Network


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

Hotspot


Dive into the research topics where Timothy C. Mueser is active.

Publication


Featured researches published by Timothy C. Mueser.


Journal of Biological Chemistry | 2007

Crystal structure of bacteriophage T4 5' nuclease in complex with a branched DNA reveals how flap endonuclease-1 family nucleases bind their substrates.

Juliette M. Devos; Stephen J. Tomanicek; Charles E. Jones; Nancy G. Nossal; Timothy C. Mueser

Bacteriophage T4 RNase H, a flap endonuclease-1 family nuclease, removes RNA primers from lagging strand fragments. It has both 5′ nuclease and flap endonuclease activities. Our previous structure of native T4 RNase H (PDB code 1TFR) revealed an active site composed of highly conserved Asp residues and two bound hydrated magnesium ions. Here, we report the crystal structure of T4 RNase H in complex with a fork DNA substrate bound in its active site. This is the first structure of a flap endonuclease-1 family protein with its complete branched substrate. The fork duplex interacts with an extended loop of the helix-hairpin-helix motif class 2. The 5′ arm crosses over the active site, extending below the bridge (helical arch) region. Cleavage assays of this DNA substrate identify a primary cut site 7-bases in from the 5′ arm. The scissile phosphate, the first bond in the duplex DNA adjacent to the 5′ arm, lies above a magnesium binding site. The less ordered 3′ arm reaches toward the C and N termini of the enzyme, which are binding sites for T4 32 protein and T4 45 clamp, respectively. In the crystal structure, the scissile bond is located within the double-stranded DNA, between the first two duplex nucleotides next to the 5′ arm, and lies above a magnesium binding site. This complex provides important insight into substrate recognition and specificity of the flap endonuclease-1 enzymes.


Proceedings of the National Academy of Sciences of the United States of America | 2001

Bacteriophage T4 gene 41 helicase and gene 59 helicase-loading protein: A versatile couple with roles in replication and recombination

Charles E. Jones; Timothy C. Mueser; Kathleen C. Dudas; Kenneth N. Kreuzer; Nancy G. Nossal

Bacteriophage T4 uses two modes of replication initiation: origin-dependent replication early in infection and recombination-dependent replication at later times. The same relatively simple complex of T4 replication proteins is responsible for both modes of DNA synthesis. Thus the mechanism for loading the T4 41 helicase must be versatile enough to allow it to be loaded on R loops created by transcription at several origins, on D loops created by recombination, and on stalled replication forks. T4 59 helicase-loading protein is a small, basic, almost completely α-helical protein whose N-terminal domain has structural similarity to high mobility group family proteins. In this paper we review recent evidence that 59 protein recognizes specific structures rather than specific sequences. It binds and loads the helicase on replication forks and on three- and four-stranded (Holliday junction) recombination structures, without sequence specificity. We summarize our experiments showing that purified T4 enzymes catalyze complete unidirectional replication of a plasmid containing the T4 ori(uvsY) origin, with a preformed R loop at the position of the R loop identified at this origin in vivo. This replication depends on the 41 helicase and is strongly stimulated by 59 protein. Moreover, the helicase-loading protein helps to coordinate leading and lagging strand synthesis by blocking replication on the ori(uvsY) R loop plasmid until the helicase is loaded. The T4 enzymes also can replicate plasmids with R loops that do not have a T4 origin sequence, but only if the R loops are within an easily unwound DNA sequence.


Acta Crystallographica Section D-biological Crystallography | 2004

A preliminary solubility screen used to improve crystallization trials: crystallization and preliminary X-ray structure determination of Aeropyrum pernix flap endonuclease-1

Brandon K. Collins; Stephen J. Tomanicek; Natasha Lyamicheva; Michael W. Kaiser; Timothy C. Mueser

Crystallization of protein and protein complexes is a multi-parametric problem that involves the investigation of a vast number of physical and chemical conditions. The buffers, salts and additives used to prepare the protein will be present in every crystallization condition. It is imperative that these conditions be defined prior to crystal screening since they will have a ubiquitous involvement in the crystal-growth experiments. This study involves the crystallization and preliminary analysis of the flap endonuclease-1 (FEN-1) DNA-repair enzyme from the crenarchaeal organism Aeropyrum pernix (Ape). Ape FEN-1 protein in a standard chromatography buffer had only a modest solubility and minimal success in crystallization trials. Using an ion/pH solubility screen, it was possible to dramatically increase the maximum solubility of the protein. The solubility-optimized protein produced large diffraction-quality crystals under multiple conditions in which the non-optimized protein produced only precipitate. Only minor adjustments of the conditions were required to produce single diffraction-quality crystals. The native Ape FEN-1 crystals diffract to 1.4 A resolution and belong to space group P6(1), with unit-cell parameters a = b = 92.8, c = 80.9 A, alpha = beta = 90, gamma = 120 degrees.


