Download In the light of the haloarchaea metabolism

Survey
yes no Was this document useful for you?
   Thank you for your participation!

* Your assessment is very important for improving the workof artificial intelligence, which forms the content of this project

Document related concepts

Mitogen-activated protein kinase wikipedia , lookup

Protein wikipedia , lookup

Lac operon wikipedia , lookup

Lipid signaling wikipedia , lookup

Basal metabolic rate wikipedia , lookup

Photosynthesis wikipedia , lookup

Catalytic triad wikipedia , lookup

Restriction enzyme wikipedia , lookup

Paracrine signalling wikipedia , lookup

Enzyme inhibitor wikipedia , lookup

NADH:ubiquinone oxidoreductase (H+-translocating) wikipedia , lookup

Nitrogen cycle wikipedia , lookup

Fatty acid metabolism wikipedia , lookup

Fatty acid synthesis wikipedia , lookup

Metabolic network modelling wikipedia , lookup

Phosphorylation wikipedia , lookup

Glucose wikipedia , lookup

Digestion wikipedia , lookup

Glyceroneogenesis wikipedia , lookup

Biochemical cascade wikipedia , lookup

Metalloprotein wikipedia , lookup

Nicotinamide adenine dinucleotide wikipedia , lookup

Proteolysis wikipedia , lookup

Oxidative phosphorylation wikipedia , lookup

Glycolysis wikipedia , lookup

Evolution of metal ions in biological systems wikipedia , lookup

Enzyme wikipedia , lookup

Microbial metabolism wikipedia , lookup

Amino acid synthesis wikipedia , lookup

Biosynthesis wikipedia , lookup

Citric acid cycle wikipedia , lookup

Biochemistry wikipedia , lookup

Metabolism wikipedia , lookup

Transcript
Communicating Current Research and Educational Topics and Trends in Applied Microbiology
A. Méndez-Vilas (Ed.)
_____________________________________________________________________
In the light of the haloarchaea metabolism
M.J.Bonete, M.Camacho, R.M.Martínez-Espinosa, J. Esclapez, V.Bautista, C.Pire,
B.Zafrilla, S.Díaz, F.Pérez-Pomares and F.Llorca.
División de Bioquímica y Biología Molecular, Grupo Proteínas de extremófilos, Facultad de
Ciencias, Universidad de Alicante, Ap. 99, 03080 Alicante, Spain
Keywords: haloarchaea, carbon metabolism, nitrogen metabolism
1. Introduction.
Nitrogen (N), carbon (C), hydrogen (H) and oxygen (O) are the major elements of all organisms and are
the components of the most important biological macromolecules: proteins, lipids, carbohydrates and
nucleic acids. The reactions performed in different ways to allow the conversion of chemical species into
other compounds, make up the biological metabolic pathways which are not well known in halophilic
archaea. Using Haloferax mediterranei, Haloferax volcanii and Halobacterium salinarum as haloarchaea
models the main metabolic pathways have been characterised in the last few years.
Halophilic archaea (haloarchaea) grow in hypersaline (3–5 M NaCl) environments, including
hypersaline lakes and marine salterns [1] . Acetate could have a role in the nutrition of natural
communities of halophilic Archaea, as it is produced from glycerol (the main carbon source in nature) in
hypersaline lakes by some species of halophiles [2]. With regard to carbon metabolism it has been
characterised the glyoxylate cycle, a metabolic pathway by which organisms can synthesize
carbohydrates from C2 compounds, in Hfx. volcanii [3].
It has been demonstrated that nitrate, nitrite or ammonium can be used by some halophilic archaea for
growth in presence of oxygen. This pathway is known as “assimilatory nitrate reduction” and it was
characterised first in Hfx. mediterranei. However, under anoxic conditions, nitrate and nitrite are the final
electron acceptor in a respiratory process denominated denitrification.
In our research group we have characterised some of the enzymes involved in different metabolic
pathways.
2. Carbon metabolism in haloarchaea
Organisms of the three domains of life need polysaccharides as a carbon source to support their
heterotrophic growth. In general, the utilization of polysaccharides involves their extracellular
hydrolysis, the intake of oligosaccharides by specific transporters and their intracellular hydrolysis to
generate hexoses and pentoses. Subsequently, these monosaccharides are oxidized via a well-conserved
set of central metabolic pathways.
The glycolytic pathway and the citric acid cycle are the main energy producing pathways in cells.
Although these pathways show a great variability between organisms, many of the core enzymes of both
glycolysis and the citric acid cycle seem to be conserved in the majority of organisms. Glycolysis is
responsible for sugars metabolism, in particular glucose, and its conservation in the evolution of the
organisms suggests that glucose is a major source of energy for many organisms. Glucose is found in
nature primarily in a polymeric state as cellulose, starch and glycogen.
2.1. Amylases.
170
©FORMATEX 2007
Communicating Current Research and Educational Topics and Trends in Applied Microbiology
A. Méndez-Vilas (Ed.)
_____________________________________________________________________
Starch is an important source of energy especially for microorganisms. It is composed exclusively of Dglucose units that are linked by α-1,4- or α-1,6-glycosidic bonds. The two components of starch are
amylose (15–25%), a linear polymer consisting of α-1,4-linked glucopyranose residues, and amylopectin
(75–85%), a branched polymer containing, in addition to α-1,4 glycosidic linkages, α-1,6-linked branch
points occurring every 17–26 glucose units. Selective hydrolysis of glycosidic bonds is therefore crucial
for energy uptake, and other biological functions like cell wall expansion and degradation, and turnover
of signalling molecules. Enormous amounts of glucose are available to organisms that can hydrolyse
glucose polymers efficiently, that show the significance that glucose polymers have for life [4].
Because of the complex structure of starch, cells require an appropriate combination of intracellular
and extracellular enzymes for its conversion to oligosaccharides and smaller sugars, such as glucose and
maltose. There are basically four groups of starch-converting enzymes [5]:
- Endoamylases: these enzymes are able to cleave α-1,4-glycosidic bonds in the inner part of the
amylose or amylopectin chain. A well studied endoamylase is α-amylase (EC 3.2.1.1) which end
products reaction are oligosaccharides with a α-configuration and α-limit dextrins.
- Exoamylases: there are enzymes that only cleave α-1,4 glycosidic bonds such as β-amylase (EC
3.2.1.2) or cleave both α-1,4 and α-1,6 glycosidic bonds like amyloglucosidase or glucoamylase (EC
3.2.1.3) and α-glucosidase (EC 3.2.1.20). Exoamylases act on the external part of amylose or
amylopectin and thus produce only glucose (glucoamylase and α-glucosidase), or maltose and β-limit
dextrin (β-amylase).
- Debranching enzymes: to this group belong enzymes that exclusively hydrolyze α-1,6 glycosidic
bonds, like isoamylase (EC 3.2.1.68) and pullulanase type I (EC 3.2.1.41). Its degradation products are
