
Citation: Carsten Conradi, Elisenda Feliu, Maya Mincheva. On the existence of Hopf bifurcations in the sequential and distributive double phosphorylation cycle[J]. Mathematical Biosciences and Engineering, 2020, 17(1): 494-513. doi: 10.3934/mbe.2020027
[1] | LanJiang Luo, Haihong Liu, Fang Yan . Dynamic behavior of P53-Mdm2-Wip1 gene regulatory network under the influence of time delay and noise. Mathematical Biosciences and Engineering, 2023, 20(2): 2321-2347. doi: 10.3934/mbe.2023109 |
[2] | Qiuyan Zhang, Lingling Liu, Weinian Zhang . Bogdanov-Takens bifurcations in the enzyme-catalyzed reaction comprising a branched network. Mathematical Biosciences and Engineering, 2017, 14(5&6): 1499-1514. doi: 10.3934/mbe.2017078 |
[3] | Dan Liu, Shigui Ruan, Deming Zhu . Stable periodic oscillations in a two-stage cancer model of tumor and immune system interactions. Mathematical Biosciences and Engineering, 2012, 9(2): 347-368. doi: 10.3934/mbe.2012.9.347 |
[4] | Pei Yu, Xiangyu Wang . Analysis on recurrence behavior in oscillating networks of biologically relevant organic reactions. Mathematical Biosciences and Engineering, 2019, 16(5): 5263-5286. doi: 10.3934/mbe.2019263 |
[5] | Juenu Yang, Fang Yan, Haihong Liu . Dynamic behavior of the p53-Mdm2 core module under the action of drug Nutlin and dual delays. Mathematical Biosciences and Engineering, 2021, 18(4): 3448-3468. doi: 10.3934/mbe.2021173 |
[6] | Yuan Ma, Yunxian Dai . Stability and Hopf bifurcation analysis of a fractional-order ring-hub structure neural network with delays under parameters delay feedback control. Mathematical Biosciences and Engineering, 2023, 20(11): 20093-20115. doi: 10.3934/mbe.2023890 |
[7] | Shunyi Li . Hopf bifurcation, stability switches and chaos in a prey-predator system with three stage structure and two time delays. Mathematical Biosciences and Engineering, 2019, 16(6): 6934-6961. doi: 10.3934/mbe.2019348 |
[8] | Wenlong Wang, Chunrui Zhang . Bifurcation of a feed forward neural network with delay and application in image contrast enhancement. Mathematical Biosciences and Engineering, 2020, 17(1): 387-403. doi: 10.3934/mbe.2020021 |
[9] | Jinna Lu, Xiaoguang Zhang . Bifurcation analysis of a pair-wise epidemic model on adaptive networks. Mathematical Biosciences and Engineering, 2019, 16(4): 2973-2989. doi: 10.3934/mbe.2019147 |
[10] | Mohammad Sajid, Sahabuddin Sarwardi, Ahmed S. Almohaimeed, Sajjad Hossain . Complex dynamics and Bogdanov-Takens bifurcations in a retarded van der Pol-Duffing oscillator with positional delayed feedback. Mathematical Biosciences and Engineering, 2023, 20(2): 2874-2889. doi: 10.3934/mbe.2023135 |
Protein phosphorylation cycles consist of three proteins, a substrate S and two enzymes K and F. One enzyme, the kinase K, attaches phosphate groups to the substrate and hence phosphorylates the substrate while the other, the phosphatase F, removes phosphate groups and hence dephosphorylates the substrate. Protein phosphorylation cycles are important mechanisms of post translational modification of a protein and as such an integral part of intracellular signaling and control [1]. Often phosphorylation and dephosphorylation follow a sequential and distributive mechanism as depicted in Figure 1a: in each encounter of S and either K or F exactly one binding site is (de)phosphorylated. If either phosphorylation or dephosphorylation follows a processive mechanism, then at least two binding sites are (de)phosphorylated in each encounter of S and either K or F (cf. Figure 1b & 1c). Here we study the sequential and distributive phosphorylation of a protein S at two binding sites as depicted in Figure 1a. Such a study of the behavior of an important biochemical module is of particular interest in the light of studies elucidating the complex behavior of signaling pathways composed of such modules [2].
