in copyright - non-commercial use permitted rights ...27450/... · peroxynitrite, but it is also...

125
Research Collection Doctoral Thesis Peroxynitrous acid and vitamin C Author(s): Kurz, Christophe Robert Publication Date: 2004 Permanent Link: https://doi.org/10.3929/ethz-a-004834723 Rights / License: In Copyright - Non-Commercial Use Permitted This page was generated automatically upon download from the ETH Zurich Research Collection . For more information please consult the Terms of use . ETH Library

Upload: lamdan

Post on 21-Aug-2018

213 views

Category:

Documents


0 download

TRANSCRIPT

Research Collection

Doctoral Thesis

Peroxynitrous acid and vitamin C

Author(s): Kurz, Christophe Robert

Publication Date: 2004

Permanent Link: https://doi.org/10.3929/ethz-a-004834723

Rights / License: In Copyright - Non-Commercial Use Permitted

This page was generated automatically upon download from the ETH Zurich Research Collection. For moreinformation please consult the Terms of use.

ETH Library

Diss. ETHNr. 15620

Peroxynitrous acid and vitamin C

Abhandlung zur Erlangung des Titels

DOKTOR DER NATURWISSENSCHAFTEN

der

Eidgenössischen Technischen Hochschule Zürich

Vorgelegt von

Christophe R. Kurz

Dipl. Chem. ETH

Geboren am 26. Mai 1972

von Vechingen (BE)

Angenommen auf Antrag von

Prof. Dr. Willem H. Koppenol, Referent

Prof. Dr. Christian Schöneich, Koreferent

Dr. Reinhard Kissner, Koreferent

2004

1

2

To see what everyone has seen

and think what no one has thought.

By Albert Szent-Györgyi

The investigations reported in this thesis have been supported by the ETH and

the Swiss Nationalfonds.

3

4

CONTENTS

1. Danksagung 9

2. Objectives of the thesis 11

3.A Summary 13

B Zusammenfassung 17

References 21

4. Introduction 25

4.1 Peroxynitrite 25

4.1.1 History 25

4.1.2 Properties of peroxynitrite 26

4.1.3 Synthesis of peroxynitrite 29

4.1.4 Physiological relevance of peroxynitrite 33

4.2 Ascorbate 35

4.2.1 History 35

4.2.2 Properties of vitamin C 36

4.2.3 Synthesis of vitamin C 38

4.2.4 Resorption and excretion 41

4.2.5 Physiological relevance of vitamin C 41

4.2.6 Vitamin C as a scavenger 43

4.3 References 45

5. Material and Methods 53

5.1. Reagents 53

5.2. Stopped-flow spectrophotometrie studies 54

5.2.1 Equipment 54

5.2.2 Electrochemistry 54

5.2.3 Monohydroascorbate 55

5.2.4 Monohydroascorbate and carbon dioxide 55

5

5.3 Cyclic voltammetry 55

5.4 EPR 56

5.5 Ion chromatography 56

5.6 Simulation by Feldberg method 57

5.7 Simulation ofkinetic reactions 57

5.8 References 58

6. On the Chemical and Electrochemical One-Electron Reduction of

Peroxynitrous acid 61

Abstract 61

Results 62

Discussion 64

Appendix 69

References 72

Supplemental information 75

7. Rapid scavenging of Peroxynitrous Acid by Monohydroascorbate 77

(Free Radical Biol. Med. 35:1529-37; 2003)

Abstract 77

Results 78

Discussion 84

Speculations as to the structures of the intermediates 90

References 92

8. Mechanism of the Reaction of Peroxynitrite with L-Ascorbate

studied at physiological pH 95

Abstract 95

Results 95

Discussion 101

References 107

6

9. The Reaction of Peroxynitrite with Carbon Dioxide and Vitamin C..

109

Abstract 109

Results 109

Discussion 113

References 121

10. Curriculum vitae 123

7

8

1. DANKSAGUNG

Der Dank gilt vielen Leuten. Sie alle haben mich während meiner Arbeit

unterstützt, sei dies in Ratschlägen, aufstellenden Worten oder gar

Lösungsvorschlägen.

Herzlich bedanken möchte ich mich bei Herrn Prof. Dr. W. H. Koppenol

welcher mir die Doktorarbeit ermöglichte und mir den Freiraum gab meinen

Forscherdrang auszuleben.

Vielen Dank auch an Herrn Prof. Dr. Christian Schöneich welcher sich für die

Übernahme des Koreferates bereiterklärte.

Herzlichen Dank auch an Herrn Dr. Reinhard Kissner, der mich während der

ganzen Dissertationszeit unterstützt hat und stets geduldig bereit war, meine

vielen Fragen zu Beantworten.

Ebenfalls Bedanken möchte ich mich bei Herrn Dr. Thomas Nauser und Frau

PD Dr. Susanna Herold welche mir mit ihrem Wissen helfend zur Seite standen.

Danke auch an alle Kollegen und Kolleginnen der Gruppe Koppenol und

Herold, welche mich unterstützten, und ich viele lustige Stunden verbringen

konnte: Alina, Anastasia, Changyuan, Francesca, Gabi, Jolanta, Martin, Patrick,

Peter, Renate, Shiva, Xiuqiong.

Zum Schluss möchte ich mich bei den vielen Heinzelmännchen bedanken, die

das Chemiegebäude unterhalten. Sei es das Putzpersonal, die Schalterdamen, der

Hausdienst oder die Mensatruppe. Allen sei hier gedankt.

9

10

2. OBJECTIVES OF THE THESIS

Peroxynitrite is a potent cytotoxic species produced in vivo. It is known to

oxidise and nitrate proteins, lipids and DNA in vitro. One crucial parameter in

this context is the reduction potential of peroxynitrous acid, which appears to be

the more reactive form of peroxynitrite. The reduction potential allows to

estimate the ability of peroxynitrous acid to react with other compounds which,

in turn, permits to predict which compounds would be good scavengers of

peroxynitrite. Therefore we investigated the reduction potential of peroxynitrous

acid.

Another objective was to determine whether monohydroascorbate, a

physiological antioxidant, protects proteins and lipids against peroxynitrite.

Monohydroascorbate is considered to be a very efficient scavenger of free

radicals. It has been that monohydroascorbate protects very well against

peroxynitrite, but it is also known that the rate constant of the reaction between

peroxynitrous acid and ascorbate is very low (Bartlett, D.; Church, D. F. ;

Bounds, P. L.; Koppenol, W. H., The kinetics of oxidation of L-ascorbic acid by

peroxynitrite, Free Radical Biol. Med. 18:85-92; 1995, Squadrito, G. L.; Jin, X.;

Pryor,W. A., Stopped-flow kinetics of the reaction of ascorbic acid with

peroxynitrite, Arch. Biochem. Biophys. 322:53-59; 1995). These observations

are in contradiction to each other, and for this reason we reinvestigated the

reaction of peroxynitrous acid with monohydroascorbate.

11

12

3.A SUMMARY

Micromolar concentrations of both nitrogen monoxide and superoxide are

produced by activated macrophages during the immune response, which very

likely leads to formation of peroxynitrite given the diffusion-controlled rate of

1.6xl010M *s *

[1]. Among the reactive oxygen species, peroxynitrite and its

conjugated acid are of biological importance. Peroxynitrous acid and/or its

deprotonated form oxidizes thiols [2-5], nitrates tyrosines [6-11] and tryptophan

[12,13], and initiates lipid peroxidation [14,15]. Both one-electron oxidations

and oxygen-transfer have been reported [16-18].

The reduction potential of the ONOOH, H+ / N02", H20 couple is a crucial

parameter to predict whether peroxynitrous acid can oxidize a compound.

Values of 2.14 V and 2.0 V were published by Merényi and Lind [19], and

Koppenol and Kissner [20], respectively, and refer to standard conditions. These

values are based on determinations and estimates of the standard Gibbs energy

of formation of peroxynitrous acid. To our knowledge, no direct electrochemical

determination has been reported in the literature. The oxidation of the

peroxynitrite anion has been investigated electrochemically and a value of

0.51±0.02 V has been reported for E°(ONOOVONOO ) [21].

Therefore, we investigated in one part of this thesis the reduction potential of

peroxynitrous acid. Peroxynitrous acid was reduced by cathodic linear sweep

voltammetry at a gold electrode and by iodide at pH 3.2 to 5.6. The cathodic

reduction wave was identified by measuring its decay in time, which was the

same as with observations by optical spectroscopy. The iodide oxidation was

followed by optical measurement of the triiodide formation. Both reductions

show one-electron stochiometry, with the product naa = 0.23 ± 0.04 and an

iodine yield of around 0.5 equivalents per equivalent of peroxynitrous acid. The

13

voltammetric reduction was irreversible up to scan rates of 80 Vs_1. Both

reductions were pH independent in the range studied. The voltammetric

reduction is most likely an irreversible elemental reaction followed by a

chemical decay which cannot be observed directly. Because of the pH

independence, we conclude that both reductions have a common short-lived

intermediate, namely [HOONO]*-. This type of radical ion has been identified in

the reduction of nitrite with hydrated electrons in pulse radiolysis experiments.

The pH independence of the one-electron reductions also suggests that estimates

of the one-electron standard reduction potential of peroxynitrous acid are in

error. See chapter 6.

A lot of efforts have been done to identify a suitable peroxynitrite scavenger in

vivo.

Monohydroascorbate has already been investigated in 1995 by Bartlett et al. [22]

and Squadrito et al. [23]. They published a rate constant for the reaction of

235 M'V1 at 25°C. From the observation of ascorbyl radicals it was concluded

that a one-electron oxidation had taken place.

However many experiments show that the protective ability of

monohydroascorbate against peroxynitrite-mediated damage is much higher

than could be expected from this rate constant [24-27]. The published small rate

constant of the reaction between peroxynitrite and monohydroascorbate is in

contradiction to these observations. Therefore, we reinvestigated in one part of

this thesis the reaction between peroxynitrite and monohydroascorbate.

Data recorded from the reaction between monohydroascorbate and peroxynitrite

suggest a complex mechanism. The peroxynitrite anion does not react with

monohydroascorbate. For the initial reaction of peroxynitrous acid with

monohydroascorbate, a second-order rate constant of kj of 5 x 105±1 M_1s_1 and

Oil 1

an equilibrium constant K of 1 x 10 M were determined. We observed three

intermediates of which the third has an absorption maximum at 345 nm and a

14

pKa of 7.6 ± 0.2. Both the protonated and the deprotonated forms of this third

intermediate can react with a second monohydroascorbate to form nitrite,

monohydroascorbate, and dehydroascorbate. In addition, there is a

monohydroascorbate-independent decay path that yields nitrate and

monohydroascorbate. Therefore, at low monohydroascorbate concentrations, the

third intermediate decays to nitrate and monohydroascorbate, while at

monohydroascorbate concentrations greater than 4 mM, this third intermediate

reacts with a second monohydroascorbate to form nitrite, dehydroascorbate, and

monohydroascorbate.

EPR experiments indicate that the yield of the ascorbyl radical is 0.24 % relative

to the initial peroxynitrous acid concentration, and that this small amount of

ascorbyl radicals is formed concomitantly with the decrease of the absorption at

345 nm. Thus, the ascorbyl radical is not a primary reaction product. Under the

conditions of these experiments, no homolysis of peroxynitrous acid to nitrogen

dioxide and hydroxyl radical is observed. See chapter 7 and 8.

The more reactive peroxynitrous acid is not the most toxic form of peroxynitrite

in vivo. Even more toxic is the peroxynitrite anion when it reacts with carbon

dioxide to form l-carboxylato-2-nitrosodioxidane at a rate constant of

3 x K^lVrV1 at 25°C [28,29]. It would be useful to know whether the rate

constant of the reaction between peroxynitrous acid and monohydroascorbate is

fast enough to prevent the reaction of the peroxynitrite anion with carbon

dioxide or whether it even interacts with the products of the reaction with carbon

dioxide.

In the presence of carbon dioxide, peroxynitrite reacts catalytically with

monohydroascorbate at a very fast rate. Low monohydroascorbate

concentrations prevent the formation of the carbonate radical anion and nitrogen

dioxide, the intermediates produced from the reaction of peroxynitrite with

carbon dioxide. Moreover, at low monohydroascorbate concentrations only

15

small amounts of ascorbyl radicals, relative to the initial monohydroascorbate

concentration, are produced. If the monohydroascorbate concentration is higher

than that of peroxynitrite, the ascorbyl radical concentration increases. Under

these conditions, an alternative pathway, which involves the reaction with a

second monohydroascorbate molecule, is observed. Stopped-flow experiments

indicate that at pH 7, the reaction between peroxynitrous acid and

monohydroascorbate competes with the reaction between peroxynitrite anion,

monohydroascorbate, and carbon dioxide. At pH 6, the main pathway involves

the reaction of monohydroascorbate with peroxynitrous acid. See chapter 8.

A role as peroxynitrite scavenger under physiological conditions seems very

likely.

16

3.B ZUSAMMENFASSUNG

Aktivierte Makrophagen produzieren mikromolare Konzentrationen von

Stickstoffmonoxid und Superoxid, welche mit einer diffusionskontrollierten

Geschwindigkeit von

I.öxIO^mV1 miteinander reagieren [1]. Unter den reaktiven

Sauerstoffmetaboliten ist Peroxynitrit und seine Säure von biologischem

Interesse. Persalpetrige Säure und/oder ihre deprotonierte Form oxidieren Thiole

[2-5], nitrieren Tyrosine [6-11] und tryptophane [12,13], und lösen

Lipidperoxydation aus [14,15]. Es wurden Einelektronen- und Sauerstofftransfer

Beobachtet [16-18].

Das Reduktionspotential ist ein kritischer Parameter für die Voraussage, ob eine

andere Substanz oxidiert werden kann oder nicht. Für persalpetrige Säure

wurden Werte von 2.14 V und 2.0 V durch Merényi und Lind [19], respektive

durch Koppenol und Kissner [20] publiziert. Die Berechnungen beruhen auf

Standardbedingungen. Beide Werte basieren auf gemessenen und geschätzten

Werten für die freie Bildungsenergie von Persalpetriger Säure. Nach unseren

Kenntnissen wurde keine direkte Bestimmung publiziert. Die Oxidation von

Peroxynitritanion wurde elektrochemisch von Amatore et al. untersucht und er

publizierte einen Wert von 0.51 ± 0.02 V für E°(ON0070NOO ) [21].

Deswegen untersuchten wir elektrochemisch das Reduktionspotential der

Persalpetrigen Säure. Persalpetrige Säure wurde durch Iodid, pH 3.2 und 5.6,

und kathodische Linearvorschub-Voltammetrie, an einer Goldelektrode bei

pH 3.2 und 5.6, reduziert. Die kathodische Reduktionswelle wurde durch ihre

Zerfallszeit identifiziert, welche identisch mit der am UV-Vis

Spektrophotometer gemessenen ist. Die Iodidoxidation wurde optisch, durch die

Messung von Triiodid, verfolgt. Beide Reduktionen zeigen eine Ein-Elektron

Stöchiometrie mit dem Produkt naa = 0.23 + 0.04 und einer Iod Ausbeute von

17

etwa 0.5 Equivalenten pro Equivalent Persalpetriger Säure. Die voltammetrische

Reduktion war bis zu einer Vorschubgeschwindigkeit von 80 Vs"1 irreversibel.

Beide Reduktionen waren im gemessenen pH-Bereich pH-unabhängig. Bei der

gemessenen voltammetrischen Reduktion handelt es sich höchstwahrscheinlich

um eine irreversible Elementarreaktion, welche von einem chemischen Zerfall

gefolgt ist. Der Zerfall konnte nicht direkt beobachtet werden. Wegen der pH-

Unabhängigkeit schlussfolgern wir, dass beide Reduktionen das gleiche,

kurzlebige Zwischenprodukt, nämlich [ONOO]"2 /[ONOOH]" ,besitzen. Ein

analoges Radikal wurde in der Pulsradiolyse, durch die Reaktion von Nitrit mit

hydrierten Elektronen, identifiziert. Die gefundene pH-Unabhängigkeit weist

darauf hin, dass die berechneten Einelektronen-Standardreduktionspotentiale

fehlerhaft sind.

Grosse Anstrengungen wurden unternommen um einen passenden Peroxynitrit

Scavenger für die in vivo Anwendung zu finden.

Ascorbat wurde bereits im Jahre 1995 durch Bartlett el al. [22] und Squadritto et

al. [23] untersucht. Beide publizierten eine Geschwindigkeitskonstante von

235 M_1s_1 für 25°C. Die Beobachtung von Ascorbylradikalen wurde als eine

Ein-Elektron Oxidations-Reaktion gedeutet.

Jedoch viele Experimente zeigen dass Ascorbat viel besser vor Peroxynitrit

verursachten Schäden schützt, als durch die Geschwindigkeitskonstante erwartet

werden könnte [24-27]. Die publizierte Geschwindigkeitskonstante für die

Reaktion zwischen Peroxynitrit und Ascorbat von 235 M_1s_1 ist sogar im

Widerspruch zu den experimentellen Beobachtungen. Deswegen untersuchten

wir erneut die Reaktion zwischen Peroxynitrit und Vitamin C.

Die gemessenen Daten der Reaktion zwischen Peroxynitrit und Ascorbat deuten

auf einen komplexen Mechanismus. Peroxynitritanion reagiert nicht mit

Ascorbat. Für die Initialreaktion zwischen Persalpetriger Säure und Ascorbat

18

wurde eine zweite-Ordnungs Geschwindigkeitskonstante kj von

5 x 105±1 M^V1, mit einer Gleichgewichtskonstante K von 1 x 103±1 M^V1

bestimmt. Wir beobachteten drei Zwischenprodukte, wovon das Dritte ein

Absorbtionsmaximum bei 345 nm und einen pKa von 7.6 + 0.2 besitzt. Beide,

die protonierte und deprotonierte Form, können mit einem zweiten

Ascorbatmolekül reagieren und ergeben schlussendlich Nitrit, Ascorbat und

Dehydroascorbat. Zusätzlich existiert ein Ascorbat unabhängiger Zerfall,

welcher Nitrat und Ascorbat produziert. Deswegen zerfällt das dritte

Zwischenprodukt bei tiefen Ascorbat Konzentrationen zu Nitrat und Ascorbat

während bei Ascorbat Konzentrationen grösser als 4 mM, das Zwischenprodukt

mit einem zweiten Ascorbat reagiert und Nitrit, Ascorbat und Dehydroascorbat

ergibt.

EPR Experimente zeigen eine relative Ascorbylradikal Ausbeute von 0.24 %

gegenüber der ursprünglich eingesetzten Peroxynitrit Konzentration. Diese

kleine Menge Ascorbylradikal wird gleichzeitig mit der Abnahme der 345 nm

Absorbtion gebildet. Daher kann das Ascorbylradikal nicht das primäre

Reaktionsprodukt sein. Unter den experimentellen Bedingungen konnte keine

Homolyse von Persalpetriger Säure zu Stickstoffmonoxid und Hydroxylradikal

beobachtet werden. Siehe Kapitel 7 und 8.

Die reaktivere Form, Persalpetrige Säure, ist nicht die gefährlichste Form von

Peroxynitrit in vivo. Die viel gefährlichere Form von Peroxynitrite ist das

Peroxynitritanion wenn es mit Kohlendioxid reagiert und das schädliche 1-

carboxylato-2-nitrosodioxidane mit einer Geschwindigkeitskonstante von

3 x 104 M'V1 bei 25°C bildet [28,29]. Es wäre sehr hilfreich zu wissen, ob die

Geschwindigkeitskonstante der Reaktion zwischen Persalpetriger Säure und

Ascorbat schnell genug ist, um die Reaktion von Peroxynitritanion mit

Kohlendioxid zu verhindern oder ob Ascorbat sogar selber mit den Produkten

der Reaktion mit Kohlendioxid interagiert.

19

In Anwesenheit von Kohlendioxid reagiert Peroxynitrit in einer sehr schnellen

katalytischen Reaktion mit Ascorbat. Tiefe Ascorbat Konzentrationen

vermeiden bereits die Bildung von Carbonatradikalanion und Stickstoffdioxid,

Zerfallsprodukte der Reaktion zwischen Peroxynitritanion und Kohlendioxid.

Ausserdem werden bei tiefen Ascorbat Konzentrationen (kleiner als die

Peroxynitrit Konzentration) nur wenige Ascorbylradikale, relativ zur Ascorbat

Konzentration, gebildet. Ist die Ascorbat Konzentration grösser als diejenige

von Peroxynitrit, steigt die Ascorbylradikal Konzentration. Unter diesen

Bedingungen wird ein zweiter Reaktionsweg sichtbar, welcher die Reaktion mit

einem zweiten Ascorbat Molekül beinhaltet. Stopped-Flow Experimente bei

pH 7 weisen darauf hin, dass die Reaktion zwischen Persalpetriger Säure und

Ascorbat mit der Reaktion zwischen Peroxynitritanion, Ascorbat und

Kohlendioxid konkurriert. Bei pH 6 kann die Hauptreaktion bereits der Reaktion

von Persalpetriger Säure mit Ascorbat zugewiesen werden. Siehe Kapitel 9.

Ascorbat als Peroxynitrit Scavenger unter physiologischen Bedingungen scheint

sehr wahrscheinlich.

20

REFERENCES

[1] Nauser, T.; Koppenol, W. H. The rate constant of the reaction of

superoxide with nitrogen monoxide: Approaching the diffusion limit. J.

Phys. Chem. A 106:4084-4086; 2002.

[2] Koppenol, W. H.; Moreno, J. J.; Pryor, W. A.; Ischiropoulos, H.;

Beckman, J. S. Peroxynitrite, a cloaked oxidant formed by nitric oxide

and superoxide. Chem. Res. Toxicol. 5:834-842; 1992.

[3] Radi, R.; Beckman, J. S.; Bush, K. M.; Freeman, B. A. Peroxynitriteoxidation of sulfhydryls. J. Biol. Chem. 266:4244-4250; 1991.

[4] Gatti, R. M.; Radi, R.; Augusto, O. Peroxynitrite-mediated oxidation of

albumin to the protein- thiyl free radical. FEBSLett. 348:287-290; 1994.

[5] Kalyanaraman, B.; Karoui, H.; Singh, R. J.; Felix, C. C. Detection of thiylradical adducts formed during hydroxyl radical- and peroxynitrite-mediated oxidation of thiols - A high resolution ESR spin-trapping study

atQ-band(35 GRz).Anal. Biochem. 241:75-81; 1996.