maltotriose and linear oligosaccharides.
- Transferases: these starch-converting enzymes cleave an α-1,4 glycosidic bond of the donor
molecule and transfer part of this donor to a glycosidic acceptor with the formation of a new glycosidic
bond. Enzymes such as amylomaltase (EC 2.4.1.25) and cyclodextrin glycosyltransferase (EC 2.4.1.19)
form a new α-1,4 glycosidic bond while branching enzyme (EC 2.4.1.18) forms a new α-1,6 glycosidic
bond. Cyclodextrin glycosyltransferase produces a series of non-reducing cyclic dextrins, α−, β- and γcyclodextrins, that contain six, seven and eight glucose units respectively. The branching enzyme
produces cyclic glucans from amylose and amylopectin.
Most of the starch-converting enzymes belong to the α-amylase family or family 13 glycosyl
hydrolases according to the classification of Henrissat [6]. Family GH13, is the largest sequence-based
family of glycosyl hydrolases. This family groups together a great number of enzymes with different
activities but all of them share the following characteristics:
• They act on α-glycosidic bonds hydrolyzing them (i) to produce α-anomeric mono- or
oligosaccharides (hydrolysis); (ii) to form α-1,4 or 1,6 glycosidic linkages (transglycosylation);
(iii) or a combination of both activities.
• They adopt a (β/α)8 barrel structure that contains the catalytic site residues.
• They share four highly conserved regions in their primary sequence which contain the amino
acids that form the catalytic site, and some amino acids that are essential for the stability of the
(β/α)8 barrel [7].
Structurally, the GH13 enzymes are multidomain proteins characterized by a conserved structural core
composed of three domains often designated as domains A, B and C. Domain A is the catalytic domain
in the form of a (β/α)8 barrel, a barrel of eight parallel β-strands surrounded by eight helices, the socalled domain A [8]. Domain B is a loop of variable length between the third β-strand and third helix of
the (β/α)8 barrel. This loop has an irregular structure that varies from enzyme to enzyme. The active site
is found in a cleft between domains A and B where a triad of catalytic residues performs catalysis.
Domain C is a C-terminal extension following the catalytic domain characterized by a Greek key
structure. This domain is made up of β-strands and is thought to stabilise the catalytic domain by
shielding hydrophobic residues of domain A from the solvent. It has also been suggested that domain C
©FORMATEX 2007
171
Communicating Current Research and Educational Topics and Trends in Applied Microbiology
A. Méndez-Vilas (Ed.)
_____________________________________________________________________
may aid in substrate binding. In addition to this conserved core, some members of family GH13 have
additional β-sheet domains after domain C, domains D and E. While no role has been assigned to domain
D, domain E is believed to be important for binding sugar or granular starch.
Many starch-degrading proteins have been identified in various organisms, but relatively few studies
are focused on archaeal α-amylases. In recent studies it has been purified and biochemically
characterised a haloarchaeal α-amylase from Hfx. mediterranei by our research group [9]. The behaviour
of this enzyme is very close to that reported for the amylases from Hbt. salinarum [10], and
Natronococcus amylolyticus [11], and the moderately halophilic bacteria Halomonas meridiana [12].
The enzymatic hydrolysis and modification of polysaccharides is of great interest in the field of
biotechnology (food, chemical, and pharmaceutical industries) [13]. α-Amylases are used in the starch
liquefaction process that converts starch into fructose and glucose syrups. They are also used as a partial
replacement for the expensive malt in the brewing industry, to improve flour in the baking industry, and
to produce modified starches for the paper industry. In addition to this, they are used to remove starch in
the manufacture of textiles and as additives to detergents for both washing machines and automated dishwashers.
2.2 Glucose dehydrogenase
Different pathways are involved in the catabolism of glucose. The classical Embden-Meyerhof
(EM) pathway, or glycolysis, is well-conserved in bacteria and eukaryotes. In archaea, the best study
about EM pathway has been carried out in the thermophilic microorganism Pyrococcus furiosus. These
studies have revealed a modified EM pathway which presents novel enzymes involved in the catabolism
of glucose. In addition, a new regulation site for energy metabolism and a unique mode of ATP
regeneration have been postulated [14-16]. The Entner-Doudoroff (ED) is also widely distributed in
nature. In bacteria consists of nine enzymes, and starts with the phosphorylation and oxidation of
glucose. The main difference between both pathways is found at the number of ATP obtained per
glucose molecule, two ATP molecules are generated in the EM whereas the ED pathway only yields a
single ATP [17].
172
©FORMATEX 2007
Communicating Current Research and Educational Topics and Trends in Applied Microbiology
A. Méndez-Vilas (Ed.)
_____________________________________________________________________
Figure 1: Catabolism pathway of glucose. Entner-Doudoroff pathway is represented as continuous line, modified
ED pathway of halophilic archaea is shown in broken line and non-phosphorylated ED pathway of thermophilic
microorganisms is represented as dotted line.
In some archaea, glucose is metabolized by way of modifications of the ED pathway (Figure 1). The first
one is the non-phosphorylated ED pathway, which was discovered in extreme thermoacidophiles like
Thermoplasma acidophilum and Sulfolobus solfataricus [18,19]. The second one was found in halophilic
archaea (and some bacteria) and it is the modified ED pathway, in which the phosphorylation step is
postponed. Glucose is oxidized to gluconate via NADP+-dependent glucose dehydrogenase (Figure 1).
Evidence for the operation of this pathway has been demonstrated for species of Halobacterium,
Haloferax and Halococcus [20].
Extremely halophilic archaea are considered a rather homogeneous group of heterotrophic
microorganisms predominantly using amino acids as their source of carbon and energy. However, it has
been shown that some halophilic archaea are able to use not only amino acids but different sugars as
well. Hfx. mediterranei for example grows in a minimal medium containing glucose as the only source
of carbon [21] using a modified ED pathway. The first step of this pathway is catalyzed by NAD(P)+ dependent glucose dehydrogenase (GlcDH). This protein is a dimeric enzyme with a molecular weight of
39 kDa per subunit. It shows dual cofactor specificity, although it displays a marked preference for
NADP+ over NAD+. The protein requires a zinc ion for catalysis and sequence analysis revealed that this
enzyme belongs to the zinc-dependent medium-chain alcohol dehydrogenase superfamily [22]. The
©FORMATEX 2007
173
