Elsevier

Free Radical Biology and Medicine

Volume 33, Issue 6, 15 September 2002, Pages 774-797
Free Radical Biology and Medicine

Review article
Structure and function of xanthine oxidoreductase: where are we now?

https://doi.org/10.1016/S0891-5849(02)00956-5Get rights and content

Abstract

Xanthine oxidoreductase (XOR) is a complex molybdoflavoenzyme, present in milk and many other tissues, which has been studied for over 100 years. While it is generally recognized as a key enzyme in purine catabolism, its structural complexity and specialized tissue distribution suggest other functions that have never been fully identified. The publication, just over 20 years ago, of a hypothesis implicating XOR in ischemia-reperfusion injury focused research attention on the enzyme and its ability to generate reactive oxygen species (ROS). Since that time a great deal more information has been obtained concerning the tissue distribution, structure, and enzymology of XOR, particularly the human enzyme. XOR is subject to both pre- and post-translational control by a range of mechanisms in response to hormones, cytokines, and oxygen tension. Of special interest has been the finding that XOR can catalyze the reduction of nitrates and nitrites to nitric oxide (NO), acting as a source of both NO and peroxynitrite. The concept of a widely distributed and highly regulated enzyme capable of generating both ROS and NO is intriguing in both physiological and pathological contexts. The details of these recent findings, their pathophysiological implications, and the requirements for future research are addressed in this review.

Introduction

Xanthine oxidoreductase (XOR) is a complex molybdoflavoenzyme that is readily available from cows’ milk, where it forms a major component of the milk fat globule membrane (MFGM) [1], [2]. Largely because of this availability, XOR has been known for 100 years and studied in its essentially pure form for over 60 years [3]. The enzyme has consequently become a model for structural and mechanistic studies of molybdoenzymes in general, involving the application of electron spin resonance, x-ray absorption fine structure, and many other physicochemical techniques [4], [5]. The last few years have seen a dramatic increase in molecular information, with determination of its cDNA and gene sequences and publication of its three-dimensional structure.

XOR is generally recognized as the terminal enzyme of purine catabolism in man, catalyzing the hydroxylation of hypoxanthine to xanthine and of xanthine to urate. Until recently, little was known of the human enzyme, although inherited XOR deficiency, xanthinuria, has long been recognized, and, puzzlingly, known to be asymptomatic [6], [7]. Some 20 years ago, attention was directed toward the physiological and pathological significance of XOR by Granger, McCord, and colleagues, who proposed a role for XOR-derived reactive oxygen species (ROS) in ischemia-reperfusion (IR) injury [8], [9], [10]. Their hypothesis, which generated several hundred publications, undoubtedly helped to stimulate interest in XOR as a source of ROS, not only in many pathological states, but also in signal transduction generally. As recognition of the pathophysiological involvement of XOR increases, so does the need to understand its enzymology, kinetics, and control. As noted above, our structural knowledge of XOR has markedly increased in recent years. Moreover, a great deal of new information has become available concerning species and tissue variations of the enzyme, its substrate specificity, and regulation. For these reasons, it seems to be appropriate to review recent developments and to attempt to inter-relate them.

In general, the present survey focuses on works published in the last 10 years, and this is reflected in the reference list, which, wherever possible, covers earlier work by quoting more recent, especially review, articles. In view of the potential clinical relevance of much of the work described, emphasis is given to studies using mammalian, particularly human, enzyme and tissues.

Section snippets

Assays in tissues

XOR activity has been detected in all species examined, including bacteria [11], [12]. In mammalian tissues, activity is widely distributed, with highest levels being found in liver and intestine [12]. There is, however, considerable species variation, as exemplified by the wide range of levels in blood [12] and heart [13], [14]. In humans, apart from liver and intestine, most tissues show little XOR activity [12], [13], [14], [15], [16], [17]. XOR mRNA levels were also found to be highest in

Mammalian and avian XOR

The best-characterized form of XOR is undoubtedly that purified from bovine milk [1], [3]. XOR has also been purified from rat [42], [43], chicken [44], [45], and turkey [46] livers and from mouse mammary gland [47]. While these preparations are less well characterized than that from bovine milk, their properties are generally similar to those of the latter. In all cases, XOR occurs as a homodimer of approximately 300 kDa; each subunit contains four redox centers; viz. a molybdenum cofactor

Structure

The structure of the genetic loci coding for human [77] and mouse [78] XOR have been elucidated and the respective genes have been assigned to chromosomes 2p22 [79] and 17 [78]. These mammalian genes contain 36 exons and the exon-intron structure is highly conserved, differing greatly from those of the Drosophila or Calliphora genes, which are more compact, with only four or five exons [80].

