ArticlePDF Available

Crystallization and preliminary X-ray analysis of binary and ternary complexes of Haloferax mediterranei glucose dehydrogenase

Authors:

Abstract and Figures

Haloferax mediterranei glucose dehydrogenase (EC 1.1.1.47) belongs to the medium-chain alcohol dehydrogenase superfamily and requires zinc for catalysis. In the majority of these family members, the catalytic zinc is tetrahedrally coordinated by the side chains of a cysteine, a histidine, a cysteine or glutamate and a water molecule. In H. mediterranei glucose dehydrogenase, sequence analysis indicates that the zinc coordination is different, with the invariant cysteine replaced by an aspartate residue. In order to analyse the significance of this replacement and to contribute to an understanding of the role of the metal ion in catalysis, a range of binary and ternary complexes of the wild-type and a D38C mutant protein have been crystallized. For most of the complexes, crystals belonging to space group I222 were obtained using sodium/potassium citrate as a precipitant. However, for the binary and non-productive ternary complexes with NADPH/Zn, it was necessary to replace the citrate with 2-methyl-2,4-pentanediol. Despite the radical change in conditions, the crystals thus formed were isomorphous.
Content may be subject to copyright.
crystallization communications
Acta Cryst. (2005). F61, 743–746 doi:10.1107/S1744309105019949 743
Acta Crystallographica Section F
Structural Biology
and Crystallization
Communications
ISSN 1744-3091
Crystallization and preliminary X-ray analysis of
binary and ternary complexes of Haloferax
mediterranei glucose dehydrogenase
Julia Esclapez,
a
K. Linda Britton,
b
Patrick J. Baker,
b
Martin Fisher,
b
Carmen Pire,
a
Juan Ferrer,
a
Marı
´a Jose
´Bonete
a
and
David W. Rice
b
*
a
Departamento de Agroquı
´mica y Bioquı
´mica,
Facultad de Ciencias, Universidad de Alicante,
Ap. 99 Alicante 03080, Spain, and
b
Krebs
Institute for Biomolecular Research, Department
of Molecular Biology and Biotechnology,
The University of Sheffield, Sheffield S10 2TN,
England
Correspondence e-mail: d.rice@sheffield.ac.uk
Received 6 May 2005
Accepted 23 June 2005
Online 8 July 2005
Haloferax mediterranei glucose dehydrogenase (EC 1.1.1.47) belongs to the
medium-chain alcohol dehydrogenase superfamily and requires zinc for
catalysis. In the majority of these family members, the catalytic zinc is
tetrahedrally coordinated by the side chains of a cysteine, a histidine, a cysteine
or glutamate and a water molecule. In H. mediterranei glucose dehydrogenase,
sequence analysis indicates that the zinc coordination is different, with the
invariant cysteine replaced by an aspartate residue. In order to analyse the
significance of this replacement and to contribute to an understanding of the
role of the metal ion in catalysis, a range of binary and ternary complexes of the
wild-type and a D38C mutant protein have been crystallized. For most of the
complexes, crystals belonging to space group I222 were obtained using sodium/
potassium citrate as a precipitant. However, for the binary and non-productive
ternary complexes with NADPH/Zn, it was necessary to replace the citrate with
2-methyl-2,4-pentanediol. Despite the radical change in conditions, the crystals
thus formed were isomorphous.
1. Introduction
Extremely halophilic archaea are found in highly saline environments
such as natural salt lakes or saltern pools. These microorganisms
require between 2.5 and 5.2 Msalt for optimal growth (Kamekura,
1998) and they can balance the external salt concentration by accu-
mulating intracellular KCl close to saturation. The biochemical
machinery of these microorganisms has therefore been adapted in the
course of evolution to be able to function at salt concentrations at
which most biochemical systems cease to function.
Comparison of the amino-acid compositions and structures of
halophilic proteins and their mesophilic counterparts has shown that
a significant difference in the characteristics of the surface of halo-
philic proteins is an excess of acidic over basic residues (Lanyi, 1974;
Bo
¨hm & Jaenicke, 1994; Dym et al., 1995; Frolow et al., 1996; Britton et
al., 1998, 2005).
The extremely halophilic archaeon Haloferax mediterranei (ATCC
33500/R4) is able to grow in a minimal medium with glucose as the
sole carbon source, which is catabolized by a modified Enter–
Doudoroff pathway. Glucose dehydrogenase (GlcDH) catalyses the
first step in this pathway, the oxidation of -d-glucose to gluconic
acid, preferentially using NADP
+
as a coenzyme. Sequence analysis
has shown that GlcDH belongs to the zinc-dependent medium-chain
alcohol dehydrogenase (MDR) superfamily (Bonete et al., 1996; Pire
et al., 2001). Biochemical studies have established that the protein is
dimeric, with a subunit molecular weight of 39 kDa, and have
confirmed the requirement of zinc for catalysis. In previous work, the
crystallization of recombinant GlcDH in the presence of NADP
+
has
been reported under conditions which closely mimic those experi-
enced by the enzyme in the cell of the halophile (Ferrer et al., 2001);
more recently, the structure has been determined to 1.6 A
˚resolution
(Britton et al., 2005). This structure has revealed that the surface of
the enzyme is not only decorated with acidic residues, but also
displays a significant reduction in the fraction of exposed hydro-
phobic surface compared with non-halophilic glucose dehydro-
genases. This reduction in hydrophobic surface predominately arises
from the loss of the exposed alkyl component of lysine side chains as
a result of the reduction of lysine content in the enzyme (Britton et
#2005 International Union of Crystallography
All rights reserved
al., 2005). Moreover, genome comparisons have shown that there
appears to be a general reduction in the frequency of lysine in
halophilic proteins (Kennedy et al., 2001).
