ArticlePDF Available

Yeast Phenomics: An Experimental Approach for Modeling Gene Interaction Networks that Buffer Disease

Authors:

Abstract and Figures

The genome project increased appreciation of genetic complexity underlying disease phenotypes: many genes contribute each phenotype and each gene contributes multiple phenotypes. The aspiration of predicting common disease in individuals has evolved from seeking primary loci to marginal risk assignments based on many genes. Genetic interaction, defined as contributions to a phenotype that are dependent upon particular digenic allele combinations, could improve prediction of phenotype from complex genotype, but it is difficult to study in human populations. High throughput, systematic analysis of S. cerevisiae gene knockouts or knockdowns in the context of disease-relevant phenotypic perturbations provides a tractable experimental approach to derive gene interaction networks, in order to deduce by cross-species gene homology how phenotype is buffered against disease-risk genotypes. Yeast gene interaction network analysis to date has revealed biology more complex than previously imagined. This has motivated the development of more powerful yeast cell array phenotyping methods to globally model the role of gene interaction networks in modulating phenotypes (which we call yeast phenomic analysis). The article illustrates yeast phenomic technology, which is applied here to quantify gene X media interaction at higher resolution and supports use of a human-like media for future applications of yeast phenomics for modeling human disease.
Content may be subject to copyright.
Genes 2015, 6, 24-45; doi:10.3390/genes6010024
genes
ISSN 2073-4425
www.mdpi.com/journal/genes
Article
Yeast Phenomics: An Experimental Approach for Modeling
Gene Interaction Networks that Buffer Disease
John L. Hartman IV 1,*, Chandler Stisher 1, Darryl A. Outlaw 1, Jingyu Guo 1, Najaf A. Shah 1,
Dehua Tian 1, Sean M. Santos 1, John W. Rodgers 1 and Richard A. White 2
1 Department of Genetics, University of Alabama at Birmingham, 730 Hugh Kaul Human Genetics
Building, 720 20th Street South, Birmingham, AL 35294, USA; E-Mails: cstisher@uab.edu (C.S.);
doutlaw@uab.edu (D.A.O.); guo.jingy@gmail.com (J.G.); najafali@gmail.com (N.A.S.);
dehuat@gmail.com (D.T.); ssantos@uab.edu (S.M.S.); jwrodger@uab.edu (J.W.R.)
2 Department of Statistics and Michael Smith Laboratories, University of British Columbia,
3182 Earth Sciences Building, 2207 Main Mall, Vancouver, BC V6T-1Z4, Canada;
E-Mail: rickw@stat.ubc.ca
* Author to whom correspondence should be addressed; E-Mail: jhartman@uab.edu;
Tel.: +1-205-996-4195; Fax: +1-205-975-5689.
Academic Editors: Karen E. Nelson, John Burn, Nicholas J. Schork, James R. Lupski and
Pabulo H. Rampelotto
Received: 20 August 2014 / Accepted: 12 January 2015 / Published: 6 February 2015
Abstract: The genome project increased appreciation of genetic complexity underlying
disease phenotypes: many genes contribute each phenotype and each gene contributes multiple
phenotypes. The aspiration of predicting common disease in individuals has evolved from
seeking primary loci to marginal risk assignments based on many genes. Genetic interaction,
defined as contributions to a phenotype that are dependent upon particular digenic allele
combinations, could improve prediction of phenotype from complex genotype, but it is
difficult to study in human populations. High throughput, systematic analysis of S. cerevisiae
gene knockouts or knockdowns in the context of disease-relevant phenotypic perturbations
provides a tractable experimental approach to derive gene interaction networks, in order to
deduce by cross-species gene homology how phenotype is buffered against disease-risk
genotypes. Yeast gene interaction network analysis to date has revealed biology more
complex than previously imagined. This has motivated the development of more powerful
yeast cell array phenotyping methods to globally model the role of gene interaction networks
in modulating phenotypes (which we call yeast phenomic analysis). The article illustrates
OPEN ACCESS
Genes 2015, 625
yeast phenomic technology, which is applied here to quantify gene X media interaction at
higher resolution and supports use of a human-like media for future applications of yeast
phenomics for modeling human disease.
Keywords: yeast phenomics; yeast models of human disease; cell proliferation phenotypes
or cell proliferation parameters (CPPs); gene interaction networks; quantitative high throughput
cell array phenotyping (Q-HTCP); genetic buffering; cystic fibrosis (CF); human-like (HL)
yeast media; ammonium toxicity; recursive expectation-maximization clustering (REMc)
1. Introduction and Perspective
1.1. Buffering of Phenotypes: Yeast Phenomic Analysis Reveals Gene Interaction Networks
Responsible for Phenotypic Variability
It is increasingly recognized that the phenotypic effects of environmental or genetic perturbation
depend upon the functional/allelic status of interacting loci [1,2]. We consider genetic buffering to
underlie phenotypic stability/variability within a population and to derive from interaction between sets
of gene variants and environmental factors, and that sets of buffering genes represent functional networks
that distinctly modulate each phenotype [3]. In S. cerevisiae, genome-wide analysis shows that genes
interact extensively [4,5]. Thus, in yeast one can experimentally define functional gene networks in terms
of their capacity to buffer, or stabilize phenotypes. Buffering networks mask functional genetic variants
subject to natural selection, and thus comprise a reservoir within populations for the complex expression
of phenotypes [3]. Gene interaction networks are constrained evolutionarily across species and diverse
phenotypes [6–9]. Therefore yeast gene interaction networks can provide insight to buffering and
variable expression of disease when yeast phenomic experiments are designed within a cellular context
analogous to human biology [10].
By yeast phenomic analysis, we mean systematic, comprehensive, quantitative analysis of
gene interaction using the comprehensive collection of yeast gene knockout/knockdown (YKO/KD)
strains [10–14]. The YKO/KD library has been widely used to identify genes that affect yeast cell
proliferation (also called fitness) on compounds of interest, i.e., gene-drug interaction [15–18]. It has
also been used for studying gene-gene interaction, where a mutation of interest is introduced to the
background of the entire YKO/KD library, and the effect of interaction between loci on fitness can be
assessed genome-wide [19,20]. The Boone laboratory is analyzing all pairwise interactions in this way,
toward a complete interaction map for a eukaryotic cell [21], and the Weissman laboratory has examined
all possible combinations for selected sets of genes, called E-MAPs (epistatic mini array profiles) [22–24].
These and other studies have demonstrated, on a genomic scale, networks of interdependent genes that
produce phenotypes. Thus gene interaction networks will be an essential component of complete models
of any cell, organism, or disease [25].
We anticipate gene interaction networks will be specific to (and can potentially define) distinct
diseases. Such disease-buffering networks will also exhibit context-specificity with respect to environmental
inputs. Most large-scale gene-gene interaction studies have been carried out pairwise in a single, or
Genes 2015, 626
limited number of media. However, genes interact with environmental factors [14,26] and in more than
pairs [27], and we are just beginning to learn about the dynamic nature of yeast gene interaction networks
with respect to more than two genetic or environmental perturbations. These observations of complexity
point to a need for greater phenotypic resolution to develop quantitative models. In this article, we
discuss our effort to develop technology to resolve yeast gene interaction networks more quantitatively
so that the YKO/KD collection can be used to model disease buffering networks more precisely.
Considering the evolutionarily conserved nature of gene interaction [28,29], prior success in using yeast
as a genetic model for human disease, and advances in technology for phenomic analysis in yeast, we
build the rationale moving forward for more extensive efforts to construct yeast phenomic models of
disease buffering networks.
1.2. The Need for Quantitative Phenotyping to Experimentally Derive Buffering Networks
To enable phenomics, we have improved methodology for quantifying yeast gene interaction on a
genomic scale. It remains technologically challenging to collect and analyze genetic interaction data due
to the combinatorial explosion inherent to such networks [30]. Most automated YKO/KD phenotyping
has been done either by microarray hybridization of total genomic DNA harvested at different time
points from liquid cultures of pooled mutants to compare relative fitness [17,31,32], or by pinning small
amounts of cell paste from mutant cultures arrayed on agar media and measuring the area of
outgrowth of the spherical culture at an endpoint [33,34]. A few studies have also performed large-scale
phenotyping by time series analysis of liquid culture arrays [14,35,36]. The quantitative high throughput
cell array phenotyping (Q-HTCP) methodology we have been developing is based on the classic method
of dilution and spotting of liquid cell suspension to agar (normally analyzed qualitatively and reported
as an image), with the modification that kinetic growth curves are obtained by serial imaging and image
analysis [16]. We discovered that growth curves generated by kinetic analysis of cell array images, if fit
to a logistic growth function, yield cell proliferation parameters (CPPs) useful for measuring gene
interaction [10,37]. In this article we illustrate improved resolution for quantifying gene interaction with
Q-HTCP data, which reveals gene X media interaction and suggests a human-like (HL) yeast media
could reduce false positive results when validating yeast phenomic results in human cells.
1.3. Are Gene Interaction Networks That Buffer Human Disease Evolutionarily Conserved?
Fundamental processes shared by eukaryotic cells such as cell cycle control and protein secretion are
genetically conserved across evolution [38,39]. The importance of such processes in disease is evident,
but whether digenic inputs involving allelic variants that interact with respect to yeast phenotypes are
predictive for expression of human disease is only beginning to be explored. A purpose of this article is
to advocate for the use of S. cerevisiae to create experimental phenomic models of gene interaction to
investigate genetic buffering of human disease.
There are multiple examples suggesting that yeast can serve as useful models of human disease. One
example is neuronal degeneration, where disease-related human proteins have been expressed in yeast
to discover yeast genes that modulate toxicity, with subsequent validation in animal models of
neurodegeneration [40–45]. Another disease model investigates the gene interaction network influencing
biogenesis of the CFTR-F508 gene product, the main cause of cystic fibrosis (CF). A yeast homolog
Genes 2015, 627
of CFTR was constructed with a mutation of the conserved disease-relevant F508 residue (Yor1-F670)
to screen the YKO/KD library for modifiers. Conservation of gene interaction was demonstrated by
comparing the Yor1-F670 phenomic screen results to the literature reporting their effects on CFTR-F508
biogenesis (when knocked down by RNA interference) [10].
In addition to modifiers of Mendelian disease, such as CF, and multifactorial diseases like
neurodegeneration, yeast phenomics holds promise for modeling organismal processes, including aging
and mitochondrial dysfunction, which are relevant to a wide variety of human disease [46–49].
Numerous other genetic models of human disease are being developed, and these span across yeast and
other model organisms [50]. A great advantage of yeast models of human disease is the relative ease of
genome-wide phenotypic analysis, nevertheless translation of these models typically necessitates a
reductionist approach, focusing on validation of a few individual genes. Thus, an important future
direction is integrative, systems level modeling of disease buffering networks.
1.4. Experimental Resources and Technology for Yeast Phenomic Analysis
To quantify pair-wise gene interaction, phenotypic measures are needed for the wild-type and mutant
cell, in the perturbed and unperturbed context [16]. The YKO/KD strain collection provides a genomic
set of mutants for systematic analyses of gene interaction. Perturbations can take the form of additional
gene mutations introduced by the synthetic genetic array method [4], small molecules, or environmental
variations. A null hypothesis, predictive of phenotype, is required so that “interaction” can be quantified
as departure from expectation [51]. The power and resolution to analyze gene interaction networks is a
function of the precision, accuracy, and quantitative resolution of phenotypic data.
To advance quantitative analysis of yeast mutant libraries, we have developed an automated workflow
with cell-array printing, time-lapse imaging, image analysis, growth-curve fitting, and quantification of
gene interaction [10,16,37,52]. Cell-array imaging can be performed manually with a commercial grade
scanner (with built-in transparency unit) or using a new imaging robot, which can be integrated with a
robotic incubator (we use the Cytomat 6001 from Thermo Fisher Scientific, Asheville, NC, USA). The
robotic Q-HTCP system has a culture capacity of 72,576 (189 × 384-cultures arrays), exceeding commercial
systems for growth-curve analysis by over 500-fold [30]. While single time point analysis of colony
outgrowth area is higher throughput for breadth of global interaction analysis [33,34], Q-HTCP is more
quantitative for greater resolution in specific disease models [10].
