Free access
Review
28 March 2018

Kinetic Modeling of Virus Growth in Cells

SUMMARY

When a virus infects a host cell, it hijacks the biosynthetic capacity of the cell to produce virus progeny, a process that may take less than an hour or more than a week. The overall time required for a virus to reproduce depends collectively on the rates of multiple steps in the infection process, including initial binding of the virus particle to the surface of the cell, virus internalization and release of the viral genome within the cell, decoding of the genome to make viral proteins, replication of the genome, assembly of progeny virus particles, and release of these particles into the extracellular environment. For a large number of virus types, much has been learned about the molecular mechanisms and rates of the various steps. However, in only relatively few cases during the last 50 years has an attempt been made—using mathematical modeling—to account for how the different steps contribute to the overall timing and productivity of the infection cycle in a cell. Here we review the initial case studies, which include studies of the one-step growth behavior of viruses that infect bacteria (Qβ, T7, and M13), human immunodeficiency virus, influenza A virus, poliovirus, vesicular stomatitis virus, baculovirus, hepatitis B and C viruses, and herpes simplex virus. Further, we consider how such models enable one to explore how cellular resources are utilized and how antiviral strategies might be designed to resist escape. Finally, we highlight challenges and opportunities at the frontiers of cell-level modeling of virus infections.

INTRODUCTION

A major challenge in biology is to predict how organisms will behave based on how they interact with their environments. This is hard because essential behaviors, such as how organisms reproduce or develop, depend on sensing and responding to diverse environmental factors, often involving the activation and expression of multiple genes as well as coordinated interaction among multiple gene products. To address this challenge, it may help to start small by targeting the simplest organisms, whose growth and development are encoded by the shortest genomes, involving a manageable number of essential genes and interactions.
Why focus on viruses? Many viruses can readily be cultured, facilitating their detailed study, and the relatively simple life cycles of many viruses have for many decades been amenable to molecular-level dissection and characterization. More broadly, viruses can affect the behavior of ecosystems and play significant roles in human health and disease. Studies on viruses that infect bacteria, the bacteriophages, played a key role in seminal discoveries of molecular biology. For example, early studies on phage T4 by Hershey and Chase provided compelling evidence of the role of nucleic acids, not proteins, as the material that encodes genetic information. Later work on phage T4 by Brenner, Jacob, and Meselson led to our mechanistic understanding of protein translation, while Jacob and Monod's studies on phage lambda provided the earliest understanding of how the expression of genes could be regulated (1, 2). The one-step growth method for phages developed by Delbrück and Ellis set a foundation for the first quantitative study of the virus growth cycle (1), and still further work on animal viruses has contributed to the understanding of how eukaryotic cells regulate their growth and how loss of such regulation can lead to cancer (3).
Every genome in nature encodes multiple processes, and genomes of viruses are no exception. In an appropriate environment of a living cell, the release of a genome from an invading virus can take command, directing material and energy resources away from cellular processes and toward the synthesis of components that are essential for virus growth, i.e., viral mRNA, viral proteins, viral genomes, and lipids of viral membranes. Assembly of these and other parts produces progeny virus particles that, upon release by the cell, may then infect other susceptible cells. For many viruses, including viruses that infect microbes, plants, animals, and humans, essential molecular processes of intracellular development have been elucidated. However, despite the relatively short lengths of virus genomes, the networks of reactions that define virus growth and their interactions with their host cells remain complex. These networks can contain multiple positive and negative feedbacks that make it challenging to predict how perturbations to viral or host cellular functions, either by genetic engineering or by the presence of drugs that specifically target viral or cellular functions, will quantitatively influence virus production or cell survival. To begin to systematically address this challenge, one may build mathematical models to account for the essential processes (Fig. 1). Here we review efforts aimed at modeling how viruses reproduce within cells, and we describe how such quantitative models have begun to provide insights into the integrated behavior of virus growth. Further, we show how such models can serve as a basis for the development of new antiviral strategies and shed light on the evolution of viruses.
FIG 1
FIG 1 Modeling integrates diverse data to predict virus growth in cells. Descriptions of the molecular functions encoded by the viral genome are used to write equations that describe how levels of viral mRNA, protein, and genomes change over the course of infection. The equations integrate kinetic and other biochemical and biophysical data as parameter values. They are typically solved computationally by numerical integration to yield predicted concentrations of intermediates and final product (virus) levels over the course of infection.
To set the stage, we use the remainder of the introduction to provide some background on virus genomes and the process of virus growth within cells. Furthermore, we suggest how the integration of knowledge on viruses at the molecular and cellular levels can be enriched by perspectives of physical scientists, mathematicians, and engineers.
Virus genome sizes span several orders of magnitude. The largest known viral genome, which belongs to Pandoravirus (2,770 kb) (4), is larger than the smallest genome of a free-standing bacterium, which belongs to Mycoplasma genitalium (580 kb) (5). The smallest viral genomes are less than 10 kb long, and they include genomes for hepatitis B virus (HBV) (3 kb) and phage Qβ (4 kb). Most known virus genomes fall in the size range of 5 to 100 kb and encode up to several hundred proteins (Fig. 2). The larger virus genomes, such as those of Epstein-Barr virus, smallpox, and Pandoravirus, are composed of DNA. In contrast, many of the smaller virus genomes, such as those of human immunodeficiency virus type 1 (HIV-1), influenza virus, and hepatitis C virus (HCV), are made of RNA. These genomes may be composed of one or more segments of either single-stranded or double-stranded DNA or RNA. Despite these biochemical differences in the genomes of DNA and RNA viruses, all viruses must make mRNA, and all viruses use the host cell translation machinery to synthesize viral proteins. Viruses that carry single-stranded RNA genomes may possess positive- or negative-sense RNA, and a positive-sense RNA genome may immediately serve as an mRNA template for the translation of viral proteins. Figure 3 shows how the genomes of different viruses employ different viral or host polymerase activities to synthesize RNA and, ultimately, protein. Viruses that carry double-stranded DNA (dsDNA) genomes, such as Epstein-Barr virus, use the host cellular RNA polymerase (RNA Pol) for transcription of viral mRNA, while other dsDNA viruses, such as smallpox virus, use their own viral DNA-dependent RNA polymerase. Viruses that carry positive-sense RNA genomes, such as hepatitis C virus and poliovirus, may immediately use their genomes as mRNA templates to direct the synthesis of virus proteins, while viruses that carry negative-sense RNA genomes, such as influenza or Ebola virus, employ a viral RNA-dependent RNA polymerase to make positive-sense RNA, which may then serve as a template for protein synthesis. Double-stranded RNA viruses, such as rotavirus, which can cause diarrhea in children, also employ a viral RNA polymerase to make viral mRNA. Retroviruses, most notably HIV-1, carry two copies of a positive-sense RNA genome, but this RNA is not translated at the outset of infection. Instead, the retroviral genome is reverse transcribed by a viral RNA-dependent DNA polymerase to make viral DNA that is integrated into the host cellular DNA, which later serves as the template to make viral mRNA. A more extensive description of the diverse strategies that viruses use to express their genome-encoded functions and to replicate their genomes is available elsewhere (6).
FIG 2
FIG 2 Most virus genomes encode fewer than 100 proteins. Virus genomes are relatively small, with most being fewer than 100 kb long, encoding about 100 proteins; 1 kb of sequence encodes about one protein. The first genome of any organism to be sequenced was that of phage phiX174 (5.39 kb), completed in 1977.
FIG 3
FIG 3 To replicate, all viruses must make mRNA and protein. Single-stranded positive-sense RNA genomes, designated (+)ssRNA, are of the same sense as mRNA, so they can immediately serve as templates for protein synthesis.
Infection cycles follow a series of basic steps that are shared by many viruses (Fig. 4). These steps generally include (i) binding of the virus particle to the cell surface, (ii) entry of the viral genome into the cytoplasm or nucleus, (iii) transcription of viral mRNA, (iv) translation of viral proteins, (v) replication of viral genomes, (vi) assembly and packaging of virus particles, and (vii) release of progeny virus particles into the extracellular environment. Some viruses, notably HIV-1 and various herpesviruses, can establish a state of latent infection in which their host cells do not actively produce virus progeny but instead turn off most genes while retaining the ability to switch to productive infection in response to certain changes that result in conditions that have been selected to enable viral persistence over a range of resource-rich or -poor host environments. Quantitative and mechanistic models have been developed to study various partial facets of the virus infection cycle, including virus particle adsorption to cells (79), virus entry (1013), transcription (1418), RNA splicing (19, 20), protein translation (2123), genome replication (2426), assembly of virus particles (2729), packaging of virus particles (3032), and release or budding of progeny virus particles from cells (33). We do not review this extensive literature here; instead, our focus is on models that have sought to be more comprehensive, integrating quantitative descriptions of most of the essential steps in the virus infection cycle. Many books provide a useful introduction and serve as references on viruses and their growth. Among the most accessible to the nonspecialist is Sompayrac's introduction to pathogenic viruses (34). A more-detailed reference in which strategies of specific viruses are used to illustrate broader concepts of virology is that by Flint and colleagues (35), and a more comprehensive reference, which provides detailed reviews of specific virus families, is Fields Virology (36).
FIG 4
FIG 4 Each step of virus reproduction can be described by a differential equation. (1) Binding. A free virus particle initially adsorbs to the surface of a living host cell, a process that is usually mediated by proteins on the surface of the virus particle and specific receptor proteins on the surface of the host cell. The equation expresses the binding event as a mass action process that depends on levels of free (unbound) virus and free (unoccupied) receptors. Further, the equation describes how levels of bound virus fall as they enter the cell. Here, for simplicity, we neglect the possible dynamics of the receptor, which can recycle to the surface or be internalized. (2) Entry. The genome of the virus, often accompanied by copackaged viral proteins, is delivered into the host cell, where it gains access to the protein synthesis machinery and other resources of the cell. The equation accounts for the appearance of genomes that are supplied by viral entry as well as the increase in genome level owing to replication and their depletion owing to their decay. (3) Transcription. The virus genome is used as a template to produce different viral mRNA transcripts. Different mRNAs (denoted by the subscripted variable i in the equation) are accounted for by potentially different rates of transcription (promoter strengths), and they depend on genome levels. For positive-sense RNA viruses, the transcription step is not included because the viral genome serves as the template for translation. (4) Translation. The viral mRNAs recruit the protein synthesis machinery of the host cell to produce different viral proteins. (5) Assembly. Virus proteins self-assemble to form aggregates (procapsids). (6) Encapsidation. The packaging of progeny viral genomes into viral procapsids yields intact viral progeny. The equation accounts for the dependence of ith protein synthesis on the level of ith mRNA and also for depletion of free proteins by processes that assemble them into capsids and processes that cause them to decay. (7) Release. Viral progeny are liberated from the host cell, and encounters between the viral progeny and other susceptible host cells initiate further rounds of growth. Concurrent processes (not shown, for simplicity) include a diverse range of cellular responses to infection, such as activation of cellular defenses (restriction or clustered regularly interspaced short palindromic repeat [CRISPR]-Cas responses in bacteria or interferon-mediated innate immune signaling in mammalian host cells), induction of cell suicide (apoptosis) responses, and shutdown of host biosynthetic functions.
For molecular biologists and biophysicists, we show how molecular mechanisms deduced from diverse experiments may be integrated into models that enable an integrated perspective of their behavior. In short, we show how models can be used to explore how individual molecules and their interactions contribute to the overall development of a minimal organism. For the physical scientist, engineer, mathematician, or computer scientist, we highlight a central role for creating mathematical and computable descriptions of the relevant molecular processes as a systematic way to account for their contributions. For the experimental biologist, we emphasize a key role for the development of quantitative experiments (with frequent sampling and repeat measures) that permit estimations of parameters and their variability, an essential process for validation and refinement of the models. For biomedical scientists, we show that such models can enable exploration and identification of potentially robust antiviral strategies or means to maximize yields of virus particles for vaccine or gene therapeutic applications. Finally, for ecological and evolutionary biologists, we make the case that the ability to calculate the growth rate or fitness of an organism can provide a means to probe questions on how genes interact and their effects on the design, robustness, and adaptability of organisms. A summary of the kinetic models of virus growth established over the last 25 years is provided in Table 1.
TABLE 1
TABLE 1 Intracellular kinetic models of virus growth
Class Virus Model(s) available Reference(s)
Binding/entry Transcription Genome replication Translation Particle assembly or release Gene interaction Antiviral Wet-lab experiments
dsDNA Phage T7     6769, 77, 80, 83
  Herpesvirus       18
  Baculovirus             119
  Baculovirus     23
ssDNA Phage M13       37, 38
(+)ssRNA Phage Qβ       47, 52
  Semliki Forest virus         10
  Poliovirus         102, 54
  Hepatitis C virus           128
  General         63
(−)ssRNA Vesicular stomatitis virus   116, 155
  Influenza virus     97
  Influenza virus       94
ssRNA-RT HIV-1       91
  HIV-1     19, 20, 88, 89
  HIV-1         16
dsDNA-RT Hepatitis B virus       125

EARLY MODELS OF VIRUS GROWTH

The earliest models of virus growth focused on phages, the viruses that infect bacteria. Here we present these models in chronological order, starting with phage T4 and moving on to RNA phages, with an emphasis on phage Qβ. We then take on phage T7, which served as a foundation for diverse modeling studies that examined antiviral strategies, mechanistic inference from mRNA-protein expression patterns, effects of genome organization and host physiology on phage fitness, and computational modeling of epistatic interactions. We conclude the section with the model for phage M13, which was recently developed by Fisk and coworkers (37, 38) but is included in this section to maintain continuity with the focus on phage models.

Bacteriophage T4, a Double-Stranded DNA Virus

The first models of intracellular virus replication, which were published by Barricelli and coworkers in 1971, focused on bacteriophage T4, a double-stranded DNA virus (3941). These earliest computer simulations were used to test mechanisms of phage recombination when two or more parent phage strains coinfect the same host cell and create a progeny phage that is a cross of the parent strains. The models used number strings to represent locations and characteristics of experimentally determined markers on the genomes of the different strains (41), and they showed that one could feasibly store such strings in the available memory of an early computer, enable stored strings to interact (recombine), store the resulting progeny strings, and provide progeny genomes as the readout. In short, these works showed that it was possible to simulate recombination between virus genomes by using an early scientific computer (IBM model 7094).
The models were mechanistic to the extent that they accounted for how different genomes might physically interact to create crossover events. They were also mechanistic in the implementation of rules that allowed genomes to amplify during growth and permitted such amplification processes to continue until only a viable number or experimentally observed burst size of phage progeny was obtained (41). However, the models were nonmechanistic in the sense that arbitrary rules were applied for cases where mechanisms were unknown, such as the regulation of DNA synthesis. Experimental data incorporated into these models included genomic locations of markers from classical genetic crosses and mutagenesis experiments compiled from the phage T4 literature.

