BIOMARKERS

Molecular Biopsy of Human Tumors

- a resource for Precision Medicine *

347 related articles for article (PubMed ID: 11961107)

  • 41. Evolutionary rate variation in two conifer species, Taxodium distichum (L.) Rich. var. distichum (baldcypress) and Cryptomeria japonica (Thunb. ex L.f.) D. Don (Sugi, Japanese cedar).
    Kusumi J; Tsumura Y; Tachida H
    Genes Genet Syst; 2015; 90(5):305-15. PubMed ID: 26687861
    [TBL] [Abstract][Full Text] [Related]  

  • 42. Nonsynonymous, synonymous and nonsense mutations in human cancer-related genes undergo stronger purifying selections than expectation.
    Chu D; Wei L
    BMC Cancer; 2019 Apr; 19(1):359. PubMed ID: 30991970
    [TBL] [Abstract][Full Text] [Related]  

  • 43. Protein evolutionary rates correlate with expression independently of synonymous substitutions in Helicobacter pylori.
    Sällström B; Arnaout RA; Davids W; Bjelkmar P; Andersson SG
    J Mol Evol; 2006 May; 62(5):600-14. PubMed ID: 16586017
    [TBL] [Abstract][Full Text] [Related]  

  • 44. Inferring natural selection operating on conservative and radical substitution at single amino acid sites.
    Suzuki Y
    Genes Genet Syst; 2007 Aug; 82(4):341-60. PubMed ID: 17895585
    [TBL] [Abstract][Full Text] [Related]  

  • 45. Towards realistic codon models: among site variability and dependency of synonymous and non-synonymous rates.
    Mayrose I; Doron-Faigenboim A; Bacharach E; Pupko T
    Bioinformatics; 2007 Jul; 23(13):i319-27. PubMed ID: 17646313
    [TBL] [Abstract][Full Text] [Related]  

  • 46. The evolutionary implications of knox-I gene duplications in conifers: correlated evidence from phylogeny, gene mapping, and analysis of functional divergence.
    Guillet-Claude C; Isabel N; Pelgas B; Bousquet J
    Mol Biol Evol; 2004 Dec; 21(12):2232-45. PubMed ID: 15317878
    [TBL] [Abstract][Full Text] [Related]  

  • 47. Molecular evolution of a small gene family of wound inducible Kunitz trypsin inhibitors in Populus.
    Talyzina NM; Ingvarsson PK
    J Mol Evol; 2006 Jul; 63(1):108-19. PubMed ID: 16755353
    [TBL] [Abstract][Full Text] [Related]  

  • 48. Synonymous rates at the RpII215 gene of Drosophila: variation among species and across the coding region.
    Llopart A; Aguadé M
    Genetics; 1999 May; 152(1):269-80. PubMed ID: 10224259
    [TBL] [Abstract][Full Text] [Related]  

  • 49. Mitochondrial phylogenomics of early land plants: mitigating the effects of saturation, compositional heterogeneity, and codon-usage bias.
    Liu Y; Cox CJ; Wang W; Goffinet B
    Syst Biol; 2014 Nov; 63(6):862-78. PubMed ID: 25070972
    [TBL] [Abstract][Full Text] [Related]  

  • 50. A likelihood approach for comparing synonymous and nonsynonymous nucleotide substitution rates, with application to the chloroplast genome.
    Muse SV; Gaut BS
    Mol Biol Evol; 1994 Sep; 11(5):715-24. PubMed ID: 7968485
    [TBL] [Abstract][Full Text] [Related]  

  • 51. Rates of synonymous substitution and base composition of nuclear genes in Drosophila.
    Moriyama EN; Gojobori T
    Genetics; 1992 Apr; 130(4):855-64. PubMed ID: 1582562
    [TBL] [Abstract][Full Text] [Related]  

  • 52. Synonymous and nonsynonymous polymorphisms versus divergences in bacterial genomes.
    Hughes AL; Friedman R; Rivailler P; French JO
    Mol Biol Evol; 2008 Oct; 25(10):2199-209. PubMed ID: 18667439
    [TBL] [Abstract][Full Text] [Related]  

  • 53. Spatial covariation of mutation and nonsynonymous substitution rates in vertebrate mitochondrial genomes.
    Broughton RE; Reneau PC
    Mol Biol Evol; 2006 Aug; 23(8):1516-24. PubMed ID: 16705079
    [TBL] [Abstract][Full Text] [Related]  

  • 54. Three genome-based phylogeny of Cupressaceae s.l.: further evidence for the evolution of gymnosperms and Southern Hemisphere biogeography.
    Yang ZY; Ran JH; Wang XQ
    Mol Phylogenet Evol; 2012 Sep; 64(3):452-70. PubMed ID: 22609823
    [TBL] [Abstract][Full Text] [Related]  

  • 55. The problem of counting sites in the estimation of the synonymous and nonsynonymous substitution rates: implications for the correlation between the synonymous substitution rate and codon usage bias.
    Bierne N; Eyre-Walker A
    Genetics; 2003 Nov; 165(3):1587-97. PubMed ID: 14668405
    [TBL] [Abstract][Full Text] [Related]  

  • 56. Partitioning the variation in mammalian substitution rates.
    Smith NG; Eyre-Walker A
    Mol Biol Evol; 2003 Jan; 20(1):10-7. PubMed ID: 12519900
    [TBL] [Abstract][Full Text] [Related]  

  • 57. The molecular clock revisited: the rate of synonymous vs. replacement change in Drosophila.
    Zeng LW; Comeron JM; Chen B; Kreitman M
    Genetica; 1998; 102-103(1-6):369-82. PubMed ID: 9720289
    [TBL] [Abstract][Full Text] [Related]  

  • 58. Estimating synonymous and nonsynonymous substitution rates under realistic evolutionary models.
    Yang Z; Nielsen R
    Mol Biol Evol; 2000 Jan; 17(1):32-43. PubMed ID: 10666704
    [TBL] [Abstract][Full Text] [Related]  

  • 59. Molecular evolution of the teosinte branched gene among maize and related grasses.
    Lukens L; Doebley J
    Mol Biol Evol; 2001 Apr; 18(4):627-38. PubMed ID: 11264415
    [TBL] [Abstract][Full Text] [Related]  

  • 60. The effect of tandem substitutions on the correlation between synonymous and nonsynonymous rates in rodents.
    Smith NG; Hurst LD
    Genetics; 1999 Nov; 153(3):1395-402. PubMed ID: 10545467
    [TBL] [Abstract][Full Text] [Related]  

    [Previous]   [Next]    [New Search]
    of 18.