#PAGE_PARAMS# #ADS_HEAD_SCRIPTS# #MICRODATA#

A de novo approach to inferring within-host fitness effects during untreated HIV-1 infection


Authors: Christopher J. R. Illingworth aff001;  Jayna Raghwani aff004;  David Serwadda aff006;  Nelson K. Sewankambo aff006;  Merlin L. Robb aff008;  Michael A. Eller aff008;  Andrew R. Redd aff009;  Thomas C. Quinn aff010;  Katrina A. Lythgoe aff004
Authors place of work: Department of Genetics, University of Cambridge, Cambridge, United Kingdom aff001;  Department of Applied Mathematics and Theoretical Physics, University of Cambridge, Cambridge, United Kingdom aff002;  School of Chemical and Biological Sciences, Queen Mary University of London, London, United Kingdom aff003;  Big Data Institute, Li Ka Shing Centre for Health Information and Discovery, Nuffield Department of Medicine, University of Oxford, Oxford, United Kingdom aff004;  Department of Zoology, Peter Medawar Building, University of Oxford, Oxford, United Kingdom aff005;  Rakai Health Sciences Program, Kalisizo, Uganda, School of Public Health, Makerere University, Kampala, Uganda aff006;  School of Medicine, Makerere University, College of Health Sciences, Kampala, Uganda aff007;  U.S. Military HIV Research Program, Walter Reed Army Institute of Research, Silver Spring, Maryland, United States of America aff008;  Henry M. Jackson Foundation for the Advancement of Military Medicine, Bethesda, Maryland, United States of America aff009;  Department of Medicine, Johns Hopkins School of Medicine, Johns Hopkins University, Baltimore, Maryland, United States of America aff010;  Laboratory of Immunoregulation, Division of Intramural Research, NIAID, NIH, Baltimore Maryland, United States of America aff011
Published in the journal: A de novo approach to inferring within-host fitness effects during untreated HIV-1 infection. PLoS Pathog 16(6): e32767. doi:10.1371/journal.ppat.1008171
Category: Research Article
doi: https://doi.org/10.1371/journal.ppat.1008171

Summary

In the absence of effective antiviral therapy, HIV-1 evolves in response to the within-host environment, of which the immune system is an important aspect. During the earliest stages of infection, this process of evolution is very rapid, driven by a small number of CTL escape mutations. As the infection progresses, immune escape variants evolve under reduced magnitudes of selection, while competition between an increasing number of polymorphic alleles (i.e., clonal interference) makes it difficult to quantify the magnitude of selection acting upon specific variant alleles. To tackle this complex problem, we developed a novel multi-locus inference method to evaluate the role of selection during the chronic stage of within-host infection. We applied this method to targeted sequence data from the p24 and gp41 regions of HIV-1 collected from 34 patients with long-term untreated HIV-1 infection. We identify a broad distribution of beneficial fitness effects during infection, with a small number of variants evolving under strong selection and very many variants evolving under weaker selection. The uniquely large number of infections analysed granted a previously unparalleled statistical power to identify loci at which selection could be inferred to act with statistical confidence. Our model makes no prior assumptions about the nature of alleles under selection, such that any synonymous or non-synonymous variant may be inferred to evolve under selection. However, the majority of variants inferred with confidence to be under selection were non-synonymous in nature, and in most cases were have previously been associated with either CTL escape in p24 or neutralising antibody escape in gp41. We also identified a putative new CTL escape site (residue 286 in gag), and a region of gp41 (including residues 644, 648, 655 in env) likely to be associated with immune escape. Sites inferred to be under selection in multiple hosts have high within-host and between-host diversity although not all sites with high between-host diversity were inferred to be under selection at the within-host level. Our identification of selection at sites associated with resistance to broadly neutralising antibodies (bNAbs) highlights the need to fully understand the role of selection in untreated individuals when designing bNAb based therapies.

Keywords:

Haplotypes – Natural selection – Alleles – HIV-1 – Antibodies – Species diversity – Viral evolution – Evolutionary immunology

Introduction

In the absence of effective antiretroviral therapy, HIV-1 evolves rapidly during infection. A key driver of evolution is the influence of the host immune system; cytotoxic CD8+ T-cells (CTLs) and neutralising antibodies (nAbs) impose selection on the virus, leading to the emergence of immune escape mutations[1]. However, other factors also influence viral evolution. For example, the host-specific nature of the immune response leads to the accumulation of mutations which are deleterious to the virus upon transmission to a new host. During the course of a new infection such variants are often lost, in particular where they occur at sites which in general are under strong purifying selection[2,3]. Selection may further act for protein or RNA secondary structure[4,5].

The complex nature of selection has led to multiple studies evaluating how the viral genotype may be both constrained and shaped during the course of evolution. These include the use of techniques for in vitro mutagenesis, and analyses of viral sequence data, evaluated at the level of population consensus or through deep sequencing exploring within-host variation at one or more time points during infection. For example, mutagenesis of HIV-1 proteins has allowed the measurement in vitro of the effect of specific mutations [6]. The development of technologies for high-throughput mutagenesis has enabled such measurements to be made across very large sets of potential mutations[79]. Mathematical methods combining such results have been used to generate an overview of fitness costs and epistatic effects for the virus[10]. Measurements of this form provide a base-level estimation of the general fitness landscape of the virus, although the extent to which in vitro data captures the behaviour of the virus in a human host may be limited.

Many years of study of HIV-1 have led to the collection of consensus genome sequence data for a large number of individual infections[11]. Such data have allowed techniques such as the fitting of maximum entropy models, which characterise the fitness costs of non-consensus variants in regions of the viral genome[1214]. While the evolution of HIV-1 occurs in within-host environments that differ between individual hosts, these models provide something of a mean picture of the viral response across an averaged, within-host environment [15,16]. In these models the extent of conservation at a particular genetic locus indicates the extent to which purifying selection acts upon the majority allele[17].

Short-read deep-sequencing data has provided valuable insights into how fitness effects shape the evolution of HIV-1. Studies can be broadly categorised into those that consider purifying selection, and those that consider positive selection. Purifying (or negative) selection represents the process by which deleterious variants are purged from a population. Over time the frequency of a variant under purifying selection evolves in a statistically predictable way towards an equilibrium state via mutation-selection balance[18]. Exploiting this fact, allele frequencies observed over time during single untreated infections[2], or at single time points within multiple infections, have been used to estimate the magnitude of selection and the mutation rate acting upon distinct regions of the genome[17,19].

Positive selection represents the process by which favourable variants are driven towards fixation. As with purifying selection population genetic methods can be adopted for the inference of fitness effects. For example, a series of models have been developed for the inference of HIV-1 escape rates from CTL responses. Whereas earlier approaches to this problem considered viral escape from a single CTL response[1922], more recent studies have considered the multiple immune responses that arise successively during infection [2325]. Under such circumstances, interference between beneficial viral mutations affects the population dynamics [26]. Therefore, accounting for this clonal interference is critical if the role of selection is to be correctly inferred[2729].

Studies assessing fitness effects in within-host HIV-1 infection have often focused upon the earliest stages of infection when strong selection on CTL escape mutations typically dominates the viral population dynamics[23,29]; in this circumstance, we can model evolution as a competition between a relatively small number of viral genotypes [23,30]. Later in infection, where escape mutations are less strongly beneficial, and where synonymous diversity has had longer to accumulate [2], the potential for hitchhiking and clonal interference is greater, such that variants observed at high frequency are less certain to have evolved under positive selection. In this circumstance, distinguishing selected from non-selected variants is a substantial challenge. To address this, we here present a de novo approach for inferring selection from HIV-1 sequence data in which any variant allele may, in theory, be detected as under selection. We adopt a parsimonious approach, assigning selection to the smallest set of variants required to explain the observed multi-locus sequence data under a likelihood model. Applied to targeted sequence data from a substantial cohort of 34 untreated individuals living in Uganda[31], we determine how selection drives viral evolution. In the presence of pervasive interference between alleles in linkage disequilibrium with one another, our consideration of data from a large number of individuals is fundamental in providing statistical confidence in the assignment of selection. Specifically, the repeated inference of selection at the same locus in different individuals enhances the power of our study to elucidate how selection during individual infections shapes genetic diversity at the population level.

Results

We applied an evolutionary inference method to deep-sequencing data spanning multiple years of infection from 34 untreated individuals living in Rakai, Uganda, enabling us to infer positive selection acting on part of the gp41 region of env (324 base pairs) and the p24 region of gag (387 base pairs)[31].

Extent of selection

An initial application of our method found extensive evidence of positive selection in the viral genome, with 74 (out of 387) nucleotide sites in p24 and 81 (out of 324) nucleotide sites in gp41 being inferred to evolve under positive selection in at least one individual (S1 Table and S1 Fig). Our method explicitly accounts for linkage disequilibrium between alleles observed in the sequencing data[27,32]; a potential remains for interactions between observed alleles in the targeted sequence region, and non-observed alleles in flanking regions of the genome (Fig 1). To estimate the effect of these interactions on our results, we ran simulations to replicate the dynamics of real infections. In a multi-locus system where all alleles under selection are observed (i.e. in which the full data of the system are available), our approach performs very well, identifying the majority of variants under selection with very few false positive inferences. However, when non-observed selected alleles interact with observed alleles via linkage disequilibrium (i.e. where only partial data of the system is available), our approach is prone to generating false positive inferences of selection (S1 Text, S3 and S4 Tables). Data from our simulation study allowed us to set statistical criteria via which we could combine inferences from multiple patients and confidently identify sites under selection, despite the presence of false positive calls.

