Network


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

Hotspot


Dive into the research topics where William R. Widger is active.

Publication


Featured researches published by William R. Widger.


FEBS Letters | 2005

Thermodynamic and EPR studies of slowly relaxing ubisemiquinone species in the isolated bovine heart complex I

Tomoko Ohnishi; J. E. Johnson; Takahiro Yano; Russell LoBrutto; William R. Widger

Previously, we investigated ubisemiquinone (SQ) EPR spectra associated with NADH‐ubiquinone oxidoreductase (complex I) in the tightly coupled bovine heart submitochondrial particles (SMP). Based upon their widely differing spin relaxation rate, we distinguished SQ spectra arising from three distinct SQ species, namely SQNf (fast), SQNs (slow), and SQNx (very slow). The SQNf signal was observed only in the presence of the proton electrochemical gradient ( Δ μ H + ) , while SQNs and SQNx species did not require the presence of Δ μ H + . We have now succeeded in characterizing the redox and EPR properties of SQ species in the isolated bovine heart complex I. The potentiometric redox titration of the g z,y,x = 2.00 semiquinone signal gave the redox midpoint potential (E m) at pH 7.8 for the first electron transfer step [E m1(Q/SQ)] of −45 mV and the second step [E m2(SQ/QH2)] of −63 mV. It can also be expressed as [E m(Q/QH2)] of −54 mV for the overall two electron transfer with a stability constant (K stab) of the SQ form as 2.0. These characteristics revealed the existence of a thermodynamically stable intermediate redox state, which allows this protein‐associated quinone to function as a converter between n = 1 and n = 2 electron transfer steps. The EPR spectrum of the SQ species in complex I exhibits a Gaussian‐type spectrum with the peak‐to‐peak line width of ∼6.1 G at the sample temperature of 173 K. This indicates that the SQ species is in an anionic Q ▪− state in the physiological pH range. The spin relaxation rate of the SQ species in isolated complex I is much slower than the SQ counterparts in the complex I in situ in SMP. We tentatively assigned slow relaxing anionic SQ species as SQNs, based on the monophasic power saturation profile and several fold increase of its spin relaxation rate in the presence of reduced cluster N2. The current study also suggests that the very slowly relaxing SQNx species may not be an intrinsic complex I component. The functional role of SQNs is further discussed in connection with the SQNf species defined in SMP in situ.


Journal of Molecular Evolution | 1997

Conserved Gene Clusters in Bacterial Genomes Provide Further Support for the Primacy of RNA

Janet L. Siefert; Kirt Martin; Fadi Abdi; William R. Widger; George E. Fox

Abstract. Five complete bacterial genome sequences have been released to the scientific community. These include four (eu)Bacteria, Haemophilus influenzae, Mycoplasma genitalium, M. pneumoniae, and Synechocystis PCC 6803, as well as one Archaeon, Methanococcus jannaschii. Features of organization shared by these genomes are likely to have arisen very early in the history of the bacteria and thus can be expected to provide further insight into the nature of early ancestors. Results of a genome comparison of these five organisms confirm earlier observations that gene order is remarkably unpreserved. There are, nevertheless, at least 16 clusters of two or more genes whose order remains the same among the four (eu)Bacteria and these are presumed to reflect conserved elements of coordinated gene expression that require gene proximity. Eight of these gene orders are essentially conserved in the Archaea as well. Many of these clusters are known to be regulated by RNA-level mechanisms in Escherichia coli, which supports the earlier suggestion that this type of regulation of gene expression may have arisen very early. We conclude that although the last common ancestor may have had a DNA genome, it likely was preceded by progenotes with an RNA genome.