At the active site of the prototypical MDR superfamily member,
horse liver alcohol dehydrogenase (HLADH), the essential catalytic
zinc ion is coordinated by three protein ligands (Cys46, His67 and
Cys174), with the coordination shell of the zinc being completed by a
water molecule (Eklund et al., 1982). The mechanism of enzymes of
the MDR superfamily is commonly proposed to involve the exchange
of the zinc-bound water molecule with the hydroxyl of the substrate,
from which a proton is then removed by a base to generate an
alkoxide intermediate, which subsequently collapses to a carbonyl
with concomitant reduction of NAD(P)
+
. In this mechanism, the zinc
is proposed to remain tetrahedrally coordinated throughout (Eklund
et al., 1982; Ehrig et al., 1991; Ramaswamy et al., 1999). In a recent
alternative proposal, it has been suggested the zinc ion cycles
between different four- and five-coordinate intermediates, the iden-
tities of which are not yet clear (Makinen et al., 1983; Kleifeld et al.,
2003).
Of the three protein ligands to the zinc in enzymes of the MDR
superfamily, the first cysteine (Cys46 in HLADH) and the histidine
(His67 in HLADH) are very strongly conserved in the sequence. In
some family members, the second cysteine (Cys174 in HLADH) is
replaced by a glutamate [for example, Glu155 in Thermoplasma
acidophilum GlcDH (John et al., 1994) and Glu153 in rat sorbitol
dehydrogenase (Johansson et al., 2001)]. This sequence pattern is also
seen in the structure of H. mediterranei GlcDH, where a glutamate
residue (Glu64) occupies the equivalent position to the second
cysteine (Britton et al., 2005). However, the structure of H. medi-
terranei GlcDH has shown that there is an additional difference, with
an aspartate residue (Asp38) occurring in a structurally equivalent
position to the highly conserved first cysteine of this motif (Britton et
al., 2005). This is a sequence change rarely seen in the MDR family,
but that has also been observed in the sequence of other halophilic
glucose dehydrogenases [for example, the Halobacterium sp1 (Ng et
al., 2000) and Haloferax volcanii GlcDHs; http://zdna2.umbi.umd.edu]
and raises the intriguing question as to whether the presence of this
aspartate residue in the active site is a halophilic adaptation. In order
to investigate the role of the metal and its ligands in catalysis and
to enhance our understanding of the reaction mechanism in
H. mediterranei GlcDH, we have explored a range of crystallization
conditions of complexes of both the wild-type and the mutant D38C
enzymes. In this paper, we report the preliminary crystallographic
analysis of these various enzyme–substrate complexes.
2. Materials and methods
2.1. Site-directed mutagenesis, expression and purification
The gene encoding the halophilic GlcDH was cloned into pGEM-
11Zf(+) and site-directed mutagenesis was performed using the
GeneEditor in vitro site-directed mutagenesis system (Promega). The
protocol supplied with the kit was followed except that the length of
the DNA-denaturation stage was increased from 5 min at room
temperature to 20 min at 310 K. The expression, renaturation and
purification of the recombinant wild-type and mutant proteins were
performed as described previously (Pire et al., 2001). The purified
D38C GlcDH was dialyzed against 50 mMphosphate buffer pH 7.3
containing 2 MNaCl at 277 K overnight. Prior to crystallization,
protein samples were concentrated to approximately 20 mg ml
1
using a Vivaspin concentrator (30 kDa molecular-weight cutoff).
2.2. Crystallization and diffraction data collection
Crystals of the wild-type and D38C mutant of GlcDH grew after
4–6 d in the presence of different substrates using the hanging-drop
vapour-diffusion method, mixing small volumes (2–3 ml) of protein
sample with an equal volume of a precipitant solution at 290 K. For
the wild-type protein, crystals of the free enzyme and the binary
complex with NADP
+
, using sodium citrate as the precipitate, have
already been reported (Ferrer et al., 2001), but using these conditions
crystals of the binary complex with NADPH could not be obtained
either in the presence or absence of zinc. However, crystals of the
NADPH–Zn complex could be produced by preparing the protein
with 1 mMNADPH and 1 mMZnCl
2
and using 67–72%(v/v)
2-methyl-2,4-pentanediol (MPD) in 100 mMHEPES pH 7.5 as the
precipitant. The D38C mutant protein behaved in the same manner
as the wild type, with crystals of the apoenzyme and various NADP
+
complexes forming in the presence of citrate and the NADPH
complexes in the presence of MPD. For the D38C protein, the protein
sample was mixed with 1 mMNADP
+
or 1 mMNADPH, 1 mM
ZnCl
2
,10mMglucose and 10 mMgluconate as appropriate. For the
free enzyme and the complexes containing NADP
+
the precipitant
used was 1.4–1.6 Msodium citrate in 100 mMHEPES pH 7.0 and for
the complexes containing NADPH the precipitant was 62–72%(v/v)
MPD in 100 mMHEPES pH 7.5. Crystals of D38C GlcDH were also
grown in the presence of 2 MKCl, 1 mMZnCl
2
and 1 mMNADP
+
using 1.4–1.6 Mpotassium citrate as precipitant in 100 mMHEPES
buffer pH 7.0.
