Hostname: page-component-848d4c4894-pftt2 Total loading time: 0 Render date: 2024-05-08T20:27:27.236Z Has data issue: false hasContentIssue false

Boolean modelling in plant biology

Published online by Cambridge University Press:  20 December 2022

Aravind Karanam
Affiliation:
Department of Physics, University of California, San Diego, La Jolla, California 92093, USA
Wouter-Jan Rappel*
Affiliation:
Department of Physics, University of California, San Diego, La Jolla, California 92093, USA
*
Author for correspondence: W.-J. Rappel, E-mail: rappel@physics.ucsd.edu

Abstract

Signalling and genetic networks underlie most biological processes and are often complex, containing many highly connected components. Modelling these networks can provide insight into mechanisms but is challenging given that rate parameters are often not well defined. Boolean modelling, in which components can only take on a binary value with connections encoded by logic equations, is able to circumvent some of these challenges, and has emerged as a viable tool to probe these complex networks. In this review, we will give an overview of Boolean modelling, with a specific emphasis on its use in plant biology. We review how Boolean modelling can be used to describe biological networks and then discuss examples of its applications in plant genetics and plant signalling.

Type
Review
Creative Commons
Creative Common License - CCCreative Common License - BY
This is an Open Access article, distributed under the terms of the Creative Commons Attribution licence (https://creativecommons.org/licenses/by/4.0/), which permits unrestricted re-use, distribution, and reproduction in any medium, provided the original work is properly cited.
Copyright
© The Author(s), 2022. Published by Cambridge University Press in association with The John Innes Centre

1. Introduction

Many cellular processes in biology are controlled by a large number of components that are part of complex signalling networks (Kitano, Reference Kitano2002). Examples include the pathways controlling cell polarity, cell motility, cell division and differentiation as well as the gene networks that underlie a myriad of biological processes. The biological function in question frequently arises out of the connections and dependencies among physical and chemical processes which may be relatively simple and well understood. Technological advances in the last few decades have contributed to a proliferation of data at the level of individual genes and metabolites (International Human Genome Sequencing Consortium, 2001; Tyers & Mann, Reference Tyers and Mann2003), paving the way for synthesising the knowledge to produce a systems-level understanding.

Models of biological networks attempt to recast the systems in a mathematical form and their level of detail depends on the amount of available data as well as its requirements (Karlebach & Shamir, Reference Karlebach and Shamir2008). In its optimal form, quantitative modelling can replace often laborious experiments by carrying out in silico experiments during which one or more components of the pathway or the interactions between components are altered. Even if this is not possible, modelling can often reveal the role of a particular component in the pathway and can, thus, predict the effect of removing or making it constitutively active.

Constructing models for biological pathways requires knowledge about their topology. In other words, one needs to know whether component A affects component B. This is equivalent to answering the question whether B is downstream or upstream of A. Furthermore, the ‘sign’ of the interaction between these components is required: does A activate (corresponding to a positive interaction) or inhibit (negative interaction) B? Ideally, one would also like to know the strength of the interaction: how much will B increase or decrease when A is present?

For pathways in which all connections and strengths are known, it is possible to construct a mathematical model that represents concentrations of the pathway components as continuous quantities that can take on all positive values. This type of model can provide significant insights, particularly for small systems made up of a handful of simple reactions where all the interactions are known (Ouellet et al., Reference Ouellet, Laidler and Morales1952; Pollard, Reference Pollard1986). For these systems, parameters like rate constants, dissociation and association constants can be inferred by monitoring the formation of the product or the decay of the substrate. Often, however, and especially for pathways that contain many components, it is not possible to quantify the type of interaction and the strength between the different components. After all, quantifying this for, say, A and B typically requires a systematic variation of the level A and measuring the response in B. This type of experiment is not always possible for all components and models with a large number of unknown parameters, and interactions can quickly lose their predictive and mechanistic value.

An alternative to creating continuous models is to construct Boolean models (Schwab et al., Reference Schwab, Kühlwein, Ikonomi, Kühl and Kestler2020). In a Boolean model, each element (alternatively called a node) can only take on one of two values: 0 and 1. The dynamics of these nodes are no longer determined by solving equations that involve rate constants but are updated using logical operations. These operations encode the connections between the different components using the elementary logical functions: identity, and, or and not. A Boolean network is then obtained by connecting a number of such nodes in a meaningful manner. Despite these significant simplifications, Boolean networks have been shown to be able to provide insights into genetic networks (Herrmann et al., Reference Herrmann, Groß, Zhou, Kestler and Kühl2012; Kauffman, Reference Kauffman1969b; Shmulevich et al., Reference Shmulevich, Dougherty and Zhang2002b; Thomas, Reference Thomas1973), protein networks (Bornholdt, Reference Bornholdt2008) and cellular regulatory networks (Lau et al., Reference Lau, Ganguli and Tang2007; Li et al., Reference Li, Long, Lu, Ouyang and Tang2004). More importantly, from a practical point of view, Boolean models have several advantages: they can be simulated relatively quickly, even on daily-use desktop computers, and several software packages are freely available (Karanam et al., Reference Karanam, He, Hsu, Schulze, Dubeaux, Karmakar, Schroeder and Rappel2021; Schwab et al., Reference Schwab, Kühlwein, Ikonomi, Kühl and Kestler2020); modifying the network and simulating variants of the original network are easy tasks, and thus Boolean modelling can be used to (a) generate hypotheses that can be tested by experiments and (b) systematically explore variants of a network that ‘predict’ or lead to an observed phenotype. These ideas have been explored in several studies as will be described later (Karanam et al., Reference Karanam, He, Hsu, Schulze, Dubeaux, Karmakar, Schroeder and Rappel2021; Maheshwari et al., Reference Maheshwari, Du, Sheen, Assmann and Albert2019).

In this review, we focus on Boolean modelling in plant biology. We start with a brief overview of Boolean logic and how one can deduce a Boolean network from rate equations as well as from experimental data. We then discuss software packages that can be used to simulate Boolean networks, after which we discuss applications of Boolean modelling to gene regulatory networks (GRNs) in plants. We then review how Boolean modelling can be used to probe the pathways in guard cells that lead to stomatal closure in response to the plant hormone abscisic acid (ABA) and carbon dioxide ( $\mathrm {CO}_{2}$ ), and end with a brief conclusion and outlook.

2. Boolean logic and networks

In this section, we will first describe in more detail how Boolean equations are evaluated, provide a simple example, and show how truth tables are a convenient way to analyse and comprehend small Boolean networks. We will then describe how a Boolean network can be constructed from experimental data and describe the various updating schemes developed for this type of network. We will also show how one can translate rate equations into Boolean equations and finish by discussing available software for the simulation of Boolean networks.

2.1. Truth tables

As we mentioned in Section 1, the nodes in a Boolean network can only take on values 0 (OFF) and 1 (ON). The ON state of a variable corresponds to high activity or concentration and the OFF state corresponds to low activity or concentration. The interactions between the nodes are given by a combination of the logical functions and, or and not acting on the input nodes that feed into the output node. The future state of the output node (say at time $t+1$ ) is obtained by evaluating its corresponding Boolean function that takes the current states (say at time t) of its input nodes as inputs. To simplify notation, we write the output node and the update rule together as an equation, commonly known as the update equation of the output node. We do not explicitly specify time because (a) the update rules do not change with time and (b) the states of the input nodes specify, through the update equation, the state of the output node in the succeeding time step only.

Fig. 1. (a) Comparison of the output of a continuous model [equation (1)] and a Boolean model [equation (2)] for the activation of a gene. In the former, the output can take on any value between 0 and 1 and depends on the model parameters, whereas in the latter, the output is either 0 or 1 and is independent of parameters. (b–d) Truth tables of elementary Boolean functions. (b) Identity gate, which copies the value of the input to the output; not gate, which copies the inverted value of the input to the output. (c) or and and gates, which take two inputs. (d) An example of a Boolean function that is a combination of the elementary functions. The output X can be determined by evaluating the parts recursively.

As a simple example, consider the activation of gene B by a transcription factor A. In this case, when the concentration of A is high, the gene is ON, whereas when it is low, B is OFF. This process can be mathematically expressed using an ordinary differential equation, which describes the rate of change of B, $dB/dt$ , as a function of the concentration of A. In its simplest form, this differential equation is written as

$$ \begin{align*}\frac{dB}{dt} = f(A) - \gamma B.\end{align*} $$

Here, $\gamma $ is a degradation constant, determining how B is removed, and the function $f(A)$ describes how the production rate of the gene depends on the transcription factor concentration A. This function is often taken to be a Hill function $f(A)=\beta A^n/(A^n+K^n)$ , with n the (integer) Hill coefficient, $\beta $ the maximum production rate and K the activation coefficient. If we take n to be very large, we can approximate $f(A)$ to be a so-called step function: $f(A)=0$ if $A<K$ and $f(A)=\beta $ if $A\ge K$ . Thus, when $A<K$ , B will be 0, while for $A>K$ , the time dependence of B is found by solving the differential equation

(1) $$ \begin{align} \frac{dB}{dt} = \beta - \gamma B. \end{align} $$

The steady-state value, achieved after a long time, can be found by setting the left-hand side of this equation to zero, resulting in $B\,{=}\,\gamma /\beta $ . Furthermore, assuming that A is set above the threshold value K at $t=0$ , the solution of this equation can be found to be ${B(t)=\frac {\gamma }{\beta }(1-e^{-\gamma t})}$ . This solution is shown in Figure 1a, where we plot B as function of time for different values of the degradation constant and using $\gamma /\beta =1$ for simplicity. When the transcription factor is turned ON, B approaches its steady-state value at a timescale that depends on $\gamma $ . In this description of gene activation, B can take on all possible values between 0 and 1.

Consider, on the other hand, a simplification of the model in which A and B can only take on values of 0 or 1 and in which the presence of A causes an instantaneous rise in B from 0 to 1. This model can be simply formulated without any parameters by a Boolean equation, which defines how the value of B is updated given the value of A. This equation can be compactly written as

(2) $$ \begin{align} B^*=A, \end{align} $$

where we have adopted the convention that the variable with an asterisk is being updated. In other words, if $A=0$ , then B is updated to 0, independent of its current state. If $A=1$ , on the other hand, B is updated to 1, again independent of its current state. The time course of this Boolean equation is shown in red in Figure 1a, where A is changed from 0 to 1 at $t=0$ . In contrast to the differential equation, B is immediately turned ON when A is set to 1.

The above example is very simple and does not involve any of the elementary Boolean functions. To illustrate these functions, let us now consider the nodes A and B as the input nodes and C as the output node. A useful way to characterise the logical operations is to construct the so-called truth tables, which list the output values for all possible combinations of input values. The truth tables for the elementary logical functions are listed in Figure 1b,c. For example, the and function ( $C^*=A$ and B) only returns $C=1$ if both A and B are ON and will return 0 for all other input combinations. Similarly, an or gate returns an output of 1 if at least one or both of the inputs is 1. Obviously, the identity gate copies the current state of the only input node to the future state of the output node. Similarly, the not gate only has a single node as input and it inverts the current state of the node. It can be shown that by compounding these three elementary gates it is possible to encode all Boolean functions (Mano & Kime, Reference Mano and Kime1997), including some commonly encountered ones in electronics, such as xor (exclusive or). In Figure 1d, we show the truth table of the compound Boolean function $X^*$ = (not A) and (B or C). This table also illustrates how the output node X is updated by evaluating parts of the function recursively.

2.2. Update rules

Once a Boolean network is constructed, the nodes are updated following a particular update scheme. This is a choice the investigator needs to make because a Boolean model contains neither a natural timescale nor a specified order in which the reactions of the model take place. In the existing literature on Boolean models, three types of update schemes have been used: synchronous, asynchronous and probabilistic update schemes (Schwab et al., Reference Schwab, Kühlwein, Ikonomi, Kühl and Kestler2020). In synchronous Boolean models, all the components are updated at the same time, that is, the states of all the nodes at time step $t+1$ are determined by their states at time step t (Espinosa-Soto et al., Reference Espinosa-Soto, Padilla-Longoria and Alvarez-Buylla2004; Fauré et al., Reference Fauré, Naldi, Chaouiya and Thieffry2006; Garg et al., Reference Garg, Di Cara, Xenarios, Mendoza and De Micheli2008; Remy et al., Reference Remy, Ruet, Mendoza, Thieffry, Chaouiya, Priami, Ingólfsdóttir, Mishra and Nielson2006). This also means that the evolution of a synchronous Boolean model is deterministic: a particular input will always result in the same output.