1.5. Examples of Yeast Phenomic Modeling of Disease in Our Laboratory
In accord with this special issue, we discuss technology, current applications and speculate that the
application of yeast phenomic modeling for human disease research is the tip of an iceberg, where the
primary challenge is to identify phenotypes for which experimental derivation of gene interaction
networks in yeast discovers gene modules relevant to variable disease expression in humans [53–58].
Cystic fibrosis is a model we are developing to investigate whether yeast phenomic analysis could
reveal a gene interaction network relevant to a Mendelian human disorder [10]. There are two unique
sources of information to validate the yeast phenomic model of CFTR-F508, one being the extensive
siRNA literature involving targets that rescue the processing defect in human cell models, and another
being genomic data from the CF GWAS consortium, which is studying large cohorts of CF patients [59].
Genes 2015, 628
Thus CF represents a promising test case for the paradigm of evolutionarily conserved yeast gene
interaction networks that buffer/modulate the expression of human disease.
In contrast to CF, a genetically tractable disease due to highly penetrant loss of function mutations at
a single locus, cancer and aging represent polygenic and complex disease. To model cancer we are studying
the genetic buffering networks of ribonucleotide reductase (RNR) and target of rapamycin (TOR), which
are evolutionarily conserved regulators of DNA replication and cell growth, respectively [16,60]. RNR
and TOR are involved in tumorigenesis and progression and comprise important targets for development
of chemotherapeutic agents [61–64]. With the hypothesis that the integration of these networks provides
a master level of cell cycle regulation (i.e., DNA and protein synthesis), an unexpected connection
between TOR and RNR suggested by this model is threonine catabolic flux, which we believe could be
regulated by TOR signaling and that we’ve found to be a mechanism for buffering depletion of dNTP
pools due to loss of RNR function [16,60]. Threonine catabolism was also found to be important for
mouse embryonic stem cell survival due to increased need for DNA replication and also a role in histone
modification [65–67]. A third model is aging, which represents a disease-associated cellular process that
can be interrogated by phenomic analysis of the YKO/KD libraries. We are using Q-HTCP to measure
chronological lifespan (CLS) by growth curve analysis of stationary phase cultures that are periodically
rescued to fresh media so that change in the viable percentage of cells can be estimated during the aging
process [68], as a function of every gene and different nutrient conditions.
In summary, every cellular process has a disease correlate and vice-versa. Gene interaction, though
rare on a percentage basis, is frequent in aggregate and contributes greatly to disease expression. Yeast
phenomics enables experimental derivation of gene interaction networks in a comprehensive and
quantitative manner that is unparalleled for modeling genotype-phenotype complexity.
1.6. Development of a Human-Like (HL) Media for Yeast Phenomic Studies
With success of yeast phenomic disease models, homologues of yeast genes will be increasingly
tested for conserved gene interaction in their human cell context. To improve the positive predictive
value of such models, we sought to reduce the potential for interaction due to uncontrolled differences
in culture media for yeast and human cells. We used Q-HTCP to analyze the YKO/KD library for gene
interaction in standard media vs. a new media we designed to more closely resemble what is used for
human cell culture. This investigation was motivated in part by work from the Botstein laboratory
showing high potassium is required in standard defined (“dropout”) yeast media to help buffer toxicity
from ammonium sulfate [69], an ingredient omitted from human tissue culture media. We also observed
ammonium toxicity, and reduced other ingredients to create a yeast media that resembles human cell
culture media but supports normal growth of the YKO/KD reference strain (BY4741). We characterized
the HL media by genome-wide Q-HTCP analysis of the YKO/KD collection to identify gene X media
interaction. Our observations of deletion strains with differential cell proliferation across media supports
the possibility that HL media could improve the translational relevance of yeast phenomic screens. We
also demonstrate in some cases that the growth inhibitory effect of small molecules depends on media.
Although beyond the scope of the paper to formally assess all of these factors in the context of an actual
human disease model, our results thus far suggest careful consideration of media type is useful for assay
optimization and interpretation.
Genes 2015, 629
2. Methods
2.1. Yeast Media and Strains
The YKO/KD library was obtained from Research Genetics (Huntsville, AL, USA) and Open
Biosystems (Huntsville, AL, USA). The genetic background for the YKO/KD library was BY4741
(S288C MATa ura3-0 his3-1 leu2-0 met17-0). Yeast media was YP (10 g/L Yeast Extract, 20 g/L
peptone) with either 2% dextrose (YPD) or 3% glycerol/3% ethanol (YPEG) as carbon source. The Cold
Spring Harbor (CSH) synthetic complete (SC) dropout media [70] was also used with either carbon
source as was the “human-like” (HL) media. The recipe for HL media was derived from the CSH SC
media with the following modifications: ammonium sulfate was removed (the normal 5 g/L was reduced
to 0.5 g/L in HL + AS media). Potassium phosphate was reduced from 1 to 0.5 gm/L. Magnesium sulfate
was reduced from 0.5 to 0.05 g/L. PABA was removed from the amino acid powder and inositol was
dropped from 0.0734 to 0.025 g/L. Leucine was reduced from 0.367 to 0.1468 g/L. The potassium
phosphate and magnesium sulfate modifications were introduced by ordering custom yeast nitrogen base
(without ammonium sulfate) from Sunrise Science (San Diego, CA, USA). The HL media recipe was
partly guided by comparison with RPMI and DMEM media components.
2.2. Quantitative High Throughput Cell Array Phenotyping (Q-HTCP)
A Caliper Sciclone 3000 liquid handling robot was used for cell array printing (384-culture arrays),
integrated with a custom imaging robotic system and a Cytomat 6001, having capacity for 189 arrays
(Thermo Fisher Scientific, Asheville, NC, USA). Images were taken every 2–3 hours and analyzed
as previously described to obtain cell proliferation parameters [10], using the logistic equation,
G(t) = K/(1 + e r(tl)), assuming G(0) < K, where G(t) is the image intensity of a spotted culture vs.
time, K is the final carrying capacity, r is the maximum specific growth rate, and l is the time that
maximal absolute growth rate occurs, when G(t) = K/2 [51].
2.3. Quantification of Gene Interaction
For the genome-wide screen, cell proliferation phenotypes (CPPs) were used to calculate gene X
media interaction in the following way: the CPP for each deletion strain was adjusted by its difference
compared to the median CPP among 384 cultures of the reference stain in YP (YPD or YPEG), and by
the difference of the median CPP of the reference strain in YP with the respective test media. After the
normalizations, the difference between the deletion strain on CSH or HL media vs. YP media was taken
as the interaction value. YP-dextrose was the control for all dextrose media and YP-ethanol/glycerol for
all ethanol-glycerol media. For the drug x gene x media interaction analysis, CPPs were obtained in the
same way and control arrays containing the same media without drug were subtracted from the respective
drug gradient arrays.
2.4. Recursive Expectation Maximization Clustering (REMc)
Clustering was performed as previously described [12], followed by hierarchical clustering and heat
map generation applying an R script (http://www.r-project.org/) to each REMc cluster. REMc was
Genes 2015, 630
performed with a 16-dataset matrix, including the ORF-deletion effect (“shift” value) on YP media and
the interaction values for defined media, for both K and L parameters. A custom java code that utilizes
Weka (www.cs.waikato.ac.nz/ml/weka) was used to generate clusters. After REMc, a python script is
used to format the clusters for analysis by the command line version of the Gene Ontology (GO) Term
Finder (GTF) downloaded from http://search.cpan.org/dist/GO-TermFinder/ [71]. GTF searches for
enrichment of Gene Ontology terms in each cluster by comparing the ratio of genes assigned to a term
within a cluster to the total number of genes tested in the experiment having that term.
3. Results
3.1. Using Growth Curve Parameters as High-Resolution, Quantitative Phenotypes
A fundamental challenge of phenomics for all organisms is quantifying phenotypes with respect to
gene interaction on a genomic scale. Every disease has multiple phenotypes with multiple different genes
contributing to each one. There is a functional spectrum among different alleles for each gene, and the
influence of a given allele on the phenotype depends on combination with particular alleles at different
loci [1]. Given this genetic complexity of phenotypic expression, yeast offers key advantages for mapping
gene interaction as comprehensively and quantitatively as possible with respect to both environment and
other genes: (1) Much of the overall fitness is encapsulated by the phenotype of cell proliferation, which
lends comprehensiveness; and (2) Cell proliferation is a continuous trait that’s straightforward to
quantify, where analysis with a logistic growth function resolves fitness into three components, providing
additional resolution for phenomic analysis and gene interaction network construction (Figure 1).
Though phenotypes are more complex in humans than yeast, it is possible to extrapolate across
species between potentially any phenotype, based upon gene interaction networks that function similarly
across evolution [6–9]. Cell proliferation is under strong selection in yeast and involves evolutionarily
conserved genes, which may be involved in the production of different phenotypes subject to natural
selection for other reasons across evolution. For example, in the yeast model of CFTR-F508, gene
interactions that influence Yor1-F670, also influence CFTR-F508 biogenesis in human cell lines [10].
In yeast, “F biogenesis” can be selected for by cell proliferation on oligomycin, whereas in humans it
is assayed by chloride transport. Although the cellular and organismal phenotypes associated with
Yor1-F670 are different from those of CFTR-F508, the network of gene interactions affecting
biogenesis of the respective proteins is similar [10]. Hardly unique to ABC transporters, quantifying cell
proliferation of yeast mutant arrays provides a powerful strategy for broadly analyzing eukaryotic gene
networks that influence a variety of human disease.
Q-HTCP is an automated method for collecting over 70,000 growth curves per experiment. Time
series image data is taken from miniature lawns of proliferating agar cell culture arrays each representing
a different defined mutant [10,16,37,52]. These data fit closely to a logistic growth function [37],
yielding independent quantitative CPPs with which to calculate gene interaction (Figure 1A).
Cell proliferation parameters represent fundamentally distinct phenotypes under differential natural
selection and thus regulated by different gene interaction networks buffering each parameter in a
context/perturbation-specific manner.
Genes 2015, 631
Figure 1. Cell proliferation parameters (CPPs) for quantifying gene interaction are obtained
from fitting Q-HTCP data with a logistic growth function. The equation and an example
growth curve are shown in panel A; CPPs are phenotypes for measuring gene interaction: K
is the carrying capacity (final growth density, quantified as average pixel intensity), L is the
time (h) it takes for a culture to reach K/2, and R is the maximum specific rate. Each
parameter is independent of the other, because G(0) is treated as unknown. The CPPs for
each graph are given in the respective table to illustrate variation in each parameter.
Of the three parameters, we have found the L parameter to be especially useful (Figure 1C).
Oligomycin growth inhibition is evident best by L in the Yor1-F508 model for CFTR-F508 [10].
Similarly, in the chronological lifespan (CLS) model of aging, where the yeast phenotype is duration of
post-mitotic survival after entry into stationary phase [72], L best reflects longevity, which is quantified
as the percentage of cells with persistent ability to regrow upon transfer to fresh media [73]. Q-HTCP
can be used for phenomic analysis of CLS by collecting growth curves at weekly intervals for the entire
genomic collection of YKO/KD strains. If other growth parameters, i.e., rate, carrying capacity and lag
remain consistent across the aging process, changes in L would estimate CLS by a rightward shift of the
growth curve over time, reflecting reduction in the number of colony forming units vs. age. Other gene-
or disease-related processes might be better modeled with a different parameter; for example,
mitochondrial function could be modeled by change in K, resulting from a relative decline in growth
after the diauxic shift (Figure 1D).
