IN THE 1880s Elie Metchnikoff observed specialized phagocytic cells ingesting bacteria, and recognized the importance of phagocytosis as a defense mechanism in multicellular organisms.1 Neutrophils are one of the professional phagocytes in humans. They ingest bacteria into intracellular compartments called phagosomes, where they direct an arsenal of cytotoxic agents. Metchnikoff noted that “what substances within the phagocyte harm and destroy the microbes is quite undecided.” One hundred years on, Mims stated that “we are still profoundly ignorant of the ways in which polymorphs attempt to kill and then to digest the great variety of microorganisms that are ingested.”2 Our understanding is gradually increasing, but there are still a number of questions to be answered.

It was recognized at an early stage that cytoplasmic granules containing digestive and antibacterial compounds are emptied into the phagosome.3 Later, it was discovered that phagocytosing neutrophils undergo a burst of oxygen consumption4,5 that is caused by an NADPH oxidase complex that assembles at the phagosomal membrane. As reviewed by others,6-8 electrons are transferred from cytoplasmic NADPH to oxygen on the phagosomal side of the membrane, generating first superoxide plus a range of other reactive oxygen species. This oxidative burst is essential for killing of a number of microorganisms, as shown by the susceptibility to infections of individuals with chronic granulomatous disease (CGD), a genetic disease in which the NADPH oxidase is inactive.9-11 

Much is known about the reactive oxygen species released into the extracellular surroundings when neutrophils respond to soluble stimuli. However, the enzymatic and chemical reactions involved in oxidant production are dependent on environmental conditions, which may vary markedly between the phagosome and the extracellular medium. Knowledge of the biochemistry within the phagosome is limited by its inaccessibility to standard detectors and scavengers. Consequently, the oxidant species directly responsible for killing bacteria are still open to speculation. This review focuses on what is known about the chemical composition of the phagosome, the nature and amount of the oxidants generated inside, and on recent information that helps clarify the importance of myeloperoxidase-derived oxidants in killing.

Superoxide and hydrogen peroxide.

A variety of soluble and particulate stimuli induce extracellular superoxide production.5,12-14 Most of the oxygen consumed can be accounted for as hydrogen peroxide,15,16 which is formed from dismutation of the superoxide radical.7However, hydrogen peroxide is bactericidal only at high concentrations,17,18 and exogenously generated superoxide does not kill bacteria directly.19-21 Therefore, a variety of secondary oxidants have been proposed to account for the destructive capacity of neutrophils (Fig 1). Table 1 provides a summary of their properties.

Fig. 1.

Possible oxidant generating reactions with stimulated neutrophils. NOS, nitric oxide synthase; MPO, myeloperoxidase.

Fig. 1.

Possible oxidant generating reactions with stimulated neutrophils. NOS, nitric oxide synthase; MPO, myeloperoxidase.

Close modal
Table 1.

Properties of Reactive Oxygen Species

Superoxide:  Mild oxidant and reductant with limited biological activity; reduces some iron complexes to enable hydroxyl radical production by the Fenton reaction; inactivates iron/sulfur proteins and releases iron; limited membrane permeability. 
Hydrogen peroxide:  Oxidizing agent; reacts slowly with reducing agents such as thiols; reacts with reduced iron and copper salts to generate hydroxyl radicals; reacts with heme proteins and peroxidases to initiate radical reactions and lipid peroxidation; membrane permeable.  
Hydroxyl radical: Extremely reactive with most biological molecules; causes DNA modification and strand breaks, enzyme inactivation, lipid peroxidation; very short range of action; generates secondary radicals, eg, from bicarbonate, chloride.  
Singlet oxygen: Electronically excited state of oxygen; reacts with a number of biological molecules, including membrane lipids to initiate peroxidation.  
Hypochlorous acid:  Strong nonradical oxidant of a wide range of biological compounds, but more selective than hydroxyl radical; preferred substrates thiols and thioethers; converts amines to chloramines; chlorinates phenols and unsaturated bonds; oxidizes iron centers; crosslinks proteins; membrane permeable; in equilibrium with chlorine gas at low pH and hypochlorite at high pH.  
Chloramines:  Milder and longer lived oxidants than HOCl; react with thiols, thioethers, iron centers; variable toxicity dependent on polarity and membrane permeability; chloramines of α-amino acids break down slowly to potentially toxic aldehydes.  
Nitric oxide:  Reacts very rapidly with superoxide to give peroxynitrite; reaction with oxygen favored at high oxygen tension; forms complexes with heme proteins; inactivates iron/sulfur centers; forms nitrosothiols.  
Peroxynitrite: Unstable short lived strong oxidant with properties similar to hydroxyl radical; hydroxylates and nitrates aromatic compounds; reacts rapidly with thiols: breaks down to nitrate; interacts with bicarbonate to alter reactivity. 
Superoxide:  Mild oxidant and reductant with limited biological activity; reduces some iron complexes to enable hydroxyl radical production by the Fenton reaction; inactivates iron/sulfur proteins and releases iron; limited membrane permeability. 
Hydrogen peroxide:  Oxidizing agent; reacts slowly with reducing agents such as thiols; reacts with reduced iron and copper salts to generate hydroxyl radicals; reacts with heme proteins and peroxidases to initiate radical reactions and lipid peroxidation; membrane permeable.  
Hydroxyl radical: Extremely reactive with most biological molecules; causes DNA modification and strand breaks, enzyme inactivation, lipid peroxidation; very short range of action; generates secondary radicals, eg, from bicarbonate, chloride.  
Singlet oxygen: Electronically excited state of oxygen; reacts with a number of biological molecules, including membrane lipids to initiate peroxidation.  
Hypochlorous acid:  Strong nonradical oxidant of a wide range of biological compounds, but more selective than hydroxyl radical; preferred substrates thiols and thioethers; converts amines to chloramines; chlorinates phenols and unsaturated bonds; oxidizes iron centers; crosslinks proteins; membrane permeable; in equilibrium with chlorine gas at low pH and hypochlorite at high pH.  
Chloramines:  Milder and longer lived oxidants than HOCl; react with thiols, thioethers, iron centers; variable toxicity dependent on polarity and membrane permeability; chloramines of α-amino acids break down slowly to potentially toxic aldehydes.  
Nitric oxide:  Reacts very rapidly with superoxide to give peroxynitrite; reaction with oxygen favored at high oxygen tension; forms complexes with heme proteins; inactivates iron/sulfur centers; forms nitrosothiols.  
Peroxynitrite: Unstable short lived strong oxidant with properties similar to hydroxyl radical; hydroxylates and nitrates aromatic compounds; reacts rapidly with thiols: breaks down to nitrate; interacts with bicarbonate to alter reactivity. 
Hydroxyl radicals and singlet oxygen.

Whether the hydroxyl radical is a major component of the neutrophil bactericidal arsenal has been controversial.22-26 There have been a large number of studies of isolated neutrophils, some of which have presented evidence for hydroxyl radical production.27-30 However, assays for this extremely reactive species rely on measuring secondary products and the use of inhibitors. They often lack specificity and reactions attributed to the hydroxyl radical may be caused by other oxidants such as superoxide or hypochlorous acid (HOCl).23,31 

There are two potential mechanisms for hydroxyl radical production by neutrophils: the superoxide-driven Fenton reaction between hydrogen peroxide and an appropriate transition metal catalyst, and the reaction of HOCl with superoxide. The most definitive investigations of the Fenton mechanism have used spin traps to establish that neutrophils do not have an endogenous transition metal catalyst and that release of lactoferrin inhibits the reaction by complexing iron.25,32Myeloperoxidase limits the reaction further, even if iron is available, by consuming hydrogen peroxide.33 The overall conclusion is that the cells generate insignificant amounts of hydroxyl radical by this mechanism.23-25 This reaction may be more significant in vivo if target cells or molecules could provide iron to the neutrophils. Although most biological forms of iron are not catalytically active, neutrophils have been shown to produce hydroxyl radicals in the presence of proteolytically degraded transferrin25,34-36 or iron complexed to thePseudomonas aeruginosa siderophore pyochelin.37,38However, intracellular iron is not necessarily available and no enhanced hydroxyl radical production was observed when neutrophils ingested Staphylococcus aureus that had been preloaded with iron.35 

Recently, more sensitive spin-trapping methods have detected myeloperoxidase-dependent hydroxyl radical formation by isolated neutrophils,25,39 presumably from HOCl and superoxide.40 Very little of the oxygen consumed by the cells has been measured as hydroxyl radicals, and whether this is sufficient to play a role in cytotoxicity is yet to be proven.

