Báo cáo khóa học: Modulation of activity of NADH oxidase from Thermus thermophilus through change in flexibility in the enzyme active site induced by Hofmeister series anions ppt
ModulationofactivityofNADHoxidase from
Thermus thermophilus
through changeinflexibilityintheenzymeactivesite induced
by Hofmeisterseries anions
Gabriel Z
ˇ
olda
´
k
1
, Mathias Sprinzl
2
and Erik Sedla
´
k
1
1
Department of Biochemistry, Faculty of Sciences, P. J. S
ˇ
afa
´
rik University, Kos
ˇ
ice, Slovakia;
2
Laboratorium fu
¨
r Biochemie,
Universita
¨
t Bayreuth, Germany
The conformational dynamics ofNADHoxidase from
Thermus thermophilus was modulated bythe Hofmeister
series ofanions (H
2
PO
4
–
, SO
4
2–
, CH
3
COO
–
, Cl
–
, Br
–
,I
–
,
ClO
4
–
, SCN
–
) inthe concentration range 0–3
M
. Both cha-
otropic and kosmotropic anions, at high concentration,
inhibit theenzymeby different mechanisms. Chaotropic
anions increase the apparent Michaelis constant and decre-
ase the activation barrier ofthe reaction. Kosmotropic ani-
ons have the opposite effect. Anionsfromthe middle of the
Hofmeister series do not significantly affect theenzyme acti-
vity even at high concentration. We detected no significant
changes in ellipticity ofthe aromatic region inthe presence
of theanions studied. There is a decreased Stern–Volmer
quenching constant for FAD fluorescence quenching in the
presence of kosmotropic anions and an increased quench-
ing constant inthe presence of chaotropic anions. All of
this indicates that activesiteflexibility is important in the
function ofthe enzyme. The data demonstrate that both the
high rigidity oftheactivesiteinthe presence of kosmotropic
anions, and its high flexibilityinthe presence of chaotropic
anions have a decelerating effect on enzyme activity. The
Hofmeister seriesofanions proved to be suitable agents for
altering enzymeactivitythrough changes inflexibilityof the
polypeptide chain, with potential importance in modulating
extremozyme activity at room temperature.
Keywords: activation; conformational dynamics; flavopro-
teins; NADH oxidase; Thermus thermophilus.
The native conformation of an enzyme is produced by the
complex interaction of van der Waals interactions, hydro-
gen bonds and ionic interactions. These interactions
produce stability oftheenzyme under physiological condi-
tions and prevent deleterious conformational changes from
perturbations inthe environment that would cause deacti-
vation. These interactions, however, must not result in
protein rigidity because theenzymeactivesite requires
flexibility for optimal catalytic function. The balance of
these two tendencies is sensitively adjusted for the physio-
logical conditions at which theenzyme works. Examples
of such adjustments are enzymes from hyperthermophiles
and psychrophiles which have optimal activity at high
(> 80 °C) and low (< 20 °C) temperatures, respectively
[1,2]. Enzymes from thermophiles are almost inactive at
room temperature because of polypeptide and side chain
rigidity inducedby higher-order interactions within secon-
dary and tertiary structures. Psychrophilic enzymes are
inactive at room temperature because the high flexibility of
their polypeptide and side chains results in partial/local or
complete unfolding ofthe tertiary structure. Modulation of
the balance between the rigidity and flexibilityof the
polypeptide and side chains can be achieved by changing the
solvent properties. Stabilization of psychrophilic enzymes
without affecting their activity, or activation of thermophilic
enzymes without affecting their stability, is interesting for
both basic and applied protein chemistry.
The use of chaotropic agents (urea, guanidinium
hydrochloride) to activate different enzymes has been
reported in several papers [3–8]. Thechangein activity
resulted from conformational changes inthe tertiary and
secondary structure ofthe enzymes studied. We have
shown recently that it is possible to activate NADH
oxidase fromThermusthermophilus with urea without
affecting the global stability oftheenzyme at room
temperature [8a]. NADHoxidase (EC 1.6.99.3) from
T. thermophilus is a dimeric flavoprotein containing one
molecule of FMN or FAD per 25 kDa monomer, which
catalyzes hydride transfer fromNADH to an acceptor
such as FAD, ferricyanide and oxygen [9]. It belongs to the
flavin reductase/nitroreductase family, which has similar
broad substrate specificity, fold and quaternary structure
[10,11]. Localization oftheactivesiteofNADHoxidase at
the edge ofthe dimeric interface (Fig. 1) is in agreement
with the fact that theactive sites of enzymes are usually the
most labile part oftheenzyme structure [12]. Perturbation
of either the static or dynamic state oftheactivesite may
lead to significantly changed activity. Previous studies in
our laboratory have indicated that activation of NADH
oxidase is not achieved by conformational change but is a
result ofthe increased dynamics ofthe polypeptide/side
chain intheenzymeactive site. To substantiate these
observations and analyze the role of dynamics in enzyme
Correspondence to E. Sedla
´
k, Department of Biochemistry, Faculty
of Sciences, P. J. S
ˇ
afa
´
rik University, Moyzesova 11, 041 54 Kos
ˇ
ice,
Slovakia. Fax: + 421 55 622 21 24, Tel.: + 421 55 622 35 82,
E-mail: sedlak_er@saske.sk
Enzyme: NADHoxidase (EC 1.6.99.3).
(Received 26 August 2003, revised 8 October 2003,
accepted 28 October 2003)
Eur. J. Biochem. 271, 48–57 (2004) Ó FEBS 2003 doi:10.1046/j.1432-1033.2003.03900.x
activity, we have investigated the effect ofthe Hofmeister
series ofanions on theactivityofNADHoxidase from
T. thermophilus.
The crystal structure provides information about the
flexibility of a given structure by comparison of temperature
B factors. Temperature B factors are atomic mean square
displacements obtained fromthe intensity ofthe diffractive
spots [13]. The absolute value ofthe B factor is dependent
on the refinement method and the conditions of crystalliza-
tion [14]. It is therefore only correct to compare B factors
within a particular structure, although such data must also
be handled with caution. Data fromthe crystals are
averaged over crystal space and time, therefore they reflect
crystal defects, static disorders and other parameters [15].
NADH oxidase has an overall low temperature factor for
the whole structure (% 23 A
˚
2
) [9] that is in accordance with
the high stability ofthe protein conformation (Fig. 1). The
flavin moiety, with a low B factor, indicates rigidity and
strong binding to the protein matrix. Trp47, the only
tryptophan residue located in close proximity to the flavin
cofactor (within 10 A
˚
), is almost parallel to the isoalloxazine
ring, but the elevated temperature factor indicates it has
high flexibility. The indole ring is stabilized through
hydrophobic interactions (side chains of Ala46, Leu49,
Phe120, Ala121, Ala122, Met123) from helix F. The crystal
structure of a homologous nitroreductase in various states
shows that binding ofthe substrate (nicotinic acid) is
accompanied bytheinduced fit of helix F and helix E [10].
Rearrangement of helix F during the binding event results in
a changeinthe torsional angle of several residues.