Mathematical models of both processive and distributive phosphorylation have been extensively studied and it is known that they admit complex dynamics (see e.g. [1,3,4,5] and the many references therein). The mass action model of the sequential and distributive phosphorylation cycle depicted in Figure 1a is arguably one of the – mathematically – best studied and challenging systems of post translational modification: both multistationarity (the existence of at least two positive steady states) and bistability (the existence of two locally stable positive steady states) have been established (cf. for example [6,7] for multistationarity, [8] for bistability). In fact it has been shown that this mass action model admits at most three positive steady states [10,9].
In contrast purely processive phosphorylation cycles have a unique stable steady state [11], while the mixed cycles depicted in Figure 1b and 1c have a unique steady state that might not be stable, and admit oscillations [12,13]. Distributive steps therefore seem to be involved with the emergence of oscillations, in particular as in more involved combinations of distributive and processive steps, oscillations have been reported as well [1,14].
Interestingly, oscillations have not been reported for the cycle depicted in Figure 1a, despite considerable effort by different research groups. One avenue to establish oscillations is to determine values of rate constants and concentration variables where a supercritical Hopf bifurcation occurs, see e.g., [15]. In [16] Hopf bifurcation points have been located in the parameter space of a variety of models from the systems biology literature. In [16,Section 5.36] a simplification of the mass action model of Figure 1a is examined. The authors provide steady state concentration values and rate constants of a candidate Hopf bifurcation point. (The rate constants are provided indirectly since they can be computed from the 'convex parameters' the authors use, cf. Section 3.2). Applying Theorem 1, however, shows that at this point no Hopf bifurcation occurs: the candidate point satisfies the conditions on the Hurwitz determinants but it fails the condition on the constant coefficient of the characteristic polynomial (in our case the last nonzero coefficient of the characteristic polynomial). Thus it cannot be a point of Hopf bifurcation, for more details see the discussion in Section 6. In summary, to the best of our knowledge, for the mass action model of the phosphorylation cycle in Figure 1a nor simplifications of it, neither Hopf bifurcations nor oscillations have been reported to date.
In this paper we work towards solving the problem of existence or non-existence of Hopf bifurcations for the double phosphorylation network. We follow the strategy outlined in [12] and as in [16] we use convex parameters (see Section 3.2). As in [16] we consider simplifications of the mass action model derived from Figure 1a. Specifically we consider all four mass action networks derived from Figure 1a that contain only two (out of four possible) enzyme-substrate complexes. This is done for two reasons: first, the four networks are biochemically interesting and hence worth studying by themselves (cf. Remark 5). Second, while it is trivial to see that three of the models do not admit Hopf bifurcations, the fourth model displays a nice structure that can be explored, even when all parameters treated as unknown, by performing symbolic computations. Treatment of the full model is currently out of reach due to the computational complexity.
The paper is organized as follows: in Section 2 we introduce the notation. We further recall convex parameters and a criterion for purely imaginary roots in Section 3. In Section 4 we motivate the mass action models studied in this paper. In Section 5 we prove that for none of the four mass action models Hopf bifurcations are possible: for each mass action model we show that there do not exist parameter values such that a Hopf bifurcation occurs (this is Theorem 5.1). The proof of the theorem relies on large symbolic computations that we present in the Maple Supplementary File 'SupplMat1.mw' (see also 'SupplMat1.pdf' for a pdf version). In Section 6 we comment in more detail on the candidate point for Hopf bifurcation presented in [16].
