ArticlePDF Available

Marla SS, Lee J, Groves JT. Peroxynitrite rapidly permeates phospholipid membranes. Proc Natl Acad Sci U S A.

Authors:

Abstract and Figures

Peroxynitrite (ONOO-) is a potent oxidant implicated in a number of pathophysiological processes. The activity of ONOO- is related to its accessibility to biological targets before its spontaneous decomposition (t1/2 approximately 1 s at pH 7.4, 37 degrees C). Using model phospholipid vesicular systems and manganese porphyrins as reporter molecules, we demonstrated that ONOO- freely crosses phospholipid membranes. The calculated permeability coefficient for ONOO- is approximately 8.0 x 10(-4) cm.s-1, which compares well with that of water and is approximately 400 times greater than that of superoxide. We suggest that ONOO- is a significant biological effector molecule not only because of its reactivity but also because of its high diffusibility.
Content may be subject to copyright.
Proc. Natl. Acad. Sci. USA
Vol. 94, pp. 14243–14248, December 1997
Chemistry
Peroxynitrite rapidly permeates phospholipid membranes
SUDHAKAR S. MARLA*, JINBO LEE*, AND JOHN T. GROVES
Department of Chemistry, Princeton University, Princeton, NJ 08544
Edited by Ronald Breslow, Columbia University, New York, NY, and approved October 14, 1997 (received for review August 22, 1997)
ABSTRACT Peroxynitrite (ONOO
2
) is a potent oxidant
implicated in a number of pathophysiological processes. The
activity of ONOO
2
is related to its accessibility to biological
targets before its spontaneous decomposition (t
1y2
' 1satpH
7.4, 37°C). Using model phospholipid vesicular systems and
manganese porphyrins as reporter molecules, we demon-
strated that ONOO
2
freely crosses phospholipid membranes.
The calculated permeability coefficient for ONOO
2
is '8.0 3
10
24
cmzs
21
, which compares well with that of water and is
'400 times greater than that of superoxide. We suggest that
ONOO
2
is a significant biological effector molecule not only
because of its reactivity but also because of its high diffus-
ibility.
Peroxynitrite (ONOO
2
) has emerged as an important member
of the family of reactive oxygen and nitrogen species (1–5)
since the recognition of its rapid formation from nitric oxide
(NO
2z
) and superoxide anion (O
2
2z
) (6) The production of
ONOO
2
in vivo has been demonstrated in the macrophage
immune response (7, 8) and under conditions of oxidative
stress such as ischemiayreperfusion (9–11). The reactions of
ONOO
2
with biological substrates are known to include the
nitration of tyrosine residues in proteins (12) and the oxidation
of redox metal centers (13, 14), DNA (15, 16), lipids (17),
sulfhydryls (18), and methionine (19). In light of this reactivity,
ONOO
2
has been implicated in a number of pathological
conditions including neurological disorders (20–23) such as
Alzheimer disease and amyotrophic lateral sclerosis, in ath-
erosclerosis (24, 25), and a variety of conditions precipitated
by endothelial injury (26). Furthermore, since nitration of
tyrosine has been shown to block tyrosine phosphorylation, a
key event in signal transduction cascades, the role of ONOO
2
as a signal molecule has been under investigation (27, 28). It
also has been demonstrated that ONOO
2
nitrates and inac-
tivates manganese superoxide dismutase in chronic rejection
of human renal allografts, which was proposed to be a general
mechanism for the amplification of ONOO
2
oxidative damage
(29). Given the short lifetime of ONOO
2
(t
1y2
' 1 s at pH 7.4,
37°C) (30), a diffusion distance of '100
m
m has been estimated
in physiological buffers (31). However, cells are compartmen-
talized into membrane-protected organelles (32). Thus, an
important determinant of toxicity or signaling effectiveness of
ONOO
2
will lie in its invasiveness and ability to access
biological targets. Here, we demonstrate that ONOO
2
can
diffuse freely across phospholipid membrane bilayers to react
with target substrates. Thus, the significance of ONOO
2
as a
biological effector molecule will derive not only from its
reactivity but also its diffusibility.
MATERIALS AND METHODS
Materials. 5,10,15,20-Tetrakis(N-methyl-49-pyridyl)por-
phinatomanganese(III) chloride [Mn(III)TMPyP(Cl)] and
Fe(III)TMPyP(Cl) were purchased from Midcentury (Posen,
TX). Peroxymonosulfate (HSO
5
2
or OXONE), hypochlorite
(OCl
2
), and 1-O-octyl-
b
-D-glucopyranoside were obtained
from Aldrich. Sephadex G25 was obtained from Pharmacia.
Anhydrous monosodium phosphate, anhydrous disodium
phosphate, Sepharose 4B, dimyristol
L-
a
-phosphatidylcholine
(DMPC), and Liposome Kit Positive (lyophilized powder
containing 63
m
mol of L-
a
-phosphatidylcholine, 18
m
mol of
stearyl amine, and 9
m
mol of cholesterol) were purchased from
Sigma. (
a
,
b
,
a
,
b
-Meso-tetrakis[O-((3
b
-hydroxy-5-cholenyl)-
amido)phenyl]-porphyrinato)manganese(III) chloride
[Mn(III)ChP(Cl)] was synthesized according to literature
methods (33). The sodium salt of ONOO
2
was prepared from
the reaction of acidic H
2
O
2
with sodium nitrite following a
published procedure (34). All solvents were analytical grade.
Water used in all the experiments was distilled and deionized
(Milli-Q, Millipore).
Vesicles Preparation. Small unilamellar vesicles (SUV) were
prepared by using the method of ultrasonication of DMPC thin
films containing Mn(III)ChP (35). The porphyrin was replaced
by retinoic acid (40
m
M) or tocopherol (80
m
M) in the lipid thin
films for SUV used in the Fe(III)TMPyP protection experi-
ments. Buffer solutions were added to the dry lipid thin films
and sonicated using a probe tip sonicator (Branson model 250)
for '5 min at 20–22°C. SUV generated were allowed to
equilibrate for 1y2 h before centrifugation to remove partic-
ulate matter. The final concentration of DMPC in the vesicular
solutions was 1.67 mM. The Mn(III)ChP concentration was
varied between 2 and 8
m
M.