Linkage disequilibrium between alleles in the virus.
Fig. 1. Linkage disequilibrium between alleles in the virus.
Here positive selection at locus A affects the behaviour of alleles at locus B due to linkage. Our method of inference corrects for such effects if A is observed, whereas a single-locus method could lead to incorrectly inferring selection at B. However, in the system shown linkage disequilibrium also exists between alleles at A and B and those at the unobserved locus C. Such effects cannot be accounted for by our approach. We used a simulation-based approach to estimate the importance of such effects, so as to account for their influence upon the results generated by our method.

Strength and time of onset of selection

Inferred variants generally evolved under weak selection. In constructing a distribution of the strength of selection among selected variants, we first assessed the degree of precision with which this statistic could be inferred at each locus. The uncertainty in a given estimate depends upon the extent to which data are available. For example, when a variant emerges and fixes between two time points, only a lower bound on the strength of selection can be inferred[33]. We therefore generated confidence intervals for each inferred magnitude of selection using a likelihood-based method, retaining only variants for which the upper and lower bounds of this interval differed by no more than an order of magnitude. Distributions of fitness effects compiled from these variants showed that most of the identified alleles under selection experienced very weak selective effects, with long tails of alleles evolving under strong positive selection (Fig 2). Application of our method to simulated data highlighted an undercalling of very weakly selected variants, and an underestimation of the magnitude of selection affecting the strongest variants (S2 Fig), however the overall inferred distribution of selection coefficients was not statistically different to the ‘true’ distribution used to generate simulations (S3 Fig). Further statistical details and notes on these simulations are provided in S1 Text, S4, S5 and S6 Figs.

Histograms of inferred strengths of selection and times in days of onset of selection across 34 individuals.
Fig. 2. Histograms of inferred strengths of selection and times in days of onset of selection across 34 individuals.
The maximum likelihood estimates are shown in each case, for the subset of the data where the upper and lower bound confidence intervals on the strength of selection are within an order of magnitude of each other. Bars indicate variants at nucleotide sites, and are coloured according to whether they are within AA positions associated with differences in susceptibility to CTLs (pink) or susceptibility to NAbs (blue). Where a codon is implicated in both a NAb and CTL response, for clarity it is coloured blue (see S2 Table).

In p24, around 65% of inferred selected variants had a strength of selection of less than 5% per generation, while in gp41 approximately half of variants were under this threshold. In our model, selection of 5% per generation would cause, in the absence of interference effects, a change in allele frequency from 5% to 95% in a period of just under eight months. Most selection acting upon variants was inferred to kick in during the first year of infection and almost all within two years, and we found no evidence of a correlation between the strength and time of onset of selection (S7 Fig). Our approach underestimates the proportion of variants under weak selection; very weak selection would not produce a change in the population sufficient to be identified from the data. The extreme beneficial end of the distribution may also be under-represented because variants which fixed within individuals before the first sample was collected cannot be identified from our data. Finally, the finite period of time over which sequencing was performed could restrict the inference of more lately selected variation because a selected variant arising later in infection would have less time to affect the composition of the population in an observable way.

Distribution of selected variants among individuals

Using the simulated data, we identified a statistical threshold at which we could robustly identify specific sites in the genome containing variants under selection despite the presence of false positive inferences of selection. A unique aspect of this dataset is the large number of untreated individuals included in the study. Where alleles at the same locus were inferred to be under selection in multiple individuals, a statistical approach was used to infer loci at which, under conservative assumptions, at least one of the inferred variants is genuinely under selection. Taking into account different patterns for nonsynonymous and synonymous mutations, we estimated per-site false-positive rates for nonsynonymous and synonymous mutations, which were subsequently used to identify with statistical confidence sites that were under selection in at least one individual in our dataset. The process used for estimation is described in full in S1 Text. Taking all 34 individuals into consideration, we calculated that in p24 we could be confident that a site is under selection in at least one individual if mutations were inferred to be under selection in at least five individuals, and/or if nonsynonymous mutations were inferred to be under selection in two or more individuals, while in gp41 we could be confident if at least five mutations and/or three nonsynonymous mutations were inferred to be under selection. This is a conservative approach, and will exclude sites genuinely under selection in only one or a few individuals, for example sites associated with escape from rare HLA alleles or NAbs, or other less common forms of selection.

Applying these criteria, we identified 11 specific nucleotide sites, representing 10 amino acid (AA) positions, under selection in p24, and likewise we identified 11 such sites, representing 8 AA positions, in gp41. All but four mutations at these AA positions represented nonsynonymous changes (see Table 1). Occasionally two different codons at the same AA position were found to be subject to selection in a single individual, but in general repeated inferences of selection at an AA position occurred in distinct individuals (S1 Table). The positions of the identified mutations in their respective protein structure are shown in Fig 3.

Locations of selected positions on the protein structures of p24 and gp41.
Fig. 3. Locations of selected positions on the protein structures of p24 and gp41.
Protein structures with pdb identities and 3J34 and 6NIJ were used to plot these figures [103,104]. Parts of proteins that were covered by sequence data are shown in gray, with the remainder of the protein in white van der Waals representation. Amino acid positions identified by our analysis are shown in red. The gp41 structure does not show the entirety of the sequenced region, with amino acids at positions 670 and 674 also identified in our analysis, but not in the protein structure. Other nearby proteins are shown in white surface representation. Figures were created using the VMD software package [105].
Tab. 1. Summary of amino acid positions containing sites under selection.
Summary of amino acid positions containing sites under selection.

Using the Los Alamos HIV database (http://www.hiv.lanl.gov), for all of the AA positions in our study, we determined whether they have previously been associated with changes in CTL susceptibility or NAbs (see Methods, Fig 4, S2 Table). In p24, 43% of sites have previously been associated with changes in CTL susceptibility and/or compensatory mutations, whilst 69% of the variants we inferred to be under selection were associated with CTL and/or compensatory mutations. For the 10 AA positions identified as almost certainly under selection, nine have previously been associated with changes in CTL susceptibility and/or compensatory mutations, and lie in epitope regions recognised by multiple HLA alleles present in the Ugandan population[6668]. The remaining codon (residue 286 in HXB2 gag) lies within CTL epitopes recognised by human leukocyte antigen (HLA) alleles that are relatively common in the Ugandan population (A1101 and B27[66]) and therefore mutations at this site probably affect sensitivity to CTLs, or possibly compensatory mutations. In gp41, 44% of AA sites have previously been associated with NAbs, whilst 61% of the variants we inferred to be under selection were associated with NAbs. Of the eight AA positions identified as being almost certainly under selection, five (residues 620, 624, 655, 674 and 677 of env) have previously been associated with NAbs, including broadly neutralising antibodies (bNAbs). These are associated with the gp120/g41 interface and the membrane-proximal external region (MPER). The other three codons (residues 641, 644 and 648 of env) all lie within an epitope region that is targeted by the monoclonal antibody HGF24, which has been shown to neutralise some viruses isolated from the African continent[58]. Although we cannot rule out other sources of selection in env, given the position of these codons on the genome, we believe selection due to humoral immune pressure is more likely. As such, all of the sites we identified as under selection in p24 probably affect susceptibility to CTLs or associated compensatory mutations, whereas all of the sites under selection in gp41 probably affect susceptibility to NAbs or associated compensatory mutations.

Amino Acid (AA) positions containing nucleotide sites that with confidence were inferred to be under selection in at least one individual.
Fig. 4. Amino Acid (AA) positions containing nucleotide sites that with confidence were inferred to be under selection in at least one individual.
Codon positions are in relation to the HXB2 reference sequence and the vertical y-axis gives the number of times that codon was inferred to be under selection across 34 individuals. Occasionally the same AA position was inferred to be under selection twice in the same individual. Pink: AA positions associated with changes in susceptibility to CTLs or compensatory mutations; Blue: AA positions associated with susceptibility to NAbs. Selection at Gag residue 286 possibly reflects CTL escape in one individual and reversion in another. Selection at Env residues 641, 644, and 648 possibly reflect selection associated with NAbs; these are all within an epitope region targeted by the neutralising antibody HGF24 in some African Isolates[58]. Where an AA position is implicated in CTL and NAb responses, for clarity they are coloured in blue.

Direction of selection

Since mutations that enable the virus to evade host immune responses are often costly in terms of viral replication[69,70], sites harbouring escape mutations are likely to evolve towards the population-level consensus when transmitted to a new host since the selective pressure of a specific immune response is removed, but costs associated with the mutation remain [2,31,7174]. This hypothesis is supported by the observation that allelic substitutions during the course of untreated infection occur towards population consensus much more frequently than expected by chance [2,31]. If analysis is restricted to sites where substitutions are observed, this bias is confined mainly to nonsynonymous substitutions[31], which is expected if immune escape mutations are generally nonsynonymous.

Since the HLA types of the 34 individuals in our study are unknown, we followed previous studies by classifying variants we confidently inferred to be under selection as escapes or reversions depending on the population-level consensus[2,71]. Specifically, we classified selected variants as escapes if they resulted in an AA change away from the subtype-specific population-level consensus in Uganda during the period the individuals were being sampled (see Methods for full details). Conversely, inferred selected variants were classified as reversions if they resulted in change towards the subtype-specific population-level consensus. Of the 32 selected nonsynonymous mutations identified in p24, 13 were classified as escapes, 14 as reversions, and five as neither. The similar number of escapes and reversions observed is expected if CTL immune escape mutations are costly in non-HLA matched hosts, with escape in one individual followed by reversion in the next. The observation of one escape and one reversion in residue 286 of gag supports our prediction that this locus likely affects CTL susceptibility. Of the 40 nonsynonymous mutations identified in gp41, 18 were classified as escapes, 11 as reversions, and 11 as neither. The fact that over a quarter of AA changes are towards population consensus suggests that in many cases antibody escape mutations are deleterious in hosts without a matching antibody, but the stereotypical pattern of adaptation in one individual followed by reversion in another found at CTL immune epitopes is likely more complex for antibody-escape evolution, where escape mutations are sometimes, but not always costly in the absence of an antibody response [75,76]. Among the three residues 641, 644 and 648 of env, five of the nine nonsynonymous variants are escapes, three are reversions, and two are neither, supporting our prediction that these sites affect susceptibility to host (probably NAb) immune responses.