An asynchronous Boolean model orders the updates of the nodes one after another in either a pre-determined or a stochastic manner. There are a number of ways to implement this scheme (Bonzanni et al., Reference Bonzanni, Garg, Feenstra, Schütte, Kinston, Miranda-Saavedra, Heringa, Xenarios and Göttgens2013; Fauré et al., Reference Fauré, Naldi, Chaouiya and Thieffry2006; Saadatpour et al., Reference Saadatpour, Albert and Albert2010; Thomas, Reference Thomas1991). For instance, one can follow a random order asynchronous update rule wherein all the nodes are updated exactly once but in a random order in each iteration (also called time step). This can be done by generating a random permutation of $\{1, 2, \ldots n\}$ , where n is the number of nodes, at the beginning of each iteration. Alternatively, one can follow a general asynchronous update rule in which the element that is updated is randomly drawn from the sequence $\{1, 2, \ldots , n\}$ . Thus, some nodes can get updated, by pure chance, twice or more before another node gets its turn. These two update methods will result in outcomes that are stochastic. This is in contrast to the deterministic asynchronous method in which nodes are updated using a fixed sequence (Aracena et al., Reference Aracena, Goles, Moreira and Salinas2009; Mortveit & Reidys, Reference Mortveit and Reidys2007) or at pre-determined time steps set by the rates of the corresponding reaction.

For biological applications, the synchronous update scheme is most likely not appropriate; it is rare that all components in a network change their value at the same time and that all processes take the same duration of time to be completed. Asynchronous updating can in principle implement data on timing and kinetics. However, these type of data are not always available, in which case it is unclear which type of asynchronous updating rule should be used. For a comparison between synchronous and asynchronous update schemes and its consequences, we refer to a study by Fauré et al. (Reference Fauré, Naldi, Chaouiya and Thieffry2006). This study applied both schemes to a model for the mammalian cell cycle and also proposed a hybrid scheme, combining both synchronous and asynchronous updating.

A third method of updating a Boolean model, also resulting in stochasticity, is through the use of so-called probabilistic Boolean networks (Shmulevich et al., Reference Shmulevich, Dougherty, Kim and Zhang2002a). In this updating method, each node in the network has a set of update equations to choose from. At the beginning of a time step, an equation for each node is randomly chosen, after which the nodes are updated synchronously. It thus combines a rule-based determinism for Boolean networks with stochasticity arising from the uncertainty from the choice of the update equation. For a review of this type of Boolean model, including its applications, we refer to Pal et al. (Reference Pal, Datta, Bittner and Dougherty2005) and Trairatphisan et al. (Reference Trairatphisan, Mizera, Pang, Tantar, Schneider and Sauter2013).

2.3. Translating rate equation models into Boolean models

To see how a signalling network may be encoded using Boolean logic, let us examine one of the simplest three-component systems that can give rise to oscillations (Novák & Tyson, Reference Novák and Tyson2008). This network is shown in Figure 2a and has only three components X, Y and Z. The network is wired such that X activates Y, Y activates Z and Z inhibits X. This is shown in the figures, where activation is indicated by arrows ( $\rightarrow $ ) and inhibition by a line and a perpendicular bar ( $\dashv $ ). This system can be translated into mathematical equations in which the concentration of the components can take on arbitrary positive values. The resulting set of ordinary differential equations is written as

(3) $$ \begin{align} \frac{dX}{dt} &= k_1 S/(1 + Z^p) - k_{-1} X, \nonumber \\ \frac{dY}{dt} &= k_2 X - k_3 Y/(K_m + Y), \nonumber \\ &\qquad\qquad \frac{dZ}{dt} = k_4 (Y - Z). \end{align} $$

In these equations, $k_1$ ,…, $k_4$ and $k_{-1}$ are the activation and degradation rates, respectively, $K_m$ is a dissociation constant, p is an integer representing the non-linear inhibition of X and S is an input signal (Novák & Tyson, Reference Novák and Tyson2008). Simulating these equations for particular sets of parameters results in an oscillatory state as shown in Figure 2b.

Fig. 2. Examples of Boolean networks. (a) Example of an oscillatory network. Arrows indicate activation and flat-edge symbols indicated inhibition. (b) The components of the network in panel (a) as a function of time, modelled using rate equations [parameters taken from Novák & Tyson (Reference Novák and Tyson2008): $k_1$ =0.1, $k_2$ =0.2, $k_3$ =0.1, $k_4$ =0.05, $k_{-1}$ =0.1, S=2, $K_m$ =0.01, p=4]. (c) Truth table for synchronous updating of the network shown in panel(a) (d) Modified network in which Y depends on X and Z. (e) Truth tables for synchronous updating of the network shown in panel (d). (f,g) State space and dynamics, represented by arrows, for asynchronous updating of the networks shown in panels (a,c). Fixed point attractors are indicated by red dots while the oscillatory cycle is shown by the red arrows.

To write this network in terms of Boolean operators, it is simplest to examine the diagram of Figure 2a. Note, however, that there are also more systematic ways to derive Boolean networks from ordinary differential equations (Davidich & Bornholdt, Reference Davidich and Bornholdt2008; Stötzel et al., Reference Stötzel, Röblitz and Siebert2015). This diagram can be translated into the following set of Boolean operators: $Y^*=X$ , $Z^*=Y$ and $X^*=$ not Z. We can then perform simulations of this Boolean network using synchronous update rules. As mentioned in Section 2.1, for synchronous updating of small networks, it is most convenient to construct the truth table. The table for this diagram is displayed in Figure 2c, which shows that it also exhibits oscillatory cycles. Specifically, starting at $(X,Y,Z)=(0,0,0)$ , the sequence is $(0,0,0)\rightarrow (1,0,0) \rightarrow (1,1,0) \rightarrow (1,1,1) \rightarrow (0,1,1) \rightarrow (0,0,1) \rightarrow (0,0,0),$ while $(1,0,1) \rightarrow (0,1,0) \rightarrow (1,0,1)$ is also a cycle.

As a second example, let us consider the previous signalling network but now changed such that the activation of Y depends on both X and Z. This can be easily incorporated by changing the rate equation for Y into

(4) $$ \begin{align} \frac{dY}{dt} = k_2 XZ - k_3 Y/(K_m + Y) \end{align} $$

while keeping the equations for X and Z unchanged. Now, there are two possible solutions: the oscillatory state, similar to the one shown in Figure 2b, and a stationary state given by $Y=Z=0$ and $X= k_1 S/k_{-1}$ . The latter is stable and the resulting state of the system depends on the initial conditions.

The Boolean network corresponding to this slightly altered network is shown in Figure 2d. The only difference between this and the previous network is that the Boolean update equation for Y is now written as $Y^*=X$ and Z. The truth table for this network, corresponding to the synchronous update scheme, is given in Figure 2e. This table reveals that (1,0,0) is a fixed point of the system: once in this state, the network will remain in it indefinitely. Note, however, that this fixed point is only reached for certain initial conditions [(0,0,0), (1,0,0), (0,0,1), (0,1,1), and (1,1,1) to be precise]. Thus, as in the continuous version of the network, the binary Boolean network displays a steady-state solution in which both Y and Z are zero and in which X has a non-zero value. Obviously, for the continuous system, this value depends on the model parameters while for the parameterless Boolean network it is simply one. As in the continuous system, the Boolean network also exhibits an oscillatory state: $(0,1,0)\rightarrow (1,0,1) \rightarrow (0,1,0)$ , which is reached from initial conditions (0,1,0), (1,0,1) and (1,1,0).

Let us now examine these two Boolean networks using asynchronous update rules. In this case, each element can be changed independently. Since our networks contain only three elements, this process can be visualised using the cubes shown in Figure 2f,g, where each node represents a particular state of the system and the edges represent transitions between the states. Here, the arrows indicate the transition between the different nodes according to the rules of the Boolean network. The dynamics of the Boolean network can then be determined by following these arrows.

For the network of Figure 2a, it is easy to see that the asynchronous update scheme also results in the same oscillatory cycle as the synchronous update scheme. This cycle is shown in Figure 2f by the red arrows. Contrary to the synchronous update scheme, however, the asynchronous update scheme for the second network does not exhibit an oscillatory state. For this update scheme, regardless of the initial conditions, the network always transitions to the same node (1,0,0). Thus, the steady state of the system corresponds to a fixed point, indicated by the red dot in Figure 2g. Finally, we should also point out that is possible to go “backwards” and transform a Boolean model into a continuous model (Wittmann et al., Reference Wittmann, Krumsiek, Saez-Rodriguez, Lauffenburger, Klamt and Theis2009). The resulting ordinary differential model could then be used to provide quantitative information regarding, for example, the concentrations of network components.

2.4. Encoding a Boolean network from experiments

The task of encoding a Boolean network based on experimental data is not trivial. It requires the identification of the relevant components (nodes in the network) as well as the correct update rules and thus requires biochemical, genetic and pharmacological data. While identifying components is typically not that difficult, determining the interactions between these components is challenging since the number of possible update equations grows exponentially in the number of nodes in the network (Demongeot et al., Reference Demongeot, Elena and Sené2008). Furthermore, to define these interactions requires careful consideration of experimental data. This task is especially difficult since available experimental information is generally incomplete. To elaborate, consider a node in a Boolean network with n nodes upstream. To formulate the update equation unambiguously, we need the response of the node for the whole set of $2^n$ inputs. When such information is available, formulating the equation is straightforward (Karanam et al., Reference Karanam, He, Hsu, Schulze, Dubeaux, Karmakar, Schroeder and Rappel2021). Generally, however, such extensive data are unavailable and simplifying assumptions about the nature of interactions are required.

Fig. 3. Inference rules for the construction of Boolean networks. Experimental data are synthesised to be represented in graphs with the least number of nodes and edges, that is, as a sparse representation. This sometimes requires an introduction of an intermediary node, as in graphs 1 and 3, but when additional information becomes available, the graph can in fact simplify, as in going from graph 1 to graph 2. For further details, see text (from Li et al., Reference Li, Assmann and Albert2006).

A classical algorithm to infer a Boolean network, called REVerse Engineering ALgorithm (reveal)(Liang et al., Reference Liang, Fuhrman and Somogyi1998), computes quantities encountered in information theory (Cover, Reference Cover1999), such as joint entropy and mutual information. The advantage of reveal over earlier methods is that one only needs a small fraction of all possible input–output relations to obtain a Boolean network with a very small error rate. The method is, of course, exact when one uses all the $2^n$ input–output relations for a network of n nodes. To include a more realistic scenario in which one allows for noise in gene regulation, either inherent or caused by measurement techniques, the so-called Best-Fit Extension method (Lähdesmäki et al., Reference Lähdesmäki, Shmulevich and Yli-Harja2003; Shmulevich et al., Reference Shmulevich, Yli-Harja, Astola and Core2001) can be employed.

We highlight here another approach used in constructing a large network, following an extensive literature search, to model guard cell dynamics in Arabidopsis in response to ABA (Albert et al., Reference Albert, Acharya, Jeon, Zañudo, Zhu, Osman and Assmann2017; Li et al., Reference Li, Assmann and Albert2006). Mathematically, this approach relies on developing a graph with the smallest number of nodes and edges consistent with all established qualitative relationships (Aho et al., Reference Aho, Garey and Ullman1972). It formulates a number of inference-based rules, shown schematically in Figure 3. In the first graph, experimental data have identified that component A promotes B (and is not a direct biochemical reaction) but also that C promotes the interaction between A and B. In that case, it is assumed that there is an intermediary node ( $IN$ ) of the $A-B$ pathway and that C acts on this intermediary node. If it is also known that A promotes C, then this intermediary node can be identified as C (graph 2 in Figure 3). Finally, if A inhibits B and C inhibits the interaction between A and B, then the logical rule can be interpreted as A promotes an intermediary node $IN$ , which inhibits B, while C inhibits $IN$ (graph 3 in Figure 3). Using these rules, it was shown that the developed network was able to capture existing experimental data (Albert et al., Reference Albert, Acharya, Jeon, Zañudo, Zhu, Osman and Assmann2017; Li et al., Reference Li, Assmann and Albert2006). We will come back to this network in Section 4.1.

