Download The Handbook of Redox Biochemistry- ESA, Inc.

Survey
yes no Was this document useful for you?
   Thank you for your participation!

* Your assessment is very important for improving the workof artificial intelligence, which forms the content of this project

Document related concepts

Two-hybrid screening wikipedia , lookup

Artificial gene synthesis wikipedia , lookup

Enzyme wikipedia , lookup

Citric acid cycle wikipedia , lookup

Oxidative phosphorylation wikipedia , lookup

Butyric acid wikipedia , lookup

Proteolysis wikipedia , lookup

Point mutation wikipedia , lookup

Nitric oxide wikipedia , lookup

Metabolism wikipedia , lookup

15-Hydroxyeicosatetraenoic acid wikipedia , lookup

Deoxyribozyme wikipedia , lookup

Specialized pro-resolving mediators wikipedia , lookup

Biochemistry wikipedia , lookup

Biosynthesis wikipedia , lookup

Nucleic acid analogue wikipedia , lookup

Evolution of metal ions in biological systems wikipedia , lookup

Gaseous signaling molecules wikipedia , lookup

Hepoxilin wikipedia , lookup

Metalloprotein wikipedia , lookup

Radical (chemistry) wikipedia , lookup

Transcript
THE HANDBOOK OF REDOX
BIOCHEMISTRY
Ian N. Acworth, D.Phil., Oxon.
This book is dedicated to Emma Louise and Kimberly Ann.
I would like to thank the following people. For their help with editing – Debbie
Aldrich, Kim Acworth, John Waraska, Bruce Kristal and Paul Gamache. For their
support – Scott Freeto, Bruce Bailey, Wayne Matson, Walter DiGiusto, M.
Bogdanov, and many other members of ESA, Inc. For his advice, willingness to
help, and for the Sunrise Free Radical Schools – Prof. Garry Buettner. For
continued help – Dr. Ken Hensley, Dr. K. Williamson, and Prof. R. Floyd
(Oklahoma Medical Research Foundation).
I would like to acknowledge all researchers in this field without whose work this
handbook could not have been completed.
ESA, Inc.
22 Alpha Road
Chelmsford, MA 01824-4171 USA
(978) 250-7000
Sales (800) 959-5095
Fax (978) 250-7090
www.esainc.com
An ISO 9001 Company
ESA Analytical, Ltd.
Brook Farm, Dorton
Aylesbury, Buckinghamshire
HP18 9NH
England, UK
01844 239381
Fax 01844 239382
ii
Preface
It has been several years since I co-wrote “The Handbook of Oxidative
Metabolism” with my colleague Dr. Bruce Bailey, showing the use of
electrochemical approaches in the study of free radical production,
macromolecular damage and antioxidant protection. Since 1995, thousands of
copies of the Handbook have been requested. It has been translated into other
languages. It has even been used as a basic course work at university. But the
field has moved rapidly and the original Handbook is now dated.
I have now updated the old Handbook and renamed it “The Handbook of Redox
Biochemistry” for reasons explained in the text. Although now greatly expanded it
is, by necessity, selective in content. For readers wanting a more in depth view of
the whole field I refer them to the excellent books by Halliwell and Gutteridge,
Gilbert and Colton and several others mentioned in the reference section
accompanying each chapter.
I would very much appreciate having any errors, omissions or new findings
brought to my attention. I can be contacted on: [email protected]
Ian Acworth
August 2003
iii
CONTENTS
Frontis
Chapter 1. Introduction.
Oxygen Toxicity – From Microbes To Man.
Why Is Oxygen Toxic?
Free Radical Pro-Oxidants.
Reactive Oxygen Species, Reactive Nitrogen Species And Other
Pro-Oxidants
How Do Aerobic Organisms Survive Even When Pro-Oxidants Are Being Continuously
Produced?
Why Use Electrochemical Detection?
Conclusions.
References.
Chapter 2. The Chemistry Of Reactive Species.
Oxygen And The Reactive Oxygen Species (ROS).
1. Oxygen.
Properties.
Formation.
Chemical Reactions And Biological Significance
2. Ozone.
Properties.
Formation.
Chemical Reactions And Biological Significance.
Measurement.
3. Singlet Oxygen.
Properties.
Formation.
Chemical Reactions And Biological Significance
Measurement.
4. Superoxide (Radical Anion).
Properties.
Formation.
Electron Transport Chains.
Immune Defense.
Enzymes Reactions.
Oxygen-Heme Interaction
Metal-Catalyzed Auto-Oxidation
Chemical Reactions
Biological Significance.
The Pro.
The Con.
Control.
Measurement.
Auto-Oxidation And Redox Cycling
5. Hydrogen Peroxide.
Properties.
Formation.
Chemical Reactions And Biological Significance.
Measurement.
6. The Hydroxyl Free Radical.
i-ix
2
8
9
17
18
23
24
25
37
37
37
37
38
42
42
42
42
45
46
46
46
48
49
49
49
50
50
54
54
54
54
55
56
56
58
59
59
59
63
63
63
64
66
67
iv
Properties.
Formation.
Chemical Reactions And Biological Significance.
Measurement.
EPR.
HPLC.
Nitrogen And The Reactive Nitrogen Species (RNS).
1. Nitrogen.
Properties.
Formation.
Chemical Reactions.
2. The Oxides Of Nitrogen.
2.1 Nitric Oxide.
Physical Properties.
Formation.
Chemical Reactions And Biological Significance.
Measurement.
2.2 The Nitroxyl Anion And Nitrosonium Ion.
2.3 Peroxynitrite.
Properties.
Formation.
Chemical Reactions And Biological Significance.
Measurement.
2.4 Nitrosoperoxycarbonate And Nitrocarbonate.
2.5 Nitrogen Dioxide, The Nitronium Cation, And Nitrite.
Properties.
Formation.
Chemical Reactions And Biological Significance.
Measurement.
2.6 The Higher Oxides Of Nitrogen – Dinitrogen Trioxide,
Dinitrogen Tetroxide And Dinitrogen Pentoxide.
Properties.
Formation.
Chemical Reactions And Biological Significance.
Measurement.
2.7 S-Nitrosothiols.
Properties.
Formation.
Chemical Reactions And Biological Significance.
Measurement.
Halogenated Reactive Species (RHS).
1. Chlorine And Hypochlorous Acid.
Properties.
Formation.
Chemical Reactions And Biological Significance.
Measurement.
2. Nitrosyl Chloride, Nitryl Chloride And Related Compounds.
Properties.
Formation.
Chemical Reactions And Biological Significance.
Measurement.
Sulfur, Thiols And Thiyl Radicals (Some Reactive Sulfur Species [RSS]).
Properties.
Chemical Reactions And Biological Significance.
Measurement.
Carbonyl Compounds.
67
67
67
70
70
72
82
82
82
82
82
84
85
85
86
89
94
97
98
98
98
100
106
106
107
107
107
107
109
110
110
110
110
111
112
112
112
113
114
114
114
114
114
116
118
118
118
118
119
120
121
121
121
125
126
v
Properties.
Formation.
Chemical Reactions And Biological Significance.
Measurement.
The Pro-Oxidant Activity Of Low Molecular Weight Compounds And Other Xenobiotics
References.
126
126
128
129
131
134
Appendix 2.1 Background To Electrode Potentials.
Thermodynamics Of Reversible Cells.
Standard Electrode Potentials.
Some Comments On SEPs.
Coupled Redox Reactions.
References.
150
150
153
155
158
158
Appendix 2.2 Background To Kinetics.
First-Order Processes.
Second-Order and Pseudo-First-Order Processes.
Some Published Second-Order Rate Constants.
Measurement Of Reaction Order And Reaction Rates.
References.
159
160
160
161
165
165
Appendix 2.3 Background To The White Blood Cell.
Granulocytes.
Lymphocytes.
Monocytes.
167
168
168
169
Chapter 3. Damage And Repair.
DNA
Introduction.
The DNA Molecule.
DNA Damage.
The Consequences Of Oxidative DNA Damage.
Repair Of ROS/RNS-Induced Damage.
Base Excision Repair.
Nucleotide Excision Repair.
Mitochondrial DNA Repair.
Single Strand DNA Damage And PARP Activation.
What Do The Levels Of DNA Adducts Mean?
Steady State Levels.
Total Adduct Levels.
Measurement Of DNA Damage.
Gas- And Liquid-Chromatography-Mass Spectrometry.
HPLC.
Postlabeling assays.
Immunochemical detection.
The Measurement Of 8-Hydroxy-2’deoxyguanosine In Urine.
DNA Damage In Health And Disease.
Amino Acids And Proteins.
Introduction
Protein Molecular Structure
Pro-oxidants And Protein Damage.
The Indirect Pathway.
The Direct Pathway.
Oxidative Damage To Tyrosine.
Protein Repair And Degradation.
Amino Acid And Protein Damage In Aging And Disease.
171
171
171
175
185
187
188
190
190
190
191
191
198
203
205
206
208
209
210
213
214
214
215
218
218
222
226
230
234
vi
Measurement Of Amino Acid And Protein Damage.
Whole Protein.
Protein Hydrolysates.
Measurement Of Free Modified Amino Acids And Modified
Residues In Whole Proteins And Protein Hydrolysates
1. Protein Carbonyls.
2. Methionine Sulfoxide.
3. 2-Oxohistidine.
4. Tyrosine Markers.
3-Nitrotyrosine.
3-Chlorotyrosine.
Dityrosine.
Other Tyrosine Oxidation Products.
Lipids.
Introduction.
Structure Of Biological Membranes.
Lipid Damage.
The Role Of Metals In Lipid Peroxidation.
Lipid Oxidation Products.
Malondialdehyde.
4-Hydroxyalkenals.
Other Reactive Carbonyls.
Cholesterol Oxidation.
The Isoprostanes.
Lipid Repair.
Lipid Damage And Disease.
Measurement Of Lipid Damage.
Diene Conjugates.
TBAR.
Carbohydrates.
Introduction.
Ribose And Deoxyribose Damage.
Glycation, Glyoxidation, Advanced Glycation End Products
(AGES) And Age-Related Pigments.
References.
Appendix 3.1 Typical DNA Extraction And Hydrolysis.
DNA Extraction Procedure.
DNA Hydrolysis Procedure.
Chapter 4. Protection.
Introduction.
Enzymes.
Catalases.
Peroxidases.
The Biological Significance Of Catalase And Glutathione
Peroxidase.
Glutathione-S-Transferase.
Heme Oxygenases.
Superoxide Dismutases.
The Catabolism Of Nitric Oxide.
Sequestration Of Metal Ions.
The Metabolism Of Iron And Copper.
Iron And Copper Species As Pro-Oxidants.
Measurement Of Iron And Copper.
Low Molecular Weight Molecules.
237
237
240
241
241
241
241
242
242
246
248
248
248
248
249
251
256
257
260
262
264
264
266
268
269
270
272
272
276
276
277
278
280
305
305
307
309
310
315
317
319
320
320
321
323
324
326
329
330
330
vii
Water-soluble antioxidants.
Albumin.
Ascorbic Acid.
Antioxidant Properties.
Pro-oxidant Properties.
Measurement.
Thiols.
1. Glutathione.
Biological Roles Of Glutathione.
Protection.
Detoxification And Bioactivation.
Cofactor.
Storage Of Cysteine In A Non-Toxic
Form.
Amino Acid Transport.
Regulation.
Compartmentalization.
Conditions And Diseases Affecting
Glutathione.
Measurement Of Glutathione And Its Disulfide.
2. Homocysteine.
3. Miscellaneous Endogenous Sulfur-Containing
Compounds.
Uric Acid.
Formation.
Xanthine Oxidase And Tissue Injury.
Antioxidant And Pro-Oxidant Activities.
Measurement.
Fat-Soluble Antioxidants.
Carotenoids
Carotenoids And Disease.
Antioxidant And Pro-Oxidant Activities Of
Carotenoids.
Retinoids.
The Biological Activity Of The Retinoids.
Antioxidant And Pro-Oxidant Activities Of The Retinoids.
Measurement Of Carotenoids And Retinoids.
Quinones And Hydroquinones.
Coenzyme Q (Ubiquinone, Ubiquinol).
Biology Of Coenzyme Q.
Antioxidant And Pro-Oxidant Activities Of
Coenzyme Q
Measurement Of Coenzyme Q.
Plastoquinone.
Vitamin K.
Pyrroloquinoline Quinone.
Tocopherols
Biology Of Tocopherols
Antioxidant, Pro-Oxidant And Other Reactions Of The
Tocopherols
Tocopherol And Disease.
Measurement Of Tocopherols And Their Metabolites.
Other Endogenous And Exogenous Metabolites Proposed As Antioxidants.
Bile Pigments.
Biogenic Amines.
Estrogen.
330
330
340
340
344
345
346
346
347
347
349
350
351
351
351
352
352
352.
355
358
359
359
360
361
363
363
363
364
365
367
368
369
369
372
372
373
375
377
378
378
380
380
382
383
387
387
391
391
391
395
viii
Histidine Derivatives.
Indoles And Related Compounds.
α-Ketoacids.
α-Lipoic, Dihydrolipoic Acids And Analogs.
Melanins
Melatonin.
Phytochemicals.
Simple Phenolic Acids.
Flavonoids.
Phytoestrogens.
Resveratrol.
Phytic Acid.
Sulfur-Containing Compounds.
Pteridines.
Antioxidant Therapy.
Enzymes.
Chelators.
Low Molecular Weight Molecules.
Estimating The Total Antioxidant Capacity.
Antioxidants As Food Preservatives.
References.
396
397
398
399
401
402
406
406
410
415
418
419
419
420
421
421
422
424
432
438
441
Index.
480
ix
Chapter 1
Introduction
Mention a pure science such as chemistry or biology and most people will have a
fair idea about the subject matter. Unfortunately, for those interested in studying
the effects of reactive species on living organisms, no succinct and accurate
descriptor of this field exists. Several general titles have been used over the
years including free radical biology, redox chemistry and redox biology, yet none
of them do justice to this complex, multi-disciplined field. While free radical
biology ignores the fact that many chemical species being studied are not free
radicals, redox chemistry implies a disregard for any biological aspects. Oxidative
metabolism has been used but this is usually associated with energy metabolism.
Although still not perfect I prefer the term Redox Biochemistry. I will discuss free
radicals and redox reactions in greater detail below.
WWW.ESAINC.COM
1
OXYGEN TOXICITY – FROM MICROBES TO MAN
Oxygen is toxic to aerobic (and anaerobic) organisms, yet paradoxically oxygen
is essential for their survival. Today terrestrial aerobes (both animals and plants)
have successfully adapted to live in an atmosphere composed of approximately
21% oxygen and can survive minor fluctuations in the level of respired oxygen
without disastrous consequences. True anaerobes, on the other hand, tolerate
oxygen poorly, and some cannot survive even a brief exposure to atmospheric
oxygen (Table 1.1).
Anaerobes were the first living organisms on the planet. These evolutionary
simple organisms show a wide range of oxygen tolerance. Strict or obligate
anaerobes will only grow if oxygen is absent. While some obligate anaerobes are
killed almost immediately following exposure to oxygen (aerophobic) (e.g.,
Clostridia species) others can survive for many days but cannot reproduce (e.g.,
Bacteroides fragilis). Another group of organisms, microaerophiles actually
require some oxygen for growth but cannot survive when exposed to
atmospheric oxygen concentrations. Most bacteria that reduce nitrate (producing
nitrite, nitrous oxide or nitrogen) are called facultative anaerobes as they are not
affected by exposure to oxygen and in fact will preferentially use oxygen, rather
than nitrate, during respiration.
Anaerobes can be found in any environment where oxygen levels are decreased
to less toxic levels including muds and other sediments; bogs and marshes;
polluted waters; certain sewage-treatment systems; rotting material; deep
underground areas such as oil pockets; the sources of springs; decaying teeth
and gangrenous wounds; the colon; and inappropriately canned foods. Rather
than using oxygen during respiration (they usually lack terminal cytochromes that
transfer electrons to oxygen) they use other electron acceptors such as ferric
ions, sulfate or carbon dioxide which become reduced to ferrous ions, hydrogen
sulfide and methane, respectively, during the oxidation of NADH (reduced
nicotinamide adenine dinucleotide is a major electron carrier in the oxidation of
fuel molecules) (Figure 1.1). Oxygen is toxic to anaerobes as it can affect the
organism’s internal homeostasis by altering its reductive capacity, consuming
compounds such as NAD(P)H, thiols and other chemicals essential for
biosynthetic reactions and inactivating key enzymes.
Although anaerobes had free range during the early stages of the evolution of
living organisms, this was eventually curtailed by the success of oxygenproducing photosynthetic plants. With the levels of oxygen rising in the
atmosphere, anaerobes had three choices, adapt, find niches where oxygen
would not penetrate, or die. Organisms eventually evolved that not only survived
in an oxygen-enriched atmosphere but prospered. Evidence suggests that the
atmospheric oxygen levels have fluctuated markedly over time, increasing from
15-18% in the late Devonian to as high as 35% in the late Carboniferous and
early Permian periods. This hyperoxia has been suggested to be one of the
WWW.ESAINC.COM
2
possible causes of the mass extinction of terrestrial vertebrates (Graham et al.
1995). Atmospheric oxygen finally stabilized at today’s level (at least to date).
1) Sulfate Reduction (e.g., Desulfovibrio (water- logged soils), Desulfomaculum (spoilage of canned foods),
Desulfomonas (intestines), Archaeglobus (a thermophile)):
Step 1: Sulfate is activated.
ATP + SO42-
Adenosine phosphosulfate (APS) + PPi
Step 2: A hydrogenase splits molecular hydrogen. Reduction of APS produces sulfite.
APS + H2
SO32- + AMP + H2O
Cyt c3
Step 3: Electrons derived from hydrogen reduce sulfite to hydrogen sulfide
SO32- + 6H + + 6e -
H2S + H2O + 2OH-
2) Methanogenesis pathway (e.g., Methanebacterium thermoautotrophicum
Formylmethanofuran
Carbon Dioxide
H2O
H4MPT
Methanofuran
+ 2H + + 2e -
N5-Formyl-5,6,7,8-tetrahydro
methanopterin
Methanofuran
H2O
N5, N10-Methylenetetrahydro
methanopterin
F420
F420H2
)
N5, N10-Methenyltetrahydro
methanopterin
F420H2
F420
Methyl-Coenzyme M
5-Methyl-5,6,7,8-tetra
hydromethanopterin
HTP
CoM
H4MPT
CoM-S-S-HTP
Methane
CO2 + 8H + + 8e - = CH4 + 2H 2O
Figure 1.1 Anaerobic Metabolism.
WWW.ESAINC.COM
3
Glucose
ATP
-ATP
ADP
Glycogen
Glucose 6-phosphate
Fructose 6-phosphate
ATP
-ATP
ADP
Fructose 1,6-bisphosphate
Dihydroxyacetone phosphate
2 x Glyceraldehyde 3-phosphate
2NAD+ + 2Pi
NADH + H+
2NADH
NAD+
2 x 3-Phosphoglyceroyl phosphate
ADP
2ADP
Substrate-level
phosphorylation
ATP
Glycerol
+2ATP
2ATP
2 x 3-Phosphoglycerate
Fatty Acids
Triglycerides
2 x 2-Phosphoglycerate
2 x Phosphoenolpyruvate
2ADP
Anaerobic
Glycolysis
2 xLactate
2ATP
2 x Pyruvate
2NAD+
2NADH
2CO2
+2ATP
Some
amino
acids
Σ = +2ATP
2 x Ethanal
2NADH
Anaerobic
Fementation
To Tricarboxylic Acid Cycle
2NAD+
2 x Ethanol
Figure 1.2 The Glycolytic Pathway And The Production Of ATP.
WWW.ESAINC.COM
4
Group
Oxidizing
Conditions
Growth
Reducing
Conditions
No growth
AerobeFacultative
Growth
Growth
Not required but
better if oxygen is
present
Anaerobeaerophobic
(obligate;
strict)
Death
Growth
Harmful
Anaerobeaerotolerant
(moderate)
Growth
Growth
Not required but
better if oxygen is
present
Microaerophile
Growth if
oxygen level is
not too high
Growth if
oxygen level
not too low
Required but at only
low levels
AerobeObligate
Effect of Oxygen
Example
Essential
Many bacteria,
most fungi,
algae, protozoa,
all higher plants
and animals
Bacteria such as
enteric and
pathogenic
species; some
protozoa, yeasts
(e.g.,
Saccharomyces)
and fungi
Many bacteria
some protozoa.
Bacteroides,
Clostridia,
Fusobacterium,
Methanobacteriu
m, and
Ruminococcus
Bacteroides
fragilis,
Treponema
pallidum
Campylobacter
jejuni
Table 1.1 The Effects Of Environment And Oxygen On Growth Of Aerobes
And Anaerobes
Facultative aerobes (Table 1.1) can survive in the presence or absence of
oxygen. They obtain their energy either by oxidative phosphorylation or
fermentation and do not require oxygen for synthesis. When oxygen is lacking
this group of organisms can oxidize some organic compounds (which act as both
electron donors and acceptors) with a small release of energy, in a process
called fermentation. A variety of compounds can be fermented including most
sugars, many amino acids, some organic acids, purines, pyrimidines and a
variety of miscellaneous products. The energy is captured as two molecules of
adenosine triphosphate (ATP) in a process termed substrate level
phosphorylation. ATP is the cell’s immediate energy providing molecule and is
used for growth, movement, and in biochemical processes e.g., biosynthesis and
maintenance of ionic gradients. The stepwise breakdown of glucose into
pyruvate is called glycolysis and occurs in both facultative and obligate aerobes
(Figure 1.2). In fermentation, pyruvate produced by glycolysis is converted to
ethanol or lactate (Figure 1.3). In the presence of oxygen however, glycolysis is
WWW.ESAINC.COM
5
followed by aerobic respiration and pyruvate is completely oxidized to carbon
dioxide and water (Figure 1.4). The oxidation of pyruvate takes place in a series
of steps called the tricarboxylic acid cycle (TCA) (also called the citric acid or
Krebs cycle) that occurs in the mitochondrion (Figure 1.4). During aerobic
respiration the oxidation of glucose generates 36 molecules of ATP. Two ATP
molecules are generated by substrate level phosphorylation (part of the cytosolic
glycolytic pathway) and two are produced by substrate level phosphorylation
occurring in the mitochondrion. However, the vast majority, thirty-two ATP
molecules, are produced by mitochondrial oxidative phosphorylation when
electrons are transferred from NADH or flavin adenine dinucleotide (reduced)
(FADH2) to oxygen by a series of electron carriers. Thus it can be seen that
aerobic respiration generates much more energy than anaerobic processes. For
example, if pyruvate is completely oxidized by the TCA cycle then yeast will be
able to form 19 times more energy from a given amount of glucose when growing
aerobically than when growing anaerobically.
ACETALDEHYDE
GLUCOSE
Glycolysis
ETHANOL
LACTATE
PYRUVATE
ANAEROBIC
Oxidative
Phosphorylation
(TCA/electron
transport)
AEROBIC
CO2 + H2O
Figure 1.3 The Metabolic Fate Of Pyruvate.
Obligate aerobes (e.g., higher plants and animals) use oxygen in respiration and
for the biosynthesis of a variety of biomolecules. All higher organisms are
obligate aerobes but they can make use of both anaerobic and aerobic
processes. For example, many tissues such as the red blood cell, the cornea of
the eye, the skin, the kidney medulla and type IIb (fast twitch-glycolytic) skeletal
muscle fibers make use of anaerobic glycolysis. Here the two molecules of ATP
produced by the anaerobic conversion of glucose to lactate is sufficient to supply
most of these tissues’ normal energy needs. However, as the average human
requires more than 40kg/day of ATP, and as much as 0.5kg/minute when
undergoing strenuous exercise, anaerobic respiration simply cannot keep pace
with this demand. Rather, higher organisms must obtain the vast majority of their
energy from aerobic respiration, and that is why oxygen is essential for their
survival.
WWW.ESAINC.COM
6
Glycolysis
Pyruvate
CoA + NAD+
Amino Acids
xNADH, xFADH2
NADH + CO2
Acetoacetyl CoA
Acetyl CoA
Fatty Acids
3ATP
O2
Oxaloacetate
Citrate
NADH
H2O
NAD+
Malate
Isocitrate
H2O
NAD+
2ATP
Fumarate
O2
FADH2
FAD
Amino
Acids
a-ketoglutarate
(2-oxoglutarate)
H2O
Succinate
H2O
Succinyl
CoA
GTP + CoASH
ADP
NADH + CO2
NAD+
3ATP
O2
NADH + CO2
GDP + Pi
Substrate-level
Phosphorylation
ATP
GDP
H2O
3ATP
O2
Electron transport chain
Figure 1.4 The Tricarboxylic Acid Cycle.
Obligate aerobes are very oxygen sensitive. A total lack of oxygen is referred to
as anoxia and rapidly results in cell death. For example, brain damage can result
from perhaps as little as three minutes of anoxia. An acute decrease in respired
oxygen leads to hypoxia, a situation where oxygen is still delivered to the tissue,
but at a rate insufficient to maintain normal cellular processes. The effects of
hypoxia depend upon the tissue and the degree and duration of the hypoxic
event. For example, the brain is a very aerobic tissue and is exquisitely sensitive
to oxygen tension. In higher animals an acute reduction in arterial oxygen tension
WWW.ESAINC.COM
7
leads to altered mental function, analgesia and loss of muscle coordination
(Blass and Gibson (1979); Gibson and Blass (1976); Gibson et al. (1978; 1981)).
A more marked drop can result in unconsciousness, progressive depression of
the central nervous system, circulatory failure and death.
Ischemia is a consequence of mechanical disruption of blood flow to a tissue
resulting in decreased oxygen, glucose and ATP levels. For example, the
occlusion of essential blood vessels to the heart (a consequence of
atherosclerosis and/or blood clots) results in ischemia. This leads to myocardial
damage and heart attack. It has been estimated that irreversible myocardial
damage can occur after about 20 minutes of ischemia (Sobel (1974)). The
affected tissue eventually dies.
Exposure to elevated levels of oxygen results in hyperoxia and is deleterious to
aerobic microorganisms, plants and animals. The growth of aerobic bacteria is
inhibited following exposure to pure oxygen. Plants show decreased chloroplast
development and leaf damage when exposed to oxygen levels above normal.
Animals exposed to 100% oxygen show a variety of symptoms depending upon
the duration of exposure (Crapo et al. (1980); Francica et al. (1991)). Humans
suffer chest soreness, coughing and sore throats following several hours of
exposure to pure oxygen. Longer periods cause alveolar damage, edema and
permanent irreversible lung damage. Hyperoxia also leads to damage to most of
the major organs. Unfortunately, earlier this century unintentional retinal damage
and blindness (retrolental fibroplasia) was caused to premature babies when they
were maintained on high oxygen levels in their incubators. Fortunately, the level
of oxygen to which premature babies are exposed is now more carefully
monitored. It should be noted, however, that hyperoxia can also be beneficial.
For example, hyperbaric oxygen is used to treat gangrene because of its toxicity
to the obligate anaerobes that cause it.
Correct oxygen tension is important to deep sea divers, astronauts, mountain
climbers, athletes going from low to high elevations and those undergoing
general anesthesia. Oxygen tension is also important in preventing the growth of
harmful anaerobic pathogens in canned and bottled foods and beverages.
WHY IS OXYGEN TOXIC?
Over the years, several theories have been put forward to explain oxygen’s
toxicity. This subject was reviewed recently by Gilbert (1999) so only an overview
will be presented here.
•
One early hypothesis as to oxygen’s toxicity was that oxygen exerted its
action through enzyme inhibition. For example, oxygen can inhibit
nitrogenase and the first enzyme in the dark reactions of photosynthesis,
WWW.ESAINC.COM
8
•
•
•
•
ribulose 1,5-bisphosphate decarboxylase, and at high concentrations
some thiol-containing enzymes (Haugaard (1946); Stadie et al. (1944)).
However, enzyme inhibition is far too slow and limited to explain oxygen’s
toxic effect, and not all enzymes are affected by oxygen.
Abundant evidence showed that irradiation caused DNA damage and
cancer through a free radical mechanism and that oxygen had a
sensitizing effect (von Sonntag (1991) and references therein).
In the mid 1950s Gerschman and Gilbert proposed that oxygen, itself a
diradical, may exert its toxic action through the formation of free oxygen
radicals. These could then damage biologically important macromolecules
such as DNA, proteins and lipids (see Gerschman (1981); Gerschman et
al. (1954); and reviews by Gilbert (1999); Halliwell and Gutteridge (1993)).
This breakthrough proposal, however, was initially strongly criticized by
researchers who proposed that free radicals were far too reactive to exist
in any great quantity in biological materials. These objections were finally
laid to rest by the detection of free radicals both in dry biological tissues
and in living organisms by electron spin resonance (Commoner et al.
(1954, 1957)).
In 1954 Harman developed his free radical theory of aging that postulated
that “a single common process, modifiable by genetic and environmental
factors, was responsible for the aging and death of all living things”
(Harman (1956; 1992a,b)). His theory proposed that the accumulating
irreversible damage to biologically important macromolecules over time
led to disease and aging.
Free radicals were further implicated by the discovery of the enzyme
superoxide dismutase (SOD). Fridovich theorized that the superoxide
radical anion was the major toxic form of oxygen and that SOD protected
against it (Fridovich (1983, 1986a,b); McCord and Fridovich (1969)). The
superoxide theory of oxygen toxicity, though not completely correct, was
responsible for a great deal of experimental work and a better
understanding of the field as a whole (reviewed in Halliwell and Gutteridge
(1993)).
We now know that oxygen mediates its toxic effects through a variety of
compounds, not just free radicals, many of which contain other atoms in addition
to oxygen. The properties of these species will be dealt with in Chapter 2.
FREE RADICAL PRO-OXIDANTS.
The term radical originally used by chemists referred to an ionic group that had
either positive or negative charges associated with it (e.g., carbonate, sulfate
etc.). A free radical is now defined as an atom or molecule that has one or more
unpaired electrons (i.e., electrons that occupy atomic or molecular orbitals by
themselves) and is capable of independent existence. In the strictest sense the
free of free radical, is redundant. It may come as some surprise that oxygen is a
WWW.ESAINC.COM
9
free radical (in fact a diradical) as are metals that have incomplete 3d shells (e.g.,
transition metals and their various oxidation (valency) states) (Table 1.2).
Scandium
1s22s22p63s2 3p63d14s2
Sc3+
(NT)
…3p6
Titanium
1s22s22p63s23p63d24s2
Ti2+
…3p6d2
Ti3+
…3p6d1
Vanadium
1s22s22p63s23p63d34s2
V2+
…3p6d3
V3+
…3p6d2
Chromium
x
1s22s22p63s23p63d44s2
1s22s22p63s23p63d54s1
Cr2+
Mn2+
…3p6d4
…3p6d5
Cr3+
Mn3+
…3p6d3
…3p6d4
Iron
1s22s22p63s23p63d64s2
Fe2+
…3p6d6
Fe3+
…3p6d5
Cobalt
1s22s22p63s23p63d74s2
Co2+
…3p6d7
Co3+
…3p6d6
Nickel
1s22s22p63s23p63d84s2
Ni2+
…3p6d8
Copper
1s22s22p63s23p63d104s1
Cu2+
…3p6d9
Cu+
(NT)
…3p6d10
Zinc (NT)
1s22s22p63s23p63d104s2
Zn2+
(NT)
…3p6d10
Table 1.2 The Electronic Configuration Of The Atoms Of First Transition
Series And Some Of Their Ions. (NT – non-transition. Note that NT compounds are also
non-radicals.)
Free radicals can be formed when a non-radical either gains or loses a single
electron (Table 1.3). Free radicals can be formed during homolytic fission of
covalent bonds. The energy required to cause bond dissociation can be brought
about by several different processes, including exposure to heat or
electromagnetic radiation, or by chemical reaction. Remember that covalent
bonds are formed when two atoms share electrons (usually one from each atom).
During homolytic fission one electron of the bonding pair is retained by atom A,
while the other is retained by atom B forming the free radicals A• and B•,
respectively. During homolysis of water, for example, the hydroxyl free radical
(HO•) and the hydrogen atom (H•) are produced. Radical reactions are much
more common in the gas phase and at high temperatures, e.g., combustion.
Readers should be aware that many radical reactions found in the literature
(especially chemistry texts) may be for gas phase reactions and are not always
applicable to biological systems. Having said this, gas phase free radical
chemistry is extremely important to those investigating the effects of atmospheric
pollution and cigarette smoke on biological systems.
WWW.ESAINC.COM
10
1. Heat.
Radicals produced during combustion or by heating in absence of oxygen,
e.g., C—C, C—H bonds typically require 450-600oC
2. Electromagnetic radiation.
Including ionizing irradiation (e.g., x-rays, γ-rays) and photolysis (e.g., UV
absorption)
3. Redox reactions.
Radicals are produced in reactions involving one-electron transfer:
• inorganic ions (e.g., ArN2+ + Cu+ → Ar• + N2 + Cu2+; Sandmeyer reaction)
• metals (e.g., H2O2 + Fe2+ → Fe3+ + HO• + OH-, Fenton reaction)
• electrolysis (e.g., 2RCO2- - e- → 2RCO2• → → R—R; Kolbe synthesis)
• hydroquinone-semiquinone-quinone systems (e.g., production of superoxide
from oxygen by ubiquinol/ubiquinone redox couple)
4. Enzymatic.
Radicals are produced by the action of peroxidases (e.g., horseradish peroxidase)
or oxidases (e.g., xanthine oxidase)
5. Chemical.
By the reaction of hydroxyl free radical with a variety of substrates
By the reaction of peroxynitrite with a variety of substrates
As part of enzyme catalyzed reactions
By reactions involved in the generation of O2•- during mitochondrial respiration
By the reaction of oxygen with other radicals:
• Production of lipid peroxyl radical when oxygen reacts with an alkyl radical
• Production of peroxynitrite radical when oxygen reacts with nitric oxide
By thermal decomposition of azo initiators (R-N=N-R):
• 2,2’-azo-bis(2-amidinopropane) dihydrochloride [AAPH] for aqueous systems
• 2,2’-azo-bis(2,4-dimethylvaleronitrile) [AMVN] for lipophilic systems
By thermal decomposition of organic peroxides:
• Di-tert-butyl peroxide
• Dibenzoyl peroxide
6. Ultrasound.
Also called sonochemical production. Primary radicals (e.g., H• and HO•) are
produced due to pyrolysis of molecules located within collapsing cavitation
microbubbles, while secondary radicals are formed by hydrogen abstraction or
addition of primary radicals to other molecular species
7. Lithotripsy.
Radicals are produced when high-energy shock waves are used to destroy solid
objects, e.g., kidney stones
8. Lyophilization.
Radicals can be produced by freeze-drying/thawing processes
WWW.ESAINC.COM
11
Table 1.3 Free Radicals Can Be Produced In A Variety Of Ways. This
table summarizes both in vitro and in vivo approaches for free radical production. (Crum et
al. (1987); Doss and Swartz (1984); Fuciarelli et al. (1995); Halliwell and Gutteridge
(1999); Heckly and Dimmick (1967); Hendrickson et al. (1970); Kondo et al. (1993); Misik
and Riesz (1999); Misik et al. (1996, 1999); Morgan et al. (1988); Ostrowski (1969); Seel
et al. (1991); Suhr et al. (1994); Vreugdenhil et al. (1991); Worthington et al. (1997)).
A wide variety of radicals can exist (Table 1.4). Like any other chemical, radicals
show a broad spectrum of physical and chemical properties. Some are stable
and unreactive, whereas others react extremely rapidly. Some are hydrophobic
while others hydrophilic. Radicals may share certain common characteristics and
can be grouped together as presented in the following table. Unfortunately, as
will be readily apparent such classification is not perfect as some radicals can
belong to more than one category. For example, some sigma radicals are also
carbon-centered monoradicals.
•
•
•
•
•
•
•
•
•
•
•
Radical
σ (sigma)
π (pi)
Monoradicals
Polyradicals
Carbon centered
Oxygen centered
Sulfur centered
Nitrogen centered
Reducing
Oxidizing
Metal
Examples
H• (hydrogen atom), R• (carbon-centered radical), R3C•
Ascorbyl•, Tocopheryl•, NAD•
R•, R3C•, NO•
O2 (a diradical)
R•, R3C•
LO2•
RS•, RSO2•
NO•, R2•NO, •NO2
CO2•-, PQ•HO•, LO2•
Cu2+, Fe2+, Fe3+
Table 1.4 Different Types Of Radicals. R is used as an abbreviation for
an alkyl group, L represents a lipid (e.g., fatty acid). Based on an original by G.R.
Buettner.
Of all the radicals that can be formed sigma (σ) radicals (e.g., the methyl radical,
CH3•) are generally much more reactive than pi (π) (e.g., the tocopherol-derived
radical, tocopheryl•) as their lone electron cannot be spread throughout the
molecule (delocalized). π-Radicals are generally less reactive than σ ones
because the lone electron is not confined to just one atom, but is delocalized
through the conjugated π-bond system (Sykes (1975)). A physiological
consequence is that σ-radicals play an important role in initiating lipid
peroxidation while chain-breaking antioxidants prevent lipid peroxidation by
reacting with the σ-radicals forming a much less energetic and less dangerous πradical species.
WWW.ESAINC.COM
12
Most free radicals have very short lifetimes. Without stabilizing features (e.g.,
delocalization or steric hindrance) they decompose rapidly, often in the absence
of external agents. Decomposition is usually through:
1. Unimolecular reactions (e.g., fragmentation or rearrangement),
2. Bimolecular reactions between radicals including dimerization (e.g., the
formation of peroxynitrite from nitric oxide and the superoxide radical
anion or the formation of hydrogen peroxide from two hydroxyl free
radicals) or disproportionation (e.g., the formation of hydrogen peroxide
and oxygen from two hydroperoxyl radicals1) which can involve electron or
hydrogen atom transfer or
3. Bimolecular reactions between radicals and other molecules (e.g.,
addition, displacement, or atom [often H] abstraction). Further information
can be found in good chemistry texts.
Phase
Example
Initiation
Fe2+ + H2O2 → Fe3+ + HO• + OHL—H + HO• → L• + H2O
Propagation
L• + O2 → LO2•
LO2• + L—H → LO2H + L•
Termination
L• + L• → L—L (dimerization)
LO2• + L• → L—L + O2
2LO2• → non-radical products
2C2H5• → C2H6 + C2H4 (disproportionation)
reduced oxidized
Table 1.5 The Three Phases Of Chain Reactions.
(L represents a lipid undergoing peroxidation.)
In biological systems the most infamous free radical cascade is the lipid
peroxidation chain reaction (Table 1.5). Here a single initiation process can lead
to the destruction of many poly-unsaturated fatty acid molecules. Unfortunately,
not only does this affect membrane fluidity and thus many biochemical
processes, but it can also lead to the production of cytotoxic carbonyl breakdown
products (Chapter 3). Lipid peroxidation is also the major process responsible for
food spoilage. Like any other chain reaction, lipid peroxidation consists of three
phases termed a) initiation, b) propagation and c) termination. Biological systems
are equipped with several mechanisms designed to prevent lipid peroxidation.
Such processes include prevention of radical formation (inhibiting initiation) or
1
Note during disproportionation one species is reduced while the other is oxidized. involving
electron or hydrogen atom transfer),
WWW.ESAINC.COM
13
interception of fatty acid radicals once formed (inhibiting propagation). Biological
systems are also capable of repairing damage that occurs.
Several techniques can be used to measure free radicals. Electron paramagnetic
resonance [EPR] (also called electron spin resonance, or ESR) is a very useful
technique and is the only way to directly measure radicals. EPR makes use of
the fact that the unpaired electron in a free radical has spin (either +1/2 or –1/2)
and thus behaves as a small magnet (i.e., is paramagnetic). When placed in an
external magnetic field the unpaired electron can align itself, either parallel or
antiparallel, to that field (i.e., the free electron only has two possible energy
levels). Exposure to electromagnetic radiation of the correct energy will move the
electron from the lower energy level to a higher excited one. Thus an absorption
spectrum is obtained which can be used for quantitation as well as gaining
information about the environment surrounding the free radical (see Halliwell and
Gutteridge (1993)).
Direct EPR methods have a sensitivity limit of 0.1nmol/L and have been used
extensively for in vitro work (e.g., to study the mechanism of enzyme action) but
are often not selective enough for most in vivo work. Some researchers are,
however, developing these techniques. Many free radicals are too reactive (e.g.,
HO•) and have too short a half-life for direct EPR methods. This can be overcome
by using spin-trap agents that react with the free radical to produce a longer-lived
species that is still paramagnetic (Figure 1.5). Interestingly, spin traps are also
proving to be beneficial in the treatment of diseases thought to involve oxidative
stress where they probably act to scavenge damaging free radicals. For
example, α-phenyl-tert-butylnitrone (PBN) is being used at pharmacological
levels to decrease ischemia-reperfusion injury in brain (Floyd (1990), Folbegrova
et al. (1995)) and dog heart (Bolli et al. (1988)); reduce the size of liver edema in
carbon tetrachloride intoxicated rats (Towner et al. (1993)); reduce the mortality
associated with endotoxic shock in rodents (Miyajima and Kotake (1997) and
references therein) and prolong the life span of the senescence-accelerated
mouse model (Edamatsu et al. (1995)).
The correct choice of a spin-trap agent is important. The ideal spin-trap should
readily and specifically react with the radical of interest. It must also produce an
adduct of sufficient longevity which possesses a characteristic EPR spectrum. It
should never decompose during experimentation producing free radicals (see
Halliwell and Gutteridge (1993)). Further limitations are placed upon a spin-trap
by biological systems. The ideal reagent must not be toxic and should readily
pass though any biological barrier (e.g., the blood-brain barrier) to reach the site
of free radical production. A major problem with some spin-trap adducts is that
they can be reduced in vivo by cellular reducing agents such as ascorbic acid
and thiols, resulting in the production of diamagnetic (non EPR active) species.
WWW.ESAINC.COM
14
CH3
CH3
N=O
CH3
+
R
CH3
Radical
(less stable)
tert -Nitrosobutane
+
CH=N
O
CH3
O
N
R
CH3
Spin-trap adduct
(more stable)
O
+
HO
C(CH3)3
CH-N
C(CH3)3
OH
α−Phenyl- tert- butylnitrone
Spin-trap adduct
PBN
PBN-OH
Figure 1.5 Spin Traps React with Free Radicals to Produce Paramagnetic
Products that can be Measured using EPR.
A different approach to spin trapping is radical scavenging. Here the free radical
reacts with an aromatic scavenging agent (e.g., salicylic acid). The aromaticradical adduct can then be quantified using HPLC-based techniques. This
approach is much more versatile than spin trapping as neither the scavenging
agent nor the product needs to be a radical. Scavengers are usually less toxic
than spin traps. Furthermore, as scavenging agents and products are
electrochemically active they can be measured at biologically relevant levels
using HPLC with electrochemical detection (see ESA Application Notes: 70-1749
Hydroxyl Free Radical Measurement; 70-4820 Alternative Method for Hydroxyl
Free Radical Measurement). The use of aromatic scavenging agents will be
revisited in Chapter 2.
WWW.ESAINC.COM
15
Reactive Oxygen
Species (ROS)
Reactive Nitrogen
Species
(RNS)
Other
A) Free Radicals
Alkoxyl
LO•
Nitric Oxide (monoxide)
NO•
Hydroperoxyl
HO2•
Nitrogen Dioxide
NO2•
Hydroxyl
Peroxyl
Superoxide
HO•
LO2•
O2-
Peroxynitrite radical
ONO2•
Hydrogen
Peroxide
H2O2
Alkyl Peroxynitrite
LO2NO-
Lipid Peroxides
Oxygen
LO2H
O2
Chloramine
Dinitrogen Pentoxide
NH2Cl
N2O5
Ozone
O3
Dinitrogen Tetroxide
N2O4
Singlet Oxygen
1
Dinitrogen Trioxide
N2O3
Nitrate
NO3-
Nitrite
Nitrocarbonate
Nitronium (Nitryl)
Nitrosonium (Nitrosyl)
Nitrosoperoxycarbonate
Nitrosonium Chloride
Nitroxyl
Nitronium Chloride
Peroxynitrite
Taurine
monochloramine
Thionitrites
(S-nitrosothiols)
NO2O2NOCO2NO2+
NO+
ONO2CO2NOCl
NONO2Cl
ONO2SO3(CH3)2NHCl
RSNO
Carboncentered
Radicals
Disulfide
Radical
Hydrogen Atom
Thiyl Radical
e.g., CCl3•
RSSR•H•
RS•
B) Non Radicals
Singlet Oxygen
∆g
O2
1 +
Σg
O2
Aldehydes
(e.g., 4hydroxynonenal)
Disulfide
Hypohalous
Acid
Hypothiocyanic
Acid
Malondialdehyde
Transition
metal ions
RCHO
RSSR
e.g., HOCl
and HOBr
HOSCN
CHOCH2CHO
e.g., Fe2+,
Fe3+
Table 1.6 The Different Pro-Oxidants And Other Species Of Importance To
Biological Systems. (L – alkyl; 1∆g and 1Σg+ represent the two forms of singlet oxygen; X• – a
radical species).
WWW.ESAINC.COM
16
REACTIVE OXYGEN SPECIES, REACTIVE NITROGEN SPECIES
AND OTHER PRO-OXIDANTS.
Although often referred to as free radicals, many of the compounds of interest to
the field of redox biochemistry are not free radicals and include many non-radical
species (Table 1.6). The term reactive (or reduced) oxygen species (ROS) is also
commonly used despite the fact that not all of the oxidizing species are reactive
(e.g., the hydroxyl free radical is typically ten million times more reactive and
much less selective than hydrogen peroxide), or are produced by the reduction of
oxygen (e.g., ozone and singlet oxygen are not reduced forms of oxygen).
Furthermore, the use of the term ROS does not take into account that many
species contain nitrogen, chlorine or sulfur. Reactive nitrogen species (RNS) is
commonly used to distinguish those compounds that contain nitrogen in addition
to oxygen, again with disregard for the variation in reactivity between members of
the group. As no suitable descriptors can be found, I will use the word prooxidant.
Pro-oxidant Species
Ferryl species
Hydrogen peroxide
Hydrogen peroxide and tyrosine
radicals
Hydrogen peroxide
Hydrogen peroxide
Hydrogen peroxide, phenoxyl radicals
Hydrogen peroxide
Hydrogen peroxide
Hydrogen peroxide
Hydrogen peroxide, superoxide and
nitric oxide
Comments
Essential to catalytic activity of cytochrome P450
and peroxidases.
The explosive oxidation of hydroquinone by
hydrogen peroxide in the presence of catalase and
peroxidase is used to generate a hot defensive
spray by the bombardier beetle.
Required for the production of thyroxine by the
thyroid peroxidase enzyme.
Estrogen-induced uterine peroxidase activity plays a
role in estrogen catabolism and may confer
bactericidal activity too.
Involved in the bioluminescence of several animal
species.
Involved in the formation of lignin. Oxidation and
polymerization of tyrosine and phenylalanine
residues catalyzed by peroxidases bound to the
plant cell wall.
With peroxidases are used by fungi to degrade
lignin.
Involved in fruit ripening.
Fertilization of sea urchin eggs causes the rapid
uptake of oxygen and production of hydrogen
peroxide that is used by a peroxidase to produce
tyrosyl radicals from tyrosine residues. These
radicals readily dimerize to dityrosine cross-linking a
fertilization membrane that prevents further
spermatozoa from entering the egg.
Redox regulation of gene expression, signal
transduction and intracellular redox signaling.
Activation of a transcription factor such as SoxS
WWW.ESAINC.COM
17
leads to the stimulation of transcription thereby
permitting bacteria to gain resistance to oxidants,
antibiotics and immune cells that generate nitric
oxide.
Nitric oxide can activate the Ras oncoprotein by Snitrosylation of essential cysteine residue,
stimulating GTPase activity and downstream
signaling through activation of extracellular signal
regulated kinase (ERK kinase).
Hydrogen peroxide produced by a plasma
membrane-bound NAD(P)H oxidase is activated by
insulin and may act as an intracellular signal for this
hormone promoting uptake of glucose and
preventing triglyceride hydrolysis in adipocytes.
Platelet derived growth factor uses hydrogen
peroxide as intracellular messenger.
Lipid peroxides and carbonyl
metabolites
Lipid centered radicals
Nitric oxide
Nitric oxide
Nitric oxide, (nitrosothiols) – endothelialderived relaxing factor
ROS, RNS, HOBr, HOCl, Cl2
Tyrosine, tryptophan, glycine and thiyl
radicals
Vitamin K hydroquinone and
semiquinone
Possibly act as antifungal and antibacterial agents
protecting damaged plants from infection.
Prostaglandin and leukotriene metabolism.
Retrograde neurotransmitter.
Bone synthesis, degradation and remodeling.
Blood pressure regulation.
Immune system – defense.
Essential to catalytic activity of several enzymes
such as ribonucleoside diphosphate reductase and
pyruvate dehydrogenase.
Required for carboxylation of glutamate to γcarboxylglutamic acid by microsomal glutamic acid
carboxylase. Important in blood clotting.
Table 1.7 Pro-oxidants Are Beneficial Too. (Halliwell and Gutteridge (1999) and
references therein; and other references at the end of this chapter).
HOW DO AEROBIC ORGANISMS SURVIVE EVEN WHEN PROOXIDANTS ARE BEING CONTINUOUSLY PRODUCED?
The cells of aerobes are constantly being exposed to pro-oxidants.
Consequently, their DNA, proteins, and lipids are continuously being damaged.
During evolution one option would have been to prevent the formation of prooxidant species. This, however, would be virtually impossible to achieve in an
oxygen-enriched environment as pro-oxidants are unavoidable side reactions of
other important biochemical processes. Instead nature accepted that prooxidants would be produced so protective mechanisms evolved to repair and
replace damaged molecules. In addition we are equipped with a suite of
antioxidant defenses designed to prevent the formation of pro-oxidants, or to
WWW.ESAINC.COM
18
intercept and destroy them if formed. Interestingly, aerobes also make good use
of pro-oxidants as messengers, signals and defense molecules (Table 1.7).
Under normal conditions the production of pro-oxidants is presumed to be in
balance with antioxidant defenses. However, the overproduction of pro-oxidants
and/or decreased antioxidant protection can lead to tissue damage and disease.
Thus, in individuals with a genetic predisposition or for those exposed to
environmental stressors such as cigarette smoke, sunlight and pollution, the prooxidant/antioxidant balance can be upset (Figure 1.6). The overproduction of prooxidant species or the failure of antioxidant defenses results in a condition called
oxidative stress, a causal, or at least ancillary, factor in the pathology of many
diseases (Sies (1985, 1997)).
Oxidative Balance
Antioxidants
Antioxidants
Oxidants
Foods
Foods
Vitamins-H
Vitamins-H22O
O Sol.
Sol.
Fat
Fat Soluble
Soluble Vit.
Vit.
Dietary
Dietary Sup.
Sup.
Small
Small Molecules
Molecules
Enzymes
Enzymes
Smoking
Cell Activity
Pollutants
Radiation
UV Light
Cell
Damage
Activation
State
of
Oxidative
Stress
Cell
Repair
Deactivation
Cellular Injury
Figure 1.6. Oxidative Balance Between Pro-Oxidant And Antioxidant
Species. Normally The Production Of Oxidants Is Matched By
Antioxidant Defenses. Under Some Circumstances Oxidant
Production Can Overwhelm These Defenses Resulting In Oxidative
Stress, Cellular Damage And Disease.
A continuously growing list of diseases and conditions, especially those involving
inflammation, are reported to be associated with oxidative stress (Table 1.8). It is
interesting to note that a number of these diseases are being treated by
manipulation of antioxidant levels or by the use of drugs with antioxidant activity
(Sies (1991)).
WWW.ESAINC.COM
19
Disease/condition
Abetalipoproteinaemia
Active pulmonary sarcoidosis
Adult respiratory distress syndrome
AIDS/HIV
Aging
Alcohol related diseases
Alzheimer’s disease
Amyotrophic lateral sclerosis
Apoptosis
Arthritis
Asbestosis
Asthma
Atherosclerosis
Autoimmune diseases (general)
Autoimmune vasculitis
Batten’s disease
Behcet's disease
Bloom’s syndrome
Bone disease (general)
Bronchopulmonary dysplasia
Cancer
Cardiovascular disease
Cataracts
Chediak-Higashi syndrome
Chronic granulomatous disease
Crohn’s disease
Reference
Mimo (1992)
Calhoun et al. (1988)
Ballmer et al. (1994); Choi and Alam (1996)
Baruchel and Wainberg (1992); Dobmeyer et al.
(1997); McLemore et al. (1998); Revillard (1991)
Balin and Allen (1986); Beckman and Ames (1998);
Benzi and Moretti (1995); Bohr and Anson (1995);
Cutler (1991); Harman (1988, 1992a,b); Hocman
(1981); Knight (1995); Lebel and Bondy (1992);
Leibovitz and Siegel (1980); Nohl (1993); Papa and
Skulachev (1997); Scarfiotti et al. (1997); Simic (1992);
Sohal (1993); Suzuki (1993)
Goebel and Schneider (1981); Guemouri et al. (1993);
Lieber (1997); Thome et al. (1997); Zhao et al. (1996)
Beal (1997); Behl et al. (1994); Choi (1995);
Markesbery (1997); McIntosh et al. (1997); Swerdlow
et al. (1997); Volicer and Crino (1990)
Beal (1997); Chou (1997); Migheli et al. (1994)
Monti et al. (1992); Samali et al. (1996); Slater et al.
(1995); Stoian et al. (1996)
Biemond et al. (1988); Greenwald (1991); Kaur et al.
(1996); Moulton (1996); Schiller et al. (1996);
Stichtenoth and Frolich (1998)
Kamp and Weitzman (1997); Kamp et al. (1992); Lenz
et al. (1996); Rom et al. (1987)
Hilterman et al. (1997); Smith et al. (1997)
Bankson et al. (1993); Devaraj and Jialal (1996);
Gambhir and Gambhir (1997); Napoli (1997)
Bashir et al. (1993); Yoshida and Gershwin (1993)
Bashir et al. (1993); Belch et al. (1989); Bruce et al.
(1997)
Clausen et al. (1988); Garg et al. (1982)
Ohno et al. (1997); Pronai et al. (1990)
Emerit and Cerutti (1981)
Ralston (1997)
Banks et al. (1998)
Bankson et al. (1993); Borek (1993); De Flora et al.
(1991); Emerit (1994); Hochstein and Atallah Klotz
(1998); Hocman (1981); Oberley and Buettner (1979);
Oberley and Oberley (1997); Ockner et al. (1993);
Palmer and Paulson (1997); Pryor (1997); Slaga
(1995); Troll (1991); Trush and Kensler (1991);
Weinberg (1996)
De Meyer and Herman (1997); Marin and RodriguezMartinez (1997); Welch and Loscalzo (1994)
Bhuyan et al. (1986); Niwa and Iizawa (1994); Varma
et al. (1984, 1995); Walsh and Patterson (1991); Zigler
and Hess (1985)
Falloon and Gallin (1986); Quie (1997); Volkman et al.
(1984)
Umeki (1994); Volkman et al. (1984)
Allgayer (1991); Baldassano et al. (1993); Curran et al.
(1991); Kimura et al. (1997); McKenzie et al. (1996);
Rachmilewitz et al. (1997); Solis-Herruzo et al. (1993)
WWW.ESAINC.COM
20
Cystic fibrosis
Diabetes
Down’s syndrome
Duchenne’s muscular dystrophy
Exercise
Favism
Friedreich’s ataxia
Gastritis
Gerstmann-Straussler Syndrome
Glomerular injury
Gout
Guillain Barre syndrome
Hashimoto’s thyroiditis
Hemolytic diseases
Hepatitis
Huntington’s disease
Hutchinson-Gilford syndrome
Hypercholesterolaemia
Hypersensitivity pneumonitis
Idiopathic hemochromatosis
Inborn errors of metabolism
Infectious mononucleosis
Inflammation (general)
Inflammatory bowel disease
Brown et al. (1994, 1995, 1996); Graseman et al.
(1998); Percival et al. (1995); Portal et al. (1995);
Winklhofer-Roob (1994); Worlitzsch et al. (1998)
Dandona et al. (1997); Giugliano et al. (1995);
Semenkovich and Heinecke (1997); Wolff et al. (1991)
Brugge et al. (1992); Kedziora and Bartosz (1988); Lott
(1982); Reiter et al. (1996)
Burr et al. (1987); Dioszeghy et al. (1989); Haycock et
al. (1996); Ragusa et al. (1997)
Fielding and Meydani (1997); Higuchi et al. (1985); Ji
(1996); Lawson et al. (1997); Leeuwenburgh et al.
(1994); Ortenblad et al. (1997); Packer (1997)
Gaetani et al. (1996); Mavelli et al. (1984); Musci et al.
(1987); Winterbourn et al. (1986)
Rotig et al. (1997)
Beno et al. (1993, 1994); Durak et al. (1994); Mannick
et al. (1996)
Migheli et al. (1994)
Rohrmoser and Mayer (1996)
Marcolongo et al. (1988); Rosen et al. (1986)
Gutowski et al. (1998)
Bagchi et al. (1990); Sugawara et al. (1988); Szabo et
al. (1996)
Fritsma (1983); Lachant and Tanaka (1986); Stack et
al. (1989); Stocks et al. (1971); Vertongen et al. (1981);
Winterbourn (1990); Yenchitsomanus and Wasi (1983)
Arthur et al. (1985); Biasi et al. (1994); Biemond et al.
(1988); Bonkovsky et al. (1997); De Maria et al. (1996);
Yu et al. (1997)
Beal (1995, 1996, 1997); Bondy (1995); Borlongan et
al. (1996); Browne et al. (1997); Shapira (1996)
Goldstein (1971)
Cohen (1995); Devaraj and Jialal (1994); Harrison and
Ohara (1995); Verhaar et al. (1998); Wennmalm (1994)
Calhoun (1991)
Britton and Brown (1985); Gutteridge et al. (1985);
Houglum et al. (1997); Selden et al. (1980); Young et
al. (1994)
Bird et al. (1995); Blau et al. (1996); Brown and Squier
(1996); Delgado and Calderon (1979); Jansen and
Wanders (1997); Kavanagh et al. (1994); Loscalzo
(1996); Moyano et al. (1997); Patel and Leonard
(1995); Pitkanen and Robinson (1996); Prohaska
(1986); Quie (1977); Welch et al. (1997); Whitin and
Cohen (1988); Yoshida et al. (1995)
Hokama et al. (1986); Niwa et al. (1984); Ritter et al.
(1994)
Billiar (1995); Chapple (1997); Cirino (1998); Connor
and Grisham (1996); Dallegri and Ottonello (1997);
Halliwell et al. (1988); Morris et al. (1995); Parke and
Parke (1996); Pyne (1994); Southorn and Powis
(1988); Stichtenoth and Frolich (1998); Trenam et al.
(1992); Weitzman and Gordon (1990); Winrow et al.
(1993); Winyard and Blake (1997)
Buffinton and Doe (1995); Macdonald (1998)
WWW.ESAINC.COM
21
Ischemia/reoxygenation injury;
reperfusion injury
Kashin-Beck disease
Keshan disease
Leprosy
Liver disease (general)
Lupus
Macular degeneration
Malaria
Motor neuron disease
Multiple sclerosis
Neuronal ceroid lipofuscinosis
Pancreatitis
Parkinson’s disease
Periodontal disease
Porphyria
Prion Diseases
Renal dialysis
Retrolental fibroplasia
Rheumatic diseases
Salmonella typhimurium infection
Septic shock
Skin inflammation
Smoking
Stroke
Transplantation
Ar' Rajab et al. (1996); Bulkley (1994); Flaherty and
Weisfeldt (1988); Gutteridge and Halliwell (1990);
Hudson (1994); Johnson and Weinberg (1993);
Maxwell (1997); McCord (1987); Szabo (1996);
Waxman (1996); Weight et al. (1996)
Peng et al. (1992); Wu and Xu (1987)
Hensrud et al. (1994); Levander et al. (1997)
Agnihotri et al. (1996); Sethi et al. (1996); Sharp and
Banerjee (1985)
Abrams et al. (1995)
Belmont et al. (1997); Benke et al. (1990); Cooke et al.
(1997); Mohan and Das (1997); Suryaprabha et al.
(1991)
Anderson et al. (1994); Nicolas et al. (1996); Van der
Hagen et al. (1996)
Delmas-Beauvieux et al. (1995); Ginsburg and Atamna
(1994); Mishra et al. (1994); Postma et al. (1996);
Vennerstrom and Eaton (1988)
Anderson et al. (1997); Donohoe and Brady (1996);
Lyras et al. (1996); Morrison (1995); Sendtner and
Thoenen (1994); Shaw et al. (1995); Wong and
Borchelt (1995); Zeman et al. (1994)
Calabrese et al. (1994); Clausen et al. (1997); Cooper
et al. (1997); Hooper et al. (1998); Langemann et al.
(1992); Nagra et al. (1997); Parkinson et al. (1997)
Garg et al. (1982); Gutteridge et al. (1983); Marklund et
al. (1981); Santavuori et al. (1989)
Sanfey (1986)
Beal (1997); Cadet and Brannock (1998); Ciccone
(1998); Di Momte et al. (1992); Fahn and Cohen
(1992); Gerlach et al. (1994); Hirsch et al. (1997);
Jenner (1996); Jenner and Olanow (1996); Koller
(1997); Owen et al. (1997); Simonian and Coyle
(1996); Youdin et al. (1988, 1990)
Ellis et al. (1998); Kimura et al. (1993); Moore et al.
(1994); Scmidt et al. (1996)
Monteiro et al. (1986, 1989); Thunell et al. (1997)
Brown et al. (1997); Wiseman and Goldfarb (1996)
Biasioli et al. (1997); Cristol et al. (1994);
Westhuyzen et al. (1995)
Anderson et al. (1994); Cunningham (1987); Johnson
et al. (1974); Southorn and Powis (1988)
Miesel et al. (1996)
Mehta et al. (1998)
Brigham (1991); Goode and Webster (1993); Keusch
(1993); Kilbourn et al. (1997); Kuhl and Rosen (1998);
Novelli (1997); Taylor and Piantadosi (1995)
Trenam et al. (1992)
Cantin and Crystal (1985); Chow (1993); Crystal
(1991); Kohlmeier and Hastings (1995); McCusker
(1992); Pryor, W.A. (1997); Rahman and MacNee
(1996)
Chang et al. (1998); Fisher and Bogousslavsky (1998);
Keli et al. (1996); Mattson (1997); Meldrum (1995)
Hernandez and Granger (1988); Keith (1993); Lehr and
WWW.ESAINC.COM
22
Ulcerative colitis
Viral infection
Werner’s syndrome
Wilson’s disease
Xeroderma pigmentosum
Messmer (1996); McCord (1985); Meyer et al. (1998);
Paller (1992); Toledo-Pereyra (1991)
Keshavarzian et al. (1997); Holmes et al. (1998);
Lundberg et al. (1994); McKenzie et al. (1996);
Ramakrishna et al. (1997); Reimund et al. (1998);
Sedghi et al. (1994)
Peterhans (1997)
Marklund et al. (1981)
Britton and Brown (1995); Carmichael et al. (1995);
Ogihara et al. (1995); Sokol et al. (1994)
Crawford et al. (1988); Runger et al. (1995);
Schallreuter et al. (1991)
Table 1.8 Diseases And Conditions Associated With Oxidative Stress.
WHY USE ELECTROCHEMICAL DETECTION?
Oxidation can be defined as a gain in oxygen, a loss of hydrogen, a loss of
protons or the loss of electrons. Conversely, reduction is the loss of oxygen, a
gain of hydrogen, or the gain of electrons. The two processes are complementary
and no oxidation process can take place without a corresponding reduction;
these complimentary reactions are typically referred to as REDuction-OXidation
or REDOX reactions. Of all the different detectors that are used in the study of
redox biochemistry, perhaps the most useful is the electrochemical detector
(ECD). This detector actually measures the flow of electrons (current) when an
electron-rich compound loses electrons to the working electrode’s surface while
this compound undergoes oxidation (conversely, electron-poor compounds can
also be measured as they accept electrons from the working electrode’s surface
while undergoing reduction). When coupled to the high resolution achievable with
high-performance liquid chromatography (HPLC) an analytical instrument is
produced that can be used to measure many different pro-oxidant, antioxidant,
and damaged species (Chapter 2 and 3). Electrochemical detection is one of the
most sensitive and selective detection techniques available for use with HPLC.
The theory behind it has been extensively reviewed elsewhere (Acworth and
Bowers (1997) and references therein; Acworth et al. (1997a,b,c; 1998)). Of all
the ECDs on the market place, ESA’s coulometric detectors are the most
sensitive and selective, and are virtually maintenance free.
ESA, Inc., offers two electrochemical detectors (Figure 1.7). The Coulochem®
detector offers a high-sensitivity DC mode along with pulsed and cyclic
capabilities. The CoulArray® is the only electrochemical detector that can work
with even the most aggressive gradients. Practical examples using these
detectors will be presented throughout this handbook.
WWW.ESAINC.COM
23
Figure 1.7 The Coulochem® III (Upper Figure) And CoulArray® Detectors.
WWW.ESAINC.COM
24
CONCLUSIONS.
Oxygen is toxic and exerts its toxicity through the production of a variety of prooxidant species. During evolution living organisms either remained anaerobic
surviving in oxygen poor conditions or became aerobic, adapting to the increased
atmospheric levels of oxygen. Aerobic organisms tolerate continued production
of pro-oxidants and have evolved mechanisms to repair or remove damaged
molecules or to prevent the formation and to intercept and deactivate the prooxidant species. Normally there is a balance between production of pro-oxidant
species and destruction by the antioxidant defenses. However, under certain
conditions this balance is upset in favor of overproduction of the pro-oxidants
leading to oxidative stress and disease. HPLC-ECD is one of the most sensitive
analytical techniques for the measurement of pro-oxidants, antioxidants, and
damage markers.
REFERENCES.
Abrams, G.A., Trauner, M., and Nathanson, M.H. (1995). Nitric oxide and liver disease. Gastroenterologist, 3, 220-233.
Acworth, I.N., Bailey, B.B., and Maher, T.J. (1998). The use of HPLC with electrochemical detection to monitor reactive
oxygen and nitrogen species, markers of oxidative damage and antioxidants: application to the neurosciences. In:
Neurochemical Markers of Degenerative Nervous Diseases and Drug Addiction. Qureshi, G., Parvez, H., Caudy, P.,
and Parvez, S. (Eds.). Progress in HPLC-HPCE, 7. VSP Publications, The Netherlands. Pp. 3-56.
Acworth, I.N., and Bowers, M. (1997a). An introduction to HPLC-based electrochemical detection: From single electrode
to multi-electrode arrays. In: Coulometric Electrode Array Detectors for HPLC. Progress in HPLC-HPCE 6. Acworth, I.
N., Naoi, M., Parvez, S., and Parvez, H. (Eds.). VSP Publications, The Netherlands. Pp. 1-48.
Acworth, I.N., McCabe, D.R., and Maher, T.J. (1997b). The analysis of free radicals, their reaction products and
antioxidants. In: Oxidants, Antioxidants and Free Radicals. Baskin, S.I., and Salem, H., (Eds.). Taylor and Francis,
Washington DC. Pp. 23-77.
Acworth, I.N., Naoi, M., Parvez, S., and Parvez, H. (Eds.). (1997c). Coulometric Electrode Array Detectors for HPLC.
Progress in HPLC-HPCE, 6. VSP Publications, The Netherlands.
Agnihotri, N. Ganguly, N.K., Kaur, S., Khullar, M., Sharma, S.C., and Chugh, K.S. (1995). Role of reactive oxygen species
in renal damage in experimental leprosy. Lepr. Rev., 66, 201-209.
Allgayer, H. (1991). Clinical relevance of oxygen radicals in inflammatory bowel disease – facts and fashion. Klin.
Wochenschr., 69, 1001-1003.
Andersen, P.M., Nilsson, P., Keranen, M.L., Forsgren, L., Hagglund, J., Karlsborg, M., Ronnevi, L.O., Gredal, O., and
Marklund, S.L. (1997). Phenotypic heterogeneity in motor neuron disease patients with CuZn- superoxide dismutase
mutations in Scandinavia. Brain, 120, 1723-1737.
Anderson, R.E., Kretzer, F.L., and Rapp, L.M. (1994). Free radicals and ocular disease. Adv. Exp. Med. Biol., 366, 73-86.
Ar' Rajab, A., Dawidson, I., and Fabia, R. (1996). Reperfusion injury. New Horiz., 4, 224-234.
Arthur, M.J., Bentley, I.S., Tanner, A.R., Saunders, P.K., Millward-Sadler, G.H., and Wright, R. (1985). Oxygen-derived
free radicals promote hepatic injury in the rat. Gastroenterology, 89, 1114-1122.
Bagchi, N., Brown, T.R., Herdegen, D.M., Dhar, A., and Sundick, R.S. (1990). Antioxidants delay the onset of thyroiditis in
obese strain chickens. Endocrinology, 127, 1590-1595.
Baldassano, R.N., Schreiber, S., Johnston, R.B. Jr., Fu, R.D., Muraki, T., and MacDermott, R.P. (1993). Crohn's disease
monocytes are primed for accentuated release of toxic oxygen metabolites. Gastroenterology, 105, 60-66.
Balin, A.K., and Allen, R.G. (1986). Mechanisms of biologic aging. Dermatol. Clin., 4, 347-358.
Ballmer, P.E., Reinhart, W.H., and Gey, K.F. (1994). Antioxidant vitamins and disease risk of a suboptimal supply. Ther.
Umsch., 51, 467-474.
Banks, B.A., Ischiropoulos, H., McClelland, M., Ballard, P.L., and Ballard, R.A. (1998). Plasma 3-nitrotyrosine is elevated
in premature infants who develop bronchopulmonary dysplasia. Pediatrics, 101,870-874.
Bankson, D.D., Kestin, M., and Rifai, N. (1993). Role of free radicals in cancer and atherosclerosis. Clin. Lab. Med., 13,
463-480.
Baruchel, S., and Wainberg, M.A. (1992). The role of oxidative stress in disease progression in individuals infected by the
human immunodeficiency virus. J. Leukoc. Biol., 52, 111-114.
Bashir, S., Harris, G., Denman, M.A., Blake, D.R., and Winyard, P.G. (1993). Oxidative DNA damage and cellular
sensitivity to oxidative stress in human autoimmune diseases. Ann. Rheum. Dis., 52, 659-666.
Beal, M.F. (1995). Aging, energy, and oxidative stress in neurodegenerative diseases. Ann. Neurol., 38, 357-366.
Beal, M.F. (1996). Mitochondria, free radicals, and neurodegeneration. Curr. Opin. Neurobiol., 6, 661-666.
WWW.ESAINC.COM
25
Beal, M.F. (1997). Oxidative damage in neurodegenerative diseases. Neuroscientist, 3, 21-27.
Beckman, K.B., and Ames, B.N. (1998). The free radical theory of aging matures. Physiol. Rev., 78, 547-581.
Behl, C., Davis, J.B., Lesley, R., and Schubert, D. (1994). Hydrogen peroxide mediates amyloid β protein toxicity. Cell, 77,
817-827.
Belch, J.J., Chopra, M., Hutchison, S., Lorimer, R., Sturrock, R.D., Forbes, C.D., and Smith, W.E. (1989). Free radical
pathology in chronic arterial disease. Free Radic. Biol. Med., 6, 375-378.
Belmont, H.M., Levartovsky, D., Goel, A., Amin, A., Giorno, R., Rediske, J., Skovron, M.L., and Abramson, S.B. (1997).
Increased nitric oxide production accompanied by the up-regulation of inducible nitric oxide synthase in vascular
endothelium from patients with systemic lupus erythematosus. Arthritis Rheum., 40, 1810-1816.
Benke, P.J., Levcovitz, H., Paupe, J., and Tozman, E. (1990). Scavengers of free radical oxygen affect the generation of
low molecular weight DNA in stimulated lymphocytes from patients with systemic lupus erythematosus. Metabolism, 39,
1278-1284.
Beno, I., Volkovova, K., Batovsky, M., and Staruchova, M. (1993). Increased mucosal antioxidant enzyme activities in
chronic gastritis and benign gastric polyps. Eur. J. Cancer Prev., 2, 461-465.
Beno, I., Volkovova, K., Staruchova, M., and Koszeghyova, L. (1994). The activity of Cu/Zn-superoxide dismutase and
catalase of gastric mucosa in chronic gastritis, and the effect of alpha-tocopherol. Bratisl. Lek. Listy., 95, 9-14.
Benzi, G., and Moretti, A. (1995). Are reactive oxygen species involved in Alzheimer’s disease? Neurobiol. Aging, 16,
661-674.
Bhuyan, K.C., Bhuyan, D.K., and Podos, S.M. (1986). Lipid peroxidation in cataract of the human. Life Sci., 38, 14631471.
Biasi, F., Chiarpotto, E., Lanfranco, G., Capra, A., Zummo, U., Chiappino, I., Scavazza, A., Albano, E., and Poli, G.
(1994). Oxidative stress in the development of human ischemic hepatitis during circulatory shock. Free Radic. Biol.
Med., 17, 225-233.
Biasioli, S., Schiavon, R., Petrosino, L., Cavallini, L., Zambello, A., De Fanti, E., and Giavarina, D. (1997). Free radicals
and oxidative stress challenge dialysis patients: effects of two different membranes. ASAIO J., 43, M766-772.
Biemond, P., Swaak, A.J., van Eijk, H.G., and Koster, J.F. Superoxide dependent iron release from ferritin in inflammatory
diseases. Free Radic. Biol. Med., 4, 185-198.
Billiar, T.R. (1995). Nitric oxide. Novel biology with clinical relevance. Ann. Surg., 221, 339-349.
Bird, S., Miller, N.J., Collins, J.E., and Rice-Evans, C.A. (1995). Plasma antioxidant capacity in two cases of tyrosinaemia
type 1: one case treated with NTBC. J. Inherit. Metab. Dis., 18, 123-126.
Blass, J.P., and Gibson, G.E. (1979). Consequences of mild, graded hypoxia. Adv. Neurol., 26, 229-250.
Blau, N. de Klerk, J.B., Thony, B., Heizmann, C.W., Kierat, L., Smeitink, J.A., and Duran, M. (1996). Tetrahydrobiopterin
loading test in xanthine dehydrogenase and molybdenum cofactor deficiencies. Biochem. Mol. Med., 58, 199-203.
Bohr, V.A., and Anson, R.M. (1995). DNA damage, mutation and fine structure DNA repair in aging. Mutat. Res., 338, 2534.
Bolli, R., Patel, B.S., Jeroudi, M.O., Lai, E.K., and McCay, P.B. (1988). Demonstration of free radical generation in
“stunned” myocardium of intact dogs with the use of the spin trap N-tert-phenyl -α-butylnitrone. J. Clin. Invest., 82, 476485.
Bondy, S.C. (1995). The relation of oxidative stress and hyperexcitation to neurological disease. Proc. Soc. Exp. Biol.
Med., 208, 337-345.
Bonkovsky, H.L., Banner, B.F., and Rothman, A.L. (1997). Iron and chronic viral hepatitis [see comments]. Hepatology,
25, 759-768.
Borek, C. (1993). Molecular mechanisms in cancer induction and prevention. Environ. Health Perspect., 101, Suppl 3,
237-245.
Borlongan, C.V., Kanning, K., Poulos, S.G., Freeman, T.B., Cahill, D.W., and Sanberg, P.R. (1996). Free radical damage
and oxidative stress in Huntington's disease. J. Fla. Med. Assoc., 83, 335-341.
Brigham, K.L. (1991). Oxygen radicals--an important mediator of sepsis and septic shock. Klin. Wochenschr., 69, 10041008.
Britton, R.S., and Brown, K.E. (1995). Genetic hemochromatosis and Wilson's disease: role for oxidant stress?
Hepatology, 21, 1195-1197.
Brown, D. R., Schulz-Schaeffer, W.J., Schmidt, B., and Kretzschmar, H.A. (1997). Prion protein-deficient cells show
altered response to oxidative stress due to decreased SOD-1 activity. Exp. Neurol., 146, 104-112.
Brown, G.K., and Squier, M.V. (1996). Neuropathology and pathogenesis of mitochondrial diseases. J. Inherit. Metab.
Dis., 19, 553-572.
Brown, R.K., and Kelly, F.J. (1994). Evidence for increased oxidative damage in patients with cystic fibrosis. Pediatr. Res.,
36, 487-493.
Brown, R.K., McBurney, A., Lunec, J., and Kelly, F.J. (1995). Oxidative damage to DNA in patients with cystic fibrosis.
Free Radic. Biol. Med., 18, 801-806.
Brown, R.K., Wyatt, H., Price, J.F., and Kelly, F.J. (1996). Pulmonary dysfunction in cystic fibrosis is associated with
oxidative stress. Eur. Respir. J., 9, 334-339.
Browne, S.E., Bowling, A.C., MacGarvey, U., Baik, M.J., Berger, S.C., Muqit, M.M., Bird, E.D., and Beal, M.F. (1997).
Oxidative damage and metabolic dysfunction in Huntington's disease: Selective vulnerability of the basal ganglia. Ann.
Neurol., 41, 646-653.
Bruce, I., McNally, J., and Bell, A. (1997). Enhanced monocyte generation of reactive oxygen species in primary systemic
vasculitis. J. Rheumatol., 24, 2364-2370.
Brugge, K.L., Nichols, S., Delis, D., Saitoh, T., and Truaner, D. (1992). The role of alterations in free radical metabolism in
mediating cognitive impairments in Down's syndrome. EXS, 62, 190-198.
Buffinton, G.D., and Doe, W.F. (1995). Altered ascorbic acid status in the mucosa from inflammatory bowel disease
patients. Free Radic. Res., 22, 131-143.
WWW.ESAINC.COM
26
Bulkley, G.B. (1994). Reactive oxygen metabolites and reperfusion injury: aberrant triggering of reticuloendothelial
function [see comments]. Lancet, 344, 934-936.
Burr, I.M., Asayama, K., and Fenichel, G.M. (1987). Superoxide dismutases, glutathione peroxidase, and catalase in
neuromuscular disease. Muscle Nerve, 10, 150-154.
Cadet, J.L., and Brannock, C. (1998). Free radicals and the pathobiology of brain dopamine systems. Neurochem. Int.,
32, 117-131.
Calabrese, V., Raffaele, R., Cosentino, E., and Rizza, V. (1994). Changes in cerebrospinal fluid levels of malondialdehyde
and glutathione reductase activity in multiple sclerosis. Int. J. Clin. Pharmacol. Res., 14, 119-123.
Calhoun, W.J. (1991). Enhanced reactive oxygen species metabolism of air space cells in hypersensitivity pneumonitis. J.
Lab. Clin. Med., 117, 443-452.
Calhoun, W.J., Salisbury, S.M., Chosy, L.W., and Busse, W.W. (1988). Increased alveolar macrophage
chemiluminescence and airspace cell superoxide production in active pulmonary sarcoidosis. J. Lab. Clin. Med., 112,
147-156.
Cantin, A., and Crystal, R.G. (1985). Oxidants, antioxidants and the pathogenesis of emphysema. Eur. J. Respir. Dis.
Suppl., 139, 7-17.
Carmichael, P.L., Hewer, A., Osborne, M.R., Strain, A.J., and Phillips, D.H. (1995). Detection of bulky DNA lesions in the
liver of patients with Wilson's disease and primary haemochromatosis. Mutat. Res., 326, 235-243.
Chang, C.Y., Lai, Y.C., Cheng, T.J., Lau, M.T., and Hu, M.L. (1998). Plasma levels of antioxidant vitamins, selenium, total
sulfhydryl groups and oxidative products in ischemic-stroke patients as compared to matched controls in Taiwan. Free
Radic. Res., 28, 15-24.
Chapple, I.L., (1997). Reactive oxygen species and antioxidants in inflammatory diseases. J. Clin. Periodontol., 24, 287296.
Choi, B.H. (1995). Oxidative stress and Alzheimer’s disease. Neurobiol. Aging, 16, 675-678.
Choi, A.M., and Alam, J. (1996). Heme oxygenase-1: function, regulation, and implication of a novel stress-inducible
protein in oxidant-induced lung injury. Am. J. Respir. Cell Mol. Biol., 15, 9-19.
Chou, S.M. (1997). Neuropathology of amyotrophic lateral sclerosis: new perspectives on an old disease. J. Formos. Med.
Assoc., 96, 488-498.
Chow, C.K. (1993). Cigarette smoking and oxidative damage in the lung. Ann. N. Y. Acad. Sci., 686, 289-298.
Ciccone, C.D. (1998). Free-radical toxicity and antioxidant medications in Parkinson's disease. Phys. Ther., 78, 313-319.
Cirino, G. (1998). Multiple controls in inflammation. Extracellular and intracellular phospholipase A2, inducible and
constitutive cyclooxygenase, and inducible nitric oxide synthase. Biochem. Pharmacol., 55, 105-111.
Clausen, J., Jensen, G.E., and Nielsen, S.A. (1988). Selenium in chronic neurologic diseases. Multiple sclerosis and
Batten's disease. Biol. Trace Elem. Res., 15, 179-203.
Cohen, R.A. (1995). The role of nitric oxide and other endothelium-derived vasoactive substances in vascular disease.
Prog. Cardiovasc. Dis., 38, 105-128.
Commoner, B., Heise, J.J., Lippincott, B.B., Norberg, R.E., Passonneau, J.V., and Townsend, J. (1957). Biological activity
of free radicals. Science, 126, 57-63.
Commoner, B., Townsend, J., and Pake, G.E. (1954). Free radicals in biological materials. Nature, 174, 689-691.
Conner, E.M., and Grisham, M.B. (1996). Inflammation, free radicals, and antioxidants. Nutrition, 12, 274-277.
Cooke, M.S., Mistry, N., Wood, C., Herbert, K.E., and Lunec, J. (1997). Immunogenicity of DNA damaged by reactive
oxygen species – implications for anti-DNA antibodies in lupus. Free Radic. Biol. Med., 22, 151-159.
Cooper, R.L. (1997). Multiple sclerosis: an immune legacy? Med. Hypotheses, 49, 307-311.
Crapo, J.D., Barry, H.A., Foscue, H.A., and Shelburne, J. (1980). Structural and biochemical changes in rat lungs
occurring during exposures to lethal and adaptive doses of oxygen. Am. Rev. Respir. Dis., 122, 123-143.
Crawford, D., Zbinden, I., Moret, R., and Cerutti, P. (1988). Antioxidant enzymes in xeroderma pigmentosum fibroblasts.
Cancer Res., 48, 2132-2134.
Cristol, J.P., Canaud, B., Rabesandratana, H., Gaillard, I., Serre, A., and Mion, C. (1994). Enhancement of reactive
oxygen species production and cell surface markers expression due to haemodialysis. Nephrol. Dial. Transplant., 9,
389-394.
Crystal, R.G. (1991). Oxidants and respiratory tract epithelial injury: pathogenesis and strategies for therapeutic
intervention. Am. J. Med., 91, 39S-44S.
Cunningham, A.S. (1987). Breast-feeding, antioxidants, and retinopathy of prematurity [letter]. Am. J. Obstet. Gynecol.,
156, 1040-1041.
Curran, F.T., Allan, R.N., and Keighley, M.R. (1991). Superoxide production by Crohn's disease neutrophils. Gut, 32, 399402.
Cutler, R.G. (1991). Antioxidants and aging. Am. J. Clin. Nutr., 53, 373S-379S.
Dallegri, F., and Ottonello, L. (1997). Tissue injury in neutrophilic inflammation. Inflamm. Res., 46, 382-391.
Dandona, P., Thusu, K., Cook, S., Snyder, B., Makowski, J., Armstrong, D., and Nicotera, T. (1996). Oxidative damage to
DNA in diabetes mellitus. Lancet, 347, 444-445.
De Flora, S., Izzotti, A., D'Agostini, F., and Cesarone, C.F. (1991). Antioxidant activity and other mechanisms of thiols
involved in chemoprevention of mutation and cancer. Am. J. Med., 91, 122S-130S.
Delgado, W., and Calderon, R. (1979). Acatalasia in two Peruvian siblings. J. Oral Pathol., 8, 358-368.
Delmas-Beauvieux, M.C., Peuchant, E., Dumon, M.F., Receveur, M.C., Le Bras, M., and Clerc, M. (1995). Relationship
between red blood cell antioxidant enzymatic system status and lipoperoxidation during the acute phase of malaria.
Clin. Biochem., 28, 163-169.
De Maria, N., Colantoni, A., Fagiuoli, S., Liu, G.J., Rogers, B.K., Farinati, F., Van Thiel, D.H., and Floyd, R.A. (1996).
Association between reactive oxygen species and disease activity in chronic hepatitis C. Free Radic. Biol. Med., 21,
291-295.
De Meyer, G.R., and Herman, A.G. (1997). Vascular endothelial dysfunction. Prog. Cardiovasc. Dis., 39, 325-342.
Devaraj, S., and Jialal, I. (1996). Oxidized low-density lipoprotein and atherosclerosis. Int. J. Clin. Lab. Res., 26, 178-184.
WWW.ESAINC.COM
27
Di Monte, D.A., Chan, P., and Sandy, M.S. (1992. Glutathione in Parkinson’s disease: A link between oxidative stress and
mitochondrial damage. Ann. Neurol., 32, S111-S115.
Dioszeghy, P., Imre, S., and Mechler, F. (1989). Lipid peroxidation and superoxide dismutase activity in muscle and
erythrocytes in adult muscular dystrophies and neurogenic atrophies. Eur. Arch. Psychiatry Sci., 238, 175-177.
Dobmeyer, T.S., Findhammer, S., Dobmeyer, J.M., Klein, S.A., Raffel, B., Hoelzer, D., Helm, E.B., Kabelitz, D., and
Rossol, R. (1997). Ex vivo induction of apoptosis in lymphocytes is mediated by oxidative stress: role for lymphocyte
loss in HIV infection. Free Radic. Biol. Med., 22, 775-785.
Donohoe, D.J., and Brady, B. (1996). Motor neuron disease: etiology, pathogenesis and treatment--a review. Ir. J. Med.
Sci., 165, 200-209.
Durak, I., Ormeci, N., Akyol, O., Canbolat, O., Kavutcu, M., and Bulbul, M. (1994). Adenosine deaminase, 5'-nucleotidase,
xanthine oxidase, superoxide dismutase, and catalase activities in gastric juices from patients with gastric cancer, ulcer,
and atrophic gastritis. Dig. Dis. Sci., 39, 721-728.
Ellis, S.D., Tucci, M.A., Serio, F.G., and Johnson, R.B. (1998). Factors for progression of periodontal diseases. J. Oral
Pathol. Med., 27, 101-105.
Edamatsu, R., Mori, A., and Packer, L. (1995). The spin trap N-tert-phenyl-α-butylnitrone prolongs the life span of the
senescence-accelerated mouse. Biochem. Biophys. Res. Comms., 211, 847-849.
Emerit, I. (1994). Reactive oxygen species, chromosome mutation, and cancer: Possible role of clastogenic factors in
carcinogenesis. Free Radic. Biol. Med., 16, 99-109.
Emerit, I., and Cerutti, P. (1981). Clastogenic activity from Bloom syndrome fibroblast cultures. Proc. Natl., Acad. Sci.
USA, 78, 1868-1872.
Fahn, S., and Cohen, G. (1992). The oxidant stress hypothesis in Parkinson's disease: evidence supporting it. Ann.
Neurol., 32, 804-812.
Falloon, J., and Gallin, J.I. (1986). Neutrophil granules in health and disease. J. Allergy Clin. Immunol., 77, 653-662.
Fielding, R.A., and Meydani, M. (1997). Exercise, free radical generation, and aging. Aging (Milano), 9, 12-18.
Fisher, M., and Bogousslavsky, J. (1998). Further evolution toward effective therapy for acute ischemic stroke [see
comments]. JAMA, 279, 1298-1303.
Flaherty, J.T., and Weisfeldt, M.L. (1988). Reperfusion injury. Free Radic. Biol. Med., 5, 409-419.
Floyd, R.A. (1990). Role of oxygen free radicals in carcinogenesis and brain ischemia. FASEB, J., 4, 2587-2597.
Folbegrova, J., Zhao, Q., Katsura, K., and Siesjo, B. (1995). N-tert-butyl-α-phenylnitrone improves recovery of brain
energy state in rats following transient focal ischemia. Proc. Natl. Acad. Sci. USA, 92, 5057-5061.
Fracica, P.J., Knapp, M.J., Piantadosi, C.A., Takeda, K., Fulkerson, W.J., Coleman, R.E., Wolfe, W.G., and Crapo, J.D.
(1991). Responses of baboons to prolonged hyperoxia: Physiology and qualitative pathology. J. Appl. Physiol., 71,
2352-2362.
Fritsma, G.A. (1983). Vitamin E and autoxidation. Am. J. Med. Technol., 49, 453.
Gaetani, G.F., Rolfo, M., Arena, S., Mangerini, R., Meloni, G.F., and Ferraris, A.M. (1996). Active involvement of catalase
during hemolytic crises of favism. Blood, 88, 1084-1088.
Gambhir, D.S., and Gambhir, J.K. (1997). Oxidized low density lipoprotein, antioxidants and coronary atherosclerosis
[editorial]. Indian Heart J., 49, 19-22.
Garg, H.S., Awasthi, Y.C., Neff, W.A., Ansari, N.H., Srivastava, S.K. (1982). Studies in neuronal ceroid lipofuscinosis:
enzymes of liver and brain tissues involved in the defense against oxidative damage. J. Neurosci. Res., 7, 305-311.
Gerlach, M., Ben-Scachar, D., Riederer, P., and Youdim, M.B.H. (1994). Altered brain metabolism of iron a cause of
neurodegenerative diseases? J. Neurochem., 63, 793-807.
Gerschman, R. (1981). Historical introduction to the “free radical theory” of oxygen toxicity. In: Oxygen and Living
Processes: An Interdisciplinary Approach. Gilbert, D.L. (Ed). Springer-Verlag, Berlin. Pp. 44-46.
Gerschman, R., Gilbert, D.L., Nye, S.W., Dwyer, P., Fenn, W.O. (1954). Oxygen poisoning and x-irradiation: A
mechanism in common. Science, 119, 623-626.
Gibson, G.E., and Blass, J.P. (1976). Impaired synthesis of acetylcholine in brain accompanying mild hypoxia and
hypoglycemia. J. Neurochem., 27, 37-42.
Gibson, G.E., Pulsinelli, W., Blass, J.P., and Duffy, T.E. (1981). Brain dysfunction in mild to moderate hypoxia. Am. J.
Med., 70, 1247-1254.
Gibson, G.E., Shimada, M., and Blass J.P. (1978). Alterations in acetylcholine synthesis and cyclic nucleotides in mild
cerebral hypoxia. J. Neurochem., 31, 757-760.
Gilbert, D.L. (1999). From the breath of life to reactive oxygen species. In: Reactive Oxygen Species in Biological
Systems. Gilbert, D.L., and Colton, C.A. (Eds.). Kluwer Academic/Plenum Publishers, New York. Pp. 3-31.
Ginsburg, H., and Atamna, H. (1994). The redox status of malaria-infected erythrocytes: an overview with an emphasis on
unresolved problems. Parasite, 1,5-13.
Giugliano, D., Ceriello, A., and Paolisso, G. (1995). Diabetes mellitus, hypertension, and cardiovascular disease: which
role for oxidative stress? Metabolism, 44, 363-368.
Goebel, K.M., and Schneider, J. (1981). Erythrocyte membrane fluidity, lipid peroxidation and lysis in alcoholic liver
disease. Acta Biol. Med. Ger., 40, 571-576.
Goldstein, S. (1971). The biology of aging. N. Engl. J. Med., 285, 1120-1129.
Goode, H.F., and Webster, N.R. (1993). Free radicals and antioxidants in sepsis. Crit. Care Med., 21, 1770-1776.
Graham, J. B., Dudley, R., Agullar, N. M., and Gans, C. (1995). Implications of the late Palaeozoic oxygen pulse for
physiology and evolution. Nature, 375, 117-120.
Grasemann, H., Ioannidis, I., Tomkiewicz, R.P., de Groot, H., Rubin, B.K., and Ratjen, F. (1998). Nitric oxide metabolites
in cystic fibrosis lung disease. Arch. Dis. Child., 78, 49-53.
Greenwald, R.A. (1991). Oxygen radicals, inflammation, and arthritis: Pathophysiological considerations and implications
for treatment. Semin. Arthritis. Rheum., 20, 219-240.
Guemouri, L., Lecomte, E., Herbeth, B., Pirollet, P., Paille, F., Siest, G., and Artur, Y. (1993). Blood activities of
antioxidant enzymes in alcoholics before and after withdrawal. J. Stud. Alcohol, 54, 626-629.
WWW.ESAINC.COM
28
Gutowski, N.J., Pinkham, J.M., Akanmu, D., Chirico, S., and Murphy, R.P. (1998). Free radicals in inflammatory
neurological disease: increased lipid peroxidation and haptoglobin levels in Guillain Barre syndrome. Ir. J. Med. Sci.,
167, 43-46.
Gutteridge, J.M.C., and Halliwell, B. (1990). Reoxygenation injury and antioxidant protection: A tale of two paradoxes.
Arch. Biochem. Biophys., 283, 223-226.
Gutteridge, J.M.C., Rowley, D.A., Griffiths, E., and Halliwell, B. (1985). Low-molecular-weight iron complexes and oxygen
radical reactions in idiopathic haemochromatosis. Clin. Sci., 68, 463-467.
Gutteridge, J.M.C., Westermarck, T., and Santavuori, P. (1983). Iron and oxygen radicals in tissue damage: implications
for the neuronal ceroid lipofuscinoses. Acta Neurol. Scand., 68, 365-370.
Halliwell, B., and Gutteridge, J.M.C. (1999). Free Radicals in Biology and Medicine. Oxford University Press.
Halliwell, B., Hoult, J.R., and Blake, D.R. (1988). Oxidants, inflammation, and anti-inflammatory drugs. FASEB J., 2, 28672873.
Harman, D. (1956). Aging: A theory based on free radical and radiation chemistry. J. Gerontol., 11, 298-304.
Harman, D. (1988). Free radicals in aging. Mol. Cell Biochem., 84, 155-161.
Harman, D. (1992a). Free radical theory of aging: history. EXS, 62, 1-10.
Harman, D. (1992b). Free radical theory of aging. Mutat. Res., 275, 257-266.
Harrison, D.G., and Ohara, Y. (1995). Physiologic consequences of increased vascular oxidant stresses in
hypercholesterolemia and atherosclerosis: Implications for impaired vasomotion [see comments]. Am. J. Cardiol., 75,
75B-81B.
Haugaard, N. (1946). Oxygen poisoning XI. The relation between inactivation of enzymes and essential sulfhydryl groups.
J. Biol. Chem., 164, 265-270.
Haycock, J.W., MacNeil, S., Jones, P., Harris, J.B., and Mantle, D. (1996). Oxidative damage to muscle protein in
Duchenne muscular dystrophy. Neuroreport, 8, 357-361.
Hensrud, D.D., Heimburger, D.C., Chen, J., and Parpia, B. (1994). Antioxidant status, erythrocyte fatty acids, and
mortality from cardiovascular disease and Keshan disease in China. Eur. J. Clin. Nutr., 48, 455-464.
Hernandez, L.A., and Granger, N. (1988). Role of antioxidants in organ preservation and transplantation. Crit. Care Med.,
16, 543-549.
Higuchi, M., Cartier, L.J., Chen, M., and Holloszy, J.O. (1985). Superoxide dismutase and catalase in skeletal muscle:
adaptive response to exercise. J. Gerontol., 40, 281-286.
Hiltermann, T.J., de Bruijne, C.R., Stolk, J., Zwinderman, A.H., Spieksma, F.T., Roemer, W., Steerenberg, P.A., Fischer,
P.H., van Bree, L., and Hiemstra, P.S. (1997). Effects of photochemical air pollution and allergen exposure on upper
respiratory tract inflammation in asthmatics. Am. J. Respir. Crit. Care Med., 156, 1765-1772.
Hirsch, E.C., Faucheux, B., Damier, P., Mouatt-Prigent, A., and Agid, Y. (1997). Neuronal vulnerability in Parkinson's
disease. J. Neural. Transm. Suppl., 50, 79-88.
Hochstein, P., and Atallah, A.S. (1988). The nature of oxidants and antioxidant systems in the inhibition of mutation and
cancer. Mutat. Res., 202, 363-375.
Hocman, G. (1981). Biochemistry of aging and cancer. Int. J. Biochem., 13, 659-672.
Hokama, Y., Dayaon, E., Iwamoto, L., Yanagisawa, R., Reichert, E., Sato, D., and Ching, C.Y. (1986). Significant
enhanced superoxide anion (O2-) production in vitro by peripheral blood monocytes of lepromatous leprosy patients
stimulated with liposome and suppression by C-reactive protein (CRP). J. Med., 17, 299-311.
Holmes, E.W., Yong, S.L., Eiznhamer, D., and Keshavarzian, A. (1998). Glutathione content of colonic mucosa: evidence
for oxidative damage in active ulcerative colitis. Dig. Dis. Sci., 43, 1088-1095.
Hooper, D.C., Spitsin, S., Kean, R.B., Champion, J.M., Dickson, G.M., Chaudhry, I., and Koprowski, H. (1998). Uric acid,
a natural scavenger of peroxynitrite, in experimental allergic encephalomyelitis and multiple sclerosis. Proc. Natl. Acad.
Sci. USA, 95, 675-680.
Houglum, K., Ramm, G.A., Crawford, D.H., Witztum, J.L., Powell, L.W., and Chojkier, M. (1997). Excess iron induces
hepatic oxidative stress and transforming growth factor beta1 in genetic hemochromatosis. Hepatology, 26, 605-610.
Hudson, K.F. (1994). A phenomenon of paradox: myocardial reperfusion injury. Heart Lung, 23, 384-393.
Jansen, G.A., and Wanders, R.J. (1997). Plasmalogens and oxidative stress: evidence against a major role of
plasmalogens in protection against the superoxide anion radical. J. Inherit. Metab. Dis., 20, 85-94.
Jenner, P. (1996). Oxidative stress in Parkinson's disease and other neurodegenerative disorders. Pathol. Biol., 44, 5764.
Jenner, P., and Olanow, C.W. (1996). Oxidative stress and the pathogenesis of Parkinson's disease. Neurology, 47,
S161-170.
Ji, L.L. (1996). Exercise, oxidative stress, and antioxidants. Am. J. Sports Med., 24, S20-24.
Johnson, K.J., and Weinberg, J.M. (1993). Postischemic renal injury due to oxygen radicals. Curr. Opin. Nephrol.
Hypertens., 2, 625-635.
Johnson, L., Schaffer, D., and Boggs, T.R. Jr. (1974). The premature infant, vitamin E deficiency and retrolental
fibroplasia. Am. J. Clin. Nutr., 27, 1158-1173.
Kamp, D.W., and Weitzman, S.A. (1997). Asbestosis: clinical spectrum and pathogenic mechanisms. Proc. Soc. Exp. Biol.
Med., 214, 12-26.
Kamp, D.W., Graceffa, P., Pryor, W.A., and Weitzman, S.A. (1992). The role of free radicals in asbestos-induced
diseases. Free Radic. Biol. Med., 12, 293-315.
Kaur, H., Edmonds, S.E., Blake, D.R., and Halliwell, B. (1996). Hydroxyl radical generation by rheumatoid blood and knee
joint synovial fluid. Ann. Rheum. Dis., 55, 915-920.
Kavanagh, T.J., Raghu, G., White, C.C., Martin, G.M., Rabinovitch, P.S., and Eaton, D.L. (1994). Enhancement of
glutathione content in glutathione synthetase – deficient fibroblasts from a patient with 5-oxoprolinuria via metabolic
cooperation with normal fibroblasts. Exp. Cell Res., 212,69-76.
Kedziora, J., and Bartosz, G. (1988). Down's syndrome: a pathology involving the lack of balance of reactive oxygen
species. Free Radic. Biol. Med., 4, 317-330.
WWW.ESAINC.COM
29
Keith, F. (1993). Oxygen free radicals in cardiac transplantation. J. Card. Surg., 8, 245-248.
Keli, S.O., Hertog, M.G., Feskens, E.J., and Kromhout, D. (1996). Dietary flavonoids, antioxidant vitamins, and incidence
of stroke: the Zutphen study. Arch. Intern. Med., 156, 637-642.
Keshavarzian, A., Haydek, J.M., Jacyno, M., Holmes, E.W., and Harford, F. (1997). Modulatory effects of the colonic
milieu on neutrophil oxidative burst: a possible pathogenic mechanism of ulcerative colitis. J. Lab. Clin. Med., 130, 216225.
Keusch, G.T. (1993). Antioxidants in infection. J. Nutr. Sci. Vitaminol., 39, S23-33.
Kilbourn, R.G., Traber, D.L., and Szabo, C. (1997). Nitric oxide and shock. Dis. Mon., 43, 277-348.
Kimura, H., Miura, S., Shigematsu, T., Ohkubo, N., Tsuzuki, Y., Kurose, I., Higuchi, H., Akiba, Y., Hokari, R., Hirokawa,
M., Serizawa, H., and Ishii, H. (1997). Increased nitric oxide production and inducible nitric oxide synthase activity in
colonic mucosa of patients with active ulcerative colitis and Crohn's disease. Dig. Dis. Sci., 42, 1047-1054.
Kimura, S., Yonemura, T., and Kaya, H. (1993). Increased oxidative product formation by peripheral blood
polymorphonuclear leukocytes in human periodontal diseases. J. Periodontal Res., 28, 197-203.
Klotz, T., Bloch, W., Volberg, C., Engelmann, U., and Addicks, K. (1998). Selective expression of inducible nitric oxide
synthase in human prostate carcinoma. Cancer, 82, 1897-1903.
Knight, J.A. (1995). The process and theories of aging. Ann. Clin. Lab. Sci., 25, 1-12.
Kohlmeier, L., and Hastings, S.B. (1995). Epidemiologic evidence of a role of carotenoids in cardiovascular disease
prevention. Am. J. Clin. Nutr., 62, 1370S-1376S.
Koller, W.C. (1997). Neuroprotective therapy for Parkinson's disease. Exp. Neurol., 144, 24-28.
Kuhl, S.J., and Rosen, H. (1998). Nitric oxide and septic shock. From bench to bedside [clinical conference]. West. J.
Med., 168, 176-181.
Lachant, N.A., and Tanaka, K.R. (1986). Antioxidants in sickle cell disease: the in vitro effects of ascorbic acid. Am. J.
Med. Sci., 292, 3-10.
Langemann, H., Kabiersch, A., and Newcombe, J. (1992). Measurement of low-molecular-weight antioxidants, uric acid,
tyrosine and tryptophan in plaques and white matter from patients with multiple sclerosis. Eur. Neurol., 32, 248-252.
Lawson, D.L., Chen, L., and Mehta, J.L. (1997). Effects of exercise-induced oxidative stress on nitric oxide release and
antioxidant activity. Am. J. Cardiol., 80, 1640-1642.
LeBel, C.P., and Bondy, S.C. (1992). Oxidative damage and cerebral aging. Progress Neurobiol., 38, 601-609.
Leeuwenburgh, C., Fiebig, R., Chandwaney, R., and Ji, L.L. (1994). Aging and exercise training in skeletal muscle:
responses of glutathione and antioxidant enzyme systems. Am. J. Physiol., 267, R439-445.
Lehr, H.A., and Messmer, K. (1996). Rationale for the use of antioxidant vitamins in clinical organ transplantation.
Transplantation, 62, 1197-1199.
Leibovitz, B.E., and Siegel, B.V. (1980). Aspects of free radical reactions in biological systems: Aging. J. Gerontol., 35,
45-56.
Lenz, A.G., Costabel, U., and Maier, K.L. (1996). Oxidized BAL fluid proteins in patients with interstitial lung diseases.
Eur. Respir. J., 9, 307-312.
Levander, O.A., and Beck, M.A. (1997). Interacting nutritional and infectious etiologies of Keshan disease. Insights from
coxsackie virus B-induced myocarditis in mice deficient in selenium or vitamin E. Biol. Trace Elem. Res., 56, 5-21.
Lieber, C.S. (1997). Role of oxidative stress and antioxidant therapy in alcoholic and nonalcoholic liver diseases. Adv.
Pharmacol., 38, 601-628.
Loscalzo, J. (1996).The oxidant stress of hyperhomocyst(e)inemia. J. Clin. Invest., 98, 5-7.
Lott, I.T. (1982). Down's syndrome, aging, and Alzheimer's disease: a clinical review. Ann. N. Y. Acad. Sci., 396, 15-27.
Lundberg, J.O., Hellstrom, P.M., Lundberg, J.M., and Alving, K. (1994). Greatly increased luminal nitric oxide in ulcerative
colitis [see comments]. Lancet, 344, 1673-1674.
Lyras, L., Evans, P.J., Shaw, P.J., Ince, P.G., and Halliwell, B. (1996). Oxidative damage and motor neuron disease
difficulties in the measurement of protein carbonyls in human brain tissue. Free Radic. Res., 24, 397-406.
MacDonald, T.T. (1998). Viral vectors expressing immunoregulatory cytokines to treat inflammatory bowel disease. Gut,
42, 460-461.
Mannick, E.E., Bravo, L.E., Zarama, G., Realpe, J.L., Zhang, X.J., Ruiz, B., Fontham, E.T., Mera, R., Miller, M.J., and
Correa, P. (1996). Inducible nitric oxide synthase, nitrotyrosine, and apoptosis in Helicobacter pylori gastritis: effect of
antibiotics and antioxidants. Cancer Res., 56, 3238-3243.
Marcolongo, R., Calabria, A.A., Lalumera, M., Gerli, R., Alessandrini, C., and Cavallo, G. (1988). The "switch-off"
mechanism of spontaneous resolution of acute gout attack. J. Rheumatol., 15, 101-109.
Marin, J., and Rodriguez-Martinez, M.A. (1997). Role of vascular nitric oxide in physiological and pathological conditions.
Pharmacol. Ther., 75, 111-134.
Markesbery, W.R. (1997). Oxidative stress hypothesis in Alzheimer's disease. Free Radic. Biol. Med., 23, 134-147.
Marklund, S., Nordensson, I., and Back, O. (1981). Normal CuZn superoxide dismutase, Mn superoxide dismutase,
catalase and glutathione peroxidase in Werner's syndrome. J. Gerontol., 36, 405-409.
Marklund, S.L., Santavuori, P., and Westermarck, T. (1981). Superoxide dismutase, catalase and glutathione peroxidase
in infantile, late infantile and juvenile neuronal ceroid- lipofuscinosis. Clin. Chim. Acta, 116, 191-198.
Mattson, M.P. (1997). Neuroprotective signal transduction: relevance to stroke. Neurosci. Biobehav. Rev., 21, 193-206.
Mavelli, I., Ciriolo, M.R., Rossi, L., Meloni, T., Forteleoni, G., De Flora, A., Benatti, U., Morelli, A., and Rotilio G. (1984).
Favism: A hemolytic disease associated with increased superoxide dismutase and decreased glutathione peroxidase
activities in red blood cells. Eur. J. Biochem., 139, 13-18.
Maxwell, S.R., and Lip, G.Y. (1997). Reperfusion injury: A review of the pathophysiology, clinical manifestations and
therapeutic options. Int. J. Cardiol., 58, 95-117.
McCord, J.M. (1987). Oxygen-derived radicals: a link between reperfusion injury and inflammation. Fed. Proc., 46, 24022406.
McCord, J.M., and Fridovich, I. (1969). Superoxide dismutase. An enzymic function for erythrocuprein (hemocuprein). J.
Biol. Chem., 244, 6049-6053.
WWW.ESAINC.COM
30
McIntosh, L.J., Trush, M.A., and Troncoso, J.C. (1997). Increased susceptibility of Alzheimer's disease temporal cortex to
oxygen free radical-mediated processes. Free Radic. Biol. Med., 23, 183-190.
McKenzie, S.J., Baker, M.S., Buffinton, G.D., and Doe, W.F. (1996). Evidence of oxidant-induced injury to epithelial cells
during inflammatory bowel disease. J. Clin. Invest., 98, 136-141.
McLemore, J.L., Beeley, P., Thorton, K., Morrisroe, K., Blackwell, W., Dasgupta, A. (1998). Rapid automated
determination of lipid hydroperoxide concentrations and total antioxidant status of serum samples from patients infected
with HIV: elevated lipid hydroperoxide concentrations and depleted total antioxidant capacity of serum samples. Am. J.
Clin. Pathol., 109, 268-273.
McCord, J.M. (1985). Oxygen-derived free radicals in postischemic tissue injury. N. Engl. J. Med., 312, 159-163.
McCusker, K.(1992). Mechanisms of respiratory tissue injury from cigarette smoking. Am. J. Med., 93, 18S-21S.
McKenzie, S.J., Baker, M.S., Buffinton, G.D., and Doe, W.F. (1996). Evidence of oxidant-induced injury to epithelial cells
during inflammatory bowel disease. J. Clin. Invest., 98, 136-141.
Mehta, A., Singh, S., and Ganguly, N.K. (1998). Impairment of intestinal mucosal antioxidant defense system during
Salmonella typhimurium infection. Dig. Dis. Sci., 43, 646-651.
Meldrum, B.S. (1995). Cytoprotective therapies in stroke. Curr. Opin. Neurol., 8, 15-23.
Meyer, K.C., Love, R.B., and Zimmerman, J.J. (1998). The therapeutic potential of nitric oxide in lung transplantation.
Chest, 113, 1360-1371.
Miesel, R., Hartung, R., and Kroeger, H. (1996). Priming of NADPH oxidase by tumor necrosis factor alpha in patients
with inflammatory and autoimmune rheumatic diseases. Inflammation, 20, 427-438.
Migheli, A., Cavalla, P., Piva, R., Giordana, M.T., and Schiffer, D. (1994). Bcl-2 protein expression in aged brain and
neurodegenerative diseases. Neuroreport, 5, 1906-1908.
Mino, M. (1992). Clinical uses and abuses of vitamin E in children. Proc. Soc. Exp. Biol. Med., 200, 266-270.
Mishra, N.C., Kabilan, L., and Sharma, A. (1994). Oxidative stress and malaria-infected erythrocytes. Indian J. Malariol.,
31, 77-87.
Miyajima, T., and Kotake, Y. (1997). Optimal time and dosage of N-tert-phenyl-α-butylnitrone (PBN) for the inhibition of
nitric oxide synthase induction in mice. Free Radic. Biol. Med., 22, 463-470.
Mohan, I.K., and Das, U.N. (1997). Oxidant stress, anti-oxidants and essential fatty acids in systemic lupus
erythematosus. Prostaglandins Leukot. Essent. Fatty Acids, 56, 193-198.
Monteiro, H.P., Abdalla, D.S., Augusto, O., and Bechara, E.J. (1989). Free radical generation during delta-aminolevulinic
acid autoxidation: Induction by hemoglobin and connections with porphyrinpathies. Arch. Biochem. Biophys., 271, 206216.
Monteiro, H.P., Abdalla, D.S., Faljoni-Alario, A., and Bechara, E.J. (1986). Generation of active oxygen species during
coupled autoxidation of oxyhemoglobin and delta-aminolevulinic acid. Biochim. Biophys. Acta., 881, 100-106.
Monti, D., Troiano, L., Tropea, F., Grassilli, E., Cossarizza, A., Barozzi, D., Pelloni, M.C., Tamassia, M.G., Bellomo, G.,
and Franceschi, C. (1992). Apoptosis--programmed cell death: A role in the aging process? Am. J. Clin. Nutr., 55,
1208S-1214S.
Moore, S., Calder, K.A., Miller, N.J., and Rice-Evans, C.A. (1994). Antioxidant activity of saliva and periodontal disease.
Free Radic. Res., 21, 417-425.
Morris, C.J., Earl, J.R., Trenam, C.W., and Blake, D.R. (1995). Reactive oxygen species and iron--a dangerous
partnership in inflammation. Int. J. Biochem. Cell Biol., 27, 109-122.
Morrison, K.E. (1995). Mechanisms in motor neuron disease: Clues from genetic studies. Mol. Med. Today, 1, 195-201.
Moulton, P.J. (1996). Inflammatory joint disease: the role of cytokines, cyclooxygenases and reactive oxygen species. Br.
J. Biomed. Sci., 53, 317-324.
Moyano, D., Vilaseca, M.A., Pineda, M., Campistol. J., Vernet, A., Poo, P., Artuch, R., and Sierra, C. (1997). Tocopherol
in inborn errors of intermediary metabolism. Clin. Chim. Acta, 263, 147-155.
Musci, G., Mavelli, I., and Rotilio, G. (1987). Evidence for superoxide generation from the autoxidation of the favisminducing aglycone divicine. Biochim. Biophys. Acta, 926, 369-372.
Nagra, R.M., Becher, B., Tourtellotte, W.W., Antel, J.P., Gold, D., Paladino, T., Smith, R.A., Nelson, J.R., and Reynolds,
W.F. (1997). Immunohistochemical and genetic evidence of myeloperoxidase involvement in multiple sclerosis. J.
Neuroimmunol., 78, 97-107.
Napoli, C. (1997). Low density lipoprotein oxidation and atherogenesis: From experimental models to clinical studies. G.
Ital. Cardiol., 27, 1302-1314.
Nicolas, M.G., Fujiki, K., Murayama, K., Suzuki, M.T., Shindo, N., Hotta, Y., Iwata, F., Fujimura, T., Yoshikawa, Y., Cho,
F., and Kanai, A. (1996). Studies on the mechanism of early onset macular degeneration in cynomolgus monkeys. II.
Suppression of metallothionein synthesis in the retina in oxidative stress. Exp. Eye Res., 62, 399-408.
Niwa, Y., and Iizawa, O. (1994). Abnormalities in serum lipids and leukocyte superoxide dismutase and associated
cataract formation in patients with atopic dermatitis. Arch. Dermatol., 130, 1387-1392.
Niwa, Y., Sakane, T., Miyachi, Y., Kanoh, T., and Somiya, K. (1984). Decrease in generation of reactive oxygen species
by neutrophils from patients with infectious mononucleosis: role of suppressor T lymphocytes. Blood, 64, 994-999.
Nohl, H. (1993). Involvement of free radicals in AGEing: A consequence or cause of senescence. Br. Med. Bull., 49, 653667.
Novelli, G.P. (1997). Role of free radicals in septic shock. J. Physiol. Pharmacol., 48, 517-527.
Oberley, L.W., and Buettner, G.R. (1979). Role of superoxide dismutase in cancer: A review. Cancer Res., 39, 11411149.
Oberley, T.D., and Oberley, L.W. (1997). Antioxidant enzyme levels in cancer. Histol. Histopathol., 12, 525-535.
Ockner, R.K., Kaikaus, R.M., and Bass, N.M. (1993). Fatty-acid metabolism and the pathogenesis of hepatocellular
carcinoma: review and hypothesis. Hepatology, 18, 669-676.
Ogihara, H., Ogihara, T., Miki, M., Yasuda, H., and Mino, M. (1995). Plasma copper and antioxidant status in Wilson's
disease. Pediatr. Res., 37, 219-226.
WWW.ESAINC.COM
31
Ohno, E., Ohtsuka, E., Watanabe, K., Kohno, T., Takeoka, K., Saburi, Y., Kikuchi, H., and Nasu, M. (1997). Behcet's
disease associated with myelodysplastic syndromes. A case report and a review of the literature. Cancer, 79, 262-268.
Ortenblad, N., Madsen, K., and Djurhuus, M.S. (1997). Antioxidant status and lipid peroxidation after short-term maximal
exercise in trained and untrained humans. Am. J. Physiol., 272, R1258-1263.
Owen, A.D., Schapira, A.H., Jenner, P., and Marsden, C.D. (1997). Indices of oxidative stress in Parkinson's disease,
Alzheimer's disease and dementia with Lewy bodies. J. Neural. Transm. Suppl., 51, 167-173.
Packer, L. (1997). Oxidants, antioxidant nutrients and the athlete. J. Sports Sci., 15, 353-363.
Paller, M.S. (1992). Free radical-mediated postischemic injury in renal transplantation. Ren. Fail., 14, 257-260.
Palmer, H.J., and Paulson, K.E. (1997). Reactive oxygen species and antioxidants in signal transduction and gene
expression. Nutr. Rev., 55, 353-361.
Papa, S., and Skulachev, V.P. (1997). Reactive oxygen species, mitochondria, apoptosis and aging. Mol. Cell Biochem.,
174, 305-319.
Parke, D.V., and Parke, A.L. (1996). Chemical-induced inflammation and inflammatory diseases. Int. J. Occup. Med.
Environ. Health, 9, 211-217.
Parkinson, J.F., Mitrovic, B., and Merrill, J.E. (1997). The role of nitric oxide in multiple sclerosis [see comments]. J. Mol.
Med., 75, 174-186.
Patel, J.S., and Leonard, J.V. (1995). Ketonuria and medium-chain acyl-CoA dehydrogenase deficiency. J. Inherit. Metab.
Dis., 18, 98-99.
Peng, A., Yang, C., Rui, H., and Li, H. (1992). Study on the pathogenic factors of Kashin-Beck disease. J. Toxicol.
Environ. Health, 35, 79-90.
Percival, S.S., Bowser, E., and Wagner, M. (1995). Reduced copper enzyme activities in blood cells of children with cystic
fibrosis. Am. J. Clin. Nutr., 62, 633-638.
Peterhans, E. (1997). Oxidants and antioxidants in viral diseases: disease mechanisms and metabolic regulation. J. Nutr.,
127, 962S-965S.
Pitkanen, S., and Robinson, B.H. (1996). Mitochondrial complex I deficiency leads to increased production of superoxide
radicals and induction of superoxide dismutase. J. Clin. Invest., 98, 345-351.
Portal, B.C., Richard, M.J., Faure, H.S., Hadjian, A.J., and Favier, A.E. (1995). Altered antioxidant status and increased
lipid peroxidation in children with cystic fibrosis [see comments]. Am. J. Clin. Nutr., 61, 843-847.
Postma, N.S., Mommers, E.C., Eling, W.M., and Zuidema, J. (1996). Oxidative stress in malaria; implications for
prevention and therapy. Pharm. World Sci., 18, 121-129.
Prohaska, J.R. (1986). Genetic diseases of copper metabolism. Clin. Physiol. Biochem., 4, 87-93.
Pronai, L., Ichikawa, Y., Nakazawa, H., and Arimori, S. (1990). Superoxide scavenging activity of leukocytes in
rheumatoid arthritis and Behcet's diseases. Tokai J. Exp. Clin. Med., 15, 93-97.
Pryor, W.A. (1997). Cigarette smoke radicals and the role of free radicals in chemical carcinogenicity. Environ. Health
Perspect., 105 Suppl 4, 875-882.
Pyne, D.B. (1994). Exercise-induced muscle damage and inflammation: A review. Aust. J. Sci. Med. Sport, 26, 49-58.
Quie, P.G. (1977). Disorders of phagocyte function: Biochemical aspects. Prog. Clin. Biol. Res., 13, 157-169.
Rachmilewitz, D., Okon, E., and Karmeli, F. (1997). Sulfhydryl blocker induced small intestinal inflammation in rats: a new
model mimicking Crohn's disease. Gut, 41, 358-365.
Ragusa, R.J., Chow, C.K., and Porter, J.D. (1997). Oxidative stress as a potential pathogenic mechanism in an animal
model of Duchenne muscular dystrophy. Neuromuscul. Disord., 7, 379-386.
Rahman, I., and MacNee, W. (1996). Role of oxidants/antioxidants in smoking-induced lung diseases. Free Radic. Biol.
Med., 21, 669-681.
Ralston, S.H. (1997). The Michael Mason Prize Essay 1997. Nitric oxide and bone: what a gas! Br. J. Rheumatol., 36,
831-838.
Ramakrishna, B.S., Varghese, R., Jayakumar, S., Mathan, M., and Balasubramanian, K.A. (1997). Circulating
antioxidants in ulcerative colitis and their relationship to disease severity and activity. J. Gastroenterol. Hepatol., 12,
490-494.
Reimund, J.M., Allison, A.C., Muller, C.D., Dumont, S., Kenney, J.S., Baumann, R., Duclos, B., and Poindron, P. (1998).
Antioxidants inhibit the in vitro production of inflammatory cytokines in Crohn's disease and ulcerative colitis. Eur. J.
Clin. Invest., 28, 145-150.
Reiter, R.J., Barlow-Walden, L., Poeggeler, B., Heiden, S.M., and Clayton, R.J. (1996). Twenty-four hour urinary excretion
of 6-hydroxymelatonin sulfate in Down syndrome subjects. J. Pineal Res., 20, 45-50.
Revillard, J.P. (1991). Oxidative stress: A marker of progression and a therapeutic target in HIV infection. Int. Conf. AIDS,
7, 73 (abstract no. TU.A.66).
Ritter, K., Kuhl, R.J., Semrau, F., Eiffert, H., Kratzin, H.D., and Thomssen, R. (1994). Manganese superoxide dismutase
as a target of autoantibodies in acute Epstein-Barr virus infection. J. Exp. Med., 180, 1995-1998.
Rohrmoser, M.M., and Mayer, G. (1996). Reactive oxygen species and glomerular injury. Kidney Blood Press. Res., 19,
263-269.
Rom, W.N., Bitterman, P.B., Rennard, S.I., Cantin, A., and Crystal, R.G. (1987). Characterization of the lower respiratory
tract inflammation of nonsmoking individuals with interstitial lung disease associated with chronic inhalation of inorganic
dusts. Am. Rev. Respir. Dis., 136, 1429-1434.
Rosen, M.S., Baker, D.G., Schumacher, H.R. Jr., and Cherian, P.V. (1986). Products of polymorphonuclear cell injury
inhibit IgG enhancement of monosodium urate-induced superoxide production. Arthritis Rheum., 29, 1473-1479.
Rotig, A., de Lonlay, P., Chretien, D., Foury, F., Koenig, M., Sidi, D., Munnich, A., and Rustin, P. (1997). Aconitase and
mitochondrial iron-sulfur protein deficiency in Friedreich ataxia. Nat. Genet., 17, 215-217.
Rotilio, G. (1984). Favism: a hemolytic disease associated with increased superoxide dismutase and decreased
glutathione peroxidase activities in red blood cells. Eur. J. Biochem., 139, 13-18.
Runger, T.M., Epe, B., and Moller, K. (1995). Repair of ultraviolet B and singlet oxygen-induced DNA damage in
xeroderma pigmentosum cells. J. Invest. Dermatol., 104, 68-73.
WWW.ESAINC.COM
32
Samali, A., Gorman, A.M., and Cotter, T.G. (1996). Apoptosis – the story so far. Experientia, 52, 933-941.
Sanfey, H., Sarr, M.G., Bulkley, G.B., and Cameron, J.L. (1986). Oxygen-derived free radicals and acute pancreatitis: A
review. Acta Physiol. Scand. Suppl., 548, 109-118.
Santavuori, P., Heiskala, H., Autti, T., Johansson, E., and Westermarck, T. (1989). Comparison of the clinical courses in
patients with juvenile neuronal ceroid lipofuscinosis receiving antioxidant treatment and those without antioxidant
treatment. Adv. Exp. Med. Biol., 266, 273-282.
Scarfiotti, C., Fabris, F., Cestaro, B., and Giuliani, A. (1997). Free radicals, atherosclerosis, aging, and related
dysmetabolic pathologies: Pathological and clinical aspects. Eur. J. Cancer Prev., 6, Suppl 1, S31-36.
Schallreuter, K.U., Pittelkow, M.R., and Wood, J.M. (1991). Defects in antioxidant defense and calcium transport in the
epidermis of xeroderma pigmentosum patients. Arch. Dermatol. Res., 283, 449-455.
Schapira, A.H. (1996). Oxidative stress and mitochondrial dysfunction in neurodegeneration. Curr. Opin. Neurol., 9, 260264.
Schiller, J., Arnhold, J., Sonntag, K., and Arnold, K. (1996). NMR studies on human, pathologically changed synovial
fluids: role of hypochlorous acid. Magn. Reson. Med., 35, 848-853.
Schmidt, A.M., Weidman, E., Lalla, E., Yan, S.D., Hori, O., Cao, R., Brett, J.G., and Lamster, I.B. (1996). Advanced
glycation endproducts (AGEs) induce oxidant stress in the gingiva: A potential mechanism underlying accelerated
periodontal disease associated with diabetes. J. Periodontal Res., 31, 508-515.
Schoenberg, M.H., and Beger, H.G., (1993). Reperfusion injury after intestinal ischemia. Crit. Care Med., 21, 1376-1386.
Sedghi, S., Keshavarzian, A., Klamut, M., Eiznhamer, D., and Zarling, E.J. (1994). Elevated breath ethane levels in active
ulcerative colitis: evidence for excessive lipid peroxidation. Am. J. Gastroenterol., 89, 2217-2221.
Selden, C., Seymour, C.A., and Peters, T.J. (1980). Activities of some free-radical scavenging enzymes and glutathione
concentrations in human and rat liver and their relationship to the pathogenesis of tissue damage in iron overload. Clin.
Sci., 58, 211-219.
Semenkovich, C.F., and Heinecke, J.W. (1997). The mystery of diabetes and atherosclerosis: time for a new plot.
Diabetes, 46, 327-334.
Sendtner, M., and Thoenen, H. (1994). Neurodegenerative disease. Oxidative stress and motorneuron disease. Curr.
Biol., 4, 1036-1039.
Sethi, N.C., Madadi, A.J., and Bhandari, S. (1996). Serum zinc, copper, magnesium, proteins and superoxide dismutase
in leprosy patients on multidrug therapy – a follow-up study. Indian J. Lepr., 68, 325-333.
Sharp, A.K., and Banerjee, D.K. (1985). Hydrogen peroxide and superoxide production by peripheral blood monocytes in
leprosy. Clin. Exp. Immunol., 60, 203-206.
Shaw, P.J., Ince, P.G., Falkous, G., and Mantle, D. (1995). Oxidative damage to protein in sporadic motor neuron disease
spinal cord. Ann. Neurol., 38, 691-695.
Sies, H. (1985). Oxidative stress: Introductory remarks. In: Oxidative Stress. Sies, H. (Ed.). Academic Press, London.
Sies, H. (1991). Oxidative stress: From basic research to clinical applications. Am. J. Med., 91, 31S-38S.
Sies, H. (1997). Oxidative stress: Oxidants and antioxidants. Exp. Physiol., 82, 291-295.
Simic, M.G. (1992). The rate of DNA damage and aging. EXS, 62, 20-30.
Simonian, N.A., and Coyle, J.T. (1996). Oxidative stress in neurodegenerative diseases. Annu. Rev. Pharmacol. Toxicol.,
36, 83-106.
Slaga, T.J. (1995). Inhibition of the induction of cancer by antioxidants. Adv. Exp. Med. Biol., 369, 167-174.
Slater, A.F., Nobel, C.S., and Orrenius, S. (1995). The role of intracellular oxidants in apoptosis. Biochim. Biophys. Acta,
1271, 59-62.
Smith, L.J., Shamsuddin, M., Sporn, P.H., Denenberg, M., and Anderson, J. (1997). Reduced superoxide dismutase in
lung cells of patients with asthma. Free Radic. Biol. Med., 22, 1301-1307.
Sobel, B.E. (1974). Biochemical and morphological changes in infarcting myocardium. In: The Myocardium: Failure and
Infarction. Braunwald, E. (Ed.).. N.P. Publishing Co. Inc., New York. Pp. 247-260.
Sohal, R.S. (1993). The free radical hypothesis of aging: an appraisal of the current status [see comments]. Aging, 5, 317.
Sokol, R.J., Twedt, D., McKim, J.M., Jr. Devereaux, M.W., Karrer, F.M., Kam, I., von Steigman, G., Narkewicz, M.R.,
Bacon,.B.R., and Britton, R.S. Oxidant injury to hepatic mitochondria in patients with Wilson's disease and Bedlington
terriers with copper toxicosis. Gastroenterolog, 107, 1788-1798.
Solis-Herruzo, J.A., Fernandez, B., Vilalta-Castell, E., Munoz-Yague, M.T., Hernandez-Munoz, I., de la Torre-Merino,
M.P., and Balsinde, J. (1993). Diminished cytochrome b content and toxic oxygen metabolite production in circulating
neutrophils from patients with Crohn's disease. Dig. Dis. Sci., 38, 1631-1637.
Southorn, P.A., and Powis, G. (1988). Free radicals in medicine. II. Involvement in human disease. Mayo Clin. Proc., 63,
390-408.
Stadie, W.C., Riggs, B.C., and Haugaard, N. (1944). Oxygen poisoning. Am. J. Med. Sci., 207, 84-114.
Stichtenoth, D.O., and Frolich, J.C. (1998). Nitric oxide and inflammatory joint diseases. Br. J. Rheumatol., 37, 246-257.
Stack, T.A. (1989). Haemolytic anemia due to oxidative stress in the preterm infant [letter]. Early Hum. Dev., 20, 79-80.
Stocks, J., Kemp, M., and Dormandy, T.L. (1971). Increased susceptibility of red-blood-cell lipids to autooxidation in
hemolytic states. Lancet, 1, 266-269.
Stoian, I., Oros, A., and Moldoveanu, E. (1996). Apoptosis and free radicals. Biochem. Mol. Med., 59, 93-97.
Sugawara, M., Kita, T., Lee, E.D., Takamatsu, J., Hagen, G.A., Kuma, K., Medeiros-Neto, G.A. (1988). Deficiency of
superoxide dismutase in endemic goiter tissue. J. Clin. Endocrinol. Metab., 67, 1156-1161.
Suryaprabha, P., Das, U.N., Ramesh, G., Kumar, K.V., and Kumar, G.S. (1991). Reactive oxygen species, lipid peroxides
and essential fatty acids in patients with rheumatoid arthritis and systemic lupus erythematosus. Prostaglandins Leukot.
Essent. Fatty Acids, 43, 251-255.
Suzuki, K. (1993). Origin of senescence: A review. Horm. Res., 39, Suppl 1, 5-8.
WWW.ESAINC.COM
33
Swerdlow, R.H., Parks, J.K., Cassarino, D.S., Maguire, D.J., Maguire, R.S., Bennett, J.P. Jr., Davis, R.E., and Parker,
W.D. Jr. (1997). Cybrids in Alzheimer's disease: A cellular model of the disease? [see comments]. Neurology, 49, 918925.
Szabo, C. (1996). The pathophysiological role of peroxynitrite in shock, inflammation, and ischemia-reperfusion injury.
Shock, 6, 79-88.
Szabo, J., Foris, G., Mezosi, E., Nagy, E.V., Paragh, G., Sztojka, I., and Leovey, A. (1996). Parameters of respiratory
burst and arachidonic acid metabolism in polymorphonuclear granulocytes from patients with various thyroid diseases.
Exp. Clin. Endocrinol. Diabetes, 104, 172-176.
Taylor, D.E., and Piantadosi, C.A. (1995). Oxidative metabolism in sepsis and sepsis syndrome. J. Crit. Care, 10, 122135.
Thome, J., Zhang, J., Davids, E., Foley, P., Weijers, H.G., Wiesbeck, G.A., Boning, J., Riederer, P., and Gerlach, M.
(1997). Evidence for increased oxidative stress in alcohol-dependent patients provided by quantification of in vivo
salicylate hydroxylation products. Alcohol Clin. Exp. Res., 21, 82-85.
Thunell, S., Andersson, D., Harper, P., Henrichson, A., Floderus, Y., and Lindh, U. (1997). Effects of administration of
antioxidants in acute intermittent porphyria. Eur. J. Clin. Chem. Clin. Biochem., 35, 427-433.
Toledo-Pereyra, L.H. (1991). Liver transplantation reperfusion injury. Factors in its development and avenues for
treatment. Klin. Wochenschr., 69, 1099-1104.
Towner, R.A., Janzen, E.G., Zhang, Y.-K., and Yamashiro, S. (1993). MRI study of the inhibitory effects of new spin traps
on in vivo CCl4-induced hepatotoxicity in rats. Free Radic. Biol. Med., 14, 677-681.
Traub, O., and Van Bibber, R. (1995). Role of nitric oxide in insulin-dependent diabetes mellitus-related vascular
complications. West. J. Med., 162, 439-445.
Trenam, C.W., Blake, D.R., and Morris, C.J. (1992). Skin inflammation: reactive oxygen species and the role of iron. J.
Invest. Dermatol., 99, 675-682.
Troll, W. (1991). Prevention of cancer by agents that suppress oxygen radical formation. Free Radic. Res. Commun., 1213 Pt 2, 751-757.
Trush, M.A., and Kensler, T.W. (1991). An overview of the relationship between oxidative stress and chemical
carcinogenesis. Free Radic. Biol. Med., 10, 201-209.
Umeki, S. (1994). Mechanisms for the activation/electron transfer of neutrophil NADPH-oxidase complex and molecular
pathology of chronic granulomatous disease. Ann. Hematol., 68, 267-277.
van der Hagen, A.M., Yolton, D.P., Kaminski, M.S., and Yolton, R.L. (1993). Free radicals and antioxidant
supplementation: a review of their roles in age-related macular degeneration. J. Am. Optom. Assoc., 64, 871-878.
Varma, S.D., Chand, D., Sharma, Y.R., Kuck, J.F. Jr., and Richards, R.D. (1984). Oxidative stress on lens and cataract
formation: role of light and oxygen. Curr. Eye Res., 3, 35-57.
Varma, S.D., Devamanoharan, P.S., and Morris, S.M. (1995). Prevention of cataracts by nutritional and metabolic
antioxidants. Crit. Rev. Food Sci. Nutr., 35, 111-129.
Vennerstrom, J.L., and Eaton, J.W. (1988). Oxidants, oxidant drugs, and malaria. J. Med. Chem., 31, 1269-1277.
Verhaar, M.C., Wever, R.M., Kastelein, J.J., van Dam, T., Koomans, H.A., and Rabelink, T.J. (1998). 5methyltetrahydrofolate, the active form of folic acid, restores endothelial function in familial hypercholesterolemia.
Circulation, 97, 237-241.
Vertongen, F., Heyder-Bruckner, C., Fondu, P., and Mandelbaum, I. (1981). Oxidative hemolysis in protein malnutrition.
Clin. Chim. Acta, 116, 217-222.
Volicer, L., and Crino, P.B. (1990). Involvement of free radicals in dementia of the Alzheimer type: A hypothesis.
Neurobiol. Aging, 11, 567-571.
Volkman, D.J., Buescher, E.S., Gallin, J.I., and Fauci, A.S. (1984). B cell lines as models for inherited phagocytic
diseases: abnormal superoxide generation in chronic granulomatous disease and giant granules in Chediak-Higashi
syndrome. J. Immunol., 133, 3006-3009.
Walsh, S., and Patterson, J.W. (1991). Effects of oxidants on lens transport. Invest. Ophthalmol. Vis. Sci., 32, 1648-1658.
Waxman, K. (1996). Shock: Ischemia, reperfusion, and inflammation. New Horiz., 4, 153-160.
Weight, S.C., Bell, P.R., and Nicholson, M.L. (1996). Renal ischemia – reperfusion injury. Br. J. Surg., 83, 162-170.
Weinberg, E.D. (1996). The role of iron in cancer. Eur. J. Cancer Prev., 5, 19-36.
Weitzman, S.A., and Gordon, L.I. (1990). Inflammation and cancer: role of phagocyte-generated oxidants in
carcinogenesis. Blood, 76, 655-663.
Wennmalm, A. (1994). Nitric oxide (NO) in the cardiovascular system: role in atherosclerosis and hypercholesterolemia.
Blood Press., 3, 279-382.
Welch, G., and Loscalzo, J. (1994). Nitric oxide and the cardiovascular system. J. Card. Surg., 9, 361-71.
Welch, G.N., Upchurch, G.R., and Loscalzo, J. (1997).Homocysteine, oxidative stress, and vascular disease [see
comments]. Hosp. Pract., 32, 81-2 and 85, 88-92.
Westhuyzen, J., Adams, C.E., and Fleming, S.J. (1996). Evidence for oxidative stress during in vitro dialysis. Nephron, 70,
49-54.
Whitin, J.C., and Cohen, H.J. (1988). Disorders of respiratory burst termination. Hematol. Oncol. Clin. North Am., 2, 289299.
Winklhofer-Roob, B.M. (1994). Oxygen free radicals and antioxidants in cystic fibrosis: the concept of an oxidantantioxidant imbalance. Acta Paediatr. Suppl., 83, 49-57.
Winrow, V.R., Winyard, P.G., Morris, C.J., and Blake, D.R. (1993). Free radicals in inflammation: second messengers and
mediators of tissue destruction. Br. Med. Bull., 49, 506-522.
Winterbourn, C.C. (1990). Oxidative denaturation in congenital hemolytic anemias: The unstable hemoglobins. Semin.
Hematol., 27, 41-50.
Winterbourn, C.C., Benatti, U., and De Flora, A. (1986). Contributions of superoxide, hydrogen peroxide, and transition
metal ions to auto-oxidation of the favism-inducing pyrimidine aglycone, divicine, and its reactions with hemoglobin.
Biochem. Pharmacol., 35, 2009-2015.
WWW.ESAINC.COM
34
Winyard, P.G., and Blake, D.R. (1997). Antioxidants, redox-regulated transcription factors, and inflammation. Adv.
Pharmacol., 38, 403-421.
Wiseman, A., and Goldfarb, P. (1996). Is there an optimal membrane protein content determined by membrane stability to
lipid peroxidation? Disease and aging consideration. Biochem. Soc. Trans., 24, 376S.
Wolff, S.P., Jiang, Z.Y., and Hunt, J.V. (1991). Protein glycation and oxidative stress in diabetes mellitus and aging. Free
Radic. Biol. Med., 10, 339-352.
Wong, P.C., and Borchelt, D.R. (1995). Motor neuron disease caused by mutations in superoxide dismutase 1. Curr.
Opin. Neurol., 8, 294-301.
Worlitzsch, D., Herberth, G., Ulrich, M., and Doring, G. (1998). Catalase, myeloperoxidase and hydrogen peroxide in
cystic fibrosis. Eur. Respir. J., 11, 377-383.
Wu, J., and Xu, G.L. (1987). Plasma selenium content, platelet glutathione peroxidase and superoxide dismutase activity
of residents in Kashin-Beck disease affected area in China. J. Trace Elem. Electrolytes Health Dis., 1, 39-43.
Yenchitsomanus, P., and Wasi, P. (1983). Increased erythrocyte superoxide dismutase activities in beta 0thalassaemia/hemoglobin E and in hemoglobin H diseases. J. Clin. Pathol., 36, 329-333.
Young, I.S., Trouton, T.G., Torney, J.J., McMaster, D., Callender, M.E., and Trimble, E.R. (1994). Antioxidant status and
lipid peroxidation in hereditary haemochromatosis. Free Radic. Biol. Med., 16, 393-397.
Yoshida, S., and Gershwin, M.E. (1993). Autoimmunity and selected environmental factors of disease induction. Semin.
Arthritis Rheum., 22, 399-419.
Yoshida, Y. Machigashira, K., Suehara, M., Arimura, H., Moritoyo, T., Nagamatsu, K., and Osame, M. (1995).
Immunological abnormality in patients with lysinuric protein intolerance. J. Neurol. Sci., 134, 178-182.
Youdim, M.B.H. (1988). Iron in the brain: Implications for Parkinson’s and Alzheimer’s diseases. Mount Sinai J. Med., 55,
97-101.
Youdim, M.B., Ben-Shachar, D., and Riederer, P. (1990). The role of monoamine oxidase, iron-melanin interaction, and
intracellular calcium in Parkinson's disease. J. Neural. Transm. Suppl., 32, 239-248.
Yu, M.W., Chiang, Y.C., Lien, J.P., and Chen, C.J. (1997). Plasma antioxidant vitamins, chronic hepatitis B virus infection
and urinary aflatoxin B1-DNA adducts in healthy males. Carcinogenesis, 18, 1189-1194.
Zeman, S., Lloyd, C., Meldrum, B., and Leigh, P.N. (1994). Excitatory amino acids, free radicals and the pathogenesis of
motor neuron disease. Neuropathol. Appl. Neurobiol., 20, 219-231.
Zhao, M., Matter, K., Laissue, J.A., and Zimmermann, A. (1996). Copper/zinc and manganese superoxide dismutases in
alcoholic liver disease: Immunohistochemical quantitation. Histol. Histopathol., 11, 899-907.
Zigler, J.S. Jr., and Hess, H.H. (1985). Cataracts in the Royal College of Surgeons rat: evidence for initiation by lipid
peroxidation products. Exp. Eye Res., 41, 67-76.
WWW.ESAINC.COM
35
Chapter 2
The Chemistry of The Reactive Species
As discussed in Chapter 1, the discovery that oxygen is toxic led to the
immediate consideration of the potential role of free radicals as the damaging
species. Data rapidly revealed that not all oxygen-based noxious compounds
were free radicals. Furthermore, pro-oxidants other than the ROS were
discovered that contained atoms in addition to oxygen. This chapter primarily
reviews the formation, reaction chemistry, and biological significance of the
various important pro-oxidants, including those based on oxygen, nitrogen,
halogens, sulfur and carbonyls. Some of the analytical approaches used to
measure them will also be discussed. The chapter concludes with an overview of
the pro-oxidant activities of a variety of xenobiotics (foreign or man-made
substances) and environmental pollutants. Remember though, that just because
a reaction can be made to occur in a test tube does not mean that such a
reaction is important biologically.
WWW.ESAINC.COM
36
OXYGEN AND THE REACTIVE OXYGEN SPECIES (ROS).
1. Oxygen.
Properties.
Oxygen (dioxygen) is a colorless and odorless diatomic gas. It was discovered
independently by Karl W. Scheele (1742-1786) in 1772 and Joseph Priestley
(1733-1804) in 1774 (see Gilbert (1999) for an excellent review). As Scheele’s
work was not published until 1777, Priestley is often credited with the discovery
of oxygen. Priestley named the new gas “dephlogisticated air” which was
eventually called “oxygene” (acid former) by Antoine Lavoisier (1743-1794).
Oxygen is particularly abundant in the earth’s crust (~54% by weight), occurs in
the atmosphere (~21% by volume, 23% by weight for dry air), and is the major
component of water’s structure (89% by weight). Oxygen has a melting point of
–219oC and a boiling point of –183oC. Oxygen is only slightly soluble in water
(~280mmol/dm3 at 25oC) — enough to support aquatic life. Oxygen is about five
times more soluble in organic solvents.
Oxygen is the first member of Group 6B of the periodic table and possesses
eight electrons with an electronic configuration of 1s2, 2s2, 2p4. Oxygen does not
possess available d orbitals so it is limited to a valency of 2. As shown in Figure
2.1, oxygen (3Σg-O2) is a diradical, possessing two unpaired electrons. Oxygen is
therefore paramagnetic and can be measured using EPR.
Formation.
Oxygen can be formed in the laboratory by:
a) thermal decomposition of metal oxides low in the electrode potential series
(e.g., Eqn 2.1);
b) thermal decomposition of higher oxides (e.g., Eqn 2.2);
c) catalytic decomposition of peroxides (Eqn 2.3);
d) the reaction between solid peroxides and water (Eqn 2.4);
e) thermal decomposition of salts containing oxygen-enriched anions (e.g.,
Eqns 2.5 and 2.6); or
f) the electrolysis of aqueous solutions of acids or alkalis. Industrially,
oxygen is obtained from the atmosphere by the liquefaction of air.
Biologically, oxygen is produced as a waste product of photosynthesis
(Eqn 2.7).
WWW.ESAINC.COM
37
Oxygen and ROS Electronic
Configuration
σ*2p
π *2p
π2p
σ 2p
GroundState
Oxygen
(3Σg-O2)
Singlet
Oxygen
(1∆gO2)
Singlet
Oxygen
(3Σg+O2)
Superoxide
Radical Anion
(O2.-)
Peroxide
Anion
(O22-)
Figure 2.1 Molecular Orbital Diagram Of Molecular Oxygen And
Some ROS (Based On An Original Figure By Halliwell And
Gutteridge (1999)).
2HgO → 2Hg + O2
2Pb3O4 → 6PbO + O2
2H2O2 → 2H2O + O2
2(Na)2O2 + H2O → 4NaOH + O2
2KNO3 → 2KNO2 + O2
2KMnO4 → K2MnO4 + MnO2 + O2
6CO2 + 6H2O → C6H12O6 + 6O2
Eqn 2.1
Eqn 2.2
Eqn 2.3
Eqn 2.4
Eqn 2.5
Eqn 2.6
Eqn 2.7
Chemical Reactions and Biological Significance.
The reactions of oxygen are often slower than would be predicted from its
electronegativity (3.5)1 which is second only to fluorine (4.0), a very reactive,
strong oxidizing agent. The reason for oxygen’s inertness is that its double bond
dissociation energy is relatively high so that reactions that require this double
1
Electronegativity is a measure of the ability of an atom to attract electrons and involves both ionization energy and
electron affinity. Pauling gave fluorine, the most electronegative element, an arbitrary value of 4.0 and related the
electronegativities of the atoms of other elements to it. The bond formed between two atoms of similar electronegativity
will be essentially covalent. An increase in the electronegativity of one atom will attract the electrons involved in the
covalent bond, causing it to be polarized. Further increases in electronegativity will result in increased polarity until the
electron pair will reside almost entirely on one atom, i.e., an electrovalent bond will be established.
WWW.ESAINC.COM
38
bond to be broken occur only at high temperatures. If this were not the case,
spontaneous combustion of animals and plants would be a very common event!
Once initiated by an external energy source most reactions involving the
breaking of oxygen’s double bond are self-sustaining, due to their exothermic
nature. The majority of oxygen reactions are found to occur at temperatures
considerably higher than room temperature, but oxygen can be made to react at
physiological temperatures by a variety of enzymes (Table 3.1). The most facile
reactions of oxygen are those in which its double bond is not completely broken,
such as in the formation of superoxide radical anion and peroxides.
An explanation for oxygen’s lack of reactivity can best be understood from its
electronic structure (Figure 3.1). Ground state oxygen has two unpaired (parallel
spin) electrons occurring in two degenerate antibonding π*2p orbitals (i.e.,
dioxygen is a triplet molecule in the ground state). For oxygen to oxidize a
chemical species in a two-electron reaction, the compound undergoing reaction
must have two unpaired (antiparallel spin) electrons to enter the π*2p orbitals of
the ground state oxygen molecule. Due to Pauli’s exclusion principle, an electron
pair would not fulfill this criterion. As a result, oxygen tends to accept electrons
singularly. This explains why molecular oxygen is kinetically unreactive with most
compounds but readily reacts with σ-radicals (such as R•) and transition metal
complexes. This is important when proposing that a compound undergoes autooxidation (see below). σ-Radicals possess an unpaired electron which can
readily enter one of oxygen’s π*2p orbitals. Readers interested in a more in-depth
discussion of the reactivity of oxygen are referred to Malmstrom (1982), and
Naqui et al. (1986).
