Skip to main content
Advertisement
  • Loading metrics

Cell cycle time series gene expression data encoded as cyclic attractors in Hopfield systems

  • Anthony Szedlak,

    Roles Conceptualization, Formal analysis, Investigation, Writing – original draft

    Affiliation Department of Physics and Astronomy, Michigan State University, East Lansing, Michigan, United States of America

  • Spencer Sims,

    Roles Formal analysis

    Affiliation Department of Physics and Astronomy, Michigan State University, East Lansing, Michigan, United States of America

  • Nicholas Smith,

    Roles Formal analysis

    Affiliation Salgomed Inc., Del Mar, California, United States of America

  • Giovanni Paternostro,

    Roles Conceptualization, Funding acquisition, Project administration, Supervision

    Affiliation Sanford Burnham Prebys Medical Discovery Institute, La Jolla, California, United States of America

  • Carlo Piermarocchi

    Roles Conceptualization, Funding acquisition, Project administration, Supervision

    carlo@pa.msu.edu

    Affiliation Department of Physics and Astronomy, Michigan State University, East Lansing, Michigan, United States of America

Abstract

Modern time series gene expression and other omics data sets have enabled unprecedented resolution of the dynamics of cellular processes such as cell cycle and response to pharmaceutical compounds. In anticipation of the proliferation of time series data sets in the near future, we use the Hopfield model, a recurrent neural network based on spin glasses, to model the dynamics of cell cycle in HeLa (human cervical cancer) and S. cerevisiae cells. We study some of the rich dynamical properties of these cyclic Hopfield systems, including the ability of populations of simulated cells to recreate experimental expression data and the effects of noise on the dynamics. Next, we use a genetic algorithm to identify sets of genes which, when selectively inhibited by local external fields representing gene silencing compounds such as kinase inhibitors, disrupt the encoded cell cycle. We find, for example, that inhibiting the set of four kinases AURKB, NEK1, TTK, and WEE1 causes simulated HeLa cells to accumulate in the M phase. Finally, we suggest possible improvements and extensions to our model.

Author summary

Cell cycle—the process in which a parent cell replicates its DNA and divides into two daughter cells—is an upregulated process in many forms of cancer. Identifying gene inhibition targets to regulate cell cycle is important to the development of effective therapies. Although modern high throughput techniques offer unprecedented resolution of the molecular details of biological processes like cell cycle, analyzing the vast quantities of the resulting experimental data and extracting actionable information remains a formidable task. Here, we create a dynamical model of the process of cell cycle using the Hopfield model (a type of recurrent neural network) and gene expression data from human cervical cancer cells and yeast cells. We find that the model recreates the oscillations observed in experimental data. Tuning the level of noise (representing the inherent randomness in gene expression and regulation) to the “edge of chaos” is crucial for the proper behavior of the system. We then use this model to identify potential gene targets for disrupting the process of cell cycle. This method could be applied to other time series data sets and used to predict the effects of untested targeted perturbations.

Introduction

Originally proposed by Conrad Waddington in the 1950s [1] and Stuart Kauffman in the 1970s [2], analysis of biological processes such as cellular differentiation and cancer development using attractor models—dynamical systems whose configurations tend to evolve toward particular sets of states—has gained significant traction over the past decade [312]. One such attractor model, the Hopfield model [13], is a type of recurrent artificial neural network based on spin glasses. It was designed with the ability to recall a host of memorized patterns from noisy or partial input information by mapping data directly to attractor states. A great deal of analytical and numerical work has been devoted to understanding the statistical properties of the Hopfield model, including its storage capacity [14], correlated patterns [15], spurious attractors [16], asymmetric connections [17], embedded cycles [18], and complex transition landscapes [19]. Due to its prescriptive, data-driven design, the Hopfield model has been applied in a variety of fields including image recognition [20, 21] and the clustering of gene expression data [22]. It has also been used to directly model the dynamics of cellular differentiation and stem cell reprogramming [23, 24], targeted inhibition of genes in cancer gene regulatory networks [25], and cell cycle across various stages of cellular differentiation [26].

Techniques for measuring large scale omics data, particularly transcriptomic data from microarrays and RNA sequencing (RNA-seq), have become standard, indispensable tools for observing the states of complex biological systems [2729]. However, analysis of the sheer variety and vast quantities of data these techniques produce requires the development of new mathematical tools. Inference and topological analysis of gene regulatory networks has garnered much attention as a method for distilling meaningful information from large datasets [3036], but simply analyzing the topology of static networks without a signaling rule (e.g. differential equations, digital logic gates, or discrete maps) fails to capture the nonlinear dynamics crucial to cellular behavior. The non-equilibrium nature of life implies that it can only be truly understood at the dynamical level, necessitating the development of new methods for analyzing time series data. As experimental methods continue to improve, more and more high-resolution time series omics and even multi-omics [37] data sets will inevitably become available. Here, we demonstrate that time series omics data (in this case, transcriptomic data) representing cyclic biological processes can be encoded in Hopfield systems, providing a new model for analyzing the dynamics of, and exploring effects of perturbations to, such systems.

The dynamics of cell cycle (CC)—the process in which a parent cell replicates its DNA and divides into two daughter cells—is both scientifically interesting and therapeutically important, and has been modeled extensively using differential equations, Boolean models, and discrete maps [3855]. Even relatively simple simulated systems such as an isolated, positively self-regulating gene subject to noise can exhibit rich dynamical behavior [56]; but like many biological processes, the proper functioning of CC requires the decentralized, coordinated action of hundreds of genes. CC thus provides researchers with a convenient case study of self-organization in a noisy environment. CC is also an upregulated process in many forms of cancer [5760], and control of CC using pharmaceutical compounds such as kinase inhibitors is a critical goal in cancer research. The combinatorics of selectively inhibiting sets of genes makes exhaustive experimental searches difficult or impractical [61]. However, network-based mathematical models such as the one presented here enable researchers to examine the effects of perturbations to complex systems [62, 63] by testing potential inhibition targets in silico. The efficacy of these predictions can then be experimentally validated or invalidated, providing new information and insights to further refine models.

The remainder of this article is structured as follows. In the Models subsection of the Results section we first discuss how periodic genes were identified in the time series gene expression data sets, and how Boolean attractors were extracted from the continuous data (explained in greater detail in the Methods section). We then introduce the Hopfield model and discuss the specific form of the coupling matrix used in this application. We discuss how to interpret the results of Hopfield simulations in the context of gene expression and cells. We also explain the objective function used by the genetic algorithm to identify potential inhibition targets, designed with the intention of disrupting CC. In the Dynamical behavior subsection, we show that this model qualitatively recreates experimental gene expression data, and we demonstrate and analyze some dynamical properties of the delayed Hopfield model, including the role played by noise. We include supplementary videos to emphasize the dynamical nature of this model. Optimal control fields for both unconstrained searches (in which any gene may be inhibited) and searches constrained to kinases are discussed. Finally, we recap our results and suggest possible improvements and generalizations to our methods in the Discussion section.