Large unilamellar vesicles (LUV) were prepared using the
detergent removal method (36) with modifications. In brief, a
mixture of
L-
a
-phosphatidylcholine (63
m
mol), stearyl amine
(18
m
mol), and cholesterol (9
m
mol) was dissolved in a
phosphate buffer (50 mM, pH 7.4) solution containing Mn(II-
I)TMPyP (200
m
M) and 1-O-octyl-
b
-D-glucopyranoside (900
m
mol). This solution was applied to a Sephadex G-25 column
(1 3 20 cm) presaturated with an eluant consisting of Mn(II-
I)TMPyP (200
m
M) in phosphate buffer (50 mM, pH 7.4). The
exclusion of the detergent on the column resulted in the
formation of LUV containing Mn(III)TMPyP (200
m
M) in the
entrapped internal volume. The vesicles were eluted from the
column by using the same solution used to presaturate the
column. The vesicles were purified further on a Sepharose 4B
column (2 3 10 cm) using phosphate buffer (50 mM, pH 7.4).
Neutral LUV without the positively charged stearyl amine
were prepared similarly. Transmission electron micrographs of
the LUV with entrapped Mn(III)TMPyP and SUV were
obtained after fixing with 0.5% OsO
4
and staining with 2%
The publication costs of this article were defrayed in part by page charge
payment. This article must therefore be hereby marked ‘‘advertisement’’ in
accordance with 18 U.S.C. §1734 solely to indicate this fact.
© 1997 by The National Academy of Sciences 0027-8424y97y9414243-6$2.00y0
PNAS is available online at http:yywww.pnas.org.
This paper was submitted directly (Track II) to the Proceedings office.
Abbreviations: ONOO
2
, peroxynitrite; Mn(III)ChP(Cl), (
a
,
b
,
a
,
b
-
meso-tetrakis[O-((3
b
-hydroxy-5-cholenyl)amido) phenyl]-porphyri-
nato)manganese(III) chloride; Mn(III)TMPyP, 5,10,15,20-tetrakis(N-
methyl-49-pyridyl)porphinatomanganese(III) chloride; HSO
5
2
, per-
oxymonosulfate; SUV, small unilamellar vesicles; LUV, large
unilamellar vesicles.
*S.S.M. and J.L. contributed equally to this work.
To whom reprint requests should be addressed. e-mail: jtgroves@
princeton.edu.
14243
uranyl acetate. Typical SUV diameters were 35 nm, and the
LUV diameters were 150 nm.
Stopped-Flow Spectrophotometry. Time-resolved UV-vis
spectra were recorded on a HI-TECH SF-61 DX2 stopped-
flow spectrophotometer (HI-TECH, Salisbury, U.K.) by using
the photo-diode array fast scan mode. The spectral resolution
was '1 nm. The stoichiometric oxidation of Mn(III)TMPyP by
ONOO
2
was achieved by stopped-flow single mixing to give
final concentrations of 5
m
M Mn(III)TMPyP and 6.7
m
M
ONOO
2
in phosphate buffer (25 mM, pH 7.4). The reaction
was followed by UV-vis (00.5 s, 100 scans, 0.005-s integration
time). The kinetic profiles of the Mn(III)TMPyP oxidation by
ONOO
2
in solution or inside the LUV were collected by using
the stopped-flow photomultiplier mode. All measurements
were performed at ambient temperatures (22–24°C). The
reactions were conducted in either a phosphate buffer (25 mM,
pH 7.4) or a Tris buffer (25 mM, pH 8.2). Typically, the
concentration of Mn(III)TMPyP was held constant (2
m
M)
while the ONOO
2
concentration was varied (50–250
m
M) to
provide pseudo first-order conditions. For experiments involv-
ing Mn(III)TMPyP entrapped in LUV, the concentration of
Mn(III)TMPyP in the bulk solution was used (typically 2–2.5
m
M). The decay of Mn(III)TMPyP and the appearance of
oxoMn(IV) complexes were monitored at 462 nm and 428 nm,
respectively. The kinetic profiles could be fitted nicely into a
single exponential to give the pseudo first-order rate constants
[k (s
21
)] under different ONOO
2
concentrations. The oxi-
dants H
2
O
2
, perchlorate (OCl
2
), and peroxymonosulfate
(HSO
5
2
) were used for the control experiments.
The reactions of the membrane-bound Mn(III)ChP with
ONOO
2
also were performed in the stopped-flow photomul-
tiplier mode by following changes at 474 nm [Mn(III)ChP] and
FIG. 1. A schematic representation of the reporter manganese porphyrins in the model vesicular systems. Mn(III)TMPyP, when entrapped inside
the vesicles, is prevented from sampling the bulk volume during the experiments. Mn(III)ChP spans the membrane with the position of the
macrocycle at the center of the membrane and porphyrin plane parallel to the membrane surface (33). The reaction of the manganese(III) porphyrin
with peroxynitrite to afford oxomanganese(IV) and NO
2
is depicted below.
14244 Chemistry: Marla et al. Proc. Natl. Acad. Sci. USA 94 (1997)
424 nm [oxoMn(IV) species]. The initial rate of oxoMn(IV)
formation was linearly dependent on the ONOO
2
concentra-
tion. This allowed an estimation of the second-order rate
constant (k) for the oxidation reaction after taking into
account the spontaneous decay rates of ONOO
2
and the
oxoMn(IV) species.
Protection of Membrane Components by Fe(III)TMPyP.
Retinoic acid (40
m
M) and tocopherol (80
m
M) were incor-
porated into SUV preparations containing DMPC (1.67 mM).
The reactions of ONOO
2
with retinoic acid were followed at
340 nm, and those with tocopherol were followed at 300 nm.
The membrane-bound substrates were mixed with 250
m
M
ONOO
2
, which resulted in a complete loss of chromophore in
both cases indicating oxidation. For the protection experi-
ments, the concentration of Fe(III)TMPyP was varied between
2 and 10
m
M. Under these conditions, the ONOO
2
concen-
tration was exhausted quickly (t
1
y
2
5 0.16 2 0.035).
RESULTS AND DISCUSSION
The membrane permeability of ONOO
2
has been examined
in a phospholipid vesicular model system by exploiting the
diagnostic reactions of manganese porphyrins we have de-
scribed recently (37, 38). Manganese porphyrins known to
occupy different regions of the membrane system have been
used to assess the accessibility of ONOO
2
to these sites. Thus,
a membrane-spanning steroidal porphyrin [Mn(III)ChP] (33,
39) and a water-soluble [Mn(III)TMPyP] porphyrin were
deployed within the hydrophobic bilayer and inside the aque-
ous interior compartment, respectively, of unilamellar vesicles
(Fig. 1). Oxidation of Mn(III) to the corresponding oxo
Mn(IV) species was followed by observing the signature
changes in the UV-vis spectra on addition of ONOO
2
to the
medium.
