14
1521-0111/86/1/2841$25.00 http://dx.doi.org/10.1124/mol.113.090183 MOLECULAR PHARMACOLOGY Mol Pharmacol 86:2841, July 2014 Copyright ª 2014 by The American Society for Pharmacology and Experimental Therapeutics Structure, Molecular Modeling, and Function of the Novel Potassium Channel Blocker Urotoxin Isolated from the Venom of the Australian Scorpion Urodacus yaschenkoi s Karen Luna-Ramírez, Adam Bartok, Rita Restano-Cassulini, Veronica Quintero-Hernández, Fredy I. V. Coronas, Janni Christensen, Christine E. Wright, Gyorgy Panyi, and Lourival D. Possani Australian Venom Research Unit and Cardiovascular Therapeutics Unit, Department of Pharmacology and Therapeutics, University of Melbourne, Parkville, Victoria, Australia (K.L.-R., C.E.W.); Departamento de Medicina Molecular y Bioprocesos, Instituto de Biotecnología, Universidad Nacional Autónoma de México, Avenida Universidad, Cuernavaca, Mexico (R.R.-C., V.Q.-H., F.I.V.C., L.D.P.); Department of Biophysics and Cell Biology, Research Center for Molecular Medicine, University of Debrecen, Debrecen, Hungary (A.B., G.P.); MTA-DE Cell Biology and Signaling Research Group, Debrecen, Hungary (G.P.); and Department of Biochemistry, La Trobe Institute for Molecular Science, La Trobe University, Melbourne, Victoria, Australia (J.C.) Received October 8, 2013; accepted April 10, 2014 ABSTRACT This communication reports the structural and functional charac- terization of urotoxin, the first K 1 channel toxin isolated from the venom of the Australian scorpion Urodacus yaschenkoi. It is a basic peptide consisting of 37 amino acids with an amidated C-terminal residue. Urotoxin contains eight cysteines forming four disulfide bridges with sequence similarities resembling the a-potassium channel toxin 6 (a-KTx-6) subfamily of peptides; it was assigned the systematic number of a-KTx-6.21. Urotoxin is a potent blocker of human voltage-gated potassium channel (K v )1.2 channels, with an IC 50 of 160 pM, whereas its affinity for other channels tested was in the nanomolar range (hK v 1.1, IC 50 5 253 nM; hK v 1.3, IC 50 5 91 nM; and hK Ca 3.1, IC 50 5 70 nM). The toxin had no effect on hK v 1.4, hK v 1.5, human ether-à-go-gorelated gene type 1 (hERG1), or human ether-à-go-golike (hELK2) channels. Multiple sequence alignments from the venom gland transcriptome showed the existence of four other new peptides similar to urotoxin. Computer modeling of urotoxins three-dimensional structure suggests the presence of the a/b-scaffold characteristic of other scorpion toxins, although very likely forming an uncommon disulfide pairing pattern. Using molecular dynamics, a model for the binding of this peptide to human K v 1.2 and hK v 1.1 channels is presented, along with the binding of an in silico mutant urotoxin (Lys25Ala) to both channels. Urotoxin enriches our knowledge of K 1 channel toxins and, due to its high affinity for hK v 1.2 channels, it may be a good candidate for the de- velopment of pharmacologic tools to study the physiologic functions of K 1 channels or related channelopathies and for restoring axonal conduction in demyelinated axons. Introduction Scorpion venoms provide a rich source of neurotoxins that bind to ion channels. These toxins have proven to be excellent tools for the identification and classification of Na 1 ,K 1 , Ca 21 , and Cl 2 channels and their families and subfamilies, for the characterization of the tissue distribution of the ion channels and for the understanding of the role of ion channels in certain types of pathologies (reviewed in Possani et al., 1999; Gati et al., 2012). Potassium channels are ubiquitous membrane proteins found in both excitable and nonexcitable cells (Shieh et al., 2000). Ninety-two K 1 channel genes have been identified in the human genome (http://www.genenames.org/genefami- lies/KCN), which makes K 1 channels the most diverse of all ion channel families regarding primary structure and physiologic function. They are involved in the regulation of various cellular processes, including cell proliferation (Wonderlin and Strobl, 1996), apoptosis (Burg et al., 2006), hormone secretion, K 1 homeostasis, neurotransmitter release, and This work was supported by grants from the Dirección General de Asuntos del Personal Academico, UNAM [IN200113-3]; the Struan Sutherland Fund, AVRU, Department of Pharmacology and Therapeutics, University of Melbourne; a scholarship from CONACyT and from the Hugh Williamson Foundation, through the Museum Victoria; the European Union, and the State of Hungary, cofinanced by the European Social Fund in the framework of the National Excellence Program [TÁMOP 4.2.4. A/2-11-1-2012-0001] and [TÁMOP 4.2.2-A-11/1/KONV-2012-0025]; and the State of Hungary [OTKA K 75904] and [OTKA NK 101337]. The three-dimensional computer model of urotoxin and the brute force dynamic simulation was supported by a Victorian Life Sciences Computation Initiative (VLSCI) grant number [VR0064] at its Peak Computing Facility at the University of Melbourne, an initiative of the Victorian Government, Australia. dx.doi.org/10.1124/mol.113.090183. s This article has supplemental material available at molpharm. aspetjournals.org. ABBREVIATIONS: a-KTx, a-potassium channel toxin; 3D, three dimensional; GFP, green fluorescent protein; HEK, human embryonic kidney cells; hELK2, human ether-à-go-golike; hERG1, human ether-à-go-gorelated gene type 1; HPLC, high-performance liquid chromatography; MD, molecular dynamics; K v , voltage-gated potassium channel; NGS, Next Generation Sequencing; PDB, Protein Data Bank; RF, remaining current fraction; TFA, trifluoroacetic acid. 28 http://molpharm.aspetjournals.org/content/suppl/2014/04/10/mol.113.090183.DC1 Supplemental material to this article can be found at: at ASPET Journals on May 11, 2020 molpharm.aspetjournals.org Downloaded from

Structure, Molecular Modeling, and Function of the Novel ...molpharm.aspetjournals.org/Content/Molpharm/86/1/28.Full.pdfmodulation of the action potential (Bauer and Schwarz, 2001)

  • Upload
    others

  • View
    2

  • Download
    0

Embed Size (px)

Citation preview

Page 1: Structure, Molecular Modeling, and Function of the Novel ...molpharm.aspetjournals.org/Content/Molpharm/86/1/28.Full.pdfmodulation of the action potential (Bauer and Schwarz, 2001)

1521-0111/86/1/28–41$25.00 http://dx.doi.org/10.1124/mol.113.090183MOLECULAR PHARMACOLOGY Mol Pharmacol 86:28–41, July 2014Copyright ª 2014 by The American Society for Pharmacology and Experimental Therapeutics

Structure, Molecular Modeling, and Function of the NovelPotassium Channel Blocker Urotoxin Isolated from the Venomof the Australian Scorpion Urodacus yaschenkoi s

Karen Luna-Ramírez, Adam Bartok, Rita Restano-Cassulini, Veronica Quintero-Hernández,Fredy I. V. Coronas, Janni Christensen, Christine E. Wright, Gyorgy Panyi,and Lourival D. PossaniAustralian Venom Research Unit and Cardiovascular Therapeutics Unit, Department of Pharmacology and Therapeutics,University of Melbourne, Parkville, Victoria, Australia (K.L.-R., C.E.W.); Departamento de Medicina Molecular y Bioprocesos,Instituto de Biotecnología, Universidad Nacional Autónoma de México, Avenida Universidad, Cuernavaca, Mexico (R.R.-C.,V.Q.-H., F.I.V.C., L.D.P.); Department of Biophysics and Cell Biology, Research Center for Molecular Medicine, University ofDebrecen, Debrecen, Hungary (A.B., G.P.); MTA-DE Cell Biology and Signaling Research Group, Debrecen, Hungary (G.P.); andDepartment of Biochemistry, La Trobe Institute for Molecular Science, La Trobe University, Melbourne, Victoria, Australia (J.C.)

Received October 8, 2013; accepted April 10, 2014

ABSTRACTThis communication reports the structural and functional charac-terization of urotoxin, the first K1 channel toxin isolated from thevenomof the Australian scorpionUrodacus yaschenkoi. It is a basicpeptide consisting of 37 amino acids with an amidated C-terminalresidue. Urotoxin contains eight cysteines forming four disulfidebridges with sequence similarities resembling the a-potassiumchannel toxin 6 (a-KTx-6) subfamily of peptides; it was assignedthe systematic number of a-KTx-6.21. Urotoxin is a potent blockerof human voltage-gated potassium channel (Kv)1.2 channels, withan IC50 of 160 pM, whereas its affinity for other channels testedwas in the nanomolar range (hKv1.1, IC505 253 nM; hKv1.3, IC50591 nM; and hKCa3.1, IC50 5 70 nM). The toxin had no effecton hKv1.4, hKv1.5, human ether-à-go-go–related gene type 1(hERG1), or human ether-à-go-go–like (hELK2) channels. Multiple

sequence alignments from the venomgland transcriptome showedthe existence of four other new peptides similar to urotoxin.Computer modeling of urotoxin’s three-dimensional structuresuggests the presence of the a/b-scaffold characteristic of otherscorpion toxins, although very likely forming an uncommondisulfide pairing pattern. Using molecular dynamics, a model forthe binding of this peptide to human Kv1.2 and hKv1.1 channelsis presented, along with the binding of an in silico mutanturotoxin (Lys25Ala) to both channels. Urotoxin enriches ourknowledge of K1 channel toxins and, due to its high affinity forhKv1.2 channels, it may be a good candidate for the de-velopment of pharmacologic tools to study the physiologicfunctions of K1 channels or related channelopathies and forrestoring axonal conduction in demyelinated axons.

IntroductionScorpion venoms provide a rich source of neurotoxins that

bind to ion channels. These toxins have proven to be excellent

tools for the identification and classification of Na1, K1, Ca21,and Cl2 channels and their families and subfamilies, for thecharacterization of the tissue distribution of the ion channelsand for the understanding of the role of ion channels in certaintypes of pathologies (reviewed in Possani et al., 1999; Gatiet al., 2012).Potassium channels are ubiquitous membrane proteins

found in both excitable and nonexcitable cells (Shieh et al.,2000). Ninety-two K1 channel genes have been identified inthe human genome (http://www.genenames.org/genefami-lies/KCN), which makes K1 channels the most diverse ofall ion channel families regarding primary structure andphysiologic function. They are involved in the regulation ofvarious cellular processes, including cell proliferation (Wonderlinand Strobl, 1996), apoptosis (Burg et al., 2006), hormonesecretion, K1 homeostasis, neurotransmitter release, and

This work was supported by grants from the Dirección General de Asuntosdel Personal Academico, UNAM [IN200113-3]; the Struan Sutherland Fund,AVRU, Department of Pharmacology and Therapeutics, University ofMelbourne; a scholarship from CONACyT and from the Hugh WilliamsonFoundation, through the Museum Victoria; the European Union, and theState of Hungary, cofinanced by the European Social Fund in the framework ofthe National Excellence Program [TÁMOP 4.2.4. A/2-11-1-2012-0001] and[TÁMOP 4.2.2-A-11/1/KONV-2012-0025]; and the State of Hungary [OTKA K75904] and [OTKA NK 101337]. The three-dimensional computer model ofurotoxin and the brute force dynamic simulation was supported by a VictorianLife Sciences Computation Initiative (VLSCI) grant number [VR0064] at its PeakComputing Facility at the University of Melbourne, an initiative of the VictorianGovernment, Australia.

dx.doi.org/10.1124/mol.113.090183.s This article has supplemental material available at molpharm.

aspetjournals.org.

