Network


Latest external collaboration on country level. Dive into details by clicking on the dots.

Hotspot


Dive into the research topics where Squire J. Booker is active.

Publication


Featured researches published by Squire J. Booker.


Science | 2011

A Radically Different Mechanism for S-Adenosylmethionine–Dependent Methyltransferases

Tyler L. Grove; Jack S. Benner; Matthew I. Radle; Jessica H. Ahlum; Bradley J. Landgraf; Carsten Krebs; Squire J. Booker

Methylation of the bacterial ribosome by two methyltransferases proceeds by an unusual radical mechanism. Methylation of small molecules and macromolecules is crucial in metabolism, cell signaling, and epigenetic programming and is most often achieved by S-adenosylmethionine (SAM)–dependent methyltransferases. Most employ an SN2 mechanism to methylate nucleophilic sites on their substrates, but recently, radical SAM enzymes have been identified that methylate carbon atoms that are not inherently nucleophilic via the intermediacy of a 5′-deoxyadenosyl 5′-radical. We have determined the mechanisms of two such reactions targeting the sp2-hybridized carbons at positions 2 and 8 of adenosine 2503 in 23S ribosomal RNA, catalyzed by RlmN and Cfr, respectively. In neither case is a methyl group transferred directly from SAM to the RNA; rather, both reactions proceed by a ping-pong mechanism involving intermediate methylation of a conserved cysteine residue.


Science | 2011

Structural Basis for Methyl Transfer by a Radical SAM Enzyme

Amie K. Boal; Tyler L. Grove; Monica I. McLaughlin; Neela H. Yennawar; Squire J. Booker; Amy C. Rosenzweig

An enzyme harnesses methyl donation and reductive cleavage of its cofactor within a single active site to methylate RNA. The radical S-adenosyl-l-methionine (SAM) enzymes RlmN and Cfr methylate 23S ribosomal RNA, modifying the C2 or C8 position of adenosine 2503. The methyl groups are installed by a two-step sequence involving initial methylation of a conserved Cys residue (RlmN Cys355) by SAM. Methyl transfer to the substrate requires reductive cleavage of a second equivalent of SAM. Crystal structures of RlmN and RlmN with SAM show that a single molecule of SAM coordinates the [4Fe-4S] cluster. Residue Cys355 is S-methylated and located proximal to the SAM methyl group, suggesting the SAM that is involved in the initial methyl transfer binds at the same site. Thus, RlmN accomplishes its complex reaction with structural economy, harnessing the two most important reactivities of SAM within a single site.


Proceedings of the National Academy of Sciences of the United States of America | 2009

Substrate positioning controls the partition between halogenation and hydroxylation in the aliphatic halogenase, SyrB2

Megan L. Matthews; Christopher S. Neumann; Miles La; Tyler L. Grove; Squire J. Booker; Carsten Krebs; Christopher T. Walsh; Bollinger Jm

The α-ketoglutarate-dependent hydroxylases and halogenases employ similar reaction mechanisms involving hydrogen-abstracting Fe(IV)-oxo (ferryl) intermediates. In the halogenases, the carboxylate residue from the His2(Asp/Glu)1“facial triad” of iron ligands found in the hydroxylases is replaced by alanine, and a halide ion (X−) coordinates at the vacated site. Halogenation is thought to result from “rebound” of the halogen radical from the X-Fe(III)-OH intermediate produced by hydrogen (H•) abstraction to the substrate radical. The alternative decay pathway for the X-Fe(III)-OH intermediate, rebound of the hydroxyl radical to the substrate radical (as occurs in the hydroxylases), reportedly does not compete. Here we show for the halogenase SyrB2 that positioning of the alkyl group of the substrate away from the oxo/hydroxo ligand and closer to the halogen ligand sacrifices H•-abstraction proficiency for halogen-rebound selectivity. Upon replacement of l-Thr, the C4 amino acid tethered to the SyrB1 carrier protein in the native substrate, by the C5 amino acid l-norvaline, decay of the chloroferryl intermediate becomes 130× faster and the reaction outcome switches to primarily hydroxylation of C5, consistent with projection of the methyl group closer to the oxo/hydroxo by the longer side chain. Competing H• abstraction from C4 results primarily in chlorination, as occurs at this site in the native substrate. Consequently, deuteration of C5, which slows attack at this site, switches both the regioselectivity from C5 to C4 and the chemoselectivity from hydroxylation to chlorination. Thus, substrate-intermediate disposition and the carboxylate → halide ligand swap combine to specify the halogenation outcome.


