Show simple item record

Kinetics and thermodynamics of metal‐binding to histone deacetylase 8

dc.contributor.authorKim, Byungchulen_US
dc.contributor.authorPithadia, Amit S.en_US
dc.contributor.authorFierke, Carol A.en_US
dc.date.accessioned2015-03-05T18:24:13Z
dc.date.available2016-05-10T20:26:27Zen
dc.date.issued2015-03en_US
dc.identifier.citationKim, Byungchul; Pithadia, Amit S.; Fierke, Carol A. (2015). "Kinetics and thermodynamics of metal‐binding to histone deacetylase 8." Protein Science 24(3): 354-365.en_US
dc.identifier.issn0961-8368en_US
dc.identifier.issn1469-896Xen_US
dc.identifier.urihttps://hdl.handle.net/2027.42/110703
dc.description.abstractHistone deacetylase 8 (HDAC8) was originally classified as a Zn(II)‐dependent deacetylase on the basis of Zn(II)‐dependent HDAC8 activity in vitro and illumination of a Zn(II) bound to the active site. However, in vitro measurements demonstrated that HDAC8 has higher activity with a bound Fe(II) than Zn(II), although Fe(II)‐HDAC8 rapidly loses activity under aerobic conditions. These data suggest that in the cell HDAC8 could be activated by either Zn(II) or Fe(II). Here we detail the kinetics, thermodynamics, and selectivity of Zn(II) and Fe(II) binding to HDAC8. To this end, we have developed a fluorescence anisotropy assay using fluorescein‐labeled suberoylanilide hydroxamic acid (fl‐SAHA). fl‐SAHA binds specifically to metal‐bound HDAC8 with affinities comparable to SAHA. To measure the metal affinity of HDAC, metal binding was coupled to fl‐SAHA and assayed from the observed change in anisotropy. The metal KD values for HDAC8 are significantly different, ranging from picomolar to micromolar for Zn(II) and Fe(II), respectively. Unexpectedly, the Fe(II) and Zn(II) dissociation rate constants from HDAC8 are comparable, koff ∼0.0006 s−1, suggesting that the apparent association rate constant for Fe(II) is slow (∼3 × 103 M−1 s−1). Furthermore, monovalent cations (K+ or Na+) that bind to HDAC8 decrease the dissociation rate constant of Zn(II) by ≥100‐fold for K+ and ≥10‐fold for Na+, suggesting a possible mechanism for regulating metal exchange in vivo. The HDAC8 metal affinities are comparable to the readily exchangeable Zn(II) and Fe(II) concentrations in cells, consistent with either or both metal cofactors activating HDAC8.en_US
dc.publisherWiley Periodicals, Inc.en_US
dc.subject.othermonovalent cationsen_US
dc.subject.othermetal‐binding mechanismen_US
dc.subject.otherfl‐SAHAen_US
dc.subject.otherhistone deacetylase 8en_US
dc.titleKinetics and thermodynamics of metal‐binding to histone deacetylase 8en_US
dc.typeArticleen_US
dc.rights.robotsIndexNoFollowen_US
dc.subject.hlbsecondlevelBiological Chemistryen_US
dc.subject.hlbtoplevelHealth Sciencesen_US
dc.description.peerreviewedPeer Revieweden_US
dc.description.bitstreamurlhttp://deepblue.lib.umich.edu/bitstream/2027.42/110703/1/pro2623.pdf
dc.identifier.doi10.1002/pro.2623en_US
dc.identifier.sourceProtein Scienceen_US
dc.identifier.citedreferenceGantt SL, Joseph CG, Fierke CA ( 2010 ) Activation and inhibition of histone deacetylase 8 by monovalent cations. J Biol Chem 285: 6036 – 6043.en_US
dc.identifier.citedreferenceHurst TK, Wang D, Thompson RB, Fierke CA ( 2010 ) Carbonic anhydrase II‐based metal ion sensing: Advances and new perspectives. Biochim Biophys Acta 1804: 393 – 403.en_US
dc.identifier.citedreferenceKasner SE, Ganz MB ( 1992 ) Regulation of intracellular potassium in mesangial cells: a fluorescence analysis using the dye, PBFI. Am J Physiol 262: F462 – F467.en_US