The solution reaction of Mn(III)TMPyP with ONOO
2
was
found to be fast and stoichiometric by rapid mixing stopped-
flow spectrophotometry (Fig. 2). The depletion of Mn(III)-
TMPyP to generate the oxoMn(IV) species followed clear
stoichiometric behavior with isosbestic points at 389, 448, 536,
and 603 nm. Kinetic measurements determined that the in-
trinsic rate of Mn(III)TMPyP oxidation by ONOO
2
was 1.8 3
10
6
M
21
zs
21
at pH 7.4. Notably, the rates for the ONOO
2
oxidation of metalloporphyrins and heme peroxidases like
myeloperoxidase are the fastest reactions reported for
ONOO
2
(14). Control experiments showed that H
2
O
2
was
incapable of oxidizing Mn(III)TMPyP to oxoMn(IV) species
under these conditions.
The rate of oxidation of Mn(III)ChP by ONOO
2
,inthe
bilayer region, was much slower (k 5 3 3 10
4
M
21
zs
21
) but
demonstrated the ability of ONOO
2
to enter and react within
the hydrophobic region of the membrane (Fig. 1). Likewise,
retinoic acid and tocopherol were oxidized readily by ONOO
2
under these conditions. Mn(III)TMPyP redox coupled with
antioxidants, which catalytically reduces ONOO
2
, has been
shown to afford a dose-dependent protection of these lipid
components (40). Fe(III)TMPyP, which has been reported to
isomerize ONOO
2
to nitrate (41), also effectively prevented
the oxidation of these lipid components by ONOO
2
. Liposo-
mal assemblies of both manganese and iron porphyrin amphi-
philes have been found to be effective as well but only when the
molecular design of the membrane anchor allowed the charged
porphyrin moiety to occupy the aqueous region of the mem-
brane interface (42).
Significantly, the reaction of ONOO
2
with Mn(III)TMPyP
entrapped within LUV resulted in the efficient production of
the oxoMn(IV) species (Fig. 3A). Similar results were obtained
with the oxidant OCl
2
. By contrast, HSO
5
2
(43) failed to react
under these conditions (Fig. 3 B and C), demonstrating that the
LUV were impermeable to this less chaotropic and less basic
anion. The high membrane permeability of ONOO
2
and OCl
2
may be attributed to the relatively high pKa of their conjugate
acids, 6.8 (18, 19) and 7.4 (44), respectively. The order
described by the Eisenman sequence I for anions crossing
artificial membranes is: SCN
2
. I
2
. NO
3
2
. Br
2
. Cl
2
.
F
2
. CH
3
OSO
3
2
' CH
3
SO
3
2
' SO
4
22
(45). Whereas HSO
5
2
behaves like the membrane-impermeable CH
3
SO
3
2
, ONOO
2
resembles NO
3
2
in its permeability. Curiously, however, the
oxidation rate of the entrapped Mn(III)TMPyP by ONOO
2
showed little dependence on pH. The second-order rate
constants at pH 7.4 (25 mM, phosphate) and at pH 8.2 (25 mM
Tris) were k
obsd
5 0.8 3 10
6
M
21
zs
21
(R 5 0.996) and k
obsd
5
1.2 3 10
6
M
21
zs
21
(R 5 0.996), respectively. We also deter-
mined that the presence of the positively charged stearyl amine
did not assist in ONOO
2
transport across membranes because
its removal to generate neutral vesicles did not change the
observed rates of the oxidation reactions. We suggest that the
hydrophobic vesicular barrier may modulate the pKa of
ONOO
2
(46), thus facilitating the transport of the neutral
HOONO species across membranes.
Comparison of rates of reaction between Mn(III)TMPyP
and ONOO
2
in solution and in the Mn(III)TMPyPyLUV
system (Fig. 3D) showed that membrane entrapment retarded
the rate of manganese oxidation by only 2-fold. Accordingly,
the barrier offered to ONOO
2
transport by the membrane is
small. This rate difference allows an estimate of the perme-
ability coefficient (P) for ONOO
2
crossing a membrane
bilayer using Fick’s Law. The permeability coefficient is de-
fined as P 5 JyDc, where J (molzcm
22
zs
21
) is the flux of
ONOO
2
across unit surface area of the membrane and Dc
(molyL) is the difference in the concentration ([ONOO
2
]
out
2
FIG. 2. Time resolved UV-vis profile of the stoichiometric oxida-
tion of Mn(III)TMPyP by ONOO
2
to generate the oxoMn(IV)
species. Stopped-flow mixing was used to give final concentrations of
Mn(III)TMPyP (5
m
M) and ONOO
2
(6.7
m
M) in 25 mM phosphate
(pH 7.4) buffer. The reaction was followed by UV-vis (00.5 s, 100
scans, 0.005-s integration time) by using a HI-TECH SF-61 DX2
rapid-mixing stopped-flow spectrophotometer (1 in 10 scans shown for
clarity). The rate constant for the oxidation of Mn(III)TMPyP by
ONOO
2
obtained under pseudo first-order conditions at pH 7.4 was
k
intrinsic
5 1.8 3 10
6
M
21
zs
21
(see Fig. 3D).
Chemistry: Marla et al. Proc. Natl. Acad. Sci. USA 94 (1997) 14245
FIG. 3. ONOO
2
is shown to cross membranes freely. Difference spectra of the UV-vis absorbances before and after addition of oxidants to
LUV—encapsulated Mn(III)TMPyP. (A) ONOO
2
(1 mM) addition resulted in an instantaneous oxidation of Mn(III) (462 nm) to the oxoMn
species (428 nm), indicating that ONOO
2
diffused rapidly across the lipid bilayer. Rapid oxidation was seen between pH 7.4 and 8.8, suggesting
that the ONOO
2
anion as well as HOONO can cross membranes (pKa ONOO
2
6.8). Similar results were obtained with OCl
2
.(B) Addition of
HSO
5
2
(1 mM) to the Mn(III)TMPyPyLUV resulted in no oxidation, verifying that the LUV prevented intermixing of the entrapped porphyrin
and HSO
5
2
in this construct. (C) Oxidation by HSO
5
2
was observed only after the LUV had been sonicated to relocate Mn(III)TMPyP into the
bulk solution. (D) Comparison of the pseudo first-order rates for the solution reaction compared with the reaction in LUV. All of the experiments
were conducted using rapid mixing stopped-flow methods as described in Materials and Methods. Reaction rates obtained from the slope of the
pseudo first-order plots were: k
intrinsic
5 1.8 3 10
6
M
21
zs
21
(R 5 0.999); and k
obsd
5 0.8 3 10
6
M
21
zs
21
(R 5 0.996).
