Theory

Exact stochastic simulation algorithm (SSA) without delays

Consider a system consisting of $N \geq 1$ chemical species, $\{X_1,\ldots, X_N\}$, undergoing $M \geq 1$ chemical reactions through reaction channels $\{R_1,\ldots,R_M\}$, each of which is equipped with a propensity function (or intensity function in the mathematics literature), $a_k(X)$. The dynamic state of this chemical system can be described by the state vector $X(t) =[X_1(t),\ldots,X_N(t)]^T$, where $X_n(t),n = 1,\ldots,N,$ is the number of $X_n$ molecules at time $t$, and $[·]^T$ denotes the transpose of the vector in the bracket.

Following Gillespie [1], the dynamics of reaction $R_k$ are defined by a state-change vector $\nu_k = [\nu_{1k} ,\ldots,\nu_{Nk}]^T$, where $\nu_{nk}$ gives the changes in the $X_n$ molecular population produced by one $R_k$ reaction, and a propensity function $a_k(t)$, which can also be dependent on the state, i.e. $a_k(t)=\hat{a}_{k}(t,X(t))$ (here for simplicity denoted as $a_k(t)$), together with the fundamental premise of stochastic chemical kinetics:

\[\begin{equation} \begin{aligned} a_k(t)\Delta t = &\text{ the probability, given } X(t)=\mathbf{x}, \\ &\text{ that one reaction }R_k \text{ will occur in the}\\ &\text{ next infinitesimal time interval }[t,t+\Delta t]. \end{aligned} \end{equation}\]

Defining the probability rate constant $c_k$ as the probability that a randomly selected combination of $R_k$ reactant molecules reacts in a unit time period, we can calculate $a_k(t)$ from $c_k$ and the molecular numbers of $R_k$ reactants at time $t$ using the method given by Gillespie.

For a chemical system in a given state $X(t)=\mathbf{x}$ at time $t$, assuming that all reactions occur instantly, Gillespie’s exact SSA answers the following two questions: (i) when will the next reaction occur? (ii) which reaction will occur? Specifically, Gillespie’s exact SSA simulates the following event in each step:

\[\begin{equation} \begin{aligned} \text{E: } & \text{no reaction occurs in the time interval }[t,t+\Delta],\\ & \text{and a reaction }R_\mu \ \text{occurs in the infinitesimal}\\ & \text{time interval }[t+\Delta,t+\Delta+\Delta t]. \end{aligned} \end{equation}\]

Based upon the fundamental premise Eq. (1), Gillespie devised an exact stochastic simulation algorithm (SSA) that involves drawing two random numbers $\Delta$ and $\mu$ which are independent and distributed according to probability density functions (PDFs) (3) and (4), respectively. $\Delta$ tells us how long we have to wait until the next reaction and $\mu$ tells us which is the next reaction:

\[\begin{equation} f_\Delta(\Delta)=a_0(t) \exp(-a_0(t)\Delta),\ \ \ \ \Delta > 0, \end{equation}\]

and

\[\begin{equation} f_\mu(\mu)={{a_\mu(t)} \over {a_0(t)}},\ \ \ \ \ \mu = 1,\ldots,M, \end{equation}\]

where $a_0(t)=\begin{matrix} \sum_{k=1}^M a_k(t) \end{matrix}$. According to the PDF Eq. (4), a realization of $\mu$ can be generated from a standard uniform random variable $r_2$, by taking $\mu$ to be the integer for which $\sum_{k=1}^{\mu-1} a_k(t) < r_2 a_0(t) \leq \sum_{k=1}^\mu a_k(t)$;based on the PDF Eq. (3), a realization of $\Delta$ can be generated from another standard uniform random variable $r_1$ as $\Delta=−\ln(r_1)/a_0(t)$. Therefore, Gillespie’s exact SSA generates a realization of $\mu$ and $\Delta$ in each step of simulation, and then updates the time and system state as $t\leftarrow t+\Delta$ and $\mathbf{x} \leftarrow \mathbf{x}+ \mathbf{\nu_\mu}$, respectively.

