ArticlePDF Available

Magnitude and sign epistasis among deleterious mutation in a positive-sense plant RNA virus

Authors:

Abstract and Figures

How epistatic interactions between mutations determine the genetic architecture of fitness is of central importance in evolution. The study of epistasis is particularly interesting for RNA viruses because of their genomic compactness, lack of genetic redundancy, and apparent low complexity. Moreover, interactions between mutations in viral genomes determine traits such as resistance to antiviral drugs, virulence and host range. In this study we generated 53 Tobacco etch potyvirus genotypes carrying pairs of single-nucleotide substitutions and measured their separated and combined deleterious fitness effects. We found that up to 38% of pairs had significant epistasis for fitness, including both positive and negative deviations from the null hypothesis of multiplicative effects. Interestingly, the sign of epistasis was correlated with viral protein-protein interactions in a model network, being predominantly positive between linked pairs of proteins and negative between unlinked ones. Furthermore, 55% of significant interactions were cases of reciprocal sign epistasis (RSE), indicating that adaptive landscapes for RNA viruses maybe highly rugged. Finally, we found that the magnitude of epistasis correlated negatively with the average effect of mutations. Overall, our results are in good agreement to those previously reported for other viruses and further consolidate the view that positive epistasis is the norm for small and compact genomes that lack genetic robustness.
Content may be subject to copyright.
ORIGINAL ARTICLE
Magnitude and sign epistasis among deleterious mutations
in a positive-sense plant RNA virus
J Lalic
´1and SF Elena1,2
How epistatic interactions between mutations determine the genetic architecture of fitness is of central importance in evolution.
The study of epistasis is particularly interesting for RNA viruses because of their genomic compactness, lack of genetic
redundancy, and apparent low complexity. Moreover, interactions between mutations in viral genomes determine traits such as
resistance to antiviral drugs, virulence and host range. In this study we generated 53 Tobacco etch potyvirus genotypes carrying
pairs of single-nucleotide substitutions and measured their separated and combined deleterious fitness effects. We found that
up to 38% of pairs had significant epistasis for fitness, including both positive and negative deviations from the null hypothesis
of multiplicative effects. Interestingly, the sign of epistasis was correlated with viral protein–protein interactions in a model
network, being predominantly positive between linked pairs of proteins and negative between unlinked ones. Furthermore, 55%
of significant interactions were cases of reciprocal sign epistasis (RSE), indicating that adaptive landscapes for RNA viruses
maybe highly rugged. Finally, we found that the magnitude of epistasis correlated negatively with the average effect of
mutations. Overall, our results are in good agreement to those previously reported for other viruses and further consolidate the
view that positive epistasis is the norm for small and compact genomes that lack genetic robustness.
Heredity (2012) 109, 71–77; doi:10.1038/hdy.2012.15; published online 11 April 2012
Keywords: epistasis; fitness landscapes; genome architecture; virus evolution
INTRODUCTION
Epistasis has been the focus of intensive research since the beginning
of genetics as a scientific discipline (Phillips, 2008). In general,
epistasis is the interaction between genes or mutations in determining
phenotypes. The direction, magnitude and prevalence of epistasis is
central to theories seeking to explain the origin of characteristics of
genetic systems, such as sex and recombination (De Visser and Elena,
2007), dominance (Bagheri and Wagner, 2004), ploidy (Kondrashov
and Crow, 1991), phenotypic plasticity (Remold and Lenski, 2004),
robustness (De Visser et al., 2003), the ruggedness of adaptive
landscapes (Weinreich et al., 2006; Poelwijk et al., 2007), or attempt-
ing to mechanistically explain dynamic biological processes such as
the accumulation of mutations in finite populations (Kondrashov,
1994), and speciation by reproductive isolation (Coyne, 1992). Very
recently, the evolutionary causes of epistasis, and not only their
evolutionary consequences, have also attracted attention (Sanjua
´nand
Nebot, 2008; De Visser et al., 2011; Macı
´aet al., 2012).
Broadly speaking, epistatic interactions can be classified as uni- or
multi-dimensional (Kondrashov and Kondrashov, 2001). Uni-dimen-
sional epistasis is defined as deviations from a linear relationship
between mean multiplicative fitness and the number of mutations
affecting fitness. By contrast, multi-dimensional epistasis includes all
the possible individual interactions among a set of mutations. Multi-
dimensional epistasis provides a more complete description of the
interactions within the fitness landscape defined by a set of mutations.
Interactions can be further classified as magnitude or as sign epistasis.
Magnitude epistasis (ME) occurs when that the fitness value
associated to a mutation, but not its sign, changes upon the genetic
background wherein it appears (Weinreich et al., 2005; Poelwijk et al.,
2007). Moreover, ME can be positive or negative, depending on
whether the fitness of the double mutant is larger or smaller than
expected under the multiplicative null model, respectively. ME is a
widespread phenomenon observed in organisms of different complex-
ity (Sanjua
´n and Elena, 2006). Sign epistasis (SE) refers to cases in
which the sign of the fitness effect of a mutation is under epistatic
control; thus, such a mutation is beneficial in some genetic back-
grounds and deleterious in others (Weinreich et al., 2005; Poelwijk
et al., 2007). In few instances where it has been sought, SE seems to be
quite common, although perhaps not as ubiquitous as ME (Weinreich
et al., 2006 Poelwijk et al., 2007; Franke et al., 2011; Kvitek and
Sherlock, 2011).
Epistasis is particularly relevant for our understanding of adaptive
evolution, as it determines the ruggedness of the adaptive landscape
(Withlock et al., 1995; Poelwijk et al., 2011) as well as the accessibility
of adaptive pathways throughout the landscape (Weinreich, 2005
Welch and Waxman, 2005; Franke et al., 2011). Evolutionary
trajectories may end at suboptimal fitness peaks due to the ruggedness
of the fitness landscape. Epistasis can therefore hamper the efficiency
of natural selection and thus slow down the rate of adaptation
(Withlock et al., 1995). Moreover, epistasis can make certain evolu-
tionary pathways towards higher fitness genotypes selectively inacces-
sible because of troughs and valleys in the fitness landscape:
intermediate genotypes have reduced fitness compared with sur-
rounding genotypes. Weinreich et al. (2005) were the first to notice
this evolutionary constraint and to postulate that such limitation
would arise only as a consequence of SE. Indeed, a particular type of
1Instituto de Biolo
´a Molecular y Celular de Plantas, Consejo Superior de Investigaciones Cientı
´ficas-UPV, Vale
`ncia, Spain and and 2Santa Fe Institute, Santa Fe NM, USA
Correspondence: Professor SF Elena, Instituto de Biologı
´a Molecular y Celular de Plantas, CSIC-UPV, Campus UPV CPI 8E, Ingeniero Fausto Elio s/n, 46022 Vale
`ncia, Spain.
E-mail: santiago.elena@csic.es
Received 3 October 2011; revised 21 February 2012; accepted 28 February 2012; published online 11 April 2012
Heredity (2012) 109, 71–77
&
2012 Macmillan Publishers Limited. All rights reserved 0018-067X/12
www.nature.com/hdy
SE known as RSE, that is, the sign of the fitness effect of a mutation is
conditional upon the state of another locus and vice versa, has been
shown to be a necessary condition for an adaptive landscape to be
rugged (Poelwijk et al., 2011).