A Single-Stranded Positive-Sense RNA Virus

In 1975, Knijnenburg and Kreisher initiated development of a computational model to describe the intracellular development of an RNA phage. Their work was motivated by an emerging mechanistic understanding from experiments on the highly coupled roles that the synthesis of viral RNA and viral proteins play in the life cycle of single-stranded RNA bacteriophages MS2 and Qβ (42, 43). Both MS2 and Qβ carry positive-sense RNA genomes which can serve as mRNA templates to synthesize phage proteins or to produce negative-sense RNA templates, which are essential intermediates in phage genome replication. First, Knijnenburg and Kreisher explored, through simulation, how the probability of ribosome binding to sites for translation of the phage coat and RNA polymerase could influence the kinetics of coat and polymerase production during infection. They accounted for the role of coat protein synthesis and accumulation in the subsequent suppression of polymerase synthesis in perhaps the earliest computational study of stochastic gene expression (44). Unfortunately, insufficient details were provided to fully understand how the simulation was implemented. In subsequent work, they expanded their initial model to account for synthesis of essential RNA and proteins of the phage (45). Their efforts were notable because the aim of their model—to better understand how the information content of the viral genome and its interactions with viral proteins contribute to the manufacture of viral progeny—defined a central long-term goal for the field. This work marked the earliest mechanistic modeling of virus replication within cells. Two assumptions were key. First, it was assumed that the kinetics of infection could be examined in a manner independent of the processes of the host cell. This independence or uncoupling was achieved by assuming that all essential cellular resources, including ribosomal subunits, replication factors, tRNAs, amino acids, and nucleoside triphosphates (NTPs), were present in excess throughout the infection cycle. Today, this assumption remains a useful simplifying step to enable one to define an initial set of equations focused on the essential functions encoded by the viral genome. A second key assumption was based on the separation of time scales between fast molecular binding release events and slow processes of template-directed polymerization. In short, the binding release processes were assumed to be sufficiently fast that the participating molecules could be treated as equilibrated with their complexes and characterized by a single equilibrium binding constant. This assumption had the effect of making the dynamics of virus intracellular development be governed primarily by the rates of viral transcription, genome replication, and protein synthesis, which also continues to be a useful starting point for the development of virus simulations.

BACTERIOPHAGE Qβ

Bacteriophage Qβ possesses a single-stranded RNA genome of 4,160 nucleotides that carries four genes. The genome is a positive-sense RNA which can serve directly as a template for translation of phage proteins. The virus-encoded proteins include a maturation/lysis protein (A2), the coat protein, a readthrough protein (A1), and a subunit of the phage RNA-dependent RNA polymerase (replicase). It is notable that two distinct proteins, the phage coat protein and A1 protein, are synthesized from a single cistron. The Qβ replicase normally uses the phage genome and negative-sense antigenomes during infection. It is also able to use a large number of phage-like shorter RNAs (typically 25 to 50 nucleotides long) as templates because they retain secondary structures essential for binding and engagement of the replicase. The Qβ infection cycle within Escherichia coli can produce about 1,000 infectious phage progeny per cell in about 45 min at 37°C, and the intracellular steps have been well characterized (43, 4648).

Kinetics of Qβ RNA Replication

Detailed in vitro kinetic studies of RNA replication by the Qβ replicase, along with analysis and computer simulations, revealed features of the dynamics that established a foundation for a model of phage Qβ growth. With an excess of high-energy monomers (NTPs), the level of RNA increases either exponentially or linearly, depending on the relative level of replicase (49, 50). When replicase concentrations exceed template concentrations, RNA growth depends only on RNA levels, so RNA levels increase exponentially. Alternatively, in the presence of excess templates (or limiting replicase), RNA levels increase linearly, a condition where one may assume a quasi-steady state at the level of the elongating RNA complex (50). These in vitro model systems have been useful for highlighting similarities between replication of short RNA templates and that of full-length phage RNA genomes. For example, formation of double-stranded RNA through the annealing of cRNA templates can deplete both RNAs serving as templates for replication, reducing overall rates of RNA synthesis (25). Further, rate-limiting processes for replication of short-chain RNA species tend to be at the stage of replicase release, requiring hundreds of seconds in vitro, as opposed to in vitro elongation, which occurs in tens of seconds (51). In contrast, for genome-length RNA, the rate-limiting process is elongation (50). For simplicity, these models omit consideration of RNA templates that simultaneously engage more than one replicase molecule, an assumption that may need to be relaxed for genomic templates.

Positive Feedback in RNA Replication during Qβ Growth

A computational model for phage Qβ intracellular growth was published in 1991 (47). In contrast with the earlier in vitro models and experiments on Qβ RNA replication, in which the replicase enzyme was readily available as an initial condition, replicase levels for the phage model were initially zero, as in actual phage Qβ infections. In the phage model, an initial phage genomic template is translated by host ribosomes to produce phage proteins. Further, the phage model accounts for the diverse competing processes that can engage the phage genomic RNA. It may be bound by ribosomes to serve as mRNA to make protein, or it may be bound by the replicase enzyme to serve as a template to make antigenomic RNA. Further, the genomic RNA may be bound by phage coat proteins during the process of phage particle assembly. However, the timing and sequence of these processes are governed by the initial conditions. In the initial absence of phage replicase and coat protein, only translation of the genomic template can take place. The synthesis of replicase coupled with the amplification of the RNA genome that encodes replicase defines an explosive positive-feedback loop that amplifies both replicase and RNA with an initially hyperbolic dependence, as follows: dx/dt = xn, where 1.5 < n < 2 and x represents the intracellular concentration of the replicase or the phage genome. This amplification is more sensitive to protein and RNA levels and more rapid than exponential amplification (n = 1). Because this amplification depends on the availability of ribosomes to make replicase, and because ribosome levels are finite, rapid amplification of RNA eventually reaches a steady state where the RNA level is balanced by the combined levels of ribosomes and replicases. Beyond the equivalence point, RNA levels increase with linear dependence on these levels. This hyperbolic-to-linear growth transition in the phage infection model mirrors the exponential-to-linear growth transition observed during the in vitro amplification of small RNA templates by replicase. Finally, accumulation of coat protein enables the coat to successfully compete against the ribosomes and replicases for binding to the genomic RNA, allowing packaging of the genomes and production of phage progeny. Over the course of the simulation, NTP resources are described as “buffered,” meaning that these building blocks for transcription and RNA replication are assumed to be unchanging. They are assumed to be supplied at the same rate as that at which they are consumed. Ribosome levels, however, are set at a finite level.

Robustness of Simulated Phage Growth with Respect to Parameter Values

The 1991 model of phage Qβ growth by Eigen and coworkers (47) focuses on the information flows associated with transcription, translation, and genome replication. The initial condition for the model is a single-phage genomic template in a cell-sized volume with a finite number of ribosomes and a fixed supply rate of substrate resources for RNA synthesis. Substrate resources for translation, such as pools of amino acid-charged tRNAs, were not specified but appear to be unbounded. The simulations terminate following packaging of genomic templates with coat proteins at a ratio of 1:180. It was noted that the overall sequence of events and behavior “proved in exhaustive trials to be rather insensitive to moderate changes in the rate constant values,” supporting the idea that the processes and interactions do not need to be finely tuned to a narrow set of parameter ranges in order to exhibit phage-like behavior. Eigen et al. found that the rate constants for RNA binding to replicase, ribosomes, and coat protein should be within a 100-fold range of each other and that coat protein binding to RNA should be balanced, i.e., favorable enough to repress translation by competing with ribosomes or replicase for RNA but not so high as to shut down replicase production and RNA synthesis. In this manner, the coat protein plays a pivotal role in mediating early functions of genomic RNA as the template for transcription and replication, with a later purpose as a packaged product of infection. The coupling of reactions during phage growth creates other compensatory effects that contribute to robustness. For example, reducing the rate of ribosome binding to phage RNA can have a detrimental effect on phage protein synthesis, but at the same time it may permit greater access of the replicase to the genomic template, an essential step for synthesis of the antigenome template (52).

Efficient Use of Host Energy Resources

Virus infections consume energy at every step of their intracellular growth. Kim and Yin sought to estimate how biochemical energy was utilized by accounting for its consumption over the course of a simulated phage infection cycle (52). An energy equivalent was defined as the energy released by hydrolysis of a phosphate group from ATP, the common currency in bioenergetics. In this work, Eigen's Qβ model was extended to account for the energy consumed in the synthesis of all phage RNA species (genome, antigenome, and mRNA) as well as the phage proteins. Synthesis of 25,000 phage particles, of which 1 to 10% are typically infectious, requires 3 × 109 energy equivalents, which is about 20% of the level needed to synthesize an E. coli host cell. Of the total energy expended, an estimated 90% is directly invested in the biosynthesis of phage progeny particles. Tenfold changes in parameter values of the model in most cases had relatively little impact on the efficiency of energy expenditures, which were directly related to 2-fold (at most) changes in phage yields. Together these results suggest that the phage has evolved to efficiently utilize available energy resources during growth, and this approach has been extended to other systems, including phage T4, influenza virus (53), and poliovirus (54).
The quantitative analysis of metabolic networks has advanced experimental and computational approaches to link cellular resource utilization with intracellular metabolic fluxes in E. coli (55). These approaches have been extended to include the metabolic demand by infections of a Qβ-like positive-sense RNA phage, MS2 (56), where biosynthetic demands by the phage growth were predicted to substantially increase metabolic activity in the pentose phosphate pathway.

Refinement of Phage Qβ Models

Models of phage Qβ growth, to date, have neglected processes of initial binding of the virus to its host cell, virus entry and release of the phage genome into the cytoplasm of the host cell, and lysis-mediated release of progeny phage from the cell. In particular, the timing of lysis directly affects phage yields and likely reflects an optimization owing to ecological factors (57, 58). While it is known that the phage lysis protein prevents cell wall biosynthesis by inhibiting an essential enzyme in the process (59), how inhibition of cell wall biosynthesis ultimately affects the timing of phage progeny release remains to be elucidated mechanistically. For other aspects of phage growth, refinements in the mechanisms of other processes, such as Qβ RNA replication (26) and particle assembly (60), may also enable refinements or improvements to current models.

Testing Antiviral Strategies

Many current antiviral drugs are directed to inhibit specific viral functions, but viruses can develop mutations that enable them to resist such treatments. One alternative strategy may be to develop defective viruses that parasitize normal virus infections by diverting essential resources for virus growth to the viral parasite (61). Such defective viruses have historically been called defective interfering particles (DIPs). In the context of Qβ infection, we employed our model to computationally explore the potential effectiveness of such a parasitic antiviral strategy (62). We simulated the behavior of minivariant 11 (MNV-11), an 87-nucleotide RNA derived from the Qβ genomic RNA that can compete with the wild-type template for the Qβ replicase. The ability of this molecular parasite to very rapidly amplify itself and compete for Qβ replicase allowed it to prevent amplification of the wild-type phage genome, based on our simulations. Other strategies against positive-strand viruses, based on RNA silencing, have also been explored by simulation and found to yield potentially diverse responses, including inhibition of virus growth, oscillatory behavior, or complete clearance of infection (63).

BACTERIOPHAGE T7

The genome of bacteriophage T7 is a double-stranded DNA of 39,937 nucleotides that carries 56 genes and infects E. coli in a lytic manner, typically producing 100 phage progeny in 40 min at 30°C. The genome enters the host cell in a gradual manner, coupled with transcription of the phage genes, initially by the E. coli RNA polymerase and later by the phage RNA polymerase, which encounters stronger promoters as more of the genome becomes accessible. Owing to the gradual nature of genome entry and transcription, T7 genes are expressed at different times depending on their relative positions in the genome.
An initial motivation of Endy et al. for developing an intracellular model for bacteriophage T7 was to better understand the behavior of phage mutants that were selected during evolution experiments. In phage cultures, wild-type phage T7 was continuously passaged on recombinant hosts that constitutively expressed an essential phage enzyme, the T7 RNA polymerase. During 30 generations, at most, phage mutants arose that carried a deletion in the gene for this essential enzyme, making them completely dependent on the host-supplied enzyme for their growth. Moreover, these mutants grew faster on the recombinant hosts than the initial wild-type phage did (64, 65). Endy et al. speculated that the faster production of mutant progeny could be attributed to the faster synthesis of shorter phage genomes and the ability to dispense with the time required to synthesize the polymerase in wild-type infections. Back-of-the-envelope calculations of the rate of T7 DNA polymerase and the transcription and translation rates needed to produce the enzyme appeared to account for the 10% faster (3 to 4 min) production of mutants than that of wild-type phage (66). Using these calculations as a basis, we assembled a more comprehensive model that accounted for the rate of phage genome entry into the host cell, synthesis of T7 mRNAs, translation of these mRNAs to produce T7 proteins, known protein-protein interactions and feedbacks on host and phage transcription, synthesis of T7 DNA genomes, and assembly of phage progeny (67). Over the course of developing the model, other applications for models of virus intracellular growth emerged. For example, simulations based on published mechanisms and data can be used to examine the consistency of the data and reveal when new results challenge the existing literature. Second, models can readily be extended to predict how different antiviral strategies will influence the simulated growth dynamics. Since the simulations can readily be modified, many potential strategies can be explored efficiently, often suggesting experiments that might not otherwise be performed or helping to prioritize planned experiments.
Several key assumptions defined the framework for the T7 model. Although Endy et al. only later became aware of Knijnenburg and Kreisher's 1983 phage model, they nevertheless employed several of the same assumptions. For example, for the T7 model it was assumed that the protein synthesis machinery of the cell (ribosomes, activated tRNAs, and proteins) was present in excess at all stages of infection, that protein-protein binding interactions were sufficiently fast to be treated as equilibrated, and that the intracellular processes were “well mixed,” neglecting any spatial heterogeneities or roles of component transport in the dynamics of growth. In contrast to Knijnenburg and Kreisher's stochastic kinetics model, however, the phage T7 model was deterministic, as it was based on a set of coupled ordinary differential and algebraic equations. Further assumptions specifically related to details of phage T7. For example, to simulate the assembly of the T7 virion particles, which had not yet been studied in detail, results for phage P22, a dsDNA phage that is morphologically similar to T7, were employed. The final model incorporated 46 parameters spanning 27 years of the phage and E. coli literature. Twenty of the parameters described rates of elongation by RNA polymerases (transcription), DNA polymerase (genome replication), and ribosomes (protein synthesis), spacing requirements for polymerases and ribosomes on their respective templates, decay rates for mRNA, proteins, and DNA, binding constants for protein-protein interactions, and other processes. Moreover, 16 parameters quantified the relative strengths and initiation efficiencies of the T7 transcriptional promoters, and 10 parameters described the stoichiometry requirements of the progeny phage particles. None of the parameters were adjusted to fit the final results. In addition, owing to a lack of mechanistic understanding of the lysis step that allows progeny T7 phage to be released from the cell, we simulated only the production of phage progeny within the cell.
The initial T7 model was able to capture essential features of the T7 growth behavior, based on a comparison with data that were independent of the model formulation. The observed timing of simultaneous shutdown of host RNA polymerase activity and expression of phage RNA polymerase activity early in the infection cycle was well captured by the model. The simulated order of appearance of phage proteins that are characteristically expressed during early, middle, and late stages of the infection cycle captured trends observed from pulse-labeling experiments, though discrepancies were notable, particularly for later structural proteins, which tended to appear earlier in the simulations than in the experiments. Finally, experimental validation of phage intracellular production was performed by lysing cells at different time points following the initiation of infection and quantifying the corresponding intracellular level of phage particles. Simulated production of wild-type T7 phage rose at a rate similar to that for wild-type phage in our experiments, but one-step growth curves appeared about 3 to 5 min earlier than those observed experimentally for a 30-min infection cycle. Simulations of a phage deletion mutant on a recombinant host that expresses the T7 RNA polymerase produced mutant phage earlier than that found in the wild-type phage simulations, consistent with experimental observations.