We also determined the direction of change for all variants inferred to be under selection, which will likely include variants under direct selection and variants changing in frequency due to hitch-hiking, and whether they represented synonymous or nonsynonymous changes (S8 Fig). We found a clear correlation between the number of times that selection was inferred at an AA position and the proportion of selected variants that were nonsynonymous (linear regression, p24 p = 0.009, r2 = 0.85; gp41 p = 0.007, r2 = 0.67). For gp41, a correlation was also found between repeated selection at an AA position and a pattern of evolution towards the population consensus (p = 0.011, r2 = 0.63); around half of inferred selection events were towards population consensus at the AA position most frequently inferred to be under selection. Although for p24 the linear regression did not reveal a significant trend (p = 0.400, r2 = 0.18), there is a distinction between AA positions selected 2 or more times, which have a high probability of being towards subtype-specific population consensus (44%), and AA positions selected once, which have only a small probability of being towards the consensus (5%). Our interpretation is that AA positions represented once in our analysis disproportionately represent mutations increasing in frequency due to hitch-hiking, with these mutations tending to be synonymous and away from population level consensus. AA positions represented multiple times, on the other hand, are more likely to represent immune escapes and reversions, and therefore tend to be nonsynonymous but with only around half of mutations away from consensus.

Comparing codon diversity at the within-host and population scale

Our data showed a strong relationship between within- and between-host sequence diversities, where the diversity of codons was measured at each AA position. Within-host diversity was measured approximately three years after seroconversion for each of the 34 individuals, and the mean calculated. Diversity at the population scale was calculated as the mean of the diversities for each of the subtypes A, D, and C, using virus sequences from a large number of individuals living in Uganda around the same time as the 34 individuals in our study (see Methods). Consistent with previous studies [2], we identified a strong relationship between measurements of sequence diversity calculated at the within-host and population scales (Fig 5, S2 Table), with all AA positions found to be highly diverse at the within-host level also highly diverse at the population level. Moreover, all but one of the AA positions containing nucleotide sites that we are confident are under selection are also diverse at the population level (the exception being residue 302 in gag). Given that most changes at these sites probably reflect escape from host immune responses, compensatory mutations, or reversions of these escapes in subsequent individuals, diversity at the population scale at these AA positions is likely maintained by the differing selection pressures faced by variants in different hosts due to different immunological backgrounds.

Within- and between-host codon diversity.
Fig. 5. Within- and between-host codon diversity.
For every AA position in our region of analysis we determined the within- and between-host codon diversity. Large markers denote AA positions in which we are confident selection is occurring in at least one individual. Markers are coloured if they are associated with changes in sensitivity to CTLs or compensatory mutations (pink) or NAbs (blue). The site confidently under selection in p24 but which is unspecified (Gag residue 286) is possibly associated with CTL escape. The sites confidently under selection in gp41 but which are unspecified (Env residues 641, 644, and 648) possibly reflect selection associated with NAbs; these are all within an epitope region targeted by the neutralising antibody HGF24 in some African Isolates[58]. Where an AA position is implicated in CTL and NAb responses, for clarity they are coloured in blue.

Not all AA positions found to be highly diverse at the population level were detected as being under selection, with a high degree of confidence, at the within-host scale. This observation could arise if codons at these positions are under selection in some individuals in the population, but we failed to observe it with confidence, either because we sampled too few individuals, selection was too weak to be detected, and/or selection drove fixation events before the first sampling time point. In addition, at unconstrained sites experiencing little or no selection, diversity might gradually accumulate at the population scale due to drift, exacerbated by small bottleneck sizes at transmission. This could explain the high levels of population diversity at positions 235 in gag and 609 in env; almost all of the diversity observed at these codon positions is due to the presence of synonymous variants.

Discussion

We developed a novel inference framework to infer the extent of selection acting upon variants which drive the evolution of within-host HIV-1 populations, considering data from the p24 region of gag, and gp41 of env, from 34 longitudinally sampled untreated individuals. A frequent assumption is that beneficial mutations will rapidly spread within individuals once they occur[77]. For example, it is well established that CTL-escape mutations accumulate and spread rapidly during acute infection [1925], though that the rate of allele fixation decreases during chronic infection[22,25]. However, estimating the extent and strength of positive selection during infection more generally is challenging due to genetic linkage among variant alleles, which makes differentiating between selected variants and variants that are increasing in frequency due to linkage with a selected variant (hitch-hiking) difficult. Our de novo approach incorporates genetic linkage and recombination. Furthermore, it is generally assumed that during untreated infection selected variants are associated with immune escape; our approach is agnostic with regards to phenotypic data, potentially allowing any polymorphic site in the genome to be identified as under positive selection. Indeed, using our approach we discovered four sites in the genome likely to be under immune pressure (both CTL and antibody) that were not listed in the Los Alamos database.

Our results indicate a pattern of weak and slow selective sweeps characterising evolution during chronic HIV-1 infection, with stronger faster selective sweeps being relatively rare. We note that where ‘weak’ selection was inferred, this was still on a scale outweighing the effects of genetic drift. Studies of the effective population size of HIV-1 have indicated a value in excess of 105 [76]; given such a value, selection of the order of 5% per generation is comfortably within a realm whereby the influence of selection dominates that of genetic drift [78]. An important caveat is that the first sampling time point for each individual in our analysis is estimated to be between 150 and 425 days since seroconversion, and therefore we will not detect variants that were under strong selection and rapidly reached fixation before the first sampling time point. Furthermore, the magnitude of the most strongly selected variants could not always be quantified; where fixation occurs entirely in the interval between two consecutive time points, no upper bound on the magnitude of selection could be fixed.

A unique aspect of our study is the large number of individuals for which we have data. Comparisons among individuals revealed AA positions which were inferred to be under selection in multiple (up to ten) individuals. Most of these sites have previously been identified as affecting sensitivity to CTLs (in p24) or NAbs (in gp41), with selected changes at these sites likely reflecting the gain or loss immune-escape mutations, or escape-related compensatory mutations (although other sources of selection cannot be ruled out). In addition, we identified four AA positions under selection in multiple individuals that haven’t previously been identified as affecting sensitivity to CTLs or NAbs (residue 286 in gag, and residues 641, 644 and 648 in env). Given the patterns of selection at these sites, reflecting evolution both away and towards population consensus, it is likely these sites also affect sensitivity to CTLs (gag) or NAbs (env). Sites under selection in multiple individuals were also found to be highly diverse at the population level. This again is consistent with a pattern where a minority of codons are repeatedly under selection, likely representing adaptation to the immunological background of some individuals, which revert upon transmission to subsequent individuals; a pattern which has been referred to as “adapt and revert”[79]. Although we do not know the HLA-type of the infected individuals in our study, the number of putative CTL escapes and reversions is consistent with the frequency of different HLA alleles in the Ugandan population.

Perhaps less expected in our analysis was the identification of AA positions that are associated with NAbs and which were found to be under selection in a large number of individuals; for one site selection was inferred in nearly a third of individuals, with another inferred in a quarter of individuals. The implication is that the same epitopes are frequently targeted by antibodies in different individuals, and with similar means of viral escape. Moreover, since around a quarter of changes at these sites are towards the subtype-specific population level consensus, many may well represent the reversion of costly antibody-escape mutations from previous individuals, supporting the observation that some but not all antibody-escape mutations are costly [70,75,8084]. These patterns can help explain why resistance to antibodies has increased over the course of the epidemic [8589], but also highlights that viral evolution at the population level in response to bNAb-based interventions is likely to be complex, involving evolutionary responses to both naturally and therapeutically induced immune responses.

Even though our framework explicitly accounts for linkage disequilibrium between observed variants, it is still vulnerable to false positive inferences of selection due to linkage disequilibrium with unobserved variants flanking the genetic regions we analysed. Although simulated data suggested that the overall distribution of fitness effects was robust to this vulnerability, our study should serve as a cautionary note; where multiple alleles evolve in linkage disequilibrium, care is needed in identifying selection with any particular allele. The large number of individuals included in our study enabled us to partly circumvent this problem by only assigning confidence that any particular nucleotide site is under selection if it is inferred to be under selection in multiple individuals. Indeed, evidence for the validity of our method is provided by the repeated observation of variants under weak selection across multiple individual infections, with these changes making biological sense under the “adapt and revert” hypothesis. Our results emphasise the role of immune escape in driving evolution during chronic infection, shaping patterns of diversity at the population level, and provides new insights that could be useful in the development of immune-based interventions, particularly in the context of viruses circulating in Africa.

Methods

In order to evaluate selection within a host, we employed a likelihood-based inference framework to infer the most parsimonious explanation of the sequence data in terms of a model of selection acting for specific nucleotides in the viral population. Some of the mathematical aspects of this framework have previously been applied in studies of the within-host evolution of the influenza virus[32,90], although the details of the model used here tailor it to HIV-1 infection. Our model explicitly accounts for linkage disequilibrium between alleles and builds upon earlier approaches for inferring selection in cases where linkage is of importance for evolution [23,27,91,92].