RNA viruses are ideal experimental systems for exploring the nature
of epistatic interactions: their compact genomes often code for
overlapping reading frames, contain functional RNA secondary
structures and encode for multi-functional proteins. Altogether, these
properties are expected to lead to strong epistasis. Indeed, recent
studies exploring uni- and multi-dimensional epistasis have provided
empirical evidences that ME is common for RNA viruses such as
Foot-and-mouth disease virus (Elena, 1999), bacteriophage f6 (Burch
and Chao, 2004), Vesicular stomatitis virus (VSV, Sanjua
´net al., 2004),
Human immunodeficiency virus type 1 (Bonhoeffer et al., 2004; Van
Opijnen et al., 2006; Parera et al., 2009; Da Silva et al., 2010; Martı
´nez
et al., 2011), Rous sarcoma virus (Sanjua
´n, 2006), or Tobacco etch virus
(TEV; De la Iglesia and Elena, 2007b), among others, as well as for
ssDNA bacteriophages such as ID11 (Rokyta et al., 2011) or fX174
(Pepin and Wichman, 2007). Furthermore, in most of these studies
positive epistasis is more abundant than negative epistasis, although
variability exists within each virus. Positive epistasis may appear as a
consequence of individual mutations having a large negative impact
on fitness such that any additional mutation that still produces a
viable virus must necessarily exert a minor impact (Elena et al., 2010).
SE, by contrast, has been detected only among compensatory
mutations for fX174 (Poon and Chao, 2006) and among pairs of
beneficial mutations for ID11 (Rokyta et al., 2011). By contrast, no
evidence of SE was found for combinations of beneficial mutations in
the RNA bacteriophage MS2 (Betancourt, 2010).
In this study we sought to characterize the patterns of multi-
dimensional epistasis for the RNA plant virus TEV (genus Potyvirus,
family Potyviridae). TEV has a single-stranded positive-sense RNA
genome of ca. 9.5 Kb that encodes for a single polyprotein that self-
processes into 10 mature peptides. An additional peptide is translated
from an overlapping ORF after þ2 frameshifting. To this end, we
generated a collection of 53 double mutants by randomly combining
20 individual mutations whose deleterious fitness effect had been
previously quantified (Carrasco et al., 2007b). The fitness of all single
and double mutants was evaluated in the primary host Nicotiana
tabacum. We characterized the statistical properties of the distribution
of epistatic interactions and found a mixture of positive and negative
effects (including some examples of synthetic lethals (SLs)). Next, we
found that RSE was the most common type of epistasis. We also
explored the negative association between the average fitness effect of
deleterious mutations and the strength of the epistastic interaction in
which they were involved. Finally, we tried to frame the observed
epistatic effects within a model of the protein–protein interaction
network (PPIN) formed by all 11 TEV proteins.
There are many novelties within our study. First, this is the first
description of extensive SE, particularly of the reciprocal type,
contributing to the architecture of fitness of an RNA virus. Second,
we contextualize epistasis in the network of interactions among viral
proteins. Third, it is the first report of epistasis for a eukaryotic virus
in its natural host rather than in in vitro cell cultures, which represent
an artificial and oversimplified environment. Last, but not least, this is
the first analysis of multidimensional epistasis for any plant pathogen.
MATERIALS AND METHODS
Virus genotypes
A subset of 20 mutants non-lethal in N. tabacum (Supplementary
Table 1) was randomly chosen from a larger collection generated in a
previous study (Carrasco et al., 2007b). Six were synonymous
mutations, whereas the rest were nonsynonymous. Plasmid pMTEV
(Bedoya and Daro
`s, 2010) was used to reconstitute the wild-type TEV
and to generate the mutant genotypes. These 20 mutations were
randomly combined to generate a set of 53 double mutants
(Supplementary Table 2) by site-directed mutagenesis using the
QuikChange II XL Site-Directed Mutagenesis Kit (Stratagene, Santa
Clara, CA, USA) as described by Carrasco et al. (2007a). The kit
incorporates PfuUltra high fidelity DNA polymerase that minimizes
the introduction of undesired mutations. The uniqueness of each
mutation was confirmed by sequencing an 800-bp fragment encom-
passing the mutated nucleotide.
Infectious RNA of each genotype was obtained by in vitro
transcription after BglII linearization of the corresponding plasmid
as described in Carrasco et al.(2007a).
Inoculation experiments
All N. tabacum plants were inoculated at an identical growth stage to
minimize variations in defense response to infection with develop-
mental stage. All inoculations were done in a single experimental
block. Nine plants per TEV genotype were inoculated by rubbing the
first true leaf with 5 mlcontaining5mgRNAin vitro transcript of the
virus and 10% Carborundum (100 mg ml 1).
Ten days post-inoculation, the whole infected plant, except the
inoculated leaf, was collected. The collected tissue was frozen in liquid
nitrogen and grounded with mortar and pestle.
RNA purification and virus quantification
An aliquot of approximately 100 mg of grounded tissue was taken and
mixed with 200 ml of e xtraction buffer (0. 2 MTri s , 0 . 2 MNaCl, 50 mM
EDTA, 2% SDS; pH 8). An equal volume of phenol:chloroform:
isoamylic alcohol (25:25:1) was added, thoroughly vortexed and
centrifuged at 14 000 g for 5 min at 25 1C. Ca. 160 ml of the upper
aqueous phase was mixed with 80 ml of a solution containing 7.5 M
LiCl an d 50 mMEDTA and incubated overnight on ice at 4 1C. The
precipitated RNA was centrifuged at 14000 g for 15 min at 41C,
washed once with 70% ice-cold ethanol, dried in a SpeedVac (Thermo
Fisher Scientific, Waltham, MA, USA) and resuspended in 30 mlof
DEPC-treated ultrapure water. RNA concentration was measured
spectrophotometrically and the samples were diluted to a final
concentration of 50 ng ml1.
Within-plant virus accumulation was measured by absolute
RTqPCR using an external standard (Pfaffl, 2004). Standard curves
were constructed using five serial dilutions of TEV RNA produced by
in vitro transcription and diluted in RNA obtained from the host plant.
RTqPCR reactions were performed in 20 ml volume using One
Step SYBR PrimeScript RTPCR Kit II (TaKaRa, Bio Inc, Otsu,
Japan) following the instructions provided by the manufacturer. The
forward TEVCP 50-TTGGTCTTGATGGCAACGTG and reverse
TEVCP 50-TGTGCCGTTCAGTGTCTTCCT primers amplify a 71
nt fragment within the TEV CP cistron. CP was chosen because it is
located in the 30end of TEV genome and hence would only quantify
complete genomes. Each RNA sample was quantified three times in
independent experiments. Amplifications were done using the ABI
PRISM Sequence Analyzer 7000 (Applied Biosystems, Carlsbad, CA,
USA). The thermal profile was: RT phase consisted of 5min at 42 1C
followed by 10 s at 95 1C; and PCR phase of 40 cycles of 5 s at
95 1C and 31 s at 60 1C. Quantification results were examined using
SDS7000 software v. 1.2.3 (Applied Biosystems).
For each genotype, a Malthusian growth rate per day was
computed as m¼(1/t)log(Qt), where Qtare the pg of TEV RNA
Epistasis in RNA genomes
JLalic
´and SF Elena
72
Heredity
per 100 ng of total plant RNA quantified at t¼10 days post-
inoculation. Absolute fitness was then defined as W¼em(Crow and
Kimura, 1970).
Estimation of epistasis among pairs of mutations
Epistasis among pair of mutations xand y,Exy, was calculated as
Exy ¼W00Wxy Wx0W0y(Kouyos et al., 2007), where W00,Wxy,Wx0,
and W0ycorrespond to the absolute fitness of the wild-type, the
double mutant and each single mutant, respectively. A value of Exy40
corresponds to the case of positive (antagonistic) epistasis, whereas a
value of Exyo0 is indicative of negative (synergistic) epistasis. Values
of Exy not significantly deviating from zero were qualified as multi-
plicative (that is, non-epistatic) mutational effects.
In all cases, reported error intervals correspond to ±1 s.e.m. All
statistical analyses were performed using IBM SPSS v. 19 (Armonk,
NY, USA).