Remarkably, the residues that are involved in substrate
binding through changes in their dihedral angles are those
with the highest temperature factor, and are mostly from
helix F. Similarly, the high B factors of helix F indicate that
it is highly flexible inNADHoxidase (Fig. 1). Stabilization
or destabilization of this helix would affect interactions with
Trp47 and thus the opening oftheactive site, which is
necessary for activity (unpublished observation). This would
indicate a mechanism of interaction ofNADH substrate
with theenzyme common to this flavoenzyme family.
The Hofmeisterseriesofanions were chosen as suitable
candidates for stabilization/destabilization of this part of
NADH oxidase. There are numerous reports on the effect
of theHofmeisterseriesof salts on folding and stability of
proteins [16–18] and enzymeactivityin both aqueous
solutions [19–26] and organic solvents [27]. It is generally
accepted that the effect of these salts on protein results from
interactions ofthe salt with the polypeptide chain (enthalpic
contribution) and, indirectly, through effects on the water
structure (entropic contribution) [28–36]. For our study, we
chose theHofmeisterseriesof anions: H
2
PO
4
–
, SO
4
2–
,
CH
3
COO
–
, Cl
–
, Br
–
,I
–
, ClO
4
–
, SCN
–
(ordered from
kosmotropic to chaotropic). Anions are more efficient than
cations in affecting the properties of polypeptide chains. The
anion–water interaction is stronger than the cation–water
interaction, thus anions have a greater effect on water
ordering. The explanation for this is the asymmetry of the
charge in a water molecule, with the negative end of the
molecule’s dipole being nearer the center than the positive
end [34,36].
The modulationofthe conformational dynamics of the
enzyme bytheHofmeisteranions enabled us to show that
both stabilization and destabilization oftheactivesite of
NADH oxidaseby kosmotropic and chaotropic anions,
respectively, inhibits its activity. Application ofthe Hof-
meister seriesofanions may be a suitable approach to
modifying properties of enzymes from extremophiles. The
work presented is the result of a continuation of our interest
in understanding the role offlexibility for catalytic efficiency
of enzymes. NADHoxidasefrom T. thermophilus is a good
candidate for such a study because theflexibilityof its
polypeptide chain is adjusted to the harsh conditions of
thermophilic bacteria. Therefore, the addition of chaotropic
agents at room temperature will not significantly perturb
the enzyme’s global structure [8a] but will modulate the
flexibility of most of its labile parts, i.e. the part of the
protein structure where theactive sites are usually located
[9].
Experimental procedures
Analytical-grade biochemicals were obtained from Merck
(Germany). Urea (high purity grade) was purchased from
Sigma. Urea concentrations were determined from refract-
ive index measurements using an Abbe Refractometer AR3-
AR6. The pH values ofthe solutions were measured with a
Sensorex glass electrode before and after measurement at
room temperature. Only the measurements at which the pH
change was less than 0.2 pH unit were used.
Protein expression and purification
The NADHoxidasefrom T. thermophilus was overpro-
duced in Escherichia coli JM 108. The purification proce-
dure for the overproduced NADHoxidase was described
Fig. 1. Homodimeric structure ofNADHoxidasefrom T. thermophilus
colored according to temperature B factor. Low B factor (< 15 A
˚
2
rigid structure) is shown by a dark blue color, intermediate B factor
(30–45 A
˚
2
) by green/yellow, and high B factor (> 60 A
˚
2
) by red. Flavin
cofactor and the closest tryptophan, Trp47, are shown. The thin line
indicates dimeric interface. The isoalloxasine ring of flavin is localized
in the rigid part ofthe homodimer, and Trp47 is localized on the most
flexible a-helix ofthe protein structure, helix F (shown within elliptical
traces).
Ó FEBS 2003 Anion effect on activityofNADHoxidase (Eur. J. Biochem. 271)49
previously [37]. The final product provided a single band on
a SDS/polyacrylamide gel stained with Coomassie Brilliant
Blue. Before use, the protein was dialyzed against 5 m
M
phosphate buffer, pH 7.0, inthe absence of FAD. The
specific activityofNADHoxidase is % 1.9 UÆmg
)1
in 5 m
M
phosphate buffer, pH 7.0.
Steady-state kinetics
All kinetic measurements were performed on a Shimadzu
UV3000 spectrophotometer. The kinetic parameters were
determined fromthe initial decrease inNADH absorbance
at 340 nm (e
340nm
¼ 6220
M
)1
Æcm
)1
), at 20 °C. Measure-
ments were performed after incubation (12 h) in 120 n
M
NADH oxidase, 5 m
M
phosphate, pH 7.0, containing
0.12 m
M
FAD and different concentrations of salts. The
reaction was started bythe addition of 180 l
M
NADH. To
determine the K
m
value the concentration ofNADH was
varied inthe range 5–200 l
M
. It is not possible to use
NADH at higher concentrations because of its large
absorbance. The data were fitted to the Michaelis equation:
m ¼ V
max
½NADH
½NADHþK
NADH
m;app
ð1Þ
where, K
m, app
is the apparent Michaelis constant and the
apparent V
max
is the maximum velocity for the catalytic
reaction. The experimental data were also plotted according
to the Lineweaver–Burk equation and analyzed by linear
regression. Similar results were obtained using both meth-
ods. The apparent k
cat
was determined as V
max
/[E]
0
, where
[E]
0
is the total concentration ofNADHoxidasein solution.
Determination ofthe Michaelis–Menten parameters has
not been possible inthe presence of some concentrations of
iodine anions because of a spectral overlap of iodine
(product ofthe peroxide and iodide) and NADH. At high
concentrations of rhodanide, perchlorate, sulfate and phos-
phate, theactivityofNADHoxidase is very low and
determination ofthe Michaelis–Menten constants has large
errors.
Temperature dependence ofenzyme activity
Enzyme activity was determined in 5 m
M
phosphate buffer
containing 0.12 m
M
FAD and 120 n
M
NADH oxidase.
Reactions were started bythe addition ofNADH to achieve
a final concentration of 0.18 m
M
NADH. Initial velocities
were measured inthe range 20–40 °C. Temperature during
measurements was kept constant by temperature controlled
water circulation around the cuvette. Temperature depend-
encies were analyzed with a simple Arrhenius equation
lnk
cat
¼
E
a
RT
þ C
1
ð2Þ
where, R is the gas constant (8.314 JÆK
)1
Æmol
)1
), E
a
is the
activation energy for the observed reaction, and C
1
is a
temperature-independent constant. At least five values were
plotted as ln (k
cat
)vs.T
)1
and analyzed by linear regression.
Coefficients of linearity were typically higher than 0.98.
From comparison ofthe Arrhenius equation and the
transition state theory, the enthalpy (DH*) and entropy
(DS*) of activation were calculated
DH
Ã
¼ E
a
À RÁT ð3Þ
T Á ln
k
cat
T
¼
T Á DS
Ã
R
þ C
2
ð4Þ
C
2
is the temperature-independent constant.
Thefreeenergyofactivation(DG*) was calculated from
the equation:
DG
Ã
¼ DH
Ã
À TDS
Ã
ð5Þ
Fluorescence emission spectroscopy
The fluorescence steady-state measurements were per-
formed on a Shimadzu RF5000 spectrofluorophotometer.