Model outline

Our model proceeds through a number of steps (see below for full details). (1) We identified variant alleles from the sequence data using a simple frequency cutoff. Measurements of the frequencies of variant alleles over time were collected into trajectories, each trajectory describing the frequency of a single allele over time. (2) Sets of alleles with similar trajectories were identified, and under the assumption that all of the differences between these similar trajectories resulted from noise in the data, the extent of noise, modelled as a single parameter, was estimated. The noise parameter defines a likelihood function for the data, the existence of which allows for fits to be made between the data and a number of models describing the evolution of the population. (3) Models of evolution at a single locus were used to assess all of the derived trajectories. Comparing the likelihoods of models describing evolution under selective neutrality, and under selection, we identified, using model selection, a subset of trajectories which potentially evolved under non-neutral selection. Such selection could arise either from intrinsic selection for or against the allele, or via linkage disequilibrium with an intrinsically selected allele elsewhere in the genome. (4) We combined the alleles present at each locus that had a potentially non-neutral trajectory into haplotypes, with these haplotypes only describing alleles at the loci identified to have potentially non-neutral trajectories (S9 Fig). (5) For each patient, the number of reads with each observed haplotype at each time of sampling was counted, forming a multi-locus dataset describing the evolution of the virus in that patient over time. (6) A series of multi-locus evolutionary models were fitted to the data from each patient, identifying the most parsimonious explanation of the data in terms of time-dependent selection acting on individual alleles in each dataset. (7) Confidence intervals for the inferred selection parameters were generated, again making use of the likelihood function. The model is thus identical to that described in a previous publication [33] with differences only in the approach to identifying noise in step two, in the models used to fit the data in step six, and in the calculation of confidence intervals for selection.

Sequencing data

For our evolutionary analysis we used previously generated deep-sequence data from 34 longitudinally sampled individuals participating in the Rakai Community Cohort study and co-enrolled in the Molecular Epidemiology Research (MER) seroconverter study. Targeted short-read deep-sequence data from the p24 region of gag (390 bp; HXB2 reference genome positions 1429–1816) and the gp41 region of env (324 bp; HXB2 7941–8264) had been sequenced using the 454 sequencing platform (Roche, Branford, CT). All individuals were untreated, with a first sampling time point around one year since seroconversion, and typically 3 or 4 subsequent time points spanning between two and seven years of infection (see Table 1 in Raghwani et al 2019). Aligned sequences can be found at https://github.com/katrinalythgoe/

RakaiHIV. Further details on the individuals, including viral loads and CD4 counts, and sequencing methods used have been given elsewhere[31,93].

Calling of variant alleles and trajectories from sequence data

Single-locus variants were identified in the data using the SAMFIRE software package[94]. Variants with a minimum allele frequency of at least 1% for at least one time point in the course of infection were identified. Variant frequencies collected over time were described in terms of trajectories. By way of notation, in a given patient we denote the trajectory


comprising the observed frequencies of allele a at locus i across all recorded times tk. Here K is the total number of points in time at which the population was observed via sequencing, which varied between 2 and 5. We note that the frequency is calculated simply as, the number of observations of allele a at locus i at time tk divided by the total number of alleles observed at locus i at time tk.

Estimating the extent of noise in the data

Noise in sequence data can arise either through the collection of an unrepresentative sample of viruses from a patient, or via errors induced in the experimental processing and sequencing of that sample [93]. We here applied a heuristic method to derive a conservative estimate of the extent of noise in the data from the data itself. This was achieved by exploiting effects caused by genetic hitch-hiking[95]. If two alleles appear uniquely upon a shared genetic background, they will initially share an identical allele frequency. Over time the allele frequencies will change in a very similar manner, differences arising over time as a result of recombination between distinct haplotypes. We thus identified putatively hitch-hiking trajectories to derive an estimate of noise in the sequence dataset.

For this analysis, we considered loci at which a minor allele frequency of at least 10% was observed in samples collected at two points in time. Loci in HIV can potentially have multiple alleles satisfying this condition. For each pair of such loci, i, and j, we found the alleles a*, b* minimising the statistic


where the minimisation was calculated over all polymorphic alleles at the loci i and j. Initially, pairs of trajectories (i,a*) and (j,b*) were denoted as being ‘similar’ if

A set of further heuristic steps was then applied to refine these sets of trajectories. On timescales close to those over which the data for this study was measured, recombination in HIV-1 has been noted as being of importance over genetic distances greater than 100 nucleotides [2]; here a distance cutoff of less than or equal to 50 nucleotides was imposed between trajectories, that is |i-j| ≤ 50.

Next, under the assumption that pairs of our variants initially arose on the same background, pairs were required to have similar frequencies at the first time of observation, requiring that


Further, so as to remove pairs of trajectories for which only one was polymorphic at a given time, it was required that the maximum ratio between minor allele frequencies did not exceed 103 for the duration of the trajectories; the framework of a Dirichlet multinomial model we subsequently use to estimate the extent of noise does not perform well on very low frequencies. Remaining sets of pairs identified in each patient were clustered into sets via an iterative process; beginning with an initial pair of trajectories (i1, a1) and (i2, a2), a trajectory (im, am) was added to the set if, following filtering, (im, am) was similar to a trajectory (in, an) already in the set. Sets of trajectories generated by the above process are shown in S10 Fig.

Having identified sets of trajectories, an inference process was used to evaluate the extent of noise in the data. Conservatively, the ‘true’ allele frequencies of each set were calculated as a simple mean of the observations, thereby assuming that all differences in frequencies result from ‘noise’ in the sequencing process. Given a set of trajectories S from a single viral population, recalling that q~ia*(tk) was calculated as nia(tk)/Ni(tk), the inferred frequency at time tk was calculated as the mean fraction of variant alleles across all loci with trajectories in S. That is,


A Dirichlet multinomial model was then parameterised across all trajectory sets, finding the value of C satisfying


Where the summations are, respectively, calculated over sets of similar trajectories S, trajectories m in each set S, and samples k in each trajectory, and where the likelihood is the Dirichlet multinomial function

in which the sums and products with index a are calculated over all alleles at a specific locus, or equivalently over all haplotypes within a population, while G indicates the gamma function.


The derived value of C provides a proxy measurement of the extent of noise in the data and was used in further likelihood calculations; we inferred the value C = 100.359. Qualitatively, this value represents the extent to which the variance of a sample of haplotypes is overdispersed in relation to a simple multinomial sample; a small value of Crepresents an increased amount of uncertainty in the data, tending towards a uniform distribution, in which samples are fully uninformative, while a high value of C represents increased uncertainty, tending towards a multinomial sample in which every read precisely identifies a haplotype in a perfectly representative sample from the population. We note that patterns of noise in genome sequence data may be substantially more complex than represented by our model; our likelihood, combined with the BIC model selection framework, provides a simple yet analytically tractable approach for the inference of selection parameters from real genome sequence data.

Identification of potentially non-neutral loci

Considering frequency data from each variant allele, ‘potentially non-neutral’ loci [92] at which significant changes in allele frequency were observed over time were identified. In this process, where q1(t) denotes the frequency of the variant allele at locus i at time t, deterministic models of evolution at a single locus were fitted to the single-locus trajectories, using the equation


for neutral (σk = 0), constant (σk = s), and time dependent models of selection, retaining trajectories for which the constant or time-dependent models of selection outperformed the neutral model. This was evaluated using the Bayesian Information Criterion (BIC) to account for the increased complexity of the models including selection [96].

Construction of haplotypes

For each individual, only potentially non-neutral loci were retained for the rest of the analysis, with each combination of alleles at these sites representing a haplotype. We converted the sequence data into a set describing the number of times each haplotype was observed in the sequence data, at each sampling time point. For example, if three non-neutral loci were identified, we might count the number of reads with the alleles G, A, and C at these loci at a given time point; the proportion of such reads would specify the observed frequency of the haplotype GAC. Only viral haplotypes which were observed in the sequencing data were considered, generally representing a small fraction of the haplotypes that could potentially exist. This approximation is equivalent to the assumption that non-observed viral haplotypes were under sufficiently strong purifying selection to prevent them from reaching an observable frequency.

Multi-locus evolutionary models

Our evolutionary models considered the effect of mutation, selection, and recombination upon the population. Within a model the frequency of the haplotype a at generation tk of the within-host viral population was specified by the frequency qa(tk), frequencies changing according to the three evolutionary processes. The model system was propagated within the space of observed haplotypes using a Wright-Fisher approach of discrete generations, with successive steps of mutation, recombination and selection.

Mutation was approximated as occurring between haplotypes that differ by a single nucleotide with rate μ per generation. Recombination was approximated as occurring in a pairwise manner between haplotypes with rate ρ per base per generation. That is, if a recombination event occurring between the loci indexed i and i+1, and involving the haplotypes a and b, were to produce the haplotype c, then in our model the new haplotype was produced at rate Δi,i+1ρqa(tk)qb(tk) where Δi,i+1 is the sequence distance between loci i and i+1.

A time-dependent model of selection acting upon haplotypes was applied, simulating the changing selection acting upon HIV during an infection. The time-dependent fitness wa of a haplotype a at time t was modelled via a hierarchical model of single-locus terms


where the sum is calculated over all loci i in the set of potentially non-neutral loci identified above. The parameter si denotes selection acting for or against all haplotypes with a variant at locus i, and the parameter I{i,a,Ti} is a binary indicator function. Here, Ti is the time at which selection begins to act upon the variant allele at locus i. The indicator function is set so that I{i,a,Ti} = 1 if t>Ti, and if the haplotype a contains the variant allele at locus i, while I{i,a,Ti} = 0 if it is true either that a does not contain the variant allele at i, or if tTi.

Selection then modifies the frequency of each haplotype according to the equation


where the sum with index b indicates a sum over all haplotypes, including a. In this manner the frequency of a haplotype changes according to the relation between its fitness and the mean fitness of the total viral population.