RESULTS
Epistasis among pairs of deleterious mutations
Figure 1 shows the relationship between observed and expected fitness
values for the set of 53 double mutant genotypes synthetized for this
study (Supplementary Table 2). The solid line represents the null
hypothesis of non-epistatic fitness effects. The observed fitness values of
20 double mutant genotypes significantly departed from this null
expectation (Supplementary Table 2; t-tests, in all cases Pp0.049).
Nine of these cases were SLs, which means that two mutations that
were viable by themselves become lethal when combined. These SLs
represent an extreme case of negative epistasis. All other significant
cases corresponded to positive epistasis. Therefore, we found variability
in the sign and strength of epistasis. However, only the nine SLs
remained significant after applying the more stringent sequential
Bonferroni correction for multiple tests of the same hypothesis (Rice,
1989). Nonetheless, for all analyses presented below, we used all 20
significant cases, unless otherwise indicated. This decision represents a
compromise between reducing the data set to only the nine SLs (which
precludes running any additional analysis) and using the whole data set
irrespective of the significance of observed fitness values.
Three double mutants contained two synonymous mutations, 22
combined one synonymous and one nonsynonymous mutation and
28 carried two nonsynonymous mutations. No differences existed,
however, in the magnitude of epistasis among these genotypic classes
(Kruskal–Wallis test: w2¼0.405, 2 df, P¼0.817).
Using the whole data set, we sought to test whether the distribution
of SL and viable mutations were homogeneous among pairs of
mutations within the same cistron or among affecting different
cistrons. In two out of nine SLs both mutations were at the same
cistron (22.2%), whereas in the case of viable double mutant
genotypes, only one genotype out of 44 had both mutations in
the same cistron (2.3%), a significant difference (w2¼5.569, 1 df,
P¼0.018) despite the small sample size. Furthermore, the average
epistasis coefficient computed for mutations within the same cistron
was 1.142±0.617, whereas it was reduced to 0.171±0.090 for
pairs of mutations affecting different cistrons. This 85.1% relaxation
in the strength of epistasis was also significant (t51 ¼2.477, P¼0.017).
Therefore, we can conclude that a tendency exists for mutations
affecting the same cistron to generate a SL phenotype and to interact
in a stronger and more negative manner, whereas mutations affecting
different viral proteins presented weaker interactions.
Statistical properties of the epistasis distribution
Figure 2 illustrates the distribution of epistasis parameters for all pairs
of point mutations analyzed. The distribution had a bimodal shape,
with SLs representing the left probability mass and the viable
genotypes being on the right side of the distribution. The average
epistasis was
E¼0:226 0:095, a value that departs from the null
hypothesis of multiplicative effects (t52 ¼2.376, P¼0.021). Further-
more, the distribution had a significant negative skewness
(g1¼1.806±0.327; t52 ¼5.515, Po0.001), that is, the tail contain-
ing negative epistasis is heavier than the Gaussian and thus asym-
metric. Similarly, the distribution was significantly leptokurtic
(g2¼1.326±0.644, t52 ¼2.058, P¼0.045), indicating that it had a
more acute peak around the mean value compared with the Gaussian.
Given that lethal mutations are largely irrelevant for evolutionary
dynamics, we sought to reanalyze the epistasis distribution
after removing SLs. The main consequence of this removal was
that the average epistasis then becomes significantly positive
(
E¼0:084 0:005; t43 ¼17.438, Po0.001). Regarding the shape of
the distribution, it still remained asymmetric with significant negative
Figure 1 Relationship between observed and expected multiplicative fitness
for 53 TEV genotypes carrying pairs of nucleotide substitutions. The solid
line represents the null hypothesis of multiplicative fitness effects.
Deviations from this line arise as a consequence of the existence of
epistatic fitness effects.
Figure 2 Distribution of epistasis. Epistasis, E, was computed as the
difference between the observed fitness of the double mutant (W00Wxy)and
the value expected from subtracting the effects of each single mutant from
the wild-type value (Wx0W0y).
Epistasis in RNA genomes
JLalic
´and SF Elena
73
Heredity
skewness (g1¼1.050±0.358; t43 ¼2.936, P¼0.005), although the
skewness parameter was 41.9% smaller than when SLs were included
in the data set. In contrast, the distribution became 77.1% more
leptokurtic (g2¼2.348±0.702, t43 ¼3.346, P¼0.002), as a conse-
quence of the removing the cases from the left tail extreme.
Pervasive RSE
We were interested in evaluating the extent to which SE was present in
our data set. Poelwijk et al. (2011) defined mathematically the
condition for SE as
jWx0W00 þWxy W0yjojWx0W00jþjWxy W0yj
Twelve out of the twenty TEV double-mutant genotypes for which we
had detected significant epistasis (Supplementary Table 2) fulfilled
this condition and thus can be classified as cases of SE. The other
eight, hence, correspond to cases of ME. Is this 3:2 proportion
expected given the observed fitness values of individual mutations
and of the double mutants? To tackle this question we applied the
above inequality to the 33 non-epistatic pairs of mutations, founding
that 26 fulfilled it, despite not being significant. A Fisher’s exact test
failed to detect significant differences among epistatic and non-
epistatic pairs fulfilling the inequality (1-tailed P¼0.124), thus
confirming that the observed proportion of ME and of SE was not
significantly enriched in the later class. Therefore, we conclude that SE
makes a major contribution (60%) to all cases of significant epistasis.
Next, we specifically evaluated the contribution of RSE to the
observed pattern of SE. According to Poelwijk et al.(2011)the
following additional condition must be met by a pair of mutations
showing SE in order to be considered as cases of RSE:
jW0yW00 þWxy Wx0jojW0yW00 jWxy Wx0j:
Herewith, this condition was fulfilled by 11 out of 12 cases of SE
(91.7%). Indeed, only synthetic lethal PC6/PC76 did not match it. As
before, given the fitness of single and double mutants, we tested
whether this extremely high prevalence of RSE among cases of SE is to
be expected. We counted the number of cases that fulfilled this second
inequality (25) among the 26 non-epistatic cases that matched the
first one. A Fisher’s exact test also showed no significant enrichment
in cases of RSE among cases of SE (1-tailed P¼0.538). Therefore, we
conclude that RSE is common in TEV genome.
Correlation between fitness effects and epistasis
It has been observed that average deleterious mutational effects and
the strength of positive epistasis are not independent parameters but,
instead, are negatively correlated (Wilke and Adami, 2001 You and
Yin, 2002; Wilke et al., 2003; Bershtein et al., 2006; Sanjua
´net al.,
2006; De la Iglesia and Elena, 2007). We sought to investigate if this
negative relationship holds for TEV. Figure 3 shows the relationship
between the mean fitness of the two mutations combined and the
estimated epistasis for all 53 double mutants. A first observation is
that two different and significant relationships exist in correspondence
to different phenotypic classes: one for the nine SLs (Spearmans
rS¼1.000, 7 df, Po0.001) and another one for the viable genotypes
(rS¼0.416, 42 df, P¼0.005). However, overall a significant
negative correlation existed after controlling for the difference within
two phenotypic classes (partial r¼0.331, 50 df, P¼0.017). The
slope for the viable genotypes was significantly smaller than the slope
for the SLs (analysis of covariance test for the homogeneity of slopes
in Figure 3: F1,49 ¼9.212, P¼0.004), suggesting that the underlying
mechanisms for the observed relationships were different for each
phenotypic class. Indeed, the correlation observed for the SLs is trivial
because it is expected based on the definition of epistasis used here.
If the observed fitness of the double mutant is Wxy ¼0, then
Exy ¼Wx0W0y¼W
2,whereW
is the geometric mean fitness
of mutations xand y. The validity of this explanation was confirmed
by the fact that linear regression throughout the origin of epistasis on
W
2for the SLs data rendered the expected slope of 1.000±0.000.
These correlations suggest that mutational effects and epistasis are
not independent traits, but instead, they may evolve hand in hand.