The fluorescence spectrum of tryptophan residues was
obtained on excitation at 295 nm. The cuvette contained
5m
M
sodium phosphate, pH 7.0, with various concentra-
tions of salts and 2.4 l
M
dimeric protein in a total volume of
2.5 mL. Fluorescence measurements were performed at
20 °C. Temperature was kept constant (± 0.3 °C) by
temperature controlled water circulation.
Quenching of FAD fluorescence
The fluorophores inNADHoxidase make it possible to
perform fluorescence quenching experiments to investigate
the dynamics ofthe environment near the fluorophore and
the accessibility ofthe fluorophores to solvent. Tryptophan
moieties are widely used in quenching experiments. NADH
oxidase contains four tryptophans at different positions,
which complicates a detailed analysis. The flavin cofactor is
another fluorophore that could be used as an intrinsic probe
quenched by externally added quenchers, e.g. iodide and
rhodanide anions. The commonly used noncharged quen-
cher acrylamide is not an efficient quencher of FAD
fluorescence. The FAD fluorescence is not affected even at
relatively high (0.2
M
) concentrations of acrylamide. Fluor-
escence quenching ofthe FAD was performed using iodide
anions (KI). Stock solution (4
M
KI in 5 m
M
phosphate
buffer, pH 7.0) was freshly prepared to avoid oxidation of
iodide [38]. Sodium dithionite could not be used inthe stock
solution of KI (inhibition of iodine formation) because of
concomitant changes inthe redox state ofthe flavin. As it is
a single population ofthe FAD, it is possible to use a simple
Stern–Volmer equation:
F
0
F
¼ 1 þ K
sm
½KIð6Þ
where, K
SV
is the Stern–Volmer quenching constant.
Comparison ofthe values of K
SV
allows us to assess the
accessibility ofthe FAD cofactor and, indirectly, the
dynamics of its environment. Fluorescence was monitored
at 525 nm after excitation by 450 nm inthe absence (F
0
)and
presence of various concentrations of KI (F). The linearity
of the experimental data (coefficient of linearity r % 0.99)
confirms the validity ofthe simple model (Eqn 6).
CD measurements
CD measurements were performed on a Jasco J-810
spectropolarimeter (Jasco, Tokyo, Japan) at 20 °Cwith
50 G. Z
ˇ
olda
´
k et al.(Eur. J. Biochem. 271) Ó FEBS 2003
27.4 l
M
NADH oxidasein 50 m
M
sodium phosphate,
pH 7.0, and at different concentrations of salt. A 1 cm path-
length cuvette was used for the aromatic region. Each
spectrum was an accumulation of 10 consecutive scans.
Results
The parameters characterizing theactivityof NADH
oxidase, i.e. the apparent rate constant, k
cat
, and the apparent
Michaelis constant, K
m
, strongly depend on the ionic
strength ofthe solution. Increasing the ionic strength from
5m
M
to 50 m
M
potassium phosphate results in a sixfold
increase inthe k
cat
value, from 1.1 to 6.6 s
)1
, and a slight
decrease inthe K
m
value, from 8.5 to 5.2 l
M
. In Table 1, it
can be seen that NADHoxidase is nonspecifically activated
by increased ionic strength, as all the salts studied at 0.5
M
induced an increase inthe k
cat
value ofthe enzyme. However,
a further increase in ionic strength enabled us to distinguish
the effects ofthe different anions. Anionsfromthe middle
part oftheHofmeister series, Br
–
, Cl
–
, CH
3
COO
–
, without
significant chaotropic or kosmotropic properties did not
affect the value of k
cat
even at high concentrations. On the
other hand, both chaotropic and kosmotropic anions caused
adecreaseinthek
cat
value with increased concentration. As
confirmed by parallel experiments with KCl and NaCl that
provided identical results within the margin of error, cations
do not have an effect on the kinetic parameters of NADH
oxidase. Figure 2 shows the relative activityoftheenzyme in
the presence of 1
M
and 2
M
salt concentrations. Whereas the
apparent k
cat
decreased inthe presence of both chaotropic
and kosmotropic anions, the apparent K
m
significantly
increased inthe presence of chaotropic anions and decreased
in the presence of kosmotropic anions (Table 1). It should be
noted that the real k
cat
(k
cat real
) is underestimated when the
substrate concentration is lower than 10 · K
m
.Fromthe
Michaelis–Menten equation (Eqn 1), we know that in the
presence of [S] ¼ 10 · K
m
the apparent k
cat
is related to
the real catalytic rate, as k
cat real
¼ 11/10 · k
cat
.Inthe
presence of high concentrations (> 0.5
M
) of chaotropic salt,
the substrate (NADH) concentration [S] is related to K
m
as
[S] ffi 3K
m
(Table 1). In this case, k
cat real
is related to k
cat
as
k
cat real
ffi 4/3k
cat
. However, at high concentrations of
chaotropic salt, the absolute value of k
cat
is % 7 times lower
than inthe presence of neutral salt. Thus k
cat real
in chaotropic
salts is related to k
cat real
in neutral salts as: k
cat real chaotrop
/
k
cat real neutral
¼ (4/3) · (1/7), i.e. significantly less than 1.
Therefore, even if K
m
increases by % 2–3-fold, the bell shape
of k
cat real
(the relative values of k
cat real chaotrop
/k
cat real neutral
)
will not be significantly affected.
As the result of decreased conformational dynamics,
enzymes from thermophiles have very low activity at low
temperatures [39]. The protein dynamics and thermal
stability are inversely related to each other [40,41]. The
dependence oftheenzymeactivity on temperature in
the presence ofthe salts was investigated to assess how the
conformational dynamics oftheactivesite is dependent on
the type of salt present (Fig. 2). Figure 3 shows the
temperature dependence ofthe relative rate constant k
cat
in the presence of 2
M
salt. For each salt, k
cat
at 20 °Cwas
Table 1. Apparent rate constant (k
cat
), Michaelis constants (K
m
) and their ratio r at various concentrations ofthe salts. Assays were performed using
120 n
M
enzyme and 0.12 m
M
FAD in 5 m
M
potassium phosphate buffer and the given concentration of salts, pH 7.0 at 20 °C. The reaction was
started bythe addition of 0.18 m
M
NADH inthe absence of salts. Apparent k
cat
¼ 1.10 ± 0.11 s
)1
, K
m
¼ 8.5 ± 0.9 l
M
and catalytic efficiency
k
cat
/K
m
¼ r ¼ 1.30 · 10
5
M
)1
Æs
)1
. Errors in determination of k
cat
and K
m
are within 10%. This value was calculated from several (2–5) independent
measurements. ND, Not determined.