Hydroxyl radicals, including those generated by ionizing radiation, kill bacteria.41,42 However, they are not as efficient as their high reactivity might suggest.41 They have a limited radius of action, so even in the confined space of the phagosome, most are likely to react with other targets before reaching the bacterium. It has been proposed that secondary products from bicarbonate or chloride might be responsible for any biological activity.41 Czapski et al43 have observed that hydroxyl radical generating systems are more toxic to bacteria in the presence of chloride, and attributed this to a reaction between the two to produce HOCl. This would suggest that any hydroxyl radical generation from HOCl and superoxide would have little additional impact on the killing process, and may actually reduce toxicity by converting the extremely bactericidal HOCl to the more reactive, but less toxic, hydroxyl radical.

Singlet oxygen could theoretically be produced by neutrophils from the reaction of hydrogen peroxide with HOCl. Although it was initially proposed to be the source of the chemiluminescence of stimulated cells,44 subsequent studies measuring specific infrared chemiluminescence have failed to detect singlet oxygen production by neutrophils.45-47 Positive results were obtained with eosinophils, which generate hypobromous acid rather than HOCl, although the conversion of oxygen consumed was only 0.4%.48Steinbeck et al47 have used a singlet oxygen trap with neutrophils, and reported a surprisingly high 19% conversion of available oxygen to the singlet form. The significance of this finding to microbicidal activity and how it can be reconciled with the chemical findings require further investigation.

Myeloperoxidase and HOCl.

Most of the hydrogen peroxide generated by neutrophils is consumed by myeloperoxidase.12,49 Myeloperoxidase is a major constituent of the azurophilic cytoplasmic granules50 and a classical heme peroxidase that uses hydrogen peroxide to oxidize a variety of aromatic compounds (RH) by a 1-electron mechanism to give substrate radicals (R)51-54(Fig 2). It is unique, however, in readily oxidizing chloride ions to the strong nonradical oxidant, HOCl.55 HOCl is the most bactericidal oxidant known to be produced by the neutrophil.5,56 Many species of bacteria are killed readily by a myeloperoxidase/hydrogen peroxide/chloride system.57 Bacterial targets include iron-sulfur proteins, membrane transport proteins,58 adenosine triphosphate (ATP)-generating systems,59 and the origin of replication site for DNA synthesis, which appears to be the most sensitive.60-62 Chloramines are generated indirectly through the reaction of HOCl with amines,63 and these are also bactericidal.64,65 Cell permeable chloramines, eg, monochloramine, can enhance the toxicity of HOCl, whereas protein chloramines have low toxicity. Other substrates of myeloperoxidase include iodide, bromide, thiocyanate, and nitrite.66-69 The corresponding hypohalous acids or nitrogen oxides that are produced vary in their bactericidal efficiency. Myeloperoxidase can also generate peroxides and hydroxylated derivatives of phenolics such as salicylate in superoxide-dependent reactions.31,70 

Fig. 2.

Reactions of myeloperoxidase. Ferric myeloperoxidase (MP3+) reacts with hydrogen peroxide to form the redox intermediate compound I, which oxidizes chloride or thiocyanate by a single 2-electron transfer to produce the respective hypohalous acids. Myeloperoxidase also oxidizes numerous organic substrates (RH) by two successive 1-electron transfers involving the enzyme intermediates compound I and compound II. Poor peroxidase substrates trap the enzyme as compound II and hypohalous acid production is inhibited unless superoxide is present to recycle the native enzyme. Superoxide can convert myeloperoxidase to compound III, which is turned over by a second reaction with superoxide. It has yet to be established whether the products of the latter reaction are compound I or MP3+ and hydrogen peroxide. Either way, the same net result is achieved.

Fig. 2.

Reactions of myeloperoxidase. Ferric myeloperoxidase (MP3+) reacts with hydrogen peroxide to form the redox intermediate compound I, which oxidizes chloride or thiocyanate by a single 2-electron transfer to produce the respective hypohalous acids. Myeloperoxidase also oxidizes numerous organic substrates (RH) by two successive 1-electron transfers involving the enzyme intermediates compound I and compound II. Poor peroxidase substrates trap the enzyme as compound II and hypohalous acid production is inhibited unless superoxide is present to recycle the native enzyme. Superoxide can convert myeloperoxidase to compound III, which is turned over by a second reaction with superoxide. It has yet to be established whether the products of the latter reaction are compound I or MP3+ and hydrogen peroxide. Either way, the same net result is achieved.

Close modal

Because myeloperoxidase has the specialized ability to oxidize chloride, it is generally considered that its function is to generate HOCl. In in vitro systems with taurine or methionine added as a trap, from 28% to 70% of the hydrogen peroxide produced by neutrophils has been detected as HOCl.71,72 However, most experimental studies are performed in media without alternative myeloperoxidase substrates. The products formed in pathophysiological situations may be more varied.

Reactive nitrogen species.

There is considerable interest in nitric oxide and peroxynitrite as potential cytotoxic agents produced by inflammatory cells.73-77 It is well documented that murine macrophages generate nitric oxide in response to cytokines,78 but results have been contradictory and mostly negative for human neutrophils isolated from peripheral blood.79-84 The prevailing view is that reactive nitrogen species are important in human inflammation, and that in vitro studies have been negative because the conditions necessary for induction have not been elucidated. Nitric oxide synthase message has recently been detected in neutrophils isolated from urine passed during infection of the urinary tract,85 and in buffy coat neutrophils after exposure to inflammatory cytokines.86 Also, because both myeloperoxidase and HOCl can oxidize nitrite,69;87neutrophils may not need their own source of nitric oxide to generate reactive nitrogen oxides. These findings suggest that nitric oxide may be a significant player in the oxidative reactions of the neutrophil in vivo, but until human neutrophils can be induced experimentally to produce nitric oxide, the relevance of it, and its reaction with superoxide to produce peroxynitrite, cannot be assessed.

The neutrophil makes tight contact with its target and the plasma membrane flows around the surface until the bacterium is completely enclosed.88 This minimizes the amount of extracellular fluid entering the phagosome with the bacterium, and means that the phagosome is initially a very small space (Fig 3). The exclusion of external medium sets up a new environment that will have an important influence on the biochemistry of oxidant production and bacterial killing. The major contributors to the chemical composition of the phagosome are the contents of the cytoplasmic granules that empty into it. Granule contents are released within seconds of ingestion and constitute a significant proportion of the phagosomal volume.3,89 There are at least four different classes of granules,90 and sequential release of the different types90,91 may provide a succession of different phagosomal environments.

Fig. 3.

Transmission electron micrograph of S aureusinside the phagosome of a human neutrophil. Arrows pointed to examples of S aureus within phagosomes (original magnification × 15,000). (Courtesy of W.A. Day, Department of Pathology, Christchurch School of Medicine.)

Fig. 3.

Transmission electron micrograph of S aureusinside the phagosome of a human neutrophil. Arrows pointed to examples of S aureus within phagosomes (original magnification × 15,000). (Courtesy of W.A. Day, Department of Pathology, Christchurch School of Medicine.)

Close modal

The large amount of degranulation into a small volume means that the initial protein concentration will be high (estimated 30% to 40% protein). This will decrease with time as the volume increases due to the osmotic influx of water associated with granule emptying and digestion of the bacterium. Ionic composition is unknown, and will depend on what is in the granules and also the activity of membrane pumps and channels that connect the phagosome to the neutrophil cytoplasm. The outward pumping of cytoplasmic chloride ions by stimulated neutrophils92 may be important for maintaining sufficient phagosomal chloride concentrations for HOCl production. Chloride is also necessary for azurophil degranulation,93and this may be a means of limiting myeloperoxidase release when chloride is depleted.