Antiviral Strategies

We initially employed our model of phage T7 intracellular growth to simulate how different antiviral strategies might influence T7 growth (67, 68). We chose to initially test antisense strategies because such approaches offered a convenient way to think mechanistically and quantitatively about the effects of targeting specific phage functions. For each antiviral strategy, the following three features were defined: the specific T7 mRNA target, the initial level of target-specific antisense RNA in the cell, and the equilibrium binding constant for the formation of the complex between the target mRNA and the antisense RNA. Here the binding constant for complex formation can be viewed as a parameter that describes the intrinsic potency of the antisense drug against its mRNA target. We expanded our model of T7 intracellular development to include antisense drugs that targeted mRNA encoding either the major coat (capsid) protein or the T7 RNA polymerase (RNA Pol). Key results suggest that drugs that target different essential components can produce qualitatively different effects on growth (Fig. 5). RNA Pol transcribes phage proteins that inhibit the function of the host transcription machinery, and this machinery is essential for expression of the RNA Pol. Thus, RNA Pol is part of a negative-feedback loop that ultimately regulates its own expression. While it is challenging to anticipate how a drug that targets a component of an early regulatory loop will ultimately affect the production of virus progeny, modeling offers a useful way to expand our intuition. Our study suggested how drug targeting of regulatory loops could offer benefits by enabling one to block established pathways of drug escape or drug resistance. However, taking a still broader perspective, it is important that novel antiviral strategies should be viewed as simply defining modified criteria for the selection of new viral escape strategies. In other terms, mutations in functions that are not directly related to the targeted function may still alter how virus growth responds to antiviral drug treatment in ways that would be challenging to anticipate. In this context, the model may serve as a foundation to test other escape strategies.
FIG 5
FIG 5 Drug targeting of viral gene regulation. The model of phage T7 intracellular development was expanded to incorporate effects of antisense drugs targeting different virus-specific functions. (A) Drug targeting of mRNA that encodes the major capsid (coat) protein of T7, a major protein component of the progeny phage. As the potency (or affinity) of the drug for this target increased, coat protein production was correspondingly reduced, ultimately inhibiting the production of virus progeny. Viruses that spontaneously generate mutations can attenuate the drug potency (or affinity between the drug and its target) and thereby reduce the inhibitory effects of drug on virus growth, as shown, for example, by the path for virus populations moving from point 1 to point 2. More rapid growth by drug-resistant viruses than by wild-type viruses allows such viruses to become enriched and enables the resulting virus population to escape from the drug. (B) When a drug targets mRNA that encodes the RNA polymerase of T7 (RNA Pol), a different behavior ensues owing to the negative feedback of RNA Pol on its own synthesis. In this case, more potent drugs do not necessarily have greater inhibitory effects on virus growth. Instead, drugs of intermediate potency can have larger inhibitory effects on virus production because they exploit the contributions of the feedback to the overall growth behavior. Mutations that reduce drug potency enable virus populations to move from point 1 to point 2, creating viruses that are more growth inhibited than the wild type. Such mutants would not be expected to become enriched over the wild-type virus. By accounting for the overall effects of drugs on such regulatory loops, the model provides an opportunity to identify potential “evolutionary traps” that select against established mechanisms of drug escape (68).

From Transcriptome to Proteome

By accounting for the synthesis of the phage mRNAs and proteins that are essential intermediates for the production of progeny particles, we simulate as a by-product the concentration-versus-time trajectory of all phage mRNAs and proteins. This information provides an idealized data set with which one may test data mining or mechanistic inference tools. For example, one may consider the simplest way to express the relationship between the concentration of a protein and its corresponding mRNA within the cell, as follows: d[Pi]/dt = ktrans[mRNAi], where [Pi] and [mRNAi] are free concentrations of the ith protein and ith mRNA, respectively, and ktrans is the rate at which the ith mRNA is translated into protein. For a fixed translation rate, plotting d[Pi]/dt versus [mRNAi] yields a line with the slope ktrans and an intercept of zero. Alternatively, one may view the free concentration of the ith protein as a process that integrates the level of the ith mRNA over time.
When the free concentration of a protein can be influenced by other processes, such as protein degradation, posttranslational modifications, or formation of complexes with other proteins or nucleic acids, additional terms appear on the right-hand side, as follows: d[Pi]/dt = ktrans[mRNAi] − kd[Pi] + f(modifications) + g(interactions), where kd is a rate constant for first-order degradation of the ith protein and f and g are functions representing other processes that may influence the concentration of the ith protein. In general, these additional processes will cause trajectories of d[Pi]/dt versus [mRNAi] to deviate from a straight line. In the context of our phage T7 model, we simulated plots of d[Pi]/dt versus [mRNAi] and found that some phage proteins produced loop-like trajectories that reflected their regulatory roles (69). Moreover, the model was used to quantify for each phage protein how it deviated from linearity in its plot of d[Pi]/dt versus [mRNAi] over the course of infection. Such deviant proteins are interesting because they represent proteins that are modified, where f(modifications) is nonzero, or proteins that interact with other proteins, where g(interactions) is nonzero. If two or more proteins interact with each other, then they would be expected to be depleted (and to deviate) in a correlated manner. We calculated extents of such correlation for all phage protein pairs and found that, indeed, the protein pairs that were computationally modeled to form complexes also shared highly correlated deviations. It will be interesting to use highly quantitative measures of mRNA and proteins from the same cells over time to explore whether such a “correlated deviation” approach can be used to infer physical interactions and formation of multiprotein complexes during infection. There is not yet a consensus on how global transcript and protein data should be treated to gain mechanistic insight. While much of the field has focused on characterizing the extent of correlation between protein and mRNA levels (70, 71), alternative approaches that combine mechanistic models of translation with mRNA and protein measurements have begun to highlight challenges in characterizing how limited and changing translation resources influence quantitative links between message and protein (69, 7275).

Genome Organization Affects Virus Fitness

Given that we have a simulation for phage T7 one-step growth, one can begin to explore alternative genome designs by changing the way the simulation is implemented. This may be done by changing the order of genes or the parameters of the model. In the case of phage T7, entry of the genome into the cell is mediated by the action of the host RNA polymerase and later by the phage RNA polymerase, with the processivity of the polymerase corresponding to the entry of phage genes. In short, genes at the left end of the genome are transcribed before genes at the right end. Moreover, because the strength of promoters increases as one moves from the left to the right end, later genes are also expressed at higher levels. By altering the positions of genes within the phage genome, one can alter their timing and level and thereby change the fitness of the phage in subtle or not-so-subtle ways. One of the not-so-subtle ways is to move the T7 RNA polymerase gene to positions where its transcription is put under the control of a T7 promoter, creating a positive-feedback loop of T7 RNA Pol on its own transcription. Predicted enhancements in phage protein synthesis and phage production were not supported by quantitative experiments on protein production by pulse-chase methods or by yields of progeny phage by one-step growth experiments (76). One reason for the discrepancy is that the simulation did not account for the contribution of nonessential genes, such as phage gene 1.7. Although it is known that 1.7 is nonessential for phage growth, control experiments established that the absence of gene 1.7 does have a detrimental effect on phage progeny yields. A better accounting of the finite resources provided by the host cell was not able to explain the differences observed between experiments and simulations (77). The significant observed differences between predicted phage with the alternative gene order and three experimentally generated and tested phages highlight the substantial work that still needs to be done to better understand the nature of the discrepancies.

Synthetic Biology Test of Model Assumptions

Current modeling of virus growth builds on simplifying assumptions that may or may not be valid. Quantitative experiments can be used to test and modify assumptions that may subsequently allow the model to account for more observations or make new predictions. An alternative approach to gaining biological insights is to alter the biology to match the simplifying assumptions of the model. For example, the genome of phage T7 has multiple overlapping genes that are not essential for phage growth. In our model, we assumed that T7 genes were not overlapping, an assumption that enabled us to model the kinetics of transcription of these genes (67). It was unknown at the time how the presence of overlapping genes might affect phage growth. To address these effects, Endy and coworkers redesigned and synthesized 30% of the T7 genome to eliminate overlapping regions. The resulting chimeric phage was able to produce viable phage progeny, though plaque sizes indicated that the growth properties of these variants were attenuated relative to those of the wild type (78). This exercise was useful in showing how simplifying modeling assumptions might be tested by “changing the biology to fit the model” instead of the more common approach of “changing the model to fit the biology.” In the specific case of phage T7, the chimeric phage showed that overlapping genes are not essential for T7 growth but likely have an impact on the overall growth or fitness of the phage.

How the Host Cell Environment Can Affect Virus Fitness

A key assumption that enables development of initial models of virus intracellular growth is that host resources are infinite. Using this assumption, one may then simulate virus growth based on the dynamics of the virus-encoded processes. These minimally include transcription of viral mRNA, translation of viral proteins, synthesis of viral genomes, and assembly of virus progeny particles. The assumption of infinite host resources is most likely to be reasonable at the earliest stages of infection, when the demand by the virus infection is defined for a small number of initial viral genomes (the ones that initiated infection). To better account for the effects of host resources on growth, phage T7 infections were performed on host cells cultured under different conditions that were set up to provide different host resources. In general, rapidly growing cells will have more biosynthetic resources, such as ribosomes for translation, than more slowly growing cells. We cultured E. coli hosts in a chemostat and controlled their growth by adjusting the dilution rate of the chemostat (79). Cells at different growth rates, spanning from 0.7 to 1.7 doublings/h, were used as hosts for synchronized one-step growth cultures of phage T7, and the resulting rise rates and eclipse times (characteristics of one-step growth) were determined. To predict how altered growth rates of host cells would influence phage growth, we employed empirical correlations established by others to indicate how cellular resources and properties, such as host RNA polymerase levels and elongation kinetics, ribosome levels and elongation rates, NTP and amino acid pools, and cell size, correlated with host growth rate. Thus, cellular growth rates set by the experimental conditions were used as inputs to the correlations to estimate cellular resource levels and kinetics, which were used as initial conditions for the T7 simulation. In experiments, faster cell growth reduced the eclipse time (time to production of initial phage progeny) and increased the rise rates of progeny production, and these behaviors were captured by the simulation. The simulation then provided an opportunity to uncouple the effects of different host resource conditions on phage growth in cases where such uncoupling would be difficult or impossible to implement in experiments. Through such studies, we identified the processing of the translation machinery, specifically the ribosome elongation rates, to be most limiting for phage production. Moreover, independent changes in simulated host resources indicated how resources of host transcription or translation would need to change in order for phage infections to become limited or “bottlenecked” by host RNA polymerase levels or ribosome levels. An assumption of the model was that host cell resources for the entire infection process could be defined adequately based on their state at the time of sampling from the chemostats. A more refined perspective will need to account for the consequences of infection on the host physiological state, accounting for the effects of potential changes in energy metabolism on levels of biosynthetic resources or, for example, the synthesis or decay of NTP or amino acid pools during infection.

Epistasis: Quantitative Assessment of Genetic Interactions

A fundamental challenge in quantitative genetics is to better understand how interactions between genes quantitatively affect the fitness of an organism. Simulations of virus growth are potentially useful here because they provide a quantifiable link between functional molecular characteristics of gene products, such as binding affinities or kinetic parameters, and the growth rate, infection productivity, or fitness of the corresponding virus. To help to understand the terminology, it is useful to consider a simple quantitative example of how mutations in two genes, call them A and B, may interact. If the wild type has a fitness of 1.0 and mutations in genes A and B give rise to mutants that have fitness levels of 0.80 and 0.70, respectively, then one may ask, what is the fitness of a mutant phage containing both mutations? If these mutations do not interact, then the double mutant will have a fitness of 0.56, just the product of the fitness values for the single variants. If these deleterious mutations interact synergistically, then the fitness of the double mutant will be <0.56, and if they interact antagonistically or by buffering their deleterious effects, then the fitness of the double mutant will be closer to that of the wild type (>0.56). In quantitative genetics one seeks to answer the following question: on average, to what extent do deleterious mutations interact synergistically, antagonistically, or not at all? This can be a challenging experiment because it requires that one be able to generate mutants having well-defined mutations and to accurately quantify the effects of these mutations on some measure of fitness. In simulations of virus growth, it is feasible to create such mutations by altering parameters that correspond to molecular functions, such as a binding constant for complex formation or a rate constant corresponding to the elongation rate of an RNA polymerase, and then to calculate the effects of such alterations on the yield of virus progeny from a cell, one measure of fitness. The rationale here is that some mutations can alter quantitative characteristics of molecules and that the corresponding parameters in the models can be changed to simulate the effects of such mutations on molecular function. If one seeks to simulate deleterious mutations, then one just needs to check that the alteration in a parameter reduces the calculated fitness. We used this approach to generate and test interactions among diverse computationally generated deleterious mutations to examine the effects on the calculated fitness of phage T7 (80). In this study, two metrics for fitness were employed: one for a resource-rich environment and one for a resource-poor one. For the rich environment, we assumed that host cells were available in unlimited supply, so the fitness of the virus depended on maximizing its production rate within a given cell. For the poor environment, we assumed that only the infected cell's resources were available, so fitness was based on maximizing the yield, without limitations on the rate. It was found that mildly deleterious mutations tended to act synergistically in resource-poor environments but antagonistically in resource-rich environments. Moreover, severely deleterious mutations tended to buffer themselves, acting antagonistically in both rich and poor environments. The work was relevant to population genetics in suggesting that the effects of synergistic interactions on fitness, while important for theory, may be challenging to detect in practice owing to their emergence when the quantitative effects of mutation on fitness are minimal. Future studies would benefit by exploring how quantitative rather than qualitative changes in environmental resources affect whether epistasis is synergistic or antagonistic. This may be achieved, for example, by using the metric for fitness based on production rate (e.g., the rich environment, as described above) and altering the growth rate of the host, creating richer or poorer environments based on higher or lower rates of host cell growth (79). Still more realistic assessments of epistasis might define the average fitness of the virus over multiple cycles of infection where one allows for fluctuations in host resources, conditions that could readily be implemented in virus growth simulations.

Robustness versus Fragility

Biological systems tend to perform robustly with respect to conditions under which they evolved but are more sensitive to environmental changes that they have not encountered (81, 82). These ideas were examined in the context of the bacteriophage T7 system by testing the effects of simulated natural mutations, implemented by changing parameters of the model over plausible ranges for natural mutations, and simulated unnatural mutations, using previously described genomic rearrangements (76, 77). In general, the simulated phage growth was robust with respect to parameter changes but relatively sensitive or fragile with respect to genomic rearrangements, which are not known to occur in nature for phage T7. Such observations were consistent with theory (83). The robust behavior of growth was also a feature of the phage Qβ growth models (47). The effects of natural and unnatural perturbations to the simulated one-step growth of phage T7 were also evaluated in rich and poor host resource environments, and it was interesting that the fitness of wild-type phage was nearly optimal under resource-poor conditions but average under resource-rich conditions (83). This finding suggests that limited host resources served as a constraint on processes of phage T7 evolution. The extent to which such findings are general remains to be tested.

BACTERIOPHAGE M13

Phage M13 is a single-stranded DNA virus with a 6.4-kb genome that encodes 11 proteins, 5 of which combine to encase a single genome in a well-defined filamentous structure. The structure is composed of four minor coat proteins (p3, p6, p7, and p9), each incorporated at 5 copies, and a major coat protein (p8) that is present at about 2,700 copies per phage particle. The well-resolved structures of the proteins and their arrangement in defining the surface of the particle, along with the facile tools for altering the gene for each protein, have in the last 15 years made phage M13 popular for controlled engineering at the nanometer scale, with diverse applications, including biosensors, batteries, and memory devices (84). Prior to these technological developments, fundamental studies established many of the molecular mechanisms associated with M13 DNA replication, mRNA processing, and mRNA degradation (85, 86). Using this and an extensive literature review, a kinetic model for M13 intracellular growth was developed that employed 81 differential equations and 64 kinetic parameters, of which 43 parameters could be estimated from experimental data (37). The model assumed unlimited resources in pools of amino acids, nucleic acids, and energy, but it avoided infinite virus growth owing to limitations on ribosomal availability. If specific data were not available, alternative approaches were found in some cases. For example, the distribution of a limited pool of ribosomes to phage mRNA should depend on the nature of the interaction of each mRNA species with the ribosome, as reflected in part by the ribosome binding sequence (RBS), but strengths of RBS sequences for M13 mRNA have not been measured experimentally. To address this limitation, an RBS calculator was used; the calculator employs an equilibrium statistical thermodynamic model to account for the effects of the RBS sequence on translation initiation rates (87).
The overall M13 intracellular growth model was able to reproduce diverse aspects of the phage's biology, including the timing and extent of phage DNA replication, the levels of mRNA and protein production, and the timing and levels of progeny phage production. Further, by observing how different parameter values could be combined to produce higher or lower simulated progeny levels, the work suggested that changes in rates of phage progeny assembly are tightly linked to levels of phage DNA and protein production. Because the phage does not kill its host cell after a single cycle of phage progeny production, phage production can continue over multiple cell divisions. Extension of the intracellular model to account for multiple cell cycles allowed for testing of the sensitivities of phage processing in establishing a persistently infected state or a cured state (38), conditions that have both been observed experimentally. More specifically, simulations suggested that p5, a protein that binds single-stranded DNA, may be involved in previously undocumented feedback loops between p1, p3, and p8, working at the level of translational attenuation. Ultimately, because these models quantitatively and mechanistically account for production of each phage component (DNA, mRNA, and protein) in the broader production of phage progeny, they provide an opportunity to explore diverse perturbations, such as mechanisms of translational control, that would be experimentally challenging to modify or elucidate independently. Here such simulations suggested that dynamic control of the amount of p5 in the infected cell plays a key role in the allocation of biosynthetic resources during M13 infection.