ABBREVIATIONS: a-KTx, a-potassium channel toxin; 3D, three dimensional; GFP, green fluorescent protein; HEK, human embryonic kidney cells;hELK2, human ether-à-go-go–like; hERG1, human ether-à-go-go–related gene type 1; HPLC, high-performance liquid chromatography; MD,molecular dynamics; Kv, voltage-gated potassium channel; NGS, Next Generation Sequencing; PDB, Protein Data Bank; RF, remaining currentfraction; TFA, trifluoroacetic acid.

28

http://molpharm.aspetjournals.org/content/suppl/2014/04/10/mol.113.090183.DC1Supplemental material to this article can be found at:

at ASPE

T Journals on M

ay 11, 2020m

olpharm.aspetjournals.org

Dow

nloaded from

Page 2: Structure, Molecular Modeling, and Function of the Novel ...molpharm.aspetjournals.org/Content/Molpharm/86/1/28.Full.pdfmodulation of the action potential (Bauer and Schwarz, 2001)

modulation of the action potential (Bauer and Schwarz,2001). Consequently, changes in the activity of K1 channelsdue to K1 channel gene mutations, drug actions, and/orregulation of channel functions by hormones and neuro-transmitters lead to diseases of the central nervous system,heart, kidney, and pancreas (Shieh et al., 2000; Restrepo-Angulo et al., 2010).Toxins that block K1 channels are minor components of

scorpion venom, but they are also found in venoms of otherorganisms (Legros et al., 1996). Based on to the alignment ofthe cysteine residues and other highly conserved amino acids,the K1 channel-blocking scorpion toxins have been classifiedinto four families: a-potassium channel toxin (a-KTx), b-KTx,g-KTx, and k-KTx (Tytgat et al., 1999; Rodríguez de la Vegaand Possani, 2004). The a-KTx is the best studied family; itsmembers are small basic peptides (up to 4 kDa) consisting of23 to 40 amino acids with the structure stabilized by three tofour disulfide bridges. To date, the a-KTx family is classifiedinto 27 subfamilies (http://www.uniprot.org/docs/scorpktx)(Tytgat et al., 1999; Goudet et al., 2002; Huys et al., 2004;Rodríguez de la Vega and Possani, 2004; Tan et al., 2006;Zhijian et al., 2006; Gurrola et al., 2012) with K1 channelaffinities varying in the micromolar to picomolar range. Thea-KTx toxins target mainly the voltage-gated potassium (Kv)channels, especially the Kv1 family members and some Ca21

-activated K1 channels (Tytgat et al., 1999; Rodríguez de laVega and Possani, 2004). The a-KTx toxin structures aresimilar, and their specificity and affinity depend on the aminoacid side chains situated on the external surface and thosepointing toward the contact surface with the channel(Rodríguez de la Vega et al., 2003; Jouirou et al., 2004). Thereare currently 131 known a-K1-channel toxins from scorpionvenom (Martin-Eauclaire and Bougis, 2012; http://www.uniprot.org/docs/scorpktx).This work describes the first K1-channel toxin from an

Urodacidae scorpion. The peptide was isolated from thevenom of the Urodacus yaschenkoi scorpion, and its primarystructure was determined by automatic Edman degradationand nucleotide sequencing, using a Next Generation Sequenc-ing approach. The peptide is a very potent blocker of thehKv1.2 channel, although it also inhibits hKv1.1, hKv1.3, andhKCa3.1 channels with lower potency, but does not block otherchannels included in this study: hKv1.4, hKv1.5, human ether-à-go-go–related gene type 1 (hERG1), and human ether-à-go-go–like (hELK2). The computer-modeled (Maestro software;Schrödinger, Portland, OR) three-dimensional (3D) structureof the toxin revealed an a/b-scaffold characteristic of thea-KTx scorpion toxins. Molecular dynamics simulationshowed that the interaction of urotoxin with hKv1.2 relies onthe hydrogen bonds between lysines and an arginine on thepeptide and acidic residues in the turret and the pore helix ofthe channel. Additionally, a simulation with mutant-urotoxin(Lys25Ala) indicated that Lys25 does not have the same rolein the binding of urotoxin to hKv1.2 as in other a-KTx-6toxins. Our results suggest that urotoxin may be useful asa pharmacologic tool to reveal the physiologic function ofhKv1.2 channels in in vitro and in vivo experiments, as a leadfor peptidomimetics for the targeting of hKv1.2 channels, andas an experimental therapeutic tool to restore axonalconduction in demyelinated axons, where inhibition of K1

channels has beneficial effects (Beraud et al., 2006; Shi andSun, 2011).

Materials and MethodsSpecimen Collection and Venom Extraction

U. yaschenkoi scorpions were collected in the semiarid and aridregions of Australia (near Broken Hill, New South Wales, Australia)by setting pitfalls along red sand dunes. The captured animals weremaintained in plastic boxes with water ad libitum and were fedfortnightly with crickets.

Venom was obtained in the laboratory by electrical stimulation inthe articulation of the telson, as previously described elsewhere(Luna-Ramírez et al., 2013). The venom was collected in EppendorfLo-bind tubes (Eppendorf AG, Hamburg, Germany), and then cen-trifuged at 14,000g for 15minutes at 4°C. The supernatant was pooledand finally freeze-dried and stored at 220°C until use. The proteinconcentration was determined by spectrophotometer Nanodrop(Thermo Scientific, Wilmington, DE) using the default program forproteins at l 5 280 nm.

Venom Separation

Initially, the soluble U. yaschenkoi venom was separated byreverse-phase high-performance liquid chromatography (HPLC).The stored venom was solubilized in water and spun at 10,000g for5 minutes, then 100 ml containing 3.0 mg of the soluble venomwas directly submitted to an analytical C18 reverse-phase column(250mm� 10mm) obtained from Vydac (Hisperia, CA). Elution of thevenom was made with a linear gradient of solution A [0.12%trifluoroacetic acid (TFA) in water] to 60% solution B (0.10% TFA inacetonitrile), run for 60 minutes, using a Waters 625 LC Systemcoupled with a Waters 996 Photodiode Array Detector at 230 nm ofabsorbance with 0.5 U sensitivity and eluted at 1 ml/min flow-rate(Waters Corporation, Milford, MA). Fractions were collected manu-ally every 5 minutes and finally freeze-dried using a Speed-VacSavant drier from ThermoFisher (San Jose, CA).

Based on our previous work (Luna-Ramírez et al., 2013), thefractions containingmainly components with molecular mass rangingfrom 4–7 kDa were assayed for their K1 channels inhibition pro-perties using cellular electrophysiology (patch-clamp technique). Theeffect of fractions 4 and 5 corresponding to retention times of 15:56–20:36 and 20:37–24:80 (Fig. 1) were screened against fourpotassium channels expressed heterologously: hKv1.1, hKv1.4,hERG1 (hKv11.1), and hELK2 (hKv12.2). Because fraction 4contained a putative K1 channel-blocking peptide, it was furtherpurified (as described later) and screened against eight K1 channels.

Repurification of Active Fraction

Fraction 4 was repurified using reverse-phase HPLC with a lineargradient of 5–30% of solution B (0.10% TFA in acetonitrile), run for60 minutes using the same conditions as described previously. Eachpeak was collected manually and then assayed once more for the K1

channel blocking potency using patch-clamp. This strategy permittedthe selection of highly purified peptides for further characterizationconcerning molecular weight and sequence determination.

Amino Acid Sequence Determination and MassSpectrometry Analysis

Amino acid sequence determination of the N-terminal segment ofthe purified peptide was obtained by automatic Edman degradationinto a protein sequencer PPSQ-31A Shimadzu (Kyoto, Japan) usingthe chemicals and procedures recommended by the provider.

The molecular weight determination of pure urotoxin was per-formed by liquid chromatography-electrospray ionization-ion trapmass spectrometry with a Thermo Electron/Finningan LCQ Ion TrapMass spectrometer (San Jose, CA). The sample (0.1–0.5 mg/ml) wasdissolved in 50% acetonitrile with 0.1% acetic acid and directlyapplied into the liquid chromatography-mass spectrometry system asdescribed earlier by our group (Batista et al., 2007).

Characterization of the Potassium Channel Blocker Urotoxin 29

at ASPE

T Journals on M

ay 11, 2020m

olpharm.aspetjournals.org

Dow

nloaded from

Page 3: Structure, Molecular Modeling, and Function of the Novel ...molpharm.aspetjournals.org/Content/Molpharm/86/1/28.Full.pdfmodulation of the action potential (Bauer and Schwarz, 2001)

Sequence Elucidation by Whole Transcriptome Using NextGeneration Sequencing

An mRNA-Seq library was generated from the venom gland ofa carefully identified specimen ofU. yaschenkoi according to Illumina’ssample preparation instructions (http://supportres.illumina.com/docu-ments/documentation/chemistry_documentation/samplepreps_truseq/truseqrna/truseq_rna_sampleprep_guide_15008136_a.pdf ). The purifiedcDNA library was used for cluster generation on Illumina’s ClusterStation and then sequenced on Illumina HiSeq 2000 following thevendor’s instructions. Typically, a paired-end sequencing run with a 101nucleotides (nt) read length was used. For each run, RNA-Seq read wasperformed. Raw sequencing intensities were then extracted, and thebases were called using Illumina’s RTA software, followed by sequencequality filtering. The extracted sequencing reads were saved as a pair offastq files for the first and second read, respectively. All raw readsgenerated from the sequencer were de novo assembled into contigs usingthe Trinity program (Grabherr et al., 2011). The information concerningproteins of scorpions was collected from the US National Center forBiotechnology Information nonredundant (nr) database. These annotatedproteinswere aligned to the assembled contigs to identify the homologousgenes in U. yaschenkoi using TBLASTN (E-value , 0.1). The nucleotidesequence of urotoxin was obtained from this library by comparisonwith the amino acid sequence obtained by Edman degradation and the

theoreticalmolecularweight of this sequencewith that directlymeasuredby mass spectrometry.

