USA has explained the vinegar as a 4% acetic acid solution that is

    synthesized from sweet or sugary substances through alcoholic fermentation. The neoclassical

    fermentation resulted in several vinegar types with different tastes, frangrances and nutritional

    values because of applying various acetic acid bacteria in vinegar making procedure. Currently

    the vinegar manufacturers are seeking for new types of vinegar using different AAB as their

    starter and tradional vinegar production has been improved using various natural substrates and

    fruits (Du Toit and Lambrechts, 2002). Acetobacter strains are the major bacteria that are dealing

    with vinegar production industrially (Sokollek et al., 1998; Kaeere et al., 2008).

  • 24

    Vinegar has been very important in the human diet since ancient times as a condiment

    and food preservative; for many centuries, acetic acid from vinegar was the strongest acid, until

    sulphuric acid was discovered around the year 1300. Although little is known about the role

    played by microorganisms in vinegar production, vinegar has been produced mainly from wine,

    alcohol and rice. Nowadays knowledge is much more advanced, above all as far as the analytical

    and industrial processes are concerned, but the microbiology of the process is still not well

    understood. At the beginning of the 21st century, the species and strains responsible for vinegar

    production are still not very clear. Nowadays, there are three different biotechnological processes

    for producing vinegar (Greenshields, 1978): the Orleans method (this is the most famous slow

    method of vinegar production. Here, barrels are filled with wine and vinegar and fermentation

    are carried out slowly by AAB, which will generally metabolize all the alcohol in 1 to 3 months),

    the German method (a very quick method also called generator method. In this method, the

    alcoholic solution to be acetified is allowed to trickle down through a tall tank or column

    (generator) packed with porous solid material on whose surface Acetobacter bacteria are

    permitted to grow) and the submerged method (a catalysed fast method involving acetator, a tank

    equipped with a variety of system that keep the mixture constantly turning, introducing air into

    the mixture to introduce oxygen to keep the bacteria working).

    1.9. Alcohol dehydrogenase

    Alcohol dehydrogenase (EC.1.1.5.5) (Gomez-Manzo et al., 2008) otherwise called

    pyrolloquinoline quinone alcohol dehydrogenases or Alcohol dehydrogenase or type III Alcohol

    dehydrogenase or membrane associated quinohaemoprotein alcohol dehydrogenase, an enzyme

    with system name alcohol:quinone oxidoreductase, belongs to quinoenzymes and requires

    quinoid cofactors (e.g., pyrroloquinoline quinone, PQQ) as enzyme-bound electron acceptors.

    They are distinct from other types of alcohol dehydrogenases because of their position in the

    cells. While majority of other alcohol dehydrogenases are located in the cytosol, otherwise called

    cytosolic NAD+/NAD(P)+-dependent alcohol dehydrogenase located in the cytoplasm,

    these family of alcohol dehydrogenase are membrane-bound. Many membrane-bound

    dehydrogenases in the periplasmic space or on the outer surface of the cytoplasmic

    membrane of acetic acid bacteria and other aerobic Gram-negative bacteria have been

    classified as PQQ- or FAD-dependent dehydrogenases (Matsushita et al., 1994). Most of the

  • 25

    enzymes are closely associated with oxidative fermentation in industry, catalyzing an

    incomplete one-step oxidation, allowing accumulation of an equivalent amount of

    corresponding oxidation products outside the cells. The active sites of individual enzymes

    face the periplasmic space (Fig 1). Apart from alcohol dehydrogenases, there are other

    membrane-bound dehydrogenases such as glucose dehydrogenase and fructose

    dehydrogenase. All the enzyme reactions are carried out by periplasmic oxidase systems

    including alcohol- and sugar-oxidizing enzymes of the organisms. D-Glucose, ethanol,

    and many other substrates are oxidized by the dehydrogenases (shown as PQQ or FAD,

    except for aldehyde dehydrogenase) that are tightly bound to the outer surface of the

    cytoplasmic membranes of the organism. These membrane-bound enzymes irreversibly

    catalyze incomplete one-step oxidation and the corresponding oxidation products

    accumulate rapidly in the culture medium or reaction mixture. The electrons (e-)

    generated by the action of these dehydrogenases are transferred to ubiquinone in the

    membrane. The reducing equivalents are further transferred to the terminal ubiquinol

    oxidase in the cytoplasmic membranes. The terminal oxidase generates an

    electrochemical proton gradient either by charge separation or by a proton pump or by

    both during substrate oxidation by the membrane-bound enzymes, allowing the

    organism to acquire bioenergy through substrate oxidation. Thus, the organisms generate

    bioenergy through the enzyme activities of PQQ- and FAD-dependent dehydrogenases.