HIV-1

HIV-1, the virus that causes AIDS, is of great interest in human health owing to the 33 million people worldwide who suffer from AIDS and the 2 million annually who die from the disease. As a retrovirus, it carries two single-stranded positive-sense RNA genomes that serve as templates for the packaged viral reverse transcriptase to synthesize a double-stranded DNA molecule, which integrates into the host genome as part of its infection cycle.
To what extent can the dynamics of HIV-1 growth within its mammalian host cell, a CD4+ T lymphocyte, be explained or accounted for by the kinetics of its underlying processes? In the case of HIV-1, following binding to the host cell and particle entry, the viral genomic RNA is released into the cytoplasm, where it is reverse transcribed to make double-stranded proviral DNA, which is then brought into the nucleus by the viral integrase enzyme and integrated into the genome of the host cell. The proviral DNA remains in a nonproductive or latent state until the cell and viral transcription processes are activated by external factors. The model of Reddy and Yin neglected the latent phase by assuming that the cell was activated, so transcription of viral messages could immediately proceed. The model accounted for splicing of mRNA and its transport from the nucleus to the cytoplasm, translation of viral proteins in the cytoplasm, feedbacks of regulatory proteins Tat and Rev on transcription, transport of viral proteins to the cell membrane, and particle assembly, budding, and maturation (88). Simulated levels of HIV-1 DNA and entering genomic RNA, which are synthesized and degraded during reverse transcription, matched well with experimental observations. Further, the kinetics of production of viral RNA genomes (full-length RNA) and translation of viral proteins was consistent with the sparse available data at the time. In the absence of mechanisms of virus particle assembly, it was assumed that virions assembled instantaneously, with the only constraint being that each virus particle must satisfy the established particle stoichiometry (e.g., 1 genome and 1,200 Gag, 80 Gag-Pol, and 280 Env proteins). This assumption resulted in a simulated production of virus progeny that preceded the observed production by about 6 h, providing an estimate for the time required for the assembly process. The model also enabled study of how perturbations to individual virus functions might influence the overall growth. If growth is particularly sensitive to small changes in a specific parameter or function, then there is a rationale for targeting drugs to that function. The simulation highlights a need to exercise caution in targeting regulatory proteins, such as Rev. When the effects of inhibiting Rev were tested, simulations suggested that doubly spliced transcripts would be enhanced, activating Tat, which would activate viral transcription overall and lead to an enhancement of viral growth. However, simulations exploring the effects of directly inhibiting Tat suggested that such interventions would always have a detrimental effect on virus growth.
More-detailed simulations enabled studies of the effects of transcript splicing on growth. Inefficient splicing of HIV-1 mRNA was generally beneficial for HIV-1 growth, but an extreme reduction in the splicing efficiency could be detrimental, suggesting the existence of a splicing efficiency that optimizes HIV-1 growth (20). When splicing causes an increase in the fraction of either Rev or Tat mRNA relative to that of the other viral mRNA pool, the outcomes are generally beneficial for HIV-1 growth. However, simulations indicated that when mutations cause either Rev or Tat mRNA to be favored over the other, the imbalance is amplified, suggesting that a balance of Rev and Tat is needed in order for HIV-1 to optimize its growth (19). Further, interactions between two feedback loops, the negative feedback of Rev on nuclear export of fully spliced transcripts and the positive feedback of Tat on overall transcription of viral mRNA, create a robust regulatory network that is able to compensate for mutations that might alter functions of components within the network (89). Such interactions between positive- and negative-feedback loops may have more general relevance to the robustness and evolvability of developmental processes in higher organisms (90).
Other approaches to modeling HIV intracellular growth have employed agent-based modeling, where different states of the cell or stages of intracellular infection are initially defined and transitions between states are expressed as rules, with probabilities of transition related to the magnitudes of experimentally determined parameters (91). Such rule-based approaches may enable accounting for compartments within the cell, such as the nucleus or assembly sites, in a manner that explicitly incorporates information about the position, size, or movement of compartments.

An Anti-HIV Strategy

As noted earlier, one may implement antiviral strategies within a model of virus growth by including additional reactions and parameters to simulate the action of an antiviral “drug” on a specific viral target. Moreover, one can test drug escape strategies by simulating how changes in virus parameters, corresponding to virus mutations, may enable virus variants to grow better than the original wild-type virus in the presence of drug. For HIV-1, an RNA interference (RNAi) strategy was proposed to computationally explore factors that would influence targeting of the Tat-mediated positive-feedback loop (92). This example is interesting because of the sequence-specific targeting by RNAi of the trans-acting responsive (TAR) element, a highly conserved RNA structure that is essential for Tat transactivation. One can imagine that mutations that would enable the virus to escape from such RNAi might also have fitness penalties on virus growth. Empirically based rules were applied to account for the effects of specific base changes on the contribution of this structure to the transcriptional feedback, and ultimately the fitness of the virus. When this strategy was carried out in experiments, it was found that no base changes were detected in the TAR element. Instead, mutations occurred in nontargeted regions that enabled indirect upregulation of transcription (93). These results, which would have been improbable to be anticipated by current methods, highlight the limitations that even quite comprehensive models have in anticipating multiple ways that viruses may find to evade antiviral strategies.

INFLUENZA A VIRUS

In 2004, Sidorenko and Reichl developed the first kinetic model for influenza A virus growth in animal cell culture with an aim to identify resource-limiting steps that might be addressed to increase virus yields for vaccine production (94). The genome of influenza A virus has eight segments that encode at least 10 virus proteins, nine of which are incorporated into the virus particle. The kinetic model accounted for steps for viral particle attachment to the cell surface, receptor-mediated entry or endocytosis of the viral particle, release of the viral ribonucleoprotein (vRNP) (RNA-protein complexes) into the cytoplasm and its transport to the nucleus, transcription of viral mRNA and its export to the cytoplasm, translation of viral proteins, replication of viral genomes, formation of viral ribonucleoprotein complexes, and budding and release of progeny virus particles from the cell surface. It was initially assumed that translation resources (ribosomes and precursors for protein synthesis) were present in excess, so levels of these components were not explicitly included in the model. Virus particles were also assumed to assemble with correct segregation of the eight RNA species and protein stoichiometries. Based on experimental observations of cell death 12 h following the initiation of infection, simulations were carried out by integrating the model to 12 h.
The model enabled one to identify potential bottlenecks in virus production and suggested strategies for increasing virus yields for vaccine production. The simulations suggested that M1, the viral matrix protein, becomes a limiting factor in the production of vRNP complexes, which was apparent as levels of M1 initially accumulated and then fell to zero as vRNP complexes were formed. Subsequently, vRNPs become a limiting factor in the budding and release of progeny virus from the infected cell. Moreover, according to the simulation, rates of virus production could be enhanced by increasing the activity of the viral polymerase, which could be achieved in practice by using stronger promoters. Likewise, simulations suggested that increasing the efficiency of translation of viral proteins would also increase the production of virus progeny. Such efficiencies could conceivably be enhanced in practice by using virus mutants that more efficiently inhibited the utilization of cellular translation resources by cellular mRNAs.
Subsequent measurements of two viral proteins (NP and M1) and virus production over the course of an infection cycle provided constraints for a population balance model of influenza virus infection (95). In this model, levels of these essential viral proteins, quantified by flow cytometry, were used to define an internal coordinate for progression of infected cells to production of virus. For an excess of added virus particles to cells (multiplicity of infection [MOI] of 3), the model was able to capture the accumulation of these proteins in infected cells during earlier stages of infection, up to about 10 h, but deviated at later times, reflecting potential limitations in amino acid resources or the onset of virus-induced apoptosis. A more-detailed rendering of the viral replication process did not reduce deviations or provide further insight into underlying mechanisms for deviations (96).
A mathematical model of influenza virus growth in cells was also published by Bazhan et al. (97). Their work describes (in Russian) the regulation of transcription, translation, replication, and assembly of the virus. An analysis of the model indicates a high sensitivity of model behavior to parameters associated with the binding of viral polymerase to virus-specific RNAs, which are essential processes for transcription and genome replication. Such essential processes might be effective targets for the development of potent antiviral drugs.

Control of Viral RNA Synthesis

In a reformulation of the Sidorenko and Reichl model, the set of original kinetic equations was reduced and further data sets from the literature were incorporated to estimate key parameters (98). In addition, mechanisms of RNA stabilization and nuclear export of RNA species were incorporated to better resolve the dynamics of viral RNA transcription and genome replication. Constraints on the model included experimental measures of virus entry, i.e., absolute and relative levels of viral messenger, replicative intermediates, and genomic RNA per cell, as well as average levels of viral progeny released per cell. The resulting model provided support for early regulation of genome replication by stabilization of viral RNA replicative intermediates. It also suggested how the viral matrix protein 1 (M1), which normally mediates export of viral genome copies from the nucleus, also might control viral RNA levels in the late phase of infection (98). Finally, the model predicted an intracellular accumulation of viral proteins and RNA toward the end of infection, providing evidence that transport processes or particle budding limits the process of virus progeny release.

Kinetics of Defective Interfering Particles

It has long been known that influenza A virus infections can produce defective virus particles that can interfere with the production of infectious virus (von Magnus phenomenon). Such particles carry deletions in one or more essential genes needed for growth, making them unable to reproduce alone. However, during coinfection with infectious virus, defective genomes divert resources to their own replication and packaging, which interferes with the production of infectious particles. Such defective interfering particles (DIPs), described above in “Testing Antiviral Strategies,” have been found in clinical isolates of influenza virus (99) and can negatively affect the manufacture of live vaccines (100). To better understand mechanistically how DIPs reproduce, the intracellular kinetic model for influenza virus was extended to include defective interfering RNAs that replicate more rapidly than full-length RNAs owing to their reduced length (101). The extended model was able to account for observed effects of DIPs on infectious virus production. Moreover, the model suggests that DIPs that specifically carry deletions in RNA segments encoding the viral polymerase can become enriched, in agreement with experimental observations. Further, the model and experimental observations suggest that other mechanisms, such as competition for viral proteins (polymerase and nucleoprotein), can also contribute to interference and DIP enrichment.

POLIOVIRUS

An interest in better understanding viral evolutionary mechanisms motivated the development of models that couple intracellular virus growth and release with extracellular spread and infection of other susceptible cells. In 2003, Krakauer and Komarova explored how selection pressures acting on virus growth within cells differed from selection pressures acting at the population level, where viral persistence depends on replication as well as transmission between susceptible host cells (102). They chose to develop a model for poliovirus owing to its relatively simple regulation, where a single positive-sense RNA genome serves as mRNA. Further, they assumed that viral fitness could be defined directly by the growth kinetics of the virus within its host cell. This model for poliovirus intracellular growth accounted for the following three main steps: first, translation of the entering viral RNA genome, producing viral proteins that include the viral RNA polymerase; second, synthesis of the positive-sense RNA genome template by employing a negative-sense RNA intermediate; and third, encapsidation of the RNA genome by viral proteins for release from the infected cell. To facilitate the mathematical analysis, it was assumed that the system could establish a steady state (or dynamic equilibrium) within the cell, where rates of production and decay of viral intermediates are balanced, resulting in constant levels of such intermediates in the cell. Further, the analysis allowed for an extracellular level of infection spread where virus released from infected cells could infect other susceptible cells. By considering only the extracellular level, higher rates of encapsidation appeared to be favored by evolution owing to higher rates of production of viral progeny. However, by also considering the intracellular level, it was found that the productive equilibrium could be established only for a bounded rate of encapsidation. Beyond a particular threshold, a productive intracellular steady state would not exist because genomic templates needed for viral replication would be removed too rapidly by encapsidation.
A key point of this work is its accounting for ways that population-level selection (at the level of cell and virus population interactions) can influence within-cell kinetics of virus production. While the assumption that steady states can arise at both the intracellular and extracellular levels enabled analysis of the model, the validity of such assumptions should be validated by experiments. Nevertheless, the work importantly shows how selection acting at multiple levels may have outcomes different from those for selection at a single level.

Optimal Resource Use within Cells

A model of poliovirus growth focusing on intracellular processes was developed to explore how principles of evolutionary ecology could be extended to virus growth (54). Life-history theory aims to reveal how different quantifiable traits of an organism must be balanced across its life span, from birth to death, in order to optimize some measure or correlate of its fitness (103). In the case of humans, material or energy resources that are spent on reproduction may not be available for nurturing or protecting offspring. How should such resources then best be allocated? In the context of a growing virus, this study of poliovirus sought to quantify how potentially limited resources of the infected cell should best be allocated over the course of the virus life cycle in the cell. The model incorporated the synthesis of viral RNA and proteins, accounting for delays of elongation by polymerase and ribosomes, respectively, as well as the time required for queuing multiple polymerases or ribosomes onto the respective template. By using the overall rate of production of viral genomes as a measure of virus growth, the model suggested that optimal virus growth would arise from an imbalance in the production of genomic and antigenomic templates, with about 40 genomic RNA molecules synthesized from each antigenomic RNA and about two antigenomic RNA molecules made from each genomic RNA, in reasonable agreement with experimental observations. It is possible that other measures of productivity, such as the total yield of viral genomes rather than their production rate, would yield similar results. The model was extended to account for potential resource limitations at the levels of ribosomes, amino acids, and nucleotides. The analysis highlighted tradeoffs between time spent on translation and that spent on transcription, because RNA synthesis is driven by products of translation and use of an RNA template for protein synthesis must be completed before it can be used further as a template for RNA synthesis. In addition, it was found that under conditions of limiting translation resources, optimal production would favor higher ratios of genomic to antigenomic RNA, while the inverse would be true under conditions of limiting transcription resources. The magnitude of asymmetries in template replication for poliovirus in this analysis may also apply for observed asymmetries in vesicular stomatitis virus (VSV), with its negative-sense genome. In future work, it would be interesting to see whether the observed trends hold in light of the costly energetics of protein synthesis relative to the less costly energetics of RNA synthesis. More broadly, while the analysis was built on assumptions that remain to be substantiated, it is useful in showing that optimization of virus growth under different resource limitations can potentially shift how resources are allocated.