We consider systems of n chemical species A1,…,An and r chemical reactions of the form
n∑i=1αijAiκj→n∑i=1βijAi,j=1,2,…r, |
where the integer numbers αij≥0, βij≥0 are the stoichiometric coefficients and κj>0 the rate constants. We use xi to denote the concentration of species Ai. Throughout this paper we will assume that all reactions are governed by mass action kinetics, that is, the reaction rate is proportional to the product of the concentrations of the reacting species raised to the power of their respective molecularities. Then the reaction rate vj(κ,x) of the j-th reaction is
vj(κ,x)=κjn∏i=1xαiji. | (1) |
With this, the above reaction network defines the following system of ordinary differential equations
˙x=Nv(κ,x), | (2) |
where N is the stoichiometric matrix. Here N is of dimension n×r and v(κ,x) of dimension r×1. The ij-th entry of N is given by the difference of the stoichiometric coefficients:
Nij=βij−αij. |
Let diag(κ) be the diagonal r×r matrix of rate constants
diag(κ)=diag(κ1,κ2,…,κr) |
where the κi coordinate of an r-dimensional vector κ is the [i,i] entry of diag(κ). Let Y be the n×r matrix whose column vectors yj contain the stoichiometric coefficients αij of the reactant of the j-th reaction. The matrix Y is sometimes called the kinetic order matrix. Given vectors x,y∈Rn, we use the notation xy=∏ni=1xyii. Then the columns of Y define the monomial vector
ψ(x)=(xy1⋮xyr) |
and v(κ,x) can be written as the product
v(κ,x)=diag(κ)ψ(x). |
For example the reaction network
S0+Kκ1⇌κ2S0Kκ3→S1+K |
consists of the three reactions
S0+Kκ1→S0K,S0Kκ2→S0+K and S0Kκ3→S1+K |
among the four species S0, K, S1 and S0K taken in that order as species A1 through A4. The stoichiometric coefficients of the first reaction are α11=α21=β41=1 and α31=α41=β11=β21=β31=0. The differences βi1−αi1 define the first column of the following stoichiometric matrix (the remaining columns are defined in a similar way):
N=[−110−1110011−1−1]. |
With mass action kinetics, the reaction rate vector is
v(k,x)=(κ1x1x2κ2x4κ3x4)=[κ1000κ2000κ3](x1x2x4x4). |
The kinetic order matrix is
Y=[100100000011]. |
In this subsection we state a criterion in terms of the principal minors of the Hurwitz matrix (Proposition 1, Theorem 1) that we will use to exclude Hopf bifurcations. The criterion follows from the results in [17]. Related results for asserting Hopf bifurcations can be found in [18,19].
For ease of notation consider an ODE system parametrized by a single parameter μ∈R:
˙x=gμ(x), |
where x∈Rs, and gμ(x) varies smoothly in μ and x. Let x∗∈Rs be a steady state of the ODE system for some fixed value μ0, that is, gμ0(x∗)=0. Furthermore, we assume that we have a smooth curve of steady states around μ0:
μ↦x(μ) | (3) |
that is, gμ(x(μ))=0 for all μ close enough to μ0 such that x(μ0)=x∗. By the Implicit Function Theorem, this curves exists if the Jacobian of gμ0(x) evaluated at x∗ is non-singular.
Let J(x(μ),μ) be the Jacobian of gμ(x) evaluated at x(μ). If, as μ varies, a single complex-conjugate pair of eigenvalues of J(x(μ),μ) crosses the imaginary axis while all other eigenvalues remain with nonzero real parts, then a simple Hopf bifurcation occurs at (x(μ),μ). In this case, a limit cycle arises. If the Hopf bifurcation is supercritical and all other eigenvalues have negative real part, then stable periodic solutions are generated for nearby parameter values.
Remark 1. As we pointed out in the informal discussion above, a simple Hopf bifurcation requires that exactly one pair of complex conjugate eigenvalues crosses the imaginary axis when the parameter μ varies. Thus, in particular, there must exist a value μ0 with steady state x(μ0), where the Jacobian J(x(μ0),μ0) has exactly one pair of purely imaginary eigenvalues. More generally, a Hopf bifurcation necessitates J(x(μ0),μ0), for some value μ0, to have a pair of purely imaginary eigenvalues.
The following proposition gives necessary and sufficient conditions for the existence of exactly one pair of purely imaginary eigenvalues in the scenario we encounter later on. It is based on Hurwitz matrices, which we define first:
Definition 1. The i-th Hurwitz matrix of a univariate polynomial p(z)=a0zs+a1zs−1+⋯+as is the following i×i matrix:
Hi=(a1a0000⋯0a3a2a1a00⋯0⋮⋮⋮⋮⋮⋮a2i−1a2i−2a2i−3a2i−4a2i−5⋯ai), |
in which the (k,l)-th entry is a2k−l as long as 0≤2k−l≤s, and 0 otherwise. Note Hs=asHs−1.