The crystals of the wild-type and D38C free enzyme grown in the
presence of sodium citrate showed a hexagonal bipyramidal
morphology (maximum dimensions 0.25 0.40 0.25 mm), whereas
crystals with a rod-like morphology (Fig. 1; maximum dimensions
0.6 0.6 0.4 mm) were obtained for all the wild-type and mutant
complexes with NADP
+
or NADPH, irrespective of whether the
precipitant was sodium citrate or MPD.
Crystals grown with sodium citrate as the precipitant were
mounted in X-ray-transparent capillaries. Preliminary data sets were
collected at 290 K by the rotation method with 1rotations per frame
using a MAR 345 detector, with double-mirror-focused Cu KX-rays
produced by a Rigaku RU-200 rotating-anode generator. The crystals
grown in the presence of MPD were flash-cooled in a cold nitrogen-
gas stream and data were collected using the same method at 100 K.
Data from the crystals of the D38C GlcDH–NADP
+
–Zn complex
crystallization communications
744 Esclapez et al. Glucose dehydrogenase Acta Cryst. (2005). F61, 743–746
Figure 1
A crystal of the binary complex of D38C GlcDH with NADPH and zinc.
grown in the presence of KCl and potassium citrate were collected at
the SRS Daresbury synchrotron on station PX14.2 at a wavelength of
0.97 A
˚with 1rotations using an ADSC Q4 detector. The data for
each crystal were processed and analysed using the HKL suite of
programs (Otwinowski & Minor, 1997) and subsequently handled
using the CCP4 suite (Collaborative Computational Project, Number
4, 1994).
3. Results and discussion
Analysis of the various data sets using the autoindexing routine of
DENZO showed that D38C GlcDH had crystallized in two different
forms, designated I and II, which are the same as those seen for the
wild-type enzyme (Ferrer et al., 2001). The crystals of the D38C free
enzyme (form I) belong to a hexagonal lattice P622, with unit-cell
parameters a=b= 89.4, c= 211.5 A
˚,== 90, = 120and a unit-
cell volume of 1.46 10
6
A
˚
3
. Given the GlcDH subunit molecular
weight of 39 kDa, the V
M
for a monomer in the asymmetric unit is
3.1 A
˚
3
Da
1
, which lies within the normal range for proteins
(Matthews, 1977). The crystals of the binary complexes and non-
productive ternary complexes (form II) were all isomorphous,
whether they were grown using citrate or MPD as the precipitant, and
diffracted to up to 1.5 A
˚resolution. These crystals belong to one of
the special pair of space groups I222 or I2
1
2
1
2
1
, with unit-cell para-
meters as shown in Table 1 and a unit-cell volume of 1.0 10
6
A
˚
3
.
The asymmetric unit appears to contain a monomer (V
M
=
3.2 A
˚
3
Da
1
). Data-collection statistics for the form I and form II
crystals are given in Table 1.
A cross-rotation function and translation function were calculated
in both I222 or I2
1
2
1
2
1
space groups on the form II data for the D38C
NADPH–Zn–glucose complex at a resolution of 20–3.0 A
˚, using a
single subunit of the wild-type H. mediterranei GlcDH as a search
model and the program AMoRe (Navaza, 1994). A clear translation-
function solution of correlation coefficient 67.6% and Rfactor 36.9%
was only seen in space group I222, indicating that this is the correct
space group. For the data from the form I crystals a similar process
was undertaken. In this case, a clear translation-function solution was
only seen in space group P6
2
22 (correlation coefficient 71.6%, R
factor 38.7%), identifying this as the correct space group.
Previously, we have obtained form II crystals from the wild-type
enzyme in complex with NADP
+
and also from the D38C mutant in
complex with NADP
+
and zinc (PDB codes 1ss0 and 1ss5, respec-
tively). All the D38C GlcDH complex crystals reported here are
isomorphous with the form II crystals. However, it is interesting to
note that the crystals of the NADP
+
complexes grow using citrate as
the precipitant, whereas those complexes containing NADPH only
grow when MPD is used as the precipitant. Despite these radically
different conditions, the crystals thus produced are isomorphous.
Analysis of these structures is now under way and given the
successful crystallization of this wide range of complexes of the wild-
type and mutant D38C GlcDH, we will be able to provide new
insights into the structure–function relationships of the enzyme and
perhaps shed light on the question of whether the aspartate residue at
position 38 in the H. mediterranei GlcDH is a halophilic adaptation or
purely serendipitous.