    Many different NAD- and FAD- dependent dehydrogenases in the cytoplasm have no

    function in oxidative fermentation and thus are not shown in Fig. 2.

  • 26

    Fig.2. Membrane-bound PQQ- and FAD-dependent primary dehydrogenases on the outer

    surface of acetic acid bacteria. (Adachi et al., 2007).

    1.9.1. Classes of alcohol dehydrogenases (EC.1.1.5.5)

    There are different classes of alcohol dehydrogenase or Pyroloquinoline quinone (PQQ)-

    dependent alcohol dehydrogenases. Among the most comprehensively studied of these enzymes

    are the three classes of PQQ-containing quinoprotein alcohol dehydrogenases; Type I are

    soluble, periplasmic enzymes containing a single Pyroloquinoline Quinone prosthetic group; this

    group includes the methanol dehydrogenase of methylotrophs. Type II dehydrogenases are

    soluble, periplasmic quinohemoproteins, having a C-terminal extension containing heme C. Type

    III dehydrogenases have similar quinohemoprotein subunits but have two additional subunits

    (one of which is a multiheme cytochrome c), bound in an unusual way to the periplasmic

    membrane (Anthony, 2004). These membrane enzymes and other quionoprotein dehydrogenases,

    their prothetic group, electron acceptors, location and organisms in which they are found are

    summarized in the Table 3 below.

  • 27

    Table 3. Summary of quinoprotein and quinohemoprotein dehydrogenase

    Enzyme Location Prosthetic

    group

    Electron

    acceptor

    Organism

    Type 1 alcohol

    dehydrogenase: eg

    -methanol

    dehydrogenase

    -ethanol dehydrogenase

    periplasm

    periplasm

    PQQ

    PQQ

    Cytochrome c

    Cytochrome c

    Methylotroph

    Pseudomononas sp

    Type 11 alcohol

    dehydrogenases

    Membrane PQQ

    Heme c

    Azurin Comamonas

    testosterone

    Pseudomonas putida

    Type 111 alcohol

    dehydrogenases

    Membrane PQQ

    4 heme c

    UQ Acetic acid bacteria

    Sorbitol dehydrogenase Membrane PQQ

    4 heme c.

    UQ Acetic acid bacteria

    Membrane glucose

    dehydrogalnse (m-GDH)

    Membrane PQQ UQ Enteric bacteria

    Acetic acid bacteria

    Acinetobacter

    calcoaceticus

    Soluble glucose

    Dehydrogenase (s-GDH)

    Periplasm PQQ ? Acenetobacter

    calcoaceticus

    Glycerol dehydrogenase Membrane PQQ UQ Acetic acid bacteria

    D-arabitol

    dehydrogenase

    Membrane PQQ UQ Acetic acid bacteria

    D-sorbitol

    dehydrogenase

    Membrane PQQ UQ Acetic acid bacteria

    Lupanine hydroxylase Periplasm PQQ heme c Cytochrome c Pseudomonas sp

    Sorbose/sorbosone

    dehydrogenase

    Periplasm PQQ Cytochrome c Acetic acid bacteria

    Methylamine Periplasm TTQ Amicyanin Methylotrophs

  • 28

    dehydrogenase

    Aromatic amine

    dehydrogenase

    Periplasm TTQ Azurin Alcaligenes

    Amine dehydrogenase Periplasm CTQ 2 heme Azurin Pseudomonas putida

    Paracoccus

    denitrificans

    UQ = ubiquinone, PQQ = pyroloquinoline quinone, TTQ = tryptophan tryptophyl quinone, CTQ

    = Cysteine Tryptophylquinone

    Source: (Matsushita et al., 2002; Davidson, 1993; Anthony, 1996; Goodwin and Anthony, 1996;

    Davidson, 2000; Choi et al., 1995; Hyun and Davidson, 1995; ; Anthony, 2000; Adachi et al.,

    1998; Hopper and Rogozinski, 1998; Asakura and Hoshino, 1999; Cozier et al., 1999; Takagi et

    al., 1999; Yoshida et al., 1999; Afolabi et al., 2001; Elias et al., 2000, 2001; Keitel et al., 2000;

    Adachi et al., 2001; Datta et al., 2001; Sugisawa and Hoshino, 2001; Chen et al., 2002; Miyazaki

    et al., 2000; Oubrie et al., 2002; Satoh et al., 2002)

