Вы находитесь на странице: 1из 9

THE JOURNAL OF BIOLOGICAL CHEMISTRY Vol. 280, No. 28, Issue of July 15, pp.

26129 26136, 2005


2005 by The American Society for Biochemistry and Molecular Biology, Inc. Printed in U.S.A.

Thiocyanate Modulates the Catalytic Activity of


Mammalian Peroxidases*
Received for publication, March 18, 2005, and in revised form, May 11, 2005
Published, JBC Papers in Press, May 13, 2005, DOI 10.1074/jbc.M503027200

Yahya R. Tahboub, Semira Galijasevic, Michael P. Diamond, and Husam M. Abu-Soud


From the Department of Obstetrics and Gynecology, The C. S. Mott Center for Human Growth and Development and the
Department of Biochemistry and Molecular Biology, Wayne State University, Detroit, Michigan 48201

We investigated the potential role of the co-substrate, overall amino acid sequence homology but differ from each
thiocyanate (SCN), in modulating the catalytic activity other with respect to their sites of expression, primary se-
of myeloperoxidase (MPO) and other members of the quences, and substrate specificities (1, 2, 8 10). MPO is a
mammalian peroxidase superfamily (lactoperoxidase 150 165-kDa molecule synthesized during myeloid differenti-
(LPO) and eosinophil peroxidase (EPO)). Pre-incubation ation that constitutes the major component of the neutrophil
of SCN with MPO generates a more complex biological azurophilic granules (11, 12). The enzyme is a homodimer
setting, because SCN serves as either a substrate or comprising of a pair of light and heavy chains derived from a
inhibitor, causing diverse impacts on the MPO heme single gene product (12, 13) with its subunits joined by a single
iron microenvironment. Consistent with this hypothe- disulfide bridge (12). The heavy chain contains an iron bound

Downloaded from http://www.jbc.org/ by guest on December 16, 2016


sis, the relationship between the association rate con- by a novel protoporphyrin IX derivative, which is covalently
stant of nitric oxide binding to MPO-Fe(III) as a function
linked to the heavy chain polypeptide (14 16). EPO is a mono-
of SCN concentration is bell-shaped, with a trough
meric molecule comprised of a light chain and a heavy chain
comparable with normal SCN plasma levels. Rapid ki-
with molecular masses of 15.5 and 50 kDa, respectively (17).
netic measurements indicate that MPO, EPO, and LPO
Compound I formation occur at rates slower than com- This enzyme is stored in eosinophil granules and catalyzes the
plex decay, and its formation serves to simultaneously formation of antimicrobial species from the oxidation of Br
catalyze SCN via 1e and 2e oxidation pathways. For and SCN (1719). LPO is a monomeric single polypeptide
the three enzymes, Compound II formation is a funda- chain with a molecular mass of 78.5 kDa (20 23). LPO has
mental feature of catalysis and allows the enzymes to been identified as an antimicrobial agent within exocrine gland
operate at a fraction of their possible maximum activi- secretions such as milk, saliva, and tears through the oxidation
ties. MPO and EPO Compound II is relatively stable and of thiocyanate by H2O2 to yield the intermediary oxidation
decays gradually within minutes to ground state upon product hypothiocyanite (OSCN) (24, 25).
H2O2 exhaustion. In contrast, LPO Compound II is un- The properties of the heme in MPO, EPO, and LPO have
stable and decays within seconds to ground state, sug- been characterized by a wide variety of spectroscopic tech-
gesting that SCN may serve as a substrate for Com- niques, including optical absorption, stopped-flow, electron
pound II. Compound II formation can be partially or paramagnetic resonance, and resonance Raman spectroscopy
completely prevented by increasing SCN concentra- (26 37). These techniques, with a combination of structure
tion, depending on the experimental conditions. Collec- analysis (16, 38 39) and advanced computer modeling (40),
tively, these results illustrate for the first time the po- have shown that the heme pockets of the mammalian peroxi-
tential mechanistic differences of these three enzymes. dases are envisioned to form the catalytic site where the step-
A modified kinetic model, which incorporates our cur- wise reduction of H2O2 takes place. The simplified mechanism
rent findings with the mammalian peroxidases classic
that governs the catalytic activity of mammalian peroxidase
cycle, is presented.
superfamily can be represented by the classic peroxidases cat-
alytic cycle. In this cycle, represented here by Equations 1 4,
Myeloperoxidase (MPO)1 and other members of the mamma- E-Fe(III) H2O2 3 E-Fe(IV)O H2O
(Eq. 1)
lian peroxidase superfamily (eosinophil peroxidase (EPO) and Compound I
lactoperoxidase (LPO)) display a crucial difference (within a
E-Fe(IV)O SCN H 3 E-Fe(III) HOSCN (Eq. 2)
wide range of biological processes) in their unique ability in
catalyzing the H2O2-dependent peroxidation of halides and E-Fe(IV)O AH2 3 E-Fe(IV)O AH
pseudohalides to produce antimicrobial agents and hypohalous Compound II (Eq. 3)
acids (17). These heme-containing enzymes share 50 70%
E-Fe(IV)O AH2 3 E-Fe(III) AH (Eq. 4)

* This work was supported by National Institutes of Health Grant H2O2 reacts rapidly and reversibly with the ground state (E-
HL066367 to (H. M. A.-S.). The costs of publication of this article were Fe(III) state) of mammalian peroxidases and generates a ferryl
defrayed in part by the payment of page charges. This article must cation radical (E-Fe(IV)O) intermediate named Com-
therefore be hereby marked advertisement in accordance with 18
U.S.C. Section 1734 solely to indicate this fact. pound I (Equation 1) (33, 41). Compound I is capable of oxidiz-
To whom correspondence should be addressed: Dept. of Obstetrics ing either halides and pseudohalides through a 2e transition,
and Gynecology, C. S. Mott Center for Human Growth and Develop- generating the ground state and the corresponding hypohalous
ment, Wayne State University, 275 E. Hancock, Detroit, MI 48201. Tel.: acid (Equation 2) or oxidizing multiple organic and inorganic
313-577-6178; Fax: 313-577-8554; E-mail: habusoud@med.wayne.edu.
1
The abbreviations used are: MPO, myeloperoxidase; EPO, eosino-
molecules (AH2) by two successive sequential 1e transitions
phil peroxidase; LPO, lactoperoxidase; E, enzyme (MPO, EPO, or LPO); generating their corresponding cation (AH) and the peroxidase
H2O2, hydrogen peroxide; NO, nitric oxide (nitrogen monoxide). intermediates Compound II (E-Fe(IV)O) and E-Fe(III), re-