dc.identifier.citedreferenceBorin ML, Goldman WF, Blaustein MP ( 1993 ) Intracellular free Na+ in resting and activated A7r5 vascular smooth muscle cells. Am J Physiol 264: C1513 – C1524.en_US
dc.identifier.citedreferenceTripp BC, Bell CB, 3rd, Cruz F, Krebs C, Ferry JG ( 2004 ) A role for iron in an ancient carbonic anhydrase. J Biol Chem 279: 6683 – 6687.en_US
dc.identifier.citedreferenceD'Souza V M, Holz RC ( 1999 ) The methionyl aminopeptidase from Escherichia coli can function as an iron(II) enzyme. Biochemistry 38: 11079 – 11085.en_US
dc.identifier.citedreferencePorter DJ, Austin EA ( 1993 ) Cytosine deaminase. The roles of divalent metal ions in catalysis. J Biol Chem 268: 24005 – 24011.en_US
dc.identifier.citedreferenceSeffernick JL, McTavish H, Osborne JP, de Souza ML, Sadowsky MJ, Wackett LP ( 2002 ) Atrazine chlorohydrolase from Pseudomonas sp. strain ADP is a metalloenzyme. Biochemistry 41: 14430 – 14437.en_US
dc.identifier.citedreferenceGattis SG ( 2010 ) Mechanism and metal specificity of zinc‐dependent deacetylases. PhD. Thesis. University of Michigan, pp 143 – 145.en_US
dc.identifier.citedreferenceWaldron KJ, Rutherford JC, Ford D, Robinson NJ ( 2009 ) Metalloproteins and metal sensing. Nature 460: 823 – 830.en_US
dc.identifier.citedreferenceRae TD, Schmidt PJ, Pufahl RA, Culotta VC, O'Halloran TV ( 1999 ) Undetectable intracellular free copper: the requirement of a copper chaperone for superoxide dismutase. Science 284: 805 – 808.en_US
dc.identifier.citedreferenceO'Halloran TV, Culotta VC ( 2000 ) Metallochaperones, an intracellular shuttle service for metal ions. J Biol Chem 275: 25057 – 25060.en_US
dc.identifier.citedreferenceShi H, Bencze KZ, Stemmler TL, Philpott CC ( 2008 ) A cytosolic iron chaperone that delivers iron to ferritin. Science 320: 1207 – 1210.en_US
dc.identifier.citedreferenceNandal A, Ruiz JC, Subramanian P, Ghimire‐Rijal S, Sinnamon RA, Stemmler TL, Bruick RK, Philpott CC ( 2011 ) Activation of the HIF prolyl hydroxylase by the iron chaperones PCBP1 and PCBP2. Cell Metab 14: 647 – 657.en_US
dc.identifier.citedreferenceFrey AG, Nandal A, Park JH, Smith PM, Yabe T, Ryu MS, Ghosh MC, Lee J, Rouault TA, Park MH, Philpott CC ( 2014 ) Iron chaperones PCBP1 and PCBP2 mediate the metallation of the dinuclear iron enzyme deoxyhypusine hydroxylase. Proc Natl Acad Sci USA 111: 8031 – 8036.en_US
dc.identifier.citedreferenceJoshi P, Greco TM, Guise AJ, Luo Y, Yu F, Nesvizhskii AI, Cristea IM ( 2013 ) The functional interactome landscape of the human histone deacetylase family. Mol Syst Biol 9: 1 – 21.en_US
dc.identifier.citedreferenceThompson RB, Maliwal BP, Fierke CA ( 1998 ) Determination of metal ions by fluorescence anisotropy exhibits a broad dynamic range. Adv Opt Biophys 3256: 51 – 59.en_US
dc.identifier.citedreferenceSuhling K, Siegel J, Lanigan PM, Leveque‐Fort S, Webb SE, Phillips D, Davis DM, French PM ( 2004 ) Time‐resolved fluorescence anisotropy imaging applied to live cells. Opt Lett 29: 584 – 586.en_US
dc.identifier.citedreferenceLevitt JA, Matthews DR, Ameer‐Beg SM, Suhling K ( 2009 ) Fluorescence lifetime and polarization‐resolved imaging in cell biology. Curr Opin Biotechnol 20: 28 – 36.en_US
dc.identifier.citedreferenceWang D, Hurst TK, Thompson RB, Fierke CA ( 2011 ) Genetically encoded ratiometric biosensors to measure intracellular exchangeable zinc in Escherichia coli. J Biomed Opt 16: 1 – 11.en_US
dc.identifier.citedreferenceHuang CC, Lesburg CA, Kiefer LL, Fierke CA, Christianson DW ( 1996 ) Reversal of the hydrogen bond to zinc ligand histidine‐119 dramatically diminishes catalysis and enhances metal equilibration kinetics in carbonic anhydrase II. Biochemistry 35: 3439 – 3446.en_US