Results

Model

Periodic gene selection.

Microarray and RNA-seq time series data sets were obtained from Eser et al. (S. cerevisiae) [64] and Dominguez et al. (HeLa, human cervical cancer) [65]. For consistency and due to its higher resolution, the S. cerevisiae data set was chosen to produce all images and movies in this article, but both data sets were analyzed. In order to encode these CC data sets into the Hopfield model, periodic genes needed to be identified, their frequencies and phases computed, and their expression converted from continuous to Boolean form. As detailed in the Methods section, decaying sinusoids were fitted to the trajectory of each gene i, and genes with sufficiently high quality fits were kept. This resulted in 379 periodic genes in S. cerevisiae and 722 periodic genes in HeLa cells. Fig 1A shows a heat map of the expression of all periodic genes detected in the Eser data set sorted by their fitted phases, and Fig 1B shows the same genes with the fitted expression curves. These fitted curves were converted from continuous values xi(t) ≥ 0 to Boolean values ξi(t) = ±1 (over/underexpressed) as shown in Fig 1C. Finally, one CC period was divided into eight uniformly spaced states with . These states, shown in Fig 1D, were used as the eight attractor patterns in the Hopfield model. p = 8 was chosen because the embedded cycle lost stability for p > 8 due to the finite capacity of the Hopfield model.

thumbnail
Fig 1. Obtaining attractors from expression data.

(A) Heat map of the expression of all detected periodic genes from [64] sorted by their fitted phases. (B) Fitted gene expression. (C) Boolean form of fitted expression, separated into periods by dashed black lines. (D) Final set of p = 8 attractors taken from one period.

https://doi.org/10.1371/journal.pcbi.1005849.g001

Note that this work is based on the simplifying assumption that cell cycle genes exhibit periodic expression, in line with other studies of cell cycle genes [6470]. This method cannot detect non-periodic cell cycle effectors, but our goal here is to model cell cycle based solely on time series gene expression data.

The Hopfield model.

The Hopfield model [13] is an Ising model whose configuration is defined by N spins σi(t) at integer time t. The state of each node (gene) takes one of two values, σi(t) = ±1 (over/underexpressed). The coupling matrix Jij defines the strength and sign of the signal sent from node j to node i, and its construction is discussed in the following subsection. The total field at node i at time t is given by (1) where ∑j Jij σj(t) is the internal field at node i due to its coupling with all nodes j and is an optional external field applied to node i representing the action of therapeutic compounds, e.g. kinase inhibitors. The dynamical update rule is given by (2) where the factor of 2 in the exponent is conventional and T is an effective temperature representing the level of noise (not a physical temperature). Biologically, this noise represents the effects of all kinds of biochemical fluctuations present in cells. Note that for hi(t) → ±∞, σi(t + 1) = ±1; for T → ∞, σi(t + 1) = ±1 with equal probability; and for T → 0, σi(t + 1) = sign (hi(t)).

The update rule from Eq 2 may be implemented in various ways. The synchronous scheme updates the state of all nodes in the system at every time step, but this is sensible only if the simulated system has a central pacemaker coordinating the activity of all nodes. A more appropriate choice for decentralized systems like gene regulatory networks is the asynchronous scheme in which the state of a randomly chosen subset of nodes is updated at each time step. Here, we use the asynchronous scheme with update probability 0.2 for each node.

Coupling matrix.

In the canonical Hopfield model, the coupling matrix is constructed to store a set of p linearly independent (i.e. distinct) Boolean patterns as point attractors, where i = 0, 1, …, N − 1 is the node index and μ = 0, 1, …, p − 1 is the pattern index. The point attractor coupling matrix is given by (3) where [15, 71] (4) With this coupling matrix and T = 0, if at some time t the configuration is given by for a small perturbation δi, then . Note that this formulation means are both attractors of the system.

A simple modification [19] to Eq 3 produces a cyclic attractor coupling matrix , constructed according to (5) At T = 0, this coupling matrix cyclically maps through the sequence of p patterns (6) or their negatives. For the remainder of this article, all attractor indexing is understood to be modulo p.

A delayed cyclic Hopfield model may be constructed by combining the point and cyclic attractor matrices into one coupling matrix, (7) for an adjustable transition strength parameter λ with 0 ≤ λ ≤ 1. If , λ ≪ 1, and T = 0, the point attractor term dominates and σi(t) = σ(t + Δt) for all Δt = 1, 2, …. If T > 0, however, stochastic fluctuations eventually push the configuration out of the basin of attraction of the μth attractor and into the (μ + 1)th basin, then eventually to the (μ + 2)th basin, and so on. The dynamics of the delayed cyclic Hopfield model are thus governed by noise-induced transitions.

Due to the sinusoidal nature of the gene expression in these CC data sets, however, the attractors are structured such that , making Qμν rank deficient and thus noninvertible. Because the definition of Jij automatically guarantees that if any sequence is an attractor, then is also an attractor, encoding the sequence (8) automatically encodes the sequence (9) In this special case of sinusoidal trajectories, the limits of summation in Eqs 35 need only run over the first four indices, μ = 0, 1, 2, 3.

Finally, to reflect the fact that real gene regulatory networks are sparse, weak edges were removed by setting all elements of the coupling matrix with |Jij| < median(|J|) to zero, where |J| is element-wise absolute value. The cyclic attractor was not preserved when using stricter thresholds due to the reduced capacity of diluted Hopfield networks [17].

Biological interpretation of the dynamics.

Extracting biological meaning from this model requires defining some convenient coarse-grained quantities. The overlap of the state vector σi(t) with the μth pattern is given by (10) where −1 ≤ mμ(t) ≤ + 1. The overlap measures the similarity between the (discretized) experimental and simulated gene expression profiles, and mμ(t) = + 1 means there is perfect agreement between the simulated cell’s expression and pattern μ.

A single configuration vector σi(t) represents the expression profile of a single cell. For many cells κ, let σik(t) be the expression of gene i in cell k. Define (11) as the overlap of cell k with attractor μ. Because the microarray and RNA-seq data used here report the gene expression averaged over many cells, it is appropriate to define the population-averaged (i.e. ensemble-averaged) expression, (12) which has .