Communicating Current Research and Educational Topics and Trends in Applied Microbiology
A. Méndez-Vilas (Ed.)
_____________________________________________________________________
activity of halophilic GlcDH was dependent on the temperature, showing its maximum between 40 and
50 ºC. The thermophilic nature of this kind of proteins has been noted in most of the halophilic enzymes,
for instance Haloarcula marismortui malate dehydrogenase. Structural studies of this enzyme have
revealed that several of the structural features conferring halophilicity are the same as those contributing
to the stability of thermophilic enzymes [23].
To study their structural properties and to extend the understanding of the molecular basis of salt
tolerance responsible for halophilic adaptation, glucose dehydrogenase from Hfx. mediterranei has been
overexpressed [24] and its structure has been solved at 1.6 Ǻ. Analysis of protein solvent interactions
reveals a highly ordered multilayered solvatation shell that can be seen to be organized into one domain
network covering much of the exposed surface accessible area to an extent not seen in almost any other
protein structure solved [25]. These studies are important not only to establish the molecular basis of
adaptation at high ionic strength mediums of this kind of proteins, but also to amplify the information
available about the halophilic enzymes involved in the catabolism of glucose.
A study of the metabolic adaptations of Hfx. volcanii by transcriptome analysis using shot-gun DNA
microarrays shown that after glucose induction transcript levels of glucose dehydrogenase and 2-keto-3desoxy-6-phosphogluconate (KDPG) aldolase were found to be increased [26]. Also induced transcript
levels were found in a gene encoding a protein with similarity to 12 transmembrane helix transport
proteins, which was proposed to be a glucose transporter, whereas transcript levels of several enzymes
involved in glucose biosynthesis are repressed. This study also reveals that the citric acid cycle, the
electron transport chain and ATP synthesis are down regulated after the shift to glucose medium. A
similar down regulation of some genes of the citric acid cycle and of respiration has been postulated in
Haloarcula marismortui [27].
It has been shown in various halophilic archaea that during exponential growth on glucose significant
amounts of acetate were formed [28]. The formation of acetate from acetyl-CoA is catalyzed by an ADPforming acetyl-CoA synthetase (ACD). This unusual synthetase was found in all acetate forming
archaea, including anaerobic hypertermophiles. In Har. marismortui during growth on glucose and on
glucose/acetate both ACD and glucose dehydrogenase activity increased parallel to phases of glucose
consumption and acetate formation and both activities decreased during growh on acetate or peptides.
Acetate formation in Haloarcula appears to be restricted to sugar metabolism under conditions when the
rate of glycolisis exceeds that of subsequent pathways.
2.3 Isocitrate DH
The citric acid cycle is the common mode of oxidative degradation in eukaryotes and prokaryotes. This
cycle, which is alternatively known as the tricarboxylic acid (TCA) cycle and the Krebs cycle, accounts
for the major portion of carbohydrate, fatty acid, and amino acid oxidation and generates numerous
biosynthetic precursors (Figure 2). The citric acid cycle is therefore amphibolic, that is, it operates both
catabolically and anabolically.
The eight enzymes of the citric acid cycle catalyse a series of well-known organic reactions that
cumulatively oxidize an acetyl group to two CO2 molecules with the concomitant generation of three
NADHs, one FADH2 and one GTP. Archaea metabolize pyruvate (which is produced by hexose
catabolism), via the citric acid cycle. However, there are differences in the enzymes used. For example,
archaea catalyse pyruvate to acetyl CoA via pyruvate oxidoreductase, having ferredoxin as electron
acceptor in halophiles and thermophiles and a deazin derivative, F420 as the electron acceptor in
methanogens [29]. This enzyme is also found in anaerobic eubacteria which might suggest that the
oxidoreductase existed before the divergence of the archaeal, eubacterial and the ancestral eukaryotic
line of descent. The pyruvate dehydrogenase (PDH) multi-enzyme complex, characteristic of respiratory
eubacteria and eukaryotes, is absent from archaea; however, thermophiles and halophiles contain
dihydrolipoamide dehydrogenase which is normally the third component in the PDH complex [30].
Therefore, it is likely that the PDH complex may have evolved soon after development of oxidative
phosphorylation since the complete complex has only been detected in respiratory organisms [31].
174
©FORMATEX 2007
Communicating Current Research and Educational Topics and Trends in Applied Microbiology
A. Méndez-Vilas (Ed.)
_____________________________________________________________________
The type of citric acid cycle seen in archaea depends on the class to which they belong. For example,
halophiles metabolize pyruvate via an oxidative citric acid cycle. Sulphur dependent thermophiles
generally fix carbon dioxide via a reductive citric acid cycle when growing autotrophically and use an
oxidative cycle when heterotrophic [32]. In methanogens there is no complete citric acid cycle; instead
there are two different routes in which 2-oxoglutarate is formed [33]. One uses an incomplete reductive
pathway where 2-oxoglutarate is synthesized from oxaloacetate, and the other uses an incomplete route
where 2-oxoglutarate is formed via citrate. It is possible that the type of citric acid cycle present in an
archaea will determine enzyme function and, hence, reflect any unique features.
Isocitrate dehydrogenase (EC 1.1.1.41 or EC 1.1.1.42) (ICDH), catalyses the oxidative
decarboxylation of isocitrate:
Isocitrate + NAD( P)+ 2-oxoglutarate + NAD(P)H + H+ + CO2
A number of isoenzymes of isocitrate dehydrogenase, distinguished by their cofactor requirement and
subcellular localization, have been identified. In Eukarya, ICDH exists in two forms: an NAD-linked
enzyme found only in the mitochondrion and displaying allosteric properties, and a non-allosteric
NADP-ICDH that is found in both the mitochondrion and cytoplasm [34,35]. Most Bacteria possess only
an NADP-ICDH [36], although NAD-linked [37] and dual-cofactor-specific (NAD and NADP) bacterial
enzymes are known [38,39].
The purification and characterisation of ICDH from the extremely halophilic Archaeon, Hfx. volcanii
have been determined. It is a dimer; strictly NADP-dependent; and its thermal stability has been studied
at 0.5 M or 3 M KCl, obtaining an activation energy for the thermal inactivation process of 360 kJ mol-1
and 610 kJ mol-1, respectively [39]. Also, the cloning of the gene encoding this ICDH from Hfx. volcanii
has been described [40]. The DNA sequence has been determined, and the derived amino acid sequence