Exact SSA for stochastic modelling coupled with delays

Delay direct method

As in the derivation of Gillespie’s exact SSA, we first need to find the probability of event Eq. (2), that is defined as $P(\Delta,\mu)\text{d}\Delta$, where $P(\Delta,\mu)$ is the joint PDF of $\Delta$ and $\mu$. Suppose that there are $d$ ongoing reactions at time $t$, which will finish at $t+T_1,\ldots,t+T_{d}$, respectively. Without loss of generality, we assume that $T_1 \leq T_2 \leq \ldots \leq T_d$. Unlike in the reaction system without delays where the propensity functions remain unchanged in the time interval $[t,t+\Delta]$, the propensity functions here change at $t+T_i,i=1,\ldots,d$, due to delayed reactions. We need to take into account such changes in propensity functions when deriving $P(\Delta,\mu)$.

As in the derivation of Gillespie’s exact SSA, $P(\Delta,\mu)\text{d}\Delta$ can be found from the fundamental premise Eq. (1) as

\[\begin{equation} P(\Delta,\mu)\text{d}\Delta=P_0(\Delta) a_\mu(t +\Delta)\text{d}\Delta, \end{equation}\]

where $P_0(\Delta)$ is the probability that no reaction will occur in the time interval $[t,t+\Delta]$, while $a_\mu(t+\Delta)\text{d}\Delta$ is the probability that a reaction $R_\mu$ occurs in $[t+\Delta,t+\Delta+\text{d}\Delta]$. Defining $T_0=0$ and $T_{d+1}=\infty$, we can find $P_0(\Delta)$ for $\Delta$ that lies in different time intervals $[T_i,T_{i+1}),i=0,\ldots,d$. If $\Delta \in [T_i,T_{i+1})$, we define the event $E_k$ as the event that no reaction occurs in the time interval $[t+T_k,t+T_{k+1}),k=0,\ldots,k=i−1$, respectively, and the event $E_i$ as the event that no reaction occurs in the time interval $[t+T_i,t+\Delta)$. Then, we can express $P_0(\Delta)$ as

\[\begin{equation} P_0(\Delta)=P(E_0,\ldots,E_i)=P(E_0) \prod_{k=1}^i P(E_k丨E_0,\ldots,E_{k-1}). \end{equation}\]

From the derivation of Gillespie’s exact SSA, we know that

\[P(E_0) = \exp (−a_0(t)T_1),\\ P_0(E_k丨E_0,\ldots,E_{k-1}) = \exp(-a_0(t+T_k)) × (T_{k+1}−T_k),k=0,\ldots,i−1,\\ P(E_i丨E_0,\ldots,E_{i-1}) = \exp(-a_0(t+T_i)(\Delta-T_i)).\]

Notice that propensity functions change at $t+T_k$ after a delayed reaction finishes, and we use $a_0(t+T_k)$ to represent the new $a_0$. The probability $P_0(\Delta)$ is then given by

\[\begin{equation} \begin{aligned} & P_0(\Delta) = \exp \bigg (-\sum_{k=0}^{i-1} a_0(t+T_k)(T_{k+1}-T_k)-a_0(t+T_i)(\Delta-T_i) \bigg ), \\ & \Delta \in [T_i,T_{i+1}), i = 0,\ldots,d, \end{aligned} \end{equation}\]

where we assume that the first term of the exponent is equal to zero when $i = 0$. Since $P_0(\Delta)$ does not depend on individual propensity functions, as shown in Eq. (7), it is seen from Eq. (5) that $\Delta$ and $\mu$ are independent random variables. Combining Eq. (5) and Eq. (7) and noticing that $a_\mu(t+\Delta)=a_\mu(t+T_i)$ for $\Delta \in [T_i,T_i+1)$, we obtain the PDF of $\Delta$ and $\mu$ as follows:

\[\begin{equation} \begin{aligned} & f_\Delta(\Delta) = a_0(t+T_i) \exp \bigg (-\sum_{k=0}^{i-1} a_0(t+T_k)(T_{k+1}-T_k) - a_0(t+T_i)(\Delta-T_i) \bigg ), \\ & \Delta \in [T_i,T_{i+1}), i = 0,\ldots,d, \end{aligned} \end{equation}\]

and

\[f_\mu(\mu)={ {a_\mu(t+T_i)} \over {a_0(t+T_i)} },\ \ \ \mu = 1,\ldots,M,\ \ \ \Delta \in [T_i,T_{i+1}).\]

It is not difficult to verify that $\int_{0}^{\infty} f_\Delta(\Delta)\, \text{d}\Delta = 1$. In simulation, $\mu$ can be generated, from a standard uniform random variable $u_1$, by taking $\mu$ to be the integer for which

\[\sum_{k=1}^{\mu-1} a_k(t+T_i) < u_1 a_0(t+T_i) ≤ \sum_{k=1}^\mu a_k(t+T_i),\]

after $\Delta$ is generated to be in the time interval $[T_i,T_{i+1})$. We next derive the method of generating $\Delta$ according to its PDF in Eq. (8).