Rather than work with a continuous vector quantity like , each cell can simply be identified as being in a discrete phenotypic state at any given time. Define the state of cell k as (13) i.e. the index of the attractor with maximum overlap, which may be interpreted as cell k’s phenotype. To better understand population-level dynamics, define the discrete probability distribution Pμ(t) as the fraction of κ cells with sk(t) = μ; that is, Pμ(t) is the probability that a randomly chosen cell is in state μ at time t. Finally, define the time-averaged distribution of states as (14) for a window of time τ.

For each data set, T, λ, and the single-node update probability were tuned to the “edge of chaos” [72] such that the cyclic attractor was preserved and the time between transitions was approximately constant, but the system was sensitive enough to perturbations that some targeted inhibitions produced noticeable changes in . See S1 Table for a list of parameters used for each data set.

Gene inhibition optimization.

In this application, the goal is to identify perturbations that halt or retard the encoded cyclic attractor. A standard genetic algorithm (GA; explained in S1 Text) was employed to identify an optimal control field that maximized a given objective function , (15) where is the control vector given by (16) for a fixed number of targets (nonzero elements) ntarg. Only negative control fields are used here to simulate the effects of targeted gene inhibition from pharmaceutical compounds. The objective function used here is , meaning that the optimal control field maximizes the time-averaged number of cells occupying a particular attractor state μ. This search was conducted across all attractors μ to determine the controllability of each attractor state. Note that while the numerical value of depends on the choice of parameters such as T, λ, and the single-node update probability, varying these parameters by ±5% gave comparable results for the optimal sets of controllers.

Dynamical behavior

Fig 2 shows the time evolution of sk(t) for a single simulated cell using the attractors derived from [64]. As expected, the system progresses cyclically through the eight attractor states. The duration of each cycle varies somewhat due to the stochasticity in the update rule from Eq 2.

thumbnail
Fig 2. Unperturbed cell state versus time.

Boxes indicate sk(t), i.e. the index of the attractor with maximum overlap at time t. The system began with the configuration and was allowed to evolve according to the Hopfield signaling rules with zero external field, mapping cyclically through the set of eight attractors. The pattern and cycle durations vary due to the system’s stochasticity.

https://doi.org/10.1371/journal.pcbi.1005849.g002

Although the gene expression for each simulated cell k has σik(t) = ±1, the population-averaged expression has , and for many cells initially synchronized with for all k, successfully recovers the experimentally observed decaying sinusoidal gene expression. Fig 3 shows a comparison between the experimental expression xi(t) from the Eser data set and the mean simulated expression with κ = 50 for i = SLD2, one of the genes responsible for initiating DNA replication in S. cerevisiae [73, 74]. The simulation time t was rescaled by eye to align the simulated and experimental curves (so that 1 minute corresponds to approximately 7.5 iterations).

thumbnail
Fig 3. Measured and simulated gene expression.

The measured expression of the S. cerevisiae gene SLD2 from [64] was scaled to the range [−1, +1] and is shown as a black beaded curve, and the population-averaged expression of the same gene as defined in Eq 12 for κ = 50 cells is shown in green, with the t axis rescaled by eye to match experimental time. Transient points (red X’s) were ignored when fitting Eq 17.

https://doi.org/10.1371/journal.pcbi.1005849.g003

Trajectories can be visualized by projecting them onto the first two principal components (PCs) of the attractor configurations. Fig 4 shows the eight attractors as stars, and a single cell trajectory (left panel) and 100 cell trajectories (right panel) with random initial states as curves with line segments colored according to sk(t) (as computed in the full N-dimensional space). Although the cells begin nearly equidistant from all , they quickly relax into encoded cycle. S1 Video shows an animation of a system of κ = 50 cells with random initial conditions projected onto the same PCs, where cells (circles) are colored according to sk(t). As with the cells shown in Fig 4, all initially random configurations eventually converge to the cycle. S2 Video shows an animation of κ = 50 cell trajectories with . As time progresses, the phases of the initially synchronized cells slowly decohere because cells stochastically and independently transition between attractors due to the finite temperature in Eq 2.

thumbnail
Fig 4. Principal component projection of unperturbed cell trajectories.

The simulated single cell (left panel) and 100 cells (right panel) began with random initial states (projected near the center of the plot), but quickly settled into the encoded cycle. Line segments were colored according to sk(t), i.e. which of the eight attractors (labeled stars) had maximum overlap at time t.

https://doi.org/10.1371/journal.pcbi.1005849.g004

S3 Video demonstrates the effect of temperature on the dynamics. 50 cells were given random initial states, and the temperature was increased and decreased in steps. Cells rarely escape the eight attractor states for T = 0.045, and one cell becomes stuck near the center in a spurious attractor (unintentional metastable states that arise from the model’s nonlinearity). At T = 0.06, fluctuations allow the cells to transition somewhat regularly through the encoded cycle, and the cell trapped in the spurious attractor eventually escapes and joins the cycle. At T = 0.09 the cells begin to noticeably diverge from the eight attractor states, but still collectively display a net clockwise flow. The noise is too great for the cells to follow the cycle for T = 0.15, but lowering the temperature again returns cells to the cycle. This illustrates the fact that the cycle is preserved only for intermediate temperatures: cells become “frozen” in intended or spurious attractor states at low temperatures, but at high temperatures the noise is too great and the couplings between genes become irrelevant to the dynamics.

Optimal control fields

The GA was used to identify some effective combinations of gene targets that slowed progress through the cyclic attractor for varying numbers of targets, ntarg. The GA found, for example, that inhibiting the set of eight S. cerevisiae genes HEK2, PRR1, QRI1, RFC4, STB1, TDA7, VPS17, and ZIM17 was sufficient to trap ∼95% of cells in the μ = 7 state. The effects of this control field on the time evolution of Pμ(t) for κ = 50 and κ = 5000 are shown in Fig 5. Cells were given random, independent initial states at t = −200 (not shown), quickly settling into the cyclic attractor with evenly distributed phases so that Pμ(0 ≤ t < 200) ≈ 1/8. The control field was activated at t = 200, causing the cells to accumulate in the μ = 7 state. The field was then disabled at t = 1000, allowing the cells to resume cycling with initially synchronized phases, as shown by the sequence of oscillations in Pμ(t > 1000). The stochastic nature of the transitions causes the cells’ phases to slowly spread so that Pμ(t → ∞) ≈ 1/8, eventually returning the system to a desynchronized state.

thumbnail
Fig 5. Cell state synchronization by targeted inhibition for 50 and 5000 cells.