Stronger mutational effects are associated with more positive inter-
actions, whereas milder effects are associated with more relaxed
positive interactions. Therefore, a reduction in the magnitude of
mutational effects translates into a relaxation of the positive epistasis.
Epistasis in the context of TEV PPIN
Mutations were grouped according to the mature protein they affect.
By doing so, we focused on the analyses of interaction among proteins
rather than among individual nucleotide residues. Rodrigo et al.
(2011) inferred the undirected PPIN shown in Figure 4 using a
compendium of physical interactions experimentally determined by
the yeast two-hybrid method. We were interested in correlating the
network properties with the characteristics of the distribution of
epistasis inferred in this study.
First, we sought to test whether the number of significant epistatic
and non-epistatic interactions was homogeneously distributed among
pairs of proteins directly linked in the PPIN graph or unlinked
(Figure 4). A Fisher’s exact test failed to reveal a significant association
(P¼0.151), thus rejecting the hypothesis that a direct interaction
between two proteins is a necessary condition to generate significant
epistasis.
Second, we explored whether the number of pairs of proteins
involved in positive and negative epistatic interactions was evenly
distributed among pairs directly connected in the PPIN and those that
are not (Figure 4). It has been argued for modularly organized PPINs
that mutations affecting independent modules would show a pattern
of positive epistasis, although PPINs organized as a single-functional
module would be more sensitive to the effect of mutations and show a
pattern of negative epistasis (Sanjua
´n and Elena, 2006; Sanjua
´nand
Nebot, 2008; Macı
´aet al., 2012). In agreement to this expectation,
Figure 3 Association between average mutational effects and the magnitude
of epistasis. Two apparent relationships exist: one for pairs of mutations
generating viable genotypes (upper cloud) and a different one associated to
the SLs (lower cloud). The regression lines are included to illustrate the
difference in the underlying relationship between epistasis and average
mutational effects between both types of phenotypes.
Epistasis in RNA genomes
JLalic
´and SF Elena
74
Heredity
we found that nine out of fourteen (64.3%) positive interactions
between linked elements, whereas only two out of six (33.3%)
interactions between unlinked elements were positive. Thus, we
conclude that mutations affecting connected elements in the PPIN
tend to be involved in more positive epistatic interactions than those
affecting non-connected components.
Finally, we hypothesized that highly linked nodes would have a
stronger tendency to be epistatic, whereas less connected nodes will be
less so. To test this hypothesis, we first computed the tendency of a
protein to be involved in significant epistasis interactions (that is,
epistasiness) for each protein as the absolute value of the average
epistasis coefficient computed across all interactions in which this
protein was involved and using the whole data set. Absolute values
were used because we tested for the tendency for involvement in
significant interactions regardless of their sign. Then, we computed
the connectivity of each node as the number of links it has in
Figure 4. A non-parametric correlation coefficient failed to detect a
significant association between these two variables (rS¼0.221, 6 df,
P¼0.599). Therefore, we conclude that the tendency of a protein to
be involved in epistatic interactions is not a necessary consequence of
the amount of interactions itself.
DISCUSSION
In this study, the distribution of epistatic interactions on fitness for a
plant RNA virus has been evaluated by constructing genotypes
carrying pairs of single-nucleotide substitutions, each having a
deleterious fitness effect. We detected cases of both positive and
negative epistasis, although positive epistases were significantly more
abundant than negative ones, such that the combined effect of
mutations is significantly less harmful than expected under the null
multiplicative model. This dominance of positive epistasis is particu-
larly frequent among mutations affecting two different proteins,
whereas, on average, mutations affecting the same protein interact
in a negative manner. These findings are in good agreement with
observations accumulated in recent years for other RNA viruses,
including retroviruses, and small ssDNA viruses (reviewed by Elena
et al., 2010), both using experimental approaches to characterize uni-
and multi-dimensional epistasis. Given this heterogeneity in viral
systems, it thus seems highly likely that positive epistasis among
deleterious mutations is a general feature of most small viruses. What
may be the mechanistic reason for this excess of positive epistasis?
Several reasons can be brought forward. First, the compactness of
viral genomes, many of which even had adopted the strategy of
overlapping genes and multifunctional proteins, necessarily implies
that the deleterious effects of different mutations can partially overlap,
hence producing positive epistasis. Indeed, this expectation is well
fulfilled by our finding that interactions are, on average, more positive
when mutations occur in two different proteins than when they both
occur in the same one. Second, positive epistasis can also be a
consequence of the existence of elements of RNA secondary structure.
Indeed, it was shown by computer simulations of RNA folding that
mutations affecting the same structural element may restore it and
thus generate positive epistasis (Wilke et al., 2003 Sanjua
´net al.,
2006). Another observation that seems to be common among
experiments of multi-dimensional epistasis in RNA viruses is the
existence of frequent cases of synthetic lethality, for example, for the
mammalian rhabdovirus VSV (Sanjua
´net al., 2004).
The dominance of positive epistasis among deleterious mutations
and the existence of frequent cases of synthetic lethality are both
fingerprints of another phenomenon: the low genetic robustness of
viral genomes. It has been postulated that epistasis and robustness are
two sides of the same coin and that negative epistasis must be a
hallmark for genetic robustness (De Visser et al., 2003, 2011; Proulx
and Phillips, 2005; Desai et al., 2007). Indeed, the observed negative
correlation between epistasis and mutational effects shown in Figure 3
provides additional support for this hypothesis and is consistent with
observations made in systems as diverse as artificial life (Wilke and
Adami, 2001; Edlund and Adami, 2004), computer simulations of
genetic systems (You and Yin, 2002; Macı
´aet al., 2012), RNA (Wilke
et al., 2003; Sanjua
´net al., 2006) and protein folding (Bershtein et al.,
2006), and in a mutation–accumulation experiment done with TEV
NIa
Pro
6K1
P3
VPg
6K2
HC
Pro
NIb
CP
PIPO
CI
P1
Figure 4 TEV PPIN inferred from yeast two-hybrid data published
elsewhere. The 11 mature peptides are indicated as nodes. Black edges
correspond to interactions for which we did not detect significant epistasis.
Red edges correspond to cases of negative epistasis and green edges
correspond to cases of positive epistasis. Double green lines correspond to
two pairs of mutations affecting the same proteins. The dashed blue line
corresponds to a case in which a first pair of mutations showed positive
epistasis (PC19/PC95) but a second pair had negative epistasis (PC22/
PC95). The PPIN was drawn using Cytoscape (Killcoyne et al., 2009).
Epistasis in RNA genomes
JLalic
´and SF Elena
75
Heredity
(De la Iglesia and Elena, 2007). The negative correlation between
epistasis and mutational effects means that the milder the
average mutational effect is, the more negative the epistatic interac-
tions between mutations will be. This results in a genotype
that is more mutationally robust against genetic perturbations. In
contrast, positive epistasis reflects strong mutational effects
and, therefore, low genetic robustness. Sanjua
´n and Elena (2006)
postulated that robustness would scale up with genetic complexity
and that it may result from the fact that more complex genetic
systems may contain more redundant structures capable of buffering
the effect of mutations. Very recently, Macı
´aet al. (2012) tested this
hypothesis by simulating the evolution of genetic circuits under
variable selection for robustness. They found that, as predicted,
negative epistasis was caused by the existence of genetic redundancy
in complex networks and not due to complexity itself, as
the correlation disappeared when the formation of redundant
structures was not allowed during the evolution of complex networks.
In this sense, RNA viruses will occupy the lower side of the
complexity spectrum and, therefore, would be highly sensitive (that
is, non-robust) to mutations.