VSV

VSV is an RNA virus that possesses a single negative-strand RNA with a genome length of 11 kb, carrying five genes. Other negative-sense single-stranded RNA viruses include Ebola, measles, and rabies viruses. Historically, VSV has served as a model virus for understanding gene regulation in negative-strand RNA viruses (104, 105), and experimentally, the low fidelity of its RNA-dependent RNA polymerase has enabled laboratory studies of virus evolution and adaptation (106108). In the biomedical arena, recombinant VSV may express engineered surface proteins of influenza A virus or HIV as a potential vaccine strategy. VSV also has the ability to discriminate and productively infect cancer cells while leaving healthy cells intact (109), opening opportunities for its use as an anticancer therapeutic (110112). Live viruses have advantages as vaccines over subunit or killed virus because they tend to elicit more potent immune responses. However, vaccination with live viruses also carries greater risk because the processes for attenuation of their growth and pathogenic properties have historically relied on serially passaged cultures that weaken growth by poorly understood and potentially reversible mechanisms. As one learns more about the detailed molecular mechanisms that influence virus growth, one may be in the position to engineer mutants that rationally attenuate growth, as described below.

Effects of Genome Organization on Virus Growth

One approach for attenuating growth in VSV has been to change the linear order of genes in its genome. The order is important because the genome (3′-N-P-M-G-L-5′) has only a single promoter for its polymerase, near its 3′ end, with attenuation sequences between genes that cause a fraction of the passing polymerase to leave the template (113). As a result, mRNA levels follow a gradient (N > P > M > G > L), with the highest expression occurring for the N gene (nucleocapsid), directly adjacent to the 3′ end, and the lowest expression for the L gene (large protein of the polymerase), adjacent to the 5′ end. In order for the virus to grow, all five proteins must be expressed and available for incorporation into progeny virus particles. One might well expect the rate of progeny production, which is one measure of virus fitness, to be optimized in some way. For a virus, optimal growth may be obtained by using available host resources to maximize the production of virus progeny (yield) or the rate at which progeny are created. In the case of VSV, high levels of nucleocapsid protein (N) are needed to bring about a switch in viral RNA synthesis from transcription to genome replication, so it is plausible that the timing and level of N protein are optimized in the wild-type virus. Moreover, altering the timing and level of N protein production from its wild-type expression pattern might well move it away from optimal growth and detrimentally influence the virus growth or fitness. To test this idea, positional mutants of VSV were created, with the N gene in the N2, N3, and N4 mutants positioned at progressively greater distances from the 3′ promoter (114). Yields of virus from infected cells were correspondingly perturbed, with the highest yields from N1 (wild type) and the lowest yields from N4. Expanding these gene-order variants to consider all permutations of five genes would define 5! or 120 possible versions. Additional mutants that retained the N1 and L5 positions of the wild type but accounted for all six perturbations of the internal three genes, P, M, and G, were also created and characterized with respect to gene expression and growth (115). We sought to better understand how gene order influences virus fitness by developing a kinetic model for VSV intracellular growth (116). Our model accounted for diverse features of VSV, including transcriptional regulation by intergenic attenuation, the switch in viral polymerase activity from transcription to genome replication, effects of different promoter strengths on the balance of full-length genomic and antigenomic RNA templates, diversion of the cellular translation machinery to synthesis of viral proteins, and the stoichiometry of RNA and protein components in VSV progeny particles. The model correctly predicted the observed experimental ranking of N gene mutants (N1 > N2 > N3 > N4) and suggested that optimal growth depends on a balance of promoter strengths for genome and antigenome synthesis. Further, we used the model to explore how the expression of other genes, such as the antigenic surface glycoprotein (encoded by the G gene), may entail tradeoffs with virus fitness. More recently, the model was extended to predict the growth of all 120 gene-order permutations, offering a means to study how fitness might depend on genome organization (117). The simulated virus fitness was found to be most sensitive to permutations that changed levels of the L and N genes, which are the least and most expressed genes of the wild-type virus, respectively. The roles of gene order and transcriptional regulation were also probed by use of this model. By computationally deleting the intergenic attenuations that regulate how transcriptional resources are distributed, one could test how the 120 gene-order variants changed. Their growth yields or fitness levels were drastically narrowed, from 6,000- to 20-fold, and many variants produced higher progeny yields than those of the wild type. Among the gene-order mutants, the wild type emerged as a fitness winner only in the presence of intergenic attenuation, suggesting that in the natural evolution of VSV this mode of regulation preceded or coevolved with the fixation of the wild-type gene order.

BACULOVIRUS

The large-scale production of recombinant proteins, particularly proteins requiring posttranslational modification, has long been implemented in insect cells infected by a recombinant baculovirus engineered to express heterologous proteins of interest (118). Baculoviruses are double-stranded DNA viruses with genomes of 80 to 180 kb that typically encode about 150 proteins. An early study highlighted two factors, the time of infection and the multiplicity of infection (MOI), for defining key tradeoffs in the production of virus and heterologous protein (119). Infection of cells during the late exponential growth phase, before they have reached their culture capacity (maximum cell concentration), can result in lower total yields of virus. However, infection of cells late, as they approach their culture capacity, can also limit virus production owing to lower biosynthetic capabilities of cells as their growth slows. If the MOI is well below 1, then cells that are not initially infected may continue to grow and become infected when the first generation of virus is released, contributing to overall higher productivity of the culture. In such scenarios, the initial virus production and release need to occur before cell growth can progress to stationary phase, when productivity drops. For these studies, the kinetics of virus production was not considered mechanistically. Instead, the focus was on three kinetic milestones: the time postinfection for extracellular virion synthesis, the time postinfection for extracellular protein (recombinant product) synthesis, and the time postinfection for cell lysis. The importance of culture time and MOI on the dynamics of cell, infected-cell, and virus populations were subsequently validated experimentally in a study that extended application of the baculovirus expression system for the production of virus-like particles (VLPs) (120). Virus-like particles are often highly immunogenic, making them potentially useful as vaccines. Infected cells were immunostained using an antibody against a key protein of the VLPs. Further, it was shown that a higher MOI could increase the rate of VLP production, reducing the time to harvest.
The baculovirus expression system has been harnessed for the production of VLPs of rotavirus, a common cause of severe diarrhea in children. In this application, three rotaviral proteins were coexpressed, and their measured viral DNA, mRNA, and protein levels were found to be consistent with models of baculovirus transcription and translation (23). As a notable aside, it was found that experimental uncertainty associated with estimating the MOI propagated exponentially in the calculation of viral DNA templates. To minimize these effects, experimental measures of early processes (e.g., binding and entry) were not used in the estimation of parameters. Instead, parameters for viral DNA replication, transcription, and translation were estimated from levels of viral DNA, mRNA, and protein measured at least 5 h after the start of infection.
Recent studies on baculovirus production have returned to the classic problem of producing high virus yields when host insect cells are at high density. A useful approach has been to explore how cell physiological changes, particularly changes that are coupled with the transition from exponential-phase cell growth to stationary-phase cell behavior, might limit the availability of resources that are essential for virus production. Global metabolic changes associated with such transitions can be estimated by combining measurements of metabolite changes with known metabolic networks and their analysis. Methods for analysis of metabolic fluxes have become increasingly standardized (121). Application of these approaches to Sf9 insect cells at high density or after infection by baculovirus suggested a depletion of intermediates within the tricarboxylic acid (TCA) cycle (122). More specifically, carbon fluxes through glycolysis and the TCA cycle were found to decrease as cell density increased, causing sharp drops in ATP production and availability of metabolic energy. Virus infections were found to have similar effects on cell metabolism, highlighting the depletion of ATP, the central currency for metabolic energy, as a key factor linking high cell density with drops in virus production. Further analysis supporting this result indicated that the drop in productivity of viruses at high cell density could not be attributed to the depletion of essential nutrients or the accumulation of inhibitory by-products. To address limitations of metabolic energy on virus production, the cell culture medium was supplemented with key depleted intermediates. Specifically, addition of pyruvate or α-ketoglutarate at the time of infection resulted in higher virus yields (up to 7-fold) during high-cell-density culture (123). In this case, metabolic flux analysis showed a strong correlation between the net rate of ATP formation and the generation of redox equivalents in the form of NADH.

HBV

HBV infection is a major cause of acute and chronic liver disease. Over 350 million people are chronically infected with HBV, and more than 150,000 people die annually of liver disease related to hepatitis B (124). The virus may remain relatively unnoticed for years, until the population of viral progeny in the patient's liver “explodes.” HBV is a difficult virus to treat, in part because of its high mutation rate. An intracellular kinetic model was developed to understand how mutation-driven transitions from chronic to acute infection might work mechanistically (125). The model accounted for the kinetics of 10 DNA, RNA, and protein components, including a pregenome-polymerase complex, in the virus and progeny viruses, and their mechanisms of production and interaction defined 18 parameters. Plausible order-of-magnitude parameter values were chosen to give simulated levels of virus progeny that approximated experimental observations; however, no parameters were estimated by independent wet-lab experiments. The model showed that patterns of gene expression that favored packaging of core particles to form progeny arrested overall replication, while patterns that favored autocatalytic amplification of core particles resulted in explosive replication. It was argued that mutations in the core or precore region of the HBV genome could bring about switching from arrested to explosive replication, corresponding to the transition from chronic to acute infection. An extension of the model explored how recycling of released virus progeny back into an infected host cell could serve as an additional source of intracellular core particles, effectively lowering the threshold for transition to explosive replication (126). To test these suggested mechanisms for switching between replication modes, experimental validation of the model parameters will be an important next step.

HCV

HCV is an enveloped positive-strand RNA virus that can cause chronic liver disease, which can lead to cirrhosis (scarring and impaired function of the liver) and liver cancer. About 130 million individuals worldwide are chronically infected with HCV, and no vaccine against HCV infection currently exists. HCV has been very challenging to study experimentally, owing in part to the lack of a laboratory system for culturing the virus. A replicon system has enabled the study of HCV genome replication in a specific liver cancer cell line (127) and served as the basis for an initial kinetic model of HCV replication (128). With guidance by the structure of the phage Qβ model, this HCV replicon model accounted for reactions that synthesize genomic and antigenomic RNA strands, a ribosome-genome complex to synthesize the viral polyprotein, cleavage of the polyprotein to produce the viral polymerase, and intermediates in the replication process formed by the polymerase and genomic and antigenomic templates. The model also allowed for spatial compartmentalization, with translation occurring in the cytoplasm, replication occurring within a vesicular membrane structure (VMS), and the viral genomic template and polymerase able to move between the cytoplasm and the VMS. Like the models for other positive-sense RNA viruses, e.g., phage Qβ and poliovirus, an imbalance in the rates of HCV replication favors production of HCV genomes over antigenomes, at a 10-to-1 ratio. Further, the model was used to explore the role of replication compartmentalization in the VMS. Specifically, could observed steady-state levels of HCV RNA be attained without the VMS? This hypothetical question was addressed by simplifying the model to allow both protein translation and genome replication to occur in the cytoplasm. The single-compartment model showed that one could attain steady-state RNA levels that were consistent with experimental results. However, the corresponding predicted levels of the viral polymerase (NS5B) were 4 orders of magnitude lower than observed levels. To address this discrepancy, ribosome levels were increased, but this adjustment then caused the steady-state RNA levels to move significantly out of the observed range. In short, this analysis indicated that the VMS may plausibly serve to restrain viral amplification and perhaps limit associated host cell damage.
To explore antiviral strategies against HCV, Mishchenko and colleagues developed an HCV replicon model which included reactions to target specific HCV functions (129). The model was based on much of the biology of the Dahari model (128) but also included inhibitors of the HCV protease (NS3), the HCV polymerase (NS5B), and a host protein (hVAP-33) that is essential for assembly of the HCV replication complex. By simulating the effects of drugs of different potencies on the steady-state level of HCV genomic RNA, this model showed that direct targeting of the polymerase or the host factor would have a greater inhibitory effect on viral replication than that of targeting the HCV protease. Moreover, testing of combined treatments that target both the protease and the polymerase showed no enhancement of inhibition, at least for weak inhibitors of these functions.
Taking a similar approach to his modeling of hepatitis B virus, Nakabayashi developed an intracellular kinetic model for HCV. As with HBV, he found two major patterns of replication, one arrested and one explosive, depending on the distribution of replication resources (130).

HSV-1

The global prevalence of adult carriers of herpes simplex virus type 1 (HSV-1) was estimated to be 67% in 2012 (131), reflecting the stability and efficient transmission of a virus with major impacts on public health. HSV-1, which encodes at least 84 proteins in productively infected cells (132), expresses its genes based on their timing, which can be divided into three classes: immediate early, early, and late (133). To better understand the role of regulatory feedbacks on the temporal order of gene expression, genome replication, and protein synthesis in the production of virus progeny, Nakabayashi and Sasaki developed an intracellular kinetic model (18). An initial version of the model that accounted for viral DNA genomes, mRNAs from each of the three classes of expression, their translation to produce their corresponding proteins, viral assembly, and degradation rates of all nucleic acid, protein, and virus species was simplified by lumping processes of transcription and translation and neglecting rates of species degradation. Analysis of the simplified model revealed two modes of growth, either explosive or arrested, whose characteristics were similarly investigated by Nakabayashi et al. in the models of hepatitis B and hepatitis C. In the case of HSV-1, the explosive or arrested growth behavior depended on the relative expression of early versus late gene products. Higher expression levels of early gene products supported genome replication and a positive-feedback loop that explosively amplified viral genomes, while greater expression of late gene products yielded more envelope and structural proteins that depleted free genomes by packaging them into virus progeny particles. Under conditions for explosive growth, where genome replication is favored over genome packaging, one could estimate a waiting time for appearance of the explosive growth, which depended inversely on the initial level (or dose) of viral genomes. In short, higher initial levels of viral infection could result in shorter waiting times for explosive growth. The work further considered scenarios in which a virus with arrested growth behavior could accumulate mutations in early or late gene promoters that shift the balance of gene expression in favor of explosive growth, effectively suggesting how diverse wait times for explosive growth may arise. Although the model did not explicitly address issues or mechanisms of viral latency—when the viral genome is maintained in a nonreplicative state but poised to transition into a lytic or productive growth state—it is plausible that analysis of the transition from arrested to explosive growth may offer insights into the transition from latency to lytic growth. Finally, the simplified model was extended to account for potential limited intracellular resources needed for synthesis of viral DNA, RNA, and proteins. Such models, combined with advances in our mechanistic understanding of viral genes, may help to elucidate how cellular resources are distributed to viral functions over the course of infection (134).

FRONTIERS FOR MODELING VIRUS GROWTH IN CELLS

In this final section, we suggest frontiers for modeling virus growth, touching on opportunities and challenges that delve into extending model building to other viruses, the environment of the host cell and its interactions with viral processes, modeling the variability of single-virus or single-cell behaviors, and the multiscale nature of virus infections as they propagate over multiple cycles.

Other Viruses

Many viruses that have been studied in great depth at the molecular level, as featured, for example, in Fields Virology, define opportunities for development of new kinetic models of virus growth in cells. Studies of more recently emerging viruses, for which less may be known, such as Ebola virus or Zika virus, may still benefit from existing models. Just as the model for phage Qβ, a single-stranded positive-sense RNA phage, served as a scaffold to build the initial replicon model for HCV, another single-stranded positive-sense RNA virus (128), other existing models may serve as scaffolds for related viruses. Ebola virus is a single-stranded negative-sense RNA virus that likely shares many similarities with VSV, for which a model is available (116). In addition, Zika virus, like HCV, is a single-stranded positive-sense RNA virus that encodes a single polyprotein, which subsequently self-cleaves to supply essential replication and structural proteins for virus growth. Rates or other parameters that quantify the magnitude of molecular interactions or reactions will always be valuable, if not essential, for model building. Databases for modeling, such as the “B10 NUMB3R5” effort (135), are useful for providing estimates of quantitative parameters of cellular functions essential for virus growth, such as protein synthesis. We hope that interest in modeling virus growth will motivate the growth of databases for virus-specific parameters, which individually will inform specific virus models but collectively may provide insight into the diversity of viral kinetics.