Cells were initialized with random states at t = −200 (not shown) and allowed to relax into the cyclic attractor so that Pμ(0 ≤ t < 200) ≈ 1/8. A set of eight genes was inhibited with an external control field over the range 200 ≤ t < 1000, fixing most cells near the μ = 7 state. After removing the control field, the cells resumed moving through the cycle with initially synchronized phases that slowly broaden. Eventually the system returns to a desynchronized state, Pμ(t → ∞) ≈ 1/8.

https://doi.org/10.1371/journal.pcbi.1005849.g005

The effects of this control field can also be visualized using a PC projection as shown in Fig 6 and S4 Video. The same set of κ = 50 trajectories from Fig 5 was projected onto the attractors’ PCs, with cells colored according to sk(t). The control field manages to fix most cells near the μ = 7 state, but as shown in the t = 910 panel in Fig 6, fluctuations occasionally push individual cells out of the μ = 7 basin and back into the cycle.

thumbnail
Fig 6. Principal component projection of 50 cell trajectories.

The trajectories used to make the κ = 50 panel of Fig 5 were projected onto the the first two principal components (PCs) of the attractor array (labeled stars). Cells (circles) are colored according to the closest attractor as computed by Eq 13. When the external field is activated, most cells become trapped in the μ = 7 state, although occasionally cells break from the group and complete another circuit before becoming trapped again. After the external field is removed, the cells eventually return to a desynchronized state. See S4 Video for an animation of these trajectories.

https://doi.org/10.1371/journal.pcbi.1005849.g006

Further searches were constrained to inhibiting between 1 and 38 kinases in HeLa cells and between 1 and 12 kinases in S. cerevisiae. For nearly all numbers of targets, the GA found μ = 2 (M phase) to be the most controllable attractor in HeLa cells and μ = 7 (S phase) in S. cerevisiae, as shown in S1 and S2 Figs, respectively. In HeLa cells, for example, an optimal score of was obtained for ntarg = 4 (0.55% of the 722 periodic genes) when inhibiting the genes AURKB, NEK1, TTK (all involved in the formation and operation of the spindle apparatus), and WEE1 (a G2/M checkpoint regulator involved with DNA damage repair). The effects of inhibiting AURKB and TTK are well documented. AURKB is targetable by the kinase inhibitor barasertib, found to be effective in combating small-cell lung cancer in vitro and in xenografts [75]. TTK (also known as MPS1) is targetable by the kinase inhibitors Mps1-IN-1 and AZ3146. Mps1-IN-1 has been found to be an effective tool to inhibit proliferation in an array of cancer cell lines, including HeLa cells [76]. AZ3146 has been effective against hepatocellular carcinoma (liver cancer) in vitro [77] and pancreatic ductal adenocarcinoma in vitro [78]. siRNAs and shRNAs have also been used to inhibit TTK in pancreatic ductal adenocarcinoma [78].

Discussion

Above we presented a delayed cyclic Hopfield model designed to store CC time series gene expression data from synchronized S. cerevisiae and HeLa cells, and the behaviors of both individual cells and populations of cells were studied. The dynamics of populations of cells successfully recreated the experimental gene expression data, including the slow decoherence of initially synchronized cells due to the stochastic transitions between attractors. Optimal control fields that freeze or stall the cyclic attractor by inhibiting only a small number of genes were identified. These predictions could be experimentally tested using kinase inhibitors, micro RNAs, short interfering RNAs, or knockout studies and confirmed simply by counting the number of cells as a function of time compared to a control with no inhibitors.

Admittedly, there are several limitations to this model. Additionally, although using the temporal ordering of the time series gene expression samples provides more information about potentially causal relationships than static samples, the Hopfield model is ultimately an effective model that builds gene-gene couplings from pairwise correlations in gene expression, thereby capturing direct, indirect, and spurious relationships between genes. Independently derived network information with experimentally confirmed molecular regulatory interactions could perhaps be used to refine the construction of the coupling matrix. Other data sets that explore a more diverse set of expression states other than a simple cycle could also reduce the number of spurious couplings. We are currently developing these kinds of non-cyclic models and time series data sets in a separate project.

Our approach can be generalized and improved in many ways. This incarnation of the model causes simulated cells to continuously undergo CC with no G0 (resting) phase. Adding a relatively stable G0 attractor between the M and G1 phases could cause cells to pause between cycles. A GA search could then be conducted to find the best sets of inhibition targets to freeze cells in the G0 state, or to find the best sets of targets to stimulate entry into CC, mimicking the effects of environmental signals such as growth factors.

We chose to discretize the continuous gene expression data using a traditional two-state model, which assumes that each gene is either fully activated or fully deactivated. Using a multi-level Hopfield model [79] could better reflect the continuous nature of gene expression data and potentially improve the search results. This model can also incorporate additional omics information, e.g. proteomics and metabolomics, simply by increasing the number of nodes in the system. We plan to explore this option as more multi-omics time series data sets become available. Single-cell experimental techniques and analytical tools are also rapidly improving in quality, decreasing in cost, and gaining in popularity [8082], and using techniques like pseudo-temporal ordering [83] could allow the Hopfield model to encode single-cell RNA-seq data as well.

Although the above simulated populations of cells exhibit intriguing dynamical and statistical properties, they behave as completely homogeneous, non-interacting particles. The importance of cell-cell communication and interactions in populations of cells has been demonstrated in a variety of systems including bacterial quorum sensing [84] and community spatial patterning [85], neuron synchronization in circadian rhythm [86], and various forms of cancer [8791]. As with many nonlinear systems, even seemingly minor changes can produce dramatically different outcomes. More complex extensions to our model could incorporate cell-cell communication by, for example, adding couplings between known signaling molecules and receptors between different cells, and could even allow for interactions between heterogeneous cell types. This would increase the computational complexity of the model, but could better reflect the underlying biology.

Methods

Gene expression fitting

In order to encode these CC data sets into the Hopfield model, periodic genes needed to be identified, their frequencies and phases computed, and their expression converted from continuous to Boolean form. SciPy’s Trust Region Reflective method [92] was used to identify genes i with periodic expression xi(t) by fitting to the form (17) for amplitude ai, decay rate bi, angular frequency ωi, phase ϕi, and asymptotic mean expression xi0. The first several time points were ignored to avoid fitting the parameters of Eq 17 to chemically perturbed (transient) states. A gene was labeled periodic if the maximum relative uncertainty in its parameters from the fit, (18) was less than the thresholds defined in S1 Table. (Note that because the HeLa data covers only a little more than one period, many of the bi’s were approximately zero. This resulted in extremely large despite the remaining variables having small uncertainty. The δbi/bi term was thus ignored for the HeLa data.) Once all frequencies {ωi} for periodic genes were computed, the frequency was fixed to the mean frequency and the fits were recomputed for each periodic gene using the form (19) thus producing the final set of continuous phases {ϕi}. Fig 1A shows a heat map of the expression of all periodic genes detected in the Eser data set sorted by their fitted phases, and Fig 1B shows the same genes with the fitted expression curves. These fitted curves were converted from continuous values xi(t) ≥ 0 to Boolean values ξi(t) = ±1 (over/underexpressed) by assigning (20) as shown in Fig 1C. Finally, one CC period was divided into eight uniformly spaced states . These states, shown in Fig 1D, were used as attractors in the Hopfield model.