Electrophysiology

Cell Culture. Cell lines were cultured in Dulbecco’s modifiedEagle medium supplemented with 10% fetal bovine serum andmaintained at 37°C in 5% CO2 atmosphere and 95% humidityatmosphere. For transient expression of hKv1.1 and hKv1.4, Chinesehamster ovary cells were cotransfected with plasmids of interestalong with plasmid for green fluorescent protein (GFP) by usingLipofectamine (Life Technologies, Carlsbad, CA). Transient expres-sion of hKv1.2: Cos7 cells were transfected with pCMV6 vectorcontaining the GFP-tagged hKv1.2 gene (OriGene Technologies,Rockville, MD) using Lipofectamine 2000 reagent (Life Technologies).Human embryonic kidney (HEK) tsA201 cells were transientlytransfected with the hKCa3.1 gene in pEGFP-C1 (gift of Dr. HeikeWulff, University of California, Davis, CA) or with the hKv1.5 gene inpYFP-C1 plasmid (gift of Dr. Antonio Felipe, University of Barcelona,Barcelona, Spain) using Lipofectamine 2000. Cells were culturedunder standard conditions. Currents were measured 1 to 2 days aftertransfection.

The hERG1 and hELK2 channels were stably expressed in Chinesehamster ovary and HEK cells, respectively (Redaelli et al., 2010). The

Fig. 1. Purification of urotoxin. HPLC sep-aration of 3 mg of soluble venom fromU. yaschenkoi in a C18 reversed-phasecolumn equilibrated with solution A (waterin 0.12% TFA), using a gradient from 0 to60% solution B (acetonitrile in 0.10% TFA)over 60 minutes. (Inset) HPLC repurifi-cation of urotoxin (peak marked with anarrow) was made with a linear gradient of5–30% of solution B (0.10% TFA in acetoni-trile) over 60 minutes. Inset shows pureurotoxin.

30 Luna-Ramírez et al.

at ASPE

T Journals on M

ay 11, 2020m

olpharm.aspetjournals.org

Dow

nloaded from

Page 4: Structure, Molecular Modeling, and Function of the Novel ...molpharm.aspetjournals.org/Content/Molpharm/86/1/28.Full.pdfmodulation of the action potential (Bauer and Schwarz, 2001)

hKv1.1, hKv1.4, hERG1 plasmids and hELK2 cell line were a gift fromE. Wanke, Universitá di Milano Bicocca, Italy.

Kv1.3 currents were measured in human peripheral T lymphocytesisolated from healthy male volunteers. Mononuclear cells were se-parated by Ficoll-Hypaque density gradient centrifugation. Collectedcells were washed twice with Ca21- and Mg21-free Hanks’ balancedsalt solution containing 25 mM HEPES buffer, pH 7.4. Cells werecultured in a 5% CO2 incubator at 37°C in 24-well culture platesin RPMI 1640 medium supplemented with 10% fetal calf serum,100 mg/ml penicillin, 100 mg/ml streptomycin, and 2 mM L-glutamineat 0.5 � 106/ml density for 2 to 6 days. Lymphocytes were activatedwith 2.5, 5, or 10 mg/ml phytohemagglutinin A (Sigma-Aldrich Kft,Budapest, Hungary) in the culturing medium.

Solutions. For recording hKv1.1 and hKv1.4 currents, the stan-dard extracellular solution contained (mM): 130 NaCl, 5 KCl, 2 CaCl2,2 MgCl2, 10 HEPES, 5 D-glucose, pH 7.40, whereas the pipettesolution contained (in mM): 130 K1-aspartate, 10 NaCl, 2 MgCl2, 10EGTA-KOH, 10 HEPES-KOH at pH 7.3 and nominal free Ca21

concentration of 50 nM (Redaelli et al., 2010). For recording hKv1.2,hKv1.3, hKv1.5, and hKCa3.1, the extracellular (bath) solutioncontained (in mM): 145 NaCl, 5 KCl, 1 MgCl2, 2.5 CaCl2, 5.5 glucose,10 HEPES, and 0.1 mg/ml bovine serum albumin, pH 7.35. For therecordings of hKv1.2, hKv1.3, and hKv1.5 currents, the pipette-fillingsolution contained (in mM) 140 KF, 2 MgCl2, 1 CaCl2, 10 HEPES, and11 EGTA, pH 7.2; for the recording of hKCa3.1 currents, it contained150 K-aspartate, 5 HEPES, 10 EGTA, 8.7 CaCl2, 2 MgCl2, pH 7.2.This latter pipette-filling solution contained 1 mM free Ca21

concentration to fully activate the KCa3.1 current (Grissmer et al.,1993).

The hERG1 and hELK2 currents were recorded in a high K1

extracellular solution ([K1]o 5 40 mM) where NaCl was replaced byan equimolar amount of KCl and using the K1-aspartate–basedpipette-filling solution (see earlier). This experimental conditionprovided the best signal-to-noise relation for hERG1 and hELK2channels. Due to the low amount of material in fraction 4 of thevenom, the high K1 extracellular solution was also used to analyzethe effect of the venom fractions on the hKv1.1 and hKv1.4 currents.Toxin fractions from concentrated stocks in distilled water werediluted directly into the extracellular solution. Osmolarity of theextracellular solution was between 302 and 308 mOsM, and that ofthe intracellular solutions were ∼295 mOsM.

Patch-Clamp Recordings and Data Analysis. Currents weremeasured using whole-cell or outside-out patch configuration involtage clamp mode using Axopatch 200A and MultiClamp 700Bamplifiers and Digidata 1200/1440A digitizers. For data analysis, thepClamp9/10 software package was used. For recording of hKv1.2,hKv1.3, and hKv1.5 currents, the cells were held at 2100 mV holdingpotential and depolarized to150 mV for 15 milliseconds (hKv1.3), 200milliseconds (hKv1.2), or 40 milliseconds (hKv1.5) every 15 seconds tofully activate the currents. For recording of hKv1.1 and hKv1.4currents, cells were held at290mV holding potential and depolarizedat 160 mV for 200 or 300 milliseconds followed by a repolarizing stepat –50 mV (100 milliseconds). The time between pulses was 2.7seconds. For recording of hERG1 tail currents, the cells were held at280 mV, depolarized at 160 mV for 500 milliseconds to activate thecurrent, and themembrane potential was stepped back at2120mV torecord the inward tail current, pulses were delivered every 5 seconds.The hELK2 currents were evoked by voltage steps to2120mV for 500milliseconds from a holding potential of 130 mV. The hKCa3.1currents were evoked by 150-millisecond voltage ramps from 2120 to140 mV every 10 seconds from a holding potential of 2100 mV.

Glass micropipettes were pulled from GC 150 F-15 borosilicatecapillaries with a resistance of 3–5 MV in the bath solution. Whennecessary, 80–90% of cell capacitance and series resistance errors werecompensated for prior to each voltage clamp protocol to decrease thevoltage errors to less than 5% of the protocol pulse. The effect of thetoxin in a given concentration is displayed as remaining currentfraction (RF 5 I/I0, where I is the current amplitude measured in the

presence of the toxin upon reaching block equilibrium, I0 is the currentamplitude measured in the toxin-free control bath solution). Datapoints of the concentration-response curves are averages of three to fiveindependent measurements where the error bars represent the S.E.M.The two parameter Hill equation was fitted on the points

RF 5 IC50H=ðIC50

H 1 ½Tx�H;

where IC50 is the half-maximum inhibition dose, H is the Hillcoefficient, and [Tx] is the toxin concentration. The limited amount ofnatural urotoxin did not allow determination of the IC50 in thenanomolar range for Kv1.3 and KCa3.1, so we used the Lineweaver–Burk analysis, where 1/RF calculated from three to five independentmeasurements was plotted as a function of toxin concentration anda straight line was fitted to the points, where IC50 5 1/slope.

Modeling

Homology Model of Urotoxin. The primary structure ofurotoxin shares high sequence similarity with a-KTx-6 family mem-bers containing the toxin 2 superfamily region (according to BlastP)and the eight cysteine-binding motif. To model its 3D structure, weused a template based on the most similar toxin to urotoxin that hasa crystal or solution nuclear magnetic resonance structure. A searchfor similar sequences using the BLAST program against the ProteinData Bank Proteins (PDB) database revealed 57.9% identity with thesequence of spinoxin (PDB ID 1V56), whose 3D structure has beensolved. The homologymodel wasmadewithMaestro software (version9.3; Schrödinger, New York, NY), and the amidation of the lastresidue was made using the VMD psfgen plugin (http://www.ks.uiuc.edu/Research/vmd/).

Molecular Dynamics Simulation of Urotoxin Binding tohKv1.1 and hKv1.2 Channels. Homology models of the poredomains of hKv1.1 and hKv1.2 channels from crystal structure of ratKv1.2 (PDB ID 3LUT) (Chen et al., 2010) were constructed using themethods of Chen et al. (2011). For hKv1.1, sequence gi|119395748was used and for hKv1.2 gi|4826782. Molecular dynamics simulationof urotoxin binding to the hKv1.1 and hKv1.2 homology models wasmade using Maestro software in the super computer AVOCA (IBMBlue Gene/Q; University of Melbourne). Channel-toxin complexeswere refined with molecular dynamics (MD) simulations to identifythe interacting residues of urotoxin with hKv1.1 and hKv1.2 channels.Each channel was embedded in a 1-palmitoyl-2-oleoyl-sn-glycero-3-phosphocholine (POPC) bilayer (�80 lipids/leaflet) and a box ofexplicit TIP3P water (�87,000 molecules). Approximately 12 K1,88 Cl2, and 74 Na1 ions were added, giving an overall cationconcentration of 0.2 M. First, the system was equilibrated for 25nanoseconds where the C-a atoms in the toxin and the whole channelwere kept rigid to ensure that the membrane had good contact withthe protein channel. Subsequently, a 5-nanosecond equilibration withno restraints was performed. Finally, two sets of 10 � 50-nanosecondbrute force MD simulations were performed at 310 K using NAMDv2.9 (http://www.ks.uiuc.edu/Research/namd/). Recognition residuesand interaction contacts for the binding were identified during thistime of simulation. The CHARMM27 force field was used. Visual-izations of molecules are in VMD software.

In Silico Generation of the Lys25Ala Mutant Urotoxin andMD Simulation. Residue lysine-25 was mutated in silico to alanine(Lys25Ala) with the Mutator plugin from VMD and the 500-nanosecond brute force dynamics simulation was performed as de-scribed earlier.