Molecular Cell | 2016

Allele-specific reprogramming of cancer metabolism by the long non-coding RNA, CCAT2

Roxana S. Redis; Luz E. Vela; Weiqin Lu; Juliana Ferreira de Oliveira; Cristina Ivan; Cristian Rodriguez-Aguayo; Douglas Adamoski; Barbara Pasculli; Ayumu Taguchi; Yunyun Chen; Agustín F. Fernández; Luis Valledor; Katrien Van Roosbroeck; Samuel Chang; Maitri Y. Shah; Garrett Kinnebrew; Leng Han; Yaser Atlasi; Lawrence H. Cheung; Gilbert Y. Huang; Paloma Monroig; Marc S. Ramirez; Tina Catela Ivković; Long Van; Hui Ling; Roberta Gafà; Sanja Kapitanović; Giovanni Lanza; James A. Bankson; Peng Huang

Altered energy metabolism is a cancer hallmark as malignant cells tailor their metabolic pathways to meet their energy requirements. Glucose and glutamine are the major nutrients that fuel cellular metabolism, and the pathways utilizing these nutrients are often altered in cancer. Here, we show that the long ncRNA CCAT2, located at the 8q24 amplicon on cancer risk-associated rs6983267 SNP, regulates cancer metabolism in vitro and in vivo in an allele-specific manner by binding the Cleavage Factor I (CFIm) complex with distinct affinities for the two subunits (CFIm25 and CFIm68). The CCAT2 interaction with the CFIm complex fine-tunes the alternative splicing of Glutaminase (GLS) by selecting the poly(A) site in intron 14 of the precursor mRNA. These findings uncover a complex, allele-specific regulatory mechanism of cancer metabolism orchestrated by the two alleles of a long ncRNA.


Journal of Biological Chemistry | 1996

The Antibiotic Bicyclomycin Affects the Secondary RNA Binding Site of Escherichia coli Transcription Termination Factor Rho

Attila Magyar; Xiangdong Zhang; Harold Kohn; William R. Widger

The interaction of Rho and the antibiotic bicyclomycin was probed using in vitro transcription termination reactions, poly(C) binding assays, limited tryptic digestions, and the bicyclomycin inhibition kinetics of ATPase activity in the presence of poly(dC) and ribo(C)10. The approximate I50 value for the bicyclomycin inhibition of transcription termination at Rho-dependent sites within a modified trp operon template was 5 μM. At antibiotic concentrations near the I50 value, bicyclomycin inhibition of Rho-dependent transcripts was accompanied by the appearance of a new set of transcripts whose size was midway between the Rho-dependent transcripts and the readthrough transcripts. Bicyclomycin did not inhibit poly(C) binding to Rho. In the presence of poly(dC), bicyclomycin showed a reversible mixed inhibition of the ribo(C)10-stimulated ATPase activity. The extrapolated Ki for bicyclomycin was 2.8 μM without ribo(C)10 and increased to 26 μM in the presence of ribo(C)10. Correspondingly, the Km(app) for ribo(C)10 without bicyclomycin was 0.8 μM and with bicyclomycin was 5 μM at infinite inhibitor concentration. The data suggested that the antibiotic binds to Rho, influencing the secondary RNA binding (tracking) site on Rho and slows the tracking of Rho toward the bound RNA polymerase.


Current Drug Targets - Infectious Disorders | 2005

The molecular basis for the mode of action of bicyclomycin

Harold Kohn; William R. Widger

Bicyclomycin (1) is a clinically useful antibiotic exhibiting activity against a broad spectrum of Gram-negative bacteria and against the Gram-positive bacterium, Micrococcus luteus. Bicyclomycin has been used to treat diarrhea in humans and bacterial diarrhea in calves and pigs and is marketed by Fujisawa (Osaka, Japan) under the trade name Bicozamycin. The structure of 1 is unique among antibiotics, and our studies document that its mechanism of action is novel. Early mechanistic proposals suggested that 1 reacted with nucleophiles (e.g., a protein sulfhydryl group) necessary for the remodeling the peptidoglycan assembly within the bacterial cell wall. We, however, showed that 1 targeted the rho transcription termination factor in Escherichia coli. The rho protein is integral to the expression of many gene products in E. coli and other Gram-negative bacteria, and without rho the cell losses viability. Rho is a member of the RecA-type ATPase class of enzymes that use nucleotide contacts to couple oligonucleotide translocation to ATP hydrolysis. Bicyclomycin is the only known selective inhibitor of rho. In this article, we integrate the evidence obtained from bicyclomycin structure-activity studies, site-directed mutagenesis investigations, bicyclomycin affinity labels, and biochemical and biophysical measurements with recent X-ray crystallographic images of the bicyclomycin-rho complex to define the rho antibiotic binding site and to document the pathway for rho inhibition by 1. Together, the structural and functional studies demonstrate how 1, a modest rho inhibitor, can disrupt the rho molecular machinery thereby leading to a catastrophic effect caused by the untimely overproduction of proteins not normally expressed constitutively, thus leading to a toxic effect on the cells.