has been compared with known ICDH sequences. The protein deduced from the nucleotide sequence is
composed of 419 amino acids with a molecular mass of 45 837 Da. The molecular weight of halophilic
enzymes should be treated with some caution in SDS-PAGE because they electrophorese more slowly
because of their highly negative intrinsic charges. For halophilic ICDH, this subunit molecular mass was
62 000 Da. Also, the enzyme shows 12 α-helical and 12 β-sheet conformations in the polypeptide. Other
typical studies of halophilic enzymes have been investigated with ICDH. The salt-dependent stability of
the recombinant dimeric ICDH from the halophilic Archaeon Hfx. volcanii has been studied in various
conditions [41]. The enzyme stability is found to be mainly sensitive to cations and very little (or not)
sensitive to anions. Divalent cations induce a strong shift of the active/inactive transition towards low
salt concentration.
Dehydrogenases discriminate between nicotinamide coenzymes through interactions established
between the protein and the 2′-phosphate of NADP+ and the 2′- and 3′-hydroxyls of NAD+ [42]. In NAD
binding site, the introduction of positively charged residues changes the preference of an NADdependent enzyme to neutralize the negatively charged 2′-phosphate of NADP+. In the halophilic ICDH,
the objective was to switch the coenzyme specificity by altering five conserved amino acids by sitedirected mutagenesis (Arg291, Lys343, Tyr344, Val350 and Tyr390). The five mutants of ICDH were
overexpressed in E. coli as inclusion bodies [43] and each recombinant ICDH protein was refolded and
purified, and its kinetic parameters were determined. Coenzyme specificity did not switch until all five
amino acids were substituted.
2.4 Isocitrate lyase and malate synthase
In general, the Krebs cycle functions similarly in bacteria and eukaryotic systems, but major
differences are found among bacteria. One difference is that in obligate aerobes, L-malate may be
oxidized directly by molecular O2 via an electron transport chain. In other bacteria, only some Krebs
cycle intermediate reactions occur because 2-oxoglutarate dehydrogenase is missing.
©FORMATEX 2007
175
Communicating Current Research and Educational Topics and Trends in Applied Microbiology
A. Méndez-Vilas (Ed.)
_____________________________________________________________________
A modification of the Krebs cycle is the commonly called “glyoxylate cycle”, which has been detected
in the three domains of living organisms. It is a metabolic pathway by which organisms can synthesize
carbohydrates from C2 compounds. This cycle functions similarly to the Krebs cycle but lacks many of
the Krebs cycle enzyme reactions (Figure 2). The glyoxylate cycle involves five enzymes, three of which
also participate in the citric acid cycle: citrate synthase, aconitase, and malate dehydrogenase. The two
other enzymes, isocitrate lyase (ICL) and malate synthase (MS), are unique to this cycle.
Figure 2. Krebs and glyoxylate cycles.
Isocitrate lyase (EC 4.1.3.1) catalyzes the cleavage of D-isocitrate to glyoxylate and succinate; and
malate synthase (EC 4.1.3.2) catalyzes the formation of L-malate from glyoxylate and acetyl CoA. The
net function of glyoxylate cycle is the formation of a C4 compound from two acetyl CoA (C2) molecules
in each cycle. One primary function of the glyoxylate cycle is to replenish the tricarboxylic and
dicarboxylic acid intermediates that are normally provided by the Krebs cycle. A pathway whose
primary purpose is to replenish such intermediate compounds is called anaplerotic.
In microorganisms growing on acetate, isocitrate can be metabolized either by the citric acid cycle or
by the glyoxylate bypass [44]. In Escherichia coli, this branch point is regulated by reversible
phosphorylation and inactivation of NADP+-isocitrate dehydrogenase, but the regulation of isocitrate
flux in eukaryotic microorganisms remains to be established. The growth of E. coli on acetate requires
the operation of the glyoxylate bypass in which ICL must compete with ICDH for isocitrate. However,
ICDH has a much higher affinity for isocitrate than does ICL. Inactivation of a large part of the ICDH
allows intracellular isocitrate to rise to a concentration at which ICL can operate, and as soon as the
glyoxylate bypass is not required, ICDH is rapidly reactivated [45].
Acetate could have a role in the nutrition of natural communities of halophilic Archaea, as it is
produced from glycerol (the main carbon source in nature) in hypersaline lakes by some species of
halophiles. Hfx. volcanii is a halophilic archaeon able to grow in minimal medium with acetate as the
sole carbon source. ICL activity was detected in this organism when it was grown on a medium with
acetate as the main carbon source [46]. We demonstrated that the activities of glyoxylate cycle key
enzymes are necessary in Hfx. volcanii to synthesize precursors for carbohydrates only when the carbon
source is a C2 compound, such as acetate [3]. Those activities are not needed when the carbon source is a
compound with three or more carbons, such as lactate. No activity could be detected when lactate was
the carbon source, but high activities were measured when acetate was.
We have isolated and sequenced the ace (acetate assimilation) gene operon, comprising the glyoxylate
cycle key enzymes isocitrate lyase and malate synthase genes (icl or aceA and ms or aceB), from the
176
©FORMATEX 2007
Communicating Current Research and Educational Topics and Trends in Applied Microbiology
A. Méndez-Vilas (Ed.)
_____________________________________________________________________
halophilic archaeon Hfx. volcanii. This is the first time that these genes are sequenced in an organism
from the domain Archaea. Phylogenetic analysis of the sequenced genes revealed that isocitrate lyase
shows a significant identity with isocitrate lyases from Eukarya and Bacteria, but it is not more closely
related to eukaryal or bacterial enzymes, and that malate synthase from Hfx. volcanii has very little
identity with any other known protein. This enzyme forms a new class of malate synthases. Afterwards,
transcriptional analysis has showed that both genes are cotranscribed into a single mRNA molecule. The
genes were transcribed only when the acetate was the carbon source, indicating transcriptional regulation
[47]. Moreover, there are two sets of palindromic sequences in the promoter region; they are possibly
involved in binding of transcriptional regulator(s) (repressor or, activator proteins). This is an interesting
point because there are only a few studies on transcriptional regulation in Archaea, which are not enough
to form a global idea of how the process take place.
3. Nitrogen metabolism
Nitrogen (N) is a major element of all organisms and it accounts for 6% approximately of their dry mass
on average. In the nature, N can be found in different redox states from +5 (as nitrate) to -3 (as
ammonia). Redox reactions are performed in different ways by several organisms, and the reactions in