Oxygen reacts with most metals except the less reactive ones (e.g., silver and
gold). Lithium forms oxides (Eqn 2.8). Sodium forms oxides and, in excess
oxygen, peroxides (Eqn 2.9). The remaining group 1A elements form
superoxides (e.g., KO2). For transition metals, the oxidation state of the metal in
the product depends upon the reaction conditions and the complexation of the
transition metal, as discussed in Chapter 2 and in greater detail below. The oneelectron reduction of oxygen by transition metal complexes is extremely
important to redox biochemistry. All non-metals, with the exception of the noble
gases and halogens, react directly with oxygen (Eqn 2.10). Oxides of the
halogens and the heavier noble gases can be prepared only by indirect means.
4Li + O2 → 2Li2O
2Na + O2 → Na2O2
S + O2 → SO2
WWW.ESAINC.COM
Eqn 2.8
Eqn 2.9
Eqn 2.10
39
Enzyme
Oxidases that reduce
dioxygen to hydrogen
peroxide
Comments
Flavin-containing Oxidases
a) Simple - D- and L-amino acid oxidase; Glucose
oxidase;
Some monoamine oxidases;
“Old Yellow Enzyme”;
Lactate oxidase
b) Metalloflavoproteins Xanthine oxidase;
Aldehyde oxidase
Reference
Malmstrom (1982)
Metal-containing Oxidases
Galactose oxidase;
Some monoamine oxidases;
benzylamine oxidases
Oxidases that reduce
dioxygen to water
Blue Oxidases
Ascorbate oxidase;
Ceruloplasmin;
Laccase
Malmstrom (1982)
Cytochrome Oxidase
Oxygenases
Dioxygenases
a) Flavin-containing: Few exist
b) Metal-containing:
Heme – Tryptophan-2,3-dioxygenase;
Indoleamine-2,3-dioxygenase
Non-heme – Lipoxygenase
Malmstrom (1982);
White and Coon
(1980)
Monoxygenases
a) Flavin-containing:
p-Hydroxybutylate hydroxylase
b) Metal-containing:
Heme – Cytochrome P450
Non-heme – Tyrosinase;
Dopamine-β- hydroxylase
Table 2.1 Some Enzymes That Utilize Oxygen.
Oxygen is converted to its allotrope2 ozone (O3) by silent (non-thermal) electrical
discharge (Eqn 2.11). Formation of ozone is an endothermic process so that
thermal energy produced during sparking would decompose it.
2
Allotropy is the ability of a substance to exist in two or more physical forms.
WWW.ESAINC.COM
40
3O2 → 2O3
O2 + 4H+ + 4e-
Eqn 2.11
Eqn 2.12
→ 2H2O
Oxygen is transformed into water through successive additions of electrons,
which produce a series of reduced intermediates collectively known as the
reactive oxygen species (ROS) (Figure 2.2). The four-electron, four-proton
reduction of oxygen to water is biologically very important as it is the terminal
reaction of aerobic respiration (Eqn 2.12). Unfortunately, respiration is not perfect
and electrons can “leak” from the electron transport chains, resulting in the
formation of potentially damaging ROS.3 Although mitochondrial respiration is
one of the major producers of ROS, it is by no means the only source (see
below). The steady-state levels of the ROS under biological conditions are
typically kept low with hydrogen peroxide, superoxide, and hydroxyl free radical
levels being 10-7-10-9, 10-10-10-11 and 10-15- 10-20M, respectively (Chance et al.
(1979); Floyd (1997)). Even so it is estimated that a 70kg human would be
expected to produce more than 1kg of superoxide annually!
'∆g O 2
HO2
SINGLET
OXYGEN
HYDROPEROXYL
H+
pKa = 4.5
e-
2/3O3
OZONE
O2
DIOXYGEN
-330mV
2H+ + e-
Ο2−
+940mV
SUPEROXIDE
H+ + e-
Η2Ο2
H+ + e-
H2 O
ΗΟ
+380mV
+2330mV
HYDROGEN
HYDROXYL
PEROXIDE
FREE RADICAL
Η 2Ο
WATER
SUPEROXIDE
DISMUTASE
CATALASE
Figure 2.2 The Relationship Between Oxygen And ROS. The
Potentials (in mV) are Electrode Potentials for the Reaction (see
Appendix 2.1)
3
Estimates for ROS production vary widely but it appears that about 2-3% of oxygen uptake by isolated mitochondria in
state 4 is reduced to H2O2 (Bovis and Chance (1973)); the production of superoxide is about 2-3nmol/min/mg protein
4
(Beal, (1997)); Escherichia coli produce 3 superoxide molecules for every 10 electrons transferred along their respiratory
chains (Imlay and Linn (1988)).
WWW.ESAINC.COM
41
The question is sometimes asked as to why oxygen and not some other
compound was chosen by nature as the terminal oxidant of the respiratory chain.
This question was elegantly answered by George (1965). He hypothesized that
“the kinetic activity of the halogens make them unsuitable as biological oxidants,
and nitrogen is too poor an oxidizing agent. Thus oxygen is the only element in
the most appropriate physical state, with a satisfactory solubility in water and with
desirable combinations of kinetic and thermodynamic properties.” Oxygen was
therefore the only option even though it came with a high price – its reduction to
pro-oxidant and biologically damaging species.
As shown in Figure 2.2 the ROS include both reduced and non-reduced forms of
oxygen. Before examining the reduced forms of oxygen in more detail, we will
first explore the chemistries of ozone and the electronically excited form of
oxygen — singlet oxygen.
2. Ozone.
Properties.
Ozone (O3) is an unstable, toxic, pale blue diamagnetic gas with a distinctive
odor. It has a melting point of –250oC and a boiling point of –112oC. It is only
slightly more soluble in water than oxygen, but unlike oxygen can also react with
it.
Formation.
Ozone is produced in the stratosphere by the action of sunlight on atmospheric
oxygen during the Chapman cycle (reviewed by Mustafa (1990) and Madronich
(1999)). The ozone layer extends from 17km above the equator (8km above the
poles) to about 50km above the Earth’s surface. This beneficial layer absorbs UV
energy and protects living species from the mutagenic effects of electromagnetic
radiation. In the lower troposphere, however, ozone is a major pollutant and one
of the main components of photochemical smog. Ozone can be formed in the
vicinity of electrical machinery and may cause health problems in poorly
ventilated areas. Ozone can be formed in the laboratory by exposing dry oxygen
gas to silent electrical discharge (Eqn 2.11).
Chemical Reactions and Biological Significance.
Ozone is thermodynamically unstable with respect to oxygen and, at high
concentrations, is dangerously explosive. It is a very reactive compound (similar
in reactivity to gaseous chlorine and fluorine) and even reacts slowly with water
to produce a variety of ROS (Table 2.2). Ozone is a strong oxidizing agent. It is
WWW.ESAINC.COM
42
only a little less reactive than the hydroxyl free radical and a much more powerful
oxidant than oxygen. Ozone can oxidize sulfide to sulfate, iodide to iodine and Fe
(II) to Fe (III). Ozone’s reaction with thiols (e.g., albumin and glutathione (GSH))
is complex. Ozone oxidizes GSH to its disulfide in a two-electron process (note
that the valency of sulfur does not change) (Eqn 2.13). It can also react with
thiols to produce compounds containing sulfur in its higher valencies. For
example, thiols or thiol anions (mercaptide ions) can act as nucleophiles and
undergo additional reactions with ozone, resulting in the formation of alkyl
sulfinates (RSO2H) and alkyl sulfonates (RSO3H) (note that sulfur’s valency has
increased from 2 to 4 and 2 to 6, respectively) (sulfur’s valency changes are
explored in greater detail below). Ozone can react with mercaptide ions in a oneelectron process, leading to the formation of the thiyl radical and ozone radical
anion (Eqn 2.14). The latter decomposes under acidic conditions and produces
hydroxyl free radicals (Eqn 2.15). The thiyl (RS•) radical can also undergo several
other reactions, including dimerization with itself or other thiols, forming the
disulfide (RSSR) and mixed disulfide (R’SSR”), respectively. It can react with the
mercaptide anion, forming the disulfide radical anion (RSSR•-) that can reduce
oxygen to superoxide (Eqn 2.16). We will explore the formation and reactions of
the thiyl radical in greater detail below.
2GSH + O3 → GSSG + O2 + H2O
RS- + O3 → RS• + O3•O3•- + H+ → HO• + O2
RSSR•- + O2 → RSSR + O2•-
Eqn 2.13
Eqn 2.14
Eqn 2.15
Eqn 2.16
Ozone readily reacts with organic compounds containing unsaturated bonds and
forms a variety of chemical species — hydrogen peroxide, carbonyl compounds,
Criegee ozonides, aliphatic radicals and hydroxy-hydroperoxides (Figure 2.3) —
depending on the environment in which the reactions are taking place (Pryor
(1993); Pryor and Church (1991)). Ozone also attacks unsaturated compounds
and produces free radicals capable of promoting lipid peroxidation.
Unlike all the other ROS, ozone is the only pro-oxidant not produced
endogenously. Ozone is extremely toxic to living organisms. For example, it can
cause respiratory problems (airway inflammation and decreased pulmonary
function), and can damage the skin of humans and animals as well as the
surface tissues of plants (Menzel (1984); Menzel and Meacher (1999); Mustafa
(1990); Runeckles (1994); Thiele et al. (1997)). Such biological damage is
complex but is thought to be due to direct oxidation, through the formation of free
radicals and by the production of other reactive intermediates. The reactions of
ozone of biological importance are summarized in Table 2.2.
WWW.ESAINC.COM
43
Reaction
Comments
In aqueous solution ozone decomposes to give a variety of
ROS including H2O2, O2•-, HO• and HO2•.
Ozone may be able to abstract H• (hydrogen atoms) from
many organic compounds leading to chain reactions in the
presence of O2.
Fatty acids with one double bond and cholesterol react to
give epoxides.
Water
Saturated compounds
Unsaturated compounds
PUFAs (polyunsaturated fatty acids) undergo lipid
peroxidation giving RO2•, RO2H, RO•, TBAR-reactive
(thiobarbituric acid reactive) substances such as MDA
(malondialdehyde), aldehydes, conjugated dienes, alkanes
(e.g., ethane, pentane), and epoxides. Many reactive
aldehydes are cytotoxic (see below).
Damage to the lipid bilayer leads to altered membrane fluidity
and permeability.
Many functional groups are oxidized including thiols, amines,
alcohols, and aldehydes. Amino acids such as cysteine,
cystine, histidine, methionine, tyrosine and tryptophan can be
oxidized. Enzymes such as those involved in prostaglandin
synthesis, cholinesterase and α1-antiproteinase can be
damaged by ozone. Other macromolecules such as structural
proteins, DNA, RNA and membranes can also be damaged.
Ozone readily reacts with antioxidants such as albumin,
ascorbic acid, GSH, tocopherol, and uric acid.
Amino acids, proteins,
enzymes, DNA and RNA
Antioxidants
Table 2.2 Some Important Reactions Of Ozone.
OH OH
R1
H H R2
DIOL
O
R1
O2
H H R2
DIRADICAL
R1
Reduction
H
O
R1
R2
O
O
O
R1
O
R1CHO
O
O
H H R2
TRIOXOLANE
NONAQUEOUS
O
O
O
H
R2
CRIEGEE
OZONIDE
O
R2
OH
H
CARBONYL
OXIDE
AQUEOUS
R2
H
O2H
R2CHO
H2O2
HYDROXYHYDROPEROXIDE
Figure 2.3 The Reaction Of Ozone With Alkenes Such As PUFAs Under
Aqueous And Non-Aqueous Conditions. Based On Mechanisms Originally
Proposed By Pryor And Church (1991).
WWW.ESAINC.COM
44
Ozone is an important atmospheric pollutant so its biological actions have been
studied in several animal and plant models. Studies have focused on the action
of respired ozone on pulmonary function. Exposure to ozone caused airway
hyperreactivity, neutrophil infiltration, increased epithelial macromolecular
permeability, and the promotion of mucus secretion that eventually culminates in
lung inflammation. The first biological fluid that comes into contact with inhaled
ozone is the respiratory tract lining fluids (RTLF) which not only serve to absorb
and detoxify ozone but also limit its passage to more vulnerable areas (e.g., the
peripheral gas exchange regions of the lung). It is extremely difficult to obtain an
RTLF sample to study possible antioxidant defenses so it is common to use
plasma (Cross et al. (1992); Van der Vleit et al. (1995a)) or skin as models
(Thiele et al. (1997)). In the plasma model ozone quickly and directly reacted with
uric acid and ascorbic acid, but only slowly reacted with protein-thiol groups. In
skin, ozone depleted both ascorbic acid and α-tocopherol while increased
production of the lipid peroxidation product, malondialdehyde. RTLF contains
high levels of glutathione (GSH) that not only reacts directly with ozone (Eqn
2.13) but also can control the toxic effects of some of its secondary products
(e.g., aldehydes and hydrogen peroxide).
The biological consequences of inhalation cannot be due totally to ozone itself.
Ozone will react primarily with RTLF antioxidants so that it will not be able to
deeply penetrate lung tissue. Pryor (1993) proposed ozone’s toxicity may be due
to a cascade mechanism whereby ozonolysis products act as messengers
capable of inducing biological changes far removed from the initial site of ozone
attack. Such products include aldehydes, hydroxyhydroperoxides, and Criegee
ozonides that can activate lipases. Lipases can then release endogenous cellular
signal transduction molecules and mediators of inflammation, such as
eicosonides and platelet-activating factors. Ozonolysis products may also be
responsible for some of the other problems caused by respired ozone such as
carcinogenesis, damage to the hematopoietic system, and altered central
nervous system functionality (Mustafa (1990)).
Measurement.
The presence of ozone can be determined by its ability to “tail” mercury (the
surface of this metal is partially converted to its oxide so that it sticks to the walls
of the vessel containing it). In the laboratory ozone can be quantified by reacting
it with acidified potassium iodide. The iodine so liberated can then be titrated with
a standard solution of sodium thiosulfate. Ozone also can be measured using
chemiluminescence-based detectors (MacDougal et al. (1998); van Heusden and
Mans (1978)).
WWW.ESAINC.COM
45
3. Singlet Oxygen.
Properties.
The input of appropriate energy can excite the unpaired electrons of the oxygen
molecule, thereby forming singlet oxygen. There are two forms of singlet oxygen
(Figure 2.1). 1∆gO2 has two electrons with opposite spins in a common π*2p
orbital; therefore, and therefore is not a free radical. It has an energy of 93.7kJ
mol-1 above the ground state. 1Σ+O2 has two electrons with opposite spins in
different π*2p orbitals. It is even more reactive than 1∆gO2 with an energy of
156.9kJ mol-1 above the ground state. In biological systems 1Σ+O2 usually decays
rapidly (t1/2=10-11s) to the 1∆gO2 state and is usually ignored. Once formed the
1
∆gO2 molecule is not long-lived (t1/2=2 x 10-6s at 37oC) due to its extreme
reactivity.
Formation.
Singlet oxygen can be formed in the laboratory by:
a) the action between hypochlorite and hydrogen peroxide (Eqn 2.17);
b) the thermal dissociation of endoperoxides (e.g., 3,3'-(1,4-naphthylidene)
dipropionate);
c) the disproportionation of superoxide and hydroperoxide;
d) decomposition of primary and secondary peroxyl radicals (Russel
reaction), and
e) the Haber-Weiss reaction (reviewed by Huie and Neta (1999)).
As discussed in greater detail below, phagocyte activation during the immune
response produces hypochlorous acid from hydrogen peroxide and chloride ions
(Harrison and Schultz (1976)) (Eqn 2.18). The subsequent reaction between
hydrogen peroxide and hypochlorite forms singlet oxygen (Eqn 2.17) is used to
kill pathogens (Kiryu et al (1999)).
OCl- + H2O2 → Cl- + H2O + 1∆gO2
Cl- + H+ + H2O2 → HOCl + H2O
WWW.ESAINC.COM
Eqn 2.17
Eqn 2.18
46
N
CH3
N
CH3
S
Methylene Blue*
O
CH3
N
N
CH3
O2
Excited State
Ground
N
R
State
(3Σg-O2)
UV
260nm
NH
HO
N
NH2
8-Hydroxy2-Deoxy-Guanosine
O
N
N
CH3
N
S
N
CH3
CH3
CH3
NH
O2
Singlet
State
(1∆gO2)
N
R
N
NH2
2-Deoxy-Guanosine
Methylene Blue
Ground State
Figure 2.4 The Production Of 1∆gO2 By Photosensitization Can Lead To
Damage Of Biologically Important Compounds (e.g., DNA Bases). Based On
Some Reactions Presented By Halliwell And Gutteridge (1989)).
Singlet oxygen is also produced through photosensitizing reactions. Here the
absorption of light of the correct energy can excite a molecule into a higher
energy state. This energy can then be transferred to an oxygen molecule in close
proximity, exciting it to its singlet state. The photosensitizer simultaneously
returns to its ground state. For example, singlet oxygen generated by the
interaction of UV light with the dye, methylene blue, can be used to explore the
chemical reactions of singlet oxygen (see Figure 2.4 above). In this case singlet
oxygen caused oxidative damage to DNA producing the very mutagenic lesion,
8-hydroxy-2’deoxyguanosine (Chapter 3). Many exogenous compounds can act
as photosensitizing agents, including dyes (e.g., acridine orange, merocyanine540, blue) and other compounds (e.g., psoralen, meso-substituted porphyrins).
Endogenous compounds can act as photosensitizing agents as well, including
porphyrins and corrins (e.g., heme), linear pyrroles (e.g., bilirubin/biliverdin),
conjugated polyenes (e.g., retinal) and flavins (e.g., FAD, FMN and riboflavin).
Some drugs (e.g., tetracycline antibiotics (Hassan and Khan (1986)) and
constituents of cosmetics may also act as photosensitizing agents (Halliwell and
Gutteridge (1989)).
Not all photosensitization damage, however, occurs through the generation of
singlet oxygen (type II mechanism). The excited photosensitizing agent itself can
inflict damage directly (type I mechanism). Furthermore, excited photosensitizing
agents (e.g., merocyanine-540) are also capable of generating other ROS, such
as superoxide and hydroxyl free radicals (e.g., Feix and Kalyanaraman (1991)).
WWW.ESAINC.COM
47
Compound
Amino acids (free and protein
bound)
• Cysteine (R-SH)
• Histidine (free and part of
carnosine)
• Methionine
• Tryptophan
Ascorbic acid
Cholesterol
Phenols (e.g., tyrosine)
Polyene
• β-Carotene
Purine
• Guanine
• 2’-Deoxy-guanosine
Pyrroles
• Bilirubin
α-Tocopherol
Consequence
Cystine (RSSR) and sulfonic acid (RSO3H)
Endoperoxide, oxohistidine, other products
Methionine sulfoxide
Hydroperoxide, dioxetene, N-formylkynurenine
Quenching, some oxidation products
5α-Hydroperoxide, minor products
Quenching, some oxidation products formed
Quenching, oxidation products
8-Hydroxyguanine
8-Hydroxy-2’deoxyguanosine
Quenching, some oxidation products
Quenching, some oxidation products (e.g., αtocopherylquinone)
Table 2.3 The Reaction Of Singlet Oxygen With Some Biologically Important
Species.
Chemical Reactions and Biological Significance.
Singlet oxygen is a much more powerful oxidizing agent than oxygen because
the spin restriction that encumbers oxygen is removed. Not surprisingly, the
typical basal singlet oxygen levels found in vivo are kept low, e.g., ~1 x 10-16 to
1 x 10-18M, for isolated hepatocytes and whole liver, respectively.
Singlet oxygen can react by two mechanisms:
1) it can transfer its excitation energy to another molecule (which in turn
becomes excited) while subsequently returning to its ground state (i.e.,
quenched); or
2) it can chemically modify another molecule. Chemical modification depends
upon the structure of the compound being attacked (Table 2.3).
Compounds containing carbon-carbon double bonds are particularly
abundant in nature (e.g., carotenoids and polyunsaturated fatty acids) and
are readily damaged by singlet oxygen (fatty acids form hydroperoxides;
phenols form endoperoxides that can undergo further decomposition;
tryptophan forms a dioxetane that then undergoes ring opening).
Singlet oxygen has both beneficial and detrimental effects. As a beneficial
molecule it, along with a variety of other pro-oxidants, plays an important role in
the active defense mechanisms of the immune system. Photosensitization
reactions are also used in disease treatment (photodynamic therapy of Herpes
WWW.ESAINC.COM
48
simplex and jaundice)(Halliwell and Gutteridge (1989)). Unfortunately, singlet
oxygen also poses a major problem for biological systems. It directly reacts with
unsaturated fatty acids (causing lipid peroxidation), and, with DNA (producing
strand breaks and mutagenic lesions) (Devasagayam et al. (1991)) (Chapter 3).
Singlet oxygen is also a problem for biological systems involved in light
transduction (e.g., the chloroplast and the eye) or in humans who are sensitive to
light (e.g., patients exhibiting porphyrias).
The major defense against singlet oxygen-induced damage appears to be
quenching by ascorbic acid (forming an unstable reactive, hydroperoxide, that
can decompose to potentially toxic compounds – L-threonolactone and oxalic
acid), carotenoids, and tocopherols (forming α-tocopherol hydroperoxide, that
decomposes to α-tocopherylquinone) (Fukuzawa et al. (1998); Halliwell and
Gutteridge (1989); Kaiser et al. (1990); Kwon and Foote (1988)). See Table 2.3
(above). Dietary flavonoids can also protect against singlet oxygen damage but
their role in vivo needs to be evaluated further (Tournaire et al. (1993)).
Measurement.
Several approaches, varying in their degree of specificity, can be used to
measure singlet oxygen levels or to assess its involvement in a reaction of
interest (Basu-Modak and Tyrrell (1993); Egorov et al. (1997); Halliwell and
Gutteridge (1989); Motohashi and Mori (1989)). These include:
•
•
•
•
•
Measurement of light emission (monomol emission at 1270nm; dimol
emission at 634 and 703nm);
The use of EPR with sterically hindered heterocyclic amines or bipyrazole
derivatives;
Measurement of novel markers produced from ß-carotene, cholesterol,
phenol and tryptopan;
The use of HPLC-ECD using the electrochemically active adduct, 2,2,6,6tetramethyl-4-piperidone-N-oxyl which is formed when 2,2,6,6-tetramethyl-4piperidone; and
The use of scavengers (e.g., azide, carnosine and diphenyl-isobenzofuran) to
inhibit singlet oxygen production.
4. Superoxide (Radical Anion).
Properties.
When oxygen is reduced in a single-electron process (Eqn 2.19), the additional
electron must enter one of oxygen’s π*2p antibonding orbitals (Figure 2.1). The
resulting molecule is both an anion and free radical, called superoxide, or, more
correctly, the superoxide radical anion. The addition of an extra electron to the
WWW.ESAINC.COM
49
oxygen molecule weakens the double bond, producing a more reactive molecule
with only one and one-half bonds. It is relatively unstable and has a half-life of
10-5s at 37oC.
O2 + e- → O2•-
Eqn 2.19
Formation.
Superoxide can be produced in the laboratory by using pulse radiolysis of
aqueous solutions, electrochemical reduction of oxygen, or from ionic salts such
as potassium- or tetramethylammonium-superoxide. There are many example of
the superoxide production in vivo including:
•
The electron transport chains. Located in mitochondria, the endoplasmic
reticulum, nuclear membrane, and chloroplasts, these along with immune
defense, are probably the most important sources of superoxide in vivo.
Mitochondria are both important sources – and important targets – of reactive
species. Acute exposure to relatively high levels of oxidants, especially in the
presence of calcium, can induce an event termed the mitochondrial
permeability transition, uncouple oxidative phosphorylation, and may
contribute to cytotoxicity via necrosis and/or apoptosis (through release of
cytochrome c, apoptosis-inducing factor, and other proteins). Longer
exposure of mitochondria to milder oxidants appears to lead to progressive
mitochondrial impairment and eventual dysfunction, possibly, in some
systems, by reducing mitochondrial DNA (mtDNA) expression. Even if
mitochondria do not undergo catastrophic failure through one of these
mechanisms, oxidant-mediated mitochondrial dysfunction may proceed due
to oxidant damage to lipids, proteins, and nucleic acids. The potential for such
oxidant-mediated damage is increased because mitochondria are also the
major source of reactive species in eukaryotes. Mitochondrial respiration
generates ROS, and their generation may be increased in damaged
mitochondria and in cells with compromised mitochondrial function. This
potential feed-forward interaction between oxidative stress and mitochondrial
dysfunction may lead to a deleterious spiral and eventual mitochondrial
collapse and cell death.
At a crude level, mitochondrial structure may be described as consisting of an
inner compartment (termed the matrix), surrounded by two lipid bilayers (the
inner and outer mitochondrial membranes). The matrix primarily houses the
elements involved in mitochondrial gene expression and energetics. The
mitochondrial gene expression system includes the mitochondrial genome,
mitochondrial ribosomes, and the transcription and translation machinery
needed to regulate and conduct gene expression as well as mtDNA
WWW.ESAINC.COM
50
replication and repair. Machinery involved in energetics includes the enzymes
of the Kreb's citric acid or TCA (tricarboxylic acid) cycle, some of the enzymes
involved in fatty acid catabolism (β-oxidation), and the proteins needed to
help regulate these systems. The inner membrane is central to mitochondrial
physiology and, as such, contains multiple protein systems of interest. These
include the protein complexes involved in the electron transport component of
oxidative phosphorylation and proteins involved in substrate and ion
transport.
Mitochondrial roles in, and effects on, cellular homeostasis extend far beyond
the production of ATP, but the transformation of energy is central to most
mitochondrial functions. For example, mitochondria play a central role in the
regeneration of antioxidants both directly, and indirectly, through the
production of reducing equivalents. Reducing equivalents are also used for
anabolic reactions. The energy produced by mitochondria is most commonly
thought of to come from the pyruvate that results from glycolysis, but it is
important to keep in mind that the chemical energy contained in both fats and
amino acids can also be converted into NADH and FADH2 through
mitochondrial pathways. The major mechanism for harvesting energy from
fats is β-oxidation; the major mechanism for harvesting energy from amino
acids and pyruvate is the TCA cycle. Once the chemical energy has been
transformed into NADH and FADH2, these compounds are fed into the
mitochondrial respiratory chain.
The mitochondrial respiratory chain consists of five proteins complexes:
NADH dehydrogenase (complex I), succinate dehydrogenase (complex II,
also part of the TCA cycle), cytochrome bc1 complex (complex III),
cytochrome c oxidase (complex IV) and the FoF1ATPase (complex V). The
components of each of these protein complexes are listed in Table 2.4. The
first four components are also referred to collectively as the mitochondrial
electron transport chain.
WWW.ESAINC.COM
51
Complex
Name (alternative
name)
Comments
Prosthetic
Group
Substrate
Binding Site
Products
I
NADH:
ubiquinone
Oxido-reductase
(NADH-Q
Reductase)
(NADH dehydrogenase)
Composed of ~39 subunits (7
coded by mitochondrial DNA;
~32 by nuclear DNA)
(Wallace [1992]). Others
report 16 subunits (MW
~500kd) (Newsholme and
Leech [1992]) or 25 subunits
(MW ~850kd) (Stryer [1988]).
FMN
Fe-S
NADH
NADH - matrix
NAD
Ubiquinone
Ubiquinone membrane
Ubiquinol
Succinate:
ubiquinone
Oxido-reductase
(Succinate-Q
Reductase)
(Succinate Dehydrogenase)
Composed of 4 subunits (MW
~140kd) all encoded by
nuclear DNA (Wallace [1992];
Stryer [1988]),
FAD
Fe-S
Ubiquinol:ferricytochrome C
Oxido-reductase
(Cytochrome bc1
complex)
(Cytochrome
reductase)
(Ubiquinone Dehydrogenase)
Composed of ~10 subunits
(MW ~250kd) with 1 subunit
encoded by mitochondrial
DNA and ~9 by nuclear DNA
(Wallace [1992]).
Ferrocyto-chrome
C:oxygen Oxidoreductase
(Cytochrome
oxidase)
FoF1 ATPase
(ATP synthase)
II
III
IV
V
+
Succinate
Succinate matrix
Fumarate
Ubiquinone
Ubiquinone
membrane
Ubiquinol
Heme b-562
Heme b-566
Heme c1
Fe-S
Ubiquinol
Ubiquinone
membrane
Ubiquinone
Cyt C-Fe
3+
Cyt C – intermembrane
space
Cyt C-Fe
2+
Composed of 6 (MW ~160kd
(Newsholme
and
Leech
[1992])) to ~13 subunits
(Wallace [1992]). Of the ~13
subunits 3 are encoded by
mitochondrial DNA and 10 by
nuclear DNA.
Heme a
Heme a3
Cu
Cyt C-Fe
2+
Cyt C Intermembrane
space
Cyt C-Fe
3+
F1: MW ~380kd. Composed
of five types of subunits (α, β,
γ,δ, ε). Contains the catalytic
site for ATP synthesis.
Located as a spherical
headpiece on matrix side.
None?
H , ADP
Oxygen
+
+
H intermembrane
space
Water
ATP
F0: MW ~66kd. Composed of
four subunits. Functions as a
transmembrane proton
channel.
Four additional subunits,
including the F1 inhibitor, are
located in the stalk between
F0 and F1
Table 2.4 Mitochondrial Respiratory Chain And ATP-Synthesizing Complex.
Under physiological conditions, electrons generally enter either through
complex I (NADH-mediated, examined in vitro using substrates such as
glutamate/malate) or complex II (FADH2-mediated, examined in vitro using
succinate) (Figures 2.5). Electrons are then sequentially passed through a
series of electron carriers. The energy released during the transfer of
WWW.ESAINC.COM
52
electrons from carrier to carrier is used to pump protons from the inner
mitochondrial matrix to the intermembrane space at three points in the chain
(complexes I, III, and IV). The progressive transfer of electrons (and resultant
proton pumping) converts the chemical energy stored in carbohydrates, lipids,
and amino acids into potential energy in the form of the proton gradient. The
potential energy stored in this gradient is used to phosphorylate ADP forming
ATP.
Two Electrons
Enter Here
MATRIX
Fumarate + 2H +
Succinate
E-FAD
E-FADH2
[(Fe-S)Red]3
Succinate-CoQ Reductase
Complex II
[(Fe-S)Ox]3
H+
H+
2e-
NADH + H+
FMN
(Fe-S)Red
Q
Cyt b [Fe 2+]
(Fe-S)Ox
Cyt c 1 [Fe2+]
Cyt c [Fe 3+]
QH2
Cyt b [Fe 3+]
(Fe-S)Red
Cyt c 1 [Fe3+]
Cyt c [Fe 2+]
Two Electrons
Enter Here
NAD+
(Fe-S)Ox
FMNH2
Q Cycle
H+
H+
QH2-Cytochrome c Reductase
Complex III
NADH-CoQ Reductase
Complex I
INTERMEMBRANE SPACE
MATRIX
H+
Cyt c [Fe 3+]
Cyt a [Fe 2+]
Cyt a 3 [Fe3+]
Cu+
O2 + 4H+
Four Electrons
Transferred Here
Cyt c [Fe 2+]
Cyt a [Fe 3+]
Cyt a 3 [Fe2+]
Cu2+
2H2O
H+
Cytochrome c Oxidase
Complex IV
INTERMEMBRANE SPACE
Figure 2.5 Components Of The Electron Transport Chain.
Mitochondrial pathways of energy production culminate in the electron
transport chain and the coupled transfer of four electrons (and four protons) to
molecular oxygen to form water. This final reaction, catalyzed by cytochrome
oxidase, is “safe,” in that the coordinate, sequential transfer of four single
electrons is rarely, if ever, associated with free radical damage. It has been
WWW.ESAINC.COM
53
generally estimated that electron leak from the respiratory chain is ~1-4% of
total oxygen consumption (Boveris and Chance (1973); Nohl and Hegner
(1978); Turrens et al. (1985)), although these estimates may be 10-fold too
high (Beckman and Ames (1998); Hansford et al. (1997)). Electron leak
predominantly occurs in complexes I and III.
Mitochondrial sources of oxidative stress other than the respiratory chain
include: monoamine oxidase (production of superoxide and hydrogen
peroxide); mitochondrial nitric oxide synthase (RNS formation); the NADH
reductase on the mitochondrial outer membrane (ROS formation); and the
flavoprotein dihydroorotate dehydrogenase (ROS formation).
•
•
•
Immune defense. The phagocytosis of pathogens by neutrophils, eosinophils
and mononuclear phagocytes involves production of superoxide through the
activation of membrane-bound NADPH oxidase complex (see below).
Enzymatic reactions. A variety of enzymes can produce superoxide
including peroxidases, oxidases and dioxygenases (Table 2.1). The ability of
xanthine oxidase to produce superoxide and the role of this enzyme in the
damage associated with ischemia-reperfusion injury is reviewed in Chapter 4.
Nitric oxide synthase, the enzyme responsible for the endogenous production
of nitric oxide from arginine can, under certain circumstances, lead to the
formation of superoxide (Griffith and Steuhr (1995)). Thus the same enzyme
can produce both precursors of the aggressive pro-oxidant, peroxynitrous
acid (see below). Cytochrome P450 (isoforms) found in the endoplasmic
reticulum of many animal and some plant tissues is a mono-oxygenase
(mixed–function oxidase) that uses oxygen and a reducing agent NADPH
(mediated by the flavoprotein NADPH-cytochrome P450 reductase (EC
1.6.2.3)) in the oxidation of many substrates, especially xenobiotics (Chapter
5). Under certain circumstances it can produce both superoxide and hydrogen
peroxide.
Oxygen-heme interaction. The binding of oxygen to the heme ring of
deoxyhemoglobin (or deoxymyoglobin) forms oxyhemoglobin (or
oxymyoglobin), a Fe (II)-oxygen complex (Eqn 2.20). Sometimes this complex
decomposes with the production of superoxide and methemoglobin
(containing Fe (III) state) at about 3%/day (Eqn 2.21) (Halliwell and
Gutteridge (1989)). Methemoglobin is scavenged by reduction to hemoglobin
by methemoglobin reductase in a NADPH-based mechanism.
Fe2+(heme) + O2 → Fe2+(heme)—O2 ↔ Fe3+(heme)—O2Fe2+(heme)—O2 → Fe3+(heme) + O2•-
•
Eqn 2.20
Eqn 2.21
Metal-catalyzed auto-oxidation of carbohydrates, thiols, monoamines
and other endogenous metabolites and redox cycling of quinones (see
below).
WWW.ESAINC.COM
54
Chemical Reactions.
In comparison to many other pro-oxidants, superoxide is not that reactive (typical
second order rate constants4 of 103 to 106 M-1s-1]). However, it does show
selective reactivity towards some molecules including other σ-radicals (e.g., nitric
oxide).
I will limit the chemistry of superoxide to aqueous conditions. Superoxide can act
both as a weak reducing and oxidizing agent. Superoxide can reduce Fe (III) to
Fe (II) and Cu (II) to Cu (I) (e.g., as part of the superoxide dismutation reaction).
The Reduction of Fe (III) to Fe (II) is biologically important as Fe (II) has the
potential to take part in the Fenton reaction, ultimately leading to the production
of hydroxyl free radicals. In this example, the Fenton reaction is being promoted
by superoxide so it is known as the superoxide-Fenton, or Haber-Weiss, reaction
(we will return to the Fenton and Haber-Weiss reactions below). Some
researchers have challenged the role of superoxide in the promotion of the
Fenton reaction in vivo. For example, based on reaction kinetics, superoxide is
much more likely to take part in dismutation by SOD than to play a role in the
Haber-Weiss reaction (Wardman and Candeias (1996)).
Superoxide can oxidize Fe (II) (or Cu (I)), ascorbic acid and compounds
containing a thiol group (for an example see Eqn 2.22)). Whether superoxide
oxidizes or reduces iron is dependent upon the experimental conditions, whether
the iron is free or bound, and what iron-chelator is present (Miller et al., (1990)).
Metal2+ + O2•- + 2H+ → Metal3+ + H2O2
Eqn 2.22
Superoxide is a weak base (pKa 4.5-4.8) and accepts protons to form the
hydroperoxyl radical (HO2•) (Eqn 2.23). Under physiological conditions only
about 1% of superoxide is protonated. Acidic-conditions promote hydroperoxyl
radical formation. Therefore, when the pH is decreased (e.g., in the lysozome, in
the microenvironment of biological membranes, and following acidosis, ischemia,
and prolonged exercise), the chemistry of the hydroperoxyl radical becomes
more important. The hydroperoxyl radical is a relatively long-lived species and is
more reactive (a stronger oxidizing and reducing agent) than superoxide.
Furthermore, the hydroperoxyl radical is lipophilic (i.e., it can readily pass through
membranes) and, unlike superoxide, is a promoter of lipid peroxidation (Chapter
3). Superoxide, due to its charge, cannot pass through membranes unless a
carrier is present. Superoxide can enter the erythrocyte by using the anion
transporter through which chloride and bicarbonate anions normally pass.
4
See Appendix 2.2 for typical rate constants.
WWW.ESAINC.COM
55
O2•- + H+ ↔ HO2•
Base
Acid
Eqn 2.23
Superoxide spontaneously dismutates (disproportionates) into hydrogen peroxide
(Eqn 2.24). The rate of dismutation is pH sensitive and is most rapid under acidic
conditions. Under physiological conditions, the rate of dismutation is found to be
approximately 105 M-1s-1 so that any reaction involving superoxide must be in
competition with dismutation. Therefore, any reaction producing superoxide will
also be producing hydrogen peroxide.
2 O2•- + 2H+ → 2H2O2 + O2
Eqn 2.24
Superoxide readily reacts with other radicals. The reaction between superoxide
and nitric oxide, which forms peroxynitrite (ONO2-), is rapid, with a typical
second-order rate constant of 107-109 M-1s-1 (Pryor and Squidrito (1995); Radi et
al. (1991a,b)) (Eqn 2.25). The importance of peroxynitrite formation is discussed
further below. Superoxide also reacts with hypochlorous acid, forming hydroxyl
free radicals (Eqn 2.26).
O2•- + NO• → ONO2O2•- + HOCl → HO• + Cl- + O2
Eqn 2.25
Eqn 2.26
Biological Significance.
Superoxide can be a benefit or a detriment to the living organism. Superoxide
helps the body in its defense against invading pathogens. However, the
unwanted production of superoxide is a problem causing enzyme inhibition,
release of redox active iron, and increasing oxidative stress. We now explore
these extremes by giving two examples, the use of superoxide in destroying
pathogens, and the problems of superoxide production in the brain.
The Pro: Superoxide is a major pro-oxidant and precursor for many of the other
aggressive cytotoxic species used by the defense system to control pathogens.
Once stimulated neutrophils, eosinophils, monocytes (with the exception of
macrophages), and B lymphocytes show increased oxygen consumption, often,
but incorrectly referred to as the respiratory burst (Babior (1978); Badwey et al.
(1979); Robinson and Badwey (1995)) (See Appendix 2.3). The increase in
oxygen consumption is associated with increased glucose flux through the
pentose phosphate pathway leading to increased NADPH production. It is
caused by the activation of a membrane-bound NADPH oxidase complex
WWW.ESAINC.COM
56
responsible for the reduction of oxygen to superoxide (Eqn 2.27). Leukocyte
NADPH oxidase is a highly complex protein whose components are distributed
between the cytosol and membranes of a variety of organelles, including the
plasma membrane, secretory vesicles, and granules (Babior (1999)). The active
enzyme is composed of heterodimeric flavohemoprotein cytochrome b558, and
two guanine nucleotide-binding proteins (one is a cytosolic Rac2 protein of the
Rho family; the other a membrane bound Rap1 protein of the Ras family) (Babior
(1999)). Upon stimulation the various components of the NADPH oxidase
complex come together and specifically organize within the membrane so that
NADPH oxidation occurs on the cytosolic side while oxygen reduction occurs on
the extracellular side. NADPH oxidation involves a flavin that either reduces
oxygen to superoxide directly or passes its electron to oxygen via cytochrome
b558. The passage of electrons across the membrane appears to be
accompanied by an outward movement of protons through membrane channels,
in order to maintain electroneutrality. Readers interested in an in-depth
discussion of enzyme activation, deactivation, and electron transport by the
oxidase are referred to Babior (1999).
NADPH + 2O2 → NADP+ + H+ + 2O2•-
Eqn 2.27
Superoxide can then produce other pro-oxidants. Once dismutated to hydrogen
peroxide hydroxyl free radicals can be produced by the Fenton reaction (see
below). Hydrogen peroxide is also used by myeloperoxidase (MPO) in the
production of hypochlorous acid and by eosinophil myeloperoxidase (EPO) for
synthesis of hypobromous acid (see below). Phagocytes also contain an
inducible-form of nitric oxide synthase that can produce nitric oxide in large
amounts. Nitric oxide reacts with superoxide to form peroxynitrite, another potent
pro-oxidant in the cells’ armamentarium (see below).
WWW.ESAINC.COM
57
AlcDH
CH 2 CH=NH
CH 2 CHO
AldDH
HO
HO
H2O
NH 3
OH
OH
CATECHOLIMINE
CATECHOLALDEHYDE
H2O 2
MAO
O2
CO 2 H
CH 2 OH
OH
CATECHOLACID
HO
OH
CATECHOLALCOHOL
SUBSTITUTED
CATECHOLACIDS
SUBSTITUTED
CATECHOLALCOHOLS
HO
CH 2 CH 2 NH 2
AlcDH
AldDH
O2
HO
O 2- + H +
OH
CATECHOLAMINE
H2O 2
H2O 2
MAO
O2
CH 2 CH 2 NH 2
O2
CH 2 CH 2 NH 2
+
HO
Nu
HO
AUTO-OXIDATION
OH
OH
5-substituted6-substitutedCATECHOLAMINE CATECHOLAMINE
SEMIQUINONE
O2
ROS
O 2- + H +
O
O
QUINONE
H2O 2
CATECHOL
AUTO-OXIDATION
(METAL INDUCED)
HO
HO
OTHER
INTERMEDIATES
+
OH
CH 2 CH 2 NH 2
CH 2 CH 2 NH 2
Nu
O
H2O 2
Catecholamine
Metabolism
N
H O2
LEUCOCHROME
(e.g., nucleophiles:
Nu: cysteine,
homocysteine,
glutathione,
protein thiols; free
radicals: HO .)
O
O
HO
O 2- + H +
H2O 2
N
H
O2
O
O 2- + H +
O
N
H
-O
+
N
H
DOPACHROME
H2O 2
Figure 2.6 The Catabolism And Auto-Oxidation Of Catecholamines Are
Intimately Involved With ROS Production. This Example Shows The
Metabolism Of Dopamine. (Adapted From Acworth et al. (1998a)).
The Con: The oxidation of the monoamine neurotransmitters is very interesting
because not only can they produce superoxide and other ROS but also a number
of biologically active and potentially toxic molecules. Together, these compounds
are being proposed to cause neuronal oxidative stress that may be one of the
mechanisms that eventually lead to neuronal degeneration. The catecholamine
neurotransmitters are notoriously unstable – either in the presence of transition
metals or when exposed to a basic pH – and undergo metal-induced “autooxidation,” producing reactive semiquinone intermediates, quinones, and ROS
(Bindoli et al. (1992); Miller et al. (1990, 1996)) (Figure 2.6). The reactive
intermediates can undergo intramolecular cyclization to form cytotoxic
aminochromes (e.g., dopachrome), polymerize to form neuromelanin, or react
with a variety of nucleophiles (e.g., cysteine) to produce a spectrum of potentially
neurotoxic compounds (Acworth et al. (1998a)).
Neuromelanin is a complex polymer (of oxidized catecholamine residues) bound
to lipofuscin granules. It is capable of binding Fe (III) and reducing it to
biologically available Fe (II) capable of producing hydroxyl free radicals (Bindoli
et al. (1989); Graham (1978); Graham et al. (1978) (Chapter 4). This finding, in
WWW.ESAINC.COM
58
conjunction with the discovery that neuromelanin can also generate other ROS,
has led some to hypothesize that it may play a role in progression of Parkinson’s
disease (Gerlach et al. (1994)). The brown/black skin melanins (eumelanins) are
pigments formed by the oxidation and polymerization of tyrosine, are devoid of
bound iron, and are actually ROS scavengers. Pheomelanins are either yellow or
red-brown pigments found in the skin and hair of redheaded people. These are
less effective radical scavengers and may even degrade with the formation of
superoxide upon exposure to strong light.
Control.
The cellular level of superoxide is maintained by the enzyme superoxide
dismutase (SOD). Several forms of SOD exist in higher animals and these will be
discussed further in Chapter 5. Together, these enzymes keep the cellular levels
of superoxide <10-11M (rat liver cytosol) to 10-10M (liver and heart mitochondria).
Measurement.
Superoxide can be monitored using a number of approaches (Halliwell and
Gutteridge (1999); Livovich and Scheeline (1997); McNeil et al. (1992); Riley et
al. (1991); Shoaf et al. (1991); Suzaki et al. (1994)). These include: measurement
of its spectrum using EPR at low temperatures; measurement of its absorbance
at 245nm; the use of differential pulse polarography; chemiluminescence
detection; the use of reporter molecules (e.g., cytochrome c, dianisidine,
epinephrine, luminol, nitroblue tetrazolium, and tetranitromethane); and
voltammetric detection.
Auto-oxidation and Redox Cycling Reactions.
Many biologically relevant compounds are reported to react spontaneously with
oxygen in a one- or two-electron process, producing superoxide and hydrogen
peroxide, respectively. These include carbohydrates (ascorbic acid, glucose,
glyceraldehydes, and glycoxidation processes), catechols, cysteine, hemoglobin
and myoglobin, lipids (cholesterol, polyunsaturated fatty acids, and lipid
peroxidation processes), and monoamines (Burkitt and Gilbert (1991); Ford et al.
(1993); Kachur et al (1998); Kon (1978); Mansouri and Perry (1987); Miyata et al.
(1998); Pryor et al. (1976); Saez et al. (1982); Sevanian and McLeod (1987);
Smith (1987); Thornalley et al. (1984); Tomoda et al. (1981); Wolff and Dean
(1987); Wolman (1975)). This process is called auto-oxidation and in the strictest
sense can be defined as the “spontaneous oxidation in air of a compound in a
process that does not require a catalyst” (Miller et al. (1990)).
WWW.ESAINC.COM
59
But do these auto-oxidation reactions really occur in vivo? It is unlikely (Miller et
al. (1990); Reilly and Aust, (1999)). Firstly, the one electron reduction of oxygen
is a thermodynamically unfavorable reaction (Eo’=-330mV) due to the energy
needed to add an extra electron to the partially filled π* orbitals of the triplet
dioxygen molecule (Reilly and Aust, (1999)). Therefore, the reduction of oxygen
will occur only if it is coupled with energetically favorable processes that can
drive the reduction reaction. Since the only biological molecules capable of
reducing dioxygen are the reduced flavins, the “auto-oxidation” of the compounds
mentioned above could not possibly produce superoxide and hydrogen peroxide.
Secondly, although the reduction of dioxygen to hydrogen peroxide by ascorbate
is favorable thermodynamically it is hindered kinetically due to spin restrictions
(Reilly and Aust, (1999)).
Ascorbic acid, lipids, thiols, etc. can promote the reduction of dioxygen, but only
in the presence of a transition metal catalyst. The transition metals are
characterized by incompletely filled 3d orbitals and depending upon their
complexation, can exist in a variety of spin states. Therefore, such redox-active
metal complexes can react with oxygen to form a superoxo-metal complex,
thereby reducing the triplet nature of the oxygen molecule, and relieving the spin
restriction for the reaction between oxygen and biomolecules (Reilly and Aust
(1999)). Compounds such as ascorbate can reduce Fe (III) in a one-electron
process, producing a radical species (ascorbyl radical) and Fe (II). The Fe (II)
can then react with oxygen (part of the Haber-Weiss reaction) producing
superoxide and Fe (III), and eventually leading to the formation of hydrogen
peroxide and hydroxyl free radicals; the radical species is no longer spinrestricted and can either reduce oxygen directly or form an addition reaction with
it (Reilly and Aust (1999)). The cycle continues until all the reductant is used up
and iron can no longer be reduced.
In redox cycling the reductant is continuously regenerated, thereby providing
substrate for the “auto-oxidation” reaction. For example, partially oxidized
compounds can be enzymatically reduced, enabling the auto-oxidative
generation of superoxide and other ROS to start again. Several enzymes (e.g.,
NADPH-cytochrome P450 reductase, NADPH-cytochrome b5 reductase [EC
1.6.2.2] NADPH-ubiquinone oxidoreductase [EC 1.6.5.3], and xanthine oxidase
[EC 1.2.3.2]), can reduce quinones into semiquinones in a single electron
process. The semiquinone can then reduce dioxygen to superoxide during its
oxidation to a quinone (Figure 2.7).
WWW.ESAINC.COM
60
OH
O2 - + H+
O2
O
Metal
OH
O2 - + H+
O2
O
Metal
OH
e.g. NADPHCytochrome P450
Reductase
e.g. NADPH-
O
Cytochrome P450
Reductase
DT-Diaphorase
Figure 2.7 The Involvement Of NADPH-Cytochrome P450 Reductase
And DT-Diaphorase In Redox Cycling.
A number of xenobiotics can undergo redox cycling, in part accounting for their
beneficial or detrimental activity in biological systems. Such compounds include
the bipyridyl herbicides (diquat and paraquat), which produce ROS and release
redox active iron from ferritin; the diabetogenic agent, alloxan; antibiotics (e.g.,
actinomycin D, mitomycin C and streptonigrin); antitumor drugs (e.g.,
anthracyclines, etoposides, tirapazamine, diaziridinylbenzoquinones); and the
hydroxylated metabolites of the antimalarial drug primaquine (Butler (1998);
Halliwell and Gutteridge (1999); Newsholme and Leech (1992); Vasquez-viva
and Augusto (1992)) (Figure 2.8).
Redox cycling is thought to play a role in carcinogenesis. For example, the
naturally occurring estrogen metabolites (the catecholestrogens) have been
implicated in hormone-induced cancer, possibly as a result of their redox cycling
and production of ROS (Yager and Liehr (1996)). Furthermore, the banned
synthetic estrogen, diethylstilbestrol, is believed to exert its carcinogenicity
through the production of ROS by redox cycling (Liehr et al. (1986); Wyllie and
Liehr (1997)). Roy et al. (1991) reported that diethylstilbestrol causes the
production of the mutagenic lesion 8-hydroxy-2’deoxyguanosine (Chapter 3).
Redox cycling can also cause DNA strand breakage. For example, redox cycling
of 2,5-dihydroxypyridine, a metabolite of 3-hydroxypyridine found in cigarette
smoke, can cause DNA strand scission (Kim and Novak (1990)).
WWW.ESAINC.COM
61
Cytochrome c (Fe 2+)
O2
H2 O
N
H3 C
Cyt c
Oxidase
N
CH3
O2 -
Oxidized Paraquat
Cytochrome c (Fe 3+)
H3C
N
NADPH
P450
Reductase
N
CH3
O2
Reduced Paraquat
NADPH
O2 -
NADPH
P450
Reductase
H3 C
N
N
CH3
Oxidized Paraquat
2H+, 2e-
O
O
Thioredoxin?
HN
HN
NH
O
NH
O
O
O
H
O
Alloxan
OH
Dialuric Acid
ROS
O2
Figure 2.8 The Redox Cycling Of Paraquat And Alloxan. Paraquat
undergoes one electron oxidation producing a paraquat radical and superoxide.
The paraquat radical can be reduced by either the electron transport chain
(cytochrome c) or by NADPH cytochrome P450 reductase in a process requiring
NADPH. In the islets of Langerhans of the pancreas alloxan is thought to
undergo a two-electron reduction by thioredoxin. In the presence of metals
dialuric acid undergoes oxidation with the production of superoxide, hydrogen
peroxide and hydroxyl free radicals.
In many cases redox cycling is deleterious to the organism and must be
prevented. DT diaphorase [(EC 1.6.99.2) also called NAD(P)H dehydrogenase
(quinone), NAD(P)H oxidoreductase, quinone reductase or azo-dye reductase] is
a flavoprotein that uses NADH or NADPH to reduce quinones, quinoneimines,
and nitrogen oxides in a two-electron process (Cadenas (1995)) (Figure 2.7). The
action of DT-diaphorase is to prevent redox cycling by removing quinones,
thereby preventing their partial reduction by other enzymes and generation of
superoxide. DT-diaphorase is a Phase II detoxifying enzyme that can be induced
WWW.ESAINC.COM
62
in a number of tissues by a wide variety of compounds, including dithiolethiones
and isothiocyanates (Chapter 4). Not only does DT-diaphorase inactivate
xenobiotics, it also plays a role in the activation of a number of quinonecontaining chemotherapeutic prodrugs (Rauth et al. (1997)).
Redox cycling is also essential to aerobic respiration and components of the
electron transfer chain (cytochromes and coenzyme Q10) redox cycle because
electrons are passed from NADH to the terminal electron acceptor, oxygen.
Unlike the examples given above, however, the redox cycling associated with
aerobic respiration is more tightly controlled and only a minor proportion of
electrons “leak”, producing ROS.
5. Hydrogen Peroxide.
Properties.
Hydrogen peroxide (H2O2) is a pale blue, viscous liquid with a melting point of
-0.9oC and a boiling point of +150oC. It is stable in the absence of reducing
agents. In the presence of such contaminants its half-life is of the order of
minutes to hours at 37oC under aqueous conditions, depending upon its
concentration and conditions. Hydrogen peroxide is formed in the single-electron
reduction of superoxide or the two-electron reduction of oxygen (Figure 2.2).
During single-electron reduction of superoxide, the extra electron enters the
remaining partially filled π*2p orbital (Figure 2.1). Consequently, the resulting
peroxide anion (O22-) has its π*2p orbitals completely filled. The peroxide anion is
not a radical and is therefore diamagnetic. It has a relatively weak, single
oxygen-oxygen bond. The peroxide anion exists only under extremely basic
conditions, so under physiological conditions it is protonated and exists as
hydrogen peroxide.
Formation.
Hydrogen peroxide is made in the laboratory by acidification of ionic peroxides
(e.g., barium peroxide). Industrially, it is made either by the catalytic reduction of
2-butylanthraquinone to 2-butylanthraquinol – which is then oxidized with oxygen
enriched air to hydrogen peroxide — or the oxidation of 2-propanol with oxygen
under slight pressure. Hydrogen peroxide is produced in vivo by the two-electron
reduction of oxygen or by superoxide dismutation (see above). As a result,
superoxide produced by the electron transport chains – cytochrome P450,
phagocytosis, etc. – will always produce hydrogen peroxide. It is also formed by
several oxidases (e.g., by monoamine oxidase as shown in Figure 2.5).
WWW.ESAINC.COM
63
Chemical Reactions and Biological Significance.
Hydrogen peroxide is thermodynamically unstable with respect to oxygen and
water and is readily decomposed by heat or by contact with finely divided solids
(e.g., manganese (IV) oxide and metals) and traces of alkali (Eqn 2.28).
Homolytic fission (e.g., by irradiation) yields the hydroxyl free radical (Eqn 2.29).
High levels of H2O2 (10-5-10-8M) have been reported in water obtained from UVprotected water stills. This can pose a problem for the measurement of the
hydroxyl free radical, especially if redox active metals are present (see below).
2H2O2 (g) → O2 (g) + H2O (g)
∆Go(298)= -126kJ mol-1 H2O2
H2O2 → 2HO•
Eqn 2.28
Eqn 2.29
Hydrogen peroxide can act as both a weak oxidizing and reducing agent. For
example, it acts as a weak oxidizing agent, converting sulfide, Fe (II) and iodide
ions into sulfate ions, Fe (III) and iodine, respectively (Eqn 2.30). Strong oxidizing
agents (e.g., silver oxide, and acidified potassium permanganate) force H2O2 to
assume the role of reducing agent (Eqn 2.31).
H2O2 + 2I- + 2H+ → 2H2O + I2
Ag2O + H2O2 → 2Ag + H2O + O2
Eqn 2.30
Eqn 2.31
Like superoxide, hydrogen peroxide is not particularly reactive (the second order
rate constants are typically 101 to 105 M-1s-1). Under physiological conditions, the
reactions of H2O2 are mainly confined to its oxidizing ability. It can oxidize thiols
and by so doing, inactivate enzymes that contain an essential thiol group
(Chapter 3). As hydrogen peroxide is fairly stable and can readily pass through
membranes it can react with biological molecules far removed from its site of
production (Makino et al. (1994)).
A significant problem for living organisms is the consequence of the reaction
between hydrogen peroxide and oxidizable metals, the Fenton reaction. The
Fenton reaction originally described the oxidation of an α-hydroxy acid (tartaric
acid) to an α-keto acid in the presence of hydrogen peroxide (or hypochlorite)
and low levels of iron salts (Fenton (1876, 1894)).
Although the Fenton reaction is often presented as a straightforward equation
(Eqn 2.32) this is a gross simplification because many reactions are possible
(e.g., Eqns 2.33 to 2.35). For example, when the Fenton reaction is carried out in
the presence of HCl, alkenes are chlorohydroxylated (Sawyer et al. (1995)).
Readers wanting a more comprehensive explanation of the Fenton reaction are
WWW.ESAINC.COM
64
referred to the following references Goldstein et al. (1993); Halliwell and
Gutteridge (1990, 1992); Koppenol (1993); Liochev (1999); and Wardman and
Candeias (1996). Although hydroxyl free radicals are thought to be the major prooxidant species formed there remains considerable controversy about whether
they exist in a free form (Wardman and Candeias (1996) and references therein).
Pro-oxidant metal species have also been proposed as the pro-oxidant species
(Buxton and Mulazzani (1999)). For example, the ferryl radical (e.g., Fe2+-O),
where iron is in its IV valency, may also be formed in conjunction with the
hydroxyl free radical (Eqns 2.36 and 2.37). However, it is doubtful that ferryl
radicals are the primary pro-oxidant species formed in vivo (Halliwell and
Gutteridge (1999); Koppenol (1993)). Similarly, a perferryl species (e.g., Fe2+-O2
⇔ Fe3+-O2•-) may be formed when Fe (III) reacts with superoxide (e.g., as part of
the Haber-Weiss reaction); however, it is unlikely to be the major reactive
species (Eqn 2.38) (Halliwell and Gutteridge (1999)). Qian and Buettner (1999)
have challenged these ideas, suggesting that an unknown “Fe2+ + O2” species
was indeed capable of initiating free radical oxidations. Their finding was based,
in part, on the fact that the Fenton reaction has only a small rate constant (103105 M-1s-1), while the reaction of Fe (II) with superoxide is much greater (106-107
M-1s-1). Therefore, under physiological conditions, the latter reaction will
effectively limit the availability of Fe (II) to take part in the Fenton reaction. Qian
and Buettner reported that when the [oxygen]/[hydrogen peroxide] ratio <10 the
Fenton reaction dominated, but when this ratio >100 (under physiological
conditions this ratio ~1000), then the Fenton reaction played only a subservient
role to the “Fe2+ + O2” species.
Oxidant + reduced metal → oxidized metal + superior oxidant
e.g., H2O2 + Fe2+ → Fe3+ + OH- + HO•
HO• + Fe2+ → OH- + Fe3+
HO• + H2O2 → H2O + H+ + O2•O2•- + Fe3+ → O2 + Fe2+ (part of the Haber-Weiss reaction)
H2O2 + Fe2+ → FeOH3+ + OHor H2O2 + Fe2+ → FeO2+ + H2O
FeO2+ + H2O2 → Fe2+ + H2O + O2
O2•- + Fe3+ ↔ [Fe3+-O2- ↔ Fe2+-O2] ↔ O2 + Fe2+
Eqn 2.32
Eqn 2.33
Eqn 2.34
Eqn 2.35
Eqn 2.36
Eqn 2.37
Eqn 2.38
Currently, many researchers are not convinced of iron’s role in pro-oxidant
production and suggest that with the exception of iron-related disorders, there is
little or no direct proof that iron plays an important role in the Fenton reaction in
vivo (Chapter 4).
Several metals besides iron are capable of undergoing changes in oxidation
status (e.g., copper, chromium, vanadium, etc.) and can reduce hydrogen
peroxide to hydroxyl free radicals. Whether they are involved in Fenton-like
reactions in vivo is still a matter of debate (Masarwa et al. (1988)). Interestingly,
WWW.ESAINC.COM
65
Cu (I) salts can react with hydrogen peroxide to form both hydroxyl free radicals
and also the powerful oxidizing agent Cu (III) (Eqn 2.39). This strongly oxidizing
form of copper can be also formed by the action of Cu (II) ions with superoxide.
H2O2 + Cu+ → Cu2+ + HO• + OH-
Eqn 2.39
Hydrogen peroxide is beneficial too. Like superoxide, it plays an important role in
the immune response. It does so both directly by inhibiting key enzymes within
the pathogen and indirectly as the “safe” precursor to the hydroxyl free radical
(Chapter 4). Hydrogen peroxide is also essential for the synthesis of thyroxine in
the thyroid gland (Dupuy et al. (1991)).
The typical steady-state cellular hydrogen peroxide concentration is estimated to
be 10-7-10-9M in the liver and 10-5M in the human eye lens. These concentrations
represent a balance between hydrogen peroxide production and destruction. Its
level is primarily controlled by two groups of enzymes, the catalases (Eqn 2.40)
and glutathione peroxidases (Eqn 2.41) (see Chapter 4).
2H2O2 → 2H2O + O2
H2O2 + 2GSH → 2H2O + GSSG
Eqn 2.40
Eqn 2.41
Measurement.
In the laboratory, hydrogen peroxide can be measured using chemical titration
with acidified potassium permanganate, but this approach is not selective and is
too insensitive for its measurement in vivo. Hydrogen peroxide can be measured
in biological systems by peroxidase-based methods with fluorometric detection
(Corbett (1989) and by HPLC-chemiluminescence methods (Yamamoto and
Ames (1987)). It can also be determined by the measurement of evolved oxygen
using an oxygen electrode following the addition of catalase (Halliwell and
Gutteridge (1999)), or evolved 14CO2 using scintillation counting when it reacts
with labeled 2-oxoglutarate (Varma (1989)). Hydrogen peroxide is
electrochemically active and can be measured voltammetrically in “real time,”
using either a platinum-disk (Yokoyama et al. (1998)) or enzyme-modified
electrodes (Livovich and Scheeline (1997); Tatsuma et al. (1992, 1994)).
WWW.ESAINC.COM
66
6. The Hydroxyl Free Radical.
Properties.
The hydroxyl free radical (HO•) is the most reactive ROS formed in vivo. It has a
half-life of 10-9 to 10-10 s and shows typical second-order rate constants of 109 to
1010 M-1s-1. The hydroxyl free radical is formed by the single electron reduction of
the peroxide ion. During this process the extra electron enters the empty σ*2p
molecular orbital. The single oxygen-oxygen bond of the peroxide ion is
weakened and cleaves, forming the hydroxyl free radical and hydroxide ion
(Figure 2.2). The addition of two electrons to the peroxide ions also cleaves the
oxygen-oxygen bond but, in this case, two oxide (O2-) ions are formed.
Formation.
The hydroxyl free radical can be formed by a number of processes including the
Fenton reaction, the Haber-Weiss reaction, and the homolytic fission of water
molecules (e.g., by ionizing radiation). It can also be produced by the
decomposition of ozone under aqueous conditions (Table 2.2) (Hoigne and
Bader (1975)), the microsomal ethanol oxidizing system (part of the endoplasmic
reticulum), and the reaction between the superoxide radical anion and
hypochlorous acid (Eqn 2.26) (Candeias et al. (1993)). Typical steady-state
levels in vivo are ~10-20M.
Chemical Reactions and Biological Significance.
The hydroxyl free radical is extremely reactive. It will react with most, if not all,
compounds found in the living cell (including DNA, proteins, lipids and a host of
small molecules). The hydroxyl free radical is so aggressive that it will react
within 5 (or so) molecular diameters from its site of production. The damage
caused by it, therefore, is very site specific (Pryor (1986)). The reactions of the
hydroxyl free radical can be classified as hydrogen abstraction, electron transfer,
and addition (Figure 2.9).
•
•
•
Hydrogen abstraction, which typically occurs with aliphatic compounds,
causes lipid peroxidation and DNA damage (Chapter 3).
Electron transfer produces secondary radicals of varying reactivity, such
as the carbonate radical anion.
Aromatic compounds typically react with the hydroxyl free radical by
addition. The products that are formed depend upon the species being
attacked and the reaction conditions. For example, the fast addition of the
hydroxyl
free
radical
to
benzene
produces
the
unstable
hydroxycyclohexadienyl radical. This can regain aromatic stability by
either dimerization or oxidation (Kaur and Halliwell (1994b)). The action of
WWW.ESAINC.COM
67
the hydroxyl free radical with a substituted benzene can produce a
spectrum of products. For example, the reaction between salicylic acid (2hydroxybenzoic acid) and the hydroxyl free radical produces 2,3- and 2,5dihydroxybenzoic acid and the decarboxylation product, phenol (Figure
2.9). Tyrosine undergoes dimerization with the production of dityrosine or
oxidation forming 3,4-dihydroxyphenylalanine (Chapter 3). Aromatic
hydroxylation is favored by the presence of oxygen, Fe (III), and Cu (II)
ions, whereas decarboxylation is favored by a lack of these compounds
(Halliwell and Gutteridge (1999)). Consequently, under physiological
conditions aromatic hydroxylation tends to be the predominant reaction.
The reaction between the hydroxyl free radical and an aromatic compound
is referred to as scavenging, and is sometimes used to trap this prooxidant prior to detection (Chapter 1 and see below). Readers should be
aware that some of the addition reactions of the hydroxyl free radical are
mimicked by peroxynitrite (see below).
The formation of the hydroxyl free radical can be disastrous for living organisms.
Unlike superoxide and hydrogen peroxide, which are mainly controlled
enzymatically, the hydroxyl free radical is far too reactive to be restricted in such
a way – it will even attack antioxidant enzymes. Instead, biological defenses
have evolved that reduce the chance that the hydroxyl free radical will be
produced and, as nothing is perfect, to repair damage. Redox active metals are
chelated (Chapter 4); hydrogen peroxide is catabolized enzymatically. The repair
of damaged molecules (e.g., enzymatic repair processes) will be discussed
further in Chapter 3. Even though low molecular weight antioxidants readily react
with hydroxyl free radicals it is doubtful that they play an important role in
controlling its level (Chapter 4). Remember, for an antioxidant to be effective it
would have to occur at the site of hydroxyl free radical production and be at
sufficient (probably unphysiological) concentration to compete with all the other
chemical species for reaction with this pro-oxidant.
WWW.ESAINC.COM
68
O2
a) ABSTRACTION
O2
C
OH
H
H
C2H5
C
C2H5
HO
OH
OH
C
C2H5
H
Peroxyl Radical
X2
H
Propanol
H
Hydroxypropyl
Radical
C2H5
C
OH
C2H5
C
OH
H
Hexan-3,4-Diol
b) ADDITION
OH
O
OH
OH
HO
O
H
R
R
R
Tyrosyl
Tyrosine
H
X2
-H2O
H
R
O
H
R
Di-Tyrosine
Radical
(Keto form)
OH
OH
R
R
Di-Tyrosine
c) ELECTRON TRANSFER
(Enol form)
HO
+ Cl -
Cl
HO
+ CO 3-
CO3- + OH -
+
OH -
Figure 2.9 Some Reactions Of The Hydroxyl Free
Radical. The Hydroxyl Free Radical Can React With
Molecules By Hydrogen Abstraction (A), Electron
Transfer (B) Or Addition (C). In This Figure,
Abstraction Forms A Carbon-Based Radical Capable
Of Reacting With Another Radical (E.G., Oxygen Or
Even Itself [Dimerization]) Through The Formation Of A
σ-Bond. Addition Forms A π-Radical That Can Regain
Aromatic Stability Through Dimerization To Dityrosine.
Electron Transfer Can Produce Very Reactive Radicals
Such As The Chlorine Radical (Cl•) And Carbonate
Radical Anion (CO3•-).
WWW.ESAINC.COM
69
Measurement.
A variety of approaches, differing in their specificity, sensitivity, applicability and
ease of use are used in to detect the hydroxyl free radical (See Halliwell and
Gutteridge (1999) and references therein). So far EPR and HPLC-based
approaches have proven to be the most useful.
EPR: Under special circumstances, EPR can be used directly to measure
hydroxyl free radicals (Halliwell and Gutteridge (1999)). However, it is more
common to use a spin trap, as discussed in Chapter 1. The use of EPR and spin
traps to measure the hydroxyl free radical is well documented in vitro but care
must be taken when interpreting data as the spin trap adduct themselves may
produce ROS (Finkelstein et al. (1980); Floyd (1983); Kaur et al. (1981);
Yamazaki and Piette (1987)). The measurement of the hydroxyl free radical in
vivo is particularly challenging. Interestingly, Dugan et al. (1995) coupled in vivo
microdialysis sampling procedures with on-line EPR to study free radical
production following focal cerebral ischemia-reperfusion in spin trap-treated
animals. Although this approach showed promise, the authors did caution about
spin trap stability, toxicity, and the spontaneous formation of the spin trap-HO•
adduct.
Some spin trap adducts are also electrochemically active. Consequently, HPLCECD is being used currently to overcome the sensitivity and quenching problems
associated with EPR (Floyd et al. (1984a,b), Iwahashi (1996); Motohashi and
Mori (1989); Stronks et al. (1984); Towel and Kalyanaraman (1991)). Fast-scan
voltammetry is also being used to explore the reaction mechanisms of some spin
traps (Baur et al. (1996)).
WWW.ESAINC.COM
70
COOH
COOH
OH
OH
OH
+ CO2
O2N
NO2
2-Hydroxy-5-NitroBenzoic Acid
NO2
2-Hydroxy-3-Nitro
Benzoic Acid
2-Nitrophenol
CO.Glucuronate
OH
ONO2H
Salicyl acyl
glucuronide
Glucuronidation
COOH
O.Glucuronate
COOH
Salicyl phenolic
glucuronide
OH
CONHCH 2COOH
OH
Salicylic acid
Salicyluric acid
HO
Cytochrome
P450
COOH
COOH
OH
OH
OH
+ CO2
OH
OH
2,3-Dihydroxybenzoic acid
(~49%)
Catechol
(~11%)
HO
2,5-Dihydroxybenzoic acid
(~40%)
Figure 2.10 Metabolism Of Salicylate And Its Reaction With The
Hydroxyl Free Radical And Peroxynitrite. (The % represent the
abundance of the products during the reaction of salicylic acid with the
hydroxyl free radical in vitro (Kaur and Halliwell (1994b)).
WWW.ESAINC.COM
71
HPLC: The other major approach to studying the formation of the hydroxyl free
radical in vitro and in vivo is based on HPLC separation. As with EPR, the
hydroxyl free radical is too reactive to be measured directly. It must first be
trapped in a stable form that is also amenable to HPLC analysis so such
methods are indirect. The production of the hydroxyl free radical is inferred from
the abundance of a product formed when this radical is scavenged, by either an
endogenous substrate or an administered reagent (see reviews by Acworth et al.
(1997, 1998a,b); Halliwell and Gutteridge (1999)). See Table 2.5.
An issue with many of the HPLC-based approaches used is that often only the
reaction products are quantified. A change in the level of a product is assumed to
reflect a change in radical production. Obviously, this is not always the case as a
change in the level of a marker could entirely be due to altered availability of the
scavenging agent (e.g., hepatic metabolism or excretion). A better approach is to
simultaneously measure both the scavenging agent and products, thereby
permitting normalization of the data.5
Target
Endogenous
Markers
Product
Comments
Creatinine
Creatol
and methyl-guanidine
Products are used as markers of
oxidative stress.
2’-Deoxy-cytidine
5-Hydroxy2’deoxycytidine
2’-Deoxyguanosine
8-Hydroxy 2’-deoxyguanosine
Histidine
2-Oxohistidine
(free and protein
Can be measured by GC-MS or
HPLC with UV or ECD. HPLC-ECD is
ideally suited to measure low tissue
levels (Chapter 4). This marker is not
exclusive to the hydroxyl free radical.
Urinary levels may be of use as a
global oxidative stress marker.
Can be measured using ELISA,
HPLC-ECD, GC-MS, GC-MS-MS and
TLC-32P approaches (Chapter 4).
HPLC-ECD is ideally suited to
measure low tissue levels. Not
specific as it can be formed by the
action of singlet oxygen also. Tissue
DNA requires careful hydrolysis to
prevent artificial formation. It is
extremely difficult to measure in urine.
Can be measured using HPLC-based
methods. Exists as both free and
References
Aoyagi et al.
(1998a,b);
Nakamura et al.
(1996); Yokozawa
et al. (1997)
Wagner et al.
(1992)
Schneider et al.
(1989); Floyd et al.
(1986)
Lewisch and
Levine (1995);
5
One exception is microdialysis. This approach is used to monitor analyte levels in the living organism in real time. As the
scavenger is perfused through a microdialysis probe directly into the tissue it will be unaffected by peripheral metabolism.
Consequently, data need not be normalized.
WWW.ESAINC.COM
72
bound)
Methionine
Methionine sulfoxide
(free and protein
bound)
L-Phenyl-alanine
(L-Phe)
o-, m-Tyrosine
(free and protein
bound)
L-Tyrosine
3,4- (and 2,4)-LDOPA)
(free and protein
bound)
L-Tyrosine
Dityrosine
(free and protein
bound)
protein bound forms (Chapter 4). The
protein bound form needs to be
hydrolyzed before analysis by HPLC.
Can be measured using HPLC-based
methods (Chapter 4). It exists as both
free and protein bound forms. The
protein bound form needs to be
hydrolyzed before analysis by HPLC.
It can be formed by hydroxyl free
radicals and other ROS.
Both target and product can be
measured using HPLC-UV, but this
approach may not be adequate for
most tissue analyses. Improved
sensitivity can be obtained using
either HPLC-fluorescence or HPLCECD following derivatization (e.g.,
OPA/βME). All tyrosine isomers are
electrochemically active and can be
measured directly using HPLC-ECD.
Protein bound targets require
hydrolysis before analysis by HPLC.
Both target and product can be
measured using HPLC-UV, but may
not be adequate for most tissue
analyses. Improved sensitivity can be
obtained using either HPLCfluorescence or HPLC-ECD following
derivatization (e.g., OPA/βME). All
tyrosine isomers are
electrochemically active and can be
measured directly using HPLC-ECD.
Protein bound targets require
hydrolysis before analysis by HPLC.
Free 3,4-isomer is also formed
enzymatically by tyrosine hydroxylase
and this will limit the use of this assay
in catecholaminergic tissue. The 2,4isomer is only a minor product and is
not commercially available. Protein
bound 3,4-L-DOPA requires
hydrolysis before analysis by HPLC.
Can be measured using GC-MS, TLC
and HPLC with UV, fluorescence or
ECD (Chapter 4). HPLC-ECD is most
often used due to the very low tissue
level of this marker. May be formed
from other pathways including
reaction of tyrosine with either
hypochlorous acid or peroxynitrous
acid. Used as a marker for hydrogen
peroxide stress. Protein bound form
requires hydrolysis before analysis by
HPLC.
WWW.ESAINC.COM
Uchida and
Kawakishi (1990,
1993)
Levine et al.
(1996); Li et al.
(1995a,b); Vogt
(1995)
Ishimitsu et al.
(1986); Nair et al.
(1995); Sontag et
al. (1997)
Hensley et al.
(1997)
Giulivi and Davies
(1994); Heinecke
et al. (1993);
Huggins et al.
(1993);
Ischiropoulos et al.
(1992);
Leeuwenburgh et
al. (1997); van der
Vleit (1995);
Vissers and
Winterbourne
(1991)
73
Exogenously
Administered
Agent and In
Vitro Examples
4-Amino-salicylic
acid (4HAS)
5-Amino-salicylic
acid (5HSA)
N-Acetyl-4-aminosalicylic acid;
dihydroxy-4aminobenzoic acids,
and many other
analytes
N-Acetyl-5-aminosalicylic acid;
dihydroxy-5aminobenzoic acids,
and many other
analytes
Aniline
o- and pAminophenol
Dimethylsulfoxide
DMPO (5,5dimethylpyrroline-Noxide)
Methanesulfinic acid
4-Hydroxybenzoic acid
(4HBA)
3,4-DHBA (minor
amounts of 2,4isomer formed)
2-Methyl-2nitroso-propane
(t-nitrosobutane)
Depends on system
being studied
PBN (α-phenyl tert-butylnitrone
Aminoxyl and other
adducts depending
on the system being
investigated
Catechol, resorcinol
Phenol
DMPO-OH, DMPOOH2
Can be measured using EPR and
HPLC-based approaches. Can
produce complex chromatograms.
Allgayer et al.
(1992)
Can be measured using HPLC with
UV or ECD. Chromatograms may be
complex due to the number of
analytes produced. This agent has
several metabolic effects including:
reduction of leukotriene production,
inhibition of interleukin-1 release,
inhibition of prostaglandin synthetase
or lipoxygenase, and interference of
antibody production.
Can be measured using HPLC with
UV or ECD. Aniline is toxic and is not
practical for biological experiments.
So far only in vitro studies using
HPLC-UV have been reported
Can be measured using EPR but the
product can be quenched in vivo.
HPLC-ECD can overcome the
problems associated with EPR and
offers better sensitivity.
Similar in reactivity to salicylate but
with less physiological activity. Only
one major product formed permitting
lower LODs than Sal (where signal is
split between two products). Intestinal
microbes readily form 4HBA, which
may be problematic, if gut is
damaged. For microdialysis perfusion
experiments the presence of metals in
the fluid path must be minimized in
order to prevent the spontaneous
production of 3,4-DHBA.
So far only in vitro studies using
HPLC-UV have been reported.
Fischer and Klotz
(1994); Palumbo
et al. (1997)
This is usually measured using EPR.
It can be measured using HPLC-ECD
but the chromatography can be
complicated.
Can be measured using HPLC-UV or
ECD but is not practical for biological
experiments.
WWW.ESAINC.COM
Radzik et al.
(1983)
Fukui et al. (1993)
Floyd et al.
(1984a)
Acworth et al.
(1998b);
Bogdanov
(1998a,b,c);
Montgomery et al.
(1995);
Ste-Marie et al.
(1996)
Hiraoka et al.
(1989, 1990);
Inami et al. (1986,
1987)
Chen et al. (1990;
1994); Cheng et
al. (1993); Stronks
et al. (1984)
Floyd et al.
(1984b); Radzik et
al. (1983)
74
D-Phe
o-, m- D-Tyrosine
See L-Phe above. D- and L-forms
cannot be resolved unless a chiral
column is used. D-Phe to L-Phe
isomerization will deplete target
molecule thereby affecting its
availability for reaction with the
hydroxyl free radical (Acworth et al.
(1997)). This approach may not be
practical for study of the central
nervous system as L-Phe can affect
dopamine synthesis and release in
brain (During et al. (1988))
See L-Phe above. May not be
practical for study of the central
nervous system as L-Phe can affect
dopamine synthesis and release in
brain (During et al. (1988)). The rate
of reaction is slower than for Sal. May
be a useful marker of food irradiation.
Tyrosine isomers may also be formed
by the action of peroxynitrite on LPhe.
L-Phe
o-, m-Tyrosine
isomers
4-POBN (α-(4pyrisyl-1-oxide)N-tertbutylnitrone)
4-POBN radical
adduct
Usually measured using EPR. HPLCECD permits the study of reaction
mechanisms. This is an in vitro
method only.
Salicylic acid
(Sal)
2,3- and 2,5-DHBA
Can be measured using GC-MS and
HPLC with a variety of detection
systems including UV, ECD, ECD
with UV, ECD with fluorescence, and
MS. The 2,3-isomer better reflects
hydroxyl free radical production as
the 2,5- isomer is formed by
cytochrome P450 (Figure 2.10). The
2,5-DHBA isomer is also formed due
to singlet oxygen activity
(Kalyanaraman et al. (1993)). Sal has
physiological affects albeit at higher
concentrations. Perfusion through a
microdialysis probe can lead to
spontaneous production of DHBA
isomers. Both Sal and the DHBAs can
bind ferric iron thereby perturbing the
iron-dependent generation of the
hydroxyl free radical. DHBAs may be
formed by the action of peroxynitrite
on Sal. Peroxynitrite also reacts with
Sal to form the marker, 2-hydroxy-5nitrobenzoate (Skinner et al. (1996)).
See Figure 2.11.
WWW.ESAINC.COM
Kaur and Halliwell
(1994a,b)
Gelvan et al.
(1992); Kaur and
Halliwell
(1994a,c); Kaur et
al. (1988, 1996);
Halliwell and Kaur
(1997)); Ishimitsu
et al. (1984);
Karam and Simic
(1988); Liu (1993),
Ramezanian et al.
(1996); Sontag et
al. (1997); Sun et
al. (1993); van der
Vliet et al. (1994)
Cheng et al.
(1993); Iwahashi
(1996);
Stoyanovsky et al.
(1999)
Acworth et al.
(1997); Bickford et
al. (1999); Blandini
et al. (1999); Floyd
et al. (1984b);
Halliwell and Kaur
(1997); Liu et al.
(1999); Luo and
Lehotay (1997);
McCabe et al.
(1997);
Tabatabaei and
Abbott (1999)
75
L-Tyrosine
3,4-Dihydroxyphenylalanine
L-Tryptophan
Complex mixture of
metabolites
Can be measured using HPLC with
UV or ECD. This has been used for in
vitro studies only.
Can be measured using HPLC with
UV or ECD. This has been used for in
vitro studies only.
Ramezanian et al.
(1996)
Maskos et al.
(1992)
Table 2.5 Endogenous Markers And Exogenous Agents Used To Study Hydroxyl
Free Radical Formation. Based Upon Acworth et al. (1998a).
The major problem with measuring endogenous markers is interpreting what
their levels really represent. Many of these markers are not produced exclusively
by the hydroxyl free radical and can be formed by other pro-oxidants (e.g., singlet
oxygen and peroxynitrite). Furthermore, as the hydroxyl free radical will react
with any compound it encounters, the measurement of just one endogenous
marker is likely to underestimate the total production of this pro-oxidant.
Currently, some endogenous markers are being proposed as a useful measures
of total “oxidative stress” e.g., 8-hydroxy-2’deoxyguanosine in urine (Chapter 3).
The use of exogenous scavengers also has limitations. For an exogenous
scavenger to be effective, enough of it must get to the site of hydroxyl free radical
production in order to compete with the other compounds capable of reacting
with this radical. Remember the radical scavenger is unlikely to react with all
hydroxyl free radicals produced so, like endogenous markers, it will
underestimate the total production of this pro-oxidant. Consequently, scavengers
are usually given at high doses (typically hundreds of milligrams per kilogram of
body weight). This may be a problem if the scavenger is toxic, suffers from
distribution problems or possesses adverse biological activity. The ideal
scavenger must be non-toxic, have limited or no biological activity, readily reach
the site of hydroxyl free radical production (i.e., pass through barriers such as the
blood-brain barrier), react rapidly with the free radical, be specific for this radical,
and neither the scavenger nor its product(s) should undergo further metabolism.
As it may be appreciated, no scavenger has successfully fulfilled all of these
criteria.
Of all of the approaches outlined in Table 2.5, the use of salicylate and 4hydroxybenzoic acid as hydroxyl free radical-scavenging agents are by far the
most common (reviewed in Acworth et al. (1997; 1998a,b)). See ESA Application
Notes for further experimental details (70-1749 Measurement of The Hydroxyl
Free Radical; 70-4820 Alternate Method for the Measurement of The Hydroxyl
Free Radical). Salicylic acid reacts rapidly (five times faster than phenylalanine
and 22 times faster than guanosine) with the hydroxyl free radical, producing
readily quantifiable products (2,3- and 2,5-dihydroxybenzoic acids (Figure 2.10)).
It readily distributes throughout the body, even passing through the blood-brain
barrier, making it a useful tool to study central metabolism. Unfortunately it also
suffers from some problems:
WWW.ESAINC.COM
76
•
•
•
It possess physiological activity if used at too high a concentration, so the
amount administered in vivo must be chosen carefully;
A growing body of evidence suggests that hydroxylated products are formed
as a consequence of peroxynitrite attack (see below). One approach to
distinguish between true hydroxyl free radical-dependent aromatic
hydroxylation and that involving peroxynitrite is to measure aromatic nitration
products along with the DHBA isomers (Halliwell and Kaur (1997)); and
A number of publications now report a significant level of spontaneous
formation of DHBAs when salicylic acid is dissolved in microdialysis perfusion
medium. This situation can be further exacerbated by metals in the flow path
(e.g., the syringe needle and metal in the probe itself) (Acworth et al. (1998b);
McCabe et al. (1997a,b); Montgomery et al. (1995)). Whether conditions do
exist in which salicylic acid can be effectively used as a hydroxyl free radicalscavenger, while coupled to microdialysis perfusion, is yet to be determined.
WWW.ESAINC.COM
77
I. post amphetamine/saline rat striatal tissue sample
II. post amphetamine/SAL rat striatal tissue sample
III. 650 nM standard
2,3-DHBA
2,5-DHBA
A
II
I
III
I. post amphetamine/saline rat striatal tissue sample
II. post amphetamine/SA rat striatal tissue sample
III. 360 µM standard
SA
B
II
III
I
Figure 2.11 Chromatograms Showing The Simultaneous Measurement Of
DHBAs And Salicylic Acid For The Detection Of The Hydroxyl Free Radical
In Rat Striatal Tissue. The DHBAs were selectively detected on the first coulometric
electrode (A; +250mV, 100nA) while salicylic acid (SA) was measured on the second electrode
(B; +750mV, 20µA gain).
The isocratic system consisted of a pump, an autosampler, a thermal chamber and a Coulochem
III detector.
LC Conditions:
Column:
Mobile Phase:
DHBA-250.
50mM Sodium acetate, 50mM Citric Acid,
25% Methanol, 5% Isopropanol, pH 2.5 with Phosphoric Acid.
Flow Rate:
0.5mL/min.
Temperature:
Ambient.
Injection Volume:
10 µL.
Guard Cell, Model 5020
EGC = +775mV
Analytical Cell, Model 5010
E1 = +250mV; E2 = +750mV
See 70-1749 Measurement of The Hydroxyl Free Radical for more details.
WWW.ESAINC.COM
78
A variety of HPLC-based procedures have been used to measure salicylic acid
and DHBAs (Acworth et al. (1997, 1998b); McCabe et al. (1997a)). UV detection
tends to be too insensitive for routine biological work, requiring use of large
amounts of salicylic acid in order to render the DHBAs detectable by HPLC-UV.
Unfortunately, high doses can exacerbate physiological problems. HPLC-ECD is
much more sensitive and selective and has been used to supplement the UV
approach. In this example the upstream ECD is used for the sensitive
measurement of the DHBAs while the downstream UV detector measures the
greater abundance of salicylic acid (Jen et al. (1998); Sloot and Gramsbergen
(1995)). Analytical approaches requiring two detectors are cumbersome,
expensive, and unnecessary. HPLC-ECD can also be used to measure both
precursor and product. Both salicylic acid and DHBAs can be measured
simultaneously on a single amperometric thin-layer electrode but, due to the high
applied potential necessary to measure salicylate, detection can suffer from
noise and co-elutions (Floyd et al. (1984b, 1986); Kaur and Halliwell (1994b)).
Perhaps a better approach to overcome these problems is to use a dual
coulometric detector that makes use of the inherent differences in the
electrochemical behavior of the DHBAs and salicylate (McCabe et al. (1997a,b)
(Figure 2.11). A major advantage of the high selectivity and sensitivity of this
approach is that less salicylic acid has to be administered to the animal, thereby
minimizing possible physiological side effects. A similar coulometric approach
this time using 4-hydroxybenzoic acid as the scavenging agent can also be used
to measure the presence of hydroxyl free radicals (Figure 2.12). This has several
advantages over the salicylic method as discussed in Table 2.5.
WWW.ESAINC.COM
79
I. basal microdialysis sample
II. post 4-HBA microdialysis sample
III. 10 nM standard
5-HIAA
DOPAC
3,4-DHBA
A
I
II
B
4-HBA
III
I. basal microdialysis sample
II. post 4-HBA microdialysis sample
III. 10 µM standard
II
III
I
Figure 2.12 Chromatograms Showing The Simultaneous Measurement Of
3,4-DHBA And 4-Hydroxybenzoic Acid (4-HBA) For The Detection Of The
Hydroxyl Free Radical In Rat Brain Microdialysis Samples. 3,4-DHBA and
neurotransmitter metabolites were selectively detected on the first coulometric electrode (A;
+150mV, 10nA) while 4-HBA was measured on the second electrode (B; +700mV, 10µA gain).
The isocratic system consisted of a pump, an autosampler, a thermal chamber and a Coulochem
III detector.
LC Conditions:
Column:
Mobile Phase:
Flow Rate:
Temperature:
Injection Volume:
Analytical cell, Model 5011:
Super ODS (4.6 x 50mm; 2µm) TosoHaas.
100mM Sodium Phosphate Buffer (pH2.8) Containing
Methanol (6.5% v/v).
1.0mL/min.
29oC .
20µL.
E1 = +150mV; E2 = +700mV
See 70-4820 Alternative Method for the Measurement of The Hydroxyl Free Radical for more
details.
WWW.ESAINC.COM
80
3-hydroxybenzoic acid
2,4-dihydroxybenzoic acid
2,3-dihydroxybenzoic acid
4-hydroxybenzoic acid
2,5-dihydroxybenzoic acid
10.0
3,5-dihydroxybenzoic acid
Response (µA)
15.0
resorcinol
3-nitrotyrosine
catechol
3,4-dihydroxybenzoic acid
tyrosine
meta-tyrosine
20.0
hydroquinone
ortho-tyrosine
homogentisic acid
3-chlorotyrosine
In recent papers the chromatography used to measure the DHBAs on a single
thin-layer electrode was extended to simultaneously measure a variety of
monoamines and metabolites in brain and CSF tissues (e.g., Sloot and
Gramsbergen (1995)). When dealing with such complex chromatography it is
vitally important to fully characterize each eluting peak (both chromatographically
and voltammetrically) to ensure its authenticity, and to avoid a possible co-elution
or misidentification. Unfortunately, such approaches are usually incredibly
tedious and time-consuming, unless coulometric electrode array detection is
used. Beal used a gradient coulometric array method to measure salicylic acid,
the DHBAs, 3-nitrotyrosine, 3-aminotyrosine and twenty-four neurochemicals
simultaneously (Beal et al. 1990, 1995). Recently, Acworth et al. (1998b)
developed a coulometric array method capable of resolving a number of possible
markers of oxidative stress (Figure 2.13).
5.0
0.0
mV
830
810
670
630
570
500
450
400
R1
R2
10.0
20.0
30.0
40.0
50.0
60.0
Retention time (minutes)
Figure 2.13 Isocratic Coulometric Array Chromatogram Showing The
Simultaneous Measurement Of Several ROS And RNS Markers And
Precursors (standards at 10µg/mL on column each). Dityrosine elutes just after tyrosine but is
not shown for clarity.
WWW.ESAINC.COM
81
The isocratic system consisted of a pump, an autosampler, a thermal chamber and a CoulArray
detector.
LC Conditions:
Column:
Mobile Phase:
TSKgel ODS-80TM (4.6 x 250mm; 5µm) TosoHaas.
20mM Sodium Phosphate Buffer (pH3.2) Containing
Methanol (8% v/v).
1.0mL/min.
31oC .
20µL.
+400, +450, +500, +570, +630, +670, +810 and +830mV
Flow Rate:
Temperature:
Injection Volume:
Array Potentials:
NITROGEN AND THE REACTIVE NITROGEN SPECIES (RNS).
1. Nitrogen.
Properties.
Nitrogen is a colorless and odorless diatomic gas that occurs in the atmosphere
to the extent of about 78% by volume. Nitrogen has a melting point of -210oC and
a boiling point of -196oC.
Nitrogen is the first member of Group 5B of the periodic table and possesses
seven electrons with an electronic configuration of 1s2, 2s2, 2p3. Unlike oxygen,
nitrogen does not possess unpaired electrons (Figure 2.14) and is therefore
considered diamagnetic. Nitrogen does not possess available d orbitals so it is
limited to a valency of 3. It can show a range of oxidation states from -3
(ammonia) to +6 (nitrate radical).
Formation.
Nitrogen can be formed in the laboratory by the oxidation of ammonia (Eqn 2.42).
Industrially, nitrogen is obtained from the atmosphere by liquefaction of air.
Biologically, nitrogen is produced as part of the nitrogen cycle.
2NH3 + 3CuO → N2 + 3H2O + 3Cu
Eqn 2.42
Chemical Reactions.
Chemically, nitrogen is fairly inert, due to the very large N≡N bond energy (946kJ
mol-1) but it can be forced to react if conditions are correct. For example, it can
WWW.ESAINC.COM
82
form ionic nitrides with electropositive elements (e.g., lithium) (Eqn 2.43),
covalent nitrides with non-metals (e.g., carbon and boron), and — under extreme
temperature and pressure and in the presence of a catalyst — it can be reduced
to ammonia (Eqn 2.44). In the presence of oxygen, nitrogen can produce nitric
oxide when sparked (Eqn 2.45): this takes place in the atmosphere during
lightning flashes.
6Li + N2 → 2(Li+)3N3N2 + 3H2 → 2NH3
N2 + O2 → 2NO•
Eqn 2.43
Eqn 2.44
Eqn 2.45
Nitrogen and Nitric Oxide
Electronic Configuration
σ *2p
π *2p
π2p
σ 2p
GroundState
Nitrogen
Nitric
Oxide
(NO)
Nitrosonium
Cation
(NO+ )
Triplet
Nitroxyl
Anion
(NO-)
Singlet
Nitroxyl
Anion
(NO-)
Figure 2.14 Molecular Orbital Diagram Of Molecular Nitrogen
And Nitric Oxide.
Unlike oxygen, in which a variety of oxidases and oxygenases (Table 2.1) make
use of this gas in biochemically important reactions, the high energy of the N≡N
bond renders nitrogen biochemically inert. Only one enzyme, the microbial
nitrogenase complex, “fixes” nitrogen. This enzyme catalyzes the reduction of
nitrogen to ammonia at a great energetic cost (Eqn 2.46). One subunit of the
nitrogenase complex is a strong reducing agent with an Eo’= -0.4V. Ammonia can
then be assimilated, by the action of glutamate dehydrogenase and glutamine
synthetase, into the nitrogen cycle through the production of amino acids.
WWW.ESAINC.COM
83
N2 + 6e- + 12ATP +12H2O → 2NH4+ +12ADP + 12PI + 4H+
Eqn 2.46
2. The Oxides of Nitrogen.
In the field of oxidative metabolism, it is the oxides of nitrogen that are the most
important. Readers are referred to Beckman (1996a) for an excellent review. At
first sight the chemistry of and the interrelationships between the nitrogen oxides
may appear pretty daunting. However, it should be remembered that many of the
reactions described in a typical chemistry textbook are for those obtained in the
gas phase and for high concentrations of reactants. These reactions are usually
less relevant to biological systems. Readers should be aware that many articles
in the literature fail to make such a distinction. However, gas phase reactions are
important in exposure to air pollution (e.g., ozone and a variety of nitrogen oxides
are formed in the atmosphere by lightning discharge and irradiation; nitrogen
oxides are produced by the internal combustion engine and in tobacco smoke).
The relationship between the different RNS is presented in Figure 2.15. The
dimeric nitrogen oxides (N2O2 and N2O4) and acid anhydrides (N2O3 and N2O5)
are usually formed only at higher concentrations of nitric oxide and/or nitrogen
dioxide and are unlikely to be formed from the low concentrations of nitric oxide
(typically 10-400nM) and nitrogen dioxide usually found in biological systems.
The one exception is with the immune system which, when activated, can
produce large quantities of nitric oxide (Hibbs et al. (1988)).
WWW.ESAINC.COM
84
Reactive Nitrogen Species (RNS)
R-SH
NO+
Nitrosonium
Cation
NONitroxyl
Anion
Citrulline
-e-
NO2++ HONitronium
Cation
-RS
+RSH
+eO 2NO
Nitric Oxide
Arginine
RSNO
Nitrosothiol
+H+
O=NOO
Peroxynitrite -H+
CO 2/
HCO 3-
HOCl
NOCl
Nitrosyl
Chloride
O2NOCO 2Nitrocarbonate
NO2Nitrate
(III)
O=NOOH
Peroxynitrous
Acid
ONO 2CO 2Nitrosoperoxycarbonate
O=NO---OH
Caged Pair
NO3- + H+
Nitrate
(V)
NO3Nitrate
(V)
HOCl
ClONO + ClNO2
Nitryl
Chlorine
Nitrite
Chloride
Figure 2.15 The Relationship Between The Different RNS.
2.1 Nitric Oxide.
Physical Properties.
Nitric oxide (NO•) is a colorless monomeric gas that can also exist as a blue
liquid and blue solid consisting mainly of centrosymmetric dimers. Nitric oxide is
quite stable in pure water and can be dissolved to 1.93mM at 25oC and at a
partial pressure of 1 atm. The solubility of nitric oxide at physiological ionic
strength and temperature is 1.55mM. Its solubility in membranes is approximately
6-7 fold higher than in the aqueous phase. Under physiological conditions the
half-life of nitric oxide is only a few seconds (see below).
Nitric oxide has a single unpaired electron in its π*2p antibonding orbital (Figure
2.14) and is therefore paramagnetic. This unpaired electron also weakens the
overall bonding seen in diatomic nitrogen molecules so that the nitrogen and
oxygen atoms are joined by only 2.5 bonds. The structure of nitric oxide is a
resonance hybrid of two forms (Figure 2.16).
•
The loss of an electron (from the π*2p antibonding orbital) produces the
nitrosonium ion (NO+), a molecule isoelectronic to nitrogen and carbon
WWW.ESAINC.COM
85
•
monoxide. The triple nitrogen-oxygen bond in the nitrosonium ion is much
more stable (bond distance 0.114nm; bond energy 1048kJ mol-1) than the
2.5 bonds in nitric oxide (bond distance 0.120nm; bond energy 627kJ mol1
) (Beckman (1996a)). The nitrosonium ion forms several ionic salts (e.g.,
nitrosonium perchlorate [NO+ClO4-] and nitrosonium hydrogen sulfate
[NO+HSO4-]).
When nitric oxide gains one electron, the nitroxyl anion (NO-) is formed
which is isoelectronic to oxygen. Like oxygen, the nitroxyl anion has two
unpaired electrons of parallel spin in two π*2p molecular orbitals in its
lowest energy configuration (triplet state) (Figure 2.14). The nitroxyl anion
also exists in a singlet state (where the two electrons form an antiparallel
spin pair residing in a single π*2p molecular orbital). As expected this state
is much higher in energy (~87.9kJ mol-1) than the triplet molecule
(Standbury (1989)).
N
N
O
O
Figure 2.16 The Resonance Forms Of Nitric
Oxide.
Formation.
Nitric oxide can be produced in the laboratory by the action of 50% nitric acid on
copper metal (Eqn 2.47). Nitric oxide so produced is contaminated with nitrogen
dioxide, but can be purified by passing it through a concentrated iron (II) sulfate
solution. Nitric oxide can then be liberated from FeSO4.NO by heating it in the
absence of air. Nitric oxide is also formed during the electrical discharge of
nitrogen (Eqn 2.45).
3Cu + 8HNO3 → 3Cu2+ + 6NO3- + 4H2O + 2NO•
Eqn 2.47
A wide selection of NO-donor reagents now exist and can be used to generate
nitric oxide in test systems (Table 3.5). These NO-donors vary in stability, pHand oxygen-sensitivity, water solubility, and contamination. For example, some
reagents can also produce nitrosonium ions, nitroxyl anions, and other nitrogen
oxides. Care must be exercised when interpreting data obtained using these
donors (reviewed by Feelisch and Stamler (1996)).
WWW.ESAINC.COM
86
Class
Metal nitrosyls
Inorganic NO donors
Nitroxyl generating compounds
Hydroxylamine
N-Hydroxyguanidines
O-Nitro and O-Nitroso compounds
S-Nitro and S-Nitroso compounds
N-Nitroso compounds
Diazeniumdiolates (NONOates)
C-Nitro and C-nitroso compounds
Heterocyclic NO donors
Examples
Nitroprusside, dinitrosyl-iron (II)
complexes, nitrosyl complexes of ironsulfur clusters, and nitrosyl complexes of
other transition metals
Acidified nitrite, nitrosonium salts and
nitrosyl halides, peroxynitrite, and
sodium azide
Angeli’s salt, Piloty’s acid, cyanamide,
and sodium nitroxyl
Organic nitrates, and organic nitrites
Thionitrates, and thionitrites
N-Nitrosamines, N-hydroxy-Nnitrosamines, N-nitrosamides, Nnitrosoguanidines, N-nitrosohydrazines,
and N-nitrosimines
Oxadiazoles, oxatriazoles, and
sydnonimines
Table 2.6 Examples Of Nitric Oxide-Donor Molecules.
In living organisms nitric oxide is produced enzymatically. Microbes can generate
nitric oxide by the reduction of nitrite or oxidation of ammonia. In mammals nitric
oxide is produced by stepwise oxidation of L-arginine catalyzed by nitric oxide
synthase (NOS). Nitric oxide is formed from the guanidino nitrogen of the Larginine in a reaction that consumes five electrons and requires flavin adenine
dinucleotide (FAD), flavin mononucleotide (FMN) tetrahydrobiopterin (BH4), and
iron protoporphyrin IX as cofactors (Figure 2.17). The primary product of NOS
activity may be the nitroxyl anion that is then converted to nitric oxide by electron
acceptors (see below).
WWW.ESAINC.COM
87
.
NH2
N
H2N
O
OH
H2N
H2N
NH
O2
H2O
O2
NH
NH
H2O
+
NADPH
NADP+
0.5 NADPH
NH3
O
L-Arginine
O
0.5 NADP+
NH3
NH3
O
O
O
N
O
O
N G-HydroxyL-Arginine
L-Citrulline
Figure 2.17 The Formation Of Nitric Oxide From Arginine.
To date, all sequenced NOS cDNAs show homology with the cytochrome P450
reductase family. Based on molecular genetics there appears to be at least three
distinct forms of NOS:
•
•
•
A Ca2+/calmodulin-requiring constitutive enzyme (c-NOS; ncNOS or type I)
(which produces relatively low levels of nitric oxide and is important in
neurotransmission, maintenance of vascular tone, and inhibition of platelet
aggregation);
A calcium-independent inducible enzyme (i-NOS; type II), which is
primarily involved in the mediation of the cellular immune response; and
A second Ca2+/calmodulin-requiring constitutive enzyme found in aortic
and umbilical endothelia (ec-NOS or type III) (Michel et al. (1996)).
The important roles of these different NOS isoforms have been reviewed
elsewhere (Feldman, et al. (1993); Kostka (1992); Snyder and Bredt (1992)).
Recently, a mitochondrial form of the enzyme, which appears to be similar to the
endothelial form, has been found in brain and liver tissue (Bates et al. (1995);
Giulivi et al. (1998); Tatoyan and Giulivi (1998)). Although the exact role of nitric
oxide in the mitochondrion remains elusive, it may play a role in the regulation of
cytochrome oxidase (Giulivi (1998)).
Nitric oxide appears to regulate its own production through a negative feedback
loop (Griscavage et al. (1995)). The binding of nitric oxide to the heme prosthetic
group of NOS inhibits this enzyme, thereby preventing the production of more
nitric oxide. Interestingly, c-NOS and ec-NOS are much more sensitive to this
regulation than i-NOS. This suggests that, in the brain, nitric oxide can regulate
its own synthesis and therefore the neurotransmission process. Furthermore,
inhibition of ec-NOS will prevent the cytotoxicity associated with excessive nitric
WWW.ESAINC.COM
88
oxide production. Conversely, the insensitivity of i-NOS to nitric oxide will enable
high levels of nitric oxide to be produced for cytotoxic effects. Several
endogenous inhibitors of NOS (mainly the guanidino-substituted derivatives of
arginine) occur in vivo as a result of post-translational modification of proteincontained arginine residues by S-adenosylmethionine (Kostka (1992)). For
example, the dimethylarginines (NG,NG-dimethyl-L-arginine and NG,N’G-dimethylL-arginine) occurs in tissue proteins, plasma, and urine of humans (Tsikas et al.
(1998)).6 These NOS inhibitors are thought to act as both regulators of NOS
activity and reservoirs of arginine for the synthesis of nitric oxide.
Numerous studies have reported that NOS can form peroxynitrite, suggesting
that both its precursors, nitric oxide and superoxide, can be produced by this
enzyme (see Miller et al. (1997) and references therein). This ability can be
explained by examining of the structure of the enzyme. c-NOS, for example,
consists of a flavin-containing reductase domain and an heme-containing
oxygenase domain linked together by a sequence of amino acids that contains a
calmodulin binding site. Binding of calmodulin brings these two domains together
and allows the transfer of NADPH-derived electrons from the reductase domain
to the oxygenase domain, resulting in the conversion of arginine to citrulline and
the concomitant formation of nitric oxide. Under certain circumstances, such as a
deficiency in arginine, or in the presence of NOS inhibitors, the activated hemeoxygenase complex dissociates and forms superoxide. Ec-NOS can also
produce superoxide in a process modulated by the tetrahydrobiopterin cofactor
(Vasquez-Vivar et al. (1998)). Recently, i-NOS was also shown to produce
superoxide, but mainly by the action of the flavin-binding site of the reductase
domain of this enzyme (Xia et al. (1998)).
Chemical Reactions and Biological Significance.
Much has been written about the chemistry of nitric oxide, particularly its
reactions in the gas phase. For example, it is a reducing agent and is readily
oxidized by oxygen to nitrogen dioxide (Eqn 2.48) and by chlorine in the
presence of a catalyst to nitrosyl chloride (NOCl) (Eqn 2.49). Interestingly, both
nitrogen dioxide and nitrosyl chloride can also be formed under aqueous
conditions in vivo, albeit by very different mechanisms.
2NO•(g) + O2(g) → 2NO2•(g)
2NO•(g) + Cl2(g) → 2NOCl(g)
Eqn 2.48
Eqn 2.49
Before we turn to the reactions of nitric oxide under physiological conditions, it
must remembered that, although many of nitric oxides reactions can occur in the
6
These inhibitors can be measured using HPLC fluorescence following OPA derivatization (Meyer et al. (1997)).
WWW.ESAINC.COM
89
aqueous phase, its reactivity is greatly accelerated within the hydrophobic interior
of biological membranes. It has been calculated that even though membrane
makes up about 3% of the total tissue volume, 90% of the reaction of nitric oxide
with oxygen occurs within this compartment (Liu et al. (1998)). Thus the
membrane is an important site for nitric oxide chemistry.
There are two major aspects to nitric oxide chemistry. First, it can undergo single
electron oxidation and reduction reactions producing nitrosonium and nitroxyl,
respectively. These will be dealt with in greater detail in their own section below
in recognition of their importance to redox biochemistry. Second as it has a single
unpaired electron in its π*2p molecular orbital it will react readily with other
molecules that also have unpaired electrons, such as free radicals and transition
metals (Fukuto and Wink (1999)).
Examples of the reaction of nitric oxide with radical species include:
•
•
•
•
•
Nitric oxide will react with oxygen to form the peroxynitrite (nitrosyldioxyl)
radical (ONO2•) and with superoxide to form the powerful oxidizing and
nitrating agent, peroxynitrite anion (ONO2-) (Eqn 3.24). Peroxynitrite
causes damage to many important biomolecules and has been implicated
in a variety of diseases (see below);
Nitric oxide reacts with thiyl radicals to form nitrosothiols that are important
in the regulation of blood pressure (see below);
Nitric oxide reacts with alkylperoxyl radicals and thereby terminates lipid
peroxidation (Chapter 3). The alkyl peroxynitrites (RO2NO) formed may be
cytotoxic decomposing to nitrogen dioxide and other pro-oxidants;
Nitric oxide can react rapidly (second order rate constant 10-9 M-1s-1) with
tyrosyl radicals forming 3-nitrosotyrosine and/or 4-O-nitrosotyrosine. This
reaction can affect the activity of enzymes that utilize tyrosyl radicals in
their mechanisms. It is a common misconception that nitric oxide will react
with tyrosine to form 3-nitrotyrosine, a compound currently being used as
a marker of RNS activity. Nitric oxide will not react with tyrosine directly.
Furthermore, it has yet to be proven that nitrosotyrosine formed as above
can be oxidized to 3-nitrotyrosine in biological systems.
Nitric oxide rapidly reacts with oxyhemoglobin producing nitrate — the
primary route of its destruction in vivo (see below).
Of great importance to redox biochemistry is the reaction between nitric oxide
and transition metal complexes (Cooper (1999)). During this reaction a “ligand”
bond is formed (the unpaired electron of nitric oxide is partially transferred to the
metal cation), resulting in a nitrosated (nitrosylated) complex. For example, such
complexes can be formed with free iron ions, iron bound to heme or iron located
in iron-sulfur clusters (Salerno (1996)). Ligand formation allows nitric oxide to act
as a signal, activating some enzymes while inhibiting others (Table 2.7). Thus,
the binding of nitric oxide to the Fe (II)-heme of guanylate (guanalyl) cyclase
[GTP-pyrophosphate lyase: cyclizing] is the signal transduction mechanism by
WWW.ESAINC.COM
90
which nitric oxide acts as a messenger molecule (Eqn 2.50). Guanylate cyclase
exists as cytosolic and membrane-bound isozymes. The soluble form (sGC)
occurs as a heterodimer (MW 150kD), which also contains 1mol heme. The
binding of nitric oxide to the iron of the heme molecule activates sGC, which then
converts GTP to cyclic-GMP. Cyclic GMP can then activate protein kinase C,
phosphodiesterase, and ion channels, thereby amplifying the original nitric oxide
signal. These actions, as well as the subsequent decrease in intracellular calcium
levels, mediates some of the biological effects of nitric oxide.
E(inactive)-Fe2+(heme) + NO• → E(active)-Fe2+(heme)-NO
Eqn 2.50
When produced in appropriate amount and periodicity, nitric oxide fulfills several
significant biological roles. It is now regarded as a general short-lived secondary
messenger (Bredt and Snyder (1994)) and, on the whole, as a beneficial
molecule. First, it plays a role in blood pressure regulation as the endotheliumderived relaxing factor (EDRF) (Moncada et al. (1991)). Second, it acts as a
retrograde neurotransmitter implicated in the formation of long-term memory
(Schuman and Madison (1991)). Third, it contributes to the regulation of bone
metabolism (Ralston (1997) Fourth, nitric oxide has a major function in the
immune system and is produced by both macrophages and neutrophils. It inhibits
the activity of key enzymes in the pathogen by forming transition metal
complexes and its product, peroxynitrite, acts as a pro-oxidant with cytotoxic
actions (Hibbs, et al., (1988); Marletta et al. (1988)). Finally, it may play a role as
an antioxidant; however, this hypothesis is controversial:
•
•
•
•
•
Nitric oxide dissolves into membranes and can intercept lipid-based
radicals, preventing lipid peroxidation processes (Hogg and Kalyanaraman
(1999)). However, the products may be toxic and unstable resulting in the
release of nitrogen dioxide and other RNS;
Nitric oxide can detoxify the hydroxyl free radical by forming nitrous acid.
However, this can occur only at unphysiological levels of nitric oxide (mM
range) in order for it to compete with other compounds for reaction with
hydroxyl free radical. Furthermore, nitric oxide can stimulate hydroxyl free
radical production by reducing Fe (III) to Fe (II) that can then take part in
the Haber-Weiss reaction;
The reaction between nitric oxide and oxygen produces nitrogen dioxide,
while its reaction with superoxide produces peroxynitrite, two very reactive
pro-oxidant species;
Nitric oxide is converted to a nitrosonium ion-like species capable of
attacking important biomolecules (see below); and
Nitric oxide may inhibit glutathione peroxidase, preventing detoxification of
lipid hydroperoxides (Chapter 4).
WWW.ESAINC.COM
91
Thiol
Metal
Extracellular
Albumin
GSH
Membrane
Adenylyl cyclase (type I)
Calcium ATPase
G Proteins
KCa+ Channel
NADPH Oxidase
NMDA Receptor
Protein Kinase C
Cytosolic/
mitochondrial
Actin
Alcohol Dehydrogenase
Aldolase
Aldehyde
Dehydrogenase
GAP-43
Glyceraldehyde 3phosphate
Dehydrogenase
γ-Glutamylcysteinyl
Synthetase
Glutathione
Glutathione peroxidase
SNAP-25
Tissue Plasminogen
Activator
2,3-Indolamine
Aconitase
Complex I
Complex II
Complex IV
Cyclo-oxygenases
Cytochrome P450
Guanylate Cyclase
Dioxygenase
Hemoglobin
NO-Synthase
Tryptophan Hydroxylase
Nuclear
AP-1
OMDM Transferase
NF-κB
SoxRS
Tyrosyl/
Tryptophyl
Radicals
Cytochrome-c
Peroxidase
Prostaglandin H
Synthase
Ribonucleotide
Reductase
Table 2.7 Some Biologically Important Targets Of Nitrogen Oxides. (From
Eiserich et al. (1995); Kuhn and Arthur (1996); Moncada and Higgs (1993); Raddi (1996); Stamler
(1994); and See Table 3.11). OMDM – O-methylguanine DNA methyltransferase. See references
and Glossary for definitions of abbreviations used.
Unfortunately, nitric oxide also represents a major problem for biological systems
(Gross and Wolin (1995)). Uncontrolled production of nitric oxide (e.g., overactivity of the immune system or uncontrolled central release following ischemia)
can directly damage enzymes and other proteins. Nitric oxide also promotes
ADP-ribosylation of proteins (Brune et al. (1994)), can directly damage DNA
(Nguyen et al. (1992); Wink et al. (1991)) and may deplete cellular antioxidants
(d’Ischia and Novellino (1996); Gorbunov et al. (1996)). The detrimental action of
nitric oxide can be further exacerbated by the formation of peroxynitrite (see
below). Cells have developed different mechanisms to protect themselves
against chronic nitric oxide exposure. For example, E. coli activate the soxRS
WWW.ESAINC.COM
92
response to prevent damage by nitric oxide-producing macrophages (Nunoshiba
et al. (1993)). Exposure of mammalian cells, including neurons, causes adaptive
resistance possibly mediated through the induction of heme oxygenase 1
(Chapter 4) (Bishop et al. (1999); Hartsfield et al. (1997); Takahashi et al.
(1997)). Although the exact mechanism of heme oxygenase is hard to pinpoint, it
could either be direct, through the reaction of nitric oxide with the heme moiety of
the enzyme (similar to that described for nitric oxide destruction in blood below),
or indirect, through the production of the antioxidant bilirubin. Consequently,
heme oxygenase 1 may play a role in regulating nitric oxide-dependent age- and
disease-related neurodegeneration.
NO
Hb-Fe2+-O2
O2
Oxyhemoglobin
Hb-Fe2+O2NO
Hb-Fe2+
Hemoglobin
NADPHdependent
reduction
Hb-Fe3+
NO3-
Methemoglobin
Figure 2.18 The Metabolism Of Nitric Oxide To Nitrate.
The
•
•
destruction
of
nitric
oxide
proceeds
via
two
mechanisms:
First, in an oxygenated solution, nitric oxide undergoes a complex
series of reactions involving many different ROS, culminating in the
formation of nitrite (Eqn 2.51) (Beckman (1996a)). Nitrate is also
formed (Eqn 2.52 to 2.54). Overall excess nitrite is produced. Under
aqueous conditions the oxidation of nitric oxide can produce a wide
variety of RNS.
Second, in the blood system in which high levels are typically
produced, nitric oxide can be oxidized to nitrate following its binding to
oxyhemoglobin (or oxymyoglobin) (Figure 2.18). During this process
WWW.ESAINC.COM
93
the Fe (II) contained within the heme ring is oxidized to the Fe (III)
form, producing methemoglobin (cf. Eqns 3.20 and 3.21). Hemoglobin
is regenerated in the erythrocyte by the enzymatic reduction of
methemoglobin with NADPH. Interestingly, hemoglobin can be Snitrosylated in the lung and can release nitric oxide during arterialvenous transit (Gow and Stamler (1998); Jia et al. (1996)). Hemoglobin
not only regulates nitric oxides level but it may also act as a carrier of
nitric oxide and therefore a regulator of blood pressure.
4NO• + O2 + 2H2O ↔ 4HNO2
2NO• + O2 ↔ 2NO2•
2NO2• ↔ N2O4
N2O4 + H2O ↔ NO2- + NO3- + 2H+
Eqn 2.51
Eqn 2.52
Eqn 2.53
Eqn 2.54
Measurement.
Many approaches are used to measure nitric oxide both directly and indirectly
and these, along with their advantages and disadvantages, are presented in
Table 2.8.
Approach
Bioassays e.g., aortic strips
Comments
Indirect. Not specific for nitric oxide. Not
quantitative. Inconvenient.
Reference
Cocks, et al. (1985);
Furchgott (1984);
Griffith, et al. (1984);
Martin, et al. (1985);
Rapaport and Murad
(1983); Rubanyi, et al.
(1985)
cGMP radioimmunoassay
Indirect. Measures formation of [32P]cGMP from [α-32P]-GTP when nitric oxide
activates guanylyl cyclase. Determined by
liquid scintillation counting. cGMP levels
increased by NO-donors and decreased
by NOS inhibitors. In some cases it lacks
selectivity and qualitative information. Can
be made more nitric oxide-specific by the
use of reporter cells such as RFL-6 cells.
Does not always work reliably.
Forstermann and Ishi
(1996); Ignarro et al.
(1984); Palacios et al.
(1989)
Chemiluminescence with
hydrogen peroxide-luminol,
or ozone
Direct. Requires photomultiplier tube (an
ozone generator is required for ozone
approach). Failure to extract all nitric
oxide from the biological sample into the
gas phase will underestimate its level.
Biological samples may foam during the
Aoki, T. (1990); Hampl
et al. (1996); Kojima et
al. (1997)
WWW.ESAINC.COM
94
Citrulline levels.
[14C]-Citrulline formation
from [14C]-arginine
procedure that can cause problems with
the detector (may be overcome using
microdialysis). Humidity can affect
sensitivity.
Indirect.
• Measured as OPA/βME derivative
using HPLC-ECD in brain
microdialysis samples. Citrulline
levels were increased following kainic
acid and decreased following NOS
inhibition. Other biochemical
pathways may also form citrulline.
• Measured as OPA/βME derivative
using HPLC-fluorescence in
endothelial cells. Other biochemical
pathways may also form citrulline.
Indirect.
• Measured in microdialysis samples.
Does not take into account loss of
label when other pathways
metabolize citrulline. Only uses
simple column extraction so signal
from [14C]-Citrulline may be
contaminated. Does not take into
account loss of radiolabel during
sample processing.
• Thin layer chromatography of
endothelial cell extracts. Does not
take into account loss of label when
other pathways metabolize citrulline.
•
Moncada, and
Palmer (1990);
Wang and Maher
(1992)
•
Hecker and Billiar
(1996)
•
Bhardwaj et al.
(1995)
•
Hecker and Billiar
(1996)
EPR
Direct or indirect. Lacks sensitivity to
measure free nitric oxide in vivo. The
angular momentum of the radical electron
can couple with the angular momentum of
the nitric oxide molecule to diminish the
paramagnetism of nitric oxide. Useful in
examining endogenous nitric oxidecomplexes. EPR can be extended by the
use of nitric oxide-spin traps including Nmethyl-D-glucamine
dithiocarbamate/ferrous ions, nitroxides,
deoxyhemoglobin, and hemoglobin. May
require freeze quenching of samples.
Cheletrophic biradical traps are showing
some promise.
Arroya et al. (1990);
Henry et al. (1993);
Henry and Singel
(1996); Korth and Weber
(1996); Kosaka and
Shiga (1996); Lai and
Komarov (1994); Singel
and Lancaster (1996)
Electrochemical
Direct. Various voltammetric probes exist.
Probes differ in sensitivity, selectivity and
kinetically.
Burlet and Cespuglio
(1997); Canini et al.
(1997); Desvignes et al.
(1997); Friedemann et
al.(1996); Malinski, and
Taha (1992); Rivot et al.
(1997); Shibuki (1990);
Indirect. Measurement of nitrite.
WWW.ESAINC.COM
95
Strehlitz et al. (1996);
Yao et al. (1995)
Fluorescent Probes
Mass Spectrometry
Nitrite/nitrate
Direct. Fluorescent probes, such as 4,5diaminofluorescein or 2,3-diamino
naphthalene, permit real time imaging of
nitric oxide production within cells.
Direct or indirect. Useful for measuring
nitric oxide in the gas phase. Probably not
the best approach for in vivo studies.
Indirect.
• Nitric oxide is readily converted in
vivo to nitrite and then nitrate (nitrate
is normally two orders of magnitude
higher). Total nitridergic involvement
can be estimated following a) nitrate
reduction to nitrite by cadmium or
nitrate reductase, b) reaction with
Greiss reagent and c) monitoring
spectrophotometrically at 540nm.
Kojima et al. (1998);
Nakatsubo et al. (1998)
Payne et al. (1996)
Muscara and Nucci
(1996); Pratt et al.
(1995); Preik-Steinhoff
and Kelm (1996); Salter
et al. (1996); Schmidt
and Kelm (1996);
Stratford (1999);
Yamada and Nabeshima
(1997)
Problems include incomplete reduction of
nitrate, influence of diet, bacterial
metabolism production unrelated to the
arginine-nitric oxide pathway.
•
•
•
•
HPLC-UV following reduction and
Greiss reaction (nitrite and nitrate)
HPLC-coulometric ECD (nitrite only)
and UV (nitrate only)
Chemiluminescence (nitrite and
nitrate)
Voltammetric nitrite-sensor (nitrite
only)
Approaches that measure both nitrite and
nitrate do so for lack of sensitivity.
Techniques measuring nitrite alone may
be more representative of nitric oxide
production, but care must be exercised
when interpreting data.
Spectrophotometric
Indirect. Nitric oxide trapped by
oxyhemoglobin and measured using a
spectrophotometer. Maybe problematic
for routine use in vivo. Coupling this
approach to microdialysis has been used
successfully to measure central NO•
levels in vivo.
Balcioglu and Maher
(1993); Feelisch et al.
(1996) and references
therein; Zou and Cowley
(1997)
Table 2.8 The Different Approaches Used To Measure Nitric Oxide (see also
Kishnani and Fung (1996)).
WWW.ESAINC.COM
96
2.2 Nitroxyl Anion And The Nitrosonium Ion.
As discussed above nitric oxide can take part in redox reactions. Chemically, it
can act as both an oxidizing agent (becoming reduced to the nitroxyl anion) (Eqn
2.55) and a reducing agent (becoming oxidized to the nitrosonium ion) (Eqn 2.56)
(Hughes (1999)). The nitroxyl anion and nitrosonium ion show distinct
chemistries and their biological significance is still being evaluated (Crow and
Beckman (1995), Stamler et al. (1992b)).
NO• + e- → NONO• - e- → NO+
Eqn 2.55
Eqn 2.56
Under physiological conditions nitric oxide is a moderate oxidizing agent
(Eo=+390mV). As discussed above (see also Figure 2.14), the nitroxyl anion can
exist in a triplet and singlet state. The triplet nitroxyl anion can act as a oneelectron reductant, thereby reforming nitric oxide. It can react with other radicals.
For example, it reacts with oxygen (second order rate constant of 3.4 x 107
M-1s-1) forming peroxynitrite (Eqn 2.57) (Huie and Padmaja (1993)). It also reacts
reversibly with transition metal ions. For example, it is similar in size and shape
to superoxide and readily reacts with copper ions found in Cu,Zn-superoxide
dismutase forming nitric oxide (Eqn 2.58) (Beckman (1996a)). Indeed, some
have proposed that NOS does not generate nitric oxide but rather the nitroxyl
anion; the latter is then converted to the former by SOD and other electron
acceptors (Hobbs et al. (1994); Schmidt et al. (1996b)). The reaction is freely
reversible so nitric oxide can readily be reduced to the nitroxyl anion as well
(Beckman (1996a)). In solution the nitroxyl radical readily undergoes a series of
reactions producing nitrite and nitrous oxide (N2O) (a simplified reaction is shown
Eqn 2.59). The singlet nitroxyl anion is more energetic than the triplet form. As it
does not have an unpaired electron it does not react with oxygen (Hughes
(1999); Stanbury (1989)). It also shows different chemical reactivity.
NO- + O2 → ONO2NO- + Cu2+(SOD) ↔ NO• + Cu+(SOD)
NO- + 2NO• → N2O + NO2-
Eqn 2.57
Eqn 2.58
Eqn 2.59
The nitrosonium ion is a strong oxidizing agent (E0=1210mV) so its direct
formation from nitric oxide is unlikely to occur under physiological conditions.
However, a nitrosonium-like species can be formed in vivo when nitric oxide
reacts with transition metal complexes (Eqns 2.60). For example, the formation of
nitrosothiols by the reaction presented in Eqn 2.61 will not occur unless an
electron acceptor (e.g., nitrogen dioxide or transition metals) is present (Eqn
2.62). Although this may be important in the synthesis of nitrosothiols (below), it
WWW.ESAINC.COM
97
can be a major problem if this were to affect a thiol group critical to an enzyme’s
function (Laval and Wink (1994); Stamler (1994)). In general nitrosation reactions
can be a major problem for any living organism, as they can generate a variety of
reactive and potentially toxic products. We will return to nitrosation reactions
when we discuss nitrous acid below.
NO• + Fe3+(heme) → Fe2+(heme)–NO
NO• + RSH → RSNO + H+ + eFe2+heme)–NO + RSH → Fe2+(heme) + RSNO + H+
Eqn 2.60
Eqn 2.61
Eqn 2.62
The extent of nitrosation in vivo can be estimated by the measurement of a
variety of nitrosated amino acid products (e.g., nitrosoproline, Nnitrosothiazolidine-4-carboxylic acid and N-nitroso-2-methylthiazolidine-4carboxylic acid) in urine using GC with a thermal energy analyzer ((Ohshima and
Bartsch (1999)).
2.3 Peroxynitrite.
Properties.
Peroxynitrous acid (ONO2H) has a pKa of 6.8. Dilute basic solutions of
peroxynitrite (~200mM) are relatively stable and are yellow in color. Under these
conditions peroxynitrite can be kept safely at -20oC for many weeks. At
physiological pH, the unstable and highly reactive peroxynitrous acid is formed.
Peroxynitrite also forms highly colored, relatively stable salts (e.g., the
tetramethylammonium salt is a yellow-orange solid). These salts are free from
other ROS and RNS and should be used when accurate determination of
reaction stoichiometry is critical. Peroxynitrite occurs in both cis- and transisomers (Beckman (1996a)).
Formation.
For readers interested in a review of the discovery of peroxynitrite see Beckman
(1996a). The peroxynitrite anion can be produced in the laboratory by several
methods (Table 2.9) (Saha et al. (1998); Uppu et al. (1996a)). The most common
synthesis of peroxynitrite comes from acidified nitrite and hydrogen peroxide
(Eqns 2.63 and 2.64) (Koppenol et al. (1996)). Peroxynitrite and its salts are now
commercially available.
HNO2 + H+ <=> H2O + NO+
NO+ + H2O2 → ONO2H + H+
WWW.ESAINC.COM
Eqn 2.63
Eqn 2.64
98
•
Reaction of ozone with azide ions.
•
Auto-oxidation of hydroxylamine.
•
1.
2.
3.
Reaction of hydrogen peroxide with:
Acidified nitrite or nitrous acid.
Alkyl nitrites (water-soluble or water-insoluble).
Nitric oxide.
• Reaction of nitric oxide with:
1. Solid potassium superoxide.
2. Tetramethylammonium superoxide.
•
Photolysis of solid potassium nitrate.
Table 2.9 Different In Vitro Methods For The
Synthesis Of Peroxynitrite.
Peroxynitrite can be formed in vivo by at least three possible reactions (Beckman
(1996a):
•
•
•
The reaction between nitric oxide and the superoxide radical anion (Eqn
2.25) is, without a doubt, the major source of peroxynitrite production in
vivo. This reaction proceeds at a near diffusion-limited rate (6.7 x 109
M-1s-1) which is approximately 3-6 times the rate at which superoxide is
dismuted by SOD. Thus, both nitric oxide and superoxide can modulate
the effects of the other. For example, superoxide can block the
hypotensive effects of nitric oxide by diverting it to form peroxynitrite. SOD
can increase the hypotensive effects of nitric oxide by decreasing the
availability of superoxide. Due to the synthesis of peroxynitrite, nitric oxide
can be considered to “detoxify” superoxide (Feigl (1988); Kanner et al.
(1991); Rubanyi et al. (1991)).7 As discussed above, these beneficial
antioxidant properties are far outweighed by the formation of peroxynitrite,
a toxic and highly reactive pro-oxidant;
Peroxynitrite can also be formed by the reaction between the nitroxyl
anion (e.g., formed by reduced SOD) and oxygen (Eqn 2.57); and
The reduction of the peroxynitrite radical (Eqn 2.65).
ONO2• + O2•- → ONO2- + O2
Eqn 2.65
7
The probability that one chemical species will attack another depends not only on the rate of the reaction but also on
the concentration of the target (Crow and Beckman (1995)). The multiple of the reaction rate and target concentration is
called the target area, and under normal conditions, the target area of SOD exceeds the target area of nitric oxide by
about 30 fold. However, during pathological conditions such as reperfusion following ischemia, the target area of nitric
oxide can exceed that of SOD such that under these conditions peroxynitrite is preferentially produced.
WWW.ESAINC.COM
99
Although peroxynitrite can be formed by the reaction between the nitrosonium ion
and hydrogen peroxide it is highly unlikely to occur under physiologically
conditions.
Chemical Reactions and Biological Significance.
The chemistry of peroxynitrite has recently been the topic of several reviews and
numerous papers (e.g., Beckman (1996); Beckman et al. (1994); Daiber et al.
(1998); Groves (1999); Pryor and Squadrito (1995); Squadrito and Pryor (1998)).
The reactivity of peroxynitrite not only depends upon the pH of the reaction and
which chemical species are present, but also on the fact that peroxynitrite exists
in vivo in the cis-isomer. The cis-isomer is much more reactive than the transisomer; the latter readily isomerizes to nitrate without further reaction. Since the
barrier for isomerization is about 110kJ mol-1, cis-trans isomerization is unlikely to
occur in vivo. Peroxynitrite is therefore locked in the more reactive form
(Beckman
(1996a)).
A
summary
of
the
many
reactions
of
peroxynitrite/peroxynitrous acid with various biomolecules is presented in Table
2.10.
Molecule
Damaged
DNA
Comments
•
•
•
•
Lipids
Proteins/
enzymes/
amino acids
•
•
•
•
•
•
Forms oxazolone, 8-oxo-2’d-adenosine and 8OH2’d-guanosine when adenine and guanine
nucleosides incubated with peroxynitrite.
Oxazolone and 8-oxo-2’d-adenosine levels are
elevated when double stranded DNA is
incubated with peroxynitrite.
Forms 8-nitroguanine when guanine is treated
with peroxynitrite in vitro.
Causes DNA strand breaks.
Activates the DNA repair enzyme,
poly(ADP)ribosyltransferase.
Damages 2-deoxyribose.
Promotes lipid peroxidation.
Forms F2-isoprostanes during oxidation of
human low-density lipoproteins.
Nitrated tyrosine can alter protein conformation
and activity.
Nitrated tyrosine can disable tyrosine
phosphorylation regulatory mechanism and
target proteins for degradation.
Oxidizes thiols. The two-electron oxidation
pathway, mediated by peroxynitrite anion,
predominates over the one-electron pathway
mediated by peroxynitrous acid and its
derivatives. In vivo, nitrocarbonate favors the
one-electron oxidation of thiols to thiyl radicals.
These are involved in chain reactions that
ultimately oxidize thiols to disulfides. This can
WWW.ESAINC.COM
Reference
Beckman et al.
(1990); Chabot et al.
(1997); Douki and
Cadet (1996); Szabo
et al. (1997); Tamir et
al. (1996); Yermilov
et al. (1995a,b);
Zingarelli et al.
(1996).
Moore et al. (1995);
Radi et al. (1991);
Rubbo et al. (1994).
Alvarez et al. (1996);
Beckman (1996);
Beckman et al.
(1992); Bouton et al.
(1997); Briviba et al.
(1998); Cooper et al.
(1998); Crow et al.
(1995);
Frears et al. (1996);
Galli et al. (1998);
Gow et al. (1996);
Halliwell (1997); Kaur
et al. (1997);
100
Ascorbic acid,
bilirubin and uric
acid
effect the activity of enzymes requiring
participation of thiol groups in their reaction
mechanisms (e.g., alcohol dehydrogenase, and
glyceraldehyde-3-phosphate dehydrogenase).
Interestingly, cysteine and GSH are
peroxynitrite scavengers.
• Procollagenase activation. May affect the
extracellular matrix and lead to disease (e.g.,
arthritis).
• Inactivation of the inhibitor of metalloproteinase1. Metalloproteinases are a group of enzymes
including collagenases, gelatinases and
stromelysins that are critical in controlling
connective tissue remodeling.
• Inactivates α1-antiproteinase, the major inhibitor
of serine proteases such as elastase.
• Inactivates E. coli Mn/Fe-SOD but not bovine
CuZn-SOD.
• Inhibits glutathione peroxidase by oxidation of
selenol to selenocysteine in enzyme’s active
site.
• Inhibits GTP binding to Rac2 by tyrosine
oxidation.
• Inactivates sarcoplasmic reticulum calciumATPase.
• Converts xanthine dehydrogenase to xanthine
oxidase, an enzyme capable of generating
ROS.
• Modulates iron regulatory protein.
• Inhibits mitochondrial electron transport.
• Forms protein carbonyls.
• Reacts with methemoglobin to generate globinbound free radical species that may play a role
in vascular disease.
• Tryptophan produces protein tryptophan
radicals, nitrotryptophan, N-formylkynurenine or
hydropyrroloindole depending upon reaction
conditions.
• Tyrosine produces protein tyrosyl radicals, 3nitrotyrosine, dityrosine and 3,4-DOPA
depending upon reaction conditions (Figure
3.9). These products can be formed by
reactions not involving peroxynitrite (see
Chapter 5).
• Phenylalanine produces tyrosine isomers and 4nitrophenylalanine.
• Oxidation of methionine to methionine sulfoxide
(Figure 2.19) and/or ethylene.
Ascorbic and uric acids scavenge peroxynitrite but
produce ascorbyl and uric acid radicals that then
have to be removed. Uric acid also forms a
nitrosated/nitrated adduct that can act as a nitric
oxide donor. Bilirubin does not react directly with
peroxynitrite but is effective at scavenging
secondary oxidants formed during the oxidation
WWW.ESAINC.COM
Ischiropoulos and AlMehdi (1995);
Ischiropoulos et al.
(1992); Kato et al.
(1997); Kong et al.
(1996): Mohr et al.
(1994); Moreno and
Pryor (1992);
Muijsers et al. (1997);
Okamoto et al.
(1997); Padmaja et
al. (1998); Pietraforte
and Minetti
(1997a,b); Pryor et
al. (1994); Quijano et
al. (1997); Radi et al.
(1991a); Ramezanian
et al. (1996); Rohn et
al. (1999); Sakuma et
al. (1997); Scorza
and Minetti (1998);
Souza and Radi
(1998); van der Vleit
et al. (1994, 1995);
Viner et al. (1996);
Whiteman and
Halliwell (1996).
Bartlett et al. (1995);
Minetti et al. (1998);
Skinner et al. (1998);
Vasquez-Vivar et al.
(1996).
101
Flavonoids
Inorganic anions
Monoamines
Salicylic acid
Tocopherol
process.
Quercetin, rutin and epigallocatechin gallate are
peroxynitrite scavengers.
Reacts with carbonate, cyanide, iodide, and
thiocyanate. The reaction with sulfite produces an
intermediate that prolongs peroxynitrite reactivity
and may play a role in sulfite’s neurotoxicity. This
intermediate eventually decays to sulfate and
nitrate.
Nitrates catecholamines in vitro forming 6nitrocatecholamines (nitration may be due to
