Network


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

Hotspot


Dive into the research topics where Jane M. Sayer is active.

Publication


Featured researches published by Jane M. Sayer.


The EMBO Journal | 2002

Structures of HIV‐1 reverse transcriptase with pre‐ and post‐translocation AZTMP‐terminated DNA

Stefan G. Sarafianos; Arthur D. Clark; Kalyan Das; Steve Tuske; Jens J. Birktoft; Palanichamy Ilankumaran; Andagar R. Ramesha; Jane M. Sayer; Donald M. Jerina; Paul L. Boyer; Stephen H. Hughes; Eddy Arnold

AZT (3′‐azido‐3′‐deoxythymidine) resistance involves the enhanced excision of AZTMP from the end of the primer strand by HIV‐1 reverse transcriptase. This reaction can occur when an AZTMP‐terminated primer is bound at the nucleotide‐binding site (pre‐translocation complex N) but not at the ‘priming’ site (post‐translocation complex P). We determined the crystal structures of N and P complexes at 3.0 and 3.1 Å resolution. These structures provide insight into the structural basis of AZTMP excision and the mechanism of translocation. Docking of a dNTP in the P complex structure suggests steric crowding in forming a stable ternary complex that should increase the relative amount of the N complex, which is the substrate for excision. Structural differences between complexes N and P suggest that the conserved YMDD loop is involved in translocation, acting as a springboard that helps to propel the primer terminus from the N to the P site after dNMP incorporation.


Journal of Biological Chemistry | 2002

Preferential Misincorporation of Purine Nucleotides by Human DNA Polymerase η Opposite Benzo[a]pyrene 7,8-Diol 9,10-Epoxide Deoxyguanosine Adducts

Dominic Chiapperino; Heiko Kroth; Irene H. Kramarczuk; Jane M. Sayer; Chikahide Masutani; Fumio Hanaoka; Donald M. Jerina; Albert M. Cheh

Human DNA polymerase η was used to copy four stereoisomeric deoxyguanosine (dG) adducts derived from benzo[a]pyrene 7,8-diol 9,10-epoxide (diastereomer with the 7-hydroxyl group and epoxide oxygen trans (BaP DE-2)). The adducts, formed by either cis or trans epoxide ring opening of each enantiomer of BaP DE-2 by N 2 of dG, were placed at the fourth nucleotide from the 5′-end in two 16-mer sequence contexts, 5′∼CG*A∼ and 5′∼GG*T. polη was remarkably error prone at all four diol epoxide adducts, preferring to misincorporate G and A at frequencies 3- to more than 50-fold greater than the frequencies for T or the correct C, although the highest rates were 60-fold below the rate of incorporation of C opposite a non-adducted G. Antito syn rotation of the adducted base, consistent with previous NMR data for a BaP DE-2 dG adduct placed just beyond a primer terminus, provides a rationale for preferring purine misincorporation. Extension of purine misincorporations occurred preferentially, but extension beyond the adduct site was weak withV max/K m values generally 10-fold less than for misincorporation. Mostly A was incorporated opposite (+)-BaP DE-2 dG adducts, which correlates with published observations that G → T is the most common type of mutation that (+)-BaP DE-2 induces in mammalian cells.


Advances in Experimental Medicine and Biology | 1991

Covalent Bonding of Bay-Region Diol Epoxides to Nucleic Acids

Donald M. Jerina; Anju Chadha; Albert M. Cheh; Mark E. Schurdak; Alexander W. Wood; Jane M. Sayer

Although the solution chemistry of diol epoxides is now fairly well understood, a great deal remains to be elucidated regarding their reaction in the presence of DNA. Not only DNA but also small molecules are capable of sequestering diol epoxides in aqueous solutions with equilibrium constants on the order of 10(2)-10(4) M-1. In the case of DNA, at least two major families of complexes are presently recognized, possibly the result of groove binding vs. intercalation. As is the case for diol epoxides free in solution, the complexed diol epoxides undergo solvolysis to tetraols and in some cases possibly to keto diols as well. Fractionation between covalent bonding and solvolysis from within the complex(s) is determined more by the nature of the parent hydrocarbon from which the diol epoxide is derived than any other factor. Studies of a wide variety of alkylating and arylating agents have show that practically every potentially nucleophilic site on DNA can serve as a target for modification. In the case of the diol epoxides, practically all of the modification occurs at the exocyclic amino groups of the purine bases. In contrast to the diol epoxides, other epoxides such as those derived from aflatoxin B1, vinyl chloride, propylene, 9-vinylanthracene, and styrene preferentially bind to the aromatic ring nitrogens N-7 in guanine and N-3 in adenine (cf. Chadha et al., 1989). Molecular modeling as well as the spectroscopic evidence suggests that the hydrocarbon portion of the diol epoxides lies in the minor groove of DNA when bound to the exocyclic 2-amino group of guanine and in the major groove when bound to the exocyclic 6-amino group of adenine. Detailed conformational analysis of adducted DNA should prove to be extremely valuable in developing mechanistic models for the enzymatic processing of chemically altered DNA. At present, the critical lesion or lesions responsible for induction of neoplasia remains obscured by the large number of apparently noncritical adducts which form when polycyclic hydrocarbon diol epoxides bond to DNA.