Phagosomal pH is under tight control. The oxidation of cytoplasmic NADPH to NADP+ and H+, and the transfer of reducing equivalents across the membrane to phagosomal oxygen, results in acidification of the cytoplasm.94 The dismutation of the superoxide anion, with its associated consumption of protons, would make the phagosome considerably alkaline. There is a transient increase in pH to 7.8 to 8.0 in the first few minutes after phagosome formation.95,96 However, activation of the oxidase is accompanied by activation of an Na+/H+antiport, an H+-ATPase, and an H+ conductance mechanism97 so that proton pumping from the cytoplasm into the phagosome restricts this increase and the pH decreases to approximately 6.0 after an hour.95,96 

Taking into account the physical and chemical characteristics discussed above, what is known about the oxidants produced and the ability of myeloperoxidase to function in the phagosome? During phagocytosis, neutrophils consume a similar amount of oxygen as with strong soluble stimuli, yet release only small amounts of superoxide or hydrogen peroxide in the surroundings.14,98,99 However, there is convincing cytochemical evidence that superoxide100,101 and hydrogen peroxide13,102,103 are generated intraphagosomally and around ingested bacteria. In the presence of heme enzyme inhibitors, hydrogen peroxide detected in the medium can account for most of the oxygen consumed.104,105 

On the assumption that ingestion of 15 to 20 bacteria gives maximal oxygen consumption, we have calculated that superoxide should be formed in the phagosomal space at the extraordinarily high rate of 5 to 10 mmol/L per second.106 Based on granule numbers, the myeloperoxidase released should reach a concentration of 1 to 2 mmol/L. Generation of large amounts of HOCl would be expected. However, the enzymology of myeloperoxidase is complex (Fig 2)49 and the efficiency of HOCl production is strongly dependent on conditions. Activity is decreased at high pH and at high hydrogen peroxide and chloride concentrations.107,108 Numerous physiological and pharmacological compounds that act as poor peroxidase substrates and reversibly inactivate the enzyme also inhibit HOCl production.109,110 It is likely that these substrates could modulate HOCl production in vivo. Superoxide reacts with myeloperoxidase107 to form a complex (Compound III) that lies outside the normal catalytic cycle. Superoxide can also reactivate myeloperoxidase that has become reversibly inhibited through compound II formation.108 

We have developed a kinetic model of the phagosome, incorporating the known reactions of myeloperoxidase, hydrogen peroxide and superoxide, and their rate constants, to address how myeloperoxidase acts in the phagosomal environment (manuscript in preparation). Predictions from the model are consistent with direct spectral observation107 that superoxide initially reacts with the myeloperoxidase to convert it to compound III. To see significant peroxidase activity or HOCl generation, the compound III must turn over. Although this has been proposed to occur via reaction with hydrogen peroxide,108 this mechanism is much too slow to give any significant HOCl production. For myeloperoxidase to continue to function after the first few seconds, a reaction between compound III and superoxide must be invoked. Such a reaction has been proposed,111 and studies with purified myeloperoxidase provide further evidence for it.31 Myeloperoxidase can then handle the high rates of formation of superoxide and hydrogen peroxide such that neither builds up beyond micromolar concentrations, and the majority of the oxygen consumed is converted to HOCl. This system appears to be reasonably robust, with realistic variations in superoxide flux, myeloperoxidase release, phagosomal volume, and hydrogen peroxide scavenging by the cytoplasm making little difference to the efficiency of HOCl formation.

Until recently, evidence that HOCl is formed in the phagosome has been indirectly based on the incorporation of 36Cl or radiolabeled iodide into organic material during the ingestion of bacteria.112-115 More definitive evidence has come from recent measurements of chlorotyrosine and chlorinated fluorescein as specific markers of HOCl production. Hazen et al116 trapped tyrosine within red blood cell ghosts and showed that it became chlorinated when the ghosts were phagocytosed. In a related study, we have recovered ingested bacteria from neutrophil phagosomes and shown that protein hydrolysates contain chlorotyrosine that was not present in the isolated neutrophils or bacteria.117 Hurst et al have recently followed up earlier studies of bleaching of fluorescein attached to ingested latex beads118 to show that this is caused by chlorination.119 They calculated that at least 12% of the oxygen consumed was converted to HOCl within the phagosome.

The kinetic modeling has enabled assessment of why it might be advantageous for the neutrophil to produce superoxide rather than hydrogen peroxide directly. If superoxide is removed from the system, we find that the HOCl production becomes sensitive to fluctuations in oxidant flux or the amount of myeloperoxidase released into the phagosome. Under some conditions HOCl production is enhanced but without superoxide to regenerate the native enzyme from compound II, myeloperoxidase becomes prone to inhibition by electron donors that readily reduce compound I but not compound II. We speculate that substrates such as tryptophan and nitrite could be present in the phagosome and impair HOCl production by this mechanism. So for the neutrophil to maintain its myeloperoxidase activity without stringent environmental requirements, there would be a clear advantage in generating superoxide.

Experiments have not been performed with appropriate substrates to establish whether myeloperoxidase-derived oxidants other than HOCl are produced intraphagosomally. However, studies using an antibody against nitrotyrosine suggest that a nitrating agent can be formed when bacteria are ingested by cytokine-treated buffy coat neutrophils.86 

Oxidative and nonoxidative mechanisms.

Efficient control of a multitude of microorganisms is so important for host survival that the neutrophil does not rely on a single antimicrobial weapon. This review concentrates on oxidative mechanisms, but as discussed elsewhere,120-122 this is complemented by nonoxidative killing by granule proteins that are released into the phagosome. The mechanism that predominates may vary depending on the microbial species, its metabolic state, and the prevailing conditions.61 

Optimal killing of many species of bacteria requires products from the oxidative burst. This is best exemplified in CGD, where affected individuals have an impaired or completely absent oxidative burst and suffer from recurrent and life-threatening infections.9,10The strains of bacteria that are killed poorly in vitro are responsible for the infections that are characteristic of CGD.10 Normal neutrophils tested in anaerobic environments, or in the presence of the NADPH oxidase inhibitor diphenyleneiodonium, are also impaired in their ability to kill these bacteria.123-126 Other species are killed normally, either because they are catalase-negative and able to supply an alternative source of hydrogen peroxide,127,128or because they can be disposed of effectively by nonoxidative mechanisms.

Myeloperoxidase and HOCl.

Myeloperoxidase appears critical for oxidative killing in experimental systems. Neutrophils isolated from the blood of myeloperoxidase-deficient individuals kill a variety of microorganisms poorly,129-131 and inhibitors of myeloperoxidase such as azide, cyanide, and salicylhydroxamic acid impair killing by normal cells.106,130,132,133 Neutrophil cytoplasts that lack granule enzymes but generate hydrogen peroxide only kill bacteria if they are coated with myeloperoxidase before ingestion.134 

Measurements of rates of killing of S aureus by neutrophils isolated from human blood reinforce the importance of myeloperoxidase.106,126 Inhibition of the oxidative burst with diphenyleneiodonium, or removal of oxygen, decreases the rate constant for killing by 80%, enabling separation of the oxidative and nonoxidative components (Fig 4). Killing rates are substantially decreased in the presence of the myeloperoxidase inhibitors azide and 4-aminobenzoic acid hydrazide, and with myeloperoxidase-deficient neutrophils. Only the oxidative component is affected, and is six times slower when myeloperoxidase is not active. These results indicate that, at least with S aureus, the normal mechanism for oxidative killing uses myeloperoxidase. Direct killing by hydrogen peroxide, or other alternative oxidative mechanisms, are poor substitutes.

Fig. 4.

Rate constants for killing of S aureus by human neutrophils. Opsonized bacteria were mixed with neutrophils in a 1:1 ratio. Numbers of extracellular and viable intracellular bacteria were measured at 0, 10, 20, and 30 minutes, and from these independent first-order rate constants for phagocytosis and killing were measured. Superoxide dismutase was conjugated to IgG (IgG-SOD) and attached to the bacteria through binding to the protein A on their surface. ABAH, the myeloperoxidase inhibitor 4-aminobenzoic acid hydrazide. The shaded area represents the contribution of nonoxidative killing measured in the presence of diphenyleneiodonium (DPI) or anaerobically (N2). The data are taken from Hampton,117 and show the mean and SD of at least three experiments.

Fig. 4.

Rate constants for killing of S aureus by human neutrophils. Opsonized bacteria were mixed with neutrophils in a 1:1 ratio. Numbers of extracellular and viable intracellular bacteria were measured at 0, 10, 20, and 30 minutes, and from these independent first-order rate constants for phagocytosis and killing were measured. Superoxide dismutase was conjugated to IgG (IgG-SOD) and attached to the bacteria through binding to the protein A on their surface. ABAH, the myeloperoxidase inhibitor 4-aminobenzoic acid hydrazide. The shaded area represents the contribution of nonoxidative killing measured in the presence of diphenyleneiodonium (DPI) or anaerobically (N2). The data are taken from Hampton,117 and show the mean and SD of at least three experiments.