[6] Ischiropoulos, H.; Zhu, L.; Chen, J.; Tsai, M.; Martin, J. C; Smith, CD.;

Beckman, J. S. Peroxynitrite-mediated tyrosine nitration catalyzed by

superoxide dismutase. Arch. Biochem. Biophys. 298:431-437; 1992.

[7] Ramezanian, M. S.; Padmaja, S.; Koppenol, W. H. Hydroxylation and

nitration of phenolic compounds by peroxynitrite. Chem. Res. Toxicol.

9:232-240; 1996.

[8] Beckman, J. S. Oxidative damage and tyrosine nitration from

peroxynitrite. Chem. Res. Toxicol. 9:836-844; 1996.

[9] Gow, A.; Duran, D.; Thorn, S. R.; Ischiropoulos, H. Carbon dioxide

enhancement of peroxynitrite-mediated protein tyrosine nitration. Arch.

Biochem. Biophys. 333:42-48; 1996.

[10] Lymar, S. V.; Jiang, Q.; Hurst, J. K. Mechanism of carbon dioxide-

catalyzed oxidation of tyrosine by peroxynitrite. Biochemistry 35:7855-

7861; 1996.

[11] Van der Vliet, A.; Eiserich, J. P.; O'Neill, C. A.; Halliwell, B.; Cross, C.

E. Tyrosine modification by reactive nitrogen species: A closer look.

Arch. Biochem. Biophys. 319:341-349; 1995.

[12] Alvarez, B.; Rubbo, H.; Kirk, M.; Barnes, S.; Freeman, B. A.; Radi, R.

Peroxynitrite-dependent tryptophan nitration. Chem. Res. Toxicol. 9:390-

396; 1996.

21

[13] Padmaja, S.; Ramezanian, M. S.; Bounds, P. L.; Koppenol, W. H.

Reaction of peroxynitrite with L-tryptophan. Redox. Rep. 2:173-177;1996.

[14] Radi, R.; Beckman, J. S.; Bush, K. M.; Freeman, B. A. Peroxynitrite-induced membrane lipid peroxidation: The cytotoxic potential of

superoxide and nitric oxide. Arch. Biochem. Biophys. 288:481-487; 1991.

[15] Darley-Usmar, V. M.; Hogg, N.; O'Leary, V. J.; Wilson, M. T.; Moncada,S. The simultaneous generation of superoxide and nitric oxide can initiate

lipid peroxidation in human low density lipoprotein. Free Radical Res.

Commun. 17:9-20; 1992.

[16] Al-Ajlouni, A.; Gould, E. S. Electron transfer. 133. Copper catalysis in

the sulfite reduction of peroxynitrite. Inorg. Chem. 36:362-365; 1997.

[17] Masumoto, H.; Kissner, R.; Koppenol, W. H.; Sies, H. Kinetic study of

the reaction of ebselen with peroxynitrite. FEBS Lett. 398:179-182; 1996.

[18] Maurer, P.; Thomas, C. F.; Kissner, R.; Rüegger, H.; Greter, O.;

Röthlisberger, U.; Koppenol, W. H. Oxidation of nitrite by peroxynitrousacid. J. Phys. Chem. A 107:1763-1769; 2003.

[19] Merényi, G.; Lind, J. Thermodynamics of peroxynitrite and its C02-adduct. Chem. Res. Toxicol. 10:1216-1220; 1997.

[20] Koppenol, W. H.; Kissner, R. Can ONOOH undergo homolysis. Chem.

Res. Toxicol. 11:87-90; 1998.

[21] Amatore, C; Arbault, S.; Bruce, D.; de Oliveira, P.; Erard, M.;

Vuillaume, M. Characterization of the electrochemical oxidation of

peroxynitrite: Relevance to oxidative stress bursts measured at the singlecell level. Chem. Eur. J. 7:4171; 2001.

[22] Bartlett, D.; Church, D. F.; Bounds, P. L.; Koppenol, W. H. The kinetics

of the oxidation of L-ascorbic acid by peroxynitrite. Free Radical Biol.

Med. 18:85-92; 1995.

[23] Squadrito, G. L.; Jin, X.; Pryor, W. A. Stopped-flow kinetic study of the

reaction of ascorbic acid with peroxynitrite. Arch. Biochem. Biophys.322:53-59; 1995.

[24] Whiteman, M.; Halliwell, B. Protection against peroxynitrite-dependent

tyrosine nitration and al-antiproteinase inactivation by ascorbic acid. A

comparison with other biological antioxidants. Free Radical Res. 25:275-

283; 1996.

22

[25] Lee, J.; Hunt, J. A.; Groves, J. T. Rapid decomposition of peroxynitrite by

manganese porphyrin-antioxidant redox couples. Bioorg. Med. Chem.

Lett. 7:2913-2918; 1997.

[26] Balavoine, G. G. A.; Geletii, Y. V. Peroxynitrite scavenging by different

antioxidants. Part I: Convenient assay. Nitric Oxide: Biol. Chem. 3:40-54;1999.

[27] Kirsch, M.; de Groot, H. Ascorbate is a potent antioxidant against

peroxynitrite-induced oxidation reactions - Evidence that ascorbate acts

by re-reducing substrate radicals produced by peroxynitrite. J. Biol.

Chem. 275:16702-16708; 2000.

[28] Lymar, S. V.; Hurst, J. K. Rapid reaction between peroxonitrite ion and

carbon dioxide: Implications for biological acitivity. J. Am. Chem. Soc.

117:8867-8868; 1995.

[29] Arteel, G. E.; Briviba, K.; Sies, H. Protection against peroxynitrite. FEBS

Letters 445:226-230; 1999.

23

24

4. INTRODUCTION

4,1 Peroxynitrite

4.1.1 HistoryN O Peroxynitrite was mentioned for the first time in 1901 by Bayer

O O et al. [1]. They discovered that a mixture of nitrous acid and

t/ hydrogen peroxide had unusual strong oxidizing properties. In

1907 Raschig proposed for the first time the formula HN04 for "Übersalpetrige

Säure" [2] which was used until Gleu and Roell published in 1929 the correct

formula ONOOH [3]. In 1952 Halfpenny and Robinson suggested, based on

hydroxylated and nitrated products, that peroxynitrous acid undergoes homolytic

fission to give hydroxyl radicals and nitrogen dioxide [4]. In 1954 Anbar and

Taube performed isotopic labelling studies with hydrogen peroxide and nitrous

acid at different pH. At high nitrite concentrations they found a quantitative

transfer of one oxygen to nitrite and nitrate, which is not expected for partial

homolysis [5]. Several years later, in 1968, Hughes and Nicklin claimed that

their results did not confirm the theory of peroxynitrite homolysis as well [6].

The lack of interest in peroxynitrite is reflected in the fact that only ca. 40 papers

on peroxynitrite had appeared until 1990. The history of peroxynitrite, up to

1992, was reviewed by Edwards et al. [7].

In 1990 it was postulated that peroxynitrite could be biologically relevant.

Beckman et al.[8] proposed for the first time that peroxynitrite can be produced

in vivo by the reaction of nitrogen monoxide with superoxide. The rate constant

of this reaction has been recently determined to be 1.6 x lO^M^s-1 [9]. This

rate constant is 7 times higher than the superoxide dismutase-catalysed

dismutation of superoxide (2.3 x 109 M *s *; pH-, buffer- and ionic strength

dependent) [10]. It leads to the conclusion that, in vivo, most superoxide could

react with nitrogen monoxide to form peroxynitrite whenever superoxide

dismutase is not present at concentrations 7 times higher then that of nitrogen

25

monoxide. Nitrogen monoxide concentrations are usually below 1 uM. Higher

concentrations are found at inflammatory and where nitrogen monoxide is used

as neurotransmitter. Indications that peroxynitrite is formed in vivo by the

reaction of nitrogen monoxide with superoxide are that superoxide dismutase

extends the biological lifetime of nitrogen monoxide by removing superoxide,

and that a decreased nitration of tyrosine is observed in the presence of

superoxide dismutase [11]. Nitrotyrosine is considered to be marker for

peroxynitrite in vivo. But peroxynitrite is not the only substrate which nitrates

tyrosine. Nitrogen dioxide can also nitrate proteins.

Beckman et al. concluded in 1990 from some scavenger studies with

peroxynitrite that 'a strong oxidant with reactivity similar to hydroxyl radical'

was formed [8]. Since 1952 it is being debated whether peroxynitrous acid can

homolyse to nitrogen dioxide and the hydroxyl radical [12-15] or not [16-19].

However, this question is more of academic interest than of relevance for

biological studies.

Peroxynitrite formed in vivo is more rapidly consumed by reactions with

carbon dioxide, thiols and metalloproteins than it decays via the isomerisation

pathway [20]. The rate constant for the reaction of peroxynitrite with carbon

dioxide is 5.8 ± 0.2 xlO4 jvrV1 at 37°C [21]. This rate constant, combined with

an in vivo concentration of about 1 mM of carbon dioxide, results in a

peroxynitrite half live of approximately 12 ms. Further information about

l-carboxylato-2-nitrosodioxidane and its physiological relevance is given in

chapter 4.1.4.

4.1.2 Properties of peroxynitrite

The pKa of peroxynitrous acid depends on the kind and concentration of the

buffer. Values in the range of 6.5 (1 mM phosphate) to 8.6 (0.1 M borate) have

been found [22]. The high pKa of peroxynitrite found in borate buffer can be

explained that borate is a Lewis acid. For diluted phosphate buffer a pKa value

26

of 6.8 is mostly used in the literature. Further, the decay of peroxynitrite is

strongly influenced by temperature and the presence of transition metals which

catalyse the isomerisation of peroxynitrite. However, in aqueous alkaline

solutions, peroxynitrite anion is stable for many weeks if it is kept frozen in the

dark [23]. There are a couple of possible decay pathways of peroxynitrite. The

term "isomerisation" is used for the rearrangement of peroxynitrous acid to

nitrate (reaction1.3), "homolysis" is used for reaction (1.4), "dimerisation" for

reaction (1.5) and "decay" for any disappearance.

The following table shows

and decay of peroxynitrite:

NO* + Of

ONOO- + H+

HOONO

HOONO

HOONO + ONOO-

[HOONO/ONOO]-

2 ONOO"

ONOO"

u 4.5xl09M1s1 [24]

k5 2.5xl05M~y1 [17]

t.5 25 s"1 [17]

k6 0.25 s"1 [17]

k7 2xl04M~y1 [17]

hv [22]

ki î.ôxio^ivrV1 [9]

k-i 1.7x10 V1 [25]

k2 lxio^ivrV1 [26]

k-2 1.6x10V1 [22]

k3 1.2 s"1 [22]

k4 1x102 s"1 [16]

Table 1: Some rate constants and properties of peroxynitrite.

27

some reaction pathways of the formation, equilibria

. \ - ONOO (1-1)

k2

k-2

k-4

k-5

HOONO (1-2)

^— N07+H+ (1.3)

k4_*OH + N02 (1-4)

[HOONO/ONOO]- (1.5)

k^» 02 + 2N02 (1-6)

^2— 02 + 2N02 (1-7)

^— N03- (1.8)

From quantum mechanical calculations the bond order between the nitrogen and

the first peroxide oxygen is approximately 1.5 [27,28]. This bond order gives the

possibility of the existence of two different structures of peroxynitrite anion.

Until now only the eis structure of peroxynitrite was found spectroscopicaly

[27,29]. Ab initio quantum chemical calculations showed that the eis isomer is

about 12 kJ/mol more stable than the trans isomer and that a barrier of

88-100 kJ/mol limits the isomerisation between the eis and the trans form

[27,28].

N-O^ -N-0

O

trans-peroxynitrite cis-peroxynitriteThe concentration of peroxynitrite is usually determined by UV/Vis-

spectroscopy. The peroxynitrite anion has a characteristic absorption with a

maximum at 302 nm with an extinction coefficient of 1705 M_1cm_1[30].

Peroxynitrous acid has an absorption maximum at 345 nm with a smaller

extinction coefficient of 80 M^cm"1 and an absorption shoulder at 250 nm with

an extinction coefficient of 600 M^cm"1 [31].

28

1800

1600

1400

1200

§ 1000

§ 800

600

400

200

0

w

1 »

1 <

*- U! »

/

/ \

i i»

1 ^

,: :,

1 v

.—^ \

-^v *

-^-—"^ ^

-^ ^

^: **i— *».

250 300 350

A,/nm

400 450

Fig. 1. Normalized spectra of 500 uM peroxynitrite (dotted line) and peroxynitrous acid (solid

line). The spectra were recorded by R. Kissner (ETH Zürich)

4.1.3 Synthesis of peroxynitrite

Peroxynitrite can be synthesised in a few different ways and its purity depends

on the respective method. Several syntheses are described in Gmelin[23] and in

Methods in Enzymology [32]. The most common ones are briefly described

below.

Biosynthesis

Peroxynitrite is formed if both reactants, superoxide and nitrogen monoxide, are

present simultaneously, for example in endothelial cells, activated macrophages,

neuronal cells, neurophils and smooth muscle cells. To decrease the formation

of peroxynitrite to 50%, superoxide dismutase should be present in the

concentration 7 times higher than that of nitrogen monoxide, (see chapter 4.1.1)

29

Chemical synthesis

Autoxidation ofhydroxylamine in alkaline solution [33]

NH2OH +02 +0H- H202 +NO" + H20

N0-+02 ONOO"

In this simple method oxygen is bubbled through an alkaline solution of

hydroxylamine, which leads to the formation of peroxynitrite. The typical yield

of peroxynitrite is about 25%, based on the starting concentrations of

hydroxylamine. The by-product, hydrogen peroxide, can be removed by

treatment with manganese dioxide. Unfortunately, nitrite, nitrate and

hydroxylamine remain.

Reaction between ozone and sodium azide [34,35]

N3" + 203 ONOO-+ N20 + 02

The ozonisation of azid ions in an alkaline solution yields peroxynitrite at

concentrations of up to 80 mM and the solution is free from hydrogen peroxide.

The disadvantage of this method is that the solution contains unreacted azide.

Reaction between hydrogen peroxide and alkyl nitrites [36,37]

H202 + OH" + RONO ONOO" + ROH + H20

The reaction of alkaline hydrogen peroxide with an alkyl nitrite gives

peroxynitrite with a nearly quantitative yield. The synthesis can be carried out

by two different methods.

Method A [37] employs a water-soluble alkylnitrite such as 2-ethoxyethyl

nitrite. However, the solution contains the alcohol at about the same

concentration as peroxynitrite.

Method B [36] is carried out in a two-phase system, containing alkaline

hydrogen peroxide in water and water-insoluble alkyl nitrite. Method B allows

to obtain concentrations of peroxynitrite up to 1 M, and the solution contains

unreacted hydrogen peroxide, nitrite and alkali.

30

Reaction between HN02 and H202 [1,38]

H202 +HN02 H20 + HOONO

HOONO +OH" ONOO"

The easiest and the most widely used method of peroxynitrite production is

based on the nitrosation of H202 in acid solutions, which generates the

peroxynitrous acid which is deprotonated by immediate addition of excess of

alkali. The disadvantage of this method is that nitrate is produced during the

reaction and therefore only contaminated peroxynitrite samples can be obtained.

For the first time this method was modified by Reed et al. [38]. Peroxynitrite

was synthesised in a quenched flow reactor from reagents pushed by a

compressed gas into a T junction. The yield of contaminations dépende on the

mixing temperature and the reaction time. The second modification was applied

by Beckman et al. [39]. Instead of a compressed gas a vacuum was utilised for

reagents mixing, resulting in a final yield of peroxynitrite of about 90 %.

Solid synthesisfrom K02 and NO'[40]

A more complicated method to obtain peroxynitrite is the solid synthesis. Its

advantage is much lower nitrate and nitrite contamination. In this method solid

potassium superoxide is diluted with quartz sand and after flushing the reaction

vessel with argon, nitrogen monoxide is added. The obtained solid peroxynitrite

is mixed with a small amount of manganese dioxide to eliminate hydrogen

peroxide which is produced by the remnants of superoxide upon dissolving in

cooled sodium hydroxide. The solution contains also low concentrations of

nitrite and nitrate.

Synthesis oftetramethylammoniumperoxynitrite in liquid ammonia [41]

In an inert and dry atmosphere tetramethylammonium hydroxide is mixed with

potassium superoxide and shaken under vacuum for three days. Then dry

31

ammonia is added and the tetramethylammonium superoxide, which is

dissolved, is transfered through a filter into another flask. After purification,

tetramethylammonium superoxide is exposed to nitrogen monoxide. Then

ammonia is slowly removed within 2 days (in vacuo). The obtained solid

tetramethylammonium peroxynitrite is of high purity and can be stored at

-70 °C for months.

Peroxynitrite productionfrom SIN-1 [42]

SIN-1 (3-morpholino-sydnonimine) is used for production of very small

concentrations of peroxynitrite over longer time. Initially SIN-1 is oxidised to

SIN-1"+ and generates superoxide from 02. In the next step SIN-1"+ releases a

nitrogen monoxide molecule and becomes inactive. The produced superoxide

and nitrogen monoxide react with each other forming peroxynitrite. It has to be

taken into account that superoxide and nitrogen monoxide are produced

consecutively and therefore they have also the possibility to react with other

compounds instead forming peroxynitrite.

0^ 0-

N

N

N

xcr^NH

SIN-1

O

NO*

N-+\

N

CH

XCN

Scheme 1: Peroxynitrite production by SIN-1SIN-1'

• +SIN-1C

32

4.1.4 Physiological relevancy of peroxynitrite

Peroxynitrite has been proposed to be produced in several cell types, such as

endothelial cells[43], macrophages[44], neuronal cells[45], neutrophils[46] and

smooth muscle cells[47].

At physiological pH peroxynitrite and its protonated form are present at a

ratio of 4 to 1 and therefore isomerise at 37 °C with a rate constant of 1.1 s"1

[22,48].

The unstable protonated form (ONOOH) is most probably more cytotoxic

than the stable, anionic form. But the peroxynitrite anion reacts with carbon

dioxide (physiological concentration of the latter compound is about 1 mM in

blood plasma) with a rate constant of 3 x 104 M'V1 at 25°C, [49] and produces

l-carboxylato-2-nitrosodioxidane(ONOOC02 ) which is proposed to be mainly

responsible for the toxicity of peroxynitrite. In the absence of other reactants,

ONOOC02 undergoes homolysis to nitrogen dioxide and trioxocarbonate(#l-)

radicals with a yield of 4 - 30 % [12,50-52] and isomerises to nitrate and carbon

dioxide.

The cytotoxic reactions of peroxynitrite, e.g. triggering of DNA strand

breakage, inhibition of lipid peroxidation, direct inhibition of mitochondrial

respiratory chain enzymes and other oxidative protein modifications show how

dangerous peroxynitrite in vivo can be.

The mentioned cytotoxic pathways of peroxynitrite show how important it is

to find a substance to prevent these reactions.

33

34

4.2 Vitamin C

HO L-Ascorbic acid (Vitamin C, 2-oxo-L-threo-hexono-l,4-

4.2.1 History

lactone-2,3-ene-diol) was named after its ability to preventH^^H^OH

q a serious illness called scurvy, one of the deficiency

diseases known longest to humankind. The name stems

from German (Antiscorbut- Vitamin). Between

1536 (Jacques Cartier) and 1601 (Sir James Lancaster) sailors discovered that a

diet including lemons, sauerkraut and vegetables can prevent this illness [53]. In

1747, James Lind carried out probably the first controlled clinical experiment

proving that lack of a compound in the diet was responsible for scurvy and that

oranges and lemons were the most effective against this disease [53]. In 1907,

Hoist discovered that guinea pigs are susceptible to scurvy what gave scientists

an opportunity to perform animal studies [54]. In 1920, vitamin C was isolated

from lemons by Zilva, but he did not identify this compound [55]. In 1928

Szent-Györgyi isolated vitamin C, ascribed the vitamin property to this

compound and named it hexuronic acid [56].

The structure of vitamin C was clarified in 1933 by Haworth and Hirst [57],

who proposed to rename vitamin C as ascorbic acid to reflect its antiscorbutic

properties. One year later Reichstein and Grüssner [58,59] published a method

of synthesis of vitamin C, which is used until today. In 1937 Sir Walter Norman

Haworth received the Nobel price in chemistry for investigations of

carbohydrates and vitamin C. In the same year Albert Szent-Györgyi Von

Nagyrapolt received the Nobel price in medicine for his discoveries in the field

of biological oxidation processes, with special reference to vitamin C. In 1965,

vitamin C was renamed as ascorbic acid or L-ascorbic acid by IUPAC-IUB

Comission on Biochemical Nomenclature [60].

The ability of vitamin C to prevent oxidation is used by the food processing

industry to inhibit food degradation and to obtain functional food of high

35

quality. The world-wide production of ascorbic acid is estimated to be around

60'000tperyear[61].

Ascorbate is known to be an excellent antioxidant, but it can also exhibit the

pro-oxidant properties e.g. the ascorbate-driven Fenton reaction [62].

Furthermore, ascorbate has the ability to re-reduce oxidised compounds.

The reaction between peroxynitrite and ascorbate was studied for the first

time in 1995 by Bartlett et al. [63] and one year later by Squadrito et al. [64].

The rate constant for the reaction is 235 M^s-1 at 25°C. From the observation of

ascorbyl radicals it was concluded that a one-electron oxidation had taken place.

4.2.2 Properties ofvitamin C

Vitamin C has two acidic protons with pKa values of 4.04 and 11.34 [65].

Monohydroascorbate, the single protonated form of vitamin C, is unstable in

water in presence of metal and oxygen. Even lowest traces of metals, nM range,

catalyse the oxidation by air and produce hydrogen peroxide and hydroxyl

radicals [66,67]. It is not possible to inhibit the metal catalysed autoxidation of

vitamin C by employing chelators as edta, dtpe, etc. In some cases the catalytic

effect of metals can be even enhanced [66,68]. Multiple literature reports about

the degradation of DNA and various cells culture, including cancer cells, can

probably be attributed to the formation of partially reduced oxygen species in

presence of traces of transition metal ions in the reaction solution or cell culture

media[69]. Most of the ascorbate is oxidised in a one-electron step process

producing ascorbyl radicals which disproportionate, in a second-order reaction

to ascorbate and dehydroascorbate. The ascorbyl radical, which is neither

strongly oxidising nor strongly reducing (E° = -0.174 V for Dasc/Asc" [65]), is

rather unreactive. The weak reactivity of ascorbyl radicals is mostly responsible

for the antioxidant effects of ascorbate: strongly oxidising radicals react with

ascorbate and the much less reactive ascorbyl radicals are formed [69].