This paper is available on line at http://www.jbc.org 26129


26130 Differences in the Mechanisms of Mammalian Peroxidases
spectively (Equations 3 and 4) (33, 41 43). Compound II is a 412 and 432 nm, respectively. For NO binding to MPO, measurements
longer lived intermediate whose decay to ground state is con- were carried out by rapidly mixing equal volumes of the enzyme solu-
tions (0.86 M) with buffer solution supplemented with increasing con-
sidered to be the rate-limiting step during steady-state cataly-
centrations of NO. The reactions for NO binding to MPO-Fe(III) were
sis (33, 44). Acceleration in Compound II formation and decay monitored under anaerobic conditions at wavelengths determined
has been noted with a series of organic and inorganic sub- based on the spectral changes that occur upon NO binding to enzyme,
strates (28, 30, 33, 41 43) and physiological reductants like as indicated. The time course of absorbance change was fit to either
nitric oxide (NO) and superoxide (O2. ) (28, 30, 33, 41 43). single (Y 1 ekt) or double exponential (Y Aek1t Bek2t)
Thiocyanate binds close enough to the heme iron of peroxi- functions as indicated. Signal-to-noise ratios for all kinetic analyses
were improved by averaging at least 6 8 individual traces. In some
dase to influence both its reduction potential and binding af-
experiments the stopped-flow instrument was attached to a rapid scan-
finity to small molecules, such as NO and CO, when these ning diode array device (Hi-Tech) designed to collect a multiple number
molecules bind to the enzymes heme center (16, 28 30, 38, 39). of complete spectra (200 800 nm) at specific time ranges. The detector
Recent x-ray crystal structure analysis and a variety of spec- was automatically calibrated relative to a holmium oxide filter, as it has
troscopic measurements have shown that SCN binding occurs spectral peaks at 360.8, 418.5, 446.0, 453.4, 460.4, 536.4, and 637.5 nm,
at distal heme cavity and serves as either a substrate or inhib- which were used by the software to correctly align pixel positions with
wavelength. Rapid scanning stopped-flow experiments involved mixing
itor, causing a diverse impact on the catalytic characteristic
solutions of MPO (3 M) with buffer solutions containing 40 M H2O2 in
behavior of mammalian peroxidases and function (16, 28 30, the absence and presence of increasing SCN concentrations at 10 C.
38, 39). In some cases, these co-substrates bind to the peroxi- SpectroscopyOptical spectra were recorded on a Cary 100 Bio UV-
dase heme iron and form the corresponding low spin, six- visible spectrophotometer at 25 C. Anaerobic spectra were recorded
coordinate complex that inhibits the catalytic activity of the using septum-sealed quartz cuvettes that could be attached through a
enzyme (45). quick-fit joint to an all-glass vacuum system. MPO samples were made
anaerobic by several cycles of evacuation and equilibrated with cata-
Despite the potential significance of mammalian peroxidases
lyst-deoxygenated N2. Separate buffer solutions were evacuated, gassed
to both human health and disease, little is known about the

Downloaded from http://www.jbc.org/ by guest on December 16, 2016


with N2, and anaerobically transferred either to the stopped-flow in-
factors that govern their substrate selectivity and specificity. strument or to anaerobic cuvettes using gas-tight syringes. Cuvettes
Previous kinetic studies with mammalian peroxidases have were maintained under N2 or NO positive pressure during spectral
been limited in scope by focusing primarily on the direct reac- measurements.
tion of SCN with Compound I (36, 37, 44, 46, 47). In the Solution PreparationA fresh saturated stock of NO was prepared
under anaerobic conditions. Anaerobic 0.2 M sodium phosphate buffer
present studies, we utilize a combination of optical absorbance,
solutions, pH 7.0, containing various concentrations of NO were pre-
rapid kinetic measurements, and NO binding studies to assess pared by mixing different volumes of buffer saturated with NO gas at
the distinct conformational changes that occur upon SCN 21 C with an anaerobic buffer solution. A saturating concentration of
binding and its influence on the catalytic activity of MPO and NO at 21 C is 2 mM.
other members of the mammalian peroxidase superfamily. Our H2O2-selective Electrode MeasurementsH2O2 measurements were
results indicate that SCN binding regulates the catalytic ac- carried out using a H2O2-selective electrode (Apollo 4000 free radical
analyzer; World Precision Instruments, Sarasota, FL). Experiments
tivity of MPO, EPO, and LPO, which appears to result from
were performed at 25 C by immersing the electrode in 10 ml of 0.2 M
significant electronic and/or conformational alterations in the sodium phosphate buffer, pH 7.0, under air. 20 M H2O2 was added to
catalytic sites of these three enzymes that are caused by SCN. a continuously stirred buffer solution during which the rise and fall in
H2O2 concentration was continuously monitored. Where indicated, 50
EXPERIMENTAL PROCEDURES l of MPO (or EPO or LPO) (200 nM final) was added to the reaction
MaterialsNO gas was purchased from Matheson Gas Products, Inc. mixture. To determine the effect of SCN on H2O2 consumption by
and used without further purification. For each experiment, a fresh mammalian peroxidases, similar experiments were repeated by adding
saturated stock of NO was prepared under anaerobic conditions. The 50 l of the enzyme solution (200 nM final) pre-incubated with 40 M
extent of nitrite/nitrate (NO2/NO3) build-up in NO preparations over SCN to the reaction mixture.
the time course used for the present studies was 11.5% (per mole of
RESULTS
NO), as determined by anion exchange high performance liquid chro-
matography under anaerobic conditions (48). All other reagents and
Effects of SCN on NO Binding to MPO Heme IronTo
materials were of the highest purity grades available and obtained from assess the influence of SCN on the catalytic activity of MPO,
Sigma or the indicated source.
the rates of NO binding to the Fe(III) form of MPO have been
Enzyme PurificationMPO was purified from detergent extracts of
human leukocytes as described (49). Trace levels of contaminating determined in the presence of increasing SCN concentrations.
eosinophil peroxidase were then removed by passage over a sulfopropyl- Investigations were carried out under two different circum-
Sephadex column (50). The purity of isolated MPO was established by stances, either a fixed amount of NO and varying concentra-
demonstrating a Reinheitszahl (RZ) value of 0.85 (A430/A280) by tions of SCN or a fixed amount of SCN and varying levels of
means of SDS-PAGE analysis with Coomassie Blue staining and in-gel NO. As shown in Fig. 1, the plots of the apparent rate constants
tetramethylbenzidine peroxidase staining to confirm no observable con-
for NO binding as a function of NO concentration were linear,
taminating eosinophil peroxidase activity (51). EPO was purified from
porcine whole blood as described previously (52). The purity of EPO was consistent with a simple one-step mechanism. As shown in
confirmed by demonstrating an RZ of 0.9 (A415/A280) via SDS-PAGE Equation 5,
analysis with Coomassie Blue staining and in-gel tetramethylbenzidine
kon
peroxidase staining (53). Enzyme concentrations were determined spec-
SCN-MPO-Fe(III) NO
O SCN-MPO-Fe(III)-NO (Eq. 5)
trophotometrically utilizing extinction coefficients of 89,000 and
koff
112,000 M1 cm1 per heme of MPO (51) and EPO (54, 55), respectively.
LPO was obtained from Worthington Biochemical Corporation (Lake-
the positive intercepts confirm that NO binds to MPO-Fe(III)
wood, NJ) and used without further purification. Purity was confirmed
by demonstrating an RZ of 0.75 (A412/A280) by means of SDS-PAGE by a reversible process. These kinetic parameters suggest that
analysis with Coomassie Blue staining. SCN modulates the affinity of MPO-Fe(III) toward NO. The
Stopped-flow and Absorbance MeasurementsRapid kinetic meas- second-order combination rate constants (kon) calculated from
urements were performed using a dual syringe stopped-flow instrument the slopes were plotted as a function of SCN concentration
obtained from Hi-Tech Ltd. (model SF-61). Measurements were carried and showed a bell-shaped relationship (Fig. 2), with a mini-
out at 10 C following rapid mixing of equal volumes of an H2O2-
mum centered at biologically relevant levels of SCN (60 M).
containing buffer solution and a peroxidase solution. MPO Compound I
and Compound II formation and decay to ground state were monitored Collectively, these results indicate that SCN binds within the
at 435 and 455 nm, respectively. EPO and LPO Compound I and enzyme system and modulates the affinity of MPO-Fe(III)
Compound II formation and decay to ground state were monitored at toward NO.
Differences in the Mechanisms of Mammalian Peroxidases 26131