Journal of Biological Chemistry | 2004

Bacteriophage T4 32 Protein Is Required for Helicase-dependent Leading Strand Synthesis When the Helicase Is Loaded by the T4 59 Helicase-loading Protein

Charles E. Jones; Timothy C. Mueser; Nancy G. Nossal

In the bacteriophage T4 DNA replication system, T4 gene 59 protein binds preferentially to fork DNA and accelerates the loading of the T4 41 helicase. 59 protein also binds the T4 32 single-stranded DNA-binding protein that coats the lagging strand template. Here we explore the function of the strong affinity between the 32 and 59 proteins at the replication fork. We show that, in contrast to the 59 helicase loader, 32 protein does not bind forked DNA more tightly than linear DNA. 32 protein displays a strong binding polarity on fork DNA, binding with much higher affinity to the 5′ single-stranded lagging strand template arm of a model fork, than to the 3′ single-stranded leading strand arm. 59 protein promotes the binding of 32 protein on forks too short for cooperative binding by 32 protein. We show that 32 protein is required for helicase-dependent leading strand DNA synthesis when the helicase is loaded by 59 protein. However, 32 protein is not required for leading strand synthesis when helicase is loaded, less efficiently, without 59 protein. Leading strand synthesis by wild type T4 polymerase is strongly inhibited when 59 protein is present without 32 protein. Because 59 protein can load the helicase on forks without 32 protein, our results are best explained by a model in which 59 helicase loader at the fork prevents the coupling of the leading strand polymerase and the helicase, unless the position of 59 protein is shifted by its association with 32 protein.


Analytical Biochemistry | 1991

Use of dual wavelength spectrophotometry and continuous enzymatic depletion of oxygen for determination of the oxygen binding constants of hemoglobin

Todd M. Larsen; Timothy C. Mueser; Lawrence J. Parkhurst

A small stopped-flow cuvette was built into a computer-controlled Cary 210 spectrophotometer. The enzymatic depletion of oxygen in solutions of hemoglobin and myoglobin was initiated by flowing the hemeproteins with the enzyme against a solution of the hemeproteins containing the appropriate substrate. The deoxygenation was homogeneous throughout the solution. Oxygen activity was calculated at each instant of time from the fractional saturation of Mb, determined from observations at the Hb/HbO2 isosbestic wavelength. Fractional saturation of Hb was determined from absorbances at the Mb/MbO2 isosbestic wavelength. The spectrophotometer cycled between these two wavelengths during the deoxygenation. The deoxygenation of HbO2 was largely complete in 20-25 min, whereas the deoxygenation of MbO2 was allowed to proceed for about 1 h. This procedure eliminates equilibration of Hb solutions with a gas phase and replaces oxygen electrode readings with spectrophotometric sensing by Mb, providing essentially instantaneous determinations of oxygen activity and hence 250-500 or more independent data points per run. The Mb and Hb data vectors require several manipulations to correct for small relative displacements in time and for small non-isosbestic effects. Detailed consideration of the enzyme kinetics allowed oxygen activities to be determined in regions where Mb is a poor sensor. Studies of HbO2 deoxygenation as a function of wavelength show that the determination of the four Adair constants requires in addition the determination of three spectroscopic parameters. Values of the apparent Adair constants, determined without these spectroscopic parameters, depend strongly on the monitoring wavelength.