2.5. Dynamics of Boolean networks

Often, the goal of modelling is to determine the steady state of the system. That is to say, what is the outcome of the system for long times? Any deterministic Boolean model, when simulated for long enough time, converges to a limit cycle or an attractor. A limit cycle is a subset of the states of the network over which the state of the system repeats over and over in a cyclical fashion. The length of the limit cycle is the number of states in the limit cycle. An attractor is a state of the system whose ‘future’ state is identical to the current state; the system gets locked-in once it reaches an attractor state. We have already seen examples of these two possible outcomes when discussing the networks presented in Figure 2. The set of states that converges to a particular attractor constitutes its so-called basin of attraction. Since the evolution of the network is deterministic, no two basins of attraction share a common element; they are said to be disjoint. Likewise, a limit cycle—along with transient states that feed into it—is disjoint with the next one. Thus, the entire state space can be carved up into disjoint basins of attraction and basins of limit cycles and each trajectory of the system is a subset of the basin its initial state belongs to. Determining the number of attractors, together with their basins of attraction, is an active area of research in the mathematical field of graph theory and we direct the interested reader to several recent studies (Aracena et al., Reference Aracena, Richard and Salinas2017; Krawitz & Shmulevich, Reference Krawitz and Shmulevich2007; Mori & Mochizuki, Reference Mori and Mochizuki2017; Veliz-Cuba & Laubenbacher, Reference Veliz-Cuba and Laubenbacher2012).

Defining and analysing the basins of attraction for non-deterministic Boolean networks (e.g., using random order asynchronous or general asynchronous update methods) is not as straightforward since the trajectory of the system is no longer deterministic. Furthermore, the set of attractors and limit cycles found using asynchronous updating can be different from the ones found using synchronous updating. This was highlighted by Saadatpour et al. (Reference Saadatpour, Albert and Albert2010), which carried out a comparative study of the dynamics and steady states of the ABA-induced stomatal closure network under synchronous and the three aforementioned asynchronous update schemes. To enumerate the fixed points and limit cycles, they reduced the system using Markov chains (Ross, Reference Ross2014) and by simplifying the Boolean update equations. They found that both types of update schemes exhibited a fixed point. However, for synchronous updating, they found large basins of attractions for two limit cycles. These limit cycles, and their basins of attractions, were not found using asynchronous updating, unless strict limitations regarding the timing of several processes were implemented.

2.6. Software tools

Once a network and the update schemes are defined, a Boolean network can be simulated to obtain the trajectory and steady states of the system, to visualise the network, and to determine the activity levels of its components. A large number of computational tools have been developed to simulate Boolean networks on personal computers, as reviewed recently by Schwab et al. (Reference Schwab, Kühlwein, Ikonomi, Kühl and Kestler2020). In addition, software packages are available to determine and computationally identify the attractors of a Boolean network (Rozum et al., Reference Rozum, Gómez Tejeda Zañudo, Gan, Deritei and Albert2021; Rozum et al., Reference Rozum, Deritei, Park, Gómez Tejeda Zañudo and Albert2022). Most of these packages, but not all (Klamt et al., Reference Klamt, Saez-Rodriguez and Gilles2007), are open source and can thus be freely used. Some of these tools, however, do not use a graphical interface, which makes it more challenging to construct and visualise the network (Albert et al., Reference Albert, Thakar, Li, Zhang and Albert2008; Garg et al., Reference Garg, Di Cara, Xenarios, Mendoza and De Micheli2008; Helikar & Rogers, Reference Helikar and Rogers2009; Klarner et al., Reference Klarner, Streck and Siebert2017; Müssel et al., Reference Müssel, Hopfensitz and Kestler2010; Paulevé, Reference Paulevé, Feret and Koeppl2017; Stoll et al., Reference Stoll, Viara, Barillot and Calzone2012). Other packages only allow synchronous updating and can thus not implement an asynchronous update scheme (Bock et al., Reference Bock, Scharp, Talnikar and Klipp2014; Terfve et al., Reference Terfve, Cokelaer, Henriques, MacNamara, Goncalves, Morris, van Iersel, Lauffenburger and Saez-Rodriguez2012). Finally, some packages are only able to run a single initialisation at a time, which means that probing a large set of initial conditions, especially valuable for large scale networks with asynchronous updating, is challenging (Gonzalez et al., Reference Gonzalez, Naldi, Sanchez, Thieffry and Chaouiya2006; Schwab & Kestler, Reference Schwab and Kestler2018).

We have recently developed Boolink, a simulation platform for Boolean networks that is based on a graphical user interface and is completely open-source (Karanam et al., Reference Karanam, He, Hsu, Schulze, Dubeaux, Karmakar, Schroeder and Rappel2021). Specifically, the software allows users to define the nodes and connections in the Boolean network, visualise the network as a tree, set various simulation parameters including the number of time steps and initial conditions, plot the activity of a few chosen nodes, and to analyse the trajectory of the system as a whole. Boolink is written in Python and C++, and the source code is freely available from the GitHub repository, https://github.com/rappel-lab/boolink-gui, along with its documentation, to use, modify and distribute. We have also packaged the software as a Docker container (Merkel, Reference Merkel2014), which is a self-contained system that comes with all the software dependencies and runs straight out of the box. In its original presentation, Boolink was only able to simulate a Boolean network using the physiologically relevant asynchronous update scheme. Recently, however, we have extended Boolink to include the ability to simulate networks using a synchronous update scheme.

3. Boolean networks and gene regulation in plants

Creating a regulatory framework based on the available data on gene expression is essential to understanding gene expression. A network that is inferred from gene expression data is termed GRN (Emmert-Streib et al., Reference Emmert-Streib, Dehmer and Haibe-Kains2014). Several methods have been developed to construct GRNs from available data, including Boolean models, information-theory-based models, and machine-learning-based models. These methods and their application and suitability in different contexts is an active field of research and we refer the interested reader to several review articles (Chai et al., Reference Chai, Loh, Low, Mohamad, Deris and Zakaria2014; Delgado & Gómez-Vela, Reference Delgado and Gómez-Vela2019; Fiers et al., Reference Fiers, Minnoye, Aibar, Bravo González-Blas, Kalender Atak and Aerts2018; Zhao et al., Reference Zhao, He, Tang, Zou and Guo2021). Here, we limit our discussion to Boolean models.

The first application of Boolean modelling was carried out by Kauffman when he described a genetic network (Kauffman, Reference Kauffman1969a). In a Boolean gene network, a gene is either turned ON (i.e., has value 1) or turned OFF (with value 0), while the topology of the network specifies how and if a gene interacts with other genes. In plants, Boolean networks have been applied to a number of genetic networks. We will discuss here three different examples: Boolean models (a) for flower development, (b) for induced systemic resistance (ISR) induced by microbes and (c) for the root stem cell niche (SCN). These models have introduced modifications to the simple implementations of Boolean networks described so far. These modifications will be discussed as the systems are introduced.

3.1. Flower development

One of the first examples of Boolean modelling studied early flower development in the model plant Arabidopsis thaliana (Mendoza & Alvarez-Buylla, Reference Mendoza and Alvarez-Buylla1998). In this model, 12 genes were considered and the topology of the network was determined based on experimental data. The model was slightly more involved than the simple Boolean implementation we described in Section 2 in that the modified model is known as a threshold Boolean model. Each node of the model still takes binary values (0 or 1) but interactions between any pair of nodes are encoded by weights between them; excitatory interactions carry a weight of $+1$ whereas inhibitory interactions carry a weight of $-1$ . The update equation of a node in this model is not Boolean but algebraic, consisting of the sum of weighted interaction terms. When a node is updated, the sum of all of its interactions with other nodes is calculated. If the sum exceeds the threshold of the node, then the node is updated to 1; if not, to 0.

The update scheme for this model is in-between the synchronous and asynchronous update schemes as described before, and is termed semi-synchronic, and block-sequential and block-parallel in later iterations (Aracena et al., Reference Aracena, Goles, Moreira and Salinas2009; Demongeot & Sené, Reference Demongeot and Sené2020). Instead of updating all the nodes at once or one after another in some order, the nodes in the model are grouped into blocks. All the nodes in a block are updated at once, and the blocks themselves are updated sequentially. This method makes use of qualitative experimental data, such as the order of activation of different parts of the genetic network.

The model was found to have six attractors, four of which were consistent with the gene expression patterns observed in A. thaliana (Mendoza & Alvarez-Buylla, Reference Mendoza and Alvarez-Buylla1998). One of the remaining two was not able to flower and the sixth one, while not observed, could be induced experimentally (Mendoza & Alvarez-Buylla, Reference Mendoza and Alvarez-Buylla1998). Since this Boolean threshold model was published, several studies have further analysed its dynamics. These studies revealed that it is possible to reduce its complexity while maintaining its steady-state behaviour (Demongeot et al., Reference Demongeot, Goles, Morvan, Noual and Sené2010; Ruz et al., Reference Ruz, Goles and Sené2018) and highlighted the crucial role of the plant hormone gibberellin in normal flower development (Demongeot et al., Reference Demongeot, Goles, Morvan, Noual and Sené2010).

Fig. 4. Boolean modelling of gene networks. (a) Example of a putative network that maintains the SCN in Arabidopsis. Arrows indicated activation and flat-edge symbols correspond to repression. For the definition of the different components, see Velderraín et al. (Reference Velderraín, Martínez-García, Álvarez-Buylla, Kaufmann and Mueller-Roeber2017). (b) Attractors of the Boolean network shown in panel (a). Green represents an active and red represents an inactive gene. The labels at the top of the diagram represent the attractors and correspond to the phenotypes observed in experiments (CEI, cortex-endodermis initials; CEP, columella epidermis initials; QC, quiescent center; VAS, vascular initials) (from Velderraín et al., Reference Velderraín, Martínez-García, Álvarez-Buylla, Kaufmann and Mueller-Roeber2017). (c) Modified network based on novel experimental and computational results (from Velderraín et al., Reference Velderraín, Martínez-García, Álvarez-Buylla, Kaufmann and Mueller-Roeber2017).

3.2. Induced systemic resistance

Recently, the ISR in A. thaliana plants triggered by beneficial microbes was investigated using Boolean modelling (Timmermann et al., Reference Timmermann, González and Ruz2020). ISR is an important defense mechanism of plants against harmful pathogens (Pieterse et al., Reference Pieterse, Zamioudis, Berendsen, Weller, Van Wees and Bakker2014) and the study investigated how the bacterium Paraburkholderia phytofirmans PsJN can trigger ISR and protection from the bacterial pathogen Pseudomonas syringae DC3000 (Timmermann et al., Reference Timmermann, Armijo, Donoso, Seguel, Holuigue and González2017, Reference Timmermann, Poupin, Vega, Urrutia, Ruz and González2019). It used the temporal experimental expression patterns of eight key genes following inoculation of PsJN and asked which threshold Boolean network was able to reproduce the time series data. Parameters of the model, including the weights among the nodes and their threshold values, were fitted to experimental data using an algorithm called differential evolution (Storn & Price, Reference Storn and Price1997), which belongs to a class of fitting algorithms called genetic algorithms (Ruz et al., Reference Ruz, Timmermann and Goles2015). The study inferred 1,000 networks from the data. One of these networks was chosen and pruned using biological reasoning. The robustness of the pruned network was then tested by determining how mutations of fundamental genes affected the ISR response. These virtual mutation experiments produced responses that were consistent with available experimental data (Timmermann et al., Reference Timmermann, González and Ruz2020). Additionally, the study found that the pruned consensus network is robust because it requires an unlikely event of a triple mutation to the network before the ISR is lost. Furthermore, the authors argue that, in the presence of errors in gene expression data, the differential evolution algorithm used to derive the GRN fared better than classical algorithms to infer Boolean networks, including REVEAL (Liang et al., Reference Liang, Fuhrman and Somogyi1998) and best fit extension (Lähdesmäki et al., Reference Lähdesmäki, Shmulevich and Yli-Harja2003).