ResultsIsolation of Urotoxin and Molecular Weight Determination

A sample containing 3.0 mg of protein from the solublevenom of the scorpion U. yaschenkoi was routinely separatedby HPLC, as previously shown elsewhere (Luna-Ramírez

Characterization of the Potassium Channel Blocker Urotoxin 31

at ASPE

T Journals on M

ay 11, 2020m

olpharm.aspetjournals.org

Dow

nloaded from

Page 5: Structure, Molecular Modeling, and Function of the Novel ...molpharm.aspetjournals.org/Content/Molpharm/86/1/28.Full.pdfmodulation of the action potential (Bauer and Schwarz, 2001)

et al., 2013). The fraction collection from the HPLC wasperformed every 5 minutes of elution time to reduce thenumber of fractions to be screened in electrophysiology.Figure 1 shows the HPLC separation of the soluble venom,where the peak at minute 18 (within fraction 4) indicates theelution of the peptide described below. Themain component offraction 4 (labeled with an arrow) was collected separatelyand finally obtained in pure form (see Fig. 1 inset), usinga gradient of 5–30% solution B [0.10% TFA in acetonitrile]),over 60 minutes. The pure component eluted at 27.3 minutes(see Fig. 1 inset) and was homogeneous and shown byelectrospray ionization-mass spectrometry to contain a pep-tide with molecular mass of 4012.75 Da, corresponding ad-equately to the mass fingerprinting made previously by ourgroup (Luna-Ramírez et al., 2013). This peptide is referred toas urotoxin throughout the paper.

Sequence Elucidation

Urotoxin was sequenced by automatic Edman degradationgiving a partial N-terminal amino acid sequence fromresidues number 1 up to 19 (GDIKXSGTRQXWGPXKKQT).Because the peptide was not previously reduced and alkylatedfor Edman degradation, the X residues were assumed to becysteines, which was later confirmed using nucleotidesequence analysis. Using Next Generation Sequencing(NGS) analysis, we unveiled the full sequence of this toxin(Supplemental Fig. 1). From the NGS study, four additionalnew peptides similar to urotoxin were detected (isotigs),whichmay be isoforms or bona fide similar peptides present inthe venom. Alternatively, different genes may encode thesame peptide, which has to be confirmed a posteriori. Thesefive similar transcripts were analyzed as previously describedelsewhere (Luna-Ramírez et al., 2013) to obtain the maturepeptide (Fig. 2). Briefly, with the sequences provided by NGSanalysis having hits with K1 channel toxins, the ExPASyTranslate tool (http://web.expasy.org/translate/) was used tofind the open reading frame; later Blastp (http://blast.ncbi.nlm.nih.gov/Blast.cgi) was used to confirm that the translatedsequence shares homology with K1 channel toxins, andfinally, the mature peptide was found with SignalP (http://www.cbs.dtu.dk/services/SignalP/).The experimental molecular weight determined by mass

spectrometry of the purified peptide, showed a molecularmass of 4012.75 Da. The translated sequence gave a theoret-ical expected molecular mass of 4070.7 Da for the foldedpeptide, meaning there was a difference of 57.95 Da, whichcorresponds to the molecular mass of glycine. This is con-sistent with the C-terminal amidation of urotoxin. Accordingto the known processes that occur during expression ofscorpion toxins (Becerril et al., 1993), the last residue of thepeptide (glycine) is eliminated during the process of matura-tion, and the amino group of the glycine is used for amidation

of the previous residue (resulting in this case in valinamide).Thus, we conclude that the complete amino acid sequence ofmature urotoxin is GDIKCSGTRQCWGPCKKQTTCTNSKC-MNGKCKCYGCV*(G), where * means that valine (V) isamidated. The sequence analysis indicates that urotoxinbelongs to the a-KTX-6 subfamily. Toxins from this subfamilyshare important features in their primary structure such asa central lysine (K25 or K23 in other toxins) that is criticalfor inhibition of Shaker channels, an essential dyad KC-N atpositions 25–28, and the presence of eight conservedcysteines to stabilize the 3D structure of the toxins bydisulfide bonds.

Pharmacologic Effects of Urotoxin

The in vitro pharmacologic effects of urotoxin were de-termined on eight K1 channels using electrophysiology. Amongthese, five channels were members of the voltage-gated Shakerfamily (hKv1.1, hKv1.2, hKv1.3, hKv1.4, and hKv1.5); otherchannels included in the study were hERG1 and hELK2 andthe intermediate conductance calcium-activated potassiumchannel hKCa3.1. The initial assays were conducted using10 nM toxin concentrations. When positive results were found,additional concentrations were assayed ideally up to 1 mM.Lack of available material limited the use of higher concen-trations in several assays as well as the broadening of theselectivity studies over a wider range of ion channels.First, fraction 4 (elution time 15 to 20 minutes; Fig. 1) was

tested against hKv1.1, hKv1.4, hERG1, and hELK2. Figure 3shows the electrophysiologic recordings obtained upon appli-cation of 50 mg/ml protein to hKv1.1, hKv1.4, hERG1, andhELK2 channels. Because fraction 4 contains several compo-nents, we cannot specify the exact molar concentration used.This fraction induced an almost complete inhibition of thehKv1.1 current, whereas no effect was observed at identicaltoxin concentrations on other potassium channel typesincluded in the screening (Fig. 3).Upon purification of urotoxin to homogeneity (Fig. 1), it was

assayed against the hKv1.1, hKv1.2, hKv1.3, hKv1.5, andhKCa3.1 channels to provide more information about itsselectivity. Electrophysiologic recordings (Fig. 4A) show that1 mM of pure urotoxin inhibited hKv1.1 channels. Theconcentration-response curve (remaining current fractionrecorded after application of different toxin concentrations)is plotted in Fig. 4B, where the solid line is the best fitresulting in an IC50 and Hill coefficient (H) of 253 nM and 1.1,respectively. The inhibition of the hKv1.1 current in thepresence of 1 mM urotoxin was fast and reversible (Fig. 4C).As the hKv1.1 channel is closely related to hKv1.2 and

hKv1.3 (Wang et al., 1994; Xie et al., 2010; Robbins andTempel, 2012), urotoxin was additionally tested in the lattertwo channels. Urotoxin, at 10 nM, fully blocked the whole cellhKv1.2 current (Fig. 5A), and the block was fully reversible

Fig. 2. Multiple alignment of urotoxin with similarpeptides found in the Next Generation Sequencinganalysis. The amino acid sequence of urotoxin wascompared with similar sequences resulting from thewhole transcriptome analysis made with the Illuminaplatform. The percentage of identity (% I) between thesesequences is shown; for each peptide, the cysteines arehighlighted in gray.

32 Luna-Ramírez et al.

at ASPE

T Journals on M

ay 11, 2020m

olpharm.aspetjournals.org

Dow

nloaded from

Page 6: Structure, Molecular Modeling, and Function of the Novel ...molpharm.aspetjournals.org/Content/Molpharm/86/1/28.Full.pdfmodulation of the action potential (Bauer and Schwarz, 2001)

when the recording chamber was washed using toxin-freesolution (Fig. 5C). The concentration-response relationship ofthe block of hKv1.2 channels gave an IC50 of 160 pM (H5 1.1)(Fig. 5E). Figure 5, B and D, shows that 10 nM urotoxinquickly and reversibly inhibits the hKv1.3 current with anIC50 of 91 nM (Fig. 5F). These data mean that urotoxin is∼560-fold selective for hKv1.2 over hKv1.3 and ∼1600-foldwhen compared with hKv1.1. Based on the classification ofGiangiacomo et al. (2004), urotoxin is selective for the hKv1.2over the hKv1.1 and hKv1.3 channels. Due to its uniquebinding geometry to K1 channels (see below), whether urotoxininhibits the toxin-resistant hKv1.5 channel was also tested.Figure 6D shows that 10 nM urotoxin does not inhibit the hKv1.5channel (RF 5 1.0, S.E.M. 5 0.003, N 5 5).Finally, the blocking potency of urotoxin was tested on

hKCa3.1 channels for two reasons. First, several peptides that

inhibit Kv1.2 also block this channel, such as maurotoxin(Regaya et al., 2004; Wulff and Castle, 2010) and charybdo-toxin (Grissmer et al., 1994). Second, because urotoxin in-hibits hKv1.3, and the selectivity of a peptide for hKv1.3 overthe other K1 channel of human T cells, hKCa3.1, is of greatinterest due to the potential therapeutic application of thesepeptides.Figure 6A shows that urotoxin inhibits the hKCa3.1

current with a lower potency than charybdotoxin; thelatter peptide was used as a positive control during therecording of the hKCa3.1 current in HEK cells by use ofa pipette-filling solution allowing full activation of thechannels (see Materials and Methods). Figure 6B showsthat 10 nM urotoxin reversibly inhibits the hKCa3.1current. The Lineweaver–Burk analysis (Fig. 6C) resulted inan IC50 5 70 nM, which is two orders of magnitude higher

Fig. 3. Fraction 4 of the whole venom induced almost complete inhibition of the hKv1.1 voltage-gated K+ current (A), but has no effect on the hKv1.4 (B),hERG1 (C), or hELK2 (D) K+ currents. Currents were recorded in the absence (control) or in the presence of 50 mg/ml of fraction 4 after the blockequilibrium (fraction 4). High potassium extracellular solution was used in the hERG1 and hELK2 currents determination. (A) The whole-cell hKv1.1currents were elicited in Chinese hamster ovary (CHO) cells by a voltage pulse to +60 mV (300 milliseconds), followed by a repolarizing step to250 mV(100 milliseconds) from a holding potential of 290 mV. The time between successive pulses was 2.7 seconds. (B) The whole-cell hKv1.4 currents wereelicited using a similar protocol, by a depolarization step to +40 mV (300milliseconds); the time between pulses was 5 seconds. (C) The whole-cell hERG1tail currents were recorded in CHO cells at 2120 mV (450 milliseconds), preceded by a depolarizing step to +60 mV (500 milliseconds) from a holdingpotential of 280 mV. The time between successive pulses was 5 seconds. A 250-millisecond segment of the inward tail currents is shown for clarity (D).The hELK2 currents were evoked by voltage steps to2120 mV for 500 milliseconds from a holding potential of +30 mV. The 300-millisecond segments ofthe inward tail currents at 2120 mV are shown for clarity.