14246 Chemistry: Marla et al. Proc. Natl. Acad. Sci. USA 94 (1997)
[ONOO
2
]
in
). The ratio of these concentrations is directly
related to k
obsd
and k
intrinsic
, the observed and intrinsic rate
constants for the reaction of ONOO
2
with Mn(III)TMPyP we
have measured (Eq. 1).
[ONOO
2
]
in
[ONOO
2
]
out
5
k
obsd
k
intrinsic
5 0.48 [1]
In the steady state, the membrane crossing rate of ONOO
2
was
taken to be equal to the rate of the oxidation of the entrapped
porphyrin inside the LUV, reflecting the barrier offered by the
lipid bilayer to the diffusion of ONOO
2
(Eq. 2). Thus, P is
expressed as in Eq. 3, where A is the surface area of vesicles,
and V
in
is the volume entrapped in LUV.
d
dt
@ONOO
2
]
flux
5 k
intrinsic
z[MNTMPyP]
in
z[ONOO
2
]
in
[2]
P 5
d
dt
[ONOO
2
]
flux
z
V
in
AzDc
5
k
intrinsic
z[MnTMPyP]
in
z[ONOO
2
]
in
zV
in
AzDc
[3]
Calculated values for A 5 4.0 3 10
6
cm
2
yL and V
in
5 10 mlyL
derive from the observed radius of the vesicles and afforded a
value for P 5 8.0 3 10
24
cmzs
21
at pH 7.4. Simulations using
this permeability coefficient agreed well with the experimental
data. A simple, sequential reaction model was assumed, taking
into consideration the ONOO
2
diffusion across phospholipid
membrane and subsequent reaction with the entrapped
Mn(III)-TMPyP.
We also have considered the recently reported reaction
between ONOO
2
and CO
2
to generate an unstable interme-
diate, presumably ONOOCO
2
2
(k ' 1 3 10
4
M
21
zs
21
) (47, 48).
Because the physiological concentration of HCO
3
2
is relatively
high ('25 mM) and is in equilibrium with CO
2
, this reaction
has been suggested to be an important pathway for ONOO
2
and a determinant of toxicity. However, the rate of diffusion
for ONOO
2
crossing lipid bilayers determined here is at least
30 times faster than the ONOO
2
reaction with the physiolog-
ical concentration of CO
2
(see Table 1). Preliminary experi-
ments showed that the oxidation of manganese by ONOO
2
in
the Mn(III)TMPyPyLUV system remained unaffected by 25
mM HCO
3
2
, indicating further that the transmembrane diffu-
sion of ONOO
2
is faster than its reaction with CO
2
. Therefore,
we conclude that the transmembrane diffusion of ONOO
2
is
considerably faster than all documented reaction rates of
ONOO
2
with its biological targets.
Significantly, the permeability coefficient P obtained for
ONOO
2
in this way is close to that reported for water (49) and
is '400 times faster than that of superoxide (50, 51) (Table 1).
Accordingly, the availability of superoxide across biological
membranes is limited both by its relatively slow diffusion and
the high concentrations of superoxide dismutase within the
cell (25, 52, 53). The high membrane permeability observed
here for ONOO
2
may have important consequences regarding
its biochemistry and pathophysiology. Thus, in marked con-
trast to superoxide ion (54), ONOO
2
can be expected to have
free access to cell interiors across membranes and over dis-
tances of cellular dimensions, necessary properties for a bio-
logical effector molecule.
We thank Dr. M. K. Stern, Monsanto Chemical Co., for stimulating
discussions. Support of this research by the National Institutes of
Health (GM36928) is acknowledged gratefully.
1. Beckman, J. S. & Koppenol, W. H. (1996) Am. J. Physiol. 271,
C1424–C1437.
2. Beckman, J. S. (1996) Chem. Res. Toxicol. 9, 836844.
3. Rubbo, H., Darley-Usmar, V. & Freeman, B. A. (1996) Chem.
Res. Toxicol. 9, 809820.
4. Radi, R. (1996) Chem. Res. Toxicol. 9, 828835.
5. Khan, A. U. & Wilson, T. (1995) Chem. Biol. 2, 437–445.
6. Huie, R. E. & Padmaja, S. (1993) Free Radical Res. Commun. 18,
195–199.
7. Ischiropoulos, H., Zhu, L. & Beckman, J. S. (1992) Arch.
Biochem. Biophys. 298, 446451.
8. Evans, T. J., Buttery, L. D. K., Carpenter, A., Springall, D. R.,
Polak, J. M. & Cohen, J. (1996) Proc. Natl. Acad. Sci. USA 93,
9553–9558.
9. Beckman, J. S., Beckman, T. W., Chen, J., Marshall, P. A. &
Freeman, B. A. (1990) Proc. Natl. Acad. Sci. USA 87, 1620–1624.
10. Carreras, M. C., Pargament, G. A., Catz, S., Poderosso, J. J. &
Boveris, A. (1994) FEBS. Lett. 341, 65–68.
11. Huang, Z., Huang, P. L., Panahian, N., Dalkara, T., Fishman,
M. C. & Moskowitz, M. A. (1994) Science 265, 1883–1885.
12. Ischiropoulos, H., Zhu, L., Chen, J., Tsai, M., Martin, J. C., Smith,
C. D. & Beckman, J. S. (1992) Arch. Biochem. Biophys. 298,
431–437.
13. Castro, L., Rodriguez, M. & Radi, R. (1994) J. Biol. Chem. 269,
29409–29415.
14. Floris, R., Piersma, R., Yang, G., Jones, P. & Wever, R. (1993)
Eur. J. Biochem. 215, 767–775.
15. King, P. A., Anderson, V. E., Edwards, J. O., Gustafson, G.,
Plumb, R. C. & Suggs, J. W. (1992) J. Am. Chem. Soc. 114,
5430–5432.
16. King, P. A., Jamison, E., Strahs, D., Anderson, V. E. & Brenow-
itz, M. (1993) Nucleic Acids Res. 21, 2473–2478.
17. Radi, R., Beckman, J. S., Bush, K. M. & Freeman, B. A. (1991)
Arch. Biochem. Biophys. 288, 481–487.
18. Radi, R., Beckman, J. S., Bush, K. M. & Freeman, B. A. (1991)
J. Biol. Chem. 266, 42444250.
19. Pryor, W. A., Jin, X. & Squadrito, G. L. (1994) Proc. Natl. Acad.
Sci. USA 91, 11173–11177.
20. Lipton, S. A., Chol, T.-B., Pan, Z.-H., Lei, S. Z., Chen, H.-S.,
Sucher, N. J., Loscalzo, J., Singel, D. J. & Stamler, J. S. (1993)
Nature (London) 364, 626631.