The cumulative distribution function of $\Delta$ can be found from Eq. (8) as

\[\begin{aligned} & F_\Delta(\Delta)=1 - \exp \bigg (-\sum_{k=0}^{i-1} a_0(t+T_k)(T_{k+1}-T_k)-a_0(t+T_i)(\Delta-T_i) \bigg ), \\ & \Delta \in [T_i,T_{i+1}), i = 0,\ldots,d. \end{aligned}\]

Then, we can generate $\Delta$ from a standard uniform random variable $u_2$, by taking $\Delta=F_\Delta^{−1}(u_2)$, where $F_\Delta^{−1}(\cdot)$ represents the inverse of $F_\Delta(\Delta)$. More specifically, we can obtain $\Delta$ as follows:

Find $T_i$ such that $F_\Delta(T_i) ≤ u_2 ≤ F_\Delta(T_{i+1})$, then calculate $\Delta$ from

\[\begin{aligned} & \Delta = T_i + {{-\ln (1-u_2)-\begin{matrix} \sum_{k=0}^{i-1} a_0(t+T_k)(T_{k+1}-T_k) \end{matrix} } \over {a_0(t+T_i)}} \\ & \Delta \in [T_i,T_{i+1}). \end{aligned}\]

Since we need $T_1,\ldots,T_d$ to generate $\Delta$ and $\mu$, we define an array of data structures, named Tstruct, whose $i$th $(i=1,\ldots,d)$ cell stores $T_i$ and the index, $\mu_i$, of the reaction that $T_i$ is associated with. The reaction index $\mu_i$ is needed during the generation of $\Delta$, when we update the propensity functions affected by the reaction that is delayed but finishes at $t+T_i$. During simulation, we need to generate $\Delta$ and $\mu$, maintain Tstruct, and then update the state vector $X(t)$. We refer to delay direct algorithm for details.

Delay rejection method

Bratsun et al. [3] and Barrio et al. [4] used an algorithm for computing the initiation times that is exactly like the original Gillespie Algorithm except that if there is a stored delayed reaction set to finish within a computed time step, then the computed time step is discarded, and the system is updated to incorporate the stored delayed reaction. The algorithm then attempts another step starting at its new state. This algorithm is called the Rejection Method. We briefly summarized this algorithm in delay rejection method.

The rejection algorithm essentially generates $\Delta$ in the event (2) using a rejection method in an iterative fashion: in the i-th iteration, it generates a $\Delta_i$ according to an exponential PDF with parameter $a_0(t+T_{i−1})$, where we have denoted the time to next event generated in the i-th iteration as $\Delta_i$. If $\Delta_i < T_i - T_{i−1}$, then we have $\Delta = T_{i-1} + \Delta_i$ and the algorithm continues simulation to generate $\mu$; otherwise, it rejects $\Delta_i$, updates the state vector $X(t+T_i)$, calculates $a_k(t+T_i),k=1,\ldots,M$, and goes to the next iteration. If $\Delta$ is determined in the (i+1)-th iteration, where i is a non-negative integer, then we have $\Delta \in [T_i,T_{i+1})$ and i delayed reactions finished in the time interval $[t,t+\Delta)$.

We point out the following equivalence:

Note

For a given integer $i$, the event $\Delta \in [T_i,T_{i+1})$ in the delay direct method is equivalent to the event $\Delta_1,\Delta_2,\ldots,\Delta_i$ are rejected and the time to the next event $\Delta_{i+1}$ is accepted, i.e. $\Delta_{i+1}+T_i=\Delta$ in the delay rejection method.

From the iterative procedure of generating $\Delta$ described above, we can find $P_0(\Delta)$ that the rejection method algorithm produces. Specifically, if $\Delta \in [T_i,T_{i+1})$, we have

\[P(E_0)=P(\Delta_1 > T_1),\\ P(E_k丨E_0,\ldots,E_{k-1}) = P(\Delta_{k+1} > T_{k+1} - T_k), k=1,\ldots,i−1,\]

because $\Delta_k,k=1,\ldots,i$, are rejected. Since $\Delta_{i+1}$ is accepted, at least one reaction will occur in the time interval $[t+T_i,t+\Delta)$, if $\Delta_{i+1} < \Delta −T_i$. Thus, $P(E_i丨E_0,\ldots,E_{i-1}) = 1−P(\Delta_{i+1} < \Delta - T_i) = P(\Delta_{i+1} > \Delta - T_i)$. Therefore, for the rejection method, $P_0(\Delta)$ in Eq. (6) can be written as

\[\begin{equation} P_0(\Delta) = P(\Delta_{i+1} > \Delta - T_i) \prod_{k=1}^i P(\Delta_k > T_k - T_{k-1}). \end{equation}\]