peroxynitrite radical or other oxides of nitrogen).
Now found in vivo and may act as a potential signal
molecule linking the actions of norepinephrine and
nitric oxide.
Peroxynitrite can hydroxylate salicylate forming 2,3and 2,5-dihydroxybenzoic acid and nitrate salicylate
forming 5-nitrosalicylic acid. Thus peroxynitrite can
interfere with assays for hydroxyl free radical
measurement (Figure 3.11).
Nitrates γ-tocopherol to form 5-nitro-γ-tocopherol.
Fiala et al. (1996);
Haenen et al. (1997).
Groves (1999) and
references therein;
Huie and Nita (1999)
and references
therein; Reist et al.
(1998)
d’Ischia and
Costantini (1995);
Shintani et al. (1996)
Halliwell and Kaur
(1997); Kaur et al.
(1997); Narayan et
al. (1997);
Ramezanian et al.
(1996).
Christen et al. (1997)
Table 2.10 Many Molecules React With Peroxynitrite.
Peroxynitrite shows five distinct reactions pathways:
1. Following protonation, the decomposition of peroxynitrous acid produces a
species showing both hydroxyl free radical- and nitrogen dioxide-like reactivity
(see Figure 2.19). Although originally it was hypothesized that both free
hydroxyl free radicals and nitrogen dioxide were produced by a homolytic
cleavage of peroxynitrite (Eqn 2.66), evidence now suggests that neither of
these free radicals is actually formed to any extent in vivo. Rather, both
hydroxyl free radical and nitrogen dioxide-like activity may exist within the
same molecule, the “radical ends” proposed by Crow and Beckman (1995).
Such reactivity can be explained by either a caged-pair (geminate pair) or
activated conformer (ONO2H*) of peroxynitrous acid (Pryor and Squadrito
(1995)).8 The attack of a molecule by one end of peroxynitrite will form
nitrated products, while the attack by the other end will lead to hydroxylation.
ONO2- + H+ → HO•…….•NO2
Eqn 2.66
For some of its reactions, peroxynitrite can be regarded as a more stable form
of the hydroxyl free radical and is capable of transporting this reactive species
8
*
-1
The vibrationally active state ONO2H is 71kJ mol above the energy level of the ground state.
WWW.ESAINC.COM
102
to places far removed from its site of production. Remember that the hydroxyl
free radical reacts as soon as it is produced. The half-life of peroxynitrite,
however, is on the order of ms to s time scale under physiological conditions
(this may be somewhat reduced by its reaction with endogenous thiols and
bicarbonate). Furthermore cell membranes offer no significant barrier to
peroxynitrite diffusion (Groves (1999) and references therein). Thus when
compared to the hydroxyl free radical, peroxynitrite can diffuse a considerable
distance in vivo before reacting and causing damage (Table 2.11).
Species
Hydrogen
Peroxide
Steady State
Level
10-9 to 10-8M
“Half-Life”#
Years – when
pure; minutes to
hours in presence
of reducing
agents; unknown
in vivo
Diffusion
Distance
Freely diffusible
over great
distances unless
intercepted
Reference
Cadenas
(1999)**
Hydroxyl free
radical
10-20M
10-9s
3nm
Hutchinson
(1957)
Nitric oxide
~10-8 to 10-7M
(e.g.,
in
cell
signaling)
30s buffer
700µm
1s heart*
130µm
Crow and
Beckman
(1995); Kelm
and Schrader
(1988)
Unknown
1s buffer
100µm
9 x 10-3s
“biological
conditions”
9µm
10-5s
1-4µm
intracellular
Peroxynitrous acid
Superoxide
10-11 to 10-10 M
Radi et al.
(1991a,b); Zhu
et al. (1992).
Beckman (1996)
15µm vascular
Table 2.11 Comparison Of Steady-State Level, Half-Life And Diffusion Distance
Between Different ROS And RNS (See Beckman (1996) and references therein for further
details. See also Boveris, A., and Cadenas, E. (1997) and Pryor, W.A. (1986). *Blood free, isolated
perfused heart. **Personal communications. #Half-life is used here although, in the strictest sense, it
should only be applied to species whose decay is first order.
2. Peroxynitrite reacts with metals to produce a potent nitrating agent similar to
the nitronium ion (heterolytic cleavage of peroxynitrite) (Eqn 2.67).
ONO2- + M2+ → NO2+ + OH- + M+
WWW.ESAINC.COM
Eqn 2.67
103
The importance of this reaction in vivo is not clear, as the amount of free metals
required are well above biological levels (Crow and Beckman (1995)). The
copper atom in CuZn-SOD can, however, react with peroxynitrite to form a
species capable of nitrating phenol whereas free copper and chelated copper
cannot. This suggests that SOD is acting as more than a species capable of
chelating copper. SOD thus appears to play two critical roles in peroxynitrite
chemistry in vivo. First it can regulate peroxynitrite formation by affecting the
availability of superoxide. Second, it can enhance the peroxynitrite’s nitration
pathway. In the absence of a compound capable of undergoing nitration (e.g.,
phenol) CuZn-SOD undergoes self-nitration, without affecting its activity
(Beckman (1996)). It has been proposed that mutations of CuZn-SOD associated
with familial amyotrophic lateral sclerosis permit easier access of peroxynitrite
into the enzyme’s active site (Beckman et al. (1993)). The production of a pronitration species may then explain the increased protein nitration and free
nitrotyrosine found in affected neurons (Abe et al. (1995); Bruijn et al. (1997)).
3. Peroxynitrite causes the oxidation of sulfur-containing groups (e.g., the
oxidation of methionine produces methionine sulfoxide) (Figure 2.19).
4. Peroxynitrite reacts with carbon dioxide/bicarbonate and forms nitrocarbonate
or other potent nitrating species (see below). Due to the abundance of carbon
dioxide/bicarbonate in biological systems, this redirection of peroxynitrite’s
reactivity is very important in vivo.
5. Peroxynitrite indirectly leads to the nitrosation of nucleophiles possibly
through an intermediate X-N=O (where X is –OONO2, -NO2, or –CO3-) (Uppa
et al. (1998)).
The level of peroxynitrite is mainly controlled by the availability of its precursors,
superoxide and nitric oxide (Briviba et al. (1999) and references therein) and by
the action of scavengers. Depending upon the tissue, SOD, nitric oxide synthase,
and oxyhemoglobin all appear to play a major role in controlling peroxynitrite
formation. A number of antioxidant scavengers such as ascorbic acid, uric acid,
and thiols (GSH, cysteine and methionine) may also affect peroxynitrite levels
(Table 2.10 and references therein). In order for a scavenger to be effective it
must react with peroxynitrite in a bimolecular fashion and rapidly enough to
compete with carbon dioxide. Although ascorbic acid is much too slow to be an
effective antioxidant, thiols rapidly react with peroxynitrite (Briviba et al. (1999)).
Some synthetic compounds (e.g., ebselen and iron (III) porphyrin) are extremely
reactive towards peroxynitrite and are of potential use in antioxidant treatment
(Chapter 4) (Squidrito and Pryor (1998)). Interestingly, glutathione peroxidase
(Chapter 4) is even more effective than ebselen at protecting against
peroxynitrite-mediated oxidation and nitration reactions (Sies et al. (1997)).
WWW.ESAINC.COM
104
Peroxynitrite is implicated in a variety of diseases and conditions including
Alzheimer’s, apoptosis, atherosclerosis, cystic fibrosis, endotoxic shock, gastritis,
idiopathic pulmonary fibrosis, inflammation, ischemia/reperfusion injury,
pneumonia, respiratory distress syndrome, rheumatoid arthritis, sepsis, and viral
infection (Halliwell (1997) and references therein; Kaur and Halliwell (1994c);
Kooy et al. (1994); Moriel and Abdalla (1997); Saleh et al. (1997); Smith et al.
(1997); Szabo (1996); van der Veen et al. (1997)). Care should be exercised
though as many reports use the presence of 3-nitrotyrosine as an indicator of the
involvement of peroxynitrite. Unfortunately, 3-nitrotyrosine is difficult to measure
accurately and can also be formed by mechanisms not involving peroxynitrite
(Chapter 4).
A)
ONO2H
NO2
OH
OH
O
-H2O
OH
O
NO2
H
H
R
R
OH
NO2
H
NO2
R
R
R
B)
OH ONO2H HO /NO2
OH
OH
+H2O
OH2
OH
-H+, -H
OH
H
H
R
R
R
R
NO2
OH
OH
NO2
H
-H
+
NO2
H
R
R
C)
O
NH2
OH
S
CH3
O
H
O
N
O
-NO2-
O
NH2
OH
OH
S
CH3
-H+
O
NH2
OH
O
S
CH3
Figure 2.19 Mechanisms For The Reaction Of Peroxynitrite With
Substituted Phenols (A And B) And The Oxidation Of Methionine To
Methionine Sulfoxide (C). Reaction A shows the generally accepted reaction sequence for
the nitration of substituted aromatic compounds. Reaction B takes into account that aromatic
compounds can be hydroxylated or nitrated by peroxynitrite and is based on that proposed by
Pryor and Squadrito (1995). Reaction C shows the oxidation and change in valency of sulfur in
methionine.
WWW.ESAINC.COM
105
Measurement.
Although peroxynitrite levels can be determined using spectrophotometric
techniques by measuring its absorbance at 302nm (Beckman et al. (1994)), the
lack of specificity and sensitivity may render this approach inadequate for most in
vivo investigations. Other methods indirectly measure ONO2-/ONO2H levels by
determining 3-nitrotyrosine concentrations and use either a qualitative
immunological approach or quantitative HPLC techniques. These are discussed
in greater detail in Chapter 3 so will not be dealt with here.
2.4 Nitrosoperoxycarbonate and Nitrocarbonate.
It has been known for many years that peroxynitrite is less stable in carbonate
than in phosphate buffers (Keith and Powell (1969)). Peroxynitrite reacts with
carbon dioxide in a second-order process with a rate constant of 3 x 104 M-1s-1 at
25oC (Lymar and Hurst (1995a,b); Uppa et al. (1996b)). In biological systems
carbon dioxide is in equilibrium with bicarbonate. Because bicarbonate is present
at a concentration typically >25mM (normal carbon dioxide levels are ~1.3mM),
its involvement in peroxynitrite-mediated reactions will be significant. Under
aqueous conditions the most likely reaction is hydrolysis producing carbonate
and nitrate (Vesela and Wilhelm (2002)) thereby preventing peroxynitritemediated damage including the inhibition of the oxidation of thiols,
oxyhemoglobin and cytochrome c2+, and prevention of aromatic oxidation
(hydroxylation) (Denicola et al. (1996); Gow et al. (1996a); Lemercier et al.
(1997); Radi et al. (1999); Uppa and Pryor (1996); Uppa et al. (1996b); Zhang et
al. (1997)). However, in the non-polar environment of membranes reactions will
include aromatic nitration (e.g., production of 3-nitrotyrosine from tyrosine and 8nitroguanine from guanine) and oxidative damage.
Several compounds have been suggested to mediate peroxynitrite/carbon
dioxide reactions. The unstable nitrosoperoxycarbonate anion (ONOOCO2-) and
its product, the more stable but still reactive nitrocarbonate (O2NOCO2-) anion,
are the most likely candidates. Alternatively, the weak peroxo O-O bond in
nitrosoperoxycarbonate could undergo either homolytic cleavage, to produce the
very reactive carbonate radical anion (CO3•-) and nitrogen dioxide (Eqn 2.68), or
heterolytic cleavage, to produce carbonate and nitronium ions (Eqn 2.69) (Bonini
et al. (1999); Denicola et al. (1996); Lymar and Hurst (1995a); Uppa et al.
(1996b)).
ONOOCO2- → CO3•- + NO2•
ONOOCO2- → CO32- + NO2+
WWW.ESAINC.COM
Eqn 2.68
Eqn 2.69
106
Lymar and Hurst (1996) have suggested that nitrosoperoxycarbonate may serve
two important biological functions. First, it acts as a scavenger of peroxynitrite
that, due to its instability, will limit the area of damage caused by this pro-oxidant
(Vasela and Wilhelm (2002)). Second, it may be a superior microbicide to
hydrogen peroxide, as it will not be deactivated by microbial catalase. Again,
unlike hydrogen peroxide that can diffuse a long way from its generation site and
cause damage in areas remote from the site of production, the activity of
nitrosoperoxycarbonate will be limited to the area in which it is produced.
2.5 Nitrogen Dioxide, The Nitronium Cation and Nitrite.
Properties.
Nitrogen dioxide (NO2•) exists as a dense, poisonous, dark brown gas, a pale
yellow liquid with a boiling point of 22oC and, at low temperatures, a pale yellow
solid composed almost entirely of its dimer dinitrogen tetroxide (N2O4). The
nitronium cation (NO2+) is a linear molecule and is isoelectronic to carbon
dioxide. It can be isolated as its stable but very reactive perchlorate salt. Nitrite
(NO2-) is the salt of the weak and unstable acid, nitrous acid (HNO2). With a pKa
of ~3.5, nitrous acid formation is favored by acidic pH.
Formation.
Nitrogen dioxide is made in the laboratory by the reaction of copper with
concentrated nitric acid, by oxidation of nitric oxide, and the thermal
decomposition of metallic nitrates (Eqn 2.70). Biologically nitrogen dioxide is
primarily formed by the oxidation of nitric oxide and possibly by the oxidation of
nitrite by peroxidases (Eqns 2.71-2.73) (Klebanoff (1993)). Nitrogen dioxide is
very soluble in water producing both nitric and nitrous acids (Eqn 2.74).
2Pb(NO3)2 → 2PbO + 4NO2• + O2
Peroxidase + H2O2 → Peroxidase Compound I + H2O
Peroxidase Compound I + NO2- → Peroxidase Compound II + NO2•
Perxoidase Compound II + NO2- → Peroxidase + NO2•
2NO2• + H2O → HNO3 + HNO2
Eqn 2.70
Eqn 2.71
Eqn 2.72
Eqn 2.73
Eqn 2.74
Chemical Reactions and Biological Significance.
Like nitric oxide, nitrogen dioxide is a radical that shows redox behavior. It can
undergo single electron reduction9 to nitrite (NO2-) (Eqn 2.75) or single-electron
9
It can also undergo a two-electron reduction to nitric oxide (e.g., with the oxidation of hydrogen sulfide to sulfur, iodide to
iodine, and sulfur dioxide to sulfate).
WWW.ESAINC.COM
107
oxidation to the nitronium cation (NO2+) (Eqn 2.76). Like nitric oxide it also can
take part in radical-radical interactions (e.g., Eqn 2.77).
NO2• + e- → NO2NO2• - e- → NO2+
NO2• + LO2• → LO2NO2
Eqn 2.75
Eqn 2.76
Eqn 2.77
Nitrogen dioxide is a strong one-electron oxidant (Eo’=+0.99V) that in turn is
reduced to nitrite. Nitrite is further metabolized to nitrate by oxyhemoglobin (Eqn
2.78). The methemoglobin so formed is reduced to hemoglobin enzymatically
using NADPH (see above). A similar reaction occurs when humans consume
foods high in nitrite – used as an antioxidant “cured meats.” This can lead to
methemoglobinemia in humans (Eqn 2.79) (Beckman (1996a)) (Anon. (1992)).
Excessive nitrite consumption can even be fatal (Chilcote et al. (1977); Ellis et al.
(1992)).
4HbO2 + 4NO2- + 4H+ → 4metHb + 4NO3- + O2 + 2H2O
NO2- + 2H+ + Fe2+ → Fe3+ + H2O + NO2•
Eqn 2.78
Eqn 2.79
At the low concentrations typically found in vivo nitrogen dioxide readily initiates
free radical oxidation of proteins and unsaturated lipids (inducing lipid
peroxidation through hydrogen atom abstraction) (Eqn 2.80) (Beckman (1996a)).
At higher concentrations nitrogen dioxide will rapidly react with the radical
produced in (Eqn 2.80) forming organic nitro-derivatives (e.g., nitro-lipid adducts
are produced from lipid radicals and 5-nitro-γ-tocopherol from γ-tocopherol
radicals) (Eqn 2.81) (Cooney et al. (1993, 1995); Huie (1994)). Nitrogen dioxide
can also react with itself or with nitric oxide forming higher oxides (see below).
Nitrogen dioxide can react with the superoxide radical anion to produce
peroxynitrate (O2NO2-). Peroxynitrate shows similar 2-electron oxidation behavior
to peroxynitrite, but different reactivities to carbon dioxide, pH stability and
decomposition pathways (Goldstein, et al., (1998); Olson et al., (2003)). Whether
peroxynitrate is biologically important is debatable as one of its precurorsors,
nitrogen dioxide, is biochemically scarce. Peroxynitrate could possibly play a role
in the phagosome.
γ-Tocopherol-H + NO2• → γ-Tocopheryl• + NO2- + H+
NO2• + γ-Tocopheryl• → γ-Tocopherol-NO2
Eqn 2.80
Eqn 2.81
Under acidic conditions nitrite forms nitrous acid – the latter is in equilibrium with
the nitrosonium ion (Eqn 2.82); consequently, many of the reactions reported
WWW.ESAINC.COM
108
specifically for nitrous acid are a result of this ion, e.g., the nitrosation of amines.
Primary amines produce unstable intermediates that undergo a series of
reactions consisting of N-nitrosation, diazotization, and decomposition. The final
product is dependent upon whether the structure of the original amine is aliphatic
or aromatic in nature. Aliphatic primary amines will lead to the formation of
reactive carbonium ions (Eqn 2.83) while aromatic primary amines form
deaminated products. For example, guanine is deaminated to form xanthine that
can disrupt base-pairing in the DNA molecule (Eritja et al. (1986); Nguyen et al.
(1992); Wink et al. (1991)). Secondary amines (such as dimethylamine and
morpholine) form relatively stable but cytotoxic nitrosamines (Eqn 2.84). The
nitrosonium ion reacts with water and superoxide, producing nitrite and
peroxynitrite, respectively. Thiols produce S-nitrosothiols (Eqn 2.85).
HNO2 + H+ ↔ H2O—NO+
R-NH2 + NO+ → R-NH-N=O + H+ → R-N=N-OH →
R-N2+ + OH- → R+ + N2
R2NH + NO+ → R2NNO + H
RSH + NO+ → RSNO + H+
Eqn 2.82
Eqn 2.83
Eqn 2.84
Eqn 2.85
The nitronium cation is a strong oxidizing agent (Eo’=1600mV). It is also an
aggressive electrophile, readily taking part in electrophilic substitution reactions
of aromatic systems, in which the formation of the carbon-nitrogen bond is the
rate-determining step (Eqn 2.86). In the laboratory, it is formed by the reaction
between concentrated nitric and sulfuric acids (Eqn 2.87). In biological systems,
it appears that a nitronium-like species is produced when peroxynitrite reacts with
metal ions at physiological pH (Beckman et al. (1992); Ischiropoulos et al.
(1992b); Koppenol et al. (1992)).
Ar-H + NO2+ → Ar-NO2 + H+
HNO3 + 2H2SO4 → NO2+ + H3O+ + 2HSO4-
Eqn 2.86
Eqn 2.87
Measurement.
Atmospheric nitrogen dioxide can be determined using chemiluminescent or
voltammetric methods (Goldman and Macrae (1994)). Measurement of aqueous
levels of nitrogen dioxide is difficult due to its reactivity. Nitrogen dioxide can be
determined by measuring of nitrate using voltammetric, reduction, and
spectrophotometric methods (Eaton et al. (1995)) (Table 2.8). Nitrite can be
determined using a variety of methods (Table 2.8).
WWW.ESAINC.COM
109
2.6 The Higher Oxides of Nitrogen – Dinitrogen Trioxide, Dinitrogen
Tetroxide and Dinitrogen Pentoxide
Properties
Dinitrogen trioxide (N2O3), the acid anhydride of nitrous acid, exists as an
unstable blue liquid and solid. Dinitrogen tetroxide (N2O4) is in equilibrium with
nitrogen dioxide and at its freezing point (-11.2oC) is pale yellow, due to the
presence of 0.1% nitrogen dioxide. Dinitrogen pentoxide (N2O5), the acid
anhydride of nitric acid, exists as a colorless non-ionic gas and an ionic solid
composed of nitronium nitrate (NO2+NO3-).
Formation.
The formation of these higher oxides requires the interaction of two nitrogen
oxides. Dinitrogen trioxide is formed by the reaction between nitric oxide and
nitrogen dioxide (Eqn 2.88). Dinitrogen tetroxide is formed when two molecules
of nitrogen dioxide react together (Eqn 2.89). Dinitrogen pentoxide is ultimately
formed from nitric oxide and nitrogen dioxide (Beckman (1996a)).
NO• + NO2• <=> N2O3
2NO2• <=> N2O4
Eqn 2.88
Eqn 2.89
Chemical Reactions and Biological Significance.
As nitrogen dioxide is not particularly abundant in vivo it is more likely that it will
react preferentially with nitric oxide to form dinitrogen trioxide than dimerize to
form dinitrogen tetroxide. In the presence of superoxide, the production of any of
these dimers will be in direct competition with the production of peroxynitrite and
peroxynitrate (see above).
Dinitrogen trioxide is unstable and readily decomposes to nitric oxide and
nitrogen dioxide. As the acid anhydride of the unstable nitrous acid, it reacts with
water to produce nitrite (Eqn 2.90). It is a strong (two-electron) oxidizing and
nitrosating agent.
N2O3 + H2O → 2HNO2 <=> 2H+ + 2NO2-
WWW.ESAINC.COM
Eqn 2.90
110
Dinitrogen tetroxide ionizes under aqueous conditions producing the nitrosonium
ion (Eqn 2.91). It is a strong (two-electron) oxidizing, nitrosylating and nitrating
agent (Eqn 2.92). Dinitrogen pentoxide is a strong oxidizing and nitrating agent.
N2O4 <=> NO+ + NO3N2O4 + 2H+ + 2e- → 2HNO2 (Eo=+1.07V)
Eqn 2.91
Eqn 2.92
As all but nitric oxide occurs at low levels, the biological significance of these
compounds is questionable. The one exception is during phagocytosis where
high levels of nitric oxide and ROS can result in the formation of dinitrogen
trioxide and dinitrogen tetroxide which, in turn are then used to kill pathogens.
Measurement.
Spectrophotometric methods exist for the measurement of dinitrogen oxides
(Feelisch and Stamler (1996)). The measurement of the low levels of these
compounds found in biological systems is difficult.
NH2
NH2
O
O
S-N=O
S-N=O
OH
OH
S -Nitrosohomocysteine
O=N S
HOOC
S -Nitrosocysteine
O=N-S
O
H
N
COOH
O
N
H
S-N=O
NH2
S -Nitrosoglutathione
S -Nitrosoprotein
Figure 2.20 Some S-Nitrosothiols Reported In Vivo.
WWW.ESAINC.COM
111
2.7 S-Nitrosothiols.
Properties.
S-Nitrosothiols (also called thionitrite, sulfenyl nitrites, and thionitrous acid esters)
are highly colored solids and liquids. In general, tertiary S-nitrosothiols are more
stable than primary ones. Several S-nitrosated low molecular weight thiols and
proteins-thiols have been found in vivo (Figure 2.20). Due to their instability only
a few nitrosothiols have been isolated in solid form. Of these, nitroso-albumin can
be kept lyophilized or in solution for several months. Glutathione nitrosothiol can
be kept desiccated in the dark at –20oC for several months and is commercially
available. Cysteinyl nitrosothiol is relatively unstable in solution (seconds to
hours) depending upon temperature, pH, oxygen pressure, and presence of
redox-active species, nucleophiles and trace metals in the solution. In general Snitrosothiol stability is favored by acidic pH.
Formation.
S-Nitrosothiols were first synthesized by Tasker and Jones (1909). Although
often cited in literature, S-nitrosothiols cannot be formed by the reaction between
nitric oxide and a thiol unless a strong electron acceptor is present (Beckman
(1996a)). There are several possible routes in vivo for S-nitrosothiol production
(Crow and Beckman (1995)):
•
•
•
•
•
Thiols can auto-oxidize (in the presence of metal catalysts), forming thiyl
radicals that are capable of reacting with nitric oxide thereby forming
nitrosothiols (Eqn 2.93 and 2.94).
Nitrosothiols can be produced by the action of a nitrosating species, such
as nitrogen dioxide or the nitrosonium ion (from N2O3) – with thiols or thiyl
radicals (Eqns 2.95 and 2.96).
They can be formed by the metal-induced nitrosation of thiols (Eqns 2.97
and 2.98).
Nitrosothiols can be produced by the action of peroxynitrite on thiols (Eqn
2.99).
S-nitrosothiols can be formed by the direct nucleophilic nitrosation of thiols
by peroxynitrite with elimination of hydrogen peroxide (van der Vleit et al.
(1998)).
Of these mechanisms the one catalyzed by a transition metal is probably the
most important biologically, as it occurs at a higher rate and efficiency than all the
other pathways (Crow and Beckman (1995)).
RSH + O2 → RS• + O2•- + H+
RS• + NO• → RSNO
WWW.ESAINC.COM
Eqn 2.93
Eqn 2.94
112
RSH + NO2• → RS• + H+ + NO2(then RS• + N2O3 → RSNO + NO2•)
RSH + N2O3 → RSNO + HNO2
NO• + Fe3+ → Fe2+….NO+
Fe2+….NO+ + RSH → Fe2+ + H+ + RSNO
RSH + ONO2- → RSNO + O2•- + H+
Eqn 2.95
Eqn 2.96
Eqn 2.97
Eqn 2.98
Eqn 2.99
Chemical Reactions and Biological Significance.
S-Nitrosothiols are of great interest to biochemists because they release nitric
oxide under physiological conditions and thus mimic many of the biological
effects reported for nitric oxide. S-nitrosothiols can be regarded as slow releasing
nitric oxide-reservoirs, capable of prolonging the activity of nitric oxide (Keaney et
al. (1993); Stamler et al. (1992a)).
The mechanism(s) by which nitric oxide is released from an S-nitrosothiol under
biological conditions is not clear and still remains controversial. The homolytic
cleavage of the S-N bond is unlikely (Eqn 2.100) even though the reaction is
favored by the formation of a disulfide from two thiyl radicals (Eqn 2.101). The
most likely biological mechanism appears to require additional thiol (or reduced
transition metal) in order to promote the reductive release of nitric oxide from Snitrosothiols (Crow and Beckman (1995)) (Eqn 2.102). An alternate mechanism
requires the formation of a disulfide radical anion (Eqn 2.103) (Beckman
(1996a)).
RSNO → RS• + NO•
2RSNO → RSSR + 2NO•
2RSNO + 2R1SH → 2NO• + 2RSH + R1SSR1
RSNO + RSH → RS•--SR + NO• + H+
Eqn 2.100
Eqn 2.101
Eqn 2.102
Eqn 2.103
S-Nitrosoproteins, formed by post-translational nitrosation of protein cysteinyl
groups, have been proposed to possess a signaling function distinct from nitric
oxide’s ability to directly stimulate cGMP production (see above) (Lipton et al.
(1993); Sucher and Lipton (1991)). Examples include:
•
•
•
The S-nitrosation of a thiol in the active site of cathepsin B that inhibits the
action of this thiol-proteinase;
S-nitrosation of the NMDA receptor infers neuroprotection;
Nitrosation of tissue type plasminogen activator endows this enzyme with
vasodilatory and platelet-inhibitory properties; and
WWW.ESAINC.COM
113
•
Inhibition of glyceraldehyde dehydrogenase by ADP-ribosylation is
possibly mediated by a nitrosothiol intermediate (see Upchurch et al.
(1995) and references therein).
In recognition of the involvement of nitrosothiols in biochemical reactions,
Stamler and Feelisch (1996) have suggested that S-nitrosation may represent a
novel cell regulatory mechanism perhaps as important as phosphorylation.
Measurement.
S-Nitrosothiols have been measured using direct spectroscopy, infrared
spectroscopy, nuclear magnetic resonance, chemiluminescence, colorimetric
assays, CZE-absorbance, HPLC-UV following derivatization, HPLC-ECD,
voltammetry, electrospray ionization mass spectrometry, LC-MS, and EPR of
released nitric oxide (Akaike et al. (1997); Ewing and Janero (1998); Fang et al.
(1998); Kluge et al. (1997); Kostka and Park (1999); Pfeiffer et al. (1998);
Samouilov and Zweier (1998); Tsikas et al. (1999); Vukomanovic et al. (1998);
Wink et al. (1999); also reviewed by Stamler and Feelisch (1996)).
REACTIVE HALOGENATED SPECIES (RHS).
1. Chlorine and Hypochlorous Acid.
Properties.
Chlorine (Cl2) is a greenish-yellow poisonous, diatomic gas that is moderately
soluble in water (reacting to form hydrochloric and hypochlorous acids), but much
more soluble in organic solvents. Chlorine has a melting point of –102oC and a
boiling point of –34.6oC. It is the second member of Group 7B of the periodic
table – the halogens – and has 17 electrons with an electronic configuration of
1s2, 2s2, 2p6, 3s2, 3p5. Hypochlorous acid is unstable and cannot be isolated in
pure form. It commonly occurs as a dilute aqueous solution.
Formation.
Chlorine is prepared in the laboratory by the action of potassium permanganate
on concentrated hydrochloric acid or by heating the latter with manganese
dioxide. Chlorine is also produced when bleaching powder is treated with dilute
acids. Hypochlorous acid is formed when chlorine is passed into water or cold
dilute sodium hydroxide solution (Eqn 2.104).
Cl2 + 2OH- ⇔ OCl- + Cl- + H2O
WWW.ESAINC.COM
Eqn 2.104
114
Both chlorine and hypochlorous acid can be produced by biological systems
(Dunford (2000)). These reactive oxidants produced by phagocytes constitute
part of the host defense mechanism but also play an active role in inflammation.
As discussed above, phagocytosis leads to the production of superoxide by
activation of the NADPH-oxidase complex. The hydrogen peroxide produced by
superoxide dismutation is then used by myeloperoxidase (MPO) for the
production of a variety of bactericidal and cytotoxic species (Kettle and
Winterbourn (1991)).
MPO is a green, heme-containing, non-specific glycoprotein (composed of two
subunits (60,000 and 12,000 Daltons)) that is secreted by activated phagocytes.
It is the most abundant protein found in neutrophils (comprising 5% of their dry
weight) and plays a major role in immune defense mechanisms (Klebanoff
(1988); Weiss and LoBuglio (1982); Weiss and Ward (1982)). MPO uses
hydrogen peroxide to oxidize a variety of halides (and pseudohalides) to their
corresponding hypohalous acids (Eqn 2.105) (Nauseef et al. (1988); van Dalen et
al. (1997)). It is the only human enzyme so far discovered that reacts with
physiological levels of chloride to produce the highly toxic hypochlorous acid
(HOCl) (Eqn 3.18) (Foote et al. (1981); Harrison and Schultz (1976)) – the active
ingredient found in household bleach! MPO is also capable of oxidizing nitrite,
producing the pro-oxidant cytotoxic species, nitrogen dioxide and nitryl chloride
(Eiserich et al. (1997); Klebanoff (1993); van der Vleit et al. (1997)). 3Chlorotyrosine and 3,5-dichlorotyrosine are being used as markers of
hypochlorous acid activity (see Chapter 3 and Figure 2.13).
R- (Cl-, Br-, I-, SCN-) + H2O2 → HOR (HOCl, HOBr, HOI, HOSCN) + H2O
Eqn 2.105
Other reactions of MPO include the novel superoxide-dependent hydroxylation of
a variety of substrates in a process that does not require hydrogen peroxide and
is unaffected by hydroxyl radical scavengers (Kettle and Winterbourn (1994)),
and the one electron oxidation of tyrosine producing tyrosyl radicals. These
radicals are very reactive and can polymerize to form dityrosine and other
addition products (e.g., trityrosine, pulcherosine and isodityrosine) (Jacob et al.
(1996)) (Chapter 3). Differences in reaction products (halogenation, nitration and
oxidation) formed intaphagosomally vs. extracellularly were reviewed by Jiang
and Hurst (1997).
Interestingly, eosinophils contain a similar enzyme eosinophil peroxidase (EPO)
that shows different substrate specificity to MPO (Mayeno et al. (1989)). EPO
preferentially uses bromide to generate a brominating agent (hypobromous acid
(HOBr)), even though physiological levels of bromide (20-100µM) are 1000 fold
less than chloride (140mM). Remarkably, at least 25-30% of the oxygen used by
WWW.ESAINC.COM
115
stimulated eosinophils is directed towards generation of halogenating species
(Mayeno et al. (1989); Weiss et al. (1986)). The major in vivo markers of
hypobromous acid production appear to be 3-bromotyrosine, 3,5-dibromotyrosine
(Wu et al. (1999a)) and a variety of novel brominated-oxysterols. These products
are probably formed by the action of reactive intermediates such as bromamine
(NH2Br) or N-bromamine derivatives (RNHBr) with tyrosine (Hazen and Wu
(1998)). Like MPO, EPO can oxidize nitrite to other RNS, can form tyrosyl
radicals, dityrosine, trityrosine, pulcherosine and 3NT from tyrosine (McCormick
et al. (1998); Wu et al. (1999b)).
Chemical Reactions and Biological Significance.
Chlorine is hydrolyzed by water and bases forming chloride and hypochlorite
(Eqn 2.106). Consequently, under aqueous conditions chlorine gas is in
equilibrium with hypochlorous acid (Eqn 2.106). The production of chlorine will
therefore be favored by acidic pH and chloride ions (Hazen et al. (1996a)). Both
chlorine and hypochlorous acid are extremely oxidizing. Like hydrogen peroxide,
hypochlorous is a poor one-electron (Eqn 2.107) but a strong two-electron
oxidizing agent (Eqn 2.108) (Koppenol (1994)). Chlorine is a powerful twoelectron oxidizing agent (Eqn 2.109).
Cl2 + H2O ⇔ HOCl + Cl- + H+
HOCl + H+ + e- = H2O + Cl•
HOCl + H2O + 2e- = H+ + Cl- + 2OHCl2 (aq) + 2e- → 2Cl-
Eqn 2.106
Eo=-460V
Eqn 2.107
o
E ’=+1.08V Eqn 2.108
Eo=+1.40V Eqn 2.109
Hypochlorous acid can give rise to hydroxyl free radicals by taking part in a
“Fenton-like” reaction with Fe (II) (Eqn 2.110). The hydroxyl free radical can also
be produced by the reaction between hypochlorous acid and superoxide (Eqn
2.26) (Wardman and Candeias (1996)). In fact, this reaction is seven orders of
magnitude faster than the production of hydroxyl free radicals from hydrogen
peroxide (Wardman and Candeias (1996)).10 Interestingly, hydroxyl free radicals
react with chloride ions to produce the hypochlorite radical (Eqn 2.111). These
decompose to produce highly reactive chlorine atoms (Eqns 2.112) (Saran et al.
(1999)). Chlorine atoms can abstract hydrogen atoms from a variety of molecules
(Eqn 2.113) or react with chloride, eventually forming chlorine molecules (Eqns
2.114 and 2.115).
10
Similarly, the pseudohalogen equivalent of hypochlorous acid hypothiocyanate (HOSCN) can take part in “Fenton-like”
reactions. Hypothiocyanate is formed from thiocyanate (SCN ) by the action of hydrogen peroxide. Thiocyanate, in turn, is
a constituent of saliva (Wardman and Candeias (1996) and references therein).
WWW.ESAINC.COM
116
Fe2+ + HOCl → Fe3+ + HO• + ClHO• + Cl- → HOCl•HOCl•- + H+ → Cl•
R-H + Cl• → R• + HCl
Cl• + Cl- → Cl2•2Cl2•- → Cl2 + 2Cl-
Eqn 2.110
Eqn 2.111
Eqn 2.112
Eqn 2.113
Eqn 2.114
Eqn 2.115
The pattern of reaction products formed by the hypochlorous acid/chlorine
system is dependent, in part, on the pH of the reaction medium. For example,
under neutral pH (favoring hypochlorous acid) aliphatic amines are chlorinated to
chloramines (Thomas et al. (1982); Weil and Morris (1949); Weiss et al. (1982)).
Several chloramines can be formed in vivo. Monochloramine, formed by the
action of hypochlorous acid on ammonia (Eqn 2.116), is a lipophilic, short-lived
oxidant that can oxidize thiols, ascorbate, and other compounds. Taurine-Nmonochloramine, formed by the action of hypochlorous acid on taurine (Eqn
2.117), is a hydrophilic, long-lived oxidant that shows limited reactivity. Although
this compound is capable of producing monochloramine (Eqn 2.118) (Grisham et
al. (1984); Weiss et al. (1983)) its intracellular production appears to protect
neutrophils. Chloramine derivatives can decompose to produce reactive carbonyl
compounds (e.g., the chloramine derivative of free serine decomposes to
produce glycoaldehyde, free threonine derivative produces both 2hydroxypropanal and acrolein, while the free tyrosine derivative produces phydroxy-phenylacetaldehyde) (Anderson et al. (1997); Hazen et al. (1996b,
1997)).
HOCl + NH3 → NH2Cl
HOCl + -SO3(CH2)2NH2 → H2O + -SO3(CH2)2NHCl
SO3(CH2)2NHCl + NH3 → NH2Cl + -SO3(CH2)2NH2
Eqn 2.116
Eqn 2.117
Eqn 2.118
Of all the amino acid residues found in a protein only lysine has a free amine
group that can take part in reaction with hypochlorous acid. This protein-bound,
lysine-derived chloramine then undergoes an intermolecular reaction resulting in
the conversion of a nearby tyrosine residue into 3-chlorotyrosine (Domigan et al.
(1995)). Under acidic conditions (favoring chlorine production), direct chlorination
of the tyrosine’s aromatic ring takes place (Hazen et al. (1996b)). The formation,
reactions and detection of 3-chlorotyrosine are discussed further in Chapter 3.
Hypochlorous acid can react with hydrogen peroxide to produce singlet oxygen
(Eqn 2.18). However, as the production of singlet oxygen is not favored by acidic
or neutral pH, the biological importance of this reaction is yet to be established.
The myeloperoxidase-hydrogen peroxide-chloride system also oxidizes thiol
groups; converts methionine residues to methionine sulfoxide; converts cysteine
to cysteine sulfinic acid; bleaches (decolorizes) heme proteins; reacts with ironsulfur centers; produces chlorohydrins from unsaturated fatty acids and
WWW.ESAINC.COM
117
cholesterol; and destroys carotenoids (Albrich et al. (1981); Carr et al. (1997);
Heinecke et al. (1994); Panasenko et al. (1997a); Winterbourn et al. (1992)).
Circulating hypochlorous acid can inhibit α1-antiprotease by oxidizing an
essential methionine residue (Clark et al. (1981)). The extent of inactivation of
this enzyme is dependent upon the site of hypochlorous acid production and the
levels of extracellular antioxidants. For example, the antioxidants albumin and
ascorbic acid can readily react with hypochlorous acid, thereby minimizing its
toxicity. Thus α1-antiprotease present in tissues rich in these antioxidants (e.g.,
blood) will be protected from hypochlorous acid inactivation. Conversely, in
tissues with low levels of albumin and ascorbate (e.g., inflamed rheumatoid
joints) α1-antiprotease will be inactivated. Although taurine (2-aminoethanesulfonic acid) has been proposed as an antioxidant, it reacts with
hypochlorous acid to form taurine-N-chloramines species that are less oxidizing
than hypochlorous acid, but yet are still capable of inactivating α1-antiprotease
(see above). Hypochlorous acid can also directly affect antioxidant enzymes.
Low levels of hypochlorous acid readily inactivate glutathione peroxidase while
moderate levels are required for inhibition of catalase. On the other hand,
superoxide dismutase reacts only slowly with hypochlorous acid, suggesting that
this enzyme may play an important role in controlling hypochlorous acid toxicity
(Aruoma and Halliwell (1987)). The release of hypochlorous acid has been
implicated in the pathogenesis of diseases ranging from atherosclerosis to
ischemia-reperfusion injury and cancer (Hazell et al. (1996); Hazen et al. (1996b)
and references therein).
Measurement.
Atmospheric chlorine can be measured using ion chromatography with
conductivity (NIOSH (1994)). Aqueous chlorine can be measured using
voltammetric and colorimetric approaches (Eaton et al. (1995)). Chlorinated
adducts can be measured using a variety of techniques including GC- and
HPLC-based approaches (Chapter 3).
2. Nitrosyl Chloride, Nitryl Chloride, and Related Compounds.
Properties.
Nitrosyl chloride (NOCl) and nitryl chloride (ClNO2) can be considered formally as
the halide salts of the nitrosonium and nitronium ions, respectively. Nitrosyl
chloride is an orange-yellow colored gas with a melting point of –62oC and boiling
point of –6oC. Nitryl chloride is a colorless gas with a melting point of –145oC and
boiling point of –15oC.
WWW.ESAINC.COM
118
Formation.
Nitrosyl chloride can be made in the laboratory by mixing chlorine and nitric oxide
gases. Nitrosyl chloride (NOCl) is also the active ingredient of aqua regia. This
can be made in the laboratory by mixing concentrated hydrochloric and nitric
acids (3:1 v/v).
Recently, evidence suggests that nitrosyl chloride and nitryl chloride (nitronium
chloride) (ClNO2) can be formed in biological systems and may act as possible
cytotoxic agents. Koppenol (1994) hypothesized that neutrophils and
macrophages might form the highly reactive compound, nitrosyl chloride. Based
upon the energetics of possible reactions, he concluded that the reaction
between hypochlorous acid and peroxynitrite is most likely to produce nitrosyl
chloride (Eqn 2.119).
HOCl + ONO2H → NOCl + O2 + H2O (∆Go’= -79.5kJ mol-1)
Eqn 2.119
Based on the work of Kono (1995), Eiserich et al. (1996) showed that the
reaction between nitrite and hypochlorous acid produces a reactive species
capable of producing 3-nitrotyrosine, 3-chlorotyrosine, and dityrosine from
tyrosine. Initially the reactive specie(s) responsible for these reactions was
thought to be nitryl chloride or the cis- or trans- isomers of chlorine nitrite
(ClONO) (Figure 2.15). Eiserich et al. (1996) hypothesized that such reactive
species can be formed as shown in Eqn 2.120. This is in contrast to the generally
accepted reaction (Eqn 2.121).11 Recent evidence suggests that the pro-oxidant
species formed by the action of MPO on nitrite and hypochlorous acid are nitryl
chloride and nitrogen dioxide (Eiserich et al. (1998)).
2HOCl + 2NO2- → ClONO + ClNO2 + 2OHHOCl + NO2- → HCl + NO3-
Eqn 2.120
Eqn 2.121
Chemical Reactions and Biological Significance.
Both nitrosyl chloride and nitryl chloride react with water to produce a variety of
products including nitrous acid, nitric acid, nitric oxide, and chloride. Nitrosyl
11
This suggests that the approaches that use the measurement of nitrite as an estimate of nitric oxide production in
tissues or fluids from patients with acute or chronic inflammation may be in error due to the removal of nitrite by its
reaction with hypochlorous acid.
WWW.ESAINC.COM
119
chloride is a very reactive substance and its presence in aqua regia enables this
reagent to even dissolve gold and platinum.
Nitrosyl chloride is an electrophilic nitrosating agent and can be regarded as a
carrier of the nitrosonium ion. Thus many of its biological reactions are due to
production of the nitrosonium ion (see above). It does not appear, however to
cause aromatic nitration (Whiteman et al., (2003)). Nitryl chloride promotes lipid
peroxidation and has been suggested to play a role in low-density lipoprotein
modification (Panasenko et al. (1997b)) and DNA damage (Spencer et al.,
(2000)).
Measurement.
Unfortunately, the direct measurement of the low levels of nitrosyl chloride and
nitryl chloride is difficult, especially in vivo (Feelisch and Stamler (1996)).
3d
3p
3s
2p
2s
Sulfur - valency 4
Sulfur - valency 2
1s2, 2s 2, 2p 6, 3s 2, 3p 4
1s2, 2s 2, 2p 6, 3s 2, 3p 3, 3d 1
Sulfur - valency 6
1s2, 2s 2, 2p 6, 3s 2, 3p 3, 3d 2
Figure 2.21 The Electronic Configuration Of Sulfur’s Different Valencies.
WWW.ESAINC.COM
120
SULFUR, THIOLS, AND THIYL RADICALS
(SOME REACTIVE SULFUR SPECIES [RSS]).
Properties.
Sulfur is a yellow solid. The two major allotropes of sulfur consist of S8 molecules
in which single bonds unite the sulfur atoms into puckered octagonal rings.
Rhombic sulfur is the form of sulfur most commonly encountered while
monoclinic sulfur exists at temperatures >95.6oC. This type of allotropy, in which
a definite transition point exists where two forms become equally stable, is called
enantiotropy. Sulfur has a melting point of +118.95oC (monoclinic) and boiling
point of 444.6oC.
Sulfur is the second member of Group 6B of the periodic table. Unlike oxygen,
the first member of Group 6B that only has a valency of 2, sulfur due to its vacant
3d orbitals can show valencies of 2, 4 and 6 (Figure 2.21). Each of these show
various oxidation states. Examples showing the different valencies of sulfur are
presented in Figure 2.22.
Chemical Reactions and Biological Significance.
Sulfur and its compounds show a wide variety of reactions and these can be
found in any good chemical text. Some of these reactions important to the field of
redox biochemistry have been recently reviewed elsewhere (Stamler and Slivka
(1996)). Possibly the best known of all sulfur-based reactions in this field involves
thiols (sulfhydryls). Thiols share many of the chemical characteristics of the
corresponding alcohols but also show many unique reactions. This is due to the
physical properties of the S-H bond, which determines thiol reactivity and
chemistry. The S-H bond is longer and weaker than the O-H bond which affects
its pKa (thiols tend to react as RS- and show nucleophilic activity) and renders it
more easily oxidized. Thiols are therefore very good reducing agents.
The oxidation behavior of a thiol is more complex than the corresponding alcohol.
Alcohol oxidation to aldehydes and acids involves a change in the oxidation state
of the carbon to which the alcohol group is attached (the valency of carbon and
oxygen cannot change) (Eqn 2.122). Oxidation of sulfur is much more
complicated. Mild oxidation of thiols results in the formation of a disulfide (the
valency of sulfur does not change) (Eqn 2.123) and not the corresponding
aldehyde or acid. Under more vigorous conditions the sulfur atom itself can be
oxidized into its higher valencies (e.g., a thiol is first oxidized to a sulfinate and
finally a sulfonate) (Figure 2.22). For example, hypochlorous acid oxidizes
methionine (sulfur valency 2) into methionine sulfoxide (sulfur valency 4) (Eqn
2.124). Ozone can also oxidize sulfur into its highest valency (6) (see above).
WWW.ESAINC.COM
121
R-CH2OH → R-CHO + 2H+ + 2e2R-SH → R-SS-R + 2H+ + 2eR-S-CH3 + HOCl → R-SO-CH3 + HCl
Eqn 2.122
Eqn 2.123
Eqn 2.124
A) Valency 2
HO
γ-Glu-Cys-Gly
S CH3
METHIONINE
γ-Glu-Cys-Gly
S-NITROSOHOMOCYSTEINE
HO
S
S
SH
O
S
NH2
γ-Glu-Cys-Gly
CYSTEINE
GLUTATHIONE
DISULFIDE
DIALLYLSULFIDE
SH HS
HO
S
NH2
NH2
GLUTATHIONE
S-NO
O
O
SH
HO
SH
S
O
HOMOCYSTEINE
DIALLYLDISULFIDE
CO2H
H
NH2
LIPOIC ACID (Reduced)
B) Valency 4
C) Valency 6
O
HO
S CH3
O
NH2
METHIONINESULFOXIDE
γ-Glu-Cys-Gly
γ− Glu-Cys-Gly
SO2H
SO3H
GLUTATHIONESULFINIC ACID
GLUTATHIONESULFONIC ACID
SO3H
H2 N
TAURINE
Figure 2.22 Compounds Illustrating The Different Valencies Of
Sulfur.
The thiol-disulfide redox couple is very important to oxidative metabolism
(Chapter 4). For example, GSH is a reducing cofactor for glutathione peroxidase,
an antioxidant enzyme responsible for the destruction of hydrogen peroxide (Eqn
2.125). The importance of the antioxidant role of the thiol-disulfide redox couple
is discussed further in Chapter 4.
2GSH + H2O2 → GSSG + H2O
WWW.ESAINC.COM
Eqn 2.125
122
Thiols and disulfides can readily undergo exchange reactions, forming mixed
disulfides (Eqn 2.126). Thiol-disulfide exchange is biologically very important. For
example, GSH can react with protein cystine groups and influence the correct
folding of proteins (Hwang et al. (1992); Zingler (1985)). GSH may also play a
direct role in cellular signaling through thiol-disulfide exchange reactions with
membrane bound receptor proteins (e.g., the insulin receptor complex),
transcription factors (e.g., nuclear factor κB), and regulatory proteins in cells
(Powis et al. (1995)). Conditions that alter the redox status of the cell can have
important consequences on cellular function.
R1-SH + R2-SS-R3 → R1-SS-R2 + R3-SH
R
R
R
-2H+, -2e-
R'SH
OH
R'S
O
OH
OH
OH
O
Catechol
Catechol-quinone
5-Thiol Substituted Catechol
NH2
NH2
NH2
H2N
Eqn 2.126
CO2H
H2N
CO2H
HO
S
S
S
HO
OH
NH2
S
HO
CO2H
OH
NH2
HO
CO2H
OH
2-S-CysteinylDopamine
5-S-CysteinylDopamine
2,5-bi-S-CysteinylNorepinephrine
NH2
NH2
H3C
H
GS
H
S-G
HO
S-Protein
HO
OH
OH
5-S-GlutathionylDopamine
Protein-Dopamine
5-S-Adduct
OH
H
HO
OH
2-Glutathionyl4-Hydroxy-Estradiol
Figure 2.23 Formation And Examples Of Biologically Relevant
Thiol Adducts
WWW.ESAINC.COM
123
Thiols are good nucleophiles and readily react with carbonyl compounds to form
hemithioacetals and thioacetals. Consequently, thiols form conjugates with a
variety of biochemicals including purines, estrogens, and monoamine
neurotransmitters (Figure 2.23). Some of these thiol adducts are potentially toxic
and their biological implications will be discussed further in Chapter 4.
Unfortunately, thiols can also undergo one-electron oxidation to produce
chemically reactive thiyl radicals (Eqn 2.127 and 2.128) (D’Aquino et al. (1994);
Hartmann et al. (1999); (Saez et al. (1982)). This can be formed by:
•
•
•
•
•
•
•
•
•
Enzymes (e.g., glutathione reductase, lipoamide dehydrogenase;
peroxidases, (horseradish and lactoperoxidase) thioredoxin reductase,
xanthine oxidase);
Metal-containing proteins (e.g., myoglobin);
Components of the electron transport pathway;
Redox-active metals (e.g., iron, copper);
Hydrogen peroxide;
Other free radicals;
Thermolysis, radiolysis and photolysis of disulfides;
Some sulfur-containing drugs (e.g., penicillamine). This may account for the
autoimmune side effects of some of these compounds (Halliwell and
Gutteridge (1999)). The hemolytic action of diphenyl disulfide is possibly
mediated, at least in part, by thiyl radical production; and
The zinc fingers of some DNA-binding regulatory proteins (Sarkar (1995)).
HO• (RO• or RO2•) + R’SH → H2O (ROH, or RO2H) + R’S•
RSH + Cu2+ → Cu+ + H+ + RS•
Eqn 2.127
Eqn 2.128
Although thiyl radicals are less reactive than hydroxyl free radicals, they still
show considerable reactivity and readily take part in electron transfer, hydrogen
transfer, and addition reactions that can be problematic for the cell (Figure 2.24).
The reaction between the glutathiyl radical and the glutathiolate anion (GS-) or
GSH produces the strongly reducing glutathione disulfide radical (GSSG•-) (Eqns
2.129 and 2.130) that can reduce oxygen to superoxide (Eqn 2.132). Thiyl
radicals can also lead to singlet oxygen production (Wefers and Sies (1983)).
Conversely, thiyl radicals may also be beneficial to the cell. Both free- and
protein contained-thiyl radicals are essential in detoxification of more potent prooxidants (Kalyanaraman (1995)).
2GSH + H2O2 → 2GS• + 2H2O
(e.g., Horseradish peroxidase or lactoperoxidase)
GS• + GSH → (GSSG)•- + H+
WWW.ESAINC.COM
Eqn 2.129
Eqn 2.130
124
(GSSG)•- + O2 → GSSG + O2•-
Eqn 2.131
Thiyl radicals can be detoxified by superoxide dismutase in mammalian cells and
by a thiol-specific enzyme in bacterial cells (Kalyanaraman (1995) and
references therein).
GSH + R
hydrogen
abstraction
GSSG + O2 -
RSH + NAD + H+)
(e.g., NADH + RS
electron
transfer
GSOO
GSNO
RH
addition
Photolysis
O2
O2
GSH
1/2 GSSG
One-electron
Oxidation*
Thermolysis
Photolysis
Radiolysis
GS-
GS
GSSG addition
R-CH=CH-R
R
R
GS
GS- + R +
electron
abstraction
R-CH-CH-R
addition
(e.g., leukotrienes, carotenoids, retinoids)
GSSG + R electron
transfer
*One-electron oxidation can be promoted by: the hydroxyl free radical,
superoxide, copper II (and other metal ions), hydrogen peroxide/iron II, alkyl radical,
alkyl peroxyl radical, ozone and peroxynitrite.
Figure 2.24 Production And Reaction Of Thiyl Radicals. GSH Is
Used As An Example. (D’Aquino et al. (1994); Forni and Wilson (1986a,b);
Hartmann et al. (1999); Karoui et al. (1996); Monig et al. (1987); Mortenson et al.
(1997); Pryor (1984); Quijano et al. (1997); Scorza and Minetti (1998); Singh et al.
(1996); Vasquez-Vivar et al. (1996)).
Measurement.
Thiols and disulfides have been determined using a variety of approaches
including colorimetric, spectrophotometric, enzymatic, and HPLC-based
techniques (see glutathione, Chapter 5). Thiyl radicals can be studied using
optical spectroscopy and EPR in conjunction with spin traps such as PBN and
DMPO (Davies et al. (1987); Kalyanaraman (1995); Kalyanaraman et al. (1996)).
WWW.ESAINC.COM
125
CARBONYL COMPOUNDS.
The carbonyl group (C=O) is one of the most important structures in organic
chemistry. It is found in a wide variety of compounds including aldehydes
(RCHO), ketones (R2CO), amides (RCONH2), carboxylic acids (RCO2H), esters
(RCO2R), acid anhydrides ((RCO)2O) and acid chlorides (RCOCl). To cover the
chemistry and biochemistry of all carbonyl compounds found in biological
systems is way beyond the scope of this handbook. We will instead concentrate
on some biologically significant reactive aldehydes.
Properties.
The simplest aliphatic aldehyde is formaldehyde (HCHO), a pungent-smelling
gas that is extremely soluble in water. Short chain (carbon length 2 to 9) aliphatic
aldehydes are clear liquids with distinct odors. Their solubility in water decreases
with increasing chain length.
Formation.
Aldehydes can be formed in the laboratory by a variety of approaches, including
careful oxidation of a primary alcohol, Rosenmund’s synthesis (chemical
reduction of an acid chloride), ozonolysis of alkenes, or by heating formic acid
and a carboxylic acid with manganous oxide at 300oC.
A number of reactive aldehydes are found in vivo (Figure 2.25) and are:
•
•
•
•
Derived from the diet (e.g., glucose can be considered as a reactive
aldehyde and is capable of undergoing glycation and glyoxidation)
Formed as part of normal metabolism (e.g., the production of
catecholaldehyde by the action of monoamine oxidase on
catecholamines).
Produced by the action of reactive species on amino acids (e.g., the
formation of acrolein, glycolaldehyde and 2-hydroxypropanal when
hypochlorous acid reacts with amino acids).
Result
from
lipid
peroxidation
(e.g.,
4-hydroxynonenal
and
malondialdehyde) and this will discussed further in Chapter 3.
Cytotoxic aldehydes are also found in automobile exhaust, cigarette smoke,
cooking emissions, and drinking water. They are also produced during the
metabolism of some anticancer drugs (e.g., oxazaphosphorines) (Ghilarducci
and Tjeerdema (1995); Jakab (1977); Sladek (1987)).
WWW.ESAINC.COM
126
NH 2
HO
HOCl
O
O
HO
OH
Glycolaldeyhde
Serine
NH 2
OH
OH
HOCl
O
CH
+
H2O
OH
Threonine
H2 N
O
O
2-Hydroxypropanal
Acrolein
CO 2H
CHO
CH 2
CH 2
HOCl
OH
OH
4-Hydroxyphenyl
acetaldehyde
Tyrosine
R
CHO
OH
4-Hydroxynonenal
CO 2H
Poly-unsaturated
Fatty Acid
H
O=C
H
C=O
Malondialdehyde
H2 O2
O2 , H 2 O
O
NH
NH
O
4,9-Diazadodecanedial
O
Spermine/
Spermidine
NH 2
Serum Amine
Oxidase
Acrolein
O
H2N
3-Aminopropanal
HN
O
N-(4-Aminobutyl)3-aminopropanal
CH 2CH 2NH 2
O2
MAO
HO
OH
Catecholamine
CH 2CH=NH
H2 O2
HO
OH
H2O
Catecholimine
CH 2CHO
NH 3
HO
OH
Catecholaldehyde
Figure 2.25 Some Reactive Carbonyls Formed In Vivo.
WWW.ESAINC.COM
127
Chemical Reactions and Biological Significance.
The chemical reactions of aldehydes will not be dealt with in depth here.
Aldehydes are fairly reactive and can undergo reduction (to alcohols), oxidation
(to acids), or nucleophilic addition with alcohols (forming hemi-acetals and
acetals), thiols (forming hemi-thioacetals and thioacetals), hydrogen cyanide
(forming cyanohydrin) and hydrogen sulfite (forming hydrogen sulfonic acid
salts). Unsaturated aldehydes where the carbon-carbon double bond is located in
the 2-position (e.g., 2-alkenals – acrolein and 4-hydroxynonenal) are much more
reactive than the corresponding saturated aldehydes. Acrolein will therefore
show the chemical reactions typical for compounds containing a carbonyl and a
double bond. 4-Hydroxynonenal has three functional sites, and will undergo
reactions typical of the carbonyl, double bond and alcohol groups. 2-Alkenals
undergo Michael addition with nucleophiles, such as thiols and amines, and lead
to the formation of cyclic hemi-acetals and protein adducts, respectively (Figure
2.26) (Sayre et al. (1993)).
SR
SR
R
O
R
O
OH
Michael Adduct
RSH
Cyclic Hemiacetal
Protein
Protein-NH2
R
O
OH
NH
O
R
OH
OH
4-Hydroxy-2-nonenal
Michael Adduct
Protein-NH2
R
R
NH
N
O
OH
Schiff Base
Protein
Protein
OH
-H2O
R
R
N
Protein
N
Protein
Pyrrole
Figure 2.26 The Reaction Of The Cytotoxic Lipid Peroxidation Product
4-Hydroxy-2-Nonenal With Thiols (RSH) Or Proteins (Protein-NH2).
WWW.ESAINC.COM
128
Aldehydes can react with nucleophiles that contain an acidic proton. A
subsequent elimination is then possible, leading to complete substitution. A
typical example of this is the Schiff base (imine) formation between an aldehyde
and a primary amine (e.g., amino acid, purine or pyrimidine) (Eqn 2.132). The
formation of Schiff bases is important to the activity of many enzymes. Here the
ε-amine group of a lysine residue located on the enzyme reacts with carbonyl
substrate (e.g., aldose, keto acid or pyridoxal phosphate). Examples of enzymes
that make use of Schiff bases include aldolase, aminotransferases, lysyl oxidase,
phosphorylase, rhodopsin, and transaldolases – the production of unwanted
Schiff bases however is a major problem for the cell. It is the first step in several
non-enzymatic reactions. For example, the reaction between aldehydes and DNA
bases can lead to the formation of base adducts and, in the case of 4hydroxynonenal and malondialdehyde, to DNA-DNA and DNA-protein crosslinks. Formation of these metabolites can interfere with normal DNA replication
(Chapter 3).
RICHO + RIINH2 ⇔ RICH=NRII + H2O
Eqn 2.132
Proteins can react with reducing (aldehyde-containing) sugars in a nonenzymatic, non-oxidative process termed glycation. The irreversible autooxidation of glycated products, glycoxidation, leads to the formation of advanced
glycation end products (AGEs) such as carboxymethyl lysine and pentosidine
(Figure 2.27). Proteins can also react with reactive carbonyls (e.g., 2-alkenals
form fluorescent pyrrole derivatives) called “advanced lipid peroxidation end
products” (ALPs) or “advanced lipoxidation end products” (ALEs) (Figure 3.23)
(Sayre et al. (1993, 1996)). AGEs and ALPs are associated with aging and the
oxidative stress pathophysiology of neurodegenerative diseases, diabetes, and
uremia (Calingasan et al. (1999); Miyata et al. (1999); Xu and Sayre (1998)).
Patients presenting with carbonyl overload are said to be under “carbonyl stress”
(Miyata et al. (1999)).
The biological importance of the reactions between aldehydes, DNA and proteins
is discussed in Chapter 3.
Measurement.
The presence of aldehydes can be determined by their ability to restore the
magenta color to Schiff’s reagent (sulfur dioxide treated rosanaline
hydrochloride). Alternatively, as they are reducing agents, their presence is
indicated by the production of a silver mirror when treated with Tollen’s reagent
(ammoniacal silver nitrate) or a precipitate of red Cu (I) oxide when treated with
Fehling’s solution (an alkaline complex of copper tartrate).
WWW.ESAINC.COM
129
Aldehydes present in biological samples can be quantified using the
thiobarbituric acid test (TBAR), GC, GC-MS, and following derivatization (e.g.,
2,4-dinitrophenylhydrazine), by HPLC with UV, fluorescence or electrochemical
detection (Chapter 3).
A) GLYCATION
R
R
R
(CH 2 )4
(CH 2 )4
(CH 2 )4
NH
NH
NH
CH
CH
Lysine
-H 2 O
H
H
R
(CH 2 )4
NH 2
O
H
C
C
OH
(CHOH) 3
CH 2 OH
C
OH
C
(CHOH) 3
Amadori
Rearrangement
OH
(CHOH) 3
CH 2
C
O
(CHOH) 3
CH 2 OH
CH 2 OH
CH 2 OH
Schiff Base
(Imine)
Eneaminol
Fructose-Lysine
(Amadori Product)
Glucose
HOCH 2
CHO
N
(CH 2 )4
CH
H 2N
CO 2 H
Pyrraline
R
B) GLYCOXIDATION
C) OTHER AGEs
(CH 2 )4
NH
R
CH 2
(CH 2 )4
O
Arginine
Imidazolone
Methylglyoxal
Lysine
Imidazole
Carboxymethyl
Lysine (CML)
Metal n+
CH 2
C
CO 2 H
[O 2 ]
NH
Glyoxal
Arginine
(CHOH) 3
CH 2 OH
Lysine
Fructose-Lysine
(Amadori Product)
NH
Arginine
CH NH 2
CH NH 2
N
NH
CO 2 H
CO 2 H
N
N
N
Pentosidine
(CH 2 )4
(CH 2 )4
H 3C
+
N
N
(CH 2 )4
(CH 2 )4
CH
H 2N
CO 2 H
Methylglyxoyllysine
Dimer
CH
H 2N
CO 2 H
Glyoxallysine
Dimer
Figure 2.27 A) Glycation Reactions, B) Glycoxidation Reactions and
C) Some Additional AGEs.
WWW.ESAINC.COM
130
THE PRO-OXIDANT ACTIVITY OF LOW MOLECULAR WEIGHT
COMPOUNDS AND OTHER XENOBIOTICS.
A number of xenobiotic compounds owe their biological activity to their ability to
generate pro-oxidants. However, this ability is also responsible for their
undesirable side effects and unwanted toxicity. Table 2.12 presents a brief, nonexhaustive list of several classes of pro-oxidant xenobiotic compounds. Readers
wanting a more comprehensive review of the toxicology of these compounds are
referred to Halliwell and Gutteridge (1999), Kehrer (1993) and Sies and de Groot
(1992).
Compound Class
or Drug Class
Acrylates
Antibiotics
Example
References
Bone cement
Cephalosporin,
chloramphenicol,
gentamicin,
rifamycin,
tetracyclines
Vale et al. (1997)
Halliwell and
Gutteridge (1999);
Muller et al. (1998)
Antihypertensives
Hydralazine
Sodium nitroprusside
Artimisinin
Methotrexate
Anthracyclins and other quinone containing
drugs (daunomycin, doxorubicin, mitomycins,
steptonigrin)
Cisplatin
Metal chelators (bleomycins, tallysomycin)
Protein antitumor antibiotics (macromycin,
neocarzinostatin)
Other - tirapazamine
Benzene and aniline and their derivatives
Rauhala et al. (1998)
Antimalarial agents
Antirheumatics
Antitumor agents
•
•
•
•
•
Aromatic
hydrocarbons
Aromatic
polyphenols (also
quinones and
hydroxylatedmethoxylated
compounds)
Aromatic quinones
and derivatives
Bipyridylium
compounds
Diabetogenics
L-DOPA, dopamine, etoposides,
6-hydroxydopamine,
Anthraquinones, benzoquinones,
indolequinones and naphthaquinones and
their derivatives
Diquat, paraquat
Alloxan, streptozotocin
WWW.ESAINC.COM
Meshnick (1998)
Gressier et al. (1994)
Baliga et al. (1998);
Evans et al. (1998);
Gewirtz (1999);
Keizer et al. (1990);
Muller et al. (1998);
Sinha and Politi
(1990)
Brennan and Schiestl
(1997); Shen et al.
(1996)
Halliwell and
Gutteridge (1999);
Woodgate et al.
(1999)
Everett et al. (1998);
Gatto et al. (1996);
Fukushima et al.
(1993); Halliwell and
Gutteridge (1999);
Yamada and
Fukushima (1993)
Halliwell and
131
•
•
•
Gutteridge (1999)
El-Bachs, et al.
(1999); Yu et al.
(1999)
Halliwell and
Gutteridge (1999);
Henderson et al.
(1999)
Crebelli et al. (1999);
Halliwell and
Gutteridge (1999)
Halliwell and
Gutteridge (1999);
Sugiyama (1992)
Halliwell and
Gutteridge (1999);
Munday (1989)
•
•
Parman et al. (1999)
Daniel et al. (1995);
Lund and Aust
(1991); Vallyathan
(1994)
Hiramoto et al. (1995)
Avent et al. (1996);
Behl et al. (1996);
McNaught et al.
(1998); Naoi et al.
(1998)
Halliwell and
Gutteridge (1999)
Dopaminergics
Apomorphine, cocaine
Ethanol
Halogenated
hydrocarbons
Bromobenzene, carbon tetrachloride,
chloroform, dibromoethane, halothane
Heavy metals
Chromium, lead, manganese, mercury,
titanium, vanadium
Hemolytic agents
Immunomodulator
Mineral dusts
Aminothiols, thiophenols
Favism agents (convicine, vicine)
Hydrazines (acetylphenylhydrazine,
phenylhydrazine, iproniazid, isoniazid)
• Quinones (juglone, lawsone, plumbagin and
menadione)
Thalidomide
Silicates e.g., asbestos
Mushroom toxins
Neurotoxins
Nitro-aromatics
Organic
hydroperoxides
and peroxides
Photosensitizing
agents
Quinoneimines
Benzenediazonium salts
Amphetamine and derivatives,
6-hydroxydopamine, isoquinolines, haloperidol
metabolites, MPTP
•
•
•
•
Fungal agents (sporidesmin)
Radiosensitizing agents/hypoxic cell
sensitizers (chloramphenicol, furazolidone,
metronidazole, misonidazole, nitrofurantoin,
nitrofurazone)
Benzoyl peroxide, cumene hydroperoxide, tertbutyl hydroperoxide
Dyes (e.g., indocyanine green)
Furocoumarins (psoralens)
Porphyrins
Quinines and antimalarials (e.g., chloroquine,
primaquine and quinacrine)
E.g., acetaminophen metabolite
Halliwell and
Gutteridge (1999);
Sestili et al. (1998)
Baumler et al. (1999);
Bonnett and
Berenbaum (1989);
Moreno (1986);
Potapenko (1991);
Spikes (1998)
Halliwell and
Gutteridge (1999)
Table 2.12 The Pro-Oxidant Activity Of Some Low Molecular Weight
Compounds And Other Xenobiotics. The separation of compounds based upon their
compound class or drug class is not perfect as some compounds may fall into more than one
category. For example, some compounds are antiobiotics possess a quinones structure and are
antitumor agents.
WWW.ESAINC.COM
132
The mechanism by which a pro-oxidant xenobiotic produces oxidative stress is
dependent, in part, upon its chemical structure. A number of xenobiotics can
undergo redox cycling producing ROS. For example, Figure 3.24 shows redox
cycling of bipyridyl herbicide paraquat and the diabetogenic agent alloxan. Other
compounds that can redox cycle include antibiotics (e.g., actinomycin D,
mitomycin C and streptonigrin); antitumor drugs (e.g., anthracyclines, etoposides,
tirapazamine, diaziridinylbenzoquinones, and EO9); and the hydroxylated
metabolites of the antimalarial drug primaquine (Butler (1998); Halliwell and
Gutteridge (1999); Newsholme and Leech (1992); Vasquez-viva and Augusto
(1992)).
The generation of ROS by redox cycling is only one possible explanation for the
action of many drugs. Rifamycin not only owes its activity to ROS generation but
also to its ability to block bacterial RNA synthesis as well. Quinones (and/or
semiquinones) can also form adducts with nucleophiles, especially thiols (Figure
2.6; Figure 2.23; Chapter 4). These adducts may act as toxins directly or
indirectly through the inhibition of key enzymes (e.g., by reacting with essential
cysteinyl residues) or the depletion of GSH.
To go through all the mechanisms of action for the list of compounds presented
in Table 2.11 is beyond the scope of this book. Rather we have chosen to give
four brief examples:
•
The bipyridyl herbicides (e.g., paraquat and diquat) are toxic to both plants
and animals (reviewed by Halliwell and Gutteridge (1999)). In animals
paraquat is converted to its radical by microsomal NADPH cyctochrome
P450 reductase (or cytochrome c oxidase). The paraquat radical can then
reduce oxygen to superoxide (Figure 2.8). The ROS produced by this
redox cycling readily explains why paraquat causes such damage to
tissues especially the lungs (pulmonary edema and alveolar inflammation).
•
MPTP (1-methyl-4-phenyl-1,2,3,4-tetrahydropyridine), a neurotoxin first
discovered in 1979 as a contaminant in a “designer drug” preparation of a
meperidine analog, causes a Parkinson’s syndrome, a consequence of its
damage to the dopaminergic nigrostriatal pathway. For its toxic action,
MPTP must first be converted to its neurotoxic pyridinium analyte MPP+
through oxidation by glial monoamine oxidase (MAO B). MPP+, is then
actively accumulated by dopaminergic neurons by the dopamine
transporter. Although MPP+ shows some structural similarity to paraquat it
does not appear to exert its pro-oxidant action through redox cycling.
Instead MPP+ exerts its toxicity by inhibiting the mitochondrial electron
transport chain at complex I. This not only prevents the production of the
essential metabolite, ATP but also produces ROS and leads to oxidative
stress. Other neurotoxic drugs include a pyridinium metabolite of the
neuroleptic agent, haloperidol, that may account for some of haloperidol’s
unwanted neurotoxic side effects, and the isoquinolinium metabolites of the
WWW.ESAINC.COM
133
isoquinolines, that some suggest might be endogenous Parkinson-causing
agents. Interestingly, these pyridinium and isoquinolinium compounds are
structurally similar to MPP+.
•
Halogenated hydrocarbons exert their pro-oxidant activity through the
formation of carbon-centered radicals and/or corresponding peroxy
radicals (through their reaction with oxygen) (Halliwell and Gutteridge
(1999)). We shall concentrate on tetrachloromethane (CCl4; TCM) (a
constituent of dry cleaning fluid) as a representative example. TCM is an
active pro-oxidant, affecting many organs but particularly the liver. The
first step in TCM activation is the production of the trichloromethyl radical
(CCl3•) by microsomal P450 (other enzymes can also promote this
reaction). The trichloromethyl radical can then follow three pathways. First,
it can directly attack macromolecules forming covalent bonds. Second, it
can abstract hydrogen atoms directly from lipids and initiate lipid
peroxidation. Third, it can react with oxygen to form the reactive
trichloromethylperoxy radical (CCl3O2•) that effectively initiates lipid
peroxidation (Chapter 3) and can covalently bind to important
macromolecules.
•
Finally, some pro-oxidants act through the production of singlet oxygen
(above). Such compounds include a diverse group of photosensitizing
agents, some of which are used as sensitizing agents in photodynamic
therapy.
REFERENCES.
Abe, K., Pan, L.H., Watanabe, M., Kato, T., and Itoyama, Y. (1995). Induction of nitrotyrosine-like immunoreactivity in the
lower motor neuron of amyotrophic lateral sclerosis. Neurosci. Lett., 199, 152-154.
Acworth, I.N., McCabe, D.R., and Maher, T.J. (1997). The analysis of free radicals, their reaction products, and
antioxidants. In: Oxidants, Antioxidants, and Free Radicals. Baskin, S.I., and Salem, H. (Eds.). Taylor and Francis,
Washington, DC. Pp. 23-77.
Acworth, I.N., Bailey, B.B., and Maher, T.J. (1998a). The use of HPLC with electrochemical detection to monitor reactive
oxygen and nitrogen species, markers of oxidative damage and antioxidants: application to the neurosciences. In:
Neurochemical Markers of Degenerative Nervous Diseases and Drug Addiction. Qureshi, G., Parvez, H., Caudy, P.,
and Parvez, S. (Eds.). Progress in HPLC-HPCE, 7. VSP Publications, The Netherlands. Pp. 3-56.
Acworth, I.N., Bogdanov, M.B., McCabe, D.R., and Beal, M.F. (1998b). Improved methods for estimation of hydroxyl free
radical levels in vivo based on liquid chromatography with electrochemical detection. Meth. Enzymol., 300, 297-313.
Akaike, T., Inoue, K., Okamoto, T., Nishino, H., Otagiri, M., Fujii, S., and Maeda, H. (1997). Nanomolar quantification and
identification of various nitrosothiols by high performance liquid chromatography coupled with flow reactors of metals
and Griess reagent. J. Biochem., 122, 459-466.
Albrich, J.M., McCarthy, C.A., and Hurst, J.K. (1981). Biological activity of hypochlorous acid: Implications for microbicidal
mechanisms of leukocyte myeloperoxidase. Proc. Natl. Acad. Sci. USA, 78, 210-214.
Aleryani, S., Milo, E., Rose, Y., and Kostka, P. (1998). Superoxide-mediated decomposition of S-nitrosothiols. J. Biol.
Chem., 273, 6041-6045.
Allgayer, H., Hofer, P., Schmidt, M., Bohne, P., Kruis, W., and Gugler, R. (1992). Hydroxyl and fatty acid radical
scavenging by aminosalicylates. Biochem. Pharmacol., 43, 259-262.
Alvarez, B., Rubbo, H., Kirk, M., Barnes, S., Freeman, B., and Radi, R. (1996). Peroxynitrite-dependent tryptophan
nitration. Chem. Res. Toxicol., 9, 390-396.
Anderson, M.M., Hazen, S.L., Hsu, F.F., and Heinecke, J.W. (1997). Human neutrophils employ the myeloperoxidasehydrogen peroxide-chloride system to convert hydroxy-amino acids into glycoaldehyde, 2-hydroxypropanal, and
acrolein. J. Clin. Invest., 99, 424-432.
Anon. (1974). Nitrates and nitrites in food. Med. Lett. Drugs Ther., 16, 75-76.
WWW.ESAINC.COM
134
Aoki, T., (1990). Continuous flow determination of nitrite with membrane separation/chemi-luminescence detection.
Biomed. Chromatogr., 4, 128-130.
Ayoyagi, K., Akiyama, K., Kuzure, Y., Takemura, K., Nagase, S., Ienaga, K., Nakamura, K., Koyama, A., and Narita M.
(1998a). Synthesis of creatol, a hydroxyl radical adduct of creatinine and its increase by puromycin aminonucleoside in
isolated rat hepatocytes. Free Radic. Res., 29, 221-226.
Aoyagi, K., Nagase, S., Koyama, A., Narita, M., and Tojo, S. (1998b). Products of creatinine with hydroxyl radical as a
useful marker of oxidative stress in vivo. Methods Mol. Biol., 108, 157-164.
Arroya, C.M., Forray, C., El-Fakahany, E.E., and Rosen, G.M. (1990). Receptor-mediated generation of an EDRF-like
intermediate in a neuronal cell line detected by spin-trapping techniques. Biochem. Biophys. Res. Commun., 170,
1177-1183.
Aruoma, O.I., and Halliwell, B. (1987). Action of hypochlorous acid on the antioxidant protective enzymes superoxide
dismutase, catalase and glutathione peroxidase. Biochem. J., 248, 973-976.
Babbs, C.F., and Griffin, D.W. (1989). Scatchard analysis of methane sulfinic acid production from dimethyl sulfoxide: a
method to quantify hydroxyl radical formation in physiologic systems. Free Radic. Biol. Med., 6, 493-503.
Babior, B.M. (1978). Oxygen dependent microbial killing by phagocytes. New Engl. J. Med., 298, 659-668; 721-725.
Badwey, J.A., Curnutte, J.T., and Karnovsky, M.L. (1979). The enzymes of granulocytes that produce superoxide and
peroxide. New Engl. J. Med., 300, 1157-1160.
Balcioglu, A., and Maher, T.J. (1993). Determination of kainic acid-induced release of nitric oxide using a novel
hemoglobin trapping technique with microdialysis. J. Neurochem., 61: 2311-2313.
Bartlett, D., Church, D.F., Bounds, P.L., and Koppenol, W. (1995). The kinetics of the oxidation of L-ascorbic acid by
peroxynitrite. Free Radic. Biol. Med., 18, 85-92.
Basu-Modak, S., and Tyrrell, R.M. (1993). Singlet oxygen: a primary effector in the ultraviolet A/near-visible light induction
of the human heme oxygenase gene. Cancer Res., 53, 4505-4510.
Bates, T.E., Loesch, A., Burnstock, G., and Clark, J.B. (1995). Immunocytochemical evidence for a mitochondrially
located nitric oxide synthase in brain and liver. Biochem. Biophys. Res. Commun., 213, 896-900.
Baur, J. E., Wang, S., and Brandt, M. C. (1996). Fast-scan voltammetry of cyclic nitroxide free radicals. Anal. Chem., 68,
3815-3821.
Beal, M.F. (1997). Oxidative damage in neurodegenerative diseases. Neuroscientist, 3, 21-27.
Beauchamp, C., and Fridovich, I. (1970). A mechanism for the production of ethylene from methional. The generation of
the hydroxyl radical by xanthine oxidase. J. Biol. Chem., 245, 4641-4646.
Beckman, K.B., and Ames, B.N. (1998). Free radical theory of aging matures. Physiol. Rev., 71, 547-581.
Beckman, J.S. (1996a). The physiological and pathological chemistry of nitric oxide. In: Nitric Oxide: Principals and
Actions. Lancaster, J. (Ed.). Academic Press, Inc., New York. Pp. 1-82.
Beckman, J.S. (1996b). Oxidative damage and tyrosine nitration from peroxynitrite. Chem. Res. Toxicol., 9, 836-844.
Beckman, J.S., Beckman, T.W., Chen, J., Marshall, P.A., and Freeman, B.A. (1990). Apparent hydroxyl radical production
by peroxynitrite: implications for endothelial injury from nitric oxide and superoxide. Proc. Natl. Acad. Sci. USA, 87,
1620-1624.
Beckman, J.S., Carson, M., Smith, C.D., and Koppenol, W.H. (1993). ALS, SOD, and peroxynitrite. Nature, 364, 584.
Beckman, J.S., Chen, J., Ischiropoulos, H., and Crow, J.P. (1994). Oxidative chemistry of peroxynitrite. Meth. Enzymol.,
233, 229-240.
Beckman, J.S., Ischiropoulos, H., Zhu, H., van der Woerd, M., Smith, C., Chen, J., Harrison, J., Martin, J.C., and Tsai, M.
(1992). Kinetics of superoxide dismutase and iron catalyzed nitration of phenolics by peroxynitrite. Arch. Biochem.
Biophys., 298, 438-445.
Bhardwaj, A., Northington, F.J., Koehler, R.C., Stiefel, T., Hanley, D.F., and Traystman, R.J. (1995). Adenosine modulates
n-methyl-d-aspartate-stimulated hippocampal nitric oxide production in vivo. Stroke, 26, 1627-1633.
Bindoli, A., Rigobello, M.P., and Deeble, D.J. (1992). Biochemical and toxicological properties of the oxidation products of
catecholamines. Free Radic. Biol. Med., 13, 391-405.
Bindoli, A., Rigobello, M.P., and Galzinga, L. (1989). Toxicity of aminochromes. Toxicol. Lett., 48, 3-20.
Bishop, A., Marquis, J.C., Cashman, N.R., and Demple, B. (1999). Adaptive resistance to nitric oxide in motor neurons.
Free Radic. Biol. Med., 26, 978-986.
Bogdanov, M.B., Ferrante, R.J., Kuemmerle, S., Klivenyi, P., and Beal, M.F. (1998a). Increased vulnerability of 3nitropropionic acid in an animal model of Huntington’s disease. J. Neurochem., 71, 2642-2644.
Bogdanov, M.B., Ramos, L.E., Xu, X., and Beal, M.F. (1998b). Elevated hydroxyl radical generation in vivo in an animal
model of amyotrophic lateral sclerosis. J. Neurochem., 71, 1321-1324.
Bogdanov, M.B., Ramos, L.E., Martinou, J-C., and Beal, M.F. (1998c). Oxidative stress is attenuated in mice
overexpressing BCL-2. J. Neurochem. Accepted for publication.
Bonini, M.G., Radi, R., Ferrer-Sueta, G., Ferreira, A.M., and Augusto, O. (1999). Direct EPR detection of the carbonate
radical anion produced from peroxynitrite and carbon dioxide. J. Biol. Chem., 274, 10802-10806.
Bors, W., Michel, C., and Saran, M. (1979). On the nature of biochemically generated hydroxyl radicals. Studies using the
bleaching of p-nitrosodimethylaniline as a direct assay method. Eur. J. Biochem., 95, 621-627.
Bouton, C., Hirling, H., and Drapier, J.-C. (1997). Redox modulation of iron regulatory proteins by peroxynitrite. J. Biol.
Chem., 272, 19969-19975.
Boveris, A., and Cadenas, E. (1997). Cellular sources and steady-state levels of reactive oxygen species. In: Oxygen,
Gene Expression, and Cellular Function. Lung Biology in Health and Disease. Clerch, L.B., and Massaro, D.J. (Eds.).
Volume 105, Marcel Dekker, New York. Pp. 1-25);
Bovis, A., and Chance, B. (1973). The mitochondrial generation of hydrogen peroxide. Biochem. J., 134, 707-716.
Bredt, D.S., and Snyder, S. (1994). Nitric oxide: A physiologic messenger molecule. Ann. Rev. Biochem., 63, 175-195.
Briviba, K., Devasagayam, T.P., Sies, H., and Steenken, S. (1993). Selective para hydroxylation of phenol and aniline by
singlet molecular oxygen. Chem. Res. Toxicol., 6, 548-553.
WWW.ESAINC.COM
135
Briviba, K., Kissner, R., Koppenol, W.H., and Sies, H. (1998). Kinetic study of the reaction of glutathione peroxidase with
peroxynitrite. Chem. Res. Toxicol., 11, 1398-1401.
Briviba, K., Klotz, L.-O., and Sies, H. (1999). Defenses against peroxynitrite. Meth. Enzymol., 301, 301-311.
Bruijn, L.I., Beal, M.F., Becher, M.W., Schulz, J.B., Wong, P.C., Price, D.L., and Cleveland, D.W. (1997). Elevated free
nitrotyrosine levels, but not protein-bound nitrotyrosine or hydroxyl radicals, throughout amyotrophic lateral sclerosis
(ALS)-like disease implicate tyrosine nitration as an aberrant in vivo property of one familial ALS-linked superoxide
dismutase 1 mutant. Proc. Natl. Acad. Sci. USA, 94, 7606-7611.
Brune, D., Dimmeler, S., Molina, V., and Lapetina, E.G. (1994). Nitric oxide: A signal for ADP-ribosylation of proteins. Life
Sci., 54, 61-70.
Burkitt, M.J., and Gilbert, B.C. (1991). The autoxidation of iron(II) in aqueous systems: The effects of iron chelation by
physiological, non-physiological and therapeutic chelators on the generation of reactive oxygen species and the
inducement of bio-molecular damage. Free Radic. Res. Commun., 14, 107-123.
Burlet, S., and Cespuglio, R. (1997). Voltammetric detection of nitric oxide (NO) in the rat brain: Its variations throughout
the sleep-wake cycle. Neurosci. Lett., 226, 131-135.
Butler, J. (1998). Redox cycling antitumor drugs. In: DNA and Free Radicals: Techniques, Mechanisms and Applications.
Aruoma, O.I., and Halliwell, B. (Eds.). OICA, London, 131-159.
Buxton, G.V., and Mulazzani, Q.G. (1999). Free radicals as a source of uncommon oxidation states of transition metals.
In: Metal Ions in Biological System, 36. Sigel, A., and Sigel, H. (Eds.). Marcel Dekker, Inc., New York. Pp. 103-123.
Cadenas, E. (1995). Antioxidant and pro-oxidant functions of DT-diaphorase. Biochem. Pharmacol., 49, 127-140.
Calingasan, N.Y., Uchida, K., and Gibson, G.E. (1999). Protein-bound acrolein: A novel marker of oxidative stress in
Alzheimer’s disease. J. Neurochem., 72, 751-756.
Candeias, L.P., Patel, K.B., Stratford, M.R.L., and Wardman, P. (1993). Free hydroxyl radicals are formed from the
reaction between the neutrophil-derived species, superoxide anion and hypochlorous acid. FEBS Letts. 333, 151-153.
Canini, F., Bourdon, L., Cespuglio, R., and Buguet, A. (1997). Voltammetric assessment of brain nitric oxide during
heatstroke. Neurosci. Lett., 231, 67-70.
Carr, A.C, Vissers, M.C., Domigan, N.M., and Winterbourn, C.C. (1997). Modification of red cell membrane lipids by
hypochlorous acid and hemolysis by preformed chlorohydrins. Redox Rep., 3, 263-271.
Chabot, F., Mitchell, J., Quinlan, G., and Evans, T. (1997). Characterization of the vasodilator properties of peroxynitrite
on rat pulmonary artery: role of poly (adenosine 5’-diphosphoribose) synthase. Br. J. Pharmacol., 121, 485-490.
Chance, B., Sies, H., and Boveris, A. (1979). Hydroperoxide metabolism in mammalian organs. Physiol. Rev., 59, 527605.
Chen, G., Griffin, M., Poyer, J.L., and McCay, P.B. (1990). HPLC procedure for the pharmacokinetic study of spin-trapping
agent, alpha-phenyl-N-tert-butyl nitrone. Free Radic. Biol. Med., 9, 93-98.
Chen, G., Janzen, E.G., and Bray, T.M. (1994). Identification of 3-MI-derived N-centered radicals obtained from incubation
of 3-MU with microsomal-NADPH system by EPR-HPLC spin trapping. Free Radic. Biol. Med., 17, 19-25.
Cheng, H.-Y., Liu, T., Feuerstein, G., and Barone, F.C. (1993). Distribution of spin-trapping compounds in rat blood and
brain: In vivo microdialysis determination. Free Radic. Biol. Med., 14, 243-250.
Chilcote, R.R., Williams, B., Wolff, L.J., and Baehner, R.L. (1977). Sudden death in an infant from methemoglobinemia
after administration of "sweet spirits of nitre". Pediatrics, 59, 280-282.
Christen, S., Woodall, A.A., Shigenaga, M.K., Southwell-Keely, P.T., Duncan, M.W., and Ames, B.N. (1997). γ-Tocopherol
traps mutagenic electrophiles such as NOx and complements α-tocopherol: Physiological implications. Proc. Natl. Acad.
Sci., 94, 3217-3222.
Clark, R.A., Stone, P.J., El Hag, A., Calore, J.D., and Franzblau, C. (1981). Myeloperoxidase-catalyzed inactivation of α1protease inhibitor by human neutrophils. J. Biol. Chem., 256, 3348-3353.
Cocks, T.M., Angus, J.A., Campbell, J.H., and Campbell, G.R. (1985). Release and properties of endothelium-derived
relaxation factor (EDRF) from endothelial cells in culture. J. Cell Physiol., 123, 310-320.
Cooney, R.V., Franke, A.A., Harwood, P.J., Hatch-Pigott, V., Custer, L.J., and Mordan, L.J. (1993). γ-Tocopherol
detoxification of nitrogen dioxide: Superiority to α-tocopherol. Proc. Natl. Acad. Sci, USA, 90, 1771-1775.
Cooney, R.V., Hardwood, P.J., Franke, A.A., Narala, K., Sundstrom, A-K., Berggren, P.-O., and Mordan, L.J. (1995).
Product of γ-tocopherol reaction with NO2 and their formation in rat insulinoma (RINm5F) cells. Free Radic. Biol. Med.,
19, 259-269.
Cooper, C.E. (1999). Nitric oxide and iron proteins. Biochim. Biophys. Acta, 1411, 290-309.
Cooper, C.E., Torres, J., Sharpe, M.A., Wilson, M.T., and Svistunenko, D.A. (1998). Peroxynitrite reacts with
methemoglobin to generate globin-bound free radical species. Implications for vascular disease. Adv. Exp. Med. Biol.,
454, 195-202.
Corbett, J.T. (1989). The scopoletin assay for hydrogen peroxide: A review and a better method. J. Biochem. Biophys.
Meth., 256, 297-308.
Cross, C.E., Motchnik, P.A., Bruener, B.A., Jones, D.A., Kaur, H., Ames, B.N., and Halliwell, B. (1992). Oxidative damage
to plasma constituents by ozone. FEBS Lett. 298, 269-272.
Crow, J.P., and Beckman, J.S. (1995). Reactions between nitric oxide, superoxide, and peroxynitrite: footprints of
peroxynitrite in vivo. Adv. Pharmacol., 34, 17-43.
Crow, J.P., Beckman, J.S., and McCord, J.M. (1995). Sensitivity of the essential zinc-thiolate moiety of yeast alcohol
dehydrogenase to hypochlorite and peroxynitrite. Biochem., 34, 3544-3552.
Daiber, A., Mehl, M., and Ullrich, V. (1998). New aspects in the reaction mechanisms of phenol with peroxynitrite: The role
of phenoxy radicals. Nitric Oxide, 2, 259-269.
D’Aquino, M., Bullion, C., Chopra, M., Devi, D., Devi, S., Dunster, C., James, G., Komuro, E., Kundo, S., Nike, E., Raza,
F., Robertson, F., Sharma, J., and Willson, R. (1994). Sulfhydryl free radical formation enzymatically by sonolysis, by
radiolysis, and thermally: Vitamin A, curcumin, muconic acid, and related conjugated olefins as references. Meth.
Enzymol., 233, 34-46.
WWW.ESAINC.COM
136
Davies, M.J., Forni, L.G., and Shuter, S.L. (1987). Electron spin resonance and pulse radiolysis studies on the spin
trapping of sulfur-centered radicals. Chem. Biol. Interact., 61, 177-188.
Denicola, A., Freeman, B.A., Trujillo, M., and Radi, R. (1996). Peroxynitrite reaction with carbon dioxide/bicarbonate:
Kinetics and influence on peroxynitrite-mediated oxidations. Arch. Biochem. Biophys., 333, 49-58.
Desvignes, C., Robert, F., Vachette, C., Chouvet, G., Cespuglio, R., Renaud, B., and Lambas-Senas, L. (1997).
Monitoring nitric oxide (NO) in rat locus coeruleus: Differential effects of NO synthase inhibitors. Neuroreport, 8, 13211325.
Devasagayam, T.P., Steenken, S., Obendorf, M.S., Schulz, W.A., and Sies, H. (1991). Formation of 8-hydroxy (deoxy)
guanosine and generation of strand breaks at guanine residues in DNA by singlet oxygen. Biochemistry, 30, 62836289.
D’Ischia, M., and Costantini, C. (1995). Nitric oxide-induced nitration of catecholamine neurotransmitters: A key to
neuronal degeneration. Bioorg. Med. Chem., 3, 923-927.
D’Ischia, M., and Novellino, L. (1996). Nitric oxide-induced oxidation of α-tocopherol. Bioorg. Med. Chem., 4, 1747-1753.
Domigan, N.M., Charlton, T.S., Duncan, M.W., Winterbourn, C.C., and Kettle, A.J. (1995). Chlorination of tyrosyl residues
in peptides by myeloperoxidase and human neutrophils. J. Biol. Chem., 270, 16542-16548.
Douki, T., and Cadet, J. (1996). Peroxynitrite mediated oxidation of purine bases of nucleosides and isolated DNA. Free
Radic. Res., 24, 369-380.
Dugan, L.A., Lin, T-S., He, Y.-Y., Hsu, C.Y., and Choi, D.W. (1995). Detection of free radicals by microdialysis/spin
trapping EPR following focal cerebral ischemia-reperfusion and a cautionary note on the stability of 5,5-dimethyl-1pyrroline N-oxide (DMPO). Free Rad. Res. 23, 27-32.
Dunford, H.B. (2000). Peroxidase-catalyzed halide ion oxidation. Redox. Rep., 5, 169-171
Dupuy, C., Virion, A., Ohayon, R., Kamiewski, J.M., Deme, D., and Pommier, J. (1991). Mechanism of hydrogen peroxide
formation catalyzed by NADPH oxidase in thyroid plasma membrane. J. Biol. Chem., 266, 3739-3743.
During, M.J., Acworth, I.N., and Wurtman, R.J. (1988). Phenylalanine administration influences dopamine release in the
rat’s corpus striatum. Neurosci. Lett., 93, 91-95.
Dzwigaj, S., and Pezerat, H. (1995). Singlet oxygen-trapping reaction as a method of (1)O2 detection: Role of some
reducing agents. Free Radic. Res., 23, 103-115.
Eaton, A.D., Clesceri, L.S., and Greenberg, A.E. (Eds.). (1995). Standard Methods for The Examination of Water and
Wastewater. American Public Health Association, Washington, DC.
Egorov, S., Yu, K., Kurella, E.G., Boldyrev, A.A., and Krasnovsky, A.A. (1997). Quenching of singlet molecular oxygen by
carnosine and related antioxidants. Monitoring 1270-nm phosphorescence in aqueous media. Biochem. Mol. Biol. Int.,
41, 687-694.
Eiserich, J.P., Butler, J., Van Der Vleit, A., Cross, C.E., and Halliwell, B. (1995). Nitric oxide rapidly scavenges tyrosine
and tryptophan radicals. Biochem. J., 310: 745-749.
Eiserich, J.P., Cross, C.E., Jones, A.D., Halliwell, B., and van der Vleit, A. (1997). Formation of nitrating and chlorinating
species by reaction of nitrite with hypochlorous acid. J. Biol. Chem., 271, 19199-19208.
Ellis, M., Hiss, Y., and Shenkman, L. (1992). Fatal methemoglobinemia caused by inadvertent contamination of a laxative
solution with sodium nitrite. Isr. J. Med. Sci. 28, 289-291.
Eritja, R., Horowitz, D.M., Walker, P.A., Ziehler-Martin, J.P., Boosalis, M.S., Goodman, M.F., Itakura, K., and Kaplan, B.E.
(1986). Synthesis and properties of oligonucleotides containing 2'-deoxynebularine and 2'-deoxyxanthosine. Nucleic
Acids Res., 14, 8135-8153.
Ewing, J.F., and Janero, D.R. (1998). Specific S-nitrosothiol (thionitrite) quantification as solution nitrite after vanadium
(III) reduction and ozone-chemiluminescent detection. Free Radic. Biol. Med., 25, 621-628.
Fang, K., Ragsdale, N.V., Carey, R.M., MacDonald, T., and Gaston, B. (1998). Reductive assay for S-nitrosothiols:
Implications for measurements in biological systems. Biochem. Biophys. Res. Commun., 252, 535-540.
Feelisch, M., and Stamler, J.S. (1996). Donors of Nitrogen Oxides. In: Methods in Nitric Oxide Research. (Feelisch, M.,
and Stamler, J.S. (Eds.). John Wiley and Sons Ltd. London. Pp 71-115.
Feelisch, M., Kubitzek, D., and Weringloer, J. (1996). The oxyhemoglobin assay. In: Methods in Nitric Oxide Research.
(Feelisch, M., and Stamler, J.S. (Eds.). John Wiley and Sons Ltd. London. Pp. 455-478.
Feigl, E.O. (1988). EDRF-a protective factor. Nature, 331, 490-491.
Feix, J.B., and Kalyanaraman, B. (1991). Production of singlet oxygen-derived hydroxyl radical adducts during
merocyanine-540-mediated photosensitization: Analysis of ESR-spin trapping and HPLC with electrochemical
detection. Arch. Biochem. Biophys., 291, 43-51.
Feldman, P.L., Griffith, O.W., and Stuehr, D.J. (1993). The surprising life of nitric oxide. C&EN, December, 26-38.
Fenton, H.J.H. (1876). On a new reaction of tartaric acid. Chem. News, 9, 190.
Fenton, H.J.H. (1894), Oxidation of tartaric acid in the presence of iron. J. Chem. Soc., 65, 899-910.
Fiala, E., Sodum, R., Bhattacharya, M., and Li, H. (1996). (-)-Epigallocatchin gallate, a polyphenolic tea antioxidant,
inhibits peroxynitrite-mediated formation of 8-oxodeoxyguanosine and 3-nitrotyrosine. Experientia, 52, 922-926.
Finkelstein, E., Rosen, G.M., and Rauckman, E.J. (1980). Spin trapping of superoxide and hydroxyl radical: practical
aspects. Arch. Biochem. Biophys., 200, 1-16.
Fischer, C., and Klotz, U. (1994). Radical-derived oxidation products of 5-aminosalicylic acid and N-acetyl-5aminosalicylic acid. J. Chromatogr. B. 661, 57-68.
Floyd, R.A. (1983). Hydroxyl free-radical spin-adducts in rat brain synaptosomes observations on the reduction of the
nitroxide. Biochem. Biophys. Acta., 756, 204-216.
Floyd, R.A. (1997). Workshop. Oxygen Society Meeting, San Francisco.
Floyd, R.A., Henderson, R., Watson, J., and Wong, P.K. (1986). Use of salicylate with high-pressure liquid
chromatography and electrochemical detection (LCED) as a sensitive measure of hydroxyl free radicals in adriamycin
treated rats. J. Free. Rad. Biol. Med., 2, 13-18.
Floyd, R.A., Lewis, C.A., and Wong, P.K. (1984a). High-pressure liquid chromatography-electrochemical detection of
oxygen free radicals. Meth. Enzymol., 105, 231-327.
WWW.ESAINC.COM
137
Floyd, R.A., Watson, J.J., and Wong, P.K. (1984b). Sensitive assay of hydroxyl free radical formation utilizing high
pressure liquid chromatography with electrochemical detection of phenol and salicylate hydroxylation products. J.
Biochem. Biophys. Meth., 10, 221-235.
Floyd, R.A., Watson, J.J., Wong, P.K., Altmiller, D.H., and Rickard, R.C. (1986). Hydroxyl free radical adduct of
deoxyguanosine: Sensitive detection and mechanisms of formation. Free Rad. Res. Comm., 1, 163-172.
Foote, C.S., Goyne, T.E., and Lehrer, R.I. (1981). Assessment of chlorination by human neutrophils. Nature, 301, 715716.
Ford, P.C., Wink, D.A., and Stanbury, D.M. (1993). Autoxidation kinetics of aqueous nitric oxide. FEBS Lett., 326, 103.
Forni, L.G., and Willson, R.L. (1986a). Thiyl free radicals and the oxidation of ferrocytochrome c. Direct observation of
coupled hydrogen atom and electron transfer. Biochem. J., 240, 905-907.
Forni, L.G., and Willson, R.L. (1986b). Thiyl and peroxyl free radicals and NADH. Direct observation of one-electron
oxidation. Biochem. J., 240, 897-903.
Forstermann, U., and Ishi, K. (1996). Measurement of cyclic GMP as an indicator of nitric oxide production. Feelisch, M.,
and Stamler, J.S. (Eds.). John Wiley and Sons Ltd., London. Pp. 555-566.
Frears, E., Zhang, Z., Blake, D., O’Connell, J., and Winyard, P. (1996). Inactivation of tissue inhibitor of metalloproteinase1 by peroxynitrite. FEBS Lett., 381, 21-24.
Fridovich, I. (1989). Superoxide dismutase: An adaptation to a paramagnetic gas. J. Biol. Chem., 264, 7761-7767.
Friedemann, M.N., Robinson, S.W., and Gerhardt, G.A. (1996). o-Phenylenediamine-modified carbon fiber electrodes for
the detection of nitric oxide. Anal. Chem., 68, 2621-2628.
Fukuzawa, K., Inokami, Y., Tokumura, A., Terao, J., and Suzuki, A. (1998). Singlet oxygen scavenging by alphatocopherol and beta-carotene: Kinetic studies in phospholipid membranes and ethanol solution. Biofactors, 7, 31-40.
Fukui, S., Hanasaki, Y., and Ogawa, S. (1993). High-performance liquid chromatographic determination of hydroxyl
radicals. J. Chromatogr., 630, 187-193.
Fukuto, J.M., and Wink, D.A. (1999). Nitric oxide (NO): Formation and biological roles in mammalian systems. In: Metal
Ions in Biological System, 36. Sigel, A., and Sigel, H. (Eds.). Marcel Dekker, Inc., New York. Pp. 547-595.
Furchgott, R.F. (1984). The role of endothelium in responses of vascular smooth muscle to drugs. Ann. Rev. Pharmacol.
Toxicol., 24, 175-197.
Galli, F., Rovidati, S., Ghibelli, L., and Canestrari, F. (1998). S-nitrosylation of glyceraldehyde-3-phosphate
dehydrogenase decreases the enzyme affinity to the erythrocyte membrane. Nitric Oxide, 2, 17-27.
Gee, P., and Davison, A.J. (1989). Intermediates in the aerobic autoxidation of 6-hydsroxydopamine: Relative importance
under different reaction conditions. Free Radic. Biol. Med., 6, 271-284.
Gelvan, D., Moreno, V., Gassmann, W., Hegenauer, J., and Saltman, P. (1992). Metal-ion-directed site specificity of
hydroxy radical detection. Biochem. Biophys. Acta, 1116, 183-191.
George, P. (1965). In: Oxidase and Redox Related Systems. King, T.E., Mason, H.S., and Morrison, M. (Eds.). Wiley,
New York. Pp. 3-36.
Gerlach, M., Ben-Shachar, D., Riederer, P., and Youdim, M.B.H. (1994). Altered brain metabolism of iron as a cause of
neurodegenerative diseases. J. Neurochem., 63, 793-807.
Ghilarducci, D.P., and Tjeerdema, R.S. (1995). Fate and effects of acrolein. Rev. Environ. Contam. Toxicol., 144, 95-146.
Gilbert, D.L. (1999). From the breath of life to reactive oxygen species. In: Reactive Oxygen Species in Biological
Systems. Gilbert, D.L., and Colton, C.A. (Eds.). Kluwer Academic/Plenum Publishers, New York. Pp. 3-31.
Giulivi, C. (1998). Functional implications of nitric oxide produced by mitochondria in mitochondrial metabolism. Biochem.
J., 332, 673-679.
Giulivi, C., and Davies, K.J.A. (1994). Dityrosine: A marker for oxidatively modified proteins and selective proteolysis.
Meth. Enzymol., 233, 363-371.
Giulivi, C., Poderoso, J.J., and Boveris, A. (1998). Production of nitric oxide by mitochondria. J. Biol. Chem., 273, 1103811043.
Gleu, K., and Hubold, R. (1935). Die einwirkung von wasserstoffsuperoxyd auf saltpetrige saure. Persalpetridge saure. Z.
Anorg. Allg. Chem. 179, 233-266.
Goldman, A.P., and Macrae, D.J. (1994). Nitrogen dioxide measurement in breathing systems. Lancet, 343, 1850.
Goldstein, S., Czapski, G., Lind, J., and Merenyi, G. (1988). Mechanism of decomposition of peroxynitric ion. Evidence for
the formation of superoxide and nitrogen dioxide radical. Inorg. Chem., 37, 3943-3947.
Goldstein, S., Meyerstein, D., and Crapski, G. (1993). The Fenton reagents. Free Radic. Biol. Med., 15, 435-554.
+
Gorbunov, N.V., Osipov, A.N., Sweetland, M.A., Day, B.W., Elsayed, N.M., and Kagan, V.E. (1996). NO redox paradox:
Direct oxidation of α-tocopherol-mediated oxidation of ascorbate. Biochem. Biophys. Res. Commun., 219, 835-841.
Gow, A., and Stamler, J. (1998). Reactions between nitric oxide and hemoglobin under physiological conditions. Nature,
391, 169-173.
Gow, A., Duran, D., Malcolm, S., and Ischiropoulos, H. (1996). Effects of peroxynitrite-induced modifications on tyrosine
phosphorylation and degradation. FEBS Lett., 385, 63-66.
Gow, A., Duran, D., Thom, S.R., and Ischiropoulos, H. (1996). Carbon dioxide enhancement of peroxynitrite-mediated
protein tyrosine nitration. Arch. Biochem. Biophys., 333, 42-48.
Graham, D. (1978). Oxidative pathways for catecholamines in the genesis of neuromelanin and cytotoxic quinones. Mol.
Pharmacol., 14, 633-643.
Graham, D., Tiffany, S.M., Bell, W.R. and Gutknecht, W.F. (1978). Autoxidation versus covalent binding of quinones as
mechanisms of toxicity of dopamine, 6-hydroxydopamine, and related compounds toward C1300 neuroblastoma cells
in vitro. Mol. Pharmacol., 14, 644-653.
Griffith, O.W., and Stuehr, D.J. (1995). Nitric oxide synthases: Properties and catalytic mechanisms. Ann. Rev. Physiol.,
57, 707-736.
Griffith, T.M., Edwards, D.H., Lewis, M.J., Newby, A.C., and Henderson, A.H. (1984). The nature of endothelium-derived
vascular relaxant factor. Nature, 308, 645-647.
WWW.ESAINC.COM
138
Griscavage, J.M., Hobbs, A.J., and Ignarro, L.J. (1995). Negative modulation of nitric oxide synthase by nitric oxide and
nitroso compounds. Adv. Pharmacol., 34, 215-234.
Grisham, M.B., Jefferson, M.M., Melton, D.F., and Thomas, E.L. (1984). Chlorination of endogenous amines by isolated
neutrophils. Ammonia-dependent bacteriocidal, cytotoxic, and cytolytic activities of the chloramines. J. Biol. Chem.,
259, 10404-10413.
Gross, S.S., and Wolin, M.S. (1995). Nitric oxide: Pathophysiological mechanisms. Ann. Rev. Physiol., 57, 737-769.
Groves, J.T. (1999). Peroxynitrite: Reactive, invasive and enigmatic. Curr. Opin. Chem. Biol., 3, 226-235.
Haenen, G., Paquay, J., Korthouwer, R., and Bast, A. (1997). Peroxynitrite scavenging by flavonoids. Biochem. Biophys.
Res. Commun., 236, 591-593.
Halliwell, B. (1997). What nitrates tyrosine? Is nitrotyrosine specific as a biomarker of peroxynitrite formation in vivo?
FEBS Lett., 411, 157-160.
Halliwell, B., and Gutteridge, J.M.C. (1981). Formation of thiobarbituric-acid-reactive substance from deoxyribose in the
presence of iron salts: the role of superoxide and hydroxyl radicals. FEBS Lett., 128, 347-352.
Halliwell, B., and Gutteridge, J.M.C. (1990). Role of free radicals and catalytic metal ions in human disease: An overview.
Meth. Enzymol., 186, 1-85.
Halliwell, B., and Gutteridge, J.M.C. (1992). Biologically relevant metal ion-dependent hydroxyl radical generation. FEBS
Lets., 307, 108-112.
Halliwell, B., and Gutteridge, J.M.C. (1999). Free Radicals in Biology and Medicine. Clarendon Press, Oxford.
Halliwell, B., Gutteridge, J.M.C., and Aruoma, O.I. (1987). The deoxyribose method: a simple "test-tube" assay for
determination of rate constants for reactions of hydroxyl radicals. Anal. Biochem., 165, 215-219.
Halliwell, B., and Kaur, H. (1997). Hydroxylation of salicylate and phenylalanine as assays for hydroxyl radicals: A
cautionary note visited for the third time. Free Rad. Res., 27, 239-244.
Halliwell, B., Hu, M.-L., Louie, S., Duvall, T., Tarkington, B., Motchnik, P., and Cross, C. (1992). Interaction of nitrogen
dioxide with human plasma. FEBS Lett., 313, 62-66.
Hampl, V., Walters, C.L., and Archer, S.L. (1996). Determination of nitric oxide by the chemiluminescence reaction with
ozone. In: Methods in Nitric Oxide Research. Feelisch, M., and Stamler, J.S. (Eds.). John Wiley and Sons Ltd., London.
Pp. 309-318.
Hansford, R.G., Hogue, B.A., and Mildaziene, V. (1997). Dependence of H2O2 formation by rat heart mitochondria on
substrate availability and donor age. J. Bioenerg. Biomem., 29, 89-95.
Harrison, J.E., and Schultz, J. (1976). Studies on the chlorinating activity of myeloperoxidase. J. Biol. Chem., 251, 13711374.
Hartmann, H.-J., Sievers, C., and Weser, U. (1999). Thiyl radicals in biochemically important thiols in the presence of
metal ions. In: Metal Ions in Biological System, 36. Sigel, A., and Sigel, H. (Eds.). Marcel Dekker, Inc., New York. Pp.
389-413.
Hartsfield, C.L., Alam, J., Cook, J.L., and Choi, A.M. (1997). Regulation of heme oxygenase-1 gene expression in
vascular smooth muscle cells by nitric oxide. Am. J. Physiol., 273, L980-988.
Hassan, T., and Khan, A.U. (1986). Phototoxicity of the tetracyclines: Photosensitized emission of singlet delta oxygen.
Proc. Natl. Acad. Sci. USA, 83, 4604-4606.
Hazell, L.J., Arnold, L., Flowers, D., Waeg, G., Malle, E., and Stocker, R. (1996). Presence of hypochlorite-modified
proteins in human atherosclerotic lesions. J. Clin. Invest., 97, 1535-1544.
Hazen, S.L., and Heinicke, J.W. (1997). 3-Chlorotyrosine, a specific marker of myeloperoxidase-catalyzed oxidation, is
markedly elevated in low density lipoprotein isolated from human atherosclerotic intima. J. Clin. Invest., 99, 2075-2081.
Hazen, S.L., and Wu, W. (1998). 3-Bromo and 3,5-dibromo-tyrosine: Potential markers for eosinophil-dependent injury in
vivo. Free Radic. Biol. Med., 25, Supplement 1, 74.
Hazen, S.L., Gaut, J.P., Hsu, F.F., Crowley, J.R., d’Avignon, A., and Heinecke, J.W. (1997). pHydroxyphenylacetaldehyde, the major product of L-tyrosine oxidation by the myeloperoxidase-H2O2-chloride system of
phagocytes, covalently modifies ε-amino groups of protein lysine residues. J. Biol. Chem., 272, 16990-16998.
Hazen, S.L., Hsu, F.F., and Heinecke, J.W. (1996b). p-Hydroxyphenylacetaldehyde is the major product of L-tyrosine
oxidation by activated human phagocytes. J. Biol. Chem., 271, 1861-1867.
Hazen, S.L., Hsu, F.F., Mueller, D.M., Crowley, J.R., and Heinicke, J.W. (1996a). Human neutrophils employ chlorine gas
as an oxidant during phagocytosis. J. Clin. Invest., 98, 1283-1289.
Hecker, M., and Billiar, T.R. (1996). Detection of L-arginine, related metabolites and guanidino compounds. In: Methods in
Nitric Oxide Research. Feelisch, M., and Stamler, J.S. (Eds.). John Wiley and Sons Ltd., London. Pp. 260-270.
Heinecke, J.W., Li, W., Daehnke, H.L., and Goldstein, J.A. (1993). Dityrosine, a specific marker of oxidation, is
synthesized by the myeloperoxidase-hydrogen peroxide system of human neutrophils and macrophages. J. Biol.
Chem., 268, 4069-4077.
Heinecke, J.W., Li, W., Mueller, D.M., Bohrer, A., and Turk, J. (1994). Cholesterol chlorohydrin synthesis by the
myeloperoxidase-hydrogen peroxide-chloride system: Potential markers for lipoproteins oxidatively damaged by
phagocytes. Biochem., 33, 10127-10136
Henry, Y.A., and Singel, D. (1996). Metal-nitrosyl interactions in nitric oxide biology probed by electron paramagnetic
resonance spectroscopy. In: Methods in Nitric Oxide Research. Feelisch, M., and Stamler, J.S. (Eds.). John Wiley and
Sons Ltd., London. Pp. 357-372.
Henry, Y., Lepoivre, M., Drapier, J-C., Ducrocq, C., Boucher, J-C., and Guissani, A. (1993). EPR characterization of
molecular targets for NO in mammalian cells and organelles. FASEB J., 7, 1124-1134.
Hensley, K., Maidt, M.L., Pye, Q.N., Stewart, C.A., Wack, M., Tabatabaie, T., and Floyd, R.A. (1997). Quantitation of
protein-bound 3-nitrotyrosine and 3,4-dihydroxyphenylalanine by high-performance liquid chromatography with
electrochemical array detection. Anal. Biochem., 251, 187-195.
Hibbs, J.B., Taintor, R.R., Vavrin, Z., and Rachlin, E.M. (1988). Nitric oxide: A cytotoxic activated macrophage effector
molecule. Biochem. Biophys. Res. Commun., 157, 87-94.
WWW.ESAINC.COM
139
Hiraoka, W., Kuwabara, M., and Sato, F. (1989). OH-induced free radicals in purine nucleoside monophosphates: ESR
and spin-trapping. Int. J. Radiat. Biol., 55, 51-58.
Hiraoka, W., Kuwabara, M., Sato, F., Matsuda, A., and Ueda, T. (1990). Free-radical reactions induced by OH-radical
attack on cytosine-related compounds: A study by a method combining ESR, spin trapping and HPLC. Nucleic Acids
Res., 18, 1217-1223.
Hobbs, A.J., Fukuto, J.M., and Ignarro, L.J. (1994). Formation of free nitric oxide from L-arginine by nitric oxide synthase:
Direct enhancement of generation by superoxide dismutase. Proc. Natl. Acad. Sci. USA, 91, 10992-10996.
Hogg, N., and Kalyanaraman, B. (1999). Nitric oxide and lipid peroxidation. Biochim. Biophys. Acta, 1411, 378-384.
Hoigne, J., and Bader, H. (1975). Ozonation of water: Role of hydroxyl radicals as oxidizing intermediates. Science, 190,
792.
Huggins, T.G., Wells-Knecht, M.C, Detorie, N.A., Baynes, J.W., and Thorpe, S.R. (1993). Formation of o-tyrosine and
dityrosine in proteins during radiolytic and metal-catalyzed oxidation. J. Biol. Chem., 268, 12341-12347.
Hughes, M.N. (1999). Relationships between nitric oxide, nitroxyl ion, nitrosonium cation and peroxynitrite. Biochim.
Biophys. Acta, 1411, 263-272.
Huie, R.E. (1994). The reaction kinetics of NO2•. Toxicol., 89, 193-196.
Huie, R.E., and Neta, P. (1999). Chemistry of reactive oxygen species. In: Reactive Oxygen Species in Biological
Systems. Gilbert, D.L., and Colton, C.A. (Eds.). Kluwer Academic/Plenum Publishers, New York. Pp. 33-73.
Huie, R.E., and Padmaja, S. (1993). The reaction rate of nitric oxide with superoxide. Free Radic. Biol. Med., 18, 195-199.
Hutchinson, F. (1957). The distance that a radical formed by ionizing radiation can diffuse in a yeast cell. Radiat. Res., 7,
473-483.
Hwang, C., Sinskey, A.J., and Lodish, H.F. (1992). Oxidized redox state of glutathione in the endoplasmic reticulum.
Science, 257, 1496-1502.
Igarashi, T., Sakurai, K., Oi, T., Obara, H., Oyha, H., and Kamada, H. (1999). New sensitive agents for detecting singlet
oxygen by electron spin resonance spectroscopy. Free Radic. Biol. Med., 26, 1339-1345.
Ignarro, L.J., Burke, T.M., Wood, K.S., Wolin, M.S., and Kadowitz, P.J. (1984). Association between cGMP accumulation
and acetylcholine-elicited relaxation of bovine intrapulmonary artery. J. Pharmacol. Exp. Ther., 228, 682-690.
Imlay, J.A., and Linn, S. (1988). DNA damage and oxygen radical toxicity. Science, 240, 1302-1309.
Inami, O., Kuwabara, M., Endoh, D., and Sato, F. (1986). OH-induced free radicals in uridine studied by a method
combining ESR, spin-trapping, and liquid chromatography. Radiat. Res., 108, 1-11.
Inami, O., Kuwabara, M., and Sato, F. (1987). OH-induced free radicals in 3’-UMP and poly(U): Spin-trapping and radical
chromatography. Radiat. Res., 112, 36-44.
Ischiropoulos, H., and Al-Mehdi, A.B. (1995). Peroxynitrite-mediated oxidative protein modifications. FEBS Letts., 364,
279-282.
Ischiropoulos, H., Zhu, L., Chen, J., Tsai, H.M., Martin, J.C., Smith, C.D., and Beckman, J.S. (1992b). Peroxynitritemediated tyrosine nitration catalyzed by superoxide dismutase. Arch. Biochem. Biophys., 298, 431-437.
Ishimitsu, S., Fujimoto, S., and Ohara, A. (1986). Hydroxylation of phenylalanine by the hypoxanthine-xanthine oxidase
system. Chem. Pharmaceut. Bull., 32, 4645-4649.
Ishimitsu, S., Fujimoto, S., and Ohara, A. (1986). Determination of m-tyrosine and o-tyrosine in human serum by highperformance liquid chromatography with fluorometric detection. J. Chromatogr., 378, 222-225.
Iwahashi, H. (1996). Quantitative detection of reduced, radical and oxidized forms of α-(4-pyridyl-1-oxide)-N-tertbutylnitrone radical adduct using high-performance liquid chromatography with electrochemical detection. J.
Chromatogr. A, 753, 235-242.
Jacob, J.S., Cistola, D.P., Hsu, F.F., Muzaffar, S., Mueller, D.M., Hazen, S.L., and Heinecke, J.W. (1996). Human
phagocytes employ the myeloperoxidase-hydrogen peroxide system to synthesize dityrosine, trityrosine, pulcherosine,
and isodityrosine by a tyrosyl radical-dependent pathway. J. Biol. Chem., 271, 19950-19956.
Jahnke, L.S. (1999). Measurement of hydroxyl radical-generated methane sulfinic acid by high-performance liquid
chromatography and electrochemical detection. Anal. Biochem., 269, 273-277.
Jakab, G.J. (1977). Adverse effect of a cigarette smoke component, acrolein, on pulmonary antibacterial defenses and on
viral-bacterial interactions in the lung. Am. Rev. Respir. Dis., 115, 33-38.
Jen, J., Leu, M., and Yang, T.C. (1998). Determination of hydroxyl radicals in an advanced oxidation process with salicylic
acid trapping and liquid chromatography. J. Chromatogr. A, 796, 283-288.
Jia, L., Bonaventura, C., Bonaventura, J., and Stamler, J.S. (1996). S-nitrosohemoglobin: A dynamic activity of blood
involved in vascular control. Nature, 380, 221-226.
Jiang, Q., and Hurst, J.K. (1997). Relative chlorinating, nitrating, and oxidizing capabilities of neutrophils determined with
phagocytosable probes. J. Biol. Chem., 272, 32767-32772.
Jourd’heuil, D., Mai, C.T., Laroux, F.S., Wink, D.A., and Grisham, M.B. (1998). The reaction of S-nitrosoglutathione with
superoxide. Biochem. Biophys. Res. Commun., 246, 525-530.
Kachur, A.V., Tuttle, S.W., and Biaglow, J.E. (1998). Autoxidation of ferrous ion complexes: A method for the generation
of hydroxyl radicals. Radiat. Res., 150, 475-482.
Kaiser, S., Mascio, P.D., Murphy, M.E., and Sies, H. (1990). Quenching of singlet molecular oxygen by tocopherols. In:
Antioxidants in Therapy and Preventive Medicine. Emerit et al. (Eds.). Plenum Press, New York. Pp. 117-124.
Kalyanaraman, B. (1995). Thiyl radicals in biological systems: Significance or trivial? Biochem. Soc. Symp., 61, 55-63.
Kalyanaraman, B., Karoui, H., Singh, R.J., and Felix, C.C. (1996). Detection of thiyl radical adducts formed during
hydroxyl radical- and peroxynitrite-mediated oxidation of thiols – a high resolution ESR spin-trapping study at Q-band
(35 GHz). Anal. Biochem., 241, 75-81.
Kalyamaraman, B., Ramanujam, S., Singh, R.J., Joseph, J., and Feix, J.B. (1993). Formation of 2,5-dihydroxybenzoic
1
acid during the reaction between O2 and salicylic acid: Analysis of ESR oximetry and HPLC with electrochemical
detection. J. Am. Chem. Soc., 115, 4007-4012.
Kanner, J., Harel, S., and Granit, R. (1991). Nitric oxide as an antioxidant. Arch. Biochem. Biophys., 289, 130-136.
WWW.ESAINC.COM
140
Karam, L.R., and Simic, M.G. (1988). Detecting irradiated foods: Use of hydroxyl radical biomarkers. Anal. Chem., 16,
1117A-1119A.
Karoui, H., Hogg, N., Frejaville, C., Tordo, P., and Kalyanaraman, B. (1996). Characterization of sulfur-centered radical
intermediates formed during the oxidation of thiols and sulfite by peroxynitrite. ESR-spin trapping and oxygen uptake
studies. J. Biol. Chem., 271, 6000-6009.
Kato, Y., Kawakishi, S., Aoki, T., Itakura, K., and Osawa, T. (1997). Oxidative modification of tryptophan residues exposed
to peroxynitrite. Biochem. Biophys. Res. Commun., 234, 82-84.
Kaur, H., and Halliwell, B. (1994a). Aromatic hydroxylation of phenylalanine as an assay for hydroxyl radicals:
Measurement of hydroxyl radical formation from ozone and in blood from premature babies using improved HPLC
methodology. Anal. Biochem. 220, 11-15.
Kaur, H., and Halliwell, B. (1994b). Detection of hydroxyl radicals by aromatic hydroxylation. Meth. Enzymol. 233, 67-82.
Kaur, H., and Halliwell, B. (1994c). Evidence for nitric oxide-mediated oxidative damage in chronic inflammation.
Nitrotyrosine in serum and synovial fluid from rheumatoid patients. FEBS Letts. 350, 8-12.
Kaur, H., Edmonds, S.E., Blake, D.R., and Halliwell, B. (1996). Hydroxyl radical generation by rheumatoid blood and
knee-joint synovial fluid. Ann. Rheumatic Dis., 55, 915-920.
Kaur, H., Fagerheim, I., Grootveld, M., Puppo, A., and Halliwell, B. (1988). Aromatic hydroxylation of phenylalanine as an
assay for hydroxyl radicals: Application to activated human neutrophils and to heme protein leghemoglobin. Anal.
Biochem., 172, 360-367.
Kaur, H., Leung, K.H.W., and Perkins, M.J. (1981). J. Chem. Soc. Chem. Commun., 142-144.
Kaur, H., Whiteman, M., and Halliwell, B. (1997). Peroxynitrite-dependent aromatic hydroxylation and nitration of salicylate
and phenylalanine. Is hydroxyl radical involved. Free Radic. Res., 26, 71-82.
Keaney, J.F., Simon, D.I., Stamler, J.S., Jaraki, O., Scharfstein, J., Vita, J.A., and Loscalzo, J. (1993). NO forms an
adduct with serum albumin that has endothelium-derived relaxing factor-like properties. J. Clin. Invest., 91, 1582-1589.
Keith, W.G., and Powell, R.E. (1969). J. Chem. Soc., A. 90.
Kelm, M., and Schrader, J. (1988). Nitric oxide release from the isolated guinea pig heart. Eur. J. Pharmacol., 155, 317321.
Kettle, A.J., and Winterbourn, C.C. (1991). The influence of superoxide on the production of hypochlorous acid by human
neutrophils. Free Rad. Res. Comm., 12-13, 47-52.
Kettle, A.J., and Winterbourn, C.C. (1994). Superoxide-dependent hydroxylation by myeloperoxidase. J. Biol. Chem., 269,
17146-17151.
Kim, S.G., and Novak, R.F. (1990). Role of P450IIE1 in the metabolism of 3-hydroxypyridine, a constituent of tobacco
smoke: Redox cycling and DNA strand scission by the metabolite 2,5-dihydroxypyridine. Cancer Res., 50, 5333-5339.
Kiryu, C., Makiuchi, M., Miyazaki, J., Fujinaga, T., and Kakinuma, K. (1999). Physiological production of singlet molecular
oxygen in the myeloperoxidase-hydrogen peroxide-chloride system. FEBS Lett., 443, 154-158.
Kishnani, N.S., and Fung, H.-L. (1996). Nitric oxide generation from pharmacological nitric oxide donors. Meth. Enzymol.,
268, 259-265.
Klebanoff, S.J. (1988). In: Inflammation: Basic Principles and Clinical Correlates. Gallin, J.I., Goldstein, I.M., and
Snyderman, R. (Eds.). Raven Press, New York. Pp. 391-443.
Klebanoff, S.J. (1993). Reactive nitrogen intermediates and antimicrobial activity: Role of nitrite. Free Radic. Biol. Med.,
14, 351-360.
Klein, S.M., Cohen, G., and Cederbaum, A.I. (1981). Production of formaldehyde during metabolism of dimethyl sulfoxide
by hydroxyl radical generating systems. Biochem., 20, 6006-6012.
Kluge, I., Gutteck-Amsler, U., Zollinger, M., and Do, K.Q. (1997). S-Nitrosoglutathione in rat cerebellum: Identification and
quantification by liquid chromatography-mass spectrometry. J. Neurochem., 69, 2599-2607.
Kojima, H., Kikuchi, K., Hirobe, M., and Nagano, T. (1997). Real-time measurement of nitric oxide production in rat brain
by the combination of luminol-H2O2 chemiluminescence and microdialysis. Neurosci. Letts., 233, 157-159.
Kojima, H., Nakatsubo, N., Kikuchi, K., Kawahara, S., Kirino, Y., Nagoshi, H., Hirata, Y., and Nagano, T. (1998). Detection
and imaging of nitric oxide with novel fluorescent indicators: Diaminofluoresceins. Anal. Chem., 70, 2446-2453.
Kon, S.H. (1978). Biological autoxidation. I. Decontrolled iron: An ultimate carcinogen and toxicant: An Hypothesis. Med.
Hypotheses, 4, 445-471.
Kong, S.-K., Yim, M., Stadtman, E., and Chock, P. (1996). Peroxynitrite disables the tyrosine phosphorylation regulatory
mechanism: Lymphocyte-specific tyrosine kinase fails to phosphorylate nitrated cdc2(6-20)NH2 peptide. Proc. Natl.
Acad. Sci., USA, 93, 3377-3382.
Kono, Y. (1995). The production of nitrating species by the reaction between nitrite and hypochlorous acid. Biochem. Mol.
Biol. Int., 36, 275-283.
Kooy, N.W., Royall, J.A., Ye, Y.Z., Kelley, D.R., and Beckman, J.S. (1994). Evidence for in vivo peroxynitrite production in
human acute lung injury. Am. Rev. Respir. Dis., 151, 1250-1254.
Koppenol, W.H. (1993a). A thermodynamic appraisal of the radical sink hypothesis. Free Radic. Biol. Med., 14, 91-94.
Koppenol, W.H. (1993b). The centennial of the Fenton reaction. Free Radic. Biol. Med., 15, 645-651.
Koppenol, W.H. (1994). Thermodynamic considerations on the formation of reactive species from hypochlorite,
superoxide and nitrogen monoxide. Could nitrosyl chloride be produced by neutrophils and macrophages? FEBS Lett.,
347, 5-8.
Koppenol, W.H., Kissner, R., and Beckman, J.S. (1996), Synthesis of peroxynitrite: To go with the flow or on solid
grounds? Meth. Enzymol., 269, 296-302.
Koppenol, W.H., Moreno, J.J., Pryor, W.A., Ischiropoulus, H., and Beckman, J.S. (1992). Peroxynitrite a cloaked oxidant
from superoxide and nitric oxide. Chem. Res. Toxicol., 5, 834-842.
Korth, H.-G., and Weber, H. (1996). Detection of nitric oxide with nitric oxide-trapping agents. In: Methods in Nitric Oxide
Research. Feelisch, M., and Stamler, J.S. (Eds.). John Wiley and Sons Ltd., London. Pp. 383-391.
WWW.ESAINC.COM
141
Korth, H.G., Ingold, K.U., Sustmann, R., de Groot, H., and Sies, H. (1992). Tetramethyl-o-chinodiethan (NOCT-!) das
erste mitgleid einer familie massgescheiderter chletroper singanger fur Stickstoffmonoxide. Angew. Chem., 104, 915917.
Kosaka, H., and Shiga, T. (1996). Detection of nitric oxide by electron spin resonance using hemoglobin. In: Methods in
Nitric Oxide Research. Feelisch, M., and Stamler, J.S. (Eds.). John Wiley and Sons Ltd., London. Pp. 373-391.
Kostka, P. (1992). Free Radicals (Nitric Oxide). Anal. Chem., 67, 411R-416R.
Kostka, P., and Park, J.K.J. (1999). Fluorometric detection of S-nitrosothiols. Meth. Enzymol., 301, 227-235.
Kuhn, D.M., and Arthure, R.E. (1996). Inactivation of brain tryptophan hydroxylase by nitric oxide. J. Neurochem., 67,
1072-1077.
Kwon, B-M., and Foote, C.S. (1988). Chemistry of singlet oxygen. 50. Hydroperoxide intermediates in the
photooxygenation of ascorbic acid. J. Am. Chem. Soc. 110, 6582-6583.
Lai, C-S., and Komarov, A.M. (1994). Spin-trapping of nitric oxide produced in vivo in septic shock mice. FEBS Lett., 345,
120-124.
Laval, F., and Wink, D.A. (1994). Inhibition by nitric oxide of the repair protein, O6-methylguanine-DNA-methyltransferase.
Carcinogen., 15, 443-447.
Leeuwenburgh, C., Rasmussen, J.E., Hsu, F.F., Mueller, D.M., Pennathur, S., and Heinecke, J.W. (1997). Reactive
nitrogen intermediates promote low density lipoprotein oxidation in human atherosclerotic intima. J. Biol. Chem., 272,
3520-3526.
Lemercier, J.-N., Padmaja, S., Cueto, R., Squadrito, G.L., Uppa, R.M., and Pryor, W.A. (1997). Carbon dioxide
modulation of hydroxylation and nitration of phenol by peroxynitrite. Arch. Biochem. Biophys., 345, 160-170.
Levine, R.L., Mosoni, L., Berlett, B.S., and Stadtman, E.R. (1996). Methionine residues as endogenous antioxidants in
proteins. Proc. Natl. Acad. Sci. USA, 93, 15036-15040.
Lewisch, S.A., and Levine, R.L. (1995). Determination of 2-oxohistidine by amino acid analysis. Anal. Biochem., 231, 440446.
Li, S., Schoneich, C., and Borchardt, R.T. (1995a). Chemical pathways of peptide degradation. VIII. Oxidation of
methionine in small model peptides by pro-oxidant/transition metal ion systems: Influence of selective scavengers for
reactive oxygen intermediates. Pharmacol. Res., 12, 348-355.
Liehr, J.G., Ulubelen, A.A., and Strobel, H.W. (1986). Cytochrome P-450-mediated redox cycling of estrogens. J. Biol.
Chem., 261, 16865-16870.
Liochev, S.I. (1999). The mechanism of “Fenton-like” reactions and their importance for biological systems. A biologist’s
view. In: Metal Ions in Biological System, 36. Sigel, A., and Sigel, H. (Eds.). Marcel Dekker, Inc., New York. Pp. 1-39.
Lion, Y., Demelle, M., and Van de Vorst, A. (1976). New method of detecting singlet oxygen production. Nature, 263, 442443.
Lipton, S.A., Choi, Y.B., Pan, Z.H., Lei, S.Z., Chen, H.S., Sucher, N.J., Loscalzo, J., Singel, D.J., and Stamler, J.S. (1993).
A redox-based mechanism for the neuroprotective and neurodestructive effects of nitric oxide and related nitrosocompounds [see comments]. Nature, 364, 626-632.
Liu, D. (1993). Generation and detection of hydroxyl radical in vivo in rat spinal cord by microdialysis administration of
Fenton’s reagents and microdialysis sampling. J. Biochem. Biophys. Meth., 27, 281-291.
Liu, X., Miller, M.J.S., Joshi, M.S., Thomas, D.D., and Lancaster, J.R. (1998). Accelerated reactions of nitric oxide with O2
within the hydrophobic interior of biological membranes. Proc. Natl. Acad. Sci. USA, 95, 2175-2179.
Livovich, V., and Scheeline, A. (1997). Amperometric sensors for simultaneous superoxide and hydrogen peroxide
detection. Anal. Chem., 69, 454462.
Luo, X., and Lehotay, D.C. (1997). Determination of hydroxyl radicals using salicylate as a trapping agent by gas
chromatography-mass spectrometry. Clin. Biochem., 30, 41-46.
Lymar, S.V., and Hurst, J.K. (1995a). Rapid reaction between peroxynitrite ion and carbon dioxide: Implications for
biological activity. J. Am. Chem. Soc., 117, 8867-8868.
Lymar, S.V., and Hurst, J.K. (1995b). Role of compartmentalization in promoting toxicity of leukocyte-generated strong
oxidants. Chem. Res. Toxicol., 8, 833-840.
Lymar, S.V., and Hurst, J.K. (1996). Carbon dioxide: Physiological catalyst for peroxynitrite-mediated cellular damage or
cellular protectant? Chem. Res. Toxicol., 9, 845-850.
McCabe, D.R., Maher, T.J., and Acworth, I.N. (1997a). Improved method for the estimation of hydroxyl free radical levels
in vivo based on liquid chromatography with electrochemical detection. J. Chromatogr. B. 691, 23-32.
McCabe, D.R., Acworth, I.N., Maidt, M.L., and Floyd, R.A. (1997b). A sensitive and selective method for the determination
th
of tissue 8-hydroxy-2’deoxyguanosine using HPLC with electrochemical array detection. Presented at The 88 Annual
Meeting of The American Association for Cancer Research. San Diego CA.
McCord, J.M., and Day, E.D. (1978). Superoxide-dependent production of hydroxyl radical catalyzed by iron-EDTA
complex. FEBS Lett., 86, 139-142.
McCormick, M.L., Buettner, G.R., Lewis, T.S., Heinecke, J.W., and Britigan, B.E. (1998). Eosinophil peroxidase forms
dityrosine via generation of tyrosyl radicals. Free Radic. Biol. Med., Supplement 1, S46.
McNeil, C.J., Greenough, K.R., Weeks, P.A., and Self, C.H. (1992). Electrochemical sensors for direct reagentless
measurement of superoxide production by human neutrophils. Free Radic. Biol. Med., 17, 399-406.
MacDougal, C.S., Rigas, M.L., Ben-Jebria, A., and Ultman, J.S. (1998). A respiratory ozone analyzer optimized for high
resolution and swift dynamic response during exercise conditions. Arch. Environ. Health, 53, 161-174.
Madronich, S. (1999). Stratospheric ozone and its effects on the biosphere. In: Reactive Oxygen Species in Biological
Systems. Gilbert, D.L., and Colton, C.A. (Eds.). Kluwer Academic/Plenum Publishers, New York. Pp. 317-334..
Makino, N., Mochizuki, Y., Bannai, S., and Sugita, Y. (1994). Kinetic studies on the removal of extracellular hydrogen
peroxide by cultured fibroblasts. J. Biol. Chem., 269, 1020-1025.
Malinski, T., and Taha, Z. (1992). Nitric oxide release from a single cell measured in situ by a porphyrinic-based
microsensor. Nature, 358, 676-678.
Malmstrom, B.G. (1982). Enzymology of oxygen. Ann. Rev. Biochem., 51, 21-59.
WWW.ESAINC.COM
142
Mansouri, A., and Perry, C.A. (1987). Hemoglobin autoxidation at physiological concentrations. Hemoglobin, 11, 353-371.
Marletta, M.A., Yoon, P.S., Iyengar, R., Leaf, C.D., and Wishnok, J.S. (1988). Macrophage oxidation of L-arginine to nitrite
and nitrate: Nitric oxide is an intermediate. Biochem., 27, 8706-8711.
Martin, W., Villani, G.M., Jothianandan, D., and Furchgott, R.F. (1985). Selective blockade of endothelium-dependent and
glyceryl trinitrate-induced relaxation by hemoglobin and by methylene blue in the rabbit aorta. J. Pharmacol. Exp. Ther.,
232: 708-716.
Masarwa, M., Cohen, H., Meyerstein, D., Hickman, D.L., Bacak, A., and Espenson, J.H. (1988). Reactions of low-valent
+
2+
transition metal complexes with hydrogen peroxide. Are they “Fenton-like” or not? 1. The case of Cu (aq) and Cr (aq). J.
Am. Chem. Soc., 110, 4293-4297.
Maskos, Z., Rush, J.D., and Koppenol, W.H., (1992). The hydroxylation of tryptophan. Arch. Biochem. Biophys. 296, 514520.
Menzel, D.B. (1984). Ozone: An overview of its toxicity in man and animals. J. Toxicol. Environ. Health., 13, 183-204.
Menzel, D.B., and Meacher, D.M. (1999). Ozone and nitrogen dioxide. In: Reactive Oxygen Species in Biological
Systems. Gilbert, D.L., and Colton, C.A. (Eds.). Kluwer Academic/Plenum Publishers, New York. Pp. 335-366.
Meyer, J., Richter, N., and Hecker, M. (1997). High-performance liquid chromatographic determination of nitric oxide
synthase-related arginine derivatives in vitro and in vivo. Anal. Biochem., 247, 11-16.
Michel, T., Xie, Q-W., and Nathan, C. (1996). Molecular biological analysis of nitric oxide synthases. In: Methods in Nitric
Oxide Research. Feelisch, M., and Stamler, J.S. (Eds.). John Wiley and Sons Ltd., London. Pp. 161-175.
Miller, D.M., Buettner, G.R., and Aust, S.D. (1990). Transition metals as catalysts of “autoxidation” reactions. Free Radic.
Biol. Med., 8, 95-108.
Miller, J.W., Selhub, J., and Joseph, J.A. (1996). Oxidative damage caused by free radicals produced during
catecholamine autoxidation: Protective effects of O-methylation and melatonin. Free Radic. Biol. Med., 21, 241-249.
Miller, R.T., Martasek, P., Roman, L.J., Nishimura, J.S., and Masters, B.S. (1997). Involvement of the reductase domain
of neural nitric oxide synthase in superoxide anion production. Biochem., 36, 15277-15284.
Minetti, M., Mallozzi, C., Di Stasi, A.M., and Pietraforte, D. (1998). Bilirubin is an effective antioxidant of peroxynitritemediated protein oxidation in human blood plasma. Arch. Biochem. Biophys., 352, 165-174.
Miyata, T., Inagi, R., Asahi, K., Yamada, Y., Horie, K., Sakai, H., Uchida, K., Kurokawa, K. (1998). Generation of protein
carbonyls by glycoxidation and lipoxidation reaction with autoxidation products of ascorbic acid and polyunsaturated
fatty acids. FEBS Lett., 437, 24-28.
Miyata, T., van Ypersele de Strihou, C., Kurokawa, K., and Baynes, J.W. (1999). Alterations in nonenzymatic biochemistry
in uremia: Origin and significance of “carbonyl stress” in long term uremic complications. Kidney Int., 55, 389-399.
Moan, J., and Wold, E. (1979). Detection of singlet oxygen production by ESR. Nature, 279, 450-451.
Mohr, S., Stamler, J.S., and Brune, B. (1994). Mechanism of covalent modification of glyceraldehyde-3-phosphate
dehydrogenase at its active site thiol by nitric oxide, peroxynitrite and related nitrosating agents. FEBS Lett., 348, 223227.
Moncada, S., and Higgs, A. (1993). The L-Arginine-nitric oxide pathway. N. Engl. J. Med., 329, 2002-2011.
Moncada, S., and Palmer, R.M.J. (1990). The L-Arginine-nitric oxide pathway in the vessel wall. In: Nitric Oxide from LArginine: A Bioregulatory System. Moncada, S., and Higgs, E.A. (Eds.). Elsevier, Amsterdam. Pp. 19-33.
Moncada, S., Palmer, R.M.J., and Higgs, E.A. (1991). Nitric oxide: physiology, pathophysiology, and pharmacology.
Pharmacol. Rev., 43, 109-141.
Monig, J., Asmus, K.D., Forni, L.G., and Willson, R.L. (1997). On the reaction of molecular oxygen with thiyl radicals: A reexamination. Int. J. Radiat. Stud. Phys. Chem. Med., 52, 589-602.
Montgomery, J., Ste-Marie, L., Boismenu, D., and Vachon, L. (1995). Hydroxylation of aromatic compounds as indices of
hydroxyl radical production: A cautionary note revisited. Free Radic. Biol. Med., 19, 927-933.
More, K.P., Darley-Usmar, V., Morrow, J., and Roberts, L.J. (1995). Formation of F2-isoprostanes during oxidation of
human low-density lipoprotein and plasma by peroxynitrite. Circ. Res., 77, 335-341.
Moreno, J.J., and Pryor, W.A. (1992). Inactivation of α-1-proteinase inhibitor by peroxynitrite. Chem. Res. Toxicol., 5, 425432.
Moriel, P., and Abdalla, S. (1997). Nitrotyrosine bound to β-VLDL-apoproteins: A biomarker of peroxynitrite formation in
experimental atherosclerosis. Biochem. Biophys. Res. Commun., 232, 332-335.
Mortensen, A., Skibsted, L.H., Sampson, J., Rice-Evans, C., and Everett, S.A. (1997). Comparative mechanisms and
rates of free radical scavenging by carotenoid antioxidants. FEBS Lett., 24, 91-97.
Motohashi, N., and Mori, I. (1989). High-performance liquid chromatography-electrochemical detection of singlet oxygen
by reaction with 2,2,6,6-teramethyl-4-piperidon. J. Chromatogr., 465, 417-421.
Muijsers, R., Folkerts, G., Henricks, P., Sadeghi-Hashjin, G., and Nijkamp, F. (1997). Peroxynitrite: A two-faced
metabolite of nitric oxide. Life Sci., 60, 1833-1845.
Muscara, M., and de Nucci, G. (1996). Simultaneous determination of nitrite and nitrate anions in plasma, urine and cell
culture supernatants by high-performance liquid chromatography with post-column reactions. J. Chromatogr. B., 686,
157-164.
Mustafa, M.G. (1990). Biochemical basis of ozone toxicity. Free Radic. Biol. Med., 9, 245-265.
Nair, U.J., Nair, J., Friesen, M.D., Bartsch, H., and Ohshima, H. (1995). Ortho- and meta-tyrosine formation from
phenylalanine in human saliva as a marker of hydroxyl radical generation during betel quid chewing. Carcinogen. 16,
1195-1198.
Nakamura, K., Ienaga, K., Nakano, K., Nakai, M., Nakamura, Y., Hasegawa, G., Sawada, M., Kondo, M., Mori, H., and
Kanatsuna, T. (1996). Diabetic renal failure and serum accumulation of creatinine oxidative metabolites creatol and
methylguanidine. Nephron, 73, 520-525.
Nakatsubo, N., Kojima, H., Sakurai, K., Kikuchi, K., Nagoshi, H., Hirata, Y., Akaike, T., Maeda, H., Urano, Y., Higuchi, T.,
and Nagano, T. (1998). Improved nitric oxide detection using 2,3-diaminonaphthalene and its application to the
evaluation of novel nitric oxide synthase inhibitors. Biol. Pharm. Bull., 21, 1247-1250.
WWW.ESAINC.COM
143
Naqui, A., Chance, B., and Cadenas, E. (1986). Reactive oxygen intermediates in biochemistry. Ann. Rev. Biochem., 55,
137-166.
Narayan, M., Berliner, L., Merola, A., Diaz, P., and Clanton, T. (1997). Biological reactions of peroxynitrite: Evidence for
an alternative pathway of salicylate hydroxylation. Free Radic. Res., 27, 63-72.
Nauseef, W.M., Olsson, I., and Arnljots, K. (1988). Biosynthesis and processing of myeloperoxidase – a marker for
myeloid cell differentiation. Eur. J. Haematol., 40, 97-110.
Newsholme, E.A., and Leech, A.R. (Eds.). (1992). Biochemistry for the Medical Sciences. John Wiley and Sons, New
York.
Nguyen, T., Brunson, D., Crespi, C.L., Penman, B.W., Wishnok, J.S. and Tannenbaum, S.R. (1992). DNA damage and
mutation in human cells exposed to nitric oxide in vitro. Proc. Natl. Acad. Sci. USA, 89, 3030-3034.
NIOSH (1994). Chlorine. Method 6011.
Nunoshiba, T., deRojas-Walker, T., Wishnok, J.S., Tannenbaum, S.R., and Demple, B. (1993). Activation by nitric oxide of
an oxidative-stress response that defends E. coli against activated macrophages. Proc. Natl. Acad. Sci. USA, 90, 99939997.
Okamoto, T., Akaike, T., Negano, T., Miyajima, S., Suga, M., Ando, M., Ichimori, K., and Maeda, H. (1997). Activation of
human neutrophil procollagenase by nitrogen dioxide and peroxynitrite: A novel mechanism for procollagenase
activation involving nitric oxide. Arch. Biochem. Biophys., 342, 261-274.
O’Donnell, V.B., Chumley, P.H., Hogg, N., Bloodsworth, A., Darley-Usmar, V.M., and Freeman, B.A. (1997). Nitric oxide
inhibition of lipid peroxidation: Kinetics of reaction with lipid peroxyl radicals and comparison with alpha-tocopherol.
Biochem., 36, 15216-15223.
Ohshima, H., and Bartsch, H. (1999). Quantitative estimation of endogenous N-nitrosation in humans by monitoring Nnitrosoproline in urine. Meth. Enzymol., 301, 40-49.
Olson, L.P., Bartlberger, M.D., and Houk, K.N. (2003). Peroxynitrate and peroxynitrite: A complete basic set onvestigation
of similarities and differences between these Nox species. J. Am. Chem. Soc., 125, 3999-4006.
Padmaja, S., Squadrito, G.L., and Pryor, W.A. (1998). Inactivation of glutathione peroxidase by peroxynitrite. Arch.
Biochem. Biophys., 349, 1-6.
Palacois, M., Knowles, R.G., Palmer, R.M.J., and Moncada, S. (1989). Nitric oxide from L-arginine stimulates the soluble
guanylate cyclase in adrenal glands. Biochem. Biophys. Res. Comm., 165, 802-809.
Palumbo, G., Carlucci, G., Mazzeo, P., Frieri, G., Pimpo, M.T. and Fanini, D. (1995). Simultaneous determination of 5aminosalicylic acid, acetyl-5-aminosalicylic acid and 2,5-dihydroxybenzoic acid in endoscopic intestinal biopsy samples
in humans by high-performance liquid chromatography with electrochemical detection. J. Pharmaceut. Biomed. Anal.
14, 175-180.
Panasenko, O.M., Briviba, K., Klotz, L.O., and Sies, H. (1997a). Oxidative modification and nitration of human low-density
lipoproteins by the reaction of hypochlorous acid with nitrite. Arch. Biochem. Biophys., 343, 254-259.
Panasenko, O.M., Panasenko, O.O., Briviba, K., and Sies, H. (1997b). Hypochlorite destroys carotenoids in low density
lipoproteins thus decreasing their resistance to peroxidative modification. Biochem. (Mosc.), 62, 1140-1145.
Payne, W.J., Le Gall, J., and Berlier, Y. (1996). Use of mass spectrometry for the detection of nitrogen oxides. In:
Methods in Nitric Oxide Research. Feelisch, M., and Stamler, J.S. (Eds.). John Wiley and Sons Ltd., London. Pp. 393402.
Pfeiffer, S., Schrammel, A., Schmidt, K., and Mayer, B. (1998). Electrochemical determination of S-nitrosothiols with a
Clark-type nitric oxide electrode. Anal. Biochem., 258, 68-73.
Pietraforte, D., and Minetti, M. (1997a). One-electron oxidation pathway of peroxynitrite decomposition in human blood
plasma: Evidence for the formation of protein tryptophan-centered radicals. Biochem. J., 321, 743-750.
Pietraforte, D., and Minetti, M. (1997b). Direct ESR detection of peroxynitrite-induced tyrosine-centered protein radicals in
human blood plasma. Biochem. J., 325, 675-684.
Pou, S., Nguyen, S., Gladwell, T., and Rosen, G. (1995). Does peroxynitrite generate hydroxyl radical? Biochim. Biophys.