21. Beckman, J. S., Carson, M., Smith, C. D. & Koppenol, W. H.
(1993) Nature (London) 364, 584.
22. Good, P. F., Werner, P., Hsu, A., Olanow, C. W. & Perl, D. P.
(1996) Am. J. Pathol. 149, 21–28.
23. Wiedau-Pazos, M., Goto, J. J., Rabizadeh, S., Gralla, E. B., Roe,
J. A., Lee, M. K., Valentine, J. S. & Bredesen, D. E. (1996)
Science 271, 515–518.
24. Hogg, N., Darley-Usmar, V. M., Graham, A. & Moncada, S.
(1993) Biochem. Soc. Trans. 21, 358–362.
25. White, C. R., Brock, T. A., Chang, L., Crapo, J., Briscoe, P., Ku,
D., Bradley, W. A., Gianturco, S. H., Gore, J., Freeman, B. A. &
Tarpey, M. M. (1994) Proc. Natl. Acad. Sci. USA 91, 1044–1048.
26. Haddad, I. Y., Pataki, G., Hu, P., Beckman, J. S. & Matalon, S.
(1994) J. Clin. Invest. 94, 2407–2413.
27. Kong, S.-K., Yim, M. B., Stadtman, E. R. & Chock, P. B. (1996)
Proc. Natl. Acad. Sci. USA 93, 3377–3382.
28. Berlett, B. S., Friguet, B., Yim, M. B., Chock, P. B. & Stadtman,
E. R. (1996) Proc. Natl. Acad. Sci. USA 93, 1776–1780.
29. MacMillan-Crow, L. A., Crow, J. P., Kerby, J. D., Beckman, J. S.
& Thompson, J. A. (1996) Proc. Natl. Acad. Sci. USA 93,
11853–11858.
Table 1. Permeability across lipid bilayers
P (cmzs
21
) k
diff
5
AzP
V
in
(s
21
)
H
2
O 2.3 3 10
23
920*
ONOO
2
8.0 3 10
24
320
O
2
2z
2.1 3 10
26
0.84*
A, membrane surface area (cm
2
yL); P, permeability coefficient
(cmzs
21
); k
diff
, rate of ONOO
2
exchange across the membrane; V
in
,
entrapped volume (mLyL).
*These values were calculated by using A and V
in
parameters of our
LUV system and published values of permeability coefficients (49
51).
Chemistry: Marla et al. Proc. Natl. Acad. Sci. USA 94 (1997) 14247
30. Pfeiffer, S., Gorren, A. C. F., Schmidt, K., Werner, E. R.,
Hansert, B., Bohle, D. S. & Mayer, B. (1997) J. Biol. Chem. 272,
3465–3470.
31. Beckman, J. S. (1996) in Nitric Oxide: Principles and Actions, ed.
Lancaster, J. (Academic, San Diego), pp. 43–44.
32. Alberts, B., Bray, D., Lewis, J., Raff, M., Roberts, K. &
Watson, J. D. (1989) Molecular Biology of the Cell (Garland,
New York).
33. Groves, J. T. & Neumann, R. (1989) J. Am. Chem. Soc. 111,
2900–2909.
34. Pape´e, H. M. & Petriconi, G. L. (1964) Nature (London) 204,
142–144.
35. Huang, C. (1969) Biochemistry 8, 344–352.
36. Mimms, L. T., Zampighi, G., Nozaki, Y., Tanford, C. & Reynolds,
J. A. (1981) Biochemistry 20, 833–840.
37. Groves, J. T., Lee, J. & Marla, S. S. (1997) J. Am. Chem. Soc. 119,
62696273.
38. Groves, J. T. & Marla, S. S. (1995) J. Am. Chem. Soc. 117,
9578–9579.
39. Groves, J. T. & Neumann, R. (1987) J. Am. Chem. Soc. 109,
5045–5047.
40. Lee, J., Hunt, J. A. & Groves, J. T. (1997) Bioorg. Med. Chem.
Lett. 7, 2913–2918.
41. Stern, M. K., Jensen, M. P. & Kramer, K. (1996) J. Am. Chem.
Soc. 118, 8735–8736.
42. Hunt, J. A., Lee, J. & Groves, J. T. (1997) Chem. Biol. 4, 845–858.
43. Bernadou, J., Fabiano, A.-S., Robert, A. & Meunier, B. (1994)
J. Am. Chem. Soc. 116, 9375–9376.
44. Sanderson, R. T. (1989) in Simple Inorganic Substances, ed.
Sanderson, R. T. (Robert E. Krieger Publishing, Malabar, FL),
p. 298.
45. Wright, E. M. & Diamond, J. M. (1977) Physiol. Rev. 57, 109–156.
46. Pryor, W. A., Jin, X. & Squadrito, G. L. (1996) J. Am. Chem. Soc.
118, 3125–3128.
47. Lymar, S. V. & Hurst, J. K. (1995) J. Am. Chem. Soc. 117, 8867.
48. Lymar, S. V., Jiang, Q. & Hurst, J. K. (1996) Biochemistry 35,
7855–7861.
49. Finkelstein, A. & Cass, A. (1967) Nature (London) 216, 717–718.
50. Takahashi, M.-A. & Asada, K. (1983) Arch. Biochem. Biophys.
226, 558–566.
51. Mao, G. D. & Poznansky, M. J. (1992) FEBS. Lett. 305, 233–236.
52. Halliwell, B. & Gutteridge, J. M. C. (1990) Methods Enzymol. 186,
1–85.
53. Chevion, M. (1988) Free Radical Biol. Med. 5, 27–37.
54. Crow, J. P. & Beckman, J. S. (1995) in Advances in Pharmacology,
eds. Ignarro, L. & Murad, F. (Academic, San Diego), Vol. 35, pp.
17–43.
14248 Chemistry: Marla et al. Proc. Natl. Acad. Sci. USA 94 (1997)
Article
As investigations in the biomedical applications of plasma advance, a demand for describing safe and efficacious delivery of plasma is emerging. It is quite clear that not all plasmas are “equal” for all applications. This Perspective discusses limitations of the existing parameters used to define plasma in context of the need for the “right plasma” at the “right dose” for each “disease system.” The validity of results extrapolated from in vitro studies to preclinical and clinical applications is discussed. We make a case for studying the whole system as a single unit, in situ. Furthermore, we argue that while plasma-generated chemical species are the proposed key effectors in biological systems, the contribution of physical effectors (electric fields, surface charging, dielectric properties of target, changes in gap electric fields, etc.) must not be ignored.