Anion
0.5
M
1.0
M
1.5
M
2.0
M
k
cat
(s
)1
)
K
m
(l
M
)
r · 10
)5
(
M
)1
Æs
)1
)
k
cat
(s
)1
)
K
m
(l
M
)
r · 10
)5
(
M
)1
Æs
)1
)
k
cat
(s
)1
)
K
m
(l
M
)
r · 10
)5
(
M
)1
Æs
)1
)
k
cat
(s
)1
)
K
m
(l
M
)
r · 10
)5
(
M
)1
Æs
)1
)
SCN
–
5.36 15.60 3.43 1.00 30.55 0.33 0.78 ND ND 0.49 ND ND
ClO
4
–
5.10 15.65 3.26 0.81 44.22 0.18 0.54 ND ND 0.27 ND ND
I
–
7.18 29.34 2.45 6.97 ND ND 3.86 ND ND 1.29 ND ND
Br
–
7.61 20.90 3.64 6.75 21.71 3.11 6.22 22.51 2.76 5.36 28.14 1.91
Cl
–
5.01 13.25 3.78 7.50 13.67 5.55 6.64 13.72 4.84 6.43 13.40 4.80
CH
3
COO
–
4.82 12.86 3.75 4.82 10.05 5.80 5.36 18.94 2.83 4.95 28.12 1.76
SO
2À
4
2.68 11.25 2.40 2.19 7.23 3.03 1.07 ND ND 0.86 ND ND
H
2
PO
4
–
3.32 13.67 2.43 2.14 10.45 2.05 1.18 8.94 1.32 0.38 ND ND
Fig. 2. Relative activityofNADHoxidasefrom T. thermophilusin the
presence of 1
M
(gray histogram) and 2
M
(black) sodium or potassium
salts ofthe designated anions, in 5 m
M
phosphate buffer, pH 7.0, at
20 °C. Activity was initiated bythe addition of 0.2 m
M
NADH.
Ó FEBS 2003 Anion effect on activityofNADHoxidase (Eur. J. Biochem. 271)51
taken as the reference value. Figure 3 shows that the slope
of the observed dependencies increases according to the
position oftheanionsintheHofmeister series, inthe order
from chaotropic to kosmotropic anions. This indicates that,
in the presence of chaotropic anions (SCN
–
,ClO
4
–
)the
activation energy is temperature independent, whereas in
the presence of kosmotropic anions (SO
4
2–
,H
2
PO
4
–
)itis
strongly temperature dependent.
To determine how activation parameters are affected in
the presence of various concentrations of different salts, the
temperature dependencies ofthe rate constants were
measured at 20–40 °C (Supplementary material). Figure 4
shows a dependence of DG*, at 20 °C, on the concentration
of perchlorate, chloride and sulfate anions. Inthe range
0.5–1.0
M
salt, there is a minimum of this dependence for all
anions studied. Whereas inthe presence of chloride (neutral)
anions, the dependence achieves a local minimum, in the
case of both sulfate (kosmotropic) and perchlorate (chao-
tropic) anions, the observed minimum is global. It should be
noted that, although the observed minima are not pro-
nounced, a similar tendency of DG* is observed for all
anions, indicating that the observed dependencies are real.
A double minimum or wide minimum, inthe range 0.5–
2.0
M
salt, of DG* vs. concentration is also observed for
bromide, iodide and acetate anions, i.e. anionsfrom the
middle part oftheHofmeister series. The wide minimum in
the case of these anions also supports the relative independ-
ence of k
cat
on the salt concentration (Table 1). Only one
minimum and one relatively sharp maximum activity of
NADH oxidase is observed for both chaotropic and
kosmotropic anions. The DG* and k
cat
dependencies
correlate in this sense that the minimum of DG* is located
at a similar (same) concentration range as the maximum of
k
cat
for each given anion.
To demonstrate that the observed changes in enzyme
activity are related to conformational changes inthe active
site, we analyzed the CD spectra ofthe peptide (data not
shown) and aromatic regions (Fig. 5). The CD spectrum of
NADH oxidaseinthe aromatic region consists of a positive
Fig. 3. Dependence of relative activityofNADHoxidase from
T. thermophilus on temperature inthe presence of 2
M
sodium or
potassium salts ofthe following anions: H
2
PO
4
–
(j), SO
4
2–
(
~
), Cl
–
(d),
Br
–
(.), I
–
(e), CH
3
COO
–
(h), ClO
4
–
(,), SCN
–
(r)in5mM
phosphate buffer, pH 7.0, at 20 °C.
Fig. 4. Dependence of activation free energy (DG*) ofthe reaction
catalyzedbyNADHoxidasefromT. thermophilus at 20 °Cinthe
presence of 2
M
NaCl (d), NaClO
4
(,), or Na
2
SO
4
(
~
), in 5 m
M
phosphate buffer, pH 7.0.
Fig. 5. CD spectra ofNADHoxidasefrom T. thermophilusin the
aromatic region inthe presence of 2
M
NaCl (dashed line), NaClO
4
(dotted line), or Na
2
SO
4
(thick solid line) and inthe absence of salts (thin
solid line), in 5 m
M
phosphate buffer, pH 7.0, at 20 °C. Inset: Nor-
malized tryptophan fluorescence (excitation wavelength 295 nm) of
NADH oxidasefrom T. thermophilusinthe aromatic region in the
presence of 2
M
NaCl (dashed line), NaClO
4
(dotted line), ans Na
2
SO
4
(thick solid line) and inthe absence of salts (thin solid line), in 5 m
M
phosphate buffer, pH 7.0, at 20 °C.
52 G. Z
ˇ
olda
´
k et al.(Eur. J. Biochem. 271) Ó FEBS 2003
band at % 265 nm and negative ellipticity at 286 nm.
NADH oxidase contains four tryptophan residues in
positions 47, 52, 131, 204. Trp131 and Trp204 are
completely exposed to the solution, whereas Trp52 is rigidly
embedded inthe protein matrix. Trp47 is in a sandwich-like
position toward the flavin cofactor at a distance of about
7.7 A
˚
. This is the only tryptophan residue suitably located
for interaction with the flavin. Interestingly, this position
between Trp47 and the flavin cofactor can be achieved only
in the dimeric form oftheenzyme [9]. In accordance with
previously published CD spectra ofthe flavin oxidases [42],
the pronounced peak at 265 nm may result from an
asymmetric environment ofthe tightly bound flavin cofac-
tor and/or Trp47 intheactivesiteofthe enzyme, and Trp52.
The small negative ellipticity at 286 nm corresponds to the
signal of tryptophan residues. The CD spectrum of NADH
oxidase inthe peptide region is not significantly perturbed,
even at high ionic strength (data not shown). Similarly, the
spectrum oftheenzymeinthe aromatic region in the
presence of 2
M
anions is only slightly affected. A slight
decrease inthe positive ellipticity at % 265 nm in the
presence of perchlorate anions, i.e. a decrease in the
asymmetry ofthe tryptophan residue and/or the flavin
cofactor intheactive site, may result from dissociation of
the flavin cofactor inthe presence of nucleophilic agents
[42]. A 24 h dialysis ofNADHoxidaseinthe presence of
2
M
perchlorate anions did not cause dissociation of the
flavin cofactor (data not shown). The CD spectrum of the
enzyme inthe aromatic regions therefore probably reflects a
slight changein either the conformation or the dynamics of
the tryptophan residue intheactive site.