    1.9.1.1. The Type I Alcohol Dehydrogenase

    Methanol dehydrogenase (MDH) belongs to type 1 alcohol dehydrogenase. The MDH of

    methylotrophic bacteria oxidizes methanol to formaldehyde during growth of bacteria on

    methane or methanol (Anthony, 1982), during which its electron acceptor is a novel acidic

    cytochrome (cytochrome cL) (Anthony, 1992). MDH is also responsible for oxidation of ethanol

    to acetaldehyde during growth on ethanol. Using phenazine ethosulphate in a dye-linked assay

    system the pH optimum is about 9 and ammonia or methylamine is required as activator. MDH

    oxidizes a wide range of primary alcohols (very rarely secondary alcohols), having a high

    affinity for these substrates; for example, the Km for methanol is 5–20 M. The pH optimum for

    the reaction with cytochrome cL is 7.0, and ammonia is not usually required as activator.

    The X-ray structure has been determined for the MDH from Methylobacterium

    extorquens (Blake et al., 1994; Ghosh and Anthony, 1995; Afolabi et al., 2001), and from

    Methylophilus sp. (Xia et al., 1992; White et al., 1993; Xia et al., 1996; Xia et al., 1999; Zheng

    et al., 2001). MDH has an α2β2 tetrameric structure; each α subunit (66 kDa) contains one

    molecule of PQQ and one Ca2+ ion. The β subunit is very small (8.5 kDa), it cannot be reversibly

  • 29

    dissociated, its function is unknown and it is not present in any other quinoproteins. The large α

    subunit has a propeller fold making up a superbarrel (Fig. 3)

    Fig.3. Propeller structure of type 1 alcohol dehydrogenase (methanol dehydrogenase)

    Source: Gosh et al., 1995

    The αβ unit of MDH looking down the pseudo 8-fold axis, simplified to show only the β-strands

    of the ‘W’ motifs of the α-chain, and the long α-helix of the β-chain, but excluding other limited

    β-structures and short α-helices (Ghosh et al., 1995). The PQQ prosthetic group is in skeletal

    form and the calcium ion is shown as a small sphere. The outer strand of each ‘W’ motif is the D

    strand, the inner strand being the A strand. The ‘W’ motifs are arranged in this view in an anti-

    clockwise manner. The exceptional motif W8 is made up of strands A-C near the C-terminus,

    plus its D strand from near the N-terminus.

    The structure has several important novel features, including novel the ‘tryptophan-docking

    motifs’ that link together the eight beta sheets, and the presence in the active site of an unusual

    eight-membered disulphide ring structure formed from adjacent cysteine residues, joined by an

  • 30

    atypical non-planar peptide bond. The PQQ is sandwiched between the indole ring of Trp243

    and the disulphide ring structure (Fig. 4).

    Fig.4 The novel disulphide ring in the active site of type 1 alcohol dehydrogenase (methanol

    dehydrogenase)

    Source: Ghosh et al. 1995.

    The ring is formed by disulphide bond formation between adjacent cysteine residues. The

    PQQ is ‘sandwiched’ between this ring and the tryptophan that forms the floor of the active site

    chamber. The calcium ion is coordinated between the C-9 carboxylate, the N-6 of the PQQ ring

    and the carbonyl oxygen at C-5. This structure is seen in all the alcohol dehydrogenases but not

    in aldose dehydrogenases. The indole ring is within 15o of co-planarity with the PQQ ring and,

    on the opposite side, the two sulfur atoms of the disulphide bridge are within 3.75Å of the plane

    of PQQ. The rarity of the disulphide ring structure would suggest some special biological

    function. Reduction of the disulphide bond leads to loss of activity but oxidation in air or

    carboxymethylation of the free thiols leads to return of activity. The activity of the

    carboxymethylated derivative rules out reduction to the thiols during the catalytic cycle. The

    disulphide ring is not present in the quinoprotein glucose dehydrogenase in which electrons are

    transferred to membrane ubiquinone from the quinol PQQH2, and in which the semiquinone free

    radical is unlikely to be involved as a stable intermediate. It is possible, therefore, that this novel

    structure might function in the stabilization of the free radical PQQ semiquinone or its protection

    from solvent at the entrance to the active site in MDH (Blake et al., 1994; Avezoux et al., 1995).

  • 31

    Recent work with the quinohemoprotein (Type II) alcohol dehydrogenase suggests, however,

    that, although it does not become completely reduced, the disulphide ring is essential for intra-

    protein electron transfer in all the alcohol dehydrogenases (Oubrie et al., 2002). In addition to the

    axial interactions, many amino acid residues are involved in equatorial interactions with the

    substituent groups of the PQQ ring system (Fig. 5).