We thank the support staff at CCLRC Synchrotron Radiation
Source, Daresbury for assistance with station alignment. This work
was supported by the BBSRC, the Wellcome Trust, The Woolfson
Foundation and MCYT (BIO2002-03179) and by FPI fellowships
from Generalitat Valenciana (Spain) to JE. The Krebs Institute is a
designated BBSRC Biomolecular Sciences Centre and a member of
the North of England Structural Biology Centre.
References
Bo
¨hm, G. & Jaenicke, R. (1994). Protein Eng. 7, 213–220.
Bonete, M. J., Pire, C., Llorca, F. I. & Camacho, M. L. (1996). FEBS Lett. 383,
227–229.
Britton, K. L., Baker, P. J., Fisher, M., Ruzheinikov, S., Gilmour, J., Bonete,
M. J., Ferrer, J., Pire, C., Esclapez, J. & Rice, D. W. (2005). Submitted.
Britton, K. L., Stillman, T. J., Yip, K. S. P., Forterre, P., Engel, P. C. & Rice, D. W.
(1998). J. Biol. Chem. 273, 9023–9030.
Collaborative Computational Project, Number 4 (1994). Acta Cryst. D50, 760–
763.
Dym, O., Mevarech, M. & Sussman, J. L. (1995). Science,267, 1344–1346.
Ehrig, T., Hurley, T. D., Edenberg, H. J. & Bosron, W. F. (1991). Biochemistry,
30, 1062–1068.
Eklund, H., Plapp, B. V., Samama, J. P. & Branden, C. I. (1982). J. Biol. Chem.
257, 14349–14358.
Ferrer, J., Fisher, M., Burke, J., Sedelnikova, S. E., Baker, P. J., Gilmour, D. J.,
Bonete, M. J., Pire, C., Esclapez, J. & Rice, D. W. (2001). Acta Cryst. D57,
1887–1889.
Frolow, F., Harel, M., Sussman, J. L., Mevarech, M. & Sholman, M. (1996).
Nature Struct. Biol. 3, 452–458.
Johansson, K., El-Ahmad, M., Kaiser, C., Jo
¨rnvall, H., Eklund, H., Ho
¨o
¨g, J. O.
& Ramaswamy, S. (2001). Chem. Biol. Interact. 130–132, 351–358.
John, J., Crennell, S. J., Hough, D. W., Danson, M. J. & Taylor, G. L. (1994).
Structure,2, 385–393.
Kamekura, M. (1998). Extremophiles,2, 289–295.
crystallization communications
Acta Cryst. (2005). F61, 743–746 Esclapez et al. Glucose dehydrogenase 745
Table 1
X-ray data-collection statistics.
Values in parentheses refer to the highest resolution shell.
Protein sample Wild type D38C
Complex NADPH/Zn Free enzyme NADP
+
/Zn NADP
+
/Zn/gluconate NADPH/Zn/glucose NADPH/Zn NADPH
Source/wavelength Cu KCu KSRS† (0.97 A
˚)CuKCu KCu KCu K
Resolution (A
˚) 2.01 3.55 1.50 2.2 2.0 1.84 2.0
Highest resolution shell (A
˚) 2.08–2.01 3.63–3.55 1.55–1.50 2.27–2.22 2.05–2.0 1.89–1.84 2.05–2.00
Space group I222 P6
2
22 I222 I222 I222 I222 I222
Unit-cell parameters (A
˚)
a60.6 89.4 60.5 61.6 60.1 60.5 60.7
b107.7 89.4 109.3 112.2 106.3 108.9 108.1
c153.6 211.5 151.9 150.5 151.9 151.2 152.6
Unique reflections 32173 6269 66981 24766 30873 44001 32792
Completeness (%) 94.9 (82.5) 96.7 (100) 94.4 (93.5) 94.5 (92.9) 92.8 (89.6) 94.5 (91.5) 95.7 (99.3)
Multiplicity 3.9 (3.4) 8.74 (3.4) 5.1 (4.6) 9.34 (3.96) 8.03 (2.16) 7.1 (2.23) 5.4 (2.62)
hI/(I)i30.1 (3.7) 11.71 (2.4) 1.37 (1.01) 14.3 (3.3) 10.0 (1.8) 16.3 (1.8) 14.58 (2.41)
R
merge
0.059 (0.354) 0.10 (0.50) 0.051 (0.709) 0.10 (0.42) 0.095 (0.45) 0.047 (0.49) 0.092 (0.447)
SRS: CCLRC Daresbury Synchrotron Radiation Source. R
merge
=PhPijIðh;iÞhIðhÞij=PhPiIðh;iÞ.
Kennedy, S. P., Ng, W. V., Salzberg, S. L., Hood, L. & DasSharma, S. (2001).
Genome Res. 11, 1641–1650.
Kleifeld, O., Frenkel, A., Martin, J. M. & Sagi, I. (2003). Nature Struct. Biol. 10,
98–103.
Lanyi, J. K. (1974). Bacteriol. Rev. 38, 272–290.
Makinen, M. W., Maret, W. & Yim, M. B. (1983). Proc. Natl Acad. Sci. USA,80,
2584–2588.