Determining cell cycle phase

The approximate CC phase for each attractor μ in the HeLa data set was determined using a set of marker genes [65, 93] and over-representation analysis using a hypergeometric distribution to calculate p-values with the Benjamini-Hochberg procedure [94] to correct for multiple hypothesis testing. All genes i with and were used as the μth input set, and all detected cyclic genes were used as the background. μ = 2 showed the activation of the G2/M checkpoint marker CCNA2 and was enriched for the gene ontology (GO) term cell division (p = 3.1 × 10−2). μ = 3 showed the activation of the G2, G2/M, and M markers BUB1, CCNB1, NEK2, and PLK1, and was enriched for signaling by Rho GTPases, cell division, midbody, mitotic prometaphase, mitotic nuclear division, and small GTPase mediated signal transduction (p = 3.7 × 10−3); M phase (p = 6.2 × 10−3); resolution of sister chromatid cohesion (p = 7.6 × 10−3); mitotic anaphase, spindle pole, and RHO GTPase effectors (p = 3.0 × 10−2); and mitotic metaphase and anaphase (p = 3.8 × 10−2). μ = 6 expressed the late G1, G1/S, and S markers CCNE1, CDC6, E2F1, MCM2, and SLBP, and was enriched for DNA replication initiation (p = 5.9 × 10−4); G1/S-specific transcription, G1/S transition, DNA replication pre-initiation, M/G1 transition, and assembly of the pre-replicative complex (p = 4.3 × 10−2); and G1/S transition of mitotic cell cycle (p = 4.4 × 10−2). μ = 7 showed the activation of the G1/S and S markers CDK2 and CDKN2AIP. The database yeastgenome.org [95] was used to determine the CC phases for the Eser S. cerevisiae data set. μ = 0 was enriched for the GO term DNA replication (p = 2.02 × 10−12), indicating an attractor in the S phase of CC. μ = 2 was enriched for mitotic spindle organization (p = 2.3 × 10−3) indicating the beginning of mitosis in S. cerevisiae. μ = 6 from Eser was enriched for the GO term pre-replicative complex assembly involved in nuclear cell cycle DNA replication (p = 3.0 × 10−5), indicating an attractor at the end of G1 phase as the cells prepare for DNA replication.

Supporting information

S1 Video. 50 cell trajectories with random initial conditions.

Data was projected onto the first two principle components of the attractor array . Attractors are shown as stars, and cells are shown as circles. Cell colors are assigned using sk(t) as measured in the full N-dimensional space. All cells k were given random initial conditions σik = ±1 with equal probability for all i and k, but eventually converge to the cyclic attractor.

https://doi.org/10.1371/journal.pcbi.1005849.s001

(AVI)

S2 Video. 50 cell trajectories with identical initial conditions.

See the caption of S1 Video for an explanation of the projection and colors. All cells k were initially synchronized with , but progress through the cycle stochastically, causing the distribution of sk(t) to broaden.

https://doi.org/10.1371/journal.pcbi.1005849.s002

(AVI)

S3 Video. Effects of temperature on 50 cell trajectories.

See the caption of S1 Video for an explanation of the projection and colors. All cells were given initial random states, and the temperature was increased and decreased in steps as shown in the top panel.

https://doi.org/10.1371/journal.pcbi.1005849.s003

(AVI)

S4 Video. Principal component projection of 50 cells being synchronized.

See the caption of S1 Video for an explanation of the projection and colors. This video is an animation of the trajectories used in Figs 5 and 6.

https://doi.org/10.1371/journal.pcbi.1005849.s004

(AVI)

S1 Text. Explanation of genetic algorithm.

https://doi.org/10.1371/journal.pcbi.1005849.s005

(PDF)

S1 Table. List of parameters used for each data set.

https://doi.org/10.1371/journal.pcbi.1005849.s006

(XLSX)

S2 Table. Enrichment results for each attractor from the Eser data set.

https://doi.org/10.1371/journal.pcbi.1005849.s007

(XLSX)

S3 Table. Enrichment results for each attractor from the Dominguez data set.

https://doi.org/10.1371/journal.pcbi.1005849.s008

(XLSX)

S1 Fig. GA’s optimal control results versus the number of inhibited kinases for HeLa cells.

Each point indicates the best measured score (mean fraction of cells) for the given attractor and number of targets. μ = 2 (M phase) is generally the most controllable attractor in HeLa cells.

https://doi.org/10.1371/journal.pcbi.1005849.s009

(PNG)

S2 Fig. GA’s optimal control results versus the number of inhibited kinases for S. cerevisiae.

Each point indicates the best measured score (mean fraction of cells) for the given attractor and number of targets. μ = 7 (S phase) is generally the most controllable attractor in S. cerevisiae.

https://doi.org/10.1371/journal.pcbi.1005849.s010

(PNG)

S1 File. Raw yeast time series gene expression data.

Taken from [64].

https://doi.org/10.1371/journal.pcbi.1005849.s011

(CSV)

S2 File. Raw HeLa time series gene expression data.

Taken from [65].

https://doi.org/10.1371/journal.pcbi.1005849.s012

(XLSX)

Acknowledgments

We would like to thank George Mias (Michigan State University) and Yunyi Kang (Sanford Burnham Prebys Medical Discovery Institute) for helpful discussions and recommendations.