3.3. Root stem cell niche

The examples above applied Boolean modelling to determine the most probable network that is consistent with experimental data. In doing so, these studies found missing links or were able to determine the most critical network components. As a result, these Boolean models were often able to predict novel components or connections between components and could suggest new experiments. To further illustrate the ability of Boolean models to guide experiments, we focus here on another example of a gene network studied using Boolean modelling, the root SCN in A. thaliana (Azpeitia et al., Reference Azpeitia, Benítez, Vega, Villarreal and Alvarez-Buylla2010; Velderraín et al., Reference Velderraín, Martínez-García, Álvarez-Buylla, Kaufmann and Mueller-Roeber2017). The root SCN in A. thaliana is well studied and is located at the root apical meristem (Dolan et al., Reference Dolan, Janmaat, Willemsen, Linstead, Poethig, Roberts and Scheres1993). It consists of a so-called quiescent center (QC), comprised of four infrequently dividing cells, and, in immediate proximity, active stem cells that are called initials. The divisions of these initials result in different types of differentiated cells and in tissue growth of the plant (Dolan et al., Reference Dolan, Janmaat, Willemsen, Linstead, Poethig, Roberts and Scheres1993). The question thus arises, how can the undifferentiated cells of the QC give rise to several differentiated cell types?

Modelling, and in particular Boolean modelling, is ideal to address this question. Experimental work has identified a number of molecular and genetic components that play a role in the maintenance of the SCN (Pardal & Heidstra, Reference Pardal and Heidstra2021). Furthermore, the interactions between some, but not all of the components can also be deduced from experimental work. It is therefore possible to construct a putative wiring diagram as in Figure 4a, which shows the components along with their interactions as either arrows, indicating activation, or flat-end symbols, indicating repression. The dynamics of this network should then allow steady-state solutions with gene expression that is consistent with the different cell types of the SCN. In terms of Boolean modelling, this means that the network should display attractors corresponding to these different cell types.

Simulating the Boolean network in Figure 4a using synchronous updating revealed four different attractors as shown in Figure 4b. In this diagram, active genes are displayed in green while inactive genes are displayed in red. Each of the four attractors correspond to a different set of genes that are ON or OFF and, thus, to a different cell phenotype. For example, the gene SCR (scarecrow) is ON (and thus has a value of 1) in the phenotype corresponding to columella epidermis initial and vascular initial but is OFF and has value 0 in cortex-endodermis initial and QC (QC cell). The diagram also shows the size of the basin of attraction, expressed as the percentage of initial conditions that resulted in the attractor.

Further testing of this model and comparing the outcomes to experimental results showed that certain interactions were missing. For example, this analysis revealed the need for a repressor of WOX5 (wuschel related homebox 5) and an additional component with an inhibitory link was predicted (Azpeitia et al., Reference Azpeitia, Benítez, Vega, Villarreal and Alvarez-Buylla2010). This prediction was verified in experiments, which showed that WOX5 is negatively regulated by CLE40 (clavata-like-40) (Stahl et al., Reference Stahl, Wink, Ingram and Simon2009). Additional predictions resulted in the modified network displayed in Figure 4c, where the postulated interactions are shown in red (Azpeitia et al., Reference Azpeitia, Benítez, Vega, Villarreal and Alvarez-Buylla2010). After this study and once new experimental findings became available, this network has been modified and extended further (Azpeitia et al., Reference Azpeitia, Weinstein, Benítez, Mendoza and Alvarez-Buylla2013). These studies showed the power of Boolean modelling: once a Boolean network has been constructed, it is fairly straightforward to modify and extend it and to generate experimental predictions. These modifications and extensions are much easier to implement than in continuous models based on rate equations. In those type of models, a modification typically requires refitting and adjusting the model parameters, which can be an arduous task (Karmakar et al., Reference Karmakar, Tang, Yue, Lombardo, Karanam, Camley, Groisman and Rappel2021).

4. Boolean networks and signalling in plants

Biological signalling pathways can be very complex, containing numerous components and multiple feedback loops. Such complex pathways can also be addressed by Boolean modelling and examples include T-cell signalling (Saez-Rodriguez et al., Reference Saez-Rodriguez, Simeoni, Lindquist, Hemenway, Bommhardt, Arndt, Haus, Weismantel, Gilles, Klamt and Schraven2007), molecular pathways of neurotransmitters (Gupta et al., Reference Gupta, Bisht, Kukreti, Jain and Brahmachari2007) and cancer pathways (Fumia & Martins, Reference Fumia and Martins2013; Sherekar & Viswanathan, Reference Sherekar and Viswanathan2021). A prime example of a complex signalling network is found in plants, where the network regulating phytohormone ABA-induced stomatal closure contains a large number of interconnected components. Below, we will review studies that attempt to cast this closure pathway into a Boolean network. Furthermore, we will also discuss recent efforts to extend this signalling network to include CO $_2$ signalling.

4.1. ABA signalling network

Stomata are pores in the epidermis of leaves that regulate gas exchange, including CO $_2$ for photosynthesis and loss of water vapor. Each stomata is formed by a pair of guard cells and its aperture is modulated in response to environmental changes such as light and CO $_2$ (Assmann & Jegla, Reference Assmann and Jegla2016; Munemasa et al., Reference Munemasa, Hauser, Park, Waadt, Brandt and Schroeder2015). Furthermore, drought results in the accumulation of ABA in guard cells, which leads to stomatal closure (Hsu et al., Reference Hsu, Dubeaux, Takahashi and Schroeder2021; Raghavendra et al., Reference Raghavendra, Gonugunta, Christmann and Grill2010).

The network that underlies ABA-induced stomatal closure in A. thaliana is complex and contains a large number of components ( $>$ 80). Consequently, the number of rate constants is also very large and, not surprisingly, many are not quantified. To illustrate the complexity of the network, we reproduce in Figure 5, the network investigated by Albert et al. in a recent study (Albert et al., Reference Albert, Acharya, Jeon, Zañudo, Zhu, Osman and Assmann2017). Given the complexity and the number of components, this network is particularly suitable for Boolean approaches (Albert et al., Reference Albert, Acharya, Jeon, Zañudo, Zhu, Osman and Assmann2017; Li et al., Reference Li, Assmann and Albert2006; Maheshwari et al., Reference Maheshwari, Du, Sheen, Assmann and Albert2019; Maheshwari et al., Reference Maheshwari, Assmann and Albert2020; Waidyarathne & Samarasinghe, Reference Waidyarathne and Samarasinghe2018).

Fig. 5. Signalling network of abscisic acid (ABA)-induced stomatal closure. Arrows indicate positive interactions while filled circles indicate negative interactions. Rectangles represent nodes that are connected to other nodes. Black lines represent direct interactions and green lines represent indirect interactions. Nodes are color coded according to their function: enzymes (red), signalling proteins (green), membrane-transport related nodes (blue), and secondary messengers and small molecules (orange). For names of the components, see original publication (Albert et al., Reference Albert, Acharya, Jeon, Zañudo, Zhu, Osman and Assmann2017).

Albert and colleagues encoded the ABA-induced stomatal closure pathway into a Boolean network. This network has a single input node, representing ABA, and a single output node, representing stomatal closure. Obviously, both the input and output node were also taken to be binary: ABA is either 1 (present) or 0 (absent) while a similar logic applied to the closure node. The network shown in Figure 5 was constructed following a careful review of available experimental literature. The Boolean dynamics was simulated using an asynchronous update scheme in which nodes were updated in a randomly selected order at each time step (random order asynchronous). A large number of simulations (2,500) were performed and the percentage of closure was computed as the percentage of simulations in which the node Closure=1 in the steady state. A verification of the wild-type model, that is, without any alterations to the wiring or components, was performed by computing the percentage of closure with and without ABA present. The results are shown in Figure 6, as expected, in the absence of ABA, the stomata remained open and the percentage of closure was 0% (open circles) while in the presence of ABA the stomata close (closed circle; 100% closure).

Fig. 6. Results of a Boolean ABA network. Shown are the percentage of closure as a function of iteration step. The wild-type (WT) curves show the network response in the absence of (open circles) and presence of ABA (closed circles). Other curves show the response following simulated knockout of the component (node set to 0) in the presence of ABA. For abbreviations, see original study (Albert et al., Reference Albert, Acharya, Jeon, Zañudo, Zhu, Osman and Assmann2017).

One of the strengths of Boolean modelling is the ease with which one can alter the network and either ‘knock-out’ a node or make a node ‘constitutively active’. A virtual knock-out experiment equates to fixing one of the nodes to 0 while fixing a node’s value to 1 corresponds to making this component constitutively active. Albert et al. performed a systematic analysis of the response of the Boolean network to ABA exposure in which all internal nodes of the network were set to either 0 or 1. Some of the results of these knock-out experiments are shown in Figure 6. For example, knocking out cytosolic pH resulted in reduced sensitivity to ABA with only 35% closure, consistent with experiments (Zhang et al., Reference Zhang, Zhu, Zhang, Li, Yan, Wang, Wang, Welti, Zhang and Wang2009).

The study showed that altering single nodes in the networks resulted in either an increased, a decreased, or an unchanged sensitivity to ABA. Where possible, the results were compared to available experimental data and this comparison agreed in most cases. In the case where no experimental data were available, the computational result could be considered a prediction. Some of these predictions were subsequently tested using experiments, demonstrating the usefulness of Boolean modelling.

Even though the model was able to reproduce experimental data in more than 75% of the predictions, there were several clear discrepancies with experimental observations (Albert et al., Reference Albert, Acharya, Jeon, Zañudo, Zhu, Osman and Assmann2017). Further highlighting the strength of Boolean modelling, these discrepancies were used in a follow-up study, which aimed to improve the model (Maheshwari et al., Reference Maheshwari, Du, Sheen, Assmann and Albert2019). For this, the original model was first reduced to a smaller network with 49 nodes and 113 edges. This reduced network was shown to duplicate all results from the original network. A subsequent computational analysis of this reduced network then revealed that that inhibiting PP2C protein phosphatase ABSCISIC ACID INSENSITIVE 2 (ABI2) by cytosolic calcium was able to rectify most of the discrepancies. The proposed inhibition by calcium was also verified in experiments, highlighting the ability to iterate between model and experiment (Maheshwari et al., Reference Maheshwari, Du, Sheen, Assmann and Albert2019). This iterative quality of Boolean modelling was also evident from a recent and additional follow-up study (Maheshwari et al., Reference Maheshwari, Assmann and Albert2020). This study examined the response of the improved network in the absence of ABA or following the removal of ABA. In the first case, the stomata should remain open while in the second case, their closed state should relax to an open state. Probing the network under these conditions, it was further improved so that its response was consistent with experiments (Maheshwari et al., Reference Maheshwari, Assmann and Albert2020).