Acta, 1244, 62-68.
Powis, G., Briehl, M., and Oblong, J. (1995). Redox signaling and the control of cell growth and death. Pharmacol. Ther.
68, 149-173.
Pratt, P.F., Nithipatikom, K., and Campbell, W.B. (1995). Simultaneous determination of nitrate and nitrite in biological
samples by multichannel flow injection analysis. Anal. Biochem., 231, 383-386.
Preik-Steinhoff, H., and Kelm, M. (1996). Determination of nitrite in human blood by combination of a specific sample
preparation with high-performance anion exchange chromatography and electrochemical detection. J. Chromatogr. B.,
685, 348-352.
Pryor, J.J., Jin, X., and Squadrito, G.L. (1994). One- and two-electron oxidations of methionine by peroxynitrite. Proc.
Natl. Acad. Sci. USA, 91, 11173-11177.
Pryor, W.A. (1986). Oxy-radicals and related species: Their formation lifetimes, and reactions. Annu. Rev. Physiol., 48,
657-663.
Pryor, W.A., (1993). Ozone in all its reactive splendor. J. Lab. Clin. Med., 122, 483-486.
Pryor, W.A. (1994). Mechanisms of radical formation from reactions of ozone with target molecules in the lung. Free
Radic. Biol. Med., 17, 451-465.
Pryor, W.A., and Church, D.F. (1991). Aldehydes, hydrogen peroxide and organic radicals as mediators of ozone toxicity.
Free Radic. Biol. Med., 11, 41-46.
Pryor, W.A., and Squadrito, G.L. (1995). The chemistry of peroxynitrite: a product from the reaction of nitric oxide with
superoxide. Am. J. Physiol., 268, L699-L722.
Pryor, W.A., Squadrito, G.L., and Friedman, M. (1995). The cascade mechanism to explain ozone toxicity: The role of lipid
ozonation products. Free Radic. Biol. Med., 19, 935-941.
Pryor, W.A., Stanley, J.P., and Blair, E. (1976). Autoxidation of polyunsaturated fatty acid: II. A suggested mechanism for
the formation of TBA-reactive materials from prostaglandin-like endoperoxides. Lipids, 11, 370-379.
WWW.ESAINC.COM
144
Qian, S.Y., and Buettner, G.R. (1999). Iron and dioxygen chemistry is an important route to initiation of biological free
radical oxidations: An electron paramagnetic resonance spin trapping study. Free Radic. Biol. Med., 26, 1447-1456.
Quijano, C., Alvarez, B., Gatti, R., Augusto, O., and Radi, R. (1997). Pathways of peroxynitrite oxidation of thiol groups.
Biochem. J., 322, 167-173.
Radi, R. (1996). Reactions of nitric oxide with metalloproteins. Chem. Res. Toxicol., 9, 828-835.
Radi, R., Beckman, J.S., Bush, K.M., and Freeman, B.A. (1991a). Peroxynitrite-mediated sulfhydryl oxidation: The
cytotoxic potential of superoxide and nitric oxide. J. Biol. Chem., 266, 4244-4250.
Radi, R., Beckman, J.S., Bush, K.M., and Freeman, B.A. (1991b). Peroxynitrite-induced membrane lipid peroxidation: The
cytotoxic potential of superoxide and nitric oxide. Arch. Biochem. Biophys., 288, 481-487.
Radi, R., Denicola, A., and Freeman, B.A. (1999). Peroxynitrite reactions with carbon dioxide-bicarbonate. Meth.
Enzymol., 301, 353-367.
Radzik, D.M., Roston, D.A., and Kissinger, P.T. (1983). Determination of hydroxylated aromatic compounds produced via
superoxide-dependent formation of hydroxyl radicals by liquid chromatography/electrochemistry. Anal. Biochem. 131,
458-464.
Ralston, S.H. (1997). The Michael Mason Prize Essay 1997. Nitric oxide and bone: what a gas! Br. J. Rheumatol., 36,
831-838.
Ramezanian, M.S., Padmaja, S., and Koppenol, W.H. (1996). Nitration and hydroxylation of phenolic compounds by
peroxynitrite. Chem. Res. Toxicol. 9, 232-240.
Rapaport, R.M., and Murad, F. (1983). Agonist induced endothelium-dependent relaxation in rat thoracic aorta may be
mediated through cyclic GMP. Circ. Res., 52, 352-357.
Rauth, A.M., Goldberg, Z., and Misra, V. (1997). DT-diaphorase: Possible roles in cancer chemotherapy and
carcinogenesis. Oncol. Res., 9, 339-349.
Reist, M., Marshall, K.-A., Jenner, P., and Halliwell, B. (1998). Toxic effects of sulphite in combination with peroxynitrite on
neuronal cells. J. Neurochem., 71, 2431-2438.
Riley, D.P., Rivers, W.J., and Weiss, R.H. (1991). Stopped-flow kinetic analysis for monitoring superoxide decay in
aqueous systems. Anal. Biochem., 196, 344-349.
Rivot, J.P., Barraud, J., Montecot, C., Jost, B., and Besson, J.M. (1997). Nitric oxide (NO): In vivo electrochemical
monitoring in the dorsal horn of the spinal cord of the rat. Brain Res., 773, 66-75.
Robinson, J.M., and Badwey, J.A. (1995). The NADPH oxidase complex of phagocytic leukocytes: A biochemical and
cytochemical view. Histochem., 103, 163-180.
Rohn, T.T., Nelson, L.K., Davis, A.R., and Quinn, M.T. (1999). Inhibition of GTP binding to Rac2 by peroxynitrite: Potential
role for tyrosine modification. Free Radic. Biol. Med., 26, 1321-1331.
Roy, D., Floyd, R.A., and Liehr, J.G. (1991). Elevated 8-hydroxydeoxyguanosine levels in DNA of diethylstilbestrol-treated
Syrian hamsters: Covalent DNA damage by free radicals generated by redox cycling of diethylstilbestrol. Cancer Res.,
51, 3882-3885.
Rubanyi, G., Ho, E.H., Cantor, E.H., Lumma, W.C., and Parker Botelho, L.H. (1991). Cytoprotective function of nitric
oxide: Inactivation of superoxide radicals produced by human leukocytes. Biochem. Biophys. Res. Commun., 181,
1392-1397.
Rubanyi, G.M., Lorenz, R.R., and Vanhoutte, P.M. (1985). Bioassay of endothelium-derived relaxing factor(s): inactivation
by catecholamines. Am. J. Physiol., 249, H95-H101.
Rubbo, H., Darley-Usmar, V., and Freeman, B. (1996). Nitric oxide regulation of tissue free radical injury. Chem. Res.
Toxicol., 9, 809-820.
Rubbo, H., Radi, R., Trujillo, M., Telleri, R., Kalyanaraman, B., Barnes, S., Kirk, M., and Freeman, B.A. (1994). Nitric oxide
regulation of superoxide and peroxynitrite-dependent lipid peroxidation. J. Biol. Chem., 269, 26066-26075.
Runeckles, V.C. (1994). The impact of UV-B radiation and ozone on terrestrial vegetation. Environ. Pollut., 83, 191-213.
Sachs, C.H. and Johnson, G. (1975). Mechanisms of action of 6-hydroxydopamine. Pharmacol., 24, 1-8.
Saez, G., Thornalley, P.J., Hill, H.A., Hems, R., and Bannister, J.V. (1982). The production of free radicals during the
autoxidation of cysteine and their effect in isolated rat hepatocytes. Biochim. Biophys. Acta, 719, 24-31.
Sagone, A.L., Decker, M.A., Wells, R.M., and De Mocko, C. (1980). A new method for the detection of hydroxyl radical
production by phagocytic cells. Biochim. Biophys. Acta, 628, 90-97.
Saha, A., Goldstein, S., Cabelli, D., and Czapski, G. (1998). Determination of optimal conditions for the synthesis of
peroxynitrite by mixing acidified hydrogen peroxide with nitrite. Free Radic. Biol. Med., 24, 653-659.
Sakuma, S., Fujimoto, Y., Sakamoto, Y., Uchiyama, T., Yoshioka, K., Nishida, H., and Fujita, T. (1997). Peroxynitrite
induces the conversion of xanthine dehydrogenase to oxidase in rabbit liver. Biochem. Biophys. Res., Commun., 230,
476-479.
Saleh, D., Barnes, P., and Giaid, A. (1997). Increased production of the potent oxidant peroxynitrite in the lungs of
patients with idiopathic pulmonary fibrosis. Am. J. Respir. Crit. Care Med., 155, 1763-1769.
Salerno, J.C. (1996), Nitric oxide complexes of metalloproteins: An introductory overview. In: Nitric Oxide: Principles and
Actions. Lancaster, J. (Ed.). Academic Press, New York. Pp. 83-110.
Salter, M., Duffy, C., Garthwaite, J., and Strijbos, J.L.M., (1996). Ex vivo measurement of brain tissue nitrite and nitrate
accurately reflects nitric oxide synthase activity in vivo. J. Neurochem., 66, 1683-1690.
Samouilov, A., and Zweier, J.L. (1998). Development of chemiluminescence-based methods for specific quantitation of
nitrosylated thiols. Anal. Biochem., 258, 322-330.
Saran, M., Beck-Speier, I., Fellerhoff, B., and Bauer, G. (1999). Phagocytic killing of microorganisms by radical processes:
Consequences of the reaction of hydroxyl radicals with chloride yielding chlorine atoms. Free Radic. Biol. Med., 26,
482-490.
Sarkar, B. (1995). Metal replacement in DNA-binding zinc finger proteins and its relevance to mutagenicity and
carcinogenicity through free radical generation. Nutrition, 11, 646-649
WWW.ESAINC.COM
145
Satoh, S., Kimura, T., Toda, M., Maekawa, M., Ono, S., Narita, H., Miyazaki, H., Murayama, T., and Nomura, Y. (1997).
Involvement of L-type amino acid transporters in S-nitrosocysteine-stimulated noradrenaline release in the rat
hippocampus. J. Neurochem., 69, 2197-2205.
Sawyer, D.T., Hage, J.P., and Sobkowiak, A. (1995). Iron(II)-induced activation of 1:1 HOOH/HCl for the
chlorohydroxylation of olefins and the chlorination of hydrocarbons: Chlorinated Fenton chemistry. J. Am. Chem. Soc.,
117, 106-109.
Sayre, L.M., Arora, P.K., Iyer, R.S., and Salomon, R.G. (1993). Pyrrole formation from 4-hydroxynonenal and primary
amines. Chem. Res. Toxicol., 6, 19-22.
Sayre, L.M., Sha, W., Xu, G., Kaur, K., Nadkarni, D., Subbanagounder, G., and Salomon, R.G. (1996). Immunochemical
evidence supporting 2-pentylpyrrole formation on proteins exposed to 4-hydroxy-2-nonenal. Chem. Res. Toxicol., 9,
1194-1201.
Schmidt, H., and Kelm, M. (1996). Determination of nitrite and nitrate by the Greiss reaction. In: Methods in Nitric Oxide
Research. Feelisch, M., and Stamler, J.S. (Eds.). John Wiley and Sons Ltd., London. Pp. 491-497.
Schmidt, H., Hofmann, H., Schindler, U., Shutenko, Z., S., Cunningham, D.D., and Feelisch, M. (1996). NO NO from NO
synthase. Proc. Natl. Acad. Sci. USA, 93, 14492-14497.
Schneider, J.E., Browning, M.M., Zhu, X., Eneff, K.L. and Floyd, R.A. (1989). Characterization of hydroxyl mediated
damage to plasmid pBR322 DNA. Mutation Res,. 214, 23-31.
Schuman, E.M., and Madison, D.V. (1991). A requirement for the intercellular messenger nitric oxide in long-term
potentiation. Science, 254, 1503-1506.
Scorza, G., and Minetti, M. (1998). One-electron oxidation pathway of thiols by peroxynitrite in biological fluids:
Bicarbonate and ascorbate promote the formation of albumin disulfide dimers in human blood plasma. Biochem. J., 15,
329, 405-413.
Sevanian, A., McLeod, L.L. (1987). Cholesterol autoxidation in phospholipid membrane bilayers. Lipids, 22, 627-636.
Shibuki, K. (1990). An electrochemical microprobe for detecting nitric oxide release in brain tissue. Neurosci. Res., 9, 6976.
Shintani, F., Kinoshita, T., Kanba, S., Ishikawa, T., Suzuki, E., Sasakawa, N., Kato, R., Asai, M., and Nakaki, T. (1996).
Bioactive 6-nitronorepinephrine identified in mammalian brain. J. Biol. Chem., 271, 13561-13565.
Shoaf, A.R., Shaikh, A.U., Harbison, R.B., and Hinojosa, O. (1991). Extraction and analysis of superoxide free radicals
from whole mammalian liver. J. Biolumin. Chemilumin., 6, 87-96.
Sies, H., Sharov, V.S., Klotz, L.O., and Briviba, K. (1997). Glutathione peroxidase protects against peroxynitrite-mediated
oxidations: A new function for selenoproteins as peroxynitrite reductase. J. Biol. Chem., 272, 27812-27817.
Singel, D.J., and Lancaster, J.R. (1996). Electron paramagnetic resonance spectroscopy and nitric oxide biology. In:
Methods in Nitric Oxide Research. Feelisch, M., and Stamler, J.S. (Eds.). John Wiley and Sons Ltd., London. Pp. 342356.
Singh, A., Singh, H., Kremers, W., and Koroll, G.W. (1981). Involvement of singlet oxygen in biochemical systems. Bull.
Eur. Physiopath. Resp., 17, 31-41.
Singh, R.J., Hogg, N., Joseph, J., and Kalyanaraman, B. (1996). Mechanisms of nitric oxide release from S-nitrosothiols.
J. Biol. Chem., 271, 18596-18603.
Skinner, K.A., Wang, Z., and Parks, D.A. (1996). Detection of salicylate nitration and hydroxylation products after liver
preservation. The Third Annual Meeting of The Oxygen Society, Florida. Poster #2-39.
Skinner, K.A., White, C.R., Patel, R., Tan, S., Barnes, S., Kirk, M., Darley-Usmar, V., and Parks, D.A. (1998). Nitrosation
of uric acid by peroxynitrite. Formation of a vasoactive nitric oxide donor. J. Biol. Chem., 273, 244491-24497.
Sladek, N.E. (1987). Metabolism and action of oxazaphosphorines. In: Metabolism and Action of Anticancer Drugs. Powis,
G., and Prough, R.A. (Eds.). Taylor and Francis, Washington, DC. Pp. 48-90.
Sloot, W.N., and Gramsbergen, J.B.P. (1995). Detection of salicylate and its hydroxylated adducts 2,3- and 2,5dihydroxybenzoic acid as possible indices for in vivo hydroxyl radical formation in combination with catechol- and
indoleamines and their metabolites in CSF and brain tissue. J. Neuroscience Meth., 60, 141-149.
Smith, L.L. (1987). Cholesterol autoxidation 1981-1986. Chem. Phys. Lipids, 44, 87-125.
Smith, M.A., Richey Harris, P., Sayre, L., Beckman, J., and Perry, G. (1997). Widespread peroxynitrite-mediated damage
in Alzheimer’s disease. J. Neuroscience, 17, 2653-2657.
Snyder, S.H., and Bredt, D.S. (1992). Biological roles of nitric oxide. Scientific American, May, 68-77.
Sontag, G., Bernweiser, I., and Krach, C. (1997). HPLC with electrode array detection and its application in analytical food
chemistry. In: Coulometric Electrode Array Detectors for HPLC. Acworth, I.N., Naoi, M., Parvez, S., and Parvez H.
(Eds.). Progress in HPLC-HPCE 6, VSP Publications, The Netherlands. Pp. 75-126.
Souza, J.M., and Radi, R. (1998). Glyceraldehyde-3-phosphate dehydrogenase inactivation by peroxynitrite. Arch.
Biochem. Biophys., 360, 187-194.
Spencer, J.P., Whiteman, M., Jenner, A., and Halliwell, B. (2000). Nitrite-induced deamination and hypochlorite-induced
oxidation of DNA in intact human respiratory tract epithelial cells. Free Radic. Biol. Med., 28, 1039-1050.
Stamler, J.S. (1994). Redox signaling: Nitrosylation and related target interactions of nitric oxide. Cell, 78, 931-936.
Stamler, J.S., and Feelisch., M. (1996). Preparation and detection of S-nitrosothiols. In: Methods in Nitric Oxide Research.
Feelisch, M., and Stamler, J.S. (Eds.). J. Wiley and Sons Ltd., New York. Pp. 521-539.
Stamler, J.S., and Slivka, A. (1996). Biological chemistry of thiols in the vasculature and in vascular-related disease. Nutr.
Rev., 54, 1-30.
Stamler, J.S., Jaraki, O., Osborne, J., Simon, D.I., Keaney, J., Vita, J., Singel, D., Valeri, C.R., and Loscalzo, J. (1992).
Nitric oxide circulates in mammalian plasma primarily as an S-nitroso adduct of serum albumin. Proc. Natl. Acad. Sci.
USA, 89, 7674-7677.
Stamler, J.S., Singel, D.J., and Loscalzo, J. (1992). Biochemistry of nitric oxide and its redox-activated forms. Science,
258, 1898-1902.
Stanbury, D.M. (1989). Reduction potentials involving inorganic free radicals in aqueous solution. Adv. Inorg. Chem., 33,
69-138.
WWW.ESAINC.COM
146
Stoyanovsky, D.A., Melnikov, Z., and Cederbaum, A.I. (1999). ESR and HPLC-EC analysis of the interaction of hydroxyl
radical with DMSO: Rapid reduction and quantification of POBN and PBN nitroxides. Anal. Chem., 71, 715-721.
Strehlitz, B., Grundig, B., Schumacher, W., Kroneck, P., Vorlop, K.-D., and Kotte, H. (1996). A nitrite sensor based on a
highly sensitive nitrite reductase mediator-coupled amperometric detection. Anal. Chem., 68, 807-816.
Ste-Marie, L., Boismenu, D., Vachon, C., and Montgomery, J. (1996). Evaluation of sodium 4-hydroxybenzoate as an
hydroxyl radical trap using gas chromatography-mass spectrometry and high-performance liquid chromatography with
electrochemical detection. Anal. Biochem. 241, 67-74.
Steuhr, D.J., and Marletta, M.A. (1985). Mammalian nitrate biosynthesis: Mouse macrophages produce nitrite and nitrate
in response to E. coli lipopolysaccharide. Proc. Natl. Acad. Sci. USA, 82, 7738-7742.
Stratford, M.R.L. (1999). Measurement of nitrite and nitrate by high-performance ion chromatography. Meth. Enzymol.,
301, 259-269.
Stronks, H.J., Janzen, E.G., and Weber, R.J. (1984). Electrochemical detection of spin adduct aminoxyls (nitroxides)
separated by high performance liquid chromatography. Anal. Letts. 17, 321-328.
Sucher, N.J., and Lipton, S.A. (1991). Redox modulatory site of the NMDA receptor-channel complex: Regulation by
oxidized glutathione. J. Neurosci. Res., 30, 582-591.
Suzaki, E., Kawai, E., Kodama, Y., Suzaki, T., and Masujima, T. (1994). Quantitative analysis of superoxide anion
generation in living cells using chemiluminescence video microscopy. Biochim. Biophys Acta, 1201, 328-332.
Szabo, C. (1996). The pathophysiological role of peroxynitrite in shock, inflammation, and ischemia-reperfusion injury.
Shock, 6, 79-88.
Szabo, C., Cuzzocrea, S., Zingarelli, B., O’Connor, M., and Salzman, A. (1997). Endothelial dysfunction in a rat model of
endotoxic shock. J. Clin. Invest., 100, 723-735.
Tabatabaei, A.R., and Abbott, F.S. (1999). LC/MS analysis of hydroxylation products of salicylate as an indicator of in vivo
oxidative stress. Free Radic. Biol. Med., 26, 1054-1058.
Takahashi, K., Hara, E., Ogawa, K., Kimura, D., Fujita, H., and Shibahara, S. (1997). Possible implications of the
induction of human heme oxygenase-1 by nitric oxide donors. J. Biochem., 121, 1162-1168.
Tamir, S., Burney, S., and Tannebaum, S.R. (1996). DNA damage by nitric oxide. Chem. Res. Toxicol., 9, 821-827.
Tasker, H.S., and Jones, H.Q. (1909). The action of mercaptans on acid chlorides. Part II. The acid chlorides of
phosphorus, sulfur, and nitrogen. J. Chem. Soc., 95, 1910-1912.
Tatoyan, A., and Giulivi, C. (1998). Purification and characterization of nitric-oxide synthase from rat liver mitochondria. J.
Biol. Chem., 273, 11044-11048.
Tatsuma, T., Gondaira, M., and Watanabe, T. (1992). Peroxidase-incorporated polypyrrole membrane electrodes. Anal.
Chem., 64, 1183-1187.
Tatsuma, T., Watanabe, T., Tatsuma, S., and Watanabe, T. (1994). Substrate-purging enzyme electrodes.
Peroxidase/catalase electrodes for H2O2 with an improved upper sensing limit. Anal. Chem., 66, 290-294.
Thiele, J.J., Traber, M.G., Tsang, K., Cross, C.E., and Packer, L. (1997). In vivo exposure to ozone depletes vitamins C
and E and induces lipid peroxidation in epidermal layers of murine skin. Free Radic. Biol. Med., 23, 385-391.
Thomas, E.L., Jefferson, M.M., and Grisham, M.B. (1982). Myeloperoxidase-catalyzed incorporation of amines into
proteins: Role of hypochlorous acid and dichloramines. Biochem., 21, 6299-6308.
Thornalley, P., wolff, S., Crabbe, J., and Stern, A. (1984). The autoxidation of glyceraldehyde and other simple
monosaccharides under physiological conditions catalyzed by buffer ions. Biochim. Biophys. Acta, 797, 276-287.
Tomoda, A., Takizawa, T., Tsuji, A., and Yoneyama, Y. (1981). Kinetic analysis of myoglobin autoxidation by isoelectricfocusing electrophoresis. Biochem. J., 193, 181-185.
Tournaire, C., Croux, S., Maurette, M.T., Beck, I., Hocquaux, M., Braun, A.M., and Oliveros, E. (1993). Antioxidant activity
of flavonoids: Efficiency of singlet oxygen (1 delta g) quenching. J. Photochem. Photobiol. B., 19, 205-215.
Towel, J., and Kalyanaraman, B. (1991). Detection of radical adducts of 5,5-dimethyl-1-pyrroline N-oxide by the combined
use of high-performance liquid chromatography with electrochemical detection and electron spin resonance. Anal.
Biochem., 196, 111-119.
Tsikas, D., Junker, W., and Frolich, J.C. (1998). Determination of dimethylated arginines in human plasma by highperformance liquid chromatography. J. Chromatogr. B, 705, 174-176.
Uchida, K., and Kawakishi, S. (1990). Site-specific oxidation of angiotensin 1 by copper(II) and L-ascorbate: Conversion of
histidine into 2-imidazolones. Arch. Biochem. Biophys., 283, 20-26.
Uchida, K., and Kawakishi, S. (1993). 2-Oxohistidine as a novel biological marker for oxidatively modified proteins. FEBS
Lett., 332, 208-210.
Uchida, K., and Kawakishi, S. (1994). Identification of oxidized histidine generated at the active site of Cu, Zn-superoxide
dismutase exposed to H2O2. J. Biol. Chem., 269, 2405-2410.
Upchurch, G.R., Welch, G.N., and Loscalzo, L. (1995). S-Nitrosothiol: Chemistry, biochemistry and biological actions.
Adv. Pharmacol., 34, 343-349.
Uppa, R.M., and Pryor, W.A. (1996). Carbon dioxide: Catalysis of the reaction of peroxynitrite with ethyl acetoacetate: An
example of aliphatic nitration by peroxynitrite. Biochem. Biophys. Res. Commun., 229, 764-769.
Uppa, R.M., Lemercier, J.N., Squadrito, G.L., Zhang, H., Bolzan, R.M., and Pryor, W.A. (1998). Nitrosation by
peroxynitrite: Use of phenol as a probe. Arch. Biochem. Biophys., 358, 1-16.
Uppa, R.M., Squadrito, G.L., Cueto, R., and Pryor, W.A. (1996). Selecting the most appropriate synthesis of peroxynitrite.
Meth. Enzymol., 269, 285-295.
Uppa, R.M., Squadrito, G.L., and Pryor, W.A. (1996). Acceleration of peroxynitrite oxidations by carbon dioxide. Arch.
Biochem. Biophys., 327, 335-343.
van Dalen, C.J., Whitehouse, M.W., Winterbourn, C.C., and Kettle, A.J. (1997). Thiocyanate and chloride as competing
substrates for myeloperoxidase. Biochem. J., 327, 487-492.
Van der Veen, R., Hinton, D., Incardonna, F., and Hofman, F. (1997). Extensive peroxynitrite activity during progressive
stages of central nervous system inflammation. J. Neuroimmunol., 77, 1-7.
WWW.ESAINC.COM
147
van der Vliet, A., Eiserich, J.P., Halliwell, B., and Cross, C.E. (1997). Formation of reactive nitrogen species during
peroxidase-catalyzed oxidation of nitrite. J. Biol. Chem., 272, 7617-7625.
van der Vliet, A., Eiserich, J.P., O’Niell, C.A., Halliwell, B., and Cross, C.E. (1995). Tyrosine modification by reactive
nitrogen species: A closer look. Arch. Biochem. Biophys., 319, 341-349.
van der Vliet, A., O’Neill, C.A., Halliwell, B., Cross, C.E., and Kaur, H. (1994). Aromatic hydroxylation and nitration of
phenylalanine and tyrosine by peroxynitrite. FEBS Letts., 339, 89-92.
van der Vleit, A., O’Neill, C.A., Eiserich, J.P., and Cross, C.E. (1995). Oxidative damage to extracellular fluids by ozone
and possible protective effects of thiols. Arch. Biochem. Biophys., 321, 43-50.
van der Vleit, A., t’Hoen, P.A., Wong, P.S., Bast, A., and Cross, C.E. (1998). Formation of S-nitrosothiols via direct
nucleophilic nitrosation of thiols by peroxynitrite with elimination of hydrogen peroxide. J. Biol. Chem., 273, 3025530262.
van Heusden, S., and Mans, L.G. (1978). Alternating measurement of ambient and cabin ozone concentrations in
commercial jet aircraft. Aviat. Space Environ. Med., 49, 1056-1061.
Varma, S.D. (1989). Radio-isotopic determination of sub-nanomolar amounts of peroxides. Free Rad. Res. Commun., 5,
359-368.
Vasela, A., and Wilhelm, J. (2002). The role of carbon dioxide in free radical reactions of the oganism. Physiol. Res., 51,
335-339.
Vasquez-Vivar, J., and Augusto, O. (1992). Hydroxylated metabolites of the antimalarial drug primaquine. Oxidation and
redox cycling. J. Biol. Chem., 267, 6848-6854.
Vasquez-Vivar, J., Santos, A.M., Junqueira, V.B., and Augusto, O. (1996). Peroxynitrite-mediated formation of free
radicals in human plasma: EPR detection of ascorbyl, albumin-thiyl and uric acid-derived free radicals. Biochem. J.,
314, 869-876.
Vasquez-Vivar, J., Kalyanaraman, B., Martasek, P., Hogg, N., Masters, B.S., Karoui, H., Tordo, P., and Pritchard, K.A.
(1998). Superoxide generation by endothelial nitric oxide synthase: The influence of cofactors. Proc. Natl. Acad. Sci.
USA, 95, 9220-9225.
Viner, R.I., Huhmer, A.F., Bigelow, D.J., and Schoneich, C. (1996). The oxidative inactivation of sarcoplasmic reticulum
calcium-ATPase by peroxynitrite. Free Radic. Res., 24, 243-259.
Vissers, M.C., and Winterbourne, C.C. (1991). Oxidative damage to fibronectin. II. The effects of H2O2 and the hydroxyl
radical. Arch. Biochem. Biophys., 285, 357-364.
Vogt, W. (1995). Oxidation of methionyl residues in proteins: Tools, targets, and reversal. Free Radic. Biol. Med., 18, 6173.
Vukomanovic, D.V., Hussain, A., Zoutman, D.E., Marks, G.S., Brien, J.F., and Nakatsu, K. (1998). Analysis of nanomolar
S-nitrosothiol concentrations in physiological media. J. Pharmacol. Toxicol. Meth. 39, 235-240.
Wade, R., and Castro, C. (1990). Redox reactivity of iron III porphyrins and heme proteins with nitric oxide. Nitrosyl
transfer to carbon, oxygen, nitrogen and sulfur. Chem. Res. Toxicol., 3, 289-291.
Wagner, J.R., Hu, C.-C., and Ames, B.N. (1992). Endogenous oxidative damage of deoxycytidine in DNA. Proc. Natl.
Acad. Sci. USA, 89, 3380-3384.
Wang, S.S., and Maher, T.J. (1992). Changes in hippocampal ECF amino acid levels in anesthetized rats during kainic
acid-induced seizures: A microdialysis study. Pharmacologist, 34, 190.
Wardman, P., and Candeias, L.P. (1996). Fenton Chemistry: An introduction. Rad. Res., 145, 523-531.
Wefers, H., and Sies, H. (1983). Oxidation of glutathione by superoxide radical to the disulfide and the sulfonate yielding
singlet oxygen. Eur. J. Biochem., 137, 29-36.
Weil, L., and Morris, J.C. (1949). Kinetic studies on the chloramines. I. The rates of formation of monochloramine, Nchlormethylamine, and N-chlordimethylamine. J. Am. Chem. Soc., 71, 1664-1671.
Weiss, S.J., and Slivka, A. (1982a). Monocyte and granulocyte-mediated tumor cell destruction. A role for the hydrogen
peroxide-myeloperoxidase-chloride system. J. Clin. Invest., 69, 255-262.
Weiss, S.J., and Ward, P.A. (1982). Immune complex induced generation of oxygen metabolites by human neutrophils. J.
Immunol., 129, 309-313.
Weiss, S.J., Klein, R., Slivka, A., and Wei, M. (1982). Chlorination of taurine by human neutrophils: Evidence for
hypochlorous acid generation. J. Clin. Invest., 70, 598-607.
Weiss, S.J., Lampert, M.B., and Test, S.T. (1983). Long-lived oxidants generated by human neutrophils: Characterization
and bioactivity. Science, 222, 625-628.
Weiss, S.J., Test, S.T., Eckmann, C.M., Roos, D., and Regiani, S. (1986). Brominating oxidants generated by human
eosinophils. Science, 234, 200-203.
White, R.E., and Coon, M.J. (1980). Oxygen activation by cytochrome P-450. Ann. Rev. Biochem 49, 315-356.
Whiteman, M., and Halliwell, B. (1996). Protection against peroxynitrite-dependent tyrosine nitration and α1-antiproteinase
inactivation by ascorbic acid. A comparison with other biological antioxidants. Free Radic. Res., 25, 275-283.
Whiteman, M., Siau, J.L., and Halliwell, B. (2003). Lack of tyrosine nitration by hypochlorous acid in the presence of
physiological concentrations of nitrite. Implication for the role of nitryl chloride in tyrosine nitration in vivo. J. Biol. Chem.,
278, 8380-8384.
Wink, D.A., Kasprzak, K.S., Maragos, C.M., Elespuru, R.K., Misra, M., Dunams, T.M., Cebula, T.A., Koch, W.H., Andrews,
A.W., and Allen, J.S. (1991). DNA deaminating ability and genotoxicity of nitric oxide and its progenitors. Science, 254,
1001-1003.
Wink, D.A., Kim, S., Coffin, D., Cook, J.C., Vodovotz, Y., Chistodoulou, D., Jourd’heuil, D., and Grisham, M.B. (1999).
Detection of S-nitrosothiols by fluorometric and colorimetric methods. Meth. Enzymol., 301, 201-211.
Winterbourn, C.C., Vandenberg, J.J.M., Roitman, E., and Kuypers, F.A. (1992). Chlorohydrin formation from unsaturated
fatty acids reacted with hypochlorous acid. Arch. Biochem. Biophys., 296, 547-555.
Wolff, S.P., and Dean, R.T. (1987). Glucose autoxidation and protein modification. The potential role of autoxidative
glycosylation in diabetes. Biochem. J. 245, 243-250.
Wolmann, M. (1975). Autoxidation in vivo. Isr. J. Med. Sci., 15, 30-57.
WWW.ESAINC.COM
148
Wright, C.D., Hattori, R., Kosuga, K., Eizawa, H., Hiki, K., and Kawai, C. (1991). Purification of nitric oxide synthase from
rat macrophages. Biochim. Biophys. Acta, 160, 813-819.
Wu, W., Chen, Y., d’Avignon, A., and Hazen, S.L. (1999). 3-Bromotyrosine and 3,5-dibromotyrosine are major products of
protein oxidation by eosinophil peroxidase: Potential markers for eosinophil-dependent tissue injury in vivo. Biochem.,
38, 3538-3548.
Wyllie, S., and Liehr, J.G. (1997). Release of iron from ferritin storage by redox cycling of stilbene and steroid estrogens
metabolites: A mechanism of induction of free radical damage by estrogens. Arch. Biochem. Biophys., 346, 180-186.
Xia, Y., Roman, L.J., Masters, B.S., and Zweier, J.L. (1998). Inducible nitric oxide synthase generates superoxide form
the reductase domain. J. Biol. Chem., 273, 22635-22639.
Xu, G., and Sayre, L.M. (1998). Structural characterization of a 4-hydroxy-2-nonenal-derived fluorophore that contributes
to lipoperoxidation-dependent protein cross-linking in aging and degenerative diseases. Chem. Res. Toxicol., 11, 247251.
Yager, J.D., and Liher, J.G. (1996). Molecular mechanisms of estrogen carcinogenesis. Ann. Rev. Pharmacol. Toxicol.,
36, 203-232.
Yamada, K., and Nabeshima, T. (1997). Simultaneous measurement of nitrite and nitrate levels as indices of nitric oxide
release in the cerebellum of conscious rats. J. Neurochem., 68, 1234-1243.
2+
Yamazaki, I., and Piette, L.H. (1990). ESR spin-trapping studies on the reaction of Fe ions with H2O2-reactive species in
oxygen toxicity in biology. J. Biol. Chem., 265, 13589-13594.
Yamamoto, Y., and Ames, B.N. (1987). Detection of lipid hydroperoxides and hydrogen peroxide at picamole levels by an
HPLC and isoluminol chemiluminescence assay. Free Radic. Biol. Med., 3, 359-361.
Yao, S.J., Xu, W., and Wolfson, S.K. (1995). A micro carbon electrode for nitric oxide monitoring. ASAIO J., 41, M404409.
Yermilov, V., Rubio, J., Becchi, M., Friesen, M.D., Pignatelli, B., and Ohshima, H. (1995). Formation of 8-nitroguanine with
peroxynitrite in vitro. Carcinogen., 16, 2045-2050.
Yermilov, V., Rubio, J., and Ohshima, H. (1995). Formation of 8-nitroguanine in DNA treated with peroxynitrite in vitro and
its rapid removal from DNA by depurination. FEBS Lett., 376, 207-210.
Yermilov, V., Yoshie, Y., Rubio, J., and Ahshima, H. (1996). Effects of carbon dioxide/bicarbonate on induction of DNA
single-strand breaks and formation of 8-nitroguanine, 8-oxoguanine and base-propenal mediated by peroxynitrite.
FEBS Lett., 399, 67-70.
Yokoyama, H., Kasai, N., Ueda, Y., Niwa, R., Konaka, R., Mori, N., Tsuchihashi, N., Matsue, T., Ohya-Nishiguchi, H., and
Kamada, H. (1998). In vivo analysis of hydrogen peroxide and lipid radicals in the striatum of rats under long-term
administration of a neuroleptic. Free Radic. Biol. Med., 24, 1056-1060.
Yokozawa, T., Fujitsuka, N., Oura, H., Ienaga, K., and Nakamura, K. (1997). In vivo effect of hydroxyl radical scavenger
on methylguanidine production from creatinine. Nephron, 75, 103-105.
Zhai, X., and Ashraf, M. (1995). Direct detection and quantification of singlet oxygen during ischemia and reperfusion in
rat hearts. Am. J. Physiol., 269, H1229-H1236.
Zhang, H., Squadrito, G.L., Uppa, R.M., Lemercier, J.-N., Cueto, R., and Pryor, W.A. (1997). Inhibition of peroxynitritemediated oxidation of glutathione by carbon dioxide. Arch. Biochem. Biophys., 339, 183-189.
Zou, L., Gunn, C., and Beckman, J.S. (1992). Bactericidal activity of peroxynitrite. Arch. Biochem. Biophys., 298, 452-457.
Zingaerlli, B., Day, B.J., Crapo, J.D., Salzman, A.L., and Szabo, C. (1997). The potential role of peroxynitrite in the
vascular contractile and cellular energetic failure in endotoxic shock. Br. J. Pharmacol., 129, 259-267.
Zingaerlli, B., O’Connor, M., Wong, H., Salzman, A., and Szabo, C. (1996). Peroxynitrite-mediated DNA strand breakage
activates poly-adenosine diphosphate ribosyl synthetase and causes cellular energy depletion in macrophages
stimulated with bacterial lipopolysaccharide. J. Immunol., 156, 350-358.
Zingler, D.M. (1985). Role of reversible oxidation-reduction of enzyme thiols-disulfides in metabolic regulation. Ann. Rev.
Biochem., 54, 305-329.
Zou, A.-P., and Cowley, A.W. (1997). Nitric oxide in renal cortex and medulla. Hypertension, 29, 194-198.
WWW.ESAINC.COM
149
Appendix 2.1
Background to Electrode Potentials
As mentioned in Chapter 2, oxidation can be defined as a gain in oxygen, a loss
of hydrogen, or the loss of electrons. Conversely, reduction is the loss of oxygen,
a gain of hydrogen, or the gain of electrons. The two processes are
complimentary and no oxidation process can take place without a corresponding
reduction; these complementary reactions are typically referred to as REDuctionOXidation or REDOX reactions. Consider the reaction between hydrogen and
oxygen (Eqn 2.1.1) and Fe (II) salts with hydrogen peroxide (Eqn 2.1.2). The
substance that provides the oxygen or removes hydrogen (oxygen and hydrogen
peroxide, respectively), and so becomes reduced, is the oxidizing agent;
similarly, the substance that provides the hydrogen or removes oxygen
(hydrogen and Fe (II), respectively), and so becomes oxidized, is the reducing
agent. Sometimes redox reactions can be extremely complex (Eqn 2.1.3) so it is
often easier to determine oxidation and reduction reactions from partial equations
(Eqns 2.1.4 and 2.1.5).
2H2 + O2 → 2H2O
Fe2+ + H2O2 → Fe3+ + HO• + OHMnO4- + 8H+ + 5Fe2+ → 5Fe3+ + Mn2+ + 4H2O
MnO4- + 8H+ + 5e- → Mn2+ + 4H2O (reduction)
5Fe2+ → 5Fe3+ +5e (oxidation)
Eqn 2.1.1
Eqn 2.1.2
Eqn 2.1.3
Eqn 2.1.4
Eqn 2.1.5
One important example of oxidation-reduction processes includes the reactions
associated with aerobic respiration. Electrons from reducing agents such as
NADH are passed along the mitochondrial electron transport chains to a
terminal-oxidizing agent, oxygen. During this process the components of these
chains (various cytochromes, flavoproteins, CoQ10, etc.) undergo redox reactions
as electrons are passed from one to another. In this way, the free energy from
the oxidation NADH by oxygen (∆Go’= -220 kJ mol-1) is utilized in a series of
steps to synthesize 2.5 or 3 moles of ATP (from ADP and PI).
THERMODYNAMICS OF REVERSIBLE CELLS.
Perhaps the best way to illustrate the thermodynamics of redox processes is to
give a simple example. School children learn that dropping zinc granules or iron
filings into a solution of Cu (II) sulfate results in the surface of these metals being
replaced by a coating of copper metal (Eqn 2.1.6). If, however, copper granules
are added to a zinc sulfate solution nothing happens (Eqn 2.1.7). The reason that
copper can be liberated from its solution while zinc cannot is a result of free
WWW.ESAINC.COM
150
energy release. For the reaction of zinc granules with copper sulfate solution, the
∆Go is –213.4 kJ mol-1 at 25oC (corresponding to an equilibrium constant of 1037)
and this is a spontaneous, irreversible reaction. The reaction of copper granules
and zinc sulfate has a ∆Go of +213.4 kJ mol-1 so will not occur spontaneously.
This reaction can be forced to occur if the appropriate energy is put into the
system.
Zn + Cu2+ → Zn2+ + Cu
Cu + Zn2+ → Cu2+ + Zn
√
X
Eqn 2.1.6
Eqn 2.1.7
The free energy from a spontaneous reaction can be utilized in the form of
electrical work. The ability for a reaction to do work can be studied by setting up
an electrochemical cell (Figure 2.1.1). The overall reaction can be divided into a
separate oxidation (Eqn 2.1.8) and reduction process (Eqn 2.1.9) commonly
called a half-cell reaction. One of the two half-cell reactions takes place in each
beaker shown in Figure 2.1.1.
Electrical
Electrical work
work
Current
Current
Zn
Zn
Cu
Cu
Salt
Salt bridge
bridge
ZnSO
ZnSO44
CuSO
CuSO44
Figure 2.1.1 Electrochemical Cell For Zn/Cu Couple.
Zn – 2e- → Zn2+ (oxidation)
Cu2+ + 2e- → Cu (reduction)
WWW.ESAINC.COM
Eqn 2.1.8
Eqn 2.1.9
151
A salt bridge connects the two beakers (typically KCl gel). This allows electrical
connection between the beakers, while also preventing the direct reaction that
would result in the precipitation of copper. The circuit is completed when the zinc
and copper electrodes are connected. Electrons will flow from the zinc to the
copper electrode as zinc ions are formed and copper ions are reduced. Overtime,
as the reaction proceeds to equilibrium, ∆G falls and the amount of electrical
work obtained from the cell decreases. A familiar example of this is the dry-cell
battery. When the reaction within the battery reaches equilibrium, no voltage is
produced and the battery is “dead”.
Battery
+
-
A
B
G
Zn/Cu
Cell
Galvanometer
Figure 2.1.2 Measurement Of The EMF Of A Cell.
The determination of free energy changes of an electrochemical cell can be
obtained by measuring the electromotive force (EMF). If an external voltage is
applied to the two electrodes shown in Figure 2.1.2 so as to oppose the direction
of current of the cell, then at a certain point (A) the current flowing in the cell will
be zero (the “null point”). Under these conditions the potential difference of the
null point is the EMF. If the external voltage is further increased (B), the current
will reverse its direction as the cell reaction is reversed (i.e., Cu + Zn2+ → Zn +
Cu2+). At the null point the cell is behaving reversibly. The thermodynamics of the
cell can now be explored. The ∆G (the amount of useful work) is related to the
potential difference (E) of a reaction by Eqn 2.1.10. Measurement of the EMF (E)
of a redox reaction gives ∆G directly (as long as the number of electrons (n)
taking part in the reaction is known). For the cell shown in Figure 2.1.1, n=2 and
the maximal amount of electrical work that can be obtained using Eqn 2.1.10 is
212.3 kJ mol-1 Zn. As a decrease in free energy (-∆G) is defined to be equal to
WWW.ESAINC.COM
152
the maximal amount of electrical work that can be performed, then ∆G must be 212.3 kJ mol-1 Zn. Since the electrochemical cell contains substances in their
standard states and the temperature is at 25oC, the free energy change of the
system now becomes the standard free energy change. Therefore ∆Go =-212.3
kJ mol-1 Zn.
∆G = -nFE
Eqn 2.1.10
(where F is the Faraday constant and n is the number of
electrons transferred in the reaction).
STANDARD ELECTRODE POTENTIALS.
The half-cell potential cannot be measured directly (the very act of carrying out a
measurement would introduce another metal into the solution that would set up
its own electrode potential). However, as discussed above, the difference
between the potentials of two half cells as part of an electrochemical cell can be
measured. If one of the half-cells is a reference electrode then a series of relative
values of electrode potentials can be obtained. The typical reference electrode is
the hydrogen electrode (Eqn 2.1.11). The hydrogen electrode under standard
conditions (H2, 1 atm; H+, 1M) is arbitrarily assigned a standard electrode
potential (SEP) of 0mV. To obtain SEP values the reference and test electrodes
(e.g., Zn/Zn2+) are connected together via a salt bridge as shown in Figure 2.1.3.
It is then relatively simple to measure the electrode potential using a voltameter.
As it is inconvenient to measure them in the laboratory, SEP values can readily
be obtained from any good chemical textbook. SEPs are sometimes also referred
to as redox potentials.1
H+ + e- → 1/2H2 (Pt)
Eqn 2.1.11
Standard electrode potentials can be used to predict whether a reaction will
occur or not. Let us return again to the zinc/copper reaction. If the EMFs of the
half-cells are known, then the EMF for the total reaction can readily be
calculated, as it is the algebraic difference between the two electrode potentials.
By convention the total standard electrode potential (EoTOTAL) is the difference
between the two half reactions (EoTOTAL = EoRHS - EoLHS). So, using the SEPs for
the reaction of zinc with Cu (II) sulfate, EoTOTAL= +1.1V (Eqns 2.1.12 to 2.1.14).
Thus it is thermodynamically possible for zinc to reduce copper. This is well
1
The electrode potential (E) of a reaction when carried out under standard state conditions (i.e., mole of substance under
o
1 atmosphere of pressure) is denoted by E . In many biochemical processes there is a net uptake or release of protons as
the reaction proceeds. A 1M solution of protons has a pH of 0 which is of little use to biochemists who normally study
reactions at neutrality (~pH 7.0). To circumvent this problem the biochemical standard state can be used where all
+
-7
substances are in their standard state except H , which is present at 10 M. The biochemical standard electrode potential
o’
is denoted by E .
WWW.ESAINC.COM
153
supported by the ∆Go value for this reaction (∆Go= -213.4 kJ mol-1). For the
reaction of copper with zinc sulfate, the Eo would be –1.10V indicating that it is
thermodynamically impossible for copper to reduce zinc ions. A negative Eo
(positive ∆Go) shows that the oxidized form is favored whereas a positive
Eo (negative ∆Go) shows that the reduced state is favored. Thus using the
standard electrode potentials shown in Table 2.5 it can be seen that H2 will not
reduce Zn2+ to Zn but it will reduce Cu2+ to copper.
Eo = -(-0.76V)
Eo = +0.34V
EoTOTAL = +1.10V
Zn - 2e- → Zn2+ Oxidation
Cu2+ + 2e- → Cu Reduction
Zn + Cu2+ → Zn2+ + Cu
Eqn 2.1.12
Eqn 2.1.13
Eqn 2.1.14
Valve
Valve Voltmeter
Voltmeter
V
Salt
Salt bridge
bridge
Electron
Electron flow
flow ifif metal
metal M
M
has
has aa negative
negative electrode
electrode
potential
potential
Electron
Electron flow
flow ifif metal
metal M
M
has
has aa positive
positive electrode
electrode
potential
potential
Hydrogen
Hydrogen
Metal
Metal M
M
n+
Metal
Metal M
Mn+
Molar
Molar H+
H+
Figure 2.1.3 Apparatus For Measuring SEPs.
Simple addition and subtraction cannot always combine standard electrode
potentials. The number of electrons must also be considered. Consider the two
half-cell reactions Eqn 2.1.15 and 2.1.16. Simply subtracting Eqn 2.1.16 from
Eqn 2.1.15 cannot solve the conversion of Fe (III) to Fe (II) (Eqn 2.1.17). The
reaction would not be balanced. The answer can only be found if the number of
electrons are considered. The answer is thus obtained arithmetically [(3 x Eqn
2.1.15) – (2 x Eqn 2.1.16)] and is found to be -0.76V (Eqn 2.1.17).
WWW.ESAINC.COM
154
Fe3+ + 3e- = Fe
Fe2+ + 2e- = Fe
Fe3+ + e- = Fe2+
-0.04V
-0.44V
-0.76V
Eqn 2.1.15
Eqn 2.1.16
Eqn 2.1.17
SOME COMMENTS ON SEPS.
Now that we have a clearer understanding of SEPs, we can apply them to the
field of redox biochemistry. A wide variety of SEPs can be found in the literature
and some of them are presented in Table 2.1.1. The SEPs are placed in
descending order from the most positive (the reaction of the strongest oxidizing
agent, the chlorine atom) to the most negative (the reaction of the strongest
reducing agent, the electron). Compounds know to be antioxidants (see Chapter
4) have typical biochemical SEP values of <500mV, significantly below the Eo’
values of the reactions of the pro-oxidant species (+600 to +2310mV). Thus the
reaction of antioxidants with reactive species such as alkyl radicals (L•) and lipid
peroxyl radicals (LO2•) (Eqns 2.1.18 and 2.1.19) forms products which are much
less oxidizing (Eqns 2.1.20 and 2.1.21). Closer examination of these reactions
reveals that the Eo’ of LO2• is +400mV higher than that of L•, indicating that LO2•
is a stronger oxidizing agent. This is significant biologically as σ-radicals like L•
readily react with oxygen to form LO2•. Thus during lipid peroxidation processes
very strong oxidizing agents are produced which can serve to promote this chain
reaction.
Based solely upon the biochemical SEPs presented in Table 2.1.1 there is an
antioxidant hierarchy with the antioxidant activity of α-tocopherol being the least
and lipoic acid being the most effective. This explains why during lipid
peroxidation membrane bound α-tocopherol can be regenerated from its radical
by membrane bound coenzyme Q10 or cytosolic ascorbic acid, GSH and lipoic
acid. The biological importance of these antioxidants in lipid peroxidation is
described in Chapters 3 and 4.
LO2• + H+ + e- = LO2H
L• + H+ + e- = LH
TO• +
H+ + e- = TOH
(tocopherol)
½(Dehydroascorbate + 2H+ +
2e- = Ascorbate)
Eo’ ≅ +1000mV (strong oxidant)
Eo’ ≅ +600mV
Eo’ ≅ +500mV
Eqn 2.1.18
Eqn 2.1.19
Eqn 2.1.20
Eo’ = +40mV (weak oxidant)
Eqn 2.1.21
Another interesting observation found in Table 2.1.1 is that the Eo’ value for the
reaction H2O2 + H+ + e- = H2O + HO• is only +320mV (not much different in value
from many antioxidant reactions). Although at first glance this reaction might
appear unimportant, it is, in fact, a major problem for all aerobic organisms.
WWW.ESAINC.COM
155
Hydrogen peroxide is readily decomposed to hydroxyl free radicals (e.g., by
ferrous ions in the Fenton reaction). Once formed, the hydroxyl free radical is an
incredibly strong oxidizing agent and the HO•, H+/H2O reaction has an
Eo’=+2310mV. Just by reacting with iron, the standard electrode potential of one
reactive oxygen species (hydrogen peroxide) can be increased by ≅+2000mV,
forming one of the most oxidizing agents known!
REDOX COUPLE
One Electron Reactions
Cl• + e- = ClSO4•- + e- = SO42HO• + H+ + e- = H2O (strongest oxidant under
biochemical conditions)
NO3• + e- = NO3Cl2•- + e- = 2ClBr• + e- = BrHO• + e- = OHC2H5• + H+ + e- = C2H6
O3•- + 2H+ + e- = O2 + H2O
Br2•- + e- = 2BrRO• + H+ + e- = ROH (aliphatic)
NO2+ + e- = NO2•
ONO2- + 2H+ + e- = NO2• + H2O
Mn3+ + e- = Mn2+
CO3- + e- = CO321
Σg O2 + e- = O2•NO+ + e- = NO•
N2O4 + e- = NO• + NO3•
CH2OH + H+ + e- = CH3OH
O3 + e- = O3••
O2SO3- + e- = SO52CO2•- + H+ + e- = HCO2HO2• + H+ + e- = H2O2
Tryptophan Radical + H+ + e- = Tryptophan
NO2• + e- = NO2HRP-II/HRP (horseradish peroxidase)
Allyl• + H+ + e- = allyl-H
HRP-I/HRP-II
O2•- + 2H+ + e- = H2O2
RS• + e- = RS- (cysteine)
Phenol• + H+ + e- = Phenol
N2O3 + e- = NO• + NO2LO2• + H+ + e- = LO2H (alkylperoxyl)
1
∆g O2 + e- = O2•SO3•- + e- = SO32PUFA• (L•) + H+ + e- = PUFA-H (L-H)
HU•- + H+ + e- = UH2- (urate)
CA-O• + H+ + e- = CA-OH (catechol)
α-TO• + H+ + e- = α-TOH (tocopherol)
ONO2• + e- = ONO2Cu2+ + H+ + e- = Cu+H (SOD)
WWW.ESAINC.COM
Eo(’)/mV
STRONG
OXIDANT
+2550
+2430
+2310
+2300, +2600
+2300
+2000
+1900
+1900
+1800
+1800
+1600
+1600
+1600
+1510
+1500
+1270
+1210
+1200
+1200
+1140
+1100
+1100
+1060
+1000
+990
+970
+960
+950
+940
+920
+900
+800
+770, +1000,
+1400
+650
+630
+600
+590
+530
+500
+430
+420
156
Cl2 + e- = Cl2•Br2 + e- = Br2•NO• + e- = NO- (triplet)
FeO22+ + 2H+ + 2e- (myoglobin)
H2O2 + H+ + e- = H2O + HO•
Mn3+ + H+ + e- = Mn2+H (SOD)
Fe3+ + O2 + e- = FeO22+ (hemoglobin)
NAD• + H+ + e- = NADH
FeO22+ + 2H+ + 2e- (myoglobin)
Asc•- + H+ + e- = AscH- (ascorbate)
Fe3+ + H+ + e- = Fe2+H (SOD)
Fe3+ + e- = Fe2+ (cytochrome)
Fe3+ + O2 + e- = FeO22+ (myoglobin)
ONO2• + e- = ONO2CoQ•- + H+ + e- = CoQH2 (coenzyme Q10)
Cu2+ + e- = Cu+
Fe3+ + e- = Fe2+ (hemoglobin)
Fe3+(EDTA) + e- = Fe2+(EDTA)
Fe3+ + e- = Fe2+ (aqueous, pH 7.0)
Fe3+(citrate) + e- = Fe2+(citrate)
Fe3+(ADP) + e- = Fe2+(ADP)
Fe3+ + e- = Fe2+ (myoglobin)
HOCl + e- = Cl- + HO•
CoQ + H+ + e- = CoQ•Dehydroascorbate + H+ + e- = Ascorbate•
HOBr + H+ + e- = H2O + Br•
Fe3+ + e- = Fe2+ (ferritin)
FADH• + H+ + e- = FADH2
Riboflavin + e- = riboflavin•O2 + e- = O2•Adriamycin + e- = Adriamycin semiquinone
NO• + e- = NO- (singlet)
HOBr + e- = Br- + HO•
Fe3+ + e- = Fe2+ (transferrin)
Paraquat2+ + e- = Paraquat+
O2 + H+ + e- = HO2•
HOCl + H+ + e- = H2O + Cl•
NAD+ + e- = NAD
CH3CHO + H+ + e- = CH3C•HOH
RSSR + e- = RSSR•(e.g., cystine or glutathione disulfide)
NAD+ + H+ + e- = NADH+
CO2 + e- = CO2•H2O + e- = e-aq (hydrated electrons)
Multiple Electron Reactions
a) Two electron
O3 + 2H+ + 2e- = O2 + H2O
Cl2 (aq) + 2e- + 2ClHOBr + H+ + 2e- = H2O + BrHOCl + H+ + 2e- = H2O + ClUric diimine + 2H+ + 2e- = Uric acid
O2 + 2H+ + 2e- = H2O2
CoQ + 2H+ + 2e- = CoQH2
WWW.ESAINC.COM
+420
+410
+390
+390
+320
+310
+310
+300
+300
+282
+280
+260
+220
+200
+200, +350
+160
+140
+120
+110
+100
+100
+50
-40
-36, -230
-174
-180
-190
-240
-317
-330
-330
-350
-360
-400
(pH 7.3)
-450
-460
-460
-930
-1380
-1500
-1580
-1800
-2870
STRONG
REDUCTANT
+2075
+1400
+1090
+1080
+400
+330
+100
157
Dehydroascorbate + 2H+ + 2e- = Ascorbate
FAD+ + 2H+ + 2e- = FADH2
GSSG + 2H+ + 2e- = 2GSH
S + 2H+ + 2e- = H2S
Lipoic acid + 2H+ + 2e- = Dihydrolipoic acid
NAD+ + H+ + 2e- = NADH
b) Four electron
O2 + 4H+ + 4e- = 2H2O
+80
-180
-240
-270
-320
-320
+820
Table 2.1.1 SEPs For Different Reactions. (Based on Buettner (1993) but
extended by Acworth et al. (1997a); Koppenol (1998) and references therein;
Koppenol and Butler (1985)).
COUPLED REDOX REACTIONS.
The EMF values (E) can be considered as entirely analogous to free energy
changes (∆G). Therefore as long as the number of electrons in the reaction are
known, Eo values can be used to predict the position of equilibrium in a reaction.
In the electron transport chain of the inner mitochondrial membrane, the redox
couples (cytochromes, etc.) are arranged in order of increasing Eo’. As electrons
are passed along the chain the total energy released in the oxidation of NADH is
utilized to synthesize ATP from ADP and inorganic phosphate at several steps
along the chain. If redox couples were not used and the energy of NADH
oxidation was to be released in a single step much of it would be wasted as heat,
with much less energy being available for the synthesis of ATP.
REFERENCES
Acworth, I.N., McCabe, D.R., and Maher, T.J. (1997a). The analysis of free radicals, their reaction products and
antioxidants. In: Oxidants, Antioxidants and Free Radicals. Baskin, S.I., and Salem, H., (Eds.). Taylor and Francis,
Washington DC. Pp. 23-77.
Buettner, G.R. (1993). The pecking order of free radicals and antioxidants: lipid peroxidation, alpha-tocopherol, and
ascorbate. Arch. Biochem. Biophys. 300, 535-543.
Koppenol, W.H. (1998). The basic chemistry of nitrogen monoxide and peroxynitrite. Free Radic. Biol. Med., 25, 385-391.
Koppenol, W.H., and Butler, J. (1985). Energetics of interconversion reactions of oxy radicals. Adv. Free Radic. Biol.
Med., 1, 91-131.
WWW.ESAINC.COM
158
Appendix 2.2
Background to Kinetics
Thermodynamics enables the prediction of whether a process or reaction will
occur spontaneously. It cannot predict the rate at which such processes occur,
for this we have to turn to the field of kinetics. The importance of the difference
between thermodynamic and kinetic control of a reaction is illustrated by the
following example. The hydrolysis of ATP is thermodynamically favorable (∆Go’=
-31 kJ mol-1 at pH 7.0 and 25oC) yet a solution of ATP at pH 7.0 is fairly stable.
This is because an activation energy has to be achieved before ATP hydrolysis
will take place. A catalyst (such as ATPase) will lower the activation energy and
hence speed up (increase the rate of) the reaction, but it will not effect the
position of equilibrium. Kinetic measurements are of great interest to redox
biochemists. For example, the measurement of reaction kinetics allows us to
determine and compare the reactivity of pro-oxidants.
Order
Units
R=k1[A]
First
s-1
R=k2[A]2
Second
M-1s-1
R=k2[A] [B]
Second
M-1s-1
Table 2.2.1 Units of Reaction Order.
The rate of a chemical reaction is dependent upon the concentration of reactants
present, temperature, pressure, pH and the presence of inhibitors. For example,
the reaction rate nearly doubles for every 10oC increase in temperature. The
exact mathematical relationship between the rate of a reaction and the
concentration of reactants is determined experimentally and is called the rate
law. The order is defined as the power to which the concentration of reactant is
raised in the rate law. Thus for the reaction where a moles of A combine with b
moles of B (Eqn 2.2.1) the expression for the rate of formation (d[P]/dt) of product
(P), might be Eqn 2.2.2 where k is a constant known as the rate constant. Once
the reaction is started the concentrations of both A and B will fall and the reaction
rate will fall too. That is why reaction rates are usually measured as soon as the
reaction has started (initial rate measurement). The rate constant is an
experimental quantity and can be either integral or non-integral (Table 2.2.1).
The above reaction is ath order in A, bth order in B, with an overall order of (a+b).
Order is not the same as molecularity. Molecularity is the minimum number of
species involved in the rate-determining step (the slowest step of the reaction).
WWW.ESAINC.COM
159
aA + bB → P
d[P]/dt = k [A]a [B]b
Eqn 2.2.1
Eqn 2.2.2
FIRST-ORDER PROCESSES.
In this process the rate of the reaction depends only upon the reactant (R) Eqn
2.2.3. The rate law for this reaction (Eqn 2.2.4) upon integration yields Eqn 2.2.5
where c is a constant. If [R]0 is the initial concentration of R (t=0) then c=ln[R]0.
Substitution of this expression in Eqn 2.2.6 yields Eqn 2.2.7. A plot of [R] versus
time gives a straight line of slope –k. An example of a first order process is
radioactive decay. The units for first-order rate constants is time (s-1) (Table
2.1.1).
The half-life of a reaction is a very useful quantity. This is the time taken for [R] to
fall to half of its initial value, i.e., the time taken for R to fall from [R]0 to 0.5[R]0 is
the same as the time to go from 0.5[R]0 to 0.25[R]0. Therefore for a first order
reactions, the half-life is independent of initial concentration of R. The expression
of calculation of the half-life (t1/2) can be derived using Eqn 2.2.6 and is shown in
Eqn 2.2.7.
R → products
-d[R]/dt = k [R]
ln [R] = -kt + c
ln [R]/[R]0 = -kt
ln 2 = kt1/2
t1/2 = 0.693/k
Eqn 2.2.3
Eqn 2.2.4
Eqn 2.2.5
Eqn 2.2.6
Eqn 2.2.7
SECOND-ORDER AND PSEUDO-FIRST-ORDER PROCESSES.
Most often reactions are second-order or pseudo-first-order.
In second-order reactions the rate depends upon two molecules reacting (Eqn
2.2.8). The rate law for this process for the condition [A]0≠[B]0 is presented in Eqn
2.2.9. Following integration, a plot of ln[B]0([A]0-x)/[A]0([B]0-x) versus t gives a
straight line of slope k2([A]0-[B]0). If the initial concentrations of A and B are equal
then the rate law becomes Eqn 2.2.10. The half-life is related to the initial
concentration of A by Eqn 2.2.11. Therefore in second-order reactions the halflife is inversely proportional to the initial concentration of A, i.e., it takes twice as
long for the initial concentration of A to go from 0.5[A]0 to 0.25[A]0 as it does for
[A]0 to 0.5[A]0. The units for second-order rate constants are
(concentration-1)(time-1) (e.g., M-1s-1, l mol-1 s-1, M-1h-1 etc.) (Table 2.1.1).
WWW.ESAINC.COM
160
A + B → Products
dx/dt = k2[A] [B]
dx/dt = k2([A]0 – x)2
t1/2 = 1/k2[A]0
Eqn 2.2.8
Eqn 2.2.9
Eqn 2.2.10
Eqn 2.2.11
With pseudo-first order reactions the concentration of one of the reactants (say
B) is in excess and will remain essentially constant throughout the reaction. Thus
the rate law becomes Eqn 2.2.12. The process is apparently first order in A and
zeroth order in B. Pseudo-first order reactions are in units of time (e.g., s-1).
However, true second-order rate constants can be obtained by dividing by the
concentration.
-d[A]/dt=k[A]
Eqn 2.2.12
SOME PUBLISHED SECOND-ORDER RATE CONSTANTS.
Table 2.2.2 shows some rate constants published in literature. Be aware that rate
constants will be affected by the experimental conditions under which they are
obtained so care should be taken when comparing them to each other.
REACTIONS
HO• + General
Metabolite
HO• + Albumin
HO• + Ascorbyl•
HO• + N-Acetylcysteine
HO• + Carnosine
HO• + β-Carotene
HO• + Cysteamine
HO• + Cysteic acid
HO• + Cysteinesulfinic
acid
HO• + Deoxyguanosine
HO• + Deoxyribose
HO• + Desferrioxamine
HO• + DMSO
HO• + “double bond”
HO• + Ethanol
HO• + Glucose
HO• + Glutathione
(reduced)
HO• + Hypotaurine
RATE CONSTANT
(M-1s-1)
109 - 1010
>1010
1.0 x 1010
1.4 x 1010
2.5 x 109
<1.0 x 1011
5.9 x 109
5.3 x 107- 1.6 x 108
3.2 x 109
1.0 x 109
3.1 x 109
1.3 x 1010
6.6-7.1 x 109
REFERENCE
Buxton et al. (1988); Ross
et al. (1997)
Halliwell et al. (1995)
Bartlett et al. (1994)
Aruoma et al. (1989a)
Aruoma et al. (1989b)
Lymar et al. (1995)
Aruoma et al. (1988)
Aruoma et al. (1988)
Aruoma et al. (1988)
≥1010
1.9 x 109
1.0 x 109
8.8 x 109
Vieira et al. (1993)
Halliwell et al. (1995)
Denicola et al. (1995)
Pryor and Squadrito
(1995); Denicola et al.
(1995)
Breen and Murphy (1995)
Pryor and Squadrito (1995)
Halliwell et al. (1995)
Aruoma et al. (1989a)
5 x 109 - 1.2 x 1010
Aruoma et al. (1988)
WWW.ESAINC.COM
161
HO• + Mannitol
HO• + NO• → HNO2
HO• + NO2• → ONO2H
HO• + Phenylalanine
1.7 x 109
1.0 x 1011
4.5 x 109
1.9 x 109
HO• + Phosphate (e.g.
DNA)
HO• + R-H → R• + H2O
HO• + Salicylic Acid
<107
HO• + Taurine
HO• + α-Tocopherol
HO• + Trolox
HO• + Unsaturated fatty
acids
O2•- + N-Acetylcysteine
1.0 x 109
5.0 x 109 - 1.0 x
1010
2.4 x 106 - 1.4 x 107
1.0 x 1010
8.0 x 1010
1.0 x 109
O2•- + Carnosine
O2•- + Catechols
1.0 x 103 - 2.7 x
106
1.0 x 104 - 2.7 x
105
1.0 x 103
1.0 x 109
O2•- + Dihydrolipoic acid
O2•- + Quinones
3.3-7.5 x 105
1.0 x 109
O2•- + Taurine
O2•- + Fe3+ → Fe2+ + O2
2O2•- + H+ → H2O2 + O2
<1 x 103
1.0 x 106
5.0 x 105 - 2.4 x
109
2.4 x 105 - 8.0 x
107
O2•- + Ascorbate
O2•- + HO2• + H+ →
H2O2 + O2
O2•- + HOCl → HO• +
Cl- + O2
O2•- + NO• → ONO2O2•- + NO2• → NO2- +
O2
O2•- + O2•- + 2H+ →
H2O2 + O2
HO2• + α-Tocopherol
HO2• + HO2• → H2O2 +
O2
H2O2 + 2Cysteine →
Cystine + 2H2O
H2O2 + Cu+ → HO•
+ HO- + Cu2+
H2O2 + Fe2+ → HO•
+ HO- + Fe3+
H2O2 + Fe2+-ADP →
HO• + HO- + Fe3+ADP
7.5 x 106
3.4 x 107- 7 x 109
1.0 x 108
<0.3 - 5 x 105
2.5 x 106
8.0 x 105
Pryor and Squadrito (1995)
Rubbo et al. (1995)
Bartlett et al. (1994)
Kaur and Halliwell (1994a);
Kaur and Halliwell
(1994b)
Breen and Murphy (1995)
Breen and Murphy (1995)
Hiller et al. (1983)
Aruoma et al. (1988)
Lymar et al. (1995)
Pryor and Squadrito (1995)
Radi et al. (1991)
Aruoma et al. (1989a)
Halliwell and Gutteridge
(1989); Radi et al. (1991)
Aruoma et al. (1989b)
Halliwell and Gutteridge
(1989)
Suzuki et al. (1991, 1993)
Halliwell and Gutteridge
(1989)
Aruoma et al. (1988)
Radi et al. (1991)
Denicola et al. (1995);
Beckman (1994)
Bielski et al. (1985);
Halliwell and Gutteridge
(1989)
Aruoma et al. (1989a)
Pryor and Squadrito
(1995); Radi et al. (1991)
Alvarez et al. (1995)
Halliwell and Gutteridge
(1989)
Rubbo et al. (1995)
Halliwell and Gutteridge
(1989)
1.3 x 101
Radi et al. (1991)
4.7 x 103
Beckman (1994)
7.6 x 101
Beckman (1994)
8.0 x 102
Beckman (1994)
WWW.ESAINC.COM
162
H2O2 + Fe2+-EDTA →
HO• + HO- + Fe3+EDTA
H2O2 + ONO2H → O2•+ NO2• + H+ + H2O
NO• + Fe2+(heme)
NO• + Fe3+(heme)
NO• + Heme
NO• + HO• → HNO2
NO• + 1/2O2 → NO2•
NO• + O2•- → ONO2NO• + Tryptophan•
NO• + Tyrosine•
NO2• + Fatty acid →
NO2- + Fatty acid• +
H+
NO2• + HO• → ONO2H
NO2• + O2•- → NO2- +
O2
NO2• + Tyrosine• → 3Nitrotyrosine
NO2+ + Tyrosine → 3Nitrotyrosine + H+
ONO2- + SOD → SODCu+-O-NO2+
ONO2- + Albumin
(single thiol)
ONO2- + Cysteine (or
glutathione)
ONO2- + CO2 →
ONO2CO2ONO2- +
myeloperoxidase
ONO2- + horseradish
peroxidase
ONO2- + cytochrome c2+
ONO2- + alcohol
dehydrogenase
ONO2H + Ascorbate
ONO2H + H2O2 → O2•+ NO2• + H+ + H2O
ONO2CO2- + Tyrosine
→ Tyr• + NO2• +
HCO3AscH- + TO• → Asc•- +
TOH
2Asc•- + H+ → AscH- +
dehydro-Asc
CCl3CO2• + ascorbate
→ CCl3CO2- +
ascorbyl•
CCl3CO2• + phenol →
CCl3CO2- + phenol•
CCl3CO2• + polyphenol
5.0 x 103
Beckman (1994)
1.0 x 105
Alvarez et al. (1995)
1 x 107
1 x 102 - 1 x 107
1 x 103 - 1 x 104
1.0 x 1011
3.5 x 106 (M-2.s-1)
3.4 x 107 - 7 x 109
1-2 x 109
1-2 x 109
1.0 x 105
Pryor and Squadrito (1995)
Pryor and Squadrito (1995
Radi et al. (1991b)
Rubbo et al. (1995)
Crow and Beckman (1995)
Pryor and Squadrito
(1995); Radi et al. (1991)
Radi (1996)
Radi (1996)
Radi et al. (1991b)
4.5 x 109
1.0 x 108
Bartlett et al. (1994)
Alvarez et al. (1995)
3.0 x 109
van der Vleit et al. (1995)
1.0 x 10
Ischiropoulos et al. (1992)
1.0 x 105
Beckman (1994)
2.6 x 103
Radi et al. (1991a)
2.0-6.0 x 103
3.0 x 104
Bartlett et al. (1994); Radi
et al. (1991a)
Lymar and Hurst (1995)
2.0 x 107
Floris et al. (1993)
3.0 x 106
Floris et al. (1993)
2.0 x 105
4.0 x 105
Thompson et al. (1995)
Crow et al. (1995)
2.4 x 102
1.0 x 105
Bartlett et al. (1994)
Alvarez et al. (1995)
>2.0 x 105
Lymar et al. (1996)
2.0 x 105
Scarpa et al. (1984)
2.0 x 105
Vieira et al. (1993)
1.3 x 108
Aruoma et al. (1997)
4 x 105 - 9.5 x 107
Aruoma et al. (1997)
8.4 x 105 - 4.0 x
Aruoma et al. (1997)
8
WWW.ESAINC.COM
163
→ CCl3CO2- +
polyphenol•
CCl3CO2• + αTocopherol →
CCl3CO2- + αTocopheroxyl•
CoQH2 + TO• → CoQ•+ TOH
DNA• + GSH → DNA +
GS•
eaq- → H• + HFe3+ + e- → Fe2+
GS• + GS- → GSSG•GS• + O2 → GSO2•
GSO2• → GS• + O2
GSSG•- → GS• + GSGSSG•- + O2 → O2- +
GSSG
L• + O2 → LO2•
LO• (LO2•) + NO•
LO2• + LH → L• + LO2H
2LO2• → Non-radical
products
LO2H + Fe2+ → LO2• +
Fe3+
RCH2• + O2 → RCH2O2•
TO + PUFA-O2• →
PUFA-O2H + TO•
Tyrosine• + Tyrosine• →
Dityrosine
108
4.9 x 108
Aruoma et al. (1997)
2.0 x 105
Mukai et al. (1990)
1.0 x 107-108
Halliwell et al. (1995)
1.6 x 101
1.0 x 106
8.0 x 108
2.0 x 109
2.0 x 109
2.4 x 105
1.6 x 108
Breen and Murphy (1995)
Beckman (1994)
Buettner (1993)
Buettner (1993)
Buettner (1993)
Buettner (1993)
Buettner (1993)
3.0 x 108
1.3 x 109
1-5 x 101
1.0 x 106-7
Buettner (1993)
Rubbo et al. (1996)
Buettner (1993)
Buettner (1993)
1.0 x 103
Beckman (1994)
3.0 x 109
8.0 x 104
Neta et al. (1990)
Buettner (1993)
4.0 x 108
van der Vleit et al. (1995)
Table 2.2.2 Rate Constants for Some Redox Biochemical
Reactions. Modified from Acworth et al. (1997a).
Rate constants are very useful to redox biochemists and can be used to compare
the rates of different chemical reactions. For example both copper and iron can
take part in the generation of hydroxyl free radicals, but which of these metals is
more effective? As presented by Halliwell and Gutteridge, if equal concentrations
of hydrogen peroxide are mixed with equal concentrations of ferrous (Eqn 2.2.13)
or cuprous ions (Eqn 2.2.14), then the initial rate of hydroxyl free radical
production by the copper-based reaction would be greater than the ferrous-based
reaction by a factor of 62 (Halliwell and Gutteridge (1999)). Thus it appears that
copper is a much more effective pro-oxidant than iron under these conditions.
Furthermore they reported that if the hepatic concentrations of hydrogen
peroxide and ferrous iron were mixed then the number of hydroxyl free radicals
produced in one liter in one second would be in excess of 1013 molecules!
WWW.ESAINC.COM
164
H2O2 + Fe2+ → Fe3+ + OH- + HO•
H2O2 + Cu+ → Cu2+ + OH- + HO•
k2 = 7.61 x 102 M-1s-1 Eqn 2.2.13
k2 = 4.70 x 103 M-1s-1 Eqn 2.2.14
MEASUREMENT OF REACTION ORDER AND REACTION RATES.
The reaction order can be determined by comparing the concentrations of
reactants (or products) as a function of time using the integrated rate laws
discussed above. The determination of reaction order can often be simplified by
using the half-times method (this can be found in many physical biochemistry
texts).
As many radical reactions proceed incredibly rapidly (e.g., the second order rate
constant for HO•-based reactions are on the order of 109-10M-1s-1) and are
beyond the standard approaches used by biochemists, two special approaches
have been developed to measure their reaction rates: stopped flow and pulse
radiolysis. These have been dealt with in detail elsewhere (e.g., Halliwell and
Gutteridge (1999); Wardman (1978)).
Stopped flow procedures are usually used when reaction rates are too rapid for
the normal biochemical procedures. In this approach solutions of reactants are
housed in separate syringes connected to a quartz cell. The outlet of the cell is
connected to a third syringe that can only be filled to a predetermined volume
until its plunger is abruptly stopped from moving. The cell is connected to a
measuring device such as an absorbance detector. At the start of the experiment
the flow from the two reaction syringes is initiated and flow rate is controlled to
prevent significant reaction from taking place within the cell. When the plunger of
the third syringe reaches the end of its travel it stops suddenly. The reaction in
the cell then proceeds to completion and the rate is determined from the change
in absorbance.
With pulse radiolysis the compound under investigation is placed in a cell and
subjected to a short pulse of radiation forming radical species (see below). With
the correct conditions a reaction can be followed for microseconds or longer.
Changes in absorbance can then be used to determine rate constants.
REFERENCES
Alvarez, B., Denicola, A., and Radi, R. (1995). Reaction between peroxynitrite and hydrogen peroxide: Formation of
oxygen and slowing of peroxynitrite decomposition. Chem. Res. Toxicol., 8, 859-864.
Aruoma, O.I., Halliwell, B., Hoey, B.M., and Butler, J. (1988). The antioxidant action of taurine, hypotaurine and their
metabolic precursors. Biochem. J., 256, 251-255.
Aruoma, O.I., Halliwell, B., Hoey, B.M., and Butler, J. (1989a). The antioxidant action of n-acetylcysteine: its reaction with
hydrogen peroxide, hydroxyl radical, superoxide and hypochlorous acid. Free Radic. Biol. Med., 6, 593-597.
Aruoma, O.I., Halliwell, B., and Williamson, G. (1997). In vitro methods for characterizing potential pro-oxidant and
antioxidant actions of non-nutritive substances in plant foods. In: Antioxidant Methodology: In Vivo and In Vitro
Concepts. Aruoma, O.I., and Cuppett, S. (Eds.). AOCS Press, Champaign. Pp. 173-204.
WWW.ESAINC.COM
165
Aruoma, O.I., Laughton, M.J., and Halliwell, B. (1989b). Carnosine, homocarnosine and anserine: could they act as
antioxidants in vivo? Biochem. J., 264, 863-869.
Bartlett, D.B., Church, D.F., Bounds, P.L., and Koppenol, W.H. (1994). The kinetics of the oxidation of l-ascorbic acid by
peroxynitrite. Free Radic. Biol. Med., 18, 85-92.
Beckman, J. S. (1994). Peroxynitrite versus hydroxyl radical: the role of nitric oxide in superoxide-dependent cerebral
injury. N.Y. Acad. Sci., 738, 69-75.
Bielski, B.H.J., Cabelli, D.E., Arudi, R.L., and Ross, A.B. (1985). Reactivity of perhydroxyl/superoxide radicals in aqueous
solution. J. Phys. Chem. Ref. Data, 14, 1041-1100.
Breen, A.P., and Murphy, J.A. (1995). Reactions of the oxyl radicals with DNA. Free Radic. Biol. Med., 18, 1033-1077.
Buettner, G.R. (1993). The pecking order of free radicals and antioxidants: lipid peroxidation, alpha-tocopherol, and
ascorbate. Arch. Biochem. Biophys. 300, 535-543.
Buxton, G.V., Greenstock, C.L., Helman, W.P., and Ross, A.B. (1988). Critical review of rate constants for reactions of
.
. hydrated electrons, hydrogen atoms and hydroxyl radical ( OH/ O ) in aqueous solution. J. Phys. Chem. Ref. Data, 17,
513-886.
Crow, J.P., and Beckman, J.S. (1995). Reactions between nitric oxide, superoxide, and peroxynitrite: Footprints of
peroxynitrite in vivo. Adv. Pharmacol., 34, 17-43.
Crow, J.P., Beckman, J.S., and McCord, J.M. (1995). Sensitivity of the essential zinc thiolate moiety of yeast alcohol
dehydrogenase to hypochlorite and peroxynitrite. Biochem., 34, 3544-3552.
Denicola, A., Souza, J.M., Gatti, R.M., Augusto, O., and Radi, R. (1995). Desferrioxamine inhibition of the hydroxyl
radical-like reactivity of peroxynitrite: Role of the hydroxamine groups. Free Radic. Biol. Med., 19, 11-19.
Floris, R., Piresma, S.R., Yang, C., Jones, P., and Wever, R. (1993). Interaction of myeloperoxidase with peroxynitrite – a
comparison to lactoperoxidase, horseradish peroxidase and catalase. Eur. J. Biochem., 215, 765-775.
Halliwell, B. and Gutteridge, J.M.C. (Eds.). (1999). Free Radicals in Biology and Medicine. Oxford: Clarendon Press.
Halliwell, B., Aeschbach, R., Loliger, J., and Aruoma, O.I. (1995). The characterization of antioxidants. Fd. Chem. Toxic.,
7, 601-617.
Hiller, K.D., Hodd, P.L., and Wilson, R.L. (1983). Anti-inflammatory drugs: Protection of a bacterial virus as an in vitro
biological measure of free radical activity. Chem. Biol. Inter., 47, 293-305.
Ischiropoulos, H., Zhu, L., Chen, J., Tsai, M., Martin, J.C., Smith, C.D., and Beckman, J.S. (1992). Peroxynitrite-mediated
tyrosine nitration catalyzed by superoxide dismutase. Arch. Biochem. Biophys., 298, 431-437.
Kaur, H., and Halliwell, B. (1994a). Detection of hydroxyl radicals by aromatic hydroxylation. Meth. Enzymol., 233, 67-82.
Kaur, H., and Halliwell, B. (1994b). Aromatic hydroxylation of phenylalanine as an assay for hydroxyl radicals:
measurement of hydroxyl radical formation from ozone and in blood from premature babies using improved HPLC
methodology. Anal. Biochem., 220, 11-15.
Lymar, S.V., and Hurst, J.K. (1995). Rapid reaction between peroxynitrite ion and carbon dioxide: Implications for
biological activity. J. Am. Chem. Soc., 117, 8867-8868.
Lymar, S.V., Jiang, Q., and Hurst, J.K. (1997). Mechanism of carbon dioxide-catalyzed oxidation of tyrosine by
peroxynitrite. Biochem. In Press.
Mukai, K., Kikuchi, S., and Urano, S. (1990). Stopped-flow kinetic study of the regeneration reaction of tocopheroxyl
radical by reduced ubiquinone-10 in solution. Biochim. Biophys. Acta, 1035, 77-82.
Neta, P., Huie, R.E., and Ross, A.B. (1990). Rate constants for reactions of peroxyl radicals in fluid solutions. J. Phys.
Chem. Ref. Data, 19, 413-513.
Pryor, W.A., and Squadrito, G.L. (1995). The chemistry of peroxynitrite: A product from the reaction of nitric oxide with
superoxide. Am. J. Physiol., 268, L699-L722.
Radi, R. (1996). Reactions of nitric oxide with metalloproteins. Chem. Res. Toxicol., 9, 828-835.
Radi, R., Beckman, J.S., Bush, K.M., and Freeman, B.A. (1991a). Peroxynitrite oxidation of sulfhydryls. J. Biol. Chem.,
266, 4244-4250.
Radi, R., Beckman, J.S., Bush, K.M., and Freeman, B.A. (1991b). Peroxynitrite-induced membrane lipid peroxidation: The
cytotoxic potential of superoxide and nitric oxide. Arch. Biochem. Biophys., 288, 481-487.
Rubbo, H., Darley-Usmar, V., and Freeman, B. (1996). Nitric oxide regulation of tissue free radical injury. Chem. Res.
Toxicol., 9, 809-820.
Rubbo, H., Radi, R., Trujillo, M., Telleri, R., Kalynararaman, B., Barnes, S., Kirk, M., and Freeman, B.A. (1995). Nitric
oxide regulation of superoxide and peroxynitrite-dependent lipid peroxidation: Formation of novel nitrogen-containing
oxidized lipid derivatives. J. Biol. Chem., 269, 26066-26075.
Scarpa, M., Rigo, A., Maiorino, M., Ursini, F., and Gregolin, C. (1984). Formation of α-tocopherol radical and recycling of
α-tocopherol by ascorbate during peroxidation of phosphatidylcholine. Biochim. Biophys. Acta, 801, 215-219.
Suzuki, Y.J., Tsuchiya, M., and Packer, L. (1991). Thioctic acid and dihydrolipoic acid are novel antioxidants which
interact with reactive oxygen species. Free Radic. Biol. Med., 15, 255-263.
Suzuki, Y.J., Tsuchiya, M., and Packer, L. (1993). Antioxidant activities of dihydrolipoic acid and its structural analogs.
Free Radic. Res. Comms., 18, 115-122.
Thomson, L., Trojillo, M., Telleri, R., and Radi, R. (1995). Kinetics of cytochrome c2 – oxidation by peroxynitrite:
Implications for superoxide measurements in nitric oxide producing biological systems. Arch. Biochem. Biophys., 319,
491-497.
van der Vleit, A., Eiserich, J.P., O’Neill, C.A., Halliwell, B., and Cross, C.E. (1995). Tyrosine modification by reactive
nitrogen species. Arch. Biochem. Biophys., 319, 341-349.
Vieira, A.J.S.C., Candeias, L.P., and Steenken, S. (1993). Hydroxyl radical induced damage to the purine bases of DNA:
In vivo studies. J. Chim. Phys., 90, 881-897.
Wardman, P. (1978). Application of pulse radiolysis methods to study the reactions and structure of biomolecules. Rep.
Prog. Phys., 41, 259-265.
WWW.ESAINC.COM
166
Appendix 2.3
Background to The White Blood Cell
The purpose of the white blood cells (leukocytes) is to defend the body. As
discussed in Chapter 3 the ability for some leukocytes to produce a variety of
ROS, RNS and RHS is essential for their phagocytic activity. However, not all
leukocytes are phagocytes and not all require pro-oxidants for their biological
action. Here we briefly explore the different types of leukocytes found in the
body.
Reticulum Cell
Myeloblast
Lymphocytic
Reticulum Cell
Monoblast
Promyelocyte
Lymphoblast
Monocyte
Eosinophil
Basophil
Neutrophil
Lymphocyte
Figure 2.3.1 The Relationship Between Granulocytes, Lymphocytes And
Monocytes And Their Development From A Common Stem Cell Located In
The Bone Marrow, The Reticulum Cell.
WWW.ESAINC.COM
167
There are three different types of leukocyte – the granulocytes, lymphocytes and
monocytes – differing in their morphology, abundance and biological function.
Granulocytes can be further subdivided into neutrophils, eosinophils and
basophils. As shown in Figure 2.1.1 all lymphocytes form from a common stem
cell, the reticulum cell.
Granulocytes
These cells are the first to be mobilized to deal with injury and infection. They are
amoeboid in nature, phagocytic and are involved in controlling allergic reactions.
They are referred to as granulocytes because their cytoplasm contains numerous
small granules.
•
•
•
Basophils typically make up ~0.5% of the total leukocyte population and
have an abundance of 0-160/µL blood. They resemble polymorphonuclear
neutrophils but contain an irregular nucleus. They are called basophils as
they readily absorb basic stains. Basophils contain heparin (or an heparinlike substance), histamine and 5HT. These cells are involved in
hypersensitivity reactions.
Eosinophils typically make up 1-4% of the total leukocyte population and
have an abundance of 50-250/µL blood. Morphologically they resemble
the neutrophil but contain larger cytoplasmic granules. These cells readily
stain with eosin. They are phagocytic but less active than neutrophils.
Eosinophils carry about 1/3rd of the total amount of blood histamine. These
cells are involved in allergic and hypersensitivity responses (e.g., asthma).
Interestingly these cells posses a special myeloperoxidase capable of
promoting bromination of a variety of substrates.
Neutrophils make up ~50-70% of the total leukocyte population and have
an abundance of 1500-6000/µL blood. They are typically 12µm in
diameter and contain an irregular (polynuclear) nucleus. Neutrophils are
stained by neutral dyes. Neutrophils readily phagocytize bacteria and
other pathogens.
Lymphocytes
These cells play a major role in the immune response. They are mononuclear,
possessing a typical spherical nucleus. These cells are typically either 6-10µm
(small) or 12-15µm (large) in diameter. They are present in the blood at 15004000/µL. The B-lymphocytes are derived from the bone marrow and are antibody
producers. The T-lymphocytes are derived from the thymus and act on
pathogens either through direct contact or through the production of
lymphokines.
WWW.ESAINC.COM
168
Monocytes
This cell-type is the precursor to the macrophage. The monocyte contains a
lobulated nucleus, is typically 14-20µm in diameter and is present in the blood at
100-1000/µL. At the later stages of inflammation monocytes leave the circulation
and enter the damaged tissue. Here they differentiate forming macrophages.
These are typically 20-40µm in diameter and do not contain a lobed nucleus.
Macrophages are metabolically more mobile and active than monocytes and
readily phagocytize bacteria, dead cells and other insoluble material.
Macrophages can also be found in the lymphatic system and are also
permanently located in a variety of tissues (e.g., the Kupffer cells of the liver).
WWW.ESAINC.COM
169
Chapter 3
Damage And Repair
The pro-oxidants can react with many molecules present in the cell. Rather than
exploring all such reactions this chapter will focus on the damage that prooxidants cause to DNA, proteins, lipids and carbohydrates. Because these
molecules are important to cellular homeostasis, a variety of mechanisms have
evolved to protect them and detect any that are damaged. Affected molecules
can then be repaired, or removed and eliminated. The failure to mend or remove
damaged species is associated with a variety of diseases.
WWW.ESAINC.COM
170
DNA.
Introduction.
DNA (deoxyribonucleic acid) is the genetic code of life. The faithful translation of
DNA into messenger RNA (mRNA) and transcription of the latter into proteins is
essential for normal growth, development and reproduction. It is thus imperative
that the integrity and fidelity of a DNA molecule be maintained. Progressive
accumulation of oxidative damage to DNA is thought to be central to the
development of a number of diseases including cancer and neurodegenerative
diseases, as well as aging (Ames (1989); Loft and Poulsen (1996); Wiseman and
Halliwell (1996)). Weinberg (1989) noted that mutation of two types of critical
genes, the proto-oncogenes1 and tumor suppressor genes, occurred during the
development of human cancer. Often, a single base-pair substitution is all that is
required to activate proto-oncogenes or to inhibit tumor suppressor genes,
ultimately leading to cancer. To better understand how ROS/RNS might play a
role in causing such base modifications and other types of DNA damage, we
must first examine the structure of DNA.
The DNA Molecule.
The structure, synthesis and biological properties of the DNA molecule have
been the topic of numerous articles and textbooks (e.g., Stryer (1988)) so will not
be dealt with in depth here. DNA is a long, fragile, ribbon-like molecule. It is a
polymer composed of deoxynucleotide monomers, each of which contains a
base, a sugar and a phosphate group (Figure 3.1). The sugar-phosphate groups
are linked together to form a structural backbone, while the bases carry genetic
information. DNA utilizes four bases that contain either a purine (adenine and
guanine) or pyrimidine (cytosine and thymine) (Table 3.1). The sugar in DNA is
2’-deoxyribose.2 A nucleoside consists of a base bonded to a sugar: the C-1
atom of 2’-deoxyribose forms a β (the base lies above the plane of the sugar
ring) N-glycosidic bond with either the N-1 of pyrimidines or N-9 of purines. A
nucleotide is a phosphate ester (typically at the 5’ position) of a nucleoside.
mRNA is similar in structure to DNA but contains a ribose-phosphate backbone
and uracil instead of thymine.
In 1953 James Watson and Francis Crick showed that DNA is composed of a
double helix, the two polymeric chains being held together by hydrogen bonding
between specific base pairs located within the helix: adenine pairs with thymine
while cytosine pairs with guanine (Figure 3.2). Such specific purine-pyrimidine
1
A proto-oncogene can be present in the human genome. It may have a role in the regulation of normal cell growth and
proliferation. Involvement of these genes in a neoplastic process results from a somatic mutation that converts them into
oncogenic alleles.
2
A primed number donates an atom of the sugar molecule whereas an unprimed number refers to an atom of the base.
WWW.ESAINC.COM
171
pairing is also the consequence of steric factors imposed by the regular helical
nature of the sugar phosphate backbone of each polynucleotide chain.
NH2
NH2
BASE
R
N 7
1 N
3N
NH
9
N
3
NH
1
Pyrimidine
(Cytosine, R=H)
O
Purine
(Adenine)
NH2
NH2
N
NUCLEOSIDE
HO
N
N
N
N
HO
O
N
O
O
OH H
OH H
2'-Deoxyadenosine
2'-Deoxycytidine
NH2
N
NUCLEOTIDE
O
O
NH2
P
O
N
N
N
O
N
O
O
O
OH H
2'-Deoxyadenosine
5'-Monophosphate
(dAMP)
P
O O
O
4'
N
O
1'
OH H
2'-Deoxycytidine
5'-Monophosphate
(dCMP)
Figure 3.1 Examples Of Bases, Nucleosides And Nucleotides.
(The arrows show hot spots for oxidative damage).
WWW.ESAINC.COM
172
Base
Type
Abbreviation
Nucleoside
Adenine
Purine
A
Adenosine
Cytosine
Pyrimidine
C
Cytidine
Guanine
Purine
G(ua)
Guanosine
Thymine
Uracil
Pyrimidine
Pyrimidine
T
U
Thymidine
Uridine
DNA or
RNA
DNA,
RNA
DNA,
RNA
DNA,
RNA
DNA
RNA
Complementary
base
T
G(ua)
C
A
A
Table 3.1 Summary Of Bases, Nucleosides And Watson-Crick
Complementary Base Pair.
H2N
O
NH
N
N
N
N
R
O
Thymine
NH2
N
Adenine
O
N
N
HN
N
N
R
R
N
O
Cytosine
R
H2N
Guanine
Figure 3.2 Specific Base Pairing In The DNA
Molecule.
WWW.ESAINC.COM
173
The DNA molecule can come in a variety of sizes and shapes. For example, in
prokaryotic organisms such as E. coli the DNA occurs as a continuous (circular)
double stranded molecule. In this bacterium the DNA molecule consists of
4000kb (kilobases) and is 1360µm long and 2nm wide. The axis of the DNA helix
can also be tightly twisted into a superhelix. This is not only important in
packaging of the DNA molecule (supercoiling leads to a more compact shape)
but also affects the ability of the double helix to unwind and interact with other
molecules. The longest DNA molecule is found in the eukaryotic organism
Drosophila melanogaster. In this organism the linear DNA molecule consists of
62,000kb and is 2.1cm long. Eukaryotic DNA is structurally much more complex
than its prokaryotic counterpart. Unlike prokaryotic DNA, eukaryotic DNA is not
naked but is tightly bound to a group of small basic proteins called histones. This
nucleoprotein complex is called chromatin. Chromatin fibers consist of repeating
units composed of DNA/histone complexes termed nucleosomes and these are
chained together by “linker” DNA strands much like beads on a string. It has
been proposed that nucleosomes are further packed into a solenoidal structure
consisting of six nucleosomes per turn of the helix (Finch and Klug (1976)). The
purpose of chromatin is to store DNA in a more condensed and manageable
form. An inadvertent benefit of such storage is some degree of protection from
pro-oxidant attack.
NH2
N
N
O
NH2
N
cytosine
O
N
O
O
P
O
NH2
O
N
N
ROS
N
O
N
O
NH2
O
adenine
OH
N
O
O
O
O
cytosine
O
OH
P
HO
N
N
8-hydroxy
adenine
N
O
O
O
NH
O
O
P
OH
N
O
O
O
N
O
O
P
O
O
O
O
O
thymine
N
O
O
O
O
N
O
O
N
OH
P
N
O
O
O
Strand
Break
O
O
thymine
dimer
NH
HO
NH2
guanine
O
HN
OH
NH
OH
O
P
OH
P
N
HO
OH
NH
N
N
NH2
O
O
O
O
O
O
8-hydroxy
guanine
OH
P
O
Figure 3.3 Examples Of ROS Induced Damage To DNA. For The Sake Of Clarity
Only One Of The DNA Strands Is Shown.
WWW.ESAINC.COM
174
DNA Damage.
DNA is a fairly reactive molecule and is readily attacked by ionizing radiation,
oxidizing agents and electrophiles (Figures 3.1 and 3.3). Oxidative damage to
DNA includes covalent modification of bases (DNA adducts), production of alkali
labile sites and strand breaks, either formed directly or as a consequence of
repair processes (Aruoma and Halliwell (1998); Breen and Murphy (1995);
Dizdaroglu (1991, 1994); von Sonntag (1991)) (Table 3.2). It must be
remembered however that the vast majority of studies in this field have been
conducted in vitro and may not always be applicable to circumstances in vivo.
Damaging Species
Alkylating Agents
Hydrogen Peroxide
Hydroxyl Free
Radical
Hypohalous Acids
Consequences
Exogenous (e.g., dimethylhydrazine and Nmethyl-N-nitrosourea) and endogenous (Sadenosylmethionine) alkylating agents can yield a
variety of alkylated adducts. Depending upon the
alkylating agent a number of adducts can be
formed including 7-methylguanine, 7ethylguanine, 7-(2-hydroxyethyl) guanine, 3methylguanine, O-4 methylthymine, O-4ethylthymine and 0-6 methylguanine.
No direct effect on bases but acts as a precursor
for hydroxyl free radical production if encountering
redox-active metals.
Can cause strand scission, and protein-DNA
cross-linking, and form a complex variety of DNA
adducts and sugar-derived products (Figure 4.4,
4.5). DNA adducts include 8-hydroxy2’deoxyguanosine (8OH2’dG), 5-hydroxy2’deoxycytidine, 8-hydroxy-2’deoxyadenine,
thymine glycol isomers, and 2,6-diamino-4hydroxy-5-formamidopyridine. 8-hydroxy2’deoxyadenine may undergo further oxidation to
guanidinohydantoin.
HOCl rapidly attacks pyrimidines forming thymine
glycol isomers, 5-hydroxyuracil, 5-hydroxycytosine
5-hydroxyhydantoin, 5-chlorouracil and 5-chloro2’d-cytidine but see Whiteman et al., (2002).
Purines (bases and nucleosides) form 8-chloroand 8-bromo derivatives.
References
Cushnir et al. (1993);
Kang et al. (1992); Netto
et al. (1992); van Delft et
al. (1993); Tan et al.
(1990)
Dizdaroglu (1993);
Halliwell and Aruoma
(1991)
Breen and Murphy
(1995); Cadet et al.
(1999); Dizdaroglu
(1991); Duarte et al.
(1999); Henle et al.
(1996); Lloyd and Philips
(1999); Luo et al. (1996);
Raoul et al. (1995); von
Sontag (1991), Wagner
et al. (1992)
Birnbaum et al., (1984);
Kozumbo et al. (1992);
Masuda et al., (2001);
Shen et al., (2001);
Whiteman et al. (1997,
1999a)
Oxidation of some bases is promoted by nitrite
due to formation of nitryl chloride.
Can activate aryl xenobiotics that are capable of
causing single strand breaks.
WWW.ESAINC.COM
175
Irradiation
Direct – absorbed energy forms DNA radical
cations that can further react with water or
undergo deprotonation reactions with bases and
2’-deoxyribose forming base-centered radicals
and sugar-derived radicals, respectively.
Breen and Murphy
(1995); Dizdaroglu
(1991); Menon (1995);
von Sontag (1991)
Indirect – solvated electrons do not cause strand
breakage but can form a variety of base radicals.
Hydroxyl free radicals cause considerable
damage (see below).
More than thirteen adducts are reported to be
formed when DNA is irradiated.
Lipid Peroxidation
Products
Can also lead to the formation of DNA
hydroperoxides.
Carbonyls – 4-hydroxnonenal, malondialdehyde
and other alkyl aldehydes can form DNA adducts
(e.g., 1-N2-propano-2’deoxyguanosine;
deoxycytidine glyoxal).
Hydroperoxides can cause single strand and
double strand DNA breaks and produce a variety
of DNA adducts.
Miscellaneous
Carbonyls – acrolein and crotonaldehyde form
1,N2-propanodeoxyguanosine adducts.
Catecholestrogens can form a semiquinone
(through redox cycling) capable of forming
catecholestrogen-adducts with guanine.
Superoxide produced when catecholestrogens
recycle can react with nitric oxide to produce
peroxynitrite that causes DNA strand breakage.
Agarwal et al. (1994);
Baker and He (1991);
Cooke et al., (2003);
Douki and Ames (1994);
Esterbauer et al. (1990);
Goda and Marnet
(1991); Li et al. (1995a);
Marnet (1985; 1996);
Vaca et al. (1995); Yang
and Schaich (1996);
Wang and Liehr (1991;
1995)
Eder and Hoffman
(1993); Feinstein et al.
(1993); Gebicki and
Gebicki (1999); Liehr
(1990; 1994); Mobley et
al. (1999); (Nath et al.
(1998); Spencer et al.
(1994); Yoshie and
Ohshima (1998)
Protein hydroperoxides cause DNA-protein cross
linking.
Nitric Oxide, Nitrogen
Dioxide,
Peroxynitrite, other
RNS
The antitumor antibiotic adriamycin can also redox
cycle. Once bound to DNA it can ultimately lead to
the production of hydroxyl free radicals causing
degradation of double and single stranded DNA.
L-DOPA, dopamine and 3-O-methyl-DOPA in the
presence of copper can promote oxidative
damage to DNA.
Can cause RNS-dependent nitration, nitrosation
and deamination of bases. Nitrogen dioxide
promotes single strand breaks. Peroxynitrite can
directly oxidize and nitrate DNA bases producing
8-oxoguanine, 8-oxoguanosine, 8-oxoadenine, 8oxoadenosine, 8-nitroguanine, 2- and 8-nitro
adenine (and adenosine), and 8-nitroxanthine.
Peroxynitrite can also cause single strand breaks
WWW.ESAINC.COM
Bartsch and Frank
(1996); Bittrich et al.
(1993); Bermudez et al.
(1999); Burney et al.
(1999); Byun et al.
(1999); Dawson and
Dawson (1994);
Gorsdorf et al. (1990);
176
and promote lipid peroxidation, thereby producing
cytotoxic carbonyl compounds capable of forming
DNA-carbonyl adducts. Peroxynitrite causes
mutation in bacteria and human cells.
Ozone
Singlet Oxygen
Superoxide Radical
Anion
Causes single-strand breaks. Forms DNA adducts
(e.g., hydroxymethyluracil, thymine glycol,
8-hydroxyguanine, 8OH2’dG and other
unidentified adducts). Can induce lipid
peroxidation forming cytotoxic carbonyl
compounds (e.g., malondialdehyde, 4hydroxynonenal) capable of damaging DNA.
Selectively attacks guanine. Guanosine produces
diastereomers of 4,8-dihydro-4-hydroxy-8-oxo2’deoxyguanosine while DNA primarily forms
8OH2’dG. Photochemical ROS generating
systems induce the expression of several
eukaryotic genes including adhesion molecules,
early response genes, immunomodulatory
cytokines, matrix metalloproteins, and stress
proteins.
No direct effect on bases but can act to promote
hydroxyl free radical production through the
Haber-Weiss reaction.
Epe et al. (1996); Harris
(1995); Juedes and
Wogan (1996); Oshima
and Bartsch (1994);
Routledge et al. (1994);
Sodum and Fiala (2001);
Spencer et al. (1996);
Szabo (1996); Yermilov
et al. (1995, 1996)
Bermudez et al. (1999);
Cajigas et al. (1994);
Calderon-Garciduenas
et al. (1997); Ferng et
al. (1997); Foksinski et
al. (1999)
Epe (1993); McCabe et
al. (1997); Ravanat and
Cadet (1995); Ryter and
Tyrrell (1998) and
references therein; Van
den Akker et al. (1994)
Dizdaroglu (1993);
Halliwell and Aruoma
(1991)
Table 3.2 Oxidative Damage To DNA, Free Nucleosides and Bases.
Ionizing radiation is arguably the most routinely used approach to promote DNA
damage. It can be either absorbed directly by DNA leading to ionization of the
bases (direct effect) or react with surrounding water molecules first (indirect
effect) producing pro-oxidant species (e.g., hydroxyl free radical, hydrated
electrons [e-aq], and hydrogen atoms). In the presence of oxygen it appears that
most DNA damage is due to the production of the hydroxyl free radical.
The hydroxyl free radical, whether produced through the homolytic effects of
ionizing irradiation on water or from the Fenton reaction, adds at diffusion limited
rates to double bonds (k>1010 M-1s-1), and efficiently abstracts hydrogen atoms
from organic molecules (k=109 M-1s-1). It is much less reactive with phosphate
groups of the DNA backbone (k<107 M-1s-1). Thus, hydroxyl free radical-induced
damage to DNA is primarily a consequence of the addition to the π-bonds of
bases or hydrogen abstraction from deoxyribose3 and the methyl group of
thymine. Interestingly under normal physiological conditions neither superoxide
3
A number of modified sugar moieties are formed including both free (e.g., 2,5-dideoxypentos-4-ulose,
2,3-dideoxypentos-4-ulose, 2-deoxypentose-4-ulose and 2-deoxytetrodialose) and DNA-bound forms (e.g.,
2-deoxypentanoic acid and erythrose) (Dizdaroglu (1991, 1998)).
WWW.ESAINC.COM
177
nor hydrogen peroxide can react directly with deoxyribose or bases (Aruoma et
al. (1989); Brawn and Fridovich (1981); Lesko et al. (1980); Rowley and Halliwell
(1983)). However, the conversion of these ROS to hydroxyl free radicals or other
pro-oxidant species will result in DNA damage (Halliwell and Gutteridge (1988);
Nassi-Calo et al. (1989)). It is interesting to note that, of all the radicals found in
aerobic cells, only the hydroxyl free radical readily attacks DNA. Pryor (1988) has
hypothesized that the hydroxyl free radical is unique in that it has a rare
combination of high electrophilicity, high thermokinetic reactivity, and a
mechanism of production close to the DNA molecule.
NH2
O
O
H
N
CH3
O
NH
OH
O
H
N
O
N
O
O
HN
8OHGua
OCH3
NH
3MGua
O
O
N
N
NH
O
NH2
NH2
OH
OH
O
NH H
CH3
N
O
NH
CO2H
7MGua
H2N
TT-DIMER
NH
U
C
N
O
O
NH
5OHCt
HN
H2N
HN
N
NH
O6MGua
O
HN
OH
N
NH
N
H2N
O
NH2
NH2
N
NH2
N
FAPyAd
8OHAd
HN
NHCHO
NH
N
CH3
N
H2N
NH2
N
OH
NH
N
HN
OH
NH
2-OHAd
N
N
5OHMU
O
N
N
O
NH
N
NH2
CH2OH
N
CG
O
H2N
H
HO
A
O
H
OH
OH
NH H
TG
NH
N
HN
NH
N
8NitroG
NH2
CH3
OH
OH
H
NH
N
5HH
NH2
N
HN
NO2
H2N
OH
NH
5OH5MH
HN
HN
H
O
N
5-OH-6-HC
CO2H
CT DIMER
H
OH
H
NH H
H2N
O
NH
O
O
H2N
HN
H
HN
O
O
OH
CH3
HN
OH
H
NH H
N
HN
H
H
NH H
O
5-OH-6-HT
O
O
HN
O
H
OH
H
NH H
HN
O
H
OH
OH
NH H
O
HN
O
UG
5-OH-6-HU
H
H
NH H
5,6-DHU
HN
NH
5,6-DOHU
O
NH2
N
OH
O
NH2
FAPyGua
O
H
N
H2N
G
5,6-DHT
NHCHO
HN
NH
N
H2N
H2N
TL DIMER
O
O
O
CH3
OH
O
OH
OH
NH
CH3
HN
O
5,6-DHC
NH
T
Figure 3.4 A Selection Of Modified Bases (from Acworth et al. (1997); Dizdaroglu
(1991); Dizdaroglu et al. (1993); Yermilov et al. (1995, 1996)). 2-OHAd – 2-hydroxyadenine; 3-MGua –
3-methylguanine;
7-MGua
–
7-methylguanine;
6
O -MGua
–
O-6-methylguanine;
WWW.ESAINC.COM
5-HH
–
5-hydroxyhydantoin;
178
5-OHC – 5-hydroxycytosine; 5-OH-6-HC – 5-hydroxy-6-hydrocytosine; 5-OH-6-HT – 5-hydroxy-6-hydrothymine; 5-OH-6-HU
– 5-hydroxy-6-hydrouracil; 5-OH5MH – 5-hydroxy-5-methylhydantoin; 5-OHMU – 5-(hydroxy-methyl) uracil; 5,6-DHC – 5,6dihydroxycytosine; 5,6-DHT – 5,6-dihydrothymine; 5,6-DHU – 5,6-dihydrouracil; 5,6-DOHU – 5,6-dihydroxyuracil; 8-NitroGua
– 8-nitroguanine; 8-OHA – 8-hydroxyadenine; 8-OHGua – 8-hydroxyguanine; A – adenine; C – cytosine; CG(ua) – cytosine
glycol; CT dimer – cytosine-tyrosine dimer; FAPyAd – 4,6-diamino-5-formamido-pyrimidine; FAPyGua – 2,6-diamino-4hydroxy-5-formamido-pyrimidine; G – guanine; T – thymine; TG(ua) – thymine glycol; TL dimer – thymine-lysine dimer; TT
dimer – thymine-tyrosine dimer; U- uracil; UG(ua) – uracil glycol. Structures shown in red are electrochemically active and
can be measured by HPLC with electrochemical detection.
O
O
H
CH2OH
N
O
O
N
HN
N
N
H2N
N
R
OH
O
N
R
5-OH2'dUd
H2N
N
R
5-OH2'dCd
N
N
N
CH3
R
N
R
5,6-DiOH2'dUd
OH
N
R
NH2
NH2
OH
OH
H
R
Td-glycol
N
N
H2N
H
FAPyG
OH
H
O
NH
N
H2N
O
H
N
HN
8,5'Cyclo-2'dAdo
O
N
OH
N
O
OH
N
OH
N
O
N
N
O
8,5'Cyclo-2'dGuo
N
H
N
O
H
OH
N
N
R
8-OH2'dAdo
Figure 3.5 A Selection Of Modified DNA Nucleosides
Dizdaroglu
H
7-M2'dGuo
N
OH
N
O
N
NH2
HN
H
O
HN
O
H2N
R
O6-M2'dGuo
O
OH
N
N
N
H2N
3-M2'dGuo
NH2
N
N
HN
R
8-OH2'dG
O
O
N
N
H2N
R
5-HOMdUd
H
N
HN
OH
OCH3
CH3
N
NH
R
FAPyA
(From Acworth et al. (1997);
(1994)).
2’Td-glycol – Thymidine glycol isomers; 3-M2’dGuo – 3-methyl-2’deoxyguanosine;
5-OH2’dCd
–
5-hydroxy-2’deoxycytidine;
5-OH2’dUd
–
5-hydroxy-2’dexoyuridine;
5-OHM2’dUd
–
5-(hydroxymethyl)-2’deoxyuridine; 5,6-DiOH2’dUd – 5,6-dihydroxy2’deoxyuridine; 7-M2’dGuo – 7-methyl-2’deoxyguanosine;
8-OH2’dAdo – 8-hydroxy-2’deoxyadenosine; 8-OH2’dG – 8-hydroxy-2’deoxyguanosine; 8,5’Cyclo-2’dAdo – 8,5’-cyclo2’adenosine; 8,5’,Cyclo-2’dGuo – 8,5’-cyclo-2’deoxyguanosine (5’R- and 5’S-); FAPyA – 4-amino-5-formylamino-6-(2’6
deoxyribosyl)-aminopyrimidine; FAPyG – 2-amino-4-hydroxy-5-formylamino-6-(2’deoxyribosyl)-aminopyrimidine; O 6
M2’dGuo – O -methyl-2’deoxyguanosine; R – 2’deoxy-ribose. Structures shown in red are electrochemically active and can
be measured by HPLC with electrochemical detection. Those in blue should be electrochemically active.
WWW.ESAINC.COM
179
Early pulse-radiolytic studies suggest that about 20% of hydroxyl free radicals
produced in close proximity to DNA damage deoxyribose molecules (Breen and
Murphy (1995) and references therein). If such damage is not repaired (see
below) it can lead to cleavage of the sugar-phosphate backbone resulting in
single-strand breaks (SSB) in the DNA molecule.456 SSBs are not usually lethal
as the complementary DNA strand serves as a template, holding the severed
chain in place long enough for repair enzymes to take action. Indeed treatment of
DNA with hydroxyl free radicals derived from hydrogen peroxide resulted in SSBs
without significant cell death (Ward et al. (1985)). Double-strand breaks (DSB)
are the common result of excessive irradiation. If the DSB are located too close
to each other, this can result in severance of the DNA molecule, permanent
damage and cell death. Readers interested in learning more about the possible
mechanisms for hydroxyl radical-induced hydrogen atom abstraction and sugarphosphate cleavage are referred to Breen and Murphy (1995) and von Sonntag
(1991).
By far the most intense research in this field has been directed towards the
chemistry and biology of DNA adduct formation. Attack of the free bases and
nucleosides by pro-oxidants can yield a wide variety of adducts and DNA-protein
cross-links (Figures 3.4 and 3.5). Such attack usually occurs at the C-4 and C-8
position of purines and C-5 and C-6 of pyrimidines (Breen and Murphy (1995)). A
summary of the reaction products of the interaction between the hydroxyl free
radical and the four nitrogen bases of DNA is presented in Table 3.3.
Hydroxyl free radical-induced damage to purine bases and nucleosides can
proceed through a C-8-hydroxy N-7 radical intermediate (Table 3.3). This radical
can either undergo oxidation with the production of an 8-hydroxy purine, or
reduction, probably by cellular thiols, followed by ring opening and the formation
of FAPy (formamido-pyrimidine) metabolites (see Figure 3.6 for hydroxyl free
radical-induced damage to guanosine). Although most research has focused on
8-hydroxy-purine adducts a growing number of publications are attempting to
measure the FAPy derivative (see below).
While many potential base lesions can be formed when DNA is irradiated, the
relative abundance of such lesions varies considerably. Table 3.4 presents the
relative abundance of base lesions when either naked DS-DNA or chromatin is
exposed to irradiation. Under aqueous oxygenated conditions irradiation is
synonymous to hydroxyl free radical damage. Note that the relative abundance of
4
A clastogen is an agent capable of causing breakage of chromosomes. Superoxide may indirectly induce genotoxicity by
the formation of long-lived clastogenic factors (CFs). Emerit (1993) hypothesized that CFs induce the production of
superoxide radical anions by competent cells which in turn release more CFs thereby promoting clastogenesis. This selfsustaining process may exceed the DNA repair system and ultimately lead to cancer. Increased cancer risk and CF
formation are found in irradiated people, those with chronic inflammation, workers exposed to asbestos, HIV-infected
people and those with Fanconi’s anemia.
5
DNA strand breaks can be measured using single-cell gel electrophoresis (SCGE, comet assay) (Collins (1998);
Fairbairn et al. (1995); McKelvey-Martin et al. (1993); Piperakis et al. (1998)).
6
It has been estimated that 36,000 SSB and >40 DSB occur in each cell/day (Bernstein (1998)).
WWW.ESAINC.COM
180
base lesions is much less than for naked DS-DNA, suggesting that the
association of histones in chromatin confers some degree of protection.7 A
second observation is that 8-hydroxyguanine is the most dominant lesion in both
dsDNA and chromatin.
Precursor
Cytosine
Thymine
Adenine
Intermediate
C-6-hydroxy C-5 radical
Product
5-Hydroxyhydantoin
C-5-hydroxy C-6 radical
5,6-Dihydroxyuracil
5-Hydroxyuracil
5-Hydroxy-5,6-dihydrouracil
Deoxyribose radical
C-6-hydroxy C-5 radical
Pyrimidine-derived cyclonucleoside
6-Hydroxy-5,6-dihydrothymine
C-5-hydroxy C-6 radical
Thymine glycol isomers
N-Tartonoylurea
Methylene radical
5-Hydroxymethyldeoxyuracil
5-Formyldeoxyuracil
Deoxyribose radical
C-4-hydroxy C-5 radical
Pyrimidine-derived cyclonucleoside
Hydrolyzed and reduced to adenine
C-8-hydroxy C-4 radical
Oxidized to 8-hydroxyadenine
Reduced to FAPy-adenine
C-5’ deoxyribose radical
Guanine
C-4-hydroxy C-5 radical
Purine-derived cyclonucleoside (8.5’-cyclo2’deoxyadenosine)
Dehydrated and reduced to guanine
C-5-hydroxy C-4 radical
Dehydrated and reduced to guanine
C-8-hydroxy N-7 radical
Oxidized to 8-hydroxyguanine
Reduced to FAPy-guanine
C-5’ deoxyribose radical
Purine-derived cyclonucleoside (8.5’-cyclo2’deoxyguanosine)
Table 3.3 Oxidative Damage To DNA Bases By The Hydroxyl Free Radical
Under Oxygenated Conditions.
DNA adducts can also be formed between bases and carbonyl compounds
produced as a result of lipid peroxidation (see below). 4-Hydroxynonenal and
7
An alternative explanation is that DNA-protein cross-links can be formed that lead to decreased yields of “protein-free”
modified bases (Gajewski et al. (1990)).
WWW.ESAINC.COM
181
malondialdehyde are two extremely reactive carbonyl compounds that readily
form adducts with DNA (and RNA) bases (Figure 3.7). These compounds are
also capable of forming DNA-DNA and DNA-protein (most commonly with lysine)
cross-links.8 The chemistries of carbonyl compounds are discussed in more
detail below. Adducts can also be formed between bases and catecholestrogens
(Chapter 4).
O
N
HN
N
N
H2N
R
Guanosine
HO
O
N
HN
H2N
OH
H
N
N
8-OH2'dGuo
O
R
O
OH
+H+
Reduction
(Thiols?)
+e-
H2N
N
H
N
N
N
H2N
N
HN
OH
N
HN
R
R
+H+
+e-
Reduction
(Thiols?)
O
N
HN
O
H
N
HN
H
O
H2N
N
N
OH
H
O
R
H2N
N
N
-e-H+
Oxidation
H
N
HN
R
H2N
FAPyG
OH
N
N
R
8-OH2'dG
R = deoxyribose
Figure 3.6 Damage To 2’Deoxyguanosine By The Hydroxyl Free
Radical Produces Both 8-Hydroxy-2’deoxyguanosine (8OH2’dG)
And FAPyG. Based on Breen and Murphy (1995).
8
It has been estimated that ~37 (rat) and ~3 (human) DNA-cross links are formed in each cell/day (Bernstien (1998)).
WWW.ESAINC.COM
182
Base
C
T
A
G
Product
Naked DNA
(Radiation Yields
nmolJ-1)
Cytosine glycol
5,6-Dihydroxycytosine
5-Hydroxyhydantoin
Thymine glycol
5-Hydroxymethyluracil
5-Hydroxy-5methylhydantoin
FAPyAd
8-Hydroxyadenine
FAPyGua
8-Hydroxyguanine
25.6
43.4
-
Chromatin
(Radiation
Yields
nmolJ-1)
2.3
0.3
0.2
0.4
0.05
0.4
5.9
15.8
3.6
46.7
1.0
3.5
1.8
8.1
Table 3.4 Abundance (Radiation Yields) Of DNA Base Adducts
Formed Under Oxygenated Conditions As Determined Using GC-MS.
(From Fuciarelli et al. (1990); Gajewski et al. (1990) and see Breen and Murphy (1995)).
Nitric oxide damages DNA by two major pathways (Bartsch and Frank (1996);
Bittrich et al. (1993); Dawson and Dawson (1994); Gorsdorf et al. (1990); Harris
(1995); Oshima and Bartsch (1994); Routledge et al. (1994); Spencer et al.
(1996); Szabo (1996); Yermilov et al. (1995, 1996)). The first route involves the
nitrosation of the amines of the nucleic acid bases (Chapter 2). Primary aromatic
amines produce deaminated products, while secondary amines form N-nitroso
compounds. The second pathway involves the formation of peroxynitrite from
nitric oxide. Peroxynitrite shows complex reactivity with DNA initiating DNA
strand breakage, oxidation (e.g., formation of 8-hydroxyguanine, 8-OH2’dG, (5hydroxymethyl)-uracil, and FAPyGua), nitration (e.g., 8-nitroguanine), and
deamination of bases (Table 3.2). Peroxynitrite can also promote the production
of lipid peroxidation related active carbonyls and cause the activation of NAD+
ADP-ribosyltransferase (Szabo (1996) and references therein).
WWW.ESAINC.COM
183
O
O
H
N
N
O
O
H
N
N
HN
N
N
R
O
O
N
H
εdGuo
N
N
O
R
N
M2Guo
O
N
N
R
N
N
N
N
N
N
N
O
N
O
R
M3Cd
N
N
EdGuo
M1dGuo
OH
O
N
N
HO
NH
O
N
N
N
O
N
NH
O
H
H
H
N
N
AdGuo 1,2
N
N
N
AdGuo 3,4
N
N
H 3C
NH
H 3C
NH
N
R
M1dAdo
N
M1dCd
N
N
R
R
CdGuo 1
O
R
O
N
N
N
N
N
N
OH
O
OH
H
N
R
R
N
CdGuo 2
O
H
O
N
N
H
H
N
N
O
NH
N
N
N
N
H
O
N
O
R
HN
N
Cd-Guo-X
N
N
N
R
N
εdCd
N
N
εdAdo
R
N
N
R
M3dAdo
Figure 3.7 Some DNA Base-Carbonyl Adducts. (Douki and Ames (1994); Eder and
Hoffman (1993); Esterbauer at al. (1991) and references therein; Goda and Marnett (1991).
AdGuo 1,2 and AdGuo 3,4 – 1,N2-propano-2’deoxyguanosine isomers derived from acrolein;
CdGuo 1 and CdGuo 2 - 1,N2-propano-deoxyguanosine isomers derived from crotonaldehyde;
Cd-Guo-X – malondialdehyde cross-link of cytidine and guanosine; εdAdo – 4-hydroxynonenaldeoxyadenosine adduct; εdCd – 4-hydroxynonenal-deoxy cytidine adduct; εdGuo –
4-hydroxynonenal-deoxyguanosine adduct; EdGuo – 1,N2-ethenodeoxy-guanosine; M1dAdo –
(mono)-malondialdehyde-deoxyadenosine adduct; M1dGuo – (mono)-malondialdehydepyrimidopurinone adduct; M2Guo – (di)-malondialdehyde-deoxyguanosine adduct; M3Cd – (tri)malondialdehyde-deoxycytidine adduct; M3Ado – (tri)-malondialdehyde-deoxyadenosine adducts.
WWW.ESAINC.COM
184
NH
H2N
O
H2N
N
N
N
N
NH
N
N
O
Syn 8-Oxoguanine Adduct
Adenine Base
Figure 3.8 Mispairing Of 8-Oxoguanine (8-Hydroxyguanine) With Adenine
To Give G→T Transversion Mutations. Note That The Guanine Adduct Is
Rotated On Its N-Glycosidic Bond Into The Syn Position.
The Consequences of Oxidative DNA Damage.
There is now considerable experimental evidence reporting the mutagenicity of
ROS in prokaryotic (e.g., bacteriophage and plasmid DNA) and eukaryotic cells
(De Flora et al. (1989); Hsei et al. (1986); Meneghini (1988); Moody and Hussan
(1982); Retel et al. (1993)). Although all DNA bases can be oxidatively damaged,
it is the modification of guanine that is the most frequent (Table 3.4). 8OH2’dG is
the most abundant DNA adduct. This adduct was first discovered by Kasai and
has recently been reviewed by him (Kasai (1997)). 8OH2’dG can exist in several
tautomeric forms (e.g., 8-oxo-2’deoxyguanosine or 7,8-dihydro-8-oxo2’deoxyguanosine) that can affect its hydrogen bonding between base-pairs
(Culp et al. (1989)). These base-pair substitutions are usually found clustered
into areas called “hot spots”. As shown in Figure 3.2, guanine normally binds to
cytosine. 8OH2’dG, however, can form hydrogen bonds with adenine (Figure
3.8). The formation of 8OH2’dG in DNA can therefore result in a G→T
transversion.9 Upon replication this mutation has been found to occur with a
frequency of 0.5-1.0% in bacteria and 2.5-4.8% in simian kidney cells (Moriya
(1993); Wood et al. (1990)). 8-Hydroxyguanine was also shown to induce codon
12 activation of c-Ha-ras and K-ras in mammalian systems (Loft and Poulsen
(1996) and references therein). Altered codon 12 activity has been postulated to
9
A transversion is the replacement of a purine with a pyrimidine or a pyrimidine with a purine. A transition is the
replacement of one purine with another purine, or one pyrimidine with another pyrimidine.
WWW.ESAINC.COM
185
contribute to the initiation and progression of human cancer (Capella et al.
(1991); Seeburg et al. (1984); Yanez et al. (1987)). G→T transversions are also
the most frequent hot spot mutations found in the p53 supressor gene which is
associated with human tumors (Harris and Holstein (1993); Hollstein et al.
(1991)). Singlet oxygen and 1,2-dioxetanes preferentially induce 8OH2’dG
formation causing GC to AT base-pair substitutions (Loft and Poulsen (1996) and
references therein).
Apurinic sites10 are mispaired mainly with adenine. As is the case with 8OH2’dG ,
the frequency of such mispairing is dependent upon the adjacent base sequence
and which polymerase is present (Loft and Poulsen (1996) and references
therein). Adenine adducts (e.g., 8-hydroxy-2’deoxyadenosine) do not seem to
mispair or lead to mutations (Shibutani (1993)). Oxidation products of cytosine
(e.g., 5-hydroxycytosine, 5-hydroxyuracil) show sequence-context-dependent
mispairing in vitro resulting in C→T transitions and C→G transversions (Breen
and Murphy (1995) and references therein; Purmal et al. (1994)). Thymine glycol
was shown to cause a low frequency T→C transition in a SS-DNA bacteriophage
M13 (Basu et al. (1989)). However, this adduct was without effect on DS-DNA
from another phage, probably due to the greater ability of its endonuclease III to
repair thymine glycol lesions. Although not studied in great detail, another
thymine adduct, 5-(hydroxymethyl)-2’deoxyuridine, may be worth investigating
further. It appears to miscode during DNA replication and induces a high
incidence of mutations in Salmonella typhimurium (Patel et al. (1992); ShirnameMore et al. (1987)).
Thymine glycol, FAPyAdo, FAPyGuo and 8-hydroxyadenine appear not to induce
base-pair modifications but rather act by blocking DNA and RNA polymerases
and are thought to be lethal (Evans et al. (1993); Ide et al. (1994); Maccabee et
al. (1994); O’Connor et al. (1998)).
Several other mechanism by which ROS/RNS can lead to mutations have been
proposed. Direct mechanisms include: conformational changes in the DNA
template that reduces the accuracy of replication by DNA polymerases (Feig and
Loeb (1993); Wiseman and Halliwell (1996) and references therein), and altered
methylation of cytosine that affects gene control (Weitzman et al. (1994)). Indirect
mechanisms include oxidative damage to proteins, including DNA polymerases
and repair enzymes. Damage to lipids causes the production of mutagenic
carbonyl compounds (Wiseman and Halliwell (1996) and references therein).
Finally, Halliwell and Wiseman (1996) suggested that ROS/RNS might be
involved in misalignment mutagenesis (“slippery DNA”) which is associated with
10
Apurinic and apyrimidinic (AP) sites result from hydrolysis of a weakened N-glycosyl linkage, a consequence of adduct
formation. Such hydrolysis is also catalyzed by N-glycosylases as part of the excision-repair process (see below) and by
treatment with heat, acid, alkylating and nitrating agents (Loeb and Preston (1986); Yermilov et al. (1995)). AP sites not
only direct the incorporation of adenine opposite the lesion but can also inhibit DNA synthesis. Fortunately, such lesions
can be bypassed by DNA polymerase. It has been estimated that ~7000 depurination events occur in each cell per day
(Bernstein (1998)).
WWW.ESAINC.COM
186
cancer and several genetic diseases such as myotonic dystrophy and
Huntington’s disease (Kunkel (1993)).
Repair of ROS/RNS-induced DNA Damage.
The repair of damaged DNA is an ongoing and continuous process involving a
number of repair enzymes (Table 3.5). Damaged DNA appears to be mended by
two major mechanisms, base excision repair (BER) and nucleotide excision
repair (NER) (Croteau and Bohr (1997); Singer and Hang (1998) and references
therein). However, as isolated DNA is found to contain low levels of damaged
bases, it would appear that these repair processes are not completely effective.
Although sugar damage and double strand breaks are also critical lesions they
have been extensively reviewed so will not be dealt with here (Lieber et al.
(1996); Povirk (1996); Weaver (1996)).
ENZYME
DNA Alkyltransferase
DNA Glycosylase
ACTION
Removes alkyl group from affected base. For example,
O6-alkylguanine alkyltransferase repairs O6-methyl2’deoxyguanosine.
Simple glycosylase - removes damaged base by cleaving the Nglycosidic bond forming an apurinic (or apyrimidinic) [AP] base.
Phosphodiester bonds either side of the AP site are cleaved by
endonucleases allowing insertion of intact nucleotide.
Some glycosylases hydrolyze the N-glycosylic bond and possess
lyase activity that cleaves resulting AP site producing a 3’ terminal
and an unsaturated aldehyde that must be removed by
endonuclease IV.
DNA AP Endonuclease
Some examples of glycosylases include:
Uracil-DNA glycosylase (recognizing uracil)
Hydroxymethyluracil-DNA glycosylase (recognizing
hydroxymethyluracil)
Hypoxanthine-DNA glycosylase (recognizing hypoxanthine)
Pyrimidine hydrate-DNA glycosylase (recognizing thymine glycols,
pyrimidine hydrates, ring-fragmented pyrimidines, urea)
Formamidopyrimidine-DNA glycosylase (recognizing purines with a
fragmented imidazole ring; 8-hydroxyguanine)
3-Methyladenine-DNA glycosylase (recognizing purines methylated
at N3 or N7; pyrimidines at O2 position)
8-Oxoguanine glycosylase/lyase (recognizing 8-oxoguanine) –
OGG1 and MutM
Pyrimidine dimer-DNA glycosylase (recognizing cyclobutane
pyrimidine dimers).
Recognizes an AP site and nicks the phosphodiester bonds of DNA
strand. Damaged DNA is then removed.
Some enzymes such as endonuclease III possess both glycosylase
and endonuclease activity for repair of oxidized pyrimidines.
Endonuclease IV removes the 3’ blocking damage resulting from the
action of Class 1 endonucleases.
WWW.ESAINC.COM
187
DNA Exonuclease
DNA Helicases
DNA Ligase
DNA Polymerases
NAD+ ADPRibosyltransferase
Degrades DNA by cleaving successive nucleotide residues (or short
oligonucleotides) from either the 5’ or 3’ ends. For example,
exonuclease III cleaves 5’ to the AP site.
Unwinds DNA to facilitate separation of the two strands of the
duplex.
Joins newly synthesized DNA to the rest of the strand.
Synthesizes DNA strand (fills the gaps left by exonucleases and
endonucleases).
Also called poly(ADP)ribose synthetase (PARS) and
poly(ADP)ribose polymerase (PARP). This enzyme responds to
DNA strand breaks by promoting the poly-ADP ribosylation of
nuclear proteins. The alteration of the DNA-histone structure
enhances the activity of DNA ligase for repairing strand breaks.
Over-activity of the ribosyltransferase can be deleterious, resulting in
rapid consumption of NAD+ and the depletion of ATP, eventually
leading to cell death.
Table 3.5 Some DNA Repair Enzymes. (See Demple and Harrison (1994)).
Base Excision Repair.
BER is first started by DNA glycosylases which recognize specific base
modifications (e.g., 8OH2’dG) (Figure 3.9). For example, formamido-pyrimidineDNA glycosylase (Fpg protein) recognizes damaged purines such as 8oxoguanine and FAPyGua (and to a lesser extent adenine derivatives) (Boiteux
et al. (1992)). Damaged pyrimidines are recognized by endonuclease III, which
acts as both a glycosylase and AP endonuclease (Bohr et al. (1995); Croteau
and Bohr (1997)). Glycosylases cleave the N-glycosylic bond between the
damaged base and the sugar. There are two classes of glycosylases: simple
glycosylases (that only hydrolyze N-glycosylic bonds between inappropriate base
and deoxyribose sugar to release the base and produce an unmodified AP site)
and glycosylase/AP lyase enzymes (that also cleave the resulting AP sites at the
DNA-phosphate backbone). Following the glycosylase step, AP endonucleases
then remove the 3'-deoxyribose moiety by cleavage of the phosphodiester bonds
thereby generating a 3’-hydroxyl group that can then be extended by DNA
polymerase. The final step in mending damaged DNA is the rejoining of the free
ends of DNA by a DNA ligase (Seeberg et al. (1995)).
It appears that the presence of 8-oxoguanine modified bases in DNA is not only
a result of ROS attack on this macromolecule (Figure 3.9). Oxidized nucleosides
and nucleotides from free cellular pools can also be incorporated into DNA by
polymerases and cause AT to CG base substitution mutations (Kamiya et al.
(1992); Shibutani et al. (1991)). To prevent this bacteria contain both BER and
error avoidance mechanisms (Grollman and Moriya (1993); Mo et al. (1992);
Taddei et al. (1997)). Three enzymes are involved in error avoidance including
an 8-oxoguanine glycosylase/AP lyase (MutM or Fpg protein), adenine DNA
glycosylase (MutY) and 8-hydroxy-2’deoxyGTPase (Fowler et al., (2003)). The
MutY enzyme removes adenine that has been misincorporated opposite 8-
WWW.ESAINC.COM
188
oxoguanine in DNA while the MuT enzyme prevents incorporation of damaged
GTP into DNA (Michaels et al. (1992); Tajiri et al. (1995)). Functional homologs
of these proteins are also found in higher eukaryotes (e.g., OGG1 corresponds
to MutM; MTH1 to MutT; MYH to MutY) (Boiteux and Radicalla (1999); Croteau
and Bohr (1997); Furuichi et al. (1994); Hayakawa et al. (1995); Hazra et al.,
2001); Ide (2001); Nashimua (2002); Sekumi et al. (1993); Wani and
D’Ambrosia (1995)).
O
N
HN
N
H 2N
NH
Guanine
ROS
O
(Acid Hydrolysis)
N
HN
H 2N
Glycosylase
Diet
OH
N
NH
8-Hydroxy-guanine
OH
ROS
Transcription
ROS
OH
OH
mRNA
Modified
DNA
DNA
Modified
mRNA
(Enzymatic
Hydrolysis)
O
Endonuclease
O
N
HN
H 2N
Polymerase
N
N
HO
N
HN
OH
OH
H 2N
N
N
HO
O
OH
8-Hydroxy-2'deoxyguanosine
O
OH
OH
8-Hydroxy-guanosine
Mut-T(MTH1)
dGMP-OH
dGTP-OH
ROS
dGTP
Figure 3.9 Oxidative Damage Can Occur To DNA, mRNA, And The
Free Nucleotide Pool. This figure uses guanine as an example and illustrates
base damage and repair. Also shown in parenthesis are the in vitro methods used to
liberate bases or nucleosides for analysis.
WWW.ESAINC.COM
189
Nucleotide Excision Repair.
Bacteria and eukaryotic cells supplement BER with NER mechanisms (Lindahl
(1993); Satoh et al. (1993); Van Houten (1990)). NER involves two components,
a global repair element and a transcription coupled repair mechanism (Croteau
and Bohr (1997)). This explains why transcription coupled repair occurs more
rapidly in transcriptionally active DNA rather than in the genome as a whole, a
consequence of faster repair of lesions in the transcribed strand (Gowen et al.
(1998) and references therein; Friedberg (1996)). The integration of DNA repair
and transcription is the responsibility of the basal transcription factor TFIIH, which
contains at least two DNA repair genes (Friedberg (1996)). To date very little is
known about the role of NER in DNA repair.
Mitochondrial DNA Repair.
The mitochondrial genome consists of a circular DNA molecule (16,569 bp)
encoding 13 polypeptides, 2 ribosomal RNA and 22 transfer RNA molecules, and
each mitochondrion contains 10 copies (Wallace (1992)). The mitochondrion
genome encodes the various complexes of the electron transport chain, but
contains no genetic information for DNA repair enzymes. These enzymes must
be obtained from the nucleus. As mitochondria are continuously producing DNAdamaging pro-oxidant species, effective DNA repair mechanisms must exist
within the mitochondrial matrix in order for these organelles to function. However,
these mitochondrial repair mechanisms do appear to operate slower than in the
nucleus (Yakes and Van Houten (1997)). Fortunately mitochondria are not longlived and its been estimated that one mitochondrion turns over per cell every 15
minutes. Thus excessively damaged mitochondria will be quickly removed.
The early finding that mitochondria could not repair UV-induced pyrimidine
dimers led to the erroneous conclusion that they lack DNA repair enzymes
(Clayton et al. (1974)). Indeed accumulating evidence now suggests that
mitochondria contain many BER enzymes and are proficient at repair (Croteau
and Bohr (1997); Bohr et al., (2002)). However, mitochondria do not appear to
repair damaged DNA by NER mechanisms (Croteau and Bohr (1997); Van
Houten (1998)).
Single Strand DNA Damage and PARP Activation.
Single strand DNA breakage activates NAD+ ADP-ribosyltransferase (PARP;
Table 4.5). PARP is a protein-modifying, nucleotide-polymerizing enzyme and is
found at high levels in the nucleus (Ueta and Hayashi (1985)). Activated PARP
first cleaves NAD+ into ADP-ribose and nicotinamide, and then attaches the
ADP-ribose units to a variety of nuclear proteins including histones and its own
automodification domain. PARP then polymerizes the initial ADP-ribose
WWW.ESAINC.COM
190
modification with other ADP-ribose units to form the nucleic acid-like polymer,
poly (ADP) ribose. PARP only appears to be involved with BER and not NER. In
BER PARP does not appear to play a direct role but rather it probably helps by
keeping the chromatin in a conformation that enables other repair enzymes to be
effective (Wiseman and Halliwell (1996) and references therein). It may also
provide temporary protection to DNA molecules while it is being repaired.
Interestingly, recent conflicting evidence suggests that PARP may not be an
important DNA repair enzyme as cells from a PARP knockout mouse model have
normal repair characteristics (Wang et al. (1995)). Other possible physiological
roles for PARP include slowing of cellular metabolism as an adaptive response to
environmental conditions, regulation of gene expression and cell differentiation,
regulation of histone shuttling, and nucleosome unfolding. PARP is also involved
in the expression of the major histocompatibility complex class II gene, ras, DNA
methyltransferase gene, protein kinase C and i-NOS (Szabo (1996) and
references therein).
Activation of PARP can be dangerous to the cell. For each mole of ADP-ribose
transferred, one mole of NAD+ is consumed, and through the regeneration of
NAD+ four ATP molecules are wasted (Dawson and Dawson (1994)). Thus the
activation of PARP can rapidly deplete a cell’s energy store and even lead to cell
death. Some researchers suggest that this may be one mechanism whereby
cells with excessive DNA damage are effectively removed. However, a variety of
diseases may involve PARP overactivation including circulatory shock, CNS
injury, diabetes, drug-induced cytotoxicity, and inflammation (Szabo (1996) and
references therein).
What Do The Levels Of DNA Adducts Mean?
The measurement of oxidative base damage as an indicator of oxidative stress
can fall into two broad categories – steady state (tissue concentration) and total
levels (rate of excretion) (Poulsen and Loft (1998)). What category is measured
is dependent upon the question that is being asked. Remember that a tissue
steady-state level will not represent the total adduct formation of the whole
organism. Conversely, the rate of excretion will not be due to just one tissue, but
may be greatly or primarily influenced by one. Only direct experimental evidence
will allow tissue contribution to be determined.
Steady State Levels.
The steady state DNA adduct level reflects the balance between damage, repair,
dilution of unrepaired adducts during DNA replication as cells divide and
incorporation of adduct nucleotides during DNA replication. Many different DNA
adducts are currently being measured (Table 3.6) but 8OH2’dG is by far the most
common (Collins et al. (1996)). Both nuclear and/or mitochondrial DNA levels of
WWW.ESAINC.COM
191
8OH2’dG have been shown to be increased in a variety of diseases including
colorectal cancer, coronary heart disease, diabetes, inflammation,
neurodegeneration, and following irradiation (Beal (1997) and references therein;
Collins et al. (1998a,b); Dandona et al. (1996); Shimoda et al. (1994); Wilson et
al. (1993)). It should come as no surprise that the oxidative damage to
mitochondrial DNA appears to be greater than nuclear DNA, a consequence of
this organelle’s ability to generate ROS. This finding is complicated however by
methodological issues (below and see Beckman and Ames (1999), and Suter
and Richter (1999)).
Lesion
2-Hydroxyadenine
Species
Human
Tissue/
Cellular
Source Of
DNA
Brain region:
Temporal lobe
Frontal lobe
5-Hydroxy
cytosine
Human
Brain region:
Temporal lobe
Frontal lobe
5-Hydroxy2’deoxycytidine
5-Hydroxy2’deoxycytidine
5-Hydroxy2’deoxycytidine
Calf
Thymus
Human
Leukocyte
Rat
5-Hydroxy2’deoxyuridine
5-Hydroxy2’deoxyuridine
5-Hydroxy2’deoxyuridine
Calf
Liver
Kidney
Brain
Thymus
Human
Leukocyte
Rat
5-Hydroxymethyluracil
5-Hydroxymethyluracil
Human
Liver
Kidney
Brain
Leukocyte
Human
Lung
5-Hydroxyuracil
Calf
5,6 Dihydroxy-
Calf
Level/Range
Method
Reference
GC-MS
Lyras et al.
(1997)
GC-MS
Lyras et al.
(1997)
HPLC-UV or
HPLC-ECD
HPLC-UV or
HPLC-ECD
HPLC-UV or
HPLC-ECD
Wagner et al.
(1992)
Wagner et al.
(1992)
Wagner et al.
(1992)
HPLC-UV or
HPLC-ECD
HPLC-UV or
HPLC-ECD
HPLC-UV or
HPLC-ECD
Wagner et al.
(1992)
Wagner et al.
(1992)
Wagner et al.
(1992)
GC-MS
Djuric et al.
(1991)
Jaruga et al.
(1994)
0.13 nmol/mg DNA
control
0.14 – Alzheimer’s
0.02 nmol/mg DNA
control
0.04 – Alzheimer’s
0.67 nmol/mg DNA
control
0.5 – Alzheimer’s
0.12 nmol/mg DNA
control
0.12 – Alzheimer’s
10+2.5 fmol/µg DNA
Thymus
3.2+1.6 fmol/µg
DNA
10+3.5 fmol/µg DNA
9.9+4.4
22.6+3.4
10+4.0 to 75+0.25
fmol/µg DNA
2.1+1.8 fmol/µg
DNA
<0.5 fmol/µg DNA
<0.5
<0.5
93+19 adducts/105
pb
4-15 adducts/105 pb
normal
5-19 – cancer
0.5 nmol/mg DNA
HPLC-ECD
Thymus
10+4.0 fmol/µg DNA
HPLC-UV or
WWW.ESAINC.COM
GC-MS-SIM
Kaur and
Halliwell
(1996)
Wagner et al.
192
dihydro2’deoxyuridine
5,6 Dihydroxydihydro2’deoxyuridine
5,6 Dihydroxydihydro2’deoxyuridine
HPLC-ECD
(1992)
Human
Leukocyte
6.2+4.6 fmol/µg
DNA
HPLC-UV or
HPLC-ECD
Wagner et al.
(1992)
Rat
Liver
HPLC-UV or
HPLC-ECD
Wagner et al.
(1992)
8-Hydroxyadenine
Calf
Kidney
Brain
Thymus
8.5+3.5 fmol/µg
DNA
10.3+4.0
14.6+4.5
0.8 nmol/mg DNA
HPLC-ECD
8-Hydroxyadenine
8OH2’dG
Human
Lymphocyte
C. elegans
Cell culture
8OH2’dG
Hamster
Kidney
Kaur and
Halliwell
(1996)
Podmore et al.
(1998)
Bogdanov et
al. (1999)
Han and Liehr
(1994)
8OH2’dG
Human
Food/
Beverages
8OH2’dG
Human
Liver
Mixed diet
3000kcal/day
Tea
Coffee
Blood – red
blood cells
Kidney
dialysate
Plasma
Saliva
Sweat
Blood mononuclear
cells
8OH2’dG
Human
8OH2’dG
Human
CSF
8OH2’dG
Human
CSF
8OH2’dG
Human
CSF
8OH2’dG
Human
Placenta
8OH2’dG
Human
Brain
8OH2’dG
Human
Brain region
0.05+0.025 nmol/mg
DNA
11.24+5.36 pg/mL
GC-MS
3.6+1.2 adducts/105
pb
10.4+1.8
1.09+0.56 nmol/g
HPLC-ECD
49 pmol/L
39 pmol/L
2.1+0.31 pg/mL –
control
67.34+20.31
13.4+2.11
15.3+3.36
11.2+9.5
15.3 to 73.5 fmol/µg
DNA control
96 to 223 – IDDM1
64 to 134 – MIDDM2
64.3+20 ng/mL –
control
25.1+12 Alzheimer’s
0.98+0.03 pg/mL
1.5+0.2 pg/mL –
control
2.1+0.2 – ALS
1.2+0.2 – other
neurological
disorders
0.2 to 1.0
adducts/105 pb
1.3 to 7.8
4.0+0.8 pmol/µg
DNA – control
21.0+7.0 –
Alzheimer’s
1.25+0.13 to 2.7+0.6
adducts/105 pb –
nuclear
WWW.ESAINC.COM
HPLC-ECD
HPLC-ECD
Bogdanov et
al. (1997)
HPLC-ECD
Bogdanov et
al. (1999)
HPLC-ECD
Dandona et al.
(1996)
GC-MS
Lovell et al.
(1999).
HPLC-ECD
Bogdanov et
al. (1999)
Bogdanov et
al. (2000)
HPLC-ECD
HPLC-ECD
ELISA
Yin et al.
(1995)
GC-MS
Lovell et al.
(1999)
HPLC-ECD
Mecocci et al.
(1993)
193
15.78+8.3 to
27.34+14.7 –
mitochondrial
5.1+3.3 adducts/105
pb – normal
3.6+1.9 – cancer
tissue
3.6+1.1 adducts/105
pb – normal
5.6+2.3 – cancer
tissue
3.9+0.26
adducts/105 pb
0.33+0.08
adducts/105 pb –
control
0.51+0.25 –
smokers
5.9+3.8 adducts/105
pb – control
7.1+4.3 – smokers
1.0+0.2 adducts/105
pb
8 adducts/105 pb
control
16 to 112 –
irradiated subjects
1.5+0.2 adducts/105
pb control
3.3+1 – smokers
4.5 to 13.4
adducts/105 pb –
control
3.2 to 6.0 – cirrhotic
tissue
1.6+0.7 adducts/105
pb – normal
3.2+2.1 – inflamed
tissue
25-75 adducts/105
pb – normal
50-200 – cancer
tissue
2.2 to 3.5
adducts/105 pb
0.58 to 0.90
adducts/105 pb male
0.33 to 0.51 female
8OH2’dG
Human
Breast tissue
8OH2’dG
Human
Kidney
8OH2’dG
Human
Leukocyte
8OH2’dG
Human
Leukocyte
8OH2’dG
Human
Leukocyte
8OH2’dG
Human
Leukocyte
8OH2’dG
Human
Leukocyte
8OH2’dG
Human
Leukocyte
8OH2’dG
Human
Liver
8OH2’dG
Human
Liver
8OH2’dG
Human
Lung
8OH2’dG
Human
Lymphocyte
8OH2’dG
Human
Lymphocyte
8OH2’dG
Human
Mononuclear
leukocyte
1.16+0.4
adducts/105 pb
Polymorphonuclear
Leukocyte
1.13+0.4
adducts/105 pb
WWW.ESAINC.COM
HPLC-ECD
Nagashima et
al. (1995)
HPLC-ECD
Okamoto et al.
(1994)
HPLC-ECD
Degan et al.
(1995)
Kiyosawa et
al. (1990)
HPLC-ECD
HPLC-ECD
Lagorio et al.
(1994)
HPLC-ECD
Hanaoka et al.
(1993)
Wilson et al.
(1993)
TLC-32P,
HPLC-ECD
HPLC-ECD
Lodovici et al.,
(2000)
HPLC-ECD
Carmichael et
al. (1995)
HPLC-ECD
Shimoda et al.
(1994)
GC-MS-SIM
Jaruga et al.
(1994)
HPLC-ECD
Bashir et al.
(1993)
Collins et al.
(1998)
HPLC-ECD
HPLC-ECD
Takeuchi et al.
(1994)
194
8OH2’dG
Human
Plasma
8OH2’dG
Human
Plasma cryoprecipitates
12.9+0.7 – control
17.7+1.2 – ALS
17.8+1.2 – other
neurological
disorders
0.37+0.04 to
2.27+0.06 pmol/µg
DNA
0.61+0.05 to
1.89+0.07
10 to 100
adducts/105 pb –
control
30-180 –
hyperplastic tissue
200ng/mL
HPLC-ECD
Bogdanov et
al. (2000)
HPLC-ECD
Lunec et al.
(1994)
GC-MS
8OH2’dG
Human
Prostate
8OH2’dG
Human
Serum
8OH2’dG
Human
Sperm
8OH2’dG
Human
Uterine Tumor
8OH2’dG
Mouse
Keratinocyte
8OH2’dG
Mouse
8OH2’dG
Mouse
Liver –
maternal
Embryo
Liver
8OH2’dG
Mousehairless
Skin cells
8OH2’dG
Rat
0.32+0.009 pg/mL
0.7+0.1 pg/g
HPLC-ECD
HPLC-UV
GC-MS
HPLC-ECD
GC-MS-SIM
Olinski et al.
(1995)
ELISA
Cooke et al.
(1998)
Fraga et al.
(1991)
Foksinski et
al. (2000)
2.1+3.2 adducts/105
pb
0.81 adducts/105 bp
– control
1.24 – tumor
0.56 – lymphocytes
1.4 adducts/106 pb
HPLC-ECD
27+8 fmol/µg DNA
HPLC-ECD
~8+6
0.8 to 1.8
adducts/104 pb
4.5+0.38
adducts/105 pb
8OH2’dG
Rat
Brain
microdialysate
Muscle
microdialysate
Feces
8OH2’dG
Rat
Heart
8OH2’dG
Rat
Calf
Liver
Thymus
8OH2’dG
Rat
Liver
8OH2’dG
Rat
Liver
8OH2’dG
Rat
Liver
8OH2’dG
Rat
Liver
1.00+0.1 nmol/mg
DNA – control
1.50+0.1 –
ischemia/reperfusion
1.74+0.6
adducts/105 pb
5.97+1.4
2.87+0.48
adducts/105 pb
0.96+0.37
adducts/105 pb
2.3 to 2.7
adducts/105 pb
20 adducts/106 pb
8OH2’dG
Rat
Various
8 to 73 adducts/106
HPLC-ECD
HPLC-ECD
HPLC-ECD
HPLC-ECD
HPLC-ECD
Beehler et al.
(1992)
Liu and Wells
(1995)
Faux et al.
(1992)
HattoriNakakuki et al.
(1994)
Bogdanov et
al. (1999)
0.19+0.008
WWW.ESAINC.COM
Bogdanov et
al. (1999)
Cordis et al.
(1998)
HPLC-ECD
Adachi et al.
(1995)
HPLC-ECD
Denda et al.
(1994)
Adachi et al.
(1993)
Cattley and
Glover (1993)
Teixeira et al.
(1995)
Fraga et al.
HPLC-ECD
HPLC-ECD
195
Rat
organs
Kidney
8OH2’dG/
8-hydroxyguanosine
Human
Liver
Brain
CSF
8OH2’dG/
8-hydroxyguanosine
Human
Serum
8-hydroxy-2’
deoxyguanosine-5’monophosphate
8-Hydroxyguanine
Human
A549 cells
Human
Brain –
substantia
nigra
8-Hydroxyguanine
8-Hydroxyguanine
Human
Lymphocyte
Rat
Liver
8-Hydroxyguanosine
(RNA)
Human
Liver – RNA
Brain region:
Temporal lobe
8OH2’dG
Frontal lobe
8-Hydroxyguanosine
(RNA)
Human
Plasma
8-Hydroxyguanosine
(RNA)
8-Hydroxyguanosine
(RNA)
8-Hydroxyguanosine
(RNA)
8-Hydroxyguanosine
(RNA)
8-Hydroxyguanosine
(RNA)
Human
Plasma
Human
Serum
Human
CSF
Human
CSF
Human
Brain region:
Temporal lobe
pb
15 to 35 fmol/µg
DNA
8 to 23
6 to 15
1.46+0.83 ng/mL –
control
2.85+2.43 –
Parkinoson’s
35.7+15 ng/mL –
control
57.5+20.4 –
Parkinoson’s
0.43+0.06
adducts/106
2.3 nmol/mg DNA –
control
5.3 – Parkinson’s
disease
0.25+0.07 nmol/mg
DNA
1.06+0.55
adducts/105 pb
0.82+0.46
3.1nmol/mg DNA
control
3.5 – Alzheimer’s
1.2 nmol/mg DNA
control
0.7 – Alzheimer’s
1.42+0.59 nM
Control
1.49+0.54
Parkinsons
127+14 fmol/mL
960+150 fmol/mL
control
5000+690 diabetic
97+32pM – Control
500+213 –
Alzheimer’s
97+32 pM Control
288+129 Parkinsons
3.1 nmol/mg DNA
control
3.5 – Alzheimer’s
WWW.ESAINC.COM
HPLC-ECD
(1990)
Shigenaga
and Ames
(1991)
ELISA
Kikuchi et al.,
(2002)
ELISA
Kikuchi et al.,
(2002)
HPLC-ECD
Mei et al.,
(2003)
GC-MS
Alam et al.
(1997)
GC-MS
Podmore et al.
(1998)
Fiala et al.
(1989)
HPLC-ECD
GC-MS
Lyras et al.
(1997)
HPLC-ECD
Abe et al.,
(2003)
HPLC-ECD
Park et al.
(1992)
HPLC-ECD
Shin et al.,
(2001)
HPLC-ECD
Abe et al.
(2002)
HPLC-ECD
Abe et al.,
(2003)
GC-MS
Lyras et al.
(1997)
196
Frontal lobe
“DNA” Adducts
Human
Cervix
FAPy-Ad
Human
Brain region:
Temporal lobe
Frontal lobe
FAPy-Gua
Calf
Thymus
FAPy-Gua
Human
Lung
FAPy-Gua
Human
Brain region:
Temporal lobe
FAPy-Gua
Human
Brain –
substantia
nigra
C8-Methylguanine
Rat
Liver
Colon
N7-Methylguanine
Mouse
Kidney
Rat
Liver
Brain
Liver
Colon
O4-Ethylthymine
6
O -Methylguanine
Human
Liver
Human
Leukocyte
Liver
32
P postlabeling
Simons et al.
(1993)
GC-MS
Lyras et al.
(1997)
HPLC-ECD
Kaur and
Halliwell
(1996)
Jaruga et al.
(1994)
0.22 nmol/mg DNA
control
0.29 – Alzheimer’s
0.05 nmol/mg DNA
control
0.04 – Alzheimer’s
1.0 nmol/mg DNA
25-33 adducts/105
pb normal
50-120 – cancer
tissue
Frontal lobe
N7-Methylguanine
1.2 nmol/mg DNA
control
0.7 – Alzheimer’s
3.81+2.13
adducts/108 pb –
control
5.89+3.7 – smokers
GC-MS-SIM
GC-MS
Lyras et al.
(1997)
GC-MS
Alam et al.
(1997)
HPLC-Fl
Netto et al.
(1992)
HPLC-ECD
Tan et al.
(1990)
HPLC-Fl
Netto et al.
(1992)
HPLC-UV
and 32P postlabeling
Kang et al.
(1992)
HPLC-UV
and 32P postlabeling
Kang et al.
(1992)
10.5 nmol/mg DNA
– control
9.5 – Alzheimer’s
0.8 nmol/mg DNA
control
0.7 – Alzheimer’s
3.2 nmol/mg DNA –
control
2.4 – Parkinson’s
disease
n.d. – basal
103 µmol/mol guan stim.
n.d. – basal
139 – stimulated
13.5+0.5 to 31.5+6.5
µmol/mol DNA
10.4+1.0 to 23.6+3.0
14.8+0.6 to 18.6+1.4
387 µmol/mol guan
– basal
7445 – stimulated
671 – basal
2318 – stimulated
1.9 to 8.7
adducts/108
n.d.
1.1 to 4.2
adducts/108
WWW.ESAINC.COM
197
O6-Methylguanine
Rat
Leukocyte
Liver
Colon
0.8 to 1.6
n.d. – basal
837 µmol/mol guan stim.
n.d. – basal
99 – stimulated
3.9 to 7.5
adducts/108
O4-Methylthymine
Human
Liver
Human
Leukocyte
Brain region:
Temporal lobe
n.d.
Thymine glycol
Frontal lobe
0.2 nmol/mg DNA
control
0.35 – Alzheimer’s
HPLC-Fl
Netto et al.
(1992)
HPLC-UV
and 32P postlabeling
Kang et al.
(1992)
GC-MS
Lyras et al.
(1997)
1.4 nmol/mg DNA
control
1.35 – Alzheimer’s
Table 3.6 Some Tissue Levels Of DNA (and RNA) Adducts Reported In The
Literature. 1IDDM - insulin-dependent diabetes mellitus. 2NIDDM - non- insulin-dependent diabetes mellitus. pb – parent base
(or nucleoside); n.d. – not determined. MS-SIM – mass spectrometry with single ion monitoring.
Based on the levels presented in Table 3.6, it can be seen that the frequency of