3.2. A Human-Like Yeast Media to Increase Positive Predictive Value of Yeast Phenomic Models
Cell culture media components can modulate phenotypes, representing a form of gene x environment
interaction. To increase translational relevance of yeast phenomic models to cultured human cells, we
introduced changes to defined yeast media to resemble human cell culture media. By reducing
differences between yeast and human cell media, discovery of gene interaction from yeast phenomic
Genes 2015, 632
screens can be focused on the disease network by reducing media-specific gene interaction. The
following changes were made to the Cold Spring Harbor synthetic complete media [70] to create HL
media: (1) in the yeast nitrogen base, magnesium sulfate was reduced 90% (from 0.5 to 0.05 g/L),
potassium phosphate was reduced 50% (from 1 to 0.5 g/L), and ammonium sulfate (5 g/L) was removed
(0.5 g/L ammonium sulfate was added to HL + AS media); (2) in the complete dropout powder, inositol
was reduced from 0.0734 to 0.025 g/L, leucine was reduced from 0.367 to 0.1468 g/L, and PABA
(0.0734 g/L) was removed. The YKO/KD collection reference strain, BY4741, exhibited robust cell
proliferation parameters across these media modifications (Figure 2).
Figure 2. Cell proliferation phenotypes (CPPs) for BY4741 are similar in HL and standard
yeast media. Box plots represent the central 75% (colored box), median (bold bar), and range
(whiskers). Media types include Cold Spring Harbor (CSH), human-like (HL) with or
without 0.5 gm/L ammonium sulfate (AS), or rich yeast/peptone (YP). Each media type was
prepared with either dextrose (D) or ethanol glycerol (EG) as a carbon source. CPP
abbreviations are described in Figure 1. Values of R were multiplied by one hundred. Data
are from two different 384-culture liquid arrays (teal and lavender box plots) printed onto
each media type. The main conclusion is that cell proliferation is similar for each media,
controlled for dextrose or ethanol/glycerol as carbon source. This experiment demonstrates
feasibility to modify synthetic defined yeast media to resemble that used for human cells. It
stands to reason that similarity in the media will improve reproducibility of gene interaction
across species by reducing gene interaction with media components.
3.3. Phenomic Analysis Reveals Clusters of Gene X Media Interaction
To investigate the potential significance of HL media for yeast phenomic analysis, we collected
growth curves for all (~6000) YKO/KD strains on eight different media, alternatively using dextrose or
ethanol/glycerol as carbon sources in YP (YPD/YPEG), CSH synthetic complete (CSHD/CSHEG), or
HL without (HLD/HLEG) or with 0.5 g/L ammonium sulfate (HLD + AS/HLEG + AS). Interactions were
calculated by normalizing growth of hundreds of replicate cultures of the reference strain (BY4741) on
Genes 2015, 633
each media to growth on YP media and also normalizing by the effect of the open reading frame on
growth in the YP media (i.e., YPD was used as the reference media for CSHD and HLD, while YPEG
was used for CSHEG and HLEG). We used REMc to objectively define groups of genes with shared
patterns of media interaction, and hierarchical clustering with heat maps to best visualize patterns of
interaction (Figure 3, Supplemental Data Files 1 and 2).
REMc analysis revealed six first round clusters. Cluster 1-0-0 had reduced fitness on YPEG (column 5),
which was alleviated in defined media (columns 6–8) (Figure 3C). As typical for most phenotypes, the
majority (3816) of strains exhibited little or no interaction, as illustrated by the large cluster, 1-0-5,
(Figure 3D). YKO/KD strains failing to grow on every E/G media comprised a distinct cluster (Figure
3B at bottom, extreme negative K and positive L interaction indicates no growth in columns 5–8 and
13–16), and consisted of genes enriched for GO Terms related to respiratory function as expected (see
also cluster 1-0-2 in Supplemental Data Files 1 and 2). Cluster 1-0-1 highlights gene deletions that have
reduced fitness on defined media with non-fermentable carbon source, but not rich media or media with
glucose as the carbon source; moreover, more strains display this phenotype on HL than on CSH media
(Figure 3E). Cluster 1-0-1 was enriched for genes functioning in mitochondrial and ribosomal processes,
consistent with roles in buffering combined perturbations of carbon and nitrogen availability (see
Supplemental Data Files 1 and 2).
Within cluster 1-0-0 (Figure 3C), several of the strains have reduced fitness on CSHD media. This
set of genes was better highlighted in the second round cluster 2-0.0-1 (Figure 3F). There was no
enrichment in cellular processes annotated by gene-ontology for clusters 1-0-0 or 2-0.0-1, which is often
true for interactions modulating relatively unstudied phenotypes. However, we did find genes in cluster
2-0.0-1 that are implicated in amino acid metabolism, amino acid permease trafficking and ammonium
efflux (Figures 3G), consistent with a report from the Botstein laboratory suggesting that up-regulation
of amino acid biosynthesis and excretion is necessary to buffer ammonium toxicity [69]. In support of
this model, Lst4 is involved in trafficking of Gap1, the general amino acid permease [74]. The gap1-0
strain had no phenotype, possibly due to redundancy among amino acid permeases, and thus we
hypothesize that Lst4 could regulate a module of permeases such that its functional loss (but not
disrupting any single permease), would alter fitness in the context of ammonium toxicity introduced by
the CSH media. The growth inhibitory phenotype suspected to be due to ammonium sulfate in CSH
would be alleviated by its removal from HL media [69]. Similarly, deletion of MEP1, the high flux
ammonium transporter reduced fitness in CSH media, but not HL. Deletion of VPS17 and VPS24, which
are involved with protein sorting to the endosome and vacuole also appear to buffer CSH media, based
on their being relatively dispensable in HL media (Figure 3H), perhaps implicating their function too in
the regulation of amino acid or ammonium permeases. Another gene in this cluster, ORT1, is implicated
in ammonium toxicity in humans. ORT1 is required for arginine biosynthesis due to its function as a
mitochondrial ornithine transporter, and its human ortholog ORNT1 is causative of the recessive disease,
hyperornithinemia-hyperammonemia-homocitrullinuria syndrome (http://omim.org/). In summary, cluster
2-0.0-1 exhibits a pattern indicative of genes that buffer ammonium toxicity based on the functional
requirement for maintaining fitness in CSH but not HL media. While some genes in the cluster have
putative relationships between their known functions, others may point to novel molecular activity or
new connections between known activities [27,60] (see also Supplemental Data Files 1 and 2).
Genes 2015, 6 34
Figure 3. REMc reveals gene X media interaction modules. Q-HTCP was performed for the YKO/KD collection on the indicated media, and
interactions were calculated for defined media using YPD or YPEG as a control (see text and methods). REMc was used to mine interaction
values derived from K and L for shared patterns. (A) A color scale was used to visualize interaction values, which are positive/blue with respect
Genes 2015, 635
to the K phenotype if YKO/KD strains have higher fitness in defined than YP media, relative to the reference strain. Conversely interactions
are negative/brown with respect to L (less time required to reach K/2) if YKO/KD strains exhibit a relative increased in fitness on defined vs.
YP media; (B) The entire data matrix (root cluster) was analyzed by hierarchical clustering (before REMc). The x-axis order (media conditions
and growth parameters) is the same in all panels; (CE) The first round of REMc yielded six clusters, three of which are shown. The YP “shift”
refers to the difference between the respective YKO/KD strain and the median of the reference (see Figure 2 and panel H); (C) Cluster 1-0-0
contains 456 YKO/KD strains, many of which exhibit lower fitness (lower K and longer L) on YPEG media (column 5 and 13) that is partially
alleviated when grown on defined media (columns 6 (CSH/EG), 7 (HL/EG) and 8 (HL + AS/EG)); (D) Cluster 1-0-5 contains a majority (3861)
of YKO/KD strains that exhibit similar phenotypes on all media; (E) Cluster 1-0-1 contains genes that share reduced fitness on non-fermentable
media (columns 6–8 and 14–16). Some strains reveal this phenotype on HL media (columns 6, 8, 14, and 16), but not YPEG or CSHEG defined
media (columns 5, 7, 13 and 15); (F) As part of the REMc workflow, clusters are iteratively analyzed until terminal clusters are obtained. Cluster
2-0.0-1, obtained in the second round from cluster 1-0-0, contained strains with low fitness in CSHD (column 3), but not YPD, HLD, or HLD
+ AS (columns 1, 2 and 4). A sub-cluster (red box) contains YKO/KD strains whose functions are listed in the table (G), with the data for
particular genes discussed in the text given in (H), where the interaction for each gene on each media type is plotted for the K parameter. Green
diamonds indicate the ORF effect, or “shift” (difference between deletion strain and reference strain on YPD or YPEG, respectively) by which
data are normalized/shifted [10].
Genes 2015, 6 36
3.4. Resolving Drug-Media Interaction by Q-HTCP across Drug-Gradients
The YKO/KD collection is frequently used to determine gene-drug interaction profiles. After finding
an abundance of gene X media interaction (Figure 3), we investigated media-dependence of drug
responses. Q-HTCP was performed on different media containing the same drug concentration gradients.
Desired concentrations of drug were added to 10 mL of media and poured into a monowell plate that
was tipped at an angle (by overlapping the bottom plate on its lid by about an inch) in order for the media
to solidify as a wedge at one end. Next, the plate was laid flat and 30 mL of media without drug was
layered over the wedge, creating a diffusion gradient. 384-cell arrays (16 × 24), consisting of 8 different
strains, with 24-cultures per row, and strains arrayed in rows 1–8 repeated in rows 9–16 (to control for
evenness in the gradient), were printed onto the drug gradient plates and control media without drug.
Q-HTCP was performed and the growth curve parameter array of the control (no drug) was subtracted
from the drug gradient plate. The change in L across the gradient was compared between different media
(Figure 4). Eight strains harboring different mutants altering drug efflux (PDR mutants) or permeability
(ERG mutants) were tested to investigate whether these factors influence the phenotype associated with
particular compounds.
Media and genetic background dependence of growth phenotypes were observed for some compounds,
including hygromycin and bortezomib (Figure 4). Hygromycin selection [75] was strong in YPD
(Figure 4A) and HL media (Figure 4B), but not on standard CSH media (Figure 4D). The addition of
0.5 mg/mL ammonium sulfate to HLD media (1/10 of what is added to CSH) slightly reduced growth
inhibition by hygromycin (Figure 4C), suggesting it reduces hygromycin efficacy. Results for hygromycin
are reminiscent of reduced efficacy of G418 with ammonium sulfate [19].
Bortezomib (Velcade, Millennium Pharmaceuticals, Cambridge, MA, USA) was approved in 2003
for myeloma treatment, and is thought to inhibit the proteasome by binding to the 26S subunit, although
its anti-cancer effect is not fully understood. Bortezomib exerted little or no growth inhibitory effect on
the reference strain (BY4741) in YPD. When a drug does not inhibit growth, it could be due to low
concentration, lack of permeability, extrusion by the efficient yeast drug efflux (pleiotropic drug
resistance, PDR) system, absence of the target, failure of the drug to sufficiently inactivate the target,
the target being unrequired under the growth conditions, and/or buffering of the physiological effect of
the drug. To assess the possibility of a drug efflux or permeability mechanism, the panel of strains,
including mutants in ergosterol biosynthesis, pleotropic drug resistance, and a chimeric fusion of the
PDR1 DNA-binding domain with the transcriptional repressor domain of CYC8 (Pdr1-Cyc8) [76], were
tested with a concentration gradient of 13 μM bortezomib in YPD, YPEG, HLD, and HLEG (Figure 4E–H).
Bortezomib was found to interact with media, drug efflux, and permeability. This result confirms a
previous report that drug efflux function influences sensitivity to bortezomib [77], and we found further
enhancement of sensitivity with HL media and ethanol/glycerol as the carbon source. Not much of a
growth phenotype was observed on YPEG (Figure 4E–H), suggesting the combination of carbon and
nitrogen sources influence drug responsiveness. The drug efflux mutants we tested were different from
those previously examined [77], and we found that a chimeric protein constructed by Stepanov et al. was
effective for sensitizing to bortezomib [76]. ERG3 deletion also influenced growth inhibition by
bortezomib on HLEG media, presumably by affecting membrane permeability.
Genes 2015, 637
Figure 4. Drug sensitivity can depend upon media type, cellular efflux and permeability of
drugs, or interaction between media and efflux or permeability. The experimental description
is given in the text. The blue color scale indicates decreased fitness (increased L), with the
range (hours) indicated in parentheses beside media type labels where relevant (panels E and F).