Proceedings of the National Academy of Sciences of the United States of America | 2007

A structural gap in Dpo4 supports mutagenic bypass of a major benzo[a]pyrene dG adduct in DNA through template misalignment

Jacob Bauer; Guangxin Xing; Haruhiko Yagi; Jane M. Sayer; Donald M. Jerina; Hong Ling

Erroneous replication of lesions in DNA by DNA polymerases leads to elevated mutagenesis. To understand the molecular basis of DNA damage-induced mutagenesis, we have determined the x-ray structures of the Y-family polymerase, Dpo4, in complex with a DNA substrate containing a bulky DNA lesion and incoming nucleotides. The DNA lesion is derived from an environmentally widespread carcinogenic polycyclic aromatic hydrocarbon, benzo[a]pyrene (BP). The potent carcinogen BP is metabolized to diol epoxides that form covalent adducts with cellular DNA. In the present study, the major BP diol epoxide adduct in DNA, BP-N2-deoxyguanosine (BP–dG), was placed at a template–primer junction. Three ternary complexes reveal replication blockage, extension past a mismatched lesion, and a −1 frameshift mutation. In the productive structures, the bulky adduct is flipped/looped out of the DNA helix into a structural gap between the little finger and core domains. Sequestering of the hydrophobic BP adduct in this new substrate-binding site permits the DNA to exhibit normal geometry for primer extension. Extrusion of the lesion by template misalignment allows the base 5′ to the adduct to serve as the template, resulting in a −1 frameshift. Subsequent strand realignment produces a mismatched base opposite the lesion. These structural observations, in combination with replication and mutagenesis data, suggest a model in which the additional substrate-binding site stabilizes the extrahelical nucleotide for lesion bypass and generation of base substitutions and −1 frameshift mutations.


Journal of Biological Chemistry | 2008

Effect of the Active Site D25N Mutation on the Structure, Stability, and Ligand Binding of the Mature HIV-1 Protease

Jane M. Sayer; Fengling Liu; Rieko Ishima; Irene T. Weber; John M. Louis

All aspartic proteases, including retroviral proteases, share the triplet DTG critical for the active site geometry and catalytic function. These residues interact closely in the active, dimeric structure of HIV-1 protease (PR). We have systematically assessed the effect of the D25N mutation on the structure and stability of the mature PR monomer and dimer. The D25N mutation (PRD25N) increases the equilibrium dimer dissociation constant by a factor >100-fold (1.3 ± 0.09 μm) relative to PR. In the absence of inhibitor, NMR studies reveal clear structural differences between PR and PRD25N in the relatively mobile P1 loop (residues 79-83) and flap regions, and differential scanning calorimetric analyses show that the mutation lowers the stabilities of both the monomer and dimer folds by 5 and 7.3 °C, respectively. Only minimal differences are observed in high resolution crystal structures of PRD25N complexed to darunavir (DRV), a potent clinical inhibitor, or a non-hydrolyzable substrate analogue, Ac-Thr-Ile-Nle-r-Nle-Gln-Arg-NH2 (RPB), as compared with PR·DRV and PR·RPB complexes. Although complexation with RPB stabilizes both dimers, the effect on their Tm is smaller for PRD25N (6.2 °C) than for PR (8.7 °C). The Tm of PRD25N·DRV increases by only 3 °C relative to free PRD25N, as compared with a 22 °C increase for PR·DRV, and the mutation increases the ligand dissociation constant of PRD25N·DRV by a factor of ∼106 relative to PR·DRV. These results suggest that interactions mediated by the catalytic Asp residues make a major contribution to the tight binding of DRV to PR.