adduct formation can be as high as 7 adduct/104 parent bases, but the range is
typically on the order of 1-5 adducts/105 bases. An average of 3 adducts/105
bases translates into ~6 x 106 adducts per cell, a phenomenal amount (Helbock
et al. (1998)). Critical evaluation of published values of DNA adduct formation
has led some researchers to question artifactual production of adducts during
sample preparation and analytical procedures (see below). Even with improved
analytical procedures, 4 adducts/107 parent bases, the equivalent of 24,000
adducts per cell were reported (Helbock et al. (1998)). At present it is not clear
whether such damage is occurring in introns or exons, or in active or quiescent
genes. It is however most likely that such damage is taking place in exposed
DNA rather than the condensed DNA occurring in chromatin (e.g., Table 3.4).
Another question that still awaits an answer is what contribution dead or dying
cells make to adduct levels. As Helbock points out, apoptotic cells may increase
the overall tissue adduct level while having little deleterious biological effect.
Total Adduct Levels.
One estimation of total DNA damage is obtained from the measurement of
urinary DNA adduct “markers”. Several possible markers are now being explored
but most research has focused on 8OH2’dG (Table 3.8). This marker is
unaffected by diet and is not produced from mRNA catabolism. In addition, the
output of 8OH2’dG is significantly elevated in a variety of conditions that are
thought to be associated with increased oxidative stress (e.g., Alzheimer’s
disease, amyotrophic lateral sclerosis, cancer, cystic fibrosis, and smoking)
(Bogdanov et al. (1997)). However, not all “stressful” conditions are associated
WWW.ESAINC.COM
198
with higher levels (e.g., immediately following intense exercise) (Kasai (1997);
Loft and Poulsen (1996)).
Marker
5-Hydroxymethyl uracil
5-Hydroxy
uracil
8-Hydroxyadenine
8-Hydroxy-2’deoxyadenosine
8-Hydroxyguanine
8-Hydroxyguanine
8-Hydroxyguanosine
8-Hydroxyguanosine
Species
Human
Level
121+56 pmol/mL
Human
58+23 pmol/mL
Human
7+4 pmol/mL
Human
0.3nM
Human
583+376 pmol/mL
Human
8-Hydroxyguanosine
dGuo-Malondialdehyde
adduct
8OH2’dG
Human
138+83 nmol/24hr control
202+102 – cancer
333+125 pmol/kg/24h
2810+830
405+85 pmol/kg/24h –
pre-exercise control
310+85 – post exercise
390+85 – post vitamin
335+125 pmol/kg/24h
Rat
Human
28.54+2.2 nmol/kg/24h
0.40+0.05
HPLC-Fl
Human
CE-ECD
8OH2’dG
Human
8OH2’dG
Human
8OH2’dG
Human
13.51+5.1 nM – Control
35.3+28.0 – Cancer
4-19 µmol/mmol creatinine
– oncological patients on
radiotherapy
7.2 nmol/mmol creatinine
– control
5.7 – H. pylori infected
9 nmol/mmol creatinine
8OH2’dG
Human
Human
Rat
Human
8OH2’dG
Human
8OH2’dG
Human
12+4 nmol/mmol
creatinine – control
9+2 small cell carcinoma
responders
7+0.4 control
11+3 small cell carcinoma
non-responders
8 to 14 nmol/24h – cancer
patients
31 to 40 – post
radiotherapy
300+100 pmol/kg/24h
WWW.ESAINC.COM
Method
HPLC then
GC-MS
HPLC then
GC-MS
HPLC then
GC-MS
LC-MS/MS
Reference
Ravanat et al.
(1999)
Ravanat et al.
(1999)
Ravanat et al.
(1999)
Weimann et
al. (2001) (
HPLC then
GC-MS
GC-MS
Ravanat et al.
(1999)
Rozalski et al.
(2002)
Park et al.
(1992)
Witt et al.
(1992)
HPLC-ECD
HPLC
HPLC
CE-UV
Park et al.
(1992)
Agarwal et al.
(1994)
Mei et al.,
(2003)
Kvasnicova et
al., (2003)
ELISA
Witherell et al.
(1998)
ELISA
Cooke et al.
(1998)
Erhola et al.
(1997)
ELISA
GC-MS
Bergtold et al.
(1990)
GC-MS
Simic and
Bergtold
(1991)
199
8OH2’dG
Human
8OH2’dG
Human
8OH2’dG
Human
8OH2’dG
Human
8OH2’dG
Human
8OH2’dG
Human
8OH2’dG
Human
8OH2’dG
Human
8OH2’dG
Human
8OH2’dG
Human
8OH2’dG
Human
8OH2’dG
Human
8OH2’dG
Human
8OH2’dG
Human
8OH2’dG
Human
8OH2’dG
Human
8OH2’dG
Human
1.33+0.29 nmol/µmol
creatinine – control
1.39+0.40 nmol/µmol
creatinine – hemochromatosis patients
35+21 nmol/24hr control
36+15 – cancer
30+15 pmol/mL
1.44 nmol/mmol creatinine
– male
1.68 – female
1.63 –smokers
1.56 – non-smokers
274 pmol/kg/24h – male
264 pmol/kg/24h – female
1.47+0.02 nmol/mmol
creatinine – male
1.58+0.02 – female
5.34+0.03 – neonates
4.4+0.3 ng/mg creatinine –
control
7.2+0.7 – ALS
4.6+0.3 – other
neurological disorders
1.51+0.38 nmol/mmol
creatinine – control
2.78+1.21 – cystic fibrosis
204+133 pmol/kg/24h
0.16 to 8.23 nmol/mmol
creatinine
1.4+0.5 to 2.5+0.4
nmol/mmol creatinine
281.7+47 pmol/kg/day –
human
333+47 – rat
213+84 pmol/kg/24h –
control
320+99 – smokers
318+130 pmol/kg/24h –
control
431+168 – smokers
629+218 pmol/kg 24h –
control
15.2+5 ng/mg creatinine –
control
20.4+8 – ileal neobladder
15.2+4 – colon neobladder
4.27+0.6 ng/mg creatinine
– control
2.8+0.3 – +antiretroviral
2.03+1 µmol/mol
creatinine – spot urine
WWW.ESAINC.COM
GC-MS
Holmberg et
al. (1999)
GC-MS
Rozalski et al.
(2002)
Ravanat et al.
(1999)
Bogdanov et
al. (1997)
HPLC then
GC-MS
HPLC-ECD
HPLC-ECD
Bogdanov et
al. (1999)
HPLC-ECD
Bogdanov et
al. (2000)
HPLC-ECD
Brown et al.
(1995)
HPLC-ECD
Degan et al.
(1991)
Germadnik et
al. (1997)
Inoue et al.
(1993)
Lengger et al.
(2000)
HPLC-ECD
HPLC-ECD
HPLC-ECD
HPLC-ECD
Loft et al.
(1992)
HPLC-ECD
Loft et al.
(1994)
HPLC-ECD
Loft et al.
(1995)
Miyake et al.
(2003)
HPLC-ECD
HPLC-ECD
Paul et al.
(2003)
HPLC-ECD
Pilger et al.
(2002)
200
8OH2’dG
Human
8OH2’dG
Human
8OH2’dG
Human
8OH2’dG
Human
8OH2’dG
Human
8OH2’dG
Human
8OH2’dG
Human
8OH2’dG
Human
8OH2’dG
Human
8OH2’dG
Human
8OH2’dG
Human
8OH2’dG
Human
8OH2’dG
1.86+1.1 – 24h urine
130 to 300 pmol/kg/24h
4.46+2.0 µg/g creatinine
9.33+3.2
1.0+0.4 nmol/mmol
creatinine –
control
1.3+0.4 – smokers
1.6 to 3.7 nmol/mmol
creatinine – control
1.4 to 14.7 – post exercise
14.9+7.8 nmol/24h –
control
18+11 – cancer
950 pmol/kg/24h
3.3+1.9 µg/g creatinine –
control
2.6+0.8 – asphalt
exposure
300 to 630 pmol/kg/24h –
vegetarian
210 to 490 – Brussel
sprout diet
2.8+1.2 nmol/mmol
creatinine – smokers
3.0+1.1 – smokers + βcarotene
480 to 520 pmol/kg/24h
HPLC-ECD
HPLC-ECD
ELISA
HPLC-ECD
Shigenaga et
al. (1990)
Shimoi et al.,
(2002)
Tagesson et
al. (1992)
HPLC-ECD
Tagesson et
al. (1992)
HPLC-ECD
Tagesson et
al. (1995)
HPLC-ECD
Taylor et al.
(1995)
Toraason et
al. (2001)
HPLC-ECD
HPLC-ECD
van Poppel et
al. (1995)
HPLC-ECD
Verhagen et
al. (1995)
HPLC-ECD
Verhagen et
al. (1995)
Witt et al.
(1992)
HPLC-ECD
Yamamoto et
al. (1996)
Human
405+85 pmol/kg/24h –
control
310+85 – post exercise
747+425 pmol/kg/24h –
control
1827+1500 pmol/kg/24h
carcinoma
33+2 pmol/mL
LC-MS/MS
8OH2’dG
Human
20 pmol/mL
LC-MS/MS
8OH2’dG
Human
2 ng/mg creatinine
LC-MS/MS
8OH2’dG
Human
1 to 100nM
LC-MS/MS
8OH2’dG
~120 to 300 pmol/kg/24h
~180 to 500
~550 to 780
172+79 pmol/kg/24h
370+63
0.65 to 1.46 nmol/h
HPLC-ECD
8OH2’dG
Human
Rat
Mouse
Human
Rat
Pig
Pietta et al.,
(2003
Ravanat et al.
(1998)
Renner et al.,
(2000)
Weimann et
al. (2001)
Shigenaga et
al. (1989)
HPLC-ECD
8OH2’dG
Rat
490+70 pmol/kg/24h
GC-MS
8OH2’dG
WWW.ESAINC.COM
HPLC-ECD
HPLC-ECD
Park et al.
(1992)
Loft et al.
(1995)
Teixeira et al.
201
8OH2’dG
Rat
Rat
165+66 to 481+163
pmol/kg/24h
100 to 450 pmol/kg/24h
8OH2’dG
HPLC-ECD
HPLC-ECD
8OH2’dG
Rat
210+50 pmol/24h
HPLC-ECD
8OH2’dG
Rat
89.3+23.7 ng/mg
creatinine
HPLC-ECD
N2-DimethylGuanine
N2-EthylGuanine
Human
0.022 to 0.185 mg/24h
Human
0.0003 to 0.0007 mg/24h
GC-MS or
GC-MS-MS
GC-MS or
GC-MS-MS
N7-(2-hydroxyethyl) guanine
N2-MethylGuanine
N7-Methylguanine
Thymidine
glycol
Human
0.0006 to 0.003 mg/24h
Human
0.352+0.09 mg/24h
Human
3.03+1.46 mg/24h
Human
390-435 pmol/kg/24h
GC-MS or
GC-MS-MS
GC-MS or
GC-MS-MS
GC-MS or
GC-MS-MS
HPLC
Thymidine
glycol
Human
110-250 pmol/kg/24h
GC-MS
Thymine glycol
Human
100-174 pmol/kg/24h
HPLC
(1995)
Fraga et al.
(1990)
Shigenaga
and Ames
(1991)
Deng et al.
(1998)
De Martinis
and Bianchi
(2002)
Cushnir et al.
(1993)
Cushnir et al.
(1993)
Cushnir et al.
(1993)
Cushnir et al.
(1993)
Cushnir et al.
(1993)
Loft and
Poulsen
(1998)
Loft and
Poulsen
(1998)
Loft and
Poulsen
(1998)
Table 3.7 Some Urinary Levels Of DNA Markers Reported In The Literature.
The use of urinary DNA adduct levels to estimate the total DNA damage is not
without its problems. One potential issue with the use of 8OH2’dG as a marker is
that it can be derived by action of ROS on the free deoxy-nucleotide pool (Mo et
al. (1992); Sukumi et al. (1993)). Dephosphorylation of 8-hydroxy-dGTP by the
MutT enzyme helps to limit incorporation of this adduct into DNA by
phosphorylase, but by so doing produces free 8OH2’dG. Furthermore, some 8hydroxy-dGTP inevitably escapes MutT and will be incorporated into DNA.
Following DNA repair this too will contribute to the 8OH2’dG pool (Figure 3.9).
The contribution of oxidized dGTP to urinary 8-hydroxy-2’deoxyguanosine levels
still needs to be critically evaluated. Another issue is that mitochondrial turnover
and repair may contribute to the urinary excretion rates of DNA adducts. Finally
apoptosis may also contribute to urinary excretion rates. All of these processes
can lead to overestimation of the amount of DNA damage repaired each day.
Conversely, several factors may lead to an underestimate of DNA repair. These
include:
WWW.ESAINC.COM
202
i)
ii)
iii)
8OH2’dG is susceptible to oxidation in vivo;
as yet unknown salvage pathways may operate on this nucleotide; and
in mammals the actual products of 8OH2’dG repair have not been
definitively identified (Helbock et al. (1998)). Another confounding
factor for the measurement of all urinary DNA adducts is the
assumption that the analytical procedure being used is valid and that it
is unequivocally and accurately identifying the analyte of interest. As
will be discussed below, this is not always the case.
Another proposed marker of DNA damage, 8-hydroxyguanine, is less useful as
its level is affected by diet. Furthermore it is not specific to DNA as it can be
formed by damage to both DNA and mRNA. Recently, 8-hydroxyguanosine is
being determined as a marker of mRNA damage (Figure 3.9 and Table 3.6).
Measurement of DNA Damage.
A variety of techniques have been used to measure DNA damage (e.g., see
Aruoma and Halliwell (1998)). Regardless of method used, extreme care must be
taken to ensure that artifactual production of adducts does not occur during the
sample extraction, preparation and analytical steps (Hofer and Moller (1998)). By
diminishing artifactual base-modification production, analytical approaches must
now possess sufficient sensitivity to measure one modification in 105 to 107
normal bases and in a few micrograms of DNA.
Two main approaches have been developed to measure DNA damage based
upon whether the DNA molecule is kept whole or hydrolyzed:11
•
Intact DNA lesions can be measured using either immunological methods
or by measuring the knicking activity of DNA repair enzymes (e.g.,
endonuclease III) in conjunction with sedimentation and gel-sequencing
techniques in order to quantify the number of strand breaks (e.g., COMET
assay – single cell gel electrophoresis) (Cadet et al. (1998); Collins et al.
(1993); Gedik et al. (1998)). DNA damage can then be visualized by
suitable staining followed by fluorescence microscopy or computer image
analysis.
•
DNA is hydrolyzed using either acid (base release) or enzymatic digestion
(producing nucleosides, nucleotides or short oligomers). Unfortunately, if
due care is not taken, both isolation and hydrolysis can cause artifactual
production of DNA adducts (reviewed by Kasai (1997)). When measuring
8-hydroxy-2’deoxyguanosine, evidence suggests that a number of
11
Although the measurement of adducts in urine does not require the use of either enzymatic or acid hydrolysis their
analysis is especially challenging due to the low level of adducts and the number and abundance of other compounds in
this sample matrix. Extensive sample cleaning procedures using solid-phase extraction or immunoaffinity columns are
often used. A novel alternative method to simplify the analysis of urine samples makes use of the ability of certain carbonbased columns to selectively retain purines (Bogdanov et al. (1999)).
WWW.ESAINC.COM
203
situations can contribute to formation of this adduct. These include organic
solvents (e.g., chloroform), phenol, oxygen, light, reagent purity (e.g.,
metal content), pH, type of buffer, certain plastics, the quantity and type of
tissue (e.g., hemolyzed blood is high in iron and can promote 8OH2’dG
under acidic conditions), the times taken for DNA isolation and hydrolysis,
and the analytical approach used (Adachi et al. (1995); Claycamp (1992);
Kasai (1997); Nicotera et al. (1994); Floyd et al. (1990)). The contribution
of these factors to artifactual adduct production continues to be evaluated
and debated. For example, while some claim that phenol may be
problematic Helbock et al. (1998) found that phenol extraction caused only
a minor increase in 8OH2’dG levels. They also showed that the use of a
“chaotropic12 sodium iodide method” to isolate DNA could lower the level
of 8OH2’dG by an order of magnitude. Similarly, Hofer and Moller (1998)
reported no effect of fresh and aged (pink) phenol and found that the
inclusion of the spin-trap TEMPO during sample preparation could prevent
artifactual production of 8OH2’dG. Taken together such observations are
likely to explain some of the previously observed differences in 8OH2’dG
levels. An up-to-date extraction and hydrolysis procedure for HPLC
analysis is presented in Appendix 3.1.
In order to address the inter-laboratory variability in the measurement of DNA
adducts, The European Standards Commission on DNA Damage (ESCODD)
was established. Over the past three years several laboratories (using different
analytical procedures) have participated in the study and have been tested on
their ability to detect 8OH2’dG (and sometimes 8OHGua) in artificial
oligonucleotides, calf thymus DNA and HeLa DNA and to show dose response
(Anon (2000, 2002); ESCODD (2002, 2003); Riis (2002)). The conclusions so far
are:
•
•
•
•
•
Over the last few years the ability to overcome artifactual production of
8OH2’dG to some extent has been improved.
The COMET assay possibly underestimates the 8OH2’dG level.
GC-MS and HPLC with amperometric detection cannot be recommended
due to artifacts.
Immunological detection, 32P-postlabelling and LC-MS-MS lack precision
and show no dose response.
HPLC with coulometric detection shows good precision, shows dose
response and is the preferred technique for 8OH2’dG detection.
12
The ability for certain substances to disrupt the structure of water promoting the solubility of certain non-polar
substances (e.g., DNA) in polar solvents (e.g., water).
WWW.ESAINC.COM
204
Gas- and Liquid- Chromatography-Mass Spectrometry.
Gas chromatography-single ion monitoring mass spectrometry (GC-SIM-MS) has
been routinely used to measure a wide variety of DNA adducts (e.g., Dizdaroglu
(1991, 1994); Lyras et al. (1998)). It is one of the few techniques that can
measure adducts formed from all four bases. After isolation DNA samples are
first hydrolyzed using acid or enzymes. Formic acid is routinely used for DNA
hydrolysis as most adducts are stable and few are formed as a result of this
treatment. However, formic acid does cause deamination and dehydration of
cytosine–derived adducts. Cytosine glycol either dehydrates to 5hydroxycytosine or dehydrates and deaminates to 5-hydroxyuracil. 5,6Dihydrocytosine,
5-hydroxy-6-hydrocytosine
and
5,6-dihydroxycytosine
deaminate to 5,6-dihydrouracil, 5-hydroxy-6-hydrouracil and 5,6-dihydroxyuracil,
respectively (Dizdaroglu et al. (1993)). Alloxan, a product of cytosine, is
decarboxylated to 5-hydroxyhydantoin.
DNA bases and adducts must be converted into their volatile derivatives (e.g.,
trimethylsilyl derivatives) before GC/MS analysis. Many different isotopically
enriched modified bases and nucleosides are now available, thus allowing
isotope-dilution mass spectrometry to be used for the analysis of several DNA
adducts (Dizdaroglu (1993b); Djuric et al. (1991b)). GC-SIM-MS routinely
achieves high (picogram) sensitivity, high selectivity and structural information.
The highest sensitivity can only be obtained when monitoring the most
characteristic ion in its mass spectrum. A few characteristic ions of the analyte
and its labeled analog must be monitored at the corresponding retention time in
order to accurately and reliably identify the analyte of interest (Dizdaroglu
(1997)). Recently Ravanat et al. (1995) reported that the conditions used for GCMS derivatization promote the formation of 8-OHGua from guanine. Furthermore,
Douki et al. (1996) found that derivatization causes the artifactual formation of 5hydroxycytosine and 8-hydroxyadenine. Taken together these findings suggest
that due care must be exercised when interpreting data on 8-OHGua obtained
using GC-MS techniques (see the ESCODD conclusions, above). This may
explain why those using GC-MS approaches report much higher levels of
guanine adducts than those using HPLC-ECD (e.g., the CSF level of 8OH2’dG is
~60,000 fold higher when measured using GC/MS than HPLC-ECD (Table 3.6)).
The use of anoxic conditions during preparation and derivatization, and the
addition of a prepurification step prior to derivatization may overcome some of
these issues (Dizdaroglu et al. (2003); Nakajima et al. (1996); Ravanat et al.
(1998)) but renders GC-MS much less routine.
The GC-MS approach has also been used to explore the specificity of
glycosylase repair enzymes. Because many base lesions can be simultaneously
measured in the same DNA sample, the base preference (which lesions are
excised and which are ignored) for a given enzyme can readily be measured
(Boiteux et al. (1992); Karakaya et al. (1997); Nackerdian et al. (1992)). This
approach can also be used to study the kinetics of excision (Karakaya et al.
WWW.ESAINC.COM
205
(1997)). Another GC-MS technique is also being used to study DNA damage:
GC/electron capture negative ion mass spectrometry. Following initial isolation by
reversed-phase HPLC, the modified bases are alkylated off-line to form the
highly electrophoretic pentafluorobenzyl derivatives. This approach offers
femtomole sensitivity.
Over the past few years liquid chromatography with mass spectrometry (LC-MS)
using electrospray interface has been applied to the measurement of various
DNA adducts (Tables 3.6 and 3.7). Like with GC-MS, the DNA must be isolated
and hydrolyzed before the adducts are measured (see below). With the “higherend” instruments used in a selected reaction monitoring mode high sensitivities
(typically low to mid pg) can be achieved. LC-MS has been used to measure the
ionization-induced decomposition of thymidine (Berger et al. (1992)), and to
study the formation of malondialdehyde-DNA adducts (Chaudhary et al. (1996);
Jajoo et al. (1992); Rouzer et al. (1997)). LC-MS (typically using triple-quads) is
beginning to show promise for the analysis of 8OH2’dG in urine and even
approaches HPLC-ECD in sensitivity (Poulsen et al. (1998); Ravanat et al.
(1998)). However, LC-MS is not devoid of problems as nucleosides can artificially
oxidize at the output of the HPLC column during the ionization process (Ravanat
et al. (1998)). Furthermore, ESCODD do not recommend LC-MS approaches
due to an inability to measure dose-dependent adduct formation (see above).
HPLC.
The measurement of DNA adducts using HPLC-based approaches first requires
isolation of DNA from the tissue and then hydrolysis of the DNA molecule (see
above). Reversed-phase HPLC permits the separation of a variety of nucleotides
and bases. Several different detectors have been used including electrochemical,
UV absorbance, fluorescence, and mass spectrometry (using different interfaces
[thermospray, electrospray, fast atom bombardment and atmospheric pressure
ionization approaches] and spectrometers [single quads, triple quads and ion
traps] (Cadet and Weinfeld (1993); Douki et al. (2003); Poulsen et al., (2003)).
They differ in selectivity, sensitivity, amount of DNA required, the ability to
determine structure and ease of use. HPLC-UV or HPLC-photodiode array
approaches are generally too insensitive to measure the low levels of adducts in
typical DNA samples (1-100µg) (Cathcart et al. (1984)). HPLC-fluorescence was
used to study the formation of DNA photoproducts (e.g., the reactions of
furocoumarins, 3-carbethoxypsoralens and pyrimidones) (Cadet and Weinfeld
(1993)) but not oxidized DNA adducts, probably due to their weak native
fluorescence.
HPLC-ECD is selective and sensitive (approximately 103-104 times greater than
absorbance-based approaches), possesses good dynamic range (essential if
adducts and precursor base/nucleoside are to be measured simultaneously),
requires simple sample preparation and no derivatization, and is easy to operate.
WWW.ESAINC.COM
206
Unfortunately, not all adducts are electrochemically active and unlike some MSbased analytical techniques HPLC-ECD does not offer structural information. The
use of HPLC-ECD to measure DNA adducts, the advantages of coulometric
detection over amperometric detection and the benefits of coulometric array
detection have been reviewed elsewhere (Acworth et al. (1997; 1998); Poulsen
et al. (2003)). A number of HPLC-ECD methods are currently being used (e.g.,
references in Tables 3.6 and 3.7). From the original work of Floyd et al. (1986),
who first used an HPLC single electrode ECD approach to measure 8-hydroxy2’deoxyguanosine, the number of DNA adducts that can be well resolved and
measured electrochemically has been considerably expanded. This is due in part
to the use of gradient chromatography and improved voltammetric resolution
achieved with coulometric arrays (Acworth et al. (1997)) (Figure 3.10) (see also
ESA Application Note 70-5970 DNA, Nucleosides and Bases). Although there
are previous reports of simultaneously measuring numerous adduct standards on
a single electrode system, poor chromatography and a lack of voltammetric
resolution severely limit the applicability of such an approach to the
measurement of hydrolyzed DNA samples (Kaur and Halliwell (1996)). HPLCECD still remains one of the most accessible and preferred methods for DNA
adduct analysis.
[520 mV]
[340 mV]
uric acid
[460 mV]
1.0 [400 mV]
2'deoxyadenosine
adenosine
thymidine
2'deoxyguanosine
guanosine
inosine
8-hydroxyadenine
adenine
thymine
2'deoxycytidine
xanthine
uridine
hypoxanthine
guanine
[580 mV]
8-hydroxy-2'deoxyguanosine
[640 mV]
8-hydroxyguanine
[0 mV]
5-hydroxy-2'deoxycytidine
2.0
5-hydroxyuracil
Response (µA)
3.0
cytidine
5-methylcytosine
cytosine
uracil
4.0
[280 mV]
[220 mV]
[160 mV]
0.0 [100 mV]
0.0
10.0
20.0
30.0
40.0
Retention time (minutes)
Figure 3.10 Separation and Detection of DNA Adducts, Nucleosides and
Bases Using Gradient HPLC With Coulometric Electrode Array And UVAbsorbance Detection (100ng each on column). (The UV channel is labeled 0mV).
Electrochemically Active Compounds Can Be Found In Figures 3.5 and 3.6.
WWW.ESAINC.COM
207
The gradient HPLC system consisted of two pumps, a PEEK pulse damper, a high-pressure
mixer, a refrigerated autosampler, a thermostated organizer module, a twelve channel
CoulArray and a dual channel UV detector.
LC Conditions:
Column:
Mobile Phase A:
Mobile phase B:
TosoHaas TSK-GEL ODS-80TM, (4.6 x 250mm; 5µm).
50mM lithium acetate, pH 4.0 (with phosphoric acid).
50 mM lithium acetate-acetonitrile, pH 4.2 (with phosphoric acid); 85:15;
(v/v).
Gradient Conditions:
0-15 mins 0% B; 15 to 40 mins 50% B; 40 to 45 mins 100% B;
45 mins 0% B; 45 to 60 mins 0% B.
Flow Rate:
1.0 mL/min.
Temperature:
31oC.
Injection Volume
20µL.
Applied Potentials:
100 to 700 mV in 60mV increments (vs. Pd).
Wavelength:
260nm (0.01 AUFS).
See also ESA Application Note 70-5970 DNA, Nucleosides and Bases for more details.
An interesting HPLC-ECD method for the measurement of DNA oxidation was
recently developed (Beckman et al. (1998)). DNA is first isolated and is then
treated with E. coli repair enzyme, formamidopyrimidine glycosylase, to release
8-hydroxyguanine. This adduct can then be easily measured using HPLC-ECD
following separation from its parent DNA by ultrafiltration. This approach has the
advantages of minimal sample treatment (thereby minimizing DNA oxidation) and
the elimination of other bases and adducts from the sample resulting in simpler
chromatography. Issues with this technique include the efficiency of adduct
liberation and the lack of commercially available glycosylase.
Postlabeling Assays.
These assays include
(Cadet et al. (1992)).
32
P postlabeling and chemical postlabeling methods
Randerath and colleagues (1981) first developed the 32P postlabeling procedure
to study carcinogen-DNA adducts. Isolated DNA is first digested enzymatically to
produce nucleoside 3’ monophosphates (or short oligomers) that are then
enzymatically radiolabeled by incubating with 32P-ATP and phage T4
polynucleotide kinase. Radiolabeled bases and adducts can then be separated
using 2-D thin layer chromatography, polyacrylamide gel electrophoresis or
HPLC (Gorelick (1993)). This approach enables high sensitivity measurement of
a variety of DNA adducts (Keith and Dirheimer (1995); Poirier and Weston
(1996)). For example, 5-hydroxmethylyuracil can be measured at the level of one
modification per 107 normal bases in 1µg of DNA (Cadet et al. (1992)).
Unfortunately this approach suffers from several disadvantages including the use
of radioactive substances, artifact problems and is not suited for high sample
throughput (Cadet et al. (1997)). The issues of this approach, including method
development and its use in the study of DNA damage, have been reviewed
WWW.ESAINC.COM
208
elsewhere (Cadet and Weinfeld (1993); Cadet et al. (1992); Marnett and
Burcham (1993)).
Using 32P postlabeling Randerath found a number of putative adducts in DNA
extracts of tissues obtained from untreated animals (Randerath et al. (1986)).
These spots were termed I-compounds as they occurred indigenously. Icompounds accumulate in an age-dependent, highly reproducible manner and
their pattern is found to affected by gender, tissue and diet. I-compounds appear
to arise via the interaction of DNA with endogenous reactants formed in the
course of normal metabolism (see Marnett and Burcham (1993) and references
therein). They exhibit a wide range of polarities suggesting that they are
structurally diverse (Randerath et al. 1990)). Although their exact structure
remains unknown, current evidence suggests that some I-compounds may
contain DNA-lipid peroxidation adducts (Li et al. (1995b) and references therein).
The biological role of the I-compounds remains controversial. Although many
DNA adducts are found to increase the probability of mutation which may
eventually lead to development of cancer, the levels of I-compounds do not
positively correlate with cancer. Their levels do show diurnal variation suggesting
that they are regulated. Consequently, some researchers have suggested that Icompounds may play a role in regulation of gene expression and proliferation
(Marnett and Burcham (1993) and references therein). Others dispute this role
and hypothesize that altered I-compound levels may just be the consequence of
changes in cytochrome P450 activity.
Chemical postlabeling follows a different procedure. Once separated using
HPLC, nucleotides and nucleosides resulting from enzymatic hydrolysis can then
be chemically postlabeled using acetic anhydride. The resulting acetylated
nucleosides can then be resolved using another HPLC system (Frenkel et al.
(1991)). However, with only picomole sensitivity this approach is not competitive
with the other approaches described in this section. A second chemical method
has nucleoside 5’-monophosphates reacting with the fluorogenic agents dansyl
chloride/fluorescein isothiocyanate. This approach is capable of measuring 1
adduct per 106 normal nucleosides in a 100µg DNA sample (Sharma et al.
(1990)).
Immunochemical Detection.
This approach makes use of the fact that antibodies raised to a specific DNA
adduct can then be utilized to measure such lesions in DNA (see Cadet and
Weinfeld (1993) and references therein; Herbert and Lunec (1998)). The major
advantage of immunochemical detection is one of sensitivity (subfemtomol level),
but this can only be achieved if the antibodies have been correctly generated.
This approach can suffer from poor selectivity due to either inappropriate
generation of the antibody (e.g., using DNA with a number of different lesions) or
WWW.ESAINC.COM
209
cross reactivity with compounds containing similar chemical structure. The
generation of monoclonal antibodies can improve the selectivity of this approach.
The most widely used immunodetection method is probably the enzyme-linked
immunoabsorbent assay (ELISA). Here, secondary antibodies that are covalently
cross-linked to an enzyme, such as peroxidase or alkaline phosphatase, detect
primary antibodies bound to the antigen. Such binding can then be visualized by
applying a substrate that yields a product that can be measured using
absorbance techniques. Alternatively, a secondary antibody covalently bound to
a fluorogen (e.g., fluorescein isothiocyanate) can be used and measured using
fluorescence approaches. Radioimmunoassay involves the coupling of a
radioisotope to the primary antibody. Binding can then be measured using
scintillation counting or phosphor imaging.
The Measurement of 8OH2’dG in Urine.
The accurate, reliable and routine measurement of urinary 8OH2’dG levels, as
an estimate of total oxidative stress, has been the focus of many laboratories.
Unfortunately, this is not as straight forward as it may first appear. The
measurement of 8OH2’dG in urine represents a considerable challenge. This
adduct is polar in nature and occurs at extremely low levels in an extremely
complex and variable matrix. The complicated nature of the urine makes it
necessary to process the sample prior to analysis to help minimize matrix
interferences and to concentrate the adduct. A variety of approaches have been
used including:
•
•
•
•
•
•
Solid phase extraction (SPE) (Brown et al. (1995); Faux et al. (1994);
Lunec et al. (1994); Shigenaga et al. (1989; 1990); Vigue et al. (1993));
SPE in conjunction with an immunoaffinity column (Fraga et al. (1990);
Park et al. (1992); Shigenaga and Ames (1991); Shigenaga et al. (1994));
triple column switching (Loft et al. (1992, 1995); Verhagen et al. (1995))
prior to HPLC-ECD analysis;
SPE and triple column switching (Lagorio et al. (1994); Tagesson et al.
(1995)) prior to HPLC-ECD analysis;
SPE and trimethylsilylation (Lunec et al. (1994)) prior to GC-MS analysis;
Evaporation, acetylation and pentafluorobenzylation (Teixeira et al.
(1993)) prior to GC-MS analysis; and
SPE, evaporation, acetylation, pentafluorobenzylation, preparative HPLC
and evaporation (Teixeira et al. (1995)) prior to GC-MS analysis.
In general, the sample processing for HPLC-ECD is much simpler than those for
GC-MS and are less likely to lead to artifactual adduct levels (see above).
ode
Although the original methods using SPE followed by HPLC-ECD were quite
simple to perform in the laboratory, the results given by a single electr
WWW.ESAINC.COM
210
detector may be erroneous (Brown et al. (1995); Faux et al. (1994); Lunec et al.
(1994); Shigenaga et al. (1989; 1990); Vigue et al. (1993)). Using a coulometric
array detector we have found that an 8OH2’dG peak that appears to be “pure” on
a single electrode system is actually due to the co-elution of several analytes
(Acworth et al. (1997); Bogdanov et al. (1999)) (Figure 3.11). This readily
illustrates the danger of basing peak purity solely upon matching retention times
of unknown samples to external standards. Methods using no or limited sample
preparation, HPLC-single electrode detection and claiming to measure urinary
8OH2’dG levels should be viewed with extreme caution.
One method that may eventually prove useful for isolating and analyzing 8hydroxy-2’deoxyguanosine from urine is SPE followed by an immunoaffinity
column and HPLC-ECD (Fraga et al. (1990); Park et al. (1992); Shigenaga and
Ames (1991); Shigenaga et al. (1994)). Here a monoclonal antibody specific for
8OH2’dG is used to clean the urine sample. The method appears to be quite
specific and sensitive for the determination of 8OH2’dG. However, the lifetime of
the columns may be a problem, with decreased binding efficiency occurring over
time (Shigenaga et al. (1994)). We have observed that a coelution occurs when
this column irreversibly ages.
8-OH2´DG In Urine
Conventional
8-OH-2´DG
1.0
0.5
0.0
CoulArray
8-OH-2´DG
What appears to be
a single 8OH2’dG
peak by
conventional HPLCECD is actually
found to be a
co-elution of several
metabolites by
CoulArray detection!
8
7
6
5
4
3
9
10
11
12
13
14
15
16
Time (min)
Figure 3.11 Analysis of 8OH2’dG In The Same Urine Sample
Using A Conventional Single Electrode Detector (Top
Chromatogram) And An Electrode Array Detector (Bottom
Chromatogram).
WWW.ESAINC.COM
211
Some approaches to sample preparation involve the use of column switching.
Here, a standard is injected onto one HPLC column and the retention time of the
analyte noted. In subsequent sample injections, a valve is toggled just before the
analyte elutes, resulting in a “heart-cut” that contains the peak of interest. The
resulting ”slug” of sample flows onto a second “trapping” column or sample loop,
and finally the valve is toggled back to its original position after the analyte
elutes. The “slug” containing the analyte is transferred onto a third analytical
column and then finally into a coulometric detector (Loft et al. (1992, 1995);
Verhargen et al. (1995)). This method is extremely complex; requires two
detectors, two or three columns, and a switching valve; and has as an extremely
long run time, thereby limiting sample throughput to only a few samples per day.
Other researchers have combined column-switching techniques with an SPE
procedure prior to injecting the sample onto the first column (Lagorio et al.
(1994); Tagesson et al. (1995)). This technique speeds up the analytical run time
because the lipophilic compounds that elute for several hours remain on the
disposable SPE cartridge.
A different approach to column switching was recently been developed and used
to measure 8OH2’dG in a variety of body fluids including urine, sweat, plasma
and CSF (Bogdanov et al. (1998, 1999, 2003). This approach uses in-line porous
graphite columns, made from carbon chosen for its purine-binding properties, to
selectively clean urine samples before analysis by ECD. Urine diluted in basic
buffer first passes through a C18 column. The band containing 8OH2’dG is then
trapped onto a carbon column. The carbon column is first washed to remove
interfering analytes and then exposed to a mobile phase containing a competitive
non-EC active compound (adenosine) to displace bound 8OH2’dG. 8OH2’dG
and the few compounds also bound to the carbon column are finally resolved on
a second C18 column and measured using ECD. This approach is reproducible,
highly selective and sensitive (~500fg on column), routine and allows up to forty
samples to be analyzed each day (Figure 3.12). During method development
Matson (1998) observed that the precipitate that sometimes forms in urine
samples over time is capable of binding 8OH2’dG. This precipitate (probably
composed of uric acid and other small molecules) is readily solubilized in the
basic dilution buffer. Procedures that ignore this precipitate may therefore not be
measuring “true” urinary 8OH2’dG levels. This carbon column-based approach
has great applicability to a variety of other assays. All that is required is a
graphite column showing selective binding for the analyte of interest and a nonEC active displacer molecule. Currently this approach is being expanded to
measure other proposed markers of oxidative stress. For example, with minor
modification to the chromatographic conditions and by using different displacing
molecules, the method can be adapted for the measurement of
5-hydroxycytidine, 8-hydroxyguanine, 8-hydroxyadenine, 7-methylguanine, or
8-nitroguanine. Plasma 3-nitrotyrosine and 3-chlorotyrosine can also be
determined but require 3-nitrobenzoic acid as the displacing molecule, and a
different HPLC-chemistry (see Bogdanov et al. (2003) for greater detail).
WWW.ESAINC.COM
212
DNA Damage in Health and Disease.
There have been many reports measuring steady state (DNA levels e.g., Table
3.6) and total DNA damage (urine e.g., Table 3.7). Loft and Poulsen (1996) noted
that while “there is good agreement between different laboratories regarding the
values of urinary excretion of repair products, the values obtained from DNA
isolated from tissues or cells differ by several orders of magnitude, some of
which may be due to the choice of analytical method”. The possible contribution
of sample preparation and analysis to DNA adduct levels is still undergoing
evaluation (see ESCODD above).
Urinary
8OH2’dG
I - 8OH2’dG standard, 10ng/ml
II - urine from an ALS patient
III-urine from a control subject
Figure 3.12 The Measurement Of 8OH2’dG In Control Urine (III)
And In Urine Obtained From An ALS Patient (II) Using The
Carbon Column Switching Procedure (Bogdanov et al. (1999,
2003).
There appears to be a direct correlation between 24hr oxygen consumption and
the urinary excretion rate of 8OH2’dG and thymidine glycol (Loft and Poulsen
(1996) and references therein). This is probably due to increased production of
pro-oxidants by mitochondria associated with increased metabolic rate. If this is
true, elevated urinary adduct levels should be seen in other conditions where the
basal metabolic rate is increased, such as with exercise. Data from the few
exercise studies available suggest that this is indeed the case. It is still not clear
whether urinary adduct levels increase following short periods of exercise (Loft
WWW.ESAINC.COM
213
and Poulsen (1996) and references therein), but they are definitely elevated
following severe activity such as marathon running (Alessio (1993)) or prolonged
and repetitive bouts of exercise (Poulsen et al. (1996)).
Conversely, short-term caloric restriction which leads to a reduced metabolic rate
is found to be associated with lowered steady state adduct levels (Chung et al.
(1992); Djuric et al. (1992); McCarter (1995); McCarter and McGee (1989)).
Furthermore, at least one study reported a lowered urinary excretion rate of 8hydroxy-2’deoxyguanosine by 40-50% following energy restriction for ten days
(Simic and Bertgold (1991)). However, energy restriction by 20% was without
affect (Loft et al. (1995); Wicric et al. (1995)). Aging is also associated with
decreased rates of metabolism. Although there have been numerous reports on
the association of DNA damage with aging so far there has not been a true
systematic study (Ames et al. (1993); Holmes et al. (1992); Lee and Wei (1997);
Perez-Camp et al. (1998); Randerath et al. (1992); Wei (1998)). Evidence to date
suggests that the rate of damage decreases with age but that steady state levels
appear to increase, possibly the result of failing repair mechanisms (Loft and
Poulsen (1996) and references therein).
Many studies have measured oxidative DNA modifications in relation to a variety
of disorders including autoimmune diseases (e.g., rheumatoid arthritis and
systemic lupus erythematosus), cancer, chronic hepatitis, cystic fibrosis,
inflammatory bowel disease, metal storage diseases (e.g., Wilson’s disease),
and Fanconi anemia (reviewed by Loft and Poulsen (1996); Marnett and
Burcham (1993); Wiseman and Halliwell (1996)). In general, many of these
diseases are associated with an increased rate of oxidative DNA modification or,
in some cases, deficient repair. In most cases a causal relationship between
DNA damage and cancer in humans still remains elusive. Increasing evidence,
however, suggests a role for a mutant p53 tumor suppressor gene in some
human cancers (Greenblatt et al. (1994); Husgafvel-Pursiainen et al. (1995);
Semenza and Weasel (1997); Soussi (1996)).
AMINO ACIDS AND PROTEINS.
Introduction.
Amino acids are the basic building block of proteins. Approximately 22 amino
acids are commonly found in living organisms. They differ in chemical reactivity,
charge, shape, size, and hydrogen bonding capacity. Free amino acids play
several important roles in the body. They take part in intermediary metabolism
(e.g., glycine, alanine, aspartate and glutamate), act as neurotransmitters (e.g.,
glutamate, aspartate and glycine), and are the precursor of monoamine
neurotransmitters (tyrosine is converted to catecholamines, histidine to
histamine, and tryptophan to serotonin), hormones (e.g., thyroxin), peptides (e.g.,
GSH, substance P, and insulin) and proteins. Amino acids possess a chiral
WWW.ESAINC.COM
214
center. All amino acids used by eukaryotes are in the L-conformation (Figure
3.13).
Figure 3.13 Stereo Pair of L-Tyrosine.
In proteins, the α-carboxyl group of one amino acid is joined via an amide
(peptide) bond to the α-amino group of another amino acid. Many amino acid
units can be joined together to form a polypeptide chain. Most proteins typically
contain between 50 and 200 amino acid residues (5-22 kDa). Proteins are a
diverse family of molecules, but they are all composed from the basic set of 22
amino acids (remember that there may be some post-translational modifications).
The sequence of amino acids in a protein is ultimately determined by the
sequence of DNA bases in the DNA molecule. It is the sequence of amino acids
that gives a protein its biochemical and physical properties. Proteins play many
important roles including: enzymatic catalysis, transport, storage, coordinated
motion, mechanical support, immune protection, generation and transmission of
nerve impulses and control of growth and differentiation (Stryer (1988)). Proteins
are very susceptible to oxidative damage that can affect their physiological
function. To better understand how pro-oxidants can damage proteins, we must
first examine their structure.
Protein Molecular Structure.
The physical (e.g., shape, solubility, and strength) and biochemical (e.g., enzyme
activity, antigen recognition, and biomechanical contraction) properties of a
protein are dependent upon its structure. It is interesting to note that although a
protein initially exists as a linear code in DNA, transcription and translation
WWW.ESAINC.COM
215
produces a molecule with three-dimensional structure. The folding of the
polypeptide backbone and stability of the resulting form is dependent upon the
electrostatic and hydrophobic interactions between amino acid residues, and
stabilization by disulfide (cystine) bridges. There are at least four levels of
structure applied to proteins.
•
•
•
•
The primary structure refers to the amino acid sequence and location of
disulfide bridges.
Secondary structure refers to the spatial arrangement of amino acid
residues close to each other in the linear sequence. These often are
repeated and can give rise to periodic structures such as the α-helix, the
β-pleated sheet and the collagen helix.
Tertiary structure refers to the spatial arrangement of amino acids far
apart in the linear sequence but which may ultimately locate close to each
other when the protein is correctly folded.
Some proteins are made of more than one independent polypeptide chain
each folded into a subunit. Quaternary structure refers to the spatial
arrangement of such subunits. These subunits can either be identical
(e.g., all 180 coat proteins of the tomato bushy stunt virus are identical) or
different (e.g., hemoglobin consists of four dissimilar subunits). The
interfaces between subunits are important in transmission of information
such as substrate binding.
Higher structural levels have now been recognized and include super secondary
structure (clusters of secondary structures, e.g., βαβ repeats) and domains
(compact folded structures linked together by flexible polypeptide segments).
The correct folding of a protein is essential for it to function properly. This is
because the first step in the action of a protein is its binding to another
molecule.13 This ability is only possible because proteins can form
complementary surfaces and clefts. These structures are produced from the wide
selection of amino acid side chains permitting a protein to form hydrogen,
electrostatic and van der Waal’s bonds with its substrate. The correct
conformation, therefore, permits all essential residues, regardless of location in
the linear sequence, to play a role in the three-dimensional shape of the surface
or cleft (active site) of the protein molecule. Consequently, no other
macromolecule group is capable of recognizing and interacting with so many
diverse molecular structures. The formation of the active site is essential for the
recognition of a substrate by an enzyme. Correct folding also influences the
ability of an enzyme to bind a cofactor or metal, regulation of its activity by
phosphorylation, and binding of allosteric modulators. It should not be surprising
that oxidative damage, either resulting in incorrect folding or alteration of the
structure of essential amino acids acid residues lining the active site or other key
sites in the protein molecule, can markedly effect protein function. In order to
13
Examples of protein binding include control of gene expression, determination of self from non-self, assembly of viral
protein coats, binding of a substrate by a receptor, and the binding of a substrate and cofactor by an enzyme.
WWW.ESAINC.COM
216
maintain cellular homeostasis damaged proteins must either be repaired or
destroyed.
Numerous in vitro experiments on protein folding show that all the information
required for the formation of the native, three-dimensional structure of a protein is
encoded in its amino acid sequence. For example, Alfinsen (1973) reported that
a denatured protein is capable of regaining its tertiary structure. For many years
in vivo folding was also assumed to be a similar autonomous process unaffected
by other cellular components such as proteins. However, this idea was
challenged with the discovery of some new proteins, the molecular chaperones
(see Schwarz et al. (1996) and references therein). Chaperones play an active
role in protein folding and prevention of aggregation, both in vivo and in vitro.
Chaperones help to direct proteins towards repair or degradation processes
thereby ensuring cell survival. Unlike enzymes that are actively involved in
folding (e.g., peptidyl-prolyl cis-trans isomerases and disulfide oxidoreductase)
molecular chaperones affect folding processes nonspecifically and can react with
a large number of proteins that expose non-native structures. Many chaperones
were originally identified as heat shock proteins (Hsps). Hsps or, more correctly,
stress proteins, are produced when cells are exposed to stressors including heat
stress (~5oC above normal temperature), oxidative stress, reperfusion-ischemia
injury, heavy metals, mutant proteins, anticancer drugs, and apoptotic agents,
and are increased in bacterial and viral infections (Arrigo (1998); Benjamin and
McMillan (1998); Freeman et al. (1999); Lappa and Sistonen (1997)). This Hsp
response has been implicated in the protection of cells from different forms of
injury and to the improvement of cell survival following injury. To date the
induction of heat shock response has been reported for a variety of diseases
including inflammation, myocardial ischemia, and cystic fibrosis (Leppa and
Sistonen (1997); Strickland et al. (1997); Thomas et al. (1995)). Hsps have also
been proposed to:
1) Transiently bind and delay folding of nascent polypeptide chains until
synthesis is complete;
2) Maintain these chains in a suitable conformation for passage across
organelle membranes;
3) Prevent aggregation;
4) Actively disassemble clathrin-coated vesicles;
5) Hold steroid aporeceptor complexes in ligand competent states; and
6) Assist in degrading damaged proteins by promoting ubiquitinylation and
proteasome lysis (from Benjamin and McMillan (1998)). The pathway by
which damaged proteins are removed is discussed below.
So far six families of Hsps exist, classified according to their molecular weights:
Hsp100 (100-110 kDa), Hsp90, Hsp70, Hsp60 (the chaperonin system), Hsp40
and the small Hsps (15-30 kDa) (including heme oxygenase and α,β-crystallin).
The different Hsps families appear to play different roles in the protein folding
process. For example, Hsp70, the most abundant group in eukaryotic cells, is an
WWW.ESAINC.COM
217
ATP-dependent chaperone that is responsible for binding to the nascent
polypeptide chain before its release from the ribosome thereby preventing
incorrect folding of the incomplete polypeptide. Interestingly, small Hsps, which
are characterized by an in vitro ATP-independent chaperone activity, not only
enhance the survival of cells exposed to oxidative stress by decreasing ROS
levels in a glutathione-dependent manner but also interfere with apoptosis (Arrigo
(1998) and references therein). Readers interested in a more in depth discussion
of the interaction and roles of the different Hsps families are referred to Benjamin
and McMillan (1998) and Buchner (1996).
Pro-oxidants and Protein Damage.
Pro-oxidants can damage proteins by both indirect and direct mechanisms. The
effects of protein oxidation on protein function and enzyme activity are
summarized in Table 3.8.
The Indirect Pathway.
This (mutation) pathway does not involve oxidative damage to the protein per se.
Rather this process involves oxidative damage to the DNA molecule encoding
the protein. Thus pro-oxidants can cause changes in the base sequence of the
DNA molecule. If such base modification is in a coding region of DNA (exon) and
not corrected, the DNA molecule may be transcribed incorrectly. Translation of
the mutant mRNA can result in a mutant protein containing a wrong amino acid in
its primary sequence. If this modified amino acid occurs in an essential part of
the protein (e.g., the active site of an enzyme or a portion that alters folding), the
function of that protein may be impaired. Fortunately, unlike modified DNA that
can pass from cell to cell during mitosis thereby continuing the production of
mutant protein, damage to a protein is non-replicating and stops with its
destruction.
Enzyme/Protein
Aconitase, branchedchain amino acid
dehydrogenase,
complex 1,
dehydratases, and 6phosphogluconate
dehydrogenase
Alcohol
Dehydrogenase
Reactive
Species/
Treatment
Hydrogen
peroxide/
superoxide
Peroxynitrite
Modification/
Comments
Reference
Damage to iron-sulfur
clusters leads to
enzyme inhibition
Bunik and Sievers
(2002); Liochev
and Fridovich
(1994)
Disruption of zincthiolate center leads
to release of zinc and
inactivation of the
enzyme
Crow et al. (1995)
WWW.ESAINC.COM
218
D-Amino acid oxidase
Chlorination
Angiotensin II
Peroxynitrite
Bovine serum
albumin, collagen,
hemoglobin,
myoglobin
ROS
generating
systems
Ca2+-ATPase
(sarcoplasmic
reticulum)
ROS/RNS
Catalase
Singlet oxygen
Hypochlorous
acid
Creatine kinase
Superoxide
Peroxynitrite
Cu/Zn superoxide
dismutase
ROS
Cytochrome P450
Radical
intermediates
Fructose
bisphosphatase
Galactose oxidase
Formation of 3chlorotyrosine in the
active site leads to
inhibition
Formation of
essential 3nitrotyrosine residue
reduces
vasoconstrictive
properties
Protein
fragmentation,
usually at glycine
sites
Dityrosine
Thiol oxidation and 3nitrotyrosine
formation lead to
inhibition
Uncertain
modification leads to
inhibition
Ronchi et al. (1980)
Ducrocq et al.
(1998)
Dean et al. (1997)
and references
therein.
Giulivi and Davies
(1994)
Klebl et al. (1998);
Viner et al. (1996,
1999)
Gantchev and van
Lier (1995)
Aruoma and
Halliwell (1987)
Heme degradation
and inhibition
Uncertain
modification leads to
inhibition
Modification of
tryptophan and
tyrosine residues
Formation of 2oxohistidine from
histidine inactivates
enzyme
Self-inactivation.
Mechanism to be
defined
RNS
3-Nitrotyosine
formation inhibits
enzyme
Hydrogen
peroxide
Hydrogen
peroxide/HRP
Oxidation of essential
thiol causes inhibition
Formation of
dityrosine activates
the enzyme
Halliwell and
Gutteridge (1999)
Stachowiak et al.
(1998)
Lewisch and Levine
(1995) and
references therein
Dean et al. (1997)
Daiber et al.,
(2000); Lin et al.,
(2003); Vernia et
al., (2001)
Halliwell and
Gutteridge (1999)
Verweij et al.
(1982)
Ozone
Formation of
dityrosine activates
the enzyme
WWW.ESAINC.COM
219
Glucose-6-phosphate
dehydrogenase
4-hydroxynonenal
Schiff base/Michael
addition leads to
inactivation
Oxidation of ironsulfur clusters lead to
inactivation
Uchida and
Stadtman (1993)
Glutamine
phosphoribosylpyrophosphate
amidotransferase
Glutamine synthetase
Metal
catalyzed
oxidation
Formation of 2oxohistidine from
histidine in metal
binding region
causes inactivation
Deactivated but only
in the absence of
GSH
Farber and Levine
(1986); Lewisch
and Levine (1995)
and references
therein
Halliwell and
Gutteridge (1999)
Glutathione
peroxidase
Superoxide
4-Hydroxynonenal
Inhibits by binding to
an essential lysine
Bosch-Morell et al.
(1999)
Hypochlorous
acid
Mechanism leading
to inhibition remains
to be clarified
Aruoma and
Halliwell (1987)
Peroxynitrite
Essential selenol
residue oxidized to
selenocysteine
3-nitrotyrosine
formation leads to
inhibition
Oxidation of essential
thiol causes inhibition
Briviba et al.
(1998); Padmaja et
al. (1998)
Francescutti et al.
(1996)
Schiff base/Michael
addition with lysine
inhibits enzyme
Uchida and
Stadtman (1993)
Peroxynitrite
S-nitrosylation of
essential thiol
residues
4-hydroxynonenal
Schiff base/Michael
addition with histidine
Galli et al. (1998);
Mohr et al. (1994);
Souza and Radi
(1998)
Uchida and
Stadtman (1992)
ROS
Dityrosine formation
Giulivi and Davies
(1994)
Ozone
Dityrosine formation
ROS
o-Tyrosine, dityrosine
Verweij et al.
(1982)
Wells-Knecht et al.
ROS
Glutathione reductase
Peroxynitrite
Glyceraldehyde-3phosphate
dehydrogenase
Hydrogen
peroxide
4-hydroxynonenal
Dean et al. (1997)
and references
therein
Halliwell and
Gutteridge (1999)
Hydroxyl free
radical, singlet
oxygen
Insulin
Lens proteins
WWW.ESAINC.COM
220
Lipoxygenase
Low density
lipoprotein
Mn-superoxide
dismutase
Lipid peroxyl
radicals/hydro
peroxides
Peroxynitrite
Hypochlorous
acid
Peroxynitrite
Phosphatidylinositol 3kinase
Peroxynitrite
Prostaglandin
endoperoxide
synthase
Tetranitromethane
Inactivation in the
presence of iron
Nitration and
oxidation. 3nitrotyrosine
formation
α1-Proteinase inhibitor
Hydrogen
peroxide
Leeuwen-burgh et
al. (1997)
Hazen and
Heinecke (1997)
3-Chlorotyrosine
Nitration and
oxidation of critical
tyrosine residues
inactivates enzyme
Nitration of the p85
regulatory subunit of
this enzyme
Inhibition by 3nitrotyrosine
formation
Prostacyclin synthase
Peroxynitrite
(1993)
Cucurou et al.
(1991)
Inhibition by 3nitrotyrosine
formation
Formation of
methionine sulfoxide
from critical
methionine residue
leads to inactivation
MacMillan-Crow et
al. (1996; 1998,
1999b)
Hellberg et al.
(1998)
Goodwin et al.
(1998); Mehl et al.,
(1999); Schmidt et
al., (2003);
Shimokawa et al.
(1990); Zou et al.
(1997, 1999)
Dean et al. (1997)
Hydroxyl free
radical
A variety of amino
acids are damaged
leading to inactivation
Kwon et al. (1990)
Tetranitromethane
Formation of 3nitrotyrosine leads to
inactivation
Kinase cannot act on
-substrates
containing 3nitrotyrosine
Fragmentation and
inhibition
Fetse and Gan
(1981)
Protein kinase
substrates
Peroxynitrite
Ribulose-1,5bisphosphate
carboxylase
Sedoheptulose
bisphosphatase
Serpin (neutrophil
cytosolic serineproteinase inhibitor
Oxygen
species
Superoxide Dismutase
Peroxynitrite
Hydrogen
peroxide
Hydrogen
peroxide
Oxidation of essential
thiol inhibits enzyme
Formation of
methionine sulfoxide
from critical
methionine residue
leads to inactivation
3-Nitrotyrosine
formation and
WWW.ESAINC.COM
Gow et al. (1996a)
Ishida et al. (1999)
Halliwell and
Gutteridge (1999)
Thomas et al.
(1991)
Macmillan-Crow
and Cruthirds
221
inhibition
Surfactant protein A
α-Synuclein
Tyrosine hydroxylase
Peroxynitrite
RNS
Peroxynitrite
3-Nitrotyrosine
3-Nitrotyrosine and
possibly dityrosine
may be causative
or protective factor in
Parkinson’s disease
Nitration and
oxidation products
lead to enzyme
inactivation. This has
now been shown to
be oxidation of an
essential thiol group
(2001); Quijano et
al., (2001)
Greis et al. (1996)
Duda et al. (2000);
Ischiropoulis (2003)
Yamin et al. (2003).
Ischiropoulos et al.
(1995); Kuhn et al.
(1999)
Table 3.8 Oxidation Can Affect Enzyme Activity And Other Protein Function.
The Direct Pathway.
This (post-translational) pathway involves the action of a pro-oxidant on a protein
resulting in modification of amino acid residues, the formation of carbonyl
adducts, cross-linking and polypeptide chain fragmentation. Such changes often
result in altered protein conformation and/or activity.
Dakin first studied protein oxidation and showed that it resulted in the formation
of carbonyl compounds such as carboxylic acids, or aldehydes with the same or
one less carbon atom as the parent amino acid e.g., glycine produces glyoxal,
glyoxylic acid, formaldehyde and formic acid (Dakin (1906, 1908)). This finding
now appears to be true for most amino acids. Consequently, proline and arginine
are converted into glutamic semialdehyde, histidine into 2-oxohistidine, and
lysine into lysyl carbonyl. Stadtman (1990, 1991, 1993) observed findings similar
to Dakin and reported that proteins will produce a variety of carbonyl products
when exposed to metal-based systems (metal/ascorbate and metal/hydrogen
peroxide) in vitro. For example, histidine yields aspartate, asparagine and 2oxoimidazoline, while proline produces glutamate, pyroglutamate, 4hydroxyproline isomers, 2-pyrrolidone and γ-aminobutyric acid (Stadtman (1993)
and references therein). Metal-based systems and other pro-oxidant conditions
can oxidize methionine to its sulfoxide (Brot and Weissbach (1991); Chao et al.
(1997)). Carbonyls can also be formed by the action of hypohalous acids on αamino acids (Chapter 2).
Carbonyl formation is not the only oxidative modification of amino acids and
many other reactions can take place forming a wide variety of modified amino
acid residues including tyrosine adducts and amino acid hydroperoxides (Table
3.9) (Figure 3.14). In the presence of oxygen proteins can undergo radical chain
reactions (Dean et al. (1997) and references therein). Alkoxyl radicals are more
WWW.ESAINC.COM
222
effective at promoting protein “peroxidation” chain reactions, than peroxyl
radicals; the latter play a more important role in lipid peroxidation processes.
Amino acid hydroperoxides once formed can then react with metals producing
free radicals thereby propagating protein chain reactions or they can be reduced
to non-reactive hydroxides.
Residue
Amino Acids
Product
Carbonyls
Arginine
Arginine
Glutamic semialdehyde
5-Hydroxy-2aminovaleric acid
Mono- and di-chlorinated
Cystine, oxy acids
Arginine
Cysteine
Cysteine
Cysteine
Glutamate
Glycine
Histidine
Cysteine/4hydroxynonenal adduct
S-nitrosylation
Glutamic acid
hydroperoxide
Aminomalonic acid
2-Oxohistidine
Methionine
Aspartate, asparagine
Histidine/4hydroxynonenal adduct
Isoleucine hydroperoxides, isoleucine
hydroxides, carbonyl
compounds
Leucine hydroperoxides,
Leucine hydroxides, αketoisocaproic acid,
isovaleric acid,
isovaleraldehyde,
carbonyl compounds
Nε-(carboxymethyl)lysine
2-Aminoadipicsemialdehyde
Lysine/4-hydroxynonenal
adduct
Lysine hydroperoxides,
lysine hydroxides and
carbonyl compounds
Methionine sulfoxide
Phenylalanine
o- and m-Tyrosine
Phenylalanine
4-Nitrophenylalanine
Histidine
Isoleucine
Leucine
Lysine
Lysine
Lysine
Lysine
Reference
Hazen et al. (1996, 1998a,b); Nosworthy and
Allsop (1956); Rowbottom (1995)
Amici et al. (1989); Climent et al. (1989)
Ayala and Cutler (1996a,b)
Zhang et al., (2001)
Armstrong (1990); Takahashi and Goro (1990);
von Sonntag (1990)
Uchida and Stadtman (1994)
Galli et al. (1998); Mohr et al. (1994)
Gebicki and Gebicki (1993); Simpson et al.
(1992)
Copley et al. (1992); Van Buskirk et al. (1984)
Lewisch and Levine (1995); Uchida and
Kawakishi (1990, 1993)
Farber and Levine (1986); Creeth et al. (1983)
Uchida and Stadtman (1994)
Gebicki and Gebicki (1993); Simpson et al.
(1992)
Dean et al. (1996); Fu et al. (1995a); Gebicki
and Gebicki (1993); Simpson et al. (1992);
Stadtman and Berlett (1991)
Dunn et al. (1990); Glomb and Monnier (1995)
Szweda and Stadtman (1992)
Uchida and Stadtman (1994)
Gebicki and Gebicki (1993); Simpson et al.
(1992); Trelstad et al. (1981)
Levine et al. (1996); Li et al. (1995c,d); Vogt
(1995)
Ishimitsu et al. (1986); Kaur and Halliwell
(1994); Liu (1993); Leeuwenburgh et al. (1997);
Nair et al. (1995); Ramezian et al. (1996);
Sontag et al. (1997); van del Vliet (1995)
Huggins et al. (1993); van der Vleit et al. (1994)
WWW.ESAINC.COM
223
Proline
Tyrosine
Cis/trans-4hydroxyproline
γ-Aminobutyric acid
Glutamate,
pyroglutamate
Glutamic semialdehyde
5-Hydroxy-2-amino
valeric acid, proline
hydroperoxides, proline
hydroxides and carbonyl
compounds
2-Pyrrolidone
Protein/4hydroxynonenal adducts
Glycolaldehyde
2-Amino-3-ketobutyric
acid
2-Hydroxypropanal and
acrolein
5- and 6-Nitrotryptophan
N-Formylkynurenine
Kynurenine oxindole3alanine
3-, 4, 5-, 6- and 7Hydroxy-tryptophan
3-Chlorotyrosine
Tyrosine
Tyrosine
L-DOPA
Dityrosine
Tyrosine
3-Nitrotyrosine
Tyrosine
Tyrosine
3,5-Dinitrotyrosine
p-Hydroxyphenylacetaldehyde
Valine hydroperoxides,
valine hydroxides and
carbonyl compounds
Proline
Proline
Proline
Proline
Proline
Protein
Serine
Threonine
Threonine
Tryptophan
Tryptophan
Tryptophan
Valine
Uchida and Kawakishi (1989)
Uchida et al. (1990)
Cooper et al. (1985); Creeth et al. (1983);
Uchida et al. (1990)
Amici et al. (1989)
Ayala and Cutler (1996); Gebicki and Gebicki
(1993); Simpson et al. (1992); Trelstad et al.
(1981)
Uchida et al. (1990)
Uchida and Stadtman (1994)
Anderson et al. (1997)
Taborsky (1973)
Anderson et al. (1997)
Alvarez et al. (1996); Padmaja et al.(1996)
Guptasarma et al. (1992); Maskos et al. (1992);
Neuzil and Stocker (1993)
Armstrong and Swallow (1969); Guptasarma et
al. (1992); Maskos et al. (1992)
Hazen et al. (1996a); Kettle (1996);
Leeuwenburgh et al. (1997)
Gieseg et al. (1993)
Giulivi and Davies (1994); Heinecke et al.
(1993); Huggins et al. (1993); Ischiropoulos et
al. (1992); Leeuwenburgh et al. (1997); van der
Vleit (1995); Vissers and Winterbourne (1991)
Beal et al. (1995); Fukuyama et al. (1996);
Hensley et al. (1997); Kamisaki et al. (1996);
Kaur and Halliwell (1994); Leeuwenburgh et al.
(1997); Maruyama et al. (1996); SalmanTabcheh et al. (1995); Schulz et al. (1995);
Shigenaga et al. (1997); Skinner et al. (1997)
Lin et al., (2000); Yi et al. (1997)
Hazen et al. (1996b)
Fu et al. (1995a,b)
Table 3.9 Many Amino Acids Can Be Modified By ROS, RNS And Other
Reactive Species.
Amino acid residues can also be modified following their reaction with
carbohydrates or other carbonyl compounds that can be produced when lipids,
and proteins are attacked by pro-oxidants. For example, malondialdehyde and 4hydroxynonenal produced by lipid peroxidation can readily form Schiff bases and
WWW.ESAINC.COM
224
H
N
H
O
H2 N
CO2H
R1
R1
N
H
S=O
CH3
H2N H CO2H
2-Oxohistidine
OH
NO2
Methionine Sulfoxide
m-Tyrosine
4-Nitrophenylalanine
R1
R1
R1
R1
R1
OH
OH
NO2
OH
OH
OH
OH
o-Tyrosine
Dityrosine
3-Nitrotyrosine
2,4-Dihydroxy
phenylalanine
R1
R1
O2N
R1
HO
R1
HO
OH
NH
OH
HO
NH
NH
HO
3,4-Dihydroxy
phenylalanine
5-Nitrotryptophan
CO2H
Hydroxytryptophan
Dihydroxytryptophan
O
R
R
R1
R1
NH
O
NH
Oxindolylalanine
Hydroxyhexahydro
pyrroloindolecarboxylic
acid
COOH
HO
CH3
H2N
HOCH2
H2N
O
COOH
OH
O
HN
O
R1
=
NH H
Thymine
tyrosine dimer
CO2H
H2N
O
H2 N
H
CO2H
Glutamic acid
semialdehyde
Lysylcarbonyl
NH
HN
CHO
OHC
NH
CytosineTyrosine dimer
"3"-Hydroxyvaline
H
COOH
OH
O
H
"2"-Hydroxyvaline
H
COOH
CH3
CH3
"1"-Hydroxyvaline
NH2
N
H
COOH
H
Indole derivatives
NH2
H
HOCH2
H
CH3
N-Formylkynurenine/
kynurenine
NH2
NH2
NH
NHR
NH
H2N
O
OH
H2N
Thymine
lysine dimer
NH2
H
CO2H
Figure 3.14 A Variety Of Modifications Can Be Formed When
Amino Acids Or Proteins Are Exposed To ROS, RNS Or Other
Reactive Compounds.
WWW.ESAINC.COM
225
Michael additions with the amine side chain of lysine residues and can lead to
inter- and intra-molecular protein cross-links (Chapter 3 and below) (e.g., Uchida
et al. (1997)). Such modification can lead to enzyme inactivation e.g., glucose-6phosphate dehydrogenase and glyceraldehyde-3-phosphate dehydrogenase are
inhibited by 4-hydroxy-nonenal (Table 3.8).
CO2H
H2N
CO2H
H2 N
CO2H
CO2H
H2N
H2N
CO2H
CO2H
H2N
H2N
CO2H
H2 N
H2N
O
OH
OH
OH
OH
OH
OH
CO2H
OH
DITYROSINE
TRITYROSINE
PULCHEROSINE
CO2H
CO2H
CO2H
CHO
H2 N
H2 N
H2N
H2N
O
CO2H
NO
Cl
OH
OH
3,4-L-DOPA
O2H
Cl
Cl
Br
NO2
OH
OH
CO2H
H2N
H2 N
H2N
H2N
3-NITROSOTYROSINE
CO2H
CO2H
CO2H
CO2H
H2 N
3-CHLOROTYROSINE
4-HYDROXYPHENYLACETALDEHYDE
ISODITYROSINE
OH
OH
O
OH
OH
OH
3-NITROTYROSINE
3-BROMOTYROSINE
3,5-DICHLOROTYROSINE
TYROSINE1-PEROXIDE
CO2H
CO2H
OH
CO2H
HO
H2 N
NH2
NH
O2H
O
TYROSINE3-PEROXIDE
NO2
OH
O
CYCLOTYROSINE
PEROXIDE ADDUCT
3-NITRO-4-HYDROXY
PHENYLACETIC ACID
CO2H
NO2
OH
3-NITRO-4-HYDROXY
PHENYLETHYLAMINE
NO2
OH
3-NITRO-4-HYDROXY
PHENYLLACTIC ACID
Figure 3.15 A Variety Of Modified Tyrosine Residues Can Be
Formed Under Oxidizing Conditions.
Oxidative Damage to Tyrosine.
Protein oxidation can lead to chain fragmentation. Garrison developed his
“peptide α-amidation” pathway to help explain how protein oxidation can result in
polypeptide chain breakage and protein fragmentation (Garrison (1987) and
references therein). Tyrosine is very susceptible to pro-oxidant modification. A
variety of metabolites can be produced depending upon which pro-oxidant is
present and the reaction conditions (Figure 3.15). As many modified tyrosine
WWW.ESAINC.COM
226
residues are currently being used as markers of pro-oxidant activity, it is worth
exploring their production and importance more fully.
3-Nitrotyrosine, both free and protein-bound, are often used as indicators of
increased nitric oxide activity. However, this can be erroneous, as nitric oxide
does not react particularly well with non-radical species (Chapter 2). Nitric oxide
does react extremely rapidly with other radicals e.g., with tyrosyl radical (formed
when tyrosine reacts with oxidants such as the hydroxyl free radical and
peroxynitrite) producing both carbon- and oxygen-nitrosotyrosines.14 Although
the conversion of 3-nitrosotyrosine to 3-nitrotyrosine might possibly be promoted
by ROS, it still remains to be proven. Current evidence suggests that 3nitrotyrosine is a better indicator of peroxynitrite production (Table 2.10; Figure
2.19) but even this has been recently challenged (Pfeiffer and Mayer (1998)).
This situation is further complicated as 3-nitrotyrosine can also be formed by
several other, albeit minor, pathways. These include the reactions between:
•
•
•
•
•
nitrogen dioxide and the tyrosyl radical;
acidified nitrite, hydrogen peroxide and tyrosine;
nitryl chloride and tyrosine;
nitrite, hydrogen peroxide, myeloperoxidase and the tyrosine; and
nitrite, hypochlorous acid and tyrosine (see Chapter 2)
(See Brennan et al., (2002); Dalber et al. (1998); Eiserich et al. (1996);
Halliwell (1997); van der Vliet et al. (1997)).
Thus 3-nitrotyrosine is probably best regarded as a biomarker for nitrating
species in general rather than for any one specific RNS. 3-Nitrotyrosine is now
commonly used as a marker of oxidative stress, its free and bound levels being
increased in a variety of disease states (Tables 3.10). Remember though that
both free and protein-bound 3-nitrotyrosine can react with hypochlorous acid and
this may lead to an underestimation of 3-nitrotyrosine at sites of chronic
inflammation and possibly explain the discrepancies in the level of this analyte
reported in literature (Whiteman and Halliwell (1999b)). Protein-bound levels may
also be underestimated due to the action of protein nitrases (Kamisaki et al.
(1998); Kuo et al. (1999)).
Protein nitration is a fascinating area of research with many questions still yet to
be answered. It is not clear why protein-bound tyrosine residues are nitrated by
peroxynitrite more efficiently than free tyrosine molecules (Crow (1999)). Why is
it that only a small fraction of the total protein pool is susceptible to nitration?
14
This reaction can also take place with an essential tyrosyl radical in the active site of ribonucleotide reductase and
results in inhibition of this enzyme (Lepoivre et al. (1994)). Many enzymes contain such an intrinsic radical essential to
their catalytic process (e.g., on tyrosine, tryptophan, glycine or thiol residues) (Pedersen and Finazzi (1993)). Another
example is pyruvate dehydrogenase which catalyses the conversion of pyruvate to acetyl-CoA by and uses both carbonand sulfur-centered radicals (Halliwell and Gutteridge (1993)). Unfortunately, under some conditions the reaction of
protein radicals with molecular oxygen (and possibly other species) can lead to cleavage of the polypeptide chain,
resulting in enzyme inactivation (Dean et al. (1997) and references therein).
WWW.ESAINC.COM
227
Although many proteins have several tyrosine residues, why are only a few of
them capable of being nitrated (e.g., neurofilament L has twenty tyrosines only
four of which are nitrated; manganese superoxide dismutase has nine tyrosine
residues but only three are nitrated) (Crow (1999) and references therein)?
Interestingly, it takes the nitration of just one, key tyrosine residue to disrupt
function (Crow et al. (1997); MacMillan-Crow et al. (1996, 1998)).
Disease/Condition
Adult respiratory distress syndrome
Aging
Alzheimer’s disease
Amyotrophic lateral sclerosis (sporadic
and familial)
Atherosclerosis
Autoimmune uveitis (experimental)
Bronchopulmonary dysplasia
Carbon monoxide poisoning
Celiac disease
Diabetes
Endotoxemia
Huntington’s disease model
Idiopathic pulmonary fibrosis
Inclusion-body myositis
Inflammation – experimental allergic
encephalomyelitis
Inflammation – myocardial
Inflammation – rheumatoid arthritis
Inflammatory bowel disease
Ischemia
Multiple sclerosis
Organ preservation and transplantation
Organ rejection, acute and chronic
Parkinson’s disease
Perennial nasal allergy
Pneumonia (influenza virus-induced)
Preeclampsia
Septic shock/renal failure
Smoking
Ulcerative colitis
Reference
Haddad et al. (1994); Kooy et al. (1995)
Uttenthal et al. (1998)
Good et al. (1996); Smith et al. (1997); Su et
al. (1997)
Abe et al. (1997); Beal et al. (1997); Bruijn et
al. (1997); Chou et al. (1996a,b); Ferrante et
al. (1997); Toghi et al. (1999)
Beckmann et al. (1994); Buttery et al. (1996)
Wu et al. (1997)
Banks et al. (1998)
Gow et al. (1996b); Ischiropoulos et al.
(1996); Thom et al. (1997)
Ter Steeg et al. (1998)
Suarez-Pinzon et al. (1997)
Kristof et al. (1998);Wizemann et al. (1994)
Beal et al. (1995)
Saleh et al. (1997)
Yang et al. (1996)
Cross et al. (1997); Okuda et al. (1997)
Ishayama et al. (1997); Kooy et al. (1997)
Halliwell (1995); Kaur and Halliwell (1994)
Miller et al. (1995); Singer et al. (1996)
Forman et al. (1998); Ischiropoulos et al.
(1995)
Bagasra et al. (1995)
Skinner et al. (1997)
MacMillan-Crow et al. (1996)
Hantraye et al. (1996)
Sato et al. (1998)
Akaike et al. (1996)
Myatt et al. (1996)
Fukuyama et al. (1997)
Petruzzelli et al. (1997)
Kimura et al. (1998)
Table 3.10 Altered Levels Of 3-Nitrotyrosine Are Found In A Variety Of
Diseases And Conditions.
Although free 3-nitrotyrosine is often regarded as the final product of RNSinduced damage in vivo, this may be true only for certain biological
compartments (e.g., cells). In other locations free 3-nitrotyrosine, whether it is
WWW.ESAINC.COM
228
formed by nitration of free tyrosine or released from protein following proteolysis,
can be further metabolized to 3-nitrophenylacetic acid and 3-nitrophenyllactic
acid. For example, both free 3-nitrotyrosine and its metabolite 3-nitrophenylacetic
acid were found to be elevated in patients with ALS (Beal et al. (1996)). Urine
does not contain any appreciable amount of 3-nitrotyrosine, and only contains its
metabolites (Ohshima et al. (1990, 1991); Shuker et al. (1993); Tabrizi-Fard et al.
(1999); Wishnol et al. (1993)). Urinary levels of 3-nitrophenylacetic acid are
primarily derived from the nitration of circulating 4-hydroxyphenylacetic acid and
not from 3NT metabolism (Mani et al., (2003)). Urinary 3-nitrophenylacetic acid
should not be used as an indicator of 3NT production.
3-Chlorotyrosine is currently being used as an indicator of the production of
reactive chlorine species. Although hypochlorous acid, formed by the action of
myeloperoxidase15 on hydrogen peroxide and chloride, is the main chlorinating
agent produced in vivo, several others (e.g., chlorine radicals, nitryl chloride and
trans-chlorine nitrite) may also be involved, but to a much lesser extent (Chapter
2). Thus, while 3-chlorotyrosine is thought to be a reasonable marker of
hypochlorous acid production, care must be exercised in always assuming that
all chlorination results from endogenous production of hypochlorous acid.16
Elevated 3-chlorotyrosine levels are associated with atherosclerotic lesions and
inflammation (Hazen and Heinecke (1997); Hazen et al. (1997); Heinecke et al.
(1998)).
3-Chlorotyrosine is not the only product when chlorinating species react with
tyrosine as 3,5-dichlorotyrosine and p-hydroxy-phenylacetaldehyde can also be
produced (Figure 2.25). p-hydroxy-phenylacetaldehyde is a very reactive
carbonyl and can readily damage proteins by forming a Schiff base with the εamine moiety of lysine residues (Chapter 2).
Dityrosine (3,3’-bityrosine, m,m’-bityrosine) is formed when two tyrosyl radicals
combine (Figure 2.9) and was first described following the oxidation of tyrosine
with peroxidase and hydrogen peroxide (Gross and Sizer (1959)). This approach
is specific for the production of dityrosine and will not cause the formation of oand m-tyrosine isomers. Dityrosine protein cross-link formation was later found to
occur normally and was shown to be responsible for the insolubility and elastic
properties of some proteins (Aesbach et al. (1976)). Dityrosine and other tyrosine
polymers (e.g., trityrosine and pulcherosine) are also found in lower organisms
where they help to strengthen structural proteins (e.g., in the hardened
fertilization envelope of sea urchins and the cuticle collagen of Ascaris) (Nomura
et al. (1990)).
15
Myeloperoxidase can directly react with tyrosine by two mechanisms. First, it reacts with free tyrosine to form tyrosine
radicals that can then combine to form dityrosine and tyrosine polymers (Jacob et al. (1996)). Secondly, it can form
tyrosine peroxides that are metastable and may contribute to neutrophil- or monocyte-mediated tissue injury (Winterbourn
et al. (1997)).
16
It is estimated that about 2% of hypochlorous acid generated by neutrophils leads to the production of 3-chlorotyrosine
(Kettle (1996)).
WWW.ESAINC.COM
229
Both free and protein-bound dityrosine residues can be formed in reactions
involving the hydroxyl free radical, peroxynitrite, hydrogen peroxide/metal, and
irradiation (UV and gamma), while free dityrosine can also be formed by the
action of certain enzymes (e.g., myeloperoxidase) (Eiserich et al. (1998); Giulivi
and Davies (1994) and references therein; Heinecke et al. (1993); Huggins et al.
(1993); Lymar et al. (1996); Michon et al. (1997); Savenkova et al. (1994);
Sharma and Jane (1998); Winterbourn et al. (1997); Yasmin et al. (1997)). The
production of tyrosyl radicals will not automatically lead to protein dityrosine
formation. Protein dityrosine cross-links will only be produced if two tyrosyl
radicals are allowed to encounter and not be formed if protein tyrosyl radicals are
located too far apart from each other.
Dityrosine confers stability to a protein making it less susceptible to proteolysis
and acid hydrolysis and, as it does not undergo further metabolism can be used
as a quantitative index of protein oxidation. Free dityrosine is more reactive and
may, under special circumstances be further metabolized. For example, it is
readily oxidized by both compounds I and II of peroxidases (rate constant with
compound I is 1 x 105 M-1s-1) producing trityrosine, polytyrosine and other
oxidation products (Marquez and Dunford (1995) and references therein).
Both free and protein-bound dityrosine are found in a variety of human tissues.
Changes in their levels are currently being used as a marker for metal catalyzed
oxidation in vivo and in vitro, as a measure of total index of oxidative stress and
as an indicator of oxidative damage involving phagocytes (Giulivi and Davies
(1993); Heinecke et al. (1993); Huggins et al. (1993); Leeuwenburgh et al.(1999);
Salman-Tabcheh et al. (1993)). For example, the level of protein bound
dityrosine is markedly elevated in LDL isolated from human atherosclerotic
lesions, compared to circulating LDL levels (Leeuwenburgh et al. (1997a)).
Protein bound dityrosine is found to be increased in aging (e.g., it is abundant in
lipofuscin granules) while free dityrosine can be formed following ischemiareperfusion (Abdelrahim et al. (1997); Kato et al. (1998); Wells-Knecht et al.
(1993); Yasmin et al. (1997)). Interestingly, dityrosine cross-linking in cardiac and
skeletal muscle in aging rats is attenuated by caloric restriction (Leeuwenburgh
et al. (1997b)).
Protein Repair and Degradation.
Proteins, like DNA, can be rendered non-functional following damage. Although
affected proteins can be replaced by de novo synthesis this is energetically
expensive so cells have developed repair mechanisms including:
•
•
The restoration of a protein into its correct, active conformation (see the
section on chaperones above);
Enzymatic repair to directly reverse some forms of amino acid residue
damage. Such mechanisms are capable of repairing proline isomerization,
WWW.ESAINC.COM
230
and reversing isoaspartyl and methionine sulfoxide formation (Visick and
Clarke (1995)). Two processes are important in protecting methionine.
First, naturally occurring antioxidants can reduce the methionine
intermediate that is initially formed when methionine is damaged by prooxidant species. Second, the enzyme methionine sulfoxide reductase can
reform methionine from methionine sulfoxide in a process that probably
uses NADPH and thioreductase. Methionine sulfoxide reductase plays an
important role in reactivating the oxidized α1-proteinase inhibitor and
preventing the formation of methionine sulfoxide in the lens of the eye.
Excessive oxidative damage to the lens may be one of the many
processes that can overwhelm the eye’s antioxidant defenses eventually
leading to cataract formation (cataracts are found to contain significant
amounts of methionine sulfoxide).
The cyclic process of methionine sulfoxide formation and methionine
regeneration has led to the suggestion that methionine residues located
on the surface of a protein may constitute an important antioxidant
defense mechanism protecting the protein from more harmful oxidation
(Levine et al. (1996)).
Damaged proteins that cannot be repaired undergo proteolysis (Visick and
Clarke (1995)).
In normal human subjects about 300g of tissue protein is catabolized daily and
replaced by newly synthesized protein. Since six ATP molecules are used for
each amino acid residue added to the growing polypeptide chain, this turnover,
accounting for 15-20% of the basal metabolic rate, is energetically expensive.
Protein turnover is biologically important and can vary enormously from protein to
protein. Proteins with especially short half-lives include enzymes that are
important in regulating metabolic pathways (e.g., hepatic phosphoenol-pyruvate
carboxykinase). Changes in the rate of synthesis of a regulatory enzyme will
rapidly alter its concentration and hence the flux through the pathway. Rapid
degradation not only allows control of metabolic flux, but also prevents any
chance of the enzyme being reactivated inappropriately. This explains why
selective protein degradation always plays an important regulatory role in timing
controls (e.g., cell cycle progression and various signal transduction pathways).
A second group of proteins with short half-lives are the abnormal proteins
resulting from errors in translation, and oxidative damage (including conformation
changes and oxidation of amino acid residues). The rate of hydrolysis of
abnormal proteins is dependent upon the amount of oxidation present. Proteins
with limited oxidation are degraded at a greater rate than those that are more
markedly damaged probably a consequence of marked changes in a protein’s
structure rendering it poorly digestible.
WWW.ESAINC.COM
231
AMP + PPi
E1-SH + ATP
E2-SH
Ub-CO2S-E1
E1-SH
Ub-CO2-
nx
Ub-CO2S-E2
E3
DUB
Protein
Ub-Protein
Protein
E2-SH
Poly-Ub-Protein
26S Proteasome
ATP
Ub
Oxidized Protein
20S Proteasome
PA700
Peptides
Oxidized
Peptides
Figure 3.16 Degradation Of Normal Proteins By The Ubiquitin-Proteasome
(26S) System. Oxidatively Damaged Proteins Are Not Processed By The
26S Proteosome But Rather By Its Proteolytic Core, The 20S Proteosome.
This Process Does Not Require Protein Ubiquitinylation Prior To
Degradation. (DUB - deubiquitination enzymes; E2 - ubiquitin conjugating enzyme; E3 ubiquitin-protein ligase; Ub – ubiquitin).
There are at least two pathways for protein degradation in eukaryotic cells,
lysosomal and non-lysosomal. Lysosomes contain at least four proteinases
(including cathepsins B, D, and E) and several peptidases (e.g., dipeptidyl
peptidases) permitting complete protein degradation within this organelle (Bohley
and Seglen (1992); Dean (1979); Dice and Terlecky (1990)). Lysosomes can
also form autophagic vacuoles capable of engulfing and digesting whole
organelles such as mitochondria. The non-lysosomal pathway is the most
important proteolytic pathway for many short-lived proteins and involves tagging
WWW.ESAINC.COM
232
them with ubiquitin prior to digestion by multisubunit complexes (the
proteasomes) located in the cytosol and nucleus (Couz et al. (1996); Hilt and
Wolf (1996)).
Ubiquitin, a small protein (~8.5kDa) found in all eukaryotic cells, plays a wide
variety of regulatory roles including gene expression, ribosome biosynthesis,
receptor expression and ubiquitin-mediated proteolysis (Hochstrasser (1996);
Schwartz and Ciechanover (1992)). The tagging of a protein by ubiquitin requires
ATP hydrolysis to form an isopeptide bond between the ε-amino group of a lysine
residue on the protein and the carboxyl terminal glycine of ubiquitin. Three
enzymes are involved in this process. The first enzyme, E1, activates ubiquitin in
a process requiring ATP. The second enzyme (ubiquitin conjugating enzyme, E2)
takes activated ubiquitin from E1 and transfers it to a damaged protein in a
process requiring the third enzyme (ubiquitin-protein ligase) E3 (Haas and
Siepmann (1997); Hochstrasser (1995)) (Figure 3.16). A protein tagged for
destruction usually acquires several molecules of ubiquitin. How a damaged
protein is recognized by the ubiquitination system is unclear but may be
controlled, in part, by chaperones (Benjamin and McMillan (1998); Raboy et al.
(1991); Sherman and Golberg (1996)). The role of deubiquitination (DUBs)
enzymes in regulation of protein turnover is still being evaluated (Hochstrasser
(1995); Wilkinson (1997)).
Ubiquinated proteins are degraded by the 26S proteasome complex (Figure
3.16) located in the cytoplasm, nucleus and endoplasmic reticulum, but not in the
mitochondria. This complex is composed of a core proteinase known as the 20S
proteasome and a pair of regulatory complexes (multisubunit proteins [ATPases]
known as protein activators or PA700s) that are attached to both ends of the
complex (Driscoll (1994); Tanaka (1998)). Although the exact steps in protein
degradation are not fully known they include: binding of the multi-ubiquinated
protein by its ubiquitin chains to the chain-binding subunits of PA700; a series of
ATP-dependent unfolding and translocation steps that feed the unfolded protein
into the central channel of the 20S proteasome; cleavage of substrate into small
peptides; and finally disassembly of the ubiquitin chains, that can then be reused
(based on Hochstrasser (1995)). Dysfunction of the ubiquitin-proteosome
pathway has been implicated in the pathogenesis of several human diseases
including cystic fibrosis, Angelman’s syndrome, and neurodegeneration (Scwartz
and Ciechanover (1999); Alves-Rodrigues et al. (1998)).
The turnover rate of a normal protein appears to be determined, in part, by its
amino-terminal residue. Proteins can be categorized into three groups depending
upon whether the amino terminal is stabilizing (half-life >20 hours; alanine,
glycine, methionine, serine, threonine and valine), destabilizing (half-life 7-30
minutes; glutamate, glutamine, isoleucine, proline and tyrosine) or highly
destabilizing (half-life <3 minutes; arginine, aspartate, leucine, lysine and
phenylalanine) (Varshavsky (1997)). The exact mechanism driving the rate of
WWW.ESAINC.COM
233
ubiquitination of these proteins remains elusive but research is centered upon the
E3 enzyme.
Oxidatively modified proteins are mainly degraded by the 20S proteosome
located in the cytosol, while oxidatively damaged soluble histones and DNAbound histones are catabolized by the 20S proteosome located in the nucleus.
Both of these are ATP- and ubiquitin-independent processes (Grune and Davies
(1997); Grune et al. (1997); Ullrich et al. (1999)). In this way the association of
damaged protein, through cross-linking and increased surface hydrophobicity,
into potentially lethal protein aggregates is prevented. The 20S proteosome
recognizes increased protein surface hydrophobicity (aromatic residues and
bulky aliphatic residues) caused by changes in a protein’s secondary and tertiary
structure caused by pro-oxidant damage (Grune et al. (1997); Pacifici et al.
(1993)). The activity of 20S proteosome is greatly affected by the amount of
protein damage. While the complex has little problem dealing with moderately
damaged proteins, extensive protein damage can lead to inhibition of the
complex and the build up of modified protein (Grune et al. (1998); Ullrich et al.
(1999)).
Amino Acid and Protein Damage in Aging and Disease.
An open question in the field of aging is whether protein oxidation is an important
aspect of aging or whether it is just one consequence. Abundant evidence shows
that protein oxidation products such as protein carbonyls and protein-containing
age pigments (e.g., lipofuscin) do accumulate with age (Halliwell and Gutteridge
(1999); Stadtman (1988)). This is especially true for long-lived proteins, such as
those in the lens, where oxidized proteins accumulate over time. Aging is also
accompanied by a decrease in the activity of key metabolic enzymes such as
glutamine synthetase, glucose-6-phosphate dehydrogenase and cytosolic neutral
protease activity. Unfortunately, there is no direct evidence that altered activity is
a consequence of protein oxidation. Treatment of rats with the spin-trap agent,
PBN, was found to prevent the age-related increase in protein carbonyl
production, loss of enzyme activity and loss of behavioral performance (Carney
et al. (1991); Stadtman et al. (1992) and references therein). Although these
findings are encouraging, Dean et al. (1997) have suggested that the levels of
PBN used were too low to have any antioxidant effects. Transgenic Drosophila
overexpressing catalase and superoxide dismutase lived longer and were more
active than those overexpressing just one of these enzymes (Orr and Sohal
(1994); Sohal et al. (1995)).
The increased pool of damaged protein seen with aging can be explained either
by an overproduction of oxidized protein overwhelming the proteolytic process
and/or decreased activity of these enzymes. The latter can occur at several
levels, including damage to genes encoding proteolytic enzymes (DNA damage
also accumulates with aging), damage to the proteolytic enzymes themselves,
WWW.ESAINC.COM
234
and oxidation-induced changes in substrate rendering it less susceptible to
proteolytic attack. However, as the precise mechanisms governing proteolytic
activity remain unresolved, it may be too soon to link changes in these systems
to the accumulation of oxidized protein in aging (Stadtman (1992)).
A variety of diseases including atherosclerosis, cataracts, diabetes, inflammation
and neurodegeneration are also associated with increased protein oxidation.
These have been reviewed elsewhere (Dean et al. (1997); Halliwell and
Gutteridge (1999)). Altered levels of one protein oxidation “marker”, 3nitrotyrosine, has been reported to be increased in a variety of diseases and
conditions (Table 3.10). A selection of potential markers and the effect of disease
on their levels are presented in Table 4.11. The role of glycation and
glycoxidation reactions in diabetes is discussed in greater detail below.
Analyte
3,4-L-DOPA
Species
Human
Tissue
LDL protein
3,4-L-DOPA
Rat
Glial cells inculture
Carbonyls
(protein bound)
Human
Mixed
tissues
3-Chlorotyrosine
(protein bound)
Dityrosine
(Free)
Human
Aorta
Human
Ventricular
fluid
Level
6 adducts/104 Tyr control
14/104 – atherosclerotic
0 – control
1 adduct/103 Tyr after
interleukin 1β treatment
1nmol/mg protein – control
<8nmol/mg – diseased brain
tissue
0.8 adducts/104 Tyr control
4.2/104 – atherosclerotic
3.5 adducts/103 Tyr control
12/103 – Alzheimer’s
0.2 adducts/103 Tyr control
~3/103 – Alzheimer’s
1-3 adducts/106 Tyr
2 adducts/106 Tyr –
control
5/104 – plaque
5 adducts/105 Tyr – control
(Protein bound)
Dityrosine
(bound)
Dityrosine
(bound)
Human
Hippocampus
Lens protein
Human
LDL protein
Dityrosine
(bound)
Rats
Cat
Human
Protein
(mito)
Protein
(cytosolic)
Urine
Urine
Urine
Human
CSF
Human
Plasma
1.4+0.7nmol/L – control
9.0+0.7nmol/L – ALS
11.4+5.4 nmol/L – AD
31+6 nmol/L
Human
Brain - gray
0.285+0.26 to 0.959+0.02
(free)
Dityrosine
3-Nitro-4hydroxyphenylacetic
acid
3-Nitrotyrosine
(free)
3-Nitrotyrosine
(free)
3-Nitrotyrosine
7 adducts/105 Tyr – control
0.5nmol/mmol creatinine
3289-11,803ng/day
0-7.9 µg/24hr
WWW.ESAINC.COM
Reference
Dean et al. (1997)
Hensley et al.
(1997)
Levine et al.
(1994); Lyras et al.
(1996)
Leeuwenburgh et
al. (1997)
Hensley et al.
(1998)
Wells-Knecht et al.
(1993)
Dean et al. (1997)
Leeuwenburgh et
al. (1999)
Marvin et al. (2003)
Ohshima et al.
(1990,1991)
Tohgi et al.
(1999a,b)
Kamisaki et al.
(1996)
Maruyama et al.
235
(free)
nmol/g
Brain - white
3-Nitrotyrosine
(free)
Human
Serum
3-Nitrotyrosine
(free)
3-Nitrotyrosine
(free)
3-Nitrotyrosine
(free)
Human
Human
Synovial
fluid
Plasma
Human
Urine
3-Nitrotyrosine
(free)
3-Nitrotyrosine
(free)
Mouse
Brain
Mouse
Brain
3-Nitrotyrosine
(free)
3-Nitrotyrosine
3-Nitrotyrosine
(protein bound)
Mouse
Spinal cord
Cat
Human
Urine
Plasma
proteins
3-Nitrotyrosine
(protein bound)
Human
3-Nitrotyrosine
(protein bound)
Human
Polymorphonuclear
leukocyte
proteins
LDL
3-Nitrotyrosine
(protein bound)
Human
LDL
3-Nitrotyrosine
(protein bound)
Human
Plasma
protein
Leukocyte
protein
3-Nitrotyrosine
(protein bound)
Rat
Plasma
proteins
3-Nitrotyrosine
(protein bound)
Rat
3-Nitrotyrosine
Human
Peritoneal
exudate
proteins
Plasma
0.276+0.25 to
0.962+0.02 nmol/g
0 – Control
0.18+0.07 to 0.49+0.27
µmol/L –
arthritis
0 – Control
0.49+0.26 µmol/L – arthritis
n.d. – control
28+12 µmol/L – renal failure
0 – control
0 to 5.8 µg 3-nitro-4hydroxyphenylacetic acid/
24hr – control
0 to 7.9 µg 3-nitro-4hydroxyphenylacetic acid/
24hr – smokers
2.0+0.1ng/mg protein
3.0+0.5 adducts/103 Tyr –
control
6.0+1.5 – SOD transgenic
20+5 adducts/103 Tyr –
control
<58ng/day
7+1.2 adducts/103 Tyr –
control
12.2+1.4 – stimulated
0 – control
21.3+1.2 adducts/103 Tyr –
stimulated
<10pmol/mg LDL protein –
control
<10pmol/mg – plaque
9+7 µmol/mol Tyr – control
840+140 – atherosclerotic
intima
7+1 adducts/103 Tyr –
control
12+1 – phorbol ester
stimulated
0 – control
14+1 – phorbol ester
stimulated
0.37+0.32 adducts/106 Tyr –
control
12.46+3.13 – stimulated
0 – control
14.11+2.33 adducts/106 Tyr
– stimulated
WWW.ESAINC.COM
(1996)
Kaur and Halliwell
(1994)
Kaur and Halliwell
(1994)
Fukuyama et al.
(1996)
Ohshima et al.
(1990)
Schulz et al. (1995)
Beal et al. (1995)
Bruijn et al. (1997)
Marvin et al. (2003)
Salman-Tabcheh et
al. (1995)
Salman-Tabcheh et
al. (1995)
Dean et al. (1997)
Leeuwenburgh et
al. (1997)
Salmen-Tabcheh et
al. (1995)
Shigenaga et al.
(1997)
Shigenaga et al.
(1997)
Skinner et al.
236
Free
Protein
bound
3-Nitrotyrosine
Free
Protein
bound
3-Nitrotyrosine
Free
Protein
bound
3-Nitrotyrosine
Protein
bound
o-Tyrosine
protein bound
o-Tyrosine
protein bound
m-Tyrosine
protein bound
0
2.3 adducts/106 Tyr
Human
Rat
Ventricular
fluid
2 adducts/103 Tyr control
4/103 Alzheimer’s
Hippocampus
Liver
0.2 adducts/103 Tyr control
~1.5/103 Alzheimer’s
15.7+0.3 adducts/106 Tyr
(1997)
Hensley et al.
(1998)
Skinner et al.
(1997)
9.5+1.1
Rat
Glial cells inculture
<0.2 adducts/103 Tyr
Hensley et al.
(1997)
Human
Lens
Human
LDL
0.3 to 0.9 adducts/103
phenylalanine
62 and 35pmol/mg protein –
control
105 and 175pmol/mg protein
– plaques
Wells-Knecht et al.
(1993)
Dean et al. (1997)
o-Tyrosine
protein bound
Rat
Cat
0.7 adducts/103
phenylalanine
0.6 adducts/103
phenylalanine
157-250ng/day
Leeuwenburgh et
al. (1999)
o-Tyrosine
Protein
(mito)
Protein
(cytosolic)
Urine
Marvin et al. (2003)
Table 3.11 A Selection Of Reports Measuring Amino Acid And Protein
Oxidation Markers. AD – Alzheimer’s Disease; ALS – Amyotrophic Lateral
Sclerosis.
Measurement of Amino Acid and Protein Damage.
Many modified amino acids can be formed during oxidation processes (Figure
3.14) but protein carbonyls and modified tyrosine residues have garnered most
attention. The analytical procedures used to measure protein oxidative damage
tend to fall into two categories – those that use whole proteins and those that
measure amino acid residues following protein hydrolysis.
Whole Protein.
A variety of techniques can be used to measure amino acid modifications in
whole protein either in situ or following isolation (e.g., Viera et al. (1999)). For
isolated proteins the choice of technique is dependent upon the purity of the
sample. For relatively clean samples (and for in vitro studies using purified
proteins) the abundance of some modified residues can be determined using UV
WWW.ESAINC.COM
237
detection. However, this approach is severely limited as only a few amino acid
residues show UV absorbance. Fluorescence is often used as an indirect
measure of protein damage (Jones and Lunec (1987)). Oxidative changes in
tyrosine, tryptophan and cysteine residues are associated with protein
aggregation and the induction of a characteristic fluorescence (excitation 360nm,
emission 454nm). Although it is still unclear exactly which modifications are being
measured, this technique is being used to study the role of ROS/RNS-induced
protein modification in diseases such as diabetes and arthritis (Jones and Lunec
(1987)).
Some analytical approaches require a degree of sample preparation before the
amount of protein damage can be quantified e.g., the use of polyclonal
antibodies. First, proteins in complex biological samples can be immobilized on
nitrocellulose and extensively washed prior to detection. The immobilized protein
can then be exposed to polyclonal (or monoclonal) antibodies raised to a specific
modified residue. Subsequent exposure to radiolabeled (or fluorogenic labeled)
immunoglobulin G permits the measurement of oxidatively modified protein using
beta scanning (or fluorescence scanning) (Crow and Ischiropoulos (1996); Ye et
al. (1996)). This approach is sensitive and fairly selective but only measures
total, not individual protein modifications.
To examine which specific proteins are being modified, more advanced
separation methods must be used. One-dimensional electrophoresis using a
sodium dodecyl sulfate-polyacrylamide gel (SDS-PAGE) can separate thiolreduced proteins based on their relative masses. Protein bands in the gel can
then be visualized using Coomassie blue or silver stain. This approach is quick,
sensitive (about 0.1µg with Coomassie blue and 0.02µg with silver stain) and can
distinguish between proteins differing by only 2% of their mass. Some proteins
such as glycoproteins and membrane proteins, however, can migrate
anomalously. Specific modified residues can be determined following Western
blotting and exposure of the resulting blot to radiolabeled or fluorogenically
labeled antibodies specifically raised to the modified residue of interest. Boundlabel can then be visualized using autoradiography, beta scanning or
fluorescence scanning.
WWW.ESAINC.COM
238
Figure 3.17 A Two-Dimensional Gel Showing Resolution Of Many
Proteins In A Rat Fibroblast Lysate. I would like to acknowledge Dr. M.
Lopez for supplying this gel image.
Perhaps the best approach to study the proteome is two-dimensional
electrophoresis (see Lopez (1997, 1998a, 1998b) and references therein). Here
proteins are initially separated based upon their charge (isoelectric focusing) and
then in the second dimension on their molecular weight (SDS-PAGE). A typical
two-dimensional gel is shown in Figure 3.17. Protein spots can be visualized
using different stains (see above). Following blotting, individual protein spots can
then be further characterized. For example, the protein sequence can be
determined using tryptic digest followed by HPLC with Edman degradation
chemistry. Protein mass can be measured using matrix-assisted laser desorption
ionizing time of flight mass spectrometry (MALDI-tof-MS). Finally, modified
residues can be determined using the antibody-based methods described above.
WWW.ESAINC.COM
239
Protein Hydrolysates.
This method is used to liberate residues from the modified protein prior to
analysis by a variety of analytical techniques such as GC- and HPLC-based
approaches. Two broad hydrolytic methods are used – acidic and enzymatic
hydrolysis. Both procedures are currently being used although neither approach
is perfect. Like with DNA adduct measurement described above, I urge
researchers to pay careful consideration to the isolation and hydrolysis
procedures.
Acid hydrolysis typically involves heating the lyophilized protein under vacuo at
110oC in 6N hydrochloric acid for 12-24hr. Phenol and/or benzoic acid (~0.1-1%)
are typically included to prevent artifactual generation of tyrosine adducts
(Heinecke et al. (1998); Kettle (1998)). The advantage of acid hydrolysis is that it
is straightforward and the protein is fully hydrolyzed to individual residues.
Unfortunately acid hydrolysis suffers from several disadvantages. The protein
must be extensively washed prior to hydrolysis in order to remove nitrite, nitrate
and chloride ions. Under acidic conditions these can cause artifactual formation
of tyrosine adducts (Heinecke et al. (1998); Kettle (1998); Shigenaga (1999);
Shigenaga et al. (1997) and references therein). The acid used for hydrolysis
must also be devoid of contaminating nitrite, nitrate and chloride. If hydrochloric
acid is used for protein hydrolysis then a strong vacuum must be maintained
during hydrolysis to avoid artifactual generation of 3-chlorotyrosine. This can be
avoided by using hydrobromic acid. Hydrobromic acid, however, is unsuitable for
measurement of bromo-tyrosine adducts. For the routine analysis of halogenated
tyrosine residues methane sulfonic acid or other non-halogenated volatile acids
are perhaps the best choice.
Another major problem with acid hydrolysis is that this process can destroy
tyrosine residues, thereby affecting the tyrosine adduct/tyrosine ratio.
Furthermore, if acid hydrolyzed protein is to be analyzed using HPLC, the pH of
the hydrolysate must be buffered so as not to expose the analytical column to
detrimental acidic pH conditions. An alternative approach is to use a volatile acid
that can be removed under a stream of air or nitrogen (Hazen (1998)).
With enzymatic hydrolysis a protein sample is typically incubated with a
proteolytic enzyme (e.g., proteinase K or pronase E) at 50oC for 12-16hr. This
approach avoids the problems of acid hydrolysis but has several issues of its
own. Enzymatic hydrolysis may not go to completion, producing tyrosine adductcontaining peptide fragments. Some proteolytic enzymes contain both tyrosine
and 3-nitrotyrosine that can be liberated upon autodigestion. Care must be
exercised in the correct choice and source of enzyme. It is also recommended
that enzyme be extensively dialyzed before use (Shigenaga et al. (1997)).
WWW.ESAINC.COM
240
Measurement of Free Modified Amino Acids and Modified Residues
in Whole Proteins and Protein Hydrolysates.
1. Protein Carbonyls.
Measurement of protein carbonyls is a commonly used method to measure
oxidative damage to proteins. Protein carbonyls are usually determined using
Schiff-base conjugation with 2,4-dinitrophenylhydrazine (DNPH) followed by
spectrophotometric, HPLC-UV, or immunochemical techniques (Ayene et al.
(1993); Fung and Grosjean (1981); Harris et al. (1994, 1995); Hensley et al.
(1995); Legler et al. (1985); Levine et al. (1994); Oliver et al. (1987); Smith et al.
(1991); Winterbourne and Buss (1998)). The limit of detection for the HPLC-UV
approach is typically 100pmol on column but this may not be sufficient to
measure the low carbonyl levels typically found under basal and even some
pathological conditions.17 Typical tissue levels vary from 0.73+0.63 nmol/mg
(human lumbar controls) to >4.0 nmol/mg (human brain) (see Evans et al., (1998)
and references therein. Unfortunately, the DNPH approach cannot effectively
distinguish between protein oxidation and post-translational modifications such
as nonenzymatic glycation. Furthermore, processes not involving oxidative
damage can also form protein carbonyls (Cao and Cutler (1995)). For example,
α,β-unsaturated alkenals formed during lipid peroxidation can react with protein
thiols forming stable covalent thioether adducts carrying carbonyl groups. The
formation of Schiff bases between a lysine residue and a reducing sugar may,
upon Amadori rearrangement, also yield carbonyl-containing ketamine protein
conjugates. Protein carbonyl measurement, its limitations and issues, is critically
reviewed by Evans et al. (1998)).
2. Methionine sulfoxide.
Methionine sulfoxide can be measured in whole protein using 13C NMR (Cohen
et al. (1979)) and electrophoretic methods (Amiconi et al. (1985) or in hydrolyzed
protein using GC- and HPLC-based approaches (Chao et al. (1997); Maier et al.
(1995)).
3. 2-Oxohistidine.
2-Oxohistidine (2-imidazolone) in proteins can be determined using automated
Edman protein sequencing and mass spectrometry or in hydrolyzed proteins
using HPLC-ECD and HPLC-fluorescence of the OPA derivative (Lewisch and
Levine (1995, 1998); Uchida and Kawakishi (1993)). Due care must be exercised
during acid hydrolysis of proteins as 2-oxohistidine is unstable and will
17
HPLC-ECD can give lower limits of detection for free carbonyls but the application of this approach to whole proteins is
limited (Chiavari and Bergamini (1985); Goldring et al. (1993)).
WWW.ESAINC.COM
241
decompose forming aspartate, ammonia and other products (Lewisch and Levine
(1998)). This can be prevented by inclusion of the reducing agent dithiothreitol
during sample processing.
4. Tyrosine Markers.
Many analytical procedures are used to measure both free and protein-bound
modified tyrosine residues. The extent of protein modification can be measured
in situ, in whole protein or protein hydrolysates (Table 3.11). Protein hydrolysis is,
however, fraught with methodological problems that can lead to artifactual
production of modified tyrosine residues (see below). Out of all the oxidized
residues that can be formed the measurement of modified tyrosine residues is
probably one of the most common. This is due partly to the fact that they are
considered to be “global” reporter molecules, capable of forming different
products with ROS, RNS and oxidizing chlorine species, and partly because their
measurement is relatively straightforward.
3-Nitrotyrosine.
The extent of protein nitration can be determined in situ using immunohistological
approaches on frozen and fixed tissues (e.g., Viera et al (1999)). Measurement
of nitration of whole proteins is difficult to determine quantitatively. Current
methods use immunochemical or UV detection (Beckman et al. (1994); Crow and
Beckman (1995); Crow and Ischiropoulos (1996); MacMillan-Crow et al. (1999);
Salman-Tabcheh et al. (1995); Viera et al. (1999); Ye et al. (1996)).
Immunochemical methods are generally limited by antibody quality and
visualization methods, and are often poorly reproducible, cumbersome, costly,
suffer from matrix effects and slow throughput (Hensley et al. (1997); Viera et al.
(1999)). Direct UV approaches are limited to relatively pure samples and are
insensitive (~1µmol on column). Due to chromatographic issues, HPLC-UV
detection is best performed on protein-tryptic digests (the limit of detection is
~0.1nmol on column).
Protein-bound 3-nitrotyrosine is more conveniently measured following
hydrolysis. 3-Nitrotyrosine, whether free or from hydrolyzed proteins, can be
measured using a variety of analytical methods including GC-thermal energy
analysis, GC-MS, LC-MS, and HPLC-based approaches (Althaus et al., (2000);
Crowley et al. (1998); Greis et al. (1996); Herce-Pagliai et al. (1998);
Leeuwenburgh et al. (1998); Ohshima et al. (1990); van der Vleit (1999); Yi et al.
(1997)) (Table 3.11). Of all the HPLC-based techniques presented in Table 3.12,
HPLC-UV is too insensitive for most tissue work and as 3-nitrotyrosine is not
fluorogenic it must be converted to a fluorophore for HPLC-fluorescence
analysis. This can be achieved by reducing 3-nitrotyrosine chemically to 3-
WWW.ESAINC.COM
242
aminotyrosine (using sodium dithionite18 [sodium hydrosulfite] or sodium
borohydride) (Sokolovsky et al. (1967)) which is fluorescent, or by derivatizing it
with a fluorogenic agent prior to HPLC separation.
HPLC Procedure
UV
Comments
Sensitivity may be a problem especially for
basal adduct measurement.
Fluorescence
3-Nitrotyrosine is chemically reduced to
3-aminotyrosine using dithionite. 3-aminotyrosine can be measured directly with
fluorescence.
Measured following derivatization with
phenylisothiocyanate.
Electrochemical dual amperometric
electrode detection
Electrochemical single amperometric
electrode detection,
Measured as the 4-fluoro-7-nitrobenzo-2-oxa1,3-diazole derivative. Requires precolumn
derivatization. Limit of detection 22pg on
column. In samples 3NT appears to elute in a
crowded area of the chromatogram.
Upstream electrode reduces 3NT to 3-aminotyrosine that can then be detected at a lower
oxidative potential (than is required for the
measurement of 3-nitrotyrosine) at the down
stream electrode. The reductive potential of –
2000mV in the presence of oxygen will
generate high currents that will severely
damage the working electrode. Reduction
efficiency may vary over time. Although this
approach can be used to measure in vitro
protein nitration, it may be unsuitable for
measurement of basal tissue and protein
levels in vivo. This approach cannot be used
to measure other tyrosine adducts (only 3nitrotyrosine can be reduced at the upstream
electrode) unless a higher potential is applied
to the downstream electrode.
3-Nitrotyrosine is measured directly at
1000mV on a single glassy carbon working
electrode. Chromatographic issues (the
References
Althaus et al.
(1997); Crow and
Ischiropoulos
(1996); Kaur and
Halliwell (1994);
Salman-Tabcheh et
al. (1995); van der
Vleit et al. (1995,
1996)
Crow and
Ischiropoulos
(1996)
Ischiropoulos and
Al-Mehdi (1995)
Kamisaki et al.
(1996)
Althaus et al.
(1997)
Kaur et al. (1998)
18
It should be remembered that dithionite reduction is very sensitive to pH, should preferably be used buffered, and must
always be used in excess.
WWW.ESAINC.COM
243
oxidative
Electrochemical amperometric or
coulometric detection,
reductive
Electrochemical amperometric or
coulometric detection,
OPA/βME derivatization
Electrochemical photolysis followed by
amperometric detection
Electrochemical dual coulometric
electrode detection,
oxidative
Electrochemical dual coulometric
electrode detection,
oxidative with on-line
chemical reduction
Electrochemical coulometric electrode
array detection
authors report that several endogenous
compounds in brain samples co-elute with the
3NT peak) makes the reliable detection of
3-nitrotyrosine challenging.
Although 3-nitrotyrosine can be measured
using reductive potentials, this approach is not
to be recommended. Unless oxygen is totally
removed from the system 1) excessive noise
makes routine measurement of 3-nitrotyrosine
difficult, and 2) excessive current will limit the
life of the working electrode.
Pre-column derivatization of amino acid with
OPA/βME is often used to render inert amino
acids electrochemically active. Tyrosine (and
its derivatives) is already electrochemically
active and no increase in sensitivity is found
upon derivatization.
A novel approach using a stroboscopic
photolytic unit to convert 3-nitrotyrosine to
L-DOPA is described. Generated L-DOPA is
detected on amperometric working electrodes
placed downstream from the photolytic unit.
Although this approach may prove useful for
measurement of higher levels of 3NT, its
applicability to biological samples is not clear.
Photolytic units usually suffer from UVinduced fragility of the reactor coil and this still
needs to be evaluated. The dead-volume of
the reactor coil can also compromise
chromatography.
Direct measurement at +750mV on the
downstream electrode, while the upstream
electrode removes contaminants at +500mV.
Anon*
Anon*
Liu et al. (1998a,b)
Maruyama et al.
(1996)
Direct measurement at +850mV on the
downstream electrode, while the upstream
electrode removes contaminants at +600mV.
Skinner et al.
(1997)
Extensive sample preparation permits
sensitive measurement of 3-nitrotyrosine as
its N-acetyl-3-aminotyrosine derivative.
Shigenaga (1999);
Shigenaga et al.
(1997)
Low picogram levels are typically measured
using these approaches.
An in-line Jone’s reductor placed prior to the
analytical cell permits on-line reduction of
3-nitrotyrosine to 3-aminotyrosine and
detection of the latter using electrochemical
oxidation. Instability of the Jones reductor may
compromise detection.
Arrays of up to 16 electrodes coupled with
gradient chromatography permit the sensitive,
selective and simultaneous measurement of
3-nitrotyosine and other tyrosine derivatives.
Analytes identified by their retention time and
WWW.ESAINC.COM
Matson (1998);
Ohshima et al.
(1999)
Beal et al. (1995);
Brujn et al. (1997);
Crow (1999);
Ferrante et al.
(1997); Hensley et
244
GC/MS
LC-MS/MS
voltammetric behavior. See Figure 2.13.
al. (1997, 1998);
Maruyama et al.
(1996); Schulz et
al. (1995)
Chemical reduction and derivatization (as
developed by Shigenaga (1999)) followed by
an oxidizing-reducing-re-oxidizing array
permits extremely selective and sensitive
detection of 3-nitrotyrosine as its N-acetylaminotyrosine derivative.
Bose et al. (1999)
Low picogram levels are typically measured
using this approach.
Free 3-nitro-, 3-chloro- and 3-bromotyrosine
derivatives. Excellent sensitivity and selectivity
when tyrosine isotopamer is used.
Excellent sensitivity and selectivity (e.g.,
monitoring daughter ion (m/z 133.1) but
difficult to operate and expensive.
Gaut et al., (2002);
Morton et al. (2003)
Althaus et al.
(2000); Marvin et
al. (2003)
Table 3.12 A Selection Of HPLC-Based Approaches Capable Of Measuring
3-Nitrotyrosine. Anon* preliminary experimentation at ESA Inc.
HPLC-ECD is perhaps the most practical, straightforward method for the
sensitive and routine measurement of 3-nitrotyrosine. A variety of HPLC-ECD
approaches have been developed to measure 3-nitrotyrosine directly or following
chemical (e.g., Figure 3.18) or electrochemical reduction (Table 3.12). Reduction
by dithionite is also used to verify analyte identity – treatment of the sample with
dithionite should, if the 3-nitrotyrosine peak is authentic, completely reduce the
height of its peak in the chromatogram. Kaur et al. (1998) concluded that the use
of dithionite to show peak authenticity can still be problematic as they found a
peak that eluted close to 3-nitrotyrosine that was also capable of being reduced
by dithionite. Perhaps a better approach would be the use of gradient HPLC and
coulometric array detection to effect better separation and qualify analytes based
on their voltammetric signature (Hensley et al. (1997, 1998)). Using this
technique coupled to in vivo microdialysis, McCabe et al. (1997) reported that
peripherally administered 3-nitrotyrosine was capable of passing through the
blood-brain barrier and entering the brain (see Application Note 70-3993
Measurement of 3-Nitrotyrosine). Passage of 3-nitrotyosine through this
protective barrier was by way of the large neutral amino acid carrier as
coadministration of valine significantly blunted its passage (Acworth et al. (1987,
1997b)) (Figures 3.19 and 3.20). These findings suggest that central 3nitrotyrosine need not always be derived from activation of RNS pathways in the
brain, but may be secondary to peripheral production resulting from chronic
diseases.
WWW.ESAINC.COM
245
O
O
Pronase E
hydrolysis
NO2
OH
Protein Pellet
NO2
Acetic Anhydride
Extensive Washing
to remove nitrite
10min, 25 oC
CO 2-
CO 2-
O
ProteinBound 3NT
NH
NH3+
N-, O-Diacetyl
3-Nitrotyrosine
3-Nitrotyrosine
Formic Acid/
Ethyl Acetate
Ethyl Acetate
Evaporated to
dryness, 30min,
37 oC
0.3M NaOH
30min, 37 oC
OH
OH
HPLC-ECD Analysis
NO2
NH2
Filter
CO 2O
NH
N-Acetyl
3-Aminotyrosine
100mM Dithionite
10min, 25 oC
Then HCl
CO 2O
NH
N-Acetyl
3-Nitrotyrosine
Shigenaga et al. (1997) Proc.
Natl. Acad. Sci. USA, 74.
Figure 3.18 Extensive Sample Clean-Up And Chemical
Conversion Of 3-Nitrotyrosine To N-Acetyl-3-Aminotyrosine
Leads To Improved Chromatographic Separation And Lower
Detection Limits By HPLC-ECD.
3-Chlorotyrosine.
Proteins containing 3-chlorotyrosine can be measured using immunostaining
procedures (Hazell et al. (1996)). Free residues and those liberated from protein
can be measured using MALDI-TOF-MS (Domigan et al. (1995)), GC-MS (Hazen
et al. (1996, 1997); van der Vleit et al. (1999), HPLC-absorbance (Domigan et al.
(1995); Eiserich et al. (1996)), HPLC-fluorescence of 1-nitroso-2-naphthol
derivatized amino acids (Kettle (1996)) and HPLC-ECD (Acworth et al. (1998);
Crow (1999)). To date there have been relatively few studies measuring tyrosine
chlorination under conditions of oxidative stress. However, the level of
3-chlorotyrosine is elevated in proteins undergoing phagocytosis, exposed to
inflammatory conditions and obtained from atherosclerotic lesions (Hazell et al.
(1996); Hazen et al. (1996, 1997)).
WWW.ESAINC.COM
246
Microdialysis of Rat Striatum after i.v.
injection of 3-nitrotyrosine (10 mg/kg)
Figure 3.19 Passage Of 3-Nitrotyrosine Through The
Blood
Brain
Barrier
Following
Its
Peripheral
Administration (10mg/Kg. I.V.). (With permission of ESA, Inc.)
3NT Concentration (ng/mL)
Passage of 3NT through the BBB
and Inhibition by Valine
120
3NT i.v. (n=3)
100
3NT i.v. & Valine
i.p. (n=2)
80
60
40
20
0
-60 -40 -20
0
20
40
60
80 100 120 140 160 180 200
Time (min)
Figure 3.20 The Passage Of 3-Nitrotyrosine Through The
Blood-Brain Barrier Is Blocked By Valine A Competitive
Inhibitor At The LNAA Transporter. (With permission of ESA,
Inc.)
WWW.ESAINC.COM
247
The isocratic HPLC system consisted of a pump, an autosampler, a thermal chamber and an
eight channel CoulArray detector.
LC Conditions:
Column:
TSKgel ODS-80TM (TosoHaas) (4.6 x 250mm: 5µm)
Mobile Phase:
20mM Sodium phosphate buffer, 8% methanol (v/v), pH3.2
Flow Rate:
1.0mL/min
Temperature:
31oC
Injection Volume
20µL
Applied Potentials:
+400, +450, +500, +570, +630, +670, +810, +830mV vs. Pd reference.
See Application Note 70-3993 Measurement of 3-Nitrotyrosine for further details.
Dityrosine.
Dityrosine, free or liberated from proteins, can be measured using a variety of
approaches including TLC, GC-MS and HPLC with either UV, fluorescence or
ECD (Abdelrahim et al. (1997); Acworth et al. (1998); Aesbach et al. (1976);
Leeuwenburgh et al. (1997a,b); Malencik et al. (1996)). See Figure 2.13.
Other Tyrosine Oxidation Products.
Tyrosine isomers are readily measured using HPLC-ECD (Chapter 2) or GC-MS
(van der Vleit (1999)). p-Hydroxyphenylacetaldehyde can be measured directly
using HPLC-UV absorbance (Hazen et al. (1996)) but due to its extreme
reactivity is best trapped using a Schiff base. The Schiff base
p-hydroxyphenylacetaldehyde adduct can then be measured using GC-MS
(Hazen et al. (1997)). The measurement of dityrosine and other polymers is
discussed above. Tyrosine peroxides can be measured using HPLC-UV
detection (Winterbourn et al. (1997)). Brominated tyrosine derivatives can be
measured using GC-MS, LC-MS and HPLC-ECD (Wu et al. (1999)).
LIPIDS.
Introduction.
Lipids are water-insoluble (hydrophobic) biomolecules that are highly soluble in
organic (lipophilic) solvents. Lipids consist of a wide variety of organic
compounds showing great structural diversity, from the simple long chain fatty
acids, through terpenes, to the more complex steroids and waxes. Lipids have a
variety of biochemical roles: They act as highly concentrated energy stores (the
triacylglycerols or fats), fuel molecules (e.g., fatty acids), signal molecules (e.g.,
prostaglandins) and components of membranes.
WWW.ESAINC.COM
248
Structure of Biological Membranes.
Membranes serve to define a cell’s shape and separate it from the extracellular
environment, and have been reviewed elsewhere (Halliwell and Gutteridge
(1999); Stryer (1998)). They are fluid-like structures that act as highly selective
permeability barriers. Lipophilic compounds tend to pass through the membrane
unimpeded while hydrophilic compounds require specific protein gates and
channels. Membranes have been likened to a sea of lipids with protein islands
floating in (intrinsic proteins) or on (extrinsic proteins) that sea. Membrane lipids
are generally regarded as being inert and play merely a structural role while
proteins are more active acting as gates, channels, receptors, energy
transducers and enzymes. However, it is now clear that lipids can also play a
more active role: some membrane lipids are the reservoir of arachidonic acid, the
precursor of prostaglandins and other bioactive molecules. In actuality,
membrane lipids are far from being inert and are of great interest to researchers
in the field of redox biochemistry. Membranes readily undergo lipid peroxidation
processes that can affect membrane fluidity and, in turn, membrane protein
function, and can give rise to several cytotoxic species (see below).
The three major kinds of membrane lipids are phospholipids, glycolipids, and
cholesterol. Phospholipids are either based on glycerol or sphingosine (Figures
3.21). Phosphoglycerides consist of a glycerol backbone with its C1 and C2
alcohol groups esterified with fatty acids and its C3 alcohol group esterified with
phosphoric acid. The phosphoric acid head group is also esterified with one of a
number of small aliphatic alcohols. These alcohols include serine, choline
inositol, ethanolamine and glycerol. The structural diversity of phosphoglycerides
is a result of their fatty acid esters and alcoholic head group. Fatty acids tend to
be between 14 and 24 carbon atoms long and can be saturated or unsaturated
(usually in the cis isomer). The unsaturated fatty acid is usually attached to C2 of
glycerol. The chain length and degree of saturation affect membrane fluidity,
while the charge and size of the head group affect binding of extrinsic proteins.
Sphingomyelin is the only phospholipid found in membranes that is not derived
from glycerol. It consists of a sphingosine backbone esterified with a fatty acid
(via an amine-alcohol ester) and a phosphorylcholine head group (Figure 3.21).
Animals also contain glycolipids (sugar-containing lipids) and cholesterol.
Glycolipids are derived from sphingosine (Figure 3.22). Cerebroside consists of a
sphingosine backbone, a fatty acid amine-alcohol ester and a glucose or
galactose head group directly attached to the primary alcohol group of
sphingosine. Gangliosides have the same basic structure but can have a
branched-chain of as many as seven sugar residues. Glycolipids are located on
the extracellular side of the plasma membrane and are involved in intercellular
recognition, an important aspect of the immune system. The sterol cholesterol is
only found in eukaryotes and then primarily in the plasma (not organelle)
membrane (Figure 3.22). Cholesterol affects membrane fluidity and architecture.
WWW.ESAINC.COM
249
O
CH 3CH 2
CH 2(CH 2)6CO 2H
OH
Palmitic acid
Linolenic acid
O
OH
O
OH
Oleic acid (cis)
OH
Trans 9-Octadecanoic acid
Fatty Acids
OH
NH 2
OH
HO
HO
OH
N(CH 3)3
Choline
Ethanolamine
OH
OH
HO
NH 2
O
Serine
OH
Inositol
Headgroups
CH 2OH
HO
CH 3(CH 2)12
NH 3+
HO
OH
HO
Glycerol
Sphingosine
Backbone
O
O
O
P
O
O
N(CH 3)3
CH 2O(CH 2)2NH 3+
CH 3(CH 2)12
N
OH
O
O
O
O
O
A Sphingomyelin
Phosphatidylcholine
Phospholipids
O
O
O
HO
O
O
Diacyl glycerol
O
N(CH 3 )3
O
O
O
P
O
HO
O
O
O
O
O
O
O
P
Diacyl phosphatidate
O
Lysophosphatidylcholine
Other
Figure 3.21 The Basic Building Blocks Of Phospholipids And Related
Species.
WWW.ESAINC.COM
250
C8H17
(or Galactose)
CH2O-Glucose
CH3(CH2)12
N
OH
HO
Cholesterol
O
O
R
Cerebroside
(A Glycolipid)
Figure 3.22 The Structures Of Cholesterol And Cerebroside.
All phospholipids are amphipathic (containing both hydrophobic and hydrophilic
regions) and, when exposed to water, will spontaneously form a bimolecular
sheet in a self-assembly process. The reason that phospholipids readily form
sheets rather than micelles is that their two fatty acid side chains are too bulky to
fit into the interior of a micelle. The formation of a sheet over a micelle is
biologically very important. Micelles are limited in size to <200µm, whereas
bilayers can form much larger structures (typically millimeters). The sheet
consists of a hydrophobic core composed of fatty acid side chains along with the
bulk of the cholesterol molecule held together by hydrophobic interactions (the
driving force for self-assembly). The hydrophilic head groups and the 3-hydroxyl
group of cholesterol face the aqueous phase and are held together by
electrostatic charges and hydrogen bonding.
Lipid Damage.
Lipid damage is probably not a familiar topic to most people but the
consequences have been known for years. Foods high in fats (e.g., meats and
dairy products) undergo oxygen-dependent deterioration leading to rancidity.19
Lipid peroxidation, the primary form of lipid damage found in biological systems,
can broadly be defined as “oxidative deterioration of polyunsaturated lipids”
(PUFAs) (Tappel (1979)). Lipid peroxidation is a particular problem for biological
membranes as they contain high levels of PUFAs. Lipid peroxidation causes a
number of problems for the cell. It decreases membrane fluidity, increases
membrane porosity, inactivates membrane-bound enzymes and produces a
range of toxic breakdown products (e.g., Chen et al., (1995) and references
therein). Normally membrane lipid peroxidation is prevented by a variety of
antioxidant mechanisms (Chapter 4) but under certain conditions, cell “rancidity”
does occur. Indeed increased lipid peroxidation is associated with a variety of
human diseases.
19
Removal of oxygen (e.g., canning), refrigeration and the inclusion of antioxidants prevent food rancidity (see Chapter 4).
WWW.ESAINC.COM
251
Like any free radical-based reaction, lipid peroxidation has three phases:
initiation, propagation and termination (Chapter 1). Initiation under peroxide free
conditions starts with the abstraction of a hydrogen atom from a methylene group
contained within a polyunsaturated molecule resulting in the formation of a
carbon-centered radical (Eqn 3.1) (typical steady-state levels in vivo are 10-17 to
10-18M). The greater the number of double bonds in the system, the greater the
chance that a hydrogen atom will be abstracted.
—CH2— - H• → —CH•—
Eqn 3.1
HO2• + L—H → L• + H2O2
Eqn 3.2
HO2• + L—OOH → H2O2 + LO2•
Eqn 3.3
Initiation can be induced by irradiation or exposure to a variety pro-oxidant
species. Some pro-oxidants are without activity.
•
•
•
•
•
•
•
Irradiation (one reason why irradiation of foods high in fats is not
recommended);
Exposure to the hydroxyl free radical;
Exposure to iron/oxygen complexes (e.g., iron/ATP, iron/DNA,
hemoglobin, myoglobin and cytochrome c);
Exposure to peroxynitrite (Halliwell and Chirico (1993); Radi et al. (1991));
Superoxide cannot enter the membrane so does not directly initiate lipid
peroxidation. Superoxide can, however, lead to formation of hydroxyl free
radicals following its dismutation to hydrogen peroxide (which can also
take part in the Fenton reaction) or it can stimulate hydroxyl free radical
production by reducing Fe (III) to Fe (II) (the Haber-Weiss reaction)
(Chapter 3). Under acidic conditions superoxide forms the lipophilic
hydroperoxyl radical that can promote lipid peroxidation in isolated PUFAs
(Eqn 3.2). Whether this reaction also occurs in membranes is unclear at
present (Aikens and Dix (1991); Halliwell and Chirico (1993)). The
hydroperoxyl radical can react with lipid hydroperoxides forming the lipid
hydroperoxyl radical that can then attack other PUFAs, thereby stimulating
the lipid peroxidation process (Eqn 3.3).
Hydrogen peroxide can enter the membrane but does not appear to
initiate lipid peroxidation directly.
Singlet oxygen does not initiate lipid peroxidation by hydrogen abstraction.
Rather it reacts directly with PUFAs, forming lipid hydroperoxides.
WWW.ESAINC.COM
252
•
•
Ozone adds across carbon-carbon double bonds that can result in a
diradical capable of hydrogen abstraction or decompose to produce
reactive aldehydes (Chapter 2).
Initiation can also be promoted by a variety of toxic compounds. For
example, carbon tetrachloride (CCl4) can be converted to a trichloromethyl
radical by cytochrome P450. The trichloromethyl radical and its peroxyl
radical are very reactive and can readily abstract a hydrogen atom from a
PUFA (Huie and Neta (1999)). Chloroform (CHCl3) is much less reactive
than carbon tetrachloride, probably because it requires more energy to
undergo homolytic fission.
Once formed, the carbon-centered radical can suffer several fates, but the most
likely under aerobic conditions is molecular rearrangement followed by reaction
with oxygen to give lipid peroxyl radicals (Figure 3.23) (typical steady-state levels
in vivo ~2 x 10-9M). These can react with a variety of molecules such as proteins,
DNA and even other lipid peroxyl radicals (see Huie and Neta (1999) for an
excellent review on the chemistry of organic peroxyl radicals) but in the
membrane they are most likely to encounter other PUFAs from which they can
abstract a hydrogen atom, thereby propagating the lipid peroxidation chain
reaction. Thus just one initiation can lead to the formation of over one hundred
lipid hydroperoxides.
After initiation, the length of propagation is dependent upon the fatty acid
composition, oxygen concentration, the amount of protein in the membrane and
the presence of antioxidants. As the rate of formation of lipid peroxides
increases, the chance that a lipid peroxyl radical encounters a protein increases.
Such lipid-protein interaction will lead to termination of the chain reaction. The
presence of chain breaking antioxidants will also end the lipid peroxidation chain
reaction (Chapter 4). As discussed in Chapter 2 nitric oxide is a potent terminator
of the lipid peroxidation process and reacts to form a variety of oxidized products
that can be measured using HPLC-UV and LC-MS (Figure 3.249) (Freeman et al.
(1995); O’Donnell et al. (1999a, b); Rubbo et al. (1994)). Some of these species
have biological activity (e.g., cell signaling and anti-inflammatory properties)
(Coles et al., (2002); Freedman (2002); Lim et al., (2002)).
Finally, lipid hydroperoxides can decompose to yield a variety of decomposition
products including reactive carbonyls (e.g., malondialdehyde and 4hydroxynonenal), polymers and alkanes (Figure 3.24 and below).
WWW.ESAINC.COM
253
POLYUNSATURATED FATTY ACID
(FREE OR PART OF TRIGLYCERIDE,
DIGLYCERIDE OR PHOSPHOLIPID)
CO 2H
R
HYDROGEN ABSTRACTION
-H
(e.g., HO, ONO2-, RO2)
R
CO 2H
LIPID CARBON-CENTERED RADICAL
R
CO 2H
LIPID CARBON-CENTERED RADICAL
MOLECULAR
REARANGEMENT
O2
OXYGEN UPTAKE
CYCLIZATION-POLYMERIZATION
O
O
CO 2H
R
LIPID PEROXY RADICAL
CYCLIZATION- FRAGMENTATION
(e.g. MALONDIALDEHYDE
4-HYDROXY-2-NONENAL
AND OTHER ALDEHYDES)
PUFA
Eqn 3.5
ABSTRACTION OF A
HYDROGEN ATOM
FROM OTHER FATTY
ACIDS CAUSES
AUTOCATALYTIC
CHAIN-REACTION
H
H
α-TOCOPHEROL
CHAIN BREAKING ANTIOXIDANTS
(e.g., TOCOPHEROL) REACT HERE TO
STOP LIPID PEROXIDATION FROM
SPREADING.
α-TOCOPHERYL
RADICAL
PUFA
H
O
O
REDOX-ACTIVE
METALS
CO 2H
R
LIPID HYDROPEROXIDE
ENZYMATIC LIPID
REPAIR (FIGURE 3.29)
Eqn 3.4
OH
CO 2H
R
EXCRETED
LIPID HYDROXIDE
HYDROGEN ATOM
ABSTRACTION FROM PUFA
PROTECTION BY
CHAIN BREAKING ANTIOXIDANTS
O
CO 2H
R
LIPID ALKOXYL RADICAL
Figure 3.23 Initiation, Propagation And Termination Reactions Of Lipid
Peroxidation. (Based on Acworth et al. (1997) and Halliwell and Gutteridge (1999)).
WWW.ESAINC.COM
254
OONO
OH
(CH2)n COOH
HYDROXYLNITROSOPEROXO FATTY ACID
OONO
O2H
(CH2)n COOH
HYDROPEROXONITROSOPEROXO FATTY ACID
ONO
(CH2)n COOH
NITRITO FATTY ACID
O2H
(CH2)n COOH
HYDROXPEROXO FATTY ACID
OONO
(CH2)n COOH
NITROSOPEROXO FATTY ACID
OTHER
ALKENES
RNS
(CH2)nCOOH
PUFA
R
CHO
H
O=C
H
C=O
OH
4-HYDROXY
ALKENAL
MALONDIALDEHYDE
OTHER
ALDEHYDES
Figure 3.24 Lipid Peroxidation Of PUFAs Can Lead To A Variety Of
Products.
WWW.ESAINC.COM
255
The Role of Metals in Lipid Peroxidation.
Metals can stimulate lipid peroxidation through decomposition of lipid