    .

    This Figure also shows Asp303, which is likely to act as the catalytic base, and Arg331

    which may also be involved in the mechanism. The equatorial interactions of the

    quinohemoprotein alcohol dehydrogenase (QH-ADH) are almost identical to these, an important

    exception being that Arg331 is replaced by a lysine (Chen et al., 2002; Oubrie et al., 2002), as is

    also the case in glucose dehydrogenase. These are exclusively hydrogen-bond and ion-pair

    interactions. Although the number of polar groups involved might indicate at first sight that the

    environment of the PQQ is polar, this is not the case. Oxygen of the 9-carboxyl forms a salt

    bridge with Arg109 and both groups are shielded from bulk solvent by the disulphide. The

    carboxyl group of Glu155 and a 2-carboxyl oxygen of PQQ are also shielded from solvent and it

    is probable that at least one is protonated, their interaction thus being stabilized through

    Fig. 5. The equatorial interactions of PQQ and the coordination of Ca2+ in the active site of type 1

    alcohol dehydrogenase (methanol dehydrogenase)

    Source: Ghosh et al., 1995

  • 32

    hydrogen bond formation. The active site contains a single Ca++ ion whose coordination sphere

    contains PQQ and protein atoms, including both oxygens of the carboxylate of Glu177 and the

    amide oxygen of Asn261. The PQQ atoms include the C5 quinone oxygen, one oxygen of the C7

    carboxylate and, surprisingly, the N6 ring atom which is only 2.45Å from the metal ion (Fig. 5).

    The C4 and C5 oxygen atoms, which become reduced during the catalytic cycle, are hydrogen

    bonded to Arg331, which also makes hydrogen bonds between its NH2 and the carboxylate of

    Asp303 which is the most likely candidate for the base required by the catalytic mechanism.

    Ethanol Dehydrogenase of Pseudomonas species (QEDH) is also a type 1 alcohol

    dehydrogenase. This ethanol dehydrogenase (QEDH), induced during growth on ethanol of

    Pseudomonas or Rhodopseudomonas, is similar to MDH (Mutzel and Gorisch, 1991; Toyama et

    al., 1995; Keitel et al., 2000). It uses a specific cytochrome c550 as electron acceptor (Schobert

    and Gorisch, 1999), although this shows no sequence identity to cytochrome cL, the electron

    acceptor for MDH. Like MDH, QEDH has a high pH optimum, requires ammonia or

    alkylamines as activator in the dye-linked assay system (ferricyanide is not used as electron

    acceptor), and is able to oxidize a wide range of alcohol substrates including secondary alcohols,

    but it differs in its very low affinity for methanol; the Km for ethanol is about 15 M and that for

    methanol is about 1000 times higher. QEDH is homodimeric, the subunits being 65 kDa; it thus

    differs from MDH in lacking a small subunit.

    Unlike MDH, PQQ dissociates from QEDH after removal of Ca2+ with EDTA, this

    process being reversible after reconstitution in the presence of Ca2+ and PQQ (Mutzel and

    Goerisch, 1991). It is possible that the additional disulphide bridge in the subunit of MDH and

    the complex with the small subunit may lead to a stronger stabilization of the native

    conformation of the enzyme.

    The X-ray structure of the enzyme from Pseudomonas aeruginosa shows that, apart from

    differences in some loops, the folding pattern is very similar to the large (α) subunit of MDH

    (Keitel et al., 2000).

    There are different loops in the vicinity of the active site and several rather flexible loops

    protrude from the molecule surface and partly occupy the space filled by the small subunit of

    MDH. The PQQ is located in the center of the superbarrel, coordinated to a calcium ion. Most

    amino acid residues that make contact with the PQQ and the Ca2+ are similar to those in MDH.

  • 33

    The main differences in the active site region are a bulky tryptophan residue in the active-site

    cavity of MDH, which is replaced by a phenylalanine and a leucine side-chain in the QEDH, and

    a leucine residue right above the PQQ in MDH which is replaced by a tryptophan side-chain in

    QEDH. Both amino acid exchanges appear to have an important influence, causing the different

    substrate specificities of these otherwise very similar enzymes. Docking calculations suggest that

    one of the tryptophans must be able to change its orientation in order to accommodate the higher

    primary alcohols in the active site (Keitel et al., 2000). In addition to the Ca2+ ion in the active-

    site cavity, QEDH contains a second Ca2+-binding site at the N terminus, which contributes to its

    stability. Although the localization of the interaction surfaces between the subunits is identical in

    QEDH and MDH, the residues and the interactions involved are not conserved.