Dehydroascorbate is unstable and decomposes rapidly via complex pathway to

36

oxalic and L-threonic acid (under aerobic conditions) or to carbon dioxide and

furfural (under anaerobic conditions) [61]. The optical isomer of L-Ascorbic

acid, D-(-)-Ascorbic acid has practically no vitamin activity because it is not

transported and can not interact with enzymes.

ability of L-ascorbate are the reduction potentials.

HAsc" + H+ 0-9)

Asc2-+H+ 0-10)

HAsc" 0-11)

Asc'"+H+ 0-12)

DAsc + HAsc" 0-13)

HAsc" 0-14)

pKi 4.04 [70]

pK2 11.7 [70]

pK3 -0.45 [72]

E°'i 0.282 V [71]

E°'2 0.05 V [65]

ki 3.5xl07M^s^

atpH5.8

[73]

Table 2: Some properties of vitamin C.

Ascorbic acid (H2Asc) has an absorption maximum at 243.5 nm with an

extinction coefficient of 10000 M^cm1 [74]. Monohydroascorbate (HAsc)

absorbs at 265 nm with an extinction coefficient in the range of 14500 -

14700 M^cm1 [66,74] In this thesis a value of 14500 M^cm1 was taken.

Ascorbyl radical has the highest absorption at 360 nm with an extinction

Of interest for the antioxidant

HAscT =

H2Asc

pK^

"COI

Asc*"+e +H+ !»

pK3HAsc" =s===

2 Asc*- + H+ ki

E°'DAsc + 2e" + H+ ^

37

coefficient of 5000 M ^m 1at pH 5.8 [75]. Dehydroascorbate itself has no

absorption, but products of its decomposition are yellow.

16000 -i

j

14000 A

12000 / \

-r 10000 / \

| / \

-£ 8000 / \^ 6000 / \

4000 / \ /"'" \

2000 y \ _y'

o 4-^^—'"'-•?"•" *--x-—i 1^=-——

220 270 320 370 420

A,/nm

Fig. 2. Normalized spectra of 800 uM ascorbyl radical (dotted line) and 50 urn

monohydroascorbate (solid line). The ascorbyl radical spectrum was recorded after pulse

radiolysis. Both spectra were obtained at pH 5.8 in 0.05 M phosphate buffer.

4.2.3 Synthesis of vitamin C

Vitamin C can not be synthesised by some organism, which lack the enzyme

1-gulonolactone oxidase. Among those species are primates, fruit-eating bats,

guinea pigs...

38

Biosynthesis

The biosynthesis of ascorbic acid in animal liver [76]:

CHO CHO

H-

HO-

H-

H-

-OH

-H

-OH

-OH

CH2OH

D-glucose

HO OH

Ascorbic acid

H-

HO-

H-

H-

-OH

-H

-OH

-OH

COOH

D-glucuronic acid

-gulonolactone*oxidase

Missing enzyme in

primates, guinea pigand bats

COOH

HO-

HO-

H-

HO-

-H

-H

-OH

-H

CH2OH

L-gulonic acid

HO

\

H^

H^ OH

%..(rf)H hlO>^=0

L-gulonic acid y-lactone

39

Industrial synthesis[61]

by Reichstein and Griissner[59]

HO.

OH

-o.

où IOH

D-glucose

H2/Ni

H-

HO-

H-

H-

tf»tô

n^°

>\°S*\c£°

0^'^

CH2OH

—OH

—H

—OH

—OH

CH2OH

D-sorbitol

H2S04

acetone

2,3:4,6-di-0-isopropylidene-

2-keto-L-gulonic acid

O

.0^0

b-iO Q X>H

2,3:4,6-di-0-isopropylidene-a-L-sorbofuranose

HO

HO OH

L-ascorbic acid

Gluconobacter oxydans

family

40

Recycling (in human neutrophils) [77]

TT A _

oxidants likeGSSG -*^ ^-*- HAsc ^

GSH — DAsc -*r ^-»* reduced oxidants

It is still unclear if dehydroascorbate is reduced by Dehydroascorbic Acid

Reductase, Protein Disulfide Isomerase or Thiol Transferase.

4.2.4 Resorption and excretion

Ingested ascorbic acid passes unchanged the stomach and is taken up in the

duodenum and transported by an active sodium-dependent mechanism. The rate

of resorption decreases with increasing ascorbate concentration [78].

The largest pools of ascorbate are the liver and the muscles tissue. About

90-95 % of ascorbic acid in the human body is present in the L-ascorbate form,

and the remaining fraction in the oxidised L-dehydroascorbate form (this ratio is

reversed in diabetic persons). Vitamin C is reabsorbed in the kidney, so a

deficiency of vitamin C is prevented. If the ascorbate concentration in plasma

exceeds over 1.2-1.8 mg/100 ml, kidneys stop reabsorption and ascorbic acid

passes to the urine. The formation of oxalic acid by vitamin C is of special

interest. About 40% of the excreted oxalic acid stems from vitamin C. This may

lead to nephrolithiasis. It was found that additional intake of vitamin C increases

the excretion of oxalic acid only insignificantly in non-risk persons of

nephrolithiasis.

4.2.5 Physiologic relevance of vitamin C

Vitamin C is an essential nutrient for the maintenance of human health; it is well

known to act as an intracellular and extracellular scavenger of free radicals and

41

oxidants. Therefore high concentrations of vitamin C are found in eyes and

lungs. In eyes radicals are produced by light-induced homolysis. In lungs

ascorbate protects from oxygen-derived radicals. Physiological concentrations of

ascorbate in various cells and tissues are of about 30 to 150 uM in human

plasma [79],1 mM in human lung and neutrophils [80], 1.3 mM in the liver [81],

7 mM in leukocytes [83] and 10 mM in neurons [82].

L-ascorbate is known to stimulate a number of enzymes, e.g. dioxygenase,

act as a cofactor in hydroxylation, be essential for iron resorption, play an

important role in the proper working of the nervous and endocrine system and

be essential in collagen synthesis, but the detailed mechanisms of numerous

actions of ascorbate are still unclear. Vitamin C takes also part in the redox

cycle of vitamin E, the main antioxidant present in the lipid part of the cells.

Ascorbic acid dissolved in the aqueous medium provokes a fast recycling of Vit.

E through the boundary layer between the aqueous medium and micelles. One-

electron oxidized form of Vit. E, a-tocopherol radical, react with vitamin C with

a second order rate constant of about 106 M-1s-1 [84,85].

The recycling of vitamin E at the expense of vitamin C found in vivo may

explain why vitamin E deficiency occurs very seldom.

Many reports deal with the relationship between vitamin C and the common

cold. A high vitamin C level in the body does not decrease the incidence of the

common cold. However, it consistently decreases the duration of common cold

episodes and the severity of the symptoms.

Therapeutically vitamin C is used for wound healing and enhances also

remarkably the body's immune response in fighting infections. The best known

health effect of ascorbate is the prevention and therapy of scurvy, a disease of

connective tissue, the symptoms of which are gum bleeding, amyotrophia and

tiredness. It has been suggested that L-ascorbate is helpful in preventing cancer

and cardiovascular diseases by protecting the DNA. A daily supplement of

ascorbate to enhance iron-uptake from food is also recommended in the

42

treatment of anemia. As mentioned earlier, the combination of iron, ascorbate

and air lead to radical production [71], therefore an additional intake of

ascorbate with iron should only be applied to persons with anemia. The

proposed daily dose of ascorbate is 120 mg /day (WHO, Aug. 2001)

Some high risk groups, such as smokers, elderly people, diabetics and

alcoholics have a lower vitamin C level than the average of the population.

Sportsmen, sick and stressed people also require a higher daily intake of

vitamin C.

High doses of ascorbic acid over long period of time (years) do not lead to

any side effects [61,78].

4.2.6. Vitamin C as a scavenger

A scavenger was defined by IUPAC in 1997 as follows: A substance that reacts

with (or otherwise removes) a trace component or traps a reactive reaction

intermediate (as in the scavenging of radicals or free electrons in radiation

chemistry). [86]

Ascorbate is known as a very good scavenger for free radicals. Further, it is

often applied in the treatment of the common cold and other diseases. Up to now

there is no substance which is cheaper, better tolerated by the body, and has less

nuisance problems associated with its administration than sodium ascorbate,

NaC6H606H, intravenously and intramuscularly, or ascorbic acid, C6H606H2,

orally.

Is L-ascorbate also a good scavenger for peroxynitrous acid? Peroxynitrous

acid can react with DNA, lipids, thiols and tyrosine residues in proteins. It is of

interest if L-ascorbate is able to scavenge peroxynitrite. A rate constant of

235 M'V1 has been reported in the literature for the reaction between

peroxynitrite and monohydroascorbate. However, many experiments show that

the protective ability of monohydroascorbate against peroxynitrite-mediated

damage is much higher than it can be expected from this rate constant [87].

43

What is necessary for monohydroascorbate to be a good scavenger for

peroxynitrite? The reduction potential of monohydroascorbate should be lower

than that one of peroxynitrous acid and the reaction of the scavenger

(monohydroascorbate) with peroxynitrous acid should be at least 10 times

higher than that of peroxynitrite with any other compound.

At physiological pH peroxynitrite reacts mainly with carbon dioxide, which

is present in a concentration of about 1.3 mM, with a second-order rate constant

of 3 x 104 jyrV1 at 25°C [49,88]. The resulting rate k' is therefore 39 s"1 (k' =

kx[C02]). In order to call monohydroascorbate a scavenger for peroxynitrite, the

physiological first-order rate k' should be higher than 390 s"1. (k' = kx[HAsc ]).

The concentration of monohydroascorbate varies from 100 uM to 10 mM (see

section 4.2.6.). Therefore the second order rate constant should be higher than

3.9 x 104M~1s~1 or even then 3.9 x 106M~1s~1, depending on the ascorbate

concentration.

Thus, in regard to this discrepancy between experimental findigs and kinetic

the reaction of monohydroascorbate with peroxynitrite was reinvestigated.

Results are described in chapter 7 and 8.

44

43 References

[1] Baeyer, A.; Villiger, V. Ueber die salpetrige Säure. Ber. Dtsch. chem.

Ges. 34:755-762; 1901.

[2] Raschig, F. Vorlesungsversuche aus der Chemie der anorganische

Stickstofverbindungen. Ber. Dtsch. chem. Ges. 40:4580-4588; 1907.

[3] Gleu, K.; Roell, E. Die Einwirkung von Ozon auf Alkaliazid.

Persalpetrige Säure I. Zeitschr. anorg. allgem. Chemie 179:233-266;1929.

[4] Halfpenny, E.; Robinson, P. L. Pernitrous acid. The reaction between

hydrogen peroxide and nitrous acid, and the properties of an

intermediate product. J. Chem. Soc. 928-938; 1952.

[5] Anbar, M.; Taube, H. Interaction of nitrous acid with hydrogen peroxideand with water. J. Am. Chem. Soc. 76:6243-6247; 1954.

[6] Hughes, M. N.; Nicklin, H. G. The chemistry of pernitrites. Part I.

Kinetics of decomposition of pernitrous acid. J. Chem. Soc. (A) 450-

452; 1968.

[7] Edwards, J. O.; Plumb, R. C. The chemistry of peroxonitrites. Prog.

Inorg Chem. 41:599-635; 1993.

[8] Beckman, J. S.; Beckman, T. W.; Chen, J.; Marshall, P. A.; Freeman, B.

A. Apparent hydroxyl radical production by peroxynitrite: Implicationsfor endothelial injury from nitric oxide and superoxide. Proc. Natl.

Acad. Sei. USA 87:1620-1624; 1990.

[9] Nauser, T.; Koppenol, W. H. The rate constant of the reaction of

superoxide with nitrogen monoxide: Approaching the diffusion limit. J.

Phys. Chem. A 106:4084-4086; 2002.

[10] Fielden, E. M.; Roberts, P. B.; Bray, R. C; Lowe, D. J.; Mautner, G. N.;

Rotilio, G.; Calabrese, L. The mechanism of action of superoxidedismutase from pulse radiolysis and electron paramagnetic resonance.

Biochem. J. 139:49-60; 1974.

[11] Beckman, J. S.; Ye, Y. Z.; Anderson, P. G.; Chen, J.; Accavitti, M. A.;

Tarpey, M. M.; White, C. R. Extensive nitration of protein tyrosines in

human atherosclerosis detected by immunohistochemistry. Biol. Chem.

Hoppe-Seyler 375:81-88; 1994.

45

[12] Coddington, J. W.; Hurst, J. K.; Lymar, S. V. Hydroxyl radical

formation during peroxynitrous acid decomposition. J. Am. Chem. Soc.

121:2438-2443; 1999.

[13] Gerasimov, O. V.; Lymar, S. V. The yield of hydroxyl radical from the

decomposition of peroxynitrous acid. Inorg. Chem. 38:4317-4321; 1999.

[14] Merényi, G.; Lind, J.; Goldstein, S.; Czapski, G. Peroxynitrous acid

homolyzes into OH and N02 radicals. Chem. Res. Toxicol. 11:712-713;1998.

[15] Richeson, C. E.; Mulder, P.; Bowry, V. W.; Ingold, K. U. The complex

chemistry of peroxynitrite decomposition: New insights. J. Amer. Chem.

Soc. 120:7211-7219; 1998.

[16] Koppenol, W. H.; Kissner, R. Can ONOOH undergo homolysis. Chem.

Res. Toxicol. 11:87-90; 1998.

[17] Kissner, R.; Koppenol, W. H. Product distribution of peroxynitrite decayas a function of pH, temperature, and concentration. J. Am. Chem. Soc.

124:234-239; 2002.

[18] Maurer, P.; Thomas, C. F.; Kissner, R.; Rüegger, H.; Greter, O.;

Röthlisberger, U.; Koppenol, W. H. Oxidation of nitrite by

peroxynitrous acid. J. Phys. Chem. A 107:1763-1769; 2003.

[19] Goldstein, S.; Czapski, G. Indirect oxidation of ferrocyanide - Evidence

against the formation of hydroxyl radicals. Nitric Oxide: Biol. Chem.

1:417-422; 1997.

[20] Koppenol, W. H. 100 Years of peroxynitrite biochemistry and 11 years

of peroxynitrite biochemistry. Redox. Rep. 6:339-341; 2001.

[21] Denicola, A.; Freeman, B. A.; Trujillo, M.; Radi, R. Peroxynitritereaction with carbon dioxide/bicarbonate: Kinetics and influence on

peroxynitrite-mediated oxidations. Arch. Biochem. Biophys. 333:49-58;1996.

[22] Kissner, R.; Nauser, T.; Bugnon, P.; Lye, P. G.; Koppenol, W. H.

Formation and properties of peroxynitrite studied by laser flash

photolysis, high pressure stopped flow and pulse radiolysis. Chem. Res.

Toxicol. 10:1285-1292; 1997.

[23] ONOO" Peroxynitrite. Gmelin Handbook of Inorganic and

Organometallic Chemistry; Nitrogen. Berlin: Springer-Verlag;1996:344-358.

46

[24] Logager, T.; Sehested, K. Formation and decay of peroxynitrous acid: a

pulse radiolysis study. J. Phys. Chem. 97:6664-6669; 1993.

[25] Merényi, G.; Lind, J. Free radical formation in the peroxynitrous acid

(ONOOH)/peroxynitrite (ONOO ) system. Chem. Res. Toxicol. 11:243-

246; 1998.

[26] Eigen, M. Proton Transfer and general acid base catalysis. In Claesson,S. eds. Fast reactions and primary processes in chemical kinetics.

Stockholm: Almqvist&Wiksell; 1967:245-253.

[27] Tsai, J.-H. M.; Harrison, J. G.; Martin, J. C; Hamilton, T. P.; van der

Woerd, M.; Jablonsky, M. J.; Beckman, J. S. Role of conformation of

peroxynitrite anion (ONOO") in its stability and toxicity. J. Am. Chem.

Soc. 116:4115-4116; 1994.

[28] Krauss, M. Electronic structure and spectra of the peroxynitrite anion.

Chem. Phys. Lett. 222:513-516; 1994.

[29] Wörle, M.; Latal, P.; Kissner, R.; Nesper, R.; Koppenol, W. H.

Conformation of peroxynitrite: Determination by crystallographic

analysis. Chem. Res. Toxicol. 12:305-307; 1999.

[30] Bohle, D. S.; Hansert, B.; Paulson, S. C; Smith, B. D. Biomimetic

synthesis of the putative cytotoxin peroxynitrite, ONOO", and its

characterization as a tetramethylammonium salt. J. Am. Chem. Soc.

116:7423-7424; 1994.

[31] Kissner, R. Absorption spectra of peroxynitrite. 2002. Personal

Communication

[32] Uppu, R. M.; Squadrito, G. L.; Cueto, R.; Pryor, W. A. Selecting the

most appropriate synthesis of peroxynitrite. Methods Enzymol. 269:285-

295; 1996.

[33] Hughes, M. N.; Nicklin, H. G. Autoxidation of hydroxylamine in

alkaline solution./. Chem. Soc. (A) 164-168; 1971.

[34] Pryor, W. A.; Cueto, R.; Jin, X.; Koppenol, W. H.; Ngu-Schwemlein,M.; Squadrito, G. L.; Uppu, P. L.; Uppu, R. M. A practical method for

preparing peroxynitrite solutions of low ionic strength and free of

hydrogen peroxide. Free Radical Biol. Med. 18:75-83; 1995.

[35] Uppu, R. M.; Squadrito, G. L.; Cueto, R.; Pryor, W. A. Synthesis of

peroxynitrite by azide-ozone reaction. Methods Enzymol. 269:311-321;1996.

47

[36] Uppu, R. M.; Pryor, W. A. Biphasic synthesis of high concentrations of

peroxynitrite using water-insoluble alkyl nitrite and hydrogen peroxide.Methods Enzymol. 269:322-329; 1996.

[37] Leis, J. R.; Pena, M. E.; Rios, A. A novel route to peroxynitrite anion. J.

Chem. Soc. Chem. Commun. 1993:1298-1299; 1993.

[38] Reed, J. W.; Ho, H. H.; Jolly, W. L. Chemical synthesis with a quenchedflow reactor. Hydroxytrihydroborate and peroxynitrite. J. Am. Chem.

Soc. 96:1248-1249; 1974.

[39] Beckman, J. S.; Chen, J.; Ischiropoulos, H.; Crow, J. P. Oxidative

chemistry of peroxynitrite. Methods Enzymol. 233:229-240; 1994.

[40] Koppenol, W. H.; Kissner, R.; Beckman, J. S. Syntheses of

peroxynitrite: To go with the flow or on solid grounds. Methods

Enzymol. 269:296-302; 1996.

[41] Bohle, D. S.; Glassbrenner, P. A.; Hansert, B. Synthesis of pure

tetramethylammonium peroxynitrite. Methods Enzymol. 269:302-311;1996.

[42] Singh, R. J.; Hogg, N.; Joseph, J.; Konorev, E.; Kalyanaraman, B. The

peroxynitrite generator, SIN-1, becomes a nitric oxide donor in the

presence of electron acceptors. Arch. Biochem. Biophys. 361:331-339;1999.

[43] Palmer, R. M. J.; Ashton, D. S.; Moncada, S. Vascular endothelial cells

synthesize nitric oxide from L-arginine. Nature 333:664-666; 1988.

[44] Ischiropoulos, H.; Zhu, L.; Beckman, J. S. Peroxynitrite formation from

macrophage-derived nitric oxide. Arch. Biochem. Biophys. 298:446-451;1992.

[45] Lafon-Cazal, M.; Culcasi, M.; Gaven, F.; Pietri, S.; Bockaert, J. Nitric

oxide, superoxide and peroxynitrite: Putative mediators of NMDA-

induced cell death in cerebral granule cells. Neuropharmacology32:1259-1266; 1993.

[46] Carreras, M. C; Pargament, G. A.; Catz, S. D.; Poderoso, J. J.; Boveris,A. Kinetics of nitric oxide and hydrogen peroxide production and

formation of peroxynitrite during the respiratory burst of human

neutrophils. FEBS Lett. 341:65-68; 1994.

[47] Schulz, R.; Smith, J. A.; Lewis, M. J.; Moncada, S. Nitric oxide

synthase in cultured endocardial cells of the pig. Br. J. Pharmacol.

104:21-24; 1991.

48

[48] Padmaja, S.; Kissner, R.; Bounds, P. L.; Koppenol, W. H. Hydrogen

isotope effect on the isomerization of peroxynitrous acid. Helv. Chim.

Acta 81:1201-1206; 1998.

[49] Lymar, S.V.; Hurst, J. K. Rapid reaction between peroxonitrite ion and

carbon dioxide: Implications for biological acitivity. J. Am. Chem. Soc.

117:8867-8868; 1995.

[50] Bonini, M. G.; Radi, R.; Ferrer-Sueta, G.; Ferreira, A. M. D.; Augusto,O. Direct EPR detection of the carbonate radical anion produced from

peroxynitrite and carbon dioxide. J. Biol. Chem. 274:10802-10806;1999.

[51] Meli, R.; Nauser, T.; Latal, P.; Koppenol, W. H. Reaction of

peroxynitrite with carbon dioxide: Intermediates and determination of

the yield of C03' and N02\ J. Biol. Inorg Chem. 7:31-36; 2002.

[52] Lymar, S. V.; Hurst, J. K. C02-catalyzed one-electron oxidation by

peroxynitrite: Properties of the reactive intermediate. Inorg. Chem.

37:294-301; 1998.

[53] Moser, U.; Bendich, A. Vitamin C. In Machlin, L. J. eds. Handbook ofVitamins. New York: Marcel Dekker; 1991:195-225.

[54] Hoist, A.; Fröhlich, T. Experimental studies relating to ship beri-beri

and scurvy. II On the etiology of scurvy. J. Hyg. (Cambridge) 7:634-

671; 1907.

[55] Zilva, S. S. The influence of aeration on the stability of the antiscorbutic

factor. The Lancet 197:478; 1921.

[56] Szent-Györgyi, A. Observations on the function of peroxidase systemsand the chemistry of the adrenal cortex. Biochem. J. 22:1387-1409;1928.