Matthews, B. W. (1977). The Proteins, 3rd ed., Vol. 3, edited by H. Neurath &
R. Hill, pp. 468–477. New York: Academic Press.
Navaza, J. (1994) Acta Cryst. A50, 157–163.
Ng, W. V. et al. (2000). Proc. Natl Acad. Sci. USA,97, 12176–
12181.
Otwinowski, Z. & Minor, W. (1997). Methods Enzymol. 276, 307–
326.
Pire, C., Esclapez, J., Ferrer, J. & Bonete, M. J. (2001). FEMS Microbiol. Lett.
200, 221–227.
Ramaswamy, S., Park, D. H. & Plapp, B. V. (1999). Biochemistry,38, 13951–
13959.
crystallization communications
746 Esclapez et al. Glucose dehydrogenase Acta Cryst. (2005). F61, 743–746
... These changes arise as a direct consequence of linked movements of the zinc ion, a zinc-bound bound water molecule, and the substrate during progression through the reaction. These results provide evidence for the molecular basis of proton traffic during catalysis, a structural explanation for pentacoordinate zinc ion intermediates, and a unifying view for the observed patterns of metal ligation in the MDR family (Esclapez et al., 2005;Baker et al., 2009). ...
... Although the D38C GlcDH is less active, it has been crystallized in the presence of several combinations of products and substrates. This fact has allowed us to describe many aspects of the mechanism of the zinc-dependent MDR superfamily (Esclapez et al., 2005;Baker et al., 2009). ...
... volcanii, and Sul. solfataricus (SsoGDH-1), have been reported (265)(266)(267)(268)(269)(270)(271)(272). GDHs belong to the medium-chain dehydrogenase/reductase (MDR) superfamily, displaying a two-domain organization comprised of a central nucleotide binding domain flanked at the N and C termini by protein regions, which together make up the catalytic domain (273) (Fig. 13A). ...
... The active site is located in a deep cleft between both domains. The catalytic domain contains one structural zinc as well as a catalytic zinc ion near the active site, as described for GDHs (see above) but with slightly modified coordination sites (268,269,272,350). Although only the apo struc-ture of AraDH has been reported, structural and sequence similarities as well as substrate modeling suggest a reaction mechanism similar to that described for GDH (see above) (269,272,350). ...
Chapter
IntroductionEvolution of metabolic pathwaysArchaeal metabolismGlycolytic pathwaysRegulation of carbohydrate metabolismConclusions
... volcanii, and Sul. solfataricus (SsoGDH-1), have been reported (265)(266)(267)(268)(269)(270)(271)(272). GDHs belong to the medium-chain dehydrogenase/reductase (MDR) superfamily, displaying a two-domain organization comprised of a central nucleotide binding domain flanked at the N and C termini by protein regions, which together make up the catalytic domain (273) (Fig. 13A). ...
... The active site is located in a deep cleft between both domains. The catalytic domain contains one structural zinc as well as a catalytic zinc ion near the active site, as described for GDHs (see above) but with slightly modified coordination sites (268,269,272,350). Although only the apo struc-ture of AraDH has been reported, structural and sequence similarities as well as substrate modeling suggest a reaction mechanism similar to that described for GDH (see above) (269,272,350). ...
Article
SUMMARY The metabolism of Archaea, the third domain of life, resembles in its complexity those of Bacteria and lower Eukarya. However, this metabolic complexity in Archaea is accompanied by the absence of many "classical" pathways, particularly in central carbohydrate metabolism. Instead, Archaea are characterized by the presence of unique, modified variants of classical pathways such as the Embden-Meyerhof-Parnas (EMP) pathway and the Entner-Doudoroff (ED) pathway. The pentose phosphate pathway is only partly present (if at all), and pentose degradation also significantly differs from that known for bacterial model organisms. These modifications are accompanied by the invention of "new," unusual enzymes which cause fundamental consequences for the underlying regulatory principles, and classical allosteric regulation sites well established in Bacteria and Eukarya are lost. The aim of this review is to present the current understanding of central carbohydrate metabolic pathways and their regulation in Archaea. In order to give an overview of their complexity, pathway modifications are discussed with respect to unusual archaeal biocatalysts, their structural and mechanistic characteristics, and their regulatory properties in comparison to their classic counterparts from Bacteria and Eukarya. Furthermore, an overview focusing on hexose metabolic, i.e., glycolytic as well as gluconeogenic, pathways identified in archaeal model organisms is given. Their energy gain is discussed, and new insights into different levels of regulation that have been observed so far, including the transcript and protein levels (e.g., gene regulation, known transcription regulators, and posttranslational modification via reversible protein phosphorylation), are presented.
... The high protein concentrations obtained has allowed us to carry on crystallization assays. In particular , the glucose dehydrogenase from H. mediterranei has been crystallized by the hanging-drop method using sodium citrate as the precipitant [4,5]. ...
... substituted by a tyrosine residue (Tyr-50) in TgGDH (Fig. 5). Although Cys-39 is highly conserved among members of the MDR superfamily, 30) Tyr-50 is the characteristic of Thermoproteus GDHs. All of the residues of TgGDH described above are conserved in TtGDH. ...