4.2. CO2 signalling pathway

The Boolean network shown in Figure 5 was recently extended to include CO $_2$ regulation of stomatal movements (Karanam et al., Reference Karanam, He, Hsu, Schulze, Dubeaux, Karmakar, Schroeder and Rappel2021). This extension is possible since several elements of CO $_2$ signalling overlap with those of ABA signalling (Hsu et al., Reference Hsu, Takahashi, Munemasa, Merilo, Laanemets, Waadt, Pater, Kollist and Schroeder2018; Merilo et al., Reference Merilo, Jalakas, Kollist and Brosché2015; Zhang et al., Reference Zhang, Takahashi, Hsu, Kollist, Merilo, Krysan and Schroeder2020). Elevated CO $_2$ levels result in the closure of stomatal pores and thus affect the water use efficiency and yield of crop plants (Dubeaux et al., Reference Dubeaux, Hsu, Ceciliato, Swink, Rappel and Schroeder2021; Engineer et al., Reference Engineer, Hashimoto-Sugimoto, Negi, Israelsson-Nordström, Azoulay-Shemer, Rappel, Iba and Schroeder2016; Zhang et al., Reference Zhang, De-oliveira Ceciliato, Takahashi, Schulze, Dubeaux, Hauser, Azoulay-Shemer, Tõldsepp, Kollist, Rappel and Schroeder2018). In an initial attempt, the ABA network was complemented with a CO $_2$ branch as shown in Figure 7a. This branch consisted of CO $_2$ as an input node and, through several intermediary nodes, inhibited the GHR1 (guard cell hydrogen peroxide resistant 1) node in the ABA network. As in the ABA studies, Boolean dynamics was implemented using asynchronous updating using randomly selected nodes (random order asynchronous) and results were averaged over a large number of realisations (Karanam et al., Reference Karanam, He, Hsu, Schulze, Dubeaux, Karmakar, Schroeder and Rappel2021). The response of ABA under high and low CO $_2$ conditions, modelled by setting the node CO $_2$ to either 1 or 0, is shown in Figure 7b. Note that here the conductance level is used instead of the percent closed stomata. This conductance level is simply computed as 1-Closure. The simulations predicted that the introduction of ABA leads to a decrease of conductance level from 1 to 0 when CO $_2$ =1, corresponding to fully closed stomata. For CO $_2$ =0, the exposure to ABA resulted in a reduction of the conductance level from 1 to 0.5.

Fig. 7. CO $_2$ and ABA-induced stomatal closure model. (a) Extended network showing the new CO $_2$ branch in blue and the existing ABA network, shown in Figure 5, as a box. Only several of ABA components are shown. (b) Predicted stomatal conductance levels of the network in panel (a) for both CO $_2$ =0 (red) and CO $_2$ =1 (blue), before and after the application of ABA. (c) Modified CO $_2$ and ABA-induced stomatal closure network, with modifications represented by orange links. (d) Predicted stomatal conductance levels using the network shown in panel (c) for both CO $_2$ =0 (red) and CO $_2$ =1 (blue), before and after the application of ABA. For names of the components, see original publication (Albert et al., Reference Albert, Acharya, Jeon, Zañudo, Zhu, Osman and Assmann2017) (from Karanam et al., Reference Karanam, He, Hsu, Schulze, Dubeaux, Karmakar, Schroeder and Rappel2021)

These predictions were subsequently tested in experiments by determining ABA-mediated stomatal closure under high and low CO $_2$ conditions (Karanam et al., Reference Karanam, He, Hsu, Schulze, Dubeaux, Karmakar, Schroeder and Rappel2021). These gas-exchange experiments were conducted by applying ABA to the transpiration stream of excised intact leaves (Ceciliato et al., Reference Ceciliato, Zhang, Liu, Shen, Hu, Liu, Schäffner and Schroeder2019). The experiments revealed that the application of ABA under both conditions resulted in a reduced conductance level (Karanam et al., Reference Karanam, He, Hsu, Schulze, Dubeaux, Karmakar, Schroeder and Rappel2021). In contrast to the simulation results of Figure 7b, however, the steady-state stomatal conductance prior to ABA application was different for high and low CO $_2$ . Specifically, it was higher for low CO $_2$ , demonstrating that CO $_2$ induces stomatal conductance reduction. Taken together, these experimental results suggested that both CO $_2$ and ABA reduce stomatal conductance and that they have additive responses (Karanam et al., Reference Karanam, He, Hsu, Schulze, Dubeaux, Karmakar, Schroeder and Rappel2021).

Again showing the ability to iterate between experiments and modelling, the network was modified to account for the experimental results. The updated network topology is shown in Figure 7c where the added links are displayed in orange. These links were partially motivated by experimental data. For example, adding CO $_2$ dependence on calcium signalling was motivated by experiments that showed that cytosolic calcium is involved in CO $_2$ -induced stomatal closure (Schulze et al., Reference Schulze, Dubeaux, Ceciliato, Munemasa, Nuhkat, Yarmolinsky, Aguilar, Diaz, Azoulay-Shemer, Steinhorst, Offenborn, Kudla, Kollist and Schroeder2021; Schwartz et al., Reference Schwartz, Ilan and Grantz1988; Webb et al., Reference Webb, McAinsh, Mansfield and Hetherington1996). The response of the updated network to ABA application is shown in Figure 7d and is now consistent with experimental results. Importantly, in the absence of ABA, the updated network has a steady state that depends on the CO $_2$ level. Furthermore, application of ABA resulted in absolute conductance changes that were similar for both conditions (Karanam et al., Reference Karanam, He, Hsu, Schulze, Dubeaux, Karmakar, Schroeder and Rappel2021).

5. Conclusion and outlook

This article summarises recent attempts in modelling genetic networks and signalling pathways using Boolean models. We have focused on plants, where this type of modelling has been used in a wide variety of studies. We have attempted to outline how one can transform a model composed of ordinary differential equations into a Boolean model and have described methods to synthesise experimental data into logical equations. Furthermore, we have shown how different update rules can result into different outcomes and have pointed readers to available software for Boolean modelling. We have also reviewed a number of genetic and signalling networks in plants that have been investigated using Boolean models. In doing so, we have strived to highlight the advantages and promises of Boolean modelling.

Most importantly, Boolean models offer a simple yet effective way to model the steady states of reaction networks in biological systems made of a large number of components. They only require knowledge about the components of the network and the nature of the connectivity between components but not detailed knowledge about kinetics or rates. Boolean models are therefore especially useful in cases where the current knowledge of interactions is only qualitative, when the kinetic and rate parameters are not precisely determined, or when the network is too large to be simulated in a reasonable time. We should point out, however, that constructing a Boolean network is not the only way to simplify a complex mathematical model. Especially when a model falls into the class of so-called sloppy models, in which many of the parameters are loosely constrained (Brown & Sethna, Reference Brown and Sethna2003; Gutenkunst et al., Reference Gutenkunst, Waterfall, Casey, Brown, Myers and Sethna2007), it is sometimes possible to systematically reduce the number of equations while still maintaining the predictive value of the model (Lombardo & Rappel, Reference Lombardo and Rappel2017; Transtrum & Qiu, Reference Transtrum and Qiu2014).

One of the main strengths of Boolean modelling is the ease with which one can generate experimental predictions. As we have discussed here, these predictions have resulted in the identification of critical components or interactions in both genetic networks in plants [e.g., negative regulation of WOX5 by CLE40 (Azpeitia et al., Reference Azpeitia, Benítez, Vega, Villarreal and Alvarez-Buylla2010)] as well as in signalling networks [e.g., calcium inhibition of ABI2 in the ABA signalling pathway (Maheshwari et al., Reference Maheshwari, Du, Sheen, Assmann and Albert2019)]. These predictions can then be tested in experiments, further specifying the underlying networks and improving our understanding of the biological mechanisms. Boolean modelling, however, is not well suited to address detailed kinetics of networks and pathways. After all, time does not explicitly occur in a Boolean model and most applications therefore focus on determining steady states of the system. Furthermore, Boolean networks cannot be used to model graded outcomes such as concentration of network components that are different from 0 and 1. In order to achieve that, hybrid models have been developed in which the set of states is expanded beyond $\{0,1 \}$ and the update equations contain a combination of algebraic and Boolean terms (Glass & Kauffman, Reference Glass and Kauffman1973). These hybrid models have been used to model the interaction of immune cells and pathogens (Thakar et al., Reference Thakar, Saadatpour-Moghaddam, Harvill and Albert2009), gene networks underlying flower development in Arabidopsis (Mendoza & Xenarios, Reference Mendoza and Xenarios2006), and light-induced stomatal opening in plants (Sun et al., Reference Sun, Jin, Albert and Assmann2014). Despite these limitations, we expect that Boolean modelling will continue to play a prominent role in deciphering the structure and mechanisms of biological networks in general and of plant networks in particular.

Financial support

This research was funded in part by a grant from the National Science Foundation (MCB-1900567).

Conflict of interest

The authors declare no conflict of interest.

Authorship contributions

W.-J.R. conceived the outline of the review and coordinated the writing process. A.K. extended the Boolink software. W.-J.R. and A.K. wrote the article.

Data availability statement

No new data or code is presented in this article.