Fluorescence is the other very sensitive method of
monitoring changes inthe environment close to the
fluorophores. As shown inthe inset of Fig. 5, the presence
of 2
M
chloride or 2
M
sulfate causes a changein fluores-
cence as compared with low ionic strength. The fluorescence
of NADHoxidase is decreased by % 40% inthe presence of
2
M
perchlorate anions. Interestingly, a similar decrease in
the fluorescence ofNADHoxidase was also observed at the
concentration of urea at which activation ofthe enzyme
occurred (unpublished observation). A decrease in fluores-
cence further confirms that the flavin cofactor does not
dissociate fromthe enzyme. Close localization of Trp47 and
the flavin cofactor causes resonance energy transfer, result-
ing in partial quenching ofthe tryptophan fluorescence;
therefore, dissociation ofthe flavin would be accompanied
by an increase in tryptophan fluorescence.
The results presented indicate a close relationship
between enzymeactivity and the stability/conformational
flexibility oftheactive site. The diminished enzyme activity
in the presence of a high concentration (> 1
M
)of
kosmotropic and chaotropic anions probably reflects high
stability/rigidity and too much flexibilityoftheactive site,
respectively. We observed that NADHoxidase from
T. thermophilus at room temperature is activated 2.5-fold
in the presence of 1.0–1.5
M
urea. This activation is
probably caused by increased conformational dynamics of
the side chains intheactivesiteinthe presence of urea. If
this suggestion of a role for flexibilityinenzymeactivity is
correct, NADH oxidase, inthe presence of kosmotropic
anions (H
2
PO
4
–
, SO
4
2–
) should be activated at higher urea
concentrations than inthe presence of neutral and chao-
tropic anions. The experiments presented in Fig. 6 support
this suggestion. Inthe presence of phosphate and sulfate
anions, NADHoxidase is more than 2.5 and 3.5 times more
active, respectively, inthe presence of urea than without
urea (Fig. 6). No activation, but relatively strong inhibition,
was observed inthe presence ofthe chaotropic anions,
ClO
4
–
and SCN
–
, as a result of increased concentrations of
urea (Fig. 6). No significant effect of urea (up to 2
M
)onthe
ellipticity ofNADHoxidaseinthe aromatic region in 2
M
sulfate, chloride and perchlorate anions (data not shown)
further indicates that changes inNADHoxidaseactivity are
not the result of pronounced conformational change but are
probably due to changes inthe dynamics of protein
structure.
Finally, the effect of anion-induced changes in the
dynamics ofthe FAD microenvironment was further studied
by FAD fluorescence quenching using KI (Fig. 7). The
quenching of FAD fluorescence monitored at 525 nm in the
presence ofanionsoftheHofmeisterseries strongly indicates
a changed flexibilityofthe flavin cofactor environment. The
effect of rhodanide, iodide and bromide anions on the
dynamics oftheenzymeactivesite was not investigated
because these anions very efficiently quench FAD fluores-
cence. As acrylamide is a weak quencher of FAD fluores-
cence, we used the efficient quenching property of iodide
anions for these measurements. The maximum concentra-
tion of KI in quenching experiments was 0.15
M
, i.e. the
concentration of iodide anions at which no significant
conformational changeinNADHoxidase was observed.
Monitoring FAD fluorescence quenching is more advanta-
geous than monitoring tryptophan fluorescence quenching
because there is only one flavin cofactor and it is located in
the activesiteofNADH oxidase. The slope ofthe depend-
encies of F
0
/F vs. quencher inthe presence of 2
M
chaotropic
anions is significantly higher than inthe presence of 2
M
neutral anions. The higher quenching constant (Eqn 6)
indicates an increase inthe dynamics ofthe flavin cofactor in
the presence ofthe chaotropic anions. Analogously, a
Fig. 6. Dependence ofthe relative activityofNADHoxidase from
T. thermophilus on [urea], inthe presence of 2
M
NaH
2
PO
4
(j),
Na
2
SO
4
(
~
), NaCl (d), KI (e), NaClO
4
(,), or KSCN (r)in5m
M
phosphate buffer, pH 7.0, 20 °C.
Ó FEBS 2003 Anion effect on activityofNADHoxidase (Eur. J. Biochem. 271)53
decrease inthe slope ofthe dependencies inthe presence of
kosmotropic salts, compared with neutralsalts, indicates that
the activesiteofNADHoxidase is more rigid.
Discussion
NADH oxidasefrom T. thermophilus has, like other
enzymes from thermophiles, low activity at room tempera-
ture. We have recently shown that theenzyme is activated in
the presence of a relatively low concentration (% 1
M
)of
chaotropic agents such as urea and guanidinium hydro-
chloride (unpublished observation). The observed activation
was not due to a conformational change but was a result of
increased conformational dynamics intheactive site. The
tightly bound structural water between Trp47 and the flavin
cofactor [9] was probably released inthe presence of
chaotropic agents, and theactivesiteoftheenzyme opened,
facilitating the arrival ofthe substrate and leading to an
increased rate constant and an increased Michaelis constant.
To test this suggestion, we investigated the effect of anions
of theHofmeister series. TheHofmeisterseriesof anions
can be divided into chaotropic anions, which salt-in the
peptide groups, and kosmotropic anions, with a tendency to
salt-out nonpolar groups [32]. The difference inthe effect of
chaotropic and kosmotropic anions is also due to a charge
density that affects anion interactions with water molecules
[34]. The combination of these effects led to the relatively
surprising bell shaped dependence ofNADH oxidase
activity vs. 1
M
and 2
M
anions, ordered according to the
Hofmeister series (Fig. 2). In fact, reports dealing with the
effect oftheHofmeisterseriesofanions on enzyme activity
usually show a monotone trend, i.e. enzymes are activated
by chaotropic or kosmotropic anions and inhibited by the
opposite anions [21–24]. Analysis ofthe bell shaped curve
showed that the decrease inthe rate constant inthe presence
of chaotropic anions corresponded to an increase in the
apparent K
m
, whereas the decrease in k
cat
in the presence of
kosmotropic anions corresponded to the decrease in K
m
(Table 1). The apparent Michaelis constant measures the
binding affinity oftheenzyme for the substrate and can also
be used as an indirect measure of either inherent flexibility of
an enzyme molecule [43] or the conformational state of the
active/binding site.
The salts used, even at 2
M
, did not significantly affect the
CD spectrum ofNADHoxidaseinthe peptide and aromatic
regions. This indicates (a) a strong interaction ofthe flavin
cofactor with the protein matrix even in conditions that lead
to the dissociation ofthe cofactor from certain mesophi-
lic flavin oxidases [42], and, more importantly, (b) the
unchanged conformational state oftheenzyme under the
conditions studied. The different dynamics ofthe enzyme
active siteinthe presence of kosmotropic and chaotropic
anions is indicated by: (a) a strong dependence of k
cat
vs.
temperature in kosmotropic anions, and a nearly independ-
ent k
cat
vs. temperature in chaotropic anions (Fig. 3) and (b)
positive and negative activation entropy in kosmotropic and
chaotropic anions, respectively. Moreover, a decrease in
tryptophan fluorescence inthe presence of perchlorate
anions and slight changes inthe CD spectra (Fig. 5) indicate
increased dynamics ofthe tryptophan residue inthe active
site ofthe enzyme, similar to results inthe presence of
% 1.0
M
urea. An analogous decrease in ellipticity in the
aromatic region accompanied by changes in tryptophan
fluorescence ofthe nonhomologous flavoprotein flavodoxin
from Desulfovibrio vulgaris, inthe presence of phosphate
anions, was interpreted as an increase inthe dynamics of the
tryptophan residue inthe vicinity ofthe flavin cofactor [44].