Close modal

Although HOCl stands out as the prime candidate for the lethal agent produced by myeloperoxidase, there is currently insufficient evidence to exclude other products of the enzyme. We recently observed that the fraction of tyrosyl residues converted to chlorotyrosine in phagocytosed S aureus (0.5% ± 0.2%, SEM of 10 experiments) was similar to that for S aureus treated with a lethal amount of HOCl (Fig 5). This suggests that enough HOCl is generated in the phagosome for it to be responsible for killing. A similar conclusion was reached by Jiang et al119 measuring fluorescein chlorination. Inhibition of killing of Candida pseudohyphae by scavengers of HOCl and chloramines also supports the involvement of chlorinated oxidants.135 However, more direct evidence is necessary to confirm this role for HOCl.

Fig. 5.

Chlorotyrosine formation and loss of viability for S aureus exposed to reagent HOCl. Bacteria (1 × 108/mL) were treated with a range of concentrations of HOCl and then analyzed for tyrosine and chlorotyrosine content,165 and the number of remaining viable colony-forming units. The results are taken from Hampton.117 The means and SD of three experiments are reported.

Fig. 5.

Chlorotyrosine formation and loss of viability for S aureus exposed to reagent HOCl. Bacteria (1 × 108/mL) were treated with a range of concentrations of HOCl and then analyzed for tyrosine and chlorotyrosine content,165 and the number of remaining viable colony-forming units. The results are taken from Hampton.117 The means and SD of three experiments are reported.

Close modal
Role of superoxide.

Neutrophils must generate superoxide to kill oxidatively. Its role could simply be as a precursor of hydrogen peroxide, or it could participate directly in the killing process. Distinguishing between these possibilities experimentally is complicated by the difficulty of getting sufficient superoxide dismutase (SOD) into the phagosome to scavenge all the superoxide generated. Adding SOD to phagocytosing neutrophils136 or modifying the expression of SOD in target bacteria137-142 has generally had little effect, but this could be because the SOD did not gain access to the phagosome. The few studies where this has been achieved indicate a direct role for superoxide in killing. Johnston et al136 showed that the killing of S aureus was impeded when SOD-coated latex beads were co-ingested with the bacteria. The accessibility problem has also been overcome by attaching SOD to the surface of S aureus.106 The SOD was covalently crosslinked to IgG that then bound to protein A in the cell wall. The bacteria were ingested normally, but the rate constant for killing was decreased by 30% (Fig 4). This represents a decrease in rate of oxidative killing to almost a half. SOD had no effect in the presence of peroxidase inhibitors, which suggests that it acts on a myeloperoxidase-dependent process.

The effect of SOD could be explained on the basis of its inhibiting hydroxyl radical production.136 If the route to hydroxyl radicals was via superoxide and HOCl, this could also explain the apparent involvement of a myeloperoxidase-dependent process. However, as argued above, the hydroxyl radical is unlikely to be a major player in the phagosome. An alternative explanation, which is consistent with the modeling studies of oxidant production, is that superoxide prevents reversible inactivation of myeloperoxidase, thereby optimizing killing by HOCl. More direct analyses are needed before firm conclusions can be drawn on the mechanism.

In the context of superoxide having a direct role in killing, it is of interest that Mycobacterium tuberculosis,143,Nocardia asteroides,144,Helicobacter pylori,145 and Actinobacillus pleuropneumoniae146 all secrete SOD. Antibodies to the superoxide dismutase of N asteroides enhanced both bacterial killing by neutrophils147 and clearance upon inoculation of mice.148 It is possible that this surface-associated superoxide dismutase could slow down intraphagosomal killing and be a factor in their pathogenicity.

Although myeloperoxidase deficiency affects at least 1 in 4,000 people, these people are not unduly prone to infections.10 Only occasional increased susceptibility to Candida infection has been noted, and doubts have even been raised about whether myeloperoxidase has a role in bacterial killing.6,149 This contrasts dramatically with CGD, where the NADPH oxidase is absent. In CGD, common pathogens including S aureus cause life-threatening problems. Yet in vitro tests show markedly impaired oxidative killing for both types of neutrophil. On this basis it would be reasonable to expect individuals with CGD and myeloperoxidase deficiency to be similarly compromised in their ability to handle certain microorganisms. The key question is: what compensates for the defect in oxidative killing and prevents infections in myeloperoxidase deficiency?

The usual explanation is that an alternative oxidative killing mechanism operates as a backup. Myeloperoxidase-deficient neutrophils do consume more oxygen than normal130,150 and show extended superoxide and hydrogen peroxide production,150,151 along with increased phagocytosis152 and degranulation.153 These changes can be attributed to a lack of myeloperoxidase-dependent autoinactivation of neutrophil functions. One possibility is that sufficient hydrogen peroxide builds up in the absence of myeloperoxidase to kill directly or via hydroxyl radicals.154 However, myeloperoxidase-deficient cells release only slightly more hydrogen peroxide than normal, because of consumption by catalase,150 and since the hydroxyl radical production that has been detected in neutrophils is myeloperoxidase-dependent39 it should be diminished in deficient cells. We found that oxidative killing of S aureus by normal cells in the presence of azide was no better than with myeloperoxidase-deficient neutrophils, which accumulate less peroxide.106 Indeed, the difference in oxidative killing between cells lacking myeloperoxidase and NADPH-oxidase activity was so slight as to raise the possibility of whether there is a significant oxidative component independent of myeloperoxidase. The nonoxidative killing capacity of myeloperoxidase-deficient cells may be slightly enhanced,106,132 and it is possible to select in vitro conditions where these cells kill normally.61 However, CGD cells also kill normally under these conditions.

In our opinion, any slow oxidative killing that has been measured in vitro with myeloperoxidase-deficient cells does not provide a convincing explanation for the benign nature of myeloperoxidase deficiency and there is a need to look beyond killing by isolated neutrophils. One consideration is that NADPH oxidase is expressed in a number of inflammatory cells, including macrophages and eosinophils,155 whereas only neutrophils and monocytes have myeloperoxidase. CGD will affect a wider spectrum of cells than myeloperoxidase deficiency and this could contribute to its greater severity. Another possibility is that cytokines encountered by neutrophils as they move to a site of inflammation, or attachment to the endothelium, activate processes that assist killing. Both can enhance the oxidative burst.156,157 They may also activate neutrophils to express nitric oxide synthase.85,86 If so, a plausible scenario would be for peroxynitrite, generated from superoxide and nitric oxide, to act as a backup defense that abrogated the need for myeloperoxidase. Peroxynitrite might also be produced if nitric oxide from adjacent endothelial or mononuclear cells gained access to the neutrophil phagosome.

Alternatively, an aspect of pathogen clearance other than killing ability may distinguish the two enzyme deficiencies. One proposal is that neutrophil oxidants, but not myeloperoxidase, are critical for digestion rather than killing.158 A crucial phase of inflammation is the removal of neutrophils along with their ingested bacteria. Neutrophils become apoptotic once they have undergone phagocytosis, and oxidase products are implicated in the process.159,160 A critical step is the expression of surface markers such as phosphatidylserine that target the cells for ingestion and removal by macrophages.161 We have recently found that normal but not CGD neutrophils expose phosphatidylserine after stimulation with phorbol myristate acetate (Fadeel et al, manuscript submitted). However, myeloperoxidase-deficient cells or cells treated with azide exposed as much phosphatidylserine as normal cells (M.B. Hampton, C.C. Winterbourn, in preparation). Thus, the process requires hydrogen peroxide generation but not myeloperoxidase-derived oxidants. This mechanism could explain the different outcomes in myeloperoxidase-deficiency and CGD. Clearance of myeloperoxidase-deficient neutrophils by macrophages would be normal, even if their bacteria were killed more slowly. In contrast, CGD neutrophils would not be ingested, and their accumulation could give rise to the characteristic granulomas of the disease. A mouse model of chronic granulomatous disease has recently been developed.162-164 Neutrophils from these animals were defective not only in killing but also in their ability to dispose of dead microorganisms. Further studies with gene knockout models should help to test the proposals outlined above and bridge the gap between in vitro studies and clinical profiles.

In the century since Metchnikoff observed phagocytic cells ingesting bacteria, considerable progress has been made toward understanding the mechanisms involved in killing. However, there is still controversy and disagreement among researchers over some fundamental issues. HOCl appears as the most likely mediator of oxygen-dependent bacterial killing in the neutrophil phagosome. Chlorinated markers indicate that HOCl is generated in lethal amounts; however, analysis of the enzymology of myeloperoxidase has shown that a number of other reactions may occur, and it is not known whether the specific prevention of HOCl production affects bacterial killing. Superoxide is integral to many of the activities, and the ability of superoxide dismutase to inhibit killing suggests that superoxide is important in the physiological function of myeloperoxidase. Elucidating the biochemistry of the phagosome may ultimately lead to an understanding of how some pathogens can survive in such a harsh environment, and will assist in the development of therapies to attenuate the inflammatory pathologies where neutrophils unleash their destructive potential against host tissue.