References

Aho, A. V., Garey, M. R., & Ullman, J. D. (1972). The transitive reduction of a directed graph. SIAM Journal on Computing, 1, 131137.CrossRefGoogle Scholar
Albert, I., Thakar, J., Li, S., Zhang, R., & Albert, R. (2008). Boolean network simulations for life scientists. Source Code for Biology and Medicine, 3, 18.CrossRefGoogle ScholarPubMed
Albert, R., Acharya, B. R., Jeon, B. W., Zañudo, J. G. T., Zhu, M., Osman, K., & Assmann, S. M. (2017). A new discrete dynamic model of ABA-induced stomatal closure predicts key feedback loops. PLoS Biology, 15, e2003451.CrossRefGoogle ScholarPubMed
Aracena, J., Goles, E., Moreira, A., & Salinas, L. (2009). On the robustness of update schedules in Boolean networks. Biosystems, 97, 18.CrossRefGoogle ScholarPubMed
Aracena, J., Richard, A., & Salinas, L. (2017). Number of fixed points and disjoint cycles in monotone Boolean networks. SIAM Journal on Discrete Mathematics, 31, 17021725.CrossRefGoogle Scholar
Assmann, S. M., & Jegla, T. (2016). Guard cell sensory systems: Recent insights on stomatal responses to light, abscisic acid, and CO2 . Current Opinion in Plant Biology, 33, 157167.CrossRefGoogle Scholar
Azpeitia, E., Benítez, M., Vega, I., Villarreal, C., & Alvarez-Buylla, E. R. (2010). Single-cell and coupled GRN models of cell patterning in the Arabidopsis thaliana root stem cell niche. BMC Systems Biology, 4, 119.CrossRefGoogle ScholarPubMed
Azpeitia, E., Weinstein, N., Benítez, M., Mendoza, L., & Alvarez-Buylla, E. R. (2013). Finding missing interactions of the Arabidopsis thaliana root stem cell niche gene regulatory network. Frontiers in Plant Science, 4, 110.CrossRefGoogle ScholarPubMed
Bock, M., Scharp, T., Talnikar, C., & Klipp, E. (2014). BooleSim: An interactive Boolean network simulator. Bioinformatics, 30, 131132.CrossRefGoogle ScholarPubMed
Bonzanni, N., Garg, A., Feenstra, K. A., Schütte, J., Kinston, S., Miranda-Saavedra, D., Heringa, J., Xenarios, I., & Göttgens, B. (2013). Hard-wired heterogeneity in blood stem cells revealed using a dynamic regulatory network model. Bioinformatics, 29, i80i88.CrossRefGoogle ScholarPubMed
Bornholdt, S. (2008). Boolean network models of cellular regulation: Prospects and limitations. Journal of the Royal Society Interface, 5, S85S94.CrossRefGoogle ScholarPubMed
Brown, K. S., & Sethna, J. P. (2003). Statistical mechanical approaches to models with many poorly known parameters. Physical Review E, 68, 021904.CrossRefGoogle ScholarPubMed
Ceciliato, P. H. O., Zhang, J., Liu, Q., Shen, X., Hu, H., Liu, C., Schäffner, A. R., & Schroeder, J. I. (2019). Intact leaf gas exchange provides a robust method for measuring the kinetics of stomatal conductance responses to abscisic acid and other small molecules in Arabidopsis and grasses. Plant Methods, 15, 110.CrossRefGoogle ScholarPubMed
Chai, L. E., Loh, S. K., Low, S. T., Mohamad, M. S., Deris, S., & Zakaria, Z. (2014). A review on the computational approaches for gene regulatory network construction. Computers in Biology and Medicine, 48, 5565.CrossRefGoogle ScholarPubMed
Cover, T. M. (1999). Elements of information theory. Wiley.Google Scholar
Davidich, M., & Bornholdt, S. (2008). The transition from differential equations to Boolean networks: A case study in simplifying a regulatory network model. Journal of Theoretical Biology, 255, 269277.CrossRefGoogle ScholarPubMed
Delgado, F. M., & Gómez-Vela, F. (2019). Computational methods for gene regulatory networks reconstruction and analysis: A review. Artificial Intelligence in Medicine, 95, 133145.CrossRefGoogle ScholarPubMed
Demongeot, J., Elena, A., & Sené, S. (2008). Robustness in regulatory networks: A multi-disciplinary approach. Acta Biotheoretica, 56, 2749.CrossRefGoogle ScholarPubMed
Demongeot, J., Goles, E., Morvan, M., Noual, M., & Sené, S. (2010). Attraction basins as gauges of robustness against boundary conditions in biological complex systems. PLoS One, 5, e11793.CrossRefGoogle ScholarPubMed
Demongeot, J., & Sené, S. (2020). About block-parallel Boolean networks: A position paper. Natural Computing, 19, 513.CrossRefGoogle Scholar
Dolan, L., Janmaat, K., Willemsen, V., Linstead, P., Poethig, S., Roberts, K., & Scheres, B. (1993). Cellular organisation of the Arabidopsis thaliana root. Development, 119, 7184.CrossRefGoogle ScholarPubMed
Dubeaux, G., Hsu, P.-K., Ceciliato, P. H., Swink, K. J., Rappel, W.-J., & Schroeder, J. I. (2021). Deep dive into CO2-dependent molecular mechanisms driving stomatal responses in plants. Plant Physiology, 187, 20322042.CrossRefGoogle ScholarPubMed
Emmert-Streib, F., Dehmer, M., & Haibe-Kains, B. (2014). Gene regulatory networks and their applications: Understanding biological and medical problems in terms of networks. Frontiers in Cell and Developmental Biology, 2, 38.CrossRefGoogle ScholarPubMed
Engineer, C. B., Hashimoto-Sugimoto, M., Negi, J., Israelsson-Nordström, M., Azoulay-Shemer, T., Rappel, W.-J., Iba, K., & Schroeder, J. I. (2016). CO2 sensing and CO2 regulation of stomatal conductance: Advances and open questions. Trends in Plant Science, 21, 1630.CrossRefGoogle ScholarPubMed
Espinosa-Soto, C., Padilla-Longoria, P., & Alvarez-Buylla, E. R. (2004). A gene regulatory network model for cell-fate determination during Arabidopsis thaliana flower development that is robust and recovers experimental gene expression profiles. The Plant Cell, 16, 29232939.CrossRefGoogle ScholarPubMed
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, e124e131.CrossRefGoogle ScholarPubMed
Fiers, M. W., Minnoye, L., Aibar, S., Bravo González-Blas, C., Kalender Atak, Z., & Aerts, S. (2018). Mapping gene regulatory networks from single-cell omics data. Briefings in Functional Genomics, 17, 246254.CrossRefGoogle ScholarPubMed
Fumia, H. F., & Martins, M. L. (2013). Boolean network model for cancer pathways: Predicting carcinogenesis and targeted therapy outcomes. PLoS One, 8, e69008.CrossRefGoogle ScholarPubMed
Garg, A., Di Cara, A., Xenarios, I., Mendoza, L., & De Micheli, G. (2008). Synchronous versus asynchronous modeling of gene regulatory networks. Bioinformatics, 24, 19171925.CrossRefGoogle ScholarPubMed
Glass, L., & Kauffman, S. A. (1973). The logical analysis of continuous, non-linear biochemical control networks. Journal of Theoretical Biology, 39, 103129.CrossRefGoogle ScholarPubMed
Gonzalez, A. G., Naldi, A., Sanchez, L., Thieffry, D., & Chaouiya, C. (2006). GINsim: A software suite for the qualitative modelling, simulation and analysis of regulatory networks. Biosystems, 84, 91100.CrossRefGoogle ScholarPubMed
Gupta, S., Bisht, S. S., Kukreti, R., Jain, S., & Brahmachari, S. K. (2007). Boolean network analysis of a neurotransmitter signaling pathway. Journal of Theoretical Biology, 244, 463469.CrossRefGoogle ScholarPubMed
Gutenkunst, R. N., Waterfall, J. J., Casey, F. P., Brown, K. S., Myers, C. R., & Sethna, J. P. (2007). Universally sloppy parameter sensitivities in systems biology models. PLoS Computational Biology, 3, e189.CrossRefGoogle ScholarPubMed
Helikar, T., & Rogers, J. A. (2009). ChemChains: A platform for simulation and analysis of biochemical networks aimed to laboratory scientists. BMC Systems Biology, 3, 115.CrossRefGoogle ScholarPubMed
Herrmann, F., Groß, A., Zhou, D., Kestler, H. A., & Kühl, M. (2012). A Boolean model of the cardiac gene regulatory network determining first and second heart field identity. PLoS One, 7, e46798.CrossRefGoogle ScholarPubMed
Hsu, P.-K., Dubeaux, G., Takahashi, Y., & Schroeder, J. I. (2021). Signaling mechanisms in abscisic acid-mediated stomatal closure. The Plant Journal, 105, 307321.CrossRefGoogle ScholarPubMed
Hsu, P.-K., Takahashi, Y., Munemasa, S., Merilo, E., Laanemets, K., Waadt, R., Pater, D., Kollist, H., & Schroeder, J. I. (2018). Abscisic acid-independent stomatal CO2 signal transduction pathway and convergence of CO2 and ABA signaling downstream of OST1 kinase. Proceedings of the National Academy of Sciences of the United States of America, 115, E9971E9980.Google ScholarPubMed
International Human Genome Sequencing Consortium. (2001). Initial sequencing and analysis of the human genome. Nature, 409, 860921.CrossRefGoogle Scholar
Karanam, A., He, D., Hsu, P.-K., Schulze, S., Dubeaux, G., Karmakar, R., Schroeder, J. I., & Rappel, W.-J. (2021). Boolink: A graphical interface for open access Boolean network simulations and use in guard cell CO2 signaling. Plant Physiology, 187, 23112322.CrossRefGoogle ScholarPubMed
Karlebach, G., & Shamir, R. (2008). Modelling and analysis of gene regulatory networks. Nature Reviews Molecular Cell Biology, 9, 770780.CrossRefGoogle ScholarPubMed
Karmakar, R., Tang, M.-H., Yue, H., Lombardo, D., Karanam, A., Camley, B. A., Groisman, A., & Rappel, W.-J. (2021). Cellular memory in eukaryotic chemotaxis depends on the background chemoattractant concentration. Physical Review E, 103, 012402.CrossRefGoogle ScholarPubMed
Kauffman, S. (1969a). Homeostasis and differentiation in random genetic control networks. Nature, 224, 177178.CrossRefGoogle ScholarPubMed
Kauffman, S. A. (1969b). Metabolic stability and epigenesis in randomly constructed genetic nets. Journal of Theoretical Biology, 22, 437467.CrossRefGoogle ScholarPubMed
Kitano, H. (2002). Computational systems biology. Nature, 420, 206210.CrossRefGoogle ScholarPubMed
Klamt, S., Saez-Rodriguez, J., & Gilles, E. D. (2007). Structural and functional analysis of cellular networks with CellNetAnalyzer. BMC Systems Biology, 1, 113.CrossRefGoogle ScholarPubMed
Klarner, H., Streck, A., & Siebert, H. (2017). PyBoolNet: A python package for the generation, analysis and visualization of Boolean networks. Bioinformatics, 33, 770772.CrossRefGoogle ScholarPubMed
Krawitz, P., & Shmulevich, I. (2007). Basin entropy in Boolean network ensembles. Physical Review Letters, 98, 158701.CrossRefGoogle ScholarPubMed
Lähdesmäki, H., Shmulevich, I., & Yli-Harja, O. (2003). On learning gene regulatory networks under the Boolean network model. Machine Learning, 52, 147167.CrossRefGoogle Scholar
Lau, K.-Y., Ganguli, S., & Tang, C. (2007). Function constrains network architecture and dynamics: A case study on the yeast cell cycle Boolean network. Physical Review E, 75, 051907.CrossRefGoogle ScholarPubMed
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, 47814786.CrossRefGoogle ScholarPubMed
Li, S., Assmann, S. M., & Albert, R. (2006). Predicting essential components of signal transduction networks: A dynamic model of guard cell abscisic acid signaling. PLoS Biology, 4, e312.CrossRefGoogle ScholarPubMed
Liang, S., Fuhrman, S., & Somogyi, R. (1998). Reveal, a general reverse engineering algorithm for inference of genetic network architectures. Biocomputing, 3, 1829.Google Scholar
Lombardo, D. M., & Rappel, W.-J. (2017). Systematic reduction of a detailed atrial myocyte model. Chaos: An Interdisciplinary Journal of Nonlinear Science, 27, 093914.CrossRefGoogle ScholarPubMed
Maheshwari, P., Assmann, S. M., & Albert, R. (2020). A guard cell abscisic acid (ABA) network model that captures the stomatal resting state. Frontiers in Physiology, 11, 927.CrossRefGoogle ScholarPubMed
Maheshwari, P., Du, H., Sheen, J., Assmann, S. M., & Albert, R. (2019). Model-driven discovery of calcium-related protein-phosphatase inhibition in plant guard cell signaling. PLoS Computational Biology, 15, e1007429.CrossRefGoogle ScholarPubMed
Mano, M. M., & Kime, C. R. (1997). Logic and computer design fundamentals. Prentice-Hall, Inc. Google Scholar
Mendoza, L., & Alvarez-Buylla, E. R. (1998). Dynamics of the genetic regulatory network for Arabidopsis thaliana flower morphogenesis. Journal of Theoretical Biology, 193, 307319.CrossRefGoogle Scholar
Mendoza, L., & Xenarios, I. (2006). A method for the generation of standardized qualitative dynamical systems of regulatory networks. Theoretical Biology and Medical Modelling, 3, 118.CrossRefGoogle ScholarPubMed
