John W. Drake's research while affiliated with National Institute of Environmental Health Sciences and other places

What is this page?


This page lists the scientific contributions of an author, who either does not have a ResearchGate profile, or has not yet added these contributions to their profile.

It was automatically created by ResearchGate to create a record of this author's body of work. We create such pages to advance our goal of creating and maintaining the most comprehensive scientific repository possible. In doing so, we process publicly available (personal) data relating to the author as a member of the scientific community.

If you're a ResearchGate member, you can follow this page to keep up with this author's work.

If you are this author, and you don't want us to display this page anymore, please let us know.

Publications (109)


Rates of Spontaneous Mutation: Insights Gained Over the Last Half Century
  • Chapter

January 2016

·

18 Reads

·

1 Citation

John W. Drake

Nikolai Timofeeff-Ressovsky understood well the need for both explicit theory and quantitation in biology. His adventures with Karl Zimmer and Max Delbrück and the somewhat romantic portrayal of those ideas by Erwin Schrödinger contributed notably to the development of population genetics and led to the modern theory of mutation. A central mystery in Timofeeff’s time was the size and composition of the gene, which he probed by the methods of radiation mutagenesis. A subsequent central mystery has been whether order may somehow underlie the apparent chaos of mutation rates . Although the first hints of order appeared in the late 1960s, the robustness of certain formulations of mutation rates did not become apparent until the 1990s. It is now clear that each of four major groups of organisms has its own characteristic rate of spontaneous mutation. The riboviruses hover at the edge of mutational meltdown, the retroelements live a few-fold less dangerously, the DNA-based microbes maintain a very small genomic rate (except in special circumstances), and the higher eukaryotes seem to have adopted a rate only a few-fold higher than the DNA-based microbes (with remarkable consequences over the course of a sexual generation). The evolutionary forces driving these characteristic rates are poorly understood.

Share

Confounders of mutation-rate estimators: Selection and phenotypic lag in Thermus thermophilus

August 2013

·

25 Reads

·

12 Citations

Mutation Research/Fundamental and Molecular Mechanisms of Mutagenesis

In a recent description of the rate and character of spontaneous mutation in the hyperthermophilic bacterium Thermus thermophilus, the mutation rate was observed to be substantially lower than seen in several mesophiles. Subsequently, a report appeared indicating that this bacterium maintains an average of about 4.5 genomes per cell. This number of genomes might result in a segregation lag for the expression of a recessive mutation and might therefore lead to an underestimate of the rate of mutation. Here we describe some kinds of problems that may arise when estimating mutation rates and outline ways to adjust the rates accordingly. The emphasis is mainly on differential rates of growth of mutants versus their parents and on various kinds of phenotypic lag. We then apply these methods to the T. thermophilus data and conclude that there is as yet no reliable impact on a previously described rate.


PFU production dynamics of a wt Qß population during infection of E. coli cells. Circles represent relative free PFU densities. Triangles represent relative total (free + intracellular) PFU densities. Separate symbols at each time represent different observations. Free and total PFU densities were determined through one-step and intracellular growth assays, respectively, as detailed in §4. The solid and dashed black curves represent sigmoid curves independently fitted to the one-step (R² = 0.9395, d.f. = 19) and intracellular (R² = 0.8057, d.f. = 51) growth curves, respectively. See text for additional details.
Evolution of Qß PFU production in Treatments (lines TA, TB and TC) and Controls (lines CA, CB and CC) during 26 serial passages. (a) PFU yield obtained for each of the lines at each passage. The bottom half of the graph shows the logarithms of the estimated amounts of PFUs added to each experimental culture at the beginning of each passage, while the top half displays the logarithms of the estimated amounts of PFUs produced after the time allowed for infection (55 min for treatments and 80 min for controls). (b) Evolution of the amounts of PFUs produced by the treatment lines relative to the corresponding amounts produced by the controls at each passage. Different symbols represent different treatment lines as indicated in figure 2a. The dashed curve represents an exponential curve fitted to the data (R² = 0.6432, d.f. = 73). See text for additional details.
Relationship between infectivity and adsorption in Qß. Different symbols represent observations obtained for different populations: the clear circles, squares and rhomboids stand for TA25, TB25 and TC25, respectively, while the black upright and inverted triangles stand for Controls25 and WT, respectively. Adsorption rate constants have been multiplied by 10⁸. Infectivity values represent arcsine transformations of the original percentage values. Correlation analysis between these two parameters indicated, as the figure suggests, a significant positive correlation between them (see text for further details).
PFU stability (expressed as percentage) of the wt and evolved Qß populations. Bars represent means ± s.e.m. (n = 6).
Relationship between PFU stability and latent period in Qß. Different symbols represent PFU stability and latent period means obtained for different populations: the clear circles, squares and rhomboids stand for TA25, TB25 and TC25, respectively, while the black upright and inverted triangles stand for Controls25 and WT, respectively. For each symbol, bars stand for the s.e.m. of the corresponding parameter (PFU stability and latent period). The numbers of observations per population are the same as those indicated in table 1 and figure 4 for latent period and PFU stability, respectively. PFU stability values represent arcsine transformations of the original percentage values. The dashed black curve represents the nonlinear regression of PFU stability on latent period (nonlinear lines regression, R² = 0.6463, d.f. = 28).