Supported by the Health Research Council of New Zealand.

Address reprint requests to Christine C. Winterbourn, PhD, Department of Pathology, Free Radical Research Group, Christchurch School of Medicine, PO Box 4345, Christchurch, New Zealand; e-mail: ccw@chmeds.ac.nz.

1
Metchnikoff
E
Immunity in Infective Diseases.
1968
Johnson Reprint Corp
New York, NY
2
Mims
CA
The pathogenesis of infectious disease.
1987
Academic
San Diego, CA
3
Hirsch
JG
Cohn
ZA
Degranulation of polymorphonuclear leucocytes following phagocytosis of microorganisms.
J Exp Med
112
1960
1005
4
Sbarra
AJ
Karnovsky
ML
The biochemical basis of phagocytosis I. Metabolic changes during the ingestion of particles by polymorphonuclear leukocytes.
J Biol Chem
234
1959
1355
5
Iyer
GYN
Islam
MF
Quastel
JH
Biochemical aspects of phagocytosis.
Nature
192
1961
535
6
Segal
AW
Abo
A
The biochemical basis of the NADPH oxidase of phagocytes.
Trends Biochem Sci
18
1993
43
7
Babior
BM
Kipnes
RS
Curnutte
JT
Biological defense mechanisms: The production by leukocytes of superoxide, a potential bactericidal agent.
J Clin Invest
52
1973
741
8
Chanock
SJ
Benna
JE
Smith
RM
Babior
BM
The respiratory burst oxidase.
J Biol Chem
269
1994
24519
9
Smith
RM
Curnutte
JT
Molecular basis of chronic granulomatous disease.
Blood
77
1991
673
10
Johnston
RB
Inherited disorders of phagocyte killing
Scriver
CR
Beaudet
AL
Sly
WS
Valle
D
The Metabolic Basis of Inherited Disease.
1989
2779
McGraw-Hill
New York, NY
11
Segal
AW
The electron transport chain of the microbicidal oxidase of phagocytic cells and its involvement in the molecular pathology of chronic granulomatous disease.
J Clin Invest
83
1989
1785
12
Klebanoff
SJ
Phagocytic cells: Products of oxygen metabolism
Gallin
JI
Goldstein
IM
Snyderman
R
Inflammation: Basic Principles and Clinical Correlates.
1992
451
Raven
New York, NY
13
Robinson
JM
Badwey
JA
The NADPH oxidase complex of phagocytic leukocytes: A biochemical and cytochemical view.
Histochem Cell Biol
103
1995
163
14
Thomas
MJ
Hedrick
CC
Smith
S
Pang
J
Jerome
WG
Willard
AS
Shirley
PS
Superoxide generation by the human polymorphonuclear leukocyte in response to latex beads.
J Leukoc Biol
51
1992
591
15
Roos
D
Eckmann
CM
Yazdanbakhsh
M
Hamers
MN
de Boer
M
Excretion of superoxide by phagocytes measured with cytochrome c entrapped in resealed erythrocyte ghosts.
J Biol Chem
259
1984
1770
16
Makino
R
Tanaka
T
Iizuka
T
Ishimura
Y
Kanegasaki
S
Stoichiometric conversion of oxygen to superoxide anion during the respiratory burst in neutrophils.
J Biol Chem
261
1986
11444
17
Hyslop
PA
Hinshaw
DB
Scraufstatter
IU
Cochrane
CG
Kunz
S
Vosbeck
K
Hydrogen peroxide as a potent bacteriostatic antibiotic: Implications for host defense.
Free Radic Biol Med
19
1995
31
18
Imlay
JA
Linn
S
Bimodal pattern of killing of DNA-repair-defective or anoxically grown Escherichia coli by hydrogen peroxide.
J Bacteriol
166
1986
519
19
Klebanoff
SJ
Role of the superoxide anion in the myeloperoxidase-mediated antimicrobial system.
J Biol Chem
249
1974
3724
20
Babior
BM
Curnutte
JT
Kipnes
RS
Biological defense mechanisms. Evidence for the participation of superoxide in bacterial killing by xanthine oxidase.
J Lab Clin Med
85
1975
235
21
Rosen
H
Klebanoff
SJ
Bactericidal activity of a superoxide anion-generating system. A model for the polymorphonuclear leukocyte.
J Exp Med
149
1979
27
22
Samuni
A
Black
CDV
Krishna
CM
Malech
HL
Bernstein
EF
Russo
A
Hydroxyl radical production by stimulated neutrophils reappraised.
J Biol Chem
263
1988
13797
23
Cohen
MS
Britigan
BE
Hassett
DJ
Rosen
GM
Do human neutrophils form hydroxyl radical? Evaluation of an unresolved controversy.
Free Radic Biol Med
5
1988
81
24
Britigan
BE
Coffman
TJ
Buettner
GR
Spin trapping evidence for the lack of significant hydroxyl radical production during the respiration burst of human phagocytes using a spin adduct resistant to superoxide-mediated destruction.
J Biol Chem
265
1990
2650
25
Rosen
GM
Pou
S
Ramos
CL
Cohen
MS
Britigan
BE
Free radicals and phagocytic cells.
FASEB J
9
1995
200
26
Miller RA, Britigan BE: Role of oxidants in microbial pathophysiology. Clin Microbiol Rev 1, 1997
27
Tauber
AI
Babior
BM
Evidence for hydroxyl radical production by human neutrophils.
J Clin Invest
60
1977
374
28
Weiss
SJ
Rustagi
PK
LoBuglio
AF
Human granulocyte generation of hydroxyl radical.
J Exp Med
147
1978
316
29
Rosen
H
Klebanoff
SJ
Hydroxyl radical generation by polymorphonuclear leukocytes measured by electron spin resonance spectroscopy.
J Clin Invest
64
1979
1725
30
Davis
WB
Mohammed
BS
Mays
DC
She
Z
Mohammed
JR
Husney
RM
Sagone
AL
Hydroxylation of salicylate by activated neutrophils.
Biochem Pharmacol
38
1989
4013
31
Kettle
AJ
Winterbourn
CC
Superoxide-dependent hydroxylation by myeloperoxidase.
J Biol Chem
269
1994
17146
32
Winterbourn
CC
Lactoferrrin-catalysed hydroxyl radical production. Additional requirement for a chelating agent.
Biochem J
210
1983
15
33
Winterbourn
CC
Myeloperoxidase as an effective inhibitor of hydroxyl radical production: Implications for the oxidative reactions of neutrophils.
J Clin Invest
78
1986
545
34
Klebanoff
SJ
Waltersdorph
AM
Prooxidant activity of transferrin and lactoferrin.
J Exp Med
172
1990
1293
35
Cohen
MS
Britigan
BE
Chai
YS
Pou
S
Roeder
TL
Rosen
GM
Phagocyte-derived free radicals stimulated by ingestion of iron-rich Staphylococcus aureus: A spin-trapping study.
J Infect Dis
163
1991
819
36
Britigan
BE
Edeker
BL
Pseudomonas and neutrophil products modify transferrin and lactoferrin to create conditions that favor hydroxyl radical formation.
J Clin Invest
88
1991
1092
37
Coffman
TJ
Cox
CD
Edeker
BL
Britigan
BE
Possible role of bacterial siderophores in inflammation—Iron bound to the pseudomonas siderophore pyochelin can function as a hydroxyl radical catalyst.
J Clin Invest
86
1990
1030
38
Elzanowska
H
Wolcott
RG
Hannum
DM
Hurst
JK
Bactericidal properties of hydrogen peroxide and copper or iron-containing complex ions in relation to leukocyte function.
Free Radic Biol Med
18
1995
437
39
Ramos
CL
Pou
S
Britigan
BE
Cohen
MS
Rosen
GM
Spin trapping evidence for myeloperoxidase-dependent hydroxyl radical formation by human neutrophils and monocytes.
J Biol Chem
267
1992
8307
40
Candeias
LP
Patel
KB
Stratford
MRL
Wardman
P
Free hydroxyl radicals are formed on reaction between the neutrophil-derived species superoxide and hypochlorous acid.
FEBS Lett
333
1993
151
41
Wolcott
RG
Franks
BS
Hannum
DM
Hurst
JK
Bactericidal potency of hydroxyl radical in physiological environments.
J Biol Chem
269
1994
9721
42
Samuni
A
Czapski
G
Radiation induced damage in Escherichia coli B: The effects of superoxide radicals and molecular oxygen.
Radiat Res
76
1978
624
43
Czapski
G
Goldstein