A stronger temperature dependence of k
cat
in the presence of
kosmotropic anions indicates the presence of an energy
barrier, i.e. the difference between the basic and transition
states. On the other hand, the near independence of k
cat
on
temperature inthe presence of chaotropic anions indicates
that theanions have a similar effect for temperature because
the energy difference between the basic and transition states
is small. This is in agreement with findings that chaotropic
anions destroy the natural hydrogen-bonded network of
water with effects similar to increased temperature or
pressure [31], with a probable effect on the dynamics of
the polypeptide/side chains of enzymes.
A noteworthy observation is the linear dependence of
activation enthalpy on activation entropy, a phenomenon
known as entropy/enthalpy compensation, inthe reaction
catalyzed byNADHoxidaseinthe presence of salts in the
concentration range 0.5–1.5
M
(Fig. 8). It is apparent from
these data that chaotropic (SCN
–
, ClO
4
–
,I
–
) and kosmo-
tropic (SO
4
2–
,H
2
PO
4
–
) anions are localized at opposite ends
of the linear dependence, and neutral anions (Br
–
, Cl
–
,
CH
3
COO
–
) are inthe middle ofthe dependence. This also
indicates that chaotropic and kosmotropic salts have
Fig. 7. Dependence of FAD fluorescence inthe presence of 2
M
NaH
2
PO
4
(j), Na
2
SO
4
(
~
), NaCl (d), CH
3
COONa (h), or NaClO
4
(,) on concentration of iodide anions expressed as dependence of F
0
/F vs.
concentration of iodide anions. Fluorescence was monitored at 525 nm
after excitation at 450 nm inthe absence (F
0
) and presence of various
concentrations of KI (F). Steeper dependence indicates that, in the
presence ofthe given salt, the accessibility ofthe flavin cofactor or
efficiency of quenching of FAD fluorescence is higher than depend-
encies with less steep slopes. The numerical value ofthe slope of the
dependencies is an expression ofthe Stern–Volmer quenching constant
(Eqn 6; coefficients of linearity for all of displayed dependencies were
r ‡ 0.99). The K
SV
values for the salts studied were 5.97 ± 0.26
M
)1
for NaH
2
PO
4
, 5.96 ± 0.36
M
)1
for Na
2
SO
4
, 14.77 ± 0.26
M
)1
for
NaCl, 11.76 ± 0.16
M
)1
for CH
3
COONa, and 25.06 ± 0.60
M
)1
for
NaClO
4
. All measurements were performed in 5 m
M
phosphate buffer,
pH 7.0, at 20 °C.
54 G. Z
ˇ
olda
´
k et al.(Eur. J. Biochem. 271) Ó FEBS 2003
different effects on the dynamic state ofthe enzyme. As the
enzyme catalyzes the same reaction inthe presence of
chaotropic and kosmotropic anions, the negative value of
activation entropy in chaotropic salts indicates a higher
flexibility ofthe basic state compared with the transition
state. In other words, the difference between the activation
parameters oftheenzymein chaotropic and kosmotropic
anions is always negative as is the case ofthe difference in
activation parameters between psychrophilic and mesophilic
enzymes [45].
Activation ofNADHoxidaseby urea inthe presence of
the anions studied is dependent on their position in the
Hofmeister series. There is strong activation by urea in
the presence of kosmotropic anions, slight activation in the
presence of neutral anions, and deactivation inthe presence
of chaotropic anions. These observations indicate that the
active siteofNADHoxidase is more stable inthe presence
of kosmotropic anions than inthe presence of chaotropic
anions. The quenching experiments ofthe flavin cofactor
fluorescence (Fig. 7) strongly support this interpretation
and strongly indicate that the anion-induced changes in the
activity ofNADHoxidase are due to a changein flexibility
oftheenzymeactivesite.
These results show that anions nonspecifically activate
NADH oxidase at low concentrations (< 0.5
M
). This is in
accordance with the positive electrostatic potential from the
protein near the flavin cofactor which is a common feature
of homologous flavoenzymes [11]. Nonspecific changes in
the exact nature ofthe contacts within related groups by the
anions may then change (in our case activate) theenzyme at
low (< 0.5
M
) concentrations of salt. At higher concentra-
tions of salt, the effect oftheanions is different and depends
on their position intheHofmeister series. Whereas anions
from the middle oftheHofmeisterseries do not affect
activity, both chaotropic and kosmotropic anions inhibit
NADH oxidase. Changes in K
m
, Stern–Volmer quenching
constants, activation entropy and fluorescence, along with
slight changes in CD spectra, and localization ofthe active
site inthe region with an increased temperature B factor
(Fig. 1) strongly suggest that the mechanism of inhibition of
chaotropic and kosmotropic anions includes a modulation
of flexibilityin comparison with the optimal dynamics
the active site. Kosmotropic anions stabilize and increase
the rigidity oftheenzymeactivesite and thus slow the
catalytic rate k
cat
. On the other hand, chaotropic anions
destabilize and increase theflexibilityoftheenzyme active
site. The increased flexibilityinthe substrate-binding site
leads to the increase in K
m
, i.e. a decrease inthe affinity of
theenzymeforthesubstrate.Thedecreaseink
cat
,
however, can be only partially explained bythe observed
increase in K
m
. The main reason for a decreased k
cat
in
the presence of chaotropic anions is probably increased
dynamics intheactivesite which perturbs the proper
position ofthe donor/substrate and the acceptor/flavin
cofactor inthe hydride transfer and/or other side chains
with an active role inthe catalytic siteofNADH oxidase.
Modulation ofthe conformational dynamics by the
Hofmeister seriesofanions therefore offers a simple
strategy for activation of enzymes from thermophiles and
psychrophiles.
Proteins from thermophiles are stabilized by a combina-
tion of strategies [46]. An important one is the presence of
optimized ionic pairs on the protein surface (electrostatic
interaction), i.e. where theactive sites of enzymes are
localized [46–48]. Perturbation weakens some ofthe ionic
interactions and may affect the mobility ofthe polypeptide/
side chain on the protein surface. This would have a positive
impact on theenzymeactivity without a significant effect on
the stable hydrophobic protein core.
On the other hand, enzymes from psychrophiles contain a
highly charged region in order to improve solvent interac-
tions with a hydrophilic surface [50]. Shielding of these
interactions by a suitably chosen salt fromthe Hofmeister
series, or another osmolyte, may stabilize the protein
structure at increased temperature without deleterious
effects on enzyme activity.