Numerical interaction values are printed in each cell. Gray color indicates that no growth
curve was observed (complete growth inhibition at the particular drug concentration). The
cell array data was rearranged in the heat maps to assist visualization (rows 1/9, 2/10…8/16
were grouped). (AD) Gradient plates were poured with 500 g/mL hygromycin in (A) YPD,
(B) HLD, (C) HLD+AS, and (D) CSHD; (EH) Gradient plates were poured with 5 μg/mL
bortezomib in (E) HLD, (F) HLEG, (G) YPD, and (H) YPEG.
In summary, gradient experiments assist optimization of drug screening conditions by surveying a
wide range of concentration in multiple different media and with multiple strains. Identification of a
media type that confers enhanced sensitivity can reduce the drug cost for a genomic screen. Likewise,
drug efflux or permeability mutations could be introduced into the YKO/KD libraries if needed [19,20],
as in the case of bortezomib. Furthermore, gradient plate Q-HTCP can be used to quantify known
gene-drug interactions (e.g., target or buffering genes) to help optimize a range of concentrations for a
phenomic screen [16]. The identification of drug X media interaction (e.g., differential drug sensitivity
of the wild type genetic background on fermentable vs. non-fermentable media) may point to comparative
Genes 2015, 638
phenomic screens to discover differential gene interaction networks informative about the impact of
biological context (e.g., Warburg effect) on drug resistance networks.
4. Discussion
Can Yeast Phenomic Models Aid Construction of Quantitative Genetic Networks that Predict Disease?
The genomic collection of S. cerevisiae yeast knockout and knockdown mutants has been a boon for
surveying genotype-phenotype complexity [21]. Prior to construction of the YKO/KD library, the
vastness of gene interaction space was unappreciated. Q-HTCP was developed as a cell array-based
growth curve technology so that the large number of mutant phenotypes can be better resolved and gene
interaction more precisely quantified, which is complementary to higher throughput, less quantitative
methods. The ability to explore gene interaction in greater detail, with increased sensitivity and
specificity comes with the constraint of more focused sets of questions and/or models, but may lead to
better understanding of complex gene networks in the context of specific disease models [10].
Most high-throughput yeast phenotyping methods measure fitness endpoints [30]. In contrast,
Q-HTCP obtains kinetic growth, likened to OD of liquid culture, but with much greater throughput, so
that fitness can be further analyzed in terms of distinct components of fitness including carrying capacity,
maximum specific rate, and time to reach half-carrying capacity [37]. For example, carrying capacity
(final growth density) of cultures on glucose is affected by the respiratory function of yeast, because
inefficient respiration results in reduced biomass accumulation following the diauxic shift (i.e., during
growth without glycolysis), but may not affect other logistic growth parameters. By contrast, YKO/KD
strains may exhibit differences in lag time if they carry mutations disrupting functions that influence efflux
of the growth inhibitory compound. This phenotype would be captured by the time to half-carrying
capacity (the L parameter) even if the carrying capacity or the maximum specific rate were unaffected [10].
Specific classes of mutants could be detected by considering the interaction pattern across all parameters:
for example cell cycle checkpoint mutants might be identified by a short L with an accelerated maximum
specific rate (due to checkpoint failure) and a reduced carrying capacity (due to cell death) [78,79]. These
are just a few examples of the overall theme, which is that genetic buffering of fitness can be further
resolved with Q-HTCP to increase the resolution of gene interaction networks, both quantitatively and
in terms of different CPPs.
Aspects of Q-HTCP development in our laboratory currently include: (1) the use of commercial liquid
handling robotics for cell array printing; (2) development of imaging methods and image analysis
software to convert images to growth curves [16]; (3) use of a logistic growth model to fit Q-HTCP data
so that CPPs can be used for quantifying gene interaction (Figure 1); (4) development of approaches to
optimize screening conditions for phenomic studies, such as media modification (Figures 2 and 3) and
gradient array analysis (Figure 4); (5) incorporation of the synthetic genetic array method to carry out
phenomic screens of gene x gene interaction [20]; and (6), development of data mining tools including
REMc [12]. Future development of the technology will focus on its application to additional disease
models and integration of yeast gene interaction networks with other omic data to more fully understand
disease expression (Figure 5).
Genes 2015, 639
Figure 5. Yeast phenomic models can serve to discover gene interaction networks underlying
expression of disease phenotypes. Disease-buffering gene interaction networks are derived
experimentally by disease-relevant perturbation of the yeast gene knockout/knockdown mutant
library followed by measurement of gene interaction with Q-HTCP. The resulting gene networks
are queried for evolutionary conservation to generate hypotheses regarding their relevance to
human phenotypes. Hypotheses can be tested in translational models, leading to model
validation or refinement, with iterative yeast phenomic analysis for further refinement.
We suspect human genotype-phenotype complexity originates substantially from cellular processes
shared in common with yeast. Thus yeast phenomic models of human disease could inform the basic
biology of gene interaction networks at the same time as variable disease expression, functioning for
both “basic” science [80] and “translational” insight intended for validation in higher eukaryotes [50,81].
The experimental convenience and eukaryotic relevance of the YKO/KD collections, together with the
development and application of Q-HTCP techniques provide a powerful opportunity to discover gene
interaction networks underlying disease biology.
5. Conclusions
Work with the YKO/KD library has revealed eukaryotic gene interaction networks to be extensive,
revealing vast phenotypic complexity even in a single-cell organism. Here we report, even when cell
proliferation of the reference strain is similar between different media, hundreds of the YKO/KD strains
have growth phenotypes, thus revealing extensive gene X media interaction. The fact that animals have
additional organismal complexity is perhaps the strongest argument for using S. cerevisiae to model
gene interaction networks [82]. Cellular functions and the genes carrying them out are conserved from
yeast to human [83]. Thus gene interactions that further modulate these processes may also be usefully
modeled across species. Many technical and biological factors will impact the implementation and
translational success of any particular yeast phenomic disease model. Cell culture media is one such
factor and the use of HL yeast media may increase positive predictive value for disease translation in
some human cell models. In any case, awareness of the potential for gene x media interaction is likely
to be beneficial. A challenge not addressed in this article, but critical to the success of yeast phenomics
for modeling disease-relevant gene interaction networks, is development of computational methods that
leverage phenomic and other omic data to extract biological insight [84]. As yeast phenomics and other
gene interaction techniques to study human disease advance, principles for genetic buffering in humans
Genes 2015, 640
will emerge and give rise to iterative models employing human data and yeast experiments [3]. An
innovative approach in this regard is the Resilience Project, which seeks to identify, from within the
non-diseased human populations, buffering loci that harbor the capacity to reduce phenotypic
manifestations of disease-associated mutations [85]. The development of a workflow to identify buffering
networks in people could leverage yeast phenomic models to help mine for gene interaction networks
underlying disease expression in human populations. Such understanding will advance disease diagnosis
and provide new targets for personalized/precision management of disease phenotypes.
Acknowledgments
The authors thank the funding agencies, which enabled the work, including Howard Hughes Medical
Institute Physician Scientist Early Career Award 57005927, American Cancer Society Research Scholar
Grant 10-066-01-TBE, National Institutes of Health R01 AG043076, and NIH P30 grants DK072482
and DK056336. The authors thank Brett McKinney for advice on computational aspects of work,
Alexander Stepanov and Mary-Ann Bjornsti for sharing the Pdr1-Cyc8 fusion strain, and Stephen Friend
and Eric Schadt for discussions about the Resilience Project.
Author Contributions
John L. Hartman conceptualized and wrote the manuscript. Chandler Stisher and Darryl A. Outlaw
performed the Q-HTCP experiments. Najaf A. Shah and John L. Hartman considered application of the
logistic model. John W. Rodgers and Richard A. White performed the Q-HTCP analysis and gene
interaction modeling. Dehua Tian, Jingyu Guo, John L. Hartman, and Sean M. Santos developed REMc
and performed analysis.
Conflicts of Interest
John L. Hartman has a financial interest in Spectrum PhenomX, LLC, which aims to commercialize
Q-HTCP technology. All other authors declare no competing interests.
References
1. Lupski, J.R.; Belmont, J.W.; Boerwinkle, E.; Gibbs, R.A. Clan genomics and the complex
architecture of human disease. Cell 2011, 147, 32–43.
2. Badano, J.L.; Katsanis, N. Beyond mendel: An evolving view of human genetic disease
transmission. Nat. Rev. Genet. 2002, 3, 779–789.
3. Hartman, J.L., 4th; Garvik, B.; Hartwell, L. Principles for the buffering of genetic variation. Science
2001, 291, 1001–1004.
4. Dixon, S.J.; Costanzo, M.; Baryshnikova, A.; Andrews, B.; Boone, C. Systematic mapping of
genetic interaction networks. Annu. Rev. Genet. 2009, 43, 601–625.
5. Costanzo, M.; Baryshnikova, A.; Myers, C.L.; Andrews, B.; Boone, C. Charting the genetic
interaction map of a cell. Curr. Opin. Biotechnol. 2011, 22, 66–74.
6. Lee, I.; Lehner, B.; Vavouri, T.; Shin, J.; Fraser, A.G.; Marcotte, E.M. Predicting genetic modifier
loci using functional gene networks. Genome Res. 2010, 20, 1143–1153.
Genes 2015, 641
7. McGary, K.L.; Park, T.J.; Woods, J.O.; Cha, H.J.; Wallingford, J.B.; Marcotte, E.M. Systematic
discovery of nonobvious human disease models through orthologous phenotypes. Proc. Natl. Acad.
Sci. USA 2010, 107, 6544–6549.
8. Woods, J.O.; Singh-Blom, U.M.; Laurent, J.M.; McGary, K.L.; Marcotte, E.M. Prediction of
gene-phenotype associations in humans, mice, and plants using phenologs. BMC Bioinform. 2013,
doi:10.1186/1471-2105-14-203.
9. Hwang, S.; Kim, E.; Yang, S.; Marcotte, E.M.; Lee, I. Morphin: A web tool for human
disease research by projecting model organism biology onto a human integrated gene network.
Nucleic Acids Res. 2014, 42, W147–W153.
10. Louie, R.J.; Guo, J.; Rodgers, J.W.; White, R.; Shah, N.; Pagant, S.; Kim, P.; Livstone, M.;
Dolinski, K.; McKinney, B.A.; et al. A yeast phenomic model for the gene interaction network
modulating cftr-f508 protein biogenesis. Genome Med. 2012, doi:10.1186/gm404.
11. Ratnakumar, S.; Hesketh, A.; Gkargkas, K.; Wilson, M.; Rash, B.M.; Hayes, A.; Tunnacliffe, A.;
Oliver, S.G. Phenomic and transcriptomic analyses reveal that autophagy plays a major role in
desiccation tolerance in Saccharomyces cerevisiae. Mol. Biosyst. 2011, 7, 139–149.
12. Guo, J.; Tian, D.; McKinney, B.A.; Hartman, J.L., 4th. Recursive expectation-maximization
clustering: A method for identifying buffering mechanisms composed of phenomic modules. Chaos
2010, doi:10.1063/1.3455188.
13. Kahraman, A.; Avramov, A.; Nashev, L.G.; Popov, D.; Ternes, R.; Pohlenz, H.D.; Weiss, B.
PhenomicDB: A multi-species genotype/phenotype database for comparative phenomics. Bioinformatics
2005, 21, 418–420.
14. Warringer, J.; Ericson, E.; Fernandez, L.; Nerman, O.; Blomberg, A. High-resolution yeast phenomics
resolves different physiological features in the saline response. Proc. Natl. Acad. Sci. USA 2003, 100,
15724–15729.
15. Parsons, A.B.; Brost, R.L.; Ding, H.; Li, Z.; Zhang, C.; Sheikh, B.; Brown, G.W.; Kane, P.M.;
Hughes, T.R.; Boone, C. Integration of chemical-genetic and genetic interaction data links bioactive
compounds to cellular target pathways. Nat. Biotechnol. 2004, 22, 62–69.
16. Hartman, J.L., 4th; Tippery, N.P. Systematic quantification of gene interactions by phenotypic array
analysis. Genome Biol. 2004, doi:10.1186/gb-2004-5-7-r49.