Biochemical and Biophysical Research Communications | 1989

Continuous spectrophotometric assay for retroviral proteases of HIV-1 and AMV.

Nashaat T. Nashed; John M. Louis; Jane M. Sayer; Ewald M. Wondrak; Peter T. Mora; Stephen Oroszlan; Donald M. Jerina

Ac-Lys-Ala-Ser-Gln-Asn-Phe(NO2)-Pro-Val-Val-NH2 (peptide I) and Thr-Phe-Gln-Ala-Phe(NO2)-Pro-Leu-Arg-Glu-Ala (peptide II) undergo hydrolysis between the p-nitrophenylalanyl and prolyl residues catalyzed by the proteases of HIV-1 and AMV, respectively. The specific hydrolyses of peptides I and II are accompanied by a decrease in their uv absorption at 269 nm (delta epsilon = 1000) and an increase at 316 nm (delta epsilon = 600). The use of microspectrophotometric cells allows continuous uv measurements on a volume (60 to 120 microliters) comparable to that required for the HPLC point assay currently used. At the highest substrate concentration possible under the assay conditions, good first-order kinetics were observed with both proteases, and the values of Vmax/Km were obtained.


Proceedings of the National Academy of Sciences of the United States of America | 2011

Inhibition of autoprocessing of natural variants and multidrug resistant mutant precursors of HIV-1 protease by clinical inhibitors.

John M. Louis; Annie Aniana; Irene T. Weber; Jane M. Sayer

Self-cleavage at the N terminus of HIV-1 protease from the Gag-Pol precursor (autoprocessing) is crucial for stabilizing the protease dimer required for onset of mature-like catalytic activity, viral maturation, and propagation. Among nine clinical protease inhibitors (PIs), darunavir and saquinavir were the most effective in inhibiting wild-type HIV-1 group M precursor autoprocessing, with an IC50 value of 1–2 μM, 3–5 orders of magnitude higher than their binding affinities to the corresponding mature protease. Accordingly, both group M and N precursor–PI complexes exhibit Tms 17–21 °C lower than those of the corresponding mature protease–PI complexes suggestive of markedly reduced stabilities of the precursor dimer–PI ensembles. Autoprocessing of group N (natural variant) and three group M precursors bearing 11–20 mutations associated with multidrug resistance was either weakly responsive or fully unresponsive to inhibitors at concentrations up to a practical limit of approximately 150 μM PI. This observation parallels decreases of up to 8 × 103-fold (e.g., 5 pM to 40 nM) in the binding affinity of darunavir and saquinavir to mature multidrug resistant proteases relative to wild type, suggesting that inhibition of some of these mutant precursors will occur only in the high μM to mM range in extreme PI-resistance, which is an effect arising from coordinated multiple mutations. An extremely darunavir-resistant mutant precursor is more responsive to inhibition by saquinavir. These findings raise the questions whether clinical failure of PI therapy is related to lack of inhibition of autoprocessing and whether specific inhibitors can be designed with low-nM affinity to target autoprocessing.


Advances in Experimental Medicine and Biology | 1986

Reactivity and Tumorigenicity of Bay-Region Diol Epoxides Derived from Polycyclic Aromatic Hydrocarbons

Donald M. Jerina; Jane M. Sayer; S. K. Agarwal; Haruhiko Yagi; Wayne Levin; A. W. Wood; A. H. Conney; Donna Pruess-Schwartz; W. M. Baird; M. A. Pigott; Anthony Dipple