Within cases of significant epistatic interactions, we found a large
contribution of SE relative to the contribution of ME. This represents
the first description of SE for an RNA virus, as previous studies of
multi-dimensional epistasis in RNA viruses did not explicitly look for
SE (for example, Bonhoeffer et al., 2004; Sanjua
´net al., 2004; Sanjua
´n,
2006; Van Opijnen et al., 2006) or simply failed to find them
(Betancourt, 2010). In contrast, SE has been shown to be common
during adaptation of b-lactamase to cefotaxime (Weinreich et al.,
2006; Salverda et al., 2011), in evolution experiments compensating
for the cost of antibiotic resistance in bacteria (Schrag et al., 1997;
Maisnier-Patin et al., 2002) and viruses (Molla et al., 1996; Cong et al.,
2007; Martı
´nez-Picado and Martı
´nez, 2009), in experimental evolu-
tion of asexual Saccharomyces cerevisiae (Kvitek and Sherlock, 2011),
and in multi-dimensional tests of epistasis in Aspergillus niger (Franke
et al., 2011). All but one cases of SE detected in TEV corresponded to
RSE, perhaps making this observation even more interesting. This
type of epistasis is particularly relevant from the perspective of
describing fitness landscapes. Poelwijk et al. (2011) have shown that
the existence of multiple adaptive peaks in a fitness landscape, that is,
ruggedness, requires RSE. Furthermore, Kwitek and Sherlock (2011)
experimentally confirmed that RSE caused the ruggedness of a fitness
landscape. The ruggedness of adaptive landscapes is critical to predict
whether evolving populations may reach the global optima or may get
stuck into suboptimal peaks (Weinreich, 2005; Withlock et al., 1995).
Our finding of a predominance of RSE suggests that the fitness
landscape for TEV, and maybe for other RNA viruses, must be highly
rugged.
In conclusion, the results reported here, together with previous
findings, contribute to the perspective that viral genomes are
dominated by positive epistasis, which may result from their
compactness and lack of genetic redundancy. In addition, we provide
the first direct proof that SE, in particular RSE, contributes in a large
extent to the architecture of viral fitness. The high frequency of RSE
suggests that adaptive landscapes for RNA viruses maybe highly
rugged. This ruggedness may impose harsh constraints on the
often-invoked but not empirically grounded limitless adaptability
of RNA viruses.
DATA ARCHIVING
Data have been archived at Dryad: doi: 10.5061/dryad.bq4pp7f9.
CONFLICT OF INTEREST
The authors declare no conflict of interest.
ACKNOWLEDGEMENTS
We thank Francisca de la Iglesia and A
`ngels Pro
`sper for their excellent technical
assistance, Ste
´phanie Bedhomme and Mark P Zwart for the discussion and
Mario A Fares for statistical advice. Jose
´ADaro
`s generously gifted us the
pMTEV plasmid. This research was supported by the Spanish Ministry of
Science and Innovation grant BFU2009-06993 to SFE. JL was supported by the
JAE program from CSIC.
Bagheri HC, Wagner GP (2004). Evolution of dominance in metabolic pathways. Genetics
168: 1716–1735.
Bedoya LC, Daro
`s JA (2010). Stability of Tobacco etch virus infectious clones in plasmid
vectors. Virus Res 149: 234–240.
Bershtein S, Segal M, Bekerman R, Tokuriki N, Tawfik DS (2006). Robustness-
epistasis link shapes the fitness landscape of a randomly drifting protein. Nature
444: 929–932.
Betancourt AJ (2010). Lack of evidence for sign epistasis between beneficial mutations in
an RNA bacteriophage. JMolEvol71: 437–443 .
Bonhoeffer S, Chappey C, Parkin NT, Whitcomb JM, Petropoulos CJ (2004). Evidence for
positive epistasis in HIV-1. Science 306: 1547–1550.
Burch CL, Chao L (2004). Epistasis and its relationship to canalization in the RNA virus
f6. Genetics 167: 559–567.
Carrasco P, Daro
`s JA, Agudelo-Romero P, Elena SF (2007a). A real-time RT-PCR assay for
quantifying the fitness of Tobacco etch virus in competition experiments . J Virol Meth
139: 181–188.
Carrasco P, de la Iglesia F, Elena SF (2007b). Distribution of fitness and virulence effects
caused by single-nucleotide substitutions in Tobacco etch virus.JVirol81:
12979–12984.
Cong M, Heneine W, Garcı
´a-Lerma JG (2007). The fitness cost of mutations associated
with Human immunodeficien cy virus type 1 drug resistance is modulated by mutational
interactions. JVirol81: 3037–3041.
Coyne JA (1992). Genetics and speciation. Nature 355:511515.
Crow JF, Kimura M (1970). An Introduction to Population Genetics Theory.Harperand
Row: New York.
Da Silva J, Coetzer M, Nedellec R, Pastore C, Mosier DE (2010). Fitness epistasis and
constraints on adaptation in a Human immunodeficiency virus type 1 protein region.
Genetics 185: 293–303.
Desai MM, Weissman D, Feldman MW (2007). Evolution can favor antagonistic epistasis.
Genetics 177: 1001–1010.
De la Iglesia F, Elena SF (2007). Fitness declines in Tobacco etch virus upon serial
bottleneck transfers. JVirol81: 4941–49 47.
De Visser JAGM, Elena SF (2007). The evolution of sex: empirical insights into the roles of
epistasis and drift. Nat Rev Genet 8: 139–149.
De Visser JAGM, Hermis son J, Wagner GP, Ancel-Meyers L, Bagheri-Chaich ian H,
Blanchard JL et al. (2003). Perspective: Evolution and detection of genetic robustness.
Evolution 57: 1959–1972.
De Visser JAGM, Cooper TF, Elena SF (2011). The causes of epistasis. Proc R Soc B 10:
3617–3624.
Edlund JA, Adami C (2004). Evolution of robustness in digital organisms. Artif Life 10:
167–179.
Elena SF (1999). Little evidence for synergism among deleterious mutations in a
nonsegmented RNA virus. J Mol Evol 49:703707.
Elena SF, Sole
´RV, Sardanye
´s J (2010). Simple genomes, complex interactions: epistasis
in RNA virus. Chaos 20: 026106.
Franke J, Klo
¨zer A, de Visser JAGM, Krug J (2011). Evolutionary accessibility of
mutational pathways. PLoS Comp Biol 7: e1002134.
Killcoyne S, Carter GW, Smith J, Boyle J (2009). Cytoscape: a community-based
framework for network modeling. Meth Mol Biol 563: 219–239.
Kondrashov AS (1994). Muller’s ratchet under epistatic selection. Genetics 136:
1469–1473.
Kondrashov AS, Crow JF (1991). Haploidy or diploidy: which is better. Nature 351:
314–315.
Kondrashov FA, Kondrashov AS (2001). Multidimensional epistasis and the disadvantage
of sex. Proc Natl Acad Sci USA 98: 12089–12092.
Kouyos RD, Silander OK, Bonhoeffer S (2007). Epistasis between deleterious mutations
and the evolution of recombination. Trends Ecol Evol 6: 308–315.
Kvitek DJ, Sherlock G (2011). Reciprocal sign epistasis between frequently experimentally
evolved adaptive mutations causes a rugged fitness landscape. PLoS Genet 7:
e1002056.
Macı
´aJ,Sole
´RV, Elena SF (2012). The causes of epistasis in genetic networks. Evolution
66:586596.
Maisnier-Patin S, Berg OG, Lijas L, Andersson DI (2002). Compensatory adaptation to the
deleterious effect of antibiotic resistance in Salmonella typhimurium.Mol Microbiol
46:355366.
Epistasis in RNA genomes
JLalic
´and SF Elena
76
Heredity
Martı
´nez JP, Bocharov G, Ignatovich A, Reiter J, Dittmar MT, Wain-Hobson S et al. (2011).
Fitness ranking of individual mutants drives patterns of epistatic interactions in HIV-1.
PLoS ONE 6:e18375.
Martı
´nez-Picado J, Martı
´nez MA (2009). HIV-1 reverse transcriptase inhibitor
resistance mutations and fitness: a view from the clinic and ex vivo.Virus Res 134:
104–123.