Article
A hyperthermophilic archaeon was isolated from a terrestrial hot spring on Kodakara Island, Japan and designated as Thermoproteus sp. glucose dehydrogenase (GDH-1). Cell extracts from cells grown in medium supplemented with glucose exhibited NAD(P)-dependent glucose dehydrogenase activity. The enzyme (TgGDH) was purified and found to display a strict preference for d-glucose. The gene was cloned and expressed in Escherichia coli, resulting in the production of a soluble and active protein. Recombinant TgGDH displayed extremely high thermostability and an optimal temperature higher than 85 °C, in addition to its strict specificity for d-glucose. Despite its thermophilic nature, TgGDH still exhibited activity at 25 °C. We confirmed that the enzyme could be applied for glucose measurements at ambient temperatures, suggesting a potential of the enzyme for use in measurements in blood samples.
... The high protein concentrations obtained has allowed us to carry on crystallization assays. In particular , the glucose dehydrogenase from H. mediterranei has been crystallized by the hanging-drop method using sodium citrate as the precipitant [4,5]. ...
Article
Full-text available
who generously supported the meeting. Meeting
... This has been described as a mechanism of adaptation to hypersaline environments because the negative charges contributed to the solubility and stability of the proteins (Mevarech et al. 2000). The increased number of acidic amino acids results in a decrease in the content of Lys and aliphatic residues, and an increase in hydrophobic residues Gly, Ala and Val (Madern et al. 2000;Esclapez et al. 2005). ...
Article
A cyclodextrin glycosyltransferase (CGTase, EC 2.4.1.19) was successfully isolated and characterized from the halophilic archaeon Haloferax mediterranei. The enzyme is a monomer with a molecular mass of 77 kDa and optimum activity at 55°C, pH 7.5 and 1.5 M NaCl. The enzyme displayed many activities related to the degradation and transformation of starch. Cyclization was found to be the predominant activity, yielding a mixture of cyclodextrins, mainly α-CD, followed by hydrolysis and to a lesser extent coupling and disproportionation activities. Gene encoding H. mediterranei CGTase was cloned and heterologously overexpressed. Sequence analysis revealed an open reading frame of 2142 bp that encodes a protein of 713 amino acids. The amino acid sequence displayed high homology with those belonging to the α-amylase family. The CGTase is secreted to the extracellular medium by the Tat pathway. Upstream of the CGTase gene, four maltose ABC transporter genes have been sequenced (malE, malF, malG, malK). The expression of the CGTase gene yielded a fully active CGTase with similar kinetic behavior to the wild-type enzyme. The H. mediterranei CGTase is the first halophilic archaeal CGTase characterized, sequenced and expressed.
... Comparisons of the amino acid sequence of the halophilic GlcDH against the sequences of other glucose dehydrogenases belonging to the MDR superfamily, have shown an increase of acidic over basis residues in the halophilic enzyme, which has also been observed in other halophilic proteins. In addition, we have developed an heterologous overexpression system for the cloned gene [16] , which have allowed us to crystallize the protein under conditions that closely mimic those experienced by the enzyme in the cell of the halophile [17,18]. More recently, the structure of the Hfx. ...
Article
Generally, halophilic enzymes present a characteristic amino acid composition, showing an increase in the content of acidic residues and a decrease in the content of basic residues, particularly lysines. The latter decrease appears to be responsible for a reduction in the proportion of solvent-exposed hydrophobic surface. This role was investigated by site-directed mutagenesis of glucose dehydrogenase from Haloferax mediterranei, in which surface aspartic residues were changed to lysine residues. From the biochemical analysis of the mutant proteins, it is concluded that the replacement of the aspartic residues by lysines results in slightly less halotolerant proteins, although they retain the same enzymatic activities and kinetic parameters compared to the wild type enzyme.
Article
Full-text available
Despite being the subject of intensive investigations, many aspects of the mechanism of the zinc-dependent medium chain alcohol dehydrogenase (MDR) superfamily remain contentious. We have determined the high-resolution structures of a series of binary and ternary complexes of glucose dehydrogenase, an MDR enzyme from Haloferax mediterranei. In stark contrast to the textbook MDR mechanism in which the zinc ion is proposed to remain stationary and attached to a common set of protein ligands, analysis of these structures reveals that in each complex, there are dramatic differences in the nature of the zinc ligation. These changes arise as a direct consequence of linked movements of the zinc ion, a zinc-bound bound water molecule, and the substrate during progression through the reaction. These results provide evidence for the molecular basis of proton traffic during catalysis, a structural explanation for pentacoordinate zinc ion intermediates, a unifying view for the observed patterns of metal ligation in the MDR family, and highlight the importance of dynamic fluctuations at the metal center in changing the electrostatic potential in the active site, thereby influencing the proton traffic and hydride transfer events.