Acta Crystallographica Section D-biological Crystallography | 2006

Assessment of a preliminary solubility screen to improve crystallization trials: uncoupling crystal condition searches

Aude Izaac; Constance A. Schall; Timothy C. Mueser

The utility of a preliminary solubility screen has been assessed on ten test proteins. It is proposed that maximizing the protein solubility prior to crystal setups is likely to improve crystal growth. In crystallization setups, drops of a protein solution are mixed with various crystallization solutions which are then allowed to equilibrate. The protein solutions usually contain a salt and buffer which are present as a constant in all crystal screens. The propensity for crystallization, driven by three components of sparse-matrix screens, the buffers, salts and precipitating agents, could potentially be masked by the components of the protein solution. Ten test proteins were dissolved in a standard buffer (100 mM NaCl, 50 mM Tris-HCl pH 7.5) and in customized optimal buffers determined to maximize solubility. The proteins were then subjected to the Index (Hampton Research) 96-well sparse-matrix crystal screen and to a precipitant/precipitant-additive screen described here. Five of the ten proteins studied showed twofold to fourfold increases in the saturation level from standard to optimal buffer, two showed slight improvement and three showed a slight decrease. Microcrystals were obtained for all proteins and optimal buffer increased the appearance of crystals for eight of the ten proteins.


Journal of Biological Chemistry | 2017

Direct Evidence That an Extended Hydrogen Bonding Network Influences Activation of Pyridoxal 5′-Phosphate in Aspartate Aminotransferase

Steven Dajnowicz; Jerry M. Parks; Xiche Hu; Korie Gesler; Andrey Kovalevsky; Timothy C. Mueser

Pyridoxal 5′-phosphate (PLP) is a fundamental, multifunctional enzyme cofactor used to catalyze a wide variety of chemical reactions involved in amino acid metabolism. PLP-dependent enzymes optimize specific chemical reactions by modulating the electronic states of PLP through distinct active site environments. In aspartate aminotransferase (AAT), an extended hydrogen bond network is coupled to the pyridinyl nitrogen of the PLP, influencing the electrophilicity of the cofactor. This network, which involves residues Asp-222, His-143, Thr-139, His-189, and structural waters, is located at the edge of PLP opposite the reactive Schiff base. We demonstrate that this hydrogen bond network directly influences the protonation state of the pyridine nitrogen of PLP, which affects the rates of catalysis. We analyzed perturbations caused by single- and double-mutant variants using steady-state kinetics, high resolution X-ray crystallography, and quantum chemical calculations. Protonation of the pyridinyl nitrogen to form a pyridinium cation induces electronic delocalization in the PLP, which correlates with the enhancement in catalytic rate in AAT. Thus, PLP activation is controlled by the proximity of the pyridinyl nitrogen to the hydrogen bond microenvironment. Quantum chemical calculations indicate that Asp-222, which is directly coupled to the pyridinyl nitrogen, increases the pKa of the pyridine nitrogen and stabilizes the pyridinium cation. His-143 and His-189 also increase the pKa of the pyridine nitrogen but, more significantly, influence the position of the proton that resides between Asp-222 and the pyridinyl nitrogen. These findings indicate that the second shell residues directly enhance the rate of catalysis in AAT.


Journal of Biological Chemistry | 2012

Models for the Binary Complex of Bacteriophage T4 Gp59 Helicase Loading Protein GP32 SINGLE-STRANDED DNA-BINDING PROTEIN AND TERNARY COMPLEX WITH PSEUDO-Y JUNCTION DNA

Jennifer M. Hinerman; J. David Dignam; Timothy C. Mueser

Background: Accessory proteins assist replicative helicases single-stranded binding protected DNA. Results: Small angle x-ray scattering molecular envelopes allows modeling of the gp59-gp32-DNA complex. Conclusion: The core and A-domains of gp32 protein interact with the C-terminal domain of gp59 protein near the identified binding site for gp41 helicase. Significance: Directional loading of replicative helicases requires structure-specific recognition of the DNA replication fork. Bacteriophage T4 gp59 helicase assembly protein (gp59) is required for loading of gp41 replicative helicase onto DNA protected by gp32 single-stranded DNA-binding protein. The gp59 protein recognizes branched DNA structures found at replication and recombination sites. Binding of gp32 protein (full-length and deletion constructs) to gp59 protein measured by isothermal titration calorimetry demonstrates that the gp32 protein C-terminal A-domain is essential for protein-protein interaction in the absence of DNA. Sedimentation velocity experiments with gp59 protein and gp32ΔB protein (an N-terminal B-domain deletion) show that these proteins are monomers but form a 1:1 complex with a dissociation constant comparable with that determined by isothermal titration calorimetry. Small angle x-ray scattering (SAXS) studies indicate that the gp59 protein is a prolate monomer, consistent with the crystal structure and hydrodynamic properties determined from sedimentation velocity experiments. SAXS experiments also demonstrate that gp32ΔB protein is a prolate monomer with an elongated A-domain protruding from the core. Fitting structures of gp59 protein and the gp32 core into the SAXS-derived molecular envelope supports a model for the gp59 protein-gp32ΔB protein complex. Our earlier work demonstrated that gp59 protein attracts full-length gp32 protein to pseudo-Y junctions. A model of the gp59 protein-DNA complex, modified to accommodate new SAXS data for the binary complex together with mutational analysis of gp59 protein, is presented in the accompanying article (Dolezal, D., Jones, C. E., Lai, X., Brister, J. R., Mueser, T. C., Nossal, N. G., and Hinton, D. M. (2012) J. Biol. Chem. 287, 18596–18607).