Of mammalian and avian XORs, cDNA sequences have been reported for rat [81], mouse [82], chicken [83],

Ischemia-reperfusion injury

Granger and colleagues [8], [9], [10] focused attention on XOR by proposing a key role for the enzyme in the pathogenesis of ischemia-reperfusion (IR) injury. Their hypothetical mechanism, outlined in Fig. 4, can be briefly summarized as follows. In the course of ischemia, transmembrane ion gradients are dissipated, allowing elevated cytosolic concentrations of calcium. This, in turn, activates a protease that irreversibly converts XDH, predominant in vivo, into XO. Concurrently, cellular ATP

Circulating XOR

The presence of circulating XOR in mammals has long been recognized, and many different assays have been developed for its determination, particularly in humans, in which levels are particularly low. Most of these assays measure enzymic activity, and involve quantification of product by a range of means, including radioactivity (urate or xanthine) [127], [128], [129], fluorescence (isoxanthopterin) [130], [131], [132], chemiluminescence (superoxide) [133] and UV-visible spectroscopy (urate)

Pathological roles of XOR in the vasculature

Recent research into IR injury has focused on the vasculature, especially on leukocyte-endothelial cell interactions, which are relevant to many vascular diseases [122], [125], [126].

As noted in “Distribution,” XOR is present both in the cytoplasm and on the outer surface of endothelial cells [37], [38]. In the vasculature, such extracellular endothelial XOR could well be supplemented by circulating enzyme, levels of which can vary greatly in certain pathological states (see “Circulating XOR”).

Regulation

As noted in “Distribution,” XOR activity in human tissues is generally low, compared with other mammalian species. Clearly, enzyme activity is subject to control at several levels, more than one of which appear to be involved in this case. Thus, Xu and colleagues [195] have recently shown that, relative to mice, both transcription rates and core promoter activity of the human gene are repressed. Analysis of human XOR promoter activity in different cell types showed both repressor and activator

Enzymology

It has recently been established that XOR can act as a source of NO. Initial reports described the detection of NO, by chemiluminescence, following incubation of XOR with NADH and nitrate [227], [229] or nitrite [228], [229], [230] under hypoxic conditions. The anaerobic XOR-catalyzed reduction of inorganic nitrite to NO was subsequently investigated in detail by Godber et al. [182], who showed xanthine to be a more effective reducing substrate than NADH in this reaction. Steady state kinetic

Conclusion

As noted in the “Introduction,” interest in XOR received a major boost in the 1980s, following publication of its hypothetical involvement in IR injury. In fact, and not only for this reason, a great deal of new information concerning the enzyme has been gained in the last 20 years. Nevertheless, as will be clear from this review, many uncertainties remain.

Studies of the tissue and cellular distribution of XOR are contradictory and potential reasons for this are discussed in “Distribution.”

Acknowledgements

I am most grateful to Dr. Arwen Pearson for the simulated structures of Fig. 3 and to Jonanthan Gledhill for help with other figures. Thanks are also due to Prof. Robert Eisenthal for discussions and careful reading of the manuscript.

References (265)

  • T. Ikegami et al.

    The presence of desulfo-xanthine dehydrogenase in purified and crude enzyme preparations from rat liver

    Arch. Biochem. Biophys.

    (1986)
  • K.V. Rajagopalan et al.

    Purification and properties of chicken liver xanthine dehydrogenase

    J. Biol. Chem.

    (1967)
  • T. Nishino et al.

    The nicotinamide adenine dinucleotide-binding site of chicken liver xanthine dehydrogenase

    J. Biol. Chem.

    (1989)
  • W.F. Cleere et al.

    Avian xanthine dehydrogenases I. Isolation and characterisation of the turkey liver enzyme

    Comp. Biochem. Physiol.

    (1975)
  • J.L. McManaman et al.

    Mouse mammary gland xanthine oxidoreductasepurification, characterisation and regulation

    Arch. Biochem. Biophys.

    (1999)
  • J. Hunt et al.

    Studies of the reductive half-reaction of milk xanthine dehydrogenase

    J. Biol. Chem.

    (1994)
  • R.C. Wahl et al.

    Evidence for the inorganic nature of the cyanolyzable sulfur of molybdenum hydroxylases

    J. Biol. Chem.

    (1982)
  • D. Furth-Walker et al.

    Regulation of xanthine oxidase activity and immunologically detectable protein in rats in response to dietary protein and iron

    J. Nutr.

    (1987)
  • T. Nishino et al.

    Studies on chicken liver xanthine dehydrogenase with reference to the problem of non-equivalence of FAD moieties

    Biochim. Biophys. Acta

    (1975)
  • W.A. Frederiks et al.

    The proportion of xanthine oxidase activity of total xanthine oxidoreductase activity in situ remains constant in rat liver under various (patho)physiological conditions

    Hepatology

    (1996)
  • V. Massey et al.

    On the mechanism of inactivation of xanthine oxidase by cyanide

    J. Biol. Chem.

    (1970)
  • H. Komai et al.

    The preparation and properties of deflavo xanthine oxidase

    J. Biol. Chem.

    (1969)
  • T.A. Krenitsky et al.