Merilo, E., Jalakas, P., Kollist, H., & Brosché, M. (2015). The role of ABA recycling and transporter proteins in rapid stomatal responses to reduced air humidity, elevated CO2, and exogenous ABA. Molecular Plant, 8, 657659.CrossRefGoogle ScholarPubMed
Merkel, D. (2014). Docker: Lightweight Linux containers for consistent development and deployment. Linux Journal, 2014, 2.Google Scholar
Mori, F., & Mochizuki, A. (2017). Expected number of fixed points in Boolean networks with arbitrary topology. Physical Review Letters, 119, 028301.CrossRefGoogle ScholarPubMed
Mortveit, H., & Reidys, C. (2007). An introduction to sequential dynamical systems. Springer Science & Business Media.Google Scholar
Munemasa, S., Hauser, F., Park, J., Waadt, R., Brandt, B., & Schroeder, J. I. (2015). Mechanisms of abscisic acid-mediated control of stomatal aperture. Current Opinion in Plant Biology, 28, 154162.CrossRefGoogle ScholarPubMed
Müssel, C., Hopfensitz, M., & Kestler, H. A. (2010). BoolNet—An R package for generation, reconstruction and analysis of Boolean networks. Bioinformatics, 26, 13781380.CrossRefGoogle ScholarPubMed
Novák, B., & Tyson, J. J. (2008). Design principles of biochemical oscillators. Nature Reviews Molecular Cell Biology, 9, 981991.CrossRefGoogle ScholarPubMed
Ouellet, L., Laidler, K. J., & Morales, M. F. (1952). Molecular kinetics of muscle adenosine triphosphatase. Archives of Biochemistry and Biophysics, 39, 3750.CrossRefGoogle ScholarPubMed
Pal, R., Datta, A., Bittner, M. L., & Dougherty, E. R. (2005). Intervention in context-sensitive probabilistic Boolean networks. Bioinformatics, 21, 12111218.CrossRefGoogle ScholarPubMed
Pardal, R., & Heidstra, R. (2021). Root stem cell niche networks: It’s complexed! Insights from Arabidopsis. Journal of Experimental Botany, 72, 67276738.CrossRefGoogle ScholarPubMed
Paulevé, L. (2017). Pint: A static analyzer for transient dynamics of qualitative networks with IPython interface. In Feret, J. & Koeppl, H. (Eds.), International conference on computational methods in systems biology (pp. 309316). Springer.Google Scholar
Pieterse, C. M., Zamioudis, C., Berendsen, R. L., Weller, D. M., Van Wees, S. C., & Bakker, P. A. (2014). Induced systemic resistance by beneficial microbes. Annual Review of Phytopathology, 52, 347375.CrossRefGoogle ScholarPubMed
Pollard, T. D. (1986). Rate constants for the reactions of ATP-and ADP-actin with the ends of actin filaments. The Journal of Cell Biology, 103, 27472754.CrossRefGoogle ScholarPubMed
Raghavendra, A. S., Gonugunta, V. K., Christmann, A., & Grill, E. (2010). ABA perception and signalling. Trends in Plant Science, 15, 395401.CrossRefGoogle ScholarPubMed
Remy, E., Ruet, P., Mendoza, L., Thieffry, D., & Chaouiya, C. (2006). From logical regulatory graphs to standard petri nets: Dynamical roles and functionality of feedback circuits. In Priami, C., Ingólfsdóttir, A., Mishra, B. & Nielson, H. R. (eds.), Transactions on computational systems biology VII (pp. 5672). Springer.CrossRefGoogle Scholar
Ross, S. M. (2014). Introduction to probability models. Academic Press.Google Scholar
Rozum, J. C., Deritei, D., Park, K. H., Gómez Tejeda Zañudo, J., & Albert, R. (2022). pystablemotifs: Python library for attractor identification and control in Boolean networks. Bioinformatics, 38, 14651466.CrossRefGoogle ScholarPubMed
Rozum, J. C., Gómez Tejeda Zañudo, J., Gan, X., Deritei, D., & Albert, R. (2021). Parity and time reversal elucidate both decision-making in empirical models and attractor scaling in critical Boolean networks. Science Advances, 7, eabf8124.CrossRefGoogle ScholarPubMed
Ruz, G. A., Goles, E., & Sené, S. (2018). Reconstruction of Boolean regulatory models of flower development exploiting an evolution strategy. In 2018 IEEE congress on evolutionary computation (CEC) (pp. 18). IEEE.Google Scholar
Ruz, G. A., Timmermann, T., & Goles, E. (2015). Reconstruction of a GRN model of salt stress response in Arabidopsis using genetic algorithms. In 2015 IEEE conference on computational intelligence in bioinformatics and computational biology (CIBCB) (pp. 18). IEEE.Google Scholar
Saadatpour, A., Albert, I., & Albert, R. (2010). Attractor analysis of asynchronous Boolean models of signal transduction networks. Journal of Theoretical Biology, 266, 641656.CrossRefGoogle ScholarPubMed
Saez-Rodriguez, J., Simeoni, L., Lindquist, J. A., Hemenway, R., Bommhardt, U., Arndt, B., Haus, U.-U., Weismantel, R., Gilles, E. D., Klamt, S., & Schraven, B. (2007). A logical model provides insights into T cell receptor signaling. PLoS Computational Biology, 3, e163.CrossRefGoogle ScholarPubMed
Schulze, S., Dubeaux, G., Ceciliato, P. H. O., Munemasa, S., Nuhkat, M., Yarmolinsky, D., Aguilar, J., Diaz, R., Azoulay-Shemer, T., Steinhorst, L., Offenborn, J. N., Kudla, J., Kollist, H., & Schroeder, J. I. (2021). A role for calcium dependent protein kinases in differential CO2 and ABA controlled stomatal closing and low CO2 induced stomatal opening in Arabidopsis. New Phytologist, 229, 27652779.CrossRefGoogle ScholarPubMed
Schwab, J. D., & Kestler, H. A. (2018). Automatic screening for perturbations in Boolean networks. Frontiers in Physiology, 9, 431.CrossRefGoogle ScholarPubMed
Schwab, J. D., Kühlwein, S. D., Ikonomi, N., Kühl, M., & Kestler, H. A. (2020). Concepts in Boolean network modeling: What do they all mean? Computational and Structural Biotechnology Journal, 18, 571582.CrossRefGoogle ScholarPubMed
Schwartz, A., Ilan, N., & Grantz, D. A. (1988). Calcium effects on stomatal movement in Commelina communis L: Use of EGTA to modulate stomatal response to light, KCl and CO2 . Plant Physiology, 87, 583587.CrossRefGoogle Scholar
Sherekar, S., & Viswanathan, G. A. (2021). Boolean dynamic modeling of cancer signaling networks: Prognosis, progression, and therapeutics. Computational and Systems Oncology, 1, e1017.CrossRefGoogle Scholar
Shmulevich, I., Dougherty, E. R., Kim, S., & Zhang, W. (2002a). Probabilistic Boolean networks: A rule-based uncertainty model for gene regulatory networks. Bioinformatics, 18, 261274.CrossRefGoogle ScholarPubMed
Shmulevich, I., Dougherty, E. R., & Zhang, W. (2002b). From Boolean to probabilistic Boolean networks as models of genetic regulatory networks. Proceedings of the IEEE, 90, 17781792.CrossRefGoogle Scholar
Shmulevich, I., Yli-Harja, O., Astola, J., & Core, C. (2001). Inference of genetic regulatory networks under the best-fit extension paradigm. In IEEE-EURASIP workshop on nonlinear signal and image processing. Citeseer.Google Scholar
Stahl, Y., Wink, R. H., Ingram, G. C., & Simon, R. (2009). A signaling module controlling the stem cell niche in Arabidopsis root meristems. Current Biology, 19, 909914.CrossRefGoogle ScholarPubMed
Stoll, G., Viara, E., Barillot, E., & Calzone, L. (2012). Continuous time Boolean modeling for biological signaling: Application of Gillespie algorithm. BMC Systems Biology, 6, 118.CrossRefGoogle ScholarPubMed
Storn, R., & Price, K. (1997). Differential evolution—A simple and efficient heuristic for global optimization over continuous spaces. Journal of Global Optimization, 11, 341359.CrossRefGoogle Scholar
Stötzel, C., Röblitz, S., & Siebert, H. (2015). Complementing ODE-based system analysis using Boolean networks derived from an Euler-like transformation. PLoS One, 10, e0140954.CrossRefGoogle ScholarPubMed
Sun, Z., Jin, X., Albert, R., & Assmann, S. M. (2014). Multi-level modeling of light-induced stomatal opening offers new insights into its regulation by drought. PLoS Computational Biology, 10, e1003930.CrossRefGoogle ScholarPubMed
Terfve, C., Cokelaer, T., Henriques, D., MacNamara, A., Goncalves, E., Morris, M. K., van Iersel, M., Lauffenburger, D. A., & Saez-Rodriguez, J. (2012). CellNOptR: A flexible toolkit to train protein signaling networks to data using multiple logic formalisms. BMC Systems Biology, 6, 114.CrossRefGoogle ScholarPubMed
Thakar, J., Saadatpour-Moghaddam, A., Harvill, E. T., & Albert, R. (2009). Constraint-based network model of pathogen–immune system interactions. Journal of the Royal Society Interface, 6, 599612.CrossRefGoogle ScholarPubMed
Thomas, R. (1973). Boolean formalization of genetic control circuits. Journal of Theoretical Biology, 42, 563585.CrossRefGoogle ScholarPubMed
Thomas, R. (1991). Regulatory networks seen as asynchronous automata: A logical description. Journal of Theoretical Biology, 153, 123.CrossRefGoogle Scholar
Timmermann, T., Armijo, G., Donoso, R., Seguel, A., Holuigue, L., & González, B. (2017). Paraburkholderia phytofirmans PsJN protects Arabidopsis thaliana against a virulent strain of Pseudomonas syringae through the activation of induced resistance. Molecular Plant-Microbe Interactions, 30, 215230.CrossRefGoogle ScholarPubMed
Timmermann, T., González, B., & Ruz, G. A. (2020). Reconstruction of a gene regulatory network of the induced systemic resistance defense response in Arabidopsis using Boolean networks. BMC Bioinformatics, 21, 116.CrossRefGoogle ScholarPubMed
Timmermann, T., Poupin, M. J., Vega, A., Urrutia, C., Ruz, G. A., & González, B. (2019). Gene networks underlying the early regulation of Paraburkholderia phytofirmans PsJN induced systemic resistance in Arabidopsis. PLoS One, 14, e0221358.CrossRefGoogle ScholarPubMed
Trairatphisan, P., Mizera, A., Pang, J., Tantar, A. A., Schneider, J., & Sauter, T. (2013). Recent development and biomedical applications of probabilistic Boolean networks. Cell Communication and Signaling, 11, 125.CrossRefGoogle ScholarPubMed
Transtrum, M. K., & Qiu, P. (2014). Model reduction by manifold boundaries. Physical Review Letters, 113, 098701.CrossRefGoogle ScholarPubMed
Tyers, M., & Mann, M. (2003). From genomics to proteomics. Nature, 422, 193197.CrossRefGoogle ScholarPubMed
Velderraín, J. D., Martínez-García, J. C., & Álvarez-Buylla, E. R. (2017). Boolean dynamic modeling approaches to study plant gene regulatory networks: Integration, validation, and prediction. In Kaufmann, K. & Mueller-Roeber, B. (eds.), Plant gene regulatory networks (pp. 297315). Springer.CrossRefGoogle ScholarPubMed
Veliz-Cuba, A., & Laubenbacher, R. (2012). On the computation of fixed points in Boolean networks. Journal of Applied Mathematics and Computing, 39, 145153.CrossRefGoogle Scholar
Waidyarathne, P., & Samarasinghe, S. (2018). Boolean calcium signalling model predicts calcium role in acceleration and stability of abscisic acid-mediated stomatal closure. Scientific Reports, 8, 116.CrossRefGoogle ScholarPubMed
Webb, A. A. R., McAinsh, M. R., Mansfield, T. A., & Hetherington, A. M. (1996). Carbon dioxide induces increases in guard cell cytosolic free calcium. The Plant Journal, 9, 297304.CrossRefGoogle Scholar
Wittmann, D. M., Krumsiek, J., Saez-Rodriguez, J., Lauffenburger, D. A., Klamt, S., & Theis, F. J. (2009). Transforming Boolean models to continuous models: Methodology and application to T-cell receptor signaling. BMC Systems Biology, 3, 121.CrossRefGoogle ScholarPubMed
Zhang, J., De-oliveira Ceciliato, P., Takahashi, Y., Schulze, S., Dubeaux, G., Hauser, F., Azoulay-Shemer, T., Tõldsepp, K., Kollist, H., Rappel, W.-J., & Schroeder, J. I. (2018). Insights into the molecular mechanisms of CO2-mediated regulation of stomatal movements. Current Biology, 28, R1356R1363.CrossRefGoogle Scholar
Zhang, L., Takahashi, Y., Hsu, P.-K., Kollist, H., Merilo, E., Krysan, P. J., & Schroeder, J. I. (2020). FRET kinase sensor development reveals SnRK2/OST1 activation by ABA but not by MeJA and high CO2 during stomatal closure. eLife, 9, e56351.CrossRefGoogle Scholar
Zhang, Y., Zhu, H., Zhang, Q., Li, M., Yan, M., Wang, R., Wang, L., Welti, R., Zhang, W., & Wang, X. (2009). Phospholipase D $\alpha$ 1 and phosphatidic acid regulate NADPH oxidase activity and production of reactive oxygen species in ABA-mediated stomatal closure in Arabidopsis. The Plant Cell, 21, 23572377.CrossRefGoogle Scholar
Zhao, M., He, W., Tang, J., Zou, Q., & Guo, F. (2021). A comprehensive overview and critical evaluation of gene regulatory network inference technologies. Briefings in Bioinformatics, 22, bbab009.CrossRefGoogle ScholarPubMed
Figure 0