[57] Haworth, W. N.; Hirst, E. L. Synthesis of ascorbic acid. J. Soc. Chem.

Ind. London Trans. Commun. 52:645-646; 1933.

[58] Reichstein, T.; Oppenauer, R. Synthese der d- und 1-ascorbinsäure (C-Vitamin). Helv. Chim. Acta 16:1019-1033; 1933.

[59] Reichstein, T.; Grüssner, A. Eine ergibige Synthese der 1-Ascorbinsäure

(C-Vitamin). Helv. Chim. Acta 17:311-328; 1934.

[60] IUPAC Trivial names of miscellaneous compounds of importance in

biochemistry. Biochim. Biophys. Acta. :Gen. Subj. 107:1-4; 1965.

49

[61] Vitamins/Vitamin C. Ullmann's Encyclopedia of industrial chemistry.Weinheim: Wiley-VCH Verlag GMBH; 2003:218-231.

[62] Burkitt, M. J.; Gilbert, B. C. Model studies of the iron-catalysed Haber-

Weiss Cycle and the ascorbate-driven fenton reaction. Free Radical Res.

Commun. 10:265-280; 1990.

[63] Bartlett, D.; Church, D. F.; Bounds, P. L.; Koppenol, W. H. The kinetics

of the oxidation of L-ascorbic acid by peroxynitrite. Free Radical Biol.

Med. 18:85-92; 1995.

[64] Squadrito, G. L.; Jin, X.; Pryor, W. A. Stopped-flow kinetic study of the

reaction of ascorbic acid with peroxynitrite. Arch. Biochem. Biophys.322:53-59; 1995.

[65] Williams, M. H.; Yandell, J. K. Outer-sphere electron transfer reactions

of ascorbate anions. Aust. J. chem. 35:1133-1144; 1982.

[66] Buettner, G. R. In the absence of catalytic metals ascorbate does not

autoxidize at pH 7: ascorbate as a test for catalytic metals. Journal ofBiochemical and Biophysical Methods 16:27-40; 1988.

[67] Miller, D. M.; Buettner, G. R.; Aust, S. D. Transition metals as catalystsof "autoxidation" reactions. Free Radical Biol. Med. 8:95-108; 1990.

[68] Glebska, J.; Grzelak, A.; Pulaski, L.; Bartosz, G. EDTA loses its

antioxidant properties upon storage in buffer. Anal. Biochem. 311:87-89;2002.

[69] Halliwell, B.; Gutteridge, J. M. C. Antioxidant protection by low-

molecular-mass agents: compounds derived from the diet. Free Radicals

in biology andMedicine. New York: Oxford university press; 1999:200-

208.

[70] Handbook of Chemistry and Physics. Internet 3rd electronical edition.

2000. 2000. Electronic Citation

[71] Buettner, G. R.; Jurkiewicz, B. A. Catalytic metals, ascorbate and free

radicals: Combinations to avoid. Radiât. Res. 145:532-541; 1996.

[72] Davis, H. F.; McManus, H. J.; Fessenden, R. W. An ESR study of free-

radical protonation equilibria in strongly acid media. J. Phys. Chem.

90:6400-6404; 1986.

[73] Bielski, B. H. J.; Allen, A. O.; Schwarz, H. A. Mechanism of

disproportionation of ascorbate radicals. J. Am. Chem. Soc. 103:3516-

3518; 1981.

50

Ogata, Y.; Kosugi, Y. Ultraviolet spectra of L-ascorbic acid and cupricascorbate complex. Tetrahedron 26:4711-4716; 1970.

Bielski, B. H. J.; Comstock, D. A.; Bowen, R. A. Ascorbic acid free

radicals. I. Pulse radiolysis study of optical absorption and kinetic

properties. J. Am. Chem. Soc. 93:5624-5629; 1971.

Lewin, S. Vitamin C: London Academic Press; 1976.

Washko, P. W.; Wang, Y.; Levine, M. Ascorbic acid recycling in human

neurophils. J. Biol. Chem. 268:15531-15535; 1993.

Vitamin C. In Eisenbrand, G.; Schreier, P. eds. Römpp-LexikonLebensmittelchemie. Stuttgart: Thieme Verlag; 1995:899-901.

Stocker, R.; Frei, B. Endogenous antioxidant defences in human blood

plasma. In Sies, H. eds. Oxidative Stress: Oxidants and Antioxidants.

London: Academic Press; 1991:213-243.

Washko, P.; Rotrosen, D.; Levine, M. Ascorbic acid in human

neutrophils. Am. J. Clin. Nutr. 54 Suppl.:1221S-1227S; 1991.

Li, X.; Cobb, C. E.; Hill, K. E.; Burk, R. F.; May, J. M. Mitochondrial

uptake and recycling of ascorbic acid. Arch. Biochem. Biophys. 387:143-

153; 2001.

Rice, M. E.; Russo-Menna, I. Differential compartimentalization of

brain ascorbate and glutathione between neurons and glia. Neuroscience

82:1213-1223; 1998.

Griffiths, H. R.; Lunec, J. Ascorbic acid in the 21st century - more than

a simple oxidant. Environ. Toxicol. Pharmacol. 10:173-182; 2001.

Packer, J. E.; Slater, T. F.; Willson, R. L. Direct observation of free

radical interaction between vitamin E and vitamin C. Nature 278:737-

738; 1979.

Scarpa, M.; Rigo, A.; Maiorino, M.; Ursini, F.; Gregolin, C. Formation

of alpha-tocopherol radical and recycling of alpha-tocopherol byascorbate during peroxidation of phosphatidylcholine liposomes.Biochim. Biophys. Acta 801:215-219; 1984.

de Bolster, M. W. G. Glossary of terms used in bioinorganic chemistry.

PureAppl. Chem. 69:1251-1303; 1997.

51

[87] Balavoine, G. G. A.; Geletii, Y. V. Peroxynitrite scavenging by different

antioxidants. Part I: Convenient assay. Nitric Oxide: Biol. Chem. 3:40-

54; 1999.

[88] Arteel, G. E.; Briviba, K.; Sies, H. Protection against peroxynitrite.FEBS Letters 445:226-230; 1999.

52

5. MATERIALAND METHODS

5.7. Reagents

All chemicals used in experiments were of analytical grade or better. Ascorbic

acid, disodium monohydrogen phosphate, sodium dihydrogen phosphate and

sodium nitrite were purchased from Fluka (Buchs, Switzerland). Potassium

superoxide was from Aldrich (St. Louis, MO, USA) or from Aber (Karlsruhe,

Germany), phosphoric acid and sodium hydroxide from Siegfried (Säckingen,

Switzerland), carbon dioxide from Pangs (Luzern, Switzerland), sodium iodide

from Merck (Darmstadt, Germany), and nitrogen monoxide from Linde

(Unterschleissheim, Germany). Peroxynitrite was synthesized according to

Koppenol et al. [1]. The concentration of the peroxynitrite solution was

determined by measuring the absorbance at 302 nm (£302 = 1705 M^cnT1) [2].

Deionised water was purified further by a Millipore Milli-Q unit (Bedford, MA,

USA). Buffers were prepared from salts and acids. All solutions were freshly

prepared and - except the peroxynitrite solutions - were evacuated for five

minutes and then saturated with argon or an argon/carbon dioxide mixture, to

avoid oxidation of monohydroascorbate by dioxygen. The peroxynitrite

solutions were prepared by dilution with 0.010 M sodium hydroxide that was

de-aerated by evacuation for five minutes prior to its addition to the

peroxynitrite stock solution. When peroxynitrite was not mixed with

monohydroascorbate and oxygen had no influence on the results, the solutions

were not evacuated.

53

5.2. Stopped-flow spectrophotometrie studies

5.2.1. Equipment

Kinetic experiments were carried out at 25°C and 37°C and ambient pressure

with an Applied Photophysics SX 17MV, or with Applied Photophysics

SX 18MV stopped-flow spectrophotometer (Leatherhead, Surrey, GB) operated

in the symmetric mixing mode. The pH was measured at the outlet. Were not

different reported, peroxynitrite was synthesized according to Koppenol et al.

[1]. Time-dependent spectra were obtained at 25°C with an OLIS stopped-flow

instrument equipped (Bogart, GA, USA) with an OLIS RSM 1000 rapid

scanning monochromator set to collect up to 1000 spectra per second. When

experiment were made with wavelength higher than 450 nm, filters were used.

5.2.2 One-electron reduction

Kinetics experiments were carried out at 25°C and ambient pressure with an

Applied Photophysics SX 17MV (Leatherhead, Surrey, UK) stopped-flow

spectrophotometer operated in the symmetric mixing mode. The reaction was

initiated by mixing 10, 50 or 100 uM peroxynitrite in 0.01 M sodium hydroxide

with various sodium iodide concentrations (200 uM to 40 mM) in 0.2 M

phosphate buffer at pH 5.8 and at pH3.2. The changes in absorbance were

followed at 350 nm and the pH was determined at the outlet of the stopped-flow

apparatus with a glass electrode. The equilibrium constant of the triiodide

formation

I2 + F + * I3~ was determined to be 930 M"1 in the phosphate buffer

system used, somewhat different from the 710 M"1 found in the literature, [3]

which is a value obtained by extrapolation to zero ionic strength. The extinction

coefficient of I3~ at 350 nm was 28100 M_1cm_1, in good agreement with the

literature. [3]

54

5.2.3. Monohydroascorbate

The reaction was initiated by mixing a peroxynitrite solution in 0.010 M sodium

hydroxide with a monohydroascorbate solution in 0.10 M phosphate buffer

(pH 5.8, 7.2, 8.0 and 8.5). The wavelengths measured were 265 nm

(monohydroascorbate); 320 nm (for peroxynitrite); 345 nm and 360 nm for the

intermediates.

5.2.4. Monohydroascorbate and carbon dioxide

Reactions were initiated by mixing peroxynitrite in a phosphate buffer of pH 12

with monohydroascorbate in phosphate buffer of pH3. The phosphate buffer

concentrations were adjusted to the pH needed. The wavelengths measured were

320 nm (for peroxynitrite); 345 nm (for the intermediates); 360 nm (for the

intermediates; mainly ascorbyl radical); and 640 nm for trioxocarbonate(#-)

radical.

5.3. Cyclic voltammetry

Cyclovoltammograms were measured by a stopped-flow technique. The

apparatus setup is shown in chapter 6.2. (Supplemental information) in Figure 1.

One pump (Dosimat 665 titration unit from Metrohm/Herisau, Switzerland) was

filled with phosphate buffer and the second one with a 2 mM peroxynitrite

solution in 0.01 M NaOH. The pump units were activated synchronously by

computer. The titration devices were both set to a flow of 20 ml/ minute, and the

measuring cell was rinsed with twenty times its volume (Fig. 2S; chapter 6.2.).

The measurements were started by a trigger signal sent by the control unit to the

measuring computer. The control unit consisted of a computer with a 12 bit

analog to digital converter, an AMEL 2049 potentiostat and an AMEL 586

function generator (Milan, Italy). The working electrode was a gold electrode

55

from Metrohm (Herisau, Switzerland) with an active surface of 3.1 mm.The

reference electrode was a silver/silver chloride electrode from Methrom

(Herisau, Switzerland) filled with 0.1 M potassium chloride and 0.1 M sodium

nitrate as bridge electrolyte. The auxiliary electrode was a glassy carbon

electrode from Metrohm (Herisau, Switzerland).

5.4. EPR

Electron paramagnetic resonance spectra were obtained with a Bruker EMX 080

(Karlsruhe, Germany) equipped with a microwave-bridge ER 041 XG and the

dielectric mixing resonator ER4117D-MVT. Data acquisition and analysis

were carried out with ACQUISIT software (Bruker). The solutions (25 uM

peroxynitrite and 250 uM monohydroascorbate) were injected into the cavity

with a KDScientific 220 syringe pump (New Hope, PA, USA) at a flow of

1 ml/min per syringe in a ratio 1:1. The solutions were of the same composition

as those used in the stopped-flow experiments.

The signal intensity of the ascorbyl radical signal was determined by mixing

a freshly prepared cerium(IV) solution (25 uM final concentration) with

monohydroascorbate at pH 5.8 in 50 mM phosphate buffer. Only a few percent

of cerium(IV) react with monohydroascorbate, due to the competing reaction

with phosphate that produces an unidentified, inert cerium-phosphate complex

[4]. The concentration of the small amount of ascorbyl radical produced was

determined by the rate of its bimolecular decay

(3.5 x Kfivry1)^].

5.5. Ion chromatography

Nitrite and nitrate were determined on a Methrom 732 ion separation center

(Herisau, Switzerland) equipped with a Super-Sep IC anion exchanging column

56

(Methrom, Herisau, Switzerland) with a phthalate solution (2.5 mM phthalate,

5% acetonitrile, pH 4.2 (TRIS), A «130 nS/cm) as eluent. [6,7].

5.6. Simulation by Feldberg method

This method is used to simulate cyclovoltammograms to determine the formal

reduction potential. The Feldberg method simulates the concentration gradient

of the substrate to be reduced at the electrode surface. When the onset potential

is reached, the peroxynitrous acid in a small volume next to the electrode is

assumed to be completely reduced, which results in a concentration gradient

between the electrode and the bulk of the solution. To simulate this gradient and

the diffusion of the ions, the solvent around the electrode is divided into small

compartments whereby, for each calculation cycle, the concentrations are only

allowed to equilibrate between adjacent compartments. The depletion rate of

active species at the electrode surface limits and thereby determines the current

[8-10].

5.7. Simulation ofkinetic reactions

Kinetics were simulated based on sets of ordinary differential equations with an

EXCEL macro program that implements a two-step implicit Runge-Kutta

algorithm of 4th order [11]. This method is more stable than explicit integration

techniques, but requires more computation time.

57

References

[1] Koppenol, W. H.; Kissner, R.; Beckman, J. S. Syntheses of peroxynitrite:

To go with the flow or on solid grounds. Methods Enzymol. 269:296-302;

1996.

[2] Bohle, D. S.; Glassbrenner, P. A.; Hansert, B. Synthesis of pure

tetramethylammonium peroxynitrite. Methods Enzymol. 269:302-311;

1996.

[3] Meyerstein, D.; Treinin, A. Charge-transfer complexes of iodine and

inorganic anions in solution. Trans. Faraday Soc. 59:1114-1120; 1963.

[4] Oxidation state +4. In Bailar, J. C; Emeléus, H. J.; Nyholm, R.; Trotman-

Dickenson, A. F. eds. Comprehensive Inorganic Chemistry. Oxford:

Pergamon Press; 1973:97-98.

[5] Bielski, B. H. J.; Allen, A. O.; Schwarz, H. A. Mechanism of

disproportionation of ascorbate radicals. J. Am. Chem. Soc. 103:3516-3518;

1981.

[6] Kissner, R.; Koppenol, W. H. Product distribution of peroxynitrite decay as

a function of pH, temperature, and concentration. J. Am. Chem. Soc.

124:234-239; 2002.

[7] Meli, R.; Nauser, T.; Latal, P.; Koppenol, W. H. Reaction of peroxynitrite

with carbon dioxide: Intermediates and determination of the yield of C03"

andN02'.J. Biol. Inorg. Chem. 7:31-36; 2002.

[8] Feldberg, S. W. Theory of relaxation of the diffusion double layer

following coulostatic charge injection. J. Phys. Chem. 74:87-90; 1970.

58

[9] Feldberg, S. W. Digital simulation: A general method for solving

electrochemical diffusion-kinetic problems. In Bard, A. J. eds.

Electroanalytical chemistry. Marcel Dekker; 1969:199-296.

[10] Mattson, J. S.; Mark, H. B.; MacDonald, H. C. Computers in chemistry and

instrumentation: New York Marcel Dekker; 1972.

[11] Schwarz, H. R. Numerische Mathematik: Stuttgart B.G. Teubner; 1988.

59

60

6. ON THE CHEMICAL AND ELECTROCHEMICAL

ONE-ELECTRON REDUCTION OF

PEROXYNITROUS ACID

CHRISTOPHE KURZT, XIUQIONG ZENGJ, STEFAN HANNEMANNT,REINHARD KISSNERT AND WILLEM H. KOPPENOL*T

^ Laboratorium für Anorganische Chemie, Eidgenössische Technische

Hochschule Zürich (Hönggerberg), Switzerland

^Department of Chemistry at Xixi Campus, ZhejiUniversity, Hangzhou 310028,

China

Abstract— Peroxynitrous acid was reduced by cathodic linear sweep

voltammetry at a gold electrode and by iodide at pH 3.2 to 5.6. The cathodic

reduction wave was identified by measuring its decay in time, which was the

same as with observations by optical spectroscopy. The iodide oxidation was

followed by optical measurement of the triiodide formation. Both reductions

show one-electron stochiometry, with the product naa = 0.23 ± 0.04 and an

iodine yield of around 0.5 equivalents per equivalent of peroxynitrous acid. The

voltammetric reduction was irreversible up to scan rates of 80 Vs"1. Both

reductions were pH independent in the range studied. The voltammetric

reduction is most likely an irreversible elemental reaction followed by a

chemical decay which cannot be observed directly. Because of the pH

independence, we conclude that both reductions have a common short-lived

intermediate, namely [HOONO]*-, which is analogous to [HONO]*~ and

[ONO]*2~. This type of radical ion has been identified in the reduction of nitrite

with hydrated electrons in pulse radiolysis experiments. The pH independence

of the one-electron reductions also suggests that estimates of the one-electron

standard reduction potential of peroxynitrous acid, including our own, are in

error.

61

Results

Voltammetry results.

Electrochemical reduction of

peroxynitrous acid at a

polycrystalline gold electrode with a

linear sweep yielded a cathodic

wave which could be assigned

unambiguously to the substance,

since the peak current at a given

scan rate decreased with the delay

time between the formation of the

acid by mixing peroxynitrite with

dihydrogen phosphate/phosphoric

acid and the onset of the

voltammetric scan. The rate constant

for the first-order peak current

decrease was found to be (1.1

±0.1) s"1 at pH 5.6 and (1.5+0.1) s"1

at pH3.2, quite typical for the

peroxynitrous acid isomerisation at

room temperature [1]. The reduction

wave of peroxynitrous acid was

followed by further reduction waves

at more negative potential. They are

most probably caused by the

reduction of decomposition

products, possibly nitrogen dioxide

and dinitrogen tetroxide. Nitrite as

60

o

oo

o o

o o

o

o

20 H o o

o

o

10- o

©

0 -I 1 1 1

0.0 1.0 2.0 3.0 4.0

(dE/dtf2 / (Vs 1)%

Fig. 1. Peak currents vs. square root of the

scan rate. 1 mM peroxynitrite with 0.1 M

phosphoric acid/phosphate buffers, goldelectrode.

Circles: pH 3.2, diamonds: pH 5.6.

T 1 1 1 1 1 1 1 1 1 1 r

600 400 200 0 -200 -400 -600

El my vs. Ag/AgCI, [CH = 0.1 M

Fig. 2. Reduction waves of peroxynitrousacid at pH 3.2, pH 5.6 and several scan

rates. 1 mM peroxynitrous acid, 0.1 M

phosphoric acid/phosphate buffers, a:

pH 5.6, 1 Vs"1; b: pH 3.2, 1 Vs"1; c: pH 5.6,

4Vs_1; d: pH3.2, 4Vs_1; e: pH5.6,

10Vs_1;f:pH3.2, 10 Vs"1.

40

<

- 30

62

the origin was ruled out, because it yielded no cathodic wave under the same

conditions, in the absence of peroxynitrous acid. Cyclic scans up to 80 Vs-1

showed no re-oxidation wave. The negative peak potential shift for a tenfold

increase in scan rate from 1 Vs-1 and 10 Vs-1 was 190 mV at pH3.2 and

210 mV at pH 5.6. The peak currents were always directly proportional to the

initial peroxynitrous acid concentration and to the square root of the scan rate

for both of the pH values (Fig. 1). The current peaks at pH 3.2 were higher than

at pH 5.6 (Fig. 2) for the same peroxynitrous acid concentrations and scan rates,

but hardly shifted by the different pH. The variation of maximum current

strength with pH was caused by a change in ohmic resistance of the metal-

solution interface depending on the buffer composition. At pH 5.6, there is a

considerable concentration of monohydrogen phosphate present, while at

pH 3.2, dihydrogen phosphate is the only anionic species.

Stopped-flow results.

Peroxynitrous acid oxidises

iodide to iodine like other peroxides

[2,3]. However, it cannot be

determined quantitatively by o 0.e

iodometric titration, since the

relative yield of its reaction with 4

iodide depends on concentrations.

With iodide concentrations below

1 mM, but in at least five-fold excess

peroxynitrous acid, 0.5over

equivalents of iodine were produced

rather precisely. Only with iodide

dilutions below 100 uM, the

isomerisation of the peroxynitrous

oO 0.5

1.0 10.0

[I ] / mM

100.0

Fig. 3. Relative yields of iodine from the

oxidation of iodide by peroxynitrousacid, vs. total iodide concentration. 0.1

M phosphoric acid/phosphate buffers,

circles: [HOONO]imtlai = 25 uM;

squares: [HOONO]imtlai = 5 uM.

63

acid competed with the iodide oxidation, and the iodine yield dropped below 0.5

equivalents. At iodide concentrations above 1 mM, the iodine yield rose above

0.5 equivalents and approached 1 equivalent with high excesses, but never

reached the exact value (Fig. 3). The rate law of the reaction was first order in

both peroxynitrite concentration and peroxynitrous acid concentration, which

means that only these molecules are involved in the rate-determining step. The

second-order rate constant at pH3.2 was found to be (2.8+0.3) x 104M_1s_1

based on the rate of formation of iodine, which was obtained by tranforming the

I3~ absorbance vs. time trace with the law of mass action of the equilibrium

I2 + F« I3~. The reaction rate hardly increased from pH 5.6 to 3.2, the

small change can be attributed to the fact that at pH5.6 about 10% of the

peroxynitrite are in the less reactive ONOO" form.