FIG. 1. SCN modulates NO binding to MPO heme iron. Plots of


the observed rates of NO binding to MPO-Fe(III) as a function of NO FIG. 3. Formation of EPO Compound II during steady-state
and SCN concentrations. An anaerobic solution containing 0.86 M catalysis of SCN. Depicted is the absorbance change over time for the
MPO-Fe(III) supplemented with varying concentrations of SCN was reaction that was initiated by rapidly mixing a buffer solution contain-
rapidly mixed with an equal volume of sodium phosphate buffer (200 ing 40 M H2O2 with an equal volume of buffer solution containing 1.2
mM at pH 7.0) supplemented with varying concentration of NO at 10 C. M EPO and 40 M SCN at 10 C. Spectra traces were recorded at
The high concentration of the phosphate buffer is to keep the pH of the 0.027, 0.054, 0.081, 0.110, and 0.135 s after initiating the reaction.
solution unaltered after the addition of NO. The observed rates of Arrows in the panels indicate the direction of spectral change over time.
MPO-Fe(III)-NO were plotted as a function of NO concentration. Experiments were carried out under anaerobic conditions in sodium
phosphate buffer (200 mM at pH 7.0).

time that Compound II is a major component in the catalytic

Downloaded from http://www.jbc.org/ by guest on December 16, 2016


activity of MPO, EPO, and LPO and that its formation allow
these three enzymes to operate at only a fraction of their
maximum activity.
Effect of SCN on the Formation, Duration, and Decay of
MPO and EPO Compound II Formation during Steady-state
CatalysisTo take a closer look at the reaction mechanism of
the metabolism of SCN by MPO, we next investigated the
influence of SCN on the kinetics of MPO compound II buildup,
FIG. 2. Relationship between the second-order combination
duration, and decay using single wavelength stopped-flow
rate (kon) of NO binding to MPO-Fe(III) obtained from Fig. 1 as
a function of the SCN concentration used. methods. Experiments were carried out under aerobic condi-
tions following rapid mixing of H2O2 (40 M) and enzyme solu-
Formation of MPO, EPO, and LPO Compound II during the tion (0.85 M) pre-incubated with various concentrations of
Metabolism of SCNWe next utilized diode array spectropho- SCN ranging from 20 160 M. The time courses for the for-
tometry to study the effect of SCN on MPO, EPO, and LPO mation, duration, and decay of MPO steady-state catalysis
intermediate formation, duration, and decay, as these events were examined by following the absorbance changes at 434 and
occur during steady-state catalysis. Investigation was carried 455 nm. As shown in Fig. 5, we observed a rapid increase in
out by the rapid mixing of a buffer solution supplemented with absorbance at 455 nm followed by a steady state in which the
1.2 M MPO (or EPO or LPO) and 40 M SCN against an equal intermediate formation remained relatively constant, with a
volume of a buffer solution supplemented with 40 M H2O2, at final decay to the origin after MPO had oxidized all of the
10 C. Spectral analysis indicated that the majority of MPO- available substrate, H2O2. The decay of the steady-state reac-
Fe(III) (80 90%) converted immediately to an MPO species tion is biphasic. There is a rapid initial decrease in absorbance
whose Soret absorbance peak, along with the visible bands, is (Fig. 5) followed by a much slower decrease in absorbance over
identical to that reported for MPO Compound II. This species the next 500 s (data is not included in Fig. 5). As summarized
was stable and occurs immediately after initiating the reaction in Fig. 6, the addition of SCN to the MPO reaction results in
without any indication of a prior accumulation of MPO Com- dramatic effects on the duration and decay of the steady-state
pound I. Increasing SCN concentrations decreased the accumulation rate. The rate of steady-state accumulation was
amount of the MPO Compound II accumulation up to 50%, as decreased in the presence of SCN in a concentration-depend-
judged by the decrease in the amplitude of the Soret absorb- ent and saturable manner (Fig. 6A). The rate of its initial decay
ance peak at 455 nm. Similar results were obtained for EPO. increased in a linear manner as a function of SCN concentra-
Indeed, the majority of EPO-Fe(III) converted immediately to a tion (Fig. 6B), yielding a second-order combination rate con-
stable EPO Compound II without any sign of the accumulation stant (kon) of 1 104 M1 s1 calculated from the slope. In
of Compound I. Fig. 3 shows spectral traces collected at 0.027, addition, SCN influences the level of the steady-state accu-
0.054, 0.081, 0.110, and 0.135 s after initiating the reaction. mulation and stability as reflected by the amplitude and dura-
Similarly, Compound II LPO accumulation during catalysis tion of the stopped-flow traces, respectively. As the concentra-
occurs without prior accumulation of LPO Compound I, but to tion of SCN present in the reaction mixture increased, both
a lesser extent, and decays immediately to ground state upon the steady-state level and the amount of the intermediates
H2O2 exhaustion. Fig. 4, panel A shows spectra collected at generated progressively decreased (Fig. 6C). When the reac-
0.035, 0.07, 0.105, 0.14, 0.175, and 0.244 s after initiating the tions were monitored at 434 nm the direction of absorbance
reaction, whereas panel B shows spectra collected at 0.45, 0.75, change was reversed, and the signal amplitudes enhanced but
0.90, 1.05, and 2.20 s after initiating the reaction. However, otherwise proceeded with identical kinetics (data not shown).
when the similar reaction was repeated in the presence of 200 Collectively, the spectra of the intermediates initially formed
M SCN, ground state LPO-Fe(III) predominated with a small following initiation of the reaction indicate that the majority of
a portion of Compound II (35% of the total LPO); its decay to MPO exists as a Compound II and that its formation occurs
ground state occurred immediately after H2O2 exhaustion without the accumulation of Compound I. Similar results were
(data not shown). Collectively, our results indicate for the first obtained for EPO (data not shown), indicating that both MPO
26132 Differences in the Mechanisms of Mammalian Peroxidases