On the basis of previous studies[77,97100], parameters for mutation and recombination were set at μ = 3 x 10−5 per generation and ρ = 10−5 per base per generation, with a generation time of two days[101], reflecting parameters derived from studies of chronic infection. Parameters σi and Ti, and initial frequencies qa(t0), were learnt from the data according to a hierarchical model framework. An initial, neutral model contained no selection parameters, every haplotype having equal fitness. Next, single-locus selection models were considered. Such a model uses the set of parameters i, a, σi, and Ti, to describe the locus and allele at which selection acted, the magnitude of selection and the time at which selection took effect. In each model, given the loci and allele under selection, the optimal values for the magnitudes and times of selection were identified, using a simple likelihood optimisation process. The Dirichlet multinomial likelihood described above, with the inferred noise parameter C, was used in this process, matching the observed haplotype frequencies to those produced by the evolutionary model. To improve coding efficiency a model lacking mutation and recombination was used to derive reasonable starting parameters for selection and haplotype frequencies; the subsequent application of a model with the addition of mutation, then the full model with mutation and recombination gave the final likelihood and maximum likelihood parameters. Replicate calculations with different random seeds were used to validate likelihood calculations. By means of an iterative process, more complex models of selection were considered. Initially, selection parameters were added, taking the best n-locus models and adding selection at an additional locus to each one. Upon the discovery of a model for which adding selection at a further locus did not improve the model, a process involving both the addition and subtraction of parameters was initiated, to the point of discovering a model for which neither adding a further pair of selection parameters, or removing an existing pair of parameters, improved the model. The comparison of models was performed using the Bayesian Information Criterion. As a conservative step, a model with an additional pair of selection parameters was required to significantly outperform a simpler model to be accepted, this being denoted by an improvement of 10 units of BIC given a maximum likelihood set of parameters for each model. Compared to an earlier model of selection at multiple loci [28], our approach has the advantage of parsimony, inferring selection at a locus only where there is specific evidence for non-neutrality at that site. Our model of time-dependent selection accounts for the expected behaviour of the host immune system against HIV. No prior distribution of selection coefficients was assumed.

Estimating confidence intervals for selection coefficients

Confidence intervals were calculated for each parameter si inferred in the maximum likelihood calculation. Supposing the maximum log likelihood for a given inferred system to be equal to some value L, error bars were generated via a constrained exploration of the model space, in which a change in model parameters was accepted if the resulting likelihood was not greater than L-2, and for which changes to the parameter of concern, σi, were constrained so that this parameter could only change in a specific direction; forcing this parameter to increase generated an estimate, after repeated iteration, for the upper error bar of this parameter, while forcing this parameter to decrease generated an estimate of the lower error bar of this parameter.

Reporting selection coefficients

In our study we report the respective fitness advantage conferred by each beneficial mutation as a percentage per generation. This statistic, s, is calculated from an inferred selection coefficient as


Direction of selection

For each of the selected mutations identified in our analysis, we determined whether evolution was towards or away from the subtype-specific population consensus. The subtype of each gene region for each individual and the population consensus for subtypes A, D and C in Uganda during a similar period to which the individuals were sampled was previously determined[31].

Amino Acid position association with immune escape

Using the Los Alamos HIV database (http://www.hiv.lanl.gov), we determined for each of the AA positions in the p24 and gp41 regions that we analysed whether they have previously been associated with changes in CTL susceptibility or neutralising antibodies. For CTL susceptibility we used the Epitope Variant and Escape Mutation Database CTL variant search tool (https://www.hiv.lanl.gov/content/immunology/variants/variant_search.html?db=ctl). AA positions were marked as being associated with susceptibility to CTLs if susceptible and/or resistant variants were returned in the search tool. We excluded codons inferred to be susceptible to CTLs if the only evidence was high levels of diversity at the population level. For neutralising antibodies we used the Neutralizing Antibody Contacts and Features search tool (https://www.hiv.lanl.gov/components/sequence/HIV/featuredb/search/env_ab_search_pub.comp), marking codons as associated with susceptibility if variants were shown or predicted to affect neutralisation by antibodies, or binding to neutralising antibodies.

Codon diversity

We used the diversity statistic π to measure mean intra- and inter-host diversity at the codon level. Given a locus a at which the read depth is N, where na of each codon a was where ni of each of the three-nucleotide codons i have been observed, we define the diversity statistic π as


This diversity statistic has been shown to be less prone than some other metrics to biases at the intra-host scale[102], and using codons (nucleotide triplet motifs) rather than amino acids or single nucleotides means our measure incorporates synonymous and nonsynonymous diversity whilst still enabling comparison with information on sensitivity to host immune responses, which is typically given at the amino acid level. To determine the mean intra-host diversity at a given codon position, we calculated π for each individual and then took the mean for all 34 individuals. To determine the mean inter-host diversity at a given codon position we calculated π for each subtype and then took the mean for all three subtypes (A, C and D).

Application of the method to simulated data

The performance of our method was evaluated using simulated data, generated in order to be as close in nature as possible to the real data. A complete discussion of the methods used is given in S1 Text.

Supporting information

S1 Fig [tif]
Codon positions containing nucleotide sites that are inferred to be under selection.

S2 Fig [tif]
Magnitudes of selection for variants that were inferred to be under selection, or not inferred to be under selection, in simulated systems in which the full data about the evolution of the system was available to the inference code.

S3 Fig [tif]

S4 Fig [tif]
Observed (solid lines) and inferred (dashed lines) haplotype frequencies for simulated data in which all loci under selection were observed.

S5 Fig [tif]
True and inferred magnitudes and timings of selection for simulated data.

S6 Fig [tif]
Observed (solid lines) and inferred (dashed lines) haplotype frequencies for simulated data in which only data from within a fraction of a simulated region was observed.

S7 Fig [tif]
No correlation between time of onset and strength of selection.

S8 Fig [tif]
Proportion of mutations inferred to be under selection that are towards population level consenus or are nonsynonymous.

S9 Fig [tif]
Illustration of the construction of haplotypes.

S10 Fig [tif]
Sets of nucleotide trajectories that were identified as putatively hitchhiking.

S1 Table [xlsx]
Summary of results for all sites inferred to be under selection

S2 Table [xlsx]
Characteristics of all codons analysed: Sensitivity to host immunity, within- and between-host diversity, and the number of times the codon was inferred under selection

S3 Table [xlsx]
True and inferred selection parameters for a case in which sequence data describes all selected alleles within a system.

S4 Table [xlsx]
True and inferred selection parameters and times for a case in which sequence data partially describes the selected alleles within a system.

S1 Text [docx]
Calculations performed on simulated data, where targeted reads describe the complete evolution of the system, and where targeted reads do not describe the complete evolution of the system.


Zdroje

1. Rambaut A, Posada D, Crandall KA, Holmes EC. The causes and consequences of HIV evolution. Nat Rev Genet. 2004;5: 52–61. doi: 10.1038/nrg1246 14708016

2. Zanini F, Brodin J, Thebo L, Lanz C, Bratt G, Albert J, et al. Population genomics of intrapatient HIV-1 evolution. Elife. 2015;4. doi: 10.7554/eLife.11282 26652000

3. Lythgoe KA, Gardner A, Pybus OG, Grove J. Short-Sighted Virus Evolution and a Germline Hypothesis for Chronic Viral Infections. Trends Microbiol. 2017;25: 336–348. doi: 10.1016/j.tim.2017.03.003 28377208

4. Keating CP, Hill MK, Hawkes DJ, Smyth RP, Isel C, Le S-Y, et al. The A-rich RNA sequences of HIV-1 pol are important for the synthesis of viral cDNA. Nucleic Acids Res. 2009;37: 945–956. doi: 10.1093/nar/gkn1015 19106143

5. Snoeck J, Fellay J, Bartha I, Douek DC, Telenti A. Mapping of positive selection sites in the HIV-1 genome in the context of RNA and protein structural constraints. Retrovirology. 2011;8: 87. doi: 10.1186/1742-4690-8-87 22044801

6. Parera M, Fernàndez G, Clotet B, Martínez MA. HIV-1 protease catalytic efficiency effects caused by random single amino acid substitutions. Mol Biol Evol. 2007;24: 382–387. doi: 10.1093/molbev/msl168 17090696

7. Rihn SJ, Hughes J, Wilson SJ, Bieniasz PD. Uneven Genetic Robustness of HIV-1 Integrase. J Virol. 2014;89: 552–567. doi: 10.1128/JVI.02451-14 25339768

8. Haddox HK, Dingens AS, Bloom JD. Experimental Estimation of the Effects of All Amino-Acid Mutations to HIV’s Envelope Protein on Viral Replication in Cell Culture. PLoS Pathog. 2016;12: e1006114. doi: 10.1371/journal.ppat.1006114 27959955

9. Haddox HK, Dingens AS, Hilton SK, Overbaugh J, Bloom JD. Mapping mutational effects along the evolutionary landscape of HIV envelope. Elife. 2018;7. doi: 10.7554/eLife.34420 29590010

10. Hinkley T, Martins J, Chappey C, Haddad M, Stawiski E, Whitcomb JM, et al. A systems analysis of mutational effects in HIV-1 protease and reverse transcriptase. Nat Genet. 2011;43: 487–489. doi: 10.1038/ng.795 21441930

11. Rhee S-Y, Gonzales MJ, Kanor R, Betts BJ, Ravela J, Shafer RW. Human immunodeficiency virus reverse transcriptase and protease sequence database. Nucleic Acids Res. 2003;31: 298–303. doi: 10.1093/nar/gkg100 12520007

12. Ferguson AL, Mann JK, Omarjee S, Ndung’u T, Walker BD, Chakraborty AK. Translating HIV sequences into quantitative fitness landscapes predicts viral vulnerabilities for rational immunogen design. Immunity. 2013;38: 606–617. doi: 10.1016/j.immuni.2012.11.022 23521886