peroxides.20 Pure lipid peroxides are stable, but in the presence of iron or copper
ions their decomposition is greatly increased producing both alkoxyl (RO•) and
peroxyl (ROO•) radicals (Eqns 3.4 and 3.5). The reaction shown in Eqn 3.4 is
equivalent to the Fenton reaction and the alkoxyl radical is analogous to the
hydroxyl free radical. Interestingly, Fe (II) reacts with lipid hydroperoxides 20
times more rapidly than with hydrogen peroxide, and much more rapidly than Fe
(III). This difference in reactivity between Fe (II) and Fe (III) salts may explain the
variable effects of chelating agents on lipid peroxidation (Halliwell and Gutteridge
(1988)). Both alkoxyl and peroxyl radicals are unstable and can abstract a
hydrogen atom from a PUFA, thereby initiating lipid peroxidation. Alkoxyl radicals
can also undergo β-scission producing a variety of products including carbonyl
compounds, alkanes and alkenes, or rearrangement and oxygenation to give
epoxyallylic peroxyl radicals (OROO•). The latter is far more favorable than
hydrogen abstraction or β-scission so that OROO• and not RO•, as widely
assumed, would promote free radical-mediated lipid peroxidation (Girotti (1998)
and references therein).
ROOH + “Fe2+” → “Fe3+” + OH- + RO•
Eqn 3.4
ROOH + “Fe3+” → “Fe2+”+ H+ + ROO•
Eqn 3.5
The role of iron in initiating lipid peroxidation has been covered extensively
elsewhere (Halliwell and Gutteridge (1999); Sergent et al. (1999)). Iron can
theoretically initiate lipid peroxidation by production of hydroxyl free radicals by
the Fenton reaction (Chapter 2). However, it is difficult to conceive that hydroxyl
free radicals, with a half-life of only 1ns, can diffuse from the site of production
into the interior of the membrane to initiate lipid peroxidation. Furthermore,
abundant evidence shows that initiation by iron does not have to involve the
production of the hydroxyl free radical (Minotti and Aust (1989, 1992)). Other
forms of iron have also been suggested as initiators including ferryl and perferryl
species (Chapter 2) but their role in the process is not conclusive (Halliwell and
Gutteridge (1999)). Recent evidence once more suggested a role for an unknown
“Fe2+ + O2” species, possibly ferryl or perferryl in nature, that is readily capable of
initiating lipid peroxidation in unsaturated fatty acid-enriched L1210 leukemia cell
cultures (Qian and Buettner (1999)). Some evidence suggests a role for an
undefined Fe (II)-Fe (III)-oxygen complex (or the ratio of Fe (II)/Fe (III) levels)
(Minotti and Aust (1989, 1992)) but again this was challenged by Halliwell and
20
Commercially available lipid preparations always contain lipid peroxides and this is a major problem when using them to
study lipid peroxidation processes. If these lipids are exposed to iron or copper ions during the experimental procedure
then the contaminating lipid peroxides will form alkoxyl and peroxyl radicals that will stimulate lipid peroxidation thereby
frustrating the experiment (Halliwell and Gutteridge (1999)).
WWW.ESAINC.COM
256
Gutteridge (1999). Unfortunately, the exact nature of the iron/oxygen complex
that is required for initiation of lipid peroxidation still remains elusive.
Unlike the production of the hydroxyl free radical where only “free” iron can take
part in the Fenton reaction, lipid peroxidation can be activated by both free and
bound iron too. Thus iron stored in ferritin or located in heme, hemoglobin,
cytochromes and peroxidases can all promote lipid peroxidation (Halliwell and
Gutteridge (1999)). The mechanism by which these compounds promote lipid
peroxidation may or may not involve free iron and is dependent upon assay
conditions. For example, ferritin can promote liposome lipid peroxidation in a
process that is inhibited by the chelating agent desferrioxamine suggesting that
peroxidation is mediated by free iron ions released during the assay. Heme
appears to promote peroxidation by both free iron ions and the production of
radical species when lipid peroxides react with the heme ring.
Lipid Peroxidation Products.
The decomposition of lipid hydroperoxides can yield a variety of products
including hydroxylated fatty acids, alkanes, alkenes and reactive carbonyl
(aldehyde) compounds (Table 3.13) (Chapter 2). Carbonyls are regarded as
secondary toxic messengers that can travel and cause damage at sites far
removed from the initial point of insult. So far malondialdehyde, and the
hydroxyalkenals (4-hydroxynonenal and 4-hydroxyhexenal) are the most
intensively studied (Esterbauer et al. (1990)). Several other reactive aldehydes
can also be formed in vivo and these are presented in Figure 2.25.
Compound
4-Hydroxy-2nonenal
4-Hydroxy-2nonenal
4-Hydroxy-2nonenal
4-Hydroxy-2nonenal
4-Hydroxy-2nonenal
Species
Tissue
Human
Plasma
Concentration/
Range
0.28+0.34 µmol/L
Human
Plasma
0.68+0.41 µmol/L
Human
0.14+0.17 nmol/mg
Human
Low density
lipoprotein
Monocytes
Human
Plasma
Synovial fluid
4-Hydroxy-2nonenal
Human
Plasma
CSF
3.9+0.8 nmol/108 cells
0.34+0.09 µmol/L rheumatoid arthritis
0.54+0.19 - rheumatoid
arthritis
0.66+0.06 µmol/L control
0.71 to 6.03 Parkinson’s
patients
0.81+0.07 HIV-1
1.24+0.18 AIDS
Reference
Esterbauer et al.
(1990)
Selley et al. (1989)
Esterbauer et al.
(1987)
Selley et al. (1989)
Selley et al. (1992)
Selley (1997)
0.2 to 3.14 – Parkinson’s
patients
WWW.ESAINC.COM
257
Bowel
4-Hydroxy-2nonenal
4-Hydroxy-2nonenal
4-Hydroxy-2nonenal
4-Hydroxy-2nonenal
4-Hydroxy-2nonenal
4-Hydroxy-2nonenal
4-Hydroxy-2nonenal
Cholesteryl ester
hydroperoxide
Cholesterol ester
hydroperoxide
Fatty Acid
hydroperoxides
Isoprostane:
2,3-dinor-5,6dihydro-15-F2t-IP
Isoprostane:
8-epi-IPF2α
Isoprostane:
8-epi-IPF2α
Isoprostane:
8-iso-PGF2α
Rat
Isoprostane:
IPF2α
Isoprostane:
IPF2α
Isoprostane:
IPF2α
Isoprostane:
IPF2α-1
Lipid hydroperoxide
0.02 to 0.23 nmol/mg
protein – control
1.25 to 7.99 – inflamed
2.65 nmol/mg protein
Rat
Hippocampal
cells - culture
Liver
Plasma
Liver
2.82+0.53 nmol/g
0.86+0.2 µmol/L
0.48+0.17 nmol/g
Rat
Liver
0.55+0.1 nmol/g
Rat
Hepatocytes
1.3+0.5 nmol/108 cells
Rat
Microsomes
Lang et al. (1985)
Rat
Retina
0.03+0.01 nmol/mg
protein
0.64+0.64 nmol/g
Human
Plasma
0.32 µmol/L
Human
Plasma
0.3 µmol/L
Human
Plasma
0.056 µmol/L
Human
Urine
390+180 pg/mg
creatinine
Yamamoto et al.
(1987)
Yamamoto and
Ames (1987)
Yamamoto et al.
(1987)
Morrow et al.,
(2003)
Human
34.3+4.5 pg/mL
Human
Breath
(condensate)
Plasma
Human
Urine
Human
Urine
Human
Human
CSF
(ventricular)
CSF (lumbar)
Human
Urine
Human
Plasma
Lipid hydroperoxide
Human
Plasma
0.08 to 0.33 µmol/L –
control
0.13 to 0.76 –
postprandial
0.0 to 1.7 nmol/L
Lipid hydroperoxide
Human
Plasma
4.0+1.7 µmol/L
Lipid hydroperoxide
Lipid hydroperoxide
Human
Human
Plasma
Serum
0
0
Malondialdehyde
Human
Plasma
3.8 to 4.6 µmol/L
Rat
314.6+40 pg/mL
141+41 to 291+102
pg/mg creatinine
dependent upon method
510+160 pg/mg
creatinine
46+4 pg/ml control
72+7 Alheimer’s
23+1 pg/ml control
31+3 Alheimer’s
737+21 pg/mg creatinine
WWW.ESAINC.COM
Mark et al. (1997)
Yoshino et al.
(1986)
Esterbauer et al.
(1990)
Norsten-Hoog and
Cronholm (1990)
Poli et al. (1985)
Van Kuijk (1988)
Baraldi et al.,
(2003)
Vasselle et al.,
(2003)
Tsikas et al.,
(2003)
Tsikas et al.,
(1998)
Morrow et al.,
(2003)
Morrow et al.,
(2003)
Pratico et al.,
(1998)
Ursini et al. (1998)
O’Gara et al.
(1989)
Cramer et al.
(1991)
Frei et al. (1988)
Weiland et al.
(1992)
Hunter and
258
Malondialdehyde
Human
Plasma
0.94 µmol/L
Malondialdehyde
Malondialdehyde
Human
Human
Plasma
Plasma
0.6 µmol/L
3.74 µmol/L
Malondialdehyde
Human
Plasma
35.1µmol/L
Malondialdehyde
Human
Plasma
0.61 µmol/L
Malondialdehyde
Malondialdehyde
Human
Human
Plasma
Plasma
25 to 38 nmol/L
1.7 µmol/L
Malondialdehyde
Malondialdehyde
Malondialdehyde
Human
Human
Human
Plasma
Serum
Serum
0.32 µmol/L
3.4 to 4.0 µmol/L
3.3 µmol/L
Malondialdehyde
Human
Serum
0.9 to 1.88 µmol/L
Malondialdehyde
Human
Serum
3.92 µmol/L
Malondialdehyde
Malondialdehyde
Human
Human
Serum
Urine
47.2 µmol/L
0.2 to 0.8 µmol/L
Malondialdehyde
Human
Urine
Malondialdehyde
Rat
Malondialdehyde
Rat
Malondialdehyde
Rat
Malondialdehyde
Rat
Brain microdialysis
Brain microdialysis
Heart perfused
Liver
Sperm
0.019+0.012 µmol/mmol
creatinine
0.4 µmol/L
Malondialdehyde
Rat
Phosphatidyl-choline
hydroperoxide
Phosphatidyl-choline
hydroperoxide
0.02 to 0.16 µmol/L
0.068+0.016 µmol/L
perfusate
0.7 to 0.8 nmol/g
0.4 to 3.9 pmol/mg
protein
Human
Brain substantia
nigra
striatum
cerebellum
Plasma
3.23+0.25 nmol/mg
protein
3.78+0.28
5,64+0.72
0.05 to 0.43 µmol/L
Gerbil
Brain
~9pmol/mg tissue
Mohamed (1986)
Ledwozyw et al.
(1986)
Wong et al. (1987)
Yasaka et al.
(1981)
Santos et al.
(1980)
Francesco et al.
(1985)
Yeo et al. (1994)
Viinikka et al.
(1984)
Lee (1980)
Satoh (1978)
Maseki et al.
(1981)
Suematsu et al.
(1977)
Nishigaki et al.
(1981)
Aznar et al. (1983)
Tomita et al.
(1990)
Korchazhkina et
al. (2003)
Waterfall et al.
(1996)
Yang et al. (1997)
Cordis et al.
(1994)
Yeo et al. (1994)
Thiffault et al.
(1995)
Miyazawa et al.
(1988)
Zhang et al.
(1994)
Table 3.13 Levels Of Lipid Peroxidation Products Reported In The
Literature.
These analytes are routinely used as markers of lipid peroxidation. However, the
reader should be aware that these low-molecular-mass products are probably
formed as the result of multiple reactions far removed from the initial lipid
peroxidation site. Consequently recent research is focusing on some of the initial
compounds formed during the lipid peroxidation process. These include
WWW.ESAINC.COM
259
cholesterol oxidation products (see below), some oxidation metabolites of
arachidonic acid (e.g., the isoprostanes (see below), isoleukotrienes, and the
levuglandins), and the derivatives of docosahexaenoic acid (the neuroprostanes).
Malondialdehyde.
Malondialdehyde (or malonaldehyde) [MDA] is a molecule with two aldehyde
groups. In aqueous conditions it exhibits several pH- and age-dependent
structural forms including keto-enol tautomers, intramolecular and intermolecular
hydrogen bonded forms, dimers and trimers. For most biological experiments
MDA is prepared by acid hydrolysis (e.g., 1% HCl) of commercially available bisdimethyl- or bis-diethylacetal. Dilute MDA solutions can be stored at 4oC for
several days without noticeable degradation, but higher concentrations,
especially if left for long periods at room temperature, undergo aldol
condensation forming dimers and trimers.
O
O
CO 2 H
Arachidonic Acid Peroxyl Radical
O
O
CO 2 H
Cyclic Peroxide
CO 2 H
O
O
OOH
OOH
OOH
Cyclic Endoperoxide Radical
O
1,3-Dihydroperoxide
PGH 2
Hydrolysis
or heat
O
O
Hydroperoxyepidioxide
O2
O
Malondialdehyde
OOH
OOH
Polyunsaturated Aldehydes
1,4-Dihydroperoxide
Figure 3.25 Possible Mechanisms For The Production Of
Malondialdehyde From PUFAs. (Esterbauer et al. (1991); Frankel and
Neff (1983); Halliwell and Gutteridge (1999); Pryor and Stanley (1995)).
WWW.ESAINC.COM
260
As shown in Figure 3.25, MDA can be formed from several compounds including
a variety of lipid peroxidation products, by the action of human platelet
thromboxane synthetase on prostaglandins PGH2, PGH3 and PGG2, and by the
action of polyamine oxidase and amino oxidase on spermine (Figure 2.25). MDA
is found in both healthy and diseased tissues (Table 3.13) (Esterbauer et al.
(1991) and references therein) but caution must be exercised when reviewing
these levels as inappropriate analytical procedures were sometimes followed
(see below).
A) The Reaction of MDA and Lysine
CHO
N
CHO
NH CH CH CH N
1,4-Dihydropyridine3,5-dicarbaldehyde (1%)
Amino-imino-propen
Cross Link (77%)
NH C
CH CHO
Aminopropenal (22%)
B) The reactions of 4-Hydroxynonenal
Protein
R
Protein
R
Protein
+
O
R
N
O
OH
Cysteine
S
OH
Cysteine
R
N
Lysine
Lysine
OH
Lysine
Lysine
S
Histidine
R
N
Histidine
OH
Figure 3.26 Some Reactions Of A) MDA And B) 4-HydroxyNonenal. (% Reflects The Abundance Of Modified Lysine
Present In Vitro (Esterbauer et al. (1991)).
WWW.ESAINC.COM
261
Under physiological conditions (pH7.4), MDA exists as an enolate anion
(-O—C=C—CHO), and in this form it is not very reactive. Thus the claim often
expressed in literature that MDA is an extremely reactive compound is untrue.
Under more acidic conditions (pH<4), β-hydroxyacrolein (HO—C=C—CHO)
(βHA) is the predominant form. Like 4-hydroxynonenal, βHA is a very reactive
electrophile capable of reacting with nucleophiles (e.g., thiols) in a Michael type
1,4 addition. For example, glycine forms the monoglycine-βHA adduct, an
enaminal, N-propenal-amino-acetic acid (OCHCH=CHNHCHRCO2H). Under
strong acidic conditions and at a high concentration, further reaction produces
the diglycine-βHA adduct. In a well-controlled study Nair et al. (1981) examined
the reaction of several amino acids and reported that the product depended on
the amino acid. Aromatic amino acids and arginine reacted at the α-amino group
forming mono-enaminal adducts, whereas cysteine forms an adduct that contains
two cysteine and 3 MDA molecules.
Proteins are much more reactive with MDA than free amino acids. Although the
reasons are not completely clear, it may be that proteins provide a more reactive
environment and that MDA-condensation products might be the reacting species
(Esterbauer et al. (1991)). MDA preferentially reacts with the ε-amine group of
lysine and is capable of causing both intra- and inter-protein cross-links (Figure
3.26). MDA also reacts with histidine, tyrosine, arginine and methionine residues.
Damage to essential –NH2 or –SH groups can lead to inactivation of enzymes
(e.g., liver microsomal glucose-6-phosphatase is inactivated when its –SH
groups are damaged).
MDA can also react with DNA bases producing a variety of mutagenic
compounds (Figure 3.7). Deoxyguanosine is the most reactive, adenosine and
cytidine are fairly reactive, while thymidine is not reactive at all. MDA has the
potential to induce amino-imino-propen cross-links between complementary
strands of DNA and cause DNA-protein cross-linking.
MDA is metabolized in the liver to malonic acid semialdehyde. This is unstable
and spontaneously decomposes to acetaldehyde. Acetaldehyde is then
converted to acetate by aldehyde dehydrogenase and acetate to carbon dioxide
and water. Some MDA eventually ends up as acetyl-CoA. Mammalian urine also
contains enaminals derived from the hydrolysis of MDA modified proteins
(Esterbauer et al. (1991)). Urinary output of MDA in humans is typically 0.20.8µmol/L (Tomita et al. (1990)).
4-Hydroxyalkenals.
These compounds were first discovered in the early 1960s and are formed as
end products of lipid peroxidation and following hepatic metabolism of the
hepatotoxic pyrrolizidine alkaloid senecionine (Esterbauer et al. (1991)). The
WWW.ESAINC.COM
262
hydroxyalkenals are very reactive and show reactions common to unsaturated
aldehydes and alcohols (Chapter 2). They readily react with thiols (e.g.,
glutathione), many amino acids (e.g., glycine produces a pyridinium derivative)
(Figure 3.26), DNA (see above), and phospholipids that contain either
ethanolamine or serine (Guichardant et al. (1998)).
4-Hydroxynonenal (CH3(CH2)4CHOHCH=CHCHO) ([trans] 4-hydroxy-2(E)nonenal) is a cytotoxic lipophilic compound (Muller et al. (1996)) and is formed by
lipid peroxidation of n-6 fatty acids (e.g., arachidonic acid) (Van Kuijk et al.
(1990)). It has been measured in both healthy and diseased tissues (Table 4.13)
(Esterbauer et al. (1990)). For example, it is elevated by treatments that promote
lipid peroxidation, in melanoma, in thalassemic subjects, in humans exhibiting
toxic oil syndrome and in Parkinson’s disease (Selley (1998); Zarkovic (2003)). 4Hydroxynonenal also possesses diverse biological activity. It can affect cell
proliferation, act as a chemotactic agent, potentiate platelet aggregation, and
modify the expression of several genes including heat shock factor and the c-fos
proto-oncogene (Esterbauer et al. (1991) and references therein; Kreuzer et al.
(1998)).
4-Hydroxyhexenal (CH3CH2CHOHCH=CHCHO) is formed by lipid peroxidation of
n-3 fatty acids (docosahexanoeic acids), and like 4-hydroxynonenal is also highly
toxic (Van Kuijk et al. (1990)). It is capable of causing reversible structural
damage to mitochondria (mitochondrial permeability transition) at doses a billion
fold lower than 4-hydroxynonenal (Kristal et al. (1996)).
High levels (mM range) of 4-hydroxyalkenals are acutely toxic to mammalian
cells leading to depletion of GSH, decrease in protein thiols, induction of lipid
peroxidation, disturbance of calcium homeostasis, inhibition of DNA, RNA and
protein synthesis, inhibition of respiration and glycolysis, and morphological
changes (see Esterbauer et al. (1991) and references therein). Lower levels (10200µM) were found to be less severe with the disturbances being more selective
and dependent upon cell type. For example, fibroblasts die when exposed to
100µM 4-hydroxynonenal while hepatocytes readily tolerate the same exposure.
This may be due to the hepatocyte’s ability to metabolize and detoxify this
aldehyde.
4-Hydroxyalkenals are subjected to detoxification in vivo. 4-Hydroxynonenal is
catabolized in the liver to a number of metabolites including: 1,4-dihydroxy-2nonene (following the action of NADH-dependent alcohol dehydrogenase), 4hydroxy-2-nonenoic acid (aldehyde dehydrogenase), a glutathione conjugate
(GSH transferases), a mercaputurate conjugate, omega-hydroxylated products
and their GSH and mercapturate adducts (Alary et al., (2003)). Both 4hydroxyhexenal and 4-hydroxynonenal can be reduced and rendered less toxic
by aldose reductase (He et al. (1998)).
WWW.ESAINC.COM
263
Other Reactive Carbonyls.
Space is too short to explore the chemistry and biology of all of the reactive
carbonyls formed in vivo in detail so this section will be limited to the other major
cytotoxic carbonyls, the 2-alkenals, or α,β-unsaturated aldehydes. Acrolein
(CH2=CHCHO) is the simplest and most reactive member of the series. It is a
strong electrophile and can readily react with thiols and amines producing a
variety of products (e.g., dysfunctional protein-acrolein adducts) (Calingasen et
al. (1999); Uchida et al. (1998a,b). Acrolein can react with guanosine producing
mutagenic lesions (Figure 3.7). It is highly cytotoxic showing similar actions to 4hydroxynonenal (reviewed by Esterbauer et al. (1991)). Acrolein is formed in vivo
during metabolism of spermine and spermidine, by the action of hypochlorous
acid on threonine, as a result of lipid peroxidation and during metabolism of allyl
alcohol and allylamine (Figure 3.25) (Anderson et al. (1997)). Acrolein is a
common pollutant and is formed during incomplete combustion of wood, petrol,
coal and plastics. It is also found in cigarette smoke (typically 25-140µg/cigarette)
and burning oil, and is a toxic agent formed by the metabolism of the anticancer
drug cyclophosphamide (Esterbauer et al. (1991) and references therein).
Crotonal (CH3CH=CHCHO) is formed during the metabolism of the
hepatocarcinogenic cyclic nitrosamine, N-nitrosopyrrolidine while other members
of the 2-alkenals (e.g., 2-pentenal, 2-heptenal, 2-octenal and 2-nonenal) are
formed by peroxidized microsomes (Esterbauer et al. (1990)).
Cholesterol Oxidation.
Plasma membranes of eukaryotic cells are usually rich in cholesterol (typically
40-45 mol % of total lipids). Cholesterol, a neutral lipid, is involved in membrane
fluidity. On the one hand it prevents the crystallization of fatty acyl chains by
fitting between them, while on the other its bulk sterically blocks the motion of the
fatty acyl chains making them less fluid. Although the accumulation of cholesterol
in the plasma membrane of aging neurons is reported to affect nerve function, it
now appears that such increases may actually protect neurons from oxidative
damage (Joseph et al. (1996)).
Cholesterol also exists as esters formed by the reaction between a carboxylic
acid with cholesterol’s C-3 hydroxyl group. These occur at high levels in lowdensity lipoprotein (LDL), the major transporter of cholesterol in the blood.
Cholesterol esters in LDL are rich in linoleate (a polyunsaturated fatty acid) while
those stored in cells contain mainly oleate and palmitate (mono-unsaturated fatty
acids). High-density lipoproteins (HDL) also contain cholesteryl esters formed by
esterification of cholesterol which has been picked up from the plasma, resulting
from cell death and membrane turnover.
WWW.ESAINC.COM
264
C 8 H 17
HO
CHOLESTEROL
C 8 H 17
C 8 H 17
HO
HO
OOH
OH
CHOLESTEROL 5 α
HYDROPEROXIDE
CHOLESTEROL 5 α
HYDROXIDE
C 8 H 17
C 8 H 17
C 8 H 17
HO
HO
HO
OOH
HO
OH
OOH
CHOLESTEROL 6 α
HYDROPEROXIDE
CHOLESTEROL 6 β
HYDROPEROXIDE
C 8 H 17
HO
OH
CHOLESTEROL 6 α
HYDROXIDE
HO
CHOLESTEROL 7 α
HYDROPEROXIDE
OOH
CHOLESTEROL 7 β
HYDROPEROXIDE
CHOLESTEROL 6 β
HYDROXIDE
C 8 H 17
C 8 H 17
OOH
C 8 H 17
HO
OH
C 8 H 17
HO
CHOLESTEROL 7 α
HYDROXIDE
OH
CHOLESTEROL 7 β
HYDROXIDE
OH
OH
CH 3
CH 3
H 3C
H 3C
HO
HO
24 α HYDROXYCHOLESTEROL
25-HYDROXYCHOLESTEROL
HO
HO
O
C 8 H 17
C 8 H 17
C 8 H 17
HO
O
C 8 H 17
HO
O
OH
5 α CHOLESTANE3 β ,5 α ,6 β TRIOL
5,6 α EPOXY-5 α
CHOLESTAN-3 β ,7-DIOL
5,6 β EPOXY-5 β
CHOLESTAN-3 β ,7-DIOL
OH
C 8 H 17
HO
OH
OH OH
3 β HYXDROXYCHOLEST5-EN-7-ONE
5 α CHOLESTANE3 β ,5 α ,6 β ,7 β -TETRIOL
Figure 3.27 Cholesterol And Its Oxidation Products.
WWW.ESAINC.COM
265
Cholesterol (and its esters) can become oxidized during lipid peroxidation giving
rise to a variety of biologically active products that show atherogenic, cytotoxic,
and mutagenic properties and possess enzyme inhibitory properties (Figure 4.22)
(Peng and Morin (1987); Smith and Johnson (1989)). Some cholesterol oxidation
products are being used as markers of oxidative stress (Geiger et al. (1997) and
references therein; Girotti (1998); Patel et al. (1996); White et al. (1994)). These
are gaining favor over phospholipid markers as:
i) Cholesterol exists as a single molecular species in the membrane;
ii) Its oxidation products can be measured directly without the need for
potentially artifactual hydrolysis steps; and
iii) Unlike phospholipids it can be readily transfer-radiolabeled without a
requirement for transfer proteins (Girotti (1998)).
Free radical-mediated reactions mainly produce 7α-hydroperoxide and
7β-hydroperoxide epimers, with lesser amounts of 7α-hydroxy, 7β-hydroxy,
7-one, epimeric 5,6-epoxides and other species. Singlet oxygen mainly yields the
5-α-hydroperoxide, 6α-hydroperoxide and 6β-hydroperoxide and these have
been proposed as a potential marker for singlet oxygen activity. Several
chlorinated sterols including a dichlorinated sterol, and cholesterol α- and βchlorohydrin are produced when cholesterol is exposed to the myeloperoxidasechlorinating system (Hazen et al. (1996)). Chlorination appears to proceed via
chlorine rather then hypochlorous acid.
The Isoprostanes.
The isoprostanes have recently received a great deal of interest as possible
markers of oxidative stress. Isoprostanes are prostaglandin-like compounds that
are produced by free radical catalyzed lipid peroxidation, independent of the
cyclooxygenase (COX) pathway (Fam and Morrow (1993); Morrow et al. (1999);
Morrow and Roberts (1996); Pratico et al. (1997); Roberts and Morrow (1997)).21
Many isoprostanes can be formed and include those produced from the E, D and
F series of prostaglandins, in turn produced from arachidonic acid (Figures 3.28).
F-isoprostanes can be further classified as belonging to the F2, F3 or F4 series.22
It was the finding of Morrow and Roberts (1996) that F2-isoprostanes are formed
in situ in phospholipids and are released by the action of phospholipases to
circulate in the plasma, which has prompted measurement of these compounds
as indices of oxidative stress in vivo. Up to 64 isomers of F2 can exist but
probably most attention has focused on 8-epi PGF2α, due, in part, to its biological
21
Plants also form isoprostanes, called phytoprostanes, produced from linolenic acid (C18:3). These include dinor
isoprostanes (PPF(1)), and the cyclic oxylipins (E1-phytoprostane) that can be further metabolized to novel
cyclopentanone derivatives (A1- and B1-phytoprostanes) (Imbusch and Mueller (2000); Thoma et al., 2003)). PPF1 is
particular abundant in pollen (e.g., 32 µg/g in birch pollen) and may be responsible for the some of the breathing problems
associated with hayfever.
22
F3 and F4 isoprostanes are formed by the peroxidation of eicosopentaenoic acid and docosohexaenoic acid,
respectively (Nourooz-Zadeh et al. (1999) and references therein). F4 isoprostanes are sometimes refered to as
neuroprostanes.
WWW.ESAINC.COM
266
activity. F2 isoprostanes have been proposed as markers of oxidative stress due
to their stability and presence in measurable quantities in all tissues and fluids.23
Levels of F2 isoprostanes are increased in models of oxidative stress and their
levels are suppressed upon treatment with antioxidants (Morrow and Roberts
(1996)). Although proposed as potentially useful oxidative stress markers, some
studies have shown that small quantities of 8-epi PGF2α are formed naturally by
the COX enzyme. This has raised the question as to the possibility that 8-epi
PGF2α may not always be a reliable marker of oxidative stress. Other
researchers have argued that the amount of 8-epi PGF2α produced by COX is too
insignificant to be of concern (Morrow and Roberts (1996) and references
therein).
CO2 H
Arachidonic Acid
ROS - Hydrogen Abstraction
Oxygenation
Rearangement
Reduction
O
HO
OH
HO
A
HO
HO
O
E2-Isoprostane
O
HO
Isothromboxane
D2-Isoprostane
F2-Isoprostane
OH
HO
CO2 H
HO
HO
CO2 H
HO
HO
HO
I
CO2 H HO
HO
II
CO2 H
B
OH
IV
III
Isoprostane F2α
HO
CO2 H
HO
OH
9α, 11α
HO
HO
CO2 H
CO2 H
HO
OH
HO
9α, 11β
OH
9β, 11α
HO
CO2 H
HO
C
OH
8−Epi
Figure 3.28 The Formation Of Isoprostanes. Several Different Head
Groups Can Be Formed (A). Each Of These Shows Different
Regioisomers (B) Which Also Show Optical Isomers (C).
23
It is unclear whether urinary levels of isoprostanes reflect plasma levels or mainly result from local renal production.
WWW.ESAINC.COM
267
Glutathione
Peroxidase
Fatty Acid
O
OH
GSH
C O
OH
CH 2
HC
CH 2
O
O C
HC
CH 2
CH 2
C
HC
O
C
1/2 Native Lipid
Bilayer
O
O C
CH 2
OH
HC
O
CH 2
Pro-oxidant
O
C O
OH
HC
O
HC
OH
CH 2
CH 2
O
C
CH 2
CH 2
O
O C
CH 2
O
O
O
O
CH 2
O
O C
Hydroxylated
Fatty Acid
OH
Hydroperoxide
O
O
Ca2+ /PLA 2
C
O
C
O
CH 2
O
O C
HC
Acyltransferase
C
HC
O
C
O
O C
CH 2
CH 2
O
O
FA-CoA
Oxidized Fatty Acid
Removed
Oxidized Bilayer with
Disrupted Structure
CH 2
O
O C
O
C
Membrane
Repaired
EXCISION-REDUCTION REPAIR
Phospholipid
Hydroperoxide
Glutathione
Peroxidase
GSH
Ca2+ /PLA 2
Hydroxylated
Fatty Acid
OH
C
OH
CH 2
HC
O
O C
CH 2
HC
OH
CH 2
CH 2
O
O
C
O
O
O C
O
O
C
Oxidized Bilayer with
Disrupted Structure
REDUCTION-EXCISION REPAIR
Figure 3.29 Excision-Reduction And Reduction-Excision
Lipid Repair Processes (Based on Van Kuijk et al. (1987)
And Girotti (1998)). PLA2 – phospholipase A2; FA-CoA – fatty acyl
CoA.
Lipid Repair.
A variety of antioxidant defenses exist in eukaryotic cells to protect against lipid
peroxidation damage. These include enzymes (e.g., catalase, and superoxide
dismutase) and metal binding proteins (e.g., ferritin) designed to prevent,
minimize ROS formation or remove them once formed (Chapter 4). Furthermore,
membranes contain the chain-breaking antioxidant, α-tocopherol that can
intercept lipid peroxyl radicals and prevent lipid peroxidation chain reactions
(Chapter 4). Unfortunately, protection by the antioxidant defenses is not complete
WWW.ESAINC.COM
268
and lipid peroxidation can still take place. Damaged lipids must be removed if the
membrane is to remain viable.
The detoxification of membrane-bound lipid peroxides appears to involve the
activity of membrane-bound phospholipase A2 (van Kuijck et al. (1987)) and
three intracellular enzymes with peroxidatic activity, glutathione peroxidase,
phospholipid hydroperoxide glutathione peroxidase, and non-seleno GSH-Stransferase type α (Figure 3.29) (Flohe (1982); Ursini and Bindoli (1984); Ursini
et al. (1985)). Although both glutathione peroxidase and phospholipid
hydroperoxide glutathione peroxidase reduce lipid hydroperoxides into lipid
alcohols in a two-electron reaction, they differ in size, amino acid sequence, and
cellular distribution (Brigelius-Flohe et al. (1994)). They also show markedly
different substrate specificity: glutathione peroxidase can only act on unesterified
fatty acid hydroperoxides or hydrogen peroxide, whereas phospholipid
hydroperoxide glutathione peroxidase is more versatile and can act on
phospholipid peroxides contained in membranes, cholesterol, cholesteryl ester
hydroperoxides and hydrogen peroxide (Grossman and Wendel (1983); Thomas
et al. (1990a,b); Ursini and Bindoli (1984); van Kuijk et al. (1987)). The final step
in lipid repair is the re-esterification of the lysophospholipid by an acyltransferase
forming a phospholipid. As shown in Figure 4.25, two possible pathways of lipid
repair exist, excision-reduction and reduction-excision repair. Evidence suggests
that both pathways can operate in vivo (Girotti (1998) and references therein).
Lipid Damage and Disease.
Lipid peroxidation is increased when cells are damaged. Thus it should come as
no surprise that many diseases are associated with increased lipid peroxidation.
However, the question that needs to be asked is whether lipid peroxidation
causes or is just the result of a disease. The latter may still be important as, in
the case for the cytotoxic aldehydes, lipid peroxidation may play a role in disease
progression. Lipid peroxidation is directly involved in atherosclerosis (Halliwell
and Gutteridge (1999)). Oxidative modification to lipoproteins and cholesterol can
result in atherosclerosis (Chang et al. (1997); Hajjar and Haberland (1997); Morin
and Peng (1989); Patel et al. (1996); Westhuyzen (1997); White et al. (1994)). In
the oxidation hypothesis of atherosclerosis, oxidative damage to lipoprotein, and
in particular LDL, increases its ability to cause this disease by altering receptormediated uptake by cells in the intima of blood vessels. Subsequently, oxidized
LDL is then taken up by scavenger receptors on other cells such as monocytes,
macrophages and smooth muscle cells. This uncontrolled process leads to the
accumulation of lipids and the formation of foam cells, an early indicator of
atherosclerotic plaque formation (Westhuyzen (1997)). Lipid peroxidation is also
associated with brain damage resulting from reperfusion injury, and possibly
neurodegenerative diseases (Acworth et al. (1998) and references therein; Cini
et al. (1994) and references therein). Lipid peroxidation is also responsible for the
WWW.ESAINC.COM
269
production of advanced lipid end products that accumulate with aging and
disease (Sayre et al. (1997)).
Measurement of Lipid Damage.
Total peroxidation can be determined by the uptake of oxygen. A variety of
methods are used to measure lipid peroxidation products including GC-MS, LCMS, HPLC with electrochemical, fluorescence or chemiluminescence detection
(reviewed by Esterbauer et al. (1991); Frederik et al., (1987); Halliwell and
Chirico (1993); Halliwell and Gutteridge (1999); Pryor (1989); Spickett (2003);
Van Kuijk and Dratz (1987);) (Table 3.14). For example, carbon and oxygen
centered radicals can be measured using EPR/spin trap procedures. Lipid
peroxides can be measured by iodine liberation, heme degradation of the
peroxide, and COX activity. These approaches vary in their selectivity, sensitivity,
ease of use and extent of sample preparation. Unfortunately many of the
analytical methods available are not accurate and can yield ambiguous data. For
example, the iodide approach may be useful for examining the oxidation of pure
lipids, but biological tissues contain other oxidants such as hydrogen peroxide
that can also cause iodine production from iodide. Furthermore, two commonly
used approaches, diene conjugation and the thiobarbituric acid reactivity (TBAR)
test that are used to examine fatty acid oxidation and carbonyl production,
respectively, also lack specificity in some systems (see below).
Analyte
4-Hydroxynonenal
4-Hydroxynonenal modified
proteins
Carbon and oxygen centered
radicals
Cholesterol hydroperoxides
and cholesteryl ester
hydroperoxides
Cholesteryl ester
Technique
HPLC-UV of 2,4-DNPH
derivative
Reference
Esterbauer et al. (1991) and
references therein
HPLC-fluorescence of 1,3cyclohexandione derivative
Esterbauer et al. (1991) and
references therein
HPLC-coulometric detection
of DNPH derivative.
Goldring et al. (1993)
GC-HPLC-antibody
techniques
Halliwell and Chirico and
references therein (1993)
GC-MS
Esterbauer et al. (1991) and
references therein
Waeg et al. (1996)
Monoclonal antibodies
EPR combined with spin
trapping (e.g., with phenyl tbutylnitrone)
HPLC-ECD
Halliwell and Chirico (1993)
and references therein
HPLC-chemiluminescence
Yasuda and Narita (1997)
WWW.ESAINC.COM
Arai et al. (1996), Korytowski
et al. (1991)
270
hydroperoxides
Hydroxy and ketocholesterols
Fatty acid hydroperoxides
Hydrocarbon gases
GC-MS
Ferrocene derivatization/
voltammetry.
Vatassery et al. (1997)
Mulchandani and Rudolph
(1995)
HPLC-ECD of GSSG
production by GSHperoxidase following reaction
with lipid hydroperoxides
O’Gara et al. (1989)
HPLC-UV
Browne and Armstrong (1998)
GC, GC-MS
GC
Isoprostanes
GC-MS, LC-MS and
immunochemical approaches
Lipofuscin/lipopigments
Fluoresence
MDA
HPLC-UV
HPLC with pre-column
derivatization with TBA,
dansylhydrazine, 2,4-DNPH
or methylamine +
acetaldehyde. HPLC-UV,
HPLC-fluorescence
Halliwell and Chirico (1993)
and references therein
De Zwart et al.(1998); Morrow
and Roberts (1996, 1999);
Pratico et al. (1998)
Delori and Dorey (1998); Yin
and Brunk (1998)
De Zwart et al. (1998) and
references therein; Esterbauer
et al. (1991) and references
therein; Halliwell and Chirico
(1993) and references therein;
Korchazkina et al., (2003); Yeo
et al. (1994)
HPLC with post-column
derivatization with TBA.
HPLC-UV and HPLCfluorescence
Other carbonyls
Phospholipid hydroperoxide
and triacylglycerol
hydroperoxides
GC-ECD and GC-MS
GC, GC with 2,4DNPH
derivatization, HPLC with
2,4-DNPH derivatization,
TLC
COX activity, heme
degradation of peroxides,
hemoglobin/methylene blue
reaction, GSH
peroxidase/GSH reductase
(free hydroperoxides only),
iodine production.
De Zwart et al. (1997);
Halliwell and Gutteridge (1989)
and references therein
Halliwell and Chirico and
references therein (1993);
Kuijk and Dratz (1987) and
references therein; Yagi
(1998)
Chemiluminescence
Iwaoka et al. (1987)
HPLC-chemiluminescence
Yamamoto et al. (1987a,b,
1990, 1994, 1998)
HPLC-chemiluminescencePDA
Holley and Slater (1991)
WWW.ESAINC.COM
271
HPLC-UV,
chemiluminescence or
evaporative light scattering
detectors
Makinen et al. (1996) and
references therein
HPLC-ECD
Arai et al. (1997), Yamada et
al. (1987)
HPLC-polarography
Korytowski et al. (1995, 1999)
GC-MS
Total peroxidation
Measurement of oxygen
uptake using an oxygen
electrode
Halliwell and Chirico (1993)
and references therein; Kuijk
and Dratz (1987) and
references therein
Halliwell and Chirico (1993)
and references therein
Table 3.14 Methods used to measure lipid peroxidation.
Lipid peroxidation is often measured using either diene conjugation or TBARs.
However, these approaches often suffer from practical limitations:
Diene Conjugates.
These are formed during peroxidation of polyunsaturated fatty acids and are
thought to be esters of octadeca-9,11-dienoic acid, a non-peroxide isomer of
linoleic acid (Dormandy and Wickens (1987)). They absorb at 230-235nm and
thus UV absorbance by these compounds is used as an indicator of the early
stages of lipid peroxidation. HPLC-UV detection has been used to study diene
conjugates in a variety of tissues (Cawood et al. (1993)). Double-derivative
spectroscopy has improved the limit of detection of this approach but it is still
limited to pure lipid samples and measurement in biological fluids remains
challenging (Situnayake et al. (1990)).
TBAR.
This simple and cost effective test is often used to measure MDA. A biological
sample is heated under acidic conditions with thiobarbituric acid, and the pink
chromagen (Figure 3.30) is measured at 532nm (UV) or 553nm (fluorescence).
Although this approach works well for standards and clean samples, biological
samples are fraught with problems. First this method lacks selectivity and many
biologically occurring carbonyls (e.g., aldehydes formed during peroxidation, bile
acids, DNA bases, reducing carbohydrates and glycoproteins) can also produce
a chromophore absorbing at 532nm. Second, the test does not actually measure
tissue levels of MDA, rather it measures stimulated MDA levels due to
WWW.ESAINC.COM
272
decomposition of lipid peroxides induced by the acid/heating stages of the test.
Third, peroxide decomposition can further stimulate lipid peroxidation, thereby
amplifying the response. Halliwell overcomes these problems by adding the
chain-breaking antioxidant butylated hydroxytoluene to the sample before
analysis and using HPLC to separate the TBA-MDA adduct from other
chromagenic interferences (Halliwell and Chirico (1993)). A typical basal plasma
chromatogram is shown in Figure 3.31.
S
N
CHO
N
MDA
HS
N
OH
CHO
OH
HO
2H2O
N
OH
OH
N
2
THIOBARBITURIC
ACID
HS
N
OH
PRODUCT
Figure 3.30. Formation Of A Chromagen From TBA And MDA.
So far there is not one method that measures all aspects of lipid peroxidation.
The correct choice of method depends on what question is being asked. As
some methods are not selective and some are affected by the sample
preparation employed, care must be exercised when choosing between methods.
If at all possible two or more different methods should be used to answer any
question posed.
WWW.ESAINC.COM
273
Figure 3.31 Chromatogram Showing Isolation Of The MDA-TBAR Derivative
(1). See ESA Application Note 70-5033 Malondialdehyde For Further Details.
The isocratic analytical system consisted of a pump, an autosampler and a fluorescence detector.
LC conditions:
Column:
Mobile Phase:
Flow Rate:
Temperature:
Injection Volume:
Detector:
Excitation Wavelength:
Emission Wavelength:
HR-80 (4.6 x 80mm; 3µm).
40% Methanol.
1.0mL/min.
Ambient.
20µL.
Fluorescence, Model 305.
515nm.
553nm.
See ESA Application Note 70-5033 Malondialdehyde For Further Details.
Aldehydes (e.g., 4-hydroxynonenal) can also be measured using GC-based
approaches, by the measurement of their 3,5-dinitrophenylhydrazine derivatives
using HPLC-ECD (Goldring et al., (1993)) or HPLC-fluorescence, or by HPLCfluorescence of an acridine derivative formed in the Hantzsch reaction (Holley et
al., (1993)) (see Figures 3.32 and 3.33). See ESA Application Note 70-5041 4Hydroxynonenal And Other Aldehydes.
WWW.ESAINC.COM
274
O
O
R
O
(NH4) 2SO4
+
O
R
CHO
HANTZSCH
Reaction
2
1,3-Cyclohexanedione
"O"
N
H
9-Alkyl-4,5,6,7,9,10octahydroacridine-1,8-dione
O
R
O
N
9-Alkyl-4,5,6,7-hexahydroacridine-1,8-dione
Figure 3.32 Formation Of A Fluorescent Derivative When
Aldehydes Undergo the Hantzsch Reaction.
Figure 3.33 Chromatogram Of Aldehyde Standards
(5ng on column). 1 - formaldehyde; 2 - acetaldehyde;
3 - propanal; 4 - unknown; 5 - butanal; 6 - pentanal; 7 - 4HNAL;
8 - hexanal; 9 - heptanal; 10 - octanal; 11 - nonanal; 12 - decanal.
WWW.ESAINC.COM
275
The gradient analytical system consisted of two pumps, an autosampler, a fluorescence detector
and a data station.
LC Conditions
Column:
Mobile Phase A:
Mobile Phase B:
Gradient Conditions:
HR-80 (4.6 x 80mm; 3µm).
5% Tetrahydrofuran (THF).
40% THF.
Isocratic 0%B from 0 to 2min. Linear increase of phase B from 0 to 100%
from 2 to 55min. Isocratic 100% phase B from 55 to 60min. Linear
decrease of phase B from 100 to 0% from 60 to 65min. Isocratic 0%
phase B from 65 to 70min.
Flow Rate:
1.0mL/min.
Temperature:
Ambient.
Injection Volume:
20µL.
Detector:
Fluorescence, Model 305.
Excitation Wavelength: 380nm.
Emission Wavelength: 445nm.
See ESA Application Note 70-5041 4-Hydroxynonenal And Other Aldehydes.
CARBOHYDRATES.
Introduction.
Carbohydrates are aldehyde or ketone compounds with multiple hydroxyl groups.
Carbohydrates show a wide degree of structural diversity. Through the multiple
hydroxyl groups single monosaccharide units (e.g., glucose, and fructose) can
join together to form more complex polysaccharides (e.g., glycogen and starch).
Carbohydrates are biologically very important and play multiple roles in living
organisms. Carbohydrates act as energy stores, fuels and intermediates; along
with phosphate they form the backbone of DNA and RNA; they act as structural
elements in plants and bacteria; and as part of glycoproteins and glycolipids they
are involved in cell-cell recognition. Unlike DNA, proteins and lipids, where a
wealth of information has been generated, carbohydrates appear to play less of a
role in redox biochemistry. For this reason, and due to space limitations, this
section will concentrate on two areas of carbohydrate chemistry that are
important to redox biochemistry, damage of DNA and RNA sugars, and
glycation-glyoxidation reactions.
WWW.ESAINC.COM
276
O
O
HO
O
OH
PO
O
O
PO
PO
Aerobic Cleavage
From a C-5' Radical
Aerobic Cleavage
From a C-2' Radical
O
O
O
P
O
O
O
O
Base
P
O
O
O
H
HO
O
O
H
P
O
O
Base
O
H
P
O
O
O
Phosphate Release
from C-3'
Oxygen Free
Conditions
PO
Phosphate Release
from C-5'
Oxygen Free
Conditions
HO
O
O
HO
O
O
RO
O
O
O
O
HO
RO
CH3
CH3
O
O
O
O
HO
PO
Figure 3.34 ROS-Induced Damage To 2’-Deoxyribose Causes DNA
Strand Scission And The Formation Of Carbohydrate Fragments.
(Based On Breen And Murphy (1995). See also von Sonntag (1984)).
Ribose and Deoxyribose Damage.
Radiation damage and hydroxyl free radicals can abstract a hydrogen atom from
a carbohydrate molecule forming radical intermediates that can undergo further
decomposition (Figure 3.1). Damage to the sugar molecule is biologically
important as it can lead to DNA and RNA strand scission (von Sonntag (1984)). It
must be remembered though, as discussed above, that the hydroxyl free radical
WWW.ESAINC.COM
277
is much more likely to attack a base than a sugar moiety (Breen and Murphy
(1995)).
Numerous mechanisms have been proposed for DNA strand scission depending
upon which ROS is involved which hydrogen atom is abstracted from the
carbohydrate molecule and whether the reaction proceeds under aerobic or
anaerobic conditions. These pathways have been extensively reviewed
elsewhere (Breen and Murphy (1995); von Sonntag (1984)). Figure 3.34 presents
a simplified scheme of ROS-induced deoxyribose decomposition.
Carbohydrate fragments are not easily measured but several have been
determined in vitro using GC/MS approaches (Breen and Murphy (1995) and
references therein; Dizdaroglu (1991)).
Glycation, Glyoxidation, Advanced Glycation End Products (AGEs) and
Age-Related Pigments.
As discussed in Chapter 2 carbonyl compounds readily form Schiff bases with
amine groups located on proteins and DNA bases. One of the most abundant
carbonyl compounds found in man is glucose. This reducing monosaccharide
can slowly react with primary amines on proteins and DNA in a process called
non-enzymatic glycation. The first step in this process is the reversible formation
of a Schiff base (Figures 2.27). This can then be converted to an eneaminol
before undergoing an Amadori rearrangement (effectively converting the Nglycoside of the aldose into an N-glycoside of the ketose) forming a more stable
Amadori adduct (e.g., fructose-lysine) (Kikuchi et al., (2003). Non enzymatic
glycation is reversible in vitro however, enolization, dehydration, cyclization,
fragmentation and oxidation reactions24 form reactive intermediates that
ultimately lead to stable end products, termed advanced glycation end-products
(AGEs) (Fu et al. (1994); Monnier et al. (2003); Wells-Knecht et al. (1995)).
Examples of AGEs include the most abundant irreversible chemical modification,
Nε-(carboxymethyl)-lysine (CML), the protein cross-link, pentosidine, and
pyralline (Figure 2.27) (Dunn et al. (1990); Nagaraj et al. (1996); Sell et al.
(1991)). The resulting conformationally altered proteins often acquire a brown
color referred to as Maillard browning. For example, human cartilage is near
white at birth, but turns to dark brown in aged individuals.
AGEs can be determined by immunoassays and flouresence or more accurately
by GC-MS and in hydrolysates by HPLC-fluorescence (Friess et al., (2003)).
24
The exact mechanism behind the Maillard browning reaction remains elusive but several possible pathways have been
proposed. For example, Wolff and colleagues have suggested a role for superoxide, hydrogen peroxide and metalinduced hydroxyl free radical formation in a process called autoxidative glycosylation (Hunt et al., (1988); Wolff et al.
(1991)).
WWW.ESAINC.COM
278
ADVANCED
LIPID
PEROXIDATION
END PRODUCTS
(ALPEs)
SCHIFF BASE
AND
MICHAEL ADDUCT
OH
O
R
HYDROXYALKENAL
O
O
MALONDIALDEHYDE
R
SCHIFF
BASE
NH2
DIHYDROPYRIDINES
and
IMINOENAMINES
PROTEIN LYSYL GROUP
OH
MODIFICATIONS (CROSSLINK
AND NON-CROSSLINK)
O
R
REDUCING
SUGAR
SCHIFF
BASE
AMADORI
PRODUCT
ADVANCED
GLYCATION
END PRODUCTS
(AGEs)
Figure 3.35 The Formation Of AGEs And ALPs From Proteins.
AGEs are found to be increased in aging and are thought to contribute to
development of diabetic complications such as accelerated atherosclerosis and
microvascular disease. Furthermore, AGEs may be one of the major driving
forces in the development and progression of Alzheimer’s and Parkinson’s
diseases (Reddy et al., (2002); Smith et al. (1994, 1995); Yan et al. (1994)).
AGEs are thought to exert their cellular effects by binding to a receptor (RAGE)
located on the cell’s surface. RAGE consists of two parts, a novel integral
membrane protein in the immunoglobulin superfamily and a lactoferrin
polypeptide (Bucciarelli et al., (2002); Yan et al. (1994)). Binding of AGE to its
receptor increases the oxidative stress of the cell (as reflected by increased MDA
production), induces the transcription of NF-κB and induces heme oxygenase
mRNA. In this way, activation of RAGE is hypothesized to underlie diabetic
vascular disease (Wendt et al., (2003); Yan at al. (1994) and references therein).
Formation of age related pigment is the result of accumulated oxidative damage
over time (Figure 3.35). The reaction between various primary amines and lipid
peroxidation derived carbonyl compounds leads to the production of two ALPs,
lipofuscin and ceroid (Kikugawa and Beppu (1987)). Lipofuscin is the classical
age pigment of post-mitotic cells, whereas ceroid accumulates due to
pathological or experimental processes. Lipofuscin occurs as yellow-brown
irregular membrane-bound granules located in lysosomes. Lipofuscin contains
about 50% (by weight) protein, a lesser amount of lipid, <1% fluorophore(s) and
dolichol bound metals (iron, copper and aluminum). It has been hypothesized
that normally damaged membranes and proteins, a consequence of ROS/RNS
attack, are digested and recycled by lysosomes Harman (1989). However, as a
WWW.ESAINC.COM
279
result of age these processes become less efficient, resulting in deposition of
lipofuscin. As the disposal system becomes overloaded, ceroid is formed.
REFERENCES.
Abdelrahim, M., Morris, E., Carver, J., Facchina, S., White, A., and Verma, A. (1997). Liquid chromatography assay of
dityrosine in human cerebrospinal fluid. J. Chromatogr. B., 696, 175-182.
Abe, K., Pan, L.H., Watanabe, M., Konno, H., Kato, T., and Itoyama, Y. (1997). Upregulation of protein-tyrosine nitration in
the anterior horn cells of amyotrophic lateral sclerosis. Neuronal Res., 19, 124-128.
Abe, T., Isobe, C., Murata, T., Sato, C., and Tohgi, H. (2003). Alteration of 8-hydroxyguanosine concentrations in the
cerebrospinal fluid and serum from patients with Parkinson’s disease. Neurosci. Lett., 336, 105-108.
Abe, T., Tohgi, H., Isobe, C., Murata, T., and Sato, C. (2002). Remarkable increase in the concentration of 8hydroxyguanosine in cerebrospinal fluid from patients with Alzheimer’s disease. J. Neurosci. Res., 70, 447-450.
Acworth, I.N., Bailey, B.B., and Maher, T.J. (1998). The use of HPLC with electrochemical detection to monitor reactive
oxygen and nitrogen species, markers of oxidative damage and antioxidants: application to the neurosciences. In:
Neurochemical Markers of Degenerative Nervous Diseases and Drug Addiction. Qureshi, G., Parvez, H., Caudy, P.,
and Parvez, S. (Eds.). Progress in HPLC-HPCE, 7. VSP Publications, The Netherlands. Pp. 3-56.
Acworth, I.N., During, M.J., and Wurtman, R.J. (1987). Processes that couple amino acid availability to neurotransmitter
synthesis. In: Amino Acid Availability and Brain Function in Health and Disease. Huether, G. (Ed.). NATO ASI Series.
H: Cell Biology, 20. Springer-Verlag, New York. Pp117-136.
Acworth, I.N., Maher, T.J., Chaiyakul, P., and McCabe, D. (1997b). 3-Nitrotyrosine passage through the blood-brain
barrier (BBB): An in vivo microdialysis study using HPLC with coulometric array detection. Oxygen Society. San
Francisco.
Acworth, I.N., McCabe, D.R., and Maher, T.J. (1997). The analysis of free radicals, their reaction products, and
antioxidants. In: Oxidants, Antioxidants, and Free Radicals. Baskin, S.I., and Salem, H. (Eds.). Taylor and Francis,
Washington DC. Pp. 23-77.
Adachi, S., Takemoto, K., Hirosue, T., and Hosogai, Y. (1993). Spontaneous and 2-nitropropane induced levels of 8hydroxy-2’deoxyguanosine in liver DNA of rats fed iron-deficient or manganese- and copper-deficient diets.
Carcinogen., 14, 265-268.
Adachi, S., Zeisig, M., and Moller, L. (1995). Improvements in the analytical method for 8-hydroxydeoxyguanosine in
nuclear DNA. Carcinogen., 16, 253-258.
Aesbach, R., Amado, R., and Neukom, H. (1976). Formation of dityrosine cross-links in proteins by oxidation of tyrosine
residues. Biochem. Biophys. Acta, 439, 292-301.
Agarwal, S., Wee, J.J., Hadley, M., and Draper, H.H. (1994). Identification of a deoxyguanosine-malondialdehyde adduct
in rat and human urine. Lipids, 29, 429-432.
Aikens, J., and Dix, T.A. (1991). Perhydroxyl radical (HOO•) initiated lipid peroxidation. J. Biol. Chem., 266, 15091-15098.
Akaike, T., Noguchi, Y., Ijiri, S., Setoguchi, K., Suga, M., Zheng, Y.M., Dietzschold, B., and Maeda, H. (1996).
Pathogenesis of influenza virus-induced pneumonia: Involvement of both nitric oxide and oxygen radicals. Proc. Natl.
Acad. Sci. USA, 93, 2448-2453.
Alam, Z.I., Jenner, A., Daniel, S.E., Lees, A.J., Cairns, N., Marsden, C.D., Jenner, P., and Halliwell, B. (1997). Oxidative
DNA damage in the Parkinsonian brain: An apparent selective increase in 8-hydroxyguanine levels in substantia nigra.
J. Neurochem., 69, 1196-1203.
Alary, J., Gueraud, F., and Cravedi, J.P. (2003). Fate of 4-hydroxynonenal in vivo: Disposition and metabolic pathways.
Mol. Asp. Med., 24, 177-187.
Allesio, H.M. (1993). Exercise-induced oxidative stress. Med. Sci. Sports Exerc., 25, 218-224.
Althaus, J.S., Fici, G.J., Plaisted, F.J., Kezdy, F.J., Campbell, C.M., Hoogerheide, J.G., and VonVoigtlander., P.F., Protein
nitration by peroxynitrite: A method for monitoring nitric oxide neurotoxicity. Microchem. J., 56, 155-164.
Althaus, J.S., Schmidt, K.R., Fountain, S.T., Tseng, M.T., Carroll, R.T., Galatsis, P., and Hall, E.D. (2000). LC-MS/MS
detection of peroxynitrite-derived 3-nitrotyrosine in rat microvessels. Free Radic. Biol. Med., 29, 1085-1095.
Alvarez, B., Rubbo, H., Kirk, M., Barnes, S., Freeman, B., and Radi, R. (1996). Peroxynitrite-dependent tryptophan
nitration. Chem. Res. Toxicol., 9, 390-396.
Alves-Rodrigues, A., Gregori, L., and Figueiredo-Pereira, M.E. (1998). Ubiquitin, cellular inclusions and their role in
neurodegeneration. Trends. Neurosci., 21, 516-520.
Ames, B.N. (1989). Endogenous oxidative DNA damage, aging and cancer. Free Radic. Res. Comm., 7, 121-128.
Ames, B.N., Shigenaga, M.K., and Hagen, T.M. (1993). Oxidants, antioxidants, and the degenerative diseases of aging.
Proc. Natl. Acad. Sci., USA, 90, 7915-7922.
Amici, A., Levine, R.L., Tsai, L., and Stadtman, E.R. (1989). Conversion of amino acid residues in proteins and amino acid
homopolymers to carbonyl derivatives by metal-catalyzed oxidation reactions. J. Biol. Chem., 264, 3341-3346.
Amiconi, G., Scoli, F., Matarese, M., Verzili, D., and Brunori, M. (1985). Determination of methionine sulfoxide in proteins:
Comparison of a gas-chromatographic and electrophoretic method. J. Biochem. Biophys. Meths., 11, 241-249.
Anderson, M.M., Hazen, S.L., Hsu, F.F., and Heinecke, J.W. (1997). Human neutrophils employ the myeloperoxidasehydrogen peroxide-chloride system to convert hydroxy-amino acids into glycoaldehyde, 2-hydroxypropanal, and
acrolein. J. Clin. Invest., 99, 424-432.
Anfinsen, C.B. (1973). Principles that govern the folding of protein chains. Science, 181, 223-230.
WWW.ESAINC.COM
280
Anon – ESCODD (2000). Comparison of different methods of measuring 8-oxoguanine as a marker of oxidative DNA
damage. ESCODD (European Standards Committee on Oxidative DNA Damage). Free Radic. Res., 32, 333-341.
Anon- ESCODD (2002). Inter-laboratory validation of procedures for measuring 8-oxo-7,8-dihydroguanine/8-oxo-7,8-2’deoxyguanosine in DNA. Free Radic. Res., 36, 239-245.
Arai, H., Terao, J., Abdalla, D. S. P., Suzuki, T., and Takama, K. (1996). Coulometric detection in high-performance liquid
chromatographic analysis of cholesteryl ester hydroperoxides. Free Radic. Biol. Med., 20, 365-371.
Armstrong, D.A. (1990). In: Sulfur-centered Reactive Intermediates in Chemistry and Biology. Chatgilialoglu, C., and
Asmus, K.D. (Eds.). Plenum Press, New York. Pp. 121-134.
Armstrong, R.C., and Swallow, A.J. (1969). Pulse- and gamma-radiolysis of aqueous solutions of tryptophan. Radiat.
Res., 40, 563-579.
Arrigo, A.-P. (1998). Small stress proteins: Chaperones that act as regulators of intracellular redox state and programmed
cell death. Biol. Chem., 379, 19-26.
Aruoma, O.I., and Halliwell, B. (1987). Action of hypochlorous acid on the antioxidant protective enzymes superoxide
dismutase, catalase and glutathione peroxidase. Biochem. J., 248, 973-976.
Aruoma, O.I., and Halliwell, B. (Eds.). (1998). DNA and Free Radicals: Techniques, Mechanisms and Applications. OICA,
London.
Aruoma, O.I., Halliwell, B., and Dizdioglu, M. (1989). Iron ion-dependent modification of bases in DNA by the superoxide
radical generating system hypoxanthine/xanthine oxidase. J. Biol. Chem., 264, 13024-13028.
Ayala, A., and Cutler, R.G. (1996a). The utilization of 5-hydroxy-2-amino valeric acid as a possible marker of oxidized
arginine and proline residues. Free Radic. Biol. Med., 21, 65-80.
Ayala, A., and Cutler, R.G. (1996b). Comparison of 5-hydroxy-2-amino valeric acid with carbonyl group content as a
marker of oxidized protein in human and mouse liver tissues. Free Radic. Biol. Med., 21, 551-558.
Ayene, S.I., Al-Mehdi, A.B., and Fisher, A.B. (1993). Inhibition of lung tissue oxidation during ischemia/reperfusion by 2mercaptopropionylglycine. Arch. Biochem. Biophys., 303, 307-312.
Aznar, J., Santos, M.T., Valles, J., and Sala, J. (1983). Serum malondialdehyde-like material (MDA-LM) in acute
myocardial infarction. J. Clin. Pathol., 36, 712-715.
Bagasra, O., Michaels, F.H., Zheng, Y.M., Broboski, L.E., Spitsin, S.V., Fu, Z.F., Tawadros, R., and Koprowski, H. (1995).
Activation of the inducible form of nitric oxide synthase in the brains of patients with multiple sclerosis. Proc. Natl. Acad.
Sci. USA, 92, 12041-12045.
Baker, M.A., and He, S.Q. (1991). Elaboration of cellular DNA breaks by hydroperoxides. Free Radic. Biol. Med., 11, 563572.
Banks, B.A., Ischiropoulos, H., McClelland, M., Ballard, P.L., and Ballard, R.A. (1998). Plasma 3-nitrotyrosine is elevated
in premature infants who develop bronchopulmonary dysplasia. Pediatrics, 101, 870-874.
Baraldi, E., Ghiro, L., Piovan, V., Carraro, S., Ciabattani, G., Barnes, P.J., and Montuschi, P. Increased exhaled 8isoprostane in childhood asthma. Chest, 124, 25-31.
Bartsch, H., and Frank, N. (1996). Blocking the endogenous formation of N-nitroso compounds and related carcinogens.
IARC Sci. Pub., 139, 189-201.
Bashir, S., Harris, G., Denman, M.A., Blake, D.R., and Winyard, P.G. (1993). Oxidative DNA damage and cellular
sensitivity to oxidative stress in human autoimmune diseases. Ann. Rheum. Dis., 52, 659-666.
Basu, A.K., Loechler, E.L., Leadon, S.A., and Essigmann, J.M. (1989). Genetic effects of thymine glycol: Site-specific
mutagenesis and molecular modeling studies. Proc. Natl. Acad. Sci. USA, 86, 7677-7681.
Bashir, S., Harris, G., Denman, M.A., Blake, D.R., and Winyard, P.G. (1993). Oxidative DNA damage and cellular
sensitivity to oxidative stress in human autoimmune diseases. Ann. Rheum. Dis., 52, 659-666.
Beal, M.F. (1997). Oxidative damage in neurodegenerative diseases. Neuroscientist, 3, 21-27.
Beal, M.F., Ferrante, R.J., Browne, S.E., Matthews, R.T., Kowall, N.W., and Brown, R.H. (1997). Increased 3-nitrotyrosine
in both sporadic and familial amyotrophic lateral sclerosis. Ann. Neurol., 42, 644-654.
Beal, M.F., Ferrante, R.J., Henshaw, R., Matthews, R.T., Chan, P.H., Kowall, N.W., Epstein, C.J., and Schulz, J.B. (1995).
3-Nitropropionic acid neurotoxicity is attenuated in copper/zinc superoxide dismutase transgenic mice. J. Neurochem.,
65, 919-922.
Beal, M.F., Ferrante, R.J., Matthews, R.T., Kowall, N.W., and Brown, R.H. (1996). 3-Nitrotyrosine in ALS. Soc. Neurosci.,
22, 1943.
Beckman, J.S., Ye, Y.Z., Anderson, P.G., Chen, J., Accavitti, M.A., Tarpey, M.A., and White, C.R. (1994). Extensive
nitration of protein tyrosines in human atherosclerosis detected by immunohistochemistry. Biol. Chem. Hoppe Seyler,
375, 81-88.
Beckman, K.B., and Ames, B.N. (1999). Endogenous oxidative damage of mt DNA. Mutat. Res., 424, 51-58.
Beckman, K.B, Saljoughi, S., and Ames, B.N. (1998). Facile analysis of 8-oxoguanine in DNA with the use of Fapy
glycosylase coupled to HPLC-EC. Free Radic. Biol. Med., Supplement 1, S72.
Beehler, B.C., Przybyszewski, J., Box, H.B., and Kulesz-Martin, M.F. (1992). Formation of 8-hydroxydeoxyguanosine
within DNA of mouse keratinocytes exposed in culture to UVB and H2O2. Carcinogen., 13, 2003-2007.
Benjamin, I.J., and McMillan, D.R. (1998). Stress (heat shock) proteins. Molecular chaperones in cardiovascular biology
and disease. Circ. Res., 83, 117-132.
Berger, M., Cadet, J., Berube, R., Langlois, R., and van Lier, J.E. (1992). Reversed-phase high-performance liquid
chromatography-thermospray mass spectrometry of radiation-induced decomposition products of thymine and
thymidine. J. Chromatogr., 593, 133-138.
Bergtold, D.S., Berg, C.D., and Simic, M.G. (1990). Urinary biomarkers in radiation therapy and cancer. Adv. Exp. Med.
Biol., 264, 311-316.
Bermudez, E., Ferng, S.F., Castro, C.E., and Mustafa, M.G. (1999). DNA strand breaks caused by exposure to ozone and
nitrogen dioxide. Environ. Res., 81, 72-80.
Bernstien, C. (1998). Sex as a response to oxidative DNA damage. In: DNA and Free Radicals: Techniques, Mechanisms
and Applications. Aruoma, O.I., and Halliwell, B. (Eds.). OICA, London. Pp. 99-118.
WWW.ESAINC.COM
281
Bessho, T., Tano, K., Kasai, H., Ohtsuka, E., and Nishimura, S. (1993). Evidence of two DNA enzymes for repair of 8hydroxyguanine (7,8-dihydroguanine) in human cells. J. Biol. Chem., 268, 19416-19421.
Birnbaum, G.I., Lassota, P., and Shugar, D. (1984). 8-Chloroguanosine: Solid-state and solution conformations and their
biological implications. Biochem., 23, 5048-5053.
Bittrich, H., Matzig, A.K., Kraker, I., and Appel, K.E. (1993). NO2-induced DNA single strand breaks are inhibited by
antioxidative vitamins in V79 cells. Chem. Biol. Interact., 86, 199-211.
Bogdanov, M., Beal, M.F., Griffin, R., and Matson, W. (1998). Quantitative method for 8-hydroxy-2’deoxyguanosine in
biologicals. Mitochondrial Medicine Meeting, UCSD.
Bogdanov, M.B., Beal, M.F., McCabe, D.R., Griffin, R., and Matson, W. (1999). A carbon column based LCEC approach
to routine 8-hydroxy-2’deoxyguanosine measurements in urine and other biological matrices. A one year evaluation of
methods. Free Radic. Biol. Med., 27, 647-666.
Bogdanov, M.B., Brown, R.H., Matson, W., Smart, R., Hayden, D., O-Donnell, H., Beal, M.F., and Cudkowicz, M. (2000).
Increased oxidative damage to DNA in ALS patients. Free Radic. Biol. Med., 29, 652-658.
Bogdanov, M.B., Matson, W.R., and Acworth, I.N. (2003). The Use of HPLC/EC for the Measurement of Oxidative DNA
Damage. In: Critical Reviews of Oxidative Stress and Aging. Volume 1.Cutler, R.G., and Rodriguez, H. (Eds.). World
Scientific, NJ. Pp. 203-221.
Bohley, P., and Seglen, P.O. (1992). Proteases and proteolysis in the lysosome. Experientia, 48, 151-157.
Bohr, V.A., Stevnsner, T., and de Souza-Pinto, N.C. (2002). Mitochondrial DNA repair of oxidative damage in mammalian
cells. Gene, 286, 127-134.
Bohr, V.A., Taffe, B.G., and Larminat, F. (1995). DNA repair, oxidative stress and aging. In: Oxidative Stress and Aging.
Cutler, R.G., Packer, L., Bertram, J., and Mori, A. (Eds.). Birkhauser, Basel. Pp. 45-54.
Boiteux, S., and Radicella, J.P. (1999). Base excision repair of 8-hydroxyguanine protects DNA from endogenous
oxidative stress. Biochimie, 81, 59-67.
Boiteux, S., Gajewski, E., Laval, J., and Dizdaroglu, M. (1992). Substrate specificity of the E. coli Fpg protein
(formamidopyrimidine-DNA glycosylase): Excision of purine lesions in DNA produced by ionizing radiation or
photosensitization. Biochem., 31, 106-110.
Bose, R., Jilling, T., Caplan, M., Derrick, M., and Tan, S. (1999). Electrochemical detection of nitrotyrosine in fetal brains
following repetitive uterine ischemia. Poster #207. Free Radic. Biol. Med., 27, S73.
Bosch-Morell, F., Flohe, L., Marin, N., and Romero, F.J. (1999). 4-Hydroxynonenal inhibits glutathione peroxidase:
Protection by glutathione. Free Radic. Biol. Med., 26, 1383-1387.
Brawn, K., and Fredovich, I. (1981). DNA strand scission by enzymatically generated oxygen radicals. Arch. Biochem.
Biophys., 206, 414-419.
Breen, A.P., and Murphy, J.A. (1995). Reactions of oxyl radicals with DNA. Free Radic. Biol. Med., 18, 1033-1077.
Brennan, M.L., Wu, W., Fu, X., Shen, Z., Song, W., Frost, H., Vadseth, C., Narin, G., Lenkiewicz, E., Borchers M.T.,
Lusis, A.J., Lee, J.J., Lee, N.A., Abu-soud, H.M., Ischiropoulos, H., and Hazen, S.L. (2002). J. Biol. Chem., 277, 1741517427.
Brigelius-Flohe, R., Auman, K.-D., Blocker, H., Gross, G., Kiess, M., Kloppel, K-D., Maiorino, M., Roveri, A., Schuckelt, R.,
Ursini, F., Wingender, E., and Flohe, L. (1994). Phospholipid hydroperoxide glutathione peroxidase: Genomic DNA,
cDNA, and deduced amino acid sequence. J. Biol. Chem., 269, 7342-7348.
Briviba, K., Kissner, R., Koppenol, W.H., and Sies, H. (1998). Kinetic study of the reaction of glutathione peroxidase with
peroxynitrite. Chem. Res. Toxicol., 11, 1398-1401.
Brot, N., and Weissbach, H. (1991). Biochemistry of methionine sulfoxide residues in proteins. Biofactors, 3, 91-96.
Brown, R.K., McBurney, A., Lunec, J., and Kelly, F.J. (1995). Oxidative damage to DNA in patients with cystic fibrosis.
Free Radic. Biol. Med., 18, 801-806.
Browne, R.W., and Armstrong, D. (1998). Separation of hydroxy and hydroperoxy polyunsaturated fatty acids by highperformance liquid chromatography. In: Methods in Molecular Biology, 108. Armstrong, D. (Ed.). Humana Press,
Totowa, New Jersey. Pp. 147-155.
Bruijn, L.I., Beal, M.F., Becher, M.W., Schulz, J.B., Wong, P.C., Price, D.L., and Cleveland, D.W. (1997). Elevated free
nitrotyrosine levels, but not protein-bound nitrotyrosine or hydroxyl radicals, throughout amyotrophic lateral sclerosis
(ALS)-like disease implicate tyrosine nitration as an aberrant in vivo property of one familial ALS-linked superoxide
dismutase 1 mutant. Proc. Natl. Acad. Sci. USA, 94, 7606-7611.
Bucciarelli, L.G., Wendt, T., Rong, L., lalla, E, Hoffman, M.A., Goova, M.T., Taguchi, A., Yan, S.F., Stern, D.M., and
Schmidt, A.M. (2002). RAGE is a multiligand receptor of the immunoglobulin superfamily: Implications for homeostasis
and chronic disease. Cell. Mol. Life Sci., 59, 1117-1128.
Buchner, J. (1996). Supervising the fold: Functional principles of molecular chaperones. FASEB J., 10, 10-19.
Bunik, V.I., and Sievers, C. (2002). Inactivation of the 2-oxoacid dehydrogenase complexes upon generation of intrinsic
radical species. Eur. J. Biochem., 269, 5004-5015.
Burney, S., Caulfield, J.L., Niles, J.C., Wishnok, J.S., and Tannenbaum, S.R. (1999). The chemistry of DNA damage from
nitric oxide and peroxynitrite. Mutat. Res., 424, 37-49.
Buttery, L.D., Springall, D.R., Chester, A.H., Evans, T.J., Standfield, E.N., Parums, D.V., Yacoub, M.H., and Poalk, J.M.
(1996). Inducible nitric oxide synthase is present within human atherosclerotic lesions and promotes the formation and
activity of peroxynitrite. Lab. Invest., 75, 77-85.
Byun, J., Henderson, J.P., Mueller, D.M., and Heinecke, J.W. (1998). 8-Nitro-2’deoxyguanosine is generated by the
myeloperoxidase-H2O2-nitrite generating system of activated human phagocytes. Biochem., 38, 2590-2600.
Cadet, J., and Weinfeld, M. (1993). Detecting DNA damage. Anal. Chem., 65, A675-A682.
Cadet, J., Delatour, T., Douki, T., Gasparutto, D., Pouget, J.P., Ravanat, J.L., and Sauvaigo, S. (1999). Hydroxyl radicals
and DNA base damage. Mutat. Res., 424, 9-21.
Cadet, J., D’Ham, C., Douki, T., Pouget, J.-P., Ravanat, J.-L., and Sauvaigo, S. (1998). Facts and artifacts in the
measurement of oxidative base damage to DNA. Free Rad. Res., 29, 541-550.
WWW.ESAINC.COM
282
Cadet, J., Douki, T., and Ravanat, J.L. (1997). Artifacts associated with the measurement of oxidized DNA bases.
Environ. Health Perspect., 105, 1034-1039.
Cadet, J., Odin, F., Mouret, J-F., Polverelli, M., Audic, A., Giacomono, P., Richard, M-J., and Favier, A. (1992). Chemical
and biochemical postlabeling methods for singling out specific oxidative DNA lesions. Mutat. Res., 275, 343-354.
32
Cajigas, A., Gayer, M., Beam, C., and Steinberg, J.J. (1994). Ozonation of DNA forms adducts: A P-DNA labeling and
thin-layer chromatography technique to measure DNA environmental biomarkers. Arch. Environ. Health, 49, 25-36.
Calderon-Garciduenas, L., Osnaya, N., Rodriguez-Alcaraz, A., and Villarreal-Calderon, A. (1997). DNA damage in nasal
respiratory epithelium from children exposed to urban pollution. Environ. Mol. Mutagen., 30, 11-20.
Calingasan, N.Y., Uchida, K., and Gibson, G.E. (1999). Protein-bound acrolein: A novel marker of oxidative stress in
Alzheimer’s disease. J. Neurochem., 72, 751-756.
Cao, G., and Cutler, R.G. (1995). Protein oxidation and aging. Difficulties in measuring reactive protein carbonyls in
tissues using 2,4-dinitrophenylhydrazine. Arch. Biochem. Biophys., 320, 106-114.
Capella, G., Cronauer-Mitra, S., Pienado, M.A., and Peruscho, M. (1991). Frequency and spectrum mutations at codon 12
and 13 of the c-K-ras gene in human tumors. Environ. Health Perspect., 93, 125-131.
Carmichael, P.L., Hewer, A., Osborne, M.R., Strain, R.J., and Phillips, D.H. (1995). Detection of bulky DNA lesions in the
liver of patients with Wilson’s disease and primary haemochromatosis. Mutation Res., 326, 235-243.
Carney, J.M., Starke, R.P., Oliver, C.N., Landum, R.W., Cheng, M.S., Wu, J.F., and Floyd, R.A. (1991). Excess brain
protein oxidation and enzyme dysfunction in normal aging and in Alzheimer’s disease. Proc. Natl. Acad. Sci. USA, 88,
3633-3636.
Cathcart, R., Schwiers, E., Saul, R.L., and Ames, B.N. (1984). Thymine glycol and thymidine glycol in human and rat
urine: a possible assay for oxidative DNA damage. Proc. Natl. Acad. Sci. USA, 81, 5633-5637.
Cattley, R.C., and Glover, S.E. (1993). Elevated 8-hydroxydeoxyguanosine in hepatic DNA of rats following exposure to
peroxisome proliferators: Relationship to carcinogenesis and nuclear localization. Carcinogen., 14, 2495-2499.
Cawood, P., Wickens, D.G., Iversen, S.A., Braganza, J.M., and Dormandy, T.L. (1983). The nature of diene conjugation in
human serum, bile and duodenal juice. FEBS Lett., 162, 239-243.
Chang, Y.H., Abdalla, D.S., and Sevanian, A. (1997). Characterization of cholesterol oxidation products formed by
oxidative modification of low density lipoprotein. Free Radic. Biol. Med., 23, 202-214.
Chao, C.-C., Ma, Y.-S., and Stadtman, E.R. (1997). Modification of protein surface hydrophobicity and methionine
oxidation by oxidative systems. Proc. Natl. Acad. Sci. USA, 94, 2969-2974.
6
Chaudhary, A.K., Reddy, G.R., Blair, I.A., and Marnett, L.J. (1996), Characterization of an N -oxoprenyl-2’deoxyadenosine adduct in malondialdehyde-modified DNA using liquid chromatography/electrospray ionization tandem
mass spectrometry. Carcinogen., 17, 1167-1170.
Chen, J.J., Bertrand, H., and Yu, B.P. (1995). Inhibition of adenine nucleotide translocator by lipid peroxidation products.
Free Radic. Biol. Med., 19, 583-590.
Chiavari, G., and Bergamini, C. (1985). High-performance liquid chromatography of carbonyl compounds as 2,4dinitrophenylhydrazones with electrochemical detection. J. Chromatogr., 318, 427-432.
Chou, S.M., Wang, H.S., and Komai, K. (1996). Colocalization of NOS and SOD1 in neurofilament accumulation within
motor neurons of amyotrophic lateral sclerosis: an immunohistochemical study. J. Chem. Neuroanat., 10, 249-258.
Chou, S.M., Wang, H.S., and Taniguchi, A. (1996). Role of SOD-1 and nitric oxide/cyclic GMP cascade on neurofilament
aggregation in ALS/MND. J. Neurol. Sci., 139, 16-26.
Chung, M.H., Kasai, H., Nishimura, S., and Yu, B.P. (1992). Protection of DNA damage by dietary restriction. Free Radic.
Biol. Med., 12, 523-525.
Cini, M., Fariello, R.G., Bianchetti, A., and Moretti, A. (1994). Studies on lipid peroxidation in the rat brain. Neurochem.
Res., 19, 283-288.
Claycamp, H.G. (1992). Phenol sensitization of DNA to subsequent oxidative damage in 8-hydroxyguanine assays.
Carcinogen., 13, 1289-1292.
Claycamp, H.G., and Ho, K.-K. (1993). Background and radiation-induced 8-hydroxy-2’-deoxyguanosine in γ-irradiated
Escherichia coli. Int. J. Radiat. Biol., 63, 597-607.
Clayton, D.A., Doda, J.N., and Friedberg, E.C. (1974). The absence of a pyrimidine dimer repair mechanism in
mammalian mitochondria. Proc. Natl. Acad. Sci. USA, 71, 2777-2781.
Climent, I., Tsai, L., and Levine, R.L. (1989). Derivatization of gamma-glutamyl semialdehyde residues in oxidized
proteins by fluoresceinamine. Anal. Biochem., 182, 226-232.
Cohen, J.S., Yariv, J., Kalb, A.J., Jacobson, L., and Shechter, Y. (1979). 13C NMR analysis of methionine sulfoxide in
protein. J. Biochem. Biophys. Meths., 1, 145-151.
Coles, B., Bloodsworth, A., Clark, S.R., Lewis, M.J., Cross, A.R., Freeman, B.A., and O’Donnell, V.B. (2002). Circ. Res.,
91, 375-381.
Collins, A.R. (1998). The comet assay: A novel approach to measuring DNA oxidation. In: DNA and Free Radicals:
Techniques, Mechanisms and Applications. Aruoma, O.I., and Halliwell, B. (Eds.). OICA, London. Pp. 241-260.
Collins, A.R., Dusinka, M., Gedik, C.M., and Setina, R. (1996). Oxidative damage to DNA: Do we have a reliable
biomarker? Environ. Health Perspect. 104, 465-469.
Collins, A.R., Duthie, S.J., and Donson, V.L. (1993). Direct enzymic detection of endogenous oxidative base damage in
human lymphocyte DNA. Carcinogen.,14, 1733-1735.
Collins, A.R., Gedik, C.M., Olmedilla, B., Southon, S., and Bellizzi, M. (1998a). Oxidative DNA damage measured in
human lymphocytes: Large differences between the sexes and between countries, and correlations with heart disease
mortality rates. FASEB J., 12, 1397-1400.
Collins, A.R., Raslova, K., Somorovska, M., Petrovska, H., Ondrusova, A., Vohnout, B., Fabry, R., and Dusinska, M.
(1998). DNA damage in diabetes: Correlations with a clinical marker. Free Radic. Biol. Med., 25, 373-377.
Cooke, M.C., Evans, M.D., Podmore, I.D., Herbert, K.E., Mistry, N., Mistry, P., Hickenbotham, P.T., Hussieni, A., Griffiths,
H.R., and Lunec, J. (1998). Novel repair action of vitamin C upon in vivo oxidative DNA damage. FEBS Letts., 363,
363-367.
WWW.ESAINC.COM
283
Cooper, B., Creeth, J.M., and Donald, A.S. (1985). Studies on the limited degradation of mucus glycoproteins. The
mechanism of the peroxide reaction. Biochem. J., 228, 615-626.
Cooke, M.C., Mistry, N., Ahmad, J., Waller, H., Langford, L., and Lunec, J. (2003). Deoxycytidine glyoxal: Lesion induction
and evidence of repair following vitamin C. Free Radic. Biol. Med., 34, 218-225.
Copley, S.D., Frank, E., Kirsch, W.M., and Koch, T.H. (1992). Detection and possible origins of aminomalonic acid in
protein hydrolysates. Anal. Biochem., 201, 152-157.
Cordis, G.A., Bagchi, D., Maulik, N., and Das, D.K. (1994). High-performance liquid chromatographic method for the
simultaneous detection of malondialdehyde, acetaldehyde, formaldehyde, acetone and propionaldehyde to monitor the
oxidative stress in heart. J. Chromatogr. A., 661, 181-191.
Cordis, G.A., Maulik, G., Bagchi, D., Riedel, W., and Das, D.K. (1998). Detection of oxidative DNA damage to ischemic
reperfused rat hearts by 8-hydroxydeoxyguanosine formation. J. Mol. Cardiol., 30, 1939-1944.
Coux, O., Tanaka, K., and Goldberg, A.L. (1996). Structure and functions of the 20S and 26S proteasomes. Ann. Rev.
Biochem., 65, 801-847.
Creeth, J.M., Cooper, B., Donald, B.S., and Clamp, J.R. (1983). Studies of the limited degradation of mucus
glycoproteins. The effects of dilute hydrogen peroxide. Biochem. J., 211, 323-332.
Cross, A.H., Manning, P.T., Stern, M.K., and Misko, T.P. (1997). Evidence for the production of peroxynitrite in
inflammatory CNS demyelination. J. Neuroimmunol., 80, 121-130.
Crow, J.P. (1999). Measurement and significance of free and protein-bound 3-nitrotyrosine, 3-chlorotyrosine, and free 3nitro-4-hydroxyphenylacetic acid in biological samples: A high-performance liquid chromatography method using
electrochemical detection. Meth. Enzymol., 301, 151-162.
Crow, J.P., and Beckman, J.S. (1995). Reactions between nitric oxide, superoxide, and peroxynitrite: Footprints of
peroxynitrite in vivo. Adv. Pharmacol., 34, 17-30.
Crow, J.P., and Ischiropoulos, H. (1996). Detection and quantitation of nitrotyrosine residues in proteins: In vivo marker of
peroxynitrite. Meth. Enzymol., 269, 185-194.
Crow, J.P., Beckman, J.S., and McCord, J.M. (1995). Sensitivity of the essential zinc-thiolate moiety of yeast alcohol
dehydrogenase to hypochlorite and peroxynitrite. Biochem., 34, 3544-3552.
Crow, J.P., Ye, Y.Z., Kirk, M., Barnes, S., and Beckman, J.S. (1997). Superoxide dismutase catalyzes nitration of
tyrosines by peroxynitrite in the rod and head domains of neurofilament-L. J. Neurochem., 69, 1945-1953.
Crowley, J.R., Yarasheski, K., Leeuwenburgh, C., Turk, J., and Heinecke, J.W. (1998). Isotope dilution mass
spectrometric quantification of 3-ntriotyrosine in proteins and tissues is facilitated by reduction to 3-aminotyrosine. Anal.
Biochem., 259, 127-135.
Cucurou, C., Battioni, J.P., Thang, D.C., Nam, N.H., and Mansuy, D. (1991). Mechanisms of inactivation of lipoxygenases
by phenidone and BW755C. Biochem., 30, 8964-8970.
Culp, S.J., Cho, B.P., Kadlubar, F.F., and Evans, F.E. (1989). Structural and conformational analyses of 8-hydroxy-2’deoxyguanosine. Chem. Res. Toxicol., 2, 416-422.
Cushnir, J.R., Naylor, S., Lamb, J.H., and Farmer, P.B. (1993). Tandem mass spectrometric approaches for the analysis
of alkylguanines in human urine. Org. Mass Spec., 28, 552-558.
Daiber, A., Harold, S., Schoneich, C., Namgaladze, D., Peterson, J.A., and Ullrich, V. (2000). Nitration of cytochrome
P450BM-3 by peroxynitrite. Stopped-flow measurement prove ferryl intermediates. Eur. J. Biochem., 267, 6729-6739.
Daiber, A., Mehl, M., and Ullrich, V. (1998). New aspects in the reaction mechanism of phenol with peroxynitrite: The role
of phenoxy radicals. Nitric Oxide, 2, 259-269.
Dakin, H.D. (1906). The oxidation of amino-acids with the production of substances of biological importance. J. Biol.
Chem., 1, 171-176.
Dakin, H.D. (1908). The oxidation of leucine, α-amido-n-valeric acid with hydrogen peroxide. J. Biol. Chem., 4, 63-76.
Dandona, P., Thusu, K., Cook, S., Snyder, B., Makowski, J., Armstrong, D., and Nicotera, T. (1996). Oxidative damage to
DNA in diabetes mellitus. Lancet, 347, 444-445.
Dawson, T.M., and Dawson, V.L. (1994). Nitric oxide: Actions and pathological roles. Neurosci., Preview Issue. 9-19.
Dean, R.T. (1978). Lysosomes and protein degradation. Ciba Found. Symp., 75, 139-149.
Dean, R.T., Fu, S., Giesag, G., and Armstrong, S.G. (1996). In: Free Radicals: A practical Approach. Punchard, N.A., and
Kelly, F.J. (Eds.). IRL Press, Oxford. Pp. 171-183.
Dean, R.T., Fu, S., Stocker, R., and Davies, M.J. (1997). Biochemistry and pathology of radical-mediated protein
oxidation. Biochem. J., 324, 1-18.
Degan, P., Bonassi, S., De Caterina, M., Korkina, L.G., Pinto, L., Scopacasa, F., Zatterale, A., Calzone, R., and Pagano,
G. (1995). In vivo accumulation of 8-hydroxy-2’-deoxyguanosine in DNA correlates with release of reactive oxygen
species in Fanconi’s anemia families. Carcinogen., 16, 735-742.
Degan, P., Shigenaga, M.K., Park, E.-M., Alperin, P.E., and Ames, B.N. (1991). Immunoaffinity isolation of urinary 8hydroxy-2’deoxyguanosine and 8-hydroxyguanine and quantitation of 8-hydroxy-2’deoxyguanosine in DNA by
polyclonal antibodies. Carcinogen., 12, 865-871.
De Flora, S., Bennicelli, C., Zanacchi, P., D’Agnostini, F., and Camoirano, A. (1989). Mutagenicity of active oxygen
species in bacteria and its enzymatic or chemical inhibition. Mutat. Res., 214, 153-158.
Delori, F.C., and Dorey, C.K. (1998). In vivo technique for autofluorescent lipopigments. In: Methods in Molecular Biology,
108. Armstrong, D. (Ed.). Humana Press, Totowa, New Jersey. Pp. 229-243.
De Martinis, B.S., and Bianchi, M.D. (2002). Methodology for urinary 8-hydroxy-2’-deoxyguanosine analysis by HPLC with
electrochemical detection. Pharmacol. Res., 46, 129-131.
Demple, B., and Harrison, L. (1994). Repair of oxidative damage to DNA: Enzymology and Biology. Ann. Rev. Biochem.,
63, 915-948.
Denda, A., Tang, Q., Endoh, T., Tsujiuchi, T., Horiguchi, K., Noguchi, O., Mizumoto, Y., Nakae, D., and Konishi, Y. (1994).
Prevention by acetylsalicylic acid of liver cirrhosis and carcinogenesis as well as generations of 8hydroxydeoxyguanosine and thiobarbituric acid-reactive substances caused by a choline-deficient, L-amino aciddefined diet in rats. Carcinogen., 15 1279-1283.
WWW.ESAINC.COM
284
Deng, X.-S., Tuo, J., Poulsen, H.E., and Loft, S. (1998). Prevention of oxidative DNA damage in rats by brussel sprouts.
Free Rad. Res., 28, 323-333.
De Zwart, L.L., Meerman, J.H.N., Commander, J.N.M., and Vermeulen, N.P.E. (1998). Biomarkers of free radical damage
applications in experimental animals and in humans. Free Radic. Biol. Med., 26, 202-226.
De Zwart, L.L., Venhorst, J., Groot, M., Commandeur, J.N.M., Hermanns, R.C.A., Meerman, J.H.M., Van Baar, B.L.M.,
and Vermeulen, P.E. (1997). Simultaneous determination of eight lipid peroxidation degradation products in urine of
rats treated with carbon tetrachloride using gas chromatography with electron-capture detection. J. Chromatogr. B, 694,
277-287.
Dice, J.F., and Terlecky, S.R. (1990). Targeting of cytosolic proteins to lysosomes for degradation. Crit. Rev. Ther. Drug
Carrier. Syst., 7, 211-233.
Dizdaroglu, M. (1991). Chemical determination of free radical induced damage to DNA. Free Radic. Biol. Med., 10, 225242.
Dizdaroglu, M. (1993). In: DNA and Free Radicals. Halliwell, B., and Aruoma, O.I. (Eds.). Ellis Horwood, Chichester. Pp.
19-39.
Dizdaroglu, M. (1993b). Quantitative determination of oxidative base damage in DNA by stable isotope-dilution mass
spectrometry. FEBS Lett., 315, 1-6.
Dizdaroglu, M. (1994). Chemical determination of oxidative DNA damage by gas chromatography-mass spectrometry.
Meth. Enzymol., 234, 3-16.
Dizdaroglu, M. (1997). Measurement of oxidized DNA bases using gas chromatography/mass spectrometry. Workshop.
Oxygen Society, San Francisco, CA.
Dizdaroglu, M. (1998). Mechanisms of free radical damage to DNA. In: DNA and Free Radicals: Techniques, Mechanisms
and Applications. Aruoma, O.I., and Halliwell, B. (Eds.). OICA, London. Pp. 3-26.
Dizdaroglu, M., Jaruga, P., and Rodriguez, H. (2003). Oxidative DNA Damage to DNA: Mechanisms of Product Formation
and Measurement by Mass Spectrometric Techniques. In: Critical Reviews of Oxidative Stress and Aging. Volume
1.Cutler, R.G., and Rodriguez, H. (Eds.). World Scientific, NJ. Pp. 165-189.
Dizdaroglu, M., Laval, J., and Boiteux, S. (1993). Substrate specificity of the Escherichia coli endonuclease III: excision of
thymine- and cytosine-derived lesions in DNA produced by radiation-generated free radicals. Biochem., 32, 1210512111
Djuric, Z., Heilbrun, L.K., Reading, B.A., Boomer, A., Valeriete, F.A., and Martino, S. (1991). Effects of low-fat diet on
levels of oxidative damage to DNA in human peripheral nucleated blood cells. J. Natl. Cancer Inst., 83, 766-769.
Djuric, Z., Lu, M.H., Lewis, S.M., Luongo, D.A., Chen, X.W., Heilbrun, L.K., Reading, B.A., Duffy, P.H., and Hart, R.W.
(1992). Oxidative DNA damage levels in rats fed low-fat, high-fat, or caloric-restricted diets. Toxicol. Appl. Pharmacol.,
115, 156-160.
Djuric, Z., Luongo, D.A., and Harper, D.A. (1991b). Quantitation of 5-(hydroxymethyl)uracil in DNA by gas
chromatography with mass spectral detection. Chem. Res. Toxicol., 4, 687-691.
Domigan, N.M., Charlton, T.S., Duncan, M.W., Winterbourne, C.C., and Kettle, A.J. (1995). Chlorination of tyrosyl
residues in peptides by myeloperoxidase and human neutrophils. J. Biol. Chem., 270, 16542-16548.
Dormandy, T.L., and Wickens, D.G. (1987). The experimental and clinical pathology of diene conjugation. Chem. Phys.
Lipids, 45, 353-364.
2
Douki, T., and Ames, B.N. (1994). An HPLC-EC assay for 1,N -propano adducts of 2’-deoxyguanosine with 4hydroxynonenal and other α,β-unsaturated aldehydes. Chem. Res. Toxicol., 7, 511-518.
Douki, T., Delatour, T., Bianchini, F., and Cadet, J. (1996). Observation and prevention of an artifactual formation of
oxidized DNA bases and nucleosides in the GC-EIMS method. Carcinogen., 17, 347-353.
Douki, T., Ravanat, J.-L., Frelon, S., Bourdat, J.-P., and Cadet, J. (2003). HPLC-MS/MS Measurement of Oxidative Base
Damage to Isolated and Cellular DNA. In: Critical Reviews of Oxidative Stress and Aging. Volume 1.Cutler, R.G., and
Rodriguez, H. (Eds.). World Scientific, NJ. Pp. 190-202.
Driscoll, J. (1994). The role of the proteasome in cellular degradation. Histol. Histopathol., 9, 197-202.
Duarte, V., Muller, J.G., and Burrows, C.J. (1999). Insertion of dGMP and dAMP during in vitro DNA synthesis opposite
an oxidized form of 7,8-dihydro-8-oxoguanine. Nucelic Acids Res., 27, 496-502.
Ducrocq, C., Dendane, M., Laprevote, O., Serani, L., and Das, B.C. (1998). Chemical modification of the vasoconstrictor
peptide angiotensin II by nitrogen oxides (NO, HNO2, HOONO) – evaluation by mass spectrometry. Eur. J. Biochem.,
253, 146-153.
Duda, J.E., Giasson, B.I., Chen, Q., Gur, T.L., Hurtig, H.I., Stern, M.B., Gollomp, S.M., Ischiropoulos, H., Lee, V.M., and
Trojanowski, J.Q. (2000). Am. J. Pathol., 175, 1439-1445.
Dunn, J.A., Ahmed, M.U., Murtiashaw, M.H., Richardson, J.M., Walla, M.D., Thorpe, S.R., and Baynes, J.W. (1990).
Reaction of ascorbate with lysine and protein under autoxidizing conditions: Formation of Nε-(carboxymethyl)lysine by
reaction between lysine and products of autoxidation of ascorbate. Biochem., 29, 10964-10970.
Eder, E., and Hoffman, C. (1993). Identification and characterization of deoxyguanosine adducts of mutagenic β-alkylsubstituted acrolein congeners. Chem. Res. Toxicol., 6, 486-494.
Eiserich, J.P., Cross, C.E., Jones, A.D., Halliwell, B., and van der Vleit, A. (1996). Formation of nitrating and chlorinating
species by reaction of nitrite with hypochlorous acid. A novel mechanism for nitric oxide-mediated protein modification.
J. Biol. Chem., 271, 19199-19208.
Eiserich, J.P., Hristova, M., Cross, C.E., Freeman, B.A., and van der Vleit, A. (1998). Formation of nitric oxide-derived
inflammatory oxidants by myeloperoxidase in neutrophils. Nature, 391, 393-398.
Emerit, I. (1994). Reactive oxygen species, chromosome mutation, and cancer: Possible role of clastogenic factors in
carcinogenesis. Free Radic. Biol. Med., 16, 99-109.
Epe, B. (1993). In: DNA and Free Radicals. Halliwell, B., and Aruoma, O.I. (Eds.). Ellis Horwood, Chichester. Pp. 41-65.
Epe, B., Ballmaier, D., Roussyn, I., Briviba, K., and Sies, H. (1996). DNA damage by peroxynitrite characterized with DNA
repair enzymes. Nucleic Acid Res., 24, 4105-4110.
WWW.ESAINC.COM
285
Erhola, M., Toyokuni, S., Okada, K., Tanaka, T., Hiai, H., Ochi, H., Uchida, K., Osawa, T., Niemmen, M.M., Alho, H.,
Kellokumu-Lehtinen, P. (1997). Biomarker evidence of DNA oxidation in lung cancer patients: Association of urinary 8hydroxy-2'-deoxyguanosine excretion with radiotherapy, chemotherapy and response to treatment. FEBS Lett., 409,
287-291.
ESCODD (2002). Comparative analysis of baseline 8-oxo-7,8-dihydroguanine in mammalian cell DNA, by different
methods in different laboratories: An approach to consensus. Carcinogen., 23, 2129-2133.
ESCODD (2003). Measurement of DNA oxidation in human cells by chromatographic and enzymic methods. Free Radic.
Biol. Med., 34, 1089-1099.
Esterbauer, H., Eckl, P., and Ortner, A. (1990). Possible mutagens derived from lipids and lipid precursors. Mutat. Res.,
238, 223-233.
Esterbauer, H., Jurgens, G., Quehenberger, O., and Koller, E. (1987). Autooxidation of human low-density lipoprotein:
Loss of polyunsaturated fatty acids and vitamin E and generation of aldehydes. J. Lipid. Res., 28, 495-509.
Esterbauer, H., Zollner, H., and Schaur, R.J. (1990). Aldehydes formed by lipid peroxidation: Mechanisms of formation,
occurrence and determination. In: Membrane Lipid Oxidation. Vigo-Pelfry, C. (Ed.). Boca Raton, FL, CRC Press, 1. Pp.
239-283.
Esterbauer, H., Schaur, R.J., and Zollner, H. (1991). Chemistry and biochemistry of 4-hydroxynonenal, malonaldehyde
and related aldehydes. Free Radic. Biol. Med., 11, 81-128.
Evans, J., Maccabee, M., Hatahet, Z., Courcelle, J., Bockrath, R., Ide, H., and Wallace, S. (1993). Thymine ring saturation
and fragmentation products: lesion bypass, misinsertion and implications for mutagenesis. Mutat. Res., 299, 147-156.
Evans, P., Lyras, L., and Halliwell, B. (1998). Measurement of protein carbonyls in human brain tissues. Meths. Enzymol.,
300, 145-156.
Fairbairn, D.W., Olice, P.L., and O’Neill, K.L. (1995). The comet assay: A comprehensive review. Mutat. Res., 339, 37-59.
Fam, S.S., and Morrow, J.D. (2003). The isoprostanes: Unique products of arachidonic acid oxidation – a review. Curr.
Med. Chem., 10, 1723-1740.
Farber, J.M., and Levine, R.L. (1986). Sequence of a peptide susceptible to mixed-function oxidation: Probable cation
binding site in glutamine synthetase. J. Biol. Chem., 261, 4574-4578.
Faux, S.P., Francis, J.E., Smith, A.G., and Chipman, J.K. (1992). Induction of 8-hydroxydeoxyguanosine in Ah-responsive
mouse liver by iron and Aroclor 1254. Carcinogen., 13, 247-250.
Faux, S.P., Gao, M., Aw, T.C., and Braithwaite, R.A. (1994). Molecular epidemiological studies in workers exposed to
chromium-containing compounds. Clin. Chem., 40, 1454-1456.
Feig, D.I., and Loeb, L.A. (1993). Mechanisms of mutation by oxidative DNA damage: Reduced fidelity of mammalian
DNA polymerase beta. Biochem., 32, 4466-4473.
Feinstein, E., Canaani, E., and Weiner, L.M. (1993). Dependence of nucleic acid degradation on in situ free-radical
production by adriamycin. Biochem., 32, 13156-13161.
Ferng, S.F., Castro, C.E., Afifi, A.A., Bermudez, E., and Mustafa, M.G. (1997). Ozone-induced DNA strand breaks in
guinea pig tracheobronchial epithelial cells. J. Toxicol. Environ. Health., 51, 353-367.
Ferrante, R.J.,Shinobu, L.A., Schulz, J.B., Matthews, R.T., Thomas, C.E., Kowall, N.W., Gurney, M.E., and Beal, M.F.
(1997). Increased 3-nitrotyrosine and oxidative damage in mice with human copper/zinc superoxide dismutase
mutation. Ann. Neurol., 42, 326-334.
Fest, A., and Gan, J.C. (1981). Selective loss of elastase inhibitory activity of alpha 1-proteinase inhibitor upon chemical
modification of tyrosyl residues. J. Biol. Chem., 256, 6374-6380.
Fiala, E.S., Conaway, C.C., and Mathis, J.E. (1989). Oxidative DNA and RNA damage in the livers of Sprague-Dawley
rats treated with the hepatocarcinogen 2-nitropropane. Cancer Res., 49, 5518-5522.
Fischer-Nielson, A., Jeding, I.B., and Loft, S. (1994). Irradiation-induced formation of 8-hydroxy-2’-deoxyguanosine and its
prevention by scavengers. Carcinogen., 15, 1609-1612.
Fischer-Nielsen, A., Poulsen, H.E., and Loft, S. (1992). 8-Hydroxydeoxyguanosine in vitro: Effects of glutathione,
ascorbate, and 5-aminosalicylic acid. Free Radic. Biol. Med., 13, 121-126.
Finch, J.T., and Klug, A. (1976). Solenoidal model for superstructure in chromatin. 901 Proc. Natl. Acad. Sci. USA, 73,
1897-1901.
Flohe, L. (1982). Glutathione peroxidases brought into focus. In: Free Radicals in Biology. Vol. V. Pryor, W.A. (Ed.).
Academic Press, New York. Pp. 223-253.
Floyd, R.A., West, M.S., Eneff, K.L., Schneider, J.E., Wong, P.K., Tingey, D.T., and Hogsett, W.E. (1990). Conditions
influencing yield and analysis of 8-hydroxy-2'-deoxyguanosine in oxidatively damaged DNA. Anal. Biochem., 188, 155158.
Floyd, R.A., Watson, J.J., Wong, P.K., Altmiller, D.H., and Rickard, R.C. (1986). Hydroxyl free radical adduct of
deoxyguanosine: Sensitive detection and mechanisms of formation. Free Radic. Res. Comm., 1, 163-172.
Foksinski, M., Bialkowski, K., Skiba, M., Ponikowska, I., Szmurlo, W., and Olinski, R. (1999). Evaluation of 8oxodeoxyguanosine, typical oxidative DNA damage, in lymphocytes of ozone-treated arteriosclerosis patients. Mutat.
Res., 438, 23-27.
Foksinski, M., Kotzbach, R., Szymanski, W., and Olinski, R. (2000). The level of typical biomarkers in oxidative stress 8hydroxy-2’-deoxyguanosine is higher in uterine myomas than in control tissues and correlated with the size of the
tumor. Free Radic. Biol. Med., 29, 597-601.
Forman, L.J., Liu, P., Nagele, R.G., Yin, K., and Wong, P.Y. (1998). Augmentation of nitric oxide, superoxide, and
peroxynitrite production during cerebral ischemia and reperfusion in the rat. Neurochem. Res., 23, 141-148.
Fowler, R.G., White, S.J., Moore, S.C., Dunn, R.L., and Schaaper, R.M. (2003). Interations among E. Coli mutt, mutM,
and mutY damage prevention pathways. DNA Repair, 2, 159-173.
Frankel, E.N., and Neff, W.F. (1983). Formation of malonaldehyde from lipid oxidation products. Biochim. Biophys. Acta,
754, 264-270.
Fraga, C.G., Motchnik, P.A., Shigenaga, M.K., Helbok, H.J., Jacob, R.A., and Ames, B.N. (1991). Ascorbic acid protects
against endogenous oxidative DNA damage in human sperm. Proc. Natl. Acad. Sci. USA, 88, 11003-11006.
WWW.ESAINC.COM
286
Fraga, C.G., Shigenaga, M.K., Park, J.-W., Degan, P., and Ames, B.N. (1990). Oxidative damage to DNA during aging: 8Hydroxy-2’deoxyguanosine in rat organ DNA and urine. Proc. Natl. Acad. Sci. USA, 87, 4533-4537.
Francesco, V., Cesare, A., Luigi, I., Stefano, F., Andrea, G., and Francesco, B. (1985). Malondialdehyde-like material and
beta-thromboglobulin plasma levels in patients suffering from transient ischemic attacks. Stroke, 16, 14-16.
Francescutti, D., Baldwin, J., Lee, L., and Mutus, N. (1996). Peroxynitrite modification of glutathione reductase: Modeling
studies and kinetic evidence suggest the modification of tyrosine at the glutathione disulfide binding site. Protein Eng.,
9, 189-194.
Freedman, J.E. (2002). Nitrated lipids: Defining their bioactivity. Circ. Res., 91, 371-372.
Freeman, B.A., White, C.R., Gutierrez, H., Paler-Martinez, A., Tarpey, M.M., and Rubbo, H. (1995). Oxygen radical-nitric
oxide reactions in vascular disease. Adv. Pharmacol., 34, 45-69.
Freeman, M.L., Borrelli, M.J., Meredith, M.J., and Lepock, J.R. (1999). On the path to the heat shock response:
Destabilization, and formation of partially folded intermediates, a consequence of protein thiol modification. Free Radic.
Biol. Med., 26, 737-745.
Frei, B., Yamamoto, Y., Niclas, D., and Ames, B.N. (1988). Evaluation of an isoluminol chemiluminescence assay for the
detection of hydroperoxides in human blood plasma. Anal. Biochem., 174, 120-130.
Frenkel, K., Zhong, Z., Wei, H., Karkoszka, J., Patel, U., Rashid, K., Georgescu, M., and Solomon, J.J. (1991).
Quantitative high-performance liquid chromatography analysis of DNA oxidized in vitro and in vivo. Anal. Biochem.,
196, 126-136.
Friedberg, E.C. (1996). Relationships between DNA repair and transcription. Ann. Rev. Biochem., 65, 15-42.
Friess, U., Waldner, M., Wahl, H-G., Rainer, L., Haring, H-U., Voelter, W., and Schleicher, E. (2003). Liquid
chromatography-based determination of urinary free and total N(epsilon)-(carboxymethyl)lysine excretion in normal and
diabetic subjects. J. Chromatogr. B, 794, 273-280.
Fu, M.X., Wells-Knecht, K.J., Blackedge, J.A., Lyons, T.J., Thorpe, S.R., and Baynes, J.W. (1994). Glycation,
glyoxidation, and cross-linking of collagen by glucose. Kinetics, mechanisms, and inhibition of late stages of the
Maillard reaction. Diabetes, 43, 676-683.
Fu, S., Gebicki, S., Jessup, W., Gebicki, J.M., and Dean, R.T. (1995a). Biological fate of amino acids, peptide and protein
hydroperoxides. Biochem. J., 311, 821-827.
Fu, S., Hick, L.A., Sheil, M.M., and Dean, R.T. (1995b). Structural identification of valine hydroperoxides and hydroxides
on radical-damaged amino acid, peptide, and protein molecules. Free Radic. Biol. Med., 19, 281-292.
Fuciarelli, A.F., Wegher, B.J., Blakely, W.F., and Dizdaroglu, M. (1990) Yields of radiation-induced base products in DNA:
Effects of DNA conformation and gassing conditions. Int. J. Radiat. Biol., 58, 397-415.
Fukuyama, N., Takebayashi, Y., Hida, M., Ishida, H., Ichimori, K., and Nakazawa, H. (1996). Clinical evidence of
peroxynitrite formation in chronic renal failure patients with septic shock. Free Radic. Biol. Med., 22, 771-774.
Fung, K., and Grosjean, D. (1981). Determination of nanogram amounts of carbonyls as 2,4-dintrophenylhydrazones by
high-performance liquid chromatography. Anal. Chem., 53, 168-171.
Furuichi, M., Yoshida, M.C., Oda, H., Tajiri, T., Nakabeppu, Y., Tsuzuki, T., and Sekiguchi, M. (1994). Genomic structure
and chromosome location of the human mutT homologue gene MTH1 encoding 8-oxo-dGTPase for prevention of A:T
to C:G transversion. Genomics, 24, 485-490.
Gajewski, E., Rao, G., Nackerdien, Z., and Dizdaroglu, M. (1990). Modification of DNA bases in mammalian chromatin by
radiation generated free radicals. Biochem., 29, 7876-7882.
Galli, F., Rovidati, S., Ghibelli, L., and Canestrari, F. (1998). S-nitrosylation of glyceraldehyde-3-phosphate
dehydrogenase decreases the enzyme affinity to the erythrocyte membrane. Nitric Oxide, 2, 17-27.
Gantchev, T.G., and van Lier, J. (1995). Catalase inactivation following photosensitization with tetrasulfonated
metallophthalocyanines. Photochem. Photobiol., 62, 123-134.
Garrison, W.M., (1987). Reaction mechanisms in the radiolysis of peptides, polypeptides, and proteins. Chem. Rev., 87,
381-398.
Gaut, J.P., Byun, J., Tran, H.D., and Heinecke, J.W. (2002). Artifact-free quantification of free 3-chlorotyrsoine, 3bromotyrosine, and 3-nitrotyrosine in human plasma by electron capture-negative chemical ionization gas
chromatography mass spectrometry and liquid chromatography-electrospray ionization tandem mass spectrometry.
Anal. Biochem., 300, 252-259.
Gebicki, S., and Gebicki, J.M. (1993). Formation of peroxides in amino acids and proteins exposed to oxygen free
radicals. Biochem J., 289, 743-749.
Gebicki, S., and Gebicki, J.M. (1999). Crosslinking of DNA and proteins induced by protein hydroperoxides. Biochem. J.,
338, 629-636.
Gedik, C.M., Wood, S.G., and Collins, A.R. (1998). Measuring oxidative damage to DNA: HPLC and Comet assay
compared. Free Radic. Res., 29, 609-615.
Geiger, P.G., Korytowski, W., Lin, F., and Girotti, A.W. (1997). Lipid peroxidation in photodynamically stressed
mammalian cells: Use of cholesterol hydroperoxides as mechanistic reporters. Free Radic. Biol. Med., 23, 57-68.
Germadnik, D., Pilger, A., and Rudiger, H.W. (1997). Assay for the determination of urinary 8-hydroxy-2’deoxyguanosine
by high-performance liquid chromatography with electrochemical detection. J. Chromatogr. B., 689, 399-403.
Gieseg, S.P., Simpson, J.A., Charlton, T.S., Duncan, M.W., and Dean, R.T. (1993). Protein-bound 3,4dihydroxyphenylalanine is a major reductant formed during hydroxyl radical damage to proteins. Biochem., 32, 47804786.
Girotti, A.W. (1998). Lipid hydroperoxide generation, turnover and affector action in biological systems. J. Lipid Res., 39,
1529-1542.
Giulivi, C., and Davies, K.J.A. (1993). Dityrosine and tyrosine oxidation products are endogenous markers for the
selective proteolysis of oxidatively modified red blood cell hemoglobin by (the 19 S) proteasome. J. Biol. Chem., 268,
8752-8759.
Giulivi, C., and Davies, K.J.A. (1994). Dityrosine: A marker of oxidatively modified proteins and selective proteolysis.
Meth. Enzymol., 233, 363-371.
WWW.ESAINC.COM
287
Glomb, M.A., and Monnier, V.M. (1995). Mechanism of protein modification by glyoxal and glycoaldehyde, reactive
intermediates of the Maillard reaction. J. Biol. Chem., 270, 10017-10026.
Goda, Y., and Marnett, L.J. (1991). High-performance liquid chromatography with electrochemical detection for
determination of the major malondialdehyde-guanine adduct. Chem. Res. Toxicol., 4, 520-524.
Goldring, C., Casini, A.F., Maellaro, E., Bello, B D., and Comporti, M. (1993). Determination of 4-hydroxynonenal by highperformance liquid chromatography with electrochemical detection. Lipids, 28, 141-145.
Good, P.F., Werner, P., Hsu, A., Olanow, C.W., and Perl, D.P. (1996). Evidence of neuronal oxidative damage in
Alzheimer’s disease. Am. J. Pathol., 149, 21-28.
Goodwin, D.C., Gunther, M.R., His, L.C., Crews, B.C., Eling, T.E., Mason, R.P., and Marnett, L.J. (1998). Nitric oxide
trapping of tyrosyl radicals generated during prostaglandin endoperoxide synthase turnover. Detection of the radical
derivative of tyrosine 385. J. Biol. Chem., 273, 8903-8909.
Gorsdorf, S., Appel, K.E., Engeholm, C., and Obe, G. (1990). Nitrogen dioxide induces singlestrand breaks in cultured Chinese hamster cells. Carcinogen., 11, 37-41.
Gow, A.J., Duran, D., Malcolm, S., and Ischiropoulos, H. (1996a). Effects of peroxynitrite-induced protein modifications on
tyrosine phosphorylation and degradation. FEBS Lett., 385, 63-66.
Gow, A., Duran, D., Thom, S.R., and Ischiropoulos, H. (1996b). Carbon dioxide enhancement of peroxynitrite-mediated
protein tyrosine nitration. Arch. Biochem. Biophys., 333, 42-48.
Gowan, L.C., Avrutskaya, A.V., Latour, A.M., Koller, B.H., Leadon, S.A. (1998). BRCA1 required for the transcriptioncoupled repair of oxidative DNA damage. Science, 281, 10091012.
Greenblatt, M.S., Bennett, W.P., Hollstein, M., and Harris, C.C. (1994). Mutations in the p53 tumor suppressor gene:
Clues to cancer etiology and molecular pathogenesis. Cancer Res., 54, 4855-4878.
Greis, K.D., Zhu, S., and Matalon, S. (1996). Identification of nitration sites on surfactant protein A by tandem electrospray
mass spectrometry. Arch. Biochem. Biophys., 335, 396-402.
Grollman, A.P., and Moriya, M. (1993). Mutagenesis by 8-oxoguanine: an enemy within. Trends Genet., 9, 246-249.
Gross, A.J., and Sizer, I.W. (1959). The oxidation of tyramine, tyrosine, and related compounds by peroxidase. J. Biol.
Chem., 234, 1611-1614.
Grossman, A., and Wendel, A. (1983). Non-reactivity of selenoenzyme glutathione peroxidase with enzymatically
hydroperoxidized phospholipids. Eur. J. Biochem., 135, 549-552.
Grune, T., and Davies, K.J. (1997). Breakdown of oxidized proteins as a part of secondary antioxidant defenses in
mammalian cells. Biofactors. 6, 165-172.
Grune, T., Blasig, I.E., Sitte, N., Roloff, B., Haseloff, R., and Davies, K.J. (1998). Peroxynitrite increases the degradation
of aconitase and other cellular proteins by proteosome. J. Biol. Chem., 273, 10857-10862.
Grune, T., Reinheckel, T., and Davies, K.J. (1997). Degradation of oxidized proteins in mammalian cells. FASEB J., 11,
526-534.
Guichardant, M., Taibi-Tronche, P., Fay, L.B., and Lagarde, M. (1998). Covalent modifications of aminophospholipids by
4-hydroxynonenal. Free Radic. Biol. Med., 25, 1049-1056.
Guptasarma, P., Balasubramamian, D., Matsugo, S., and Saito, I. (1992). Hydroxyl radical mediated damage to proteins,
with special reference to the crystallins. Biochem., 31, 4296-4303.
Haas, A.L., and Siepmann, T.J. (1997). Pathways of ubiquitin conjugation. FASEB J., 11, 1257-1268.
Haddad, I.Y., Pataki, G., Hu, P., Galliani, C., Beckman, J.S., and Matalon, S. (1994). Quantitation of nitrotyrosine levels in
lung sections of patients and animals with acute lung injury. J. Clin. Invest., 94, 2407-2413.
Hajjar, D.P., and Haberland, M.E. (1997). Lipoprotein trafficking in vascular cells. Molecular Trojan horses and cellular
saboteurs. J. Biol. Chem., 272, 22975-22978.
Halliwell, B. (1995). Oxygen radicals, nitric oxide and human inflammatory joint disease. Ann. Rheum. Dis., 54, 505-510.
Halliwell, B. (1998). What nitrates tyrosine? Is nitrotyrosine specific as a biomarker of peroxynitrite in vivo? FEBS Lett.,
411, 157-160.
Halliwell, B., and Aruoma, O.I. (1991). DNA damage by oxygen-derived species. Its mechanism and measurement in
mammalian systems. FEBS Lett., 281, 9-19.
Halliwell, B., and Chirico, S. (1993). Lipid peroxidation: Its mechanisms, measurement and significance. Am. J. Clin. Nutr.,
57, 715S-725S.
Halliwell, B., and Gutteridge, J.M.C. (1988). Iron as a biological pro-oxidant. ISI Atlas Sci. Biochem., 1, 48-52.
Halliwell, B., and Gutteridge, J.M.C. (Eds.). (1999). Free Radicals in Biology and Medicine. Clarendon Press, Oxford.
Han, X., and Liehr, J.G. (1994). 8-Hydroxylation of guanine in kidney and liver DNA of hamsters treated with estradiol:
Role of free radicals in estrogen-induced carcinogenesis. Cancer Res., 54, 5515-5517.
Hanaoka, T., Tsugane, S., Yamano, Y., Takahashi, T., Kasai, H., Natori, Y., and Watanabe, S. (1993). Quantitative
analysis of 8-hydroxyguanine in peripheral blood cells: An application for asbestos patients. Int. Arch. Occup. Environ.
Health, 65, S215-S217.
Hantraye, P., Brouillet, E., Ferrante, R., Palfi, S., Dolan, R., Matthews, R.T., and Beal, M.F. (1996). Inhibition of neuronal
nitric oxide synthase prevents MPTP-induced Parkinsonism in baboons. Nat. Med., 2, 1017-1021.
Harman, D. (1989). Lipofuscin and ceroid formation: The cellular recycling system. Adv. Exp. Med. Biol., 266, 3-15.
Harris, C.C. (1995). Tumor suppressor genes: At the crossroads of molecular carcinogenesis and molecular
epidemiology. In: Molecular Aspects of Carcinogenesis International Meeting. Univ. York, York. U.K. Pp. 22.
Harris, C.C., and Holstein, M. (1993). Clinical implications of the p53-tumor suppressor gene. N. Engl. J. Med., 329, 13181327.
Harris, M.E., Carney, J.M., and Leedle, R.A. (1994). Detection of oxidation products in individual neurons by fluorescence
microscopy. Exper. Neurol., 129, 95-102.
Harris, M.E., Hensley, K., Butterfield, D.A., Leedle, R.A., and Carney, J.M. (1995). Direct evidence of oxidative injury
produced by the Alzheimer's beta- amyloid peptide (1-40) in cultured hippocampal neurons. Exper. Neurol., 131, 193202.
WWW.ESAINC.COM
288
Hattori-Nakakuki, Y., Nishigori, C., Okamoto, K., Imamura, S., Hiai, H., and Toyokuni, S. (1994). Formation of 8-hydroxy2’deoxyguanosine in epidermis of hairless mice exposed to near-UV. Biochem. Biophys. Res. Commun., 201, 11321139.
Hayakawa, H., Taketomi, A., Sakumi, K., Kuwano, M., and Sekiguchi, M. (1995). Generation and elimination of 8-oxo-7,8dihydro-2’deoxyguanosine 5’triphosphate, a mutagenic substrate for DNA synthesis in human cells. Biochem., 34, 8995.
Hazell, L.J., Arnold, L., Flowers, D., Waeng, G., Malle, E., and Stocker, R. (1996). Presence of hypochlorite-modified
proteins in human atherosclerotic lesions. J. Clin. Invest., 97, 1535-1544.
Hazen, S.L. (1998). Personal communications.
Hazen, S.L., and Heinecke, J.W. (1997). 3-Chorotyrosine, a specific marker of myeloperoxidase-catalyzed oxidation, is
markedly elevated in low-density lipoprotein isolated from human atherosclerotic intima. J. Clin. Invest., 99, 2075-2081.
Hazen, S.L., d’Avignon, A., Anderson, M.M., Hsu, F.F., and Heinecke, J.W. (1998a). Human neutrophils employ
myeloperoxidase-hydrogen peroxide-chloride system to oxidize alpha-amino acids to a family of reactive aldehydes.
Mechanistic studies identifying labile intermediates along the reaction pathway. J. Biol. Chem., 273, 4997-5005.
Hazen, S.L., Crowley, J.R., Mueller, D.M., and Heinecke, J.W. (1997). Mass spectrometric quantification of 3chlorotyrosine in human tissues with attomole sensitivity: A sensitive and specific marker for myeloperoxidasecatalyzed chlorination at sites of inflammation. Free Radic. Biol. Med., 23, 909-916.
Hazen, S.L., Gaut, J.P., Hsu, F.F., Crowley, J.R., d’Avignon, A., and Heinecke, J.W. (1997). PHydroxyphenylacetaldehyde, the major product of L-tyrosine oxidation by the myeloperoxidase-H2O2-chloride system of
phagocytes, covalently modifies ε-amino groups of protein lysine residues. J. Biol. Chem., 272, 16990-16996.
Hazen, S.L., Hsu, F.F., d’Avignon, A., and Heinecke, J.W. (1998b). Human neutrophils employ myeloperoxidase to
convert alpha-amino acids to a battery of reactive aldehydes: A pathway for aldehyde generation at sites of
inflammation. Biochem., 37, 6864-6873.
Hazen, S.L., Hsu, F.F., Duffin, K., and Heinecke, J.W. (1996). Molecular-chlorine generated by the myeloperoxidasehydrogen peroxide-chloride system of phagocytes converts low-density lipoprotein cholesterol into a family of
chlorinated sterols. J. Biol. Chem., 271, 23080-23088.
Hazen, S.L., Hsu, F.F., and Heinecke, J.W. (1996). P-hydroxyphenylacetaldehyde is the major product of L-tyrosine
oxidation by activated human phagocytes. J. Biol. Chem., 271, 1861-1867.
Hazen, S.L., Hsu, F.F., Mueller, D.M., Crowley, J.R., and Heinecke, J.W. (1996). Human neutrophils employ chlorine gas
as an oxidant during phagocytosis. J. Clin. Invest., 98, 1283-1289.
Hazra, T.K., Hill, J.W., Izumi, T., and Mitra, S. (2001). Multiple DNA glycosylases for repair of 8-oxoguanine and their
potential in vivo functions. Prog. Nucleic Acid Res. Mol. Biol., 68, 193-205.
He, Q., Srivastava, S., van Kuijk, F.J., and Ansari, N.H. (1998). Reduction of 4-hydroxynonenal and 4-hydroxyhexenal by
retinal aldose reductase. Biochem. Biophys. Res. Commun., 247, 719-722.
Heinecke, J.W., Hsu, F.F., Crowley, J.R., Hazen, S.L., Leeuwenburgh, C., Mueller, D.M., Rasmussen, J.E., and Turk, J.
(1998). Detecting oxidative modification of biomolecules with isotope dilution mass spectrometry: Sensitive and
quantitative assays for oxidized amino acids in proteins and tissues. Meth. Enzymol., 300, 124-144.
Heinecke, J.W., Li, W., Daehnke, H.L., and Goldstein, J.A. (1993). Dityrosine, a specific marker of oxidation, is
synthesized by the myeloperoxidase-hydrogen peroxide system of human neutrophils and macrophages. J. Biol.
Chem., 268, 4069-4077.
Helbock, H.J., Beckman, K.B., Shigenaga, M.K., Walter, P.B., Woodall, A.A., Yeo, H.C., and Ames, B.N. (1998). DNA
oxidation matters: The HPLC-electrochemical detection assay of 8-oxo-deoxyguanosine and 8-oxo-guanine. Proc. Natl.
Acad. Sci. USA, 95, 288-293.
Hellberg, C.B., Boggs, S.E., and Lapetina, E.G. (1998). Phosphatidylinositol 3-kinase is a target for protein tyrosine
nitration. Biochem. Biophys. Res. Commun., 252, 313-317.
Henle, E.S., Luo, Y., Gassmann, W., and Linn, S. (1996). Oxidative damage to DNA constituents by iron-mediated Fenton
reactions. The deoxyguanosine family. J. Biol. Chem., 271, 21177-21186.
Hensley, K., Hall, N., Subramaniam, R., Cole, P., Harris, M., Askenov, M., Askenova, M., Gabbita, S.P., Wu, J.F., Carney,
J.M., Lovell, M., Markesbery, W.R., and Butterfield, D.A. (1995). Brain regional correspondence between Alzheimer's
disease histopathology and biomarkers of protein oxidation. J. Neurochem., 65, 2146-2156.
Hensley, K., Maidt, M.L., Pye, Q.N., Stewart, C.A., Wack, M., Tabatabaie, T., and Floyd, R.A. (1997). Quantitation of
protein bound 3-nitrotyrosine and 3,4-dihydroxyphenylalanine by high-performance liquid chromatography with
electrochemical array detection. Anal. Biochem., 251, 187-195.
Hensley, K., Maidt, M.L., Yu, Z.Q., Sang, H., Markesbery, W.R., and Floyd, R.A. (1998). Electrochemical analysis of
protein nitrotyrosine and dityrosine in the Alzheimer brain indicates region-specific accumulation. J. Neurosci., 18,
8126-8132.
Henry, Y.A., and Singel, D. (1996). Metal-nitrosyl interactions in nitric oxide biology probed by electron paramagnetic
resonance spectroscopy. In: Methods in Nitric Oxide Research. Feelisch, M., and Stamler, J.S. (Eds.). John Wiley and
Sons Ltd. London. Pp. 357-372.
Herbert, K.E., and Lunec, J. (1998). Progress in the immunodetection of products of oxidative damage to DNA. In: DNA
and Free Radicals: Techniques, Mechanisms and Applications. Aruoma, O.I., and Halliwell, B. (Eds.). OICA, London.
Pp. 271-284.
Herbert, K.E., Evans, M.D., Finnegan, M.T.V., Farooq, S., Mistry, N., Podmore, I.D., Farmer, P., and Lunec, J. (1996). A
novel HPLC procedure for the analysis of 8-oxoguanine in DNA. Free Radic. Biol. Med., 20, 467-473.
Herce-Pagliai, C., Kotecha, S., and Shuker, D.E. (1998). Analytical methods for 3-nitrotyrosine as a marker of exposure to
reactive nitrogen species: A review. Nitric Oxide, 2, 324-336.
Hilt, W., and Wolf, D.H. (1996). Proteasomes: Destruction as a program. Trends Biochem. Sci., 21, 96-102.
Hochstrasser, M. (1996). Ubiquitin-dependent protein degradation. Ann Rev. Genet., 30, 405-439.
Hofer, T., and Moller, L. (1998). Reduction of oxidation during the preparation of DNA and analysis of 8-hydroxy-2’deoxyguanoisne. Chem. Res. Toxicol. In press.
WWW.ESAINC.COM
289
Holmberg, I., Stal, P., and Hamberg, M. (1999). Quantitative determination of 8-hydroxy-2’deoxyguanosine in human urine
by isotope dilution mass spectrometry: Normal levels in hemochromatosis. Free Radic. Biol. Med., 26, 129-135.
Holmes, G.E., Bernstein, C., and Bernstein, H. (1992). Oxidative and other DNA damages as the basis of aging: A review.
Mutat. Res., 275, 305-315.
Holstein, M., Didransky, D., Vogelstein, B., and Harris, C.C. (1991). p53 mutations in human cancer. Science, 253, 49-52.
Holley, A.E., and Slater, T.F. (1991). Measurement of lipid hydroperoxides in normal human blood plasma using HPLCchemiluminescence linked to a diode array detector for measuring conjugated dienes. Free Radic. Res. Comm., 15, 5163.
Holley, A.E., Walker, M.K., Cheeseman, K.H., and Slater, T.F. (1993). Measurement of n-alkanals and hydroxyalkanals in
biological samples. Free Radic. Biol. Med., 15, 281-289.
Hsei, A.W., Recio, L., Karz, K.S., Lee, C.Q., Wagner, M., and Schenley, R.L. (1986). Evidence for reactive oxygen
species inducing mutations in mammalian cells. Proc. Natl. Acad. Sci. USA, 83, 9616-9620.
Huggins, T.G., Wells-Knecht, M.C., Detoria, N.A., Baynes, J.W., and Thorpe, S.R. (1993). Formation of o-tyrosine and
dityrosine in proteins during radiolytic and metal-catalyzed oxidation. J. Biol. Chem., 268, 12341-12347.
Huie, R.E., and Neta, P. (1999). Chemistry of reactive oxygen species. In: Reactive Oxygen Species in Biological
Systems. Gilbert, D.L., and Colton, C.A. (Eds.). Kluwer Academic/Plenum Publishers, New York. Pp. 33-73.
Hunt, J.V., Dean, R.T., and Wolff, S.P. (1988). Hydroxyl radical production and autoxidative glycosylation. Glucose
autoxidation as the cause of protein damage in the experimental glycation model of diabetes mellitus and aging.
Biochem. J., 256, 205-212.
Hunter, M.I.S., and Mohamed, J.B. (1986). Plasma antioxidants and lipid peroxidation products in duchene muscular
dystrophy. Clin. Chim. Acta, 155, 122-132.
Husgafvel-Pursiainen, K., Ridanpaa, M., Anttila, S., and Vainio, H. (1995). p53 and ras gene mutations in lung cancer:
Implications for smoking and occupational exposures. J. Occup. Environ. Med., 37, 69-76.
Ide, H. (2001). DNA substrates containing defined oxidative base lesions and their application to study substrate
specificities of base excision repair enzymes. Prog. Nucleic Acid Res. Mol. Biol. 68, 207-221.
Ide, H., Tedzuka, K., Shimzu, Kimura, Y., Purmal, A.A., Wallace, S.S., and Kow, Y.W. (1994). Alpha-deoxyadenosine, a
major anoxic radiolysis product of adenine in DNA, is a substrate for E. coli endonuclease IV. Biochem., 33, 7842-7847.
Imbusch, R., and Mueller, M.J. (2000). Formation of isoprostanes F(2)-like compounds (phytoprostanes (1)) from alphalinolenic acid in plants. Free Radic. Biol. Med., 28, 720-726.
Inoue, T., Mu, Z., Sumikawa, K., Adachi, K., and Okochi, T. (1993). Effect of physical exercise on the content of 8hydroxydeoxyguanosine in nuclear DNA prepared from human lymphocytes. Jpn. J. Cancer Res., 84, 720-725.
Ischiropoulos, H., and Al-Mehdi, A. B. (1995). Peroxynitrite-mediated oxidative protein modifications. FEBS Letts., 364,
279-282.
Ischiropoulos, H. (2003). Oxidative modifications of alpha-synuclein. Ann. N. Y. Acad. Sci., 991, 93-100.
Ischiropoulos, H., al-Mehdi, A.B., and Fisher, A.B. (1995). Reactive species in ischemic rat lung injury: Contribution of
peroxynitrite. Am. J. Physiol., 269, L158-164.
Ischiropoulos, H., Beers, M.F., Ohnishi, S.T., Fisher, D., Garner, S.E., and Thom, S.R. (1996). Nitric oxide production and
perivascular nitration in brain after carbon monoxide poisoning in rats. J. Clin. Invest., 97, 2260-2267.
Ischiropoulos, H., Duran, D., and Horwitz, J. (1995). Peroxynitrite-mediated inhibition of L-DOPA synthesis in PC12 cells.
J. Neurochem., 65, 2366-2372.
Ishida, H., Makino, A., and Mae, T. (1999). Fragmentation of the large subunit of ribulose-1,5-bisphosphate carboxylase
by reactive oxygen species occurs near Gly-329. J. Biol. Chem., 274, 5222-5226.
Ishimitsu, S., Fujimoto, S., and Ohara, A. (1986). Determination of m-tyrosine and o-tyrosine in human serum by highperformance liquid chromatography with fluorometric detection. J. Chromatogr., 378, 222-225.
Ishiyama, S., Hiroe, M., Nishikawa, T., Abe, S., Shimojo, T., Ito, H., Ozasa, S., Yamakawa, K., Matsuzaki, M.,
Mohammed, M.U., Nakazawa, H., Kasajima, T., and Marumo, F. (1997). Nitric oxide contributes to the progression of
myocardial damage in experimental autoimmune myocarditis in rats. Circulat., 95, 489-496.
Iwaoka, T., Tabata, F., and Takahashi, T. (1987). Lipid peroxide detected by chemiluminescence. Free Radic. Biol. Med.,
3, 329-333.
Jacob, J.S., Cistola, D.P., Hsu, F.F., Muzaffar, S., Mueller, D.M., Hazen, S.L., and Heinecke, J.W. (1996). Human
phagocytes employ the myeloperoxidase-hydrogen peroxide system to synthesize dityrosine, trityrosine, pulcherosine,
and isodityrosine by a tyrosyl radical-dependent pathway. J. Biol. Chem., 271, 19950-19956.
Jajoo, H.K., Burcham, P.C., Goda, Y., Blair, I.A., and Marnett, L. A thermospray liquid chromatography/mass spectrometry
method for analysis of human urine for the major malondialdehyde-guanine adduct. J. Chem. Res. Toxicol., 5, 870-875.
Jaruga, P., Zastawny, T.H., Skokoswki, J., Dizdaroglu, M., and Olinski, R. (1994). Oxidative DNA base damage and
antioxidant enzyme activities in human lung cancer. FEBS Lett., 341, 59-64.
Jones, A.F., and Lunec, J. (1987). Protein fluorescence and its relationship to free radical activity. Br. J. Cancer, 8, 60-65.
Joseph, J.A., Villalobos-Molinas, R., Denisova, N.A., Erat, S., and Strain, J. (1996). Cholesterol: A two-edged sword in
brain aging. Free Radic. Biol. Med., 22, 455-462.
Juedes, M.J., and Wogan, G.N. (1996). Peroxynitrite-induced mutation spectra of pSP189 following replication of bacteria
and in human cells. Mutat. Res., 349, 51-61.
Kamisaki, Y., Wada, K., Bian, K., Balabanli, B., Davis, K., Martin, E., Behbod, F., Lee, Y.C., and Murad, F. (1998). An
activity in rat tissues that modifies nitrotyrosine-containing proteins. Proc. Natl. Acad. Sci. USA, 95, 11584-11589.
Kamisaki, Y., Wada, K., Nakamoto, K., Kishimoto, Y., Kitano, M., and Itoh, T. (1996). Sensitive determination of
nitrotyrosine in human plasma by isocratic high-performance liquid chromatography. J. Chromatogr. B., 685, 343-347.
Kamiya, H., Miura, K., Ishikawa, H., Nishimura, S., and Ohtsuka, E. (1992). c-Ha-ras containing 8-hydroxyguanine at
codon 12 induces point mutations at the modified and adjacent positions. Cancer Res., 52, 3483-3485.
Kang, H., Konishi, C., Eberle, G., Rajewsky, M.E., Kuroki, T., and Huh, N-H. (1992). Highly sensitive, specific detection of
6
4
4
O -methylguanine, O -methylthymine, and O -ethylthymine by the combination of high-performance liquid
32
chromatography prefractionation, P postlabeling, and immunoprecipitation. Cancer Res., 52, 5307-5312.
WWW.ESAINC.COM
290
Karakaya, A., Jaruga, P., Bohr, V., Grollman, A.P., and Dizdaroglu, M. (1997). Kinetics of excision of purine lesions from
DNA by Escherichia coli Fpg protein. Nucl. Acid Res., 25, 474-479.
Kasai, H. (1997). Analysis of a form of oxidative DNA damage, 8-hydroxy-2’-deoxyguanosine, as a marker of cellular
oxidative stress during carcinogenesis. Mutat. Res., 387, 147-163.
Kato, Y., Maruyama, W., Naoi, M., Hashizume, Y., and Osawa, T. (1998). Immunohistochemical detection of dityrosine in
lipofuscin pigments in the aged human brain. FEBS Lett., 439, 231-234.
Kaur, H., and Halliwell, B. (1994). Evidence for nitric oxide-mediated oxidative damage in chronic inflammation.
Nitrotyrosine in serum and synovial fluid from rheumatoid patients. FEBS Lett., 350, 9-12.
Kaur, H., and Halliwell, B. (1996). Measurement of oxidized and methylated DNA bases by HPLC with electrochemical
detection. Biochem. J., 318, 21-23.
Kaur, H., Lyras, L., Jenner, P., and Halliwell, B. (1998). Artifacts in HPLC detection of 3-nitrotyrosine in brain tissue. J.
Neurochem., 70, 2220-2223.
Keith, G., and Dirheimer, G. (1995). Postlabeling: A sensitive method for studying DNA adducts and their role in
carcinogenesis. Curr. Opin. Biotechnol., 6, 3-11.
Kettle, A.J. (1996). Neutrophils convert tyrosyl residues in albumin to chlorotyrosine. FEBS Lett., 379, 103-106.
Kettle, A.J. (1998). Detection of 3-chlorotyrosine in proteins exposed to neutrophil oxidants. Meth. Enzymol., 300, 111120.
Kikuchi, A., Takeda, A., Onodera, H., Kimpara, T., Hisanaga, K., Nobuyuki, S., Nunomura, A., Castellani, R.J., Perry, G.,
Smith, M.A., and Itoyama, Y. (2002). Systemic increase of oxidative nucleic acid damage in Parkinson’s disease and
multiple system atrophy. Neurobiol. Dis., 9, 244-248.
Kikuchi, S., Shinpo, K, Takeuchi, M., Yamagishi, S., Makita, Z., Sasaki, N., and Tashiro, K. (2003). Glycation – a sweet
tempter for neuronal death. Brain Res. Brain Res. Rev., 41, 306-323.
Kikugawa, K., and Beppu, M. (1987), Involvement of lipid oxidation products in the formation of fluorescent and crosslinked proteins. Chem. Phys. Lip., 44, 277-296.
Kimura, H., Hokari, R., Miura, S., Shigematsu, T., Hirokawa, M., Akiba, Y., Kurose, I., Higuchi, H., Fujimori, H., Tsuzuki,
Y., Serizawa, H., and Ishii, H. (1998). Increased expression of an inducible isoform of nitric oxide synthase and the
formation of peroxynitrite in colonic mucosa of patients with ulcerative colitis. Gut, 42, 180-187.
Kiyosawa, H., Suko, M., Okudaira, H., Murata, K., Miyamoto, T., Chung, M.H., Kasai, H., and Nishimura, S. (1990).
Cigarette smoking induces formation of 8-hydroxydeoxyguanosine, one of the oxidative DNA damages in human
peripheral leukocytes. Free Radic. Res. Commun., 11, 23-27.
Klebl, B.M., Ayoub, A.T., and Pette, D. (1998). Protein oxidation, tyrosine nitration and inactivation of sarcoplasmic
2+
reticulum Ca -ATPase in low-frequency stimulated rabbit muscle. FEBS Lett., 422, 381-384.
Kooy, N.W., Lewis, S.J., Royall, J.A., Ye, Y.Z., Kelly, D.R., and Beckman, J.S. (1997). Extensive tyrosine nitration in
human myocardial inflammation: Evidence for the presence of peroxynitrite. Crit. Care Med., 25, 812-819.
Kooy, N.W., Royall, J.A., Ye, Y.Z., Kelly, D.R., and Beckman, J.S. (1995). Evidence for in vivo peroxynitrite production in
human acute lung injury. Am. J. Respir. Crit. Care Med., 151, 1250-1254.
Korchazhkina, O., Exley, C., and Spncer, S.A. (2003). Measurement of reversed-phase high-performance liquid
chromatography of malondialdehyde in normal human urine following derivatization with 2,4-dinitrophenylhydrazine. J.
Chromatogr. B, 794, 353-362.
Korytowski, W., Bachowski, G., and Girotti. A.W. (1991). Chromatographic separation and electrochemical determination
of cholesterol hydroperoxides generated by photodynamic action. Anal. Biochem., 197, 149-156.
Korytowski, W., Geiger, P.G., and Girotti, A.W. (1995). High-performance liquid chromatography with mercury cathode
electrochemical detection: Application to lipid hydroperoxide analysis. J. Chromatogr. B., 670, 189-197.
Korytowski, W., Geiger, P.G., and Girotti, A.W. (1999). Lipid hydroperoxide analysis by high-performance liquid
chromatography with mercury cathode electrochemical detection. Meths. Enzymol., 300, 23-33.
Kozumbo, W.J., Agarwal, S., and Koren, H.S. (1992). Breakage and binding of DNA by reaction products of hypochlorous
acid with aniline, 1-naphthylamine, or 1-naphthol. Toxicol. Appl. Pharmacol., 115, 107-115.
Kreuzer, T., Grube, R., Wutte, A., Zarkovic, N., and Schaur, R.J. (1998). 4-Hydroxynonenal modifies the effects of serum
growth factors on the expression of the c-fos proto-oncogene and the proliferation of HELA carcinoma cells. Free
Radic. Biol. Med., 25, 42-49.
Kristal, B.S., Park, B.K., and Yu, B.P. (1996). 4-Hydroxynonenal is a potent inducer of the mitochondrial permeability
transition. 271, 6033-6038.
Kristof, A.S., Goldberg, P., Laubach, V., and Hussain, S.N.A. (1998). Role of inducible nitric oxide synthase in endotoxininduced acute lung injury. Am. J. Respir. Crit. Care Med., 158, 1883-1889.
Kuhn, D.M., Aretha, C.W., and Geddes, T.J. (1999). Peroxynitrite inactivation of tyrosine hydroxylase: Mediation by
sulfhydryl oxidation, not tyrosine nitration. J. Neurosci., 19, 10289-10294.
Kunkel, T.A. (1993). Nucleotide repeats. Slippery DNA and diseases. Nature, 365, 207-208.
Kuo, W.N., Kanadia, R.N., and Shanbhag, V.P. (1999). Denitration of peroxynitrite-treated proteins by “protein nitrases”
from dog prostate. Biochem. Mol. Biol. Int., 47, 1061-1067.
Kvasnicova, V., Samcova, E., Jursova, A., and Jelinek, I. (2003). Determination of 8-hydroxy-2’-deoxyguanosine in
untreated urine by capillary electrophoresis with UV detection. J. Chromatogr. A, 985, 513-517.
Kwon, N.S., Chan, P.C., and Kesner, L. (1990). Inactivation of alpha 1-proteinase inhibitor by Cu(II) and hydrogen
peroxide. Agents Actions, 29, 388-393.
Lands, W.E.M., Kulmacz, R.J., and Marshall, P.J. (1994). Lipid peroxide actions in the regulation of prostaglandin
synthesis. In: Free Radicals in Biology. Volume 6. Pryor, W.A. (Ed.). Academic Press, New York. Pp. 31-61.
Lagorio, S., Tagesson, C., Forastiere F., Axelson, O., and Carere, A. (1994). Exposure to benzene and urinary
concentrations of 8-hydroxydeoxyguanosine, a biological marker of oxidative damage to DNA. Occup. Environ. Med.,
51, 739-743.
Lang, J., Celetto, C., and Esterbauer, H. (1985). Quantitative determination of the lipid peroxidation product 4hydroxynonenal by high-performance liquid chromatography. Anal. Biochem., 150, 369-378.
WWW.ESAINC.COM
291
Ledwozyw, A., Michalak, J., Stephen, A., and Kadziolka, A. (1986). The relationship between plasma triglycerides,
cholesterol, total lipid and lipid peroxidation products during human atherosclerosis. Clin. Chim. Acta, 155, 275-284.
Lee, D.M. (1980). Malondialdehyde formation in stored plasma. Biochem. Biophys. Res. Commun., 95, 1633-1672.
Lee, S.W., and Wei, J.Y. (1997). Molecular interactions of aging and cancer. Clin. Geriatr. Med., 13, 69-77.
Leeuwenburgh, C., Hansen, P., Holloszy, J.O., and Heinecke, J.W. (1999). Hydroxyl radical generation during exercise
increases mitochondrial protein oxidation and levels of urinary dityrosine. Free Radic. Biol. Med., 27, 186-192.
Leeuwenburgh, C., Hansen, P., Shaish, A., Holloszy, J.O., and Heinecke, J.W. (1998). Markers of protein oxidation by
hydroxyl radical and reactive nitrogen species in tissues of aging rats. Am. J. Physiol., 274, R453-461.
Leeuwenburgh, C., Hardy, M.M., Hazen, S.L., Wagner, P., Oh-ishi, S., Steinbrecher, U.P., and Heinicke, J.W. (1997).
Reactive nitrogen intermediates promote low-density lipoprotein oxidation in human atherosclerotic intima. J. Biol.
Chem., 272, 1433-1436.
Leeuwenburgh, C., Hansen, P.A., Holloszy, J.O., and Heinecke, J.W. (1999). Oxidized amino acids in the urine of aging
rats: Potential markers for assessing oxidative stress in vivo. Am. J. Physiol., 276, R128-135.
Leeuwenburgh, C., Rasmussen, J.E., Hsu, F.F., Mueller, D.M., Pennathur, S., and Heinecke, J.W. (1997a). Mass
spectrometric quantification of markers for protein oxidation by tyrosyl radical, copper, and hydroxyl radical in low
density lipoprotein isolated from human atherosclerotic plaques. J. Biol. Chem., 272, 3520-3526.
Leeuwenburgh, C., Wagner, P., Holloszy, J.O., Sohal, R.S., and Heinecke, J.W. (1997b). Caloric restriction attenuates
dityrosine cross-linking of cardiac and skeletal muscle proteins in aging mice. Arch. Biochem. Biophys., 346, 74-80.
Lengger, C., Schoch, G., and Topp, H. (2000). A high-performance liquid chromatographic method for the determination
of 8-oxo-7,8-2’-deocyguanosine in urine from man and rat. (2000). Anal. Biochem., 287, 65-72.
Lepoivre, M., Flaman, J.M., Bobe, P., Lemaire, G., and Henry, Y. (1994). Quenching of the tyrosyl free radical of
ribonucleotide reductase by nitric oxide. Relationship to cytostasis induced in tumor cells by cytotoxic macrophages. J.
Biol. Chem., 269, 21891-21897.
Lappa, S., and Sistonen, L. (1997). Heat shock response – pathophysiological implications. Annals Med., 29, 73-79.
Lesko, S.A., Lorentzen, R.J., and Ts’o P.O.P. (1980). The role of superoxide in deoxyribonucleic acid strand scission.
Biochem., 19, 3023-3028.
Levine, R.L., Mosoni, L., Berlett, B.S., and Stadtman, E.R. (1996). Methionine residues as endogenous antioxidants in
proteins. Proc. Natl. Acad. Sci. USA, 93, 15036-15040.
Levine, R.L., Williams, J.A., Stadtman, E.R., and Shacter, E. (1994). Carbonyl assays for determination of oxidatively
modified proteins. Meth. Enzymol., 233, 346-357.
Lewisch, S.A., and Levine, R.L. (1995). Determination of 2-oxohistidine by amino acid analysis. Anal. Biochem., 231, 440446.
Lewisch, S.A., and Levine, R.L. (1998). Determination of 2-oxohistidine by amino acid analysis. Meths. Enzymol., 300,
120-124.
Li, D.H., Wang, M.Y., Dhingra, K., and Hittelman, W.N. (1995a). Discovery of a 4-hydroxynonenal-induced DNA adduct in
32
human breast tissues by P-postlabeling assay. Proc. Ann. Meet. Am. Assoc. Cancer Res., 36, A826.
Li, D.H., Wang, M., Liehr, J.G., and Randerath, K. (1995b). DNA adducts induced by lipids and lipid peroxidation products:
Possible relationship to I-compounds. Mut. Res., 344, 117-126.
Li, S., Schoneich, C., and Borchardt, R.T. (1995c). Chemical pathways of peptide degradation. VIII. Oxidation of
methionine in small model peptides by prooxidant/transition metal ion systems: Influence of selective scavengers for
reactive oxygen intermediates. Pharmacol. Res., 12, 348-355.
Li, S., Nguyen, T.H., Schoneich, C., and Borchardt, R.T. (1995d). Aggregation and precipitation of human relaxin induced
by metal-catalyzed oxidation. Biochem., 34, 5762-5772.
Lieber, M.R., Grawunder, U., Wu, X., and Yaneva, M. (1997). Tying loose ends: roles of Ku and DNA-dependent protein
kinase in the repair of double-strand breaks. Curr. Opin. Genet. Dev., 7, 99-104.
Liehr, J.G. (1990). Genotoxic effects of estrogens. Mutat. Res., 238, 269-276.
Liehr, J.G. (1994). Catechol estrogens as mediators of estrogen-induced carcinogenesis. Proc. Ann. Meet. Am. Assoc.
Cancer Res., 35, 704.
Lim, D.G., Sweeney, S., Bloodsworth, A., White, C.R., Chumley, P.H., Krishna, N.R., Schopfer, F., O’Donnell, V.B.,
Eiserich, J.P., and Freeman, B.A. (2002). Proc. Natl. Acad. Sci. USA, 99, 15941-15946.
Lin, H.L., Kent, U.M., Zhang, H., Waskell, L., and Hollenberg, P.F. (2003). Mutation of tyrosine 190 to alanine eliminates
the inactivation of cytochrome P450 2B1 by peroxynitrite. Chem. Res. Tox., 16, 129-136.
Lin, J.K., Chen, K.J., Liu, G.Y., Chu, Y.R., and Lin-Shiau, S.Y. (2000). Nitration and hydroxylation of aromatic amino acid
and guanine by the air pollutant peroxyacetyl nitrate. Chem. Biol. Interact., 127, 219-236.
Lindahl, T. (1993). Instability and decay of the primary structure of DNA [see comments]. Nature, 362, 709-715.
Liochev, S.I., and Fridovich, I. (1994). The role of O2• in the production of HO•: In vitro and in vivo. Free Radic. Biol. Med.,
16, 29-33.
Liu, H., Duda, C.T., Huang, T., Aruda, W., and Kissinger, P.T. (1998a). Optimization of post column
photolysis/electrochemistry for liquid chromatographic determination of 3-nitro-L-tyrosine. J. Chromatogr. A, 818, 69-75.
Liu, H., Huang, T., Kissinger, C.B., and Kissinger, P.T. (1998b). Comparison of detection methods for liquid
chromatographic determination of 3-nitro-L-tyrosine. J. Chromatogr. B, 713, 289-295.
Liu, R.H., and Hotchkiss, J.H. (1995). Potential genotoxicity of chronically elevated nitric oxide: A review. Mutat. Res.,
339, 73-89.
Lloyd, D.R., and Phillips, D.H. (1999). Oxidative DNA damage mediated by copper (II), iron (II) and nickel (II) Fenton
reactions: Evidence for site-specific mechanisms in the formation of double-strand breaks, 8-hydroxydeoxyguanosine
and putative intrastrand cross-links. Mutat. Res., 424, 23-36.
Lodovici, M., Casalini, C., Cariaggi, R., Michelucci, L., and Dolara, P. (2000). Levels of 8-hydroxydeoxyguanosine as a
marker of DNA damage in human leukocytes. Free Radic. Biol. Med., 28, 13-17.
Loeb, L.A., and Preston, B.D. (1986). Mutagenesis by apurinic/apyrimidinic sites. Ann. Rev. Genet., 20, 201-230.
Loft, S., and Poulsen, H.E. (1996). Cancer risk and oxidative DNA damage in man. J. Mol. Med., 74, 297-312.
WWW.ESAINC.COM
292
Loft, S., and Poulsen, H.E. (1998). Markers of oxidative damage to DNA: Antioxidants and molecular damage. Meth.
Enzymol., 300, 166-184..