dc.identifier.citedreferenceKiefer LL, Paterno SA, Fierke CA ( 1995 ) Hydrogen‐bond network in the metal‐binding site of carbonic‐anhydrase enhances zinc affinity and catalytic efficiency. J Am Chem Soc 117: 6831 – 6837.en_US
dc.identifier.citedreferenceStrahl BD, Allis CD ( 2000 ) The language of covalent histone modifications. Nature 403: 41 – 45.en_US
dc.identifier.citedreferenceChoudhary C, Kumar C, Gnad F, Nielsen ML, Rehman M, Walther TC, Olsen JV, Mann M ( 2009 ) Lysine acetylation targets protein complexes and co‐regulates major cellular functions. Science 325: 834 – 840.en_US
dc.identifier.citedreferenceWolfson NA, Pitcairn CA, Fierke CA ( 2013 ) HDAC8 substrates: histones and beyond. Biopolymers 99: 112 – 126.en_US
dc.identifier.citedreferenceKhochbin S, Verdel A, Lemercier C, Seigneurin‐Berny D ( 2001 ) Functional significance of histone deacetylase diversity. Curr Opin Genet Dev 11: 162 – 166.en_US
dc.identifier.citedreferenceGregoretti IV, Lee YM, Goodson HV ( 2004 ) Molecular evolution of the histone deacetylase family: functional implications of phylogenetic analysis. J Mol Biol 338: 17 – 31.en_US
dc.identifier.citedreferenceBlander G, Guarente L ( 2004 ) The Sir2 family of protein deacetylases. Annu Rev Biochem 73: 417 – 435.en_US
dc.identifier.citedreferenceFinnin MS, Donigian JR, Cohen A, Richon VM, Rifkind RA, Marks PA, Breslow R, Pavletich NP ( 1999 ) Structures of a histone deacetylase homologue bound to the TSA and SAHA inhibitors. Nature 401: 188 – 193.en_US
dc.identifier.citedreferenceMarek M, Kannan S, Hauser AT, Moraes Mourão M, Caby S, Cura V, Stolfa DA, Schmidtkunz K, Lancelot J, Andrade L, Renaud JP, Oliveira G, Sippl W, Jung M, Cavarelli J, Pierce RJ, Romier C ( 2013 ) Structural basis for the inhibition of histone deacetylase 8 (HDAC8), a key epigenetic player in the blood fluke Schistosoma mansoni. PLoS Pathog 9: 1 – 15.en_US
dc.identifier.citedreferenceDowling DP, Gattis SG, Fierke CA, Christianson DW ( 2010 ) Structures of metal‐substituted human histone deacetylase 8 provide mechanistic inferences on biological function. Biochemistry 49: 5048 – 5056.en_US
dc.identifier.citedreferenceLohman TM, Mascotti DP ( 1992 ) Thermodynamics of ligand‐nucleic acid interactions. Methods Enzymol 212: 400 – 424.en_US
dc.identifier.citedreferenceDowling DP, Gantt SL, Gattis SG, Fierke CA, Christianson DW ( 2008 ) Structural studies of human histone deacetylase 8 and its site‐specific variants complexed with substrate and inhibitors. Biochemistry 47: 13554 – 13563.en_US
dc.identifier.citedreferenceVannini A, Volpari C, Gallinari P, Jones P, Mattu M, Carfi A, De Francesco R, Steinkuhler C, Di Marco S ( 2007 ) Substrate binding to histone deacetylases as shown by the crystal structure of the HDAC8‐substrate complex. EMBO Rep 8: 879 – 884.en_US
dc.identifier.citedreferenceMcCall KA, Fierke CA ( 2004 ) Probing determinants of the metal ion selectivity in carbonic anhydrase using mutagenesis. Biochemistry 43: 3979 – 3986.en_US
dc.identifier.citedreferenceGantt SL, Gattis SG, Fierke CA ( 2006 ) Catalytic activity and inhibition of human histone deacetylase 8 is dependent on the identity of the active site metal ion. Biochemistry 45: 6170 – 6178.en_US
dc.identifier.citedreferenceChristianson DW, Cox JD ( 1999 ) Catalysis by metal‐activated hydroxide in zinc and manganese metalloenzymes. Annu Rev Biochem 68: 33 – 57.en_US
dc.identifier.citedreferenceAuld DS ( 2001 ) Zinc coordination sphere in biochemical zinc sites. Biometals 14: 271 – 313.en_US