Article
Full-text available
Questions about which reactive oxygen species (ROS) or reactive nitrogen species (RNS) can escape from the mitochondria and activate signals must be addressed. In this study, two parameters, the calculated dipole moment (debye, D) and permeability coefficient (Pm) (cm s−1), are listed for hydrogen peroxide (H2O2), hydroxyl radical (•OH), superoxide (O2•−), hydroperoxyl radical (HO2•), nitric oxide (•NO), nitrogen dioxide (•NO2), peroxynitrite (ONOO−), and peroxynitrous acid (ONOOH) in comparison to those for water (H2O). O2•− is generated from the mitochondrial electron transport chain (ETC), and several other ROS and RNS can be generated subsequently. The candidates which pass through the mitochondrial membrane include ROS with a small number of dipoles, i.e., H2O2, HO2•, ONOOH, •OH, and •NO. The results show that the dipole moment of •NO2 is 0.35 D, indicating permeability; however, •NO2 can be eliminated quickly. The dipole moments of •OH (1.67 D) and ONOOH (1.77 D) indicate that they might be permeable. This study also suggests that the mitochondria play a central role in protecting against further oxidative stress in cells. The amounts, the long half-life, the diffusion distance, the Pm, the one-electron reduction potential, the pKa, and the rate constants for the reaction with ascorbate and glutathione are listed for various ROS/RNS, •OH, singlet oxygen (1O2), H2O2, O2•−, HO2•, •NO, •NO2, ONOO−, and ONOOH, and compared with those for H2O and oxygen (O2). Molecules with negative electrical charges cannot directly diffuse through the phospholipid bilayer of the mitochondrial membranes. Short-lived molecules, such as •OH, would be difficult to contribute to intracellular signaling. Finally, HO2• and ONOOH were selected as candidates for the ROS/RNS that pass through the mitochondrial membrane.
Chapter
Full-text available
Xanthine oxidoreductase (XOR) is a molybdoflavin enzyme which occurs in two forms; the reduced form known as xanthine dehydrogenase (XDH, EC 1.17.1.4) and the oxidised form known as xanthine oxidase (XO, EC 1.17.3.2). In humans, it is a 293 kDa homodimer which catalyses consecutive hydroxylation steps of purine degradation. The oxidised form of the enzyme produces hydrogen peroxide and superoxide (O2•−), both of which are reactive oxygen species (ROS) that can interact with several biomolecules producing adverse reactions. XOR can also produce nitric oxide, a cardiovascular protective molecule. Overproduction of nitric oxide results in the formation of the highly reactive peroxynitrite radical. XOR-produced ROS may provide protection against infection, while at the same time can also lead to inflammation, oncogenesis, brain injury and stroke. XOR is also involved in tumour lysis syndrome in chemotherapy patients as well in ischaemia-reperfusion injury, increasing the levels of ROS in the body. Consequently, the presence of XOR in blood can be used as a biomarker for a number of conditions including oxidative stress and cardiovascular disease.
Article
Boronic acid-containing fluorescent molecules have been widely used to sense hydrogen peroxide and peroxynitrite, which are important reactive oxygen and nitrogen species in biological systems. However, it has been challenging to gain specificity. Our previous studies developed genetically encoded, green fluorescent peroxynitrite biosensors by genetically incorporating a boronic acid-containing noncanonical amino acid (ncAA), p-boronophenylalanine (pBoF), into the chromophore of circularly permuted green fluorescent proteins (cpGFPs). In this work, we introduced pBoF to amino acid residues spatially close to the chromophore of an enhanced circularly permuted red fluorescent protein (ecpApple). Our effort has resulted in two responsive ecpApple mutants: one bestows reactivity toward both peroxynitrite and hydrogen peroxide, while the other, namely, pnRFP, is a selective red fluorescent peroxynitrite biosensor. We characterized pnRFP in vitro and in live mammalian cells. We further studied the structure and sensing mechanism of pnRFP using X-ray crystallography, 11B-NMR, and computational methods. The boron atom in pnRFP adopts an sp2-hybridization geometry in a hydrophobic pocket, and the reaction of pnRFP with peroxynitrite generates a product with a twisted chromophore, corroborating the observed "turn-off" fluorescence response. Thus, this study extends the color palette of genetically encoded peroxynitrite biosensors, provides insight into the response mechanism of the new biosensor, and demonstrates the versatility of using protein scaffolds to modulate chemoreactivity.
Article
Full-text available
Marine phytoplankton produce and scavenge Reactive Oxygen Species, to support cellular processes, while limiting damaging reactions. Some prokaryotic picophytoplankton have, however, lost all genes encoding scavenging of hydrogen peroxide. Such losses of metabolic function can only apply to Reactive Oxygen Species which potentially traverse the cell membrane outwards, before provoking damaging intracellular reactions. We hypothesized that cell radius influences which elements of Reactive Oxygen Species metabolism are partially or fully dispensable from a cell. We therefore investigated genomes and transcriptomes from diverse marine eukaryotic phytoplankton, ranging from 0.4 to 44 μm radius, to analyze the genomic allocations encoding enzymes metabolizing Reactive Oxygen Species. Superoxide has high reactivity, short lifetimes and limited membrane permeability. Genes encoding superoxide scavenging are ubiquitous across phytoplankton, but the fractional gene allocation decreased with increasing cell radius, consistent with a nearly fixed set of core genes for scavenging superoxide pools. Hydrogen peroxide has lower reactivity, longer intracellular and extracellular lifetimes and readily crosses cell membranes. Genomic allocations to both hydrogen peroxide production and scavenging decrease with increasing cell radius. Nitric Oxide has low reactivity, long intracellular and extracellular lifetimes and readily crosses cell membranes. Neither Nitric Oxide production nor scavenging genomic allocations changed with increasing cell radius. Many taxa, however, lack the genomic capacity for nitric oxide production or scavenging. The probability of presence of capacity to produce nitric oxide decreases with increasing cell size, and is influenced by flagella and colony formation. In contrast, the probability of presence of capacity to scavenge nitric oxide increases with increasing cell size, and is again influenced by flagella and colony formation.
Article
Zinc is a transition metal that displays wide physiological implications ranging from participation in hundreds of enzymes and proteins to normal growth and development. In the reproductive tract of both sexes, zinc maintains a functional role in spermatogenesis, ovulation, fertilization, normal pregnancy, fetal development, and parturition. In this work, we review evidence to date regarding the importance of zinc in oocyte maturation and development, with emphasis on the role of key zinc-binding proteins, as well as examine the effects of zinc and reactive oxygen species (ROS) on oocyte quality and female fertility. We summarize our current knowledge about the participation of zinc in the developing oocyte bound to zinc finger proteins as well as loosely bound zinc ion in the intracellular and extracellular environments. These include aspects related to (1) the impact of zinc deficiency and overwhelming production of ROS under inflammatory conditions on the offset of the physiological antioxidant machinery disturbing biomolecules, proteins, and cellular processes, and their role in contributing to further oxidative stress; (2) the role of ROS in modulating damage to proteins containing zinc, such as zinc finger proteins and nitric oxide synthases (NOS), and expelling the zinc resulting in loss of protein function; and (3) clarify the different role of oxidative stress and zinc deficiency in the pathophysiology of infertility diseases with special emphasis on endometriosis-associated infertility.