FIG. 4. Absorbance changes for the


metabolism of SCN by LPO as it oc-
curs during steady-state catalysis.
Depicted is the absorbance change over
time for the reaction that was initiated by
rapidly mixing a buffer solution contain-
ing 40 M H2O2 with an equal volume of
buffer solution containing 1 M LPO and
40 M SCN using stopped-flow diode ar-
ray methods at 10 C. Panel A, spectra
collected at 0.035, 0.07, 0.105, 0.14, 0.175,
and 0.244 s after initiating the reaction.
Panel B, spectra collected at 0.45, 0.75,
0.90, 1.05, and 2.20 s after initiating the
reaction. Arrows indicate the direction of
spectral change over time.

Downloaded from http://www.jbc.org/ by guest on December 16, 2016


FIG. 5. Effect of SCN concentration on MPO Compound II
formation, duration, and decay as they occur during steady-
state catalysis. Formation, duration, and decay of steady-state catal-
ysis of MPO were monitored as a function of time at 455 nm. An aerobic
solution containing 0.86 M MPO-Fe(III) pre-incubated with different
concentrations of SCN was rapidly mixed with an equal volume of
sodium phosphate buffer (200 mM at pH 7.0) supplemented with 40 M
H2O2 at 10 C. The initial concentration of SCN in the mixtures is 20,
40, 80, and 160 M, from bottom to top.

and EPO display a similar kinetic mechanism in metabolizing


SCN.
Effect of SCN on the Formation, Duration, and Decay of
LPO Steady-state CatalysisThe time courses for the forma-
tion, duration, and decay of LPO steady-state catalysis were
examined by following the decrease in absorbance changes at
412 nm. At each SCN concentration tested, there is a rapid
decrease in absorbance followed by a steady state in which the
intermediate formation remained relatively constant and fi-
FIG. 6. Rate of MPO Compound II formation, duration, and
nally decayed exponentially to the origin after LPO had oxi- decay as a function of SCN concentration. The observed rates of
dized all of the available substrate, H2O2. However, it is ap- MPO Compound II formation (panel A), decay (panel B), and duration
parent that as the SCN concentration increased there was a (panel C) (monitored at 455 nm) observed in Fig. 5 were plotted as a
decrease in rate of formation, a decrease in the duration of the function of SCN concentration. The percent absorbance change at 455
nm is also shown for MPO steady state as a function of SCN concen-
steady state, and an increase in the decay rate (Fig. 7). When tration (panel C).
the reactions were monitored at 434 nm the direction of ab-
sorbance change was reversed, and the signal amplitudes were
reduced but otherwise proceeded with identical kinetics (data SCN in a concentration-dependent and saturable manner
not shown). This behavior suggests that the spectral changes (Fig. 8A). The rate of its decay increased in a linear manner as
observed should reflect the change in LPO Compound II during a function of SCN concentration (Fig. 8B), yielding a second-
catalysis. As summarized in Fig. 8, the addition of SCN to order combination rate constant (kon) of 7 103 M1 s1 and a
LPO reaction results in dramatic effects on the duration and first order dissociation rate constant of 1 s1 calculated from
decay of the steady-state accumulation rates. The rate of the slope and the intercept, respectively. In addition, SCN
steady-state accumulation was decreased in the presence of influences the level of the steady-state accumulation and sta-
Differences in the Mechanisms of Mammalian Peroxidases 26133

FIG. 7. Spectral changes of SCN metabolism by LPO as they


occur during steady-state catalysis at four different SCN con-
centrations. Formation, duration, and decay of the steady-state catal-
ysis of LPO were monitored as a function of time at 412 nm. An aerobic FIG. 9. A typical recording by a H2O2-selective electrode for
solution containing 0.86 M LPO-Fe(III) pre-incubated with different H2O2 consumption by MPO. Top, the addition of 20 M H2O2 followed
concentrations of SCN was rapidly mixed with an equal volume of by 0.2 M MPO to a stirred 0.2 M sodium phosphate buffer (pH 7) results
sodium phosphate buffer (200 mM, pH 7.0) supplemented with 40 M in dramatic biphasic acceleration in the rate of H2O2 consumption at
H2O2 at 10 C. The initial concentration of SCN in the mixtures is 10, 25 C. Bottom, the addition of 20 M H2O2 followed by 0.2 M MPO
20, 40, 80, and 160 M from bottom to top. pre-incubated with 40 M SCN; only the fast phase has been seen.
Tracings shown are from typical experiment performed at least three
times. The H2O2 traces were corrected by subtracting the autooxidation
of H2O2.

Downloaded from http://www.jbc.org/ by guest on December 16, 2016


that H2O2 is consumed as a substrate by MPO during steady-
state catalysis (Fig. 9, top). The first step occurs immediately
after the enzyme addition and is attributable to the formation
of MPO Compound I. The second step is much slower and is
attributable to the reaction of MPO with H2O2 after the con-
version of MPO Compound II to ground state. When the same
reaction was repeated by adding an MPO sample saturated
with SCN solution to a stirred H2O2 solution, only the first
step was observed (Fig. 9, bottom). Similar results were ob-
tained for both EPO and LPO.