References

  1. 1. Waddington CH (2014) The Strategy of the Genes, volume 20. Routledge.
  2. 2. Kauffman S (1971) Differentiation of malignant to benign cells. Journal of Theoretical Biology 31: 429–451. pmid:5556142
  3. 3. Huang S, Eichler G, Bar-Yam Y, Ingber DE (2005) Cell fates as high-dimensional attractor states of a complex gene regulatory network. Physical Review Letters 94: 128701. pmid:15903968
  4. 4. Alvarez-Buylla ER, Chaos Á, Aldana M, Benítez M, Cortes-Poza Y, et al. (2008) Floral morphogenesis: stochastic explorations of a gene network epigenetic landscape. PLOS ONE 3: e3626. pmid:18978941
  5. 5. Huang S, Ernberg I, Kauffman S (2009) Cancer attractors: a systems view of tumors from a gene network dynamics and developmental perspective. In: Seminars in Cell & Developmental Biology. Elsevier, volume 20, pp. 869–876.
  6. 6. Demicheli R, Coradini D (2010) Gene regulatory networks: a new conceptual framework to analyse breast cancer behaviour. Annals of Oncology 22: 1259–1265. pmid:21109571
  7. 7. Choi M, Shi J, Jung SH, Chen X, Cho KH (2012) Attractor landscape analysis reveals feedback loops in the p53 network that control the cellular response to DNA damage. Sci Signal 5: ra83–ra83.
  8. 8. Yuan R, Zhu X, Radich JP, Ao P (2016) From molecular interaction to acute promyelocytic leukemia: Calculating leukemogenesis and remission from endogenous molecular-cellular network. Scientific Reports 6.
  9. 9. Pusuluri ST (2017) Exploring Neural Network Models with Hierarchical Memories and Their Use in Modeling Biological Systems. Ph.D. thesis, Ohio University.
  10. 10. Udyavar AR, Wooten DJ, Hoeksema M, Bansal M, Califano A, et al. (2017) Novel hybrid phenotype revealed in small cell lung cancer by a transcription factor network model that can explain tumor heterogeneity. Cancer Research 77: 1063–1074. pmid:27932399
  11. 11. Wooten DJ, Quaranta V (2017) Mathematical models of cell phenotype regulation and reprogramming: Make cancer cells sensitive again! Biochimica et Biophysica Acta (BBA)-Reviews on Cancer 1867: 167–175.
  12. 12. Yuan R, Zhu X, Wang G, Li S, Ao P (2017) Cancer as robust intrinsic state shaped by evolution: a key issues review. Reports on Progress in Physics 80: 042701. pmid:28212112
  13. 13. Hopfield JJ (1982) Neural networks and physical systems with emergent collective computational abilities. Proceedings of the National Academy of Sciences 79: 2554–2558.
  14. 14. Peretto P (1988) On learning rules and memory storage abilities of asymmetrical neural networks. Journal de Physique 49: 711–726.
  15. 15. Kanter I, Sompolinsky H (1987) Associative recall of memory without errors. Physical Review A 35: 380.
  16. 16. Amit DJ, Gutfreund H, Sompolinsky H (1985) Storing infinite numbers of patterns in a spin-glass model of neural networks. Physical Review Letters 55: 1530. pmid:10031847
  17. 17. Derrida B, Gardner E, Zippelius A (1987) An exactly solvable asymmetric neural network model. Europhysics Letters 4: 167.
  18. 18. Nishimori H, Nakamura T, Shiino M (1990) Retrieval of spatio-temporal sequence in asynchronous neural network. Physical Review A 41: 3346.
  19. 19. Personnaz L, Guyon I, Dreyfus G (1986) Collective computational properties of neural networks: New learning mechanisms. Physical Review A 34: 4217.
  20. 20. Nasrabadi NM, Li W (1991) Object recognition by a Hopfield neural network. IEEE Transactions on Systems, Man, and Cybernetics 21: 1523–1535.
  21. 21. Cote S, Tatnall A (1997) The Hopfield neural network as a tool for feature tracking and recognition from satellite sensor images. International Journal of Remote Sensing 18: 871–885.
  22. 22. Maetschke SR, Ragan MA (2014) Characterizing cancer subtypes as attractors of Hopfield networks. Bioinformatics 30: 1273–1279. pmid:24407221
  23. 23. Lang AH, Li H, Collins JJ, Mehta P (2014) Epigenetic landscapes explain partially reprogrammed cells and identify key reprogramming genes. PLOS Comput Biol 10: e1003734. pmid:25122086
  24. 24. Jang S, Choubey S, Furchtgott L, Zou LN, Doyle A, et al. (2017) Dynamics of embryonic stem cell differentiation inferred from single-cell transcriptomics show a series of transitions through discrete cell states. eLife 6: e20487. pmid:28296635
  25. 25. Szedlak A, Paternostro G, Piermarocchi C (2014) Control of asymmetric Hopfield networks and application to cancer attractors. PLOS ONE 9: e105842. pmid:25170874
  26. 26. Hannam R, Annibale A, Kuehn R (2017) Cell reprogramming modelled as transitions in a hierarchy of cell cycles. Journal of Physics A: Mathematical and Theoretical.
  27. 27. Wang Z, Gerstein M, Snyder M (2009) RNA-Seq: a revolutionary tool for transcriptomics. Nature Reviews Genetics 10: 57–63. pmid:19015660
  28. 28. Marguerat S, Bähler J (2010) RNA-seq: from technology to biology. Cellular and Molecular Life Sciences 67: 569–579. pmid:19859660
  29. 29. Trapnell C, Roberts A, Goff L, Pertea G, Kim D, et al. (2012) Differential gene and transcript expression analysis of RNA-seq experiments with TopHat and Cufflinks. Nature Protocols 7: 562–578. pmid:22383036
  30. 30. Margolin AA, Nemenman I, Basso K, Wiggins C, Stolovitzky G, et al. (2006) ARACNE: an algorithm for the reconstruction of gene regulatory networks in a mammalian cellular context. BMC Bioinformatics 7: S7.
  31. 31. Hecker M, Lambeck S, Toepfer S, Van Someren E, Guthke R (2009) Gene regulatory network inference: data integration in dynamic models—a review. Biosystems 96: 86–103. pmid:19150482
  32. 32. Lee I, Blom UM, Wang PI, Shim JE, Marcotte EM (2011) Prioritizing candidate disease genes by network-based boosting of genome-wide association data. Genome Research 21: 1109–1121. pmid:21536720
  33. 33. Marbach D, Costello JC, Küffner R, Vega NM, Prill RJ, et al. (2012) Wisdom of crowds for robust gene network inference. Nature Methods 9: 796–804. pmid:22796662