S
Andorn
N
Aronovich
J
Radiation-induced generation of chlorine derivatives in N2O-saturated phosphate buffered saline: Toxic effects on Escherichia coli cells.
Free Radic Biol Med
12
1992
353
44
Allen
RC
Stjernholm
RL
Steele
RH
Evidence for the generation of an electronic excitation state(s) in human polymorphonuclear leukocytes and its participation in bactericidal activity.
Biochem Biophys Res Commun
47
1972
679
45
Shook
FC
On the question of singlet oxygen production in leucocytes, macrophages and the dismutation of superoxide anion
Bannister
WH
Bannister
JV
Biochemical and Clinical Aspects of Superoxide and Superoxide Dismutase.
1980
222
Elsevier/North-Holland
New York, NY
46
Kanofsky
JR
Singlet oxygen production in biological systems.
Chem Biol Interact
70
1989
1
47
Steinbeck
MJ
Khan
AU
Karnovsky
MJ
Intracellular singlet oxygen generation by phagocytosing neutrophils in response to particles coated with a chemical trap.
J Biol Chem
267
1992
13425
48
Kanofsky
JR
Hoogland
H
Wever
R
Weiss
SJ
Singlet oxygen production by human eosinophils.
J Biol Chem
263
1988
9692
49
Kettle
AJ
Winterbourn
CC
Myeloperoxidase: A key regulator of neutrophil oxidant production.
Redox Rep
3
1997
3
50
Bainton
DF
Ullyot
JL
Farquhar
MG
The development of neutrophilic polymorphonuclear leukocytes in human bone marrow.
J Exp Med
134
1971
907
51
Hurst
JK
Myeloperoxidase: active site structure and catalytic mechanisms
Everse
J
Everse
KE
Grisham
MB
Peroxidases in Chemistry and Biology.
1991
37
CRC
Boca Raton, FL
52
Dunford
HB
Free radicals in iron-containing systems.
Free Radic Biol Med
3
1987
405
53
Marquez
LA
Dunford
HB
Kinetics of oxidation of tyrosine and dityrosine by myeloperoxidase compounds I and II.
J Biol Chem
270
1996
30434
54
Heinecke
JW
Li
W
Daehnke
HL
Goldstein
JA
Dityrosine, a specific marker of oxidation, is synthesized by the myeloperoxidase-hydrogen peroxide system of human neutrophils and macrophages.
J Biol Chem
268
1993
4069
55
Harrison
JE
Shultz
J
Studies on the chlorinating activity of myeloperoxidase.
J Biol Chem
251
1976
1371
56
Klebanoff
SJ
Myeloperoxidase-halide-hydrogen peroxide antibacterial system.
J Bacteriol
95
1968
2131
57
Albrich
JM
Hurst
JK
Oxidative inactivation of Escherichia coli by hypochlorous acid. Rates and differentiation of respiratory from other reaction sites.
FEBS Lett
144
1982
157
58
Albrich
JM
Gilbaugh
JH
Callahan
KB
Hurst
JK
Effects of the putative neutrophil-generated toxin, hypochlorous acid, on membrane permeability and transport systems of Escherichia coli.
J Clin Invest
78
1986
177
59
Barrette WCJr
Hannum
DM
Wheeler
WD
Hurst
JK
General mechanism for the bacterial toxicity of hypochlorous acid: Abolition of ATP production.
Biochemistry
28
1989
9172
60
McKenna
SM
Davies
KJA
The inhibition of bacterial growth by hypochlorous acid; possible role in the bacterial activity of phagocytes.
Biochem J
254
1988
685
61
Rosen
H
Michel
BR
Redundant contribution of myeloperoxidase-dependent systems to neutrophil-mediated killing of Escherichia coli.
Infect Immun
65
1998
4173
62
Rosen
H
Orman
J
Rakita
RM
Michel
BR
VanDevanter
DR
Loss of DNA-membrane interactions and cessation of DNA synthesis in myeloperoxidase-treated Escherichia coli.
Proc Natl Acad Sci USA
87
1990
10048
63
Learn
DB
Myeloperoxidase-catalyzed oxidation of chloride and other halides: The role of chloramines
Everse
J
Everse
KE
Grisham
MB
Peroxidases in Chemistry and Biology.
1991
83
CRC
Boca Raton, FL
64
Grisham
MB
Jefferson
MM
Melton
DF
Thomas
EL
Chlorination of endogenous amines by isolated neutrophils. Ammonia-dependent bactericidal, cytotoxic and cytolytic activities of the chloramines.
J Biol Chem
259
1984
10404
65
Beilke
MA
Collins-Lech
C
Sohnle
PG
Candidacidal activity of the neutrophil myeloperoxidase system can be protected from excess hydrogen peroxide by the presence of ammonium ion.
Blood
73
1989
1045
66
Klebanoff
SJ
Myeloperoxidase: Occurrence and biological function
Everse
J
Everse
KE
Grisham
MB
Peroxidases in Chemistry and Biology.
1991
1
CRC
Boca Raton, FL
67
Thomas
EL
Fishman
M
Oxidation of chloride and thiocyanate by isolated leukocytes.
J Biol Chem
261
1986
9694
68
Van Dalen
CJ
Whitehouse
M
Winterbourn
CC
Kettle
AJ
Thiocyanate and chloride as competing substrates for myeloperoxidase.
Biochem J
327
1997
487
69
van der Vliet
A
Eiserich
JP
Halliwell
B
Cross
CE
Formation of reactive nitrogen species during peroxidase-catalyzed oxidation of nitrite: A potential additional mechanism of nitric oxide-dependent toxicity.
J Biol Chem
272
1997
7617
70
Winterbourn
CC
Pichorner
H
Kettle
AJ
Myeloperoxidase-dependent generation of a tyrosine peroxide by neutrophils.
Arch Biochem Biophys
338
1997
15
71
Foote
CS
Goyne
TE
Lehler
RI
Assessment of chlorination by human neutrophils.
Nature
301
1983
715
72
Weiss
SJ
Klein
R
Slivka
A
Wei
M
Chlorination of taurine by human neutrophils. Evidence for hypochlorous acid generation.
J Clin Invest
70
1982
598
73
Brunelli
L
Crow
JP
Beckman
JS
The comparative toxicity of nitric oxide and peroxynitrite to Escherichia coli.
Arch Biochem Biophys
316
1995
327
74
Nathan
C
Xie
Q
Nitric oxide synthases: Roles, tolls, and controls.
Cell
78
1994
915
75
Schmidt
HHHW
Walter
U
NO at work.
Cell
78
1994
919
76
Zhu
L
Gunn
C
Beckman
JS
Bactericidal activity of peroxynitrite.
Arch Biochem Biophys
298
1992
452
77
Kaplan
SS
Lancaster
JR
Basford
RE
Simmons
RL
Effect of nitric oxide on staphylococcal killing and interactive effect with superoxide.
Infect Immun
64
1996
69
78
Nathan
CF
Hibbs
JB
Role of nitric oxide synthesis in macrophage antimicrobial activity.
Curr Opin Immunol
3
1991
65
79
Denis
M
Human monocytes/macrophages: NO or no NO?
J Leukoc Biol
55
1994
682
80
Schmidt
HHHW
Seifert
R
Bohme
E
Formation and release of nitric oxide from human neutrophils and HL-60 cells induced by a chemotactic peptide, platelet activating factor and leukotriene B4.
FEBS Lett
244
1989
357
81
Carreras
MC
Pargament
GA
Catz
SD
Poderoso
JJ
Boveris
A
Kinetics of nitric oxide and hydrogen peroxide production and formation of peroxynitrite during the respiratory burst of human neutrophils.
FEBS Lett
341
1994
65
82
Krishna Rao
KM
Padmanabhan
J
Kilby
DL
Cohen
HJ
Currie
MS
Weinberg
JB
Flow cytometric analysis of nitric oxide production in human neutrophils using dichlorofluorescein diacetate in the presence of calmodulin inhibitor.
J Leukoc Biol
51
1992
496
83
Padgett
EL
Pruett
SB
Rat, mouse and human neutrophils stimulated by a variety of activating agents produce much less nitrite than rodent macrophages.
Immunology
84
1995
135
84
Yan
L
Vandivier
RW
Suffredini
AF
Danner
RL
Human polymorphonuclear leukocytes lack detectable nitric oxide synthase activity.
J Immunol
153
1994
1825
85
Wheeler
MA