Molla A, Korneyeve M, Gao Q, Vasavanonda S, Schipper PJ, Mo HM et al. (1996). Ordered
accumulation of mutations in HIV protease confers resistance to ritonavir. Nat Med 2:
760–766.
Parera M, Pe
´rez-A
´lvarez N, Clotet B, Martı
´nez MA (2009). Epistasis among deleterious
mutations in the HIV-1 protease. J Mol Biol 392:243250.
Pepin KM, Wichman HA (2007). Variable epistatic effects between mutations at host
recognition sites in fX174. Evolution 67: 1710–1724.
Pfaffl MV (2004). Quantification strategies in real-time PCR. In Bustin SA
(ed A-Z of Quantitative PCR, International University Line.LaJolla:USA,
pp 87–112.
Phillips PC (2008). Epistasis – the essential role of gene interactions in the structure and
evolution of genetic systems. Nat Rev Genet 9: 855–867.
Poelwijk FJ, Kiviet DJ, Weinreich DM, Tans SJ (2007). Empirical fitness landscapes reveal
accessible evolutionary paths. Nature 445: 383–386.
Poelwijk FJ, Tanase-Nicola S, Kiviet DJ, Tans SJ (2011). Reciprocal sign epistasis
is a necessary condition for multi-peaked fitness landscapes. J Theor Biol 272:
141–144.
Poon AFY, Chao L (2006). Functional origins of fitness effect-sizes of compensatory
mutations in the DNA bacteriophage fX174. Evolution 60: 2032–2043.
Proulx SR, Phillips PC (2005). The opportunity for canalization and the evolution of
genetic networks. Am Nat 165: 147–162.
Remold SK, Lenski RE (2004). Pervasive joint influence of epistasis and plasticity on
mutational effects in Escherichia coli. Nat Genet 36: 423 426.
Rice WR (1989). Analyzing tables of statistical tests. Evolution 43: 223–225.
Rodrigo G, Carrera J , Ruiz-Ferrer V, Del Toro FJ, Llave C, Voinnet O et al. (2011).
Characterization of the Arabidopsis thaliana interactome targeted b y viruses. Santa Fe
Institute Working Paper 11-10-049.
Rokyta DR, Joyce P, Caudle B, Miller C, Beisel CJ, Wichman HA (2011). Epistasis between
beneficial mutations and the phenotype-to-fitness map for a ssDNA virus. PLoS Genet
7: e1002075.
Salverda MLM, Dellus E, Gorter FA, Debets AJM, Van der Oost J, Hoekstra RF et al.
(2011). Initial mutations direct alternative pathways of protein evolution. PLoS Genet
7: e1001321.
Sanjua
´n R (2006). Quantifying antagonistic epistasis in a multifunctional RNA secondary
structure of the Rous sarcoma virus.JGenVirol87: 1595–1602.
Sanjua
´n R, Elena SF (2006). Epistasis correlates to genomic complexity. Proc Natl Acad
Sci USA 103: 14402–14405.
Sanjua
´n R, Forment J , Elena SF (20 06). In silico predicted robustne ss of viroids R NA
secondary structure. II. Interaction between mutation pairs. MolBiolEvol23:
2123–2130.
Sanjua
´n R, Moya A, Elena SF (2004). The contribution of epistasis to the architecture of
fitness in an RNA virus. Proc Natl Acad Sci USA 101: 15376–15379.
Sanjua
´n R, Nebot MR (2008). A network model for the correlation between epistasis and
genomic complexity. PLoS ONE 3:e2663.
Schrag SJ, Perrot V, Levin BR (1997). Adaptation to the fitness cost of antibiotic
resistance in E. coli.Proc R Soc B 264: 1287–1291.
Van Opijnen T, Boerlijst MC, Berkhout B (2006). Effects of random mutations in the
Human immunodeficiency virus type 1 transcriptional promoter on viral fitness in
different host cell environments. JVirol80: 6678– 6685.
Weinreich DM (2005). The rank ordering of genotypic fitness values predicts genetic
constraints on natural selection on landscapes lacking sign epistasis. Genetics 171:
1397–1405.
Weinreich DM, Delaney NF, DePristo MA, Hartl DL (2006). Darwinian evolution can follow
only very few mutational paths to fitter proteins. Science 312: 111–114.
Weinreich DM, Watson RA, Chao L (2005). Perspective: sign epistasis and genetic
constraint on evolutionary trajectories. Evolution 59: 1165–11 74.
Welch JJ, Waxman D (2005). The nk model and population genetics. JTheorBiol234:
329–340.
Wilke CO, Adami C (2001). Interaction between directional epistasis and average
mutational effects. Proc R Soc B 298: 1469–1474.
Wilke CO, Lenski RE, Adami C (2003). Compensatory mutations cause excess of
antagonistic epistasis in RNA secondary structure folding. BMC Evol Biol 3:114.
Withlock MC, Phillips PC, Moore FBG, Tonsor SJ (1995). Multiple fitness peaks and
epistasis. AnnuRevEcolEvolSyst26: 601–6 29.
You L, Yin J (2002). Dependence of epistasis on environment and mutation severity as
revealed by in silico mutagene sis of phage T7. Genetics 160: 1273–1281.
Supplementary Information accompanies the paper on Heredity website (http://www.nature.com/hdy)
Epistasis in RNA genomes
JLalic
´and SF Elena
77
Heredity
... (±1 SD). This value was significantly positive (one-sample t-test: t 24 = 2.760, P = 0.007), confirming previous observations that epistasis in TEV genome is predominantly positive (Lalić and Elena, 2012a). ...
... As already mentioned in the Introduction, a necessary condition for ruggedness in adaptive landscapes is the existence of reciprocal sign epistasis (Poelwijk et al. 2011). In our previous study on epistasis among pairs of random deleterious mutations in TEV genome, we found a significant enrichment for this type of interaction (Lalić and Elena, 2012a). The question now is whether this conclusion also holds for the reduced set of potentially beneficial mutations being studied here. ...
... Epistasis was prevalent (61.54%) in our dataset and predominately positive (68.75%), meaning that the absolute effect of the second mutation is smaller than that of the first. Previously, we have found significant prevalence of positive epistatic effects among deleterious mutations in TEV measured in its primary host N. tabacum (Lalić and Elena 2012a) as well as across a panel of alternative susceptible hosts (Lalić and Elena 2012b). Thus, positive epistasis is common in TEV genome among both deleterious and beneficial mutations and for both primary and alternative hosts. ...
Preprint
RNA viruses are the main source of emerging infectious diseases owed to the evolutionary potential bestowed by their fast replication, large population sizes and high mutation and recombination rates. However, an equally important parameter, which is usually neglected, is the topography of the fitness landscape, that is, how many fitness maxima exist and how well connected they are, which determines the number of accessible evolutionary pathways. To address this question, we have reconstructed the fitness landscape describing the adaptation of Tobacco etch potyvirus to its new host, Arabidopsis thaliana. Fitness was measured for most of the genotypes in the landscape, showing the existence of peaks and holes. We found prevailing epistatic effects between mutations, with cases of reciprocal sign epistasis being common at latter stages. Therefore, results suggest that the landscape was rugged and holey, with several local fitness peaks and a very limited number of potential neutral paths. The viral genotype fixed at the end of the evolutionary process was not on the global fitness optima but stuck into a suboptimal peak.
... Conversely, when some alleles are only beneficial or harmful in specific genetic backgrounds, the chronological order of genetic changes may result in a diverse population both genetically and phenotypically. Viral genomes appear to be dominated by sign epistasis (Molla et al., 1996;Cong et al., 2007;Martinez-Picado and Martínez, 2008;Lalić and Elena, 2012), indicating the significant role sign epistasis plays in viral genome evolution. ...