17. Hillenmeyer, M.E.; Fung, E.; Wildenhain, J.; Pierce, S.E.; Hoon, S.; Lee, W.; Proctor, M.;
Onge, R.P., St.; Tyers, M.; Koller, D.; et al. The chemical genomic portrait of yeast: Uncovering a
phenotype for all genes. Science 2008, 320, 362–365.
18. Chatr-Aryamontri, A.; Breitkreutz, B.J.; Heinicke, S.; Boucher, L.; Winter, A.; Stark, C.; Nixon, J.;
Ramage, L.; Kolas, N.; O’Donnell, L.; et al. The biogrid interaction database: 2013 update. Nucleic
Acids Res. 2013, 41, D816–D823.
19. Tong, A.H.; Boone, C. Synthetic genetic array analysis in Saccharomyces cerevisiae. Methods Mol.
Biol. 2006, 313, 171–192.
20. Singh, I.; Pass, R.; Togay, S.O.; Rodgers, J.W.; Hartman, J.L., 4th. Stringent mating-type-regulated
auxotrophy increases the accuracy of systematic genetic interaction screens with Saccharomyces
cerevisiae mutant arrays. Genetics 2009, 181, 289–300.
21. Costanzo, M.; Baryshnikova, A.; Bellay, J.; Kim, Y.; Spear, E.D.; Sevier, C.S.; Ding, H.; Koh, J.L.;
Toufighi, K.; Mostafavi, S.; et al. The genetic landscape of a cell. Science 2010, 327, 425–431.
Genes 2015, 642
22. Schuldiner, M.; Collins, S.R.; Thompson, N.J.; Denic, V.; Bhamidipati, A.; Punna, T.; Ihmels, J.;
Andrews, B.; Boone, C.; Greenblatt, J.F.; et al. Exploration of the function and organization of the
yeast early secretory pathway through an epistatic miniarray profile. Cell 2005, 123, 507–519.
23. Ihmels, J.; Collins, S.R.; Schuldiner, M.; Krogan, N.J.; Weissman, J.S. Backup without
redundancy: Genetic interactions reveal the cost of duplicate gene loss. Mol. Syst. Biol. 2007,
doi:10.1038/msb4100127.
24. Jonikas, M.C.; Collins, S.R.; Denic, V.; Oh, E.; Quan, E.M.; Schmid, V.; Weibezahn, J.;
Schwappach, B.; Walter, P.; Weissman, J.S.; et al. Comprehensive characterization of genes required
for protein folding in the endoplasmic reticulum. Science 2009, 323, 1693–1697.
25. Boone, C. Yeast systems biology: Our best shot at modeling a cell. Genetics 2014, 198, 435–437.
26. Warringer, J.; Blomberg, A. Automated screening in environmental arrays allows analysis of
quantitative phenotypic profiles in Saccharomyces cerevisiae. Yeast 2003, 20, 53–67.
27. Taylor, M.B.; Ehrenreich, I.M. Genetic interactions involving five or more genes contribute to a
complex trait in yeast. PLOS Genet. 2014, doi:10.1371/journal.pgen.1004324.
28. Dixon, S.J.; Fedyshyn, Y.; Koh, J.L.; Prasad, T.S.; Chahwan, C.; Chua, G.; Toufighi, K.;
Baryshnikova, A.; Hayles, J.; Hoe, K.L.; et al. Significant conservation of synthetic lethal genetic
interaction networks between distantly related eukaryotes. Proc. Natl. Acad. Sci. USA 2008, 105,
16653–16658.
29. Roguev, A.; Bandyopadhyay, S.; Zofall, M.; Zhang, K.; Fischer, T.; Collins, S.R.; Qu, H.;
Shales, M.; Park, H.O.; Hayles, J.; et al. Conservation and rewiring of functional modules revealed
by an epistasis map in fission yeast. Science 2008, 322, 405–410.
30. Blomberg, A. Measuring growth rate in high-throughput growth phenotyping. Curr. Opin. Biotechnol.
2011, 22, 94–102.
31. Winzeler, E.A.; Shoemaker, D.D.; Astromoff, A.; Liang, H.; Anderson, K.; Andre, B.; Bangham, R.;
Benito, R.; Boeke, J.D.; Bussey, H.; et al. Functional characterization of the S. cerevisiae genome
by gene deletion and parallel analysis. Science 1999, 285, 901–906.
32. Giaever, G.; Chu, A.M.; Ni, L.; Connelly, C.; Riles, L.; Veronneau, S.; Dow, S.; Lucau-Danila, A.;
Anderson, K.; Andre, B.; et al. Functional profiling of the Saccharomyces cerevisiae genome.
Nature 2002, 418, 387–391.
33. Collins, S.R.; Schuldiner, M.; Krogan, N.J.; Weissman, J.S. A strategy for extracting and analyzing
large-scale quantitative epistatic interaction data. Genome Biol. 2006, :10.1186/gb-2006-7-7-r63.
34. Baryshnikova, A.; Costanzo, M.; Kim, Y.; Ding, H.; Koh, J.; Toufighi, K.; Youn, J.Y.; Ou, J.;
San Luis, B.J.; Bandyopadhyay, S.; et al. Quantitative analysis of fitness and genetic interactions in
yeast on a genome scale. Nat. Methods 2010, 7, 1017–1024.
35. Liti, G.; Carter, D.M.; Moses, A.M.; Warringer, J.; Parts, L.; James, S.A.; Davey, R.P.;
Roberts, I.N.; Burt, A.; Koufopanou, V.; et al. Population genomics of domestic and wild yeasts.
Nature 2009, 458, 337–341.
36. Onge, R.P., St.; Mani, R.; Oh, J.; Proctor, M.; Fung, E.; Davis, R.W.; Nislow, C.; Roth, F.P.;
Giaever, G. Systematic pathway analysis using high-resolution fitness profiling of combinatorial
gene deletions. Nat. Genet. 2007, 39, 199–206.
Genes 2015, 643
37. Shah, N.A.; Laws, R.J.; Wardman, B.; Zhao, L.P.; Hartman, J.L., 4th. Accurate, precise modeling
of cell proliferation kinetics from time-lapse imaging and automated image analysis of agar yeast
culture arrays. BMC Syst. Biol. 2007, doi:10.1186/1752-0509-1-3.
38. Schekman, R. Lasker basic medical research award. Sec mutants and the secretory apparatus.
Nat. Med. 2002, 8, 1055–1058.
39. Hartwell, L.H. Nobel lecture. Yeast and cancer. Biosci. Rep. 2002, 22, 373–394.
40. Willingham, S.; Outeiro, T.F.; DeVit, M.J.; Lindquist, S.L.; Muchowski, P.J. Yeast genes
that enhance the toxicity of a mutant huntingtin fragment or alpha-synuclein. Science 2003, 302,
1769–1772.
41. Auluck, P.K.; Caraveo, G.; Lindquist, S. Alpha-synuclein: Membrane interactions and toxicity in
parkinson’s disease. Annu. Rev. Cell Dev. Biol. 2010, 26, 211–233.
42. Treusch, S.; Hamamichi, S.; Goodman, J.L.; Matlack, K.E.; Chung, C.Y.; Baru, V.; Shulman, J.M.;
Parrado, A.; Bevis, B.J.; Valastyan, J.S.; et al. Functional links between abeta toxicity, endocytic
trafficking, and Alzheimer’s disease risk factors in yeast. Science 2011, 334, 1241–1245.
43. Armakola, M.; Higgins, M.J.; Figley, M.D.; Barmada, S.J.; Scarborough, E.A.; Diaz, Z.; Fang, X.;
Shorter, J.; Krogan, N.J.; Finkbeiner, S.; et al. Inhibition of RNA lariat debranching enzyme
suppresses TDP-43 toxicity in ALS disease models. Nat. Genet. 2012, 44, 1302–1309.
44. Gitler, A.D.; Bevis, B.J.; Shorter, J.; Strathearn, K.E.; Hamamichi, S.; Su, L.J.; Caldwell, K.A.;
Caldwell, G.A.; Rochet, J.C.; McCaffery, J.M.; et al. The parkinson’s disease protein alpha-synuclein
disrupts cellular rab homeostasis. Proc. Natl. Acad. Sci. USA 2008, 105, 145–150.
45. Johnson, B.S.; McCaffery, J.M.; Lindquist, S.; Gitler, A.D. A yeast TDP-43 proteinopathy model:
Exploring the molecular determinants of TDP-43 aggregation and cellular toxicity. Proc. Natl. Acad.
Sci. USA 2008, 105, 6439–6444.
46. Couplan, E.; Aiyar, R.S.; Kucharczyk, R.; Kabala, A.; Ezkurdia, N.; Gagneur, J.; Onge, R.P., St.;
Salin, B.; Soubigou, F.; le Cann, M.; et al. A yeast-based assay identifies drugs active against human
mitochondrial disorders. Proc. Natl. Acad. Sci. USA 2011, 108, 11989–11994.
47. Vilaca, R.; Silva, E.; Nadais, A.; Teixeira, V.; Matmati, N.; Gaifem, J.; Hannun, Y.A.;
Sa Miranda, M.C.; Costa, V. Sphingolipid signalling mediates mitochondrial dysfunctions and reduced
chronological lifespan in the yeast model of niemann-pick type c1. Mol. Microbiol. 2014, 91, 438–451.
48. Zhang, H.; Singh, K.K. Global genetic determinants of mitochondrial DNA copy number.
PLOS ONE 2014, 9, e105242.
49. Rinaldi, T.; Dallabona, C.; Ferrero, I.; Frontali, L.; Bolotin-Fukuhara, M. Mitochondrial diseases
and the role of the yeast models. FEMS Yeast Res. 2010, 10, 1006–1022.
50. Gitler, A.D.; Lehmann, R. Modeling human disease. Science 2012, doi:10.1126/science.1227179.
51. Mani, R.; Onge, R.P., St.; Hartman, J.L., 4th; Giaever, G.; Roth, F.P. Defining genetic interaction.
Proc. Natl. Acad. Sci. USA 2008, 105, 3461–3466.
52. Rodgers, J.; Guo, J.; Hartman, J.L., 4th. Phenomic assessment of genetic buffering by kinetic
analysis of cell arrays. Methods Mol. Biol. 2014, 1205, 187–208.
53. Carlborg, O.; Haley, C.S. Epistasis: Too often neglected in complex trait studies? Nat. Rev. Genet.
2004, 5, 618–625.
54. Moore, J.H.; Williams, S.M. Traversing the conceptual divide between biological and statistical
epistasis: Systems biology and a more modern synthesis. Bioessays 2005, 27, 637–646.
Genes 2015, 644
55. Badano, J.L.; Leitch, C.C.; Ansley, S.J.; May-Simera, H.; Lawson, S.; Lewis, R.A.; Beales, P.L.;
Dietz, H.C.; Fisher, S.; Katsanis, N. Dissection of epistasis in oligogenic bardet-biedl syndrome.
Nature 2006, 439, 326–330.
56. Phillips, P.C. Epistasis—The essential role of gene interactions in the structure and evolution of
genetic systems. Nat. Rev. Genet. 2008, 9, 855–867.
57. Cowper-Sal lari, R.; Cole, M.D.; Karagas, M.R.; Lupien, M.; Moore, J.H. Layers of epistasis:
Genome-wide regulatory networks and network approaches to genome-wide association studies.
Wiley Interdiscip. Rev. Syst. Biol. Med. 2011, 3, 513–526.
58. Breen, M.S.; Kemena, C.; Vlasov, P.K.; Notredame, C.; Kondrashov, F.A. Epistasis as the primary
factor in molecular evolution. Nature 2012, 490, 535–538.
59. Knowles, M.R.; Drumm, M. The influence of genetics on cystic fibrosis phenotypes. Cold Spring
Harb. Perspect. Med. 2012, doi:10.1101/cshperspect.a009548.
60. Hartman, J.L., 4th. Buffering of deoxyribonucleotide pool homeostasis by threonine metabolism.
Proc. Natl. Acad. Sci. USA 2007, 104, 11700–11705.