total make up the biological N-cycle which is summarised in Figure 3. All of these reactions are
performed by bacteria, cyanobacteria, archaea and some specialized fungi. However, there is only one
example of a higher life form performing such a reaction; the remarkable exceptions are assimilatory
nitrate reduction and nitrogen fixation, which also occurs in plants.
Figure 3. NO3- is used as nitrogen source for growth under aerobic conditions thanks to the assimilatory NO3reduction, while it acts as an electron acceptor to eliminate excess of reductant power through dissimilatory NO3reduction. Dissimilatory NO3- reduction, NO3- respiration or denitrification are used equivalently in the literature
because in all cases, N is finally excreted as gas, however, dissimilatory pathway makes reference to nonassimilatory reaction not directly coupled to generation of proton-motive force. In some Enterobacteriaceae member,
NO2- is reduced to NH4+ which is then excreted; this process is known as NO3-/ NO2- ammonification. Specialised
organisms are able to oxidize either NH4+ or NO2- by using a pathway called nitrification, while other organisms
such us some planctomycetes oxidize NH4+ and utilize NO2- as respiratory electron acceptor in a pathway named
anammox. Finally, (di)nitrogen fixation allows several bacteria and archaea to reduce N2 to NH4+ to provide Nrequirements.
©FORMATEX 2007
177
Communicating Current Research and Educational Topics and Trends in Applied Microbiology
A. Méndez-Vilas (Ed.)
_____________________________________________________________________
3.1 Assimilatory nitrate reduction
Nitrate assimilation is a main process of the N-cycle and is carried out by higher plants, fungi, algae,
many bacteria, cyanobacteria and some archaea. Thanks to this pathway, NO3- is used as N source for
growth under oxygenic conditions. More than 2 x 1013 kg of N per year are removed from the
environment by NO3- assimilation. This metabolic pathway has been studied extensively in higher plants,
fungi and algae from genetic, physiological and biochemical points of view. However, physiological and
biochemical studies have been only performed in a few bacterial and archaea members. In fact, in
halophilic archaea this pathway has been only described from Hfx. mediterranei, a denitrifier
haloarchaea that has been chosen as a model to analysed N-cycle in our laboratory [48]
Assimilation of nitrate involves three pathway-specific steps: uptake, reduction to nitrite and
further reduction to ammonium. External nitrite can also be taken up and reduced directly to ammonium
(Figure 4).
L-glutamate
α-ketoglutarate
Gln
Fd
eGS
NO3-
NO3-
N as
NO2-
NiR
GOGAT??
L-glutamate
NH4+
GDH
e-
e-
Fd
L-glutamate + NAD(P) + H2O
α-ketoglutarate + NAD(P)H
Figure 4. Assimilatory nitrate reduction in Hfx. mediterranei. All the enzymes presented in
the figure has been purified and characterised except glutamate synthase.
With regard to the first step, NO3- is incorporated into the cells by high-affinity transport. These
transporters have been characterised so far in bacteria such as Klebsiella, Bacillus and Azotobacter.
There are different NO3- uptake systems: ABC-type transporters and MFS-type permeases. However,
there is no any analysis of this kind of transporters in halophilic archaea. The assimilatory nitrate operon
described in our research group has revealed that the nasB gene probably encodes a NO3- transporter of
NarK group, which is proton-motive force dependent transporter [49]. When NO3- is in the cytoplasm, it
is reduced to NO2- by assimilatory nitrate reductase (Nas, EC 1.6.6.2), a cytoplasmic molybdoenzyme
that catalyse a two-electron reduction. This enzyme is structurally and functionally different from the
dissimilatory or respiratory nitrate reductases.
Nas are traditionally classified into two classes: one class depends upon ferredoxin (Fd) or flavodoxin
as reductant, while the other depends upon NADH. Fd-Nas have been characterised as monomeric
enzymes containing one 4Fe-4S cluster and molybdenum cofactor as redox centres, while NADH
enzymes are heterodimers in most of the described cases. Nas from Hfx. mediterranei is a monomeric
enzyme with remarkable thermohalophilicity character as it is expected for a halophilic enzyme [50].
This enzyme takes electrons from ferredoxin [51], which is a small protein containing a 2Fe-2S cluster.
2Fe-2S cluster ferredoxins and flavodoxins are very low negative redox potential electron donors (Em ≈ 300 mV), so electrons are transferred from ferredoxin to the 4Fe-4S cluster of Nas (Em ≈ -190 mV) and
then from this redox centre to de Mo-cofactor (MoCo) (Em ≈ -150 mV); in this cofactor takes place the
substrate reduction. The MoCo in bacteria consists of a unique pterin complexed with Mo, forming MoMPT, and modified by the addition of a guanosine to form Mo-MGD. We have sequenced the mobA
178
©FORMATEX 2007
Communicating Current Research and Educational Topics and Trends in Applied Microbiology
A. Méndez-Vilas (Ed.)
_____________________________________________________________________
gene from Hfx. mediterranei and its product is implied in Mo-MGD synthesis, for this reason, Hfx.
mediterranei Nas could be included in the Nas bacterial group [49].
Nitrite produced by Nas is reduced to ammonium by assimilatory nitrite reductase (NiR, EC 1.7.7.1)
which is also a monomeric enzyme that catalyses a six-electron reduction. This enzyme is structurally
and functionally different from the dissimilatory or respiratory nitrite reductases. According to their
electron donor specificity, two types of NiRs have been described: Fd-dependent NiRs and NAD(P)Hdependent NiRs. The last type of NiR is usually present in fungi and most heterotrophic bacteria. NiR
from Hfx. mediterranei is an Fd-dependent enzyme that contains siroheme and a 4Fe-4S cluster as redox
centres. In this case, electrons flow from ferredoxin to the 4Fe-4S cluster of NiR and then from this
centre to the siroheme, where the NO2- takes place. At the moment, this is the only assimilatory nitrite
reductase characterised from a halophilic archaea [49, 52] (Figure 4). The resulting ammonium is finally
metabolized through central pathways such as those where glutamine synthetase-glutamate synthase
(GS-GOGAT) or glutamate dehydrogenase (GDH) play the main role (see below).
NO3- assimilation in general is regulated by NO3- and/or NO2- induction (pathway-specific control)
and NH4+ repression (general N control). Nevertheless, the regulatory proteins and systems involved in
both controls are quite different depending on the organism. Related to this point, recently we have
demonstrated that in Hfx. mediterranei, ammonium exerts a negative effect on nitrate assimilation as it
has been described from most of the organisms [53], but this negative effect was reversed in presence of
L-methionine-D,Lsulfoximine (MSX), an irreversible inhibitor of glutamine synthetase activity [54], so
assimilation of NH4+, rather than NH4+ per se, has a negative effect on assimilatory nitrate and nitrite