Journal of Biological Chemistry | 1999

Identifying the bicyclomycin binding domain through biochemical analysis of antibiotic-resistant rho proteins.

Attila Magyar; Xiangdong Zhang; Fadi Abdi; Harold Kohn; William R. Widger

Mutations M219K, S266A, and G337S in transcription termination factor Rho have been shown to confer resistance to the antibiotic bicyclomycin (BCM). All three His-tagged mutant Rho proteins exhibited similar K m values for ATP; however, the V max values at infinite ATP concentrations were one-fourth to one-third that for the His-tagged wild-type enzyme. BCM inhibition kinetics of poly(C)-dependent ATPase activity for the mutant proteins were non-competitive with respect to ATP (altering catalytic function but not ATP binding) and showed increased K i values compared with His-tagged wild-type Rho. M219K and G337S exhibited increased ratios of poly(U)/poly(C)-stimulated ATPase activity and lower apparent K m values for ribo(C)10 in the poly(dC)·ribo(C)10–dependent ATPase assay compared with His-tagged wild-type Rho. The S266A mutation did not show an increased poly(U)/poly(C) ATPase activity ratio and maintained approximately the same K m for ribo(C)10in the poly(dC)·ribo(C)10–dependent ATPase assay. The kinetic studies indicated that M219K and G337S altered the secondary RNA binding domain in Rho whereas the S266A mutation did not. Transcription termination assays for each mutant showed different patterns of Rho-terminated transcripts. Tyrosine substitution of Ser-266 led to BCM sensitivity intimating that an OH (hydroxyl) moiety at this position is needed for BCM (binding) inhibition. Our results suggest BCM binds to Rho at a site distinct from both the ATP and the primary RNA binding domains but close to the secondary RNA-binding (tracking) site and the ATP hydrolysis pocket.


Photosynthesis Research | 1991

The cloning and sequencing of Synechococcus sp. PCC 7002 petCA operon: Implications for the cytochrome c-553 binding domain of cytochrome f

William R. Widger

The genes encoding the Rieske iron-sulfur protein and cytochrome f from a unicellular, naturally transformable, photoheterotrophic cyanobacterium, Synechococcus sp. PCC 7002, formerly Agmenellum quadruplicatum, have been isolated and sequenced. The two genes were found to be on a single operon, petCA.The Synechococcus sp. PCC 7002 iron-sulfur protein contains 181 amino acids, the conserved putative iron-binding domains CTHLGCV, residues 108–114, and CPCHGS, residues 128–133, no presequence and has a 73% sequence identity to the Nostoc PCC 7906 iron-sulfur protein. The 325 amino acid apocytochrome f sequence contains a 42 amino acid presequence, a CANCH heme binding domain, residues 20–24 from the presumed start of the mature protein, and a predicted hydrophobic membrane-spanning domain, residues 250–269. The mature cytochrome f sequence has a 71.5% sequence identity with Nostoc PCC 7906 cytochrome f and possesses a large (-14) negative charge and low calculated pI of 4.47 compared to higher plant chloroplast sequences. Nine separate domains showing differences in charged residues among cyanobacteria and plants have been identified and the possibility that these domains are involved in the ionic interactions with plastocyanin or cytochrome c-553 is discussed.