61. Reichard, P. Ribonucleotide reductases: The evolution of allosteric regulation. Arch. Biochem. Biophys.
2002, 397, 149–155.
62. Pereira, S.; Fernandes, P.A.; Ramos, M.J. Mechanism for ribonucleotide reductase inactivation by
the anticancer drug gemcitabine. J. Comput. Chem. 2004, 25, 1286–1294.
63. Cerqueira, N.M.; Pereira, S.; Fernandes, P.A.; Ramos, M.J. Overview of ribonucleotide reductase
inhibitors: An appealing target in anti-tumour therapy. Curr. Med. Chem. 2005, 12, 1283–1294.
64. Thelander, L. Ribonucleotide reductase and mitochondrial DNA synthesis. Nat. Genet. 2007, 39,
703–704.
65. Wang, J.; Alexander, P.; Wu, L.; Hammer, R.; Cleaver, O.; McKnight, S.L. Dependence of mouse
embryonic stem cells on threonine catabolism. Science 2009, 325, 435–439.
66. Ryu, J.M.; Han, H.J. L-Threonine regulates G1/S phase transition of mouse embryonic stem cells
via Pi3k/Akt, MAPKs, and mTORC pathways. J. Biol. Chem. 2011, 286, 23667–23678.
67. Shyh-Chang, N.; Locasale, J.W.; Lyssiotis, C.A.; Zheng, Y.; Teo, R.Y.; Ratanasirintrawoot, S.;
Zhang, J.; Onder, T.; Unternaehrer, J.J.; Zhu, H.; et al. Influence of threonine metabolism on
S-adenosylmethionine and histone methylation. Science 2013, 339, 222–226.
68. Murakami, C.; Kaeberlein, M. Quantifying yeast chronological life span by outgrowth of aged cells.
J. Vis. Exp. 2009, doi:10.3791/1156.
69. Hess, D.C.; Lu, W.; Rabinowitz, J.D.; Botstein, D. Ammonium toxicity and potassium limitation in
yeast. PLOS Biol. 2006, doi:10.1371/journal.pbio.0040351.
70. Burke, D.; Dawson, D.; Stearns, T. Methods in Yeast Genetics; CSHL Press: Plainview, NY, USA, 2000.
71. Boyle, E.I.; Weng, S.; Gollub, J.; Jin, H.; Botstein, D.; Cherry, J.M.; Sherlock, G. Go: Termfinder—
Open source software for accessing gene ontology information and finding significantly enriched
gene ontology terms associated with a list of genes. Bioinformatics 2004, 20, 3710–3715.
72. Kaeberlein, M. Lessons on longevity from budding yeast. Nature 2010, 464, 513–519.
73. Sutphin, G.L.; Olsen, B.A.; Kennedy, B.K.; Kaeberlein, M. Genome-wide analysis of yeast aging.
Subcell. Biochem. 2012, 57, 251–289.
Genes 2015, 645
74. Roberg, K.J.; Bickel, S.; Rowley, N.; Kaiser, C.A. Control of amino acid permease sorting in the
late secretory pathway of saccharomyces cerevisiae by SEC13, LST4, LST7 and LST8. Genetics
1997, 147, 1569–1584.
75. Goldstein, A.L.; McCusker, J.H. Three new dominant drug resistance cassettes for gene disruption
in Saccharomyces cerevisiae. Yeast 1999, 15, 1541–1553.
76. Stepanov, A.; Nitiss, K.C.; Neale, G.; Nitiss, J.L. Enhancing drug accumulation in saccharomyces
cerevisiae by repression of pleiotropic drug resistance genes with chimeric transcription repressors.
Mol. Pharmacol. 2008, 74, 423–431.
77. Fleming, J.A.; Lightcap, E.S.; Sadis, S.; Thoroddsen, V.; Bulawa, C.E.; Blackman, R.K.
Complementary whole-genome technologies reveal the cellular response to proteasome inhibition
by ps-341. Proc. Natl. Acad. Sci. USA 2002, 99, 1461–1466.
78. Weinert, T.A.; Kiser, G.L.; Hartwell, L.H. Mitotic checkpoint genes in budding yeast and the
dependence of mitosis on DNA replication and repair. Genes Dev. 1994, 8, 652–665.
79. Paulovich, A.G.; Hartwell, L.H. A checkpoint regulates the rate of progression through S phase in
S. cerevisiae in response to DNA damage. Cell 1995, 82, 841–847.
80. Botstein, D. Why we need more basic biology research, not less. Mol. Biol. Cell 2012, 23,
4160–4161.
81. Chakravarti, A.; Clark, A.G.; Mootha, V.K. Distilling pathophysiology from complex disease
genetics. Cell 2013, 155, 21–26.
82. Botstein, D.; Fink, G.R. Yeast: An experimental organism for 21st century biology. Genetics 2011,
189, 695–704.
83. Dolinski, K.; Botstein, D. Orthology and functional conservation in eukaryotes. Annu. Rev. Genet.
2007, 41, 465–507.
84. Califano, A.; Butte, A.J.; Friend, S.; Ideker, T.; Schadt, E. Leveraging models of cell regulation and
gwas data in integrative network-based association studies. Nat. Genet. 2012, 44, 841–847.
85. Friend, S.H.; Schadt, E.E. Translational genomics. Clues from the resilient. Science 2014, 344,
970–972.
© 2015 by the authors; licensee MDPI, Basel, Switzerland. This article is an open access article
distributed under the terms and conditions of the Creative Commons Attribution license
(http://creativecommons.org/licenses/by/4.0/).
... To promote systems level analysis of yeast quiescence, we sought to characterize auxotrophy and media composition influences broadly, in order to build a better foundation for reproducible studies by helping explain whether the weak correlation between previous genome-wide studies of CLS in the YKO libraries might be attributable to the use of different media, auxotrophic backgrounds, and/or phenotypic assays (Smith et al. 2016). The studies described here used quantitative high throughput cell array phenotyping (Q-HTCP), a robotic method for obtaining tens of thousands of growth curves per assay, and human-like media (HL), which was designed to eliminate ammonium toxicity and with greater resemblance to human tissue culture media (Hartman IV et al. 2015). Using this approach in highly controlled experiments, all possible auxotrophic allele combinations, availability of corresponding nutrients, glucose, or ammonium sulfate, as well as perturbation of TORC1 signaling or replication stress were examined for effects on quiescence development and maintenance by frequent monitoring of colony forming capacity extending past 60 days to characterize different stages of quiescence development. ...
... We developed HL media to eliminate the effect of ammonium toxicity (Hess et al. 2006;Santos et al. 2013;Santos et al. 2012) and improve translational relevance between yeast and human cells by increasing similarity between yeast and human cell culture media (Hartman IV et al. 2015). Influences of glucose, ammonium sulfate, and selected auxotrophic backgrounds on quiescence in HL media were evidenced by quiescence profiles for BY4741, BY4742, FY4 (prototrophic parent) and BY4712 (MATa leu2∆0), using dextrose concentrations of 0.4, 2, and 5%, each with and without 0.5 gm ammonium sulfate (Fig. 3). ...
... ; https://doi.org/10.1101/2020.05.25.114801 doi: bioRxiv preprint (Hartman IV et al. 2015). Using a 384-well format, uniform aeration was achieved by inverting the culture plates (Santos and Hartman IV 2019). ...
Preprint
Full-text available
Yeast cells survive in stationary phase culture by entering quiescence, which is measured by colony forming capacity upon nutrient re-exposure. Yeast chronological lifespan (CLS) studies, employing the comprehensive collection of gene knockout strains, have correlated weakly between independent laboratories, which is hypothesized to reflect differential interaction between the deleted genes, auxotrophy, media composition and other assay conditions influencing quiescence. This hypothesis was investigated by high-throughput quiescence profiling of the parental prototrophic strain, from which the gene deletion strain libraries were constructed, and all possible auxotrophic allele combinations in that background. Defined media resembling human cell culture media promoted long-term quiescence, and was used to assess effects of glucose, ammonium sulfate, auxotrophic nutrient availability, Target of Rapamycin signaling, and replication stress. Frequent, high-replicate measurements of colony forming capacity from cultures aged past 60 days provided profiles of quiescence phenomena such as gasping and hormesis. Media acidification was assayed in parallel to assess correlation. Influences of leucine, methionine, glucose, and ammonium sulfate metabolism were clarified, and a role for lysine metabolism newly characterized, while histidine and uracil perturbations had less impact. Interactions occurred between glucose, ammonium sulfate, auxotrophy, auxotrophic nutrient limitation, aeration, TOR signaling, and/or replication stress. Weak correlation existed between media acidification and maintenance of quiescence. In summary, experimental factors, uncontrolled across previous genome-wide yeast CLS studies, influence quiescence and interact extensively, revealing quiescence as a complex metabolic and developmental process that should be studied in a prototrophic context, omitting ammonium sulfate from defined media, and employing highly replicable protocols.
... The concentration of L-serine in our standard SC media is 1 mM (47,48), so we tested the effect of supplementing an additional 1 mM or 5 mM L-serine into NR cultures at the time of inoculation ( Figure 5A and B). 5 mM L-serine significantly extended CLS while 1 mM did not. To confirm the L-serine effect was not specific to SC media, we also tested for CLS extension in a custom synthetic growth medium (HL) designed to support longevity that does not have ammonium sulfate as a nitrogen source (36,49). BY4741 had significantly longer CLS in nonrestricted HL medium compared to SC medium, and 5 mM L-serine further extended it ( Figure S4A and B). ...
... CR could also potentially make ammonium sulfate a preferred nitrogen source over the amino acids that are usually preferred under non-restricted conditions. Ammonium sulfate has been shown to reduce CLS and is actually left out of the custom HL medium designed to optimize CLS (49,78). Therefore, assimilation of the ammonium under CR could potentially extend CLS by reducing ammonium toxicity, similar to the CR-induced consumption of acetic acid (38,48). ...
... Yeast strains and media S. cerevisiae strains used in this study were BY4741 (MATa his3∆1 leu2∆0 met15∆0 ura3∆0), FY4 (MATa prototrophic), and several deletion mutants from the Euroscarf yeast knockout (YKO) collection in the BY4741 background (89). Synthetic complete (SC) growth medium was used for all experiments except for the use of custom 'human-like' HL media (36,49). A recipe for SC media with individual amino acid concentrations is provided in Supplemental Table S6 (47). ...
Article
Full-text available
Caloric restriction (CR) improves healthspan and lifespan of organisms ranging from yeast to mammals. Understanding the mechanisms involved will uncover future interventions for aging associated diseases. In budding yeast, Saccharomyces cerevisiae , CR is commonly defined by reduced glucose in the growth medium, which extends both replicative and chronological lifespan (CLS). We found that conditioned media collected from stationary phase CR cultures extended CLS when supplemented into non-restricted (NR) cultures, suggesting a potential cell non-autonomous mechanism of CR-induced lifespan regulation. Chromatography and untargeted metabolomics of the conditioned media, as well as transcriptional responses associated with the longevity effect, pointed to specific amino acids enriched in the CR conditioned media (CRCM) as functional molecules, with L-serine being a particularly strong candidate. Indeed, supplementing L-serine into NR cultures extended CLS through a mechanism dependent on the one-carbon metabolism pathway, thus implicating this conserved and central metabolic hub in lifespan regulation.
... Understanding the influence of quiescence mechanisms on organismal function is fundamental to aging biology, for example, in stem cell function, and the developmental interplay with senescence, apoptosis, and necrosis [19]. In these regards, yeast is a useful genetic model for unbiased discovery and hypothesis generation about human conditions [36,37,40]. It is well positioned to inform eukaryotic quiescence mechanisms linking nutrient signaling, quiescence, and longevity in multicellular organisms [21,90]. ...
... To promote systems level analysis of yeast quiescence, we sought to characterize auxotrophy and media composition influences broadly, in order to build a better foundation for reproducible studies by explaining how the weak correlation between previous genomewide studies of CLS in the YKO libraries is attributable to the use of different media, auxotrophic backgrounds, and/or phenotypic assays [85]. The studies described here used quantitative high-throughput cell array phenotyping (Q-HTCP), a robotic method for obtaining tens of thousands of growth curves per assay, and human-like media (HL), which was designed to eliminate ammonium toxicity and with greater resemblance to human tissue culture media [37]. Using this approach, experiments controlled for all possible auxotrophic allele combinations, availability of corresponding nutrients, glucose, or ammonium sulfate, as well as perturbation of TORC1 signaling or replication stress were conducted to examine effects on quiescence development and maintenance. ...