Returning to the ODE system ˙x=gμ(x), consider the characteristic polynomial of the Jacobian matrix J(x(μ),μ)
pμ(z):=det(zI−J(x(μ),μ))=a0(μ)zs+a1(μ)zs−1+⋯+as(μ), |
where the determinant of a matrix A is denoted by detA. For i=1,…,s, we let Hi(μ) be the i-th Hurwitz matrix of pμ(z).
We now state the criterion we will use to determine whether the characteristic polynomial has a pair of purely imaginary roots.
Proposition 1. In the setup above, let s≥2, μ0 be fixed, and let Hi(μ0) denote the i-th Hurwitz matrix of pμ0(z). We assume that
detH1(μ0)>0, …, detHs−2(μ0)>0. | (4) |
Then pμ0(z) has at most one pair of symmetric roots, and has exactly one if and only if detHs−1(μ0)=0. In this case, the pair consists of purely imaginary roots if and only if as(μ0)>0.
Proof. The first statement follows from Corollary 3.2 in [17] using (4). So, assume pμ0(z) has a pair of symmetric roots z,−z such that detHs−1(μ0)=0. By Lemma 3.3 in [17], the Routh-Hurwitz criterion for stable polynomials, and using (4), we conclude that all the other s−2 roots of pμ0(z) have negative real part. Now as(μ0) is (−1)s times the product of the roots of pμ0(z), and hence its sign agrees with the sign of (−1)s(−1)s−2z(−z)=−z2. We conclude that the pair of symmetric roots of pμ0(z) is real if as(μ0)≤0 and purely imaginary if as(μ0)>0.
As a consequence of Propostion 1, we obtain the following criterion to preclude Hopf bifurcations:
Theorem 1. Consider the dynamical system ˙x=g(x) with s≥2 and assume there exists a curve of steady states x(μ) as in the above setting. As before, let pμ(z) be the characteristic polynomial of the Jacobian J(x(μ),μ) of gμ(x) evaluated at x(μ). Further let Hi(μ) denote the i-th Hurwitz matrix of pμ(z).
If
detH1(μ)>0, …, detHs−2(μ)>0 |
for all μ and either
as(μ)≤0 whenever detHs−1(μ)=0, |
or
detHs−1(μ)≠0 for all μ, |
then the system does not undergo a (simple) Hopf bifurcation.
Proof. Follows from Remark 1 and Proposition 1.
In the following sections we will use Theorem 1 to prove the non-existence of Hopf bifurcations in subnetworks of the mass action network derived from Figure 1a. For a reaction network with n species, if the rank s of the stoichiometric matrix N is not maximal, as it is the case for our networks, then the dynamics takes place in the invariant s-dimensional linear subspaces x0+imN. This implies that 0 is a root of the characteristic polynomial of Nv(κ,x) with multiplicity n−s and hence it factors as
pκ,x(z)=zn−s(a0(κ,x)zs+a1(κ,x)zs−1+⋯+as(κ,x)). |
The polynomial a0(κ,x)zs+a1(κ,x)zs−1+⋯+as(κ,x) is the characteristic polynomial of the Jacobian of the restriction of system (2) to x+imN, and hence we apply Theorem 1 to this polynomial.
By the previous subsection, in order to determine whether a Hopf bifurcation can arise in our systems, we need to analyze the Jacobian matrix of the right-hand side of (2) for all possible values of rate constants κ and positive steady states x∗. Here we reparametrize the Jacobian matrices using so-called convex parameters. These parameters were introduced by Clarke in [20] to analyze the stability of mass action reaction systems (2), in the context of Stoichiometric Network Analysis (SNA) theory.
Let a positive steady state x∗∈Rn>0 of (2) satisfy the polynomial system N diag(κ)ψ(x∗)=0. Then the rate functions v=v(κ,x∗) satisfy the linear problem
Nv=0,v≥0. | (5) |
The vector v is referred to as a flux vector in the SNA theory [21]. The solutions v of (5) define a convex polyhedral cone called the flux cone. Convex polyhedral cones have a finite number of extreme vectors up to a scalar positive multiplication [22]. Therefore, any flux vector v can be represented as a nonnegative linear combination of its extreme vectors {E1,…,El}
v=l∑i=1λiEi=Eλ,allλi≥0, | (6) |
where E is the matrix with columns E1,…,El and λ=(λ1,…,λl).