dc.identifier.citedreferenceHernick M, Fierke CA ( 2005 ) Zinc hydrolases: the mechanisms of zinc‐dependent deacetylases. Arch Biochem Biophys 433: 71 – 84.en_US
dc.identifier.citedreferenceBecker A, Schlichting I, Kabsch W, Groche D, Schultz S, Wagner AF ( 1998 ) Iron center, substrate recognition and mechanism of peptide deformylase. Nat Struct Biol 5: 1053 – 1058.en_US
dc.identifier.citedreferenceZhu J, Dizin E, Hu X, Wavreille AS, Park J, Pei D ( 2003 ) S ‐Ribosylhomocysteinase (LuxS) is a mononuclear iron protein. Biochemistry 42: 4717 – 4726.en_US
dc.identifier.citedreferenceGattis SG, Hernick M, Fierke CA ( 2010 ) Active site metal ion in UDP‐3‐O‐((R)−3‐Hydroxymyristoyl)‐ N ‐acetylglucosamine deacetylase (LpxC) switches between Fe(II) and Zn(II) depending on cellular conditions. J Biol Chem 285: 33788 – 33796.en_US
dc.identifier.citedreferenceHernick M, Gattis SG, Penner‐Hahn JE, Fierke CA ( 2010 ) Activation of Escherichia coli UDP‐3‐O‐[(R)−3‐hydroxymyristoyl]‐ N ‐acetylglucosamine deacetylase by Fe2+ yields a more efficient enzyme with altered ligand affinity. Biochemistry 49: 2246 – 2255.en_US
dc.identifier.citedreferenceWegener D, Wirsching F, Riester D, Schwienhorst A ( 2003 ) A fluorogenic histone deacetylase assay well suited for high‐throughput activity screening. Chem Biol 10: 61 – 68.en_US
dc.identifier.citedreferenceMazitschek R, Patel V, Wirth DF, Clardy J ( 2008 ) Development of a fluorescence polarization based assay for histone deacetylase ligand discovery. Bioorg Med Chem Lett 18: 2809 – 2812.en_US
dc.identifier.citedreferenceSingh RK, Mandal T, Balasubramanian N, Cook G, Srivastava DK ( 2011 ) Coumarin‐suberoylanilide hydroxamic acid as a fluorescent probe for determining binding affinities and off‐rates of histone deacetylase inhibitors. Anal Biochem 408: 309 – 315.en_US
dc.identifier.citedreferenceKristoffersen AS, Erga SR, Hamre B, Frette O ( 2014 ) Testing fluorescence lifetime standards using two‐photon excitation and time‐domain instrumentation: rhodamine B, coumarin 6 and lucifer yellow. J Fluoresc 24: 1015 – 1024.en_US
dc.identifier.citedreferenceTakagai Y, Nojiri Y, Takase T, Hinze WL, Butsugan M, Igarashi S ( 2010 ) "Turn‐on" fluorescent polymeric microparticle sensors for the determination of ammonia and amines in the vapor state. Analyst 135: 1417 – 1425.en_US
dc.identifier.citedreferenceZheng H, Zhan XQ, Bian QN, Zhang XJ ( 2013 ) Advances in modifying fluorescein and rhodamine fluorophores as fluorescent chemosensors. Chem Comm 49: 429 – 447.en_US
dc.identifier.citedreferenceMartin MM, Lindqvist L ( 1975 ) pH‐dependence of fluorescein fluorescence. J Lumin 10: 381 – 390.en_US
dc.identifier.citedreferenceIrving H, Williams RJP ( 1948 ) Order of stability of metal complexes. Nature 162: 746 – 747.en_US
dc.identifier.citedreferenceOutten CE, O'Halloran TV ( 2001 ) Femtomolar sensitivity of metalloregulatory proteins controlling zinc homeostasis. Science 292: 2488 – 2492.en_US
dc.identifier.citedreferenceWang D, Hosteen O, Fierke CA ( 2012 ) ZntR‐mediated transcription of zntA responds to nanomolar intracellular free zinc. J Inorg Biochem 111: 173 – 181.en_US
dc.identifier.citedreferenceVinkenborg JL, Nicolson TJ, Bellomo EA, Koay MS, Rutter GA, Merkx M ( 2009 ) Genetically encoded FRET sensors to monitor intracellular Zn2+ homeostasis. Nat Methods 6: 737 – 740.en_US
dc.identifier.citedreferenceBozym RA, Thompson RB, Stoddard AK, Fierke CA ( 2006 ) Measuring picomolar intracellular exchangeable zinc in PC‐12 cells using a ratiometric fluorescence biosensor. ACS Chem Biol 1: 103 – 111.en_US