Discussion

The complete lack of re-oxidation wave and the linear dependence of the

peak current on the square root of the scan rate rule out a quasi-reversible

electrochemical reaction [4]. The remaining conceivable mechanisms are an

irreversible electrochemical reaction El5 case (a), possibly preceded by a

chemical equilibrium, CrEl5 case (b), or a reversible electrochemical reaction

followed by an irreversible chemical one, ErQ, case (c). Process (c) would show

k fRT^ino re-oxidation wave only if the kinetic parameter X = — ~~^\ ,

where kc is the

chemical rate constant based on mole fractions and v the scan rate, were larger

than 0.1 [5]. The negative peak potential shift with a tenfold increase in v should

29.6 mVbe for a truly reversible electrode reaction. Even if the reduction were

nJ

59.1 mVquasi-reversible, the shift should not exceed [6]. The experimental

value of about 200 mV is far too high to account for any concept of type (c). The

64

huge peak potential shift suggests na = 1 for the rate-determining

electrochemical step and a a smaller than 0.5. Processes (a) and (b) remain the

only suitable models. A possible preceding equilibrium (b) could be the

protonation of peroxynitrous acid,

HOONO + H+ < [H2OONO]+ (1)

an idea which is supported by the slightly increased isomerisation rate of

peroxynitrous acid at pH < 3 [7]. An estimated pKa of about 2 for equilibrium

(1) is compatible with this observation. The settling of the protonation

equilibrium would be much faster than the electrochemical reduction, and

therefore the protonation equilibrium should shift the peak potential position by

RT K Xh?oono+AE = ln;7-r with K

anaF K+l XHOono

from the value without equilibrium. K is the conditional equilibrium constant for

a given buffer pH and pKa = 2. With the assumption of a = 0.5, the peak

potentials should differ by about 200 mV for pH 3.2 and 5.6 with na = 1 or by

100 mV with na = 2, given the same scan rates. Experimentally, this was not

found, the peaks appeared at nearly identical potentials for both pH values and

the same scan rates (Fig. 2). The peak potential difference was only 15 mV at

10 Vs-1 and smaller at lower scan rates. The lack of a significant pH dependent

peak shift rules out a protonation equilibrium. Another preceding equilibrium

(b) could be the homolysis of peroxynitrous acid which is claimed to take place

at rate of one third of the isomerisation rate: [8,9]

HOONO +=+ N02* + OH* (2)

This reaction would be almost pH independent below the pKa of

peroxynitrous acid. It is, however, rather unlikely that the cathodic current

measured was caused by the reduction of free OH* radicals, since their steady-

state concentration would be very low due to the fast recombination reactions to

HN03, HOONO and H202. The shape of the cathodic peak and its potential shift

with scan rate also imply a kinetically inhibited reduction, which is not expected

65

for OH* radicals. Since there were no other potential reactants in the

experimental setup, a purely irreversible mechanism of type (a) is consistent

with the data obtained. Linear sweep peaks of reaction type (a) can be analysed

1.857 RTfor ana by means of the relation |EP - Ep/2| =

~

[10]. Averaging of ana

for 11 peaks recorded at pH 3.2 yields ana = 0.26+0.06, and for 9 peaks at

pH 5.6, ana = 0.27+0.03 was found. Plots of ln(ip) vs. Ep have slopes with

ana = 0.21+0.01 for pH 5.6 and ana = 0.22+0.02 for pH 3.2, in good agreement

with the values derived from |EP - Ep/2| .The low values suggest na = 1. The rate-

determining step in the electrochemical reduction of peroxynitrous acid is

obviously the transfer of an electron to the molecule. If the electron transfer

would yield a moderately stable product, the reaction would be at least quasi-

reversible, since a pH independent one-electron transfer is most likely an

elemental reaction. Therefore it must be followed by a very fast and irreversible

chemical decay

HOONO+ e" [HOONO]*" (3)

[HOONO]*" OH" + N02* (4)

[HOONO]*" + H+ H20 + N02* (5)

so that no quasi-reversible signal can be observed with ordinary voltammetry

equipment. The formation of [HOONO]*" as an intermediate is not that far¬

fetched as it looks at first glance: it has been shown by numerous researchers

that nitrite forms N02*2" in pulse radiolysis experiments, when it is treated with

hydrated electrons [11-16].

The iodide oxidation by peroxynitrous acid is hardly pH dependent in the pH

range where peroxynitrite is extensively protonated. The rate law for the iodide

oxidation is ~~vr = k [HOONO] [I"]. This rate law and the constant we

measured are in agreement with the earlier observations by Goldstein and

Czapski [2], but the relative yields are not compatible with the mechanism

66

postulated later [3]. If the iodide oxidation were a two-electron process as

suggested, a yield of one equivalent of iodine per equivalent of peroxynitrous

should be attainable at sufficient iodide excess, which is not the case (Fig. 3).

Since the rate law was determined consistently in all cases, the combined

observations imply that a single electron transfer to HOONO should be rate-

limiting, as it is the case in the electrochemical reduction. Therefore, it is rather

probable that the same intermediate product as in the electrochemical reduction

is formed, [HOONO]*". The radical anion would behave as desribed above. The

scheme explains easily why the first half-equivalent of I2 is formed

stochiometrically even at small iodide concentrations, while the formation of the

second half-equivalent remains incomplete even at high iodide excess, since the

reaction

N02* + I"< N02" + I* (6)

is an equilibrium with a constant of about 10" [17].

Since we found the electrochemical reduction of peroxynitrous acid to be pH

independent, its reduction potential should also not depend on pH.

Thermodynamic calculations of the one-electron reduction potential of

peroxynitrous acid by Merényi and Lind [18] and Koppenol and Kissner [19]

assumed that it is:

ONOOH + H+ + e +=* N02' + H20. (7)

Values of 2.14 V and 2.0 V were published by Merényi and Lind, [18] and

Koppenol and Kissner [19] respectively, and refer to standard conditions. These

values are based on determinations and estimates of the standard Gibbs energy

of formation of peroxynitrous acid, see Appendix. If the proton dependence is

eliminated from equation (7) to

ONOOH + e < » N02' + OH" (8)

the reduction potentials mentioned above can be recalculated. With new values

of Gibbs energy of formation of peroxynitrous acid from Merényi and co¬

workers [20], Lymar and Poskrebyshev [21] and our own value, 1.67 ± 0.03 V,

67

1.65 ± 0.02 V and 1.64 V, respectively, are obtained. For the calculation of the

reduction potential, the following values were used:

A/G°(N02', aq) = 63±1 kJ/mol [22] and AfG°(UO~, aq) = -157.3 kJ/mol [23].

These adjusted values are still insufflent to describe the oxidative properties

of peroxynitrous acid, since they are based on the assumption that N02" and HO

are in equilibrium with HOONO, which is surely not the case according to the

electrochemical results. In order to determine the correct one-electron reduction

potential of peroxynitrous acid by thermodynamic calculation, one would have

to know the Gibbs energy of formation of [HOONO]*".

68

HON02 -*—~-^ HO*, NO2

k

Appendix

Standard Gibbs energy of formation

In 2003 Merényi and coworkers adjusted their former value by studying the

equilibrium ONOOH + H20+ H+ +=+ H202 + HN02 + H+ (9)

and derived a standard Gibbs energy of formation of +(67.4 ± 1.3) kJ mol"1 for

peroxynitrite anion [20].

In the same year Lymar and Poskrebyshev published a value for the standard

Gibbs energy of formation of +(30.8 ± 1) kJ mol"1 [21]. The value was obtained

from a study of the reaction

ONOO « » NO' + 02' (10)

by pulse radiolysis in the presence of methyl viologen.

Three determinations of the

enthalpy of formation of

peroxynitrite have been+162 + 87

published; they originate from

calorimetric measurements of the

heat of isomerisation of H*,O-N00r ^— 0=NOOH

-, -, , -, , TT , f.Fie. 4. Enthalpy cycle according to Mahoney

peroxynitrite to nitrate. Heats 01n„-m 1 ; j

+u- uv +- ah^ J(1970), as completed in this publication. All

isomerisation of -38.8 ± 2 kcal values ink! mol1.

mol"1 [24], -37.8 ± 0.4 kcal mol"1 [25] and -163 ±4 kJ mol"1 [26] have been

reported. In relation to the standard enthalpy of formation of nitrate,

-205.0 kJmol"1 [27], the standard enthalpies of formation for peroxynitrite can

be calculated: -43 ±8,-47 + 2 and -42 ± 4 kJmol"1, respectively. We adopt a

value of -43 ± 4 kJ mol"1, where the uncertainty covers all published values.

With this value and the enthalpy of ionisation of peroxynitrous acid, the

thermodynamic cycle of Mahoney [28] can be completed, see Fig. 4. We note

that the enthalpy of homolysis in water, +87 kJ mol1, is quite close to that

computed for the gas phase, +94 kJ mol1 [29], while for the 0-0 bond

69

dissociation energy in the gas phase a value of +82 kJ mol1has been reported

[30].

In 1992, Koppenol et al. estimated that S°(0=NOO") is +188 J K"1 mol"1,

based on S°(N03") = +146 J K"1 mol"1 plus the difference between S°(ONOO*)

and S°(N03'), 42 J K^mol"1 [31]. This resulted in an estimated standard Gibbs

energy of formation of peroxynitrite of +42 kJ mol1, a value that precludes

homolysis. We were criticised [18], because our estimate implies that the

entropies of solution for nitrate and peroxynitrite are the same. Adoption of a

thermodynamic cycle regarding the entropy of peroxynitrite and peroxynitrous

acid [18], with incorporation of the entropy of ionisation, -16 J K"1 mol"1

resulted in an estimate of+58 ± 10 kJ mol"1 for AfG°(ONOO")aq [19]. The large

error is caused for a large part by that in determination of the enthalpy of

ionisation of peroxynitrous acid.

This estimate can be improved. For the entropy of formation of peroxynitrous

acid in various conformations in the gas phase values between +274 and

+281 JK"1 mol"1 have been derived by ab initio methods [29]. These entropies

are significantly lower than our earlier estimate of +301 J K"1 mol"1 [19],

derived from the entropy of oxoperoxonitrogen(»), +293 J K"1 mol"1 [33]. The

+ 293 0=NOO- - 0=NOOHn +278y

8

179

105

smaller entropy is_ ^ 5

probably due to the

ability of

peroxynitrous acid to +285 0=NOO

form an internal

hydrogen bond. We

adopt a value of +106 0=NOO-aq^ — 0=NOOHaq +173

J mo'

^ eFig 5. Entropy cycle [32] with revised values. All values in

average of the two eis-

isomers, from which follows a value of +106 J K"1 mol"1 for ^(ONOO" )aq, see

Fig. 5, and this, in turn, results in a AfG°(ONOO")aq of +65 kJ mol" \ However,

70

this result has still the same uncertainty as the old estimate, and the entropy of

hydration of the peroxynitrite anion is quite negative when compared with other

four-atom monovalent anions [34].

Another estimate can be obtained from a cycle centred on the Gibbs energy of

formation of peroxynitrous acid in the gas phase calculated by ab initio methods

[29]. The Gibbs energy of homolysis for cw,cw-ONOOH in the gas phase is

given as +50.6 kJ mol" \ We note that the entropy change ofthat reaction results

in a TAS term of +45.2 kJ mol"l, which is more than +(38 ± 2) kJ mol"1 as found

for similar reactions. The Gibbs hydration energies of nitrogen dioxide and

hydroxyl are known (together +3 kJ mol"l) and that of peroxynitrous acid can be

estimated as follows. The cost of forming a cavity in water is +17 kJ mol"1,

which is compensated by ca. -8 kJ mol"1 for every hydrogen bond between

solute and solvent [35]. For peroxynitrous acid we assume that it can form half

as many hydrogen bonds as hydrogen peroxide, for which a Gibbs hydration

energy of - 28 kJ mol"1 is calculated from the Henry constant of 7.74 x 104 [36].

These considerations result in a Gibbs energy of hydration of peroxynitrous acid

of - 6 kJ mol"l, and a Gibbs energy of formation of +29 kJ mol"1 (+65 kJ mol"1

for ONOO").

Although these two estimates yield the same value for the standard Gibbs

energy of formation of peroxynitrite, they both involve parameters from an ab

initio study [29] Given the assumption regarding the hydration energy of

peroxynitrous acid, we estimate the error to be smaller than the previous

estimate but still more than 5 kJ mol- 1.

Acknowledgement - This project was supported by the ETH and the Swiss

Nationalfonds.

71

References

[1] Kissner, R.; Nauser, T.; Bugnon, P.; Lye, P. G.; Koppenol, W. H.

Formation and properties of peroxynitrite studied by laser flash

photolysis, high pressure stopped flow and pulse radiolysis. Chem. Res.

Toxicol. 10:1285-1292; 1997.

Goldstein, S.; Czapski, G. Direct and indirect oxidations by peroxynitrite.

Inorg. Chem. 34:4041-4048; 1995.

Goldstein, S.; Meyerstein, D.; van Eldik, R.; Czapski, G. Spontaneousreactions and reduction by iodide of peroxynitrite and peroxynitrate:Mechanistic insight from activation parameters. J. Phys. Chem. A

101:7114-7118; 1997.

Matsuda, H.; Ayabe, Y. Zur Theorie der Randies-Sevcikschen

Kathodenstrahl-Polarographie. ZElektrochem 59:494-503; 1955.

Nicholson, R. S.; Shain, I. Single scan and cyclic methods applied to

reversible, irreversible, and kinetic systems. Anal. Chem. 36:706-723;1964.

Nadjo, L.; Saveant, J. M. Linear sweep voltammetry: kinetic control by

charge transfer and/or secondary cehmical reactions. J. Electroanal. chem.

48:113-145; 1973.

Benton, D. J.; Moore, P. Kinetics and mechanism of the formation and

decay of peroxynitrous acid in perchloric acid solutions. J. Chem. Soc. (A)3179-3182; 1970.

Merényi, G.; Lind, J.; Goldstein, S.; Czapski, G. Peroxynitrous acid

homolyzes into OH and N02 radicals. Chem. Res. Toxicol. 11:712-713;1998.

Lymar, S. V.; Khairutdinov, R. F.; Hurst, J. K. Hydroxyl radical

formation by 0-0 bond homolysis in peroxynitrous acid. Inorg. Chem.

42:5259-5266; 2003.

Bard, A. J.; Faulkner, L. R. Electrochemical Methods. John Wils & sons

Inc.; 1980:222-224.

Elliot, A. J.; McCracken, D. R.; Buxton, G. V.; Wood, N. D. Estimation

of rate constants for nearly-diffusion-controlled reactions in water at high

temperature. J. Chem. Soc. ,Faraday Trans. 86:1539-1547; 1990.

Broszkiewicz, R. K. The pulse radiolysis study ofNaN02 and NaN03

solutions. Bull. Acad. Pol. Sei.,Ser. Sei. Chim. 24:221-229; 1976.

72

[13] Fei, N. S. Determination of the rate constant of a reaction of e-aq with an

acceptor in neutral water by pulsed method. Khimiya Vysokikh Energii

4:178-180; 1970.

[14] Cercek, B. Activation energies for reactions of the hydrated electron.

Nature 223:491-492; 1969.

[15] Baxendale, J. H.; Fielden, E. M.; Keene, J. P. The pulse radiolysis of

aqueous solutions of some inorganic compounds. Proc. Roy. Soc.,Ser. A

286:320-336; 1965.

[16] Thomas, J. K.; Gordon, S.; Hart, E. J. The rates of reaction of the hydratedelectron in aqueous inorganic solutions. J. Phys. Chem. 68:1524-1527;1964.

[17] Barkett, A.; Ottolenghi, M. Laser flash photolysis of aqueous triiodide

solution. Mol. Photochem 6:253-261; 1974.

[18] Merényi, G.; Lind, J. Thermodynamics of peroxynitrite and its C02-

adduct. Chem. Res. Toxicol. 10:1216-1220; 1997.

[19] Koppenol, W. H.; Kissner, R. Can ONOOH undergo homolysis. Chem.

Res. Toxicol. 11:87-90; 1998.

[20] Merényi, G.; Lind, J.; Czapski, G.; Goldstein, S. Direct determination of

the Gibbs' energy of formation of peroxynitrous acid. Inorg. Chem.

42:3796-3800; 2003.

[21] Lymar, S.V.; Poskrebyshev, G. A. Rate of ON-OO- bond homolysis and

the Gibbs energy of formation of peroxynitrite. J. Phys. Chem. A

107:7991-7996; 2003.

[22] Stanbury, D. M. Reduction potentials involving inorganic free radicals in

aqueous solution. Adv. Inorg. Chem. 33:69-138; 1989.

[23] Ross, P. N. Oxygen. In Bard, A. J.; Parsons, R.; Jordan, J. eds. Standard

Potential in Aqueous Solution. New York: Marcel Dekker, INC; 1985:49-

66.

[24] Ray, J. D. Heat of isomerization of peroxynitrite to nitrate and kinetics of

isomerization of peroxynitrous acid to nitric acid. J. Inorg. Nucl. Chem.

24:1159-1162; 1962.

[25] Bohle, D. S.; Hansert, B.; Paulson, S. C; Smith, B. D. Biomimetic

synthesis of the putative cytotoxin peroxynitrite, ONOO", and its

characterization as a tetramethylammonium salt. J. Am. Chem. Soc.

116:7423-7424; 1994.

73

[26] Manuszak, M.; Koppenol, W. H. The enthalpy of isomerization of

peroxynitrite to nitrate. Thermochim. Acta 273:11-15; 1996.

[27] Wagman, D. D.; Evans, W. H.; Parker, V. B.; Schumm, R. H.; Halow, I.;

Bailey, S. M.; Churney, K. L.; Nuttal, R. L. Selected values for inorganicand CI and C2 organic substances in SI units. J. Phys. Chem. Ref. Data

11 (Suppl. 2)37-38; 1982.

[28] Mahoney, L. R. Evidence for the formation of hydroxyl radicals in the

isomerization of pernitrous acid to nitric acid in aqueous solution. J. Am.

Chem. Soc. 92:5262-5263; 1970.

[29] McGrath, M. P.; Rowland, F. S. Determination of the barriers to internal

rotation in ONOOX (X = H, CI) and characterization of the minimum

energy conformers. J. Phys. Chem. 98:1061-1067; 1994.

[30] Olson, L. P.; Bartberger, M. D.; Houk, K. N. Peroxynitrate and

peroxynitrite: A complete basis set investigation of similarities and

differences between these NOx species. J. Am. Chem. Soc. 125:3999-

4006;2003.

[31] Koppenol, W. H.; Moreno, J. J.; Pryor, W. A.; Ischiropoulos, H.;

Beckman, J. S. Peroxynitrite, a cloaked oxidant formed by nitric oxide

and superoxide. Chem. Res. Toxicol. 5:834-842; 1992.

[32] Koppenol, W. H.; Kissner, R. Can 0=NOOH undergo homolysis? Chem.

Res. Toxicol. 11:87-90; 1998.

[33] Benson, S. W. Thermochemical Kinetics: New York John Wiley & Sons;1976.

[34] Marcus, Y. The hydration entropies of ions and their effects on the

structure of water. J. Chem. Soc., Faraday Trans. 1 82:233-242; 1986.

[35] Schwarz, H. A.; Dodson, R. W. Equilibrium between hydroxyl radicals

and Thallium (II) and the oxidation potential of OH(aq). J. Phys. Chem.

88:3643-3647; 1984.

[36] Sander, R. Henry's Law constants. In Linstrom, P. J.; Mallard, W. G. eds.

NIST Chemistry WebBook, NIST Standard Reference Database Number

69. Gaithersburg, MD 20899: National Institute of Standards and

Technology; 2003.

74

Supporting Information

0.2 M phosphatebu

Figure IS. The equipment for stopped-flow electrochemistry consist of two

pumps, controlled by computer, an electrochemical cell, potentiostat and a

computer for the collection and evaluation of the data.

peroxynitrite

outlet

auxiliaryelectrode

0.2 M dihydrogenphosphate

Figure 2S. Details of the measuring cell.

75

76

7. RAPID SCAVENGING OF PEROXYNITROUS

ACID BY MONOHYDROASCORBATE

(Free Radical Biol. Med. 35:1529-37; 2003)

CHRISTOPHE R. KURZ, REINHARD KISSNER, THOMAS NAUSER,

DANIEL PERRIN and WILLEM H. KOPPENOL*

Laboratorium für Anorganische Chemie, Eidgenössische Technische

Hochschule Zürich (Hönggerberg), Switzerland

Abstract— The reaction of peroxynitrous acid with monohydroascorbate,

over the concentration range of 250 uM to 50 mM of monohydroascorbate at

pH 5.8 and at 25°C, was reinvestigated and the rate constant of the reaction

found to be much higher than reported earlier (Bartlett, D.; Church, D. F. ;

Bounds, P. L.; Koppenol, W. H., The kinetics of oxidation of L-ascorbic acid

by peroxynitrite, Free Radical Biol. Med. 18:85-92; 1995, Squadrito, G. L.;

Jin, X.; Pryor,W. A., Stopped-flow kinetics of the reaction of ascorbic acid

with peroxynitrite, Arch. Biochem. Biophys. 322:53-59; 1995). The new rate

constants at pH 5.8 are kx= 1 x 106M"V and k.x= 500 s"1 for 25°C, and

h= 1.5 x 106 M"V and k.x= 1 x 103 s"1 for 37°C. These values indicate that

even at low monohydroascorbate concentrations most of peroxynitrous acid

forms an adduct with this antioxidant. The mechanism of the reaction

involves formation of an intermediate, which decays to a second intermediate

with an absorption maximum at 345 nm. At low monohydroascorbate

concentrations, the second intermediate decays to nitrate and

monohydroascorbate, while at monohydroascorbate concentrations greater

than 4 mM, this second intermediate reacts with a second

monohydroascorbate to form nitrite, dehydroascorbate, and

monohydroascorbate. EPR Experiments indicate that the yield of the ascorbyl

radical is 0.24 % relative to the initial peroxynitrous acid concentration, and

77

that this small amount of ascorbyl radicals is formed concomitantly with the

decrease of the absorption at 345 nm. Thus, the ascorbyl radical is not a

primary reaction product. Under the conditions of these experiments, no

homolysis of peroxynitrous acid to nitrogen dioxide and hydroxyl radical is

observed. Aside from monohydroascorbate's ability to "repair" oxidatively

modified biomolecules, it may play a role as scavenger of peroxynitrous acid.

Results

Stoppedflow studies

Kinetic traces recorded at 345 nm, 25°C and at pH 5.8 after mixing of

peroxynitrite with different concentrations of monohydroascorbate (0.25 -

50 mM) show an increase in absorption during 20 - 40 ms, followed by

decrease. At monohydroascorbate concentrations higher than 1 mM, the rate

of this previously unobserved increase in absorption depends on

monohydroascorbate (Fig. 1), while a dependence on peroxynitrous acid is

observed over the whole concentration range (not shown).