Fig. 1. (a) Comparison of the output of a continuous model [equation (1)] and a Boolean model [equation (2)] for the activation of a gene. In the former, the output can take on any value between 0 and 1 and depends on the model parameters, whereas in the latter, the output is either 0 or 1 and is independent of parameters. (b–d) Truth tables of elementary Boolean functions. (b) Identity gate, which copies the value of the input to the output; not gate, which copies the inverted value of the input to the output. (c) or and and gates, which take two inputs. (d) An example of a Boolean function that is a combination of the elementary functions. The output X can be determined by evaluating the parts recursively.

Figure 1

Fig. 2. Examples of Boolean networks. (a) Example of an oscillatory network. Arrows indicate activation and flat-edge symbols indicated inhibition. (b) The components of the network in panel (a) as a function of time, modelled using rate equations [parameters taken from Novák & Tyson (2008): $k_1$=0.1, $k_2$=0.2, $k_3$=0.1, $k_4$=0.05, $k_{-1}$=0.1, S=2, $K_m$=0.01, p=4]. (c) Truth table for synchronous updating of the network shown in panel(a) (d) Modified network in which Y depends on X and Z. (e) Truth tables for synchronous updating of the network shown in panel (d). (f,g) State space and dynamics, represented by arrows, for asynchronous updating of the networks shown in panels (a,c). Fixed point attractors are indicated by red dots while the oscillatory cycle is shown by the red arrows.

Figure 2

Fig. 3. Inference rules for the construction of Boolean networks. Experimental data are synthesised to be represented in graphs with the least number of nodes and edges, that is, as a sparse representation. This sometimes requires an introduction of an intermediary node, as in graphs 1 and 3, but when additional information becomes available, the graph can in fact simplify, as in going from graph 1 to graph 2. For further details, see text (from Li et al., 2006).

Figure 3

Fig. 4. Boolean modelling of gene networks. (a) Example of a putative network that maintains the SCN in Arabidopsis. Arrows indicated activation and flat-edge symbols correspond to repression. For the definition of the different components, see Velderraín et al. (2017). (b) Attractors of the Boolean network shown in panel (a). Green represents an active and red represents an inactive gene. The labels at the top of the diagram represent the attractors and correspond to the phenotypes observed in experiments (CEI, cortex-endodermis initials; CEP, columella epidermis initials; QC, quiescent center; VAS, vascular initials) (from Velderraín et al., 2017). (c) Modified network based on novel experimental and computational results (from Velderraín et al., 2017).

Figure 4

Fig. 5. Signalling network of abscisic acid (ABA)-induced stomatal closure. Arrows indicate positive interactions while filled circles indicate negative interactions. Rectangles represent nodes that are connected to other nodes. Black lines represent direct interactions and green lines represent indirect interactions. Nodes are color coded according to their function: enzymes (red), signalling proteins (green), membrane-transport related nodes (blue), and secondary messengers and small molecules (orange). For names of the components, see original publication (Albert et al., 2017).

Figure 5

Fig. 6. Results of a Boolean ABA network. Shown are the percentage of closure as a function of iteration step. The wild-type (WT) curves show the network response in the absence of (open circles) and presence of ABA (closed circles). Other curves show the response following simulated knockout of the component (node set to 0) in the presence of ABA. For abbreviations, see original study (Albert et al., 2017).

Figure 6

Fig. 7. CO$_2$ and ABA-induced stomatal closure model. (a) Extended network showing the new CO$_2$ branch in blue and the existing ABA network, shown in Figure 5, as a box. Only several of ABA components are shown. (b) Predicted stomatal conductance levels of the network in panel (a) for both CO$_2$=0 (red) and CO$_2$=1 (blue), before and after the application of ABA. (c) Modified CO$_2$ and ABA-induced stomatal closure network, with modifications represented by orange links. (d) Predicted stomatal conductance levels using the network shown in panel (c) for both CO$_2$=0 (red) and CO$_2$=1 (blue), before and after the application of ABA. For names of the components, see original publication (Albert et al., 2017) (from Karanam et al., 2021)

Author comment: Boolean modelling in plant biology — R0/PR1

Comments

Dear Dr. Hamant,

My co-author Aravind Rao Karanam and I are delighted to submit our review entitled “Boolean

modeling in plant biology”. As you will recall, this submission was initiated by your kind

invitation, which you sent last year after reading our recent work on Boolean modeling. We

believe we have created a review that is of interest not only to the readership of your journal

in particular but also to the broader plant biology community.

We look forward to your response.

Best regards,

Wouter-Jan Rappel

Wouter-Jan Rappel, Ph.D.

Department of Physics

UC San Diego

9500 Gilman Drive

La Jolla, CA 92093

Review: Boolean modelling in plant biology — R0/PR2

Conflict of interest statement

Reviewer declares none.

Comments

Comments to Author: The authors have reviewed the use of Boolean network modelling in plant biology, giving a pedagogical introduction and a set of examples from the field. The review is well written and this will serve as a useful reference piece, and is very aligned with QPB’s mission. I do have some comments about the style and structure that I think may help further improve things, outlined below. For transparency, I am Iain Johnston and am happy for this review to be treated as public domain. To my mind, my most important limitations as a reviewer here are:

- I come from the same disciplinary background as the authors, and so may not be the best person to comment on how readily this article may be read by biologists unfamiliar with these ideas (although this is my major comment)

- There are several papers in the authors' set of examples with which I’m not familiar.

Awkwardly, and related, my most important point isn’t a single actionable change. At the moment the review is a very good introduction to Boolean modelling in plant biology -- *for physicists*. There are several clear instances where the target audience of the article appears to be physicists, not biologists. This is apparent in, for example, some jargon (“disjoint”, “sparsest graph”); leaning on ODEs as a pedagogical “stepping stone” to introduce an example. Some of these words -- and ideas -- will be unknown to the biologists reading the article as an introduction to the topic.

How best to address this? A glossary would be one option. Another would be to give a layman’s introduction to such terms and ideas as they arise -- though this would need more words. Other options may be possible too. But perhaps a useful exercise might be to read through the article from the perspective of a biology PhD student. They’ve done maths maybe up to the age of 18, they’ve heard of ODEs but don’t gain immediate intuition about a system from looking at them, they haven’t done graph or set theory, and are more used generally to thinking qualitatively than quantitatively. Can their hands be held more through the concepts that are bread-and-butter to physicists, but unusual to biologists?

Other points --

The authors focus on logic gates as the key architectures in BNs, with expressions like C = A AND B. A couple of thoughts:

1. The equality sign is a bit odd here (and throughout), because we’re fundamentally talking about update rules. Perhaps C -> A AND B, or C_new = A AND B would make this clearer? (as before the update rule is applied, the equality doesn’t necessarily hold)

2. Can the authors include the (albeit rather trivial) identity operation (A -> B) in their introduction? This is extremely common in biology (transcription factors acting as activators) -- and indeed comes up in their examples.

The introduction doesn’t mention parametric Boolean models like Boolean Threshold Dynamics, where edges are given (positive or negative) weights which act as coefficients for the incoming node states in a function that determines the update step. Although any static instance of such a model can (of course) be mapped to a model using logic gates alone, the parametric versions are commonly used and merit an explanation.

Fig 3 is pretty unclear as a standalone object. Can the experimental and inferred structures be distinguished (eg by colour) and the caption expanded? “For details, see text” is awkward for a reader viewing the figures as an extended abstract. Also -- are the figures taken directly from the source papers? There may be copyright issues here if so? Some other figures -- Fig 7 in particular -- aren’t as good as they could be. In Fig 7 B,D the text is tiny, lots of whitespace, and the takehome from the time series isn’t immediately apparent from the figure.

“Boolean networks and plant genetics” is an odd title, and the reference to “genetics” throughout also sit awkwardly. Most readers will think inheritance, chromosomes, breeding, mutations when reading “plant genetics”. I suspect the emphasis the authors mean is “gene *regulatory* networks”. If so, can “genetics” be replaced throughout with “gene regulation” or similar?

The point on p13-14 about the computational ease of knockouts or other manipulations probably deserves promotion to the introduction, as it’s a key and general strength of the modelling approach.

If I may, I’d like to suggest a classic citation that would point the interested reader both to more information about Boolean modelling and the broader spectrum of modelling approaches for gene regulation: https://www.nature.com/articles/nrm2503

Did the ISR case study do anything more than match experimental data? If there was additional insight that came from the modelling approach it’d be nice to hear about it.

Some Bibtex bugs mean that several internal references aren’t clear (“Sec. ”).

Review: Boolean modelling in plant biology — R0/PR3

Conflict of interest statement

Reviewer declares none.

Comments

Comments to Author: Although Boolean networks are simple models of gene regulatory networks, they help capture qualitative information about the gene regulatory processes. Also, the model presents very interesting mathematical properties. In this sense, I believe the submitted review paper is of interest, contributing to this field.

For completeness purposes, I think the following sections need to improve with the following comments:

Update rules

1-It is important to point out in this section that the number of possible updating rules is exponential. Indeed, for a network with n nodes, the number of updates is given by a recursive formula (Proposition 5) that appears in:

Demongeot, J., Elena, A., Sené, S., 2008. Robustness in regulatory networks: A multidisciplinary approach. Acta Biotheoretica 56, 27–49.

For example, for n=10, there are 102,247,563 different updating rules.

2-A popular updating rule in biological applications, like the Arabidopsis thaliana network, is the Block-Sequential updating rule which is not discussed in this section. In this case, the set of nodes for a given sequence is partitioned into blocks. The nodes in the same block are updated synchronously, but blocks follow each other sequentially. For example, in the Arabidopsis thaliana network, the updating rule is (EMF1, TFL1)(LFY, AP1, CAL)(LUG, UFO, BFU)(AG, AP3, PI)(SUP).

Encoding a Boolean network from experiments

3- A review of classical methods to infer Boolean networks from data is missing, like REVEAL (and variants), Best-Fit extension algorithm, etc.

4- A review of more recent approaches to infer Boolean networks from data using evolutionary computation is missing.

Dynamics of Boolean Networks

5- Derterming or counting the number of fixed points a Boolean network can have is an active research topic. There have been several works published in this field. Authors should consider reviewing some.

Recommendation: Boolean modelling in plant biology — R0/PR4

Comments

Comments to Author: Dear Authors

Thank-you for your submission to QPB. We have received 2 expert reviews on this manuscript providing constructive feedback on how to improve the text.

The manuscript is well written and organized, covering relevant topics on Boolean modelling towards understanding processes in plants. The comments of reviewer 1 are particularly pertinent in terms of the extent to which the target audience has background knowledge in maths and graph/set theory. An effort to simplify the language and presentation of the material would enhance the reach and penetrance of the work.

We look forward to reading this manuscript in a revised form.

Decision: Boolean modelling in plant biology — R0/PR5

Comments

No accompanying comment.

Author comment: Boolean modelling in plant biology — R1/PR6

Comments

No accompanying comment.

Review: Boolean modelling in plant biology — R1/PR7

Conflict of interest statement

Reviewer declares none.

Comments

Comments to Author: To my eyes the manuscript is substantially more accessible from a biologist’s perspective, and I believe the authors' edits have clarified several points. Happy to recommend acceptance, and thanks to the authors for a useful reference -- I will be using it!

One remaining point from me. The topic of inferring GRNs from data has now been raised. There could be (and is) a whole review article on this topic alone. I recognise that this is not the focus of this article but the methods that the authors mention in their response to Reviewer 2 seem rather sparse and dated (early 2000s) -- the field, and technology, has advanced dramatically since then. I recommend including a link to a recent review on the topic for the interested reader to follow. I notice (without endorsing it, as I haven’t read it) that this reference exists for example

https://academic.oup.com/bib/article-abstract/22/5/bbab009/6128842

Review: Boolean modelling in plant biology — R1/PR8

Conflict of interest statement

Reviewer declares none.

Comments

Comments to Author: The revised version satisfactorily addresses my previous comments, good work.

Recommendation: Boolean modelling in plant biology — R1/PR9

Comments

Comments to Author: The authors have sufficiently address the concerns of the reviewers.

Decision: Boolean modelling in plant biology — R1/PR10

Comments

No accompanying comment.

Author comment: Boolean modelling in plant biology — R2/PR11

Comments

No accompanying comment.

Recommendation: Boolean modelling in plant biology — R2/PR12

Comments

No accompanying comment.

Decision: Boolean modelling in plant biology — R2/PR13

Comments

No accompanying comment.