Characterization of the Potassium Channel Blocker Urotoxin 33

at ASPE

T Journals on M

ay 11, 2020m

olpharm.aspetjournals.org

Dow

nloaded from

Page 7: Structure, Molecular Modeling, and Function of the Novel ...molpharm.aspetjournals.org/Content/Molpharm/86/1/28.Full.pdfmodulation of the action potential (Bauer and Schwarz, 2001)

than the IC50 value for hKv1.2, confirming the preference ofurotoxin for the hKv1.2 channel and indicating similarpotential of the toxin for inhibiting hKv1.3 and hKCa3.1channels.Analyzing the kinetics of the development and the relief of

the inhibition equilibrium, we could determine the Kd valuesof urotoxin for hKv1.1, hKv1.2, and hKCa3.1 channels accordingto the following formula: Kd 5 koff/kon, where koff is (TOFF)

21,kon is ((TON)

21 2 (TOFF)21)/[Tx], [Tx] is the toxin concentra-

tion, and TOFF and TON are the time constants for thedevelopment and relief of the inhibition at a given toxinconcentration (Goldstein and Miller, 1993), respectively. Forfurther details see Supplemental Fig. 7. The calculated Kd

values were 468 nM, 190 pM, and 109 nM for the hKv1.1,hKv1.2, and hKCa3.1 channels, respectively, in good agree-ment with the IC50 values obtained from the inhibitionequilibrium. For Kv1.3, the kinetics of association anddissociation of urotoxin were too fast to be resolved due

to the inherent properties of Kv1.3 (i.e., cumulativeinactivation).

Modeling of the 3D Structure of Urotoxin

Urotoxin shares high sequence similarity with the a-KTx-6subfamily of scorpion toxins. Maurotoxin (a-KTx-6.2) andspinoxin (a-KTx-6.13) are known members of this subfamilywith which urotoxin shares 50 and 57.9% identity, re-spectively (Fig. 7). The 3D solution structures of maurotoxinand spinoxin are known (Kharrat et al., 1996; Kobayashiet al., 2003, http://www.ebi.ac.uk/pdbe-srv/view/entry/1v56/summary.html); of these, spinoxin was chosen to make thehomology modeling of urotoxin. Schrödinger’s Maestrosoftware was used for this purpose. Urotoxin is 37 residueslong, although it contains three more amino acid residuesthan spinoxin (two at the beginning and one more at theend), but the cysteines are conserved in correspondingpositions (Fig. 7). Figure 8 shows the modeled tertiary

Fig. 4. Block of hKv1.1 current by purified urotoxin. (A) Urotoxin blocks hKv1.1 currents expressed in Chinese hamster ovary cells. Currents wereelicited in a whole-cell patch-clamped cell using depolarizing pulses and recording the conditions as those described in the legend of Fig. 3. Current traceswere recorded in toxin-free solution (control) and at equilibrium block in the presence of urotoxin (1 mM urotoxin). (B) Concentration–responserelationship curve of urotoxin for hKv1.1 channels. The remaining current fraction (RF, see Materials and Methods) was determined at various toxinconcentrations (n $ 3 independent determinations at each toxin concentration). Data were plotted against concentration and fitted by using a Hillequation giving the following parameters: IC50 = 253 nM and H = 1.1. Error bars indicate S.E.M. (C) Time course of the development and the relief of theblock of macroscopic hKv1.1 currents by urotoxin. The gray bar indicates the application of 1 mM urotoxin. Peak currents were determined from tracesshown in A during repeated depolarizations and plotted as a function of time.

34 Luna-Ramírez et al.

at ASPE

T Journals on M

ay 11, 2020m

olpharm.aspetjournals.org

Dow

nloaded from

Page 8: Structure, Molecular Modeling, and Function of the Novel ...molpharm.aspetjournals.org/Content/Molpharm/86/1/28.Full.pdfmodulation of the action potential (Bauer and Schwarz, 2001)

Fig. 5. Reversible block of hKv1.2 and hKv1.3 currents by 10 nM urotoxin. Urotoxin inhibits hKv1.2, but not hKv1.3, with high affinity. (A) The hKv1.2currents were measured in a voltage-clamped COS7 cell transiently expressing the channel. Currents were evoked by 200-millisecond depolarizingpulses from a holding potential of2100 to +50mV every 15 seconds. (B) hKv1.3 currents were recorded in a voltage-clamped activated human peripherallymphocyte expressing endogenous hKv1.3 channels. The channels were activated by 15-millisecond depolarization pulses from a holding potential of2100 to +50 mV every 15 seconds. Representative traces show the K+ currents in the absence (control and wash) and in the presence of 10 nM urotoxinapplied in the extracellular bath solution, upon equilibration of the block. (C and D) Time course of the development and the relief of the block ofmacroscopic hKv1.2 (C) and hKv1.3 (D) currents by 10 nM urotoxin. Gray bars indicate the perfusion of the recording chamber with 10 nM urotoxin in thebath solution. Peak currents were determined from traces shown in A and B during repeated depolarizations (see details in A and B) and plotted asa function of time. (E) Concentration-response of the inhibition of hKv1.2 channels. Remaining current fraction (RF = I/I0, where I is the peak current inthe presence of the toxin in a given concentration and I0 is the peak current measured in the control solution) is plotted as a function of urotoxin

Characterization of the Potassium Channel Blocker Urotoxin 35

at ASPE

T Journals on M

ay 11, 2020m

olpharm.aspetjournals.org

Dow

nloaded from

Page 9: Structure, Molecular Modeling, and Function of the Novel ...molpharm.aspetjournals.org/Content/Molpharm/86/1/28.Full.pdfmodulation of the action potential (Bauer and Schwarz, 2001)

structure of urotoxin. It has one a-helix (residue T8-T19) andtwo b-sheets (residues S24-C26 and Y31-C33, respectively).Residues G1-I3 are random coil (gray); K4 is an isolatedbridge (ochre); C5-G7 turn (blue-green); T8-T19 a-helix(purple); T20 coil (gray); C21-N23 turn (blue-green); S24-C26 b-sheet (yellow); M27-G29 turn (blue-green); K30isolated bridge (ochre); C31-C33 b-sheet (yellow); Y34-C363-10-helix (blue), and V37 unarranged or coil (gray). The bestfitting for the homology model was obtained with disulfidebridges between C5-C26, C11-C31, C15-C21, and C33-C36,which resembles the cysteine pairing pattern of maurotoxin(C3-C24; C9-C29; C13-C19; C31-C34; Kharrat et al., 1996,1997; Rochat et al., 1998).Intensive work was performed for the experimental de-

termination of the disulfide pairing of urotoxin. Unfortunatelythe native toxin was extremely resistant to proteolyticdigestion, a condition necessary to open the folded structureand produce peptide fragments that would facilitate identi-fication of the disulfide arrangement. Several aliquots ofnative urotoxin were separately treated for digestion withtrypsin, a mixture of trypsin and chymotrypsin, and bothenzymes in the presence of low amounts of solvent (acetoni-trile) for 24 hours at 37°C. Selective reduction was alsoperformed before proteolytic digestion, but then the frag-ments obtained were not sufficiently clear to obtain thedisulfide pairing.

Molecular Dynamics Simulations

Binding of Urotoxin to the hKv1.2 Channel. From MDsimulation, the key residues required for binding to thehKv1.2 channel were identified (Supplemental Data: hKv1.2).The average position of the toxin throughout the simulation isabove the pore, possibly preventing the ion conduction (Fig. 9).The toxin binds to the channel mainly in the turret region andon fewer occasions in the pore domain through lysines (K) andarginine (R) interacting strongly with the acidic residuesaspartate (D) and glutamate (E) of the channel.Supplemental Figure 2 shows the bound state of the toxin-

channel complex. Some of the most recurrent toxin-channelresidue pairs that form H-bonds are shown in SupplementalFig. 3. In detail, Supplemental Fig. 3A shows urotoxin boundto the channel through K17, forming H-bonds with D363 andD379 at the same time; in fact, it is forming a hydrogennetwork with one residue of the pore region and the other onefrom the extended turret region. Supplemental Figure 3B showsanother example of binding throughout the simulation whereK17-E355 form an H-bond; on the other end of urotoxin’sa-helix, at the same time, R9 forms an H-network with D363and D379. These bindings occur on one side, in the turretregion and on the other side, in the pore helix of the channel.The interaction of K17 and E355 is dynamic: the H bond formsand breaks during the whole simulation, thereby allowingK17 to form an H-network with D363 and D379 as well.Another frequent residue pair that was observed is K30-E353and K16-D379. These H-bonds also form and break during the

simulation, indicating the active nature of the toxin-channelinteraction. In all cases, the toxin residues were found to actas hydrogen donors.The results suggest that the toxin maximizes its electro-

static interactions with the peripheral acidic residues of thechannel. Data are in agreement with the surface representa-tion of the toxin and channel, showing the toxin in blue(positive) and the channel in red (negative) electrostaticallypredisposed to interact (Supplemental Fig. 4). The net chargeof the hKv1.2 channel surface is216 whereas that of urotoxinis 17. It is important to remember that this MD simulationuses brute force dynamics instead of docking, so the toxin andthe channel conserve their flexibility and are free to moveduring the simulation.Analysis of the toxin-channel interactions shown above

reveals that the toxin binds to the channel primarily, but notexclusively, via two types of basic residues, lysines (positions16, 17, and 30) and one arginine (position 9), to glutamate andaspartate in the channel.The lysine residue at position 25 in urotoxin (or position 23

in spinoxin and maurotoxin, as well as in certain other toxins)is conserved across all a-KTx-6 scorpion toxins. Severaltheoretical docking studies have suggested that this lysineresidue protrudes into the channel selectivity filter (Yi et al.,2008; Chen et al., 2011; Chen and Chung, 2012), but in thiscase K16, K17, and R9 were the most active residues, followedby K30, K25, K32, and K4, in forming hydrogen bonds withE355, E353, and D379, and D363 in the turret and pore helixregion, respectively (Supplemental Table 1). A hydrogen bondis considered formed if the donor and acceptor atoms (nitrogenor oxygen) are within 3.2 Å of each other.All the lysines and arginine-9 in the toxin form salt bridges

with D363, D379, E353, and E355. K25 forms a salt bridgeonly with E355 or E353 (Supplemental Table 2). A saltbridge is considered formed if the distance is less than 3.2 Åbetween a side chain oxygen atom from an acidic residue anda nitrogen atom from a basic residue.MD Simulation of the Binding of Urotoxin to the

hKv1.1 Channel. During the 500-nanosecond simulation(Supplemental Data hKv1.1), K25 and K30 from urotoxinwere the most common residues involved in forming hydrogenbonds and salt bridges with D377, E351, and E353 of hKv1.1;none of these residues are in the pore helix domain. Otherresidues involved in the binding were K4, K16, and R9 fromthe toxin and D377, Y375, and Y379 from the channel(Supplemental Fig. 5; Supplemental Tables 3 and 4). Thesebonds alternate during the simulation. Based on these data,an apparently weaker interaction of urotoxin is predicted withthe hKv1.1 channel as compared with the hKv1.2 channel(5 times fewer bonds), which agrees well with the experimen-tal data where a higher affinity binding of urotoxin to hKv1.2was observed (see Figs. 4 and 5).Binding of Mutant Urotoxin (Lys25Ala) to hKv1.1 and

hKv1.2. As mentioned before, within the a-KTx subfamily,K23 has been pointed out as a key residue for the binding of

concentration. Peak currents were determined at equilibrium block from experiments shown in A. Fitting the Hill equation to the data points yieldedIC50 = 160 pM and H = 1.1. The error bars indicate the S.E.M. of three to five independent measurements. (F) The IC50 for the hKv1.3 channel wasdetermined from Lineweaver–Burk analysis, where 1/RF was plotted as a function of toxin concentration and fitting a line to the points and IC50 = 1/slope, and H = 1 was used for the Hill equation, resulting in IC50 = 91 nM on hKv1.3.