13. Barton JP, Goonetilleke N, Butler TC, Walker BD, McMichael AJ, Chakraborty AK. Relative rate and location of intra-host HIV evolution to evade cellular immunity are predictable. Nat Commun. 2016;7: 11660. doi: 10.1038/ncomms11660 27212475

14. Louie RHY, Kaczorowski KJ, Barton JP, Chakraborty AK, McKay MR. Fitness landscape of the human immunodeficiency virus envelope protein that is targeted by antibodies. Proc Natl Acad Sci U S A. 2018;115: E564–E573. doi: 10.1073/pnas.1717765115 29311326

15. Kiepiela P, Leslie AJ, Honeyborne I, Ramduth D, Thobakgale C, Chetty S, et al. Dominant influence of HLA-B in mediating the potential co-evolution of HIV and HLA. Nature. 2004;432: 769–775. doi: 10.1038/nature03113 15592417

16. Matthews PC, Leslie AJ, Katzourakis A, Crawford H, Payne R, Prendergast A, et al. HLA footprints on human immunodeficiency virus type 1 are associated with interclade polymorphisms and intraclade phylogenetic clustering. J Virol. 2009;83: 4605–4615. doi: 10.1128/JVI.02017-08 19244334

17. Theys K, Feder AF, Gelbart M, Hartl M, Stern A, Pennings PS. Within-patient mutation frequencies reveal fitness costs of CpG dinucleotides and drastic amino acid changes in HIV. PLoS Genet. 2018;14: e1007420. doi: 10.1371/journal.pgen.1007420 29953449

18. Haldane JBS. The Effect of Variation of Fitness. Am Nat. 1937;71: 337–349.

19. Fernandez CS, Stratov I, De Rose R, Walsh K, Dale CJ, Smith MZ, et al. Rapid viral escape at an immunodominant simian-human immunodeficiency virus cytotoxic T-lymphocyte epitope exacts a dramatic fitness cost. J Virol. 2005;79: 5721–5731. doi: 10.1128/JVI.79.9.5721-5731.2005 15827187

20. Asquith B, Edwards CTT, Lipsitch M, McLean AR. Inefficient Cytotoxic T Lymphocyte–Mediated Killing of HIV-1–Infected Cells In Vivo. PLoS Biol. 2006;4: e90. doi: 10.1371/journal.pbio.0040090 16515366

21. Ganusov VV, De Boer RJ. Estimating Costs and Benefits of CTL Escape Mutations in SIV/HIV Infection. PLoS Comput Biol. 2006;2: e24. doi: 10.1371/journal.pcbi.0020024 16604188

22. Ganusov VV, Goonetilleke N, Liu MKP, Ferrari G, Shaw GM, McMichael AJ, et al. Fitness costs and diversity of the cytotoxic T lymphocyte (CTL) response determine the rate of CTL escape during acute and chronic phases of HIV infection. J Virol. 2011;85: 10518–10528. doi: 10.1128/JVI.00655-11 21835793

23. Kessinger TA, Perelson AS, Neher RA. Inferring HIV Escape Rates from Multi-Locus Genotype Data. Front Immunol. 2013;4. doi: 10.3389/fimmu.2013.00252 24027569

24. van Deutekom HWM, Wijnker G, de Boer RJ. The rate of immune escape vanishes when multiple immune responses control an HIV infection. J Immunol. 2013;191: 3277–3286. doi: 10.4049/jimmunol.1300962 23940274

25. Leviyang S, Ganusov VV. Broad CTL Response in Early HIV Infection Drives Multiple Concurrent CTL Escapes. PLoS Comput Biol. 2015;11: e1004492. doi: 10.1371/journal.pcbi.1004492 26506433

26. Hill WG, Robertson A. The effect of linkage on limits to artificial selection. Genet Res. 1966;8: 269. 5980116

27. Illingworth CJR, Mustonen V. Distinguishing driver and passenger mutations in an evolutionary history categorized by interference. Genetics. 2011;189: 989–1000. doi: 10.1534/genetics.111.133975 21900272

28. Garcia V, Regoes RR. The Effect of Interference on the CD8 T Cell Escape Rates in HIV. Front Immunol. 2015;5. doi: 10.3389/fimmu.2014.00661 25628620

29. Garcia V, Feldman MW, Regoes RR. Investigating the Consequences of Interference between Multiple CD8+ T Cell Escape Mutations in Early HIV Infection. PLoS Comput Biol. 2016;12: e1004721. doi: 10.1371/journal.pcbi.1004721 26829720

30. Yang Y, Ganusov VV. Kinetics of HIV-Specific CTL Responses Plays a Minimal Role in Determining HIV Escape Dynamics. Frontiers in Immunology. 2018. doi: 10.3389/fimmu.2018.00140 29472921

31. Raghwani J, Redd AD, Longosz AF, Wu C-H, Serwadda D, Martens C, et al. Evolution of HIV-1 within untreated individuals and at the population scale in Uganda. PLoS Pathog. 2018;14: e1007167. doi: 10.1371/journal.ppat.1007167 30052678

32. Illingworth CJR. Fitness Inference from Short-Read Data: Within-Host Evolution of a Reassortant H5N1 Influenza Virus. Mol Biol Evol. 2015;32: 3012–3026. doi: 10.1093/molbev/msv171 26243288

33. Asquith B, McLean AR. In vivo CD8 T cell control of immunodeficiency virus infection in humans and macaques. Proceedings of the National Academy of Sciences. 2007. pp. 6365–6370. doi: 10.1073/pnas.0700666104 17404226

34. Yang OO, Daar ES, Jamieson BD, Balamurugan A, Smith DM, Pitt JA, et al. Human immunodeficiency virus type 1 clade B superinfection: evidence for differential immune containment of distinct clade B strains. J Virol. 2005;79: 860–868. doi: 10.1128/JVI.79.2.860-868.2005 15613314

35. Watanabe K, Murakoshi H, Tamura Y, Koyanagi M, Chikata T, Gatanaga H, et al. Identification of cross-clade CTL epitopes in HIV-1 clade A/E-infected individuals by using the clade B overlapping peptides. Microbes Infect. 2013;15: 874–886. doi: 10.1016/j.micinf.2013.08.002 23968885

36. Buckheit RW 3rd, Allen TG, Alme A, Salgado M, O’Connell KA, Huculak S, et al. Host factors dictate control of viral replication in two HIV-1 controller/chronic progressor transmission pairs. Nat Commun. 2012;3: 716. doi: 10.1038/ncomms1697 22395607

37. Sipsas NV, Kalams SA, Trocha A, He S, Blattner WA, Walker BD, et al. Identification of type-specific cytotoxic T lymphocyte responses to homologous viral proteins in laboratory workers accidentally infected with HIV-1. J Clin Invest. 1997;99: 752–762. doi: 10.1172/JCI119221 9045880

38. Hoof I, Pérez CL, Buggert M, Gustafsson RKL, Nielsen M, Lund O, et al. Interdisciplinary analysis of HIV-specific CD8+ T cell responses against variant epitopes reveals restricted TCR promiscuity. J Immunol. 2010;184: 5383–5391. doi: 10.4049/jimmunol.0903516 20363973

39. Tang Y, Huang S, Dunkley-Thompson J, Steel-Duncan JC, Ryland EG, St John MA, et al. Correlates of spontaneous viral control among long-term survivors of perinatal HIV-1 infection expressing human leukocyte antigen-B57. AIDS. 2010;24: 1425–1435. doi: 10.1097/QAD.0b013e32833a2b5b 20539088

40. Miura T, Brockman MA, Schneidewind A, Lobritz M, Pereyra F, Rathod A, et al. HLA-B57/B*5801 human immunodeficiency virus type 1 elite controllers select for rare gag variants associated with reduced viral replication capacity and strong cytotoxic T-lymphocyte [corrected] recognition. J Virol. 2009;83: 2743–2755. doi: 10.1128/JVI.02265-08 19116253

41. Chopera DR, Mlotshwa M, Woodman Z, Mlisana K, de Assis Rosa D, Martin DP, et al. Virological and Immunological Factors Associated with HIV-1 Differential Disease Progression in HLA-B*58:01-Positive Individuals. Journal of Virology. 2011. pp. 7070–7080. doi: 10.1128/JVI.02543-10 21613398

42. Koup RA, Roederer M, Lamoreaux L, Fischer J, Novik L, Nason MC, et al. Priming immunization with DNA augments immunogenicity of recombinant adenoviral vectors for both HIV-1 specific antibody and T-cell responses. PLoS One. 2010;5: e9015. doi: 10.1371/journal.pone.0009015 20126394

43. Feeney ME, Tang Y, Pfafferott K, Roosevelt KA, Draenert R, Trocha A, et al. HIV-1 viral escape in infancy followed by emergence of a variant-specific CTL response. J Immunol. 2005;174: 7524–7530. doi: 10.4049/jimmunol.174.12.7524 15944251

44. McKinnon LR, Capina R, Peters H, Mendoza M, Kimani J, Wachihi C, et al. Clade-specific evolution mediated by HLA-B*57/5801 in human immunodeficiency virus type 1 clade A1 p24. J Virol. 2009;83: 12636–12642. doi: 10.1128/JVI.01236-09 19759140

45. Payne RP, Branch S, Kløverpris H, Matthews PC, Koofhethile CK, Strong T, et al. Differential escape patterns within the dominant HLA-B*57:03-restricted HIV Gag epitope reflect distinct clade-specific functional constraints. J Virol. 2014;88: 4668–4678. doi: 10.1128/JVI.03303-13 24501417