Biochemistry | 2012

Evidence for only oxygenative cleavage of aldehydes to alk(a/e)nes and formate by cyanobacterial aldehyde decarbonylases

Ning Li; Wei-chen Chang; Douglas M. Warui; Squire J. Booker; Carsten Krebs; J. Martin Bollinger

Cyanobacterial aldehyde decarbonylases (ADs) catalyze the conversion of C(n) fatty aldehydes to formate (HCO(2)(-)) and the corresponding C(n-1) alk(a/e)nes. Previous studies of the Nostoc punctiforme (Np) AD produced in Escherichia coli (Ec) showed that this apparently hydrolytic reaction is actually a cryptically redox oxygenation process, in which one O-atom is incorporated from O(2) into formate and a protein-based reducing system (NADPH, ferredoxin, and ferredoxin reductase; N/F/FR) provides all four electrons needed for the complete reduction of O(2). Two subsequent publications by Marsh and co-workers [ Das, et al. ( 2011 ) Angew. Chem. Int. Ed. 50 , 7148 - 7152 ; Eser, et al. ( 2011 ) Biochemistry 50 , 10743 - 10750 ] reported that their Ec-expressed Np and Prochlorococcus marinus (Pm) AD preparations transform aldehydes to the same products more rapidly by an O(2)-independent, truly hydrolytic process, which they suggested proceeded by transient substrate reduction with obligatory participation by the reducing system (they used a chemical system, NADH and phenazine methosulfate; N/PMS). To resolve this discrepancy, we re-examined our preparations of both AD orthologues by a combination of (i) activity assays in the presence and absence of O(2) and (ii) (18)O(2) and H(2)(18)O isotope-tracer experiments with direct mass-spectrometric detection of the HCO(2)(-) product. For multiple combinations of the AD orthologue (Np and Pm), reducing system (protein-based and chemical), and substrate (n-heptanal and n-octadecanal), our preparations strictly require O(2) for activity and do not support detectable hydrolytic formate production, despite having catalytic activities similar to or greater than those reported by Marsh and co-workers. Our results, especially of the (18)O-tracer experiments, suggest that the activity observed by Marsh and co-workers could have arisen from contaminating O(2) in their assays. The definitive reaffirmation of the oxygenative nature of the reaction implies that the enzyme, initially designated as aldehyde decarbonylase when the C1-derived coproduct was thought to be carbon monoxide rather than formate, should be redesignated as aldehyde-deformylating oxygenase (ADO).


Journal of the American Chemical Society | 2011

Conversion of Fatty Aldehydes to Alka(e)nes and Formate by a Cyanobacterial Aldehyde Decarbonylase: Cryptic Redox by an Unusual Dimetal Oxygenase

Ning Li; Hanne Nørgaard; Douglas M. Warui; Squire J. Booker; Carsten Krebs; J. Martin Bollinger

Cyanobacterial aldehyde decarbonylase (AD) catalyzes conversion of fatty aldehydes (R-CHO) to alka(e)nes (R-H) and formate. Curiously, although this reaction appears to be redox-neutral and formally hydrolytic, AD has a ferritin-like protein architecture and a carboxylate-bridged dimetal cofactor that are both structurally similar to those found in di-iron oxidases and oxygenases. In addition, the in vitro activity of the AD from Nostoc punctiforme (Np) was shown to require a reducing system similar to the systems employed by these O(2)-utilizing di-iron enzymes. Here, we resolve this conundrum by showing that aldehyde cleavage by the Np AD also requires dioxygen and results in incorporation of (18)O from (18)O(2) into the formate product. AD thus oxygenates, without oxidizing, its substrate. We posit that (i) O(2) adds to the reduced cofactor to generate a metal-bound peroxide nucleophile that attacks the substrate carbonyl and initiates a radical scission of the C1-C2 bond, and (ii) the reducing system delivers two electrons during aldehyde cleavage, ensuring a redox-neutral outcome, and two additional electrons to return an oxidized form of the cofactor back to the reduced, O(2)-reactive form.