Host Cell Physiology and Innate Immune Responses

The availability of biosynthetic resources from a host cell is essential for virus growth, so a better understanding of how the cell's capacity for synthesis of proteins, nucleic acids, lipids, and other components of virus particles changes under different environmental conditions will need to be incorporated into models that seek to account for the dependence of virus growth on host physiology. The effects of host cellular growth rate on the intracellular kinetics of phage growth, which were probed by use of empirical correlations between cell growth and biosynthetic resources (79), have been expanded to incorporate genome-scale metabolic models of the bacterial host for the effects on phage growth (56, 136). Moreover, models of metabolic fluxes developed for mammalian cell cultures have been applied to identify conditions for higher glucose uptake rates and more efficient use of ATP, suitable for bioprocessing of influenza A virus vaccines (137, 138), and similar approaches have been applied to baculovirus growth in insect cell cultures (122).
In addition to supplying biosynthetic resources for virus reproduction, host cells also activate defensive innate immune signaling and responses, typically mediated by type I interferons (IFNs). Such signaling can trigger both autocrine and paracrine cellular responses that shut down protein synthesis that is essential for expression of essential viral functions. Different facets of such responses have been quantified and modeled mechanistically, including initial sensing of the viral dsRNA (139), induction of IFNs that activate both positive and negative feedbacks (140), and responses to different degrees of viral antagonism (141). Notably, in the case of infections by the highly pathogenic Nipah virus, modeling of the host response along with transcription-level measures of IFN and inflammatory cytokine activation provided insight by suggesting that the virus may delay suppression of inflammation and thereby enhance vascular permeability and infection spread (142).

Single-Cell and Single-Virus Tracking

To obtain a measure of viral or cellular function, most biochemical or molecular biological studies work with millions of cells in a tube or dish, for which average levels of nucleic acid, protein, intermediates, and infectious virus particles can readily be detected and quantified. However, in nature, infectious diseases are often transmitted or spread by countable numbers of infectious virus particles or cells. When the behavior of individual cells was studied in response to viral infection, the single cells exhibited infectious particle yields (or burst sizes) that were highly variable (143, 144). This variability among different infected cells can be attributed to genetic heterogeneity of the stock virus population (145) or host cells (146), resource differences linked to the stage of the cell cycle (147), or intrinsic stochastic behavior of reactions that involve small numbers of molecules (148150). To account for the stochastic behavior, and assuming a well-mixed environment, Gillespie developed an algorithm for exact stochastic simulation of the reaction kinetics (151), which broadly enabled modeling of the lysis-lysogeny decision in phage lambda (152), stochastic versus deterministic intracellular infection (153), viral capsid assembly (154), bifurcation of gene expression in HIV-1 (16), transcriptional delays and autocatalytic feedbacks in VSV (155), genome amplification for influenza A virus (156), and infection-mediated activation of innate immune signaling (157, 158).
Models of virus intracellular kinetics have often applied the following assumption without explicit mention: intracellular environments are spatially homogeneous or well mixed. This simplifying assumption allows concentrations of reacting components to be treated as solely time dependent rather than time and position dependent, so mathematically the equations are written as sets of ordinary differential equations, which are numerically solved more readily than sets of partial differential equations. Future kinetic models will have spatial and temporal dependencies, incorporating advances in single-particle tracking by use of dye-labeled or reporter-expressing virus particles which have enabled direct visualization and quantification of virus-receptor interactions, internalization, intracellular transport, genomic release, nuclear transport, and cell-to-cell transmission (159161). Different modeling approaches account for processes of intracellular particle transport by diffusion or mediated by active particle transport along microtubules (162), and an approach with relevant parameters for adeno-associated virus (AAV) has been outlined (163). More generally, approaches of statistical physics and their application to intracellular transport processes, which have been reviewed previously (164, 165), in combination with models for viral gene expression and function, will provide a more comprehensive accounting of the molecular biology and physical facets of virus intracellular growth.

Multiscale Modeling of Virus Infection Spread

To cause disease, virus infections of host cells amplify over multiple generations, requiring the intracellular processes that are the focus of this review to play out over multiple cycles of susceptible host cell infection. The viral progeny released from an initial infected host cell go on to further infect other host cells, perhaps in the same tissue of a multicellular host organism.
Models of virus infection in humans, initiated in the 1990s for the dynamics of HIV-1 in AIDS patients and informed by the predator-prey models of epidemiology (166, 167), provided the key insight that low levels of HIV-1 in patients were not only actively reproducing but also developing resistance to antiviral drug treatments (168). Similar modeling approaches have been extended to other viral infections in humans, including hepatitis C (169) and influenza A (170); related approaches have also been used to model virus infections as they selectively infect and spread in tumors as oncolytic therapies to treat cancer (171). Such within-host models, which accounted for the predator-prey dynamics of viruses and their host cells, are now serving as foundations for multiscale modeling in a top-down manner; the kinetics of processes within infected cells are of increasing interest owing to their capacity to account for explicit mechanisms of drug-target interactions and time delays in the release of virus progeny from their infected host cells (172, 173). At the same time, intracellular models of virus growth are serving as starting points for multiscale modeling in a bottom-up manner; the virus particles released from cells are tracked as they bind to and infect a new population of susceptible host cells (174, 175). When virus particles released from an infected cell are transported to neighboring or distant cells, physical processes of diffusion or fluid flow contribute to the dynamics of infection spread. Intracellular growth coupled with diffusional spread can be modeled using reaction-diffusion equations (176) or models that computationally simulate the behavior of cells on a grid as “agents” or “automata” that follow preestablished rules, using changes in the local environment (e.g., infection or immune signaling status of neighboring cells) to propagate infection spread (177180).
Viruses can readily mutate as they reproduce, so modeling approaches adapted from population genetics may be useful in accounting for how different virus or gene variants become enriched. Moreover, as viral infections spread, they also affect the dynamics of their host populations, whether they are host cells or host organisms; upon infection, those cells or organisms may recover or die. To unite the combined behavior of the virus during its transmission and growth with the fall or rise of the host populations, predator-prey models from epidemiology may be appropriate. More than 3 decades ago, May and Anderson suggested that models combining the population genetics and epidemiology perspectives would potentially account for both essential features (181). Such models can be viewed as multiscale in the sense that they account for local molecular and cellular mechanisms of amplification and genetic diversification that originate from within cells, as well as processes of within-host spread between infected and susceptible cells of a multicellular host. Further transmissions between infected and susceptible multicellular host organisms, most notably animals and humans, highlight how additional ecological and evolutionary considerations (182), including the behavior and mobility of animal and human hosts (183, 184), contribute to the dynamics of epidemics.

ACKNOWLEDGMENTS

We are indebted to the outstanding graduate students, postdocs, and undergrads in the Yin lab, who over the last 25 years have contributed their ideas and hard work toward our appreciation and understanding of viruses. We thank Paul Ahlquist, Udo Reichl, and Ophelia Venturelli for many thoughtful comments and suggestions on the manuscript.
We gratefully acknowledge support over the years from the National Science Foundation (grants BES-0087939, BES-0331337, BES-9896067, EF0313214, EIA-0130874, and EIA-0331337), the National Institutes of Health (grants AI077296, AI071197, AI091646, AI104317, T32-GM08349, and T32-HG002760), the National Library of Medicine (grant T15LM007359), the Office of Naval Research (grant N00014-98-1-0226), the Texas-Wisconsin Modeling and Control Consortium, Merck Research Laboratories, the Wisconsin Alumni Research Foundation, the Wisconsin Institute for Discovery, and the University of Wisconsin-Madison (Graduate School, Office of the Vice Chancellor for Research and Graduate Education, and the William F. Vilas Trust Estate).