Article
Full-text available
Mapping genetic variations to phenotypic variations poses a significant challenge, as mutations often combine unexpectedly, diverging from assumed additive effects even in the same environment. These interactions are known as epistasis or genetic interactions. Sign epistasis, as a specific type of epistasis, involves a complete reversal of mutation effects within altered genetic backgrounds, presenting a substantial hurdle to phenotype prediction. Despite its importance, there is a limited systematic overview of the mechanistic causes of sign epistasis. This review explores the mechanistic causes, highlighting its occurrence in signalling cascades, peaked fitness landscapes, and physical interactions. Moving beyond theoretical discussions, we delve into the practical applications of sign epistasis in agriculture, evolution, and antibiotic resistance. In conclusion, this review aims to enhance the comprehension of sign epistasis and molecular dynamics, anticipating future endeavours in systematic biology engineering that leverage the knowledge of sign epistasis.
... Normally, Dobzansky-Muller incompatibilities provide a great explanation of how reciprocal sign interactions can cause hybrid sterility [48]. Even though reciprocal sign epistasis is well acknowledged in fitness landscape studies [49], we are unable to state whether this type of genetic interaction is a rare event or not, since only a few studies tested it [50][51][52][53]. In E. coli, it was found that 40% of mutations were sign epistasis, but only one was reciprocal sign epistasis (0.02%) [50], whereas in mammals, reciprocal sign epistasis might comprise even 2.9% [53]. ...
Article
Full-text available
Although it is well known that epistasis plays an important role in many evolutionary processes (e.g., speciation, evolution of sex), our knowledge on the frequency and prevalent sign of epistatic interactions is mainly limited to unicellular organisms or cell cultures of multicellular organisms. This is even more pronounced in regard to how the environment can influence genetic interactions. To broaden our knowledge in that respect we studied gene–gene interactions in a whole multicellular organism, Caenorhabditis elegans. We screened over one thousand gene interactions, each one in standard laboratory conditions, and under three different stressors: heat shock, oxidative stress, and genotoxic stress. Depending on the condition, between 7% and 22% of gene pairs showed significant genetic interactions and an overall sign of epistasis changed depending on the condition. Sign epistasis was quite common, but reciprocal sign epistasis was extremally rare. One interaction was common to all conditions, whereas 78% of interactions were specific to only one environment. Although epistatic interactions are quite common, their impact on evolutionary processes will strongly depend on environmental factors.
... However, some examples of positive epistasis among beneficial mutations have also been observed (Chou et al., 2009;Fumasoni and Murray, 2020;Hsieh et al., 2020;Khan et al., 2011;Levin-Reisman et al., 2019). Surveys of interactions between deleterious mutations have produced numerous examples of both positive and negative epistasis (Costanzo et al., 2016;Elena and Lenski, 1997;Hall and MacLean, 2011;Jasnos and Korona, 2007;Lalić and Elena, 2012;Sanjuán et al., 2004;Van Leuven et al., 2021). ...
Article
Full-text available
As an adapting population traverses the fitness landscape, its local neighborhood (i.e., the collection of fitness effects of single-step mutations) can change shape because of interactions with mutations acquired during evolution. These changes to the distribution of fitness effects can affect both the rate of adaptation and the accumulation of deleterious mutations. However, while numerous models of fitness landscapes have been proposed in the literature, empirical data on how this distribution changes during evolution remains limited. In this study, we directly measure how the fitness landscape neighborhood changes during laboratory adaptation. Using a barcode-based mutagenesis system, we measure the fitness effects of 91 specific gene disruption mutations in genetic backgrounds spanning 8,000-10,000 generations of evolution in two constant environments. We find that the mean of the distribution of fitness effects decreases in one environment, indicating a reduction in mutational robustness, but does not change in the other. We show that these distribution-level patterns result from differences in the relative frequency of certain patterns of epistasis at the level of individual mutations, including fitness-correlated and idiosyncratic epistasis.
... Previous studies have shown that bottleneck events are generally disadvantageous [41][42][43][44][45][46] and reduce genetic variation and fitness in the virus population, a phenomenon known as Muller's ratchet [47][48][49][50][51][52][53]. The possible positive effect of bottlenecks is the removal of defective interfering viruses and the selection of evolutionary trajectories in rugged fitness landscapes [54][55][56][57][58]. In this study, the obligated effect of cell-to-cell movement on mutation repair provides new insight into Muller's ratchet. ...
Article
Full-text available
Some debilitating mutations in RNA viruses are repairable; however, the triggering factors of mutation repair remain largely unknown. In this study, multiple triggering factors of mutation repair are identified based on genetic damage to the TLS in CMV. TLS mutations in different RNAs distinctively impact viral pathogenicity and present different types of mutation repair. RNA2 relative reduction level or RNA3 sequence change resulting from TLS mutation is correlated with a high rate of mutation repair, and the TLS mutation of RNA1 fails to be repaired at the high inoculum dose. However, the TLS mutation of RNA1 can be repaired at a low dose of inoculation, particularly around the dilution end-point or in the mixed inoculation with RNA2 having a pre-termination mutation of the 2b gene, an RNAi suppressor. Taken together, TLS mutations resulting in quality or quantity defects of the viral genome or TLS mutations at low doses around the dilution end-point are likely to be repaired. Different levels of TLS mutation repair necessarily require cell-to-cell movement, therefore implying its obligated effect on the evolution of low-fitness viruses and providing a new insight into Muller’s ratchet. This study provides important information on virus evolution and the application of mild viral vaccines.
Article
Full-text available
Viruses are highly evolvable biological entities capable of wreaking havoc on our society. Therefore, a better understanding of virus evolution is important for two main reasons: (i) it will lead to better management of current diseases and prevention of future ones, and (ii) it will contribute to a better understanding of evolutionary processes and their dynamics. In order to understand the evolution of viruses as a whole, it is necessary to consider different elements that shape virus evolution. In this review, we give a general overview of the most relevant factors that determine the evolution of plant viruses. We will focus on mutation rates, epistasis, robustness, recombination, genome organization, virus-host interactions, transmission, community interactions and abiotic factors. Since this review gives a summarized overview of the most important factors in virus evolution it can be a useful starting material for anyone interested in approaching (plant) virus evolution.
Article
Full-text available
Most empirical studies of linkage disequilibrium (LD) study its magnitude, ignoring its sign. Here, we examine patterns of signed LD in two population genomic datasets, one from Capsella grandiflora and one from Drosophila melanogaster. We consider how processes such as drift, admixture, Hill-Robertson interference, and epistasis may contribute to these patterns. We report that most types of mutations exhibit positive LD, particularly, if they are predicted to be less deleterious. We show with simulations that this pattern arises easily in a model of admixture or distance biased mating, and that genome-wide differences across site types are generally expected due to differences in the strength of purifying selection even in the absence of epistasis. We further explore how signed LD decays on a finer scale, showing that loss of function mutations exhibit particularly positive LD across short distances, a pattern consistent with intragenic antagonistic epistasis. Controlling for genomic distance, signed LD in C. grandiflora decays faster within genes, compared to between genes, likely a by-product of frequent recombination in gene promoters known to occur in plant genomes. Finally, we use information from published biological networks to explore whether there is evidence for negative synergistic epistasis between interacting radical missense mutations. In D. melanogaster networks, we find a modest but significant enrichment of negative LD, consistent with the possibility of intra-network negative synergistic epistasis.