DISCUSSION

Pre-incubation of SCN with MPO and other members of the
mammalian peroxidase superfamily (e.g. EPO and LPO) causes
multiple and sequential reactions and suggests a multi-func-
tional role for SCN before and during steady-state catalysis.
Indeed, pre-incubation with SCN prior to initiating peroxida-
tion generates a more complex biological setting through its
potential capacity to bind within the enzyme systems, above
the heme prosthetic group, or directly to the heme iron, gener-
ating a six-coordinate low spin complex. SCN binds within the
MPO system, serves as a substrate or inhibitor, and modulates
the heme iron microenvironment to cause a significant alter-
ation in its catalytic site, thereby altering the enzymes affinity
toward H2O2. Rapid kinetic measurements utilizing sequential
stopped-flow methods have indicated that the direct reaction
FIG. 8. Rate of LPO steady-state formation, duration, and de- between MPO, EPO, and LPO Compound I and SCN is ex-
cay as a function of SCN concentration. The observed rates of tremely fast and occurs with second-order rate constants rang-
LPO steady-state formation (panel A), decay (panel B), and duration ing from 9.6 107 to 2 108 M1 s1 (33, 36, 37, 44, 46, 47).
(panel C) (monitored at 412 nm) observed in Fig. 7 were plotted as a Thus, the order of addition (the enzyme and H2O2 mixed first
function of SCN concentration. The percent absorbance change at 412
nm is also shown for the LPO steady state as a function of SCN and SCN second, or the enzyme and SCN mixed first and
concentration (panel C). H2O2 second) may have created a considerable degree of con-
troversy, ambiguity, and diverse results in predicting the pre-
bility as reflected by amplitude and duration of the stopped- ferred biological substrate of the mammalian peroxidase
flow traces, respectively. As the concentration of SCN present superfamily.
in the reaction mixture was increased, both the steady-state Because MPO, EPO, and LPO Compound I formation rates
level and the amount of the intermediates generated progres- are slower than the 2e oxidation of SCN, Compound I accu-
sively decreased (Fig. 8C). mulation cannot be detected during steady-state catalysis (27,
Direct H2O2 Consumption by MPO Using a H2O2-selective 33, 36, 37, 44, 46, 47). Thus, the observed absorbance changes
ElectrodeFollowing the addition of 20 M H2O2 to a continu- during SCN metabolism should reflect the alteration in Com-
ously stirred phosphate buffer, the H2O2 signal rose rapidly, pound II accumulation, duration, and decay. Rapid kinetic
achieved a maximum after 30s, and fell gradually to the measurements indicated that MPO, EPO, and LPO Compound
origin as H2O2 was depleted by autooxidation. The addition of II is the predominant species formed, providing the first direct
MPO to the reaction mixture caused an immediate rapid decay evidence for the involvement of Compound II in the catalytic
in the level of free H2O2 followed by a slow decay, indicating inhibition of mammalian peroxidases. The kinetic parameters
26134 Differences in the Mechanisms of Mammalian Peroxidases

FIG. 11. Stereo view of the MPO distal heme cavity of human
MPO with bound thiocyanate. The figure was created using the
coordinates deposited in the Protein Data Bank (accession code 1D7W).

His-95 in the native enzyme. The bent mode of the Fe-N-O may
allow a perturbation in the hydrogen bonding within the distal
FIG. 10. Modified working kinetic model for SCN interaction cavity. Formation of such a complex is accomplished with the
with mammalian peroxidases (E, enzyme MPO, EPO, or LPO). movement of the iron atom into the porphyrin ring, generating
its respective low spin, six-coordinate complex. SCN is a rel-
of Compound II formation, duration, and decay for MPO and atively bulky molecule; its binding above the heme moiety of
EPO were similar, but they differed from those obtained for MPO may constrain NO binding by either filling the space
LPO Compound II. The degree of MPO and EPO Compound II directly above the heme or by promoting a protein conforma-
complex accumulation was surprisingly high in that 80 90% of tional change that constricts the distal heme pocket (Fig. 11).

Downloaded from http://www.jbc.org/ by guest on December 16, 2016


the total enzymes were estimated to be in Compound II form The ability of SCN to modify the heme pocket of MPO and/or
and attenuated in a saturable manner to a level of saturation shield the catalytic site of the enzyme is directly reflected by
approaching 50% upon increasing SCN concentration. Be- the modulation of NO binding to the enzyme heme moiety.
cause Compound II is a catalytically inactive complex, a steady Varying the concentration of SCN displays different types of
but suboptimal rate of SCN oxidation and HOSCN synthesis impact on the MPO heme iron microenvironment. Indeed, the
is maintained that is proportional to the percentage of MPO plot of the second-order combination rate constant (kon) of NO
and EPO cycling through the 2e oxidation pathway. Indeed, binding to MPO-Fe(III) as a function of SCN concentration
the presence of SCN in the milieu resulted in a dramatic displayed a bell-shaped curve with a negative slope over the
acceleration in H2O2 consumption by both MPO and EPO, as range where SCN concentrations were 60 M and a positive
detected by continuous monitoring with a H2O2-selective elec- slope over the range where SCN concentrations were 60 M,
trode. In contrast, the LPO Compound II was unstable and with saturation 300 M. This behavior may have a broader
converted to ground state immediately after H2O2 exhaustion, link effect on MPO activity, because the initial decrease in the
indicating that SCN may serve as a 1e substrate for LPO second-order rate constant of NO combination (kon) occurs
Compound II. At a low SCN concentration, LPO Compound II within the range where SCN binds to a distal cavity located in
was also maintained at a constant level, whereas the enzyme MPO Compound I and allows the direct contact with the oxy-