The decrease in absorption, which begins 50 - 100 ms after mixing, was

measured before at higher monohydroascorbate concentrations [1,2]. The

experiments at these concentrations (Fig. 1, traces C and D) are in agreement

with the cited literature, and show a faster decrease of absorption at higher

monohydroascorbate concentrations. The rate of decrease is not affected by

the peroxynitrous acid concentration.

The insert of Fig. 1 shows that directly after mixing there is a decrease in

absorption. This process clearly had started already during the mixing and

was difficult to quantify. At 345 nm, monohydroascorbate does not absorb,

and the absorption directly after mixing and that after ca. 40 ms are

significantly higher than those expected for peroxynitrous acid alone, see

below. These observations indicate that there may be two intermediates, Im!

and Im2, of which Im2 has an absorption maximum at 345 nm.

78

0.0 0.5 1.0 1.5 2.0 2.5 3.0time /s

Fig. 1. Time course of the absorption at 345 nm after mixing

monohydroascorbate with peroxynitrous acid. 25 uM peroxynitrous acid mixed

with A, 250 uM; B, 1 mM; C, 8 mM, and D, 50 mM monohydroascorbate. All

concentrations indicated are those after mixing. All traces were recorded at

345 nm, 25°C and pH 5.8 (50 mM phosphate buffer). The traces C and D are in

agreement with earlier work [1,2] where higher ONOOH and

monohydroascorbate concentrations were used. At lower monohydroascorbateconcentrations (trace A and B) no dependence on the monohydroascorbateconcentration is observed. Each curve is an average of 12 measurements. Inset:

A fast decrease in absorption that depends on the monohydroascorbateconcentration is observed during the first 6 ms.

The transient absorption with a maximum at 345 nm (Fig. 2, spectra A and

B) has not been observed before. We note that the spectrum of the ascorbyl

radical (Fig. 2, trace C), determined by pulse radiolysis of a dinitrogen

monoxide saturated solution of 0.50 mM monohydroascorbate, shows a

maximum at 360 nm in agreement with the literature [3,4].

79

The absorption at

345 nm, 4 ms after

mixing (Fig. 2,

spectrum A), is about

50% higher than that

expected for

peroxynitrous acid at

zero time. The

extinction coefficient of

peroxynitrous acid at

that wavelength is

78 M^cm"1 (Kissner,

2001, unpublished).

Monohydroascorbate

itself does not absorb at

that wavelength. If we

separate the absorption

at 345 nm into two components, Ai and A2, where Ai is the absorption 4 ms

after mixing, and A2 the maximum absorption at about 50 ms minus A1; then

Ai increases linearly with the peroxynitrite concentration, up to a two-fold

excess (Fig. 3). A2, which is much smaller than Ai, shows a saturation effect at

a ratio of [ONOOH] : [HAsc-] = 1:1 (inset Fig. 3). These observations suggest

that a 1:1 peroxynitrite-dehydroascorbate adduct was formed.

The rates of decay of the Im2 and monohydroascorbate, observed at

345 nm and 265 nm, respectively, are similar (not shown). Both rates are first-

order in monohydroascorbate. However, Im2is formed during the first 100 ms

and its decay lags behind that of monohydroascorbate absorption which starts

do diminish directly after mixing.

80

320 340 360

1/nm

380 400

Fig. 2. Absorption spectrum of Im2 compared with that

of peroxynitrous acid and the ascorbyl radical. Absorption

spectra of Im2 4 ms (A) and 52 ms (B) after mixing 1 mM

peroxynitrite with 10 mM monohydroascorbate at pH5.8

(50 mM phosphate buffer) at 25°C. All concentrations

indicated are those after mixing. Controls: Absorption

spectrum of the ascorbyl radical (C) 50 us after irradiation

of a N2Û-saturated 500 uM monohydroascorbate solution

at pH 5.8 (50 mM phosphate buffer) with 2 MeV electrons.

Absorption spectrum of a 1 mM peroxynitrous acid

solution at pH 3 (D). Monohydroascorbate has no

absorption at 345 nm.

o9 -

o8 -

o

s? -

x

§5- o

Pi

-O

a4- O

5" On

2 -

1 -

o -

300

UM ONOOH

400 500

0 1

-- 5

n

O

s

J"SB

as

600

time /s

Fig. 3. Kinetics of the formation of Im2 as a function of the peroxynitrous acid

concentration. Ascorbic acid, 250 uM, mixed with peroxynitrous acid A, 25 uM;

B, 50 uM; C, 125 uM; D, 250 uM; E, 375 uM; F, 500 uM. Inset: Absorption after mixing

monohydroascorbate (250 uM final concentration) with variable concentrations of

peroxynitrous acid final concentrations separated into Ai (; absorption 4 ms after mixing)

and A2 (o; absorption difference between Ai and maximal absorption after about 50 ms ).

From measurements over a wide range of monohydroascorbate

concentrations, it is clear that the rates of formation and decay of Im2 are

dependent on the monohydroascorbate concentration. It is implied that the

rate of formation of the first intermediate also depends on the

monohydroascorbate concentration. As shown in Figure 1, Im2 is already

present at very low monohydroascorbate concentrations, while at higher

monohydroascorbate concentrations the rate of formation increases with a

slope of ka = 1.6 x 103 M_1s_1 (Fig. 4).

81

We do not wish to imply that the intercept at zero monohydroascorbate

concentration is real. First, under this condition there is no reaction and there

are no intermediates. Second, the points in Figure 4a are based on

absorbances between 5-15 ms, which may well contain contributions from

Iirii. The error bars only

show that the measurements i60

were reproducible. It seems

likely though that directly

after mixing there is already

a significant amount of Im2.

The decay of the

intermediate has a

monohydroascorbate-

independent, kß =0.85 s-1,

and dependent rate constant,

ytr=95M1s"1(Fig. 3b).

As shown in the inset of

Figure 1, rapid absorption

decay takes place within

4 ms after mixing. Control b)

experiments without

peroxynitrite, but with added

nitrite or nitrate, show no

transient at 345 nm, nor do

experiments in which

peroxynitrite is mixed with

0 10 40 5020 30

mM Hasc"

Fig. 4. Formation (a) and decay (b) rate constants

of Im2. Measurements were carried out at 345 nm

under pseudo-first-order condition. Conditions:

25 uM peroxynitrite, 50 mM phosphate buffer,

pH5.8, T = 25°C. The error bars represent the 95 %,

(2s) error margin. All concentrations indicated are

those after mixing.

buffer, which confirms that an intermediate is formed as a product of the

reaction of monohydroascorbate with peroxynitrous acid is formed.

82

Furthermore, we observed no influence of nitrite, nitrate, or dioxygen on the

kinetics (not shown).

EPR studies

We observed a small

signal with a g-value of

2.005 which is due to

the ascorbyl radical.

This signal was not

detected without

monohydroascorbate or

peroxynitrite. Directly

after mixing

25 uM peroxynitrite and

250 uM monohydro¬

ascorbate, no EPR

signal was seen; instead,

this signal rises for

approximately

2 seconds and

subsequently disappears

by second-order kinetics within 10 seconds (Fig. 5, trace B). Compared with

the optical kinetic trace recorded at 345 nm (Fig. 5, trace A), it is clear that

the concentration of ascorbyl radicals increases while that of Im2 decreases.

The total ascorbyl radical concentration formed during the reaction was

estimated at only 60 nM. Because the ascorbyl radical decays by second-order

kinetics [3,5], its concentration could be estimated to be 20 nM at t = 2 s.

Special care was taken to avoid contamination by dioxygen and transition

metals, which also produce ascorbyl radicals [6].

83

time /s10

Fig. 5. Comparison of EPR and UV-Vis results.

The decay at 345 nm (A) is compared with an EPR

trace, B, that shows formation and disappearance of

ca. 20 nM ascorbyl radical. Trace B increases

concomitant with the decrease of trace A. Curve A is

an average of 12 measurements and curve B an

average of 10 measurements. Both measurements

were made at initial concentrations of 25 uM

peroxynitrite and 250 uM monohydroascorbate in

50 mM phosphate buffer at pH 5.8. All concentrations

indicated are those after mixing.

Product analysis

When 25 uM peroxynitrite and increasing amounts (25 uM - 50 mM) of

monohydroascorbate are mixed, one observes that nitrite increases,

concomitant with a decrease in nitrate. At 2 mM monohydroascorbate equal

amounts of nitrite and nitrate are produced (not shown), and at concentrations

higher than 10 mM saturation is observed. At lower monohydroascorbate

concentrations, most of the peroxynitrous acid is converted to nitrate. Without

monohydroascorbate, peroxynitrous acid yields already 13 % nitrite at pH 5.8,

in agreement with the literature [7].

Discussion

Earlier work

The reaction of peroxynitrous acid with monohydroascorbate has already

been studied by Bartlett et al. [1] and by Squadrito et al. [2]. They concluded

that the reaction was first-order in peroxynitrous acid and in

monohydroascorbate with a rate constant of 235 M" V \ The rapid increase in

absorption during the first 100 ms after mixing (Fig. 1) was not seen by these

workers because vibrations, generated when the flow was stopped, prevented

accurate measurements during 50 ms after mixing. Based on the observation

of the EPR spectrum of the ascorbyl radical, Bartlett et al. [1] concluded that

this radical was a product. We found only a very small EPR signal, and the

corresponding ascorbyl concentration would amount to much less than one

percent of the peroxynitrous acid concentration. The intermediate Im2 which

exhibits an absorbance maximum at 345 nm is thus not identical to the

ascorbyl radical which has an absorption maximum at 360 nm (Fig. 2).

Above pH 5.8 the reaction proceeds more slowly, as was reported earlier

[1]. A manuscript on the reaction of peroxynitrite with monohydroascorbate

near neutral pH is in preparation.

84

ONOO" + H"1

Kr, k-6

ONOOH + HAsc^ ^^

k-i

N(V + H/^

HAsc^ + DAsc + NO?"+H7O9 T Ilj\

Ktz i

ko

[111,^1^ Illl2k_?

kj

HAsc"

345

HAsc"' + NO3" + H

Scheme 1.

Intermediates

The important findings in this paper are that the absorbance at 345 nm ca.

5 ms after mixing is higher than that expected for peroxynitrous acid alone,

that it continues to rise for ca. 0.1 s and then decreases; both processes are

monohydroascorbate-dependent. The increase in absorbance is also dependent

on the peroxynitrous acid concentration. Our attempt to fit these observations

to a reaction scheme that includes a single intermediate was not successful. It

appeared necessary to invoke a second intermediate, appearing prior to Im2,

as shown in Scheme 1. Evidence for the first intermediate is two-fold: at high

concentrations of monohydroascorbate, a decrease in absorption is observed

directly after mixing (inset Fig. 1), and, at that time, there appears to be

already a significant amount of Iirii. We have no data that would provide

information regarding the precise structures of these intermediates. As shown

in inset of Figure 3, the absorption difference A2 shows a saturation effect at

[ONOOH] > [HAsc-]. From this, as suggested by Bartlett et. al. [1], we

conclude that Im2 is a 1:1 adduct of peroxynitrous acid and

monohydroascorbate. We already mentioned that Im2 is not the ascorbyl

radical, as the EPR signal of the ascorbyl radical increases while the

concentration of Im2 decreases (Fig. 5).

Kinetics andMechanism

85

We postulate that Imi is formed very rapidly from monohydroascorbate

and peroxynitrous acid, that Im2 is formed from Imi, and that the former

decays spontaneously and by reaction with a second monohydroascorbate

molecule. Since the concentration of the second intermediate (Im2) increases

with that of peroxynitrous acid (Fig. 3), we conclude that Imi and Im2 are in

equilibrium. We observed the reaction at 345 nm, the absorbance maximum

of Im2; we assume that Imi also absorbs at this wavelength. The rates of

formation of Imi and Im2 increase with the monohydroascorbate

concentration, as shown in Figure 4a. These rates are extracted from the

increase in absorption 5 -15 ms after mixing and are subject to a considerable

error. We assume that, when the monohydroascorbate concentration

approaches zero, kx in Figure 4a is zero. We cannot verify that, because below

500 uM monohydroascorbate no significant rise is observed. The rate

constant for the increase in absorption at 345 nm at monohydroascorbate

concentrations larger than 500 uM (Fig. 4a) is 1.6 x 10 M s.

The decay of Im2, follows the following rate law:

ku,obServed =kß +kr[ascorbate] with kp =0.85 s"1 and ky =95 M_1s_1

The monohydroascorbate-independent pathway yields nitrate and

monohydroascorbate; the decay rate constant kß is smaller than the

isomerization rate constant of peroxynitrous acid at pH 5.8, which is another

indication that at least one intermediate is present. From the faster decay of

Im2 at increased monohydroascorbate concentrations and the product analysis,

we conclude that Im2 interacts with another monohydroascorbate molecule

and produces dehydroascorbate, monohydroascorbate and nitrite.

86

Rate

constantRate value Remark Compound

Absorption Initial

coefficient concentration

ki lxlO^'V a HAsc 0 M'W1250 u.M-

50 mM

k.i 5xl02 s"1 a HOONO 82 M'W1 25 uM

k2 40 s"1 a ONOO" 685 M'W1 —

k.2 5 s"1 a N03 0 M'W1 —

k3 1.2 s"1 b N02 5 M'W1 —

k4 0.85 s"1 c DAsc 0 M'W1 —

k5 lx^M'V c Imi 250 M'W1 —

k6 lxlO^-^s"1 d Im2 310 M'W1 —

k.6 1.58xl03s_1 e

Table 1. Values used for the simulation of the reaction of peroxynitrous acid

with monohydroascorbate at 345 nm and 25°C. a, determined by simulation; b,Kissner et all. [7]; c, rate constant determined by decay of Im2; d, rate constant for

protonation; e, rate constant determined by pKa [13].

Kinetic traces of the reaction between peroxynitrous acid and

monohydroascorbate were simulated for 25°C and 37°C based on Scheme 1.

These simulations were carried out for variable monohydroascorbate

concentrations at fixed concentrations of peroxynitrous acid and reasonable to

good agreement with the experimental results was found. The rate constants

and extinction coefficients used in the simulation are listed in Table 1.

Importantly, the forward rate constants for the reaction of peroxynitrous acid

with monohydroascorbate at pH 5.8 are 1 x 106 M_1s_1 and 1.5 x 106 M_1s_1 at

25°C and 37°C, respectively. The rate constants for the backward reaction are

5 x 102 s"1 and 1 x 103 s"1 at 25°C and 37°C, respectively (not shown). From

these rate constants it can be calculated that at 1 uM peroxynitrous acid and

87

500 uM monohydroascorbate every second peroxynitrous acid molecule is

scavenged by an monohydroascorbate. While the agreement between

simulation and experiment is satisfactory, we do not exclude the possibility of

improvements to this model.

Homolysis or one-electron oxidation?

In the literature contrasting views are expressed with respect to the issue of

peroxynitrous acid to nitrogen dioxide and hydroxyl radicals [7-13]. Both

hydroxyl and nitrogen dioxide radicals react with high rate constants of

7.0 x KflVrV1 [3] and 3.5 x 107 M'V [14] respectively, with

monohydroascorbate and yield the ascorbyl radical. The reaction of

peroxynitrous acid with low concentrations of monohydroascorbate takes

place during 3-4 s (Fig. 1), which is also the time span of the isomerization of

peroxynitrous acid to nitrate at 25°C.

Under these conditions we estimate from our EPR experiments that the

maximal concentration of ascorbyl radical formed during the reaction is about

20 nM. This concentration is too small to be part of the main reaction path. It

is possible that the ascorbyl radicals observed are not produced from the

reaction between monohydroascorbate and peroxynitrous acid at all, but from

the oxidation of monohydroascorbate by dioxygen, a reaction catalyzed by

transition metals present as trace contamination [6,15].

A mechanism consisting of two fast sequential one-electron oxidations,

one by peroxynitrous acid and one by nitrogen dioxide, can be excluded,

because both oxidations would yield ascorbyl radicals.

It is conceivable that the ascorbyl radical was not observed because it

reacted rapidly with another peroxynitrous acid. In that case there would have

been an increase in ascorbyl radicals at higher monohydroascorbate

concentrations. No such increase was observed.

88

Physiological relevance

Can monohydroascorbate compete with carbon dioxide at physiological

pH and temperature? The reaction of peroxynitrite with carbon dioxide

proceeds with a rate constant of 5.8 x 104 M_1s_1 at 37°C [16]. If we assume

that in a very short time 10 uM peroxynitrite were formed in a cell, then the

product distribution of rate constants and concentrations for removal via the

carbon dioxide or the monohydroascorbate pathway can be simulated by

adding the reaction of peroxynitrite with carbon dioxide to Scheme 1. A

simulation for 1 mM carbon dioxide, 100 uM monohydroascorbate (human

serum) and 10 uM peroxynitrite at pH7.3 and 37°C, shows that

monohydroascorbate diverts only ca. 3.5 % of the peroxynitrite (not shown).

At monohydroascorbate concentrations of 10 mM, as found in neurons,

peroxynitrite is trapped to an extent of 67 %. If the reaction between

peroxynitrite and carbon dioxide involves an equilibrium [17], then

monohydroascorbate will, even at low monohydroascorbate concentrations,

curtail the reaction with carbon dioxide. Our studies were carried out at pH

5.8, which may be relevant to infections, where the pH may be significantly

lower than 7.3. Under such conditions monohydroascorbate at physiological

concentrations may provide protection against peroxynitrous acid induced

oxidation and nitration [18-20]. A study of the pH dependence on the reaction

of monohydroascorbate and peroxynitrite is in progress.

89

Speculations as to the structures of the intermediates.

Some efforts to establish the structures of the intermediate were made, but

failed. Therefore we are able to suggest only structures that are based on the

known reaction pathways of peroxynitrite and monohydroascorbate.

A possible reaction scheme between monohydroascorbate and peroxynitrous

acid, including a monohydroascorbate dependent and independent decay

pathway, is drawn in Figure 1.

HO.

H*

HO

"OH h

^°^?—o+

0*)0

\ /u \\J /

/ \ - N-0

HO

HO

?>O

H

HO \ 0\

H V—-SN-0

o-o

A

+ HAsc

HO OIO

H

+ N09

HO O

HO PH ,0

Fig. 1

HAsc

DAsc

H20NO"

HAsc

H20NO"

HO O

H

O" O

90

Another possibility is that peroxynitrous acid transfer an oxygen to the C-C

double bond or to the C-0 bound.

HO O

H

o bN-0 HO o O" +HT

+ NO,

Fig. 2

HAsc"

NO3-H+

It is also possible that peroxynitrous acid adds to the double bond, as shown

in Figure 3.

We do not expect that the reaction product in figure 3 is stable, but we do not

know to what products it would decay.

HO.

H'OH

O

HO f

O

O"ON

00H

-O"

OH

,0+ H0,

OON'

OH )=0

N-0// \

o o/

H

Fig. 3

None of these three reaction mechanisms involves three intermediates, and we

must conclude that their structures remain unsolved.

91

Acknowledgement - This project was supported by the ETH and the Swiss

Nationalfonds.

We thank Dr. P.L. Bounds for helpful discussions.

Abbreviation

Imi : First intermediate of the reaction of peroxynitrous acid with

monohydroascorbate.

Im2: Second and last intermediate of the reaction of peroxynitrous

acid with monohydroascorbate.

Peroxynitrite: Means the protonated and the deprotonated form.

References

[1] Bartlett, D.; Church, D. F.; Bounds, P. L.; Koppenol, W. H. The

kinetics of the oxidation of L-ascorbic acid by peroxynitrite. Free

Radical Biol. Med. 18:85-92; 1995.

[2] Squadrito, G. L.; Jin, X.; Pryor, W. A. Stopped-flow kinetic study of

the reaction of ascorbic acid with peroxynitrite. Arch. Biochem.

Biophys. 322:53-59; 1995.

[3] Schöneshöfer, M. Pulsradiolytische Untersuchung zur Oxidation der

Ascorbinsäure durch OH-Radikale und Halogen-Radikal-Komplexe in

wässriger Lösung. Z. Naturforsch. 27b:649-659; 1972.

[4] Bielski, B. H. J.; Comstock, D. A.; Bowen, R. A. Ascorbic acid free

radicals. I. Pulse radiolysis study of optical absorption and kinetic

properties. J. Am. Chem. Soc. 93:5624-5629; 1971.

[5] Bielski, B. H. J.; Allen, A. O.; Schwarz, H. A. Mechanism of

disproportionation of ascorbate radicals. J. Am. Chem. Soc. 103:3516-

3518; 1981.

[6] Buettner, G. R.; Jurkiewicz, B. A. Catalytic metals, ascorbate and free

radicals: Combinations to avoid. Radiât. Res. 145:532-541; 1996.

92

[7] Kissner, R.; Koppenol, W. H. Product distribution of peroxynitrite

decay as a function of pH, temperature, and concentration. J. Am.

Chem. Soc. 124:234-239; 2002.

[8] Maurer, P.; Thomas, C. F.; Kissner, R.; Rüegger, H.; Greter, O.;

Röthlisberger, U.; Koppenol, W. H. Oxidation of nitrite by

peroxynitrous acid. J. Phys. Chem. A 107:1763-1769; 2003.

[9] Coddington, J. W.; Hurst, J. K.; Lymar, S. V. Hydroxyl radical

formation during peroxynitrous acid decomposition. J. Am. Chem. Soc.

121:2438-2443; 1999.

[10] Gerasimov, O. V.; Lymar, S. V. The yield of hydroxyl radical from the

decomposition of peroxynitrous acid. Inorg. Chem. 38:4317-4321;1999.

[11] Merényi, G.; Lind, J.; Goldstein, S.; Czapski, G. Peroxynitrous acid

homolyzes into OH and N02 radicals. Chem. Res. Toxicol. 11:712-

713; 1998.

[12] Richeson, C. E.; Mulder, P.; Bowry, V. W.; Ingold, K. U. The complex

chemistry of peroxynitrite decomposition: New insights. J. Amer.

Chem. Soc. 120:7211-7219; 1998.

[13] Koppenol, W. H.; Kissner, R. Can ONOOH undergo homolysis. Chem.

Res. Toxicol. 11:87-90; 1998.