    1.9.1.2. The Type II alcohol dehydrogenase

    Soluble Quinohemoprotein Alcohol Dehydrogenase (QH-ADH) of Comamonas

    testosteroni is a type II alcohol dehydrogenase. The best-known quinohemoprotein ADH is that

    isolated from Comamonas testosteroni (Groen et al, 1986; Jongejan et al., 1998; Oubrie et al.,

    2002). It has also been described in Pseudomonas putida (Toyama et al., 1995; Chen et al.,

    2002) which produces two distinct forms, having different substrate specificities; ADH-IIB is

    induced during growth on butanol and ADH-IIG induced on glycerol. This same organism also

    produces a Type I alcohol dehydrogenase, induced during growth on ethanol. The electron

    acceptor for QH-ADH is a specific blue copper protein, azurin (Matsushita et al., 1999) which is

    probably oxidized directly by the membrane oxidase. This periplasmic enzyme is a monomer (71

    kDa) containing one molecule of PQQ and a single heme C. In the dye-linked assay system the

    pH optimum is 7.7 and there is no requirement for an amine activator. Because electron transfer

    from PQQ is by way of heme C this enzyme can also be assayed using ferricyanide. It has a wide

    specificity for primary and secondary alcohols, although it is unable to oxidize methanol; it also

    oxidizes aldehydes and can accept large molecules such as steroids as substrates. This has been

    exploited for enantiospecific oxidation of industrially important precursor molecules (synthons)

    (Geerlof et al., 1994). The enzyme has been extensively characterized by EPR, NMR and

    Raman resonance spectroscopy with respect to the nature of the heme and its relationship to

    PQQ (De Jong et al., 1995a; De Jong et al., 1995b), the conclusions being supported by the X-

    ray structures of the enzymes from Comamonas (Oubrie et al., 2002) and from Pseudomonas

  • 34

    (Chen et al., 2002). QH-ADH comprises two domains connected by a long linker (23 amino

    acids) which spans the whole length of the enzyme.

    The N-terminal dehydrogenase domain has the typical β barrel with its propeller fold,

    having the active site containing PQQ and a Ca2+ ion. The C-terminal domain, located on top of

    the dehydrogenase, is a type I cytochrome c, with 5 α helical segments, which enclose the C-type

    heme which is covalently bonded to cysteine residues and which has typical histidine and

    methionine heme iron ligands. A channel leads from the periplasm to the region of the PQQ and

    a second channel contains a chain of hydrogen-bonded water molecules between the periplasm

    and the cavity between the two domains. The N-terminal dehydrogenase domain is very similar

    to the α subunit of MDH, the PQQ being located at the top of the superbarrel in a hydrophobic

    cavity that is accessible through a deep and narrow channel. It is ‘sandwiched’ between a co-

    planar tryptophan and the disulphide ring as in MDH (Fig. 4) and it has in-plane bonding

    interactions with almost exactly the same side chains as in MDH, the only significant difference

    is that Arg331 which is bonded to the O5 of PQQ in MDH (Fig. 5) is replaced by a lysine side

    chain in QH-ADH as it is in mGDH. The ligation of the Ca2+ ion with PQQ and with amino acid

    side chains is also exactly the same as in MDH (Fig. 5). QH-ADH is the only alcohol

    dehydrogenase whose X-ray structure includes the substrate, or rather a product of substrate

    oxidation. In the case of the Comamonas enzyme this is tetrahydrofuran-2- carboxylic acid,

    presumably produced from the two step oxidation of tetrahydrofurfuryl alcohol (Oubrie et al.,

    2002). The tetrahydrofuran ring makes van der Waal’s contacts with the hydrophobic walls of

    the substrate cavity. An oxygen atom of the substrate carboxylate is hydrogen bonded to the

    active site aspartate (Asp303 in MDH), and the glutamate carboxylate that coordinates to the

    Ca2+ (Glu177 in MDH), and to the two sulfur atoms of the disulphide ring. The enzyme from

    Pseudomonas putida was crystallized in the presence of isopropanol, and acetone, its oxidation

    product, was shown to be present in the active site, close to the O-4 and O-5 of PQQ and close to

    a carboxylate oxygen of the proposed active site aspartate (Asp303 in MDH; Fig. 5) (Chen et al.,

    2002). The side chains of the products lie in a cavity lined with mainly hydrophobic side chains-

    cysteines, phenylalanines, tyrosine and proline. The volume of this substrate cavity is about 120