continued to catalyze SCN oxidation and HOSCN synthesis. ferryl oxygen. Under these circumstances, Compound I appears
We observed that increasing the SCN concentration was ac- to act favorably in triggering electron transfer to the heme with
companied by a modest decrease in the rate of Compound II subsequent conversion to hypothiocyanate as a final reaction
formation, but with a significant decrease in accumulation, a product (16, 38, 39, 73). This decrease was reversed and
decrease in duration, and a significant increase in the decay reaches saturation at higher SCN concentrations, where
rate. LPO Compound II formation and the accompanying inhi- SCN binds to MPO at both the distal cavity and the proximal
bition did not occur at a higher SCN concentration. This helix sites and forms an inactive complex (SCN-E-Fe(III)(inh)).
behavior can be interpreted by either or both of the following Indeed, Blair-Johnson et al. (73) have shown that the inhibitory
statements. 1) The Compound II formation rate becomes slower complex with thiocyanate indicates replacement of chloride at
than the complex decay; and 2) the majority of the reaction the proximal helix halide binding site in addition to binding in
proceeds through a 2e oxidation pathway. the distal cavity in an orientation parallel with the heme (Fig.
A general kinetic scheme of how SCN interacts with mam- 11). It was thought that halides binding at the distal cavity site
malian peroxidase intermediates incorporated into the classic inhibit the enzyme by interfering with the deprotonation of
catalytic cycle is illustrated in Fig. 10. This working kinetic H2O2 by the adjacent distal His-95 (16, 38, 39, 73), a mecha-
model consists of two major pathways, namely the classical nism consistent with previous reports indicating that inhibi-
peroxidase cycle (pathway A) and the binding of a co-substrate tion by halides is competitive with respect to H2O2 (56 59).
to E-Fe(III), the formation of E(inh)-Fe(II) complex, and the Thus, the saturation in a bell-shaped curve occurs, at a high
effect of these bindings on the formation of Compound I (path- concentration of SCN (300 M), when the two sites are
way B) (Fig. 10). Compound I formation is rapid and occurs occupied. Therefore, the plasma levels of SCN (60 M) are
with a second-order rate constant ranging from 1 107 to 4.3 crucial in saturating the substrate site of the enzyme, as judged
107 M1 s1 (27, 33, 36, 37, 44, 46, 47). Mammalian peroxidase by the inflection point in the bell-shaped curve that is illus-
Compound I oxidizes halides and pseudohalides by a 2e oxi- trated in Fig. 2. Consequently, the plasma level of SCN may
dation process yielding the corresponding hypohalous acid and govern the catalytic reaction of MPO both in vivo and in vitro.
the ground state (E-Fe(III)) (33, 41 43). Alternatively, Com- Identifying similarities and differences in the interactions
pound I oxidizes organic and inorganic substrates by two se- between MPO and various members of the mammalian perox-
quential 1e oxidations to ground state through the formation idase superfamily yields valuable mechanistic insights into the
of Compound II, the rate-limiting step in the classic cycle of role of Compound II formation in modulating catalytic activity
peroxidases (33, 44). and function. The Compound I of MPO, EPO, and LPO simul-
NO, like CN, binds to the MPO heme iron and displaces the taneously catalyzes SCN via 2e and 1e oxidation pathways,
water molecule (W1), which is hydrogen-bonded to the distal which leads to the formation of HOSCN and sulfur-centered
Differences in the Mechanisms of Mammalian Peroxidases 26135
thiocyanate radicals (SCN), respectively (60, 61). The parti- II becomes slower than the complex decay. Collectively, the
tioning between the two pathways depends in part on the rate formation of LPO Compound II during steady-state catalysis is
of 2e versus 1e oxidation of SCN and on the concentration a fundamental feature of catalysis and functions to down-reg-
of SCN. ulate the peroxidation process by the enzyme.
It appears that as the SCN concentration increases, Com- The ability of MPO, EPO, and LPO to employ SCN as a 1e
pound I is converted rapidly and more efficiently to both substrate influences the nature of the end products of the
ground state and Compound II. This behavior is illustrated by SCN oxidation reaction. For example, 1e oxidation of SCN
the significant alteration in the intermediate distributions dur- generates a thiocyanate radical, which dimerizes to yield a
ing steady-state catalysis, as reflected by the decrease in the labile short-lived intermediate, thiocyanogen (SCN)2 (61, 63).
amplitude of Compound II formation observed in an SCN-de- This intermediate is rapidly hydrolyzed to generate a combi-
pendent and saturable manner (Fig. 6C). Thus, the plateau in nation of HOSCN and OSCN as final end products (61, 63).
the curve is governed by the rate of the 1e oxidation reaction These observations are consistent with earlier studies by Modi
of SCN by Compound I, which would then result from the et al., which suggested that LPO catalyzes SCN oxidation at
cleavage of SCN from the SCN-MPO(inh)-Fe(III) complex. pH 6.1 to produce HOSCN and OSCN (60). Alternatively,
Thus, the plot of the formation rate of Compound II plateaus at (SCN)2 is hydrolyzed to produce CN as an end product with-
a rate that should be comparable with the rate of SCN-E(inh)- out the formation of HOSCN (64, 65). In the presence of the
Fe(III) conversion to the active form, which, under these cir- plasma level of SCN, SCN may be directly oxidized to
cumstances, is the formation of E-Fe(III) and/or SCN-E-Fe(III) OSCN (60, 63, 66). It has been suggested that other products
(Fig. 10). The low conversion rate constant limits E-Fe(III) besides hypothiocyanite and CN might be formed during the
bioavailability and subsequently attenuates the turnover num- metabolism of SCN. For example, Pruitt et al. have proposed
ber of peroxidation. that hypothiocyanous acid might be oxidized by excess H2O2 or