Loft, S., Astrup, A., Buemann, B., and Poulsen, H.E. (1994). Oxidative DNA damage correlates with oxygen consumption
in humans. FASEB J., 8, 534-537.
Loft, S., Larsen, P.N., Rasmussen, A., Fischer-Nielsen, A., Bondesen, S., Kirkegaard, P., Rasmussen, L.S., Ejlersen, E.,
Tornoe, K., Bergholdt, R., and Poulsen, H.E. (1995). Oxidative DNA damage after transplantation of the liver and small
intestine in pigs. Transplant., 59, 16-20.
Loft, S., Vistisen, K., Ewertz, M., Tjonneland, A., Overvad, K., and Poulsen, H.E. (1992). Oxidative DNA-damage
estimated by 8-hydroxydeoxyguanosine excretion in man: Influence of smoking, gender and body mass index.
Carcinogen., 13, 2241-2247.
Loft, S., Wierik, E.J.M.V., van der Berg, H., and Poulsen, H.E. (1995). Energy restriction and oxidative DNA damage.
Cancer Epid. Biomarkers Prev., 4, 515-519.
Lopez, M.F. (1997). Two-dimensional electrophoresis and mass spectrometry: Deciphering the proteome. Am. Biotech.
Lab., May.
Lopez, M.F. (1998a). Proteomic databases: Roadmaps for drug discovery. Am. Biotech. Lab., Accepted for publication.
Lopez, M.F. (1998b). Gene products are where the biological action is. J. Chrom. B., Accepted for 1998 publication.
Lovell, M.A., Gabbita, S.P., and Markesbery, W.R. (1999). Increased DNA oxidation and decreased levels of repair
products in Alzheimer’s disease ventricular CSF. J. Neurochem., 72, 771-776.
Lui, L., and Wells, P.G. (1995). DNA oxidation as a potential molecular mechanism mediating drug-induced birth defects:
Phenytoin and structurally related teratogenesis initiate the formation of 8-hydroxy-2-’deoxyguanosine in vitro and in
vivo in murine maternal hepatic and embryonic tissues. Free Radic. Biol. Med., 19, 639-648.
Lunec, J., Herbert, K., Blount, S., Griffiths, H.R., and Emery, P. (1994). 8-Hydroxydeoxyguanosine: A marker of oxidative
DNA damage in systemic lupus erythematosis. FEBS Lett., 348, 131-138.
Luo, Y., Henle, E.S., and Linn, S. (1996). Oxidative damage to DNA by iron-mediated Fenton reactions. The
deoxycytidine family. (1996). J. Biol. Chem., 271, 21167-21176.
Lymar, S.V., Jiang, Q., and Hurst, J.K. (1996). Mechanism of carbon dioxide-catalyzed oxidation of tyrosine by
peroxynitrite. Biochem., 35, 7855-7861.
Lyras, L., Cairns, N.J., Jenner, A., Jenner, P., and Halliwell, B. (1997). An assessment of oxidative damage to proteins,
lipids, and DNA in brain from patients with Alzheimer’s disease. J. Neurochem., 68, 2061-2069.
Lyras, L., Evans, P.J., Shaw, P.J., Ince, P.G., and Halliwell, B. (1996). Oxidative damage and motor neuron disease.
Difficulties in the measurement of protein carbonyls in human brain tissue. Free Radic. Res. Comm., 24, 397-406.
Lyras, L., Perry, R.H., Perry, E.K., Ince, P.G., Jenner, A., Jenner, P., and Halliwell, B. (1998). Oxidative damage to
proteins, lipids and DNA in cortical brain regions from patients with dementia with Lewy bodies. J. Neurochem., 71,
302-312.
Maccabee, M., Evans, J.S., Glackin, M., Hatahet, Z., and Wallace, S.S. (1994). Pyrimidine ring fragmentation products.
Effects of lesion structure and sequence context on mutagenesis. J. Mol. Biol., 236, 514-530.
MacMillan-Crow, L.A., and Cruthirds, D.L. (2001). Invited Review: Manganese superoxide dismutase in disease. Free
Radic. Rev., 34, 325-336.
MacMillan-Crow, L.A., and Thompson, J.A. (1999). Immunoprecipitation of nitrotyrosine-containing proteins. Meth.
Enzymol., 301, 135-144.
MacMillan-Crow, L.A., and Thompson, J.A. (1999b). Tyrosine modifications and inactivation of active site manganese
superoxide dismutase mutant (Y34F) by peroxynitrite. Arch. Biochem. Biophys., 366, 82-88.
MacMillan-Crow, L.A., Crow, J.P., Kerby, J.D., Beckman, J.S., and Thompson, J.A. (1996). Nitration and inactivation of
manganese superoxide dismutase in chronic rejection of human renal allografts. Proc. Natl. Acad. Sci. USA, 93, 1185311858.
MacMillan-Crow, L.A., Crow, J.P., and Thompson, J.A. (1998). Peroxynitrite-mediated inactivation of manganese
superoxide dismutase involves nitration and oxidation of critical tyrosine residues. Biochem., 37, 1613-1622.
Maier, K.L., Lenz, A.G., Beck-Speier, I., and Costabel, U. (1995). Analysis of methionine sulfoxide in proteins. Meths.
Enzymol., 251, 455-461.
Makinen, M., Piironen, V., and Hopia, A. (1996). Postcolumn chemiluminescence, ultraviolet and evaporative lightscattering detectors in high-performance liquid chromatographic determination of triacylglycerol oxidation products. J.
Chromatogr., 734, 221-229.
Malencik, D.A., Sprouse, J.F., Swanson, C.A., and Anderson, S.R. (1996). Dityrosine: Preparation, isolation and analysis.
Anal. Biochem., 242, 202-213.
Mani, A.R., Pannala, A.S., Orie, N.N., Harry, D., Rice-Evans, C.A., and Moore, K.P. (2003). Nitration of endogenous parahydroxyphenylacetic acid and the metabolism of nitrotyrosine. Biochem. J., 374, 521-527.
Mark, R.J., Lovell, M.A., Markesbery, W.R., Uchida, K., and Mattson, M.P. (1997). A role for 4-hydroxynonenal, an
aldehydic product of lipid peroxidation, in disruption of ion homeostasis and neuronal death induced by amyloid βpeptide. J. Neurochem., 68, 255-264.
Marnett, L.J. (1985). Arachidonic acid metabolism and tumor initiation. Prostagland. Leukotri. Cancer, 2, 39-82.
Marnett, L.J. (1996). Identification, detection, and genetic effects of DNA adducts of endogenous aldehydes. Proc. Ann.
Meet. Am. Assoc. Cancer Res., 37, 646.
Marquez, L.A., and Dunford, H.B. (1995). Kinetics of oxidation of tyrosine and dityrosine by myeloperoxidase I and II. J.
Biol. Chem., 270, 30434-30440.
Maruyama, W., Hashizume, Y., Matsubara, K., and Naoi, M. (1996). Identification of 3-nitro-L-tyrosine, a product of nitric
oxide and superoxide, as an indicator of oxidative stress in the human brain. J. Chromatogr. B., 676, 153-158.
Marvin, L.F., Delatour, T., Tavvazi, I., Fay, L.B., Cupp, C., and Guy, P.A. (2003). Quantification of dityrosine, onitrotyrosine and 0-tyrosine in cat urine by LC-MS/MS. Anal. Chem., 75, 261-267.
Maseki, M., Nishigaki, I., Hagihara, M., Tomoda, Y., and Yagi, K. (1981). Lipid peroxide levels and lipid content of serum
lipoprotein fractions of pregnant subjects with or without pre-eclampsia. Clin. Chim. Acta, 115, 155-161.
WWW.ESAINC.COM
293
Maskos, Z., Rush, J.D., and Koppenol, W.H. (1992). The hydroxylation of phenylalanine and tyrosine: A comparison with
salicylate and tryptophan. Arch. Biochem. Biophys., 296, 521-529.
Masuda, M., Suzuki, T., Friesen, M.D>, Ravanat, J.L., Cadet, J., Pignatelli, B., Nishino. H., and Ohshima, H. (2001).
Chlorination of guanosine and other nucleosides by hypochlorous acid and myeloperoxidase of activated human
neutrophils. Catalysis by nicotine and trimethylamine. J. Biol. Chem., 276, 40486-40496.
Matson, W.R. (1998). Personal communications.
Matthews, R.T., and Beal, M.F. (1996). Increased 3-nitrotyrosine in brains of apo E-deficient mice. Brain. Res., 718, 181184.
McCabe, D.R., Acworth, I.N., Maidt, M.L., and Floyd, R.A. (1997). A sensitive and selective method for the determination
th
of tissue 8-hydroxy-2’deoxyguanosine using HPLC with electrochemical array detection. Presented at The 88 Annual
Meeting of The American Association for Cancer Research. San Diego CA.
McCarter, R.J. (1995). Role of caloric restriction in the prolongation of life. Clin. Geriatr. Med., 11, 553-565.
McCarter, R.J., and McGee, J.R. (1989). Transient reduction of metabolic rate by food restriction. Am. J. Physiol., 257,
E175-179.
McKelvey-Martin, V.J., Green, M.H., Schmezer, P., Pool-Zobel, B.L., and De Meo, M.P. (1993). The single cell gelelectrophoresis assay (comet assay): A European review. Mutat. Res., 288, 47-63.
Mecocci, P., MacGarvey, U., Kaufman, A.F., Koontz, D., Shoffner, J.M., Wallace, D.C., and Beal, M.F. (1993). Oxidative
damage to mitochondrial DNA shows marked age-dependent increases in human brain. Ann. Neurol., 34, 609-616.
Mehl, M., Daiber, A., Herold, S., Shoun, H., and Ullrich, V. (1999). Nitric Oxide, 3, 142-152.
Mei, N., Tamae, K., Kunugita, N., Hirano, T., and Kasai, H. (2003). Analysis of 8-hydroxydeoxyguanosine 5’monophosphate (8-OH-dGMP) as a reliable marker of cellular oxidative DNA damage after γ-irradiation. Env. Mol.
Mutagen., 41, 332-338.
Mei, S.R., Yao, Q.H., Cai, L.S., Xing, J., Xu, G.W., Wu, C.Y. (2003). Capillary electrophoresis with end-column
amperometric detection of urinary 8-hydroxy-2’-deoxyguanosine. Electrophoresis, 24, 1411-1415.
Meneghini, R. (1988). Genotoxicity of active oxygen species in mammalian cells. Mutat. Res., 195, 215-230.
Menon, C. (1995). Ionizing radiation induced DNA hydroperoxide damage and its repair. Diss. Abstr. Int., 56, 1291-1295.
Michaels, M.L., Tchou, J., Grollman, A.P., and Miller, J.H. (1992). A repair system for 8-oxo-7,8-dihydrodeoxyguanine.
Biochem., 31, 10964-10968.
Michon, T., Chenu, M., Kellershon, N., Desmadril, M., and Gueguen, J. (1997). Horseradish peroxidase oxidation of
tyrosine-containing peptides and their subsequent polymerization: A kinetic study. Biochem., 36, 8504-8513.
Miller, M.J., Thompson, J.H., Zhang, X.J., Sadowska-Krowicka, H., Kakkis, J.L., Munshi, U.K., Sandoval, M., Rossi, J.L.,
Eloby-Childress, S., and Beckman, J.S. (1995). Role of inducible nitric oxide synthase expression and peroxynitrite
formation in guinea pig ileitis. Gastroenterol., 109, 1475-1483.
Minotti, G., and Aust, S.D. (1989). The role of iron in oxygen radical mediated lipid peroxidation. Chem. Biol. Interact., 71,
1-19.
Minotti, G., and Aust, S.D. (1992). Redox cycling and lipid peroxidation. Lipids, 27, 219-226.
Miyake, E., Eto, H., Takechi, Y., Kamidono, S., and Hara, I. (2003). Increased urinary 8-hydroxy-2’-deoxyguanosine
excretion after ileal neobaldder replacement. BJU Int., 91, 657-660.
Miyazawa, T., Yasuda, K., Fujimoto, K., and Kaneda, T. (1988). Presence of phosphatidylcholine hydroperoxide in human
plasma. J. Biochem., 103, 744-749.
Mo, J.Y., Maki, H., and Sekiguchi, M. (1992). Hydrolytic elimination of a mutagenic nucleotide, 8-oxodGTP by human 18kilodalton protein: Sanitization of nucleotide pool. Proc. Natl. Acad. Sci. USA, 89, 11021-11025.
Mobley, J.A., Bhat, A.S., and Brueggemeier, R.W. (1999). Measurement of oxidative DNA damage by catechol estrogens
and analogues in vitro. Chem. Res. Toxicol., 12, 270-277.
Mohr, S., Stamler, J.S., and Brune, B. (1994). Mechanism of covalent modification of glyceraldehyde-3-phosphate
dehydrogenase at its active site thiol by nitric oxide, peroxynitrite and related nitrosating agents. FEBS Lett., 348, 223227.
Monnier, V.M., Sell, D.R., Saxena, A., Saxena, M., Subramaniam, R., Tessier, F., and Weiss, M. F. (2003). Glycoxidative
And Carbonyl Stress In Aging And Age-Related Diseases. In: Critical Reviews of Oxidative Stress and Aging. Volume
1.Cutler, R.G., and Rodriguez, H. (Eds.). World Scientific, NJ. Pp. 413-433.
Moody, C.S., and Hassan, H.M. (1982). Mutagenicity of oxygen free radicals. Proc. Natl. Acad. Sci. USA, 79, 2855-2859.
Morin, R.J., and Peng, S.K. (1989). The role of cholesterol oxidation products in the pathogenesis of atherosclerosis. Ann.
Clin. Lab. Sci., 19, 225-237.
Moriya, M. (1993). Single-stranded shuttle phagemid for mutagenesis studies in mammalian cells: 8-Oxoguanine in DNA
induces targeted C.C→T.A transversions in simian kidney cells. Proc. Natl. Acad. Sci. USA, 90, 1122-1126.
Morrow, J.D., and Roberts, L.J. (1996). The isoprostanes. Current knowledge and directions for future research. Biochem.
Pharmacol., 51, 1-9.
Morrow, J.D., and Roberts, L.J. (1999). Mass spectrometric quantification of F2-isoprostanes in biological fluids and
tissues as measure of oxidant stress. Meths. Enzymol., 300, 3-12.
Morrow, J.D., Chen, Y., Brame, C.J., Yang, J., Sanchez, S.C., Xu, J., Zackert, W.E., Awad, J.A., and Roberts, L.J. (1999).
The isoprostanes: Unique prostaglandin-like products of free radical-initiated lipid peroxidation. Drug Metab. Rev., 31,
117-139.
Morrow, J.D., Reich, E.E., Roberts, L.J., and Montine, T.J. (2003). Quantification Of Isoprostanes As An Index Of Oxidant
Stress Status In Vivo. In: Critical Reviews of Oxidative Stress and Aging. Volume 1.Cutler, R.G., and Rodriguez, H.
(Eds.). World Scientific, NJ. Pp. 383-392.
Morton, L.W., Puddy, I.B., and Croft, K.D. (2003). Comparison of nitration and oxidation of tyrosine in advanced human
carotid plaque proteins. Biochem. J., 370, 339-344.
Muller, K., Hardwick, S.J., Marchant, C.E., Law, N,S., Waeg, G., Esterbauer, H., Carpentar, K.L., and Mitchinson, M.J.
(1996). Cytotoxic and chemotactic potencies of several aldehyde components of oxidized low-density lipoprotein for
human monocyte-macrophages. FEBS Lett., 388, 165-168.
WWW.ESAINC.COM
294
Myatt, L., Rosenfield, R.B., Eis, A.L., Brockman, D.E., Greer, I., and Lyall, F. (1996). Nitrotyrosine residues in placenta.
Evidence of peroxynitrite formation and action. Hyperten., 28, 488-493.
Nackerdien, Z., Olinski, R., and Dizdaroglu, M. (1992). DNA base damage in chromatin of gamma-irradiated cultured
human cells. Free Radic. Res. Commun., 16, 259-273.
Nagashima, M., Tsuda, H., Takenoshita, S., Nagamachi, Y., Hirohashi, S., Yokota, J., and Kasai, H. (1995). 8Hydroxydeoxyguanosine levels in DNA of human breast cancer are not significantly different from those of noncancerous breast tissue by the HPLC-ECD method. Cancer Lett., 90, 157-162.
Nair, U., Nair, J., Friesen, D., Bartsch, H., and Ohshima, H. (1995). Ortho- and meta-tyrosine formation from
phenylalanine in human saliva as a marker of hydroxyl radical generation during betel quid chewing. Carcinogen., 16,
1195-1198.
Nair, V., Vietti, D.E., and Cooper, C.S. (1981). Degenerative chemistry of malondialdehyde. Structure, stereochemistry
and kinetics of formation of enaminals from reaction with amino acids. J. Am. Chem. Soc., 103, 3030-3096.
Nakajima, M., Takeuchi, T., and Morimoto, K. (1996). Determination of 8-hydroxydeoxyguanosine in human cells under
oxygen-free conditions. Carcinogen., 17, 787-791.
Nashimura, S. (2002). Involvement of mammalian OGG1(MMH) in excision of the 8-hydroxyguanine residue in DNA. Free
Radic. Biol. Med., 32, 813-821.
Nassi-Calo, L., Mello Fihlo, A.C., and Meneghini, R. (1989). O-Phenanthroline protects mammalian cells from hydrogen
peroxide-induced gene mutation and morphological transformation. Carcinogen., 10, 1055-1057.
2
Nath, R.G., Ocando, J.E., Guttenplan, J.B., and Chung, F.L. (1998). 1,N -propanodeoxyguanosine adducts: Potential new
biomarkers of smoking-induced DNA damage in human oral tissue. Cancer Res., 58, 581-584.
Netto, L.E.S., RamaKrishna, N.V.S., Kolar, C., Cavalieri, E.L., Rogan, E.G., Lawson, T.A., and Augusto, O. (1992).
8
Identification of C -methylguanine in the hydrolysates of DNA from rats administered 1,2-dimethylhydrazine. Evidence
for in vivo DNA alkylation by methyl radicals. J. Biol. Chem., 267, 21524-21527.
Neuzil, J., and Stocker, R. (1993). Bilirubin attenuates radical-mediated damage to serum albumin. FEBS Lett., 331, 281284.
Nicotera, T.M., Munson, B.R., and Fiel, R.J. (1994). Photosensitized formation of 8-hydroxy-2'-deoxyguanosine and DNA
strand breakage by a cationic meso-substituted porphyrin. Photochem. Photobiol., 60, 295-300.
Nielsen, H.B., Hanel, B., Loft, S., Poulsen, H.E, Pedersen, B.K., Diamant, M., Vistisen, K., and Secher, N.H. (1995).
Restricted pulmonary diffusion capacity after exercise is not an ARDS-like injury. J. Sports Sci., 13, 109-113.
Nishigaki, I., Hagihara, M., Tsunekawa, H., Maseki, M., and Yagi, K. (1981). Lipid peroxide levels of serum lipoprotein
fractions of diabetic patients. Biochem. Med., 25, 373-378.
Nomura, K., Suzuki, N., and Matsumoto, S. (1990). Pulcherosine, a novel tyrosine-derived, trivalent cross-linking amino
acid from the fertilization envelope of sea urchin embryo. Biochem., 29, 4525-4534.
Norsten-Hoog, C., and Cronholm, T. (1990). Analysis of aldehydic lipid peroxidation products in rat liver and hepatocytes
by gas chromatography and mass spectrometry of the oxime-t-butyldimethylsilyl derivatives. Anal. Biochem., 189, 131137.
Nosworthy, J., and Allsop, C.B. (1956). The radiolysis of aqueous solutions of tyrosine. J. Colloid Sci., 11, 565-574.
Nourooz-Zadeh, J., Liu, E.H.C., Yhlen, B., Anggard E.E., and Halliwell, B. (1999). F4 isoprostanes as specific markers of
docosahexaenoic peroxidation in Alzheimer’s disease. J. Neurochem., 72, 734-740.
O’Connor, T.R., Boiteux, S., and Laval, J. (1988). Ring-opened 7-methyl guanine residues in DNA are a block to in vitro
DNA synthesis. Nucleic Acid Res., 16, 5879-5894.
O’Donnell, V.B., Eiserich, J.P., Bloodsworth, A., Chumley, P.H., Kirk, M., Barnes, S., Darley-Usmar, V.M., and Freeman,
B.A. (1999a). Nitration of unsaturated fatty acids by nitric oxide-derived reactive species. Meth. Enzymol., 301, 454470.
O’Donnell, V.B., Eiserich, J.P., Chumley, P.H., Jablonsky, M.J., Krishna, N.R., Kirk, M., Barnes, S., Darley-Usmar, V.M.,
and Freeman, B.A. (1999b). Nitration of unsaturated fatty acids by nitric oxide-derived reactive nitrogen species
peroxynitrite, nitrous acid, nitrogen dioxide, and nitronium ion. Chem. Res. Toxicol., 12, 83-92.
O’Gara, C.Y., Maddipati, K.R., and Marnett, L.J. (1989). A sensitive electrochemical method for quantitative
hydroperoxide determination. Chem. Res. Toxicol., 2, 296-300.
Ohshima, H., Brouet, I., Friesen, M., and Bartsch, H. (1991). Nitrotyrosine as a new marker of endogenous nitrosation and
nitration. IARC Sci. Publ., 106, 443-448.
Ohshima, H., Celan, I., Chazotte, L., Pignatelli, B., and Mower, H.F. (1999). Analysis of 3-nitrotyrosine in biological fluids
and protein hydrolyzates by high-performance liquid chromatography using postseparation, on-line reduction column
and electrochemical detection: Results with various nitrating agents. Nitric Oxide, 3. In press.
Ohshima, H., Friessen, M., Brouet, I., and Bartsch, H. (1990). Nitrotyrosine as a new marker for endogenous nitrosation
and nitration of proteins. Fd. Chem. Toxicol., 28, 647-652.
Okamoto, K., Toyokuni, S., Uchida, K., Ogawa, O., Takenawa, J., Kahehi, Y., Kinoshita, H., Hattori-Nakakuki, Y., Hiai, H.,
and Yoshida, O. (1994). Formation of 8-hydroxy-2’deoxyguanosine and 4-hydroxy-2-nonenal-modified proteins in
human renal-cell carcinoma. Int. J. Cancer, 58, 825-829.
Okuda, Y., Sakoda, S., Fujimura, H., and Yanagihara, T. (1997). Nitric oxide via an inducible isoform of nitric oxide
synthase is a possible factor to eliminate inflammatory cells from the central nervous system of mice with experimental
allergic encephalomyelitis. J. Neuroimmunol., 73, 107-116.
Olinski, R., Zastawny, T.H., Foksinski, M., Barecki, A., and Dizdaroglu, M. (1995). DNA base modifications and
antioxidant enzyme activities in human benign prostatic hyperplasia. Free Radic. Biol. Med., 18, 807-813.
Oliver, C.N., Ahn, B., Moerman, E.J., Goldstein, S., and Stadtman, E.R. (1987). Age-related changes in oxidized proteins.
J. Biol. Chem., 262, 5488-5491.
Orr, W.C., and Sohal, R.S. (1994). Extension of life-span by overexpression of superoxide dismutase and catalase in
Drosophila melanogaster. Science, 263, 1128-1130.
Ohshima, H., and Bartsch, H. (1994). Chronic infections and inflammatory processes as cancer risk factors: Possible role
of nitric oxide in carcinogenesis. Mutat. Res., 305, 253-264.
WWW.ESAINC.COM
295
Pacifici, R.E., Kono, Y., and Davies, K.J. (1993). Hydrophobicity as the signal for selective degradation of hydroxyl radicalmodified hemoglobin by the multicatalytic proteinase complex, proteosome. J. Biol. Chem., 268, 15405-15411.
Padmaja, S., Ramazenian, M.S., Bounds, P.L., and Koppenol, W.H. (1996). Reaction of peroxynitrite with L-tryptophan.
Redox Report, 2, 173-177.
Padmaja, S., Squadrito, G.L., and Pryor, W.A. (1998). Inactivation of glutathione peroxidase by peroxynitrite. Arch.
Biochem. Biophys., 349, 1-6.
Park, E.-M., Shigenaga, M.K., Degan, P., Korn, T.S., Kitlzer, J.W., Wehr, C.M., Kolachana, P., and Ames, B.N. (1992).
Assay of excised oxidative DNA lesions: Isolation of 8-oxoguanine and its nucleoside derivatives from biological fluids
with a monoclonal antibody column. Proc. Natl. Acad. Sci. USA, 89, 3375-3379.
Patel, R.P., Diczfalusy, U., Dzeletovic, S., Wilson, M.T., and Darley-Usmar, V.M. (1996). Formation of oxysterols during
oxidation of low-density lipoprotein by peroxynitrite, myoglobin and copper. J. Lipid Res., 37, 2361-2371.
Patel, U., Bhimani, R., and Frenkel, K. (1992). Mechanism of mutagenicity by 5-hydroperoxymethyl-2’deoxyuridine, an
intermediate product of ionizing radiation, in bacteria. HPMdU bacterial mutagenicity and oxidation of DNA bases.
Mutat. Res., 238, 145-156.
Paul, S., Bogdanov, M.B., Matson, W.R., Metakis, L., Jacobs, J., and Beal, M.F. (2003). Urinary 8-hydroxy-2’deoxyguanosine, a metabolite of oxidized DNA, is not elevated in HIV patients on combination antiretroviral therapy.
Free Radic. Res., 37, 499-502.
Pedersen, J.Z., and Finazzi, A.A. (1993). Protein-radical enzymes. FEBS Lett., 325, 53-58.
Peng, S.K., and Morin, R.J. (1987). Effects of membrane function by cholesterol oxidation derivatives in cultured aortic
smooth muscle cells. Artery, 14, 85-99.
Perez-Campo, R., Lopez-Torres, M., Cadenas, S., Rojas, C., and Barja, G. (1998). The rate of free radical production as a
determinant of the rate of aging: Evidence from the comparative approach. J. Comp. Physiol, B, 168, 149-158.
Petruzzelli, S., Puntoni, R., Mimotti, P., Pulera, N., Baliva, F., Fornai, E., and Giuntini, C. (1997). Plasma 3-nitrotyrosine in
cigarette smokers. Am. J. Crit. Care Med., 156, 1902-1907.
Pfeiffer, S., and Mayer, B. (1998). Lack of tyrosine nitration by peroxynitrite generated in situ from nitric oxide and
superoxide. Free Radic. Biol. Med., 25, Supplement 1, S64.
Pietta, P.G., Simonetti, P., Gardana, C., Cristoni, S., Bramati, L., and Mauri, P.L. (2003). LC-APCI-MS/MS analysis of
urinary 8-hydroxy=2’-deoxyguanosine. J. Pharmaceu. Biomed. Anal., 32, 657-661.
Pilger, A., Ivancsits, S., Germadnik, D., and Rudiger, H.W. (2002). Urinary excretion of 8-hydroxy-2’deoxyguanosine
measured by high-performance liquid chromatography with electrochemical detection. J. Chromatogr. B, 778, 393-401.
Piperakis, S.M., Visvardis, E.-E., and Tassiou (1998). Comet assay for nuclear DNA damage. Meth. Enzymol, 300. 184194.
Podmore, I.D., Griffiths, H.R., Herbert, K.E., Mistry, N., Mistry, P., and Lunec, J. (1998). Vitamin C exhibits pro-oxidant
properties. Nature, 392, 569.
Poirier, M.C., and Weston, A. (1996). Human DNA adduct measurement: State of the art. Environ. Health Perspect., 104,
Suppl. 5, 883-893.
Poli, G., Diianzani, M.U., Chesseman, K.H., Slater, T.F., and Lang, J., and Esterbauer, H. (1985). Separation and
characterization of the aldehydic products of lipid peroxidation stimulated by carbon tetrachloride or ADP-iron in
isolated rat hepatocytes and rat liver microsomal suspension. Biochem. J., 227, 629-638.
Poulsen, H.E., and Loft, S. (1998). Interpretation of oxidative DNA modification: Relation between tissue levels, excretion
of urinary repair products and single cell gel electrophoresis (comet assay). In: DNA and Free Radicals: Techniques,
Mechanisms and Applications. O.I. Aruoma and B. Halliwell (Eds.). OICA International. In Press.
Poulsen, H.E., Loft, S., Jensen, B.R., Sorensen, M., Hoberg, A,-M., and Weimann, A. (2003). HPLC-ECD, HPLC-MS/MS
(Urianry Biomarkers). In: Critical Reviews of Oxidative Stress and Aging. Volume 1.Cutler, R.G., and Rodriguez, H.
(Eds.). World Scientific, NJ. Pp. 233-256.
Poulsen, H.E., Loft, S., and Vistisen, K. (1996). Extreme exercise and oxidative DNA modification. J Sports Sci., 14, 343346.
Poulsen, H.E., Loft, S., and Weimann, A. (1998). Urinary measurement of 2-oxodG (8-oxo-2’-deoxyguanosine. In:
Handbook of Clinical Analysis: In Vivo Oxidative Damage to Biomolecules. Lunec, J. (Ed.). In press.
Povirk, L.F. (1996). DNA damage and mutagenesis by radiomimetic DNA-cleaving agents: bleomycin, neocarzinostatin
and other enediynes. Mutat. Res., 355, 71-89.
Pratico, D., Barry, O.P., Lawson, J.A., Adiyaman, M., Hwang, S.W., Khanapure, S.P., Iuliano, L., Rokach, J., and
Fitzgerald, G.A. (1998). IPF2α-I: An index of lipid peroxidation in humans. Proc. Natl. Acad. Sci. USA, 95, 3449-3454.
Pratico, D., Reilly, M., Lawson, J.A., and FitzGerald, G.A. (1997). Novel indices of oxidant stress in cardiovascular
disease: Specific analysis of F2-isoprostanes. Prostaglandins and Control of Vascular Smooth Muscle Cell Proliferation.
Birkhauser Verlag Basel. Pp. 26-41.
Pryor, W.A. (1988). Why is the hydroxyl radical the only radical that commonly adds to DNA? Hypothesis: It has a rare
combination of high electrophilicity, high thermochemical reactivity, and a mode of production that can occur near DNA.
Free Radic. Biol. Med., 4, 219-223.
Pryor, W.A. (1989). On the detection of lipid hydroperoxides in biological samples. Free Radic. Biol. Med., 7, 177-178.
Pryor, W.A., and Stanley, P.A. (1975). A suggested mechanism for the production of malondialdehyde during the
oxidation of polyunsaturated fatty acids. Nonenzymatic production of prostaglandin endoperoxides during autoxidation.
J. Org. Chem., 40, 3615-3617.
Purmal, A.A., Kow, Y.W., and Wallace, S.S. (1994). Major oxidative products of cytosine, 5-hydroxycytosine and 5hydroxyuracil, exhibit sequence context-dependent mispairing in vitro. Nucleic Acid Res., 22, 72-88.
Qian, S.Y., and Buettner, G.R. (1999). Iron and dioxygen chemistry is an important route to initiation of biological free
radical oxidations: An electron paramagnetic resonance spin trapping study. Free Radic. Biol. Med., 26, 1447-1456.
Quijano, C., Hernandez-Saavedra, D., Castro, L., McCord, J.M., Freeman, B.A., and Radi, R. (2001). Reaction of
peroxynitrite with Mn-superoxide dismutase. Role of the metal center in decomposition kinetics and nitration. J. Biol.
Chem., 276, 11631-11638.
WWW.ESAINC.COM
296
Raboy, B., Sharon, G., Parag, H.A., Shochat, Y., and Kulka, R.G. (1991). Effect of stress protein degradation: Role of
ubiquitin system. Acta Biol. Hung., 42, 3-20.
Radi, R., Beckman, J.S., Bush, K.M., and Freeman, B.A. (1991). Peroxynitrite-induced membrane lipid peroxidation: The
cytotoxic potential of superoxide and nitric oxide. Arch. Biochem. Biophys., 288, 481-487.
Ramezainian, M.S., Padmaja, S., and Koppenol, W.H. (1996). Nitration and hydroxylation of phenolic compounds by
peroxynitrite. Chem. Res. Toxicol., 9, 232-240.
Randeroth, K., Li, D.H., and Randerath, E. (1990). Age-related DNA modifications (I-compounds): Modulation by
physiological and pathological processes. Mutat. Res., 238, 245-253.
Randerath, K., Li, D., Nath, R., and Randerath, E. (1992). Exogenous and endogenous DNA modifications monitored by
32
P-postlabeling: Relationships to cancer and aging. Exp. Gerentol., 27, 533-549.
32
Randeroth, K., Reddy, M.V., and Gupta, R.C. (1981). P-Labeling for DNA damage. Proc. Natl. Acad. Sci. USA, 78,
6126-6129.
Randeroth, K., Reddy, M.V., and Disher, R.M. (1986). Age- and tissue-related DNA modifications in untreated rats:
32
Detection by p-postlabeling assay and possible significance for spontaneous tumor induction and aging. Carcinogen.,
7, 1615-1617.
Raoul, S., Bardet, M., and Cadet, J. (1995). γ Irradiation of 2’-deoxyadenosine in oxygen-free aqueous solutions:
Identification and conformational features of formamidopyrimidine nucleoside derivatives. Chem. Res. Toxicol., 8, 924933.
Ravanat, J.L., and Cadet, J. (1995). Reaction of singlet oxygen with 2’deoxyguanosine and DNA. Isolation and
characterization of the mian oxidation products. Chem. Res. Toxicol., 8, 379-388.
Ravanat, J.-L., Duretz, B., Guiller, A., Douki, T., and Cadet, J. (1998). Isotope dilution high-performance liquid
chromatography-electrospray tandem mass spectrometry assay for the measurement of 8-oxo-7,8-dihydro-2’deoxyguanosine in biological samples. J. Chromatogr. B, 715, 349-356.
Ravanat, J.-L., Guicherd, P., Tuce, Z., and Cadet, J. (1999). Simultaneous determination of five oxidative DNA lesions in
human urine. Chem. Res. Toxicol. In press.
Ravanat, J.-L., Turesky, R.J., Gremaud, E., Trudel, L.J., and Stadler, R.H. (1995). Determination of 8-oxoguanine in DNA
by gas chromatography-mass spectrometry and HPLC-electrochemical detection: Overestimation of the background
level of the oxidized base by the gas chromatography-mass spectrometry assay. Chem. Res. Toxicol., 8, 1039-1045.
Reddy, V.P., Obrenovich, M.E., Atwood, C.S., Perry, G., and Smith, M.A. (2002). Involvement of Maillard reactions in
Alzheimer’s disease. Neurotox. Res., 4, 121-209.
Renner, T., Fechner, T., and Scherer, G. (2000). Fast quantification of urinary marker of oxidative stress 8-hydroxy-2’deoxyguanosine using solid-phase extraction and high-performance liquid chromatography with triple-stage quadrupole
mass detection. J. Chromatogr. B, 738, 311-317.
Retel, J., Hoebee, B., Braun, J.E.F., Lutgerink, J.T., van den Akker, E., Wanamarta, A.H., Joneje, H., and Lafleur, M.V.M.
(1993). Mutational specificity of oxidative DNA damage. Mutat. Res., 299, 165-182.
Riis, B. (2002). Comparison of results from different laboratories in measuring 8-oxo-2’-deoxyguanosine in synthetic
oligonucleotides. Free Radic. Res., 36, 649-659.
Roberts, L.J., and Morrow, J.D. (1997). The generation and actions of isoprostanes. Biochim. Biophys. Acta, 1345, 121135.
Ronchi, S., Galliano, M., Minchiotti, L., Curti, B., Rudie, N.G., Porter, D.J., and Bright, H.J. (1980). An active site-tyrosinecontaining heptapeptide from D-amino acid oxidase. J. Biol. Chem., 255, 6044-6046.
Routledge, M.N., Wink, D.A., Keefer, L.K., and Dipple, A. (1994). DNA sequence changes induced by two nitric oxide
donor drugs in the supF assay. Chem. Res. Toxicol., 7, 628-632.
Rouzer, C.A., Chaudhary, A.K., Nokubo, M., Ferguson, D.M., Reddy, G.R., Blair, I.A., and Marnett, L.J. (1997). Analysis
of malondialdehyde-2’-deoxyguanosine adduct pyrimidopurinone in human leukocyte DNA by gas
chromatography/electron capture/negative chemical ionization/mass spectrometry. Chem. Res. Toxicol., 10, 181-188.
Rowbottom, J. (1955). The radiolysis of aqueous solutions of tyrosine. J. Biol. Chem., 212, 877-885.
Rowley, D.A., and Halliwell, B. (1983). DNA damage by superoxide-generating systems in relation to the mechanism of
action of the antitumor antibiotic adriamycin. Biochim. Biophys. Acta, 761, 86-93.
Rozalski, R., Gackowski, D., Roszkowski, K., Foksinski, M., and Olinski, R. (2002). The level of 8-hydroxyguanine, a
possible repair product of oxidative DNA damage, is higher in urine of patients than in control subjects. Cancer,
Epidemiol. Biomarkers, Prev., 11, 1072-1075.
Rubbo, H., Radi, R., Trujillo, M., Telleri, R., Kalyanaraman, B., Barnes, S., Kirk, M., and Freeman, B. (1994). Nitric oxide
regulation of superoxide and peroxynitrite-dependent lipid peroxidation. J. Biol. Chem., 269, 26066-26075.
Ryter, S.W., and Tyrrell, R.M. (1998). Singlet oxygen: A possible effector of eukaryotic gene expression. Free Radic. Biol.
Med., 24, 1520-1534.
Sakumi, K., Furuichi, M., Tsuzuki, T., Kakuman, T., Kawabata, S., Maki, H., and Sekiguchi, M. (1993). Cloning and
expression of a cDNA for a human enzyme that hydrolyzes 8-oxo-dGTP, a mutagenic substrate for DNA synthesis. J.
Biol. Chem., 268, 23524-23530.
Saleh, D., Barnes, P.J., and Giaid, A. (1997). Increased production of the potent oxidant peroxynitrite in the lungs of
patients with idiopathic pulmonary fibrosis. Am. J. Respir. Crit. Care Med., 155, 1763-1769.
Salman-Tabcheh, S., Guerin, M.C., and Torreilles, J. (1995). Nitration of tyrosyl-residues from extra- and intracellular
proteins of human whole blood. Free Radic. Biol. Med., 19, 695-698.
Salman-Tabcheh, S., Rabgaoui, N., and Torreilles, J. (1993). Neutrophil-catalyzed dimerization of tyrosyl peptides. Free
Radic. Res. Commun., 19, 217-227.
Santos, M.T., Valles, J., Aznar, J., and Vilches, J. (1980). Determination of plasma malondialdehyde-like material and its
clinical application in stroke patients. J. Clin. Pathol., 33, 973-976.
Sato, M., Fukuyama, N., Sakai, M., and Nakazawa, H. (1998). Increased nitric oxide in nasal lavage fluid and nitrotyrosine
formation in nasal mucosa – indices of severe perennial nasal allergy. Clin. Exp. Allergy, 28, 597-605.
WWW.ESAINC.COM
297
Satoh, K. (1978). Serum lipid peroxide in cerebrovascular disorders determined by a new colorimetric method. Clin. Chim.
Acta, 90, 37-43.
Satoh, M.S., Jones, C.J., Wood, R.D., and Lindahl, T. (1993). DNA excision-repair defect of xeroderma pigmentosum
prevents removal of a class of oxygen free radical-induced base lesions. Proc. Natl. Acad. Sci. USA, 90, 6335-6339.
Savenkova, M.I., Mueller, D.M., and Heinecke, J.W. (1994). Tyrosyl radical generated by myeloperoxidase is a
physiological catalyst for the initiation of lipid peroxidation in low density lipoproteins. J. Biol. Chem., 269, 2092-2097.
Sayre, L.M., Zelasko, D.A., Harris, P.L.R., Perry, G., Salomon, R.G., and Smith, M.A. (1997). 4-Hydroxynonenal-derived
advanced lipid peroxidation end products are increased in Alzheimer’s disease. J. Neurochem., 68, 2092-2097.
Schmidt, P., Youhnovski, N., Daiber, A., Balan, A., Arsic, M., Bachschmidt, M., Przybylski, M., and Ullrich, V. (2003).
Specific nitration of tyrosine 430 revealed by high resolution mass spectrometry as basis for redox regulation of bovine
prostacyclin synthase. J. Biol. Chem., 278, 12813-12819.
Schulz, J.B., Matthews, R.T., Muqit, M.M.K., Browne, S.E., and Beal, M.F. (1995). Inhibition of neuronal nitric oxide
synthase by 7-nitroindazole protects against MPTP-induced neurotoxicity in Mice. J. Neurochem., 64, 936-939.
Schwartz, A.L., and Ciechanover, A. (1992). Ubiquitin-mediated protein modification and degradation. Am. J. Respir. Cell
Mol. Biol., 7, 463-468.
Schwartz, A.L., and Ciechanover, A. (1999). The ubiquitin-proteasome pathway and pathogenesis of human diseases.
Ann. Rev. Med., 50, 57-74.
Seeburg, P.H., Colby, W.W., Capon, D.J., Goeddel, D.V., and Levinson, A.D. (1984). Biological properties of human cHA-ras1 genes mutated at codon 12. Nature, 312, 71-75.
Seeberg, E., Eide, L., Bjoras, M., and Bjras, M. (1995). The base excision repair pathway. Trends Biochem. Sci., 20, 391397.
Sell, D.R., Nagaraj, R.H., Grandhee, S.K., Odetti, P., Lapolla, A., Fogarty, J., and Monnier, V.M. (1991). Pentosidine: A
molecular marker for cumulative damage to proteins in diabetes, aging and uremia. Diabetes Metab. Rev., 7, 239-251.
Selley, M.L. (1997). Determination of the lipid peroxidation product (E)-4-hydroxy-2-nonenal in clinical samples by gas
chromatography-negative-ion chemical ionization mass spectrometry of the O-pentafluorobenzyl oxime. J.
Chromatogr., 691, 263-268.
Selley, M.L. (1998). (E)-4-Hydroxy-2-nonenal may be involved in the pathogenesis of Parkinson’s disease. Free Radic.
Biol. Med., 25, 169-174.
Selley, M.L., Bartlett, M.R., McGuiness, J.A., Hapal, A.J., and Ardlie, N.G. (1989). Determination of the lipid peroxidation
product trans-4-hydroxy-2-nonenal in biological samples by high performance liquid chromatography and combined
capillary gas chromatography/negative ion chemical ionization mass spectrometry. J. Chromatogr., 488, 329-340.
Selley, M.L., Bourne, D.J., Bartlett, M.R., Tymms, K.E., Brook, A.S., Duffield, A.M., and Ardlie, N.G. (1992). Occurrence of
(E)-4-hydroxy-2-nonenal in plasma and synovial fluid of patients with rheumatoid arthritis and osteoarthritis. Ann.
Rheum. Dis., 51, 481-484.
Semenza, J.C., and Weasel, L.H. (1997). Molecular epidemiology in environmental health: The potential of tumor
suppressor gene p53 as a biomarker. Environ. Health Perspect., 105, 155-163.
Sergent, O., Morel, I., and Cillard, J. (1999). Involvement of metal ions in lipid peroxidation: Biological implications. In:
Metal Ions in Biological System, 36. Sigel, A., and Sigel, H. (Eds.). Marcel Dekker, Inc., New York. Pp. 251-287.
Sharma, M., Box, H.C., and Kelman, D. (1990). Fluorescence postlabeling assay of cis-thymidine glycol monophosphate
in X-irradiated calf-thymus DNA. Chem. Biol. Interact., 74, 107-117.
Sharma, M., and Jane, R. (1998). Isolation and analysis of dityrosine from enzyme-catalyzed oxidation of tyrosine and Xirradiated peptide and protein. Chem. Biol. Interact., 108, 171-185.
Sherman, M.Y., and Goldberg, A.L. (1996). Involvement of molecular chaperones in intracellular protein breakdown.
EXS., 77, 57-78.
Shen, Z., Mitra, S.N., Wu, W., Chen, Y., Yang, Y., Qin, J., and Hazen, S.L. (2001). Eosinophil peroxidase catalyzes
bromination of free nucleosides and double-stranded DNA. Biochem., 40, 2041-2051.
Shibutani, S. (1993). Quantitations of base substitutions and deletions induced by chemical mutagens during DNA
synthesis in vitro. Chem. Res. Toxicol., 6, 625-629.
Shibutani, S., Takeshita, M., and Grollman, P. (1991). Insertion of specific bases during DNA synthesis past the oxidation
damaged base 8-oxodG. Nature, 349, 431-434.
Shigenaga, M.K. (1999). Quantitation of protein-bound 3-nitrotyrosine by high-performance liquid chromatography with
electrochemical detection. Meth. Enzymol., 301, 27-40.
Shigenaga, M.K., and Ames, B.N. (1991). Assays for 8-hydroxy-2’-deoxyguanosine: A biomarker of in vivo oxidative DNA
damage. Free Radic. Biol. Med., 10, 211-216.
Shigenaga, M.K., Aboujaoude, E.N., Chen, Q., and Ames, B.N. (1994). Assays of oxidative DNA damage biomarkers 8oxo-2’-deoxyguanosine and 8-oxoguanine in nuclear DNA and biological fluids by high performance liquid
chromatography and electrochemical detection. Meth. Enzymol., 234, 16-33.
Shigenaga, M.K., Gimeno, C.J., and Ames, B.N. (1989). Urinary 8-hydroxy-2’deoxyguanosine as a biological marker of in
vivo oxidative DNA damage. Proc. Natl. Acad. Sci. USA, 86, 9697-9701.
Shigenaga, M.K., Lee, H.H., Blount, B.C., Christen, S., Shigeno, E.T., Yip, H., and Ames, B.N. (1997). Inflammation and
NOx-induced nitration: Assay for 3-nitrotyrosine by HPLC with electrochemical detection. Proc. Natl. Acad. Sci. USA,
94, 3211-3216.
Shigenaga, M.K., Park, J.-W., Cundy, K.C., Gimeno, C.J., and Ames, B.N. (1990). In vivo oxidative DNA damage:
Measurement of 8-hydroxy-2’-deoxyguanosine in DNA and urine by high-performance liquid chromatography with
electrochemical detection. Meth. Enzymol., 186, 521-530.
Shimoda, R., Nagshima, M., Sakamoto, M., Yamaguchi, M., Hirohashi, S., Yokota, J., and Kasai, H. (1994). Increased
formation of oxidative DNA damage, 8-hydroxydeoxyguanosine in human liver with chronic hepatitis. Cancer Res., 54,
3171-3172.
WWW.ESAINC.COM
298
Shimoi, K., Kasai, H., Yokota, N., Toyokuni, S., and Kinae, N. (2002). Comparison between high-performance liquid
chromatography and enzyme-linked immunosorbent assay for the determination of 8-hydroxy-2’-deoxyguanosine in
human urine. Cancer Epidemiol. Biomark. Prev., 11, 767-770.
Shimokawa, T., Kulmacz, R.J., DeWitt, D.L., and Smith, W.L. (1990). Tyrosine 385 of prostaglandin endoperoxide
synthase is required for cyclo-oxygenase catalysis. J. Biol. Chem., 265, 20073-20076.
Shin, C.S., Moon, B.S., Park, K.S., Kim, S.Y., Park, S.J., Ching, M.H., and Lee, H.K. (2001). Serum 8-hydroxy-guanine
levels are increased in diabetic patients. Diabetes Care, 24, 733-737.
Shirname-More, L., Rossman, T.G., Troll, W., Teebor, T.W., and Frenkel, K. (1987). Genetic effects of 5-hydroxymethyl2’-deoxyuridine, a product of ionizing radiation. Mutat. Res., 178, 177-186.
Shuker, D.E., Prevost, V., Friesen, M.D., Lin, D., Ohshima, H., and Bartsch, H. (1993). Urinary markers for measuring
exposure to endogenous and exogenous alkylating agents and precursors. Environ. Health Perspect., 99, 33-37.
Schwarz, E., Lilie, H., and Rudolph, R. (1996). The effect of molecular chaperones on in vivo and in vitro folding
processes. Biol. Chem., 377, 411-416.
Simic, M.G., and Bergtold, D.S. (1991). Dietary modulation of DNA damage in humans. Mut. Res., 250, 17-24.
Simons, A.M., Phillips, D.H., and Coleman, D.V. (1993). Damage to DNA in cervical epithelium related to smoking
tobacco. Brit. Med. J., 306, 1444-1448.
Simpson, J.A., Narita, S., Gieseg, G., Gebicki, S., Gebicki, J.M., and Dean, R.T. (1992). Long-lived reactive species on
free-radical-damaged proteins. Biochem. J., 282, 621-624.
Singer, B., and Hang, B. (1998). Role of chemical structure in determination of repair enzyme substrate specificity and
mechanism. In: DNA and Free Radicals: Techniques, Mechanisms and Applications. Aruoma, O.I., and Halliwell, B.
(Eds). OICA, London. Pp. 26-61.
Singer, I.I., Kawka, D.W., Scott, S., Weidner, J.R., Mumford, R.A., Riehl, T.E., and Stenson, W.F. (1996). Expression of
inducible nitric oxide synthase and nitrotyrosine in colonic epithelium in inflammatory disease. Gastroent., 111, 871885.
Situnayake, R.D., Crump, B.J., Zezulka, A.V., Davis, M., McConkey, B., and Thurnham, D.I. (1990). Measurement of
conjugated diene lipids by derivative spectroscopy in heptane extracts of plasma. Ann. Clin. Chem., 27, 258-266.
Skinner, K.A., Crow, J.P., Skinner, H.B., Chandler, R.T., Thompson, J.A., and Parks, D.A. (1997). Free and proteinassociated nitrotyrosine formation following rat liver preservation and transplantation. Arch. Biochem. Biophys., 342,
282-288.
Smith, C.D., Carney, J.M., Starke-Reed, P.E, Olive, C.N., Stadtman, E.R., Floyd, R.A., and Markesbery, W.R. (1991).
Excess brain protein oxidation and enzyme dysfunction in normal aging and Alzheimer’s disease. Proc. Natl. Acad. Sci.
USA, 88, 10540-10543.
Smith, L.L., and Johnson, B.H. (1989). Biological activities of oxysterols. Free Radic. Biol. Med., 7, 285-332.
Smith, M.A., Sayre, L.M., Monnier, V.M., and Perry, G. (1995). Radical AGEing in Alzheimer’s disease. Trends Neurosci.,
18, 172-176.
Smith, M.A., Richey Harris, P.L., Sayre, L.M., Beckman, J.S., and Perry, G. (1997). Widespread peroxynitrite-mediated
damage in Alzheimer’s disease. J. Neurosci., 17, 2653-2657.
Smith, M.A., Taneda, S., Richey, P.L., Miyata, S., Yan, S.D., Stern, D., Sayre, L.M., Monnier, V.M., and Perry, G. (1994).
Advanced Maillard reaction end products are associated with Alzheimer’s disease pathology. Proc. Natl. Acad. Sci.
USA, 91, 5710-5714.
Sodum, R.S., and Fiala, E.S. (2001). Analysis of peroxynitrite reactions with guanine, xanthine, and adenine nucleosides
by high-pressure liquid chromatography with electrochemical detection: C-8nitration and –oxidation, Chem. Res.
Toxicol., 14, 438-450.
Sohal, R.S., Agarwal, A., Agarwal, S., and Orr, W.C. (1995). Simultaneous overexpression of copper- and zinc-containing
superoxide dismutase and catalase retards age-related oxidative damage and increases metabolic potential in
Drosophila melanogaster. J. Biol. Chem., 270, 15671-15674.
Sokolovsky, M., Riordan, J.F., and Vallee, B.L. (1967). Conversion of 3-nitrotyrosine to 3-aminotyrsosine in peptides and
proteins. Biochem. Biophys. Res. Comm., 27, 20-25.
Sontag, G., Bernweiser, I., and Krach, C. (1997). HPLC with electrode array detection and its application in analytical food
chemistry. In: Coulometric Electrode Array Detectors for HPLC. Progress in HPLC-HPCE (6), Acworth, I.N., Naoi, M.,
Parvez, S., and Parvez, H. (Eds.). VSP Press, The Netherlands. Pp. 75-126.
Soussi, T. (1996). The p53 tumor suppressor gene: A model for molecular epidemiology of human cancer. Mol. Med.
Today, 2, 32-37.
Souza, J.M., and Radi, R. (1998). Glyceraldehyde-3-phosphate dehydrogenase inactivation by peroxynitrite. Arch.
Biochem. Biophys., 360, 187-194.
Spencer, J.P.E., Jenner, A., Aruoma, O.I., Evans, P.J., Kaur, H., Dexter, D.T., Jenner, P., Lees, A.J., Marsden, D.C., and
Halliwell, B. (1994). Intense oxidative DNA damage promoted by L-DOPA and its metabolites: Implications for
neurodegenerative disease. FEBS Letts., 353, 246-250.
Spencer, J.P.E., Wong, J., Jenner, A., Aruoma, O.I., Cross, C.E., and Halliwell, B. (1996). Base modification and strand
breakage in isolated calf thymus DNA and in DNA from human skin epidermal keratinocytes exposed to peroxynitrite or
3-morpholinosydnonimine. Chem. Res. Toxicol., 9, 1152-1158.
Spickett, C.M. (2003). Analysis Of Phospholipid Oxidation By Electrospray Mass Spectrometry. In: Critical Reviews of
Oxidative Stress and Aging. Volume 1.Cutler, R.G., and Rodriguez, H. (Eds.). World Scientific, NJ. Pp. 393-409.
Stachowiak, O., Dolder, M., Wallimann, T., and Richter, C. (1998). Mitochondrial creatine kinase is a prime target of
peroxynitrite-induced modification and inactivation. J. Biol. Chem., 273, 16694-16699.
Stadtman, E.R. (1988). Protein modification in aging. J. Gerentol., 43, B112-120.
Stadtman, E.R. (1990). Metal ion-catalyzed oxidation of proteins: Biochemical mechanisms and biological consequences.
Free Radic. Biol. Med., 9, 315-325.
Stadtman, E. (1991). Ascorbic acid and oxidative damage of proteins. Am. J. Clin. Nutr., 54, 1125S-1128S.
Stadtman, E. (1992). Protein oxidation and aging. Science, 257, 1220-1224.
WWW.ESAINC.COM
299
Stadtman, E. (1993). Oxidation of free amino acids and amino acid residues in proteins by radiolysis and by metalcatalyzed reactions. Ann. Rev. Biochem., 62, 797-821.
Stadtman, E.R., and Berlett, B.S. (1991). Fenton chemistry: Amino acid oxidation. J. Biol. Chem., 266, 17202-17211.
Stadtman, E., Starke-Reed, P.E., Oliver, C.N., Carney, J.M., and Floyd, R.A. (1992). Protein modification in aging. EXS.
62, 64-72.
Strickland, E, Qu, B.H., Millen, L., and Thomas, P.J. (1997). The molecular chaperone Hsc70 assists the in vitro folding of
the N-terminal nucleotide-binding domain of the cystic fibrosis transmembrane conductance regulator. J. Biol. Chem.,
272, 25421-25424.
Su, J.H., Deng, G., and Cotman, C.W. (1997). Neuronal DNA damage precedes tangle formation and is associated with
up-regulation of nitrotyrosine in Alzheimer’s disease brain. Brain Res., 774, 193-199.
Suarez-Pinzon, W.L., Szabo, C., and Rabinovitch, A. (1997). Development of autoimmune diabetes in NOD mice is
associated with the formation of peroxynitrite in islet beta-cells. Diabetes, 46, 907-911.
Suematsu, T., Kamada, T., Abe, H., Kikichi, S., and Yagi, K. (1977). Serum lipoperoxide level in patients suffering from
liver diseases. Clin. Chim. Acta, 79, 267-270.
Suter, M., and Rachter, C. (1999). Fragmented mitochondrial DNA is the predominant carrier of oxidized DNA bases.
Biochem., 38, 459-464.
Suzuki, J., Inoue, Y., and Suzuki, S. (1995). Changes in the urinary excretion level of 8-hydroxyguanine by exposure to
reactive oxygen-generating substances. Free Radic. Biol. Med., 18, 431-436.
Szabo, C. (1996). DNA strand breakage and activation of poly-ADP ribosyltransferase: A cytotoxic pathway triggered by
peroxynitrite. Free Radic. Biol. Med., 21, 855-869.
Szweda, L.I., and Stadtman, E.R. (1992). Iron-catalyzed oxidative modification of glucose-6-phosphate dehydrogenase
from L. mesenteroides. Structural and functional changes. J. Biol. Chem., 267, 3096-3100.
Taborsky, G. (1973). Oxidative modifications of proteins in the presence of ferrous ion and air. Effect of ionic composition
of the reaction medium on the nature of the oxidation products. Biochem., 12, 1341-1348.
Tabrizi-Fard, M.A., Maurer, T.S., and Fung, H.-L. (1999). In vivo disposition of 3-nitro-L-tyrosine in rats: Implications on
tracking systemic peroxynitrite exposure. Drug Met. Disp., 27, 429-431.
Taddei, F., Hayakawa, H., Bouton, M.-F., Cirinesi, A.-M., Matic, I., Sekiguchi, M., and Radman, M. (1997). Counteractions
by MuT protein of transcriptional errors caused by oxidative damage. Science, 278, 128-130.
Tagesson, C., Kallberg, M., Klintenberg, C., and Startkhammer, H. (1995). Determination of urinary 8hydroxydeoxyguanosine by automated coupled-column high performance liquid chromatography: A powerful technique
for assaying in vivo oxidative DNA damage in cancer patients. Eur. J. Cancer., 31A, 934-940.
Tagesson, C., Kallberg, M., and Leanderson, P. (1992). Determination of urinary 8-hydroxydeoxyguanosine by coupledcolumn high-performance liquid chromatography with electrochemical detection: A non-invasive assay for in vivo
oxidative DNA damage in humans. Toxicol. Meth., 1, 242-251.
Tajiri, T., Maki, H., and Sekiguchi, M. (1995). Functional cooperation of MutT, MutM and MutY proteins in preventing
mutations caused by spontaneous oxidation of guanine nucleotide in E. coli. Mutat. Res., 336, 257-267.
Takahashi, R., and Goto, S. (1990). Alteration of aminoacyl-tRNA synthetases with age: Heat-labilization of the enzyme
by oxidative damage. Arch. Biochem. Biophys., 277, 228-233.
Takeuchi, T., and Morimoto, K. (1993). Increased formation of 8-hydroxy-deoxyguanosine, an oxidative DNA damage, in
lymphoblasts from Fanconi’s anemia patients due to possible catalase deficiency. Carcinogen., 14, 1115-1120.
Takeuchi, T., Nakajima, M., Ohta, Y., Mure, K., Takeshita, T., and Morimoto, K. (1994). Evaluation of 8hydroxydeoxyguanosine, a typical oxidative DNA damage, in human leukocytes. Carcinogen., 15, 1519-1523.
Tan, B.H., Bencsath, F.A., and Gaubatz, J.W. (1990). Steady-state levels of 7-methylguanine increase in nuclear DNA of
postmitotic mouse tissues during aging. Mut. Res., 237, 229-238.
Tanaka, K. (1998). Molecular biology of the proteasome. Biochem. Biophys. Res. Commun., 247, 537-541.
Tappel, A.L. (1979). Measurement of and protection from in vivo lipid peroxidation. In: Biochemical And Clinical Aspects of
Oxygen. Caughey, W.S. (Ed.). Academic Press, New York. Pp. 137-154.
Taylor, A., Lipman, R.D., Jahngen-Hodge, J., Palmer, V., Smith, D., Padhye, N., Dallal, G.E., Cyr, D.E., Laxman, E., and
Shepard, D. (1995). Dietary calorie restriction in the Emory mouse: Effects of life span, eye lens cataract prevalence,
and progression, levels of ascorbate, glutathione, glucose, and glycohemoglobin, tail collagen break time, DNA and
RNA oxidation, skin integrity, fecundity, and cancer. Mech. Aging Dev., 79, 33-57.
Teixeira, A., Ferreira, M., van Dijk, W., van de Werken, G., and de Jong, A. (1995). Analysis of 8-hydroxy2’deoxyguanosine in rat urine and liver DNA by stable isotope dilution gas chromatography/mass spectrometry. Anal.
Biochem., 226, 307-319.
Ter Steeg, J.C.A., Koster-Kamphuis, L., Van Straatan, E.A., Forget, P.P., and Buurman, W.A. (1998). Nitrotyrosine in
plasma of celiac disease patients as detected by a new sandwich ELISA. Free Radic. Bio. Med., 25, 953-963.
Thiffault, C., Aumont, N., Quirion, R., and Poirier, J. (1995). Effect of MPTP and L-deprenyl on antioxidant enzymes and
lipid peroxidation levels in mouse brain. J. Neurochem., 65, 2725-2733.
Thom, S.R., Xu, Y.A., and Ischiropoulos, H. (1997). Vascular endothelial cells generate peroxynitrite in response to
carbon monoxide exposure. Chem. Res. Toxicol., 10, 1023-1031.
Thoma, I., Loeffler, C., Sinha, A.K., Gupta, M., Krischke, M., Steffan, B, and Mueller, M.J. (2003). Cyclopentanone
isoprostanes induced by reactive oxygen species trigger defense gene activation and phytoalexin accumulation in
plants. Plant J., 34, 363-375.
Thomas, J.P., Maiorino, M., Ursini, F., and Girotti, A.W. (1990a). Protective action of phospholipid hydroperoxide
glutathione peroxidase against membrane-damaging lipid peroxidation. J. Biol. Chem., 265, 454-461.
Thomas, J.P., Geiger, P.G., Maiorino, M., Ursini, F., and Girotti, A.W. (1990b). Enzymatic reduction of phospholipid and
cholesterol hydroperoxides in artificial bilayers and lipoproteins. Biochim. Biophys. Acta, 1045, 252-260.
Thomas, P.J., Qu, B.H., and Pedersen, P.L. (1995). Defective protein folding as a basis of human disease. Trends
Biochem. Sci., 20, 456-459.
WWW.ESAINC.COM
300
Thomas, R.M., Nauseef, W.M., Lyer, S.S., Peterson, M.W., Stone, P.J., and Clark, R.A. (1991). A cytosolic inhibitor of
human neutrophil elastase and cathepsin G. J. Leukocyte Biol., 50, 568-579.
Tohgi, H., Abe, T., Yamazaki, K., Murata, T., Ishizaki, E., and Isobe, C. (1999a). Remarkable increase in cerebrospinal
fluid 3-nitrotyrosine in patients with sporadic amyotrophic lateral sclerosis. Ann. Neurol., 46, 129-131.
Tohgi, H., Abe, T., Yamazaki, K., Murata, T., Ishizaki, E., and Isobe, C. (1999b). Alterations of 3-nitrotyrosine
concentrations in the cerebrospinal fluid during aging and in patients with Alzheimer’s disease. Neurosci. Letts., 269,
52-54.
Tomita, M., Okuyama, T., Hatta, Y., and Kawaj, S. (1990). Determination of free malondialdehyde by gas chromatography
with an electron-capture detector. J. Chromatogr., 526, 174-179.
Toraason, M., Hayden, C., Marlow, D., Rinehart, R., Mathias, P., Werren, D., DeBord, D.G., and Reid, T.M. (2001). DNA
strand breaks, oxidative damage, and 1-OH pyrene in roofers with coal-tar pitch dust and/or asphalt fume exposure. Int.
Arch. Environ. Health, 74, 396-404.
Trelstad, R.L., Lawley, K.R., and Holmes, L.B. (1981). Nonenzymatic hydroxylations of proline and lysine by reduced
oxygen derivatives. Nature, 289, 310-322.
Tsikas, D., Schwedhelm, E., Fauler, J., Gutzki, F., Mayatepek, E., and Frolich, J.C. (1998). Specific and rapid
quantification of 8-iso-prostaglandin F2α in urine of healthy humans and patients with Zellweger syndrome by GC-MS. J.
Chromatogr., B, 716, 7-17.
Tsikas, D., Schwedhelm, E., Suchy, M.T., Niemann, J., Gutzki, F.M., Erpenbecl, V.J., Hohfield, J.M., Surdacki, A., and
Frolich, J.C. (2003). Divergence in urinary 8-iso-PGF(2alpha) levels. Possible methodological, mechanistic and clinical
implications. J. Chromatrogr. B, 794, 237-255.
Uchida, K., and Kawakishi, S. (1989). Bioorg. Chem., 17, 330-343.
Uchida, K., and Kawakishi, S. (1990). Site-specific oxidation of angiotensin 1 by copper(II) and L-ascorbate: Conversion of
histidine into 2-imidazolones. Arch. Biochem. Biophys., 283, 20-26.
Uchida, K., and Kawakishi, S. (1993). 2-Oxohistidine as a novel biological marker for oxidatively modified proteins. FEBS
Lett., 332, 208-210.
Uchida, K., and Stadtman, E.R. (1992). Modification of histidine residues in proteins by reaction with 4- hydroxynonenal.
Proc. Natl. Acad. Sci. USA, 89, 4544-4548.
Uchida, K., and Stadtman, E.R. (1994). Quantitation of 4-hydroxynonenal protein adducts. Meth. Enzymol., 233, 371-382.
Uchida, K., Kanematsu, M., Mortimitsu, Y., Osawa, T., Noguchi, N., and Niki, E. (1999a). Acrolein is a product of lipid
peroxidation: Formation of free acrolein and its conjugate with lysine residues in oxidized low density lipoproteins. J.
Biol. Chem., 273, 16058-16066.
Uchida, K., Kanematsu, M., Sakai, K., Matsuda, T., Yasuda, Y., Miyata, T., Noguchi, N., Niki, E., and Osawa, T. (1998b).
Protein-bound acrolein: Potential markers for oxidative stress. Proc. Natl. Acad. Sci. USA, 95, 4882-4887.
Uchida, K., Kato, Y., and Kawakishi, S. (1990). A novel mechanism for oxidative cleavage of prolyl peptides induced by
the hydroxyl radical. Biochem. Biophys. Res. Commun., 169, 265-271.
Uchida, K., Sakai, K., Itakura, K., Osawa, T., Toyokuni, S. (1997). Protein modification by lipid peroxidation products:
Formation of malondialdehyde-derived N(epsilon)-(2-propenol)lysine in proteins. Arch. Biochem. Biophys., 346, 45-52.
Ueta, K., and Hayashi, O. (1985). ADP-ribosylation. Ann. Rev. Biochem., 54, 73-100.
Ullrich, O., Reinheckel T., Sitte, N., and Grune, T. (1999). Degradation of hypochlorite-damaged glucose-6-phosphate
dehydrogenase by the 20S proteosome. Free Radic. Biol. Med., 27, 487-492.
Ullrich, O., Sitte, N., Sommerburg, O., Sandig, V., Davies, K.J., and Grune, T. (1999). Influence of DNA binding on the
degradation of oxidized histones by the 20S proteosome. Arch. Biochem. Biophys., 362, 211-216.
Ursini, F., and Bindoli, A. (1984). The role of selenium peroxidases in the protection against oxidative damage of
membranes. Chem. Phys. Lipids, 44, 255-276.
Ursini, F., Maiorino, M., Gregolin, C. (1985). The selenoenzyme phospholipid hydroperoxide glutathione peroxidase.
Biochim. Biophys. Acta, 839, 62-70.
Ursini, F., Zamburlini, A., Cazzolato, G., Maiorino, M., Bon, G.B., and Sevanian, A. (1998). Postprandial plasma lipid
hydroperoxides: A possible link between diet and atherosclerosis. Free Radic. Biol. Med., 25, 250-252.
Uttenthal, L.O., Alonso, D., Fernandez, A.P., Campbell R.O., Moro, M.A., Leza, J.C., and Rodrigo, J. (1998). Neural and
inducible nitric oxide synthase and nitrotyrosine immunoreactivities in the cerebral cortex of the aging rat. Microsc. Res.
Tech., 43, 75-88.
32
Vaca, C.E., Fang, J.L., Mutanen, M., and Valsta, L. (1995). P-postlabelling determination of DNA adducts of
malonaldehyde in humans: Total white blood cells and breast tissue. Carcinogen., 16, 1847-1851.
Van Buskirk, J.J., Kirsch, W.M., Kleyer, D.L., Darklen, R.M., and Koch, T.H. (1984). Aminomalonic acid: Identification in E.
coli and atherosclerotic plaque. Proc. Natl. Acad. Sci. USA, 81, 722-725.
Van Delft, J.H.M., van Winden, M.J.M., van den Ende, A.M.C., and Baan, R.A. (1993). Determining N7-alkylguanine
adducts by immunochemical methods and HPLC with electrochemical detection. Applications in animal studies and in
monitoring human exposure to alkylating agents. Environ. Health Perspect., 99, 25-32.
Van der Vliet, A., Eiserich, J.P., Halliwell, B. and Cross, C.E. (1997). Formation of reactive nitrogen species during
peroxidase-catalyzed oxidation of nitrite. A potential additional mechanism of nitric oxide-dependent toxicity. J. Biol.
Chem., 272, 7617-7625.
Van der Vliet, A., Eiserich, J.P., Kaur, H., Cross, C.E., and Halliwell, B. (1996). Nitrotyrosine as biomarker for reactive
nitrogen species. Meth. Enzymol., 269, 175-184.
Van der Vliet, A., Eiserich, J.P., O’Niell, C.A., Halliwell, B., and Cross, C.E. (1995). Tyrosine modification by reactive
nitrogen species: A closer look. Arch. Biochem. Biophys., 319, 341-349.
Van der Vleit, A., Jenner, A., Eiserich, J.P., Cross, C.E., and Halliwell, B. (1999). Analysis of aromatic nitration,
chlorination, and hydroxylation by gas chromatography-mass spectrometry. Meth. Enzymol., 301, 471-483.
Van der Vliet, A., O’Neill, C.A.O., Halliwell, B., Cross, C.E., and Kaur, H. (1994). Aromatic hydroxylation and nitration of
phenylalanine and tyrosine by peroxynitrite. FEBS Letts., 339, 89-92.
WWW.ESAINC.COM
301
Van den Akker, E., Lutgerink, J.T., Lafleur, M.V.M., Joenje, H., and Retel, J. (1994). The formation of one-G deletions as
a consequence of single-oxygen-induced DNA damage. Mutat. Res., 309, 45-52.
Vane, J.R., Mitchell, J.A., Appleton, I., Tomlinson, A., Bishop-Bailey, D., Croxtall, J., and Willoghby, D.A. (1994). Inducible
isoforms of cyclo-oxygenase and nitric oxide synthase in inflammation. Proc. Natl. Acad. Sci. USA, 91, 2046-2050.
Van Houten, B. (1990). Nucleotide excision repair in Escherichia coli. Microbiol. Rev., 54, 18-51.
Van Houten, B. (1998). Damage and Repair of Mitochondrial DNA. Oxygen Society, Washington DC.
Van Kuijk, F.J., (1987). Sevanian, A., Handelman, G.J., and Dratz, E.A. (1987). A new role for phospholipase A2:
Protection of membranes from lipid peroxidation damage. Trends Biochem. Sci., 12, 31-34.
Van Kuijk, F.J., (1988). Thesis. See Esterbauer, H., Schaur, R.J., and Zollner, H. (1991). Chemistry and biochemistry of 4hydroxynonenal, malondialdehyde and related aldehydes. Free Radic. Biol. Med., 11, 81-128.
Van Kuijk, F.J., and Dratz, E.A. (1987). Detection of phospholipid peroxides in biological samples. Free Radic. Biol. Med.,
3, 349-354.
Van Kuijk, F.J., Holte, L.L., and Dratz, E.A. (1990). 4-Hydroxyhexenal: A lipid peroxidation product derived from oxidized
docosahexaenoic acid. Biochim. Biophys. Acta, 1043, 116-118.
van Poppel, G., Poulsen, H.E., Loft, S., and Verhagen, H. (1995). No influence of beta-carotene on oxidative DNA
damage in male smokers. J. Natl. Acad. Sci. USA, 90, 310-311.
Varshavsky, A. (1997). The N-end rule pathway of protein degradation. Genes Cells, 2, 13-28.
Vassalle, C., Botoo, N., Andreassi, M.G., Berti, S., and Biagini, A. (2003). Evidence for enhanced 8-isoprostane plasma
levels, as index of oxidative stress in vivo, in patients with coronary heart disease. Coron. Artery Dis., 14, 213-218.
Vatassery, G.T., Quach, H.T., Smith, W.E., Krick, T.P., and Ungar, F. (1997). Analysis of hydroxy and keto cholesterols in
oxidized brain synaptosomes. Lipids, 32, 101-107.
Verhagen, H., Poulsen, H.E., Loft, S., van Poppel, G., Willems, M.I., and van Bladeren, P.J. (1995). Reduction of oxidative
DNA-damage in humans by brussel sprouts. Carcinogen., 16, 969-970.
Vernia, S, Sanz-Gonzalez, S.M., and Lopez-Garcia, M.P. (2001). Involvement of peroxynitrite on early loss of p450 in
short-term hepatocytes cultures. Adv. Exp. Med. Biol., 500, 209-212.
Verweij, H., Christianse, K., and Van Steveninck, J. (1982). Ozone-induced formation of O,O’-dityrosine cross-links in
proteins. Biochem. Biophys. Acta, 701, 180-184.
Viera, L., Zu, Y., Estevez, A.G., and Beckman, J.S. (1999). Immunohistochemical methods to detect nitrotyrosine. Meth.
Enzymol., 301, 373-381.
Vigue, C.A., Frei, B., Shigenaga, M.K., Ames, B.N., Packer, L., and Brooks, G.A. (1993). Antioxidant stress and indexes
of oxidative stress during consecutive days of exercise. J. Appl. Physiol., 75: 566-573.
Viinikka, L., Vuori, J., and Ylikorkala, O. (1984). Lipid peroxides, prostacyclin and thromboxane A2 in runners during acute
exercise. Med. Sci. Sport Exer., 16, 275-277.
Viner, R.I., Ferrington, D.A., Williams, T.D., Bigelow, D.J., and Schoneich, C. (1999). Protein modification during biological
2+
aging: Selective tyrosine nitration of the SERCA2a isoform of sarcoplasmic reticulum Ca -ATPase in skeletal muscle.
Biochem. J., 340, 657-669.
Viner, R.I., Huhmer, A.F., Bigelow, D.J., and Schoneich, C. (1996). The oxidative inactivation of sarcoplasmic reticulum
2+
Ca -ATPase by peroxynitrite. Free Radic. Res., 24, 243-259.
Visick, J.E., and Clarke, S. (1995). Repair, refold, recylce: How bacteria can deal with spontaneous environmental
damage to proteins. Mol. Microbiol., 16, 835-845.
Vissers, M.C., and Winterbourne, C.C. (1991). Oxidative damage to fibronectin. II. The effects of H2O2 and the hydroxyl
radical. Arch. Biochem. Biophys., 285, 357-364.
Vogt, W. (1995). Oxidation of methionyl residues in proteins: Tools, targets, and reversal. Free Radic. Biol. Med., 18, 6173.
Von Sonntag, C. (1990). In: Sulfur-centered Reactive Intermediates in Chemistry and Biology. Chatgilialoglu, C., and
Asmus, K.D. (Eds.). Plenum Press, New York. Pp. 359-366.
Von Sonntag, C. (1991). The chemistry of free-radical-mediated DNA damage. In: Physical and Chemical Mechanisms in
Molecular Radiation Biology. Glass, W.A., and Varma, M.N. (Eds.). Plenum Press, New York. Pp. 287-321.
Waeg, G., Dimsity, G., and Esterbauer, H. (1996). Monoclonal antibodies for detection of 4-hydroxynonenal modified
proteins. Free Rad. Res., 25, 149-159.
Wagner, J.R., Hu, C.-C., and Ames, B.N. (1992). Endogenous oxidative damage of deoxycytidine in DNA. Proc. Natl.
Acad. Sci. USA, 89, 3380-3384.
Wallace, D.C. (1992). Diseases of the mitochondrial DNA. Ann. Rev. Biochem., 61, 1175-1212.
Wang, M.L., and Liehr, J.G. (1991). Detection of DNA adducts of unsaturated fatty acids (UF) hydroperoxides (UF-OOH)
32
by P-postlabeling analysis. Proc. Ann. Meet. Am. Assoc. Cancer Res., 32, A864.
Wang, Z.Q., Auer, B., Stingl, L., Berghammer, H., Haidacher, D., Schweiger, M., and Wagner, E.F. (1995). Mice lacking
ADPRT and poly(ADP-ribosyl)ation develop normally but are susceptible to skin disease. Genes Dev., 9, 510-520.
Wani, G., and D’Ambrosia, S.M. (1995). Cell type-specific expression of human 8-oxo-7,8-dihydroguanosine triphosphate
in normal breast and skin tissues in vivo. Carcinogen., 16, 277-283.
Ward, J.F., Blakely, W.F., and Joner, E.I. (1985). Mammalian cells are not killed by DNA single-strand breaks caused by
hydroxyl radicals from hydrogen peroxide. Radiat. Res., 103, 383-392.
Waterfall, A.H., Singh, G., Fry, J.R., and Marsden, C.A. (1996). Acute acidosis elevates malondialdehyde in rat brain in
vivo. Brain Res., 712, 102-106.
Wei, H., Cai, Q., Rahn, R., and Zhang, X. (1997). Singlet oxygen involvement in ultraviolet (254nm) radiation-induced
formation of 8-hydroxy-deoxyguanosine in DNA. Free Radic. Biol. Med., 23, 148-154.
Wei, Y.H. (1998). Mitochondrial DNA mutations and oxidative damage in aging and diseases: An emerging paradigm of
gerontology and medicine. Proc. Natl. Acad. Sci Counc. Repub. China, B., 22, 55-67.
Weiland, E., Schettler, V., Diedrich, F., Schuff-Werner, P., and Oellerich, M. (1992). Determination of lipid hydroperoxides
in serum iodometry and high performance liquid chromatography compared. Eur. J. Clein. Chem. Clin. Biochem., 30,
363-369.
WWW.ESAINC.COM
302
Weimann, A., Belling, D., and Poulsen, H.E. (2001). Measurement of 8-oxo-2’-deoxyguanosine and 8-oxo-2’deoxyadenosine in DNA and human urine by high performance liquid chromatography-electrospray tandem mass
spectrometry. Free Radic. Biol. Med., 30, 757-764.
Weinberg, R.A. (Ed.). (1989). Oncogenes and The Molecular Origins of Cancer. Cold Spring Harbor Press, New York.
Weitzman, S.A., Turk, P.W., Milkowaki, D.H., Kozlowski, K. (1994). Free radical adducts induce alterations in DNA
cytosine methylation. Proc. Natl. Acad. Sci. USA, 91, 1261-1264.
Wells-Knecht, M.C., Huggins, T.G., Dyer, D.G., Thorpe, S.R., and Baynes, J.W. (1993). Oxidized amino acids in lens
protein with age. Measurement of o-tyrosine and dityrosine in the aging human lens. J. Biol. Chem., 268, 12348-12352.
Wells-Knecht, M.C., Zyzak, D.V., Litchfield, J.E., Thorpe, S.R., and Baynes, J.W. (1995). Mechanism of autoxidative
glycosylation: Identification of glyoxal and arabinose as intermediates in the autoxidative modification of proteins by
glucose. Biochem., 34, 3702-3709.
Wendt, T., Tanji, N., Guo, J., Hudson, B.I., Bierhaus, A., Ramasamy, R., Arnold, B., Nawroth, P.P., D’Agati, V., and
Scmidt, A.M. (2003). Glucose, glycation and RAGE: Implications for amplification of cellular dysfunction in diabetic
nephropathy. J. Am. Soc. Nephrol., 14, 1383-1395.
Westhuyzen, J. (1997). The oxidation hypothesis of atherosclerosis: An update. Ann. Clin. Lab. Sci., 27, 1-10.
White, C.R., Brock, T.A., Chang, L.Y., Crapo, J., Briscoe, P., Ku, D., Bradley, W.A., Gianturco, S.H., Gore, J., and
Freeman, B.A. (1994). Superoxide and peroxynitrite in atherosclerosis. Proc. Natl. Acad. Sci. USA, 91, 1044-1048.
Whiteman, M., and Halliwell, B. (1999b). Loss of 3-nitrotyrosine on exposure to hypochlorous acid: Implications for the
use of 3-nitrotyrosine as a biomarker in vivo. Biochem. Biophys. Res. Commun., 258, 168-172.
Whiteman, M., Hong, H.S., Jenner, A., and Halliwell, B. (2002). Loss of oxidized and chlorinated bases in DNA treated
with ROS. Biochem. Biophys. Res. Commun., 296, 883-889.
Whiteman, M., Jenner, A., and Halliwell, B. (1997). Hypochlorous acid –induced base modifications in isolated calf thymus
DNA. Chem. Res. Toxicol., 10, 1240-1246.
Whiteman, M., Spencer, J.P., Jenner, A., and Halliwell, B. (1999a). Hypochlorous acid-induced DNA base modification:
Potentiation by nitrite: Biomarkers of DNA damage by reactive oxygen species. Biochem. Biophys. Res. Commun.,
257, 572-576.
Wicric, E.J.M.V., Leeuwen, R.E.W., van Hendricks, H.F.J., Verhagen, H., Loft, H., and Poulsen, H.E., van der Berg, H.
(1995). Moderate energy restriction does not affect parameters of oxidative stress and genotoxicity in man. J. Nutr.,
125, 2631-2639.
Wilkinson, K.D. (1997). Regulation of ubiquitin-dependent processes by deubiquitinating enzymes. FASEB J., 11, 12451256.
Wilson, V.L., Taffe, B.G., Shields, P.G., Powey, A.C., and Harris, C.C. (1993). Detection and quantification of 8hydroxydeoxyguanosine adducts in peripheral blood of people exposed to ionizing radiation. Environ. Health Prosp.,
99, 261-263.
Winterbourne, C.C., and Buss, I.H. (1998), Protein carbonyl measurement by an enzyme-linked immunosorbent assay.
Meth. Enzymol., 300, 106-111.
Winterbourn, C.C., Pichorner, H., and Kettle, A.J. (1997). Myeloperoxidase-dependent generation of a tyrosine peroxide
by neutrophils. Arch. Biochem. Biophys., 338, 15-21.
Wiseman, H., and Halliwell, B. (1996). Damage to DNA by reactive oxygen and nitrogen species: Role in inflammatory
disease and progression to cancer. Biochem. J., 313, 17-29.
Wishnok, J.S., Tannedbaum, S.R., Stillwell, W.G., Glogowski, J.A., and Leaf, C.D. (1993). Urinary markers for exposure
to alkylating or nitrosating agents. Environ. Health Perspect., 99, 155-159.
Witherell, H.L., Hiatt, R.A., Replogle, M., and Parsonnet, J. (1998). Helicobacter pylori infection and urinary excretion of 8hydroxy-2-deoxyguanosine, and oxidative DNA adduct. Cancer Epidemiol. Biomarkers Prev., 7, 91-96.
Witt, E.H., Reznick, A.Z., Viguie, C.A., Starke-Reed, P., and Packer, L. (1992). Exercise, oxidative damage and effects of
antioxidant manipulation. J. Nutr., 122, 766-773.
Wizemann, T.M., Gardner, C.R., Laskin, J.D., Quinones, S., Durham, S.K., Goller, N.L., Ohnishi, S.T., and Laskin, D.L.
(1994). Production of nitric oxide and peroxynitrite in the lung during acute endotoxemia. J. Leukoc. Biol., 56, 759-768.
Wolff, S.P., Jiang, Z.Y., and Hunt, J.V. (1991). Protein glycation and oxidative stress in diabetes mellitus and aging. Free
Radic. Biol. Med., 10, 339-352.
Wong, S.H.Y., Knight, J.A., Hopfer, S.M., Zaharia, O., Leach, C.N., and Sunderman, F.W. (1987). Lipoperoxides in
plasma as measured by liquid-chromatographic separation of malondialdehyde thiobarbituric acid adduct. Clin. Chem.,
33, 214-220.
Wood, M.L., Dizdaroglu, M., Gajewski, E., and Essigmann, J.M. (1990). Mechanistic studies of ionizing radiation and
oxidative mutagenesis: Genetic effects of a single 8-hydroxyguanine (7-hydro-8-oxoguanine) residue inserted at a
unique site in a viral genome. Biochem., 29, 7024-7032.
Wu, G.S., Zhnag, J., and Rao, N.A. (1997). Peroxynitrite and oxidative damage in experimental autoimmune uveitis.
Invest. Ophthalmol. Vis. Sci., 38, 1333-1339.
Wu, W., Chen, Y., d’Avignon, A., and Hazen, S.L. (1999). 3-Bromotyrosine and 3,5-dibromotyrosine are major products of
protein oxidation by eosinophil peroxidase: Potential markers for eosinophil-dependent tissue injury in vivo. Biochem.,
38, 3538-3548.
Yagi, K. (1998). Simple procedure for specific assay of lipid hydroperoxides in serum and plasma. In: Methods in
Molecular Biology, 108. Armstrong, D. (Ed.). Humana Press, Totowa, New Jersey. Pp. 107-110.
Yakes, F.M., and Van Houten, B. (1997). Mitochondrial DNA damage is more extensive and persists longer than nuclear
DNA damage in human cells following oxidative stress. Proc. Natl. Acad. Sci. USA, 94, 514-519.
Yamada, K., Terao, J., and Matsushita, S. (1987). Electrochemical detection of phospholipid hydroperoxides in reversedphase high-performance liquid chromatography. Lipids, 22, 125-128.
Yamamoto, F., Frei, B., and Ames, B.N. (1990). Assay of lipid hydroperoxides using HPLC with post-column
chemiluminescence detection. Meth. Enzymol., 186, 371-380.
WWW.ESAINC.COM
303
Yamamoto, T., Hosokawa, K., Tamura, T., Kanno, H., Urabe, M., and Honjo, H. (1996). Urinary 8-hydroxy-2’deoxyguanosine (8-OhdG) levels in women with and without gynocological cancer. J. Obstet. Gynaecol. Res., 22, 359363.
Yamamoto, Y. (1994). Chemiluminescence-based high-performance liquid chromatography assay of lipid hydroperoxides.
Meth. Enzymol., 233, 319-324.
Yamamoto, Y., and Ames, B.N. (1987). Detection of lipid hydroperoxides and hydrogen peroxide at picomole levels by an
HPLC and isoluminol chemiluminescence assay. Free Radic. Biol. Med., 3, 359-361.
Yamamoto, Y., Brodsky, M.H., Baker, J.C., and Ames, B.N. (1987). Detection and characterization of lipid hydroperoxides
at picomole levels by high-performance liquid chromatography. Anal. Biochem., 160, 7-13.
Yamamoto, Y., Kambayashi, Y., and Ueda, T. (1998). Assay of phospholipid hydroperoxides by chemiluminescencebased high-performance liquid chromatography. In: Methods in Molecular Biology, 108. Armstrong, D. (Ed.). Humana
Press, Totowa, New Jersey. Pp. 63-70..
Yamim, G., Uversky, V.N., and Fink, A.L. (2003). Nitration inhibits fibrillation of human alpha-synuclein in vitro by
formation of soluble oligomers. FEBS Letts., 542, 147-152.
Yan, S.D., Chen, X., Schmidt, A.M., Brett, J., Godman, G., Zou, Y.S., Scott, C.W., Caputo, C., Frappier, T., Smith, M.A.,
Perry, G., Yen, S.H., and Stern, D. (1994). Glycated tau protein in Alzheimer’s disease: A mechanism for induction of
oxidative stress. Proc. Natl. Acad. Sci. USA, 91, 7787-7791.
Yan, S.D., Schmidt, A.M., Anderson, G.M., Zhang, J., Brett, J., Zou, Y.S., Pinsky, D., and Stern, D. (1994). Enhanced
cellular oxidant stress by the interaction of advanced glycation end products with their receptor/binding proteins. J. Biol.
Chem., 269, 9889-9897.
Yanez, L., Groffen, J., and Valenzuela, D.M. (1987). C-K-ras mutations in human carcinomas occur preferentially in codon
12. Oncogene, 1, 315-318.
Yang, C.C., Alvarez, R.B., Engel, W.K., and Askanas, V. (1996). Increase of nitric oxide synthases and nitrotyrosine in
inclusion-body myositis. Neuroreport, 8, 153-158.
Yang, C.-S., Tsai, P.-J., Wu, J.-P., Lin, N.-N., Chou, S.T., and Kuo, J.-S. (1997). Evaluation of extracellular lipid
peroxidation in brain cortex of anesthetized rats by microdialysis perfusion and high-performance liquid
chromatography with fluorometric detection. J. Chromatogr. B., 693, 257-263.
Yang, M.H., and Schaich, K.M. (1996). Factors affecting DNA damage caused by lipid hydroperoxides and aldehydes.
Free Radic. Biol. Med., 20, 225-236.
Yasaka, T., Ohya, I., Matsumoto, J., Shiramizu, T., and Sasaguri, Y. (1981). Acceleration of lipid peroxidation in human
paraquat poisoning. Arch. Intern. Med., 141, 1169-1173.
Yasmin, W., Strynadka, R., and Schulz, R. (1997). Generation of peroxynitrite contributes to ischemia-reperfusion injury in
isolated rat hearts. Cardiovasc. Res., 33, 422-432.
Yasuda, M., and Narita, S. (1997). Simultaneous determination of phospholipid hydroperoxides and cholesteryl ester
hydroperoxides in human plasma by high-performance liquid chromatography with chemiluminescence detection. J.
Chromatogr. B., 693, 211-217.
Ye, Y.Z., Strong, M., Huang, Z-Q., and Beckman, J.S. (1996). Antibodies that recognize nitrotyrosine. Meth. Enzymol.,
269, 201-209.
Yeo, H.C., Helbock, H.J., Chyu, D.W., and Ames, B.N. (1994). Assay of malondialdehyde in biological fluids by gas
chromatography-mass spectrometry. Anal. Biochem., 220, 391-396.
Yermilov, V., Rubio, J., and Ohshima, H. (1995). Formation of 8-nitroguanine in DNA treated with peroxynitrite in vitro and
its rapid removal from DNA by depurination. FEBS Lett., 376, 207-210.
Yermilov, V., Yoshie, Y., Rubio, J., and Ohshima, H. (1996). Effects of carbon dioxide/bicarbonate on induction of DNA
single-strand breaks and formation of 8-nitroguanine and base-propenal mediated by peroxynitrite. FEBS Lett., 399, 6776.
Yi, D., Smythe, G.A., Blount, B.C., and Duncan, M.W. (1997). Peroxynitrite-mediated nitration of peptides:
Characterization of products by electrospray and combined gas chromatography-mass spectrometry. Arch. Biochem.
Biophys., 344, 253-259.
Yin, B., Whyatt, R.M., Perera, F.P., Randall, M.C., Cooper, T.B., and Santella, R.M. (1995). Determination of 8hydroxydeoxyguanosine by an immunoaffinity chromatography-monoclonal antibody-based ELISA. Free Radic. Biol.
Med., 18, 1023-1032.
Yin, D., and Brunk, U. (1998). Autofluorescent ceroid/lipofuscin. In: Methods in Molecular Biology, 108. Armstrong, D.
(Ed.). Humana Press, Totowa, New Jersey. Pp. 217-227..
Yoshie, Y., and Ahshima, H. (1998). Synergistic induction of DNA strand breakage by catechol-estrogen and nitric oxide:
Implications for hormonal carcinogenesis. Free Radic. Biol. Med., 24, 341-344.
Yoshino, K., Sano, M., Fujita, M., and Tomita, I. (1986). Formation of aliphatic aldehydes in rat plasma and liver due to
vitamin E deficiency. Chem. Pharm. Bull., 34, 5184-5187.
Zarkovic, N. (2003). 4-Hydroxynonenal as a bioactive marker of pathophysiological processes. Mol. Aspects Med., 24,
281-291.
Zhang, C., Reiter, C., Esiserich, J.P., Boersa, B., Parks, D.A., Beckman, J.S., Barnes, S., Kirk, M., Baldus, S., DarleyUsmar, V.M., and White, C.R. (2001). L-Arginine chlorination inhibits endotheilial nitric oxide production. J. Biol. Chem.,
276, 27159-271-65.
Zhang, J.-R., Andrus, P.K., and Hall, E.D. (1994). Age-related phospholipid hydroperoxide levels in gerbil brain measured
by HPLC-chemiluminescence and their relation to hydroxyl radical stress. Brain Res., 639, 275-282.
Zou, M., Martin, C., and Ullrich, V. (1997). Tyrosine nitration as a mechanism of selective inactivation of prostacyclin
synthase by peroxynitrite. Biol. Chem., 378, 707-713.
Zou, M., Yesilkaya, A., and Ullrich, V. (1999). Peroxynitrite inactivates prostacyclin synthase by heme-thiolate-catalyzed
tyrosine nitration. Drug. Metab. Rev., 31, 343-349.
WWW.ESAINC.COM
304
Appendix 3.1
Typical DNA Extraction and Hydrolysis
With Thanks to M.L. Maidt and R. Floyd
(Oklahoma Medical Research Foundation, OK).
DNA Extraction Procedure (non-chaotropic).
I.
Reagents.
Homogenizing buffer:
0.3 M sucrose, 0.025 M TRIS buffer, 0.002 M EDTA.
Adjust pH to 7.2 with 1.0 M HCl. Filter solution through a 0.2 µm Nylon filter into a
sterile, autoclaved bottle. Store at 4oC for up to 6 months.
RES (RNA extraction solution):
1.0 M lithium chloride, 2 M urea, 0.04 M sodium citrate, 0.005 M EDTA, 2% SDS
Adjust to pH 6.8 with 1.0 M HCl. Filter solution through 0.2 µm Nylon filter before
addition of SDS. Store at room temperature for up to 6 months.
DNAse
RNAse
Proteinase K
TE solution: 0.010 M TRIS buffer, 0.001 M EDTA
CIA solution: Chloroform:isoamyl alcohol (24:1)
3 M sodium acetate solution (pH 7.0)
95% ethanol
II.
Procedural Notes.
1.
2.
3.
Adjust pH of solutions with 1.0 M HCl or 1.0 M NaOH.
In order to minimize contamination; wear gloves during all procedures.
Homogenizing solution must be filtered through a 0.2 µm Nylon-66 filter
into a sterile bottle and stored at 4oC for up to 6 months. Discard solution if
bacterial growth is observed.
All bottles should be clean, RNA/DNA free, and stored with caps/covers
on.
4.
WWW.ESAINC.COM
305
III.
Procedure.
1.
Homogenize tissue in homogenizing buffer (at a ratio of 1:4 to 1:5; e.g.,
100 mg tissue to 0.4 - 0.5 mL buffer) for 10 - 15 sec. For softer tissues (i.e.
brain) a Teflon plunger-type homogenizer can be used. For harder tissue
(e.g., muscle, heart, intestine, etc.) a Polytron-type homogenizer can be
used to shred the tissue.
Add an equal volume of RES.
Calculate the amount of RNAse solution that needs to be added to the
sample solution in order to achieve a final concentration of 100 µg/mL.
Heat this RNAse aliquot at 70oC for 10 min to inactivate DNAse. After
heating, add the RNAse to the sample solution to achieve a final
concentration of 100 µg/mL and incubate for 30 min at 50oC.
Add sufficient proteinase K to achieve a concentration of 250 µg/mL and
incubate at 50oC for 60 min.
Add an equal volume of ClA and place tubes on rotary mixer. Mix slowly at
15 – 20 rotations/min for 15 min.
Place tubes in centrifuge and spin for 5 min at 2,000 – 3,000 g.
After centrifuging, the tube will have an aqueous portion on top that
contains the DNA/RNA mixture. The bottom layer contains the ClA. The
interface between the two layers contains proteins. Remove the aqueous
layer into a clean tube, taking care not to remove any of the proteincontaining interface.
Repeat steps 5 through 7 two more times. With each successive
extraction, the protein-containing interface should become visibly smaller.
Add a 1:15 aliquot (e.g. 10 µL to 150 µL sample) of 3M sodium acetate
(pH 7.00) and 2.5x volume of 95% (e.g. 1,250 µL to 500 µL aqueous
sample volume).
Place tube into refrigerator (4oC) for 60 min.
Place tube in centrifuge at 3,000 – 4,000 g for 10 min to spin out the
precipitated DNA.
Evaporate the sample to dryness.
2.
3.
4.
5.
6.
7.
8.
9.
10.
11.
12.
Note – Step 3 is only necessary if an RNA-free sample is desired. This step can
be omitted since the RNA, DNA and adducts that will be extracted can be
separated chromatographically on the HPLC-ECD system.
WWW.ESAINC.COM
306
2. DNA Hydrolysis Procedure.
I.
Reagents.
TE Solution: 0.010 M TRIS buffer, 0.001 M EDTA.
0.5 M Sodium acetate, pH 5.1
1.0 M Magnesium chloride
Nuclease P1: 10 mg/mL prepared in water and stored at 4oC
TRIS base: 1.0 M solution, pH 10.0
Alkaline phosphatase: 1 unit/µL solution
5.8 M acetic acid
II.
Procedure.
1.
2.
3.
4.
If not in solution, solubilize the DNA in 0.25 mL TE solution.
Add 25 µL of 0.5 M sodium acetate (pH 5.1).
Add 2.75 µL of 0.1 M magnesium chloride.
Heat sample in a boiling water bath (100oC) for 5 min to make the DNA
single stranded.
Immediately cool the sample on ice for 5-10 min.
Once cooled, add 10 µL (10 µg) of nuclease P1 from a stock solution
prepared at 1.0 mg/mL in water.
Incubate the sample at 37oC for 1 hr.
Add 8 µL of 1.0M TRIS base (pH 10.0) to adjust to pH 7.8.
Add 2 µL alkaline phosphatase (equivalent to 2 units at stock
concentration of 1 unit/µL).
Incubate sample at 37oC for 1hr.
Precipitate enzyme by adding 4 µL of 5.8 M acetic acid.
Place sample in centrifuge and spin at 3,000-4,000 g for 5 min.
Remove the aqueous portion that contains the hydrolysate and filter this
solution through a 0.2 µm spin filter (5,000 g for 10 min). The sample is
now ready for injection onto the HPLC system.
5.
6.
7.
8.
9.
10.
11.
12.
13.
Note: DNA extractor kits are also available from Wako Fine Chemical Company.
WWW.ESAINC.COM
307
Chapter 4
Protection
Living organisms are constantly being exposed to a variety of pro-oxidant species
capable of damaging many vitally important biomolecules, yet life still thrives.
This is because, during evolution, the cell developed a series of defensive
mechanisms designed to minimize the consequences of pro-oxidant action. We
have already discussed repair and destruction of damaged macromolecules in
Chapter 3. Now it is time to turn our attention to the protective role played by the
antioxidants. This chapter will present an overview of endogenous and
exogenous antioxidants, examine the use antioxidant therapy to treat disease,
present ways of estimating the total antioxidant capacity of different systems and
explore the use of antioxidants as food preservatives.
WWW.ESAINC.COM
308
INTRODUCTION.
Mention the word “antioxidant” and most people would probably think of vitamins
C or E. However, although these compounds are vitally important antioxidants,
many other molecules are used as antioxidants in vivo.1 In fact antioxidants are
made up of three different classes of molecules:
1. Enzymes (e.g., catalase and superoxide dismutase).
2. Metal sequestering proteins (e.g., ferritin and ceruloplasmin).
3. Low molecular weight (small) molecules (e.g., vitamins C (ascorbic acid) and
E (α-tocopherol).
Perhaps the best definition of an antioxidant was put forward by Halliwell and
Gutteridge (1990) who stated that an antioxidant is “any substance that, when
present at low concentrations compared to those of an oxidizable substrate,
significantly delays or inhibits oxidation of that substrate”. In this instance a
substance could be any of the three classes of antioxidants mentioned above,
while substrate refers to many biologically important molecules such as DNA,
lipids and proteins. However, this definition is not perfect. For example, the
hydroxyl free radical reacts with virtually every molecule it encounters, but it
would be ridiculous to propose that such molecules are antioxidants (see below).
Lester Packer modified Halliwell and Gutteridges’ definition and suggested that
such compounds should ideally display a range of antioxidant activities, react
with more than just one specific pro-oxidant, and be present in sufficient
concentration in vivo. This definition is appropriate for small molecule
antioxidants but does not take into account the ability for enzymes and metal
sequestering proteins to act as antioxidants.
The antioxidant or suite of antioxidants used to control oxidation is a very
complex topic. Much is dependent upon which pro-oxidant species is involved,
and where it is being produced. For example:
1) Under biological conditions, superoxide is relatively long lived. Its cellular
level is readily controlled by intracellular enzymes but in the extracellular
fluids, where enzyme activity is typically low, its level is probably either
regulated by small molecule species, or once it has been taken up by
erythrocytes, by enzymes. Control of the plasma levels of superoxide is of
critical importance due to its reaction with circulating nitric oxide. Thus by
influencing the level of superoxide, extracellular superoxide dismutase (ecSOD) can direct nitric oxide’s activity away from its normal physiological
role towards oxidation through the formation of peroxynitrite.
2) The hydroxyl free radical is much too reactive to be controlled
enzymatically. In fact, the hydroxyl free radical is just as capable of
damaging enzymes as it is any other molecule it encounters. Rather than
1
Remember that for many compounds showing antioxidant activity, such reactions are often secondary to other more
important biological functions.
WWW.ESAINC.COM
309
dealing directly with the hydroxyl free radical, by far the most effective
approach is to prevent its production in the first place. To this end, the
availability of substrates involved in hydroxyl free radical production is
tightly regulated: hydrogen peroxide is catabolized enzymatically, while
iron is sequestered in an unreactive form.
3) Prevention of damage to the hydrophobic lipid portion of membranes is
controlled by metal sequestration, enzymatic removal of pro-oxidants and
through the reaction of lipid peroxyl radicals with the fat-soluble vitamin, αtocopherol. Regeneration of α-tocopherol from its radical is accomplished
probably through reaction with cytosolic ascorbic acid or glutathione
(GSH) at the membrane-cytosol interface.
The next three sections will explore the way living organisms make use of
enzymes, metal chelators and low molecular weight molecules as antioxidants.
ENZYMES.
Many different enzymes can act as antioxidants. Enzymes can be categorized as
being either primary, reacting directly with pro-oxidants (e.g., superoxide
dismutase (SOD), and catalase), or secondary, involved in regenerating low
molecular weight antioxidant species (e.g., ascorbate dehydrogenase and
glutathione reductase). Some background information including the basic
properties of the antioxidant enzymes is presented in Table 4.1.
Enzyme
Ascorbate Dehydrogenase;
GSH-dependent
Dehydroascorbate (DHAA)
Reductase; Glutaredoxin;
NADH-dependent (DHAA)
reductase; NADPHdependent (DHAA)
Reductase
Catalase (CAT)
Comments
Secondary. Regenerate ascorbic
dehydroascorbic acid.
acid
from
Primary. Discovered by Thernard in 1818. Named
catalase by Loew in 1901. There are many forms
of CAT – most contain heme but some contain
manganese.
E. coli contain two forms of CAT:
HPI is a tetrameric protein (337,000 Daltons)
containing 2 molecules of protoheme IX per
tetramer and is associated with the periplasmic
membrane. It is inducible by hydrogen peroxide
and ascorbate. It possesses both catalytic and
peroxidative activities. HPI is increased during log
growth.
HPII
is
a
tetrameric
WWW.ESAINC.COM
protein
containing
Reference
Maellaro et al.
(1994); Park and
Levine (1996);
Rose and Bode
(1992, 1993);
Sauberlich (1994);
Wells and Xu
(1994)
Iozzo et al. (1982);
Michiels et al.
(1994); Percy
(1984); Radi et al.
(1991); Singh et al.
(1996) and
references therein
2
310
molecules of protoheme IX per tetramer and is
associated with the cytoplasm. It possesses
peroxidatic activity only, is not inducible by either
hydrogen peroxide or ascorbate and it is
increased during the stationary phase of growth.
Humans also contain at least two forms of CAT:
Ubiquitous heme-based enzyme of molecular
mass ~240,000 Daltons, consisting of four
identical
subunits
each
containing
ferriprotoporphyrin. Different forms of the enzyme
are found in peroxisomes and cytoplasm. CAT
has also been reported to be found in the
cytoplasmic granules of eosinophils and in heart
mitochondria.
It
is
abundant
in
liver
(peroxisomes),
kidney
and
erythrocytes
(cytoplasm).
In higher organisms CAT binds four molecules of
NADPH. Although the function of NADPH binding
is unclear it has been suggested to protect CAT
from hydrogen peroxide inactivation or to act as a
source of NADPH for glutathione peroxidase
under conditions of stress.
CAT concentrations are ~1.2 x 10-6, 3.8 x 10-8,
and 7.2 x 10-7M for liver, heart and heart
mitochondria, respectively.
CAT is inhibited non-specifically by cyanide and
azide, or specifically by aminotriazole, which
interacts with compound 1 (see below). Thus
aminotriazole can only inhibit CAT if hydrogen
peroxide is present to generate compound 1.
Glutathione Peroxidase
(GPx)
Manganese-CATs are found in a variety of
microbes. The enzyme from Lactobacillus
plantarum has a molecular weight of 172,000
Daltons and is composed of six subunits each
containing one Mn3+ ion. As it does not contain
heme it is not inhibited by either cyanide or azide.
The enzyme is thought to be important in
protecting microbes from damage by hydrogen
peroxide.
Primary. Discovered by Mills in 1957. This
enzyme is not found in bacteria or higher plants,
but is found in all eukaryotes. It shows distinct
tissue distribution with high amounts in liver,
moderate amounts in heart, lung and brain, and
low amounts in muscle. Liver levels of GPx are
reported to be ~2.7 x 10-6 and 1.2 x 10-6M for liver
cytosol and mitochondria, respectively. There are
five known forms:
Chu et al. (1993);
Michelson (1998);
Michiels et al.
(1994); Sies et al.
(1997); Stadtman
(1980, 1990); Ursini
et al. (1997)
GPx-1. A cytosolic enzyme of molecular mass of
76,000 to 92,000 Daltons. It consists of four
WWW.ESAINC.COM
311
identical subunits each containing selenium in the
form of a single selenocysteine residue.
Unspecific acting on hydrogen peroxide, lipid
hydroperoxides, peroxidized membranes and
DNA-hydroperoxides. Specific for its cofactor,
GSH.
Mitochondrial GPx. It has never been isolated but
may be related to thioredoxin/peroxiredoxin. Likely
to occur in the matrix.
Human plasma GPx. A tetrameric protein of
molecular mass ~88,000 Daltons that is
synthesized and secreted by the kidney. Each
subunit contains a selenium atom.
A novel GSHPx-1 enzyme. This has similar
substrate specificities to GPx-1. This cytosolic
enzyme is a tetrameric protein with a molecular
weight of 88,000 Daltons. Reduces hydrogen
peroxide, amino hydroperoxide, linoleic acid
hydroperoxide but not phosphatidylcholine
hydroperoxide.
Glutathione Reductase
(GR)
Glutathione synthetic
enzymes
Glutathione-S-Transferase
(GST)
Phospholipid
hydroperoxide
glutathione
peroxidase PH-GPx. This is a monomeric
(~20,000 Daltons), hydrophobic, seleniumcontaining enzyme that is associated with
membranes and preferentially reduces lipid
hydroperoxides (Chapter 4). It occurs in two
forms. The long form (23,000 Daltons) is a
mitochondrial GPx and is located in mitochondrial
membrane. The short form (20,000 Daltons) is
found in the cytosol. The long form may play a
critical role in prevention of apoptosis.
Secondary. GR is a flavoprotein (molecular mass
104,800 Daltons; composed of two subunits)
which reduces GSSG to GSH using NADPH
produced by the pentose phosphate pathway. The
enzyme is located in both cytoplasm and
mitochondria. GR is responsible for removal of
toxic GSSG and keeps GSH in its biologically
active form. GR can also reduce mixed disulfides
such as those formed between GSH and
coenzyme A.
Secondary. γ-Glutamylcysteine synthetase and
glutathione synthetase – see below.
Primary. A non-selenium containing GPx. It is a
dimer composed of four different subunits (22,000
to 25,000 Daltons) which combine to produce six
isozymes. The enzyme is found in liver (where it
accounts for ~10% of total soluble protein), red
blood cell and intestine. It is located in the
cytoplasm, nucleus and on the surface of the cell,
but not in mitochondria. GST does not act on
WWW.ESAINC.COM
Meister and
Anderson (1983)
Burk et al. (1978);
Jenkinson et al.
(1983); Ketterer
and Meyer (1989);
Lawrence and Burk
(1976)
312
Heme Oxygenase
Paraoxonase (PON 1)
Peroxiredoxin (Prx)
Superoxide
Dismutase (SOD)
hydrogen
peroxide
but
does
reduce
hydroperoxides. GST may function as GPx when
Se levels are low. It is also involved in
detoxification of xenobiotics through the
mercapturic acid pathway (see below). GST is
also involved in conjugation reactions of
endogenous compounds (e.g., steroids, and
prostaglandins).
Primary. Microsomal mono-oxygenase enzymes
consisting of two homologous isozymes
responsible for conversion of the pro-oxidant,
heme, to the antioxidant, bilirubin. NADPH and
oxygen are used as cofactors. HO-2 is
constitutive while HO-1 is an inducible-enzyme.
HO-1 production is stimulated up to 100 fold by
cytokines, endotoxins, glutathione depletion, heat
shock, heavy metals, prostaglandins, and by
Alzheimer’s
disease.
Inhibited
by
zinc
protoporphyrin-9 and tin-protoporphyrin. HO-1
plays a role in adaptive resistance of cells to
chronic exposure of nitric oxide.
Primary. An enzyme found in human serum
associated with high-density lipoprotein (HDL). It
protects both HDL and LDL against lipid
peroxidation and is capable of hydrolyzing both
peroxidized
phospholipids
and
cholesteryl
linoleate hydroperoxides. PON 1 also possesses
arylesterase activity capable of hydrolyzing
organophosphate insecticides and nerve gases.
PON 1 has two genetic polymorphisms important
in determining the capacity of LDL to protect HDL
and its activity towards organophosphates.
Primary. Consists of six members. Prx-3 is
located in the mitochondria, Prx-1 and Prx-2 exist
in the cytosol, and Prx-5 is found in both
mitochondria and peroxisomes.
Prx-3, also called MER5 and antioxidant protein-1
(AOP-1), scavenges hydrogen peroxide in the
cooperation with thiol, and peroxynitrite by itself.
Primary. Discovered by Fridovich and McCord in
1969. Consists of several unrelated enzymes
differing in amino acid sequence, active metal site
and cellular location.
Prokaryotes have an iron-SOD (Fe-SOD)
associated with their outer membrane and a
manganese-SOD (Mn-SOD) located in their
matrix. The Fe-SOD is a dimer with overall
molecular weight of 40,000 Daltons. The Mn-SOD
occurs as dimers or tetramers with overall
molecular weight of 40,000 and 80,000 Daltons,
respectively. E. coli also contains a hybrid SOD
consisting of one subunit of MnSOD and one
subunit of FeSOD.
WWW.ESAINC.COM
Bishop et al.
(1999); Maines
(1997); Smith et al.
(1994)
Aviram et al.
(1999); Mackness
et al. (1996, 1998)
Hattori, et al.,
(2003) and
references therein
Adachi et al.
(1992); Benov et al.
(1996);
Hjalmarsson et al.
(1987); Fridovich
(1974, 1986; 1995);
Geller and Winge
(1982); Gregory
and Dapper (1980);
Halliwell and
Gutteridge (1999);
Karlsson and
Marklund
(1988a,b); Luoma
et al. (1998);
313
Eukaryotes contain:
A copper-zinc enzyme (CuZn-SOD). This is an
acidic, very stable protein that occurs in the
cytoplasm, possibly between the inner and outer
mitochondrial membranes, in lysosomes and in
the nucleus. This blue-green colored dimeric
enzyme has molecular weight of ~32,000 Daltons
and is readily inhibited by cyanide and copper
chelators such as diethyldithiocarbamate. Liver
levels of this enzyme are ~2.4-3.7 x 10-5M.
Marklund (1984));
Marklund et al.
(1982); McCord
and Fridovich
(1969); Michalski
(1996); Michiels et
al. (1994);
Sandstrom et al.
(1992); Stralin et al.
(1995)
A manganese enzyme (Mn-SODs). This occurs in
the mitochondrial matrix. This pink tetrameric
enzyme has a molecular weight of ~88,000
Daltons. It is less stable than CuZn-SOD but is
not affected by cyanide or copper chelators. Liver
levels of this enzyme are ~0.3-1.1 x 10-5M
An extracellular copper-zinc enzyme (EC-CuZnSOD). This tetrameric, slightly hydrophobic
glycoprotein has a molecular weight of 135,000
Daltons. It is the major SOD found in extracellular
fluid (e.g., plasma, lymph and synovial fluid). It is
also found in tissues and binds to the endothelial
cells of the vascular system. It is involved in
inflammation. The enzyme shows heterogeneity
with regard to its ability to bind heparin and
therefore its ability to bind to proteoglycans found
on the cell surface and connective tissue matrix.
Class A lacks heparin binding, class B has
intermediary binding while class C binds heparin
with strong affinity. Interestingly, heparin
suppresses inflammation by releasing EC-Cu,ZnSOD. This enzyme is inhibited by cyanide, azide,
hydrogen peroxide and copper chelators.
Thioredoxin, and
Thioredoxin Reductase
An extracellular manganese enzyme (EC-MnSOD). This enzyme has a molecular weight of
150,000 and occurs as a dimer or tetramer.
Primary. Thioredoxin (TRX) is a small (~12,000
Daltons), heat-stable, dithiol, redox-active protein
widely distributed in bacterial, plant and animal
kingdoms; TRX-1 is located in the cytosol,
nucleus and can be excreted while TRX-2 only
occurs in the mitochondria. Thioredoxins contain
cysteine residues capable of undergoing
reversible oxidation. Reduced thioredoxin serves
as a hydrogen donor for ribonucleotide reductase,
protein tyrosine phosphatase and for enzymes
involved in reducing sulfate and methionine
sulfoxide. The thioredoxin/thioredoxin reductase
system also acts as a peroxidase and is capable
of scavenging lipid hydroperoxides. It is also
involved in the up-regulation of the expression of
catalase-hydroperoxidase I. Thioredoxin can also
WWW.ESAINC.COM
Arteel et al. (1999);
Bjornstedt et al.
(1997); Buchanan
et al. (1994);
Follman and
Haberlein (1995);
Gleason and
Holmgren (1988);
Holmgren (1985);
Takemoto et al.
(1998) and
references therein;
Zhong et al. (1998)
314
donate electrons to glutathione peroxidase
thereby providing a mechanism by which human
plasma glutathione peroxidase can reduce
hydroperoxides
under
conditions
of
low
glutathione levels.
Thioredoxin Peroxidase
(TPx) and peroxiredoxins
(Prx)
Thioredoxin reductase is an FAD-containing
general-purpose protein disulfide reductase which
uses NADPH as the reducing agent for efficient
reduction of TRX-disulfide to TRX. Mammalian
thioredoxin reductase, a dimeric protein
(molecular weight 57,000 Daltons) is homologous
to glutathione reductase. Interestingly, the rat
enzyme contains a selenocysteine residue and in
addition to its primary substrate, thioredoxin, has
broad substrate specificity (e.g., selenite,
organoselenium compounds, cytosolic and
plasma GPx, and peroxynitrite).
TPx is a newly discovered, non-selenium
containing enzyme that reduces hydrogen
peroxide using electrons from thioredoxin. It has
two essential conserved cysteine moieties. Its
reaction is analogous to GPx.
TPx homologs, the Prxs are designated 1-Cys Prx
as they contain only one conserved cysteine
moiety. Four Prxs have been identified, all are the
products of distinct genes and all are members of
the thioredoxin superfold family. Prx I and II are
cytosolic, Prx III is mitrochondrial, while Prx IV is
extracellular. The Prxs act on hydrogen peroxide
and phospholipid hydroperoxides. For many of
these proteins the endogenous thiol cofactor
remains unknown. These enzymes may play a
role in cell proliferation and differentiation,
protection of proteins from oxidative damage, and
intracellular signaling.
Fisher et al. (1999);
Kang et al.
(1998a,b); Lee et
al.(1999); Lim et al.
(1998); Matsumoto
et al. (1999);
Schroder and
Ponting (1998)
Table 4.1 Some Antioxidant Enzymes.
Catalases.
CAT is a very reactive enzyme that is responsible for the dismutation of hydrogen
peroxide (Eqn 4.1). The actual mechanism of CAT activity is best described by
Eqns 4.2 and 4.3. Compound I contains iron in the Protein-FeIV-O form2
(Footnote 2) (Eqn 4.4) (and see Chapter 2). It is very difficult to saturate
2
Many reactions catalyzed by heme-containing proteins involve the oxidation of heme one (compound II) or two
(compound I) equivalents above the Fe (III) state. Compound I is readily formed by catalase, cytochrome P450 monooxygenases and others (Eqn 4.2). There is still great confusion in this area with several different formulae (e.g.,
IV
IV
4+
+
4+
+
Protein• -Fe =O, Protein-Fe =O, Protein-Fe =O and Protein -Fe =O) being used to describe the structure of compound
I. Readers are referred to Everse (1998) for an excellent review on this topic.
WWW.ESAINC.COM
315
catalase with hydrogen peroxide because its maximum velocity (Vmax) for
dismutation of hydrogen peroxide is huge.
H2O2 + H2O2 → H2O + O2
CAT-FeIII + H2O2 → Compound I
(k=1.7 x 107 M-1s-1)
Compound I + H2O2 → CAT-FeIII + H2O + O2
(k=2.6 x 107 M-1s-1)
Protein-FeIII + H2O2 → Protein-FeIV-O + H2O
Eqn 4.1
Eqn 4.2
Eqn 4.3
Eqn 4.4
Similar reactions are though to take pace for manganese-CAT (Eqns 4.5 and
4.6).
CAT-Mn3+ + H2O2 + 2H+ → CAT-Mn (V)
CAT-Mn (V) + H2O2 → CAT-Mn3+ + O2 + 2H+
Eqn 4.5
Eqn 4.6
CAT can also show peroxidative reactions in the presence of hydrogen peroxide,
with compound I oxidizing methanol and ethanol to methanal and ethanal,
respectively (Thurman and Handler (1989)). The peroxidative action of CAT is
typically much less rapid (k=102 to 103 M-1s-1) than its catalytic activity. Spinach
catalase can oxidize formic acid to carbon dioxide, and nitrite into nitrate
(Halliwell and Gutteridge (1999)).
CAT activity is localized mainly to peroxisomes, fragile membrane bound
organelles of various shapes and sizes (Chandoga (1994)). Mammalian
peroxisomes are intimately involved with redox biochemistry. They are
responsible for the production of hydrogen peroxide (flavin oxidases), catabolism
of hydrogen peroxide (CAT) and superoxide (SOD), lipid metabolism
(biosynthesis and catabolism (oxidation of fatty acids and their derivatives)) and
intermediary metabolism (transaminases and dehydrogenases) (del Rio et al.
(1992); Lazarow (1987); Vamecq and Draye (1989)). Peroxisomes are essential
to normal cellular function. Peroxisome dysgenesis and/or dysfunction in their
enzymatic pathways are found in several inherited metabolic diseases with
serious clinical ramifications (e.g., Zellweger syndrome) (Keller et al. (1993)).
CAT activity can be determined using colorimetric, spectrophotometric, and
polarographic assays (Aebi (1984); Cohen et al. (1996); Dempsey et al. (1975);
Haining and Legan (1972); Sinah (1972); Thomson et al. (1978)).
WWW.ESAINC.COM
316
Peroxidases.
Peroxidases catalyze the general reaction presented in Eqn 4.7, where R
represents an alkyl group.
RH2 + H2O2 → R + 2H2O
Eqn 4.7
Glutathione Peroxidase-1 (GPx-1) is an important antioxidant enzyme that
specifically requires glutathione (GSH) as a cofactor. It is a relatively unspecific
enzyme acting upon hydrogen peroxide, free lipid hydroperoxides and other
peroxidized compounds (Eqns 4.8 and 4.9). In contrast to heme peroxidases
(e.g., CAT) GPx produces just one oxidation product. Overall GPx is much more
versatile than CAT but unlike CAT it reacts with hydrogen peroxide only slowly.
2GSH + H2O2 → GSSG (glutathione disulfide) + 2H2O
2GSH + RO2H → GSSG + ROH + H2O
Eqn 4.8
Eqn 4.9
GPx-1 cannot act on lipid hydroperoxides when they are part of a membrane,
unless the hydroperoxide is first freed by phospholipase A2 (see Figure 3.29).
Phospholipid hydroperoxide glutathione peroxidase (PH-GPx), on the other hand,
plays an important role in controlling membrane lipid peroxidation and is the only
enzyme capable of reducing fatty acid hydroperoxides while they are still part of
the membrane (Chapter 4).
The GPxs are remarkable because they require selenium. The selenium atom is
part of a selenium analog of cysteine that is covalently attached to the enzyme
and is located within its active site (Figure 4.1). (Reddy and Massaro (1983);
Zachara (1992)). Selenium is an essential micronutrient obtained from the diet.
Humans require at least 60µg/day (typical intake for developed countries is 60200µg/day) in order to remain healthy (Burk (1989); Foster and Sumar (1997)).
Selenium deficiency is associated with liver necrosis, degenerative heart disease
(Keshan disease), exudative diathesis and a failure to grow and reproduce
(Reddy and Massaro (1983); Zachara (1992)). Overdosing with selenium is
associated with increased lipid peroxidation and cellular toxicity possibly through
generation of ROS (Seko and Imura (1997)).
WWW.ESAINC.COM
317
GSH
GSSG + H+
E-Se-S-G
E-Se-
Selenosulfide
Selenolate
ROOH + H+
H2O
GSH
E-Se-OH
ROH
Selenenic
Acid
Figure 4.1 The Role Of Selenium In Glutathione
Peroxidase.
Glutathione peroxidase and possibly other selenoproteins that contain either
selenocysteine or selenomethionine can also act as peroxynitrite reductases
protecting against oxidative damage caused by this pro-oxidant (Sies et al.
(1997)).3 Unfortunately, GPx can itself be inactivated by peroxynitrite probably by
oxidation of the ionized selenol of the selenocysteine residue (Padmaja et al.
(1998)).
In order for glutathione peroxidases to function properly GSH availability must be
maintained. GSSG must therefore be continuously reduced to GSH. This is
achieved by glutathione reductase in a reaction using NADPH, (Eqn 4.10) (and
see below).
GSSG + NADPH + H+ → 2GSH + NADP+
Eqn 4.10
Several non-specific peroxidases are also found in animals. Myeloperoxidase is
located in phagocytes. As discussed in Chapter 2, it is responsible for the
production of hypochlorous acid, a strong pro-oxidant and bacteriocidal agent.
Thyroid peroxidase is the major enzyme of the thyroxine synthesis pathway and
is responsible for both the iodination of tyrosine residues of the protein
thyroglobulin and also their coupling to form iodothyronines. Lactoperoxidase is
found in milk and saliva and may be responsible for the oxidation of thiocyanate
3
Interestingly, derivatives of another Group 6B element, tellurium, also protects against peroxynitrite-mediated oxidation
and nitration reactions (Briviba et al. (1998)).
WWW.ESAINC.COM
318
(also found in milk and saliva) to hypothiocyanate, a compound toxic to some
pathogens. Hypothiocyanate may offer some protection to babies against
gastrointestinal tract infection. Other peroxidases include yeast cytochrome c
peroxidase, bacterial NADH peroxidase, plant non-specific horseradish
peroxidase, fungal non-specific chloroperoxidase and plant ascorbate
peroxidase.
For the purification of GPx and measurement of enzyme activity readers are
referred to the excellent review by Toribio et al. (1996) and to Halliwell and
Gutteridge (1999).
The Biological Importance of Catalase and Glutathione Peroxidase.
At first site it may appear that nature is being redundant in having two enzymes,
CAT and GPx, that are both capable of catabolizing hydrogen peroxide. It should
be remembered, however, that these enzymes show distinct activity and
specificity. CAT acts very rapidly and specifically with hydrogen peroxide,
whereas GPx reacts more slowly with hydrogen peroxide but is capable of
reacting with hydroperoxides and other peroxidized compounds too. Some
tissues have one enzyme present at much higher levels than the other so it
becomes de facto the enzyme responsible for controlling hydrogen peroxide
levels. For example, GPx is the principal enzyme for controlling hydrogen
peroxide levels in brain and spermatozoa.
Some tissues have both enzymes present but have them located in different
compartments. For example, liver CAT is located in peroxisomes where it is
responsible for handling hydrogen peroxide produced by the ROS generating
enzymatic pathways located within these organelles. Liver GPx is located in the
cytosol where it can destroy hydrogen peroxide produced within this
compartment. Interestingly, heart mitochondria also contain a CAT located within
the matrix that handles hydrogen peroxide produced within these organelles
(Radi et al. (1991)).
In some cases there seems to be a backup function. For example, although both
enzymes are present in red blood cells, it would appear that the low levels of
hydrogen peroxide produced by SOD is normally handled by glutathione
peroxidase. CAT is only recruited when glutathione peroxidase is overwhelmed
e.g., following administration of a drug which generates hydrogen peroxide.
Interestingly, patients suffering from an inborn error of metabolism, acatalasemia,
who possess a mutant CAT with low activity, show few, if any, symptoms.
Presumably, in such cases, GPx compensates for the low catalase activity.
WWW.ESAINC.COM
319
Glutathione-S-Transferase.
Glutathione-S-transferase (GST) is a non-selenium containing enzyme that
functions in a similar fashion to GPx. GST does not act on hydrogen peroxide but
it will reduce a variety of hydroperoxides (Eqns 4.11 and 4.12).
ROOH + GSH → GSOH + ROH
GST
GSOH + GSH → GSSG + H2O
Nonenzymatic?
Eqn 4.11
Eqn 4.12
GST is also involved in the detoxification of some xenobiotics through the
mercapturic acid pathway (see below).
Heme Oxygenases.
Heme plays several important roles in the body (e.g., as part of hemoglobin in
the red blood cell it is responsible for transporting oxygen throughout the body).
Unfortunately, heme can be a major problem too. Heme released following red
blood cell turnover, tissue damage, or turnover of heme-containing proteins, can
initiate lipid peroxidation or react with hydrogen peroxide to release redox-active
iron (Chapter 3). This iron, in turn, can take part in the Fenton reaction with
additional hydrogen peroxide. It is vitally important therefore that during heme
catabolism its potential pro-oxidant activity is prevented. This is the responsibility
of heme oxygenase (HO).
HO consists of two forms, the inducible protein HO-1 (HSP32) and the
constitutive isozyme HO-2. These enzymes, which are different gene products,
share limited sequence homology and show different regulation and tissue
distribution. Both enzymes catalyze the first and rate-limiting step in heme
degradation producing three biologically important molecules: biliverdin, carbon
monoxide and iron (Figure 4.2). Iron is a gene regulator and, as discussed
earlier, a pro-oxidant. Thus free iron produced during this process must be
effectively bound in order to prevent it from reacting with hydrogen peroxide.
Biliverdin is reduced to bilirubin by biliverdin reductase. While bilirubin is an
effective antioxidant, biliverdin is not. In scavenging two hydroperoxyl radicals
bilirubin is oxidized to biliverdin, and this can then be rapidly reduced back to
bilirubin by biliverdin reductase leading to react with more pro-oxidants. In
addition, bilirubin is an antimutagen and anti-complement agent (Marilena
(1997)). Carbon monoxide plays a similar role to nitric oxide and its reaction with
guanylyl cyclase has been implicated as a modulator of gene expression, in
retrograde neurotransmission and a modulator of vascular tone (Dulak and
Jozkowicz (2003); Durante (2002); Ingi and Ronnett (1995) and references
WWW.ESAINC.COM
320
therein; Marilena (1997)). HO-2 is expressed at high levels in discrete brain
regions and the carotid bodies.
HEME
O2 + NADPH
HEME OXYGENASE
CO
+ Fe3+
H2O + NADP+
GUANYLYL
CYCLASE
R'
R
R
R"
R
R'
O
NH
NH
N
NH
O
R"
R
BILIVERDIN
NADPH + H+
BILIVERDIN
REDUCTASE
NADP+
R'
R
O
NH
R"
R
N
R
R"
R
NH
R'
NH
O
BILIRUBIN
Figure 4.2 The Catabolism Of Heme.
Superoxide Dismutases.
Superoxide dismutase (SOD) is responsible for the dismutation of superoxide
(Eqn 4.13). It is the only enzyme known that can act on a free radical. Although
superoxide dismutation occurs naturally at physiological pH, it is much slower
(k2=~5 x 105 M-1s-1) than the enzyme catalyzed reaction (e.g., CuZn-SODbovine
k2=~1.6 x 109 M-1s-1; Mn-SODE. coli k2=1.8 x 109 M-1s-1).
WWW.ESAINC.COM
321
2O2•- + 2H+ → H2O2 + O2
Eqn 4.13
The general mechanism for superoxide dismutation involves changes in the
oxidation state of the metal within the enzyme (Eqns 4.14-4.16) (also see Hart et
al. (1999)). In this example, M(n)+ represents Cu+ or Fe2+. The Zn atom of the
CuZn-SOD appears not to take part in the reaction, but is essential in the
stabilization of the enzyme. The mechanism of action of the manganesecontaining enzyme is much more complex and will not be dealt with here.
E-M(n+1)+ + O2•- → E-M(n)+ + O2
Electron transfer
(n)+
E-M
+ O2•- + 2H+ → E-M(n+1)+ + H2O2
Proton and electron transfer
Overall: O2•- + O2•- + 2H+ → H2O2 + O2
Eqn 4.14
Eqn 4.15
Eqn 4.16
It is interesting to note that in certain circumstances both CuZn-SOD and FeSOD can be inactivated by their own reaction product, hydrogen peroxide. In the
case of CuZn-SOD it appears that hydrogen peroxide reduces Cu2+ to Cu+ which
causes reversible inactivation. At higher hydrogen peroxide levels Cu+ can then
produce hydroxyl free radicals that can oxidize an essential histidine (His118) in
the enzyme’s active site to an inactive 2-oxohistidine residue, thereby irreversibly
inhibiting the enzyme (Chapter 3). Excessive production of hydroxyl free radicals
can permanently damage the enzyme through fragmentation (see Uchida and
Kawakishi (1994) and references therein). As CuZn-SOD is very effective at
controlling superoxide levels the physiological significance of hydrogen peroxide
inhibition still remains to be clarified.
CuZn-SOD, by affecting the concentration of superoxide, plays a critical role in
the regulation of the formation of peroxynitrite from its precursors, superoxide
and nitric oxide (Chapter 2). This has led some researchers to suggest that the
raison d’être for SOD is to enhance the biological activity of nitric oxide by
preventing its diversion to peroxynitrite (Pryor and Squadrito (1995)). Thus downregulation of CuZn-SOD leads to cell death by overproduction of peroxynitrite,
while over-expression or supplementation with SOD can overcome some of the
problems associated with peroxynitrite (Troy et al. (1996)). To help explain the
importance of SOD in controlling the production of peroxynitrite Crow and
Beckman (1995) proposed the use of a “target area”. This is the probability that
some species will be attacked by a pro-oxidant, and depends upon the reaction
rate times the concentration of the species. Thus as the reaction rate of CuZnSOD with superoxide is ~2 x 109 M-1s-1 and CuZn-SOD’s typical cellular
concentration is ~10-5M then this enzyme has a target area of 2 x 10-4 s-1. Nitric
oxide reacts with superoxide at a rate of 6.7 x 109 M-1s-1 and, under physiological
conditions, is present at a concentration of 10-7M. Nitric oxide therefore has a
WWW.ESAINC.COM
322
target area 30 fold lower than SOD. However, under pathological conditions nitric
oxide can reach concentrations of 4 x 10-6M, resulting in a target area of 2.7 x
104 s-1, exceeding the target area of SOD (Crow and Beckman (1995) and
references therein).
Changes in SOD activity can lead to disease and may even play a role in aging.
Autosomal dominant mutations in CuZn-SOD has been linked to familial
amyotrophic lateral sclerosis (ALS) (Andersen et al. (1997); Beckman (1993) and
references therein; Wong and Borchelt (1995)). Interestingly transgenic animals
expressing CuZn-SOD mutants also have decreased chaperone activity (proteins
that are responsible for proper protein folding and targeting mutant proteins for
degradation – Chapter 3) (Bruening et al. (1999)). Insufficiency of chaperones
may be directly involved in loss of motor neurons in ALS.
The role of SOD in aging is less clear. Genetic manipulations that increase
CuZn-SOD activity have only slight, if any, effect on maximal life span of several
species, even though they do show increased resistance to oxidative stress
(Sohal (1997) and references therein; Warner (1994) and references therein).
Interestingly increasing the activity of both CuZn-SOD and catalase does
significantly increase life span (Sohal (1997) and references therein; Warner
(1994) and references therein). The amount of SOD present in tissues can be
determined using immunological methods (Halliwell and Gutteridge (1999)). SOD
activity can be determined using both direct- and indirect methods. Direct
methods include pulse radiolysis, stop flow spectroscopy, EPR-spin with spin
trap, far UV detection, polography and 19F NMR (Halliwell and Gutteridge (1999);
Michalski (1996)). Unfortunately these approaches are not usually practical with
crude enzyme preparations and at SOD levels typically found in biological
systems. For these reasons, indirect assays are far more common. In these
assays, superoxide is generated by some mechanism (e.g., xanthine/xanthine
oxidase, auto-oxidation reactions, or illuminated flavins) and allowed to react with
a reporter molecule (e.g., cytochrome c or nitroblue tetrazolium) which can be
monitored, usually by changes in absorbance (Halliwell and Gutteridge (1999);
Michalski (1996)). The addition of SOD will remove superoxide from the reaction
and thereby inhibit the absorbance change. Readers are referred to Halliwell and
Gutteridge (1999) and Michalski (1996) for precautions when using indirect SOD
assays.
The Catabolism of Nitric Oxide.
Unlike superoxide and hydrogen peroxide, nitric oxide is not catabolized
enzymatically. Rather, as discussed in Chapter 2, nitric oxide undergoes a series
of oxidation reactions eventually producing nitrite and nitrate. Excess nitric oxide
produced in the plasma is also oxidized to nitrate by its reaction with
oxyhemoglobin or oxymyoglobin (Figure 2.18). Nitric oxide can also be removed
by its reaction with superoxide to produce peroxynitrite.
WWW.ESAINC.COM
323
SEQUESTRATION OF METAL IONS.
As has already been discussed, transition metals are capable of undergoing
redox reactions. They therefore possess pro-oxidant activity and are capable of
initiating lipid peroxidation and hydroxyl free radical production. Of all the
transition metals encountered in vivo, iron and copper are the most abundant and
therefore the most problematic. The amounts of nickel, manganese, chromium,
vanadium and cobalt are usually incredibly low under physiological conditions
and thus do not normally pose a problem.4 Adult humans typically contain about
4.5g of iron, obtain ~1mg of iron from the diet and, when in iron-balance, excrete
~1mg. About 60% of the body’s iron is found in hemoglobin, the remainder
occurs in myoglobin, various enzymes and in the iron-transport protein,
transferrin (Table 4.2). Other pools include iron bound to iron-storage proteins
(e.g., ferritin) and to non-protein chelators (e.g., citrate, ATP, ADP and GTP).
Free iron concentrations are very low (Fe2+ ~10-8M; Fe3+ ~10-18M). Consequently
there is virtually no free iron ions in humans (iron is undetectable in plasma from
healthy individuals but can be measured in sweat (Gutteridge and Halliwell
(1999)). Human adults also contain ~80mg of copper. About 90% of circulating
copper is bound to ceruloplasmin; the rest is bound to albumin, small peptides
and histidine (Table 4.2). There is little free copper (ions) in humans (e.g., normal
human plasma is devoid of free copper, while sweat contains
2-27µM (Gutteridge and Halliwell (1999)).
One approach to prevent the pro-oxidant activity of iron and copper would be to
eliminate them entirely from the biological system. Not only would this be
physiologically impossible, but it would be fatal as well, for both iron and copper
are essential micronutrients and play a role in the activity of several enzymes, in
the immune system and the electron transport chains (Harris and Gitlin (1996);
Olivares and Uauy (1996); Percival (1998); Solomon and Lowery (1993); Stryer
(1988)). Instead, nature makes use of these metal ions but keeps them
sequestered in redox-inactive chelation complexes. For a given metal ion, the
metal’s oxidation state and the nature of its ligand affect the strength of chelation.
In general chelators that use oxygen atoms in ligating the metal prefer the
oxidized form of the metal; consequently the redox potential of the metal is
usually decreased. Chelators that use nitrogen atoms to bind the metal prefer its
reduced form; thus the redox potential is usually increased (Miller et al. (1990)).
For example, the effects of chelation on the redox potential is readily apparent
from Appendix 2.1. The Eo’ for the Fe3+/Fe2+ couple is +260mV, +110 and
-190mV for cytochrome, aqueous ions and ferritin, respectively. The importance
of binding and redox state is illustrated by iron’s interaction with ferritin and this is
discussed further below.
4
However, inadvertent exposure to elevated amounts of these metals can be biologically catastrophic.
WWW.ESAINC.COM
324
Metal
Iron
Protein
Transferrin
(lactoferrin;
melanotransferrin;
ovotransferrin
[conalbumin];
plasma
transferrin; and
uteroferrin)
Process
Transport
Comments
Group of proteins with a molecular
mass of ~80,000 Dalton capable of
tightly but reversibly binding two Fe3+
ions. Requires a counter anion
(usually bicarbonate) at each site for
tight binding of iron. Human
transferrin is usually only 30% full
under normal conditions. Occurs in
extracellular locations, but can enter
the cell for delivery of Fe3+ to ferritin
or to mitochondrial ferrochelatase for
heme biosynthesis. Also includes
ovotransferrin, lactotransferrin and
sero-transferrin.
References
Aisen and
Listowsky
(1980);
Crichton
(1990)
Iron
Ferritin
Storage
A large (~500,000 Dalton) protein
made up of 24 identical subunits and
capable of storing up to 2300-2500
iron (Fe III) ions. The ferritin core
consists of ferrihydrite (5Fe2O3.9H2O)
and phosphate. The mechanisms
underlying the neurotoxicity of
6-hydroxydopamine include its autooxidation (producing ROS) and the
reduction Fe3+ to Fe2+ (thereby
releasing active pro-oxidant Fe2+
ions).
Aisen and
Listowsky
(1980);
Crichton
(1990);
Double et al.
(1998);
Harrison and
Arosio (1996);
Theil (1987)
Iron
Hemosiderin
Storage
Derived from intralysosomal
aggregation and proteolysis of ferritin.
Thought to act as a back-up system
under conditions of iron excess.
Crichton
(1990);
Harrison and
Arosio (1996)
Iron
Neuromelanin
Storage
This complex polymer bound to
lipofuscin granules, abundant in the
nigrostriatal neuronal pathway, is
capable of acting as an iron store.
See Gerlach
et al. (1994)
and
references
therein
Copper
Ceruloplasmin
Transport
An α2-glycoprotein (~134,000
Dalton), consisting of three
homolgous (42-45,000 Dalton)
capable of binding six copper
atoms/molecule. Responsible for
binding >90% of circulating copper. It
may be able to donate copper
intracellularly for incorporation into
other copper-proteins such as
superoxide dismutase and
cytochrome oxidase. It has other
biological roles including: mobilization
of iron to transferrin; antioxidant
activity; regulation of plasma biogenic
amines; role in inflammatory
Evans (1973);
Frieden
(1986); Luza
and Speisky
(1996)
WWW.ESAINC.COM
325
response; growth promotor of certain
cells; stimulation of angiogenesis in
cornea. It also possesses ferroxidase
activity.
This protein can bind one copper
atom preferentially before complexing
with others. Occurs in portal and
general circulatory systems.
Copper
Albumin
Transport
Copper
Transcuprein
Transport
A 270,000 Dalton copper-containing
protein occurs in plasma and portal
circulation.
Copper
Metallothionein
Storage
Copper
Amino acid
Complexes
Transport/
storage
Single polypeptide chain (5000 to
6500 Dalton) containing 25-30%
cysteine residues, no disulfide bonds
or aromatic amino acid residues.
Binds a total of 11-12 copper atoms
per molecule in two sites. Under
oxidative stress redox active copper
may be released.
Found in erythrocytes and plasma
(e.g., histidine-amino acid complex).
Evans (1973);
Frieden
(1986)
Frieden
(1986); Weiss
and Linder
(1985)
Fabisiak et al.
(1999);
Frieden
(1986); Luza
and Speisky
(1996)
Evans (1973)
Table 4.2 The Binding Of Iron And Copper.
Another advantage of chelation is the ability to solubilize ions for use in biological
processes. Both Cu (II) and Fe (III) ions have very limited solubility at
physiological pH (10-9 and 10-18M, respectively). Thus without appropriate
chelation the accumulation of sufficient amounts of Fe (III) and Cu (II) ions for
normal metabolism would put excessive demands on the absorption, transport
and storage processes.
Before exploring the pro-oxidant role of iron and copper further, we must first
understand the metabolism of these metals and how nature tries to prevent
accidental release of redox active forms.
The Metabolism of Iron and Copper.
The mechanism of iron uptake and storage in mammals is a fascinating area of
study (reviewed by Meneghini (1997) and Harrison and Arosio (1996)). Under
normal conditions transferrin is in excess in relation to iron. Iron is transported in
the plasma as diferric-transferrin complex that serves to solubilize iron, protect
against iron’s pro-oxidant activity and act as a ligand for the transferrin-receptor
(TR) located in the plasma membrane (Bali et al. (1991)). The TR allows a fine
control of intracellular iron homeostasis. The Fe (III)-transferrin-TR complex is
WWW.ESAINC.COM
326
encapsulated by an endosome in which Fe (III) is reduced to the more soluble Fe
(II) form (Dancis et al. (1994); Watkins et al. (1992)). The emptied (apo)transferrin is then exported and released when the endosome fuses with the
plasma membrane (Mengheni (1997)). Inside the cell, the Fe (II) ion can follow
several routes including storage in ferritin. Ferritin chelates Fe (III) ions strongly
but only forms a weak chelation complex with Fe (II) ions.5 Thus iron entering into
ferritin as Fe (II) ions must be oxidized by the protein before being stored.
Similarly, release of iron involves the reduction of Fe (III) to Fe (II) ions. δAminolevulinic acid, ascorbic acid, nitric oxide and superoxide are capable of
reducing Fe (III) ions and releasing iron from ferritin (Oteiza et al. (1995); Reif
(1992)). It is unlikely that these compounds represent the physiological
mechanism by which iron release is controlled and this process still awaits
elucidation (Meneghini (1997)). Iron is thus stored in cells in a safe form that
cannot take part directly in the Fenton reaction. As discussed below, conditions
in which chelation is weakened may permit the metal to act as a pro-oxidant.
Apart from oxidation and subsequent storage in ferritin, Fe(II) ions can also be
transferred to sites of protein synthesis for incorporation into iron-containing
proteins, or along with Fe (III) ions, chelated by citrate, ATP, ADP (Aisen (1994).
This chelatable iron pool can affect iron regulatory protein (IRP), which is
responsible for posttranscriptional control of intracellular iron homeostasis
(Meneghini (1997) and references therein).6 Low levels of chelatable iron change
the conformation of IRP, permitting it to bind to the stem-loop structures (iron
responsive elements) of both ferritin and the transferrin receptor mRNAs. Such
binding stabilizes TR mRNA but inhibits the translation of the ferritin mRNA
(Klausner et al. (1993); Kuhn (1994); Leibold and Munro (1988)). Thus TR
synthesis is increased while ferritin synthesis is decreased in periods when
intracellular iron is low. If iron levels increase, then IRP is inactivated, TR
synthesis is decreased and ferritin production is increased. These two concerted
mechanisms help to maintain intracellular iron levels.
5
Similarly EDTA forms a stronger complex with Fe (III) [log k=25.0] than with Fe (II) [log k=14.3] as does citrate (Fe (III)
[log k=11.2], Fe (II) [log k=4.8]).
6
IRP is an interesting molecule. It is a protein with a single iron-sulfur cluster (4Fe-4S) (Figure 4.3) showing great
sequence homology with mitochondrial aconitase. Removal of the iron-sulfur forms the apo-enzyme that can bind to the
IRE of the mRNA of TR, ferritin and possibly other mRNAs too.
WWW.ESAINC.COM
327
Cys
Cys
S
S
Fe
S
S
S
Fe
S
Cys
S
Fe
Fe
OH2
Cys
S
Fe
S
Cys
Fe
S
S
S
Cys
S
Cys
[2Fe-2S] Cluster
[4Fe-4S] Cluster
CH3
HC
CH3
S
Cys
Cys
S
Fe
S
Cys
N
H3 C
N
Cys
S
OC:OCH2 CH2
SCH2
Fe
Protein
CH3
N
N
SCH2
Protein
(CH2 )2 CO2 CH3
Single Iron-Sulfur Cluster
Heme (Cytochrome C)
Figure 4.3 Structures Of Iron-Sulfur Clusters And A Heme Molecule.
Copper metabolism is different to that of iron. Copper is transported bound to
ceruloplasmin. Unlike transferrin that readily releases iron and is effectively
recycled, ceruloplasmin has to be degraded in order for it to release its copper
load. Ceruloplasmin has ferroxidase activity, oxidizing Fe (II) to Fe (III) while
simultaneously reducing oxygen to water (Eqn 4.17). This effectively competes
with spontaneous Fe (II) oxidation that is capable of producing ROS. Thus
ceruloplasmin is an antioxidant as it depresses the availability of Fe (II) that can
take part in the Fenton reaction or Fe (II) -dependent lipid peroxidation. Although
ceruloplasmin can react with hydrogen peroxide and superoxide, this is not
thought to be biologically important. Ceruloplasmin can oxidize a variety of
substrates in vitro, such as polyamines and polyphenols, but again the
physiological relevance, if any, of these reactions is unclear at present.
4Fe2+ + O2 + 4H+ → 4Fe3+ + 2H2O
WWW.ESAINC.COM
Eqn 4.17
328
The intracellular trafficking of copper in mammalian cells is highly regulated. Cells
have developed a number of mechanisms to ensure copper’s proper intracellular
transport and compartmentalization. The passage of copper into a cell begins
with its reduction by one of several plasma membrane reductases followed by its
transport across the membrane by a high-affinity copper transporter (Valentine
and Gralla (1997) and references therein). Once inside the cell several different
proteins (or metal chaperones) bind and deliver copper to specific intracellular
proteins (Valentine and Gralla (1997)). Thus intracellular free copper is
undetectable (Rae et al. (1999)). Cox17 in conjunction with the mitochondrial
proteins SCO1 and SCO2 specifically deliver copper to mitochondrial cytochrome
oxidase (Glerum et al. (1996a,b)). Lys7 (yeast) and CCS (human) deliver copper
specifically to CuZn-SOD (Cullota et al. (1997, 1999); Rothstein et al. (1999)).
Atx1 (yeast) and Hah1 (human) direct copper to a post-Golgi- compartment via a
protein Ccc2 (yeast) or Wilson disease protein (human) (Valentine and Gralla
(1997) and references therein). Similar mechanisms are now being proposed for
iron transport and compartmentalization.
Iron and Copper Species as Pro-oxidants.
Iron and copper can occur in several forms in vivo, but not all forms act as prooxidants. The role of these metals in the production of hydroxyl free radicals was