During the past decade substantial progress has been made in elucidating factors that determine the tumorigenic activity of bay-region diol epoxides, major ultimate carcinogenic metabolites derived from polycyclic aromatic hydrocarbons. Neither high nor low chemical reactivity of the diol epoxides (as measured by rates of uncatalyzed solvolysis) is required for high tumorigenic response. In contrast, aspects of molecular structure such as conformation and absolute configuration strongly influence tumorigenic activity. The role of conformation is illustrated by the observation that those diol epoxides whose hydroxyl groups are pseudoaxial are weak or inactive as tumorigens. Absolute configuration is an important determinant of biological activity of bay-region diol epoxides: in all cases studied to date, the predominantly formed (R,S)-diol-(S,R)-epoxides are generally the most tumorigenic of the four metabolically possible configurational isomers. In the course of investigating the effects of structural factors on tumorigenic activity, we identified the (4R,3S)-diol-(2S,1R)-epoxide of benzo(c)phenanthrene as the most potent tumorigen (in initiation-promotion experiments on mouse skin) of the diol epoxides studied to date. Studies of all four configurationally isomeric diol epoxides derived from benzo(c)phenanthrene led to the striking observation that these diol epoxides exhibit an exceptionally high efficiency of covalent binding, relative to hydrolysis, when allowed to react with calf thymus DNA in aqueous solution. Thus, these diol epoxides should provide an excellent tool for the detailed study of such binding. When the four isomeric benzo(c)phenanthrene diol epoxides are compared, there appears to be no simple correlation between tumorigenic response and either the extent of binding to DNA or the major types of deoxyribonucleoside adducts formed. Deoxyribonucleoside adducts of benzo(c)phenanthrene diol epoxide have also been identified from the DNA of cultured rodent embryo cells after treatment of the cells with tritium-labeled benzo(c)phenanthrene. The distribution of adducts is consistent with predominant metabolic formation of the (4R,3S)-diol-(2S,1R)-epoxide; deoxyadenosine is the major site in the cellular DNA attacked by this epoxide, just as it is in DNA in solution. Further experiments are in progress which we hope will identify more subtle aspects of the DNA binding of benzo(c)phenanthrene diol epoxides that may be uniquely correlated with their tumorigenic activity.


Chemico-Biological Interactions | 1983

Effects of pH and salt concentration on the hydrolysis of a benzo[a]pyrene 7,8-diol-9,10-epoxide catalyzed by DNA and polyadenylic acid

D.P. Michaud; S.C. Gupta; Dale L. Whalen; Jane M. Sayer; Donald M. Jerina

The time-dependent absorbance change that occurs when benzo[alpha]pyrene 7,8-diol-9,10-epoxide is added to solutions of calf thymus DNA has been shown, by an unequivocal chromatographic method, to correspond to DNA-catalyzed hydrolysis of the diol-epoxide. At 25 degrees C and mu = 0.10, the kinetics of the reaction of the diol-epoxide with polyadenylic acid or DNA are consistent with preequilibrium formation of a non-covalent complex between the diol-epoxide and the polynucleotide or DNA, followed by hydrolysis of the bound epoxide by a process that is first-order in hydronium ions. Cacodylic acid also catalyzes the hydrolysis of the epoxide bound to polyadenylic acid. The rate of the DNA-catalyzed hydrolysis exhibits little or no enantiomeric selectivity for the diol-epoxide. DNA catalyzed hydrolysis of the diol-epoxide is extraordinarily sensitive to the salt concentration in the reaction medium: the rate of hydrolysis of the bound epoxide at pH 7 is retarded by a factor of approximately 45 in the presence of 0.1 M sodium chloride compared to a 1 mM buffer containing no added salt. Thus, studies of the interactions of DNA with carcinogenic diol-epoxides must take into account the ionic environment of DNA within the cell.


Tetrahedron Letters | 1992

Epoxide and diol epoxide adducts of polycyclic aromatic hydrocarbons at the exocyclic amino group of deoxyguanosine

Barbara Zajc; Mahesh K. Lakshman; Jane M. Sayer; Donald M. Jerina

Abstract Synthesis and separation of the diastereomeric trans N 2 -2′-deoxyguanosine adducts of tetrahydrophenanthrene 3,4-epoxide and benzo[a]pyrene 7,8-diol 9,10-epoxide (benzylic hydroxyl group and epoxide oxygen trans), as well as the incorporation of the former into the pentanucleotide TpApG * pApT, are described.

Collaboration


Dive into the Jane M. Sayer's collaboration.

Top Co-Authors

Avatar

Donald M. Jerina

National Institutes of Health

View shared research outputs
Top Co-Authors

Avatar

Haruhiko Yagi

National Institutes of Health

View shared research outputs
Top Co-Authors

Avatar

Heiko Kroth

National Institutes of Health

View shared research outputs
Top Co-Authors

Avatar
Top Co-Authors

Avatar

Anthony Dipple

National Institutes of Health

View shared research outputs
Top Co-Authors

Avatar

John M. Louis

National Institutes of Health

View shared research outputs
Top Co-Authors

Avatar
Top Co-Authors

Avatar
Top Co-Authors

Avatar

Herman J. C. Yeh

National Institutes of Health

View shared research outputs
Top Co-Authors

Avatar
Researchain Logo
Decentralizing Knowledge