Proceedings of the National Academy of Sciences of the United States of America | 2013

X-ray structure of an AdoMet radical activase reveals an anaerobic solution for formylglycine posttranslational modification

Peter Goldman; Tyler L. Grove; Lauren A. Sites; Martin I. McLaughlin; Squire J. Booker; Catherine L. Drennan

Arylsulfatases require a maturating enzyme to perform a co- or posttranslational modification to form a catalytically essential formylglycine (FGly) residue. In organisms that live aerobically, molecular oxygen is used enzymatically to oxidize cysteine to FGly. Under anaerobic conditions, S-adenosylmethionine (AdoMet) radical chemistry is used. Here we present the structures of an anaerobic sulfatase maturating enzyme (anSME), both with and without peptidyl-substrates, at 1.6–1.8 Å resolution. We find that anSMEs differ from their aerobic counterparts in using backbone-based hydrogen-bonding patterns to interact with their peptidyl-substrates, leading to decreased sequence specificity. These anSME structures from Clostridium perfringens are also the first of an AdoMet radical enzyme that performs dehydrogenase chemistry. Together with accompanying mutagenesis data, a mechanistic proposal is put forth for how AdoMet radical chemistry is coopted to perform a dehydrogenation reaction. In the oxidation of cysteine or serine to FGly by anSME, we identify D277 and an auxiliary [4Fe-4S] cluster as the likely acceptor of the final proton and electron, respectively. D277 and both auxiliary clusters are housed in a cysteine-rich C-terminal domain, termed SPASM domain, that contains homology to ∼1,400 other unique AdoMet radical enzymes proposed to use [4Fe-4S] clusters to ligate peptidyl-substrates for subsequent modification. In contrast to this proposal, we find that neither auxiliary cluster in anSME bind substrate, and both are fully ligated by cysteine residues. Instead, our structural data suggest that the placement of these auxiliary clusters creates a conduit for electrons to travel from the buried substrate to the protein surface.


Journal of Biological Chemistry | 2015

Mechanistic Diversity of Radical S-Adenosylmethionine (SAM)-dependent Methylation

Matthew R. Bauerle; Erica L. Schwalm; Squire J. Booker

Radical S-adenosylmethionine (SAM) enzymes use the oxidizing power of a 5′-deoxyadenosyl 5′-radical to initiate an amazing array of transformations, usually through the abstraction of a target substrate hydrogen atom. A common reaction of radical SAM (RS) enzymes is the methylation of unactivated carbon or phosphorous atoms found in numerous primary and secondary metabolites, as well as in proteins, sugars, lipids, and RNA. However, neither the chemical mechanisms by which these unactivated atoms obtain methyl groups nor the actual methyl donors are conserved. In fact, RS methylases have been grouped into three classes based on protein architecture, cofactor requirement, and predicted mechanism of catalysis. Class A methylases use two cysteine residues to methylate sp2-hybridized carbon centers. Class B methylases require a cobalamin cofactor to methylate both sp2-hybridized and sp3-hybridized carbon centers as well as phosphinate phosphorous atoms. Class C methylases share significant sequence homology with the RS enzyme, HemN, and may bind two SAM molecules simultaneously to methylate sp2-hybridized carbon centers. Lastly, we describe a new class of recently discovered RS methylases. These Class D methylases, unlike Class A, B, and C enzymes, which use SAM as the source of the donated methyl carbon, are proposed to methylate sp2-hybridized carbon centers using methylenetetrahydrofolate as the source of the appended methyl carbon.


Biochemistry | 2009

Characterization of RimO, a New Member of the Methylthiotransferase Subclass of the Radical SAM Superfamily †

Kyung-Hoon Lee; Lana Saleh; Brian P. Anton; Catherine L. Madinger; Jack S. Benner; David F. Iwig; Richard J. Roberts; Carsten Krebs; Squire J. Booker

RimO, encoded by the yliG gene in Escherichia coli, has been recently identified in vivo as the enzyme responsible for the attachment of a methylthio group on the beta-carbon of Asp88 of the small ribosomal protein S12 [Anton, B. P., Saleh, L., Benner, J. S., Raleigh, E. A., Kasif, S., and Roberts, R. J. (2008) Proc. Natl. Acad. Sci. U.S.A. 105, 1826-1831]. To date, it is the only enzyme known to catalyze methylthiolation of a protein substrate; the four other naturally occurring methylthio modifications have been observed on tRNA. All members of the methylthiotransferase (MTTase) family, to which RimO belongs, have been shown to contain the canonical CxxxCxxC motif in their primary structures that is typical of the radical S-adenosylmethionine (SAM) family of proteins. MiaB, the only characterized MTTase, and the enzyme experimentally shown to be responsible for methylthiolation of N(6)-isopentenyladenosine of tRNA in E. coli and Thermotoga maritima, has been demonstrated to harbor two distinct [4Fe-4S] clusters. Herein, we report in vitro biochemical and spectroscopic characterization of RimO. We show by analytical and spectroscopic methods that RimO, overproduced in E. coli in the presence of iron-sulfur cluster biosynthesis proteins from Azotobacter vinelandii, contains one [4Fe-4S](2+) cluster. Reconstitution of this form of RimO (RimO(rcn)) with (57)Fe and sodium sulfide results in a protein that contains two [4Fe-4S](2+) clusters, similar to MiaB. We also show by mass spectrometry that RimO(rcn) catalyzes the attachment of a methylthio group to a peptide substrate analogue that mimics the loop structure bearing aspartyl 88 of the S12 ribosomal protein from E. coli. Kinetic analysis of this reaction shows that the activity of RimO(rcn) in the presence of the substrate analogue does not support a complete turnover. We discuss the possible requirement for an assembled ribosome for fully active RimO in vitro. Our findings are consistent with those of other enzymes that catalyze sulfur insertion, such as biotin synthase, lipoyl synthase, and MiaB.