... Tetrad analysis revealed reduced spore viability, which improved in a second backcross Online Resource 2- Table S3. Human-like (HL) yeast media was used for all experiments [37]; its composition is detailed in Online Resource 2- Table S4. When supplemented, the ammonium sulfate concentration was 0.5 g/L (in accord with the reduced potassium in HL YNB; see [41]). ...
Article
Full-text available
Yeast cells survive in stationary phase culture by entering quiescence, which is measured by colony-forming capacity upon nutrient re-exposure. Yeast chronological lifespan (CLS) studies, employing the comprehensive collection of gene knockout strains, have correlated weakly between independent laboratories, which is hypothesized to reflect differential interaction between the deleted genes, auxotrophy, media composition, and other assay conditions influencing quiescence. This hypothesis was investigated by high-throughput quiescence profiling of the parental prototrophic strain, from which the gene deletion strain libraries were constructed, and all possible auxotrophic allele combinations in that background. Defined media resembling human cell culture media promoted long-term quiescence and was used to assess effects of glucose, ammonium sulfate, auxotrophic nutrient availability, target of rapamycin signaling, and replication stress. Frequent, high-replicate measurements of colony-forming capacity from cultures aged past 60 days provided profiles of quiescence phenomena such as gasping and hormesis. Media acidification was assayed in parallel to assess correlation. Influences of leucine, methionine, glucose, and ammonium sulfate metabolism were clarified, and a role for lysine metabolism newly characterized, while histidine and uracil perturbations had less impact. Interactions occurred between glucose, ammonium sulfate, auxotrophy, auxotrophic nutrient limitation, aeration, TOR signaling, and/or replication stress. Weak correlation existed between media acidification and maintenance of quiescence. In summary, experimental factors, uncontrolled across previous genome-wide yeast CLS studies, influence quiescence and interact extensively, revealing quiescence as a complex metabolic and developmental process that should be studied in a prototrophic context, omitting ammonium sulfate from defined media, and employing highly replicable protocols.
... To address these questions, this work establishes a yeast phenomic model to understand genetic pathways that buffer doxorubicin toxicity [24][25][26][27][28][29][30], and how the Warburg effect influences the doxorubicin-gene interaction network. We conducted yeast phenomic analysis of doxorubicin-gene interaction, consisting of quantitative high throughput cell array phenotyping (Q-HTCP) of the yeast knockout and knockdown (YKO/KD) libraries, using multiple growth inhibitory concentrations of doxorubicin in either dextrose-(HLD) or ethanol/glycerol-based (HLEG) media. ...
... Some mutants appear multiple times in the library and they are treated independently in our analysis. HL yeast media, a modified synthetic complete media [27], was used with either 2% dextrose (HLD) or 3% ethanol and 3% glycerol (HLEG) as the carbon source. ...
... Overlap between these studies and ours is shown in Fig. 2p-r and in Additional file 7: Table S9-10. While many genes overlapped between the studies, differing results were also observed, possibly attributable to strain background, media conditions, and methods for scoring interactions [27,45]. To assess within-study reproducibility, we sub-cloned four colonies from glycerol stocks used in the first experiment and retested doxorubicin-gene interaction, revealing higher correlation and overall reproducibility within-study than between-study (Fig. 2s, t). ...
Article
Full-text available
Background: The influence of the Warburg phenomenon on chemotherapy response is unknown. Saccharomyces cerevisiae mimics the Warburg effect, repressing respiration in the presence of adequate glucose. Yeast phenomic experiments were conducted to assess potential influences of Warburg metabolism on gene-drug interaction underlying the cellular response to doxorubicin. Homologous genes from yeast phenomic and cancer pharmacogenomics data were analyzed to infer evolutionary conservation of gene-drug interaction and predict therapeutic relevance. Methods: Cell proliferation phenotypes (CPPs) of the yeast gene knockout/knockdown library were measured by quantitative high-throughput cell array phenotyping (Q-HTCP), treating with escalating doxorubicin concentrations under conditions of respiratory or glycolytic metabolism. Doxorubicin-gene interaction was quantified by departure of CPPs observed for the doxorubicin-treated mutant strain from that expected based on an interaction model. Recursive expectation-maximization clustering (REMc) and Gene Ontology (GO)-based analyses of interactions identified functional biological modules that differentially buffer or promote doxorubicin cytotoxicity with respect to Warburg metabolism. Yeast phenomic and cancer pharmacogenomics data were integrated to predict differential gene expression causally influencing doxorubicin anti-tumor efficacy. Results: Yeast compromised for genes functioning in chromatin organization, and several other cellular processes are more resistant to doxorubicin under glycolytic conditions. Thus, the Warburg transition appears to alleviate requirements for cellular functions that buffer doxorubicin cytotoxicity in a respiratory context. We analyzed human homologs of yeast genes exhibiting gene-doxorubicin interaction in cancer pharmacogenomics data to predict causality for differential gene expression associated with doxorubicin cytotoxicity in cancer cells. This analysis suggested conserved cellular responses to doxorubicin due to influences of homologous recombination, sphingolipid homeostasis, telomere tethering at nuclear periphery, actin cortical patch localization, and other gene functions. Conclusions: Warburg status alters the genetic network required for yeast to buffer doxorubicin toxicity. Integration of yeast phenomic and cancer pharmacogenomics data suggests evolutionary conservation of gene-drug interaction networks and provides a new experimental approach to model their influence on chemotherapy response. Thus, yeast phenomic models could aid the development of precision oncology algorithms to predict efficacious cytotoxic drugs for cancer, based on genetic and metabolic profiles of individual tumors.
... A potential advantage of so-called synthetic 2 lethality-based treatment strategies is that they could have efficacy against passenger gene mutation or compensatory gene expression, while classic targeted therapies are directed primarily at driver genes ( Figure 1A). Quantitative high throughput cell array phenotyping of the yeast knockout and knockdown libraries provides a phenomic means for systems level, high-resolution modeling of gene interaction [5][6][7][8][9], which is applied here to predict cancer-relevant drug-gene interaction through integration with cancer pharmacogenomics resources ( Figure 1B). ...
... Saccharomyces cerevisiae does not have a dCK homolog and is thus naturally resistant to gemcitabine and cytarabine. To examine the gene-drug interaction networks for gemcitabine and cytarabine in yeast, we introduced human dCK into the yeast knockout and knockdown (YKO/KD) library by the synthetic genetic array (SGA) method [27][28][29] and conducted phenomic analysis on the resulting double mutant library by quantitative high-throughput cell array phenotyping (Q-HTCP) [6][7][8][9] using multiple growth inhibitory concentrations of gemcitabine or cytarabine ( Figure 1B). Cell proliferation parameters (CPPs) obtained by Q-HTCP were used to quantify and compare drug-gene interaction for gemcitabine vs. cytarabine. ...
... Some mutants appear multiple times in the library and they are treated independently in our analysis. HLD is a modified synthetic complete medium [8] and was used with 2% dextrose (HLD) as the carbon source. Doxycycline hydrochloride (BP26535) was obtained from Fisher Scientific. ...
Article
Full-text available
Knowledge about synthetic lethality can be applied to enhance the efficacy of anticancer therapies in individual patients harboring genetic alterations in their cancer that specifically render it vulnerable. We investigated the potential for high-resolution phenomic analysis in yeast to predict such genetic vulnerabilities by systematic, comprehensive, and quantitative assessment of drug–gene interaction for gemcitabine and cytarabine, substrates of deoxycytidine kinase that have similar molecular structures yet distinct antitumor efficacy. Human deoxycytidine kinase (dCK) was conditionally expressed in the Saccharomyces cerevisiae genomic library of knockout and knockdown (YKO/KD) strains, to globally and quantitatively characterize differential drug–gene interaction for gemcitabine and cytarabine. Pathway enrichment analysis revealed that autophagy, histone modification, chromatin remodeling, and apoptosis-related processes influence gemcitabine specifically, while drug–gene interaction specific to cytarabine was less enriched in gene ontology. Processes having influence over both drugs were DNA repair and integrity checkpoints and vesicle transport and fusion. Non-gene ontology (GO)-enriched genes were also informative. Yeast phenomic and cancer cell line pharmacogenomics data were integrated to identify yeast–human homologs with correlated differential gene expression and drug efficacy, thus providing a unique resource to predict whether differential gene expression observed in cancer genetic profiles are causal in tumor-specific responses to cytotoxic agents.
... For screening and prioritization, the types of data used are more flexible and can be derived from various species, tissues, cell types and levels of mechanistic understanding. For example, data from invertebrate species like yeast and daphnia have proven very informative for comparative toxicity testing and investigation of basic biology (Garcia-Reyero et al. 2012;Hartman et al. 2015;Goldstein and King 2016;Gust et al. 2016;North et al. 2016). For regulatory assessment, we currently prefer data from either human cells or from vertebrate cells focusing on highly conserved processes in the tissue of interest, because of the importance on cell-type and tissue identity in disease (Greene et al. 2015;Gross and Ideker 2015). ...
... Due to functional interdependencies among molecular pathways, a disease is rarely a consequence of an abnormality in a single gene or even pathway, but reflects disruption of complex intracellular networks. Many genes contribute to each phenotype and each gene contributes to multiple phenotypes (Goh et al. 2007;Hartman et al. 2015;Darabos et al. 2016). Importantly, individuals or subpopulations with the same disease can have different perturbations (Schadt 2009;Barabasi et al. 2011;Ideker and Krogan 2012). ...
Article
Full-text available
Advances in the biological sciences have led to an ongoing paradigm shift in toxicity testing based on expanded application of high-throughput in vitro screening and in silico methods to assess potential health risks of environmental agents. This review examines progress on the vision for toxicity testing elaborated by the US National Research Council (NRC) during the decade that has passed since the 2007 NRC report on Toxicity Testing in the 21st Century (TT21C). Concomitant advances in exposure assessment, including computational approaches and high-throughput exposomics, are also documented. A vision for the next generation of risk science, incorporating risk assessment methodologies suitable for the analysis of new toxicological and exposure data, resulting in human exposure guidelines is described. Case study prototypes indicating how these new approaches to toxicity testing, exposure measurement, and risk assessment are beginning to be applied in practice are presented. Overall, progress on the 20-year transition plan laid out by the US NRC in 2007 has been substantial. Importantly, government agencies within the United States and internationally are beginning to incorporate the new approach methodologies envisaged in the original TT21C vision into regulatory practice. Future perspectives on the continued evolution of toxicity testing to strengthen regulatory risk assessment are provided.
... preconceived definitions of what growth should look like (Hartman et al. 2015;Banks et al. 2012;Narayanan et al. 2015). Estimation of cell densities can be enhanced after serial dilutions (Hartman et al. 2015) but at the cost of extra robotic steps that inevitably will increase errors and decrease throughput. ...
... preconceived definitions of what growth should look like (Hartman et al. 2015;Banks et al. 2012;Narayanan et al. 2015). Estimation of cell densities can be enhanced after serial dilutions (Hartman et al. 2015) but at the cost of extra robotic steps that inevitably will increase errors and decrease throughput. ...
Article
Full-text available
The capacity to map traits over large cohorts of individuals - phenomics - lags far behind the explosive development in genomics. For microbes the estimation of growth is the key phenotype because of its link to fitness. We introduce an automated microbial phenomics framework that delivers accurate, precise and highly resolved growth phenotypes at an unprecedented scale. Advancements were achieved through introduction of transmissive scanning hardware and software technology, frequent acquisition of exact colony population size measurements, extraction of population growth rates from growth curves and removal of spatial bias by reference-surface normalization. Our prototype arrangement automatically records and analyzes close to 100,000 growth curves in parallel. We demonstrate the power of the approach by extending and nuancing the known salt-defense biology in baker's yeast. The introduced framework represents a major advance in microbial phenomics by providing high-quality data for extensive cohorts of individuals and generating well-populated and standardized phenomics databases.