Journal of Biological Chemistry | 2002

Mutations in the Rho Transcription Termination Factor That Affect RNA Tracking

Yi Xu; Harold Kohn; William R. Widger

Model studies have identified 16 conserved positively charged amino acids that form a positive strip pointing toward the center hole of Rho. Fourteen residues were individually changed to either an alanine or a glycine and one to a glutamate. Residues Arg269, Arg272, Lys283, Arg296, Lys298, and Arg299 form a subdomain (locus) located N-terminal to (above) the ATP hydrolysis domain (P-loop) and mutations in these residues led to either inactive Rho or to proteins displaying decreasedk cat for poly(C)-dependent ATP hydrolysis, increased K m for ribo(C)10 activation, and decreased transcription termination efficiencies (57–77%) compared with wild-type Rho. Residues Arg347, Lys348, Lys352, and Arg353 form a subdomain (locus) C-terminal to (below) the ATP hydrolysis domain, and mutations in these residues also show a decreased k cat for poly(C)-dependent ATP hydrolysis, an increasedK m for ribo(C)10 activation, and a 50–70% decrease in transcription termination, compared with wild-type Rho. Residues Arg212 and Lys336 surround the ATP hydrolysis domain, and mutations in these residues also altered the kinetic properties of Rho. We conclude that the secondary RNA-tracking site consists of amino acids whose putative orientation faces the central hole in Rho and in part reside in two clusters of positively charged residues located above and below the ATP hydrolysis domain.


Photosynthesis Research | 2003

Cyanobacterial signature genes

Kirt Martin; Janet L. Siefert; Sailaja Yerrapragada; Yue Lu; Thomas Z. McNeill; Pedro A. Moreno; George M. Weinstock; William R. Widger; George E. Fox

A comparison of 8 cyanobacterial genomes reveals that there are 181 shared genes that do not have obvious orthologs in other bacteria. These signature genes define aspects of the genotype that are uniquely cyanobacterial. Approximately 25% of these genes have been associated with some function. These signature genes may or may not be involved in photosynthesis but likely they will be in many cases. In addition, several examples of widely conserved gene order involving two or more signature genes were observed. This suggests there may be regulatory processes that have been preserved throughout the long history of the cyanobacterial phenotype. The results presented here will be especially useful because they identify which of the many genes of unassigned function are likely to be of the greatest interest.


Plant Molecular Biology | 1992

Cloning and sequencing of the petBD operon from the cyanobacterium Synechococcus sp. PCC 7002

Susan N. Brand; Xiaolin Tan; William R. Widger

The genes encoding the photosynthetic cytochrome b6 (petB) and subunit 4 (petD) have been cloned and sequenced from the unicellular, photoheterotrophic, transformable cyanobacterium Synechococcus sp. PCC 7002, formerly designated Agmenellum quadruplicatum. The gene arrangement was found to be similar to that reported in the cyanobacterium Nostoc PCC 7906. The DNA and derived protein sequences were compared to chloroplast and the other cyanobacterial sequences. By pulsed-field electrophoresis, the petBD operon and the petCA operon, encoding the Rieske iron-sulfur protein and cytochrome f, were found to be located on separate, unlinked,Not I-digested DNA fragments. ThepetBD operon was found on the third largest Not I fragment (NC-325) while the petCA operon was found on the second largest Not I fragment (NB-370). These results suggest the two operons are not in proximity. The 1.35 kb transcript was shown to be light-regulated. Transcripts from cells grown under constant illumination showed a decrease in petB transcript levels to undetectable levels within 2 h after the cells were placed in the dark. Upon reillumination, transcript levels rose to three-fold over that seen initially under constant illumination.

Collaboration


Dive into the William R. Widger's collaboration.

Top Co-Authors

Avatar

Harold Kohn

University of North Carolina at Chapel Hill

View shared research outputs
Top Co-Authors

Avatar
Top Co-Authors

Avatar

Yuriy Fofanov

University of Texas Medical Branch

View shared research outputs
Top Co-Authors

Avatar
Top Co-Authors

Avatar
Top Co-Authors

Avatar

J. E. Johnson

University of Houston–Downtown

View shared research outputs
Top Co-Authors

Avatar
Top Co-Authors

Avatar

Mark Rojas

University of Texas Medical Branch

View shared research outputs
Top Co-Authors

Avatar
Top Co-Authors

Avatar
Researchain Logo
Decentralizing Knowledge