    Å3 which is about twice that of the Type I ethanol dehydrogenase and much larger than seen in

    the Xray structure of MDH (~18 Å3) (Chen et al., 2002). MDH is unable to oxidize secondary

    alcohols or primary alcohols with substituents on the C2 atom but it is still able to oxidized a

  • 35

    wide range of large alcohols and it is possible that the entrance to the active site might be flexible

    in order to accommodate these substrates. Because the catalytic machinery of MDH is strictly

    conserved in QH-ADH, the mechanism of alcohol oxidation is most likely identical for the two

    enzymes (Oubrie et al., 2002). The positions of the substrate products together with the results

    obtained by site directed mutagenesis of Asp303 in MDH (Afolabi et al., 2001), all suggest that

    Asp308 is the catalytic base. The mechanism for aldehyde oxidation is presumably essentially

    similar to that for alcohol oxidation: it is proposed that Asp308 abstracts a proton from a

    hydrogen-bonded water and the resulting hydroxyl ion performs a nucleophilic attack on the

    aldehyde C1 atom in concert with hydride transfer from this atom to the C5 of PQQ, to give the

    carboxylic acid product (Oubrie et al., 2002). The shortest distance between PQQ and the heme

    is 13–15 Å which is close to the maximum travel distance for electrons but the predicted rate of

    transfer through the protein is much higher than the measured rate of substrate oxidation. A

    number of paths are possible for the electron flow but they all involve the disulphide bridge and

    probably at least one water molecule (for example, see Fig. 8) (Chen et al., 2002; Oubrie et al.,

    2002). During oxidation of the reduced PQQ, protons are released into the periplasm. This is

    likely to be by way of a hydrogen bonded network involving a water filled chamber between the

    two domains, Lys335, Asp308 and Glu185 (Oubrie et al., 2002); these are equivalent to the

    MDH residues Arg331, Asp303 and Glu177 (Fig. 5). Azurin isolated from P. putida is a good

    electron acceptor for the QH-ADH, the interaction being mediated by hydrophobic forces

    (Matsushita et al., 1999). The heme is buried within the cytochrome domain except for one edge

    which is surrounded by a charge-neutral surface area which may form a binding site for azurin,

    in which one of the histidine ligands to the buried copper is exposed to the surface and is

    surrounded by a surface patch of hydrophobic residues (Chen et al., 2002).

    1.9.1.3. Type III alcohol dehydrogenase

    Membrane Associated Quinohemoprotein Alcohol Dehydrogenase of acetic acid bacteria

    is a type III alcohol dehydrogenase. This enzyme is a quinohemoprotein-cytochrome c complex

    and has only been described in the acetic acid bacteria, Acetobacter and Gluconobacter

    (Matsushita and Adachi, 1993; Matsushita et al., 1994, Goodwin and Anthony, 1998). Together

    with the membrane-bound aldehyde dehydrogenase, it is responsible for the oxidation of alcohol

    to acetic acid in vinegar production. It does not require ammonia as activator and has a pH

  • 36

    optimum of 4-6. Its substrate specificity is relatively restricted, oxidizing only a few primary

    alcohols (chain length, C2-C6) (but not methanol), or secondary alcohols and has some activity

    with formaldehyde and acetaldehyde.

    The Type III ADH has 3 subunits and is tightly bound to the periplasmic membrane,

    requiring detergent for its isolation. Translation of the gene sequences shows that all the subunits

    have N-terminal signal peptides typical of periplasmic proteins. Its natural electron acceptor is

    ubiquinone in the membrane. Subunit I (72-80 kDa) is a quinohemoprotein similar to the soluble

    Type II Quinoheamoprotein alcohol dehydrogenase, with a single molecule of PQQ and a single

    heme C. Its N-terminal region has sequence similarity to the soluble methanol dehydrogenase but

    with a C terminal extension having a single heme binding site. Subunit II (48k-53 kDa) has 3

    hemes that can be distinguished by biochemical techniques in the pure protein (Matsushita et al.,

    1996). Subunits I and II therefore have a total of 4 hemes. Most Type III ADHs have a third

    subunit (subunit III, 14-17 kDa) in which the gene is not linked to the genes encoding the other

    two subunits, and whose predicted amino acid sequence indicates that its processed size is

    greater (about 20 kDa) than that obtained by SDS -PAGE (14 kDa). The Type III ADH may be

    assayed with phenazine methosulfate, or with ferricyanide which reacts at the level of one or

    more of the heme C prosthetic groups on subunits I and II. It differs from all other ADHs in

    using short-chain ubiquinone homologues (Q1 and Q2) as electron acceptors and native

    ubiquinone (Q9 and Q10) when reconstituted in membrane vesicles (Matsushita et al., 1992).