dc.identifier.citedreferenceQin Y, Dittmer PJ, Park JG, Jansen KB, Palmer AE ( 2011 ) Measuring steady‐state and dynamic endoplasmic reticulum and Golgi Zn2+ with genetically encoded sensors. Proc Natl Acad Sci USA 108: 7351 – 7356.en_US
dc.identifier.citedreferencePetrat F, de Groot H, Rauen U ( 2001 ) Subcellular distribution of chelatable iron: a laser scanning microscopic study in isolated hepatocytes and liver endothelial cells. Biochem J 356: 61 – 69.en_US
dc.identifier.citedreferenceMeguro R, Asano Y, Odagiri S, Li C, Iwatsuki H, Shoumura K ( 2007 ) Nonheme‐iron histochemistry for light and electron microscopy: a historical, theoretical and technical review. Arch Histol Cytol 70: 1 – 19.en_US
dc.identifier.citedreferenceEsposito BP, Epsztejn S, Breuer W, Cabantchik ZI ( 2002 ) A review of fluorescence methods for assessing labile iron in cells and biological fluids. Anal Biochem 304: 1 – 18.en_US
dc.identifier.citedreferenceMacKenzie EL, Iwasaki K, Tsuji Y ( 2008 ) Intracellular iron transport and storage: from molecular mechanisms to health implications. Antioxid Redox Signal 10: 997 – 1030.en_US
dc.identifier.citedreferenceSingh RK, Lall N, Leedahl TS, McGillivray A, Mandal T, Haldar M, Mallik S, Cook G, Srivastava DK ( 2013 ) Kinetic and thermodynamic rationale for suberoylanilide hydroxamic acid being a preferential human histone deacetylase 8 inhibitor as compared to the structurally similar ligand, trichostatin a. Biochemistry 52: 8139 – 8149.en_US
dc.identifier.citedreferencePearson RG ( 1963 ) Hard and soft acids and bases. J Am Chem Soc 85: 3533 – 3539.en_US
dc.identifier.citedreferenceFersht AR ( 2000 ) Transition‐state structure as a unifying basis in protein‐folding mechanisms: contact order, chain topology, stability, and the extended nucleus mechanism. Proc Natl Acad Sci USA 97: 1525 – 1529.en_US
dc.identifier.citedreferenceVannini A, Volpari C, Filocamo G, Casavola EC, Brunetti M, Renzoni D, Chakravarty P, Paolini C, De Francesco R, Gallinari P, Steinkühler C, Di Marco S ( 2004 ) Crystal structure of a eukaryotic zinc‐dependent histone deacetylase, human HDAC8, complexed with a hydroxamic acid inhibitor. Proc Natl Acad Sci USA 101: 15064 – 15069.en_US
dc.identifier.citedreferenceSomoza JR, Skene RJ, Katz BA, Mol C, Ho JD, Jennings AJ, Luong C, Arvai A, Buggy JJ, Chi E, Tang J, Sang BC, Verner E, Wynands R, Leahy EM, Dougan DR, Snell G, Navre M, Knuth MW, Swanson RV, McRee DE, Tari LW ( 2004 ) Structural snapshots of human HDAC8 provide insights into the class I histone deacetylases. Structure 12: 1325 – 1334.en_US
dc.identifier.citedreferenceBorin ML, Goldman WF, Blaustein MP ( 1993 ) Intracellular free Na+ in resting and activated A7r5 vascular smooth‐muscle cells. Am J Physiol 264: C1513 – C1524.en_US
dc.identifier.citedreferenceWoehl EU, Dunn MF ( 1995 ) The roles of Na+ and K+ in pyridoxal‐phosphate enzyme catalysis. Coord Chem Rev 144: 147 – 197.en_US
dc.owningcollnameInterdisciplinary and Peer-Reviewed


Files in this item

Show simple item record

Remediation of Harmful Language

The University of Michigan Library aims to describe library materials in a way that respects the people and communities who create, use, and are represented in our collections. Report harmful or offensive language in catalog records, finding aids, or elsewhere in our collections anonymously through our metadata feedback form. More information at Remediation of Harmful Language.

Accessibility

If you are unable to use this file in its current format, please select the Contact Us link and we can modify it to make it more accessible to you.