36 Luna-Ramírez et al.

at ASPE

T Journals on M

ay 11, 2020m

olpharm.aspetjournals.org

Dow

nloaded from

Page 10: Structure, Molecular Modeling, and Function of the Novel ...molpharm.aspetjournals.org/Content/Molpharm/86/1/28.Full.pdfmodulation of the action potential (Bauer and Schwarz, 2001)

the toxin to Shaker channels. Because the MD results pre-sented here indicate that Lys25 of urotoxin (equivalent toLys23 in other toxins) is not the key residue for binding tothe channel, we decided to run an MD simulation witha mutant urotoxin that lacks K25 (urotoxin-K25A) (Supple-mental Data: hKv1.2_Ala and hKv1.1_Ala) to assess thevalidity of our brute force dynamic simulation and the role ofK25 in the binding.The results after 500 nanoseconds of simulation showed

that urotoxin-K25A still binds to hKv1.1 and hKv1.2 channels.Even more, the binding seems to be stronger with thismutant-urotoxin than with native urotoxin (based on thenumber of bonds made during the simulation). In addition,the binding of themutant urotoxin involved the same residuesas with native urotoxin (all the remaining Ks and R9 fromurotoxin binding with the acidic residues in the channel).Therefore, K25 in urotoxin is not the key residue for its

binding to hKv1.1 and hKv1.2 channels, and our simulationscan be considered accurate (see Supplemental Fig. 6).

DiscussionOur work describes the isolation and characterization of the

K1 channel blocker toxin urotoxin from the venom of theAustralian scorpionU. yaschenkoi. Urotoxin is a selective andpotent blocker of the hKv1.2 channel (IC50 5 160 pM); it alsoblocks hKv1.1, hKv1.3, and hKCa3.1 channels with at least400-fold lower potency, but it does not inhibit other channelsexamined in this study.Comparison of its sequence—37 amino acids including 8

cysteine residues and C-terminal amidation—with others inthe literature shows it belongs to the a-KTx-6 family (Tytgatet al., 1999). The systematic number proposed for urotoxin isa-KTx 6.21, and its assigned Genbank accession number is

Fig. 6. Urotoxin reversibly blocks hKCa3.1 channels with low affinity but does not inhibit hKv1.5. The hKCa3.1 or Kv1.5 channels were expressed intsA201 cells, and urotoxin was applied in the bath solution. (A) KCa3.1 currents were elicited every 10 seconds with 150-millisecond voltage ramps to +40from2120 mV. The holding potential was2100 mV. Representative traces show hKCa3.1 currents in the absence (control and wash) and in the presenceof 10 nM urotoxin (as indicated). As a positive control, charybdotoxin (ChTx), a known blocker of hKCa3.1 was used in the same cell (5 nMChTx). (B) Timecourse of the development and the relief of the block of hKCa3.1 currents by 10 nM urotoxin. The gray bar indicates the application of 10 nM urotoxin inthe bath solution. Peak currents were determined at a membrane potential of +40 mV and shown as a function of time. (C) The IC50 for the hKCa3.1channel was determined from Lineweaver–Burk analysis (see Materials and Methods) resulting in IC50 = 70 nM. (D) The hKv1.5 currents were elicitedby 40-millisecond depolarization pulses to +50 mV from a holding potential of 2100 mV every 15 seconds. Representative traces show that 10 nMurotoxin had no effect on hKv1.5. The inset shows the normalized peak currents in the absence or presence (gray bar) of 10 nM urotoxin in the bathsolution.

Characterization of the Potassium Channel Blocker Urotoxin 37

at ASPE

T Journals on M

ay 11, 2020m

olpharm.aspetjournals.org

Dow

nloaded from

Page 11: Structure, Molecular Modeling, and Function of the Novel ...molpharm.aspetjournals.org/Content/Molpharm/86/1/28.Full.pdfmodulation of the action potential (Bauer and Schwarz, 2001)

KC818423. This subfamily consists of a cysteine-stabilizeda/b-scaffold that comprises an a-helix connected to a double-stranded b-sheet. So far, the a-KTx-6 subfamily is repre-sented by four different scorpion families but has not yet beenfound in the venom of Buthidae scorpions; urotoxin is the firstexample from an Urodacidae scorpion.Based on sequence identity, the closest peptides to urotoxin

in the a-KTx-6 subfamily are OcKTx5, OcKTx4, and spinoxin.The a-KTx-6 peptides from Opistophthalmus carinatus(OcKTx) have not been electrophysiologically characterized;however, the predicted structure for these toxins shows thelast residue amidated as with urotoxin. OcKTx4 is the onlymember of the OcKTx family that does not have an amidatedC-terminal, probably because it lacks the G at the lastposition, which is important for amidation of the previousresidue (Zhu et al., 2004).

Peptides with ∼50–60% identity to urotoxin—Pi-1, Pi-4,hemitoxin, and maurotoxin—all inhibit rat Kv1.2 channelswith high affinity, ranging from IC50 5 8 pM to 16 nM (Rochatet al., 1998; M’Barek et al., 2003; Mouhat et al., 2004; Srairi-Abid et al., 2008). Pi-4 has no effect on rat Kv1.1 and Kv1.3channels, but a small effect is observed in rat Ca21-activated(SK channel) K1 channels, though the highest affinity for Pi-4was toward rat Kv1.2 (M’Barek et al., 2003). Pi-1 inhibits thevoltage-gated K1 channel of human T lymphocytes Kv1.3,with nanomolar affinity (IC50 5 11 nM) (Peter et al., 1998)and competes with 125I-apamin binding to the SK channels inrat brain synaptosomes with an IC50 of 50 pM (Mouhat et al.,2004). Maurotoxin, similarly to Pi-1, also inhibits Kv1.3 butalso blocks rat Kv1.1 and the Ca21-activated K1 channelsIKCa1 (KCa3.1) of human T cells (Rochat et al., 1998). Anotherpeptide in the a-KTx-6 subfamily that inhibits Kv1.2 with

Fig. 7. Multiple alignment of urotoxin with different potassium toxins from scorpions. This alignment shows the percentage of identity (% I) of severalpotassium toxins with respect to urotoxin. The conserved cysteines are shown in gray (shaded). Amidated residues are shown in bold and underlined.Glycine residues involved in amidation are shaded. OcKTx5: potassium channel toxin a-KTx 6.10 from Opistophthalmus carinatus; OcKTx4: potassiumchannel toxin a-KTx 6.9 from O. carinatus; Spinoxin: potassium channel toxin a-KTx 6.13 from Heterometrus spinifer; Pi-1: potassium channel toxina-KTx 6.1 from Pandinus imperator; Pi-4: potassium channel toxin a-KTx 6.4 from P. imperator; Hemitoxin: potassium channel toxin a-KTx 6.15 fromHemiscorpius lepturus; Maurotoxin: potassium channel toxin a-KTx 6.2 from Scorpio maurus palmatus; Pi-7: potassium channel toxin a-KTx 6.5 fromP. imperator; OcKTx1: potassium channel toxin a-KTx 6.6 from O. carinatus; OcKTx2: potassium channel toxin a-KTx 6.7 from O. carinatus; OcKTx3:potassium channel toxin a-KTx 6.8 fromO. carinatus; HsTx1: potassium channel toxin a-KTx 6.3 fromH. spinifer; Anuroctoxin: potassium channel toxina-KTx 6.12 from Anuroctonus phaiodactylus. UniProtKB/Swiss-Prot Access number is indicated in parentheses.

Fig. 8. Predicted tertiary structure of urotoxin showing thedisulfide bridges. Maestro model visualized with VMDshowing New Cartoon drawing method. Toxin starts inthe upper gray coil. Residues G1-I3 are unarranged or incoil (gray); K4 is an isolated bridge (ochre); C5-G7 turn(blue-green); T8-T19 a-helix (purple); T20 coil (gray);C21-N23 turn (blue-green); S24-C26 b-sheet (yellow); M27-G29 turn (blue-green); K30 isolated bridge (ochre);C31-C33 b-sheet (yellow); Y34-C36 3-10-helix (blue),and V37 unarranged or coil (gray).

38 Luna-Ramírez et al.

at ASPE

T Journals on M

ay 11, 2020m

olpharm.aspetjournals.org

Dow

nloaded from

Page 12: Structure, Molecular Modeling, and Function of the Novel ...molpharm.aspetjournals.org/Content/Molpharm/86/1/28.Full.pdfmodulation of the action potential (Bauer and Schwarz, 2001)

nanomolar affinity is anuroctoxin (IC50 5 5 nM) (Bagdanyet al., 2005). Anuroctoxin has the lowest identity score tourotoxin and also inhibits Kv1.3 with subnanomolar affinity.Thus, considering the pharmacologic profile of a-KTx-6toxins, urotoxin and maurotoxin share a similar pattern ofion channel inhibition, except that urotoxin is particularlyselective for Kv1.2 channel.The 3D structure of urotoxin was predicted based on

homology modeling using the solution structure of spinoxin.The 3D model of urotoxin shows the a/b scaffold with onea-helix and two b-sheets bound by four disulfide bridges withthe following proposed connectivity: 5-26, 11-31, 15-21, 33-36.The b-sheets in urotoxin, spinoxin, and maurotoxin are highlyconserved (Table 1), and each has the K23 or K25 positionedproperly to form part of the functional dyad required forbioactivity, complemented with the CYGC at the C termini toprovide the aromatic residue for dyad formation. Further-more, these three toxins have the C termini amidated. Threeamino acids (R14, K15, and G33) were reported to be re-sponsible for the nonconventional pairing of the disulfidebridges in maurotoxin (Kharrat et al., 1997). The amino acidsin equivalent positions in urotoxin are K16, K17, and G35,which support the validity of the 3D model and the cysteinepairing of urotoxin proposed in this paper. Nonetheless, itwould be desirable to confirm experimentally the disulfide

bridge pattern of urotoxin. The need for experimentalconfirmation is underlined by the discrepancy of the 3Dmodel of hemitoxin, which suggested an unusual disulfidebridging pattern, compared with the experimental data thatshowed a conventional one (Srairi-Abid et al., 2008). Thenonconventional disulfide pairing seems to be important forhigh-affinity binding of maurotoxin to rat Kv1.2 channels.Rendering the disulfide pattern to a conventional one (C1-C5; C2-C6; C3-C7; C4-C8, Pi-1–like) reduced its affinity forKv1.2 by a factor of ∼46 and simultaneously increased itsaffinity for Shaker B channels by one order of magnitude(M’Barek et al., 2003).The 3D folding, although important for the high stability of

the peptide, does not correlate directly with a specific targetmolecule (Rodríguez de la Vega et al., 2003). Even thoughmaurotoxin, spinoxin, and urotoxin share high sequence and3D folding similarities, their affinities appear to be differenttoward ion channels (Table 2). For example, maurotoxin andurotoxin both inhibit Kv1.1, Kv1.2, and Kv1.3 channels, butthe affinity of urotoxin for Kv1.2 is at least two orders ofmagnitude higher than for the other two channels. In con-trast, maurotoxin inhibits Kv1.1, Kv1.2, and Kv1.3 channelswith a smaller degree of selectivity. Specificity for a givenchannel requires a minimum a 100-fold difference in IC50