Article
Understanding how genotypes map onto phenotypes, fitness, and eventually organisms is arguably the next major missing piece in a fully predictive theory of evolution. We refer to this generally as the problem of the genotype-phenotype map. Though we are still far from achieving a complete picture of these relationships, our current understanding of simpler questions, such as the structure induced in the space of genotypes by sequences mapped to molecular structures, has revealed important facts that deeply affect the dynamical description of evolutionary processes. Empirical evidence supporting the fundamental relevance of features such as phenotypic bias is mounting as well, while the synthesis of conceptual and experimental progress leads to questioning current assumptions on the nature of evolutionary dynamics—cancer progression models or synthetic biology approaches being notable examples. This work delves into a critical and constructive attitude in our current knowledge of how genotypes map onto molecular phenotypes and organismal functions, and discusses theoretical and empirical avenues to broaden and improve this comprehension. As a final goal, this community should aim at deriving an updated picture of evolutionary processes soundly relying on the structural properties of genotype spaces, as revealed by modern techniques of molecular and functional analysis.
Article
Full-text available
Environmental conditions can affect viral accumulation, virulence and adaptation, which have implications in the disease outcomes and efficiency of control measures. Concurrently, mixed viral infections are relevant in plants, being their epidemiology shaped by within-host virus–virus interactions. However, the extent in which the combined effect of variations in abiotic components of the plant ecological niche and the prevalence of mixed infections affect the evolutionary dynamics of viral populations is not well understood. Here, we explore the interplay between ecological and evolutionary factors during viral infections and show that isolates of two strains of Pepino mosaic potexvirus coexisted in tomato plants in a temperature-dependent continuum between neutral and antagonistic interactions. After a long-term infection, the mutational analysis of the evolved viral genomes revealed strain-specific single-nucleotide polymorphisms that were modulated by the interaction between the type of infection and temperature. These results suggest that the temperature is an ecological driver of virus-virus interactions, with an effect on the genetic diversity of individual viruses that are co-infecting an individual host. This research provides insights into the effect that changes in host growth temperatures might have on the evolutionary dynamics of viral populations in mixed infections.
Article
Full-text available
Robustness is the invariance of phenotypes in the face of perturbation. The robustness of phenotypes appears at various levels of biological organization, including gene expression, protein folding, metabolic flux, physiological homeostasis, development, and even organismal fitness. The mechanisms underlying robustness are diverse, ranging from thermodynamic stability at the RNA and protein level to behavior at the organismal level. Phenotypes can be robust either against heritable perturbations (e.g., mutations) or nonheritable perturbations (e.g., the weather). Here we primarily focus on the first kind of robustness—genetic robustness—and survey three growing avenues of research: (1) measuring genetic robustness in nature and in the laboratory; (2) understanding the evolution of genetic robustness; and (3) exploring the implications of genetic robustness for future evolution.
Article
Full-text available
Understanding the mechanisms by which plants trigger host defenses in response to viruses has been a challenging problem owing to the multiplicity of factors and complexity of interactions involved. The advent of genomic techniques, however, has opened the possibility to grasp a global picture of the interaction. Here, we used Arabidopsis thaliana to identify and compare genes that are differentially regulated upon infection with seven distinct (+)ssRNA and one ssDNA plant viruses. In a first approach, we established lists of genes differentially affected by each virus and compared their involvement in biological functions and metabolic processes. We found that phylogenetically-related viruses significantly alter the expression of similar genes and that viruses naturally infecting Brassicaceae display a greater overlap in the plant response. In a second approach, virus-regulated genes were contextualized using transcriptional and protein-protein interaction networks of A. thaliana. Our results confirm that host cells undergo significant reprogramming of their transcriptome during infection, which is possibly a central requirement for the mounting of host defenses. We uncovered a general mode of action in which perturbations preferentially affect genes that are highly connected, central and organized in modules. This indicates redundancy in the information transmission from virus effectors to immune response genes and allowed us to suggest that this over-triggering of hub proteins is a mechanism to confer robustness for expression of host defenses.
Data
Full-text available
Whether evolution is erratic due to random historical details, or is repeatedly directed along similar paths by certain constraints, remains unclear. Epistasis (i.e. non-additive interaction between mutations that affect fitness) is a mechanism that can contribute to both scenarios. Epistasis can constrain the type and order of selected mutations, but it can also make adaptive trajectories contingent upon the first random substitution. This effect is particularly strong under sign epistasis, when the sign of the fitness effects of a mutation depends on its genetic background. In the current study, we examine how epistatic interactions between mutations determine alternative evolutionary pathways, using in vitro evolution of the antibiotic resistance enzyme TEM-1 β-lactamase. First, we describe the diversity of adaptive pathways among replicate lines during evolution for resistance to a novel antibiotic (cefotaxime). Consistent with the prediction of epistatic constraints, most lines increased resistance by acquiring three mutations in a fixed order. However, a few lines deviated from this pattern. Next, to test whether negative interactions between alternative initial substitutions drive this divergence, alleles containing initial substitutions from the deviating lines were evolved under identical conditions. Indeed, these alternative initial substitutions consistently led to lower adaptive peaks, involving more and other substitutions than those observed in the common pathway. We found that a combination of decreased enzymatic activity and lower folding cooperativity underlies negative sign epistasis in the clash between key mutations in the common and deviating lines (Gly238Ser and Arg164Ser, respectively). Our results demonstrate that epistasis contributes to contingency in protein evolution by amplifying the selective consequences of random mutations
Article
There has been a recent revival of interest in how genetic interactions evolve, spurred on by an increase in our knowledge of genetic interactions at the molecular level. Empirical work on genetic networks has revealed a surprising amount of robustness to perturbations, suggesting that robustness is an evolved feature of genetic networks. Here, we derive a general model for the evolution of canalization that can incorporate any form of perturbation. We establish an upper bound to the strength of selection on canalization that is approximately equal to the fitness load in the system. This method makes it possible to compare different forms of perturbation, including genetic, developmental, and environmental effects. In general, load that arises from mutational processes is low because the mutation rate is itself low. Mutation load can create selection for canalization in a small network that can be achieved through dominance evolution or gene duplication, and in each case selection for canalization is weak at best. In larger genetic networks, selection on genetic canalization can be reasonably strong because larger networks have higher mutational load. Because load induced through migration, segregation, developmental noise, and environmental variance is not mutation limited, each can cause strong selection for canalization.
Article
Article
Abstract,88 3.1. Introduction ,88 3.2. Markers of a Successful Real-Time RT-PCR Assay,88 3.2.1. RNA Extraction ,88 3.2.2. Reverse Transcription ,90 3.2.3.,Comparison of qRT-PCR with Classical End-Point Detection Method,91 3.2.4. Chemistry Developments for Real-Time RT-PCR,92 3.2.5. Real-Time RT-PCR Platforms,92 3.2.6. Quantification Strategies in Kinetic RT-PCR,92
Article
The importance of genetic interactions in the evolutionary process has been debated for more than half a century. Genetic interactions such as underdominance and epistasis (the interaction among genetic loci in their effects on phenotypes or fitness) can play a special role in the evolutionary process because they can create multiple fitness optima (adaptive peaks) separated by fitness minima (adaptive valleys). The valleys prevent deterministic evolution from one peak to another. We review the evidence that genetic interaction is a common phenomenon in natural populations. Some studies give strong circumstantial evidence for multiple fitness peaks, although the mapping of epistatic interactions onto fitness surfaces remains incompletely explored, and absolute proof that multiple peaks exist can be shown to be empirically impossible. We show that there are many reasons that epistatic polymorphism is very difficult to find, even when interactions are an extremely important part of the genetic system. When polymorphism results in the presence of multiple fitness peaks within a group of interbreeding populations, one fitness peak will quickly be nearly fixed within all interbreeding populations, but when epistatic or underdominant loci are nearly fixed, there will be no direct evidence of genetic interaction. Thus when complex landscapes are evolutionarily most important, evidence for alternative high fitness genetic combinations will be most ephemeral. Genetic interactions have been most clearly demonstrated in wide crosses within species and among closely related species. This evidence suggests that genetic interactions may play an important role in taxonomic diversification and species-level constraints. Population genetic analyses linked with new approaches in metabolic and molecular genetic research are likely to provide exciting new insights into the role of gene interactions in the evolutionary process.