Article
Full-text available
Paracrine superoxide (O2•−) and hydrogen peroxide (H2O2) signaling critically depends on these substances' concentrations, half-lives and transport ranges in extracellular media. Here we estimated these parameters for the lumen of human capillaries, arterioles and arteries using reaction-diffusion-advection models. These models considered O2•− and H2O2 production by endothelial cells and uptake by erythrocytes and endothelial cells, O2•− dismutation, O2•− and H2O2 diffusion and advection by the blood flow. Results show that in this environment O2•− and H2O2 have half-lives <60. ms and <40. ms, respectively, the former determined by the plasma SOD3 activity, the latter by clearance by endothelial cells and erythrocytes. H2O2 concentrations do not exceed the 10 nM scale. Maximal O2•− concentrations near vessel walls exceed H2O2's several-fold when the latter results solely from O2•− dismutation. Cytosolic dismutation of inflowing O2•− may thus significantly contribute to H2O2 delivery to cells. O2•− concentrations near vessel walls decay to 50% of maximum 12 μm downstream from O2•− production sites. H2O2 concentrations in capillaries decay to 50% of maximum 22 μm (6.0 μm) downstream from O2•− (H2O2) production sites. Near arterioles' (arteries') walls, they decay by 50% within 6.0 μm (4. μm) of H2O2 production sites. However, they reach maximal values 50 μm (24 μm) downstream from O2•− production sites and decrease by 50% over 650 μm (500 μm). Arterial/olar endothelial cells might thus signal over a mm downstream through O2•−-derived H2O2, though this requires nM-sensitive H2O2 transduction mechanisms.
Article
Full-text available
Ischemic stroke is a severe cerebrovascular disease with high mortality and morbidity. In recent years, reperfusion treatments based on thrombolytic and thrombectomy are major managements for ischemic stroke patients, and the recanalization time window has been extended to over 24 h. However, with the extension of the time window, the risk of ischemia/reperfusion (I/R) injury following reperfusion therapy becomes a big challenge for patient outcomes. I/R injury leads to neuronal death due to the imbalance in metabolic supply and demand, which is usually related to mitochondrial dysfunction. Mitophagy is a type of selective autophagy referring to the process of specific autophagic elimination of damaged or dysfunctional mitochondria to prevent the generation of excessive reactive oxygen species (ROS) and the subsequent cell death. Recent advances have implicated the protective role of mitophagy in cerebral ischemia is mainly associated with its neuroprotective effects in I/R injury. This review discusses the involvement of mitochondria dynamics and mitophagy in the pathophysiology of ischemic stroke and I/R injury in particular, focusing on the therapeutic potential of mitophagy regulation and the possibility of using mitophagy-related interventions as an adjunctive approach for neuroprotective time window extension after ischemic stroke.
Article
Full-text available
Cold atmospheric plasma (CAP) and plasma-treated liquids (PTLs) have recently become a promising option for cancer treatment, but the underlying mechanisms of the anti-cancer effect are still to a large extent unknown. Although hydrogen peroxide () has been recognized as the major anti-cancer agent of PTL and may enable selectivity in a certain concentration regime, the co-existence of nitrite can create a synergistic effect. We develop a mathematical model to describe the key species and features of the cellular response toward PTL. From the numerical solutions, we define a number of dependent variables, which represent feasible measures to quantify cell susceptibility in terms of the membrane diffusion rate constant and the intracellular catalase concentration. For each of these dependent variables, we investigate the regimes of selective versus non-selective, and of synergistic versus non-synergistic effect to evaluate their potential role as a measure of cell susceptibility. Our results suggest that the maximal intracellular concentration, which in the selective regime is almost four times greater for the most susceptible cells compared to the most resistant cells, could be used to quantify the cell susceptibility toward exogenous . We believe our theoretical approach brings novelty to the field of plasma oncology, and more broadly, to the field of redox biology, by proposing new ways to quantify the selective and synergistic anti-cancer effect of PTL in terms of inherent cell features.
Article
Ankylosing spondylitis (AS) is a chronic auto-immune disease, affecting the spine, sacroiliac, and sometimes peripheral joints. It is also involved with cardio-vascular risk factors due to accelerated atherosclerosis. Oxidative burst, systemic inflammation coupled with endothelial dysfunction (ED), resulting in reduced bioavailability of the vasodilator nitric oxide (NO) and an increased number of circulating endothelial cells (CECs) may correlate with disease activity and its sustenance. Hence, the study was aimed to detect and quantify CECs and assess the oxidative stress and inflammatory status in AS patients vis-à-vis healthy controls, as well as relate these parameters with AS disease activity and atherosclerotic markers in patients. Our study showed an increased frequency of endothelial cells in peripheral blood of AS patients in pro-inflammatory conditions. In AS patient population, they showed significant reduction of flow-mediated dilatation (%FMD) (p < 0.05), and increased soluble adhesion molecules such as sICAM-1 (p < 0.01) and sVCAM-1 (p < 0.05) compared to healthy controls. A marked increase in pro-inflammatory markers such as TNF-α (p < 0.01) and IL-1β (p < 0.001) and reactive free radicals (p < 0.05) along with reduced serum nitrite in AS, provided a strong pro-inflammatory milieu which positively correlated with Bath ankylosing spondylitis disease activity and functional indices (BASDAI and BASFI). The observed significant upregulation in CECs (CD45⁻/CD31⁺/CD105⁺/CD144⁺) in patients compared to healthy controls positively correlated with disease activity and duration as well as with markers of oxidative stress. Thus, chronic inflammation and oxidative burst induce loss of NO bioavailability, leading to ED. This may cause the derangement of CECs that may be considered as a prognostic biomarker for ED.