(Giangiacomo et al., 2004). This criterion is only fulfilled byurotoxin (ratio IC50 Kv1.3/IC50 Kv1.2 5 562, for IC50 Kv1.1/IC50 Kv1.2 5 1579 and for IC50 KCa3.1/IC50 Kv1.2 5 435). Incontrast, for maurotoxin the ratio is IC50 Kv1.1/IC50 Kv1.2 556, and IC50 Kv1.3/IC50 Kv1.2 5 225, whereas for spinoxinIC50 Kv1.3/IC50 Kv1.2 5 25 (Prof. P. Gopalakrishnakone,National University of Singapore, personal communication).MD simulation is a powerful tool at the molecular level for

understanding the electrophysiologic experiments we per-formed. The MD simulation shows that residue lysines 16, 17,and 30 and arginine-9 are critical for the blocking of thehKv1.2 channel, and the residues involved from the channelare the acidic residues aspartate (363 and 379) and glutamate(355 and 353). Urotoxin forms more favorable electrostaticinteractions with the outer vestibule of hKv1.2 over hKv1.1,consistent with the selectivity observed experimentally. Ourbrute force MD simulation results are in agreement with theprevious docking models that have been made with aKTx-6toxins and Shaker channels, in which “toxins have two tothree amino acids that are essential for binding usuallyarginines or lysines” (Visan et al., 2004; Chen et al., 2011;Chen and Chung, 2012). To our knowledge, this is the firstnondocking simulation made with Shaker channels and thelongest simulation so far.Based on in silico docking studies, the general inhibitory

mechanism of Kv1.2-specific toxins consists of two steps.

TABLE 1Secondary structure of urotoxin, spinoxin, and maurotoxin

Toxin a-Helix b1-Sheet b2-Sheet

Urotoxin 8-19 T-T 24-26 SKC 31-33 CKCSpinoxin 10-17 Y-T 22-24 AKC 29-31 CKCMaurotoxin 6-17 S-T 22-25 AKC 28-31 SKCK

TABLE 2Bioactivity comparison of maurotoxin, spinoxin, and urotoxin showingIC50 values at nM concentrations

Toxin/Channel Kv1.1 Kv1.2 Kv1.3 Other

Maurotoxin 45 0.8 180 SK/KCNN and KCa3.1 (1 nM)Spinoxin NA 2.5 63Urotoxin 253 0.16 91 KCa3.1 (70 nM)

KCa3.1/KCNN4, intermediate conductance calcium-activated potassium channels;NA, not active; SK/KCNN, apamin-sensitive small conductance calcium-activatedpotassium channels.

Fig. 9. Average position of urotoxin with respect to hKv1.2 during the MDsimulation. Urotoxin is shown in purple near the pore domain. Thechannel is in gray. Highlighted (in color) are the most active residuesforming H-bonds and salt bridges during the 500-nanosecond simulation.

Characterization of the Potassium Channel Blocker Urotoxin 39

at ASPE

T Journals on M

ay 11, 2020m

olpharm.aspetjournals.org

Dow

nloaded from

Page 13: Structure, Molecular Modeling, and Function of the Novel ...molpharm.aspetjournals.org/Content/Molpharm/86/1/28.Full.pdfmodulation of the action potential (Bauer and Schwarz, 2001)

During the first step, a ring of positively charged amino acids(R10, R19, K30, and K33 in Pi4) guide the recognition andcorrect positioning of the toxin by means of electrostaticinteraction with acidic residues in the channel. The sub-sequent high-affinity binding is mediated by hydrophobicforces and hydrogen bonding between the dyad (Y35 in Pi4)and the aromatic cluster of amino acids of the channel (W366,W367, Y377), whereas the dyad lysine protrudes into theselectivity filter and is stabilized there by carbonyl oxygens ofthe selectivity filter aspartate-379 (M’Barek et al., 2003). OurMD simulation supports strongly the existence of this firststep, where electrostatic interactions guide urotoxin to thepore of hKv1.2. In addition, the formation of H-bonds takesplace involving D379 from the channel and the lysines in thetoxin, emphasizing a similar mechanism to other in silicodocking simulations; however, the final position of the toxin inthe pore was beyond the limits of the simulation.Although the dyad hypothesis (Dauplais et al., 1997) is

suitable to explain the binding of many toxins to the channelpore, high-affinity binding of Tc32 to Kv1.3 was also describedin the absence of these critical residues in the toxin (Batistaet al., 2002). Furthermore, the presence of the functional dyaditself is not a prerequisite for the binding of Pi-1 to Kv1.2(Mouhat et al., 2004), where mutant toxins with substituteddyad residues ([A24, A33]-Pi-1) still displayed affinity forKv1.2. These results allow alternative explanations for thefinal step of toxin-channel interaction, which may evolve fromthe position of urotoxin in the pore predicted by the MDsimulation (Fig. 9). Nevertheless, the MD simulation withthe mutant urotoxin (K25A) supports the accuracy of oursimulations as the binding still happens in the absence ofK25. This result might suggest that urotoxin would be able toblock the hKv1.5 channel as K25 does not make unfavorableinteractions with the charged amino acid at the entrance ofthe pore. The experiment (Fig. 6D) did not confirm thishypothesis. It is important to consider that E355 is missing inKv1.5 (it is G461 in the equivalent position of Kv1.5), whichmay explain the lack of inhibition of the hKv1.5 current byurotoxin (please note E355 in Kv1.2 and Kv1.1 is important inurotoxin binding).Overall, these results demonstrate the potential for

urotoxin to become a pharmacologic tool to characterizethe hKv1.2 channel. Knowledge of the interacting surfacesbetween the channels and toxins may allow the design ofspecific drugs to control pathologies associated with K1

channels, such as demyelinating diseases (e.g., Guillain-Barrésyndrome, multiple sclerosis) and Lambert-Eaton myasthenicsyndrome (Judge and Bever, 2006; Shi and Sun, 2011).

Acknowledgments

The authors thank Timoteo Olamendi Portugal for amino acidsequence determination, Dr. Cesar Batista and Erika P. Meneses-Romero at Unidad de Proteomica at IBT, UNAM, for molecular massdetermination, and Dr. Fernando Zamudio for attempting thedisulfide bridges connectivity determination. The authors also thankDr. Mike Kupier for giving access to the ANOVA super computer atVLSCI and for advice.

Authorship Contributions

Participated in research design: Luna-Ramírez, Restano-Cassulini,Bartok, Quintero-Hernández, Wright, Panyi, Possani.

Conducted experiments: Luna-Ramírez, Bartok, Restano-Cassulini,Coronas, Quintero-Hernández, Christensen.

Contributed new reagents or analytic tools: Luna-Ramírez, Possani,Quintero-Hernández.

Performed data analysis: Luna-Ramírez, Restano-Cassulini, Bartok,Quintero-Hernández, Coronas, Christensen, Panyi, Possani.

Wrote or contributed to the writing of the manuscript: Luna-Ramírez,Quintero-Hernández, Restano-Cassulini, Bartok, Wright, Panyi,Possani.

References

Bagdány M, Batista CVF, Valdez-Cruz NA, Somodi S, Rodriguez de la Vega RC,Licea AF, Varga Z, Gáspár R, Possani LD, and Panyi G (2005) Anuroctoxin, a newscorpion toxin of the alpha-KTx 6 subfamily, is highly selective for Kv1.3 overIKCa1 ion channels of human T lymphocytes. Mol Pharmacol 67:1034–1044.

Batista CV, Gómez-Lagunas F, Rodríguez de la Vega RC, Hajdu P, Panyi G, GáspárR, and Possani LD (2002) Two novel toxins from the Amazonian scorpion Tityuscambridgei that block Kv1.3 and Shaker B K1-channels with distinctly differentaffinities. Biochim Biophys Acta 1601:123–131.

Batista CVF, Román-González SA, Salas-Castillo SP, Zamudio FZ, Gómez-LagunasF, and Possani LD (2007) Proteomic analysis of the venom from the scorpion Tityusstigmurus: biochemical and physiological comparison with other Tityus species.Comp Biochem Physiol C Toxicol Pharmacol 146:147–157.

Bauer CK and Schwarz JR (2001) Physiology of EAG K1 channels. J Membr Biol 182:1–15.

Becerril B, Vázquez A, García C, Corona M, Bolivar F, and Possani LD (1993) Cloningand characterization of cDNAs that code for Na1-channel-blocking toxins of thescorpion Centruroides noxius Hoffmann. Gene 128:165–171.

Beraud E, Viola A, Regaya I, Confort-Gouny S, Siaud P, Ibarrola D, Le Fur Y, BarbariaJ, Pellissier JF, and Sabatier JM et al. (2006) Block of neural Kv1.1 potassiumchannels for neuroinflammatory disease therapy. Ann Neurol 60:586–596.

Burg ED, Remillard CV, and Yuan JXJ (2006) K1 channels in apoptosis. J MembrBiol 209:3–20.

Chen R and Chung SH (2012) Structural basis of the selective block of Kv1.2 bymaurotoxin from computer simulations. PLoS ONE 7:e47253.

Chen R, Robinson A, Gordon D, and Chung SH (2011) Modeling the binding of threetoxins to the voltage-gated potassium channel (Kv1.3). Biophys J 101:2652–2660.

Chen X, Wang Q, Ni F, and Ma J (2010) Structure of the full-length Shaker potas-sium channel Kv1.2 by normal-mode-based X-ray crystallographic refinement. ProcNatl Acad Sci USA 107:11352–11357.

Dauplais M, Lecoq A, Song J, Cotton J, Jamin N, Gilquin B, Roumestand C, Vita C,de Medeiros CL, and Rowan EG et al. (1997) On the convergent evolution of animaltoxins. Conservation of a diad of functional residues in potassium channel-blockingtoxins with unrelated structures. J Biol Chem 272:4302–4309.