[14] Alfassi, Z. B.; Huie, R. E.; Neta, P.; Shoute, L. C. T. Temperature

dependence of the rate constants for reaction of inorganic radicals with

organic reductants. J. Phys. Chem. 94:8800-8805; 1990.

[15] Buettner, G. R. In the absence of catalytic metals ascorbate does not

autoxidize at pH 7: ascorbate as a test for catalytic metals. Journal ofBiochemical and Biophysical Methods 16:27-40; 1988.

[16] Denicola, A.; Freeman, B. A.; Trujillo, M.; Radi, R. Peroxynitritereaction with carbon dioxide/bicarbonate: Kinetics and influence on

peroxynitrite-mediated oxidations. Arch. Biochem. Biophys. 333:49-58;1996.

[17] Meli, R.; Nauser, T.; Latal, P.; Koppenol, W. H. Reaction of

peroxynitrite with carbon dioxide: Intermediates and determination of

the yield of C03' and N02'. J. Biol. Inorg. Chem. 7:31-36; 2002.

[18] Kirsch, M.; de Groot, H. Ascorbate is a potent antioxidant against

peroxynitrite-induced oxidation reactions - Evidence that ascorbate acts

93

by re-reducing substrate radicals produced by peroxynitrite. J. Biol.

Chem. 275:16702-16708; 2000.

[19] Balavoine, G. G. A.; Geletii, Y. V. Peroxynitrite scavenging bydifferent antioxidants. Part I: Convenient assay. Nitric Oxide: Biol.

Chem. 3:40-54; 1999.

[20] Whiteman, M.; Halliwell, B. Protection against peroxynitrite-dependent

tyrosine nitration and al-antiproteinase inactivation by ascorbic acid. A

comparison with other biological antioxidants. Free Radical Res.

25:275-283; 1996.

94

8. MECHANISM OF THE REACTION OF

PEROXYNITRITE WITH L-ASCORBATE STUDIED

AT PHYSIOLOGICAL PH

CHRISTOPHE R. KURZ, REINHARD KISSNER, and WILLEM H.

KOPPENOL*

Laboratorium für Anorganische Chemie, Eidgenössische Technische

Hochschule Zürich (Hönggerberg), Switzerland

Abstract— Existing data on the reaction mechanism of monohydroascorbate

with peroxynitrous acid, together with new data on the effect of pH, suggest a

complex mechanism. The peroxynitrite anion does not react with

monohydroascorbate. For the initial reaction of peroxynitrous acid with

monohydroascorbate, a second-order rate constant of kj of 5 x 105±1 M'V1 and

Oil 1

an equilibrium constant K of 1 x 10 M were determined. We observed three

intermediates of which the third has an absorption maximum at 345 nm and a

p£Ta of 7.6 ± 0.2. Both the protonated and the deprotonated forms of this third

intermediate can react with a second monohydroascorbate to form nitrite,

monohydroascorbate, and dehydroascorbate. In addition, there is a

monohydroascorbate-independent decay path that yields nitrate and

monohydroascorbate. EPR Experiments showed that ascorbyl radicals are not

involved in the reaction of peroxynitrite with monohydroascorbate.

Results

Stopped-flow studies

Kinetic traces recorded at 360 nm, 25 °C and at pH 7.2 after mixing 25 uM

peroxynitrite with different concentrations of monohydroascorbate

(0.25 - 50 mM) show an increase in absorption during 0.1 - 0.5 s, followed by

95

decrease (Fig. 1, all concentrations are those after mixing). At

monohydroascorbate concentrations higher than 1 mM, the rate of this increase

in absorption depends, like the magnitude of the absorption, on the

monohydroascorbate concentration (Fig. 1, traces C and D). The rate of increase

and the absorption intensity also depend on the peroxynitrite concentration, but

the decay rate of absorption is independent (data not shown). The absorption

immediately after mixing is significantly higher than that expected for

peroxynitrite alone at the same pH (Fig. 1, trace E). Monohydroascorbate does

not absorb at wavelengths higher than 320 nm. Both the increase and decay rates

of this absorption increase with temperature (inset Fig. 1).

0 2 4 6 8 10 12

time /s

Fig. 1. Time course for the change in absorption at 360 nm after mixing various

concentrations of monohydroascorbate with peroxynitrous acid. Peroxynitrite(25 uM) was mixed with monohydroascorbate at the following concentrations: trace

A, 250 uM; trace B, 1 mM; trace C, 8 mM, trace D, 50 mM and E, without

monohydroascorbate. All concentrations indicated are those after mixing. All traces

were recorded at 360 nm, 25 °C and pH 7.2 (50 mM phosphate buffer). At lower

monohydroascorbate concentrations (trace A and B) no dependence on the

monohydroascorbate concentration is observed. Each curve is an average of

12 measurements. Inset: 25 uM peroxynitrous acid mixed with 250 uM

monohydroascorbate at pH 7.2 at the following temperature: trace A, 25 °C; trace

a, 37 °C. Measurements were made at 360 nm.

96

From pH 5.8 to 8.5 the rates of formation and decay of the absorption at

360 nm decrease with increasing pH (Figs. 2 and 3). The absorption maxima at

3 ms and between 0.1 - 2.3 s, dependent on pH, increased at higher pH (Fig. 2).

Trace a in the inset of Figure 2 was recorded at 360 nm to diminish the

peroxynitrite interference on the absorption of the intermediate. Trace ß was

recorded at 320 nm to maximize the maximum contribution from the

peroxynitrite anion. Trace y represents the peroxynitrite decay without

monohydroascorbate at the same pH and concentration.

time /s

Fig. 2. Kinetics of the formation and decay of an intermediate as a function of

pH. Kinetics traces observed at 360 nm after mixing 250 uM monohydroascorbatewith 25 uM peroxynitrite at the following pH: trace A, pH 8.5; trace B, pH 8.0,trace C, pH 7.2, trace D, pH 5.8. Inset: Time course of the absorption after mixing25 uM peroxynitrite with 250 uM monohydroascorbate at pH8.5. The trace

a (= trace A) was recorded at 360 nm and ß at 320 nm. Trace y is peroxynitritealone measured at pH 8.5 and 320 nm.

97

Upon mixing 50 mM

monohydroascorbate

with 25 uM of

peroxynitrite in this pH

range, we observed an

immediate decrease in

absorption (Fig. 3). This

process clearly had

started already during

the mixing and can only

be observed at high

monohydroascorbate

concentrations. This

initial absorption drops

10 15 20

time /ms

25 30 35 40

Fig. 3. Kinetics traces recorded a 360 nm after mixing25 uM peroxynitrite with 50 mM monohydroascorbate at

the following pH: trace A, pH5.8; trace B, pH 7.2; trace

C, pH 8.0; trace D, pH 8.5. The fast decrease in absorption

directly after mixing is not observed at

monohydroascorbate concentrations lower than 8 mM. The

dotted lines represent the absorption calculated for

peroxynitrite at the same pH as the measurements.

during the first 5 - 7 ms and rises then again, indicating formation of another

intermediate, as shown in Figures 1 and 2.

The transient

absorption with a

maximum at 345 nm

was recorded 144 ms

after mixing 20 mM

mono-hydroascorbate

with 1 mM

peroxynitrite at pH 5.8

(Fig. 4, spectrum A)

[1]. The absorption at

wavelengths below

345 nm increases with

increasing pH, and the

320 340 360 380 400 420

Fig. 4 Absorption spectra of Im37HIm3 after mixing20 mM monohydroascorbate with 1 mM peroxynitrite at

different pH and time. Spectrum A, pH5.8 and 144 ms;

spectrum B, pH 7.2 and 144 ms; spectrum C, pH 8.0 and

192 ms; spectrum D, pH 8.5 and 432 ms.

98

transient absorption maximum can no longer be determined (Fig. 4, spectra B,

C, andD).

Control experiments without peroxynitrite, but with nitrite or nitrate added,

show no transient absorption at 345 nm, as in experiments in which

peroxynitrite is mixed with buffer only, which confirms that the intermediate is

formed as a product of the reaction of monohydroascorbate with peroxynitrous

acid. Furthermore, we observed no influence of nitrite, nitrate, or dioxygen on

the absorption kinetics (not shown).

EPR studies

We observe a small signal with a g-value of 2.005 which is due to the

ascorbyl radical. This signal is not detected when monohydroascorbate or

peroxynitrite are absent. Immediately after mixing 25 uM peroxynitrite and

250 uM monohydroascorbate (final concentrations) at pH 7.2, no EPR signal is

observed; instead, the signal intensity rises for approximately 3 seconds and

subsequently disappears within 10 seconds (Fig. 5, trace B). A comparison with

the optical kinetic trace recorded at 360 nm and pH 7.2 (Fig. 5, trace A) shows

that the concentration of ascorbyl radical increases while the absorption

decreases. The total concentration of ascorbyl radical formed during the reaction

at pH5.8 is estimated to be only 60 nM [1]. According to the amount of

ascorbyl radical found at pH 5.8 and given the retarded disproportionation of the

ascorbyl radical at higher pH, we estimate the total ascorbyl radical

concentration formed at pH 7.2 to be below 100 nM.

99

8

time /s

10 12 14

Fig. 5 Comparison of the decay at 360 nm (trace A) with an EPR trace (trace

B) of the ascorbyl radical formation and decay. Trace B, an average of

10 measurements, reaches maximum distinctly later than trace A, after an average

of 12 measurements. Measurements were made at initial concentrations of 25 uM

peroxynitrite and 250 uM monohydroascorbate in 50 mM phosphate buffer at

pH 7.2. All concentrations indicated are those after mixing.

Ion-chromatography

When 25 uM peroxynitrite is mixed with monohydroascorbate at pH 8.1, the

nitrite yield increases with increasing monohydroascorbate concentration and a

corresponding decrease of nitrate. Even at low monohydroascorbate

concentrations, the nitrite yield is already much higher than without

monohydroascorbate (not shown).

100

Discussion

Previous work

From our study of peroxynitrous acid with monohydroascorbate at pH 5.8 we

concluded that; (1) no ascorbyl radical is produced directly, (2) there exist at

least 2 intermediates of which one has an absorption maximum at 345 nm, and

(3) the second-order rate constant of the reaction of peroxynitrous acid with

monohydroascorbate is about 106M1s1 [1]. In this work, we studied the

reaction of peroxynitrite with monohydroascorbate between pH5.8 and 8.5 to

determine the mechanism and the rate constant for the reaction of peroxynitrite

with monohydroascorbate at pH 7.2.

Intermediates

When 25 uM peroxynitrite is mixed with 50 mM of monohydroascorbate, a

fast decrease in absorption during the first 5 ms is observed (Fig. 3). This initial

absorption is much higher than that of peroxynitrite alone and was observed at

all pH values at monohydroascorbate concentrations higher than 8 mM. The

initial high absorption, indicating production of intermediate Im1; decreases

during the first 5 - 7 ms to a value that is still higher than that formed in the

presence of peroxynitrite alone. The decrease in absorption is due to an

intermediate we label Im2. The following increase in absorption with an

absorption maximum at pH 5.8 of 345 nm is assigned to a third intermediate

HIm3 (Figs. 1 and 3). The intermediate Im2 is invoked to explain the absorption

minimum observed after 7-10 ms: otherwise, the absorption would be expected

to increase, decrease, or remain constant, but not show a minimum.

As shown in Figure 2, the difference between the absorption seen at mixing

time and the maximal absorbance at 360 nm increases with pH. When we

assume that only peroxynitrous acid reacts with monohydroascorbate (see

section 'Peroxynitrite anion'), we would expect that less HIm3 is formed at

higher pH (Fig. 2), but the absorption-pH profile contradicts this. To account for

101

the increase in absorption at higher pH, we introduce another species resulting

from deprotonation of HIm3 to form the more strongly absorbing Im3 .We

estimate a p^a of 7.6 ± 0.2 for the HIm3/Im3 acid-base pair by simulation.

Peroxynitrite anion

The decay of the absorption ascribed to HIm3/Im3 proceeds more slowly at

higher pH. We observed that the rate of formation of HIm3/Im3 also decreases

with increasing pH. The spectrum of the third intermediate changes as function

of pH (Fig. 4), with a clear absorption maximum of 345 nm at pH 5.8. At higher

pH, the absorption increases in the same manner as that of peroxynitrite, but

with a maximum below 320 nm. Unfortunately, monohydroascorbate begins to

absorb below 320 nm, and no information could be collected below that

wavelength. The peroxynitrite anion exhibits a maximum at 302 nm with an

absorption coefficient of 1360 M^cm"1 at 320 nm. Since, the signal intensity

after mixing is proportional to the absorption of peroxynitrous acid at each pH

(Fig. 3 and 4), we conclude that at 320 nm and, at t = 0 s, mainly the

peroxynitrite anion absorbs. Of interest is that the decrease in absorption at

320 nm is faster than the decay of peroxynitrite or that of the intermediate

measured at 360 nm (inset Fig. 2).

The observed increase in absorption at 320 nm as function of the pH

immediately after mixing suggests that there is still peroxynitrite present, even

after a few seconds. Furthermore, the relative absorption of HIm3/Im3 decreases

with increasing pH (Fig. 4), implying that only peroxynitrous acid reacts with

monohydroascorbate.

When peroxynitrite is mixed with excess monohydroascorbate at pH 13, the

yellow color of peroxynitrite remains visible for hours.

102

Homolysis or one-electron oxidation?

As we reported earlier, the ascorbyl radical is not a significant intermediate

in the reaction of peroxynitrous acid with monohydroascorbate at pH5.8 [1].

Extension of our earlier EPR experiments to pH range 5.8-8.5 showed that the

ascorbyl radical concentrations remain at about 60 nM and represent a yield of

ca. 0.24 % relative to the initial peroxynitrous acid concentration; this small

amount of ascorbyl radicals forms concomitantly with the decrease of the

absorption at 360 nm. Thus, since the ascorbyl radical is not a primary reaction

product, no one-electron oxidation has taken place, excluding homolysis of

peroxynitrous acid to nitrogen dioxide and hydroxyl radical under the conditions

of these experiments. As mentioned earlier [1], it is possible that the ascorbyl

radicals observed originate from the oxidation of monohydroascorbate by

dioxygen, a reaction that can be catalyzed by trace amounts of contaminating

transition metals[2,3].

Kinetics and mechanism

We postulate from the increased absorption immediately after mixing that

Im! is formed very rapidly from monohydroascorbate and peroxynitrous acid,

indicating that the rate constant kx for the reaction of monohydroascorbate with

peroxynitrous acid to form Imi within 3 ms is high. Imi decays within 10 ms to

form Im2, at which time there is still free peroxynitrite present in the solution, as

indicated by the absorption spectra (Fig 4). The presence of free peroxynitrite

implies that Im2 reverts via Imi to monohydroascorbate and peroxynitrous acid.

Im2 also decays further to HIm3, which is in equilibrium with its deprotonated

form Im3 .The decay rate of this third intermediate strongly depends on the

monohydroascorbate concentration. We conclude, as before [1] that HIm3 can

react with another monohydroascorbate molecule to form nitrite,

monohydroascorbate, dehydroascorbate and water, as can Im3 .Ion-

chromatography of the products obtained from even a small concentration of

103

monohydroascorbate (250 uM) at pH 8.1 shows the presence of nitrite at a

concentration higher than that expected from decomposition [4]; furthermore,

the nitrite yield increases as a function of monohydroascorbate concentration

(data not shown).

ONOO + H+ H++ Im3+

H^sc» HAsc + DAsc + N02 + OH

kl k2 k3^

yc t_ _

,

ONOOH + HAsc ^=?r imi +=*: Im2 ^^ HIm33 4

» HAsc +N03 +H

k.i k-2 k_3k.

9 k\+ HAsc

N03 + H' ^

HAsc + DAsc + N02 +H20

Scheme 1

As shown in the Figure 1, the decay of HIm3 at 250 uM and 1 mM HAsc is

similar at pH7.2, indicating that HIm3 decays via a monohydroascorbate-

independent pathway. Possibly, HIm3 decay results in the formation of

monohydroascorbate and nitrate. A second possibility is that there is a slow back

reaction from HIm3 to Im2. The rates of decay of Im3 (HIm3 and Im3 ) and

monohydroascorbate, observed at 360 nm and 265 nm, respectively, are similar

to those published earlier for reaction at pH5.8, although the

monohydroascorbate decay precedes that of Im3 [1]. Unfortunately, this

simultaneous decay of monohydroascorbate and HIm3 prevents discrimination

of these two pathways, and we are not able to determine whether HIm3

undergoes decay to monohydroascorbate and nitrate or is in equilibrium with

peroxynitrous acid and monohydroascorbate, through Im2 and Imi.

104

Rate_, L , _, , „ , Absorption Initial

A ARate value Remark Compound r~

.

A A A.

constant coetncient concentration

*1 5 x 105±1 M'V a HAsc 0 M'W1250 u.M-

50 mM

£-1 5 x 102±1 s"1 a HOONO 82 M'W1 25 uM

^2 3 ± 1 x 102 s"1 a ONOO" 685 M'W1 —

k-2 3 ± 1 x 101 s"1 a N03 0 M'W1 —

h 2 ± 1 x 102 s"1 a N02 5 M'W1

k-3 3±2x loV1 a DAsc 0 M'W1 —

k4 0.6 ± 0.3 s"1 a, b Imi -400 M'W1 —

h1.2 ±0.3 x 102

M-V a, b Im2 0 M'W1 —

k6 o.uo.imV a HIm3 460 M'W1

h 2.5 x I02s_1 a Im3 1600 M'W1 —

k-7, h lxio^V1 c

k.* 1.6 x I03s_1 d

kg 1.2 s"1 e

Table 1. Values used for the simulation of the reaction of peroxynitrous acid

with monohydroascorbate at 360 nm and 25°C. a = determined by simulation;b = Kurz et al. [1]; c = rate constant for protonation; d = rate constant determined by

pKa [5]; e = Kissner et al. [4].

In contrast to the results obtained at pH 7.2 and lower, the decay traces for

the third intermediate at pH 8.5, when 25 uM peroxynitrite is mixed with

250 uM and 1 mM monohydroascorbate, differ from one another (data not

shown). The monohydroascorbate-dependent and -independent decay rates are

not influenced by pH. We therefore invoke a monohydroascorbate dependent

path that becomes more important at higher pH. We propose that Im3 can react

105

with another monohydroascorbate molecule and form nitrite,

monohydroascorbate, dehydroascorbate and hydroxide.

We simulated the two mechanisms shown in Scheme 1, i.e. with and without

reaction k.3 for the pH range 5.8-8.5 and a monohydroascorbate concentration

range of 250 uM to 50 mM. The simulations suggest a second-order rate

constant kx of 5 x 105±1 M'V1 with a first-order rate constant k_x of 5 x 102±1 s"1

for the back reaction. Further, a pKa value of 7.6 ±0.2 of HIm3 could be

estimated.

Physiological relevance

Can the reaction rate of monohydroascorbate compete for peroxynitrite with

carbon dioxide or other compounds at physiological pH and temperature? The

reaction with carbon dioxide, which is present at around 1 mM in human

plasma, proceeds with a rate constant of 3 x 104M~1s~1 at 25 °C [6] and a

backward rate constant of 1.5 x 106 s_1 [7]. Given that, in both cases, multiple

equilibria are involved, it is not possible to answer this question. However, in

the presence of both carbon dioxide (1 mM) and monohydroascorbate(250 uM),

the ascorbyl radical is observed at pH 8 and higher, but not at lower pH (Kurz,

2004, unpublished). We propose, therefore, that under physiological conditions

(0.1 - 0.2 uM monohydroascorbate) monohydroascorbate competes with carbon

dioxide (1 mM) for peroxynitrite. Monohydroascorbate has been reported to

protect against damage caused by peroxynitrite which has been mostly ascribed

to monohydroascorbate's ability to repair[5,8-10]; here we showed that

monohydroascorbate may prevent damage by acting as scavenger of

peroxynitrous acid.

Acknowledgment - This project was supported by the ETH and the

Schweizerische Nationalfonds.

We thank Dr. P.L. Bounds for helpful discussions.

106

Abbreviation

Imi : First intermediate of the reaction of peroxynitrous acid with

monohydroascorbate. (New intermediate compared with chapter 7.)

Im2: First intermediate of the reaction of peroxynitrous acid with

monohydroascorbate. (Is equal to Imi in chapter 7)

Im3: Third and last intermediate of the reaction of peroxynitrous acid

with monohydroascorbate. (Is equal to Im2 in chapter 7)

Peroxynitrite: Means the protonated and deprotonated form.

References

[1] Kurz, C.; Kissner, R.; Perrin, D.; Nauser, T.; Koppenol, W. H. Rapid

scavenging of peroxynitrous acid by monohydroascorbate. Free Radical

Biol. Med. 35:1529-1537; 2003.

[2] Buettner, G. R. In the absence of catalytic metals ascorbate does not

autoxidize at pH 7: ascorbate as a test for catalytic metals. Journal ofBiochemical and Biophysical Methods 16:27-40; 1988.

[3] Buettner, G. R.; Jurkiewicz, B. A. Catalytic metals, ascorbate and free

radicals: Combinations to avoid. Radiât. Res. 145:532-541; 1996.

[4] Kissner, R.; Koppenol, W. H. Product distribution of peroxynitrite decayas a function of pH, temperature, and concentration. J. Am. Chem. Soc.

124:234-239; 2002.

[5] Koppenol, W. H.; Kissner, R. Can ONOOH undergo homolysis. Chem.

Res. Toxicol. 11:87-90; 1998.

[6] Lymar, S. V.; Hurst, J. K. Rapid reaction between peroxonitrite ion and

carbon dioxide: Implications for biological acitivity. J. Am. Chem. Soc.

117:8867-8868; 1995.

[7] Squadrito, G. L.; Pryor, W. A. Mapping the reaction of peroxynitrite with

C02: Energetics, reactive species, and biological implications. Chem. Res.

Toxicol. 15:885-895; 2002.

[8] Kirsch, M.; de Groot, H. Ascorbate is a potent antioxidant against

peroxynitrite-induced oxidation reactions - Evidence that ascorbate acts

107

by re-reducing substrate radicals produced by peroxynitrite. J. Biol.

Chem. 275:16702-16708; 2000.

[9] Whiteman, M.; Halliwell, B. Protection against peroxynitrite-dependent

tyrosine nitration and al-antiproteinase inactivation by ascorbic acid. A

comparison with other biological antioxidants. Free Radical Res. 25:275-

283; 1996.