Advances in Protein Chemistry | 2001

Radical mechanisms of S-adenosylmethionine-dependent enzymes.

Perry A. Frey; Squire J. Booker

Publisher Summary This chapter discusses the family of S-adenosylmethione (SAM)-dependent enzymes that make use of the 5’-deoxyadenosyl radical, with special reference to the mechanism by which SAM is cleaved reversibly at the active site. One member of the family, lysine 2,3-aminomutase, makes use of the 5’-deoxyadenosyl radical to initiate the molecular rearrangement of a substrate, and the radical is subsequently regenerated in the catalytic cycle, such that SAM functions in a catalytic role as a true coenzyme. The other family members use SAM as a substrate, and the 5’-deoxyadenosyl free radical appears to be an intermediate in an irreversible hydrogen abstraction reaction. The latter enzymes include the pyruvate formate-lyase activating enzyme, the anaerobic ribonucleotide reductase from escherichia coli , biotin synthase, and lipoic acid synthase. The chapter also considers the functions of the 5’-deoxyadenosyl radical and the mechanisms of these diverse reactions. Photolytic or thermal cleavage of adenosylcobalamin produces cob(II)alamin and the 5’- deoxyadenosyl radical through homolytic scission of the Co-C bond. Spectral and kinetic evidence indicate that this same cleavage is brought about at the active sites of adenosylcobalamin-dependent enzymes, and the ensuing 5’-deoxyadenosyl radical is thought to initiate the free radical-based reaction mechanisms catalyzed by those enzymes.


Annual Review of Biochemistry | 2016

Radical S-Adenosylmethionine Enzymes in Human Health and Disease

Bradley J. Landgraf; Erin L. McCarthy; Squire J. Booker

Radical S-adenosylmethionine (SAM) enzymes catalyze an astonishing array of complex and chemically challenging reactions across all domains of life. Of approximately 114,000 of these enzymes, 8 are known to be present in humans: MOCS1, molybdenum cofactor biosynthesis; LIAS, lipoic acid biosynthesis; CDK5RAP1, 2-methylthio-N(6)-isopentenyladenosine biosynthesis; CDKAL1, methylthio-N(6)-threonylcarbamoyladenosine biosynthesis; TYW1, wybutosine biosynthesis; ELP3, 5-methoxycarbonylmethyl uridine; and RSAD1 and viperin, both of unknown function. Aberrations in the genes encoding these proteins result in a variety of diseases. In this review, we summarize the biochemical characterization of these 8 radical S-adenosylmethionine enzymes and, in the context of human health, describe the deleterious effects that result from such genetic mutations.

Collaboration


Dive into the Squire J. Booker's collaboration.

Top Co-Authors

Avatar

Tyler L. Grove

Pennsylvania State University

View shared research outputs
Top Co-Authors

Avatar

Carsten Krebs

Pennsylvania State University

View shared research outputs
Top Co-Authors

Avatar

Nicholas D. Lanz

Pennsylvania State University

View shared research outputs
Top Co-Authors

Avatar
Top Co-Authors

Avatar

Robert M. Cicchillo

Pennsylvania State University

View shared research outputs
Top Co-Authors

Avatar

Alexey Silakov

Pennsylvania State University

View shared research outputs
Top Co-Authors

Avatar

Anthony J. Blaszczyk

Pennsylvania State University

View shared research outputs
Top Co-Authors

Avatar

Bradley J. Landgraf

Pennsylvania State University

View shared research outputs
Top Co-Authors

Avatar

David F. Iwig

Pennsylvania State University

View shared research outputs
Top Co-Authors

Avatar

Bo Wang

Pennsylvania State University

View shared research outputs
Researchain Logo
Decentralizing Knowledge