Article
Full-text available
The hyperthermophilic archaeon Sulfolobus solfataricus grows optimally above 80 degrees C and utilizes an unusual, promiscuous, non-phosphorylative Entner-Doudoroff pathway to metabolize both glucose and galactose. The first enzyme in this pathway, glucose dehydrogenase, catalyzes the oxidation of glucose to gluconate, but has been shown to have activity with a broad range of sugar substrates, including glucose, galactose, xylose, and L-arabinose, with a requirement for the glucose stereo configuration at the C2 and C3 positions. Here we report the crystal structure of the apo form of glucose dehydrogenase to a resolution of 1.8 A and a complex with its required cofactor, NADP+, to a resolution of 2.3 A. A T41A mutation was engineered to enable the trapping of substrate in the crystal. Complexes of the enzyme with D-glucose and D-xylose are presented to resolutions of 1.6 and 1.5 A, respectively, that provide evidence of selectivity for the beta-anomeric, pyranose form of the substrate, and indicate that this is the productive substrate form. The nature of the promiscuity of glucose dehydrogenase is also elucidated, and a physiological role for this enzyme in xylose metabolism is suggested. Finally, the structure suggests that the mechanism of sugar oxidation by this enzyme may be similar to that described for human sorbitol dehydrogenase.
Article
Full-text available
We report the complete sequence of an extreme halophile, Halobacterium sp. NRC-1, harboring a dynamic 2,571,010-bp genome containing 91 insertion sequences representing 12 families and organized into a large chromosome and 2 related minichromosomes. The Halobacterium NRC-1 genome codes for 2,630 predicted proteins, 36% of which are unrelated to any previously reported. Analysis of the genome sequence shows the presence of pathways for uptake and utilization of amino acids, active sodium-proton antiporter and potassium uptake systems, sophisticated photosensory and signal transduction pathways, and DNA replication, transcription, and translation systems resembling more complex eukaryotic organisms. Whole proteome comparisons show the definite archaeal nature of this halophile with additional similarities to the Gram-positive Bacillus subtilis and other bacteria. The ease of culturing Halobacterium and the availability of methods for its genetic manipulation in the laboratory, including construction of gene knockouts and replacements, indicate this halophile can serve as an excellent model system among the archaea.
Article
Full-text available
A homology-based modeling study on the extremely halophilic glutamate dehydrogenase from Halobacterium salinarum has been used to provide insights into the molecular basis of salt tolerance. The modeling reveals two significant differences in the characteristics of the surface of the halophilic enzyme that may contribute to its stability in high salt. The first of these is that the surface is decorated with acidic residues, a feature previously seen in structures of halophilic enzymes. The second is that the surface displays a significant reduction in exposed hydrophobic character. The latter arises not from a loss of surface-exposed hydrophobic residues, as has previously been proposed, but from a reduction in surface-exposed lysine residues. This is the first report of such an observation.
Article
Full-text available
The catalytic role of the active site metal-water complex in horse liver alcohol dehydrogenase (alcohol:NAD+ oxidoreductase, EC 1.1.1.1) is investigated on the basis of a comparative analysis of the pH dependence of steady-state kinetic parameters of the native and active-site-specific Co2+-reconstituted enzyme and on the basis of assignment of the coordination environment of the Co2+ by electron paramagnetic resonance methods. The pH dependence of the kinetic parameters for the oxidation of benzyl alcohol reveals two ionizations (pK1 approximately equal to 6.7; pK2 approximately equal to 10.6) that govern kcat and belong to the ternary enzyme-NAD+-alcohol complex and two ionizations (pK1' approximately equal to 7.5; pK2' approximately equal to 8.9) that govern kcat/Km and belong to the binary enzyme-NAD+ complex. The ionizations pK2 and pK2' decrease by 0.5-1 pK alpha unit upon replacement of the active site Zn2+ by Co2+. A similar metal ion dependence of pK2 and pK2' is observed for the oxidation of 2-propanol. We attribute these ionizations to a metal-bound water molecule. The zero-field splitting energy of the Co2+ in the binary enzyme-NADH complex and the ternary enzyme-NADH-CF3CH2OH complex is approximately equal to 22 cm-1, indicative of a pentacoordinate species. Binding of a water molecule to the metal ion as the fifth ligand in the ternary enzyme-NADH-CF3CH2OH complex is confirmed on the basis of magnetic interactions of H2(17)O with Co2+. The results indicate that the active site metal ion in catalytically competent ternary enzyme-coenzyme-substrate complexes is pentacoordinate and is ligated by a neutral water molecule in the physiological pH range. We suggest that the neutral metal-bound water molecule serves as the base catalyst for proton abstraction in alcohol oxidation.
Article
Full-text available
Haloarcula marismortui is an archaebacterium that flourishes in the world's saltiest body of water, the Dead Sea. The cytosol of this organism is a supersaturated salt solution in which proteins are soluble and active. The crystal structure of a 2Fe-2S ferredoxin from H. marismortui determined at 1.9 A is similar to those of plant-type 2Fe-2S ferredoxins of known structure, with two important distinctions. The entire surface of the protein is coated with acidic residues except for the vicinity of the iron-sulphur cluster, and there is an insertion of two amphipathic helices near the N-terminus. These form a separate hyperacidic domain whose postulated function to provide extra surface carboxylates for solvation. These data and the fact that bound surface water molecules have on the average 40% more hydrogen bonds than in a typical non-halophilic protein crystal structure support the notion that haloadaptation involves better water binding capacity.