Downloaded from http://www.jbc.org/ by guest on December 16, 2016


The time course for MPO Compound II decay monitored in MPO to yield two short-lived intermediate oxidants, cyanosul-
the absence of SCN, detected by monitoring the absorbance furous acid (HO2SCN) and cyanosulfuric acid (HO3SCN),
changes at 455 nm, is extremely slow and occurs with a rate which then break down in the presence of H2O2 to generate
constant of 0.008 s1 at 10 C (28 30). In the presence of SCN cyanate (OCN) (19, 67, 68).
and after the reaction had ceased, there was an initial rapid In related studies, Furtmuller et al. have demonstrated that
decay that was followed by a slower decay regenerating ferric the apparent second-order rate constants of the reactions be-
MPO. The initial rapid phase may be attributed to a rapid tween the LPO, EPO and MPO Compound I and SCN are
decay of Compound I to ground state, which increases in a extremely high, with apparent second-order constants of 2
linear manner as a function of SCN concentration with a 108, 1 108 and 9.6 107 M1 s1, respectively (36, 44, 47).
second-order rate constant of 1 104 M1s1, whereas the These results are consistent with the direct reaction between
second phase was attributed to the decay of MPO Compound II Compound I and SCN, which, to a large extent, prevents the
to ground state. Compound II does not possess suitable redox potential alteration in the heme pocket and, thereby, the for-
potential to oxidize halides and pseudohalides (33, 41). Rather, mation of the inactive form of the enzyme that occurs upon the
Compound II uses a variety of organic and inorganic com- prior incubation of SCN with the three enzymes.
pounds as substrates (33). Similar results were observed for Binding of the co-substrate to the heme moiety generating
EPO, indicating that both enzymes display similar mechanism the E-Fe(III)-SCN complex (Fig. 10) may considerably restrict
in metabolizing SCN. Van Dalen and Kettle have recently H2O2 access to the heme iron, and the slow rate of H2O2
reported similar spectral changes during the turnover of the binding observed would then result from ligand replacement
metabolism of SCN by EPO and MPO (62). Like them, we also processes. There is no data available to support the notion that
noted that the addition of GSH, a hypothiocyanite scavenger, SCN serves as a ligand in mammalian peroxidase when pres-
prevented the buildup of the Compound II intermediate; but, at ent at plasma levels. However, the presence of higher levels of
the same time GSH would not have contributed to the turnover SCN (e.g. 1 mM) may promote the formation of such a com-
of Compound II, because it did not reduce preformed MPO plex (45). Formation of E-Fe(III)-SCN can be characterized
Compound II (data not shown). Furthermore, Compound II based on its distinctive absorbance feature, which displays a
formation can be destabilized in the presence of organic and large red shift in the Soret absorbance region compared with
inorganic substrates (e.g. low levels of NO) (data not shown). native enzyme (45). Whether the slow binding of H2O2 to the
In contrast, LPO Compound II formation and the accompa- enzyme occurs through conformational changes in the heme
nying LPO inhibition did not occur under conditions where pocket geometry or because of the replacement reaction, the
SCN is in large excess. Thus, a majority of Compound I rate-limiting step becomes the binding of H2O2 to the enzymes.
quickly generates its ferric complex upon initiating the reac- Under all circumstances, the catalytic activity of the peroxi-
tion. The apparent ability of LPO to oxidize SCN through two dases will depend on multiple factors, including the concentra-
successive 1e transfers generating Compound II and LPO- tion of SCN versus H2O2, the affinity of the enzymes toward
Fe(III) is unprecedented. Increasing SCN concentration accel- H2O2 versus SCN, and the rate of SCN-E-Fe(III) and
erates the decay of the Compound II, a rate-limiting step in the E-Fe(III)-SCN breakdown (Fig. 10). Changes in heme pocket
peroxidase cycle, and enhances the overall rate of catalysis. geometry upon ligand binding have been described for cyto-
Enhancement of peroxidase catalysis due to the reduction of chrome C peroxidase (70), and the slower rates of NO binding
LPO Compound II to LPO-Fe(III) has been noted with other to the Fe(III) forms of a number of heme proteins have been
physiological reductants such as NO, superoxide, and ascorbic attributed to ligand replacement (69, 71, 72). Ligation of SCN
acid (28 30,33). That the rate of LPO Compound II formation to the distal heme would prevent access of H2O2 to the catalytic
remained relatively unchanged at the high SCN concentra- site of the enzymes.
tions studied (50 M SCN) is consistent with the reaction In summary, our current studies reveal for the first time that
being limited by the release of SCN from the SCN-LPO(inh) the formation of Compound II during steady-state catalysis is a
complex. Consistent with this interpretation, the failure of the fundamental feature of the kinetic reaction of mammalian per-
higher SCN concentration to alter the LPO spectrum ap- oxidases. Its formation during steady-state catalysis operates
peared to be linked to the fact that the formation of Compound to down-regulate the catalytic activity of MPO, EPO, and LPO.
26136 Differences in the Mechanisms of Mammalian Peroxidases
Our results also demonstrate that pre-incubation of SCN with 27. Marquez, L. A., and Dunford, H. B. (1995) J. Biol. Chem. 270, 30434 30440
28. Abu-Soud, H. M., and Hazen, S. L. (2000) J. Biol. Chem. 275, 37524 37532
mammalian peroxidases significantly affects their catalytic 29. Abu-Soud, H. M., and Hazen, S. L. (2001) Biochemistry 40, 1074710755
sites, subsequently altering heme iron reactivity, decreasing 30. Abu-Soud, H. M., Khassawneh, M. Y., Sohn, J. T., Murray, P., Haxhiu, M. A.,
and Hazen, S. L. (2001) Biochemistry 40, 11866 11875
the affinity toward H2O2, and disturbing the distribution of the 31. Hu, S., and Kincaid, J. R. (1991) J. Am. Chem. Soc. 113, 7189 7194
intermediates in a distinct and different manner. These find- 32. Sharonov, Y. A. (1995) FEBS Lett. 377, 512514
ings suggest a regulatory role for SCN in mediating particular 33. Kettle, A. J., and Winterbourn, C. C. (1997) Redox Rep. 3, 315
34. Ortiz de Montellano, P. R. (1992) Annu. Rev. Pharmacol. Toxicol. 32, 89 107
structure/function performance regardless of its bioavailability 35. Lardinois, O. M., Medzihradszky, K. F., and Ortiz de Montellano, P. R. (1999)
and function, which has the ultimate effect of promoting sub- J. Biol. Chem. 274, 3544135448
36. Furtmuller, P. G., Jantschko, W., Zederbauer, M., Jakopitsch, C., Arnhold, J.,
strate selectivity and specificity. Thus, the metabolism of and Obinger, C. (2004) Jpn. J. Infect. Dis. 57, S30 S31
SCN by mammalian peroxidases is determined by the follow- 37. Arnhold, J., Furtmuller, P. G., and Obinger, C. (2003) Redox Rep. 8, 179 186
ing four major factors: 1) the concentration of H2O2; 2) the 38. Fiedler T. J., Davey, C. A., and Fenna, R. E. (2000) J. Biol. Chem. 275,
11964 11971
concentration of SCN; 3) the stability of Compound II that 39. Davey, C. A., and Fenna, R. E. (1996) Biochemistry 35, 1096710973
formed during steady-state catalysis; and 4) the order of addi- 40. De Gioia, L., Ghibaudi, E. M., Laurenti, E., Salmona, M., and Ferrari, R. P.
(1996) J. Biol. Inorg. Chem. 1, 476 485
tion between the enzymes and their substrates (enzyme and 41. Hurst, J. K. (1991) in Peroxidases in Chemistry and Biology (Everse, J.,
H2O2 mixed first and SCN second, or enzyme and SCN Everse, K. E., and Grisham, M. B., eds) Vol. 1, pp. 37 62, CRC Press, Boca
mixed first and H2O2 second). Raton, FL
42. Marquez, L. A., Huang, J. T., and Dunford, H. B. (1994) Biochemistry 33,
14471454
REFERENCES 43. Marquez, L. A., Dunford, H. B., and Van Wart, H. (1990) J. Biol. Chem. 265,
1. Belding, M. E., Klebanoff, S. J., and Ray, C. G. (1970) Science 16, 195196 5666 5670
2. Kimura, S., and Ikeda-Saito, M. (1988) Proteins 3, 113120 44. Furtmuller, P. G., Burner, U., Regelsberger, G., and Obinger, C. (2000) Bio-
3. Jong, E. C., Henderson, W. R., and Klebanoff, S. J. (1980) J. Immunol. 124, chemistry 39, 15578 15584
1378 1382 45. Ferrari, R. P., Ghibaudi, E. M., Traversa, S., Laurenti, E., De Gioia, L., and
4. Jong, E. C., and Klebanoff, S. J. (1980) J. Immunol. 124, 1949 1953 Salmona M. (1997) J. Inorg. Biochem. 68, 1726