Smith
SD
Garcia-Cardena
G
Nathan
CF
Weiss
RM
Sessa
WC
Bacterial infection induces nitric oxide synthase in human neutrophils.
J Clin Invest
99
1997
110
86
Evans
TJ
Buttery
LDK
Carpenter
A
Springall
DR
Polak
JM
Cohen
J
Cytokine-treated human neutrophils contain inducible nitric oxide synthase that produces nitration of ingested bacteria.
Proc Natl Acad Sci USA
93
1996
9553
87
Klebanoff
SJ
Reactive nitrogen intermediates and antimicrobial activity: Role of nitrite.
Free Radic Biol Med
14
1993
351
88
Stossel
TP
The machinery of cell crawling.
Sci Am
271
1994
40
89
Rozenberg-Arska
M
Salters
MEC
van Strijp
JAG
Geuze
JJ
Verhoef
J
Electron microscopic study of phagocytosis of Escherichia coli by human polymorphonuclear leukocytes.
Infect Immun
50
1985
852
90
Borregaard
N
Lollike
K
Kjeldsen
L
Sengelov
H
Bastholm
L
Nielsen
MH
Bainton
DF
Human neutrophil granules and secretory vesicles.
Eur J Haematol
51
1993
187
91
Riches
DWH
Phagocytic cells: Degranulation and secretion
Gallin
JI
Goldstein
IM
Snyderman
R
Inflammation: Basic Principles and Clinical Correlates.
1988
363
Raven
New York, NY
92
Menegazzi
R
Busetto
S
Dri
P
Cramer
R
Patriarca
P
Chloride ion efflux regulates adherence, spreading, and respiratory burst of neutrophils stimulated by tumor necrosis factor-α (TNF) on biologic surfaces.
J Cell Biol
135
1996
511
93
Fittschen
C
Henson
PM
Linkage of azurophil granule secretion in neutrophils to chloride ion transport and endosomal transcytosis.
J Clin Invest
93
1994
247
94
Demaurex
N
Schrenzel
J
Jaconi
ME
Lew
DP
Krause
K-H
Proton channels, plasma membrane potential, and respiratory burst in human neutrophils.
Eur J Biochem
51
1993
309
95
Segal
AW
Geisow
M
Garcia
R
Harper
A
Miller
R
The respiratory burst of phagocytic cells is associated with a rise in vacuolar pH.
Nature
290
1981
406
96
Cech
P
Lehrer
RI
Phagolysosomal pH of human neutrophils.
Blood
63
1984
88
97
Nanda
A
Curnutte
JT
Grinstein
S
Activation of H+ conductance in neutrophils requires assembly of components of the respiratory burst oxidase but not its redox function.
J Clin Invest
93
1994
1770
98
Lock
R
Dahlgren
C
Characteristics of the granulocyte chemiluminescence reaction following an interaction between human neutrophils and Salmonella typhimurium bacteria.
APMIS
96
1988
299
99
Lundqvist
H
Karlsson
A
Follin
P
Sjolin
C
Dahlgren
C
Phagocytosis following translocation of the the b-cytochrome from the specific granules to the plasma membrane is associated with an increased leakage of reactive oxygen species.
Scand J Immunol
36
1992
885
100
Nathan
DG
Baehner
RL
Weaver
DK
Failure of nitro blue tetrazolium reduction in the phagocytic vacuoles of leukocytes in chronic granulomatous disease.
J Clin Invest
48
1969
1895
101
Briggs
RT
Robinson
JM
Karnovsky
ML
Karnovsky
MJ
Superoxide production by polymorphonuclear leukocytes A cytochemical approach.
Histochemistry
84
1986
371
102
Karnovsky
MJ
Cytochemistry and reactive oxygen species: A retrospective.
Histochemistry
102
1994
15
103
Briggs
RT
Karnovsky
ML
Karnovsky
MJ
Cytochemical demonstration of hydrogen peroxide in polymorphonuclear phagosomes.
J Cell Biol
64
1975
254
104
Root
RK
Metcalf
JA
H2O2 release from human granulocytes during phagocytosis. Relationship to superoxide anion formation and cellular catabolism of H2O2: Studies with normal and cytochalasin B-treated cells.
J Clin Invest
60
1977
1266
105
Test
ST
Weiss
SJ
Quantitative and temporal characterization of the extracellular hydrogen peroxide pool generated by human neutrophils.
J Biol Chem
259
1984
399
106
Hampton
MB
Kettle
AJ
Winterbourn
CC
The involvement of superoxide and myeloperoxidase in oxygen-dependent bacterial killing.
Infect Immun
64
1996
3512
107
Winterbourn
CC
Garcia
R
Segal
AW
Production of the superoxide adduct of myeloperoxidase (compound III) by stimulated neutrophils, and its reactivity with H2O2 and chloride.
Biochem J
228
1985
583
108
Kettle
AJ
Winterbourn
CC
Superoxide modulates the activity of myeloperoxidase and optimizes the production of hypochlorous acid.
Biochem J
252
1988
529
109
Kettle
AJ
Winterbourn
CC
Mechanism of inhibition of myeloperoxidase by anti-inflammatory drugs.
Biochem Pharmacol
41
1991
1485
110
Kettle
AJ
Gedye
CA
Winterbourn
CC
Superoxide is an antagonist of anti-inflammatory drugs that inhibit hypochlorous acid production by myeloperoxidase.
Biochem Pharmacol
45
1993
2003
111
Cuperus
RA
Muijsers
AO
Wever
R
The superoxidase activity of myeloperoxidase: Formation of compound III.
Biochim Biophys Acta
871
1986
78
112
Zgliczynski
JM
Stelmaszynska
T
Chlorinating ability of human phagocytosing leucocytes.
Eur J Biochem
56
1975
157
113
Klebanoff
SJ
Iodination of bacteria: A bactericidal mechanism.
J Exp Med
126
1967
1063
114
Klebanoff
SJ
Clark
RA
Iodination of human polymorphonuclear leukocytes: A re-evaluation.
J Lab Clin Med
89
1977
675
115
Segal
AW
Garcia
RC
Harper
AM
Iodination by stimulated human neutrophils. Studies on its stoichiometry, subcellular localization and relevance to microbial killing.
Biochem J
210
1983
215
116
Hazen
SL
Hsu
FF
Mueller
DM
Crowley
JR
Heinecke
JW
Human neutrophils employ chlorine gas as an oxidant during phagocytosis.
J Clin Invest
98
1996
1283
117
Hampton MB: The role of neutrophil oxidants in bacterial killing. Doctoral thesis, University of Otago, Dunedin, New Zealand, 1995
118
Hurst
JK
Albrich
JM
Green
TR
Rosen
H
Klebanoff
SJ
Myeloperoxidase-dependent fluorescein chlorination by stimulated neutrophils.
J Biol Chem
259
1984
4812
119
Jiang
Q
Griffin
DA
Barofsky
DF
Hurst
JK
Intraphagosomal chlorination dynamics and yields determined using unique fluorescent bacterial mimics.
Chem Res Toxicol
10
1997
1080
120
Lehrer
RI
Ganz
T
Antimicrobial polypeptides of human neutrophils.
Blood
76
1990
2169
121
Martin
E
Ganz
T
Lehrer
RI
Defensins and other endogenous peptide antibiotics of vertebrates.
J Leukoc Biol
58
1995
128
122
Weiss
J
Phagocytic cells: Oxygen-independent antimicrobial systems
Gallin
JI
Goldstein
IM
Snyderman
R
Inflammation: Basic Principles and Clinical Correlates.
1992
603
Raven
New York, NY
123
Mandell
GL
Bactericidal activity of aerobic and anaerobic polymorphonuclear neutrophils.
Infect Immun
9
1974
337
124
McRipley
RJ
Sbarra
AJ
Role of the phagocyte in host-parasite interactions XII. Hydrogen peroxide-myeloperoxidase bactericidal system in the phagocyte.
J Bacteriol
94
1967
1425
125
Ellis
JA
Mayer
SJ
Jones
OTG
The effect of the NADPH oxidase inhibitor diphenyleneiodonium on aerobic and anaerobic microbicidal activities of human neutrophils.
Biochem J
251
1988
887
126
Hampton
MB
Winterbourn
CC
Modification of neutrophil oxidant production with diphenyleneiodonium and its effect on neutrophil function.
Free Radic Biol Med
18
1995
633
127