reduction in Hfx. mediterranei.
3.2 Ammonium assimilation.
3.2.1 Glutamate dehydrogenases.
L-glutamate dehydrogenases (EC 1.4.1.2-4, GDH) are important enzymes in any cell due to the pivotal
position occupied by glutamate and 2-oxoglutarate in the central metabolism of nitrogen and carbon
compounds, being implicated in the two major pathways (GS-GOGAT and GDH metabolic routes).
They catalyse the interconversion between 2-oxoglutarate and L-glutamate reversibly using NAD(P)H or
both as cofactors and playing a key role since they provide a link between carbon and nitrogen
metabolism [55].
2-oxoglutarate + NAD(P)H + NH4+ + H+ ↔ L-glutamate + NADP+ + H2O
GDHs are evolutionarily conserved in the three domains, Archaea, Bacteria and Eukarya. These
enzymes are classified into three groups according to their coenzyme specificity: NAD or NADPspecific and NAD(P)-non specific dependent GDH. Generally, in prokaryotic organisms, NAD-GDH
displays a catabolic role; meanwhile NADP-GDH is a biosynthetic enzyme involved in ammonia
assimilation. Dual NAD(P)-GDHs are found in animals and higher plants, but NADP-GDHs are mostly
involved in ammonia assimilation in bacteria, fungi and algae [56-59]. Due to their ubiquous presence in
all the domains, we can found very interesting GDHs adapted to each medium including extreme ones
occupied by thermophiles, halophiles, acidophiles, etc .
Hyperthermophilic GDHs from the archaea Pyrococcus furiosus, Thermococcus litoralis,
Thermococcus profundus and Sulfolobus solfataricus are hexameric, as in mesophilic bacteria, and show
a high temperature optimum for activity [60-64]. However, the coenzyme specificity of GDH from these
hyperthermophiles resembles that of the eukaryotes rather than bacteria, since the enzyme uses both
NADH and NADPH [64]. Halophilic GDHs also displayed a high thermostability. This thermophilic
nature of halophilic enzymes has been noted in several systems [22,40,65]. Dym et al. [65] pointed out,
taking into account the crystal structure of Har. marismortui malate dehydrogenase, several of the
©FORMATEX 2007
179
Communicating Current Research and Educational Topics and Trends in Applied Microbiology
A. Méndez-Vilas (Ed.)
_____________________________________________________________________
structural features concerning halophilicity are the same as those which contribute to the stability of
thermophilic enzymes. The amino acid composition for halophilic GDHs showed a higher cumulative
amount of acidic amino acids than basic amino acids, similar to other halophilic proteins [22, 40, 65].
Hfx. mediterranei has at least two different GDHs. The NADP-GDH (EC 1.4.1.4) is an hexameric
enzyme composed of monomers with a molecular mass of 55 kDa and its activity was markedly
dependent on the salt concentration, being optimal at 1.5 M NaCl or KCl at 60 °C and pH 8.5 [58]. It
also displayed a high thermostability, requiring temperatures above 60 °C for measurable rates of
thermodenaturation. Another GDH, NAD-dependent (EC 1.4.1.2), displays a different behaviour that
may be related to the different and essential roles that GDHs should play in the metabolism of organisms
[61, 55], including extreme halophilic Archaea; although it still “raising questions” as to the role of the
distinct GDH proteins, as recently reported for Hbt. salinarum, which had at least four putative GDH
genes in its genome [66].
Bonete and col., [57,58,67,68] have isolated and characterised glutamate dehydrogenases from Hbt.
salinarum, Hfx. mediterranei, and Salinibacter ruber. These enzymes do not present dual-coenzyme
+
+
specifity, they are specific for either NAD or NADP . Hbt. salinarum (formerly Hbt. halobium) has two
different GDHs. A catabolic role is assigned to the NAD-GDH whereas the NADP-GDH may catalyse in
vivo the formation of glutamate from ammonium and 2-oxoglutarate [57,67]. Both of them displayed a
very high optimum temperature (70 ºC). GDHs from Hbt. salinarum are also remarkably resistant to
thermal denaturation, reflecting the thermophilic character of these enzymes, a characteristic also found in
other dehydrogenases from haloarchaea [22,40,58,69,70]. An important difference between these two
GDHs is the regulation of catalytic activity, NAD-GDH is stimulated in vitro by L-amino acids [22,,67].
Besides, a variety of metabolites act as inhibitors of the oxidative deamination of L-glutamate by NADGDH from Hbt. salinarum, such as TCA metabolites [71]. Other halophilic organisms also have two or
more GDHs, which differ in coenzyme specificity [72].
Protein sequence analysis of thermophilic GDHs indicates that the enzymes from P. furiosus and
Pyrococcus horikoshii are more closely related to the mesophilic bacterial GDH than to the thermophilic
archaeon S. solfataricus or to the halophilic archaeon Hbt. salinarum [64], suggesting that lateral gene
transfer between bacterial and archaeal gdhA genes may have occurred during evolution.
3.2.2 Glutamine synthetase/ Glutamate synthase
The ammonia assimilation pathway that implies glutamine synthetase (GS; EC 6.3.1.2) and glutamate
synthase (GOGAT; EC 1.4.7.1), also well-known as route GS/GOGAT, is of crucial importance since
the products L-glutamine and L-glutamate play a key role as nitrogen donors in the biosynthetic
reactions. The GS/GOGAT pathway is particularly important because it allows ammonia assimilation
into L-Glu at low intracellular ammonia concentrations and it seems that efficiently substitutes the other
glutamate biosynthetic reaction (that of glutamate dehydrogenase, GDH; EC 1.4.1.2) in these conditions.
Nevertheless, in some bacteria it has been found that GDH is active under low ammonium conditions
[73]. Glutamine synthetase, the first enzyme of the GS/GOGAT pathway, was first purified and
characterised from plants, but in the last 20 years, it has been extensively studied at both biochemical and
molecular levels in bacteria [74], cyanobacteria, yeast, mammals and some methanogenic and
hyperthermophilic archaea [75]. Three types of GS have been described based on the molecular size ant
the number of subunits. GSI is a dodecameric protein with subunit ranging between 44 and 60 kDa, and
is usually found among several groups of bacteria and archaea [76,77]. GSII is common in eukaryotes
and a few soil-dwelling bacteria as an octameric enzyme composed of subunits ranging between 35 and
50 kDa [78]. Finally, GSIII has been characterised from cyanobacteria and two anaerobic bacteria as a
hexameric enzyme composed of a 75 kDa subunits [79].
Although GS from Bacteria and Eukarya has been studied so far, only a few analysis have been carried
out with GS from members of Archaea, and just two of them correspond to a GS from haloarchaea [
80,81]. GSs from Hfx. mediterranei and Hbt. salinarum are octamers belonging to the GS type II.
However, a few GSs described from methanogenic or hyperthermophilic archaea are dodecamers of
180
©FORMATEX 2007