discussed in Chapter 2, while their ability to initiate lipid peroxidation was covered
in Chapter 3. In general many more forms of iron can stimulate lipid peroxidation
than hydroxyl free radical formation (Halliwell and Gutteridge (1999)). This is
because lipid peroxidation need not proceed through the formation of the
hydroxyl free radical (Chapter 3). The importance of iron and copper participation
in pro-oxidant reactions is a complex area that has become controversial
because of varying assay conditions being used7, the form of the metal being
studied, and which pro-oxidant activity is being measured (Halliwell and
Gutteridge (1999)). Overall it appears that under normal physiological conditions,
apart from the low molecular mass iron pool that is used for synthesis of ironcontaining proteins, redox active metals are tightly bound in vivo and do not act
as pro-oxidants. However, problems do occur under disease conditions (e.g.,
metal storage diseases, hemochromatosis, and acute porphyria) or with
overexposure to metals (e.g., through the diet or in the treatment of
thalassemias) (Britton and Brown (1995); Evans (1973); Gerlach et al. (1994);
Halliwell and Gutteridge (1999); Houglum et al. (1997); Liochev (1999); Livrea et
al. (1996); Muller-Hocker et al. (1987); Olivares and Uauy (1996)).
Superoxide can reduce Fe (III) to Fe (II) contained within a number of proteins or
free in solution. For example, aconitase, which exists in both mitochondrial and
cytosolic forms, reacts with superoxide to release iron from its 4Fe-4S cluster.
This process is reversible in vivo. But it is the effect of superoxide on ferritin,
7
Laboratory reagents can be contaminated by free “reactive” iron. For example, bottles of old saline can contain dissolved
iron, a form that could produce ROS under test conditions and lead to erroneous data (Halliwell and Gutteridge (1993)).
WWW.ESAINC.COM
329
however, that is quantitatively the most important (McCord (1998)). The Fe (II) so
formed can then take part in both Fenton reaction and lipid peroxidation.
Measurement of Iron and Copper.
Iron and copper can be measured using atomic absorbance and anodic stripping
voltammetry. However, these approaches are usually not sufficient to measure
the low levels of redox active metals in vivo.
Iron can be measured using the bleomycin method developed by Halliwell and
Gutteridge (1999) (see also Chevion et al. (1999); Gutteridge and Halliwell
(1999)). Here the ability for the antibiotic bleomycin to damage DNA is directly
proportional to the amount of labile iron ions. Other assays include EPR-based
approaches and the use of fluorescent probes such as calcein (Chevion et al.
(1999); Gutteridge and Halliwell (1999)). The latter is the only method for the
direct determination of cellular pool of labile iron.
Copper can be measured using the phenanthroline assay. Here 1,10phenathroline in the presence of oxygen, a suitable reducing agent and copper
can lead to DNA degradation. Such DNA damage can then be detected using a
TBAR approach (see Chapter 3) (Chevion et al. (1999); Gutteridge and Halliwell
(1999)).
LOW MOLECULAR WEIGHT MOLECULES.
Humans utilize a wide range of antioxidants that are either synthesized de novo
or obtained from the diet; many of the latter are the classical vitamins.
Antioxidants have diverse chemical structures, but for convenience are normally
separated based upon their solubility. This section examines the antioxidant (and
pro-oxidant) activities of both water- and fat-soluble antioxidants. Many other
endogenous and exogenous compounds also show antioxidant activity and these
will be discussed too. Within each section we will examine the antioxidants
alphabetically. This does not reflect either their level or importance.
Water-Soluble Antioxidants.
Albumin.
Albumin is a water-soluble protein abundant in serum and plasma (380-920µM
human serum 535-760µM human plasma) (Table 4.3). Albumin can bind copper
(Table 4.2) thereby preventing copper-ion dependent hydroxyl free radical
formation and lipid peroxidation. However, hydroxyl free radical formation can
continue at the metal binding site, resulting in damage to the albumin molecule.
This is probably biologically insignificant, as albumin is rapidly turned-over.
WWW.ESAINC.COM
330
Copper binding also diverts copper’s reactivity away from significant targets such
as essential sulfhydryl groups of enzymes, thus limiting oxidative damage.
Albumin transports a variety of compounds such as tryptophan, fatty acids,
bilirubin and certain drugs. It has been hypothesized that albumin protects fatty
acids from lipid peroxidation not only by interfering with copper’s pro-oxidant
effect but also by assimilating the antioxidant bilirubin (Halliwell and Gutteridge
(1990)).
Albumin is an efficacious scavenger of hypochlorous acid (Halliwell (1998); Wasil
et al. (1987)).
Compound
Species
Tissue
Range
Albumin
Human
Plasma
Allantoin
Allantoin
Human
Human
Allantoin
Human
Ascorbic acid
Ascorbic acid
Human
Human
Plasma - control
Serum - control
Serum - rheumatoid
Synovial fluid rheumatoid
Muscle
Plasma
Plasma
Plasma
Ascorbic acid
Human
Plasma
Ascorbic acid
Human
Ascorbic acid
Human
Plasma
CSF
Plasma
Ascorbic acid
Human
Plasma
10 to 110 µmol/L
21.5+4.4
65 to 164 µmol/L
neonates
27.8±1.32 µmol/L
Ascorbic acid
Human
Plasma
23.2±17.3 µmol/L
Ascorbic acid
Human
Plasma
10 to 90 µmol/L
Ascorbic acid
Human
Plasma
5.68 to 85 µmol/L
Ascorbic acid
Ascorbic acid
Human
Mouse
Saliva
Liver
0 to 3.7 mg/L total
1.00+0.05 µmol/g
Ascorbic acid
Rat
9.7 to 15.4 µmol/L
20 to 40 µmol/L
Ascorbic acid
Rat
Ascorbic acid
Rat
Brain - ECF
Ventricularmyocardium
Brain
Liver
Kidney
Lens
WWW.ESAINC.COM
370 to 530 µmol/L
neonates,
535 to 760 adults
12.4 to 20.6 µmol/L
14.1 to 25.4 µmol/L
20.3 to 45.2
7.2 to 31.3
0.03 µmol/g
11.9 µmol/L
50+20 µmol/L
61.8 µmol/L
(2.3 µmol/L DHAA)
40 to 140 µmol/L
1.8 µmol/g
1.6
0.4
30.3±7.7 ng/mg
Reference
Gopinathan et al.
(1994)
Lux et al. (1992)
Grootveld and
Halliwell (1987)
Hellsten et al.
(1997)
Koh et al. (1980)
Schorah et al.
(1996)
Koch et al.
(1980)
Nagy and Degrell
(1989)
Gopinathan et al.
(1994)
Lykkesfeldt
(1995)
Nagy and Degrell
(1989)
Levine et al.
(1996)
Capellmann et
al. (1994)
Karp (1990)
Barja de Quiroga
et al. (1991)
Tsai et al. (1996)
Schell and Bode
(1993)
Mitton and
331
Ascorbic acid
Rat
Liver
Kidney
Pancreas
Colon
Bilirubin
Human
Plasma
α-Carotene
Human
Plasma
(ascorbate + DHAA)
5.4±0.65 nmol/g
3.2±0.41
0.32±0.58
4.1±0.19
20 to 126 µmol/L –
neonates,
<20 – adults
0.055+0.053 µmol/L
α-Carotene
Human
Plasma
0.02 to 0.22 µmol/L
α-Carotene
Human
Plasma
0.08 to 0.19 µmol/L
β-Carotene
Human
Plasma
0.13 to 0.33 µmol/L
β-Carotene
Human
Plasma
0.3 to 0.6 µmol/L
β-Carotene
Human
Plasma
0.13 to 1.3 µmol/L
β-Carotene
β-Carotene
Human
Human
Plasma - neonate
Plasma
0.02 to 0.05 µmol/L
0.182+0.084 µmol/L
β-Carotene
Human
Plasma
0.07 to 0.88 µmol/L
β-Carotene
Human
Plasma
1.81+1.57 µmol/L
β-Carotene
(trans)
β-Carotene
(13-cis)
β-Carotene
(trans)
Human
Serum
0.322+0.259 µmol/L
Human
Serum
0.016+0.011 µmol/L
Human
β-Carotene
(9-cis)
Human
β-Carotene
(13-cis)
Human
γ-Carotene
Human
Serum
Liver
Adrenal
Serum
Liver
Adrenal
Serum
Liver
Adrenal
Plasma
131 to 703 nmol/L
1.4 to 7.3 nmol/g
1.8 to 4.4 nmol/g
n.d.
0.4 to 2.1 nmol/g
0.1 to 0.3 nmol/g
8 to 38 nmol/L
0.1 to 0.9 nmol/g
0.2 to 1.1 nmol/g
144 to 277 nmol/L
α-Cryptoxanthin
Human
Plasma
39 to 130 nmol/L
β-Cryptoxanthin
Human
Plasma
149 to 371 nmol/L
β-Cryptoxanthin
Human
Plasma
0.05 to 0.52 µmol/L
β-Cryptoxanthin
Cysteine
Human
Human
Plasma - neonate
Plasma
0.51 to 0.63 µmol/L
9.0+1.1 µmol/L
Cysteine
Human
Plasma
2.7±1.8 µmol/L
WWW.ESAINC.COM
Trevithick (1994)
Rose and Bode
(1995)
Gopinathan et al.
(1994)
Milne and
Botnen (1986)
Sowell et al.
(1994)
Khachik et al.
(1992)
Khachik et al.
(1992)
Motchnik et al.
(1994)
Winklhofer-Roob
et al. (1997)
Finckh (1995)
Milne and
Botnen (1986)
Sowell et al.
(1994)
Yamashita and
Yamamoto
(1997)
Stahl et al.
(1993)
Stahl et al.
(1993)
Stahl et al.
(1993)
Stahl et al.
(1993)
Stahl et al.
(1993)
Khachik et al.
(1992)
Khachik et al.
(1992)
Khachik et al.
(1992)
Sowell et al.
(1994)
Finckh (1995)
Andersson et al.
(1995)
Velury and
Howell (1988)
332
Cyst(e)ine
Human
Blood
42.0 to 123 µmol/L
Cysteine
Rat
Liver
0.050±0.032 µmol/g
DHAA
Human
Plasma
Glutathione
Human
Blood
Plasma
Glutathione
Human
Brain
Glutathione
Human
Brain
Glutathione
Human
Erythrocytes
Glutathione
Human
Erythrocyte
Glutathione
Human
Hair
Glutathione
Human
Plasma
Glutathione
Human
Plasma neonate
Glutathione
Human
Whole blood
5.8+2.7 µmol/L
5.8+5.1
849+63 µmol/L –
GSH
3.39+1.04 – GSH
0.3+0.04 µmol/g –
GSH
3.75+0.98 nmol/g –
GSSG
0.19+0.04 µmol/g –
GSH
3.26+0.65 nmol/g –
GSSG
2.02+0.1 µmol/mL
cell –
GSH
6.8+0.8 – GSSG
19.6 to 23.2
nmol/million RBC –
GSH
5+3.6 nmol/µg DNA –
GSH
0.096+0.08 – GSSG
1.0±0.8 µmol/L –
GSH
2.4 to 3.4 µmol/L
GSH
0.4 to 0.6 GSSG
0.9 to 1.7 mmol/L
Glutathione
Cow
Eye - lens
Glutathione
Guinea pig
Brain
Liver
Glutathione
Horse
Hemolysate
Glutathione
Horse
Bronchial lavage
Hemolysate
Glutathione
Mouse
Bronchial lavage
Liver
Glutathione
Mouse
Trachea
Glutathione
Rat
Bile
WWW.ESAINC.COM
6.79+0.4 µmol/g GSH
<0.048 GSSG
0.96+0.03 PSSG
1.68±0.03 nmol/mg
protein – GSH
6.69±0.23 nmol/mg
protein – GSH
720 to 1190 µmol/L –
GSH
0.9 to 1.2 – GSH
33 to 114 µmol/L –
GSSG
0.06 to 0.1 – GSSG
10+1 µmol/g – GSH
200+110 nmol/g –
GSSG
2.71±0.71 nmol/mg
protein – GSH
2.9+0.54 µmol/g –
GSH
Richie et al.
(1996)
Demaster et. al.
(1984)
Nagy and Degrell
(1989)
Michelet et al.
(1995)
Sofic et al.
(1992)
Sofic et al.
(1991)
Kuninori and
Nishiyama
(1991)
Rabenstein and
Saetre (1978)
Krien et al.
(1992)
Velury and
Howell (1988)
Smith et al.
(1996)
Rabenstein and
Saetre (1978)
Ozcimder et al.
(1991)
Mefford and
Adams (1978)
Smith et al.
(1995)
Smith et al.
(1995)
Carro-Ciampi et
al. (1988)
Lakritz et. al.
(1997)
Carro-Ciampi et
al. (1988)
333
Glutathione
Rat
Brain
Glutathione
Rat
Lens
325+168 nmol/g –
GSSG
650+205 µmol/L –
GSH
218+70 – GSSG
5.67+0.6 µmol/g –
GSH
0.284+0.03 – GSSG
1.77+0.4 µmol/g –
GSH
0.149+0.05 – GSSG
2.25±0.03 nmol/mg
protein – GSH
3.9 µmol/g – GSH
Glutathione
Rat
Lens
5.2 µmol/g – GSH
Glutathione
Rat
Liver
Glutathione
Rat
Liver
Glutathione
Rat
Liver
Kidney
Pancreas
Colon
Glutathione
Rat
Lung
Homocysteine
Human
Plasma
Homocysteine
Human
Plasma
4.07 ± 0.21 µmol/g –
GSH
4 to 6 µmol/g – GSH
20 to 650 nmol/g –
GSSG
6.5±0.39 nmol/g
3.8±0.22
0.61±0.02
0.26±0.01
1.31+0.6 µmol/mg –
GSH
0.04+0.02 – GSSG
11.3±2.96 µmol/L –
males
8.8±2.7 µmol/L –
females
1.8±1.2 µM
Homocysteine
Human
Plasma
0.14+0.03 µmol/L
Homocysteine
Human
CSF
0.210+0.028 µmol/L
Homocyst(e)ine
Homocyst(e)ine
Human
Human
Plasma
Plasma
13.4+4.8 µmol/L
12.4+2.9 µmol/L
Homocysteine
Rat
Liver
0.087±0.006 µmol/g
Hypoxanthine
Human
CSF
5.94+0.74 µmol/L
Hypoxanthine
Human
CSF
Hypoxanthine
Human
Hypoxanthine
Human
Plasma
Erythrocytes
Urine
Serum
1.8 to 5.5 µmol/L –
neonate
0.6 to 5.1 µmol/L –
adult
2.5+1 µmol/L
8.0+6.2 µmol/L
48+26 µmol/24h
1.7 to 16.9 µmol/L
Glutathione
Rat
Blood
Liver
Kidney
WWW.ESAINC.COM
Asensi et al.
(1994)
Mefford and
Adams (1978)
Iriyama et al.
(1986)
Mitton and
Trevithick (1994)
Demaster et. al.
(1984)
Harvey et al.
(1989)
Rose and Bode
(1995)
Martin and White
(1991)
Wu et. al. (1994)
Velury and
Howell (1988)
Andersson et al.
(1995)
Quinn et al.
(1997)
Wu et al. (1994).
Fermo et al.
(1992)
Demaster et. al.
(1984)
Castro-Gago et
al. (1986)
Harkness and
Lund (1983)
Boulieu et al.
(1984)
Kock et al.
334
Hypoxanthine
Rat
CSF
100 to 840 nmol/L
Hypoxanthine
Rat
CSF
80 to 310 nmol/L
Lipoic acid
Human
Plasma hydrolysates
Lutein/
zeaxanthin
Lutein
Human
Lens
58 to 208 nmol/L (ox)
160 to 703 (red)
0.02+0.001 nmol/g
Human
Plasma
135 to 265 nmol/L
Lycopene
Human
Plasma
0.65+0.36 µmol/L
Menaquinone-4
Human
Milk
2.9+2.3 nmol/L
Menaquinone-4
Human
Phylloquinone
Human
Plasma maternal
Plasma umbilical
Placenta
Milk
0.11 nmol/L
0.09 nmol/L
2.66 pmol/g
4.6+2.0 nmol/L
Phylloquinone
Human
Phylloquinone
Human
Plasma
Liver
Plasma
1.19+0.16 nmol/L
28+4 pmol/g
0.29-2.64 nmol/L
Phylloquinone
Human
Phylloquinone
Human
Plasma maternal
Plasma umbilical
Placenta
Serum
3.4 nmol/L
0.02 nmol/L
2.62 pmol/g
20+11 nmol/L
Phylloquinone
Human
Serum
0.09-1.96 nmol/L
Retinol
Human
Lens
0.133+0.014 nmol/g
Retinol
Human
Plasma
0.3 to 3.0 µmol/L
Retinol
Human
Plasma
1.29 to 2.99 µmol/L
Retinol
Retinyl esters
Mousehairless
Human
Plasma
Liver
Lens
0.67+0.05 µmol/L
14+3.5 nmol/g
21 to 25 ng/g
Retinyl esters
Human
Plasma
0.03 to 0.38 µmol/L
Retinyl palmitate
Liver
240+62 nmol/g
Cardiac mitochondria
α-Tocopherol
Mousehairless
Cow
Mouse
Rat
Chicken
Tocopherols
Human
0.02+0.01
0.37+0.02
0.33+0.04
19+4 pmol/mg*
4+0.8
79+18
35 to 99
10 to 40 µmol/L
α-Tocopherol
Plasma
Erythrocyte
Liver
Muscle
Plasma
WWW.ESAINC.COM
(1994)
Walter et al.
(1988)
Polasek et al.
(1989)
Teichert and
Preiss (1992)
Yeum et al.
(1995)
Khachik et al.
(1992)
Yamashita and
Yamamoto
(1997)
Isshiki et al.
(1988)
Hiraike et al.
(1988)
Isshiki et al.
(1988)
Usui et al. (1990)
Sadowski et al.
(1989)
Hiraike et al.
(1988)
Sann et al.
(1985)
Moussa et al.
(1989)
Yeum et al.
(1986)
Winklhofer-Roob
et al. (1997)
Khachik et al.
(1992)
Savoure et al.
(1995)
Yeum et. al.
(1995)
Sowell et al.
(1994)
Savoure et al.
(1995)
Lass and Sohal
(1999)
Murphy and
Kehrer (1987)
Farrell et al.
335
α-Tocopherol
Human
Erythrocyte
3.92 to 6.99 µmol/L
α-Tocopherol
Human
Lens
α-Tocopherol
Human
Plasma
3.65+0.39 nmol/g
3.6 to 5.9 nmol/g
10 to 34 µmol/L
α-Tocopherol
Human
Plasma
28 to 49 µmol/L
α-Tocopherol
α-Tocopherol
Human
Human
Plasma - neonate
Serum
0.65 to 1.23 µmol/L
9.98 to 22.5 µmol/L
α-Tocopherol
α-Tocopherol
Human
Human
α-Tocopherol
Mouse hairless
α-Tocopherol
Rat
27.9±10 µmol/L
20.3±0.5 µmol/L
13.4±0.3 µmol/L
5.4+0.1 nmol/g
24.2+1.1
21.9+0.6
21.2+2.9
5.4+0.2
5.6+0.1 µmol/L
α-Tocopherol
Rat
Serum
Plasma
Blood
Brain
Heart
Kidney
Liver
Skin
Erythrocyte
membrane
Lens
0.3+0.05 nmol/g
α-Tocopherol
Rat
α-Tocopherol
Rat
Liver
Muscle
Plasma
78.5±2.3 nmol/g
21.7±0.5
10+0.1 µmol/L
β-Tocopherol
γ-Tocopherol
Human
Chicken
γ-Tocopherol
Human
Plasma - neonate
Plasma
Erythrocyte
Liver
Muscle
Lens
γ-Tocopherol
Human
Plasma
8.47 to 19.0 µmol/L
2.3+1.0 pmol/mg*
0.64+0.18
19+4
8 to 25
0.62 to 1.20 nmol/g
0.85+0.14 nmol/g
0.2 to 1.3 µmol/L
γ-Tocopherol
Human
Plasma
5 to 15 µmol/L
γ-Tocopherol
Mouse hairless
δ-Tocopherol
Human
Brain
Heart
Kidney
Liver
Skin
Erythrocyte
0.01+0.02 nmol/g
0.19+0.05
0.35+0.06
0.29+0.03
0.04+0.00
0.05 to 0.13 µmol/L
α-Tocopherylquinone
Chicken
α-Tocopherylquinone
α-Tocopherylquinone
Human
Plasma
Erythrocyte
Liver
Muscle
Erythrocyte
0.21+0.08 nmol/g*
0.82+0.12
1.47+1.1
0.07 to 0.49
0.04 to 0.07 µmol/L
Rat
Plasma
0.053+0.011 µmol/L
WWW.ESAINC.COM
(1978)
Vatassery et al.
(1993)
Yeum et al.
(1995)
Winklhofer-Roob
et al. (1997)
Khachik et al.
(1992)
Finckh (1995)
Chou et al.
(1985)
Edlund (1988)
Lang et. al.
(1986)
Pods et al.
(1996)
Takeda et al.
(1996)
Mitton and
Trevithick (1994)
Lang et. al.
(1986)
Takeda et al.
(1996)
Finckh (1995)
Murphy and
Kehrer (1987)
Yeum et. al.
(1995)
Winklhofer-Roob
et al (1997)
Khachik et al.
(1992)
Pods et al.
(1996)
Vatassery et al.
(1993)
Murphy and
Kehrer (1987)
Vatassery et al.
(1993)
Takeda et al.
(1996)
336
α-Tocopherylquinone
γ-Tocopherylquinone
Rat
Chicken
α-Tocotrienol
Mouse hairless
γ-Tocotrienol
Mouse hairless
Ubiquinone total
Human
Ubiquinol-8
Rat
Ubiquinol-8
Rat
Ubiquinone-8
Rat
Ubiquinol-9
Human
Ubiquinol-9
Mouse hairless
Ubiquinol-9
Rat
Ubiquinol-9
Rat
Ubiquinol-9
Rat
Ubiquinone-9
Cow
Mouse
Pig
Erythrocyte
membrane
Plasma
Erythrocyte
Liver
Muscle
Brain
Heart
Kidney
Liver
Skin
Brain
Heart
Kidney
Liver
Skin
Plasma - control
endurance athletes
hyperthyroid
hypothyroid
hypercholesterolemic
Brain
Liver
Kidney
Plasma
Liver
Heart
Kidney
Brain
Heart
Kidney
Liver
Serum
Serum
Brain
Heart
Kidney
Liver
Skin
Brain
Heart
Kidney
Liver
Serum
Liver
Muscle
Plasma
Liver
Heart
Kidney
Cardiac mitochondria
WWW.ESAINC.COM
0.36+0.14 µmol/L
0.16+0.06 pmol/mg*
0.21+0.08
0.09+0.03
0.004 to 0.49
n.d.
0.08+0.01 nmol/g
0.06+0.04
0.10+0.04
0.24+0.20
n.d.
0.19+0.05 nmol/g
0.15+0.07
0.19+0.16
0.76+0.71
0.80+0.20 mg/L
0.58+0.17
0.27+0.13
0.62+0.11
1.15+0.15
3.02+0.9 nmol/g
13.2+2.9
1.5+0.27
82.5+13.8nmol/L
3.02+0.8 nmol/L
1.0+0.12
0.61+0.04
2.88+0.41 nmol/g
11.5+2.6
2.33+0.27
14.6+3
101+22
46.4+7.8 nmol/L
1.6+0.1 nmol/g
19+4
81+29
42+16
2.2+0.3
24.2+5.8 nmol/g
26.2+4.8
80+18
85+24
44+9 nmol/L
121.5±11.6 nmol/g
4.9±0.4
0.46+0.07 µmol/L
84.4+0.5 nmol/g
21+0.6
23+5
0.12+0.01 µmol/g
6.01+0.05
0.18+0.01
Takeda et al.
(1996)
Murphy and
Kehrer (1987)
Pods et al.
(1996)
Pods et al.
(1996)
Grossi et al.
(1992)
Wakabayashi et
al. (1994)
Okamoto et. al.
(1988)
Wakabayashi et
al. (1994)
Wakabayashi et
al. (1994)
Pods et al.
(1996)
Wakabayashi et
al. (1994)
Lang et. al.
(1986)
Okamoto et. al.
(1988)
Lass and Sohal
(1999)
337
Ubiquinone-9
Rabbit
Rat
Human
Brain
Heart
Kidney
Liver
Brain
Heart
Kidney
Liver
Skin
Muscle
0.10+0.01
5.18+0.01
1.25+0.13 µmol/g
3.14+0.4
4.15+0.25
2.26+0.3
10.2+0.5 nmol/g
245+22
302+124
46+18
7.6+1.9 nmol/L
31.9±3.2 nmol/g
24+5.8 nmol/g
240+44
126+36
152+44
593+94 nmol/L
47+4 nmol/g
254+23
156+14
165+18
0.4 to 1.0 µmol/L
Ubiquinone-9
Mouse hairless
Ubiquinone-9
Rat
Ubiquinone-9
Rat
Ubiquinone-9
Rat
Ubiquinol-10
Human
Brain
Heart
Kidney
Liver
Serum
Brain
Heart
Kidney
Liver
Plasma
Ubiquinol-10
Human
Serum
1.59+0.2 µmol/L
Ubiquinol-10
Ubiquinol-10
Human
Human
Plasma - neonate
Plasma
0.48 to 0.60 µmol/L
0.927+0.214 µmol/L
Ubiquinol-10
Human
Plasma
1.14 µmol/L (male)
0.56 (female)
50.86+16 nmol/g
creatinine (male)
57.9+24 (female)
0.6+0.1 nmol/g
2.8+0.7
11+6
1.7+0.3
0.4+0
15.2±1.4 nmol/g
Urine
Ubiquinol-10
Mouse hairless
Ubiquinol-10
Rat
Ubiquinol-10
Rat
Plasma
Liver
Heart
Kidney
Ubiquinone-10
Cow
Mouse
Pig
Rabbit
Rat
Human
Human
Cardiac mitochondria
Ubiquinone-10
Ubiquinone-10
Brain
Heart
Kidney
Liver
Skin
Liver
Plasma - neonate
Plasma
WWW.ESAINC.COM
0.11±0.011 µmol/L
13.3±1.7 nmol/g
2.2±0.09
5.1±0.9
6.51+0.02 µmol/g
0.71+0.03
5.79+0.05
4.78+0.08
0.65+0.02
0.03 µmol/L
0.041+0.012 µmol/L
Aberg et al.
(1992)
Pods et al.
(1996)
Lang et. al.
(1986)
Wakabayashi et
al. (1994)
Aberg et al.
(1992)
Motchnik et al.
(1994)
Wakabayashi et
al. (1994)
Finckh (1995)
Yamashita and
Yamamoto
(1997)
Okamata et. al.
(1988)
Pods et al.
(1996)
Lang et. al.
(1986)
Okamata et. al.
(1988)
Lass and Sohal
(1999)
Finckh (1995)
Yamashita and
Yamamoto
338
Ubiquinone-10
Human
Serum
0.82+0.5 µmol/L
Ubiquinone-10
Human
Serum
0.68 to 1.68 µmol/L
Ubiquinone-10
Ubiquinone-10
Human
Human
Ubiquinone-10
Mouse hairless
Ubiquinone-10
Rat
Ubiquinone-10
Rat
Uric acid
Human
0.28 to 0.95 µmol/L
15.5+1.2 nmol/g
132+10.7
77+7.7
64+4.8
3.4+0.5 nmol/g
21+8
31+14
n.d.
n.d.
2.7±0.5 nmol/g
2.2±0.4
21+1.7 nmol/g
19.5+2
25.5+2
25+2
258 to 621 µmol/L
273 to 485
123 to 351
Uric acid
Human
Uric acid
Human
Serum
Brain
Heart
Kidney
Liver
Brain
Heart
Kidney
Liver
Skin
Liver
Muscle
Brain
Heart
Kidney
Liver
Serum - control
Serum - rheumatoid
Synovial fluid rheumatoid
Plasma
Urine
Plasma
Uric acid
Human
Uric acid
Human
Uric acid
Human
Uric acid
Human
Muscle
Plasma
Plasma
Uric acid
Mouse
Liver
Uric acid
Rat
Liver
Kidney
Pancreas
Colon
Uric acid
Rat
CSF
0.13±0.01 nmol/g
0.40±0.01
0.24±0.01
1.1±0.06
0.1 to 3.6 µmol/L
Uric acid
Rat
CSF
3.55 to 19.0 µmol/L
Xanthine
Human
CSF
5.2+0.87 µmol/L
Xanthine
Human
CSF
0.9 to 9.1 µmol/L neonate
Plasma - control
Plasma premenopausal
females
Serum
WWW.ESAINC.COM
120 to 360 µmol/L
1500 to 3600
140 to 600 µmol/L
160 to 450 µmol/L
120 to 340
185 to 486 µmol/L
0.26 µmol/g
305 µmol/L
191 to 413 µmol/L
neonates,
180 to 420 adults
130+5 nmol/g
(1997)
Wakabayashi et
al. (1994)
Ikenoya et. al.
(1979)
Edlund (1988)
Aberg et al.
(1992)
Pods et al.
(1996)
Lang et. al.
(1986)
Aberg et al.
(1992)
Grootveld and
Halliwell (1987)
Boulieu et al.
(1984)
Benzie and
Strain (1996)
Wyngaarden and
Kelly (1976)
Kock et al.
(1994)
Hellsten et al.
(1997)
Gopinathan et al.
(1994)
Barja de Quiroga
et al. (1991)
Rose and Bode
(1995)
Walter et al.
(1988)
Polasek et al.
(1989)
Castro-Gago et
al. (1986)
Harkness and
Lund (1983)
339
Rat
Serum
Plasma
Erythrocytes
Urine
CSF
0.6 to 4.7 µmol/L adult
0.2 to 6.2 µmol/L
1.4+0.7 µmol/L
<0.5 µmol/L
68+42 µmol/24h
2.6 to 8.6 µmol/L
Xanthine
Rat
CSF
0.21 to 2.9 µmol/L
Xanthophylls
Human
Lens
11 to 25ng/g
Zeaxanthin
Human
Plasma
35 to 50 nmol/L
Xanthine
Xanthine
Human
Human
Xanthine
Kock et al.(1994)
Boulieu et al.
(1984)
Walter et al.
(1988)
Polasek et al.
(1989)
Yeum et. al.
(1995)
Khachik et al.
(1992)
Table 4.3 Levels Of Antioxidants Reported In The Literature.
Ascorbic Acid.
Most animals can synthesize ascorbic acid (vitamin C) from glucose in the liver
(Banhegyi et al. (1997)). Man, primates, guinea pigs and fruit bats, however, do
not possess L-gulonolactone oxidase, the terminal enzyme in the pathway of
ascorbic acid biosynthesis. For these mammals, ascorbic acid must be derived
from the diet. Thus diets deficient in ascorbic acid will lead to disease (e.g.,
scurvy) in these organisms.
Ascorbic acid plays several important metabolic roles (reviewed by Levine
(1986); Levine et al. (1996)). For example, it is a cofactor for enzymes involved in
both post-translational modification of collagen (prolyl and lysyl hydroxylases)
and catecholamine synthesis (dopamine-β-hydroxylase) (Sauberlich (1994),
Udenfriend et al. (1954)). Furthermore, it may act as a neuromodulator in the
brain (Grunewald (1993)). Ascorbic acid plays key roles in the regulation of
absorption of iron and its cellular metabolism (Dorey et al. (1993); Hoffman et al.
(1991); Toff and Bridges (1995)). It may also prevent stomach cancer by helping
eliminate nitrosamines derived from the diet or formed in the stomach from
secondary amines (Chapter 2) (Block (1991); Cohen and Bhagavan (1995); Dyke
et al. (1994)).
Antioxidant Properties.
Ascorbic acid is a very important antioxidant (Halliwell (1996); Rose and Bode
(1993)). Halliwell (1996) has summarized the various in vitro antioxidant
properties of ascorbic acid and an updated version is presented in Table 4.4.
Although data supporting the antioxidant role of ascorbic acid in vitro is
overwhelming, in vivo data are much more scarce. Ascorbic acid does fulfill the
WWW.ESAINC.COM
340
criteria used to establish whether a compound is a suitable antioxidant candidate
(see above and Rose and Bode (1993)):
• It is present in adequate amounts in the body (Table 4.3));
• It is versatile and readily oxidized;
• It is compartmentalized (e.g., it is particularly abundant in adrenal
chromaffin granules and neuronal monoamine vesicles where it prevents
oxidation of the monoamines);
• It is readily available;
• It is conserved by the kidneys; and
• It has tolerable toxicity.
Comments
Scavenges peroxynitrite (rate constant 2.4 x 102
M-1.s-1)
Scavenges superoxide and HO2•
(rate constants 1.0 x 104 to >105 M-1.s-1)
Scavenges hydroxyl free radicals (rate constant
>109 M-1.s-1)
Scavenges thiyl and sulphenyl radicals
Scavenges hypochlorous acid. Protects against
chloramine-dependent modifications to LDL
Scavenges ozone and nitrogen dioxide
Scavenges some flavonoid radicals but may be
oxidized by other flavonoids
Scavenges and quenches singlet oxygen
May regenerate α-tocopherol from α-tocopheroxyl
radical in membranes (and lipoproteins)
Protects plasma lipids from peroxidation induced
by activated neutrophils
Inhibits lipid peroxidation induced by hemoglobin
(or myoglobin)-H2O2 mixtures and prevents
peroxide-induced release of iron from heme
Protects neutrophils from self inflicted oxidative
stress
Reference
Bartlett et al. (1995); Vasquez-Vivar et al.
(1996); Whiteman and Halliwell (1996)
Cabelli and Bielski (1983); Halliwell and
Gutteridge (1999); Nandi and Chatterjee
(1987); Nishikimi (1975); Radi et al.
(1991)
Anbar and Neta (1987); Bartlett et al.
(1994)
Asmus (1987); Sevilla et al. (1989)
Carr et al., (2000); Folkes et al. (1995);
Halliwell et al. (1987)
Cross et al. (1994)
Bors et al. (1995)
Chou and Khan (1983)
Beyer (1994); Bisby and Parker (1995);
Buettner (1993); Cadenas et al. (1996);
Chan (1993); Esterbauer et al. (1989);
Kagan et al. (1992a,b); Liebler et al.
(1986); Mehlhorn et al. (1989); Muckai et
al. (1992); Niki et al. (1984); Packer
(1994); Packer et al. (1979); van den
Berg et al. (1990); Wefers and Sies
(1988). But see Glascott and Farber
(1998) and references therein.
Frei et al. (1989)
Rice-Evans et al. (1989)
Wang et al. (1997)
Table 4.4 Evidence Supporting The Role For Ascorbic Acid As An
Antioxidant In Vitro.
WWW.ESAINC.COM
341
HO
O
O
-H+
OH
HO
HO
HO
O
O
O
O
-e -
OH
OH
OH
HO
O
O-
SEMIDEHYDRO-LASCORBIC ACID
L-ASCORBATE ANION
L-ASCORBIC ACID
O
-e -
HO
H
OH
H
OH
O
O
O
O
O
O
HO
OH
OH
+H2O
O
O
OH
O
OH
-H2O
O
O
DEHYDRO-L-ASCORBIC
DEHYDRO-L-ASCORBIC
ACID
ACID HYDRATE
+H2 O
GLUCOSE-6PHOSPHATE
+
H
O
LACTIC ACID
OH
HO
O
O
OH
CH 2 OH
H
2,3-DIKETO-L-GULONIC ACID
L-THREONIC ACID
+
OXALIC ACID
Figure 4.4 Chemical Structure Of Ascorbate And Relationship To Its
Oxidation Products. (Antioxidant pathway is presented in red; metabolism in blue).
Furthermore, evidence suggests that ascorbic acid is depleted during oxidative
stress, e.g., in patients with rheumatoid arthritis (Blake et al. (1981); Lunec and
Blake (1985)), adult respiratory distress syndrome (Cross et al. (1990)) and
preeclampsia (Huber et al. (1997)). Recently Wang et al. (1997) reported that
ascorbic acid recycling in stimulated neutrophils is an important antioxidant
defense mechanism protecting them from damage by their own pro-oxidant
molecules.
When ascorbic acid reacts with a more aggressive radical the result is the
production of an intermediate radical (ascorbyl) of low reactivity (Figure 4.4). The
lower activity comes from the ability of ascorbate to delocalize the radical
electron throughout its π-system (Chapter 1). As can be seen from Table 2.1.1,
WWW.ESAINC.COM
342
the redox potential (Eo) of ascorbic acid is low and it is therefore a good reducing
agent (antioxidant). Consequently ascorbic acid will quench the hydroxyl free
radical, lipid peroxyl radical, uric acid radical and tocopheroxyl radical. Ascorbic
acid may be involved in the regeneration of tocopherol from the tocopheroxyl
radical formed during the prevention of lipid peroxidation (Figure 4.5). Although
there is abundant in vitro evidence for this interaction (see Table 4.4) little in vivo
evidence exists. In fact recent data suggest that ascorbic acid acts as an
antioxidant independent of α-tocopherol and reacts with radicals prior to their
reaction with α-tocopherol (Glascott and Farber (1998) and references therein).
In this way ascorbic acid prevents the loss of α-tocopherol in an indirect manner
in vivo.
MEMBRANE
CYTOSOL
CH3
LIPID-O2
O
O
HO
CH(OH)CH2 OH
R
O
H3 C
Lipid Peroxyl
Radical
O
CH3
O
Semidehydroascorbate
Radical
CH3
α-Tocopherol
CH3
O
LIPID-O2H
Lipid HydroPeroxide
R
O
H3 C
CH3
O
O
CH(OH)CH2 OH
CH3
α-Tocopheroxyl
Radical
HO
OH
Ascorbic
Acid
Figure 4.5 Proposed Interaction Between Ascorbic Acid
And Tocopherol At The Cytosol-Membrane Interface.
As shown in Figure 4.4, the ascorbyl radical can undergo a single electron
oxidation to form dehydroascorbic acid (DHAA). This reaction probably involves
disproportionation (Eqn 4.18). DHAA can then suffer three fates, hydration with
irreversible ring opening and the formation of 2,3-diketo-L-gulonic acid with the
loss of ascorbic acid, dehydration with cyclization, or reduction of DHAA with the
regeneration of ascorbic acid. As 2,3-diketo-L-gulonic acid produces oxalic acid
which is toxic to both animals and plants, the favored pathway is regeneration.
This also helps to conserve ascorbic acid. There has been some debate as to
how ascorbic acid is regenerated from the DHAA (Rose and Bode (1993) and
references therein). Evidence suggests that ascorbic acid may be regenerated
either in an enzyme-independent process using GSH (Varma and Richards
(1988); Winkler (1992)) or an enzyme-dependent process using either NADH,
WWW.ESAINC.COM
343
NADPH or GSH (Halliwell (1996); Maellero et al. (1994); Park and Levine (1996);
Rose and Bode (1992, 1993); Sauberlich (1994); Wells and Xu (1994)) (see
Table 5.1).
2 Ascorbyl• → Ascorbate + Dehydroascorbate
Eqn 4.18
Pro-oxidant Properties.
Several studies suggest that ascorbic acid may be toxic. However, these reports
usually use extremely high doses of ascorbic acid, or involve diseased patients
(e.g., iron-overloaded individuals) (Halliwell, (1996)). Although in vitro evidence
suggests that, under certain circumstances, ascorbic acid may act as a prooxidant (Baysal et al. (1989); Girotti et al. (1985); Giulivi and Cadenas (1993);
Herbert (1996); Skakagami and Satoh (1997)) there is little evidence that it may
act as a pro-oxidant in vivo.
Ascorbic acid’s pro-oxidant activity comes from the finding that it can reduce Fe
(III) to Fe (II) (Eqn 4.19) that can then take part in the Fenton reaction
(Udenfriend et al. (1954)). Bendich et al. (1986) found that, under certain
conditions, ascorbate is capable of reducing oxygen to superoxide that can then
reduce Fe (III) to Fe (II) (Eqns 4.20 and 4.21). However, based upon
thermodynamic principles, Halliwell (1996) concluded that it is unlikely that
ascorbic acid can reduce oxygen to superoxide. Whether acting directly or
through the superoxide radical ascorbate can both mobilize bound iron and
influence the Fe3+/Fe2+ redox status. This effect on iron metabolism has been
proposed to promote lipid peroxidation (Minotti and Aust (1992)). For example,
Andorn et al. (1996) reported that ascorbate could promote iron-dependent lipid
peroxidation in human brain samples. However, others could find no pro-oxidant
activity even in the presence of iron overload (Chen et al., (2000)). Whether this
pro-oxidant effect of ascorbic acid is important in vivo, however, is still a matter of
conjecture.
Fe3+ + ascorbate → Fe2+ + ascorbyl•
O2 + Ascorbate → Ascorbyl• + O2•- + H+
O2•- + Fe3+ → O2 + Fe2+
Eqn 4.19
Eqn 4.20
Eqn 4.21
Ascorbate can also react with copper ions producing hydroxyl free radicals
(Buettner (1996); Aruoma et al. (1991)). Some have even warned of the
possibility that the production of ROS could lead to gastric problems when
ascorbic acid, iron and copper are consumed as part of a multivitamin pill
(Maskos and Koppenol (1991)). Indeed, many multivitamin supplements now
come with a warning on their labels. However, recent evidence could not find any
WWW.ESAINC.COM
344
pro-oxidant activity of ascorbic acid to either plasma proteins or lipids when
subjected to the “Undenfriend system” (hydrogen peroxide + copper II + iron II)
(Suh et al., (2003)).
The consumption of ascorbic acid (>500mg/day) was reported to promote the
formation of a potentially mutagenic lesion in human DNA (8-oxoadenine),
probably through ascorbic acid’s reaction with DNA-bound metals (Podmore et
al., (1998)). Similarly, co-supplementation of human subjects with ascorbic acid
and iron was also found to promote DNA adduct formation (Rehman et al.
(1998)). Whether ascorbic acid can actually promote DNA damage in vivo has
been severely challenged by several groups (Frei (1998); Levine et al. (1998);
Poulsen et al. (1998)). However, recent evidence suggests that ascorbic acid can
promote the decomposition of lipid hydroperoxides thereby producing genotoxic
aldehydes (e.g., 4-oxo-3-nonenal, 4,5-epoxy-2(E)-decanal and 4-hydroxynonenal) that can form mutagenic DNA lesions such as etheno2’deoxyadenosine (Lee et al., (2001)). This may explain why ascorbic acid lacks
efficacy as an anticarcinogenic agent.
Measurement.
The unstable nature of ascorbic acid makes it difficult to analyze accurately
unless certain precautions are undertaken. Although a multitude of assays have
been published for the analysis of ascorbate and dehydroascorbate, many suffer
from the “four-S syndrome” - resulting from inattention to stability, sensitivity,
specificity and substance interference (Washko et al. (1992)). These issues have
been extended by Lykkesfeldt et al. (1995) to include analyte recovery,
reproducibility, choice of detection principle and column durability. It should be
remembered that ascorbic acid, the ascorbyl radical and dehydroascorbic acid
are all in equilibrium with each other, so in order to measure true analyte levels
reliably, the correct choice of analytical method should minimally disturb these
equilibria. For example, many methods include a variety of antioxidants during
sample processing and analysis (e.g., dithiothreitol, homocysteine); however,
these substances may affect the ascorbate/dehydro-ascorbate ratio and lead to
erroneous data. This situation is even more complex as dehydroascorbic acid is
relatively unstable (half-life of 6 minutes at pH 7.0 and 37oC) and is rapidly and
irreversibly lost as 2,3-diketogulonic acid (Schell and Bode (1993)).
Several analytical approaches have been used to measure ascorbic acid and
include spectrophotometric, gas chromatographic and HPLC-based techniques
and these have been critically reviewed (Lykkesfeldt et al. (1995), Washko et al.
(1992)). These authors examined the use of HPLC-ECD in detail and concluded
that coulometric detection was more reliable and offered marked improvement in
sensitivity over amperometric approaches. Several groups have used HPLCcoulometric detection to measure ascorbate either alone (Lykkesfeldt et al.
(1995); Schell and Bode (1993); Xu and Wells (1996)) or with other antioxidants
WWW.ESAINC.COM
345
(e.g., glutathione, glutathione disulfide, uric acid) simultaneously (Rose and Bode
(1995), Sofic et al. (1991)). Ascorbic acid is also routinely determined as part of a
global method for metabolic profiling using gradient HPLC and coulometric array
detection (Gamache et al. (1993); Rizzo et al. 1991)). Tissue levels of ascorbic
acid in a variety of species are presented in Table 4.1.
Dehydroascorbic acid can be measured one of three ways: directly, following
derivatization, or after reduction to ascorbic acid (Washko et al. (1992)). One of
the most promising assays is the use of HPLC-coulometric detection following
chemical reduction of dehydroascorbic acid to ascorbic acid (Dhariwal et al.
(1990); Lykkesfeldt et al. (1995)).
The ascorbyl radical has been measured directly using a spectrophotometric
approach with an absorbance wavelength of 360nm or indirectly using
semidehydroascorbate reductase and monitoring changes in the NAD+/NADH
ratio at 340nm. The ascorbyl radical can also be measured using EPR
approaches (Washko et al. (1992) and references therein).
Thiols.
The term thiol (or sulfhydryl) refers to a compound that contains an –SH group.
The chemistry of thiols and their corresponding disulfides is extensive and has
been briefly described in Chapter 2. Numerous thiols can be found in biological
systems but due to space constraints this review will be limited to the aminothiols
cysteine, homocysteine and glutathione.
1. Glutathione.
The tripeptide, glutathione, (γ-glutamyl-cysteinyl-glycine) [GSH] was first
discovered by J. de Rey-Pailhade over 100 years ago and is the most ubiquitous
peptide found in cells. In this text GSH will be used to designate glutathione and
GSSG will be used for glutathione disulfide. Often in literature GSH is
inaccurately referred to as reduced glutathione and GSSG as oxidized
glutathione (the latter should actually refer to glutathione sulfenic, sulfinic and
sulfonic acids).
GSH can be obtained from the diet or can be synthesized de novo in the liver
(Anderson (1998); Flagg et al. (1994); Jones (1995); Lu (1999)). Synthesis
primarily occurs in the cytoplasm from non-essential amino acid constituents
according to Figure 4.6. GSH is degraded by membrane bound γ-glutamyl
transpeptidase producing glutamate and cysteinyl-glycine that is then hydrolyzed
to cysteine and glycine by cysteinyl-dipeptidase. The enzymes involved in the
synthesis and catabolism of GSH are linked to form the γ-glutamyl cycle (see
below).
WWW.ESAINC.COM
346
O
O
OH
OH
O
O
H2 N
O
SH
NH2
H2 N
Cysteine
NH
Glycine
H2 N
O
SH
NH
OH
H2 N
OH
SH
OH
OH
O
NH
O
ATP
ADP, Pi
O
O
OH
ATP
γ-Glutamyl-cysteine
Glutamate
O
ADP, Pi
HO
Glutathione
(GSH)
γ-Glutamyl-cysteine
Synthetase
Glutathione
Synthetase
Figure 4.6 Synthesis Of GSH.
Mammalian tissue concentrations of GSH are typically 0.5 to 10mM, but
considerably less is found in plasma (Table 4.1). GSH is readily oxidized to
GSSG (see below). In most tissues the level of GSSG is kept low, a
consequence of glutathione reductase activity. For example, in brain the GSSG
level is typically <1% that of GSH (Cooper (1997); Cooper et al. (1980)). Some
GSH also exists as mixed disulfides (with cysteine, coenzyme A and protein
thiolates). The tissue levels of GSH are tightly regulated. For example, it is
difficult to deplete hepatic GSH below 30% of control values even following
xenobiotic challenge or prolonged starvation. On the other hand, even with
supplementation it is difficult to exceed hepatic GSH stores. Such nutritional and
hormonal regulation of glutathione homeostasis was reviewed recently (Taylor et
al. (1996)).
Biological Roles of Glutathione.
GSH plays several important roles in biological systems. It protects against the
action of some pro-oxidants and is involved in detoxification of harmful
compounds. GSH acts as a cofactor for numerous enzymes and represents a
safe storage form of cysteine. It is involved in the transport of amino acids across
membranes and in the regulation of cellular metabolism.
Protection.
GSH can readily be oxidized to its disulfide (GSSG).(Eqn 4.22). With an Eo’=240mV (Table 2.1.1), the GSH/GSSG couple is one of the most reducing
reactions found for endogenous small molecule antioxidants. The reductive
WWW.ESAINC.COM
347
capacity of GSH is utilized by glutathione peroxidase (and other peroxidases) in
the destruction of peroxides (Figure 4.7).8 The reduction of GSSG back to GSH
is energetically unfavorable so it is catalyzed by an enzyme, glutathione
reductase, that requires a strong reducing agent (NADPH) (NADP+/NADPH: Eo’=320mV). Glutathione reductase consists of two subunits, each containing FAD at
the active site. Electrons are passed from NADPH through FAD to a cystine
disulfide bridge located in the active site. The resulting two active cysteine thiols
then take part in reducing GSSG.
2GSH → GSSG + 2H+ + 2e-
Glucose
SH
2NAD
Pentose
Phosphate
Pathway
H2O2 (RO2 H)
γ -Glu-Cys-Gly
(GSH)
+
+ 2H +
2NADPH
Eqn 4.22
2
Glutathione
Peroxidase
Glutathione
Reductase
γ-Glu-Cys-Gly
S
Ribulose-5Phosphate
S
γ -Glu-Cys-Gly
2H2O (ROH + H2 O)
(GSSG)
Figure 4.7 The Use And Regeneration Of GSH.
The GSH/GSSG ratio is usually kept high in cells (typically >10:1 to >100:1),
maintaining a reducing environment. Keeping cellular GSSG levels low is
important as it can affect protein synthesis and inhibit several enzymes possibly
by forming mixed disulfides with essential protein thiols. Enzymes affected
include adenylate cyclase, chicken hepatic fatty acid synthetase, rabbit muscle
phospho-fructokinase, and phosphorylase phosphatase.9 Under periods of
oxidative stress, the liver and heart can actively transport GSSG out of their cells,
thereby preventing deleterious action. The GSH/GSSG ratio is also high in
mitochondria. Here it serves to keep transport proteins and enzymes (e.g.,
ATPases and dehydrogenases) active by maintaining essential thiol groups in a
reduced form.
8
GSH also reacts directly with a variety of free radicals producing reactive thiyl radicals that must be further metabolized
(Chapter 2).
9
Conversely, GSH can protect certain enzymes from oxidative stress through tightly regulated S-thiolation (e.g.,
glyceraldehyde-3-phosphate dehydrogenase) (Grant et al. (1999)).
WWW.ESAINC.COM
348
GSH also plays a role in the regeneration of other antioxidants. For example, it
can regenerate ascorbic acid from dehydoascorbic acid and membrane bound
α-tocopherol from the α-tocopheroxyl radical formed during inhibition of lipid
peroxidation (see below).
R-X
GSH
GSH-R
GSH Transferase
-Glutamate
Glycine-Cysteine-R
Glutamyltranspeptidase
NHCOCH3
R
S
O
OH
A Mercapturic Acid
Cysteinylglycinase -Glycine
N-Acetylase
Cysteine-R
Figure 4.8 The Detoxification Of A Xenobiotic (R-X) By The Mercapturic
Acid Pathway.
Detoxification and Bioactivation.
GSH is important in the detoxification of potentially harmful endogenous
compounds and xenobiotics (e.g., α-oxoaldehydes10 and redox active
compounds such as monoamines, catecholestrogens, polyphenols and some
drugs) (Figure 2.24). GSH is a nucleophile and readily forms S-conjugates with
electrophilic compounds in a reaction catalyzed by glutathione-S-transferases.
Glutathione-S-conjugates are further metabolized through the mercapturate
pathway (Figure 4.8) and are usually excreted in the bile. The mercapturic
pathway is also important for the metabolism of endogenous substances such as
the leukotrienes.
Unfortunately many S-conjugates retain (and some may even be bioactivated to
products that can even exceed) the electrophilic and redox properties of the
parent compound (Anders and Dekant (1998); Monks and Lau (1998)). For
example, dibromoethane can produce a highly reactive episulfonium intermediate
capable of damaging DNA (Anders and Dekant (1998) and references therein).
10
A number of potentially toxic α-oxoaldehydes (e.g., glyoxal, methylglyoxal and 4,5-dioxovalerate) are formed in vivo
during lipid peroxidation, glycation and as part of normal metabolism (e.g., metabolism of ketone bodies and
triosephosphates, and threonine catabolism). These compounds readily react with amine groups found in DNA (leading to
mutagenesis and apoptosis), RNA and proteins (leading to protein degradation and cytokine-mediated immune response
(Thornalley (1998) and references therein). They are detoxified by oxidation to aldonic acids catalyzed by cytosolic
enzymes (glyoxylase I and II) using GSH as cofactor.
WWW.ESAINC.COM
349
Excretion of the GSH-conjugate may protect the cell from the toxic effect of these
compounds, but once in the circulation, other cells can then accumulate these
toxic metabolites leading to tissue damage. Another unfortunate consequence of
GSH-mediated detoxification is the deactivation of anticancer drugs (Zhang et al.
(1998a)). Inhibition of this detoxification pathway is currently being explored as a
strategy to modulate drug resistance.
Consumption of excessive amounts of the painkiller acetaminophen
(paracetamol) can lead to hepatotoxicity. The mechanism for this action involves
the conversion of acetaminophen (and its derivative 4-ethoxyacetanilide) to a
highly reactive quinoneimine by the action of the cytochrome P450 system. The
quinoneimine exerts its toxicity by reacting with protein-thiol groups (adducts) and
by depleting GSH stores due to the formation of acetaminophen-GSH adducts.
Treatment includes supplementation with N-acetylcysteine and methionine that
act by maintaining GSH levels.
H2O2
Catabolism
Ascorbate
Dehydroascorbate
Glutaredoxin
2H2O
H2O2
GSH
GSSG
Peroxidases
NADPH
GSH Reductase
Figure 4.9 The Relationship Between Glutathione And
Ascorbic Acid (see Meister (1994) for greater details).
Cofactor.
GSH is a cofactor for many enzymes including glutathione peroxidase, and other
peroxidases, dehydrochlorinase, formaldehyde dehydrogenase, glyoxalase,
maleyl-acetoacetate isomerase, and prostaglandin endoperoxidase isomerase
(Meister (1989); Thornally (1998)). GSH is also used by dehydroascorbate
reductase, the enzyme responsible for the regeneration of ascorbic acid from its
potentially toxic metabolite dehydroascorbate (Eqns 4. 23 and 4.24) (Figure 4.9)
(Meister (1994)).
WWW.ESAINC.COM
350
Dehydroascorbate + 2GSH → Ascorbate + GSSG
Or
Eqn 4.23
2Semidehydroascorbate + 2GSH → 2Ascorbate + GSSG
Eqn 4.24
Storage of Cysteine in a Non-toxic Form.
Cysteine is excitotoxic probably through its action on the NMDA receptor. Its
sulfhydryl group readily forms hemithioacetals with aldehydes and hemithioketals
with α-ketoacids. For example, following its reaction with pyridoxal 5’-phosphate,
the hemithioacetal adduct cyclizes to form a thiazolidinone that can effectively
inhibit any enzyme using pyridoxal 5’-phosphate as a cofactor. The sulfhydryl
group of GSH is much less reactive, thus permitting cysteine to be stored in
mammalian cells at 10-100 times the level of the free amino acid (Cooper
(1997)).
Amino Acid Transport.
The enzymes for GSH synthesis and catabolism are linked forming the
γ-glutamyl cycle that may function in amino acid transport (Orlowski and Meister
(1970)). The location of γ-glutamyltranspeptidase on the cell surface is thought to
enable the translocation of amino acids across the cell membrane. The
γ-glutamyl cycle is found to be most active in tissues where amino acid transport
is high (e.g., the kidney). However, as it is energetically very expensive (requiring
the hydrolysis of 3 ATP molecules) and other less energetic amino acid
transporters are available, the importance of the γ-glutamyl cycle in amino acid
translocation remains to be elucidated (Meister (1994) and references therein).
The γ-glutamyl cycle also plays a role in the metabolism of estrogens,
leukotrienes and prostaglandins and in the detoxification of xenobiotics (see
Cooper and Kristal (1997) and references therein).
Regulation.
The regulation of cellular metabolism was first proposed in the 1950s and 1960s
when it was shown that key metabolic enzymes could be regulated by thioldisulfide exchange. More recently in vitro studies have reported that signal
transduction (protein/protein interactions) and gene transcription (protein/DNA
interactions) are dependent on the redox status of critical sulfhydryl groups
which, in turn, could be affected by GSH levels (Taylor et al. (1996) and
references therein).
WWW.ESAINC.COM
351
Compartmentalization.
GSH is primarily synthesized in the cytoplasm, yet many of its physiological
functions occur in other compartments including the endoplasmic reticulum11,
mitochondrial matrix, nucleus and extracellular space. The availability of GSH in
these different compartments is complex and dependent upon several factors
including transport, utilization, synthesis and reduction of GSSG and GSH-mixed
disulfides (Cooper and Kristal (1997); Smith et al. (1996)). For example, the
uptake of GSH by mitochondria is an energy dependent transport process
coupled to the efflux of anions, whereas the passage of GSH into the nucleus
appears to be by passive diffusion (Smith et al. (1996)).
Conditions and Diseases Affecting Glutathione.
The GSH/GSSG ratio is affected by a number of conditions and diseases
including aging, AIDS, arthritis, cancer, cardiovascular disease, Crohn’s disease,
diabetes, exercise, gluten sensitivity, nephrotoxicity, neurodegenerative disease,
oxidative stress, pre-eclampsia, pulmonary disease, and Wilson’s disease (Boda
and Nemeth (1992); Harding et al. (1994); Iantomasi et al. (1993, 1994);
Lomaestro and Malone (1995); Navarro et al. (1999); Reed (1990); Samiac et al.
(1998); Staal (1998); Summer and Esenburg (1985); Vina et al. (1996)).
Measurement of Glutathione and its Disulfide.
Many methods exist for the measurement of glutathione. There are two major
considerations when choosing a method. Firstly, not all methods differentiate
between GSH and GSSG. For example, the most widely used technique,
enzyme recycling, measures total GSH (GSH and GSSG) in a reaction involving
NADPH, 5,5’-dithiobis-(2-nitrobenzoic acid) and glutathione reductase (Tietze
(1969)). In order to measure GSSG alone the alkylating agent N-ethylmaleimide
(NEM) is added, but this can also lead to inactivation of glutathione reductase.
Excess NEM must therefore be completely removed for the assay to function. A
less laborious modification uses 2-vinylpyridine that does not inhibit the enzyme
(Griffith (1980)). This technique may not be suitable for the measurement of low
GSH tissue levels. Brigelius et al. (1983) developed an assay using 1-chloro-2,3dinitrobenzene and glutathione-S-transferase with a limit of detection of 300nM.
Secondly, extreme care must be exercised when collecting, storing and analyzing
samples as it is relatively easy to artificially alter the GSH/GSSG ratio (e.g.,
Jones et al. (1998)). Precautions include the use of serine borate to inhibit γglutamyltransferase activity, not rupturing red blood cells, and sampling at a
specific time of day to avoid possible circadian rhythms).
11
Where it is involved in protein-disulfide isomerase-dependent protein folding (Walker and Gilbert (1997)).
WWW.ESAINC.COM
352
Analyte
GSH, GSSG
Electrode
Material
Glassy carbon
Eye lens
GSH
Gold amalgam
Brain
GSH
Gold amalgam
GSH, GSSG
Gold amalgam
Liver microdialysis
perfusion
Blood
GSH, GSSG
Gold amalgam
Red blood cells
GSH, GSSG
Gold amalgam
Eye-lens
GSH, ascorbic acid
Flow-through
graphite
Flow-through
graphite
Flow-through
graphite
Human brain
Liver and peripheral
tissues
Plasma
Rose and Bode
(1995)
Acworth and Bailey
(1995)
Flow-through
graphite
Plasma
Melnyk et al. (1999)
Flow-through
graphite
Bronchoalveolar
lavage, plasma
Smith et al. (1995)
Flow-through
graphite
Flow-through
graphite
Flow-through
graphite
Flow-through
graphite
Lung
Lakritz et al. (1997)
Human substantia
nigra
Human hair
Sofic et al. (1992)
Krien et al. (1992)
Bile, kidney, lens,
liver
Carro-ciampi et al.
(1988)
Flow-through
graphite
Bile, liver
Harvey et al. (1989)
Flow-through
graphite
Bile, urine
Hill et al. (1992)
GSH, ascorbic acid, uric
acid, cysteine, allantoin
GSH, GSSG, cysteine,
homocysteine,
methionine, Nacetylcysteine
GSH, GSSG, cysteine,
cystine, cystathionine,
homocysteine,
cysteinylglycine,
methionine, homocystine
GSH, GSSG,
cystathionine, cysteine,
cystine, homocysteine,
homocystine, methionine
GSH, GSSG
GSH, GSSG
GSH, GSSG
GSH, GSSG, methionine,
N-acetylcysteine
GSH, GSSG, cysteine,
homocysteine,
cystathionines
GSH-conjugates
Tissue
Reference
Ozcimder
et
al.
(1991).
Pileblad and
Magnusson (1989)
Yang et al. (1995)
Allison and Shoup
(1983)
Yamashita and
Rabenstein (1989)
Mitton and Trevithick
(1994)
Sofic et al. (1991)
Table 4.4 HPLC-ECD Methods For The Measurement Of Glutathione,
Glutathione Disulfide etc. From Acworth et al. (1998) and updated.
A variety of HPLC techniques have also been developed. HPLC-UV requires
derivatization (e.g., with Sanger’s reagent) (Fariss and Reed (1987)) but as it
possesses poor limit of detection it may not be sensitive enough for many
WWW.ESAINC.COM
353
biological applications. HPLC-fluorescence detection requires derivatization with
OPA (Keller and Menzel (1985); Michelet et al. (1995)), monobromobimane
(Fahey and Newton (1987); Newton et al. (1981)) or dansyl chloride (Martin and
White (1991)). HPLC-ECD uses either amperometric or coulometric electrodes to
measure GSH and GSSG directly, thus avoiding typical problems associated with
derivatization procedures (e.g., derivatization efficiency, ghost peaks) (Table
4.4).12 In general, coulometric detection offers superior sensitivity and selectivity
to the dual-amperometric approach and furthermore avoids the use of toxic,
unstable mercury amalgams and the problem of complete oxygen removal prior
to reductive determination of GSSG. A typical chromatogram showing the use of
coulometric detection to measure a variety of thiols and disulfides is presented in
Figure 4.10. See also ESA Application Notes – 70-5343 Total Glutathione and
70-5043 Total Thiols. Hill et al. (1993) used a gradient coulometric array
approach to study the metabolism of a variety of S-substituted GSH conjugates
formed when animals were exposed to hydroquinone. Their method offered
excellent resolution of the conjugates, even in complex biological matrices such
as urine.
Figure 4.10 Separation Of Thiol And Disulfide Standards Using HPLC-ECD
(20ng On Column). (Reproduced with permission of Achilli and Cellerino (1996)).
12
Although this approach is exquisitely sensitive it is not without problems. The low amount of organic modifiers typically
used with reversed-phase HPLC can result in microbial growth in the system leading to noise and ghost peaks in the
chromatogram. Also, biological materials (lipids, proteins etc) can build up on the column and foul the working electrode,
causing poor chromatography and loss of sensitivity. This can be overcome by routine cleaning. Finally, trace transition
metal contamination can result in auto-oxidation of GSH. Only biologically compatible HPLC-systems should be used.
WWW.ESAINC.COM
354
2. Homocysteine.
Homocysteine is a key metabolite in both sulfur amino acid biochemistry and the
transfer of “activated” methyl groups (in regeneration of the “methyl carrier” - Sadenosylmethionine). Homocysteine is formed from methionine, and can react
with serine for the synthesis of cysteine, or converted back to methionine by
transfer of a methyl group from N5-methyltetrahydropteroyl-tri-L-glutamate (a
folate derivative) in a reaction that involves a vitamin B12 (cobalamin) derivative,
5-methyltetrahydrofolate and the enzyme methionine synthase (or tetrahydropteroylglutamate methyltransferase) (Figure 4.11). It is not surprising therefore
that plasma levels of homocysteine are used clinically to monitor folate and
cobalamin function.
NH2
O
Methionine
Adenosyl
Transferase
Adenosyl
R
CH3
Pi + PPi
ATP
Methyltransferase
S
OH
R-CH3
S-Adenosyl-Methionine (SAM)
NH2
NH2
O
S
O
S
Adenosyl
OH
CH3
OH
Methionine
S-Adenosyl-Homocysteine (SAH)
+ Methylcobalamin
+ 5-Methyltetrahydrofolate
Methionine
Synthase
H2O
NH
Adenosine
2
SAH Transferase
O
HS
OH
Homocysteine
Serine
Cystathionine β-synthase
H2O
NH2
NH2
O
S
O
OH
OH
Cystathionine
H2O
Cystathioninase
NH4+
NH2
HS
O
+
H3C
CH2
CO.CO2H
α-Ketoglutarate
OH
Cysteine
Figure 4.11 The Metabolism Of Homocysteine.
WWW.ESAINC.COM
355
Several forms of homocysteine occur in plasma, including the reduced thiol
(<0.25µM), the free disulfide (homocystine and homocysteine-cysteine mixed
disulfides; <2µM) and the protein bound disulfide (<8µM) (Ueland (1995)). The
plasma homocysteine pool is in dynamic equilibrium with the other aminothiols
found in the plasma compartment. For example, an increase in plasma
homocysteine also increases the level of free cysteine by liberating it from its
protein-bound form. Furthermore, changes in plasma homocysteine redox status
rapidly affect and are related to the redox status of other aminothiols (Ueland et
al. (1996)).
Although homocysteine can act as an antioxidant through its oxidation to its
disulfide (Baker et al. (1996); Ueland et al. (1996); Zappacosta et al., (2001)) it
can also act as a pro-oxidant capable of promoting lipid peroxidation and protein
damage (Halverson et al. (1996); Olszewski and McCully (1993)) and as a
neurotoxin (Kim and Pae (1996)). The possible physiological relevance of its prooxidant status still awaits clarification.
Homocysteine can be converted enzymatically to an intramolecular thioester, or
thiolactone. Homocysteine thiolactone readily undergoes nucleophilic addition
with primary amines to form biologically relevant homocystamide adducts (see
Ferguson et al. (1999) and references therein). For example, LDL reacts with
homocysteine thiolactone to produce LDL-homocystamide adducts that have
been implicated to increase atherogenicity of LDL (Naruszewicz et al. (1994)).
LDL-homocystamide adducts may also serve as markers of plasma
homocysteine levels and can be measured using polyclonal antibodies (Ferguson
et al. (1998)). Contrary to earlier reports, LDL-homocystamide adducts may
serve as a local antioxidant making the LDL molecule more resistant to prooxidant damage (Ferguson et al. (1999)).
Moderate hyperhomocysteinemia (plasma levels <30µM) is an important
cardiovascular risk factor (Jacobsen (1998)). Consequently, there is now a
growing interest in the possible roles of homocysteine in control of the plasma
redox thiol status in disease and oxidative stress (Frishman (1998); Jacobson
(1998); Selhub and D’Angelo (1998); Ueland (1995)).
As it is the “total” (free + free disulfide + protein-bound disulfide) level of
homocysteine that has been determined to be clinically significant, sample
pretreatment includes the use of urea to denature proteins and a reducing agent
(e.g., dithiothreitol, sodium borohydride, or tributylphosphine) to reduce
disulfides. Such sample treatment may be problematic (e.g., incomplete
reduction) or cause problems for the subsequent analytical procedure.
Homocysteine and its metabolites have been determined using a variety of
techniques, including antibody based procedures, GC-MS, radioenzymatic
methods and HPLC with UV or fluorescence detection following derivatization
with reagents such as OPA, monobromobimane, 2-chloro-1-methylpyridinium
iodide, or 7-fluorobenzo-2-oxa-1,3-diazole-4-sulfonate (Andersson et al. (1995);
WWW.ESAINC.COM
356
I.S. Penicillamine
Homocysteine
Daskalakis et al. (1996); Fermo et al. (1992, 1998); Fiskerstrand et al. (1993);
Frantzen et al. (1998); Hyland and Bottiglieri (1992); Imai et al (1983); Jacobsen
et al. (1994); Kaniowska et al. (1998); Santhosh-Kumar et al. (1994);
Shipchandler and Moore (1995)). Homocysteine has also been measured using
HPLC with coulometric (Achilli and Cellerino (1996); Martin et al. (1999)) (Figure
5.8) or amperometric (D’Ermo et al. (1999)) electrochemical detection or pulsed
amperometric detection on a gold working electrode (Evrovski et al. (1995); Wu
et al. (1994)). Recently, Bailey (1998) has developed a simplified yet highly
sensitive HPLC-ECD-based method capable of measuring the total plasma level
of homocysteine (Figure 4.12). The addition of the novel reducing agent tris(2carboxyethyl)-phosphine (a stable, water soluble and easy to handle reagent)
directly to plasma completely liberates homocysteine from all of its disulfides
without the need for the use of urea or extensive sample preparation. See ESA
Application Notes – 70-3994 Total Plasma Homocysteine and 70-4989 An
Alternate Plasma Homocysteine Method.
Figure 4.12 Chromatogram Of A Human Plasma Sample.
The isocratic system consisted of a pump, an autosampler and a Coulochem® III detector.
LC Conditions:
Column:
Guard Column:
Mobile Phase:
HR-80 C18 (4.6 x 80mm; 3 µm).
C18
0.15M Sodium Dihydrogen Phosphate, 1.0mM SDS, 10% Acetonitrile,
final pH = 2.80 (With Phosphoric Acid).
Flow Rate:
1.2mL/min.
Temperature:
Ambient.
Injection Volume:
20µL.
Cell Potentials:
EGC = +850mV.
E1 = +450mV.
E2 = +750mV.
See Application Notes – 70-3994 Total Plasma Homocysteine or 70-4989 Alternative
Homocysteine, for further details.
WWW.ESAINC.COM
357
As with GSH, when attempting to measure fractionated homocyst(e)ines,
regardless of the analytical approach being used, care must be taken so as not
to artificially alter the redox status of the tissue sample (Fermo et al. (1997)).
3. Miscellaneous Endogenous Sulfur-Containing Compounds.
A variety of sulfur containing compounds including coenzyme A, cysteamine,
cysteic acid, hypotaurine, S-adenosyl-L-methionine, pantothenic acid and
aminoethylcysteine ketimine decarboxylated dimer have been proposed as
antioxidants (Aruoma et al. (1988); Evans et al. (1997); Fontana et al. (1998);
Matarese et al. (1998); Slyshenkov et al. (1995)). However, whether these
compounds occur at sufficient levels in vivo to act as antioxidants still remains to
be clarified. It seems unlikely that taurine functions as an antioxidant in vivo as it
does not react rapidly with ROS, and its reaction product with HOCl still shows
pro-oxidant activity (Aruoma et al. (1988)). Lipoic acid, a strong antioxidant, is
discussed below.
NH2
O
N
N
N
HN
N
NH
NH
H2 N
ADENINE
GUANINE
H2O
1
NH3
H2O + O2
2H + + O2.-
N
HN
N
4 - XANTHINE DEHYDROGENASE
H2O
2
NH3
O
1 - ADENINE DEAMINASE
2 - GUANINE DEAMINASE
3 - XANTHINE OXIDASE
NH
O
N
HN
NH
3
O
NH
NH
4
O
NH
HN
3
O
O
NH
NH
4
O2, NAD+/NADH
HYPOXANTHINE
2H + + O2.-
H2O + O2
O2, NAD+/NADH
XANTHINE
URIC ACID
Figure 4.13 The Conversion Of Purines Into Uric Acid.
WWW.ESAINC.COM
358
Uric Acid.
Formation.
The purine catabolic pathway shown in Figure 4.13 is found in all animals. Uric
acid represents a metabolic branch-point and the final end product of purine
catabolism is species dependent. In man and primates, who lack uricase, uric
acid is the final product and is excreted in the urine. Mammals other than man
and primates excrete allantoin. Teleost fish produce allantoic acid; other fish and
reptiles produce urea; some marine invertebrates hydrolyze urea to ammonia
and carbon dioxide.
O
N
HN
O2
NH
N
HYPOXANTHINE
Mo VI
FeII
Mo IV
FeIII
O2.-
O
N
HN
O
NH
NH
O2
XANTHINE
O
FeII
Mo IV
FeIII
O2.-
NH
HN
O
Mo VI
O
NH
NH
URIC ACID
Figure 4.14 The Role Of Metals In Superoxide Production By
Xanthine Oxidase.
As presented in Figure 4.13, a single enzyme, xanthine oxidase, is responsible
for the conversion of hypoxanthine to xanthine and xanthine to uric acid with the
simultaneous production of superoxide as a byproduct. The mechanism of action
of xanthine oxidase is shown in Figure 4.14. This enzyme is a molybdenum-
WWW.ESAINC.COM
359
containing flavin-hydroxylase and is capable of existing as both an oxidase (using
oxygen as a cofactor and producing superoxide) and dehydrogenase (using
NAD+ and not producing superoxide). Xanthine dehydrogenase is the naturally
occurring form of the enzyme but is readily converted into the oxidase form either
irreversibly (by proteolysis) or reversibly (by sulfhydryl oxidation) (Hille and
Nishino (1995); Parks et al. (1999)). As this can occur during sample preparation
due care should be exercised when purifying the enzyme.
Apart from its role in uric acid formation, xanthine oxidase can also catalyze other
reactions. It possesses nitrite reductase activity and can reduce nitrite to nitric
oxide (Zhang et al. (1998c)). Although the physiological importance of this activity
is at present unknown, the generation of nitric oxide by xanthine oxidase may
serve as a supplement to NOS and help to redistribute blood flow following
ischemia (Zhang et al. (1998c)). Xanthine oxidase can decompose
dibromoacetontrile (a product of water disinfection) to cyanide a problem for
potable water (Mohamadin and Abdel-Naim (2003)). Xanthine oxidase also
promotes decomposition of nitrosothiols forming nitric oxide (anaerobic
conditions) and peroxynitrite (aerobic conditions) (Trujillo et al. (1998)).
Xanthine oxidase can be inhibited by analogs of uric acid, e.g., allopurinol. This
compound is called a “suicide inhibitor” because once it is oxidized by xanthine
oxidase it is converted to oxypurinol, a compound that binds tightly to
molybdenum-containing active site of the enzyme. Allopurinol is commonly used
to treat gout. Other inhibitors include pterinaldehyde, the FAD-site inhibitor,
diphenyleneiodonium and some flavonoids (Cos et al. (1998)).
Xanthine Oxidase and Tissue Injury.
There is considerable interest in the role that xanthine oxidase may play in the
damage associated with reperfusion injury (Hille and Nishino (1995); Nishino
(1994); Saugstad (1996)). It has been hypothesized that during hypoxia ATP is
depleted by conversion to hypoxanthine, while xanthine dehydrogenase (the
principle form of the enzyme) is slowly converted to the oxidase. Upon
reperfusion, xanthine oxidase aggressively converts hypoxanthine to uric acid
while simultaneously reducing oxygen to superoxide which also dismutates to
hydrogen peroxide. The resulting burst of ROS damages tissue, facilitating the
release of xanthine oxidase with the possibility of tissue damage far removed
from the site of the initial insult. Although this hypothesis is attractive it has been
challenged. For example, there is some question as to whether xanthine
dehydrogenase exists in the organs affected by reperfusion injury (see Nishino
(1994) and references therein). Another issue is whether the enzyme actually is
converted to xanthine oxidase during ischemia (see Nishino (1994) and
references therein). However, it now appears that xanthine dehydrogenase
conversion to xanthine oxidase may not be essential for ROS production, as
xanthine dehydrogenase can lead to ROS production too (Zhang et al. (1998b)).
WWW.ESAINC.COM
360
Antioxidant and Pro-oxidant Activities.
Uric acid is often erroneously regarded as a waste product. This does not make
good biochemical sense as the production of purines is an energy-dependent
process and their destruction is physiologically costly. Furthermore, >90% of uric
acid is reabsorbed from urine by the kidney. In man the circulating level of uric
acid is typically 120-450µM (see Table 4.1) which approaches the solubility limit
for this analyte. A variety of diseases are associated with uric acid’s poor
solubility. For example, when urine is unusually acidic calcium urate stones can
form in the kidney and bladder. Conditions where the solubility limit of uric acid
(~450µM) is exceeded, either by its increased production or failure of the kidney
to effect its removal, can result in the deposition of monosodium urate crystals in
joints, leading to painful inflammation and gout. In fact, a uric acid concentration
>600µM is virtually certain to cause gout. Interestingly, a phenomenal
concentration of uric acid is achieved by blood sucking insects where hemolymph
levels can reach as high as 5mM. This high level of uric acid protects the insect
from the pro-oxidant effects of digested hemoglobin (Souza et al. (1997)).
So what is the evolutionary significance of uric acid? It seems strange that
evolution has not enabled man to produce a more soluble and less problematic
product. One possible answer is that uric acid is a very good antioxidant. It has
been hypothesized that an important step in human evolution was the
replacement of ascorbic acid with uric acid as the principal circulating antioxidant
(Ames et al. (1981); Becker (1993); Cutler (1991)). Consequently, the increased
plasma concentration of uric acid enabled man to live longer and avoid the
cancers commonly associated with short-lived species. In support of this theory it
is found that man and primates live longer than prosimians where the circulating
level of uric acid is ten-fold lower. It appears that uric acid now contributes up to
60% of the measured total antioxidant capacity of plasma in healthy subjects
(Benzie (1996); Wayner et al. (1997)).
Uric acid is an effective antioxidant. Uric acid reacts with singlet oxygen, nitrogen
dioxide, alkyl peroxyl radicals and peroxynitrite13 (Halliwell and Gutteridge (1993);
Hooper et al. (1998); Skinner et al. (1998)). Uric acid, like most other
compounds, will react with hydroxyl free radicals if present at sufficient
concentration at the site of production of this pro-oxidant. The one-electron
oxidation of uric acid produces potentially damaging urate radicals (Aruoma and
Halliwell (1989); Kitteridge and Wilson (1984); Maples and Mason (1988)).
Fortunately, the uric acid radical can be converted back to uric acid by its
interaction with ascorbic acid (Maples and Mason (1988)). Uric acid can also bind
iron and copper in forms that do not react with hydrogen peroxide, and it protects
against damage from heme intermediates containing iron in IV and V valencies
13
Interestingly, peroxynitrite reacts with uric acid to produce a nitrated derivative that possesses vasoactive properties
through its release of nitric oxide (Skinner et al. (1998)).
WWW.ESAINC.COM
361
(Ames et al. (1981); Halliwell and Gutteridge (1990); Halliwell et al. (1992);
Maples and Mason (1988); Vasquez-Vivar et al. (1996)). Interestingly, uric acid
does not react with superoxide and only plays a limited protective role against
hypochlorous acid (Hu et al. (1993)).
Following reaction with pro-oxidants uric acid can decompose to give a variety of
metabolites, some of which are potentially toxic (Figure 4.15) (Hicks et al.,
(1993); Kaur and Halliwell (1990)). Possible reaction mechanisms for uric acid
breakdown have been studied using coulometric electrochemical detection
coupled to mass spectrometry (Volk, et al. (1989)). Uric acid breakdown products
can also be used as oxidative stress markers. For example, Hillered and Persson
(1995) measured parabanic acid as an indicator of oxidative stress in
microdialysis perfusates obtained from patients with severe acute brain injuries.
Similarly, the ratio of allantoin/uric acid levels have also been used as a potential
index of free radical reactions in vivo and have shown to be increased with
disease (Grootveld and Halliwell (1987); Lux et al. (1992); Moison et al. (1997)),
exercise (Hellsten et al. (1996)) and during development (Moison et al. (1997)).
H2NCONH2
NH
O
HN
UREA
CO2 H
N
HO
NH
OH
N
O
OH
O
OXONIC ACID
CYANURIC ACID
NH
HN
O
N
O
NH
NH
URIC ACID
O
NH
O
O
NH2 CONH
O
NH
ALLANTOIN
NH
O
H2NCONHCOCO2H
NH
PARABANIC ACID
OXALURIC ACID
Figure 4.15 Uric Acid Can Decompose To Give A Variety Of
Products.
Even though uric acid can show some pro-oxidant activity this is insignificant
when compared to its ability to act as an antioxidant. The question arises as to
WWW.ESAINC.COM
362
why do biological systems have both uric acid and ascorbic acid present to react
with radicals. This can be answered by looking at the electrode potential of the
various reactions. The Eo for uric acid is +590mV which is markedly higher than
that of ascorbic acid (see Table 2.1.1). Using the Nernst equation there will be
29.9 Kjmol-1 less energy if urate, rather than ascorbate, reacts with the hydroxyl
free radical (under standard conditions and a pH=7). Thus there is definitely a
major advantage when energy is released in more controllable small steps
(hydroxyl free radical/urate; urate radical/ascorbate) rather than all at once
(Benzie and Strain (1996)).
Under some conditions, excessive demands on dietary ascorbic acid for uric acid
recycling may lead to ascorbic acid depletion which will, in turn, interfere with
tocopherol regeneration (see below). This and other problems associated with
uric acid pro-oxidant activity (e.g., stimulated secretion of superoxide from
leukocytes; its ability to release iron from oxyhemoglobin) led Benzie and Strain
(1996) to question uric acid’s true biological role.
Measurement.
Uric acid can be measured using GC-MS (Chen et al. (1998)). It can also be
measured using HPLC-UV absorbance (Ames et al. (1981), Tang-Liu and
Riegelman (1982); Yang (1998)) or with better selectivity and sensitivity, by
HPLC-ECD. HPLC-ECD approaches either measure it alone (Aoki et al. (1984),
Iwamoto et al. (1983), Roch-Ramel et al. (1980)) or in conjunction with ascorbate
(Honegger et al. (1989), Irayama et al. (1984), Shirachi and Omaye (1992)) or
other metabolites (Gogia et al. (1998); Rose and Bode (1995)). Uric acid has also
been determined as part of a global method for metabolic profiling using gradient
HPLC with coulometric array detection (Rizzo et al. (1991)).
Fat-Soluble Antioxidants.
Carotenoids.
Carotenoids are primarily a group of lipophilic C40 polyisoprenoid compounds that
possess an extensive conjugated double bond system which enables them to
strongly absorb UV and/or visible light, act as antioxidants and renders them
electrochemically active. A consequence of electromagnetic radiation absorption
is that many of the carotenoids are brilliantly colored. For example, lycopene is
red (tomatoes), lutein and zeaxanthin are yellow (sweet corn), and α- and βcarotene are orange (carrots). Carotenoids can be subdivided into the
xanthophylls (oxygenated carotenoids) and carotenes (hydrocarbons). The
carotenoids are synthesized in plants and microorganisms (where they aid
photosynthesis or act as photo-protectants) and are essential nutrients for
animals. The consumption of large amounts of astaxanthin producing organisms
WWW.ESAINC.COM
363
is responsible for the pink coloration of flamingo feathers and salmon flesh. It is
also responsible for the red coloration of boiled shellfish. Over 600 carotenoids
have been identified and some of the more biologically important ones are
presented in Figure 4.16.
CH3
CH3
CH3
CH3
H3 C
CH3
CH3
CH3
CH3
CH3
CH3
CH3
CH3
CH3
CH3
α-CAROTENE
CH3
CH3
CH3
CH3
CH3
H3 C
CH3
CH3
CH3
CH3
CH3
CH3
CH3
CH3
LUTEIN
CH3
CH3
CH3
CH3
CH3
CH3
H3 C
CH3
CH3
CH3
CH3
CH3
CH3
CH3
CH3
H3 C
CH3
CH3
CH3
ZEAXANTHIN
CH3
CH3
CH3
CH3
γ-CAROTENE
CH3
OH
H3 C
CH3
CH3
HO
CH3
CH3
OH
H3 C
CH3
CH3
HO
CH3
CH3
OH
H3 C
CH3
CH3
CH3
CH3
CH3
CH3
CH3
CH3
H3 C
CH3
O
O
CH3
CH3
CH3
CH3
CH3
CH3
CANTHAXANTHIN
δ-CAROTENE
CH3
CH3
CH3
CH3
CH3
CH3
CH3
CH3
CH3
CH3
CH3
H3 C
CH3
O
CH3
CH3
HO
OH
O
CH3
CH3
CH3
CH3
CH3
ASTAXANTHIN
LYCOPENE
Figure 4.16 The Structures Of Some Biologically Important Carotenoids.
Carotenoids and Disease.
The role of the carotenoids in health and disease prevention has been reviewed
extensively elsewhere (Canfield et al. (1993); Mayne (1996); Krinsky (1993);
Omaye et al. (1997)) so will not be dealt with in detail here. Mayne (1996)
critically reviewed the role for β-carotene (and some other carotenoids) in cancer
prevention (including lung, oral, gastrointestinal, breast, prostate, cervical and
skin cancers) but concluded, “supplemental β-carotene is unlikely to be beneficial
in reducing the major cancers occurring in westernized populations”. In fact,
supplemental β-carotene was without effect (Hennekens et al. (1996)) or even
promoted cancer in subjects that smoked (De Luca and Ross (1996); Heinonen
et al. (1994); Omenn (1998); Omenn et al. (1996); Paolini et al., (2003)). Other
WWW.ESAINC.COM
364
studies associate β-carotene levels with either decreased incidence or little effect
on cardiovascular disease (reviewed by Palace et al. (1999)). For example, βCarotene did not prevent cardiovascular disease and was potentially harmful
(Lonn and Yusuf (1997); Rimm and Stampfer (1997)).
Supplemental β-carotene and canthaxanthin have been used to successfully
treat certain photosensitivity diseases such as erythropoietic protoporphyria
(Mayne (1996)). The consumption of β-carotene rich foods has been associated
consistently with decreased risk of cardiovascular disease, yet β-carotene
supplementation failed to reduce the incidence of this disease (Mayne (1996)).
Finally, dietary carotenoids have been found to be protective against various
forms of cataract (Taylor (1993)) and lutein and zeaxanthin can reduce the risk of
macular degeneration (Seddon et al. (1994)). The potential roles for lycopene in
human health and disease still awaits further evaluation (Clinton (1998); Gerster
(1997)).
Antioxidant and Pro-oxidant Activities of Carotenoids.
The antioxidant role of the carotenoids has been reviewed extensively elsewhere
(Bast et al. (1998); Byers and Perry (1992); Krinsky (1989, 1993), Paloza and
Krinsky (1992), Rousseau et al. (1992)). The antioxidant activity of carotenoids is
usually determined in vitro using solutions or model membrane systems such as
liposomes. Many of these studies measure the ability of carotenoids to inhibit
lipid peroxidation. In general, although many reports imply that carotenoids can
also act as antioxidants in vivo, there is still little direct evidence for this capability
(Krinsky (1993)). Some have even challenged the fact that carotenoids are
antioxidants (Crabtree and Adler (1997)).
The antioxidant activity of a carotenoid depends upon its structure (e.g., the
number and degree of conjugation of double bonds, degree of steric hindrance,
and presence of functional groups on the terminal rings) (Terao (1989), Miki
(1991)). For example, Miller et al. (1996) showed that lycopene (11 conjugated
double bonds, no terminal cyclohexene rings) was nearly three-times more
efficient at scavenging radicals than α-carotene (9 conjugated double bonds, a
higher degree of steric hindrance, and two terminal cyclohexene rings) and 100times more efficient than astaxanthin (11 conjugated double bonds, and electron
withdrawing carbonyl and hydroxyl groups on cyclohexene rings). Thus the
hydroxyl free radical and peroxyl radical scavenging abilities of the carotenoids
are found to be lycopene > β,β-carotene = zeaxanthin > isozeaxanthin >
astaxanthin (Woodall et al. (1997)).
Carotenoids are electron rich compounds and can readily react with electron
deficient compounds such as the ROS. The resulting charge on the carotenoid
product is delocalized through hyperconjugation, thus inferring stability and
rendering it less reactive. There have been several studies on the reaction
WWW.ESAINC.COM
365
mechanisms of carotenoids and different ROS and RNS (Liebler and McClure
(1996); Rice-Evans et al. (1997)) (Figure 4.17). For example, β-carotene reacts
with peroxyl radicals to form the peroxyl radical-β-carotene adduct
[β-carotene......RO2]• (Burton and Ingold (1984)), while reaction with NO2• forms
the β-carotene radical cation [β-carotene]•+ (Everett et al. (1996)). Under certain
conditions, both the carotenoid adduct and carotenoid radical cation can be
formed simultaneously. The resulting carotenoid adduct and the carotenoid
radical cation both undergo slow bimolecular decay to non-radical products (e.g.,
epoxides and carbonyl-containing chain cleavage products). The carotenoid
radical cation can also be rapidly scavenged by tocopherol (Mortensen and
Skibsted (1997)). Interestingly, some RNS appear to be capable of nitrating
β-carotene’s rings producing 4-nitro-β-carotene. However, as this appears to only
take place in the gas phase, formation of this adduct is biologically unimportant.
β-Carotene
R
R
[β-Carotene]+
[β-Carotene]
R
R
β-Carotene-R
SUBSTITUTION
[β-Carotene-R]
R
O2
R-β-Carotene-R
R-β-Carotene-O2
ADDITION
AUTOOXIDATION
Figure 4.17 Possible Reaction Mechanisms Of β-Carotene.
β-Carotene shows several antioxidant activities in vitro. It is a chain breaking
antioxidant and scavenges lipid peroxyl, nitrogen dioxide, thiyl, RSO2•, and the
trichloromethylperoxyl radicals (e.g., typical rate constants for the scavenging of
carbon-centered and peroxyl radicals by β-carotene are ~1.0 x 104 and ~1 to 50 x
105 M-1s-1, respectively (Appendix 2.2) (Iannone et al. (1998); Ozhogina and
Kaisikinia (1995)). It scavenges and quenches singlet oxygen (Frei and Ames
(1991); Miller et al. (1996) and references therein; Mortensen et al. (1997);
Ozhogina and Kaisikinia (1995), Rice-Evans and Diplock (1993)) and readily
reacts with peroxynitrite (Pannala et al. (1999) and references therein).
WWW.ESAINC.COM
366
Lycopene is one of the most abundant carotenoids in the Western diet (Gerster
(1997)). It is the most effective singlet oxygen quencher of all the carotenoids,
can effectively scavenge peroxyl radicals and has been reported to protect
against some human cancers (Di Macio et al. (1989); Gerster (1997); Klebanov
et al. (1998); Stahl and Sies (1996)). Lycopene is also more effective than
β-carotene at scavenging peroxynitrite (Pannala et al. (1998b)) and hypochlorous
acid (Panasenko et al. (1997)).
The antioxidant chemistry of carotenoids is intimately dependent upon the
oxygen tension (pO2). At the low pO2 typically found in tissues carotenoids act as
antioxidants but at high pO2 they can auto-oxidize and show pro-oxidant behavior
(Burton and Ingold (1984)). The exact mechanism is unclear but may involve the
addition of oxygen to the lipid peroxyl-carotene radical intermediate forming a
peroxyl radical adduct (Eqns 4.25 and 4.26) capable of promoting lipid
peroxidation. Furthermore, the formation of auto-oxidation products (e.g.,
epoxides) increases with pO2 and these can decompose to produce alkoxyl
radicals capable of promoting lipid peroxidation (Kennedy and Liebler (1992);
McClure and Liebler (1994)). The pro-oxidant actions of carotenoids in biological
systems have recently been reviewed (Palozza (1998)).
β-Carotene + ROO• → ROO-β-Carotene•
Eqn 4.25
ROO-β-Carotene• + O2 → ROO-β-Carotene-OO•
Eqn 4.26
Retinoids.
Of the 600 or so naturally occurring carotenoids, about 50 have vitamin A activity.
Provitamin A carotenoids can be converted enzymatically in the intestinal
mucosa to produce retinal and finally retinol (vitamin A1). Over 90% of the total
body reserve of vitamin A1 is stored in the liver of well nourished individuals,
primarily in stellate (Ito or fat-storing) cells. The principal storage form of vitamin
A1 is as retinyl palmitate, with oleate and stearate occurring as the next most
prevalent esters (Blomhoff et al. (1992)). Unlike the carotenoids which are
relatively safe, the retinoids are toxic and excessive consumption can lead to a
variety of diseases (Meyers et al. (1996)). Vitamin A1 is transported by specific
binding-proteins and its circulating levels are strictly regulated (Olson (1993)).
WWW.ESAINC.COM
367
The Biological Activity of the Retinoids.
The retinoids (including retinal, retinol and retinoic acid and their isomers) (see
Figure 4.18) have a variety of important biological roles including signal
transduction in the eye (Wald (1968), Wolken, (1966)), maintenance of epithelial
tissue (Argiles et al. (1989)), and regulation of the immune system response
(West et al. (1991)). Vitamin A1 is also involved in the regulation of proliferation
and differentiation of many cell types (Blomhoff et al. (1991, 1992)), probably
through the action of retinoic acid on nuclear retinoic acid receptors (Kastener et
al. (1994); Kliewer et al. (1994)). Many retinoids have teratogenic activity (Kamm,
(1982)). For example, retinoic acid can cause developmental anomalies in
prenatal systems and may play a role in coordinating cellular development
(Thaler et al. (1993); Tickle et al. (1985); Wagner et al. (1990)). Retinoids are
also involved in the establishment of the development axis of the central nervous
system (Durston et al. (1989); Sundin and Eichele (1992); Wagner et al. (1990)).
Finally, retinoic acid may also play a role in melatonin synthesis by activating
HOMT (Bernard and Klein (1995)). Vitamin A may play a protective role in
cardiovascular disease (Palace et al. (1999)).
CH2OH
CH2 OH
ALL-TRANS RETINOL
13-CIS RETINOL
CHO
ALL-TRANS RETINAL
11-CIS RETINOL
CH2 OH
CO2 H
CO2H
ALL-TRANS RETINOIC ACID
O
O
9-CIS RETINOIC ACID
4-OXO-ALL-TRANS
RETINOIC ACID
4-OXO-13-CIS
RETINOIC ACID
CO2 H
CO2H
Figure 4.18 The Structure Of Biologically Important Retinoids.
WWW.ESAINC.COM
368
Antioxidant and Pro-oxidant Activities of the Retinoids.
Like the carotenoids, the retinoids have a conjugated double bond system and
can act as antioxidants. The retinoids have half the number of double bonds of
the carotenoids and are therefore less capable of delocalizing charge. One
consequence of this is that retinoids oxidize electrochemically at a higher
potential than the carotenoids (see below). Thus, retinoids are chemically less
effective antioxidants. This may be one of the reasons why nature uses them for
functions other than as antioxidants. In fact vitamin A1 is protected from oxidation
in vivo by a variety of antioxidants including ascorbic acid, GSH and vitamin E,
thereby keeping it in its active form (Olson (1996)).
Vitamin A acts as a pro-oxidant sensitizer, exciting oxygen into destructive singlet
oxygen (Halliwell and Gutteridge (1999)). This may be a problem for the eye
which not only contains vitamin A (as rhodopsin) but is also particularly abundant
in poly-unsaturated fatty acids that can readily undergo singlet oxygen-induced
lipid peroxidation. Several studies now suggest that lipid peroxides and cytotoxic
breakdown products may lead to severe retinal damage (Halliwell and Gutteridge
(1999)). It should not be surprising, therefore, that the eye is usually well
protected by GSH and other antioxidants capable of acting against ROS-induced
damage.
Measurement of Carotenoids and Retinoids.
The tissue levels for some carotenoids and retinoids are presented in Table 4.1
and the various analytical methods used to measure them in Table 4.5.
Unfortunately, many carotenoids are extremely light and oxygen sensitive, and
can undergo rapid decomposition so due care must be exercised during sample
preparation (Wyss (1995)). Analytical methods differ in sensitivity, selectivity,
complexity and limitations. HPLC-UV is usually adequate when tissue levels are
high as in plant material. However, when tissue levels are low (as in animal
samples) concentration steps may be required. For example, Gunderson et al.
(1997) used a complex HPLC-UV method with on-line solid phase extraction
(consisting of five valves and three columns) to measure low levels of retinoic
acid in plasma. GC procedures can result in the decomposition of more labile
analytes (e.g., retinol and its esters are converted to anhydroretinol) and require
derivatization to make the analytes volatile. Although not as common as
spectrophotometric approaches, HPLC-ECD has proven very useful. Gamache
et al. (1997a) used HPLC with gradient elution and coulometric electrode array
detection to measure a variety of retinoids, carotenoids, ubiquinone and vitamin
K simultaneously in human plasma (see Figure 4.19). Human plasma data have
been verified as part of a NIST/NCI micronutrients measurement quality
assurance program: measurement reproducibility, repeatability, stability, and
relative accuracy for fat-soluble vitamin-related compounds in human sera. See
Table 4.6. Compounds with conjugated double bonds (e.g., carotenoids,
WWW.ESAINC.COM
369
retinoids, tocopherols, vitamin D2 and D3) are all electrochemically active, a
direct result of their ability to delocalize charge. Thus, the very chemical
characteristics that make a compound a good antioxidant also render it
measurable by electrochemical detection. Gamache’s method has now been
extended to include astaxanthin, β-cryptoxanthin, zeaxanthin and all-trans
retinoic acid. With a limit of detection of <10pg (on column) for most analytes,
this approach enables the direct measurement of many carotenoids and retinoids
in animal and human tissues. When coupled to the greater resolution obtained by
the use of a C30 column, the abundance of various carotene isomers in
processed carrots, human plasma and cervical tissue samples can be
determined (Ferruzzi et al. (1998); Gamache et al. (1997b; 2003)). See ESA
Application Note 70-4927 Carotenoid Isomers and Figure 4.20.
Analyte
Apocarotenoids, retinoids
Carotenoids
Carotenoids
Carotenoids
Carotenoids
Carotenoids
Retinoic acids, retinal,
retinol
Retinoic acids
Retinoic acids
Retinoid radicals
Retinoids
Retinoids
Retinoids, carotenoids,
and tocopherols,
ubiquinone, vitamin K
Retinyl esters, retinol
Vitamin A1
Vitamin A1
Vitamin A1, β-carotene,
tocopherol
Vitamin A1, carotenoids,
tocopherol(s)
Method
GC-MS
HPLC-ECD C30
column
HPLC-UV
HPLC-UV, HPLC-MS
HPLC-UV, C30
column
HPLC-UV, HPLC-IR
Capillary-LC-ECD
HPLC-PDA
HPLC-UV
HPLC-EPR, HPLCECD
HPLC-ECD
HPLC-UV
HPLC-ECD
HPLC-UV
HPLC-ECD
HPLC-UV
HPLC-UV, HPLCECD
HPLC-UV
Reference
Furr et al. (1992)
Gamache et al. (1997a); Ferruzzi et al.
(1998)
Barua and Furr (1992); Handelman et
al. (1992); Nurdin (1991); Parker et al.
(1993); Schmitz et al. (1993)
van Breemen (1996)
Bell et al. (1997); Emenhiser et al.
(1995,1996)
Stanchar et al. (1988)
Hagen et al. (1996)
Gundersen et al. (1997)
Dimitrova et al. (1996)
Iwahashi et al. (1987)
Bryan et al. (1991); MacCrehan and
Schonberger (1987b); Sakhi et al.
(1998)
Noll (1996); Stanchar and Zonta (1984)
Gamache et al. (1997b; 2003)
Got et al. (1995)
Wring and Hart (1989); Wring et al.
(1988)
Peng et al. (1987)
MacCrehan and Schonberger (1987a)
Barua and Olson (1998); Barua et al.
(1993); Kaplan et al. (1987, 1990);
Miller et al. (1984, 1985); Steghens et
al. (1997); Talwar et al. (1998); Xu et al.
(1996)
Table 4.5 A Variety Of GC- And HPLC-Based Methods Can Be Used To
Measure The Carotenoids And Retinoids.
WWW.ESAINC.COM
370
γ-Tocopherol
Retinol
Response (µA)
0.80
Retinyl
Acetate
0.60
α-Tocopherol
Lutein
Retinyl
Palmitate
δ-Tocopherol
CoQ10
Lycopene
β-Carotene
0.40
0.20
-0.00
0.0
5.0
10.0
15.0
20.0
25.0
Retention time (minutes)
Figure 4.19 HPLC-CoulArray Chromatogram Showing Simultaneous
Measurement Of Several Fat Soluble Vitamin And Antioxidant Standards
(only 6 channels are shown for clarity. See ESA Application Note 10-1176 Fat
Soluble Vitamins for further details. See also ESA Application Note – 70-4935
Multivitamins in Tablets, Infant Formula and Milk.
Trans-Retinol
γ-Tocopherol
1
1
α-Tocopherol
1
1
Trans-β-Carotene
2
Trans-Lutein
2
Trans-Lycopene
3
Retinyl Palmitate
3
CoQ10
*Units: µg/L
NIST Categories:
1
NIST SRM 968c
Level I
Level II
Range*
Range*
Level I
Mean*
RSD %
Level II
Mean*
RSD %
814-868
472-496
861
3.1
489
3.9
3770-4030
1460-1660
4013
7.5
1565
7.1
7000-7940
16030-17750
8040
2.6
17666
4.2
141-173
344-438
153
10.9
365
12.6
40-54
61-75
44
13.7
69
8.0
100-160
140-200
148
8.3
191
9.7
30
80
27
21.0
72
15.3
520
-
535
3.2
-
-
2
Certified Values; Reference Values;
ESAL Data
3
Information Values
Table 4.6 Fat Soluble Vitamin And Antioxidant Levels In Plasma Using
ESA’s CoulArray Method As Part Of NIST/NCI Micronutrients Measurement
Quality Assurance Program. (Brown and Sharpless (1995); Duewer et al., (1997 and
1999)).
WWW.ESAINC.COM
371
Carotenoid Isomers
Trans-ß-carotene
Trans-ß-carotene
Trans-∝-carotene
Trans-∝-carotene
2.0
1.0
0.0
0.0
9-cis-ß-carotene*
9-cis-ß-carotene*
1.0
3.0
Lutein
2.0
4.0
Response (µA)
Trans-∝-carotene
Trans-∝-carotene
3.0
Lutein
Lutein
Response (µA)
4.0
9-cis-∝-carotene
9-cis-∝-carotene
Cooked Carrot
Trans-ß-carotene
Trans-ß-carotene
Raw Carrot
0.0
5.0
10.0
15.0
20.0
0.0
Retention time (minutes)
5.0
10.0
15.0
20.0
Retention time (minutes)
Figure 4.20 Analysis Of A Thermally Processed Carrot.
The isocratic analytical system consisted of a pump, an autosampler, a thermostatic chamber, a
twelve-channel CoulArray® detector and a UV/vis detector placed prior to the array.
LC Conditions:
Column:
Mobile Phase:
ESA Carotenoid C30 (4.6 x 250mm; 5µm).
Methanol – Methyl-tert-butyl Ether (MTBE) – 1.0M Ammonium Acetate,
pH 4.4 (63:35:2) (v/v/v).
Flow Rate:
1.0 mL/min.
Temperature:
28 oC.
Injection Volume:
10 µL.
Applied Potentials:
100, 160, 220, 280, 340, 400, 460 and 520mV vs. Pd.
Wavelength:
450nm.
See ESA Application Note 70-4927 Carotenoid Isomers for further details.
Quinones and Hydroquinones.
Coenzyme Q (Ubiquinone, ubiquinol).
Coenzyme Q10 (CoQ10), also called ubiquinone-50 (2,3-dimethoxy-5-methyl-6decaprenylbenzoquinone), was first discovered by Crane et al. (1957), and is one
of a number of important quinones found in biological systems (Figure 4.21)
(Crane and Navas (1997); Andree et al. (1999)). Although coenzyme Q was
earlier named vitamin Q, this was deemed inappropriate as ubiquinone can be
WWW.ESAINC.COM
372
synthesized de novo in all animal tissues. Coenzyme Q exists in three
biologically relevant forms: the fully oxidized quinone, the partially reduced
semiquinone radical (semiubiquinone) and the fully reduced ubiquinol (see Figure
4.22 for the relationship between these redox forms). The term coenzyme Q will
be used where the indeterminate form is most applicable and/or to discuss both
the oxidized and reduced forms together. In addition to serving as an electron
and proton carrier in the respiratory chain of mitochondria, evidence suggests
that ubiquinone can also, under certain conditions, act as a pro-oxidant.
Ubiquinol is a very important lipid soluble antioxidant.
O
O
CH3O
CH3
CH3
CH3
CH3
CH3O
(CH2CH=CCH2)10H
CH3
(CH2CH=CCH2)n H
O
O
PLASTOQUINONE
(n= 6 to 10)
UBIQUINONE-50
O
OH
O
O
NH
CH3
CH3
CH3
CH2CH=CCH2(CH2CH2CHCH2)3H
HO
HO
N
O
O
PHYLLOQUINONE
(VITAMIN K1)
O
O
PYRROLOQUINOLINE QUINONE
Figure 4.21 Some Quinones Found In Biological Systems.
Biology of Coenzyme Q.
The biosynthesis of coenzyme Q in mammalian cells involves the interplay
between two metabolic pathways: the 4-hydroxybenzoate pathway (using
tyrosine or phenylalanine) for synthesis of the quinone moiety and the
mevalonate pathway for production of polyprenyl side-chain (Appelkvist et al.
(1994)).14 The mevalonate pathway also produces cholesterol, dolichol, and two
cytoplasmic intermediates (farnesyl-PP and gereanylgeranyl-PP) which are
capable of isoprenylating proteins. In humans the major form of coenzyme Q has
14
A variety of drugs have been used to manipulate CoQ10’s levels. The HMG-CoA reductase inhibitors, pravastatin,
lovastatin and related cholesterolemic drugs, have been used to decrease CoQ10’s levels in a variety of tissues
(Appelkvist et al. (1993); Goldstein and Brown (1990); Low et al. (1992); Willis et al. (1990)). On the other hand, the
cholesterol synthesis inhibitor squalestatin, or the peroxisome proliferators clofibrate and di(2-ethylhexyl)phthalate lead to
elevation in ubiquinone levels (Aberg et al. (1994); Thelin et al. (1994)). The importance of such CoQ10 manipulations on
the effect of oxidative stress still awaits clarification.
WWW.ESAINC.COM
373
ten isoprenyl subunits (CoQ10), whereas rats have nine (CoQ9). Under normal
conditions enough coenzyme Q is produced in the cell to satisfy its needs.
However, pathological conditions have been associated with disturbances in
coenzyme Q levels. Decreased coenzyme Q levels have been associated with
cardiomyopathy (Folkers et al. (1985); Mortensen (1993)), encephalomyopathy
(Ogasahara et al. (1989)), degenerative muscle disease (Karlsson et al. (1990))
and hepatocellular carcinoma (Eggens et al. (1989)). Elevated levels have been
found in Alzheimer’s disease (Soderberg et al. (1992)), prion disease in mice
(Ericsson and Dalner (1993)) and hyperplastic liver nodules in rats (Olsson et al.
(1991)). Whether a change in ubiquinone levels is the cause or an effect of the
disease is still under investigation.
A
OH
CH3O
CH3
CH3
CH3O
(CH2CH=CHCH2)10H
OH
-H+ - e -
UBIQUINOL-50
O
CH3O
CH3
2H + + 2e -
CH3
CH3O
O
CH3O
(CH2CH=CHCH2)10H
OH
CH3
SEMIQUINONE
-H+ - e CH3
CH3O
(CH2CH=CHCH2)10H
O
UBIQUINONE-50
B
OH
CH3O
CH3
CH3
CH3O
O2
O2 -
(CH2CH=CHCH2)10H
OH
UBIQUINOL-50
O2
O
CH3O
H2O2
CH3
CH3
CH3O
O
OH
CH3
CH3O
SEMIQUINONE
O2
CH3
CH3O
(CH2CH=CHCH2)10H
(CH2CH=CHCH2)10H
O2 -
O
UBIQUINONE-50
Figure 4.22 The Redox Relationship Between
Ubiquinone, Ubisemiquinone And Ubiquinol
(A), And The Involvement Of Ubiquinol In ROS
Production (B).
Current evidence suggests that coenzyme Q synthesis begins in the
endoplasmic reticulum and is completed in the Golgi body from where coenzyme
Q is transported to other cellular locations (Ernster and Dallner (1995) and
WWW.ESAINC.COM
374
references therein). Some coenzyme Q passes through the plasma membrane to
the blood where it is bound to serum lipoproteins. The amount of ubiquinone and
ubiquinol in rat and human is tissue dependent, with the heart showing the
greatest abundance of ubiquinone and the lung the least (Table 5.1).
Furthermore, the ubiquinol/ubiquinone ratio is also tissue specific approaching
100% in the pancreas, liver and intestine, but only 25% in the lung (Aberg et al.
(1992)). Although it is not clear by which mechanism(s) coenzyme Q is reduced
in membranes (other than the mitochondrial inner membrane), several possible
enzymes have been suggested, including microsomal NADH- (and NADPH)
cytochrome reductases as well as the NADH dehydrogenases associated with
the outer mitochondrial and plasma membranes (Ernster and Dallner (1995) and
references therein; Kishi et al.(1997); Takahashi et al. (1995, 1996a,b).
Coenzyme Q’s primary role is as part of the electron-transport (respiratory) chain
of the inner mitochondrial membrane, where it acts as an obligatory two-electron
and two-proton carrier molecule and forms a redox link between flavin
dehydrogenases and cytochromes (Chapter 2). Coenzyme Q is also found in
extra-mitochondrial redox chains where it plays a similar role (Crane et al.
(1993)).
Antioxidant and Pro-oxidant Activities of Coenzyme Q
Abundant evidence shows that ubiquinol is an antioxidant (Table 4.7) (Crane and
Navas (1997); Nohl et al.(1997)). It appears that ubiquinol acts directly to inhibit
both the initiation (by reduction of the perferryl radical) and propagation (by
reducing the lipid peroxyl radical) phases of lipid peroxidation (Beyer (1990,
1991); Ernster and Dallner (1995); Ernster and Forsmark-Andree (1993)).
Another possible antioxidant function of coenzyme Q cycle is the regeneration of
α-tocopherol formed during inhibition of the propagation phase of lipid
peroxidation. Interestingly the isoprenoid side chain can affect antioxidant activity
in membrane preparations with short chain homologs showing the greatest
activity (Kagan et al (1990)).
The fact that the coenzyme Q redox cycle involves electron transfer raises the
possibility that ROS could be generated as byproducts, therefore suggesting a
pro-oxidant role for this redox couple (Figure 4.22B) (Nohl et al. (1996)). For
measurement of the oxidation activity of coenzyme Q see Kagan et al. (1994)).
WWW.ESAINC.COM
375
Evidence for Antioxidant Activity
IN VITRO – SUBCELLULAR SYSTEMS
• Reduced CoQ6 is four times more effective than the
oxidized form in reducing the stable free radical diphenylp-picrylhydrazyl
• Reduced CoQ6 is much more effective than the oxidized
form in preventing hemoglobin induced peroxidation of
arachidonic acid emulsions
• Reduced CoQ10 inhibits ascorbate-Fe2+ -induced
peroxidation of phosphatidylcholine liposomes
• CoQ2 to CoQ10 homologues protect lipid vesicles from
peroxidation. CoQ3 protects phospholipid vesicles from
HO•-radical damage
• Reduced and oxidized CoQ6 protects mitochondrial lipids
from light induced peroxidation
• CoQ10 protects submitochondrial particles from lipid
peroxidation
• Reduced CoQ10 prevents lipid peroxidation directly and
spares α-tocopherol in liposomal membranes exposed to
pro-oxidant conditions
• Reduced CoQ10 is more effective at inhibiting liver
microsome lipid peroxidation
• Submitochondrial particles isolated from heart of
exercise-trained animals had higher levels of CoQ10 and
following succinate (to reduce oxidized CoQ10) showed
less lipid peroxidation than sedentary (age-matched)
controls
• CoQ10 protects against ROS inactivation of respiratory
chain components and inhibits membrane lipid
peroxidation
• Reduced CoQo and reduced CoQ2 readily react with nitric
oxide. Under anaerobic conditions ubiquinones and the
nitroxyl anion are formed. Under aerobic conditions the
reaction proceeds with the formation of peroxynitrite
CELLULAR SYSTEMS
• CoQ10 protects cultured cells against free radical damage
during one-electron reduction of antitumor quinones
IN VIVO SYSTEMS (intact animal and clinical studies)
• Conditions where the metabolic rate and ROS production
are increased (e.g., following endurance exercise) are
associated with elevated CoQ levels in highly aerobic
tissues
•
Administration of CoQ10 protects animals against the
effects of ischemia
•
Administration of CoQ10 to humans increases circulating
CoQ10 levels in lipoproteins and protects LDL from lipid
WWW.ESAINC.COM
References
Mellors and Tappel (1966)
Mellors and Tappel (1966)
Booth et al. (1982)
Landi et al. (1984, 1985,
1989)
Mellors and Tappel (1966)
Takeshige et al. (1980)
Frei et al. (1990)
Beyer (1988); Ernster and
Nordenbrand (1967)
Beyer (1990) and
references therein
Solaini et al. (1987)
Poderosa et al. (1999)
Powis (1989); Takahashi et
al. (1988)
Beyer (1990); Beyer et al.
(1962, 1984); Boveris et al.
(1969); Davies et al. (1982);
Lang et al. (1987);
Ogasahara et al. (1989);
Pederson et al. (1963);
Salminen and Vihko
(1983a,b); Saran and Bors
(1989)
Katagiri et al. (1986);
Kawasaki et al. (1986);
Shibata et al. (1986);
Sugiyama et al. (1980)
Mohr et al. (1992)
376
•
•
peroxidation.
Administration of CoQ10 protects against oxidative
damage resulting from the administration of carbon
tetrachloride or ethanol
Administration of CoQ to humans has been used to treat
a variety of diseases associated with altered oxidative
metabolism such as congestive heart failure and certain
neurological diseases (e.g., Kearns-Sayre syndrome)
Bertelli et al. (1986); Brattin
et al. (1985); Quinn et al.
(1980); Therman et al.
(1973); Uysal et al. (1985)
Bresolin et al. (1988);
Fisher et al. (1986); Folkers
and Yamamura (1977,
1980, 1984, 1986);
Karlsson et al. (1986);
Lenaz (1985); Mortensen et
al. (1986); Ogasahara et al.
(1985, 1986); Takahashi et
al. (1986); Trumpower
(1982); Yamamura et al.
(1980); Zeirz et al. (1989)
Table 4.7 In Vitro And In Vivo Evidence That Coenzyme Q10 Acts As An
Antioxidant.
Measurement of Coenzyme Q.
Many different approaches have been used to measure tissue levels of
ubiquinols, ubiquinones, and their derivatives including GC, GC-MS, and
chemiluminescence (Frei et al. (1988); Gurtler and Blomstrand (1971); Morimoto
et al. (1973)).15 By far the most common approach, however is HPLC. Aberg et
al. (1992) used HPLC-UV to measure the redox status of CoQ9 and CoQ10 in
human and rat tissues. Andersson (1992) used HPLC-PDA to measure the redox
status of CoQ9 and CoQ10 and the metabolite ubichromenol in pharmaceutical
preparations. Some researchers have used a combination of HPLC with both UV
and EC detectors for the simultaneous measurement of ubiquinone and
tocopherols (Lang and Packer (1987); Lang et al. (1986); Poddo et al. (1996)).
This combination of detectors was necessary as ubiquinones cannot be
measured oxidatively with ECD. To overcome the unnecessary expense of
having to use two different detector modalities, some approaches used chemical
reduction with either sodium borohydride (Okamoto et al. (1988)) or platinum
(Wakabayashi et al. (1994)) to convert the ubiquinones to their corresponding
electrochemically oxidizable ubiquinols. Thus ubiquinones can be indirectly
measured using HPLC-ECD. However, the use of chemical reducing agents not
only increases sample preparation time but also often leads to chromatographic
issues, especially when attempting to measure low analyte levels in biological
samples. Yamashita and Yamamoto (1997) simplified the chemical reduction
step by using an on-line (unspecified) reactor column placed before the
electrochemical cell. Coulometric-based assays offer a much simpler approach to
the measurement of ubiquinones. Instead of requiring chemical reduction,
ubiquinones are electrochemically reduced before being measured oxidatively.
This method permits the quantitation of both reduced and oxidized CoQ10
15
As with many antioxidants, regardless of the approach used, due care must be exercised to
prevent changes in analyte redox status during isolation and analysis.
WWW.ESAINC.COM
377
simultaneously. With two serially placed coulometric electrodes, the upstream
electrode can electrochemically reduce ubiquinones to ubiquinols with 100%
conversion efficiency. Thus the naturally occurring ubiquinols, as well as those
formed from coulometric reduction of the ubiquinones, can all be measured at the
downstream “oxidizing” electrode. Several researchers have taken advantage of
this approach including Edlund (1988); Finckh et al. (1995; 1999); Gamache et
al. (1997; 2003); Grossi et al. (1992); Kaikkonen et al. (1997); Lagendijk et al.
(1996); Motchnik et al. (1994); Tang et al., (2001, 2002).
Plastoquinone.
Plastoquinone is structurally related to ubiquinone (Figure 4.21) and, like
ubiquinone, is also involved in redox reactions. Plastoquinone is primarily found
in chloroplasts of higher plants (and is also present in other photosynthetic
organisms) where it acts as an electron acceptor and forms part of the
photosystem II complex. This complex is responsible for light-driven transfer of
electrons from water to plastoquinone forming oxygen and plastoquinol. Thus the
energy from two photons is stored in the reducing potential of plastaquinol.
Plastoquinol can then feed these electrons into a proton-pumping electron
transport chain that is linked to the photosystem I complex. The proton gradient
is ultimately used to produce ATP (Stryer (1988)). Plastaquinol can be involved in
single electron redox reactions producing semiplastoquinone or two electron
redox reactions producing plastoquinone (cf Figure 4.22). The ability for these
reactions to be coupled to ROS production is but one of the oxidative stressors
occurring in chloroplasts.
Vitamin K.
Vitamin K consists of two groups of naphthoquinones, the phylloquinones
(vitamin K1) (Figure 4.21 and the menaquinones (vitamin K2). The members of
each group differ in the length of their phytyl side-chain. Menadione (vitamin K3)
is a synthetic vitamin K analog. Vitamin K is essential in mammals and its daily
requirement is met by a combination of dietary intake (phylloquinone) and
microbial synthesis in the large intestine (menaquinone). Vitamin K is needed for
the post-translational carboxylation of certain glutamate residues in proteins to γcaboxyglutamate. This modification is extremely important in the activation of
certain blood-clotting factors (factors II (prothrombin), VII, IX and X).
The exact roles of the phylloquinones and menaquinones in oxidative metabolism
still remains unclear. There is some evidence that they (in their reduced forms)
can act as antioxidants capable of inhibiting lipid peroxidation (Fiorentini et al.
(1997); Ohyashiki et al.(1991); Tampo and Yonaha (1996); Vervoort et al.
(1997)). However, there is considerable evidence that menadione acts as a prooxidant and cytotoxin in vitro. Like many quinones it appears that several
WWW.ESAINC.COM
378
mechanisms can account for menadione’s toxicity. First, it can react with cellular
nucleophiles such as amines and thiols (e.g., GSH) leading to the formation of
aryl-conjugates (see below) (Brunmark and Cadenas (1989); Gant et al. (1988)).
Second, reaction of menadione with GSH through aryl substitution and by
oxidation of GSH to GSSG depletes cellular GSH levels (Bellomo et al. (1987)).
Third, menadione metabolism is associated with an inhibition of GSH-reductase
that also leads to the depletion of GSH (Bellomo et al. (1987)). Fourth, quinones
are readily reduced under physiological conditions by either direct reactions with
NADH (or NADPH) (Kukielka et al. (1990)) or through those involving enzymes
such as diaphorase (NAD(P)H: [quinone acceptor] oxidoreductase) (Murphy et
al. (1991); Thor et al. (1982)), in one- or two-electron processes, yielding the
semiquinone and hydroquinone, respectively (Figures 2.7 and 4.22). Metaldependent auto-oxidation reforms the quinone with concomitant regeneration of
ROS (Brunmark and Cadenas (1989); Comporti (1989); O’Brian (1991)).
Consequently, menadione promotes DNA damage (Ngo (1993); Nutter et al.
(1992)) but without production of 8OH2’dG (Fischer-Nielsen et al. (1995)). It also
causes hemolysis in a variety of species (Munday et al. (1994) and references
therein). Conversely, menadione has been reported to be beneficial, having
anticancer, antimalarial, and antileishmanial activity, and has been used as an
electron carrier in the treatment of a patient with a severe defect in complex III of
the electron transport chain (Eleff et al. (1984); Munday et al. (1994) and
references therein).
Technique
HPLC-UV
HPLC-Fl
HPLC-electrofluoresence
HPLC-ECD (single or dual
electrodes)
HPLC-ECD (coulometric
electrode array)
References
Haroon et al. (1982); Lefevere et al. (1982); Shearer
et al. (1980)
Abe et al. (1979)
Langenberg and Tjaden (1984a,b); Moussa et al.
(1989, 1994)
Haroon et al.(1984); Hart et al. (1984, 1985);
Hiroshima et al. (1981); Isshiki et al. (1988);
Kaikkonen et al. (1997); McCarthy et al. (1997);
Rawlinson et al. (1998)
Gamache et al. (1997)
Table 4.8 Some HPLC-Based Methods For Vitamin K Measurement.
A variety of analytical approaches are used to measure vitamin K in different
sample matrices including HPLC coupled with: UV, fluorescence, coulometricelectrofluoresence, amperometric, coulometric, and coulometric array detection
(Table 4.8).
WWW.ESAINC.COM
379
Pyrroloquinoline quinone.
Pyrroloquinoline quinone (2,7,9-tricarboxypyrroloquinoline quinone; methoxatin)
(PQQ) is a widely distributed redox-active cofactor and essential nutrient (Figure
4.21) (McIntire (1992)). PQQ is reported to be an essential and versatile cofactor
for a variety of enzymes including dehydrogenases and oxidases, and takes part
in a variety of hydroxylation, transamination, decarboxylation and hydration
reactions (Duine (1989); McWhirter and Klapper (1990)). Others however, using
more selective and sensitive approaches, propose that PQQ has been
misidentified and that another cofactor, topaquinone, may be present instead
(Harris (1992); Kumazawa et al. (1990)).
Free PQQ has been found in a variety of animal tissues. It can catalyze
dioxygen-superoxide interconversion, and participates in both superoxide
generation (in the respiratory burst) and scavenging (acting as an antioxidant)
(Bishop et al. (1994, 1995); Gallop et al. (1993)). In its role as antioxidant, PQQ
scavenges ROS, may convert xanthine oxidase into xanthine dehydrogenase,
spares GSH and protects neurons against the neurotoxic action of NMDA (Gallop
et al. (1993) and references therein). The mammalian enzyme PQQ reductase
also participates in mechanisms protecting tissues against oxidative stress
(Christensen (1994)). Under certain conditions it can act as a pro-oxidant
producing hydrogen peroxide during metal-induced auto-oxidation (He et al.,
(2003)).
Tocopherols.
In literature, vitamin E has become synonymous with the most abundant form
found in human tissues, α-tocopherol (2,5,7,8-tetramethyl-2-(4’, 8’, 12’-trimethyltridecyl)-6-chromanol). In fact, vitamin E is not a single compound but consists of
a group of eight naturally occurring, lipophilic molecules including: the
tocopherols (which differ in the number of methyl groups on the chromanol ring
(Figure 4.23)) and the tocotrienols (which also possess an unsaturated tail).
Furthermore, each of the tocopherols (and tocotrienols) can produce a
corresponding tocopheryl quinone (and tocotrienyl quinone) during oxidation
processes (Figure 4.24).
In the strictest sense, the use of vitamin E to represent α-tocopherol is incorrect
as the other forms of tocopherol (and tocotrienols) also show varying degrees of
biological activity (e.g., influencing membrane fluidity, controlling prostaglandin
and leukotriene synthesis, and regulating nucleic acid synthesis and gene
expression) and antioxidant capacity (Sokol (1989) and references therein). The
antioxidant and biological activities of the different forms of vitamin E show great
interspecies variability. For example, γ-tocopherol has about 50% the antioxidant
activity of α-tocopherol (Mukai (1993)) but only 10-30% of its biological activity
WWW.ESAINC.COM
380
(Bunyan et al. (1961)). On the other hand, α-tocotrienol has higher (Serbinova et
al. (1993)) or equivalent (Suarna et al. (1993)) antioxidant activity to α-tocopherol
but only 30% its biological activity (Bunyan et al. (1961)).
HO
H3 C
HO
CH 3
O
CH 3
CH 3
4'
8'
CH 3
CH 3
HO
CH 3
CH 3
2
H3C
O
O
CH 3
H3C
CH 3
CH 3
O
CH 3
CH 3
CH 3
RRR-β -TOCOPHEROL
HO
CH 3
CH 3
CH 3
CH 3
CH 3
RRR-γ -TOCOPHEROL
HO
CH 3
CH 3
CH 3
CH 3
CH 3
RRR-α -TOCOPHEROL
CH 3
CH 3
CH 3
RRR-δ-TOCOPHEROL
NO 2
CH 3
O
CH 3
CH 3
CH 3
CH 3
CH 3
5-Nitro-RRR- γ -TOCOPHEROL
CH 3
HO
HO
O
H3 C
CH 2CO 2H
CH 3
H3C
HO
O
CH 3
CH 3
4'
8'
CH 3
CH 3
CH 3
CH 3
CH 3
O
CH 3
CH 3
RRR-β -TOCOTRIENOL
CH 3
CH 3
RRR-γ -TOCOTRIENOL
CH 3
CH 3
CH 3
CH 3
CH 3
CH 3
HO
O
CH 3
H3C
RRR-α -TOCOTRIENOL
CH 3
CH 3
HO
2
CH 3
CO 2 H
TROLOX
IRF1005
CH 3
O
CH 3
CH 3
HO
CH 3
CH 3
CH 3
O
CH 3
CH 3
CH 3
RRR-δ-TOCOTRIENOL
Figure 4.23 The Structures Of Naturally Occurring Tocopherols,
Tocotrienols, 5-Nitro-γ-Tocopherol And The Synthetic Water Soluble
Analogs, Trolox™ (Hoffman-Laroche) And IRF1005.
The tocopherols show complex stereochemistry. Although the tocopherol
molecule has three centers of chirality at the 2, 4’ and 8’ carbons, it only occurs
naturally in one form, the RRR isomer (Figure 4.23). The RRR form also has two
ambo-forms (RRR and SRR) and each of these has four isomers (ambo-RRR
has the RRR, RSR, RRS and RSS isomers; ambo-SRR has the SRR, SSR, SRS
and SSS isomers). Unless otherwise stated most vitamin E supplements usually
consist of a mixture of isomers and may even totally lack the natural RRR
isomer.
WWW.ESAINC.COM
381
CH 3
CH 3
HO
-H+,
CH 3
H3 C
O
-e -
CH 3
H3 C
R
CH 3
O
O
CH 3
TOCOPHEROL
R
TOCOPHER(OX)YL
RADICAL
2H+, +2e -H2 O,
+H2 O,
-H+, -e -
CH 3
-2H+, -2e -
HO
HO
H3 C
OH
CH 3
R
CH 3
O
H3 C
CH 3
HO
CH 3
O
R
CH 3
TOCOPHERYLHYDROQUINONE
TOCOPHERYLQUINONE
Figure 4.24 Redox Reactions Of The Tocopherols.
The Biology of Tocopherols.
In man, tocopherols are essential and must be obtained from the diet. Adequate
consumption of vitamin E is mandatory for good health. The role of tocopherols in
human health and disease has been extensively reviewed (e.g., see Diplock et
al. (1989); Packer et al. (1993); Sokol (1989); Traber and Packer (1995); Traber
and Sies (1996)). Being lipophilic, the tocopherols are absorbed, processed and
transported like most other fats. This is a very complex field but fortunately has
been excellently reviewed by Traber (1994). Unlike other fat soluble vitamins, the
tocopherols have no specific plasma carrier proteins, rather they are transported
by plasma lipoproteins.16 The circulating level of tocopherol, and consequently its
biological activity (e.g., the amount of a particular tocopherol that is necessary to
alleviate various symptoms in vitamin E deficient rats) is controlled by hepatic
tocopherol binding protein (Traber (1994)). Human diets typically contain large
amounts of γ-tocopherol, mostly derived from corn and soybean oils, and these
usually exceed the intake of α-tocopherol considerably (Bieri and Evarts (1974)).
Nonetheless, the plasma concentrations of γ-tocopherol rarely reach 20% of αtocopherol levels (Traber and Kayden (1989); Traber (1994)). The tocotrienols
are present in high concentrations in palm oil. Consumption of palm oil can lead
16
Vitamins A and D are stored in the liver and can easily reach toxic levels. So far there are no reports of any toxic effects
of vitamin E following supplementation (Bendich and Machlin (1988)). Tocopherols are readily excreted through bile and in
the feces, thereby preventing excessive hepatic accumulation.
WWW.ESAINC.COM
382
to elevated, but non-sustained, increases in circulating levels of tocotrienols,
again a consequence of poor binding to tocopherol binding protein. Literature
tissue levels for the tocopherols, tocotrienols and their quinones are presented in
Table 4.3.
There has been considerable controversy regarding the biological activity of
RRR-α-tocopherol and synthetic (all rac)- α-tocopherol (Veris (1994)). Much of
the debate has centered upon blood levels following consumption of
experimental amounts of natural or synthetic α-tocopherol. Unfortunately, the
level of tocopherol in the blood may not be the best index for biological activity as
this is dependent upon several factors including: the rate of absorption, the rate
of release by the liver, the rate of transfer to other tissues and the animal model
being used. Knight and Roberts (1985) reported that, following administration of
either natural or synthetic α-tocopherol to rabbit pups, the plasma level of
tocopherol was of little value in establishing efficacy of vitamin E therapy. Traber
et al. (1990) reported that humans could discriminate between natural and
synthetic α-tocopherol administered as their esters. Acetate and succinate esters
are often used in commercial preparations as they are more stable than the free
forms. Following ingestion pancreatic enzymes hydrolyze the ester.
Consequently, almost all of the vitamin E in blood and tissue is in the free form
following ingestion of tocopherol esters. Following ingestion of RRR-αtocopheryl- and SRR-α-tocopheryl acetate there was little difference in the
circulating forms until 11 hours post administration. The circulating RRR-/SRR-αtocopherol ratio continued to increase with time and reached 4-fold after 24
hours. These researchers concluded that this discrimination was not due to
differences in absorption but rather to differences in secretion by the liver. Chung
et al. (1992) experimented with pigs. These animals are often regarded to be
more similar to humans in their metabolism than rats. They reported that RRR-αtocopherol had much more biological potency than the synthetic forms when
using the pig rather than rat as the model.
Antioxidant, Pro-Oxidant and other Reactions of the Tocopherols.
A major biological role for the tocopherols is their ability to act as potent peroxyl
radical scavengers and chain breaking antioxidants (Liebler (1993); Wolf et al.
(1998)). α-Tocopherol is the principal lipid-soluble antioxidant of both plasma and
the LDL particle (Esterbauer et al. (1990)). It occurs at low levels in membranes,
typically one molecule for every 2000-3000 lipid molecules, where it acts as a
chain breaking antioxidant, a singlet oxygen quencher, and a regulator of both
enzyme activity and membrane fluidity (Packer and Landvik (1989)). Tocopherols
can react with alkyl radicals, hydroperoxyl radicals, peroxynitrite, as well as the
hydroxyl free radical. α-Tocopherol can also react with superoxide in vitro and
form a variety of novel compounds (e.g., tocopherol dimers and epoxides)
depending on whether protic or aprotic conditions are used (Ha and Csallany
(1992)). As shown in Figures 3.23 and 4.25, tocopherol can readily intercept a
WWW.ESAINC.COM
383
lipid peroxyl radical and donate a labile proton forming a lipid hydroperoxide and
tocopheroxyl radical, thereby preventing the propagation of the chain reaction
(Chapter 3). Readers interested in the structure-activity relationship of various
natural and synthetic tocopherol species are referred to Lien et al. (1999).
MEMBRANE
LO 2
LO 2H
TOCOPHEROL
TOCOPHEROXYL
RADICAL
CYTOSOL
DIHYDROLIPOIC ACID/α -LIPOIC ACID
ASCORBATE/SEMIDEHYDROASCORBATE RADICAL
GSH/GSSG
ASCORBATE COUPLE/GSH COUPLE
ASCORBATE COUPLE/ α -LIPOIC ACID COUPLE
UBISEMIQUINONE
UBIQUINOL
INDIRECT TOCOPHEROL REGENERATION
DIRECT TOCOPHEROL REGENERATION
LO 2
LO 2H
LO 2
LO 2H
LO 2
LO 2H
LO 2
LO 2H
LO 2
LO 2H
TOCOPHEROL
TOCOPHEROXYL
RADICAL
TOCOPHEROL
TOCOPHEROXYL
RADICAL
TOCOPHEROL
TOCOPHEROXYL
RADICAL
TOCOPHEROL
Lipoamide
Dehydrogenase
DIHYDROLIPOIC ACID
GSSG
GSH
NADPH
Glutathione
Reductase
NADP+
GSH
ASCORBATE
GSSG
SEMIDEHYDROASCORBATE
RADICAL
ASCORBATE
TOCOPHEROL
GSSG
GSH
?
NAD+
SEMIDEHYDROASCORBATE
RADICAL
TOCOPHEROXYL
RADICAL
TOCOPHEROXYL
RADICAL
NADH
α -LIPOIC
ACID
NADPH
Glutathione
Reductase
NADP+
DIHYDROLIPOIC ACID
NADH
Lipoamide
Dehydrogenase
α -LIPOIC
ACID
DIHYDROLIPOIC ACID
NAD+
NADH
Lipoamide
Dehydrogenase
α -LIPOIC
ACID
NAD+
Figure 4.25 The Prevention Of Lipid Peroxidation By Tocopherol And
Possible Direct And Indirect Pathways For The Regeneration Of Tocopherol
By Water Soluble And Fat Soluble Antioxidants. (The Eventual
Regeneration Of Ascorbate, GSH And Lipoic Acid Is Explained In Text.).
Due to delocalization of charge the tocopheroxyl radical is much less energetic
than the lipid peroxyl radical and is therefore less able to react with other species
and cause damage (Chapter 1). Tocopherol is regenerated from the tocopheroxyl
radical following reaction with endogenous reducing agents which can occur
either within the membrane (e.g., ubiquinol) (Kagan et al. (1990); Maguire et al.
(1992); Stoyanovski et al. (1995)) or in the cytoplasm (e.g., GSH, ascorbic acid
WWW.ESAINC.COM
384
and dihydrolipoic acid) (Nohl and Gille (1998)). For this reason, a little bit of
tocopherol goes a long way. This may explain why the short-term absence of
vitamin E in human diet does not result in any specific deficiency disease in
normal adults. However, vitamin E deficiency is problematic for premature babies
who can be predisposed to hemolytic anemia (probably due to the fragility of the
erythrocyte membrane) and in those with diseases that affect fat metabolism,
such as abetalipo-proteinaemia and short-bowel syndrome.
The redox behavior of tocopherol is presented in Figure 4.24 and shows the
redox reactions found in vivo and during electrochemical measurement.
Tocopherol can undergo a single electron oxidation with the production of the
tocopheroxyl radical (the usual “biological” product when tocopherol reacts with
an ROS). A second single electron oxidation of the tocopheroxyl radical, or a two
electron oxidation of tocopherol, results in ring opening and the production of
tocopheryl quinone, the final oxidation product of tocopherol. Tocopheryl quinone
is a potent anticoagulant and may be responsible for some of the effects of
α-tocopherol in preventing heart attacks and strokes (Dowd and Zheng (1995)).
Evidence suggests that tocopheryl quinone can be “recycled” to tocopherol in
vivo, albeit with very poor yields (<0.8%) (Moore and Ingold (1997)). The
physiological impact of this salvage pathway has yet to be determined.
Tocopheryl quinone can be reduced in a two electron process to form tocopheryl
hydroquinone. Kohar et al. (1995) showed that tocopheryl hydroquinone can be
formed in vivo and have suggested that the tocopheryl hydroquinone/tocopheryl
quinone couple may be biologically important acting like the coenzyme Q redox
system. Recently Siegel et al. (1997) demonstrated that tocopheryl form is more
readily reduced by NAD(P)H:quinone oxidoreductase (NQO1) than ubiquinone.
This suggests that one of the physiological functions of NQO1 is the regeneration
of antioxidant forms (tocopheryl hydroquinone) of α-tocopherol.
A variety of different products can be formed when tocopherols react with RNS.
d’Ischia and Novellino (1996) examined the mechanism of reaction between nitric
oxide and α-tocopherol. Under aprotic conditions one oxidation pathway
produced some α-tocopheryl quinone and little amounts of its nitrite ester.
However, under physiological conditions they reported considerable oxidation of
α-tocopherol to α-tocopheryl quinone, formation of its nitrite ester (which is
capable of nitrosating amines), a yellow dimer and a series of related oligomers.
Similarly, nitric oxide (de Groot et al. (1993)), nitrogen dioxide (Cooney et al.
(1995)) and peroxynitrite (Hogg et al. (1993), Pannala et al. (1999); Vatassery
(1996)) have all been reported to rapidly oxidize α-tocopherol. Such depletion in
α-tocopherol levels have been suggested to underlie the cytotoxic nature of
many RNS. In support of this, Burkart et al. (1995) showed that α-tocopherol
efficiently protects eukaryotic cells from nitric oxide-induced cytotoxicity.
Gorbunov et al. (1996) found that nitric oxide can also lead to the production of
the tocopheroxyl radical and suggested that another possible cytotoxic
mechanism is the depletion of ascorbic acid resulting from α-tocopherol
regeneration.
WWW.ESAINC.COM
385
Recent research has focused on tocopherol nitration. While α-tocopherol can
react with nitrogen dioxide to form a nitrosating agent, γ-tocopherol17 reduces
nitrogen dioxide to nitric oxide without the formation of a nitrosating species
(Cooney et al. (1993)). This γ-tocopherol-mediated detoxification of nitrogen
dioxide may be important physiologically as it has been shown to prevent
nitrogen dioxide-mediated DNA strand breaks (Bittrich et al. (1993)).
γ-Tocopherol also reacts with nitrating species to form a variety of oxidation
products and adducts (Hoglen et al. (1997); Pannala et al.(1999)). For example,
based on the earlier work of Cooney et al. (1995), Christen et al. (1997) used
HPLC-ECD to show that γ-tocopherol preferentially reacts with nitrating agents
(e.g., peroxynitrite or SIN-1) forming 5-nitro-γ-tocopherol (Figure 4.23). 5-Nitro-γtocopherol is currently being used as a marker of RNS damage in vivo in a
variety of biological fluids (Hensley et al., 1999, 2000); Morton et al., (2002);
Williamson et al., (2002)).
As expected δ-tocopherol is even more reactive than γ-tocopherol, due to it
having two available positions for substitution to take place (Figure 4.23). αTocopherol cannot react in this way as its 5-position is blocked and thus
unavailable for reaction with electrophiles. This suggests that γ-tocopherol’s
principal role may be to trap membrane-soluble electrophilic nitrogen oxides
whereas α-tocopherol’s action is as a chain breaking antioxidant (Christen et al.
(1997); Wolf (1997)). Furthermore, as large doses of α-tocopherol (the form in
supplements) readily displace γ-tocopherol (the form in human diet) from plasma
and other tissues, Christen et al. (1997) have questioned the wisdom of only
having α-tocopherol in supplements. However, Goss et al. (1999) have
challenged the requirement for γ-tocopherol and reported that α-tocopherol could
prevent nitration of both γ-tocopherol and tyrosine by peroxynitrite. Their data
suggest that α-tocopherol alone is readily capable of removing any peroxynitritederived RNS and that γ-tocopherol probably only plays a role once α-tocopherol
levels are depleted.
Under certain circumstances α-tocopherol has been reported to have pro-oxidant
activity (Bowery et al. (1992). For example, in LDL in the absence of regenerating
antioxidants (e.g., ascorbic acid), the α-tocopheroxyl radical can abstract a
hydrogen atom from a nearby PUFA molecule to form a conjugated diene that
can then form a lipid peroxyl radical upon reaction with oxygen. This peroxyl
radical can then react with α-tocopherol to complete the cycle of α-tocopherolmediated lipid peroxidation. As it is unlikely that regenerating antioxidants will be
absent, the physiological significance of tocopherol’s pro-oxidant activity is yet to
be fully explored.
17
γ-Tocopherol is also capable of undergoing Michael addition with cellular thiols (such as GSH) forming 5-substituted
adducts which are potential chemotherapeutic agents (Thornton et al. (1995)).
WWW.ESAINC.COM
386
Tocopherol and Disease.
Several studies have addressed the importance of the tocopherols’ antioxidant
properties in disease prevention. For example, tocopherol appears to play a vital
role in preventing coronary diseases: there is a significant correlation between
α-tocopherol concentrations and coronary artery disease (Gey (1998); Gey et al.
(1992)); there is an inverse relationship between plasma vitamin E concentration
and the risk of angina pectoris (Riemersma et al. (1991)); and vitamin E
supplementation reduces the risk of coronary artery (Rimm et al. (1992);
Stampfer et al. (1993)) and atherosclerotic heart disease (Diaz et al. (1997)).
α-Tocopherol inhibits LDL oxidation both in vitro and in human subjects (Jialal
and Grundy (1992); Jialal et al. (1995)). Additionally, α-tocopherol may offer
additional benefits for cardiovascular disease beyond its antioxidant properties.
For example, it reduces platelet adhesion (Salonen et al. (1991)); inhibits smooth
muscle proliferation and protein kinase C activity (Ozer et al. (1993)); inhibits
agonist-induced monocyte adhesion to cultured human endothelial cells (Faruqui
et al. (1994)); and preserves endothelium-dependent vasodilation in
hypercholesterolemic rabbits (Keaney et al. (1994)). In addition to cardiovascular
problems, humans with suboptimal vitamin E intake are thought to be at
increased risk for several aging related diseases such as cancer (Bostick et al.
(1993); Gridley et al. (1992); Wald et al. (1984)), and neurological disorders (Behl
et al. (1992); (Diplock (1998); Kayden (1993); Sokol (1989); Tanyel and Mancano
(1997); Vatassery (1998)).
The potential benefits of α-tocopherol supplementation on cancer incidence is far
from clear. Although previous studies have suggested that higher intakes of
α-tocopherol may be associated with a reduced risk of lung cancer, Heinonen
and Albanes (1994) found no such correlation among male smokers. Recently,
McAll and Frei (1999) posed the question “can antioxidant vitamins materially
reduce oxidative damage in humans”. They concluded that “with the only
exception of supplemental vitamin E, and possibly vitamin C, being able to lower
lipid oxidative damage in both smokers and non-smokers, the current evidence is
insufficient to conclude that antioxidant vitamin supplementation materially
reduces oxidative damage in humans”.
Measurement of Tocopherols and Their Metabolites.
A variety of methods have been used to measure the tocopherols including
colorimetric (Diplock et al. (1996)), spectrophotometric (Vatassery and
Mortenson (1972)), fluorometric (Taylor et al. (1976); Vatassery and Mortenson
(1972)), HPLC with UV (Barua et al. (1991; 1993); Miller and Yang (1985); Xu et
al. (1996)), GC-MS (Liebler et al. (1996)), and tandem MS approaches (Walton et
al. (1988)). Several HPLC-ECD methods exist for the measurement of the
tocopherols, either alone or with other lipophilic antioxidants (Table 4.9).
WWW.ESAINC.COM
387
Analyte
Tocopherol
Tocopherols, tocotrienols,
ubiquinols, ubiquinones
Tocopherols/tocopheryl
quinones
Tocopherols/tocopheryl
quinones
Tocopherol, tocopheryl
quinones
Tocopherol, ubiquinol,
ubiquinone
Tocopherol, ubiquinol,
ubiquinone
Tocopherols
α-Tocopherol, β-carotene and
retinol
α-Tocopherol, CoQ10
Tocopherols, β-carotene,
lycopene, ubiquinol,
ubiquinone
Tocopherols, carotenoids,
ubiquinol, ubiquinone
Tocopherols, CoQ10,
retinoids, carotenoids, vitamin
K
Tocopherol, CoQ10 and
tocopherol oxidation products
Method
Reference
HPLC-dual coulometric
electrode detection.
HPLC-amperometric electrode
detection for tocopherols,
tocotrienols, ubiquinols; HPLCUV detection for ubiquinones.
Castle and Cook
(1985)
Podda et al. (1996)
HPLC-triple coulometric
electrode detection.
HPLC-dual coulometric
electrode detection.
HPLC-triple coulometric
electrode detection.
HPLC-amperometric electrode
detection for tocopherol only.
HPLC-amperometric electrode
detection with platinum catalyst
reduction.
HPLC-amperometric electrode
detection.
HPLC-UV and amperometric
electrode detection.
HPLC-dual coulometric
electrode detection.
HPLC-dual coulometric
electrode detection.
Murphy and Kehrer
(1987)
Vatassery et al.
(1993)
Takeda et al. (1996)
Lang and Packer
(1987)
Wakabayashi et al.
(1994)
Huang et al. (1986)
MacCrehan and
Schonberger (1987)
Edlund (1988)
Motchnik et al.
(1994)
HPLC-dual coulometric
electrode detection.
Gradient HPLC-coulometric
electrode array detection.
Finckh et al. (1995,
1999)
Gamache et al.
(1997, 2003)
HPLC- dual coulometric
electrode detection.
Leray et al. (1998)
Table 4.9 HPLC-ECD Methods Used To Measure Tocopher