Remark 2. (cf. [22]). (a) The vectors E1,E2,…,El need not be linearly independent.
(b) If all extreme vectors E1,…,El are unit vectors, then their choice is unique.
(c) When v=v(κ,x∗) and x∗∈Rn>0, then all components of v are positive. This might impose some restrictions on the possible values of λ in (6).
The nonnegative coefficients λ1,…,λl in (6) are often referred to as convex parameters in the literature. However, they do not account alone for all new parameters - the other group of parameters used in SNA theory are reciprocals of each positive steady state coordinate x∗k>0. They are denoted by
hk=1x∗k,k=1,…,n. | (7) |
Definition 2. A vector of convex parameters is a vector of the form (h,λ)=(h1,…,hn,λ1,…,λl)∈Rn>0×Rl≥0 such that Eλ∈Rr>0.
The convex parameters are convenient for parameterizing the Jacobian J(κ,x) evaluated at a positive steady state x=x∗. To see this, note that the (j,i)-th entry of the Jacobian of v(κ,x) evaluated at x∗ is
∂vj(κ,x)∂xi|x=x∗=αijvj(κ,x∗)x∗i=αijvj(κ,x∗)1x∗i. |
Hence, the Jacobian of Nv(κ,x) evaluated at x∗ is
J(κ,x)|x=x∗=Ndiag(v(κ,x∗))YTdiag(1x∗), |
where we use the vector notation
1x=(1x1,…,1xn)T. |
Using now (6) and (7) to write v(κ,x∗)=Eλ, the Jacobian of Nv(κ,x) evaluated at x∗ can be written as
J(κ,x)|x=x∗=J(h,λ)=Ndiag(Eλ)YTdiag(h), | (8) |
with (h,λ) a vector of convex parameters. Therefore, given a vector of rate constants κ and a corresponding positive steady state x∗, there exist convex parameters (h,λ) such that equality (8) holds.
Conversely, given convex parameters (h,λ), we define x∗=1/h and let
κ=diag(ψ(h))Eλ, |
which is a positive vector since all entries of Eλ are positive. Then, using that ψj(x)−1=ψj(x−1), we obtain v(κ,x∗)=diag(κ)ψ(x∗)=Eλ, and equality (8) holds as well. This proves the following proposition:
Proposition 2. The set of Jacobian matrices J(κ,x∗) for all κ and corresponding positive steady states x∗ agrees with the set of matrices defined by the right-hand side of (8), for all h∈Rn>0 and λ∈Rl≥0 such that Eλ∈Rr>0.
The computation of the Jacobian in convex parameters (8) appears in great detail in previous works [23,24]. In [25] it is used to detect saddle-node bifurcations. In Section 5, we use the Jacobian in convex coordinates given in (8) and apply Theorem 1 to conclude that there does not exist a point (h, λ) where a Hopf bifurcation occurs. Using the equality between the two sets of matrices in Proposition 2, this will imply that there do not exist κ and x∗ where a Hopf bifurcation occurs.
Remark 3. The coefficients of the characteristic polynomial of J(κ,x∗) are homogeneous polynomials in the convex parameters if the Jacobian matrix is parametrized as in (8).
Figure 1a contains four phosphorylation events: the phosphorylation of S0 and S1 and the dephosphorylation of S2 and S1. At the level of mass action kinetics each of these phosphorylation events can be described by the following reactions (with i=0,1):
Si+K⇌KSi→Si+1+K and Si+1+F⇌FSi+1→Si+F. |
Consequently, if all phosphorylation events depicted in Figure 1a are described at the mass action level one obtains the following reaction network:
S0+Kκ1⇌κ2KS0κ3→S1+Kκ4⇌κ5KS1κ6→S2+KS2+Fκ7⇌κ8FS2κ9→S1+Fκ10⇌κ11FS1κ12→S0+F. | (N) |
To apply Theorem 1, one has to compute Hurwitz determinants (see Section 3.1 above). These are determinants of matrices that are composed of the coefficients of the characteristic polynomial of the Jacobian of the ODEs defined by 9. In order to show whether or not there exist some values of the rate constants where a Hopf bifurcation occurs, we have to treat all rate constants as fixed but unknown. The coefficients of the characteristic polynomial may contain several hundred terms (cf. the supporting information of [7]). To facilitate the analysis we consider the following simplifications of (N):
(ⅰ) We consider only the forward reaction of the reversible reactions
Si+K⇌KSi and Si+1+F⇌FSi+1. |
This is a reasonable assumption if the rate constants for the reversible reactions are small.