Communicating Current Research and Educational Topics and Trends in Applied Microbiology
A. Méndez-Vilas (Ed.)
_____________________________________________________________________
about 600 kDa, as it has been expected from bacterial GS. So, GS from haloarchaea exhibits typical
properties of GS from eukaryotes and soil bacteria species. This fact supports the hypothesis that some
members of Archaea are quite liked to eukaryotic organisms because both of them share similar
properties at physiological and metabolic levels [82]. The results obtained from Hfx.mediterranei suggest
that GS from this haloarchaea could allow the assimilation of the ammonium produced by assimilatory
nitrite reductase [81,83], while GDH would allow the assimilation of ammonium when this nitrogen
source is present in the culture media at high concentration.
The typical bacterial glutamate synthase is NADPH dependent and it has two dissimilar enzyme
subunits α and β that forms the αβ active protomer that contains one FAD, one FMN, and three different
[Fe-S] centres: one [3Fe-4S] cluster and two [4Fe-4S] centres [84]. The plant type glutamate synthase is
dependent on reduced ferredoxin as physiological electron donor and it is formed by a single polypeptide
chain similar in sequence and cofactor content (one FMN, and one [3Fe-4S] cluster) to de α subunit of
the bacterial enzyme. The eukaryotic pyridine nucleotide dependent glutamate synthase has been purified
from yeast, lupin seeds and nodules, alfalfa (Medicago sativa) and B. mori. It uses NADH as reductant
and it is a single polypeptide chain formed by an N-terminal region similar to the bacterial α and a Cterminal region similar to the bacterial β subunit [84].
Whether Archaea contain a bacterial type, a plant type or a forth class stills needs to be verified.
Methanococcus jannaschii genome sequencing first revealed the presence of one ORF encoding a 490
residues polypeptide that appears to be formed by an N-terminal domain containing the Cys-signature
typical of two bacterial ferredoxins [4Fe–4S] clusters followed by a polypeptide mapping on the
synthase domain of NADPH-GltS α subunit [84]. A similar ORF has been found in Archeoglobus
fulgidus and appears to be conserved in other Archaea and in Thermatogales as the result of lateral gene
transfer. It has also been proposed that a fifth type of glutamate synthase may exist. An ORF encoding a
50 kDa protein of Pyrococcus sp. KOD1 with significant sequence similarity to NADPH-GltS β subunit
has been found and it was assigned the gene name of gltA [84]. The analysis of halophilic archaea
genome sequenced revealed that in Har.marismortui, Natronomonas pharaonis and Haloquadratum
walsbyi there are an open reading frame of about 1500 residues similar in sequence to the plant type
ferrodoxin dependent glutamate synthase and to the bacterial NADPH glutamate synthase α subunit, but
there are not evidence of the presence of an open reading frame corresponding to a polypeptide similar to
the bacterial NADPH glutamate synthase β subunit. In Hfx. mediterranei we have measured activity of
glutamate synthase with methylviologen as reducing agent in extracts from cells grown in ammonium
starvation (data not published) and the sequencing of the gene is carried on showing high homology with
the halophilic archaea glutamate synthase sequenced but, at the moment, the classification of Archaea
glutamate synthase and particularly of halophilic archaea glutamate synthase stills needs to be verified
by characterizing the enzyme from organisms of this class. These preliminary studies shed light on the
possibility of an active GS/GOGAT cycle in Hfx. mediterranei under ammonium starvation.
4. Nitrate respiration
Although haloarchaea generally grow heterotrophically under aerobic conditions in hypersaline
environments, they possess facultative anaerobic capabilities [85,86]. These anaerobic capabilities are
important since the high salt concentrations and elevated temperatures the organisms encounter, together
with high cell densities promoted by aerobic growth and flotation, reduce the availability of molecular
oxygen. Although haloarchaeal microorganisms frequently encounter microaerobic or even anoxic
conditions, detailed knowledge regarding the extent of haloarchaeal anaerobic growth is still only
beginning to emerge.
Acknowledgements The support by Ministerio de Educación number projects PB95-0695, PB98-0969, BIO200203179, BIO2005-08991-C02-01 are gratefully acknowledged.
©FORMATEX 2007
181
Communicating Current Research and Educational Topics and Trends in Applied Microbiology
A. Méndez-Vilas (Ed.)
_____________________________________________________________________
References
[1] W.D. Grant Philosophical Transactions Royal Society London B 359, 1249 (2004)
[2] A.Oren FEMS Microbiology Ecology 18, 75 (1995).
[3] J. A. Serrano, M. Camacho and M. J. Bonete, FEBS Letters 434, 13 (1998).
[4] M. Vihinen and P. Mäntsälä, Critical Reviews in Biochemistry and Molecular Biology 24, 329 (1989).
[5] M. van der Maarel, B. van der Veen, J. Uitdehaag, H. Leemhuis and L. Dijkhuizen, Journal of Biotechnology 94, 137
(2002).
[6] B. Henrissat, Biochemical Journal 280, 309 (1991).
[7] T. Kuriki and T. Imanaka, Journal of Bioscience and Bioengineering 87, 557 (1999).
[8] S. Janeĉek, B. Svensson and E. A. MacGregor, European Journal of Biochemistry 270, 635 (2003).
[9] F. Pérez-Pomares, V. Bautista, J. Ferrer, C. Pire, F. C. Marhuenda-Egea and M. J. Bonete, Extremophiles 7, 299 (2003).
[10] W. A. Good and P. A. Hartman, Journal of Bacteriology 104, 601 (1970).
[11] T. Kobayashi, H. Kanai, T.Hayashi, T. Akiba, R. Akaboshi,and K. Horikoshi Journal of Bacteriology, 174, 3439
(1992).
[12] M.J.Coronado, C. Vargas, E. Mellado, G.Tegos, C.Drainas, J.J.Nieto, and A.Ventosa Microbiology 146, 861 (2000).
[13] J. E. Nielsen and T. V. Borchert, Biochimica et Biophysica Acta 1543, 253 (2000).
[14] H. Sakuraba and T. Ohshima, Journal of Bioscience and Bioengineering 93 (5), 441 (2002)
[15] C. H. Verhees, S. W. M. Kengen, J. E. Tuininga, G. J. Schut, M. W. W. Adams, W. M. de Vos and J. van der Oost,
Biochemical Journal 375, 231 (2003).
[16] H. J. van de Werken, C. H. Verhees, J. Akerboom, W. M. de Vos and J. van der Oost, FEMS Microbiology Letters
260, 69 (2006).
[17] B. Siebers and P. Schönheit, Current Opinion in Microbiology 8, 695 (2005).
[18] M. DeRosa, A. Gambacorta, B. Nicolaus, P. Giardina, E. Poerio and V. Buonocore, Biochemical Journal 224, 407
(1984).
[19] N. Budgen and M. J. Danson, FEBS Letter 196, 207 (1986).
[20] M. J. Danson, in M. Kates, D. J. Kushner and A. T. Matheson (eds.) Elsevier, Amsterdam, pp 1-24 (1993).
[21] F. Rodríguez-Valera, G. Juez and D. J. Kushner, Systematic and Applied Microbiology 4, 369 (1983).
[22] M. J. Bonete, C. Pire, F. I. Llorca, and M. L. Camacho, FEBS Letter 383, 227 (1996).
[23] O. Dym, M. Mevarech and J.L. Sussman, Science 267, 1344 (1995)
[24] C. Pire, J. Esclapez, J. Ferrer and M. J. Bonete, FEMS Letter 200, 221 (2001).