    There is good evidence that electron transfer from reduced PQQ to the membrane ubiquinone

    takes place by way of the hemes on the cytochrome subunit II but that only two of them may be

    involved in this electron transfer process (Matsushita et al., 1996; Frebortova et al., 1998). It has

    been suggested that the cytochrome subunit II is firmly embedded in the membrane, that subunits

    I and III are firmly attached to each other and that this attachment helps the dehydrogenase

    subunit I couple with the cytochrome c (subunit II). This raises the question of how the

    ubiquinone in the membrane reacts with subunit II to accept electrons from its heme. Clearly part

    of the protein must be embedded in the membrane for this to occur but subunit II does not appear

    to have typical hydrophobic transmembrane helices (Kondo and Horinouchi, 1997). The Type III

    ADH thus appears to be unique in a number of ways; it requires detergent for its isolation from

    membranes and so seems to be a typical integral membrane protein, although none of the

    subunits appears (from their gene sequences) to have characteristic membrane protein structural

  • 37

    domains. Furthermore, the electron acceptor for the quinohemoprotein/cytochrome c complex is

    membrane ubiquinone, so we have the unusual situation where a c-type cytochrome precedes

    ubiquinone in the electron transport chain (Anthony, 2004).

    1.10. Pyroloquinoline Quinone (PQQ) Cofactor

    4, 5-dihydro-4, 5-dioxo-1H-pyrrolo- [2, 3- ] quinoline-2, 7, 9-tricarboxylic acid (PQQ)

    (Fig. 6) is an aromatic, tricyclic ortho-quinone that serves as the redox cofactor for several

    bacterial dehydrogenases. Among the best-known examples are methanol dehydrogenase and

    glucose dehydrogenase. PQQ belongs to the family of quinone cofactors that has been

    recognized as the third class of redox cofactors following pyridine nucleotide- and flavin-

    dependent cofactors.

    PQQ is a prosthetic group required by several bacterial dehydrogenases, including methanol

    dehydrogenase (MDH) of Gram negative methylotrophs, quinohemoprotein alcohol

    dehydrogenase and some glucose dehydrogenases. PQQ is derived from two amino acids,

    tyrosine and glutamic acid (Houck, 1991; Van Kleef, 1988) (i.e all carbon and nitrogen atoms of

    PQQ are derived from conserved tyrosine and glutamate residues), but the pathway for its

    biosynthesis is unknown.

    PQQ is an important cofactor of bacterial dehydrogenases, linking the oxidation of many

    different compounds to the respiratory chain. PQQ was the first of the class of quinone cofactors

    that have been discovered in the last 18 years and make up the prosthetic group of quinoproteins

    (Duine, 1991)

    PQQ was discovered in 1979 from a bacterium, and afterward it was reported to be in common

    foods. Because PQQ-deprived mice showed several abnormalities, such as poor development

    and breakable skin, PQQ has been considered as a candidate for vitamin. It was a mystery, that

    until 2003 it was not identified as vitamin. Since the first vitamin (now called vitamin B1) was

    discovered in 1910 by Dr. U. Suzuki, thirteen substances have been recognized as vitamins; the

    latest one was vitamin B12 found in 1948.So it takes 55 years to discover “PQQ” a previously

    identified substance as new vitamin ( Choi, 2008; Kashara and Kato, 2003).

  • 38

    1.11. Mechanisms of oxidation of alcohols in alcohol dehydrogenases,

    Two mechanisms have been proposed for the oxidation of alcohols in quinoprotein

    dehydrogenases, both of which begin with the pyroloquinoline quinone in an oxidized state.

    Initially, an addition/elimination mechanism was proposed, a suggestion that is now

    considered unlikely; rather, a hydride transfer mechanism is preferred (Oubrie et al., 1999;

    Oubrie and Dijkstra., 2000; Anthony and Williams, 2003) ( Fig. 7)

    Fig 6: Chemical structure of PQQ (4, 5-dihydro-4, 5-dioxo-1H-pyrrolo- [2, 3-f] quinoline-

    2, 7, 9-tricarboxylic acid)

    (Source:http://www.dlarborist.com/treetrends/2005/05/27/auxin_action_s.jpg)

  • 39

    Fig. 7. The current accepted reaction cycle for alcohol oxidation in quinoproteins