Acknowledgements
We thank the Fonds of Chemischen Industrie for financial support. We
are also grateful for support through grants no. D/01/02768 from the
Deutsche Akademische Austauschdienst (DAAD), and no. 1/8047/01
and 1/0432/03 fromthe Slovak Grant Agency. We thank Norbert
Grillenbeck for his technical assistance. We also thank Linda Sowdal
and Dr LeAnn K. Robinson for their invaluable editorial help in
preparing the manuscript.
References
1. Vieille, C. & Zeikus, G.J. (2001) Hyperthermophilic enzymes:
sources, uses, and molecular mechanisms for thermostability.
Microbiol. Mol. Biol. Rev. 65, 1–43.
2. Gerday, C., Aittaleb, M., Arpigny, J.L., Baise, E., Chessa, J P.,
Garsoux, G., Petrescu, I. & Feller, G. (1997) Psychrophilic
enzymes: a thermodynamic challenge. Biochim. Biophys. Acta
1342, 119–131.
3. Fan, Y X., Ju, M., Zhou, J M. & Tsou, C L. (1996) Activation
of chicken liver dihydrofolate reductase by urea and guanidine
hydrochloride is accompanied by conformational change at the
active site. Biochem. J. 315, 97–102.
4. Zhang, H J., Sheng, X R., Pan, X M. & Zhou, J M. (1997)
Activation of adenylate kinase by denaturants is due to the
Fig. 8. Enthalpy/entropy compensation of activation parameters of
reaction catalyzed byNADHoxidaseinthe presence of chaotropic
(SCN
–
, ClO
4
–
,I
–
) (dark symbols), neutral (Br
–
, Cl
–
, CH
3
COO
–
)(grey
symbols), and kosmotropic (SO
4
2–
,H
2
PO
4
–
) (white symbols) salts.
Ó FEBS 2003 Anion effect on activityofNADHoxidase (Eur. J. Biochem. 271)55
increasing conformational flexibility at its active sites. Biochem.
Biophys. Res. Commun. 238, 382–386.
5. Narayanasami, R., Nishimura, J.S., McMillan, K., Roman, L.J.,
Shea, T.M., Robida, A.M., Horowitz, P.M. & Masters, B.S.S.
(1997) The influence of chaotropic reagents on neuronal nitric
oxide synthase and its flavoprotein module. Urea and guanidine
hydrochloride stimulate NADPH–cytochrome c reductase activity
of both proteins. Nitric Oxide – Biol. Chem. 1, 39–49.
6. Das, M. & Dasgupta, D. (1998) Enhancement of transcriptional
activity of T7 RNA polymerase by guanidine hydrochloride.
FEBS Lett. 427, 337–340.
7. Inui, T., Ohkubo, T., Urade, Y. & Hayaishi, O. (1999) Enhance-
ment of lipocalin-type prostaglandin D synthase enzyme activity
by guanidine hydrochloride. Biochem. Biophys. Res. Commun.
266, 641–646.
8. Deshpande, R.A., Kumar, A.R., Khan, M.I. & Shankar, V. (2001)
Ribonuclease Rs from Rhizopus stolonifer: lowering of optimum
temperature inthe presence of urea. Biochim. Biophys. Acta 1545,
13–19.
8a. Z
ˇ
olda
´
k, G., S
ˇ
ut’a
´
k, R., Antalı
´
k, M., Sprinzl, M. & Sedla
´
k, E.
(2003) Role of conformational flexibility for enzymatic activity in
NADH oxidasefromThermus thermophilus. Eur. J. Biochem. 270,
doi:10.1046/j.1432-1033.2003.03889.x.
9. Hecht, H.J., Erdmann, H., Park, H.J., Sprinzl, M. & Schmid,
R.D. (1995) Crystal structure ofNADHoxidasefrom Thermus
thermophilus. Nat. Struct. Biol. 2, 1109–1114.
10. Lovering, A.L., Hyde, E.I., Searle, P.F. & White, S.A. (2001) The
structure of Escherichia coli nitroreductase complexed with nico-
tinic acid: three crystal forms at 1.7 A
˚
, 1.8 A
˚
and 2.4 A
˚
resolution.
J. Mol. Biol. 309, 203–213.
11. Haynes, C.A., Koder, R.L., Miller, A F. & Rodgers, D.W. (2002)
Structures of nitroreductase in three states: effects of inhibitor
binding and reduction. J. Biol. Chem. 227, 11513–11520.
12. Shoichet, B.K., Baase, W.A., Kuroki, R. & Matthews, B.W.
(1995) A relationship between protein stability and protein func-
tion. Proc. Natl Acad. Sci. USA 92, 452–456.
13. Karplus, M. & Petsko, G.A. (1990) Molecular dynamics simula-
tions in biology. Nature 399, 631–639.
14. Carugo, O. & Argos, P. (1998) Accessibility to internal cavities and
ligand binding sites monitored by protein crystallographic thermal
factors. Proteins 31, 201–213.
15. Frauenfelder, H., Petsko, G.A. & Tsernoglou, D. (1979) Tem-
perature-dependent x-ray diffraction as a probe of protein struc-
tural dynamics. Nature 280, 558–563.
16. Goto, Y. & Aimoto, S. (1991) Anion and pH-dependent con-
formational transition of an amphiphilic polypeptide. J. Mol. Biol.
218, 387–396.
17. Jelesarov, L., Du
¨
rr, E., Thoms, R.M. & Bosshard, H.R. (1998)
Salt effects on hydrophobic interaction and charge screening in the
folding of a negatively charged peptide to a coiled coil (leucine
zipper). Biochemistry 37, 7539–7550.
18. Sedla
´
k, E., Z
ˇ
olda
´
k, G., Antalı
´
k, M. & Sprinzl, M. (2002) Ther-
modynamic properties of nucleotide-free EF-Tu from Thermus
thermophilus inthe presence of low-molecular weight effectors of
its GTPase activity. Biochim. Biophys. Acta 1597, 22–27.
19. Ivell, R., Sander, G. & Parmeggiani, A. (1981) Modulation by
monovalent and divalent cations ofthe guanosine-5¢-triphospha-
tase activity dependent on elongation factor Tu. Biochemistry 20,
6852–6859.
20. Fasano, O., De Vendittis, E. & Parmeggiani, A. (1982) Hydrolysis
of GTP by elongation factor Tu can be inducedby monovalent
cations inthe absence of other effectors. J. Biol. Chem. 257, 3145–
3150.
21. Wolosiuk, R.A. & Stein, M. (1990) Modulationof spinach
chloroplast NADP-glyceraldehyde-3-phosphate dehydrogenase
by chaotropic anions. Arch. Biochem. Biophys. 279, 70–77.
22. Wondrak, E.M., Louis, J.M. & Oroszlan, S. (1991) The effect
of salt on the Michaelis Menten constant ofthe HIV-1 pro-
tease correlates with theHofmeister series. FEBS Lett. 280,
344–346.
23. Hall, D.L. & Darke, P.L. (1995) Activation ofthe herpes simplex
virus type 1 protease. J. Biol. Chem. 270, 22697–22700.
24. Nishimura, J.S., Narayanasami, R., Miller, R.T., Roman, L.J.,
Panda, S. & Masters, B.S.S. (1999) The stimulatory effects of
Hofmeister ions on the activities of neuronal nitric-oxide synthase.