[10] Balavoine, G. G. A.; Geletii, Y. V. Peroxynitrite scavenging by different

antioxidants. Part I: Convenient assay. Nitric Oxide: Biol. Chem. 3:40-54;1999.

108

9. THE REACTION OF PEROXYNITRITE WITH

CARBON DIOXIDE AND VITAMIN C.

CHRISTOPHE R. KURZ, REINHARD KISSNER, and WILLEM H.

KOPPENOL*

Laboratorium für Anorganische Chemie, Eidgenössische Technische

Hochschule Zürich (Hönggerberg), Switzerland

Abstract— In the presence of carbon dioxide, peroxynitrite reacts catalytically

with monohydroascorbate at a very fast rate. Low monohydroascorbate

concentrations prevent the formation of trioxocarboanate(#-l) radical and

nitrogen dioxide, the intermediates produced from the reaction of peroxynitrite

with carbon dioxide. Moreover, at low monohydroascorbate concentrations only

small amounts of ascorbyl radicals, relative to the initial monohydroascorbate

concentration, are produced. If the monohydroascorbate concentration is higher

than that of peroxynitrite, the ascorbyl radical concentration increases. Under

these conditions, an alternative pathway, which involves reaction with a second

monohydroascorbate molecule, is observed. Stopped-flow experiments indicate

that at pH 7, the reaction between peroxynitrous acid and monohydroascorbate

competes with the reaction between peroxynitrite anion, monohydroascorbate,

and carbon dioxide. At pH 6, the main pathway involves the reaction of

monohydroascorbate with peroxynitrous acid.

Results

Kinetic traces recorded at pH 7 after mixing 2 mM peroxynitrite with

17 mM carbon dioxide showed a fast decrease in absorption at 640 nm. (Fig. 1,

all concentrations are given after mixing). When 2 mM peroxynitrite was mixed

109

with 17 mM carbon dioxide and 50 uM, 2 mM or 10 mM monohydroascorbate,

a weaker, or no absorption was observed.

Kinetic traces s

recorded at 320 nm at5

4

pH 7 after mixing s*

3a

2 mM peroxynitrite f2

Xi

with 17 mM carbon"

i

dioxide show faster°

-i

decay in the presence of

50 uM, 2 mM, or

10 mM monohydro¬

ascorbate than in the

absence of monohydro¬

ascorbate. Interestingly,

25 50 75

time /ms

100 125 150

Fig 1 Time courses of changes in absorption at

640 nm after mixing peroxynitrite (2 mM) with carbon

dioxide (17 mM) and the following concentrations of

monohydroascorbate: trace A, 0 uM; trace B, 50 uM;

trace C, 2 mM and trace D, 10 mM. All concentrations

are given after mixing. All traces were recorded at 25°C

and pH 7.0 in 50 mM phosphate buffer.

10 20 30

time /ms

40 50 60

Fig. 2 Kinetic traces recorded at 320 nm after mixing 2 mM peroxynitrite with

17 mM carbon dioxide and monohydroascorbate at different concentrations: trace A,0 uM; trace B (x), 50 uM; trace C, 2 mM and trace D (-), 10 mM. All traces were

recorded at 25°C and pH 7.0 in 50 mM phosphate buffer.

110

if peroxynitrite was in excess over monohydroascorbate, the absorption

decreased faster compared to the trace recorded in the absence of

monohydroascorbate. Further, in the presence of monohydroascorbate the decay

rate observed at 320 nm did not change with increasing monohydro-ascorbate

concentration in the range from 50 uM to 10 mM (Fig. 2). The reaction of

peroxynitrite with monohydro-ascorbate, in the absence of carbon dioxide, is too

slow to be responsible for this acceleration in the absorption decay (Fig 3). At

pH 6 the rate of the absorption decay increased with increasing

monohydroascorbate concentration, which is significantly different from results

at pH 7 (cf. Fig 2 and 4). At pH 8, the absorption decay became faster in the

presence of 50 uM of monohydroascorbate and then slowed down with further

increase in monohydroascorbate concentration (Fig. 5). The decrease in

absorption at 320 nm changed also from first-order to second-order kinetics. The

initial absorption was higher with increasing monohydroascorbate

concentrations and the absorption starts to decrease after about 2 ms (Fig. 5).

The rate of the absorption decay, recorded at 320 nm, was maximal at pH 7 and

decreased with changes in pH (cf. Fig. 2, 4 and 5). The absorption spectra at

pH 6, 7 and 8, recorded after mixing 50 uM peroxynitrite with 1 mM carbon

dioxide and 10 mM monohydro-ascorbate have different absorption maxima

(Fig. 6). With decreasing pH the absorption maximum shifts. The spectrum

recorded at pH 8 has a maximum at 360 nm and at pH 6, at 345 nm. The

maximum also decreases with decreasing pH. The high absorption observed at

pH 8 at wavelengths below 340 nm is due to peroxynitrite anion, which absorbs

at 320 nm 15 times more than peroxynitrous acid.

Ill

5

time /s

10

Fig. 3 Kinetic traces recorded at 320 nm after mixing 2 mM peroxynitrite with

different monohydroascorbate concentrations: trace A, 0 uM; trace B, 50 uM; trace C,2 mM and trace D, 10 mM. All traces were recorded at 25°C and pH 7.0 in 50 mM

phosphate buffer.

At pH 8 the kinetic trace recorded at 360 nm after mixing 50 uM

peroxynitrite with 1 mM carbon dioxide disappeard within 0.1 s (Fig. 7). When,

in addition, 100 uM or 1 mM monohydroascorbate is present, the absorption

reaches 0.05and 0.08 in 20 and 30 ms, respectively; this increase is followed by

a second-order decay.

EPR measurements showed an ascorbyl radical spectrum after mixing 2 mM

peroxynitrite with 17 mM carbon dioxide and 10 mM monohydro-ascorbate at

pH8.

Product analysis at pH 8 and low monohydro-ascorbate concentrations

showed that nitrate is nearly quantitatively formed from peroxynitrite.

112

0 50 100 150 200 250 300

time /ms

Fig. 4 Time course of the change in absorption at 320 nm at pH 6 after mixing various

concentrations of monohydro-ascorbate with peroxynitrite and carbon dioxide.

Peroxynitrite (2 mM) was mixed with carbon dioxide (17 mM) and the followingconcentrations of monohydroascorbate: trace A, 0 uM; trace B, 50 uM; trace C, 2 mM

and trace D, 10 mM.

Discussion

l-Carboxylato-2-nitrosodioxidane was proposed to be an intermediate in the

reaction of peroxynitrite with carbon dioxide. This intermediate, the absorption

maximum of which was observed at 600 nm by Papina (unpublished; most

probably trioxocarbonate(#l-) radical), is formed with a rate constant of

3x10^^_1[1].

ONOO" + CQ2<» ONOOCO2 Products

^

As is shown in Figure 1, even equimolar concentrations of vitamin C and

peroxynitrite, in the presence of carbon dioxide, prevent the appearance of the

absorption at 640 nm. This means that either the formation of 1-caboxylato-

2-nitrosodioxidane is completely inhibited or that this intermediate is very

rapidly consumed by vitamin C.

113

2.0

1.8

1.6

1.4 H

J 1.2 H

a.

*0.8

0.6 -

0.4 -

0.2 -

0.0

0 10 15 20 25

time /ms

30 35 40 45 50

Fig.5 Time course for the change in absorption at 320 nm at pH 8 after mixingvarious concentrations of monohydroascorbate with peroxynitrite and carbon dioxide.

Peroxynitrite (2 mM) was mixed with carbon dioxide (17 mM) and the followingconcentrations of monohydroascorbate: trace A, 0 uM; trace B, 50 uM; trace C,2 mM and trace D, 10 mM.

When the reaction between peroxynitrite with carbon dioxide is observed at

320 nm, where mainly peroxynitrite absorbs (Fig. 2), the decrease of the signal

is faster in the presence of vitamin C. As reported in chapter 7, the reaction of

peroxynitrous acid with vitamin C could be fast enough to prevent the reaction

of peroxynitrite with carbon dioxide [2]. Unfortunately, there exist equilibria

between the substrate, peroxynitrous acid, monohydroascorbate, and a third

intermediate (HIm3), which also absorbs at 320 nm and therefore prevent a

straightforward conclusion.

ONOOH + Hasc" Him, - Products (2)

The kinetic traces obtained for the reaction of peroxynitrous acid with

monohydroascorbate decrease much slower in the absence of carbon dioxide and

therefore reaction 2 can not be responsible for the consumption of peroxynitrite

114

within few milliseconds. (Fig. 3) Two possibilities for this fast consumption of

peroxynitrite are proposed:

- peroxynitrous acid reacts with vitamin C whereby the first intermediate Im!

formed within a few milliseconds reacts with carbon dioxide in an

irreversible step (eq. 3),

- l-caboxylato-2-nitrosodioxidane is consumed by monohydroascorbate in a

fast irreversible reaction (eq. 4).

ONOOH + Hasc" «»

Imi+C°2

» Products (3)

ONOO" + CQ2<»

ONOOCO2+ HAsc

» Products (4)

To verify these two possibilities, rates of reaction of peroxynitrite with

carbon dioxide and monohydroascorbate were measured at pH 6. If reaction 3

would occur, the absorption at 320 nm should decrease faster at lower pH. As

shown in Figure 4, at pH 6 the absorption traces disappeared slower compared to

traces recorded at pH 7, which leads us to the conclusion that reaction 3 does not

take place.

Kinetic traces recorded at 320 nm at pH 7 show that the initial absorption is

diminished when monohydroascorbate is present, which indicates that the

reaction has already progressed within the mixing time (Fig. 2). It is of special

interest that, already at very low monohydroascorbate concentrations, 20 times

smaller than the one of peroxynitrite, peroxynitrite is consumed much faster and

the initial absorption decay does not increase further with increasing

monohydroascorbate concentration. In the absence of monohydroascorbate the

decay is much slower. These observations indicate that monohydroascorbate

does not partitcipate in the rate limiting step of this reaction. Further, from

similar kinetic experiments when either peroxynitrite or monohydroascorbate is

in excess, it can be concluded that monohydroascorbate has to react catalytically

with a fast turn over reaction rate. Therefore the following equation can be

proposed:

115

ONOO" + CQ2<ia »ONOOCO2

+ HAsc» HAsc" + Products (5)

k-a kß

Taking into account that the absorption decay at 320 nm at pH 7 is not

accelerated further with increasing monohydroascorbate concentrations and that

at monohydroascorbate concentrations20 times smaller than the one of

peroxynitrite the absorption decay reaches its rate limit, one can conclude that kp

has to be higher than ka.

The maximum of absorption spectra recorded at different pH shifts from

360 nm to 345 nm with decreasing pH (Fig. 6). The absorption spectrum with a

maximum at 360 nm, recorded at pH 8, is similar to the absorption spectrum of

ascorbyl radical (s36o = 5000 M4cm_1 [3]). EPR measurements confirmed the

presence of ascorbyl radicals in the reaction of peroxynitrite with carbon dioxide

and monohydroascorbate (data not shown).

The absorption spectrum with a maximum at 345 nm, recorded at pH 6, is

similar to the spectrum of the intermediate HIm3 recorded after mixing

peroxynitrite with monohydroascorbate in the absence of carbon dioxide (eq. 2)

HIm3 has an extinction coefficient of about 460 M"1cm"1, about 11 times

smaller than the ascorbyl radical[3]. Therefore it has to be present in higher

concentrations than the ascorbyl radical to be registered. At pH 6, a similar

absorption behaviour, independent from carbon dioxide, suggests that mainly

the reaction between peroxynitrous acid and monohydroascorbate is responsible

for the peroxynitrite decay. The high extinction coefficient of ascorbyl radicals

allow to estimate roughly the amount of ascorbyl radicals produced in the

reactions at pH 7 and pH 8. One must realise that both peroxynitrite and HIm3

absorb at 360 nm, but less than the ascorbyl radical. The ascorbyl radicals

concentrations found after mixing 50 uM peroxynitrite with 1 mM carbon

dioxide and 10 mM monohydroascorbate were about 17 uM at pH 8 and 9 uM

at pH 7. As mentioned before, at pH 6 mainly HIm3 is produced and therefore

the reaction with carbon dioxide is only of minor importance. Ascorbyl radicals

116

are observed only when the monohydroascorbate concentration is higher than

the concentration of peroxynitrite.

320 330 340 350 360 370 380 390 400

nm

Fig.6 Absorption spectra of an intermediate after mixing 50 uM peroxynitrite with

1 mM carbon dioxide and 10 mM monohydroascorbate at different pH and times. The

spectrum at pH 8 was recorded 205 ms after mixing; the spectrum at pH 7 at 25 ms

and the spectrum at pH 6 at 25 ms. Additionally, the spectrum of ascorbyl radical

(Asc*-) (11 uM), obtained by pulse radiolysis, and the spectrum of HIm3 (48 uM),obtained at pH 7, by mixing peroxynitrite with monohydroascorbate in the absence of

carbon dioxide are shown.

The ascorbyl radical yield depends, as shown in Figure 7, on the

monohydroascorbate concentration and increases with increasing pH. The pH

dependent increase in the ascorbyl radical concentration correlates with the

amount of protonated and deprotonated peroxynitrite. At higher pH, the

concentration of peroxynitrite anion is higher and therefore more peroxynitrite

can react with carbon dioxide. Many of the reagents and intermediates absorb at

320 nm, among them the ascorbyl radical, which may explain the initial peak

after 2 ms, when 2 mM peroxynitrite is mixed with 17 mM carbon dioxide and

10 mM monohydroascorbate at pH 7 and 8 (Fig. 2 and 5).

One obvious question remains: How are ascorbyl radicals produced? The

production of nitrogen dioxide and trioxocarbonate(#-l) radicals in the reaction

of peroxynitrite with carbon dioxide is well known [4-6]. Rate constants of the

117

reactions between those radicals and monohydroascorbate are sufficiently high

to produce ascorbyl radicals [7,8], but the formation of those radicals is inhibited

because monohydroascorbate scavenges l-carboxylato-2-nitrosodioxidane in a

fast catalytic reaction (Scheme 1).

2 Asc"

+ N02 + C02

+ HAsc

HAsc" + N03 + C02 - X

CCKk «

HT

H2C03hco;

OH"

H+ + ONOO" + C02 =5=^=

+ HAsc

ki

NO, + H+« k'nONOOH + HAsc"

ONOOCO,

-11i t'

Kn

ONOOC02H

h:k24 k_15

2 Asc "+ H

Asc'" +N02 + H+

+ HAsc"

DAsc +HAsc

Asc'"+CO?" +H+

-*-N02 +CO3

+ HAsc

(4% - 30%)

:$: N03" + C02

NO," +2HImj

H++ Im,+ HAsc» HAsc + DAsc + N07 + OH

17—

345 kzo» HAsc +N0, +H+Im2 -^

». HIiri

+ HAsc

HAsc +DAsc + N02 +H20

Scheme 1

Taking these results into account, it can be proposed that the catalytic

reaction between l-carboxylato-2-nitrosodioxidane and monohydroascorbate

proceeds via an intermediate X, which can react with another

monohydroascorbate molecule to form finally 2 ascorbyl radicals, carbon

dioxide and nitrite.

ONOO + CO2-*: ONOOCOo+ HAsc

» X

+ HAsc

HAsc + N03 + C02

2 Asc' +N02 + C02(6)

118

A second possibility of ascorbyl radical formation comes from the catalytic

cycle of ascorbyl radicals, as proposed in Scheme 2. At low

monohydroascorbate concentration no ascorbyl radicals would be observed

because of their fast consumption by peroxynitrite.

Different products are proposed for reactions given by Scheme 2 and

equation 6.

ONOO" + C02 «» ONOOC02"

HAsc» Asc'" + 00^"+ H+ + N02

HAsc" + DAsc

ONOO" + H20 ^H NO9

Asc'

HAsc" + OH" HAsc"

NO +02

H90n2o4 r2

N02 + N03 + 2 H

Scheme 2

For low monohydroascorbate concentrations nitrate is the main product

generated in the reaction given by equation 6, while nitrite is mainly formed in

the reaction presented by Scheme 2. Analysis showed nitrate (quantitatively to

peroxynitrite) to be the major product at pH 8, contradicting the reaction in

Scheme 2.

119

0.0 0.1 0.2 0.3 0.4 0.5

time /sFi

g 7. Comparison of the kinetic traces after mixing 50 uM peroxynitrite with 1 mM

carbon dioxide and different monohydroascorbate concentrations. Trace A, 0 uM; trace

B, 100 uM monohydroascorbate and trace C, 1 mM monohydroascorbate. All traces

were recorded at 360 nm, 25°C and pH 8.0 in 50 mM phosphate buffer.

From the confirmation of reaction 6 one can conclude that the catalytic rate

constant of l-carboxylato-2-nitrosodioxidane with monohydroascorbate kp has

to be much higher than ka. The rate constant ky has to be lower than kp, so the

intermediate has time to react with another monohydroascorbate molecule, but

both rate constants need to be high enough that monohydroascorbate can react

catalytically.

Physiological relevance

The fast reaction of peroxynitrite with carbon dioxide is assumed to be the

main reaction path of peroxynitrite in vivo. Given the fast catalytic effect of

monohydroascorbate, on the reaction of peroxynitrite with carbon dioxide, one

would expect that l-carboxylato-2-nitrosodioxidane should react mainly with

monohydroascorbate also in vivo. Even at very small concentrations of

monohydroascorbate (50 uM) all peroxynitrite is consumed. In vivo,

120

peroxynitrite will be present at smaller concentrations than

monohydroascorbate, which yields ascorbyl radicals as a product. Earlier

investigations with monohydroascorbate and peroxynitrite, without carbon

dioxide, also show a high protection ability of monohydroascorbate against

peroxynitrite induced damage [9-13]. Therefore, it can be concluded that

monohydroascorbate can very efficiently prevent peroxynitrite induced

oxidations and nitrations in vivo.

Acknowledgement - This project was supported by the ETH and the Swiss

Nationalfonds.

References

[1] Lymar, S. V.; Hurst, J. K. Rapid reaction between peroxonitrite ion and

carbon dioxide: Implications for biological acitivity. J. Am. Chem. Soc.

117:8867-8868; 1995.

[2] Kurz, C.; Kissner, R.; Koppenol, W. H. Mechanism of the reaction of

peroxynitrite with L-ascorbate studied at physiological pH. Free Radical

Biol. Med. in progress: 2004.

[3] Bielski, B. H. J.; Comstock, D. A.; Bowen, R. A. Ascorbic acid free

radicals. I. Pulse radiolysis study of optical absorption and kinetic

properties. J. Am. Chem. Soc. 93:5624-5629; 1971.

[4] Bonini, M. G.; Radi, R.; Ferrer-Sueta, G.; Ferreira, A. M. D.; Augusto, O.

Direct EPR detection of the carbonate radical anion produced from

peroxynitrite and carbon dioxide. J. Biol. Chem. 274:10802-10806; 1999.

[5] Goldstein, S.; Czapski, G.; Lind, J.; Merényi, G. Mechanism of

decomposition of peroxynitric ion (02NOO~): Evidence for the formation

of 02" and N02 radicals. Inorg. Chem. 37:3943-3947; 1998.

[6] Meli, R.; Nauser, T.; Latal, P.; Koppenol, W. H. Reaction of peroxynitritewith carbon dioxide: Intermediates and determination of the yield of

C03' andN02'. J Biol. Inorg. Chem. 7:31-36; 2002.

121

[7] Alfassi, Z. B.; Huie, R. E.; Neta, P.; Shoute, L. C. T. Temperature

dependence of the rate constants for reaction of inorganic radicals with

organic reductants. J. Phys. Chem. 94:8800-8805; 1990.

[8] Huie, R. E.; Shoute, L. C. T.; Neta, P. Temperature dependence of the rate

constants for reactions of the carbonate radical with organic and inorganicreductants. Int. J. Chem. Kinet. 23:541-552; 1991.

[9] Kirsch, M.; de Groot, H. Ascorbate is a potent antioxidant against

peroxynitrite-induced oxidation reactions - Evidence that ascorbate acts

by re-reducing substrate radicals produced by peroxynitrite. J. Biol.

Chem. 275:16702-16708; 2000.

[10] Whiteman, M.; Halliwell, B. Protection against peroxynitrite-dependent

tyrosine nitration and al-antiproteinase inactivation by ascorbic acid. A

comparison with other biological antioxidants. Free Radical Res. 25:275-

283; 1996.

[11] Balavoine, G. G. A.; Geletii, Y. V. Peroxynitrite scavenging by different

antioxidants. Part I: Convenient assay. Nitric Oxide: Biol. Chem. 3:40-54;1999.

[12] Lee, J. B.; Hunt, J. A.; Groves, J. T. Manganese porphyrins as redox-

coupled peroxynitrite reductases. J. Am. Chem. Soc. 120:6053-6061;1998.

[13] Lee, J. B.; Hunt, J. A.; Groves, J. T. Rapid decomposition of peroxynitrite

by manganese porphyrin-antioxidant redox couples. Bioorg. Med. Chem.

Lett. 7:2913-2918; 1997.

122

10. CURRICULUM VITAE

Name: Christophe Robert Kurz

Geburtsdatum: 26. Mai 1972

Zivilstand verheiratet

Bürgerort Vechigen (Be)

Ausbildungen

2001 Ausbildung in Menschen und Mitarbeiterführung bei Prof.

R. Steiger, ETH Zürich

2000--2004 Dissertation in der Arbeitsgruppe von Prof. Dr. W. H.

Koppenol am Institut für Anorganische Chemie der ETH

Zürich mit Schwergewicht Peroxynitrit.

1998 Ausbildung zum Strahlenschutz-Sachverständigen am

Paul Scherrer Institut

1993--1999 Chemiestudium an der ETH Zürich mit Spezialisierung

auf Anorganischer Chemie und Analytik.

1988--1992 Matur Typus C am Mathematisch

Naturwissenschaftlichen Gymnasium der Kantonsschule

Zürich

Zusätzliches

2002 Vorlesungen in Vortrags und Diskussionstechnik bei Prof.

R. Steiger, ETH Zürich

2001 Vorlesung in Discovery Management im Zentrum für

Betriebswissenschaften, ETH Zürich

2000-2003 Assistent im Praktikum Allgemeine Chemie I/II an der

ETH Zürich

123

124