    Xanthine oxidase from human liverpurification and characterisation

    Arch. Biochem. Biophys.

    (1986)
  • S. Abadeh et al.

    Purification and partial characterisation of xanthine oxidase from human milk

    Biochim. Biophys. Acta

    (1992)
  • S. Gardlik et al.

    A molybdopterin-free form of xanthine oxidase

    Arch. Biophys. Biochem.

    (1987)
  • S. Page et al.

    Xanthine oxidoreductase in human mammary epithelial cellsactivation in response to inflammatory cytokines

    Biochim. Biophys. Acta

    (1998)
  • P. Xu et al.

    Molecular cloning and characterization of the human xanthine dehydrogenase gene

    Genomics

    (1996)
  • G. Cazzaniga et al.

    Chromosomal mapping, isolation and characterization of the mouse xanthine dehydrogenase gene

    Genomics

    (1994)
  • P. Xu et al.

    Assignment of human xanthine dehydrogenase gene to chromosome 2p22

    Genomics

    (1994)
  • Y. Amaya et al.

    Proteolytic conversion of xanthine dehydrogenase from the NAD-dependent type to the O2-dependent type

    J. Biol. Chem.

    (1990)
  • P. Xu et al.

    Molecular cloning, tissue expression of human xanthine dehydrogenase

    Biochem. Biophys. Res. Commun.

    (1994)
  • P. Xu et al.

    Molecular cloning, tissue expression of human xanthine dehydrogenase

    Biochem. Biophys. Res. Commun.

    (1995)
  • H. Schindelin et al.

    Molybdopterin from molybdenum and tungsten enzymes

    Adv. Protein Chem.

    (2001)
  • J. Rebelo et al.

    Gene sequence and crystal structure of the aldehyde oxidoreductase from Desulfovibrio desulfuricans ATCC 27774

    J. Mol. Biol.

    (2000)
  • V. Massey et al.

    Milk xanthine dehydrogenasethe first one hundred years

    Biochem. Soc. Trans.

    (1997)
  • Bray, R. C. The inorganic biochemistry of molybdoenzymes. Q. Rev. Biophys....
  • R. Hille

    The mononuclear molybdenum enzymes

    Chem. Rev.

    (1996)
  • J.L. Johnson et al.

    Molydenum cofactor deficiency and isolated sulfite oxidase deficiency

  • H.A. Simmonds et al.

    Hereditary xanthinuria

  • J.M. McCord

    Oxygen-derived free radicals in postischemic tissue injury

    N. Engl. J. Med.

    (1985)
  • D.N. Granger et al.

    Ischemia-reperfusion injuryrole of oxygen-derived free radicals

    Acta Physiol. Scand.

    (1986)
  • T.A. Krenitisky et al.

    A comparison of the distribution and electron acceptor specificities of xanthine oxidase and aldehyde oxidase

    Comp. Biochem. Physiol.

    (1974)
  • D.A. Parks et al.

    Xanthine oxidasebiochemisty, distribution and physiology

    Acta Physiol. Scand.

    (1986)
  • J.W. De Jong et al.

    Xanthine oxidoreductase activity in perfused hearts of various species, including humans

    Circ. Res.

    (1990)
  • K. Vettenranta et al.

    Xanthine oxidase during fetal development

    Pediatr. Res.

    (1990)
  • A. Sarnesto et al.

    Organ distribution and molecular forms of human xanthine dehydrogenase, xanthine oxidase protein

    Lab. Invest.

    (1996)
  • M. Saksela et al.

    Xanthine oxidoreductase gene expression and enzyme activity in developing human tissues

    Biol. Neonate

    (1998)
  • M. Kurosaki et al.

    Tissue and cell-specific expression of mouse xanthine oxidoreductase gene in vivoregulation by bacterial lipopolysaccharide

    Biochem. J.

    (1995)
  • G. Bruder et al.

    High concentrations of antibodies to xanthine oxidase in human and animal sera. Molecular characterization

    J. Clin. Invest.

    (1984)
  • A. Kooij et al.

    High levels of xanthine oxidoreductase in rat endothelial, epithelial and connective tissue cells

    Virchows Arch. B Cell Pathol.

    (1992)
  • Cited by (690)

    View all citing articles on Scopus
    1

    Roger Harrison obtained an MA in Natural Sciences at the University of Cambridge and a PhD in Chemistry at the University of Birmingham, UK. Following postdoctoral research fellowships at the State University of New York, Buffalo, NY, USA, and at the National Institutes of Health, Bethesda, MD, USA, he was awarded an Alexander von Humboldt research fellowship at the University of Hamburg, Germany, where he spent 2 years. After a final postdoctoral research fellowship at the University of Loughborough, UK, he joined the University of Bath, where he is presently Professor of Biochemistry. For the last 12 years his research group has focused on the enzymology and immunology of xanthine oxidoreductase, with particular interest in the human enzyme, and more latterly on its catalysis of nitric oxide and peroxynitrite production.

    View full text