Apparent substrate inhibition by
L
-arginine is overcome in the
presence of protein-destabilizing agents. J. Biol. Chem. 274, 5399–
5406.
25. Fan, Y X., McPhie, P. & Miles, E.W. (2000) Regulation of
tryptophan synthase by temperature, monovalent cations, and an
allosteric ligand. Evidence from Arrhenius plots, absorption
spectra, and primary kinetic isotope effects. Biochemistry 39,
4692–4703.
26. Tanfani, F., Scire
`
, A., Masullo, M., Raimo, G., Bertoli, E. &
Bocchini, V. (2001) Salts induce structural changes in elongation
factor 1alpha fromthe hyperthermophilic archaeon Sulfolobus
solfataricus: a Fourier transform infrared spectroscopic study.
Biochemistry 40, 13143–13148.
27. Ru, M.T., Hirokane, S.Y., Lo, A.S., Dordick, J.S., Reimer, J.A. &
Clark, D.S. (2000) On the salt-induced activation of lyophilized
enzymes in organic solvents: effect of salt kosmotropicity on
enzyme activity. J. Am. Chem. Soc. 122, 1565–1571.
28. von Hippel, P.H. & Wong, K Y. (1964) Neutral salts: the gen-
erality of their effects on the stability of macromolecular con-
formations. Science 145, 577–580.
29. Breslow, R. & Guo, T. (1990) Surface tension measurements show
that chaotropic salting-in denaturants are not just water-structure
breakers. Proc. Natl Acad. Sci. USA 87, 167–169.
30. Timasheff, S.N.(1993) The control of protein stability and associ-
ation by weak interactions with water: how do solvents affect these
processes? Annu. Rev. Biophys. Biomol. Struct. 22, 67–97.
31. Leberman, R. & Soper, A.K. (1995) Effect of high salt con-
centrations on water structure. Science 378, 364–366.
32. Baldwin, R.L. (1996) How Hofmeister ion interactions affect
protein stability. Biophys. J. 71, 2056–2063.
33. Cacace, M.G., Landau, E.M. & Ramsden, J.J. (1997) The Hof-
meister series: salt and solvent effects on interfacial phenomena.
Q. Rev. Biophys. 30, 241–277.
34. Collins, K.D. (1997) Charge density-dependent strength of
hydration and biological structure. Biophys. J. 72, 65–76.
35. Saunders, A.J., Daris-Searles, P.R., Allen, D.L., Pielak, G.J. &
Erie, D.A. (2000) Osmolyte-induced changes in protein con-
formational equilibria. Biopolymers 53, 293–307.
36. Hribar, B., Southall, N.T., Vlachy, V. & Dill, K.A. (2002) How
ions affect the structure of water. J. Am. Chem. Soc. 121, 12302–
12311.
37. Park, H.J., Reiser, C.O.A., Kondruweit, S., Erdmann, H., Sch-
mid, R.D. & Sprinzl, M. (1992) Purification and characterization
of a NADHoxidasefromthe thermophile Thermus thermophilus
HB8. Eur. J. Biochem. 205, 881–885.
38. Lehrer, S.S. (1971) Solute perturbation of protein fluorescence.
The quenching ofthe tryptophyl fluorescence of model compounds
and of lysozyme by iodide ion. Biochemistry 10, 3254–3263.
39. Za
´
vodszky, P., Kardos, J., Svingor, A. & Petsko, G.A. (1998)
Adjustment of conformational flexibility is a key event in the
thermal adaptation of proteins. Proc. Natl Acad. Sci. USA 95,
7406–7411.
40. Vihinen, M. (1987) Relationship of protein flexibility to thermo-
stability. Protein Eng. 1, 477–480.
41. Tsai, A.M., Udovic, T.J. & Neumann, D.A. (2001) The inverse
relationship between protein dynamics and thermal stability.
Biophys. J. 81, 2339–2343.
56 G. Z
ˇ
olda
´
k et al.(Eur. J. Biochem. 271) Ó FEBS 2003
42. Zlateva, T., Boteva, R., Filippi, B., Veenhuis, M. & van der Klei,
I.J. (2001) Deflavination of flavo-oxidases by nucleophilic
reagents. Biochim. Biophys. Acta 1548, 213–219.
43. Fields, P.A. (2001) Review: protein function at thermal extremes:
balancing stability and flexibility. Comp.Biochem.Physiol.A129,
417–431.
44. Murray, T.A., Foster, M.P. & Swenson, R.P. (2003) Mechanism
of flavin mononucleotide cofactor binding to the Desulfovibrio
vulgaris flavodoxin. 2. Evidence for cooperative conformational
changes involving tryptophan 60 inthe interaction between the
phosphate- and ring-binding subsites. Biochemistry 42, 2317–
2327.
45. Lonhienne, T., Gerday, C. & Feller, G. (2000) Psychrophilic
enzymes: revisiting the thermodynamic parameters of acti-
vation may explain local flexibility. Biochim. Biophys. Acta 1543,
1–10.
46. Kumar, S., Tsai, C J. & Nussinov, R. (2000) Factors enhancing
protein thermostability. Protein Eng. 13, 179–191.
47. Xiao, L. & Honig, B. (1999) Electrostatic contributions to the
stability of hyperthermophilic proteins. J. Mol. Biol. 289, 1435–
1444.
48. Loladze, V.V., Ibarra-Molero, B., Sanchez-Ruiz, J.M. &
Makhatadze, G.I. (1999) Engineering a thermostable protein via
optimization of charge–charge interactions on the protein surface.
Biochemistry 38, 16419–16423.
49. Spector, S., Wang, M., Carp, S.A., Roblee, J., Hendsch, Z.S.,
Fairman, R., Tidor, B. & Raleigh, D.P. (2000) Rational modi-
fication of protein stability bythe mutation of charged surface
residues. Biochemistry 39, 872–879.
50. Feller, G., Arpigny, J.L., Narinx, E. & Gerday, Ch (1997) Mole-
cular adaptations of enzymes from psychrophilic organisms.
Comp. Biochem. Physiol. 118A, 495–499.
Supplementary material
The following material is available from http://blackwell
publishing.com/products/journals/suppmat/EJB/EJB3900/
EJB3900sm.htm
Supplementary material S1. (A–H) Activation parameters
calculated fromthe temperature dependencies ofactivity at
various concentrations of salts.
Ó FEBS 2003 Anion effect on activityofNADHoxidase (Eur. J. Biochem. 271)57
. Modulation of activity of NADH oxidase from Thermus thermophilus through change in flexibility in the enzyme active site induced by Hofmeister series anions Gabriel Z ˇ olda ´ k 1 ,. and increase the flexibility of the enzyme active site. The increased flexibility in the substrate-binding site leads to the increase in K m , i.e. a decrease in the affinity of theenzymeforthesubstrate.Thedecreaseink cat , however,. strongly indicate that the anion -induced changes in the activity of NADH oxidase are due to a change in flexibility oftheenzymeactivesite. These results show that anions nonspecifically activate NADH oxidase