The random variables $\Delta_k,k=1,\ldots,i+1$, follow an exponential distribution with parameter $a_0(t+T_{k−1})$, and thus we have

\[\begin{equation} \begin{aligned} & P(\Delta_k > T_k - T_{k-1}) = \exp(-a_0(t+T_{k-1})(T_k - T_{k-1})) \\ & k= 1,\ldots,i, \end{aligned} \end{equation}\]

and

\[\begin{equation} P(\Delta_{i+1} > \Delta −T_i) = \exp(-a_0(t+T_i)(\Delta-T_i)). \end{equation}\]

Substituting Eqs. (10) and (11) into Eq. (9), we find that $P_0(\Delta)$ in Eq. (9) is exactly the same as $P_0(\Delta)$ in Eq. (7) that is derived directly from the event (2) and the fundamental premise (1). Since the delay direct algorithm generates $\Delta$ and $\mu$ according to PDFs of $\Delta$ and $\mu$ derived from $P_0(\Delta)$ in Eq. (7), the rejection method is equivalent to the direct method and also is an exact SSA for chemical reaction systems with delays.

Other Methods

The rigorous proof of delay rejection method opens a way to transform any exact SSA method to an exact delay SSA method. If we denote dt_reaction as the time to the next reaction generated by the instantaneous reactions $\{R_1,\ldots,R_M\}$ using a given exact stochastic simulation algorithm method, for instance next reaction method [5], direct method with composition and rejection [6-7], partial propensity method [8] etc. To extend such an SSA method to a delay SSA method, it suffices to compare dt_reaction with dt_delay which is the time to the next delay reaction (the minimum delay time in the delay channels). If dt_reaction < dt_delay, we reject dt_delay, otherwise accept dt_delay. Based on the proof in delay rejection method, such an extension guarantees to the exactness of transformed delay SSA method.

References

[1] Daniel T. Gillespie, "Exact stochastic simulation of coupled chemical reactions", The Journal of Physical Chemistry 1977 81 (25), 2340-2361. https://doi.org/10.1021/j100540a008

[2] Xiaodong Cai, "Exact stochastic simulation of coupled chemical reactions with delays", The Journal of Chemical Physics 126, 124108(2007). https://doi/10.1063/1.2710253

[3] Dmitri A. Bratsun, Dmitri N. Volfson, Jeff Hasty, and Lev S. Tsimring "Non-Markovian processes in gene regulation (Keynote Address)", Proc. SPIE 5845, Noise in Complex Systems and Stochastic Dynamics III, (23 May 2005). https://doi.org/10.1117/12.609707

[4] Manuel Barrio, Kevin Burrage, André Leier, Tianhai Tian. "Oscillatory Regulation of Hes1: Discrete Stochastic Delay Modelling and Simulation", PLoS Computational Biology, 10.1371(2006). https://doi.org/10.1371/journal.pcbi.0020117

[5] David F. Anderson, "A modified Next Reaction Method for simulating chemical systems with time dependent propensities and delays", The Journal of Chemical Physics 128, 109903(2008). https://doi/10.1063/1.2799998

[6] Slepoy, Alexander, Aidan P. Thompson, and Steven J. Plimpton. "A constant-time kinetic Monte Carlo algorithm for simulation of large biochemical reaction networks." The journal of chemical physics 128, no. 20 (2008): 05B618. https://doi.org/10.1063/1.2919546

[7] Mauch, Sean, and Mark Stalzer. "Efficient formulations for exact stochastic simulation of chemical systems." IEEE/ACM Transactions on Computational Biology and Bioinformatics 8, no. 1 (2009): 27-35. https://doi.org/10.1109/TCBB.2009.47

[8] Ramaswamy, Rajesh, and Ivo F. Sbalzarini. "A partial-propensity formulation of the stochastic simulation algorithm for chemical reaction networks with delays." The Journal of chemical physics 134, no. 1 (2011): 014106. https://doi.org/10.1063/1.3521496