[25] K. L. Britton, P. J. Baker, M. Fisher, S. Ruzheinikov, D. J. Gilmour, M. J. Bonete, J. Ferrer, C. Pire, J. Esclapez and
D. W. Rice, Proceedings of the National Academy of Sciences of the United States of America 103, 4846 (2006).
[26] A. Zaigler, S.C. Schuster, and J. Soppa, Molecular Microbiology 48, 1089 (2003)
[27] C. Bräsen and P. Schönheit, FEMS Microbiology Letters 241, 21 (2004)
[28] C. Bräsen and P. Schönheit, Archives of Microbiology 175, 360 (2001)
[29] M. J. Danson and D. W. Hough, The Archaebacteria: Biochemistry and Biotechnology, edited by Danson, Hough and
Lunt, (Portland Press, London,1991) pp. 7-21.
[30] M.J. Danson Biochem. Soc. Trans. 16, 87 (1988)
[31] , M. J. Danson, R. Eisenthal, S. Hall, S. R. Kessell and D. L. Williams, Biochemical Journal. 218, 811 (1984)
[32] O. Kandler and K. O. Stetter, Zentrabl. Bakteriol. Hyg. Abt. 1 Orig. C2, 111 (1981)
[33] I. Ekiel, I. C. P. Smith and G. D. Sprott, Journal of. Bacteriology 156, 316 (1983).
[34] R. D. Chen and P. Gadal, Plant Physiol. Biochem. 28, 411 (1990).
[35] S. Gálvez and P. Gadal, Plant Scicence 105, 1 (1995).
[36] I. H. Steen, M. S. Madsen, N.-K. Birkeland and T. Lien, FEMS Microbiology Letters 160, 75 (1998).
[37] A. J. Lloyd and P. D. J. Weitzman, Biochem. Soc. Trans. 16, 411 (1988).
[38] I. H. Steen, T. Lien and N.-K. Birkeland, Archives in. Microbiology. 168, 412 (1997).
[39] M. L. Camacho, R. A. Brown, M.-J. Bonete, M. J. Danson and D. W. Hough, FEMS Microbiology Letters 134, 85
(1995).
[40] M. Camacho, A. Rodríguez-Arnedo and M.-J. Bonete, FEMS Microbiology Letters 209, 155 (2002).
[41] R. Chen, A. Greer and A. M. Dean, Proceedings of the National Academy of Sciences of the United States of America
93, 12171(1996).
[42] D. Madern, M. Camacho, A. Rodríguez-Arnedo, MJ. Bonete and G. Zaccai, Extremophiles 8, 377 (2004).
[43] A. Rodríguez-Arnedo, M. Camacho, F. Llorca and M.-J. Bonete, The Protein Journal 24, 1572 (2005).
[44] H. L. Kornberg, Biochemical Journal 99, 1 (1966).
[45] E. M. T. El-Mansi, H. G. Nimmo and W. H. Holms, Journal of General Microbiology 132, 797 (1986).
[46] A. Oren and P. Gurevich, FEMS Microbiology Letters 130, 91 (1995).
[47] J. A. Serrano and M. J. Bonete, Biochimica et Biophysica Acta-Gene Structure and Expression 1520, 154 (2001).
182
©FORMATEX 2007
Communicating Current Research and Educational Topics and Trends in Applied Microbiology
A. Méndez-Vilas (Ed.)
_____________________________________________________________________
[48] M.J. Bonete, F.C. Marhuenda-Egea, C. Pire, J. Ferrer and R.M. Martínez-Espinosa (2004) Nitrate Assimilation in
Halophilic Archaea.ncia: In “Halophilic Microorganisms”. Pág: 193-203. (A. Ventosa ed.)Springer-Verlag. Berlín,
Heidelberg, New York.
[49] B. Lledó, F.C. Marhuenda-Egea, R.M. Martínez-Espinosa and M.J. Bonete Gene 361, 80, (2005).
[50] R.M. Martínez-Espinosa; F.C. Marhuenda-Egea y M.J. Bonete FEMS Microbiology Letters 204, 381, (2001).
[51] R.M. Martínez-Espinosa, F.C. Marhuenda-Egea, A. Donaire and M.J. Bonete Biochimica et Biophysica Acta, 1623,
47, (2003).
[52] R.M. Martínez-Espinosa; F.C. Marhuenda-Egea y M.J. Bonete. FEMS Microbiology Letters 196, 113, (2001).
[53] R.M. Martínez-Espinosa, B. Lledó, F.C. Marhuenda-Egea and M.J. Bonete The effect of ammonium on assimilatory
nitrate reduction in the haloarchaeon Hfx. mediterranei. Extremophiles. In press. (2007)
[54] R.M. Martínez-Espinosa; J. Esclapez; V. Bautista and M.J. Bonete FEMS Microbiology Letters 264, 110, (2006)
[55] L.E. Smith, B.M. Austen, K.M. Blumental and F.J. Nyc, The Enzymes (vol. XI, 3rd edn). Edited by
P.D.Boyer(Academic Press, New york, 1975) pp. 293.
[56] F.M. Veronese, J.F. Nyc, Y. Degani, D.M. Brown and E.L. Smith, Jounal of Biological Chemistry 249, 7922 (1974).
[57] M.J. Bonete, M.L. Camacho and E. Cadenas, International Journal of Biochemistry 25, 1149 (1987).
[58] J. Ferrer, F. Pérez-Pomares and M. J. Bonete, FEMS Microbiology Letters. 141, 59 (1996).
[59] J. R. Ruiz, J. Ferrer, M. Camacho and M. J. Bonete, FEMS Microbiology Letters 159, 15 (1998).
[60] K. S. P. Yip, P. J. Baker, K. Britton, P. C. Engel, D. W. Rice, S.E. Sedelnikova, T. J. Stillman, A. Pasquo, R.
Chiaraluce, V. Consalvi and R. Scandurra, Acta-Crystallographica-Section D-Biological Crystallography, 51, 240
(1995).
[61] M. W. Bhuiya, H. Sakuraba, C. Kujo, N. Nunoura-Kominato, Y. Kawarabayasi, H. Kikuchi and T. Ohshima.,
Extremophiles 4, 333 (2000).
[62] I. Helianti, Y. Morita, Y. Murakami, K. Yokoyama and E. Tamiya, Applied Microbiology and Biotechnology 59, 462
(2001).
[63] S. Wang, Y. Feng, Z. Zhang, B. Zheng, N. Li, S. Cao, I. Matsui and Y. Kosugi, Archives of Biochemistry and
Biophysics, 411, 56 (2003).
[64] J. DiRuggiero, and F.T. Robb. Advances in Protein Chemistry, 48, 311 (1996).
[65] O. Dym, M. Mevarech and J.L. Sussman, Science, 267, 1344 (1995).
[66] L.M. Ingoldsby, K.F.Geoghegan, M.H. Bronagh, and P.C. Engel, Gene, 349, 237 (2005).
[67] M.J.Bonete, M.L. Camacho and E. Cadenas, International Journal of Biochemistry 18, 785, (1986).
[68] M.J.Bonete, F. Perez-Pomares, S. Díaz, ,J. Ferrer. and A. Oren,. FEMS Microbiology Letters, 226, 181, (2003).
[69] M.J.Bonete, J. Ferrer, C. Pire, M. Penades and J.L.Ruiz, Biochimie, 82, 1143 (2000).
[70] M.J.Danson, R. Eisenthal, S. Hall, S.R. Kessell. and D.L.Williams, Biochemical Journal 218, 811 (1984).
[71] M.J. Bonete, F. Perez-Pomares, J. Ferrer and M.L. Camacho. Biochimica et Biophysica Acta 1289 14 (1996)
[72] S.Díaz, F- Pérez-Pomares, C. Pire, J. Ferrer and M.J.Bonete. Extremophiles, 10: 105-115 (2006).
[73] H. Ertan, Archives of Microbiology 158, 42 (1992)
[74] L. Reitzer, Annual Review of Microbiology 57, 155 (2003)
[75] R.N. Adul Rahman, B. Jongsareejit, S. Fujiwara, T. Imanaka, Applied Environmental. Microbiology 63, 2472 (1997)
[76] D.L. Robertson, R.S. Alberte, Plant Physiology 111, 1169 (1996)
[77] J.R. Brown, Y. Masuchi, F.T. Robb, W.F. Doolittle, Journal of Molecular Evolution 38, 566 (1994)
[78] Y. Kumada, D.R. Benson, D. Hillemann, T.J. Hosted, D.A. Rocheford, C.J. Thompson, W. Wohlleben, Y. Tateno,
Proceedings of the National Academy of Sciences of the United States of America 90, 3009 (1993)
[79] J.C. Reyes and F.J. Florencio, Journal of Bacteriology 176, 1260 (1994)
[80] B. Manitz and A.W. Holldorf, Archives of Microbiology 159, 90 (1993)
[81] R.M. Martínez-Espinosa, J. Esclapez, V. Bautista, and M.J. Bonete, FEMS Microbiology Letters 264, 110 (2006)
[82] C.R. Woose, Proceedings of the National Academy of Sciences of the United States of America 99, 8742 (2002)
[83] R.M. Martínez-Espinosa, B. Lledó, F.C. Marhuenda-Egea, and M.J. Bonete, Extremophiles, 792 (2007) DOI
10.1007/s00792-007-0095-9
[84] M.A. Vanoni, and B. Curti, Archives of Biochemistry and Biophysics 433, 193 (2005)
[85] S. DasSarma and P. Arora. Halophiles, p. 458–466. In Encyclopedia of life sciences, vol. 8. Nature Publishing Group,
London, United Kingdom. (2002)
[86] B. Lledó, R.M. Martínez-Espinosa, F.C. Marhuenda-Egea and M.J. Bonete . Biochimica et Biophysica Acta 1674, 50
(2004).
©FORMATEX 2007
183