+1

Experimental selection reveals a trade-off between fecundity and lifespan in the coliphage Qß
  • Article
  • Full-text available

June 2013

·

86 Reads

·

13 Citations

Open Biology

Open Biology

Understanding virus evolution is key for improving ways to counteract virus-borne diseases. Results from comparative analyses have previously suggested a trade-off between fecundity and lifespan for viruses that infect the bacterium Escherichia coli (i.e. for coliphages), which, if confirmed, would define a particular constraint on the evolution of virus fecundity. Here, the occurrence of such a trade-off is investigated through a selection experiment using the coliphage Qß. Selection was applied for increased fecundity in three independent wild-type Qß populations, and the ability of the virions to remain viable outside the host was determined. The Qß life-history traits involved in the evolution of fecundity and the genetic changes associated with this evolution were also investigated. The results reveal that short-term evolution of increased fecundity in Qß was associated with decreased viability of phage virions. This trade-off apparently arose because fecundity increased at the expense of reducing the amount of resources (mainly time) invested per produced virion. Thus, the results also indicate that Qß fecundity may be enhanced through increases in the rates of adsorption to the host and progeny production. Finally, genomic sequencing of the evolved populations pinpointed sequences likely to be involved in the evolution of Qß fecundity.

Download






Figure S2

July 2012

·

5 Reads

One-step curves for RTIN and RTSUB. Three independent curves were obtained per RT mutant. In all cases, the first bursts appeared between 30 and 40 min after infection. If the 20 min allowed for adsorption are also considered, it means that Qß requires a minimum of 50 to 60 min to complete an infection cycle in RTH cells. Thus, if any of the progeny released in the first bursts would have immediately infected a new cell, the first second-generation bursts might have been expected 100 min after the first infection. Indeed, the curves show a slight increase of Qß density at that time. For this reason, values collected 100 min after infection (empty symbols) were not considered in the curve-fitting analyses conducted to characterize RTIN and RTSUB single-burst dynamics. Squares represent outliers that were automatically excluded from the analyses, and the resulting fitted curves are shown as dashed curves. PFU = plaque-forming units. See Table S1 for additional information. (TIF)



Citations (76)


... There are two factors worth noting that theoretically establish a minimal limit to mutation rate. First, physiological limitations of DNA repair and replication probably prevent reduction of the genomic mutation rate to zero (Drake 1990(Drake , 1993Kondrashov 1995); it could be energetically costly or physicochemically impossible to prevent all replication errors. A second factor that might prevent complete reduction of the mutation rate is the necessity of variation; ...

Reference:

The Evolution of a High Mutation Rate and Declining Fitness in Asexual Populations
Evolving Mutation Rates and Prospects for Antimutagenesis
  • Citing Chapter
  • January 1990

... As Jan Drake polymerases. A simplified sketch of this polymeraseproofreading model is presented in Figure 2. has pointed out, it was fortunate that A·T→G ·C mutations were investigated early on for the tsL141 allele, The model treats polymerization and proofreading as two possible outcomes of a series of random events, otherwise it may not have been identified as an antimutator (Drake 1992). which take place after a dNTP (right or wrong) binds to the enzyme. ...

Roots: Mutation Rates
  • Citing Article
  • February 1992

BioEssays

... The thermotolerant capacity of different microbial species induced by various horizontal gene transfer mechanisms has been reported such as conjugation [43], natural transformation [44], transduction [45], gene transfer through membrane vesicles, nanotubes and nano pods [46]. Analysis of genome mutation studies in Thermus thermophilus and Sulfolobus acidocaldarius showed a lower frequency of base substitution occurs in thermophilic when compare with the mesophilic organisms [47]. These results suggest that a stringent DNA repair mechanism is maintained in the thermophilic organisms for the stability of their genome. ...