Mandell
GL
Hook
EW
Leukocyte bactericidal activity in chronic granulomatous disease: Correlation of bacterial hydrogen peroxide production and susceptibility to bacterial killing.
J Bacteriol
100
1969
531
128
Pitt
J
Bernheimer
HP
Role of peroxide in phagocytic killing of pneumococci.
Infect Immun
9
1974
48
129
Lehrer
RI
Hanifin
J
Cline
MJ
Defective bactericidal activity in myeloperoxidase-deficient human neutrophils.
Nature
223
1969
78
130
Klebanoff
SJ
Hamon
CB
Role of myeloperoxidase mediated antimicrobial systems in intact leukocytes.
J Reticuloendothel Soc
12
1972
170
131
Kitahara
M
Eyre
HJ
Simonian
J
Atkin
CL
Hasstedt
SJ
Hereditary myeloperoxidase deficiency.
Blood
57
1981
888
132
Klebanoff
SJ
Myeloperoxidase: Contribution to the microbicidal activity of intact leukocytes.
Science
169
1970
1095
133
Humphreys
JM
Davies
B
Hart
CA
Edwards
SW
Role of myeloperoxidase in the killing of Staphylococcus aureus by human neutrophils: Studies with the myeloperoxidase inhibitor salicylhydroxamic acid.
J Gen Microbiol
135
1989
1187
134
Odell
EW
Segal
AW
The bactericidal effects of the respiratory burst and the myeloperoxidase system isolated in neutrophil cytoplasts.
Biochim Biophys Acta
971
1988
266
135
Wagner
DK
Collins-Lech
C
Sohnle
PG
Inhibition of neutrophil killing of Candida albicans pseudohyphae by substances which quench hypochlorous acid and chloramines.
Infect Immun
51
1997
731
136
Johnston RB Jr
Keele BB Jr
Misra
HP
Lehmeyer
JE
Webb
LS
Baehner
RL
Rajagopalan
KV
The role of superoxide anion generation in phagocytic bactericidal activity. Studies with normal and chronic granulomatous disease leukocytes.
J Clin Invest
55
1975
1357
137
Mandell
GL
Catalase, superoxide dismutase, and virulence of Staphylococcus aureus. In vitro and in vivo studies with emphasis on staphylococcal-leukocyte interaction.
J Clin Invest
55
1994
561
138
Schwartz
CE
Krall
J
Norton
L
McKay
K
Kay
D
Lynch
RE
Catalase and superoxide dismutase in Escherichia coli. Roles in resistance to killing by neutrophils.
J Biol Chem
258
1983
6277
139
Welch
DF
Role of catalase and superoxide dismutase in the virulence of Listeria monocytogenes.
Ann Inst Pasteur/Microbiol (Paris)
138
1987
265
140
Papp-Szabò
E
Sutherland
CL
Josephy
PD
Superoxide dismutase and the resistance of Escherichia coli to phagocytic killing by human neutrophils.
Infect Immun
61
1994
1442
141
Papp-Szabò
E
Firtel
M
Josephy
PD
Comparison of the sensitivities of Salmonella typhimurium oxyR and katG mutants to killing by human neutrophils.
Infect Immun
62
1994
2662
142
McManus
DC
Josephy
PD
Superoxide dismutase protects Escherichia coli against killing by human serum.
Arch Biochem Biophys
317
1995
57
143
Kusunose
E
Ichihara
K
Noda
Y
Kusunose
M
Superoxide dismutase from Mycobacterium tuberculosis.
J Biochem
80
1994
1343
144
Beaman
BL
Scates
SM
Moring
SE
Deem
R
Misra
HP
Purification and properties of a unique superoxide dismutase from Nocardia asteroides.
J Biol Chem
258
1994
91
145
Spiegelhalder
C
Gerstenecker
B
Kersten
A
Schiltz
E
Kist
M
Purification of Helicobacter pylori superoxide dismutase and cloning and sequencing of the gene.
Infect Immun
61
1993
5315
146
Langford
PR
Loynds
BM
Kroll
JS
Cloning and molecular characterization of Cu,Zn superoxide dismutase from Actinobacillus pleuropneumoniae.
Infect Immun
64
1997
5035
147
Beaman
BL
Black
CM
Doughty
F
Beaman
L
Role of superoxide dismutase and catalase as determinants of pathogenicity of Nocardia asteroides: Importance in resistance to microbicidal activities of human polymorphonuclear neutrophils.
Infect Immun
47
1985
135
148
Beaman
L
Beaman
BL
Monoclonal antibodies demonstrate that superoxide dismutase contributes to protection of Nocardia asteroides within the intact host.
Infect Immun
58
1990
3122
149
Thong
YH
How important is the myeloperoxidase microbicidal system of phagocytic cells?
Med Hypotheses
8
1982
249
150
Nauseef
WM
Metcalf
JA
Root
RK
Role of myeloperoxidase in the respiratory burst of human neutrophils.
Blood
61
1983
483
151
Rosen
H
Klebanoff
SJ
Chemiluminescence and superoxide production by myeloperoxidase-deficient leukocytes.
J Clin Invest
58
1976
50
152
Stendahl
O
Coble
BI
Dahlgren
C
Hed
J
Molin
L
Myeloperoxidase modulates the phagocytic activity of polymorphonuclear neutrophil leukocytes. Studies with cells from a myeloperoxidase-deficient patient.
J Clin Invest
73
1984
366
153
Dri
P
Cramer
R
Menegazzi
R
Patriarca
P
Increased degranulation of human myeloperoxidase-deficient polymorphonuclear leucocytes.
Br J Haematol
59
1985
115
154
Klebanoff
SJ
Pincus
SH
Hydrogen peroxide utilization in myeloperoxidase-deficient leukocytes: A possible microbicidal control mechanism.
J Clin Invest
50
1971
2226
155
Cross
AR
Jones
OTG
Enzymic mechanisms of superoxide production.
Biochim Biophys Acta
1057
1991
281
156
Nathan
CF
Respiratory burst in adherent human neutrophils: Triggering by colony-stimulating factors CSF-GM and CSF-G.
Blood
73
1989
301
157
Nathan
CF
Neutrophil activation on biological surfaces: Massive secretion of hydrogen peroxide in response to products of macrophages and lymphocytes.
J Clin Invest
80
1987
1550
158
Weiss
J
Kao
L
Victor
M
Elsbach
P
Respiratory burst facilitates the digestion of Escherichia coli killed by polymorphonuclear leukocytes.
Infect Immun
55
1987
2142
159
Coxon
A
Rieu
P
Barkalow
FJ
Askari
S
Sharpe
AH
von Andrian
UH
Arnaout
MA
Mayadas
TN
A novel role for the beta 2 integrin CD11b/CD18 in neutrophil apoptosis: A homeostatic mechanism in inflammation.
Immunity
5
1996
653
160
Kasahara
Y
Iwai
K
Yachie
A
Ohta
K
Konno
A
Seki
H
Miyawaki
T
Taniguchi
N
Involvement of reactive oxygen intermediates in spontaneous and CD96 (Fas/APO-1)-mediated apoptosis of neutrophils.
Blood
89
1997
1748
161
Savill
JS
Wyllie
AH
Henson
JE
Walport
MJ
Henson
PM
Haslett
C
Macrophage phagocytosis of aging neutrophils in inflammation: Programmed cell death in the neutrophil leads to its recognition by macrophages.
J Clin Invest
83
1997
865
162
Pollock
JD
Williams
DA
Gifford
MAC
Lin Li
L
Du
X
Fisherman
J
Orkin
SH
Doerschuk
CM
Dinauer
MC
Mouse model of X-linked chronic granulomatous disease, an inherited defect in phagocyte superoxide production.
Nat Genet
9
1995
202
163
Jackson
SH
Gallin
JI
Holland
SM
The p47phox mouse knock-out model of chronic granulomatous disease.
J Exp Med
182
1995
751
164
Morgenstern
DE
Gifford
MAC
Li
LL
Doerschuk
CM
Dinauer
MC
Absence of respiratory burst in X-linked chronic granulomatous disease mice leads to abnormalities in both host defense and inflammatory response to aspergillus fumigatus.
J Exp Med
185
1997
207
165
Kettle
AJ
Detection of chlorotyrosine in albumin exposed to stimulated human neutrophils.
FEBS Lett
379
1996
103
Sign in via your Institution