46. Du VY, Bansal A, Carlson J, Salazar-Gonzalez JF, Salazar MG, Ladell K, et al. HIV-1-Specific CD8 T Cells Exhibit Limited Cross-Reactivity during Acute Infection. J Immunol. 2016;196: 3276–3286. doi: 10.4049/jimmunol.1502411 26983786

47. Vollbrecht T, Brackmann H, Henrich N, Roeling J, Seybold U, Bogner JR, et al. Impact of changes in antigen level on CD38/PD-1 co-expression on HIV-specific CD8 T cells in chronic, untreated HIV-1 infection. J Med Virol. 2010;82: 358–370. doi: 10.1002/jmv.21723 20087935

48. Zhao S, Zhai S, Zhuang Y, Wang S, Huang D, Kang W, et al. Inter-clade cross-reactivity of HIV-1-specific T cell responses in human immunodeficiency virus type 1 infection in China. Curr HIV Res. 2007;5: 251–259. doi: 10.2174/157016207780076995 17346138

49. Malhotra U, Nolin J, Mullins JI, McElrath MJ. Comprehensive epitope analysis of cross-clade Gag-specific T-cell responses in individuals with early HIV-1 infection in the US epidemic. Vaccine. 2007;25: 381–390. doi: 10.1016/j.vaccine.2006.07.045 17112643

50. Kaul R, Dong T, Plummer FA, Kimani J, Rostron T, Kiama P, et al. CD8(+) lymphocytes respond to different HIV epitopes in seronegative and infected subjects. J Clin Invest. 2001;107: 1303–1310. doi: 10.1172/JCI12433 11375420

51. Aidoo M, Sawadogo S, Bile EC, Yang C, Nkengasong JN, McNicholl JM. Viral, HLA and T cell elements in cross-reactive immune responses to HIV-1 subtype A, CRF01_AE and CRF02_AG vaccine sequence in Ivorian blood donors. Vaccine. 2008. pp. 4830–4839. doi: 10.1016/j.vaccine.2008.06.097 18640166

52. Migueles SA, Laborico AC, Imamichi H, Shupert WL, Royce C, McLaughlin M, et al. The Differential Ability of HLA B*5701 Long-Term Nonprogressors and Progressors To Restrict Human Immunodeficiency Virus Replication Is Not Caused by Loss of Recognition of Autologous Viral gag Sequences. Journal of Virology. 2003. pp. 6889–6898. doi: 10.1128/jvi.77.12.6889-6898.2003 12768008

53. Blattner C, Lee JH, Sliepen K, Derking R, Falkowska E, de la Peña AT, et al. Structural delineation of a quaternary, cleavage-dependent epitope at the gp41-gp120 interface on intact HIV-1 Env trimers. Immunity. 2014;40: 669–680. doi: 10.1016/j.immuni.2014.04.008 24768348

54. Falkowska E, Le KM, Ramos A, Doores KJ, Lee JH, Blattner C, et al. Broadly neutralizing HIV antibodies define a glycan-dependent epitope on the prefusion conformation of gp41 on cleaved envelope trimers. Immunity. 2014;40: 657–668. doi: 10.1016/j.immuni.2014.04.009 24768347

55. Hraber P, Korber B, Wagh K, Giorgi EE, Bhattacharya T, Gnanakaran S, et al. Longitudinal Antigenic Sequences and Sites from Intra-Host Evolution (LASSIE) Identifies Immune-Selected HIV Variants. Viruses. 2015;7: 5443–5475. doi: 10.3390/v7102881 26506369

56. Bricault CA, Yusim K, Seaman MS, Yoon H, Theiler J, Giorgi EE, et al. HIV-1 Neutralizing Antibody Signatures and Application to Epitope-Targeted Vaccine Design. Cell Host Microbe. 2019;26: 296. doi: 10.1016/j.chom.2019.07.016 31415756

57. Huang J, Kang BH, Pancera M, Lee JH, Tong T, Feng Y, et al. Broad and potent HIV-1 neutralization by a human antibody that binds the gp41-gp120 interface. Nature. 2014;515: 138–142. doi: 10.1038/nature13601 25186731

58. Corti D, Langedijk JPM, Hinz A, Seaman MS, Vanzetta F, Fernandez-Rodriguez BM, et al. Analysis of Memory B Cell Responses and Isolation of Novel Monoclonal Antibodies with Neutralizing Breadth from HIV-1-Infected Individuals. PLoS ONE. 2010. p. e8805. doi: 10.1371/journal.pone.0008805 20098712

59. Gnanakaran S, Daniels MG, Bhattacharya T, Lapedes AS, Sethi A, Li M, et al. Genetic signatures in the envelope glycoproteins of HIV-1 that associate with broadly neutralizing antibodies. PLoS Comput Biol. 2010;6: e1000955. doi: 10.1371/journal.pcbi.1000955 20949103

60. Haim H, Strack B, Kassa A, Madani N, Wang L, Courter JR, et al. Contribution of intrinsic reactivity of the HIV-1 envelope glycoproteins to CD4-independent infection and global inhibitor sensitivity. PLoS Pathog. 2011;7: e1002101. doi: 10.1371/journal.ppat.1002101 21731494

61. Brunel FM, Zwick MB, Cardoso RMF, Nelson JD, Wilson IA, Burton DR, et al. Structure-function analysis of the epitope for 4E10, a broadly neutralizing human immunodeficiency virus type 1 antibody. J Virol. 2006;80: 1680–1687. doi: 10.1128/JVI.80.4.1680-1687.2006 16439525

62. Nelson JD, Brunel FM, Jensen R, Crooks ET, Cardoso RMF, Wang M, et al. An affinity-enhanced neutralizing antibody against the membrane-proximal external region of human immunodeficiency virus type 1 gp41 recognizes an epitope between those of 2F5 and 4E10. J Virol. 2007;81: 4033–4043. doi: 10.1128/JVI.02588-06 17287272

63. Chuang G-Y, Chuang GY., Acharya P, Schmidt SD, Yang Y, Louder MK, et al. Residue-Level Prediction of HIV-1 Antibody Epitopes Based on Neutralization of Diverse Viral Strains. Journal of Virology. 2013. pp. 10047–10058. doi: 10.1128/JVI.00984-13 23843642

64. Montero M, Gulzar N, Klaric K-A, Donald JE, Lepik C, Wu S, et al. Neutralizing epitopes in the membrane-proximal external region of HIV-1 gp41 are influenced by the transmembrane domain and the plasma membrane. J Virol. 2012;86: 2930–2941. doi: 10.1128/JVI.06349-11 22238313

65. Williams KL, Cortez V, Dingens AS, Gach JS, Rainwater S, Weis JF, et al. HIV-specific CD4-induced Antibodies Mediate Broad and Potent Antibody-dependent Cellular Cytotoxicity Activity and are Commonly Detected in Plasma from HIV-infected Humans. EBioMedicine. 2015. pp. 1464–1477. doi: 10.1016/j.ebiom.2015.09.001 26629541

66. Cao K, Moormann AM, Lyke KE, Masaberg C, Sumba OP, Doumbo OK, et al. Differentiation between African populations is evidenced by the diversity of alleles and haplotypes of HLA class I loci. Tissue Antigens. 2004. pp. 293–325. doi: 10.1111/j.0001-2815.2004.00192.x 15009803

67. Kijak GH, Walsh AM, Koehler RN, Moqueet N, Eller LA, Eller M, et al. HLA class I allele and haplotype diversity in Ugandans supports the presence of a major east African genetic cluster. Tissue Antigens. 2009. pp. 262–269. doi: 10.1111/j.1399-0039.2008.01192.x 19254258

68. González-Galarza FF, Takeshita LYC, Santos EJM, Kempson F, Maia MHT, da Silva ALS, et al. Allele frequency net 2015 update: new features for HLA epitopes, KIR and disease and HLA adverse drug reaction associations. Nucleic Acids Res. 2015;43: D784–8. doi: 10.1093/nar/gku1166 25414323

69. Walker B, McMichael A. The T-Cell Response to HIV. Cold Spring Harbor Perspectives in Medicine. 2012. pp. a007054–a007054. doi: 10.1101/cshperspect.a007054 23002014

70. Bar KJ, Tsao C-Y, Iyer SS, Decker JM, Yang Y, Bonsignori M, et al. Early Low-Titer Neutralizing Antibodies Impede HIV-1 Replication and Select for Virus Escape. PLoS Pathogens. 2012. p. e1002721. doi: 10.1371/journal.ppat.1002721 22693447

71. Herbeck JT, Nickle DC, Learn GH, Gottlieb GS, Curlin ME, Heath L, et al. Human Immunodeficiency Virus Type 1 env Evolves toward Ancestral States upon Transmission to a New Host. Journal of Virology. 2006. pp. 1637–1644. doi: 10.1128/JVI.80.4.1637-1644.2006 16439520

72. Leslie AJ, Pfafferott KJ, Chetty P, Draenert R, Addo MM, Feeney M, et al. HIV evolution: CTL escape mutation and reversion after transmission. Nature Medicine. 2004. pp. 282–289. doi: 10.1038/nm992 14770175

73. Friedrich TC, Dodds EJ, Yant LJ, Vojnov L, Rudersdorf R, Cullen C, et al. Reversion of CTL escape–variant immunodeficiency viruses in vivo. Nature Medicine. 2004. pp. 275–281. doi: 10.1038/nm998 14966520

74. Delport W, Scheffler K, Seoighe C. Frequent Toggling between Alternative Amino Acids Is Driven by Selection in HIV-1. PLoS Pathogens. 2008. p. e1000242. doi: 10.1371/journal.ppat.1000242 19096508

75. Frost SDW, Wrin T, Smith DM, Pond SLK, Liu Y, Paxinos E, et al. Neutralizing antibody responses drive the evolution of human immunodeficiency virus type 1 envelope during recent HIV infection. Proceedings of the National Academy of Sciences. 2005. pp. 18514–18519. doi: 10.1073/pnas.0504658102 16339909