Confounders of mutation-rate estimators: Selection and phenotypic lag in Thermus thermophilus
  • Citing Article
  • August 2013

Mutation Research/Fundamental and Molecular Mechanisms of Mutagenesis

... Finally, we found a correlation between survival (low decay rate m) and reproduction (high relative growth, Fig. 4B), as in previous studies with phages (correlations: De Paepe & Taddei, 2006;Dessau et al., 2012;confirmed trade-offs: Heineman & Brown, 2012;García-Villada & Drake, 2013). Interestingly, we did not find a correlation when reproduction was calculated as the multiplication rate following the classic example in De Paepe & Taddei (2006) (Fig. 4B). ...

Experimental selection reveals a trade-off between fecundity and lifespan in the coliphage Qß
Open Biology

Open Biology

... We compromised on a readily accessible title, followed by "De mutatis. . ." as a subtitle [17]. This was long my favorite title, joined later by my favorite running head, composed for a paper in the JBC that dissected a mutator DNA polymerase using fidelity assays based most on in vivo systems: In Vivo Veritas. ...

Updating the Theory of Mutation
  • Citing Article
  • October 1983

American Scientist

... Replication mode of some RNA viruses, such as vesicular stomatitis virus [12] and the poliovirus [88] is quite close to GR, with multiple rounds of RNA copying inside the cell. Other RNA viruses, such as bacteriophages Qβ [36] and φ6 [8], as well as the turnip mosaic virus [64], replicate in a way that is very reminiscent of SMR. At the same time, since many other viruses operate in a regime somewhere between those two extremes, hence, it is convenient to represent the mode of replication using a continuous parameter 0 < α ≤ 1 [85], with α 0 corresponding to SMR, and 0 < α ≤ 1 describing GR, with α = 1 being a situation, where both positive and negative strands replicate at the same rate. ...

The Three Faces of Riboviral Spontaneous Mutation: Spectrum, Mode of Genome Replication, and Mutation Rate

... When rates of spontaneous mutation are expressed per genome replication, different broad groups of organisms display characteristic values [31,32]: of the order of 0.001 for DNA-based microbes (including both viral and cellular organisms); of the order of 0.01 for higher eukaryotes; from the order of 0.01 to the order of 0.1 for retroviruses and from the order of 0.1 to the order of 1 for RNA viruses exclusive of retroviruses [47,88,3,30]. This wide range of mutation rates of RNA viruses suggests that it is inappropriate to gather RNA viruses together into a single group that is subject to different evolutionary rules than organisms with DNA genomes. ...

A Test of Kimura’s Mutation-Rate Conjecture
  • Citing Chapter
  • January 1970

NATO Security through Science Series C: Environmental Security

... It has been observed that mutation rate estimates can differ between MA experiments and specific locus methods, even with carefully chosen reporter genes (Table 1), and these differences are often significant (paired Student's ttest, two-tailed P = 0.029) (Drake 2012). These discrepancies likely arise in part because mutation rates vary among individual nucleotide types (Long et al. , 2018Kucukyildirim et al. 2016), depending on their neighboring content (Sung et al. 2012), and vary among chromosomal regions (Long et al. 2014;Dillon et al. 2015;Sung et al. 2015;Niccum et al. 2019). ...

Contrasting Mutation Rates from Specific-Locus and Long-Term Mutation-Accumulation Procedures

G3 Genes Genomes Genetics

... The T4 phage strain used in this study is highly virulent and rapidly lyses the cell after infection (it is an obligate lytic, [2]). Phage T4r (T4 rapid) lacks the lysis inhibition (LIN) genes of wild-type T4 and rapidly lyses bacteria, even multiply infected "super-infected" bacteria, with resultant relatively low titer yields compared to wild-type T4 [3]. Because of the rapid lysis of T4r infected bacteria any survival of a low number N O of bacteria can only be due to either previously acquired resistance (Darwinian random mutations) or a kind of rapid response by the stress-challenged bacteria, which can be called Lamarckian in a certain sense of the word. ...

The Bacteriophage T4 Rapid-Lysis Genes and Their Mutational Proclivities

·

Leilei Zhang

·

Frank G. Chao

·

[...]

·

John W. Drake

... Despite the information learned from structural and biochemical studies of DNA polymerases, Jan concluded that fidelities of wild-type and mutant RB69 DNA polymerases determined in vitro only partially reflected their fidelities in vivo (Bebenek et al. 2002). For example, biochemical studies of engineered RB69 DNA polymerases showed that increasing the size of the nucleotide-binding pocket reduced nucleotide insertion fidelity, but that was countered by in vivo mutation studies (Trzemecka et al. 2010) that forced reconsideration of the biochemical studies (Xia et al. 2011). Hence, there is a real and continuing need for in vivo replication fidelity experiments. ...

Reversal of a Mutator Activity by a Nearby Fidelity-Neutral Substitution in the RB69 DNA Polymerase Binding Pocket

Journal of Molecular Biology