REFERENCES

1.
Cairns J, Stent GS, Watson JD (ed). 1992. Phage and the origins of molecular biology. Cold Spring Harbor Laboratory Press, Plainview, NY.
2.
Brock TD. 1990. The emergence of bacterial genetics. Cold Spring Harbor Laboratory Press, Cold Spring Harbor, NY.
3.
McCance DJ. 1998. Human tumor viruses. American Society for Microbiology, Washington, DC.
4.
Philippe N, Legendre M, Doutre G, Coute Y, Poirot O, Lescot M, Arslan D, Seltzer V, Bertaux L, Bruley C, Garin J, Claverie JM, Abergel C. 2013. Pandoraviruses: amoeba viruses with genomes up to 2.5 Mb reaching that of parasitic eukaryotes. Science 341:281–286.
5.
Fraser CM, Gocayne JD, White O, Adams MD, Clayton RA, Fleischmann RD, Bult CJ, Kerlavage AR, Sutton G, Kelley JM, Fritchman RD, Weidman JF, Small KV, Sandusky M, Fuhrmann J, Nguyen D, Utterback TR, Saudek DM, Phillips CA, Merrick JM, Tomb JF, Dougherty BA, Bott KF, Hu PC, Lucier TS, Peterson SN, Smith HO, Hutchison CA, III, Venter JC. 1995. The minimal gene complement of Mycoplasma genitalium. Science 270:397–403.
6.
Ball LA. 2007. Virus replication strategies, p 119–140. In Knipe DM, Howley PM, Griffin DE, Lamb RA, Martin MA, Roizman B, Straus SE (ed), Fields virology, 5th ed. Lippincott Williams & Wilkins, Philadelphia, PA.
7.
Valentine RC, Allison AC. 1959. Virus particle adsorption. I. Theory of adsorption and experiments on the attachment of particles to non-biological surfaces. Biochim Biophys Acta 34:10–23.
8.
English TJ, Hammer DA. 2004. Brownian adhesive dynamics (BRAD) for simulating the receptor-mediated binding of viruses. Biophys J 86:3359–3372.
9.
Gilbert PA, Kamen A, Bernier A, Garnier A. 2007. A simple macroscopic model for the diffusion and adsorption kinetics of r-adenovirus. Biotechnol Bioeng 98:239–251.
10.
Dee KU, Hammer DA, Shuler ML. 1995. A model of the binding, entry, uncoating, and RNA synthesis of Semliki Forest virus in baby hamster kidney (BHK-21) cells. Biotechnol Bioeng 46:485–496.
11.
Dinh AT, Theofanous T, Mitragotri S. 2005. A model for intracellular trafficking of adenoviral vectors. Biophys J 89:1574–1588.
12.
Lim KI, Yin J. 2006. Dynamic tradeoffs in the raft-mediated entry of human immunodeficiency virus type 1 into cells. Biotechnol Bioeng 93:246–257.
13.
Decuzzi P, Ferrari M. 2008. The receptor-mediated endocytosis of nonspherical particles. Biophys J 94:3790–3797.
14.
Ackers GK, Johnson AD, Shea MA. 1982. Quantitative model for gene regulation by lambda phage repressor. Proc Natl Acad Sci U S A 79:1129–1133.
15.
Arnold S, Siemann M, Scharnweber K, Werner M, Baumann S, Reuss M. 2001. Kinetic modeling and simulation of in vitro transcription by phage T7 RNA polymerase. Biotechnol Bioeng 72:548–561.
16.
Weinberger LS, Burnett JC, Toettcher JE, Arkin AP, Schaffer DV. 2005. Stochastic gene expression in a lentiviral positive-feedback loop: HIV-1 Tat fluctuations drive phenotypic diversity. Cell 122:169–182.
17.
Boireau S, Maiuri P, Basyuk E, de la Mata M, Knezevich A, Pradet-Balade B, Backer V, Kornblihtt A, Marcello A, Bertrand E. 2007. The transcriptional cycle of HIV-1 in real-time and live cells. J Cell Biol 179:291–304.
18.
Nakabayashi J, Sasaki A. 2009. The function of temporally ordered viral gene expression in the intracellular replication of herpes simplex virus type 1 (HSV-1). J Theor Biol 261:156–164.
19.
Kim H, Yin J. 2005. In silico mutagenesis of RNA splicing in HIV-1. Biotechnol Bioeng 91:877–893.
20.
Kim H, Yin J. 2005. Effects of RNA splicing and posttranscriptional regulation on HIV-1 growth: a quantitative and integrated perspective. Syst Biol 152:138–152.
21.
Godefroy-Colburn T, Thach RE. 1981. The role of mRNA competition in regulating translation. IV. Kinetic model. J Biol Chem 256:11762–11773.
22.
Ray A, Walden WE, Brendler T, Zenger VE, Thach RE. 1985. Effect of medium hypertonicity on reovirus translation rates. An application of kinetic modeling in vivo. Biochemistry 24:7525–7532.
23.
Roldao A, Vieira HL, Charpilienne A, Poncet D, Roy P, Carrondo MJ, Alves PM, Oliveira R. 2007. Modeling rotavirus-like particles production in a baculovirus expression vector system: infection kinetics, baculovirus DNA replication, mRNA synthesis and protein production. J Biotechnol 128:875–894.
24.
Bourque DP, Bourque JG. 1975. A kinetic model of TMV-RNA replication based on rates of virus accumulation in vivo. Virology 63:130–134.
25.
Biebricher CK, Eigen M, Gardiner WC, Jr. 1984. Kinetics of RNA replication: plus-minus asymmetry and double-strand formation. Biochemistry 23:3186–3194.
26.
Hosoda K, Matsuura T, Kita H, Ichihashi N, Tsukada K, Yomo T. 2007. Kinetic analysis of the entire RNA amplification process by Qbeta replicase. J Biol Chem 282:15516–15527.
27.
Schwartz R, Shor PW, Prevelige PE, Jr, Berger B. 1998. Local rules simulation of the kinetics of virus capsid self-assembly. Biophys J 75:2626–2636.
28.
Zlotnick A, Johnson JM, Wingfield PW, Stahl SJ, Endres D. 1999. A theoretical model successfully identifies features of hepatitis B virus capsid assembly. Biochemistry 38:14644–14652.
29.
Mellado MC, Mena JA, Lopes A, Ramirez OT, Carrondo MJ, Palomares LA, Alves PM. 2009. Impact of physicochemical parameters on in vitro assembly and disassembly kinetics of recombinant triple-layered rotavirus-like particles. Biotechnol Bioeng 104:674–686.
30.
Purohit PK, Kondev J, Phillips R. 2003. Mechanics of DNA packaging in viruses. Proc Natl Acad Sci U S A 100:3173–3178.
31.
Yang Q, Catalano CE. 2004. A minimal kinetic model for a viral DNA packaging machine. Biochemistry 43:289–299.
32.
Forrey C, Muthukumar M. 2006. Langevin dynamics simulations of genome packing in bacteriophage. Biophys J 91:25–41.
33.
Reynwar BJ, Illya G, Harmandaris VA, Muller MM, Kremer K, Deserno M. 2007. Aggregation and vesiculation of membrane proteins by curvature-mediated interactions. Nature 447:461–464.
34.
Sompayrac L. 2002. How pathogenic viruses work. Jones and Bartlett Publishers, Boston, MA.
35.
Flint SJ, Racaniello VR, Rall GF, Skalka AM. 2015. Principles of virology, 4th ed. ASM Press, Washington, DC.
36.
Knipe DM, Howley PM, Cohen JI, Griffin DE, Lamb RA, Martin MA, Racaniello VR, Roizman B (ed). 2013. Fields virology, 6th ed. Lippincott Williams & Wilkins, Philadelphia, PA.
37.
Smeal SW, Schmitt MA, Pereira RR, Prasad A, Fisk JD. 2017. Simulation of the M13 life cycle. I. Assembly of a genetically-structured deterministic chemical kinetic simulation. Virology 500:259–274.
38.
Smeal SW, Schmitt MA, Pereira RR, Prasad A, Fisk JD. 2017. Simulation of the M13 life cycle. II. Investigation of the control mechanisms of M13 infection and establishment of the carrier state. Virology 500:275–284.
39.
Barricelli NA, Toombs R. 1971. Virus-genetic theory testing by data processing machines. III. The intepretation of radiation genetic and partial phage experiments in T4. General remarks on the theory-testing results. J Theor Biol 32:603–623.
40.
Barricelli NA, Toombs R. 1971. Virus-genetic theory testing by data processing machines. II. Fit of classical genetic T4 data. J Theor Biol 32:581–601.
41.
Barricelli NA, Nelson L, Toombs R. 1971. Virus-genetic theory testing by data processing machines. I. Basic procedures and models tested. J Theor Biol 32:559–580.
42.
Nathans D, Oeschger MP, Polmar SK, Eggen K. 1969. Regulation of protein synthesis directed by coliphage MS2 RNA. I. Phage protein and RNA synthesis in cells infected with suppressible mutants. J Mol Biol 39:279–292.
43.
Weissmann C. 1974. The making of a phage. FEBS Lett 40(Suppl):S10–S18.
44.
Knijnenburg A, Kreischer U. 1975. A new aspect of the RNA bacteriophages translation control mechanism. Biosystems 7:245–249.
45.
Knijnenburg A, Kreischer U. 1983. Discrete simulation of replication of a RNA-bacteriophage prototype system, p 267–290. In Bellmann K (ed), Molecular genetics information systems: modelling and simulation. Akademie-Verlag, Berlin, Germany.
46.
Zinder N (ed). 1975. RNA phages. Cold Spring Harbor Laboratory, Cold Spring Harbor, NY.
47.
Eigen M, Biebricher C, Gebinoga M, Gardiner W. 1991. The hypercycle. Coupling of RNA and protein biosynthesis in the infection cycle of an RNA bacteriophage. Biochemistry 30:11005–11018.
48.
Calendar R (ed). 1988. The bacteriophages. Plenum Press, New York, NY.
49.
Biebricher CK, Eigen M, Luce R. 1981. Kinetic analysis of template-instructed and de novo RNA synthesis by Q beta replicase. J Mol Biol 148:391–410.
50.
Biebricher CK, Eigen M, Gardiner WC, Jr. 1983. Kinetics of RNA replication. Biochemistry 22:2544–2559.
51.
Biebricher CK, Eigen M, Gardiner WC, Jr. 1985. Kinetics of RNA replication: competition and selection among self-replicating RNA species. Biochemistry 24:6550–6560.
52.
Kim H, Yin J. 2004. Energy-efficient growth of phage Q beta in Escherichia coli. Biotechnol Bioeng 88:148–156.
53.
Mahmoudabadi G, Milo R, Phillips R. 2017. Energetic cost of building a virus. Proc Natl Acad Sci U S A 114:E4324–E4333.
54.
Regoes RR, Crotty S, Antia R, Tanaka MM. 2005. Optimal replication of poliovirus within cells. Am Nat 165:364–373.
55.
Feist AM, Henry CS, Reed JL, Krummenacker M, Joyce AR, Karp PD, Broadbelt LJ, Hatzimanikatis V, Palsson BO. 2007. A genome-scale metabolic reconstruction for Escherichia coli K-12 MG1655 that accounts for 1260 ORFs and thermodynamic information. Mol Syst Biol 3:121.
56.
Jain R, Srivastava R. 2009. Metabolic investigation of host/pathogen interaction using MS2-infected Escherichia coli. BMC Syst Biol 3:121.
57.
Abedon ST. 1989. Selection for bacteriophage latent period length by bacterial density: a theoretical examination. Microb Ecol 18:79–88.
58.
Wang IN. 2006. Lysis timing and bacteriophage fitness. Genetics 172:17–26.
59.
Bernhardt T, Wang I, Struck D, Young R. 2001. A protein antibiotic in the phage Qbeta virion: diversity in lysis targets. Science 292:2326–2329.
60.
van den Worm SH, Koning RI, Warmenhoven HJ, Koerten HK, van Duin J. 2006. Cryo electron microscopy reconstructions of the Leviviridae unveil the densest icosahedral RNA packing possible. J Mol Biol 363:858–865.
61.
Dimmock NJ, Rainsford EW, Scott PD, Marriott AC. 2008. Influenza virus protecting RNA: an effective prophylactic and therapeutic antiviral. J Virol 82:8570–8578.
62.
Kim H, Yin J. 2004. Quantitative analysis of a parasitic antiviral strategy. Antimicrob Agents Chemother 48:1017–1020.
63.
Groenenboom MA, Hogeweg P. 2008. The dynamics and efficacy of antiviral RNA silencing: a model study. BMC Syst Biol 2:28.
64.
Yin J. 1993. Evolution of bacteriophage T7 in a growing plaque. J Bacteriol 175:1272–1277.
65.
Kong D, Yin J. 1995. Whole-virus vaccine development by continuous culture on a complementing host. Biotechnology (NY) 13:583–586.
66.
Kong D. 1995. Laboratory evolution and computer simulation of bacteriophage T7: implications for a new vaccine-development strategy. M.S. thesis. Dartmouth College, Hanover, NH.
67.
Endy A, Kong D, Yin J. 1997. Intracellular kinetics of a growing virus: a genetically structured simulation for bacteriophage T7. Biotechnol Bioeng 55:375–389.
68.
Endy D, Yin J. 2000. Toward antiviral strategies that resist viral escape. Antimicrob Agents Chemother 44:1097–1099.
69.
You L, Yin J. 2000. Patterns of regulation from mRNA and protein time-series. Metab Eng 2:210–217.
70.
Gygi SP, Rochon Y, Franza BR, Aebersold R. 1999. Correlation between protein and mRNA abundance in yeast. Mol Cell Biol 19:1720–1730.
71.
Nie L, Wu G, Culley DE, Scholten JC, Zhang W. 2007. Integrative analysis of transcriptomic and proteomic data: challenges, solutions and applications. Crit Rev Biotechnol 27:63–75.
72.
Lee PS, Shaw LB, Choe LH, Mehra A, Hatzimanikatis V, Lee KH. 2003. Insights into the relation between mRNA and protein expression patterns. II. Experimental observations in Escherichia coli. Biotechnol Bioeng 84:834–841.
73.
Mehra A, Lee KH, Hatzimanikatis V. 2003. Insights into the relation between mRNA and protein expression patterns. I. Theoretical considerations. Biotechnol Bioeng 84:822–833.
74.
de Sousa Abreu R, Penalva LO, Marcotte EM, Vogel C. 2009. Global signatures of protein and mRNA expression levels. Mol Biosyst 5:1512–1526.
75.
Mier-y-Teran-Romero L, Silber M, Hatzimanikatis V. 2010. The origins of time-delay in template biopolymerization processes. PLoS Comput Biol 6:e1000726.
76.
Endy D, You L, Yin J, Molineux IJ. 2000. Computation, prediction, and experimental tests of fitness for bacteriophage T7 mutants with permuted genomes. Proc Natl Acad Sci U S A 97:5375–5380.
77.
You L, Yin J. 2001. Simulating the growth of viruses. Pac Symp Biocomput 2001:532–543.
78.
Chan LY, Kosuri S, Endy D. 2005. Refactoring bacteriophage T7. Mol Syst Biol 1:2005.0018.
79.
You L, Suthers P, Yin J. 2002. Effects of Escherichia coli physiology on the growth of phage T7 in vivo and in silico. J Bacteriol 184:1888–1894.
80.
You L, Yin J. 2002. Dependence of epistasis on environment and mutation severity as revealed by in silico mutagenesis of phage T7. Genetics 160:1273–1281.
81.
Kirschner M, Gerhart J. 1998. Evolvability. Proc Natl Acad Sci U S A 95:8420–8427.
82.
Calson JM, Doyle J. 2000. Highly optimized tolerance: robustness and design in complex systems. Phys Rev Lett 84:2529–2532.
83.
You L, Yin J. 2006. Evolutionary design on a budget: robustness and optimality of bacteriophage T7. IEE Proc Syst Biol 153:46–52.
84.
Soto CM, Ratna BR. 2010. Virus hybrids as nanomaterials for biotechnology. Curr Opin Biotechnol 21:426–438.
85.
Cashman JS, Webster RE, Steege DA. 1980. Transcription of bacteriophage fl. The major in vivo RNAs. J Biol Chem 255:2554–2562.
86.
Goodrich AF, Steege DA. 1999. Roles of polyadenylation and nucleolytic cleavage in the filamentous phage mRNA processing and decay pathways in Escherichia coli. RNA 5:972–985.
87.
Salis HM. 2011. The ribosome binding site calculator. Methods Enzymol 498:19–42.
88.
Reddy B, Yin J. 1999. Quantitative intracellular kinetics of HIV type 1. AIDS Res Hum Retroviruses 15:273–283.
89.
Kim H, Yin J. 2005. Robust growth of human immunodeficiency virus type 1 (HIV-1). Biophys J 89:2210–2221.
90.
Tsai TY, Choi YS, Ma W, Pomerening JR, Tang C, Ferrell JE, Jr. 2008. Robust, tunable biological oscillations from interlinked positive and negative feedback loops. Science 321:126–129.
91.
Zarrabi N, Mancini E, Tay J, Shahand S, Sloot P. 2012. Modeling HIV-1 intracellular replication: two simulation approaches. Procedia Comput Sci 1:555–564.
92.
Leonard JN, Schaffer DV. 2005. Computational design of antiviral RNA interference strategies that resist human immunodeficiency virus escape. J Virol 79:1645–1654.
93.
Leonard JN, Shah PS, Burnett JC, Schaffer DV. 2008. HIV evades RNA interference directed at TAR by an indirect compensatory mechanism. Cell Host Microbe 4:484–494.
94.
Sidorenko Y, Reichl U. 2004. Structured model of influenza virus replication in MDCK cells. Biotechnol Bioeng 88:1–14.
95.
Sidorenko Y, Schulze-Horsel J, Voigt A, Reichl U, Kienle A. 2008. Stochastic population balance modeling of influenza virus replication in vaccine production processes. Chem Eng Sci 63:157–169.
96.
Sidorenko Y, Voigt A, Schulze-Horsel J, Reichl U, Kienle A. 2008. Stochastic population balance modeling of influenza virus replication in vaccine production processes. II. Detailed description of replication mechanism. Chem Eng Sci 63:2299–2304.
97.
Bazhan SI, Kashevarova NA, Khlebodarova TM, Likhoshvai VA, Kolchanov NA. 2009. A mathematical model of the intracellular reproduction of the influenza virus. Biofizika 54:1066–1080. (In Russian.)
98.
Heldt FS, Frensing T, Reichl U. 2012. Modeling the intracellular dynamics of influenza virus replication to understand the control of viral RNA synthesis. J Virol 86:7806–7817.
99.
Saira K, Lin X, DePasse JV, Halpin R, Twaddle A, Stockwell T, Angus B, Cozzi-Lepri A, Delfino M, Dugan V, Dwyer DE, Freiberg M, Horban A, Losso M, Lynfield R, Wentworth DN, Holmes EC, Davey R, Wentworth DE, Ghedin E, INSIGHT FLU002 Study Group, INSIGHT FLU003 Study Group. 2013. Sequence analysis of in vivo defective interfering-like RNA of influenza A H1N1 pandemic virus. J Virol 87:8064–8074.
100.
Frensing T. 2015. Defective interfering viruses and their impact on vaccines and viral vectors. Biotechnol J 10:681–689.
101.
Laske T, Heldt FS, Hoffmann H, Frensing T, Reichl U. 2016. Modeling the intracellular replication of influenza A virus in the presence of defective interfering RNAs. Virus Res 213:90–99.
102.
Krakauer DC, Komarova NL. 2003. Levels of selection in positive-strand virus dynamics. J Evol Biol 16:64–73.
103.
Stearns SC. 1992. The evolution of life histories. Oxford University Press, Oxford, United Kingdom.
104.
Iverson LE, Rose JK. 1981. Localized attenuation and discontinuous synthesis during vesicular stomatitis virus transcription. Cell 23:477–484.
105.
Whelan SP, Barr JN, Wertz GW. 2004. Transcription and replication of nonsegmented negative-strand RNA viruses. Curr Top Microbiol Immunol 283:61–119.
106.
Clarke D, Duarte E, Moya A, Elena S, Domingo E, Holland J. 1993. Genetic bottlenecks and population passages cause profound fitness differences in RNA viruses. J Virol 67:222–228.
107.
Novella IS, Clarke DK, Quer J, Duarte EA, Lee CH, Weaver SC, Elena SF, Moya A, Domingo E, Holland JJ. 1995. Extreme fitness differences in mammalian and insect hosts after continuous replication of vesicular stomatitis virus in sandfly cells. J Virol 69:6805–6809.
108.
Remold SK, Rambaut A, Turner PE. 2008. Evolutionary genomics of host adaptation in vesicular stomatitis virus. Mol Biol Evol 25:1138–1147.
109.
Stojdl DF, Lichty BD, tenOever BR, Paterson JM, Power AT, Knowles S, Marius R, Reynard J, Poliquin L, Atkins H, Brown EG, Durbin RK, Durbin JE, Hiscott J, Bell JC. 2003. VSV strains with defects in their ability to shutdown innate immunity are potent systemic anti-cancer agents. Cancer Cell 4:263–275.
110.
Bell JC, Garson KA, Lichty BD, Stojdl DF. 2002. Oncolytic viruses: programmable tumour hunters. Curr Gene Ther 2:243–254.
111.
Barber GN. 2004. Vesicular stomatitis virus as an oncolytic vector. Viral Immunol 17:516–527.
112.
Ahmed M, Cramer SD, Lyles DS. 2004. Sensitivity of prostate tumors to wild type and M protein mutant vesicular stomatitis viruses. Virology 330:34–49.
113.
Barr JN, Whelan SP, Wertz GW. 2002. Transcriptional control of the RNA-dependent RNA polymerase of vesicular stomatitis virus. Biochim Biophys Acta 1577:337–353.
114.
Wertz G, Perepelitsa V, Ball L. 1998. Gene rearrangement attenuates expression and lethality of a nonsegmented negative strand RNA virus. Proc Natl Acad Sci U S A 95:3501–3506.
115.
Ball LA, Pringle CR, Flanagan B, Perepelitsa VP, Wertz GW. 1999. Phenotypic consequences of rearranging the P, M, and G genes of vesicular stomatitis virus. J Virol 73:4705–4712.
116.
Lim KI, Lang T, Lam V, Yin J. 2006. Model-based design of growth-attenuated viruses. PLoS Comput Biol 2:e116.
117.
Lim KI, Yin J. 2009. Computational fitness landscape for all gene-order permutations of an RNA virus. PLoS Comput Biol 5:e1000283.
118.
Aucoin MG, Mena JA, Kamen AA. 2010. Bioprocessing of baculovirus vectors: a review. Curr Gene Ther 10:174–186.
119.
Licari P, Bailey JE. 1992. Modeling the population dynamics of baculovirus-infected insect cells: optimizing infection strategies for enhanced recombinant protein yields. Biotechnol Bioeng 39:432–441.
120.
Hu YC, Bentley WE. 2000. A kinetic and statistical-thermodynamic model for baculovirus infection and virus-like particle assembly in suspended insect cells. Chem Eng Sci 55:3991–4008.
121.
Stephanopoulos GN, Aristidou AA, Nielsen J. 1998. Metabolic engineering. Principles and methodologies. Academic Press, New York, NY.
122.
Bernal V, Carinhas N, Yokomizo AY, Carrondo MJ, Alves PM. 2009. Cell density effect in the baculovirus-insect cells system: a quantitative analysis of energetic metabolism. Biotechnol Bioeng 104:162–180.
123.
Carinhas N, Bernal V, Monteiro F, Carrondo MJ, Oliveira R, Alves PM. 2010. Improving baculovirus production at high cell density through manipulation of energy metabolism. Metab Eng 12:39–52.
124.
Villeneuve JP. 2005. The natural history of chronic hepatitis B virus infection. J Clin Virol 34(Suppl 1):S139–S142.
125.
Nakabayashi J, Sasaki A. 2011. A mathematical model of the intracellular replication and within host evolution of hepatitis type B virus: understanding the long time course of chronic hepatitis. J Theor Biol 269:318–329.
126.
Nakabayashi J. 2016. The intracellular dynamics of hepatitis B virus (HBV) replication with reproduced virion “re-cycling.” J Theor Biol 396:154–162.
127.
Lohmann V, Korner F, Koch J, Herian U, Theilmann L, Bartenschlager R. 1999. Replication of subgenomic hepatitis C virus RNAs in a hepatoma cell line. Science 285:110–113.
128.
Dahari H, Ribeiro RM, Rice CM, Perelson AS. 2007. Mathematical modeling of subgenomic hepatitis C virus replication in Huh-7 cells. J Virol 81:750–760.
129.
Mishchenko EL, Bezmaternykh KD, Likhoshvai VA, Ratushny AV, Khlebodarova TM, Yu Sournina N, Ivanisenko VA, Kolchanov NA. 2007. Mathematical model for suppression of subgenomic hepatitis C virus RNA replication in cell culture. J Bioinform Comput Biol 5:593–609.
130.
Nakabayashi J. 2012. A compartmentalization model of hepatitis C virus replication: an appropriate distribution of HCV RNA for the effective replication. J Theor Biol 300:110–117.
131.
Looker KJ, Magaret AS, May MT, Turner KM, Vickerman P, Gottlieb SL, Newman LM. 2015. Global and regional estimates of prevalent and incident herpes simplex virus type 1 infections in 2012. PLoS One 10:e0140765.
132.
Roizman B, Knipe D, Whitley R. 2006. Herpes simplex virus, p 2501–2601. In Knipe DM, Howley PM, Griffin DE, Lamb RA, Martin MA, Roizman B, Straus SE (ed), Fields virology, 5th ed. Lippincott Williams & Wilkins, Philadelphia, PA.
133.
Boehmer PE, Lehman IR. 1997. Herpes simplex virus DNA replication. Annu Rev Biochem 66:347–384.
134.
Dauber B, Poon D, Dos Santos T, Duguay BA, Mehta N, Saffran HA, Smiley JR. 2016. The herpes simplex virus virion host shutoff protein enhances translation of viral true late mRNAs independently of suppressing protein kinase R and stress granule formation. J Virol 90:6049–6057.
135.
Milo R, Jorgensen P, Moran U, Weber G, Springer M. 2010. BioNumbers—the database of key numbers in molecular and cell biology. Nucleic Acids Res 38:D750–D753.
136.
Birch EW, Ruggero NA, Covert MW. 2012. Determining host metabolic limitations on viral replication via integrated modeling and experimental perturbation. PLoS Comput Biol 8:e1002746.
137.
Wahl A, Sidorenko Y, Dauner M, Genzel Y, Reichl U. 2008. Metabolic flux model for an anchorage-dependent MDCK cell line: characteristic growth phases and minimum substrate consumption flux distribution. Biotechnol Bioeng 101:135–152.
138.
Sidorenko Y, Wahl A, Dauner M, Genzel Y, Reichl U. 2008. Comparison of metabolic flux distributions for MDCK cell growth in glutamine- and pyruvate-containing media. Biotechnol Prog 24:311–320.
139.
Sun X, Xian H, Tian S, Sun T, Qin Y, Zhang S, Cui J. 2016. A hierarchical mechanism of RIG-I ubiquitination provides sensitivity, robustness and synergy in antiviral immune responses. Sci Rep 6:29263.
140.
Qiao L, Phipps-Yonas H, Hartmann B, Moran TM, Sealfon SC, Hayot F. 2010. Immune response modeling of interferon beta-pretreated influenza virus-infected human dendritic cells. Biophys J 98:505–514.
141.
Fribourg M, Hartmann B, Schmolke M, Marjanovic N, Albrecht RA, Garcia-Sastre A, Sealfon SC, Jayaprakash C, Hayot F. 2014. Model of influenza A virus infection: dynamics of viral antagonism and innate immune response. J Theor Biol 351:47–57.
142.
Seto J, Qiao L, Guenzel CA, Xiao S, Shaw ML, Hayot F, Sealfon SC. 2010. Novel Nipah virus immune-antagonism strategy revealed by experimental and computational study. J Virol 84:10965–10973.
143.
Ellis EL, Delbrück M. 1939. The growth of bacteriophage. J Gen Physiol 22:365.
144.
Delbrück M. 1945. The burst size distribution in the growth of bacterial viruses (bacteriophages). J Bacteriol 50:131–135.
145.
Steinhauer D, de la Torre J, Meier E, Holland J. 1989. Extreme heterogeneity in populations of vesicular stomatitis virus. J Virol 63:2072–2080.
146.
Marcus PI. 1959. Symposium on the biology of cells modified by viruses or antigens. IV. Single-cell techniques in tracing virus-host interactions. Bacteriol Rev 23:232–249.
147.
Zhu Y, Yongky A, Yin J. 2009. Growth of an RNA virus in single cells reveals a broad fitness distribution. Virology 385:39–46.
148.
Delbrück M. 1940. Statistical fluctuations in autocatalytic reactions. J Chem Phys 8:120–124.
149.
Kaern M, Elston TC, Blake WJ, Collins JJ. 2005. Stochasticity in gene expression: from theories to phenotypes. Nat Rev Genet 6:451–464.
150.
Raj A, van Oudenaarden A. 2009. Single-molecule approaches to stochastic gene expression. Annu Rev Biophys 38:255–270.
151.
Gillespie DT. 1976. A general method for numerically simulating the stochastic time evolution of coupled chemical reactions. J Comput Phys 22:403–434.
152.
Arkin A, Ross J, McAdams HH. 1998. Stochastic kinetic analysis of developmental pathway bifurcation in phage lambda-infected Escherichia coli cells. Genetics 149:1633–1648.
153.
Srivastava R, You L, Summers J, Yin J. 2002. Stochastic versus deterministic modeling of intracellular viral kinetics. J Theor Biol 218:309–321.
154.
Zhdanov VP. 2004. Stochastic kinetics of reproduction of virions inside a cell. Biosystems 77:143–150.
155.
Hensel SC, Rawlings JB, Yin J. 2009. Stochastic kinetic modeling of vesicular stomatitis virus intracellular growth. Bull Math Biol 71:1671–1692.
156.
Heldt FS, Kupke SY, Dorl S, Reichl U, Frensing T. 2015. Single-cell analysis and stochastic modelling unveil large cell-to-cell variability in influenza A virus infection. Nat Commun 6:8938.
157.
Hu J, Nudelman G, Shimoni Y, Kumar M, Ding Y, Lopez C, Hayot F, Wetmur JG, Sealfon SC. 2011. Role of cell-to-cell variability in activating a positive feedback antiviral response in human dendritic cells. PLoS One 6:e16614.
158.
Rand U, Rinas M, Schwerk J, Nohren G, Linnes M, Kroger A, Flossdorf M, Kaly-Kullai K, Hauser H, Hofer T, Koster M. 2012. Multi-layered stochasticity and paracrine signal propagation shape the type-I interferon response. Mol Syst Biol 8:584.
159.
Damm EM, Pelkmans L. 2006. Systems biology of virus entry in mammalian cells. Cell Microbiol 8:1219–1227.
160.
Brandenburg B, Zhuang X. 2007. Virus trafficking—learning from single-virus tracking. Nat Rev Microbiol 5:197–208.
161.
Sun E, He J, Zhuang X. 2013. Live cell imaging of viral entry. Curr Opin Virol 3:34–43.
162.
Lagache T, Holcman D. 2008. Quantifying intermittent transport in cell cytoplasm. Phys Rev E Stat Nonlin Soft Matter Phys 77:030901.
163.
Lagache T, Dauty E, Holcman D. 2009. Physical principles and models describing intracellular virus particle dynamics. Curr Opin Microbiol 12:439–445.
164.
Lagache T, Dauty E, Holcman D. 2009. Quantitative analysis of virus and plasmid trafficking in cells. Phys Rev E Stat Nonlin Soft Matter Phys 79:011921.
165.
Bressloff PC, Newby JM. 2013. Stochastic models of intracellular transport. Rev Mod Phys 85:135–196.
166.
Ho D, Neumann A, Perelson A, Chen W, Leonard J, Markowitz M. 1995. Rapid turnover of plasma virions and CD4 lymphocytes in HIV-1 infection. Nature 373:123–126.
167.
Wei X, Ghosh SK, Taylor ME, Johnson VA, Emini EA, Deutsch P, Lifson JD, Bonhoeffer S, Nowak MA, Hahn BH, Saag MS, Shaw GM. 1995. Viral dynamics in human immunodeficiency virus type 1 infection. Nature 373:117–122.
168.
Furtado MR, Callaway DS, Phair JP, Kunstman KJ, Stanton JL, Macken CA, Perelson AS, Wolinsky SM. 1999. Persistence of HIV-1 transcription in peripheral-blood mononuclear cells in patients receiving potent antiretroviral therapy. N Engl J Med 340:1614–1622.
169.
Neumann AU, Lam NP, Dahari H, Gretch DR, Wiley TE, Layden TJ, Perelson AS. 1998. Hepatitis C viral dynamics in vivo and the antiviral efficacy of interferon-alpha therapy. Science 282:103–107.
170.
Baccam P, Beauchemin C, Macken CA, Hayden FG, Perelson AS. 2006. Kinetics of influenza A virus infection in humans. J Virol 80:7590–7599.
171.
Komarova NL, Wodarz D. 2010. ODE models for oncolytic virus dynamics. J Theor Biol 263:530–543.
172.
Dixit NM, Perelson AS. 2004. Complex patterns of viral load decay under antiretroviral therapy: influence of pharmacokinetics and intracellular delay. J Theor Biol 226:95–109.
173.
Guedj J, Dahari H, Rong L, Sansone ND, Nettles RE, Cotler SJ, Layden TJ, Uprichard SL, Perelson AS. 2013. Modeling shows that the NS5A inhibitor daclatasvir has two modes of action and yields a shorter estimate of the hepatitis C virus half-life. Proc Natl Acad Sci U S A 110:3991–3996.
174.
Haseltine EL, Rawlings JB, Yin J. 2005. Dynamics of viral infections: incorporating both the intracellular and extracellular levels. Comput Chem Eng 29:675–686.
175.
Haseltine EL, Yin J, Rawlings JB. 2008. Implications of decoupling the intracellular and extracellular levels in multi-level models of virus growth. Biotechnol Bioeng 101:811–820.
176.
Haseltine EL, Lam V, Yin J, Rawlings JB. 2008. Image-guided modeling of virus growth and spread. Bull Math Biol 70:1730–1748.
177.
Howat TJ, Barreca C, O'Hare P, Gog JR, Grenfell BT. 2006. Modelling dynamics of the type I interferon response to in vitro viral infection. J R Soc Interface 3:699–709.
178.
Bauer AL, Beauchemin CA, Perelson AS. 2009. Agent-based modeling of host-pathogen systems: the successes and challenges. Inf Sci (NY) 179:1379–1389.
179.
Yakimovich A, Gumpert H, Burckhardt CJ, Lutschg VA, Jurgeit A, Sbalzarini IF, Greber UF. 2012. Cell-free transmission of human adenovirus by passive mass transfer in cell culture simulated in a computer model. J Virol 86:10123–10137.
180.
Akpinar F, Inankur B, Yin J. 2016. Spatial-temporal patterns of viral amplification and interference initiated by a single infected cell. J Virol 90:7552–7566.
181.
May RM, Anderson RM. 1983. Epidemiology and genetics in the coevolution of parasites and hosts. Proc R Soc Lond B Biol Sci 219:281–313.
182.
Loewe L. 2009. A framework for evolutionary systems biology. BMC Syst Biol 3:27.
183.
Yin J. 2007. Chemical engineering and virology: challenges and opportunities at the interface. AIChE J 53:2202–2209.
184.
Murillo LN, Murillo MS, Perelson AS. 2013. Towards multiscale modeling of influenza infection. J Theor Biol 332:267–290.