    Source: Anthony and Williams (2003)

    The mechanism is based on a hydride transfer from the alcohol to the C-5 position of the

    pyroloquinoline quinone (Oubrie et al., 1999). The back-reaction is via a radical intermediate

    protonated at O-4 or O-5. The IUPAC numbering scheme of PQQ is also shown. Following

    substrate binding, the reaction is initiated by amino acid (Asp(11) or Glu (25)) base-catalyzed

    proton abstraction of the hydroxyl proton of the alcohol. Nucleophilic attack of the hydride from

    the substrate to the C-5 position of PQQ then occurs. Subsequently, the PQQ enolizes to form the

    quinol. The reduced PQQ is reoxidized by two sequential single electron transfers (ET) to

    cytochrome c1 in MDH, cytochrome c550 in QEDH, or the cytochrome c domain in QH-ADH

    via the intermediate free radical (Duine and Frank, 1980; Frank, et al., 1988; Dijkstra et al.,

  • 40

    (1989), a process that is thought to be mediated by the disulfide bridge (Avezoux et al., 1995, Oubrie et

    al., 2002; Chen et al., 2002).

    Information that is not usually obtained from x-ray analysis but is necessary for

    obtaining full understanding of a dehydrogenation reaction cycle concerns the protonation states

    of the single and doubly reduced species. As shown in Fig. 8, apart from one review (Duine,

    1999), reduced PQQ is usually shown protonated at both O-4 and O-5, whereas the radical is

    depicted singly protonated at either O-4 (Zheng and Bruice, 1997; Duine et al., 1984) or O-5

    (Anthony,1996; Anthony and Williams, 2003; Oubrie, 2003), although a deprotonated radical

    was recently postulated (Sato et al., 2001). Knowledge of the protonation states is crucial if the

    electron transport (ET) and proton transfer pathway in both reoxidation steps are to be

    understood, because depending on the protonation states of the initial and final molecules, the

    reaction is either simple ET or must be accompanied by the release of a proton. Furthermore,

    apart from the driving force and the reorganization energy, according to ET theory (Marcus and

    Sutin.,1985), the rate of ET is dependent on the electroninc coupling between the donor (PQQ)

    and the acceptor (heme). Therefore, a full understanding of ET kinetics in quinoproteins will

    only be possible with knowledge of both the spatial and electronic structures of the ET partners

    (Davidson, 2004). The later may be provided by electron nuclear double resonance (ENDOR)

    via determination of hyperfine coupling (hfcs) in combination with density functional theory

    (DFT) calculations (Buttner et al., 2005). These methods enable us to establish that the PQQ

    radical is deprotonated when bound in QEDH from Pseudomonas aeruginosa

    The other enzyme involved in the oxidation of ethanol is aldehyde dehydrogenase. It is

    also a NADP+ independent enzyme and located in the cytoplasmatic membrane. Its optimum pH

    is between 4 and 5, although it can catalyse the oxidation of acetaldehyde to acetate at lower pH

    values (Adachi et al., 1980). It is an enzyme that is sensitive to oxygen concentrations, and when

    these are low its activity decreases, accumulating acetaldehyde. It is also more sensitive to the

    presence of ethanol than alcohol dehydrogenase (Muraoka et al., 1983).

    1.12. In vitro and in vivo properties of Alcohol dehydrogenase

    The particulated alcohol dehydrogenase could be assayed in vitro in the presence of one

    of the following dyes as an electron acceptor; 2,6-dichlorophenolindophenol, phenazine

    methosulfate or potassium ferricyanide. NAD or NADP were not effective as an electron

  • 41

    acceptor at all (Adachi et al., 1978). Many people still believe that acetate is produced by the

    cytosolic NAD(P)-dependent alcohol dehydrogenase and keto-D-gluconate by the cytosolic

    NAD(P)-dependent D-gluconate dehydrogenase located in the cytoplasm. Such a serious

    confusion is probably caused by the confused of localization of the enzymes concerned. Before

    describing the the actions of the individual PQQ- and FAD-dependent dehydrogenases, it is

    worth clarifying the common physiological roles and localizations of PQQ-and FAD-dependent

    dehydrogenases in acetic acid bacteria and other microorganisms. At present, of the enzymes

    exploited as either PQQ-dependent or FAD-dependent dehydrogenases, aldehyde dehydrogenase

    is the only one that is known to use a molybdopterin coenzyme. Unlike the cytoplasmic

    oxidoreductases, no energy is required for substrate intake into the periplasm and pumping out

    the oxidation products across the outer membrane. M