  34. 34. Cahan P, Li H, Morris S, LummertzdaRocha E, Daley G, et al. (2014) CellNet: network biology applied to stem cell engineering. Cell 158: 903–915. pmid:25126793
  35. 35. Ong E, Szedlak A, Kang Y, Smith P, Smith N, et al. (2015) A scalable method for molecular network reconstruction identifies properties of targets and mutations in acute myeloid leukemia. Journal of Computational Biology 22: 266–288. pmid:25844667
  36. 36. Szedlak A, Smith N, Liu L, Paternostro G, Piermarocchi C (2016) Evolutionary and topological properties of genes and community structures in human gene regulatory networks. PLOS Computational Biology 12: e1005009. pmid:27359334
  37. 37. Chen R, Mias GI, Li-Pook-Than J, Jiang L, Lam HY, et al. (2012) Personal omics profiling reveals dynamic molecular and medical phenotypes. Cell 148: 1293–1307. pmid:22424236
  38. 38. Takahashi M (1966) Theoretical basis for cell cycle analysis i. labelled mitosis wave method. Journal of Theoretical Biology 13: 202–211.
  39. 39. Takahashi M (1968) Theoretical basis for cell cycle analysis: Ii. further studies on labelled mitosis wave method. Journal of theoretical biology 18: 195–209. pmid:5647131
  40. 40. Lloyd D, Lloyd AL, Olsen LF (1992) The cell division cycle: a physiologically plausible dynamic model can exhibit chaotic solutions. Biosystems 27: 17–24. pmid:1391688
  41. 41. Panetta JC, Adam J (1995) A mathematical model of cycle-specific chemotherapy. Mathematical and computer modelling 22: 67–82.
  42. 42. Swierniak A, Polanski A, Kimmel M (1996) Optimal control problems arising in cell-cycle-specific cancer chemotherapy. Cell proliferation 29: 117–139. pmid:8652742
  43. 43. Chiorino G, Metz J, Tomasoni D, Ubezio P (2001) Desynchronization rate in cell populations: mathematical modeling and experimental data. Journal of Theoretical Biology 208: 185–199. pmid:11162063
  44. 44. Tyson JJ, Novak B (2001) Regulation of the eukaryotic cell cycle: molecular antagonism, hysteresis, and irreversible transitions. Journal of theoretical biology 210: 249–263. pmid:11371178
  45. 45. Basse B, Baguley BC, Marshall ES, Joseph WR, van Brunt B, et al. (2003) A mathematical model for analysis of the cell cycle in cell lines derived from human tumors. Journal of mathematical biology 47: 295–312. pmid:14523574
  46. 46. Alarcon T, Byrne H, Maini P (2004) A mathematical model of the effects of hypoxia on the cell-cycle of normal and cancer cells. Journal of theoretical biology 229: 395–411. pmid:15234206
  47. 47. Chen KC, Calzone L, Csikasz-Nagy A, Cross FR, Novak B, et al. (2004) Integrative analysis of cell cycle control in budding yeast. Molecular biology of the cell 15: 3841–3862. pmid:15169868
  48. 48. Li F, Long T, Lu Y, Ouyang Q, Tang C (2004) The yeast cell-cycle network is robustly designed. Proceedings of the National Academy of Sciences of the United States of America 101: 4781–4786. pmid:15037758
  49. 49. Steuer R (2004) Effects of stochasticity in models of the cell cycle: from quantized cycle times to noise-induced oscillations. Journal of theoretical biology 228: 293–301. pmid:15135028
  50. 50. Fauré A, Naldi A, Chaouiya C, Thieffry D (2006) Dynamical analysis of a generic boolean model for the control of the mammalian cell cycle. Bioinformatics 22: e124–e131.
  51. 51. Ferrell JE, Pomerening JR, Kim SY, Trunnell NB, Xiong W, et al. (2009) Simple, realistic models of complex biological processes: positive feedback and bistability in a cell fate switch and a cell cycle oscillator. FEBS letters 583: 3999–4005. pmid:19878681
  52. 52. Kar S, Baumann WT, Paul MR, Tyson JJ (2009) Exploring the roles of noise in the eukaryotic cell cycle. Proceedings of the National Academy of Sciences 106: 6471–6476.
  53. 53. Barik D, Baumann WT, Paul MR, Novak B, Tyson JJ (2010) A model of yeast cell-cycle regulation based on multisite phosphorylation. Molecular Systems Biology 6: 405. pmid:20739927
  54. 54. Ferrell JE, Tsai TYC, Yang Q (2011) Modeling the cell cycle: why do certain circuits oscillate? Cell 144: 874–885. pmid:21414480
  55. 55. Luo X, Xu L, Wang J (2017) Funneled potential and flux landscapes dictate the stabilities of both the states and the flow: Fission yeast cell cycle. Advanced online publication.
  56. 56. Assaf M, Roberts E, Luthey-Schulten Z, Goldenfeld N (2013) Extrinsic noise driven phenotype switching in a self-regulating gene. Physical Review Letters 111: 058102. pmid:23952448
  57. 57. Evan GI, Vousden KH (2001) Proliferation, cell cycle and apoptosis in cancer. Nature 411: 342. pmid:11357141
  58. 58. Malumbres M, Barbacid M (2009) Cell cycle, CDKs and cancer: a changing paradigm. Nature Reviews Cancer 9: 153. pmid:19238148
  59. 59. Hanahan D, Weinberg RA (2011) Hallmarks of cancer: the next generation. Cell 144: 646–674. pmid:21376230
  60. 60. Diaz-Moralli S, Tarrado-Castellarnau M, Miranda A, Cascante M (2013) Targeting cell cycle regulation in cancer therapy. Pharmacology & Therapeutics 138: 255–271.
  61. 61. Feala JD, Cortes J, Duxbury PM, Piermarocchi C, McCulloch AD, et al. (2010) Systems approaches and algorithms for discovery of combinatorial therapies. Wiley Interdisciplinary Reviews: Systems Biology and Medicine 2: 181–193. pmid:20836021
  62. 62. Csermely P, Agoston V, Pongor S (2005) The efficiency of multi-target drugs: the network approach might help drug design. Trends in Pharmacological Sciences 26: 178–182. pmid:15808341
  63. 63. Ágoston V, Csermely P, Pongor S (2005) Multiple weak hits confuse complex systems: a transcriptional regulatory network as an example. Physical Review E 71: 051909.
  64. 64. Eser P, Demel C, Maier KC, Schwalb B, Pirkl N, et al. (2014) Periodic mRNA synthesis and degradation co-operate during cell cycle gene expression. Molecular Systems Biology 10: 717. pmid:24489117
  65. 65. Dominguez D, Tsai YH, Gomez N, Jha DK, Davis I, et al. (2016) A high-resolution transcriptome map of cell cycle reveals novel connections between periodic genes and cancer. Cell Research. pmid:27364684
