HPLC-based monitoring of products formed from hydroethidine-based fluorogenic probes--the ultimate approach for intra- and extracellular superoxide detection.

BACKGROUND Nearly ten years ago, we demonstrated that superoxide radical anion (O2⋅¯) reacts with the hydroethidine dye (HE, also known as dihydroethidium, DHE) to form a diagnostic marker product, 2-hydroxyethidium (2-OH-E(+)). This particular product is not derived from reacting HE with other biologically relevant oxidants (hydrogen peroxide, hydroxyl radical, or peroxynitrite). This discovery negated the longstanding view that O2⋅¯ reacts with HE to form the other oxidation product, ethidium (E(+)). It became clear that due to the overlapping fluorescence spectra of E(+) and 2-OH-E(+), fluorescence-based techniques using the "red fluorescence" are not suitable for detecting and measuring O2⋅¯ in cells using HE or other structurally analogous fluorogenic probes (MitoSOX(TM) Red or hydropropidine). However, using HPLC-based assays, 2-OH-E(+) and analogous hydroxylated products can be easily detected and quickly separated from other oxidation products. SCOPE OF REVIEW The principles discussed in this chapter are generally applicable in free radical biology and medicine, redox biology, and clinical and translational research. The assays developed here could be used to discover new and targeted inhibitors for various superoxide-producing enzymes, including NADPH oxidase (NOX) isoforms. MAJOR CONCLUSIONS HPLC-based approaches using site-specific HE-based fluorogenic probes are eminently suitable for monitoring O2⋅¯ in intra- and extracellular compartments and in mitochondria. The use of fluorescence-microscopic methods should be avoided because of spectral overlapping characteristics of O2⋅¯-derived marker product and other, non-specific oxidized fluorescent products formed from these probes. GENERAL SIGNIFICANCE Methodologies and site-specific fluorescent probes described in this review can be suitably employed to delineate oxy radical dependent mechanisms in cells under physiological and pathological conditions. This article is part of a Special Issue entitled Current methods to study reactive oxygen species - pros and cons and biophysics of membrane proteins. Guest Editor: Christine Winterbourn.

[1]  P. Wardman,et al.  Cytochrome C is a potent catalyst of dichlorofluorescin oxidation: implications for the role of reactive oxygen species in apoptosis. , 2001, Biochemical and biophysical research communications.

[2]  Gregor Rothe,et al.  Flow Cytometric Analysis of Respiratory Burst Activity in Phagocytes With Hydroethidine and 2′,7′‐Dichlorofluorescin , 1990, Journal of leukocyte biology.

[3]  João Wosniak,et al.  Protein disulfide isomerase overexpression in vascular smooth muscle cells induces spontaneous preemptive NADPH oxidase activation and Nox1 mRNA expression: effects of nitrosothiol exposure. , 2009, Archives of biochemistry and biophysics.

[4]  Brian J. Smith,et al.  Mutation of succinate dehydrogenase subunit C results in increased O2.-, oxidative stress, and genomic instability. , 2006, Cancer research.

[5]  Hongtao Zhao,et al.  Detection and characterization of the product of hydroethidine and intracellular superoxide by HPLC and limitations of fluorescence. , 2005, Proceedings of the National Academy of Sciences of the United States of America.

[6]  Joy Joseph,et al.  Global Profiling of Reactive Oxygen and Nitrogen Species in Biological Systems , 2011, The Journal of Biological Chemistry.

[7]  B. Kalyanaraman,et al.  Detection of 2-hydroxyethidium in cellular systems: a unique marker product of superoxide and hydroethidine , 2007, Nature Protocols.

[8]  F. Laurindo,et al.  Assessment of superoxide production and NADPH oxidase activity by HPLC analysis of dihydroethidium oxidation products. , 2008, Methods in enzymology.

[9]  B. Kalyanaraman,et al.  Cytochrome c-mediated oxidation of hydroethidine and mito-hydroethidine in mitochondria: identification of homo- and heterodimers. , 2008, Free radical biology & medicine.

[10]  D. Pimentel,et al.  S-Glutathiolation of p21ras by Peroxynitrite Mediates Endothelial Insulin Resistance Caused by Oxidized Low-Density Lipoprotein , 2006, Arteriosclerosis, thrombosis, and vascular biology.

[11]  Hongtao Zhao,et al.  Superoxide reacts with hydroethidine but forms a fluorescent product that is distinctly different from ethidium: potential implications in intracellular fluorescence detection of superoxide. , 2003, Free radical biology & medicine.

[12]  V. Neugebauer,et al.  Mitochondrial Reactive Oxygen Species Are Activated by mGluR5 through IP3 and Activate ERK and PKA to Increase Excitability of Amygdala Neurons and Pain Behavior , 2011, The Journal of Neuroscience.

[13]  B. Kalyanaraman Oxidative chemistry of fluorescent dyes: implications in the detection of reactive oxygen and nitrogen species. , 2011, Biochemical Society transactions.

[14]  A. Kabanov,et al.  Pluronic-modified superoxide dismutase 1 attenuates angiotensin II-induced increase in intracellular superoxide in neurons. , 2010, Free radical biology & medicine.

[15]  J. Joseph,et al.  Hydropropidine: a novel, cell-impermeant fluorogenic probe for detecting extracellular superoxide. , 2013, Free radical biology & medicine.

[16]  R. Stocker,et al.  Improved analysis of hydroethidine and 2-hydroxyethidium by HPLC and electrochemical detection. , 2007, Free Radical Biology & Medicine.

[17]  J. K. Hurst,et al.  Pathways for intracellular generation of oxidants and tyrosine nitration by a macrophage cell line. , 2007, Biochemistry.

[18]  D. Harrison,et al.  Measurement of reactive oxygen species in cardiovascular studies. , 2007, Hypertension.

[19]  R. Currie,et al.  Huntingtin inclusion bodies are iron‐dependent centers of oxidative events , 2006, The FEBS journal.

[20]  Michael S Janes,et al.  Selective fluorescent imaging of superoxide in vivo using ethidium-based probes , 2006, Proceedings of the National Academy of Sciences.

[21]  C. Georgiou,et al.  Superoxide radical detection in cells, tissues, organisms (animals, plants, insects, microorganisms) and soils , 2008, Nature Protocols.

[22]  Sonya Cunningham,et al.  1-Methyl-4-phenylpyridinium (MPP+)-induced apoptosis and mitochondrial oxidant generation: role of transferrin-receptor-dependent iron and hydrogen peroxide. , 2003, The Biochemical journal.

[23]  B. Kalyanaraman,et al.  Alterations in bioenergetic function induced by Parkinson’s disease mimetic compounds: lack of correlation with superoxide generation , 2012, Journal of neurochemistry.

[24]  U. Brunk,et al.  What does the commonly used DCF test for oxidative stress really show? , 2010, The Biochemical journal.

[25]  Michael S Janes,et al.  The selective detection of mitochondrial superoxide by live cell imaging , 2008, Nature Protocols.

[26]  B. Kalyanaraman,et al.  Hydroethidine- and MitoSOX-derived red fluorescence is not a reliable indicator of intracellular superoxide formation: another inconvenient truth. , 2010, Free radical biology & medicine.

[27]  A. Hale,et al.  Quantitative Regulation of Intracellular Endothelial Nitric-oxide Synthase (eNOS) Coupling by Both Tetrahydrobiopterin-eNOS Stoichiometry and Biopterin Redox Status , 2009, Journal of Biological Chemistry.

[28]  T. Vanden Hoek,et al.  Mitochondrial oxidant stress triggers cell death in simulated ischemia-reperfusion. , 2011, Biochimica et biophysica acta.