Downloaded from http://www.jbc.org/ by guest on December 16, 2016


5. Weiss, S. J., Test, S. T., Eckmann, C. M., Roos, D., and Regiani, S. (1986) 46. Dunford, H. B., and Hsuanyu, Y. (1999) Biochem. Cell Biol. 77, 449 457
Science 234, 200 203 47. Furtmuller, P. G., Jantschko, W., Regelsberger, G., Jakopitsch, C., Arnhold, J.,
6. Mayeno, A. N., Curran, A. J., Roberts, R. L., and Foote, C. S. (1989). J. Biol. and Obinger, C. (2002) Biochemistry 41, 1189511900
Chem. 264, 5660 5668 48. Thayer, J. R., and Huffaker, R. C. (1980) Anal. Biochem. 102, 110 119
7. Klebanoff, S. J., Waltersdorph, A. M., and Rosen, H. (1984) Methods Enzymol. 49. Rakita, R. M., Michel, B. R., and Rosen, H. (1990) Biochemistry 29, 10751080
105, 399 403 50. Wever, R., Plat, H., and Hamers, M. N. (1981) FEBS Lett. 123, 327331
8. Ten, R. M., Pease, L. R., McKean, D. J., Bell, M. P., and Gleich, G. J. (1989) J. 51. Agner, K. (1963) Acta Chem. Scand. 17, S332S338
Exp. Med. 169, 17571769 52. Jorg, A., Pasquier, J-M., and Kelebanoff, S. J. (1982) Biochim. Biophys. Acta
9. Ueda, T., Sakamaki, K., Kuroki, T., Yano, I., and Nagata, S. (1997) Eur. 701, 185191
J. Biochem. 243, 32 41 53. van Dalen, C. J., Whitehouse, M. W., Winterbourn, C. C., and Kettle, A. J.
10. Petrides, P. E. (1998) J. Mol. Med. 76, 688 698 (1997) Biochem. J. 327, 487 492
11. Klebanoff, S. J., and Clark, R. A. (1978) The Neutrophil: Functions and Clin- 54. Bolscher, B. G., Plat, H., and Wever, R. (1984) Biochim. Biophys. Acta 784,
ical Disorders, pp. 1 810, Elsevier Science Publishers B. V., Amsterdam 177186
12. Nauseef, W. M., and Malech, H. L. (1986) Blood 67, 1504 1507 55. Carlson, M. G., Peterson, C. G., and Venge, P. (1985) J. Immunol. 134,
13. Nauseef, W. M., Cogley, M., and McCormick, S. (1996) J. Biol. Chem. 271, 18751879
9546 9549 56. Stelmaszynska, T., and Zgliczynski, J. M. (1974) Eur. J. Biochem. 45, 305312
14. Dugad, L. B., La Mar, G. N., Lee, H. C., Ikeda-Saito, M., Booth, K. S., and 57. Zgliczynski, J. M., Selvaraj, R. J., Paul, B. B., Stelmaszynska, T., Poskitt,
Caughey, W. S. (1990) J. Biol. Chem. 265, 71737179 P. K. F., and Sbarra, A. J. (1977) Proc. Soc. Exp. Biol. Med. 154, 418 422
15. Taylor, K. T., Stroble, F., Yue, K. T., Ram, P., Pohl, J., Woods, A. S., and 58. Bakkenist, A. R. J., De Boer, J. E. G., Plat, H., and Wever, R. (1980) Biochim.
Kinkade, J. M., Jr. (1995) Arch. Biochem. Biophys. 316, 635 642 Biophys. Acta 613, 337348
16. Zeng, J., and Fenna, R. E. (1992) J. Mol. Biol. 226, 185207 59. Andrews, P. C., and Krinsky, N. I. (1982) J. Biol. Chem. 257, 13240 13245
17. Bolscher, B. G., Zoutberrg, G. R., Cuperus, R. A., and Wever, R. (1984) Bio- 60. Modi, S., Deodhar, S. S., Behere, D. V., and Mitra, S. (1991) Biochemistry 30,
chim. Biophys. Acta 784, 189 191 118 124
18. Caulfield, J. P., Korman, G., Butterworth, A. E., Hogan, M., and David, J. R. 61. Adak, S., Mazumdar, A., and Banerjee, R. K. (1997) J. Biol. Chem. 272,
(1980) J. Cell Biol. 86, 46 63 11049 11056
19. Arlandson, M., Decker, T., Roongta, V. A., Bonilla, L., Mayo, K. H., MacPher- 62. van Dalen, C. J., and Kettle, A. J. (2001) Biochem. J. 358, 233239
son, J. C., Hazen, S. L., and Slungaard, A. (2001) J. Biol. Chem. 276, 63. Aune, T. M., and Thomas, E. L. (1977) Eur. J. Biochem. 80, 209 214
215224 64. Hughes, M. N. (1975) in Chemistry and Biochemistry of Thiocyanic Acid and
20. Dull, T. J., Uyeda, C., Strosberg, A. D., Newdwin, G., and Seilhamer, J. J. Its Derivatives (Newman, A. A., ed.) pp 1 67, Academic Press, New York
(1990) DNA Cell Biol. 9, 499 509 65. Chung, J., and Wood, J. I. (1970) Arch. Biochem. Biophys. 141, 7378
21. Cals, M. M., Mailliart, P., Brignon, G., Anglade, P., and Dumas, B. R. (1991) 66. Thomas, E. L. (1981) Biochemistry 20, 32733280
Eur. J. Biochem. 198, 733739 67. Pruitt, K. M., and Tenovuo, J. (1982) Biochim. Biophys. Acta 704, 204 214
22. Ferrari, R. P., Laurenti, E., Cecchini, P. I., Gambino, O., and Sondergaard, I. 68. Pruitt, K. M., Tenovuo, J., Andrews, R. W., and McKane, T. (1982) Biochem-
(1995) J. Inorg. Biochem. 58, 109 127 istry 21, 562567
23. Carlstrom, A. (1969) Acta Chem. Scand. 23, 185202 69. Cooper, C. E. (1999) Biochim. Biophys. Acta 1411, 290 309
24. Reiter, B., and Perraudin, J. P. (1991) in Peroxidases in Chemistry and Biology 70. Barrick, D. (1995) Curr. Opin. Biotechnol. 6, 411 418
(Everse, J., Everse, K. E., and Grisham, M. B., eds) Vol. 1, pp. 143180, CRC 71. Traylor, T. G., and Sharma, V. S. (1992) Biochemistry 31, 2847 4849
Press, Boca Raton, FL 72. Abu-Soud, H. M., Wu, C., Ghosh, D. K., and Stuehr D. J. (1998) Biochemistry
25. Wolfson, L. M., and Sumner, S. S. (1993) J. Food Prot. 56, 887 892 37, 37773786
26. Hoogland, H., Dekker, H. L., van Riel, C., van Kuilenburg, A., Muijsers, A. O., 73. Blair-Johnson, M., Fiedler, T., and Fenna, R. (2001) Biochemistry 40,
and Wever, R. (1988) Biochim. Biophys. Acta 955, 337345 13990 13997
Thiocyanate Modulates the Catalytic Activity of Mammalian Peroxidases
Yahya R. Tahboub, Semira Galijasevic, Michael P. Diamond and Husam M. Abu-Soud
J. Biol. Chem. 2005, 280:26129-26136.
doi: 10.1074/jbc.M503027200 originally published online May 13, 2005

Access the most updated version of this article at doi: 10.1074/jbc.M503027200

Alerts:
When this article is cited
When a correction for this article is posted

Click here to choose from all of JBC's e-mail alerts

Downloaded from http://www.jbc.org/ by guest on December 16, 2016


This article cites 67 references, 18 of which can be accessed free at
http://www.jbc.org/content/280/28/26129.full.html#ref-list-1

Вам также может понравиться