Article
Yeast cells are extensively used as a key model organism owing to their highly conserved genome, metabolic pathways, and cell biology processes. To assist in genetic engineering and analysis, laboratory yeast strains typically harbor auxotrophic selection markers. When uncompensated, auxotrophic markers cause significant phenotypic bias compared to prototrophic strains and have different combinatorial influences on the metabolic network. Here, we used BY4741, a laboratory strain commonly used as a “wild type” strain in yeast studies, to generate a set of revertant strains, containing all possible combinations of four common auxotrophic markers ( leu2 ∆, ura3 ∆, his3 ∆1, met15 ∆). We examined the effect of the auxotrophic combinations on complex phenotypes such as resistance to rapamycin, acetic acid, and ethanol. Among the markers, we found that leucine auxotrophy most significantly affected the phenotype. We analyzed the phenotypic bias caused by auxotrophy at the genomic level using a prototrophic version of a genome‐wide deletion library and a decreased mRNA perturbation (DAmP) library. Prototrophy was found to suppress rapamycin sensitivity in many mutants previously annotated for the phenotype, raising a possible need for reevaluation of the findings in a native metabolic context. These results reveal a significant phenotypic bias caused by common auxotrophic markers and support the use of prototrophic wild‐type strains in yeast research.
Preprint
Full-text available
Knowledge about synthetic lethality can be applied to enhance the efficacy of anti-cancer therapies in individual patients harboring genetic alterations in their cancer that specifically render it vulnerable. We investigated the potential for high-resolution phenomic analysis in yeast to predict such genetic vulnerabilities by systematic, comprehensive, and quantitative assessment of drug-gene interaction for gemcitabine and cytarabine, substrates of deoxycytidine kinase that have similar molecular structures yet distinct anti-tumor efficacy. Human deoxycytidine kinase (dCK) was conditionally expressed in the S. cerevisiae genomic library of knockout and knockdown (YKO/KD) strains to globally and quantitatively characterize differential drug-gene interaction for gemcitabine and cytarabine. Pathway enrichment analysis revealed that autophagy, histone modification, chromatin remodeling, and apoptosis-related processes influence gemcitabine specifically, while drug-gene interaction specific to cytarabine was less enriched in Gene Ontology. Processes having influence over both drugs were DNA repair and integrity checkpoints and vesicle transport and fusion. Non-GO enriched genes were also informative. Yeast phenomic and cancer cell line pharmacogenomics data were integrated to identify yeast-human homologs with correlated differential gene expression and drug-efficacy, thus providing a unique resource to predict whether differential gene expression observed in cancer genetic profiles are causal in tumor-specific responses to cytotoxic agents.
Chapter
Yeasts have proven to be an invaluable model organism to explore the fundamental cellular processes and pathways conserved across eukaryotic organisms. The wide array of available genetic and genomic tools, coupled with experimental tractability, make them ideal organisms for genetic research. A significant contribution to our systemic understanding of many human diseases has been made by studies carried out in yeasts. High-throughput genetic screens have been used to identify human disease genes, as well as to dissect the molecular pathways that regulate the function of disease related proteins. Recent developments in the areas of chemical-genetics and chemical genomics have further highlighted the importance of yeasts in the simultaneous analysis of a large number of drugs, as well as facilitating the identification of their mechanism of action. In this chapter, we describe the various genetic tools that have been used by yeast researchers to increase our understanding of the basic human biology and provide insights into the molecular mechanisms underlying various human diseases.
Article
Full-text available
Quantitative high-throughput cell array phenotyping (Q-HTCP) is applied to the genomic collection of yeast gene deletion mutants for systematic, comprehensive assessment of the contribution of genes and gene combinations to any phenotype of interest (phenomic analysis). Interacting gene networks influence every phenotype. Genetic buffering refers to how gene interaction networks stabilize or destabilize a phenotype. Like genomics, phenomics varies in its resolution with there being a trade-off allocating a greater number of measurements per sample to enhance quantification of the phenotype vs. increasing the number of different samples by obtaining fewer measurements per sample. The Q-HTCP protocol we describe assesses 50,000-70,000 cultures per experiment by obtaining kinetic growth curves from time series imaging of agar cell arrays. This approach was developed for the yeast gene deletion strains, but it could be applied as well to other microbial mutant arrays grown on solid agar media. The methods we describe are for creation and maintenance of frozen stocks, liquid source array preparation, agar destination plate printing, image scanning, image analysis, curve fitting, and evaluation of gene interaction.
Article
Full-text available
Many human diseases including development of cancer is associated with depletion of mitochondrial DNA (mtDNA) content. These diseases are collectively described as mitochondrial DNA depletion syndrome (MDS). High similarity between yeast and human mitochondria allows genomic study of the budding yeast to be used to identify human disease genes. In this study, we systematically screened the pre-existing respiratory-deficient Saccharomyces cerevisiae yeast strains using fluorescent microscopy and identified 102 nuclear genes whose deletions result in a complete mtDNA loss, of which 52 are not reported previously. Strikingly, these genes mainly encode protein products involved in mitochondrial protein biosynthesis process (54.9%). The rest of these genes either encode protein products associated with nucleic acid metabolism (14.7%), oxidative phosphorylation (3.9%), or other protein products (13.7%) responsible for bud-site selection, mitochondrial intermembrane space protein import, assembly of cytochrome-c oxidase, vacuolar protein sorting, protein-nucleus import, calcium-mediated signaling, heme biosynthesis and iron homeostasis. Thirteen (12.7%) of the genes encode proteins of unknown function. We identified human orthologs of these genes, conducted the interaction between the gene products and linked them to human mitochondrial disorders and other pathologies. In addition, we screened for genes whose defects affect the nuclear genome integrity. Our data provide a systematic view of the nuclear genes involved in maintenance of mitochondrial DNA. Together, our studies i) provide a global view of the genes regulating mtDNA content; ii) provide compelling new evidence toward understanding novel mechanism involved in mitochondrial genome maintenance and iii) provide useful clues in understanding human diseases in which mitochondrial defect and in particular depletion of mitochondrial genome plays a critical role.
Article
Full-text available
Despite recent advances in human genetics, model organisms are indispensable for human disease research. Most human disease pathways are evolutionally conserved among other species, where they may phenocopy the human condition or be associated with seemingly unrelated phenotypes. Much of the known gene-to-phenotype association information is distributed across diverse databases, growing rapidly due to new experimental techniques. Accessible bioinformatics tools will therefore facilitate translation of discoveries from model organisms into human disease biology. Here, we present a web-based discovery tool for human disease studies, MORPHIN (model organisms projected on a human integrated gene network), which prioritizes the most relevant human diseases for a given set of model organism genes, potentially highlighting new model systems for human diseases and providing context to model organism studies. Conceptually, MORPHIN investigates human diseases by an orthology-based projection of a set of model organism genes onto a genome-scale human gene network. MORPHIN then prioritizes human diseases by relevance to the projected model organism genes using two distinct methods: a conventional overlap-based gene set enrichment analysis and a network-based measure of closeness between the query and disease gene sets capable of detecting associations undetectable by the conventional overlap-based methods. MORPHIN is freely accessible at http://www.inetbio.org/morphin.
Article
Full-text available
Recent research suggests that genetic interactions involving more than two loci may influence a number of complex traits. How these 'higher-order' interactions arise at the genetic and molecular levels remains an open question. To provide insights into this problem, we dissected a colony morphology phenotype that segregates in a yeast cross and results from synthetic higher-order interactions. Using backcrossing and selective sequencing of progeny, we found five loci that collectively produce the trait. We fine-mapped these loci to 22 genes in total and identified a single gene at each locus that caused loss of the phenotype when deleted. Complementation tests or allele replacements provided support for functional variation in these genes, and revealed that pre-existing genetic variants and a spontaneous mutation interact to cause the trait. The causal genes have diverse functions in endocytosis (END3), oxidative stress response (TRR1), RAS-cAMP signalling (IRA2), and transcriptional regulation of multicellular growth (FLO8 and MSS11), and for the most part have not previously been shown to exhibit functional relationships. Further efforts uncovered two additional loci that together can complement the non-causal allele of END3, suggesting that multiple genotypes in the cross can specify the same phenotype. Our work sheds light on the complex genetic and molecular architecture of higher-order interactions, and raises questions about the broader contribution of such interactions to heritable trait variation.
Article
Disruption-deletion cassettes are powerful tools used to study gene function in many organisms, including Saccharomyces cerevisiae. Perhaps the most widely useful of these are the heterologous dominant drug resistance cassettes, which use antibiotic resistance genes from bacteria and fungi as selectable markers. We have created three new dominant drug resistance cassettes by replacing the kanamycin resistance (kanr) open reading frame from the kanMX3 and kanMX4 disruption-deletion cassettes (Wach et al., 1994) with open reading frames conferring resistance to the antibiotics hygromycin B (hph), nourseothricin (nat) and bialaphos (pat). The new cassettes, pAG25 (natMX4), pAG29 (patMX4), pAG31 (patMX3), pAG32 (hphMX4), pAG34 (hphMX3) and pAG35 (natMX3), are cloned into pFA6, and so are in all other respects identical to pFA6–kanMX3 and pFA6–kanMX4. Most tools and techniques used with the kanMX plasmids can also be used with the hph, nat and patMX containing plasmids. These new heterologous dominant drug resistance cassettes have unique antibiotic resistance phenotypes and do not affect growth when inserted into the ho locus. These attributes make the cassettes ideally suited for creating S. cerevisiae strains with multiple mutations within a single strain. Copyright © 1999 John Wiley & Sons, Ltd.
Article
The discovery of genes that control cell division in yeast, and their relation to cancer, is reviewed.
Article
The functions of many open reading frames (ORFs) identified in genome-sequencing projects are unknown. New, whole-genome approaches are required to systematically determine their function. A total of 6925 Saccharomyces cerevisiae strains were constructed, by a high-throughput strategy, each with a precise deletion of one of 2026 ORFs (more than one-third of the ORFs in the genome). Of the deleted ORFs, 17 percent were essential for viability in rich medium. The phenotypes of more than 500 deletion strains were assayed in parallel. Of the deletion strains, 40 percent showed quantitative growth defects in either rich or minimal medium.
Article
THE Genetics Society of America's Edward Novitski Prize recognizes an extraordinary level of creativity and intellectual ingenuity in the solution of significant problems in genetics research. The 2014 recipient, Charles Boone, has risen to the top of the emergent discipline of postgenome systems biology by focusing on the global mapping of genetic interaction networks. Boone invented the synthetic genetic array (SGA) technology, which provides an automated method to cross thousands of strains carrying precise mutations and map large-scale yeast genetic interactions. These network maps offer researchers a functional wiring diagram of the cell, which clusters genes into specific pathways and reveals functional connections.
Article
Calculating the probability of the corresponding significance point is important for finite sample sizes. However, it is difficult to evaluate this probability when the sample sizes are moderate to large. Under these circumstances, consideration of a more accurate approximation for the distribution function is extremely important. Herein, we performed a saddlepoint approximation in the upper tails for the distribution of the sum of independent non-identically uniform random variables under finite sample sizes. Saddlepoint approximation results were compared with those for a normal approximation. Additionally, the order of errors of the saddlepoint approximation was derived. © 2014 The Authors. Statistica Neerlandica © 2014 VVS.
Article
The genetics approach to uncovering the causes of disease has focused mainly on finding the underlying primary mutations, with diseased individuals playing the leading role in this discovery. But as health care begins to focus more on preventive therapies, an emphasis on understanding how individuals remain healthy—“resilient” to disease—may provide insights into disease pathogenesis and new treatments. This view underlies “The Resilience Project” ([www.resilienceproject.me][1]), an effort to search broadly for these apparently healthy people (see the photo). There are, indeed, individuals whose genetics indicate exceptionally high risk of disease, yet they never show any signs of the disorder. What are the genetic and environmental factors that buffer disease for them? How can such information be gathered and harnessed most efficiently and effectively? [1]: http://www.resilienceproject.me