76. Richman DD, Wrin T, Little SJ, Petropoulos CJ. Rapid evolution of the neutralizing antibody response to HIV type 1 infection. Proceedings of the National Academy of Sciences. 2003. pp. 4144–4149. doi: 10.1073/pnas.0630530100 12644702

77. Zanini F, Puller V, Brodin J, Albert J, Neher RA. mutation rates and the landscape of fitness costs of HIV-1. Virus Evol. 2017;3: vex003. doi: 10.1093/ve/vex003 28458914

78. Rouzine IM, Rodrigo A, Coffin JM. Transition between stochastic evolution and deterministic evolution in the presence of selection: general theory and application to virology. Microbiol Mol Biol Rev. 2001;65: 151–185. doi: 10.1128/MMBR.65.1.151-185.2001 11238990

79. Lythgoe KA, Fraser C. New insights into the evolutionary rate of HIV-1 at the within-host and epidemiological levels. Proceedings of the Royal Society B: Biological Sciences. 2012. pp. 3367–3375. doi: 10.1098/rspb.2012.0595 22593106

80. Wei X, Decker JM, Wang S, Hui H, Kappes JC, Wu X, et al. Antibody neutralization and escape by HIV-1. Nature. 2003;422: 307–312. doi: 10.1038/nature01470 12646921

81. van Gils MJ, Bunnik EM, Burger JA, Jacob Y, Schweighardt B, Wrin T, et al. Rapid escape from preserved cross-reactive neutralizing humoral immunity without loss of viral fitness in HIV-1-infected progressors and long-term nonprogressors. J Virol. 2010;84: 3576–3585. doi: 10.1128/JVI.02622-09 20071586

82. Lynch RM, Rong R, Boliar S, Sethi A, Li B, Mulenga J, et al. The B Cell Response Is Redundant and Highly Focused on V1V2 during Early Subtype C Infection in a Zambian Seroconverter. Journal of Virology. 2011. pp. 905–915. doi: 10.1128/JVI.02006-10 20980495

83. Moore PL, Ranchobe N, Lambson BE, Gray ES, Cave E, Abrahams M-R, et al. Limited neutralizing antibody specificities drive neutralization escape in early HIV-1 subtype C infection. PLoS Pathog. 2009;5: e1000598. doi: 10.1371/journal.ppat.1000598 19763271

84. Lynch RM, Wong P, Tran L, O’Dell S, Nason MC, Li Y, et al. HIV-1 fitness cost associated with escape from the VRC01 class of CD4 binding site neutralizing antibodies. J Virol. 2015;89: 4201–4213. doi: 10.1128/JVI.03608-14 25631091

85. Bunnik EM, Euler Z, Welkers MRA, Boeser-Nunnink BDM, Grijsen ML, Prins JM, et al. Adaptation of HIV-1 envelope gp120 to humoral immunity at a population level. Nat Med. 2010;16: 995–997. doi: 10.1038/nm.2203 20802498

86. Bouvin-Pley M, Morgand M, Moreau A, Jestin P, Simonnet C, Tran L, et al. Evidence for a continuous drift of the HIV-1 species towards higher resistance to neutralizing antibodies over the course of the epidemic. PLoS Pathog. 2013;9: e1003477. doi: 10.1371/journal.ppat.1003477 23853594

87. Bouvin-Pley M, Morgand M, Meyer L, Goujard C, Moreau A, Mouquet H, et al. Drift of the HIV-1 Envelope Glycoprotein gp120 toward Increased Neutralization Resistance over the Course of the Epidemic: a Comprehensive Study Using the Most Potent and Broadly Neutralizing Monoclonal Antibodies. Journal of Virology. 2014. pp. 13910–13917. doi: 10.1128/JVI.02083-14 25231299

88. Rademeyer C, Korber B, Seaman MS, Giorgi EE, Thebus R, Robles A, et al. Features of Recently Transmitted HIV-1 Clade C Viruses that Impact Antibody Recognition: Implications for Active and Passive Immunization. PLoS Pathog. 2016;12: e1005742. doi: 10.1371/journal.ppat.1005742 27434311

89. Stefic K, Bouvin-Pley M, Essat A, Visdeloup C, Moreau A, Goujard C, et al. Sensitivity to Broadly Neutralizing Antibodies of Recently Transmitted HIV-1 Clade CRF02_AG Viruses with a Focus on Evolution over Time. Journal of Virology. 2018. doi: 10.1128/jvi.01492-18 30404804

90. Sobel Leonard A, McClain MT, Smith GJD, Wentworth DE, Halpin RA, Lin X, et al. The effective rate of influenza reassortment is limited during human infection. PLoS Pathog. 2017;13: e1006203. doi: 10.1371/journal.ppat.1006203 28170438

91. Illingworth CJR, Mustonen V. Components of selection in the evolution of the influenza virus: linkage effects beat inherent selection. PLoS Pathog. 2012;8: e1003091. doi: 10.1371/journal.ppat.1003091 23300444

92. Illingworth CJR, Fischer A, Mustonen V. Identifying selection in the within-host evolution of influenza using viral sequence data. PLoS Comput Biol. 2014;10: e1003755. doi: 10.1371/journal.pcbi.1003755 25080215

93. Redd AD, Mullis CE, Serwadda D, Kong X, Martens C, Ricklefs SM, et al. The Rates of HIV Superinfection and Primary HIV Incidence in a General Population in Rakai, Uganda. The Journal of Infectious Diseases. 2012. pp. 267–274. doi: 10.1093/infdis/jis325 22675216

94. Illingworth CJR. SAMFIRE: multi-locus variant calling for time-resolved sequence data. Bioinformatics. 2016;32: 2208–2209. doi: 10.1093/bioinformatics/btw205 27153641

95. Smith JM, Haigh J. The hitch-hiking effect of a favourable gene. Genetical Research. 1974. p. 23. doi: 10.1017/s0016672300014634 4407212

96. Schwarz G. Estimating the Dimension of a Model. Ann Stat. 1978;6: 461–464.

97. Mansky LM. The mutation rate of human immunodeficiency virus type 1 is influenced by the vpr gene. Virology. 1996;222: 391–400. doi: 10.1006/viro.1996.0436 8806523

98. Sanjuan R, Nebot MR, Chirico N, Mansky LM, Belshaw R. Viral Mutation Rates. J Virol. 2010;84: 9733–9748. doi: 10.1128/JVI.00694-10 20660197

99. Neher RA, Leitner T. Recombination rate and selection strength in HIV intra-patient evolution. PLoS Comput Biol. 2010;6: e1000660. doi: 10.1371/journal.pcbi.1000660 20126527

100. Batorsky R, Kearney MF, Palmer SE, Maldarelli F, Rouzine IM, Coffin JM. Estimate of effective recombination rate and average selection coefficient for HIV in chronic infection. Proc Natl Acad Sci U S A. 2011;108: 5661–5666. doi: 10.1073/pnas.1102036108 21436045

101. Markowitz M, Louie M, Hurley A, Sun E, Di Mascio M, Perelson AS, et al. A novel antiviral intervention results in more accurate assessment of human immunodeficiency virus type 1 replication dynamics and T-cell decay in vivo. J Virol. 2003;77: 5037–5038. doi: 10.1128/jvi.77.8.5037-5038.2003 12663814

102. Zhao L, Illingworth CJR. Measurements of intrahost viral diversity require an unbiased diversity metric. Virus Evol. 2019;5: vey041. doi: 10.1093/ve/vey041 30723551

103. Zhao G, Perilla JR, Yufenyuy EL, Meng X, Chen B, Ning J, et al. Mature HIV-1 capsid structure by cryo-electron microscopy and all-atom molecular dynamics. Nature. 2013;497: 643–646. doi: 10.1038/nature12162 23719463

104. de la Peña AT, de la Peña AT, Rantalainen K, Cottrell CA, Allen JD, van Gils MJ, et al. Similarities and differences between native HIV-1 envelope glycoprotein trimers and stabilized soluble trimer mimetics. PLOS Pathogens. 2019. p. e1007920. doi: 10.1371/journal.ppat.1007920 31306470

105. Humphrey W, Dalke A, Schulten K. VMD: Visual molecular dynamics. Journal of Molecular Graphics. 1996. pp. 33–38. doi: 10.1016/0263-7855(96)00018-5 8744570


Článek vyšel v časopise

PLOS Pathogens


2020 Číslo 6
Nejčtenější tento týden
Nejčtenější v tomto čísle
Kurzy

Zvyšte si kvalifikaci online z pohodlí domova

Hypertenze a hypercholesterolémie – synergický efekt léčby
nový kurz
Autoři: prof. MUDr. Hana Rosolová, DrSc.

Multidisciplinární zkušenosti u pacientů s diabetem
Autoři: Prof. MUDr. Martin Haluzík, DrSc., prof. MUDr. Vojtěch Melenovský, CSc., prof. MUDr. Vladimír Tesař, DrSc.

Úloha kombinovaných preparátů v léčbě arteriální hypertenze
Autoři: prof. MUDr. Martin Haluzík, DrSc.

Halitóza
Autoři: MUDr. Ladislav Korábek, CSc., MBA

Terapie roztroušené sklerózy v kostce
Autoři: MUDr. Dominika Šťastná, Ph.D.

Všechny kurzy
Přihlášení
Zapomenuté heslo

Zadejte e-mailovou adresu, se kterou jste vytvářel(a) účet, budou Vám na ni zaslány informace k nastavení nového hesla.

Přihlášení

Nemáte účet?  Registrujte se

#ADS_BOTTOM_SCRIPTS#