Author Bios

John Yin
Department of Chemical and Biological Engineering, Wisconsin Institute for Discovery, University of Wisconsin-Madison, Madison, Wisconsin, USA
John Yin earned bachelor degrees in the liberal arts and engineering from Columbia University and a Ph.D. in chemical engineering from UC-Berkeley. He pursued postdoctoral studies at the Max Planck Institute for Biophysical Chemistry in Göttingen, Germany, advancing the experimental evolution of viruses. Yin's academic career started at Dartmouth College, and in 1998, he joined the Chemical Engineering Department at the University of Wisconsin-Madison. A decade later, he and colleagues initiated a research thrust in evolutionary systems biology as a founding theme of the Wisconsin Institute for Discovery. Today he is the Vilas Distinguished Achievement Professor, pursuing single-cell measures and models of virus-host interactions and quantitative characterization of spreading infections. His research interests also include the chemical origins of information, metabolism, self-replication, and life. Yin is a semiprofessional pianist and cellist, and he enjoys sous vide methods of quantitative cooking.
Jacob Redovich
Department of Chemical and Biological Engineering, Wisconsin Institute for Discovery, University of Wisconsin-Madison, Madison, Wisconsin, USA
Jacob Redovich is an undergraduate at the University of Wisconsin-Madison, where he is studying chemical engineering and chemistry. He has held research positions in the UW-Madison Department of Chemistry, the Zhejiang University Department of Chemical and Biological Engineering in Hangzhou, China, and most recently the Yin group in the Department of Chemical and Biological Engineering at UW-Madison. Jacob has conducted research in electrocatalytic materials chemistry, biomaterials, and viral kinetics. His main research interests include kinetic modeling, data science, and genetics. He has also worked as a chemistry tutor and participated in a process engineering co-op at Monsanto.

Information & Contributors

Information

Published In

cover image Microbiology and Molecular Biology Reviews
Microbiology and Molecular Biology Reviews
Volume 82Number 2June 2018
eLocator: 10.1128/mmbr.00066-17

History

Published online: 28 March 2018

Permissions

Request permissions for this article.

Keywords

  1. DNA virus
  2. RNA virus
  3. bacteriophages
  4. biophysics
  5. computational biology
  6. computer modeling
  7. growth modeling
  8. kinetics
  9. mathematical modeling
  10. molecular biology

Contributors

Authors

John Yin
Department of Chemical and Biological Engineering, Wisconsin Institute for Discovery, University of Wisconsin-Madison, Madison, Wisconsin, USA
Jacob Redovich
Department of Chemical and Biological Engineering, Wisconsin Institute for Discovery, University of Wisconsin-Madison, Madison, Wisconsin, USA

Notes

Address correspondence to John Yin, [email protected].

Metrics & Citations

Metrics

Note:

  • For recently published articles, the TOTAL download count will appear as zero until a new month starts.
  • There is a 3- to 4-day delay in article usage, so article usage will not appear immediately after publication.
  • Citation counts come from the Crossref Cited by service.

Citations

If you have the appropriate software installed, you can download article citation data to the citation manager of your choice. For an editable text file, please select Medlars format which will download as a .txt file. Simply select your manager software from the list below and click Download.

View Options

Figures and Media

Figures

Media

Tables

Share

Share

Share the article link

Share with email

Email a colleague

Share on social media

American Society for Microbiology ("ASM") is committed to maintaining your confidence and trust with respect to the information we collect from you on websites owned and operated by ASM ("ASM Web Sites") and other sources. This Privacy Policy sets forth the information we collect about you, how we use this information and the choices you have about how we use such information.
FIND OUT MORE about the privacy policy