Gati CDC, Mortari MR, and Schwartz EF (2012) Towards therapeutic applications ofarthropod venom K1-channel blockers in CNS neurologic diseases involvingmemory acquisition and storage. J Toxicol 2012:756358.

Giangiacomo KM, Ceralde Y, and Mullmann TJ (2004) Molecular basis of a-KTxspecificity. Toxicon 43:877–886.

Goldstein SAN and Miller C (1993) Mechanism of charybdotoxin block of a voltage-gated K1 channel. Biophys J 65:1613–1619.

Goudet C, Chi CW, and Tytgat J (2002) An overview of toxins and genes from thevenom of the Asian scorpion Buthus martensi Karsch. Toxicon 40:1239–1258.

Grabherr MG, Haas BJ, Yassour M, Levin JZ, Thompson DA, Amit I, Adiconis X, FanL, Raychowdhury R, and Zeng Q et al. (2011) Full-length transcriptome assemblyfrom RNA-Seq data without a reference genome. Nat Biotechnol 29:644–652.

Grissmer S, Nguyen AN, Aiyar J, Hanson DC, Mather RJ, Gutman GA, KarmilowiczMJ, Auperin DD, and Chandy KG (1994) Pharmacological characterization of fivecloned voltage-gated K1 channels, types Kv1.1, 1.2, 1.3, 1.5, and 3.1, stablyexpressed in mammalian cell lines. Mol Pharmacol 45:1227–1234.

Grissmer S, Nguyen AN, and Cahalan MD (1993) Calcium-activated potassiumchannels in resting and activated human T lymphocytes. Expression levels, cal-cium dependence, ion selectivity, and pharmacology. J Gen Physiol 102:601–630.

Gurrola GB, Hernández-López RA, Rodríguez de la Vega RC, Varga Z, Batista CVF,Salas-Castillo SP, Panyi G, del Río-Portilla F, and Possani LD (2012) Structure,function, and chemical synthesis of Vaejovis mexicanus peptide 24: a novel potentblocker of Kv1.3 potassium channels of human T lymphocytes. Biochemistry 51:4049–4061.

Huys I, Olamendi-Portugal T, Garcia-Gómez BI, Vandenberghe I, Van Beeumen J,Dyason K, Clynen E, Zhu S, van der Walt J, and Possani LD et al. (2004) Asubfamily of acidic a-K(1) toxins. J Biol Chem 279:2781–2789.

Jouirou B, Mouhat S, Andreotti N, De Waard M, and Sabatier JM (2004) Toxindeterminants required for interaction with voltage-gated K1 channels. Toxicon 43:909–914.

Judge SIV and Bever CT, Jr (2006) Potassium channel blockers in multiple sclerosis:neuronal Kv channels and effects of symptomatic treatment. Pharmacol Ther 111:224–259.

Kharrat R, Mabrouk K, Crest M, Darbon H, Oughideni R, Martin-Eauclaire MF,Jacquet G, el Ayeb M, Van Rietschoten J, and Rochat H et al. (1996) Chemicalsynthesis and characterization of maurotoxin, a short scorpion toxin with fourdisulfide bridges that acts on K1 channels. Eur J Biochem 242:491–498.

Kharrat R, Mansuelle P, Sampieri F, Crest M, Oughideni R, Van Rietschoten J,Martin-Eauclaire MF, Rochat H, and El Ayeb M (1997) Maurotoxin, a four disulfidebridge toxin from Scorpio maurus venom: purification, structure and action onpotassium channels. FEBS Lett 406:284–290.

Legros C, Oughuideni R, Darbon H, Rochat H, Bougis PE, and Martin-Eauclaire MF(1996) Characterization of a new peptide from Tityus serrulatus scorpion venomwhich is a ligand of the apamin-binding site. FEBS Lett 390:81–84.

Luna-Ramírez K, Quintero-Hernández V, Vargas-Jaimes L, Batista CVF, WinkelKD, and Possani LD (2013) Characterization of the venom from the Australian

40 Luna-Ramírez et al.

at ASPE

T Journals on M

ay 11, 2020m

olpharm.aspetjournals.org

Dow

nloaded from

Page 14: Structure, Molecular Modeling, and Function of the Novel ...molpharm.aspetjournals.org/Content/Molpharm/86/1/28.Full.pdfmodulation of the action potential (Bauer and Schwarz, 2001)

scorpion Urodacus yaschenkoi: molecular mass analysis of components, cDNAsequences and peptides with antimicrobial activity. Toxicon 63:44–54.

M’Barek S, Mosbah A, Sandoz G, Fajloun Z, Olamendi-Portugal T, Rochat H,Sampieri F, Guijarro JI, Mansuelle P, and Delepierre M et al. (2003) Synthesisand characterization of Pi4, a scorpion toxin from Pandinus imperator that actson K1 channels. Eur J Biochem 270:3583–3592.

Martin-Eauclaire MF and Bougis PE (2012) Potassium channels blockers from thevenom of Androctonus mauretanicus mauretanicus. J Toxicol 2012:103608.

Mouhat S, Mosbah A, Visan V, Wulff H, Delepierre M, Darbon H, Grissmer S, DeWaard M, and Sabatier JM (2004) The ‘functional’ dyad of scorpion toxin Pi1 is notitself a prerequisite for toxin binding to the voltage-gated Kv1.2 potassium chan-nels. Biochem J 377:25–36.

Péter MJ, Jr, Varga Z, Panyi G, Bene L, Damjanovich S, Pieri C, Possani LD,and Gáspár R, Jr (1998) Pandinus imperator scorpion venom blocks voltage-gatedK1 channels in human lymphocytes. Biochem Biophys Res Commun 242:621–625.

Possani LD, Becerril B, Delepierre M, and Tytgat J (1999) Scorpion toxins specific forNa1-channels. Eur J Biochem 264:287–300.

Redaelli E, Cassulini RR, Silva DF, Clement H, Schiavon E, Zamudio FZ, Odell G,Arcangeli A, Clare JJ, and Alagón A et al. (2010) Target promiscuity and hetero-geneous effects of tarantula venom peptides affecting Na1 and K1 ion channels.J Biol Chem 285:4130–4142.

Regaya I, Beeton C, Ferrat G, Andreotti N, Darbon H, De Waard M, and Sabatier JM(2004) Evidence for domain-specific recognition of SK and Kv channels by MTX andHsTx1 scorpion toxins. J Biol Chem 279:55690–55696.

Restrepo-Angulo I, De Vizcaya-Ruiz A, and Camacho J (2010) Ion channels in toxi-cology. J Appl Toxicol 30:497–512.

Robbins CA and Tempel BL (2012) Kv1.1 and Kv1.2: similar channels, differentseizure models. Epilepsia 53 (Suppl 1):134–141.

Rochat H, Kharrat R, Sabatier JM, Mansuelle P, Crest M, Martin-Eauclaire MF,Sampieri F, Oughideni R, Mabrouk K, and Jacquet G et al. (1998) Maurotoxin, a fourdisulfide bridges scorpion toxin acting on K1 channels. Toxicon 36:1609–1611.

Rodríguez de la Vega RC, Merino E, Becerril B, and Possani LD (2003) Novel inter-actions between K1 channels and scorpion toxins. Trends Pharmacol Sci 24:222–227.

Rodríguez de la Vega RC and Possani LD (2004) Current views on scorpion toxinsspecific for K1-channels. Toxicon 43:865–875.

Shi R and Sun W (2011) Potassium channel blockers as an effective treatment torestore impulse conduction in injured axons. Neurosci Bull 27:36–44.

Shieh CC, Coghlan M, Sullivan JP, and Gopalakrishnan M (2000) Potassium chan-nels: molecular defects, diseases, and therapeutic opportunities. Pharmacol Rev52:557–594.

Srairi-Abid N, Shahbazzadeh D, Chatti I, Mlayah-Bellalouna S, Mejdoub H, BorchaniL, Benkhalifa R, Akbari A, and El Ayeb M (2008) Hemitoxin, the first potassiumchannel toxin from the venom of the Iranian scorpion Hemiscorpius lepturus.FEBS J 275:4641–4650.

Tan PTJ, Veeramani A, Srinivasan KN, Ranganathan S, and Brusic V (2006)SCORPION2: a database for structure-function analysis of scorpion toxins. Toxicon47:356–363.

Tytgat J, Chandy KG, Garcia ML, Gutman GA, Martin-Eauclaire MF, van der WaltJJ, and Possani LD (1999) A unified nomenclature for short-chain peptides isolatedfrom scorpion venoms: a-KTx molecular subfamilies. Trends Pharmacol Sci 20:444–447.

Visan V, Fajloun Z, Sabatier JM, and Grissmer S (2004) Mapping of maurotoxinbinding sites on hKv1.2, hKv1.3, and hIKCa1 channels. Mol Pharmacol 66:1103–1112.

Wang H, Kunkel DD, Schwartzkroin PA, and Tempel BL (1994) Localization of Kv1.1and Kv1.2, two K channel proteins, to synaptic terminals, somata, and dendrites inthe mouse brain. J Neurosci 14:4588–4599.

Wonderlin WF and Strobl JS (1996) Potassium channels, proliferation and G1 pro-gression. J Membr Biol 154:91–107.

Wulff H and Castle NA (2010) Therapeutic potential of KCa3.1 blockers: recentadvances and promising trends. Expert Rev Clin Pharmacol 3:385–396.

Xie G, Harrison J, Clapcote SJ, Huang Y, Zhang JY, Wang LY, and Roder JC (2010) Anew Kv1.2 channelopathy underlying cerebellar ataxia. J Biol Chem 285:32160–32173.

Yi H, Qiu S, Cao Z, Wu Y, and Li W (2008) Molecular basis of inhibitory peptidemaurotoxin recognizing Kv1.2 channel explored by ZDOCK and molecular dynamicsimulations. Proteins 70:844–854.

Zhijian C, Feng L, Yingliang W, Xin M, and Wenxin L (2006) Genetic mechanisms ofscorpion venom peptide diversification. Toxicon 47:348–355.

Zhu S, Huys I, Dyason K, Verdonck F, and Tytgat J (2004) Evolutionary traceanalysis of scorpion toxins specific for K-channels. Proteins 54:361–370.

Address correspondence to: Karen Luna-Ramírez, AVRU and CTU,Department of Pharmacology and Therapeutics, University of Melbourne,Parkville, Victoria 3010, Australia. E-mail: [email protected];or Lourival D. Possani, Institute of Biotechnology, UNAM, Avenida Uni-versidad, 2001, Apartado Postal 510-3 Cuernavaca Morelos 62210, Mexico.E-mail: [email protected]

Characterization of the Potassium Channel Blocker Urotoxin 41

at ASPE

T Journals on M

ay 11, 2020m

olpharm.aspetjournals.org

Dow

nloaded from