Article
Full-text available
Peroxynitrite anion (ONOO-) is a potent oxidant that mediates oxidation of both nonprotein and protein sulfhydryls. Endothelial cells, macrophages, and neutrophils can generate superoxide as well as nitric oxide, leading to the production of peroxynitrite anion in vivo. Apparent second order rate constants were 5,900 M⁻¹.s⁻¹ and 2,600-2,800 M⁻¹.s⁻¹ for the reaction of peroxynitrite anion with free cysteine and the single thiol of albumin, respectively, at pH 7.4 and 37 °C. These rate constants are 3 orders of magnitude greater than the corresponding rate constants for the reaction of hydrogen peroxide with sulfhydryls at pH 7.4. Unlike hydrogen peroxide, which oxidizes thiolate anion, peroxynitrite anion reacts preferentially with the undissociated form of the thiol group. Peroxynitrite oxidizes cysteine to cystine and the bovine serum albumin thiol group to an arsenite nonreducible product, suggesting oxidation beyond sulfenic acid. Peroxynitrous acid was a less effective thiol-oxidizing agent than its anion, with oxidation presumably mediated by the decomposition products, hydroxyl radical and nitrogen dioxide. The reactive peroxynitrite anion may exert cytotoxic effects in part by oxidizing tissue sulfhydryls.
Article
Mn(III)TMPyP reacts rapidly with the toxic oxidant peroxynitrite (ONOO−) to generate an oxoMn(IV) species, but Mn(III)TMPyP is not catalytic for ONOO− decomposition due to the slow reduction of oxoMn(IV) back to the Mn(III) oxidation state. However, when redox-coupled with biological antioxidants that efficiently reduce oxoMn(IV), Mn(III)TMPyP is transformed into an efficient “peroxynitrite reductase.”
Article
CONGENERS of nitrogen monoxide (NO) are neuroprotective and neurodestructive1– 7. To address this apparent paradox, we considered the effects on neurons of compounds characterized by alternative redox states of NO: nitric oxide (NO.) and nitrosonium ion (NO+)8. Nitric oxide, generated from NO. donors or synthesized endogenously after NMDA (N-methyl-D-aspartate) receptor activation, can lead to neurotoxicity3,4. Here, we report that NO.-mediated neurotoxicity is engendered, at least in part, by reaction with superoxide anion (O.- 2), apparently leading to formation of peroxynitrite (ONOO−), and not by NO. alone. In contrast, the neuroprotective effects of NO result from downregulation of NMDA-receptor activity by reaction with thiol group(s) of the receptor's redox modulatory site1. This reaction is not mediated by NO. itself, but occurs under conditions supporting S-nitrosylation of NMDA receptor thiol (reaction or transfer of NO+). Moreover, the redox versatility of NO allows for its interconversion from neuroprotective to neurotoxic species by a change in the ambient redox milieu. The details of this complex redox chemistry of NO may provide a mechanism for harnessing neuroprotective effects and avoiding neurotoxicity in the central nervous system.
Article
The first detection and characterization of oxomanganese(V) porphyrin complexes under ambient catalytic conditions is described. The reaction of (tetra-(N-methylpyridyl)porphyrinato)manganese(III) [Mn(III)TMPyP] with a variety of oxidants such as m-chloroperoxybenzoic acid (m-CPBA), HSO5-, and ClO- has been shown to produce the same, short-lived intermediate (1) by stopped-flow spectrophotometry. The Soret maximum of 1 was found at 443 nm, intermediate between that of oxomanganese(IV) (428 nm) and Mn(III)TMPyP (462 nm), thus facilitating its detection. The rate of formation of 1 from Mn(III)TMPyP followed second-order kinetics, first order in Mn(III) porphyrin and first order in oxidant. The rate constants have the following order:  m-CPBA (2.7 × 107 M-1 s-1) > HSO5- (6.9 × 105 M-1 s-1) ≈ ClO- (6.3 × 105 M-1 s-1). Once formed, the intermediate species 1 was rapidly converted to oxoMn(IV) (2) by one-electron reduction with a first-order rate constant of 5.7 s-1. The oxoMn(IV) species 2 was relatively stable under the reaction conditions, decaying slowly to Mn(III)TMPyP with a first-order rate constant of 0.027 s-1. The identity of 1 as an oxomanganese(V) complex was indicated by its reactivity. The one-electron reduction of 1 to oxoMn(IV) was greatly accelerated by nitrite ion (k = 1.5 × 107 M-1 s-1). However, the reaction between nitrite and oxoMn(IV) is much slower (k = 1.4 × 102 M-1 s-1). The oxoMn(V) intermediate 1 was shown to be highly reactive toward olefins, affording epoxide products. By contrast, oxoMn(IV) (2) was not capable of effecting the same reaction under these conditions. In the presence of carbamazepine (3) efficient oxygen transfer from the highly reactive oxoMn(V) (1) to the olefin (second-order rate constant of 6.5 × 105 M-1 s-1) resulted in the conversion of 1 directly back to Mn(III)TMPyP without the appearance of the stable oxoMn(IV) intermediate 2. With m-CPBA as the oxidant in the presence of H218O, the product epoxide was shown to contain 35% 18O, consistent with an O-exchange-labile oxoMn(V) intermediate. Nitrite ion inhibited the epoxidation reaction competitively by one electron reduction of the oxoMn(V) intermediate to the unreactive oxoMn(IV).
Article
Peroxynitrite is a versatile and important biological oxidant that is produced from the reaction of nitric oxide and superoxide radicals. Two mechanisms have been proposed to rationalize oxidation reactions of peroxynitrite. One assumes that HO−ONO can homolize to form the hydroxyl radical and nitrogen dioxide, and that the hydroxyl radical is the proximate oxidant in peroxynitrite systems. The second argues this homolysis is too slow to occur at ordinary temperatures and suggests an excited species, HOONO*, is the proximate oxidant. If the radical mechanism is correct, then peroxynitrite should disappear more slowly in solvents of higher viscosity. This is true because for free radical initiators undergoing single-bond homolysis:  (1) cage return is substantial and more of the cages would return to re-form HO−ONO as the viscosity of the medium increases; and (2) diffusion from the radical cage competes effectively with other cage processes. We have studied the disappearance of peroxynitrite at pH 5 and 7 in buffers with and without dioxane (as a control) or up to 30 wt % of the poly(ethylene glycol) (PEG) polymers, PEG 3350 and PEG 8000. These polyethers produce substantial changes in viscosity, raising the viscosity from about 0.89 to about 17 mPa·s. The rate constant for diffusion should decrease by about 10- to 20-fold as the viscosity increases in this interval, and the rate of diffusion from the solvent cage would be predicted to vary accordingly. However, at pH 5, where most of HOONO is undissociated, no change in the rate of disappearance of peroxynitrite is observed with increasing viscosity. At pH 7, a small increase in the observed rate constant is found, but it is likely due to the greater concentration of the undissociated HOONO in the ether-containing solvents resulting from a pKa shift. Thus, we conclude that the viscosity test does not support a free radical mechanism for the unimolecular decomposition of peroxynitrite.