  66. 66. Spellman PT, Sherlock G, Zhang MQ, Iyer VR, Anders K, et al. (1998) Comprehensive identification of cell cycle–regulated genes of the yeast saccharomyces cerevisiae by microarray hybridization. Molecular biology of the cell 9: 3273–3297. pmid:9843569
  67. 67. Whitfield ML, Sherlock G, Saldanha AJ, Murray JI, Ball CA, et al. (2002) Identification of genes periodically expressed in the human cell cycle and their expression in tumors. Molecular biology of the cell 13: 1977–2000. pmid:12058064
  68. 68. De Lichtenberg U, Jensen LJ, Fausbøll A, Jensen TS, Bork P, et al. (2004) Comparison of computational methods for the identification of cell cycle-regulated genes. Bioinformatics 21: 1164–1171. pmid:15513999
  69. 69. Rustici G, Mata J, Kivinen K, Lió P, Penkett CJ, et al. (2004) Periodic gene expression program of the fission yeast cell cycle. Nature genetics 36: 809. pmid:15195092
  70. 70. Kelliher CM, Leman AR, Sierra CS, Haase SB (2016) Investigating conservation of the cell-cycle-regulated transcriptional program in the fungal pathogen, cryptococcus neoformans. PLoS genetics 12: e1006453. pmid:27918582
  71. 71. Kohonen T (1974) An adaptive associative memory principle. IEEE Transactions on Computers 100: 444–445.
  72. 72. Kauffman SA, Johnsen S (1991) Coevolution to the edge of chaos: coupled fitness landscapes, poised states, and coevolutionary avalanches. Journal of Theoretical Biology 149: 467–505. pmid:2062105
  73. 73. Muramatsu S, Hirai K, Tak YS, Kamimura Y, Araki H (2010) CDK-dependent complex formation between replication proteins Dpb11, Sld2, Pol ϵ, and GINS in budding yeast. Genes & Development 24: 602–612.
  74. 74. Bruck I, Kaplan DL (2014) The replication initiation protein Sld2 regulates helicase assembly. Journal of Biological Chemistry 289: 1948–1959. pmid:24307213
  75. 75. Helfrich BA, Kim J, Gao D, Chan DC, Zhang Z, et al. (2016) Barasertib (azd1152), a small molecule aurora b inhibitor, inhibits the growth of sclc cell lines in vitro and in vivo. Molecular cancer therapeutics 15: 2314–2322. pmid:27496133
  76. 76. Kwiatkowski N, Jelluma N, Filippakopoulos P, Soundararajan M, Manak MS, et al. (2010) Small-molecule kinase inhibitors provide insight into mps1 cell cycle function. Nature chemical biology 6: 359–368. pmid:20383151
  77. 77. Liu X, Liao W, Yuan Q, Ou Y, Huang J (2015) Ttk activates akt and promotes proliferation and migration of hepatocellular carcinoma cells. Oncotarget 6: 34309. pmid:26418879
  78. 78. Stratford JK, Yan F, Hill RA, Major MB, Graves LM, et al. (2017) Genetic and pharmacological inhibition of ttk impairs pancreatic cancer cell line growth by inducing lethal chromosomal instability. PloS one 12: e0174863. pmid:28380042
  79. 79. Zurada JM, Cloete I, Van der Poel E (1996) Generalized Hopfield networks for associative memories with multi-valued stable states. Neurocomputing 13: 135–149.
  80. 80. Patel AP, Tirosh I, Trombetta JJ, Shalek AK, Gillespie SM, et al. (2014) Single-cell RNA-seq highlights intratumoral heterogeneity in primary glioblastoma. Science 344: 1396–1401. pmid:24925914
  81. 81. Scialdone A, Natarajan KN, Saraiva LR, Proserpio V, Teichmann SA, et al. (2015) Computational assignment of cell-cycle stage from single-cell transcriptome data. Methods 85: 54–61. pmid:26142758
  82. 82. Zheng GX, Terry JM, Belgrader P, Ryvkin P, Bent ZW, et al. (2017) Massively parallel digital transcriptional profiling of single cells. Nature Communications 8: 14049. pmid:28091601
  83. 83. Trapnell C, Cacchiarelli D, Grimsby J, Pokharel P, Li S, et al. (2014) Pseudo-temporal ordering of individual cells reveals dynamics and regulators of cell fate decisions. Nature Biotechnology 32: 381.
  84. 84. Waters CM, Bassler BL (2005) Quorum sensing: cell-to-cell communication in bacteria. Annu Rev Cell Dev Biol 21: 319–346. pmid:16212498
  85. 85. Momeni B, Brileya KA, Fields MW, Shou W (2013) Strong inter-population cooperation leads to partner intermixing in microbial communities. eLife 2: e00230. pmid:23359860
  86. 86. Kalsbeek A, Merrow M, Roenneberg T, Foster R (2012) Suprachiasmatic nucleus: cellular clocks and networks. The Neurobiology of Circadian Timing 199: 129.
  87. 87. Skog J, Würdinger T, Van Rijn S, Meijer DH, Gainche L, et al. (2008) Glioblastoma microvesicles transport RNA and proteins that promote tumour growth and provide diagnostic biomarkers. Nature Cell Biology 10: 1470–1476. pmid:19011622
  88. 88. Hong BS, Cho JH, Kim H, Choi EJ, Rho S, et al. (2009) Colorectal cancer cell-derived microvesicles are enriched in cell cycle-related mRNAs that promote proliferation of endothelial cells. BMC Genomics 10: 556. pmid:19930720
  89. 89. Renzulli JF, Del Tatto M, Dooner G, Aliotta J, Goldstein L, et al. (2010) Microvesicle induction of prostate specific gene expression in normal human bone marrow cells. The Journal of Urology 184: 2165–2171. pmid:20850816
  90. 90. Del Tatto M, Ng T, Aliotta JM, Colvin GA, Dooner MS, et al. (2011) Marrow cell genetic phenotype change induced by human lung cancer cells. Experimental Hematology 39: 1072–1080. pmid:21864488
  91. 91. Tetta C, Ghigo E, Silengo L, Deregibus MC, Camussi G (2013) Extracellular vesicles as an emerging mechanism of cell-to-cell communication. Endocrine 44: 11–19. pmid:23203002
  92. 92. Jones E, Oliphant T, Peterson P, et al. (2001–). SciPy: Open source scientific tools for Python. URL http://www.scipy.org/.
  93. 93. Whitfield ML, George LK, Grant GD, Perou CM (2006) Common markers of proliferation. Nature reviews Cancer 6: 99. pmid:16491069
  94. 94. Hochberg Y, Benjamini Y (1990) More powerful procedures for multiple significance testing. Statistics in Medicine 9: 811–818. pmid:2218183
  95. 95. Cherry JM, Hong EL, Amundsen C, Balakrishnan R, Binkley G, et al. (2011) Saccharomyces Genome Database: the genomics resource of budding yeast. Nucleic Acids Research 40: D700–D705.