Journal of Applied Crystallography | 2005

Rapid preparation of custom grid screens for crystal growth optimization

Anne Benedict Senger; Timothy C. Mueser

Initial crystallization conditions are typically discovered using commercially available sparse-matrix screens. Positive results are then refined using some type of expansion tray. For example, coarse and shallow gradients can be prepared which vary single chemical parameters around the initial conditions in a grid-screen format. Expansion trays are customarily formulated using numerous volume calculations and pipetting stock solutions into individual wells. This tedious process is plagued by pipetting errors, including differences in viscosity, small volumes with large dilutions, evaporation and poor mixing. Here we present a simple method to standardize expansion-tray formats. Instead of using independent well-volume calculations, the initial and final conditions are formulated and gradients (A/B gradients) are prepared using standardized pipetting maps. These step gradients can be prepared by adding decreasing amounts of the initial (A) solution to consecutive wells followed by the addition of the final (B) solution with increasing volume. This simple idea can be applied to both coarse and shallow grids where the pipetting errors are confined within the boundaries defined by the A and B solutions. Programmable electronic pipettes and robotic liquid handlers can be used to prepare the standardized A/B gradients rapidly, regardless of the components, eliminating reprogramming between trays.


Nature Communications | 2017

Direct visualization of critical hydrogen atoms in a pyridoxal 5′-phosphate enzyme

Steven Dajnowicz; Ryne C. Johnston; Jerry M. Parks; Matthew P. Blakeley; David A. Keen; Kevin L. Weiss; Oksana O. Gerlits; Andrey Kovalevsky; Timothy C. Mueser

Enzymes dependent on pyridoxal 5′-phosphate (PLP, the active form of vitamin B6) perform a myriad of diverse chemical transformations. They promote various reactions by modulating the electronic states of PLP through weak interactions in the active site. Neutron crystallography has the unique ability of visualizing the nuclear positions of hydrogen atoms in macromolecules. Here we present a room-temperature neutron structure of a homodimeric PLP-dependent enzyme, aspartate aminotransferase, which was reacted in situ with α-methylaspartate. In one monomer, the PLP remained as an internal aldimine with a deprotonated Schiff base. In the second monomer, the external aldimine formed with the substrate analog. We observe a deuterium equidistant between the Schiff base and the C-terminal carboxylate of the substrate, a position indicative of a low-barrier hydrogen bond. Quantum chemical calculations and a low-pH room-temperature X-ray structure provide insight into the physical phenomena that control the electronic modulation in aspartate aminotransferase.Pyridoxal 5’-phosphate (PLP) is a ubiquitous co factor for diverse enzymes, among them aspartate aminotransferase. Here the authors use neutron crystallography, which allows the visualization of the positions of hydrogen atoms, and computation to characterize the catalytic mechanism of the enzyme.

Collaboration


Dive into the Timothy C. Mueser's collaboration.

Top Co-Authors

Avatar

Nancy G. Nossal

National Institutes of Health

View shared research outputs
Top Co-Authors

Avatar

Charles E. Jones

National Institutes of Health

View shared research outputs
Top Co-Authors

Avatar

Andrey Kovalevsky

Oak Ridge National Laboratory

View shared research outputs
Top Co-Authors

Avatar

C. Craig Hyde

National Institutes of Health

View shared research outputs
Top Co-Authors

Avatar
Top Co-Authors

Avatar

Lawrence J. Parkhurst

University of Nebraska–Lincoln

View shared research outputs
Top Co-Authors

Avatar

Bijan Ahvazi

National Institutes of Health

View shared research outputs
Top Co-Authors

Avatar
Top Co-Authors

Avatar

Jerry M. Parks

Oak Ridge National Laboratory

View shared research outputs
Top Co-Authors

Avatar
Researchain Logo
Decentralizing Knowledge