Article
On the basis of the three-dimensional structure of horse liver alcohol dehydrogenase determined by X-ray crystallography, His 51 has been proposed to act as a general base during catalysis by abstracting a proton from the alcohol substrate. A hydrogen-bonding system (proton relay system) connecting the alcohol substrate and His 51 has been proposed to mediate proton transfer. We have mutated His 51 to Gln in the homologous human liver beta 1 beta 1 alcohol dehydrogenase isoenzyme which is expected to have a similar proton relay system. The mutation resulted in an about 6-fold drop in V/Kb (Vmax for ethanol oxidation divided by Km for ethanol) at pH 7.0 and a 12-fold drop at pH 6.5. V/Kb could be restored completely or partially by the presence of high concentrations of glycylglycine, glycine, and phosphate buffers. A Brønsted plot of the effect on V/Kb versus the pKa of these bases plus H2O and OH- was linear. Only secondary or tertiary amine buffers differed from linearity, presumably due to steric hindrance. These results suggest that His 51 acts as a general base catalyst during alcohol oxidation in the wild-type enzyme and can be functionally replaced in the mutant enzyme by general base catalysts present in the solvent. Steady-state kinetic constants for NAD+ and the trifluoroethanol inhibition patterns were similar between the wild-type and the mutant enzyme. Differences in the inhibition constants (Ki) of caprate and trifluoroethanol below pH 7.8 and in the pH dependence of Ki can be explained by the substitution of neutral Gln for positively charged His.
Article
Horse liver alcohol dehydrogenase was crystallized from an equilibrium mixture containing predominantly NAD+ and p-bromobenzyl alcohol. X-ray diffractometer data to a resolution of 2.9 A were collected and used to compute electron density maps with phases calculated from the isomorphous enzyme . NADH . dimethyl sulfoxide complex, which has been refined to an R value of 25.6%. The electron density maps were readily interpreted in a graphics display system. Both subunits of the dimer bind coenzyme and alcohol in essentially the same manner; there is no evidence of asymmetry between subunits. The bromophenyl group is accommodated in a large hydrophobic pocket that has the side chain of Leu-116 rotated into a different position than in the complex with dimethyl sulfoxide. The alcohol oxygen is directly ligated to the catalytic zinc atom. The zinc is tetracoordinate and there is no room for a water molecule to make the zinc pentacoordinate. A hydrogen-bonded system formed with the hydroxyl groups of the alcohol, Ser-48 and nicotinamide ribose (2'), and the imidazole of His-51 may provide a proton relay system that links the buried alcohol to solvent. The insertion of the coenzyme's hydroxyl group into this system appears to install the catalytically active species. The observed structure has the pro-R hydrogen on C1 of the alcohol pointing away from C4 of the nicotinamide ring. This is probably a nonproductive complex that easily becomes productive by a rapid rotation of the alcohol to put the pro-R hydrogen within 3 A of C4 of the nicotinamide ring and in position for a direct transfer of hydrogen. A model of the productive complex readily explains the stereospecificity of hydride transfer observed for ethanol.
Article
The archaea are a group of organisms distinct from bacteria and eukaryotes. Structures of proteins from archaea are of interest because they function in extreme environments and because structural studies may reveal evolutionary relationships between proteins. The enzyme glucose dehydrogenase from the thermophilic archaeon Thermoplasma acidophilum is of additional interest because it is involved in an unusual pathway of sugar metabolism. We have determined the crystal structure of this glucose dehydrogenase to 2.9 A resolution. The monomer comprises a central nucleotide-binding domain, common to other nucleotide-binding dehydrogenases, flanked by the catalytic domain. Unexpectedly, we observed significant structural homology between the catalytic domain of horse liver alcohol dehydrogenase and T. acidophilum glucose dehydrogenase. The structural homology between glucose dehydrogenase and alcohol dehydrogenase suggests an evolutionary relationship between these enzymes. The quaternary structure of glucose dehydrogenase may provide a model for other tetrameric alcohol/polyol dehydrogenases. The predicted mode of nucleotide binding provides a plausible explanation for the observed dual-cofactor specificity, the molecular basis of which can be tested by site-directed mutagenesis.
Article
‘Halophilic adaptation’ of proteins, i.e. the requirement for high concentrations of monovalent ions for thermodynamic stability of proteins from halophilic organisms, is not fully understood. In this work, an explanation for the halophilic behavior of dihydrofolate reductase (h-DHFR) from Halobacterium volcanii is attempted, based on a model structure derived from comparative modeling to dihydrofolate reductase from Escherichia coli. The model structure of h-DHFR shows an unique asymmetrical charge distribution over the protein surface, with positively charged amino acids centered around the active site and negative charges on the opposite side of the enzyme. This particular charge distribution and the correlated molecular dipole are functionally relevant. The negative charges on the surface form clusters which are shielded at high salt concentrations; at low salt, they repulse each other, thus destabilizing the protein. Results are in accordance with denaturation data and, thus, provide an explanation for the exceptional stability properties of h-DHFR.