(ⅱ) We consider only two of the four enzyme-substrate complexes KS0, KS1, FS2 and FS1.
There are six ways to choose two complexes out of four. Due to the symmetry of the ODE system obtained by interchanging K and F, S0 and S2, and relabeling the rate constants as appropriate, it suffices to consider the following four simplified networks derived from N:
● The network containing only KS0 and FS2:
K+S0κ1→KS0κ2→K+S1κ3→K+S2F+S2κ4→FS2κ5→F+S1κ6→F+S0. | (N1) |
● The network containing only KS0 and FS1 (mathematically equivalent to the network containing only FS2 and KS1):
K+S0κ1→KS0κ2→K+S1κ3→K+S2F+S2κ4→F+S1κ5→FS1κ6→F+S0. | (N2) |
● The network containing only KS0 and KS1 (mathematically equivalent to the network containing only FS2 and FS1):
K+S0κ1→KS0κ2→K+S1κ3→KS1κ4→K+S2F+S2κ5→F+S1κ6→F+S0. | (N3) |
● The network containing only KS1 and FS1:
K+S0κ1→K+S1κ2→KS1κ3→K+S2F+S2κ4→F+S1κ5→FS1κ6→F+S0. | (N4) |
Remark 4. The dynamics of the network obtained upon removal of an intermediate resembles that of the original network when the binding reaction occurs at a slower time scale than unbinding reactions. Specifically, under this assumption, the simplified system corresponds to the slow system arising from Tikhonov-Fenichel singular perturbation reduction of the original system, which additionally agrees with its quasi-steady-state reduction [26]. Upon removal of a reverse unbinding reaction, the assumption is that this reaction occurs at a much slower time scale than the other reactions in the network, and hence is negligible.
Remark 5. A biochemical interpretation of the simplification leading to the networks (N1) – (N4) can be obtained by comparing it to the well-known Michaelis-Menten approximation: we view our simplification as similar in spirit but with different focus and hence concurrent. To see this recall that to simplify a reaction network based on the Michaelis-Menten approximation, a mass action network of the form
Si+K⇌KSi→Si+1+K | (9) |
is replaced by a network of the form
SivMM→Si+1, |
where vMM is the familiar Michaelis-Menten kinetics
vMM=kcatK0[Si]Km+[Si], |
with catalytic constant kcat, total enzyme concentration K0, Michaelis-Menten constant Km, and where [⋅] denotes the concentration. This is a standard practice of model simplification, for example if the condition formulated by Briggs and Haldane holds [27]; see [28,29] for the mathematical study of this reduction. In [30] it is shown that the dynamical system that arises from the Michaelis-Menten approximation of the full network 9 does not exhibit periodic orbits.
A particular feature of the Michaelis-Menten kinetic vMM is that it is approximately linear in [Si] for small values of [Si] and approximately constant for very large values. Thus the simplification based on the Michaelis-Menten approximation covers, for example, the saturation of available enzyme with substrate very well. However, it does not capture the influence of varying concentrations of the free enzyme K.
In our simplification we replace a mass action network of the form (9) by a mass action network of the form
Si+Kκ→Si+1+K, |
where, according to the law of mass action the reaction rate is
vMA=κ⋅[Si]⋅[K]. |
Clearly, vMA is linear in [Si], hence for small values of Si and values of [K] close to its total concentration K0 the two reaction rates vMA and vMM behave similar. However, vMA is not able to reproduce the saturation of enzyme by the substrate. But it does capture the influence of varying concentrations of free enzyme [K]. Hence for some values of [K] and [Si] our simplification behaves similar to the one based on the Michaelis-Menten approximation, but not everywhere. Moreover, our simplification covers the influence of varying concentrations of K.
In summary, the simplification based on the Michaelis-Menten approximation on the one hand is well suited to describe the saturation of the enzyme with substrate but does not account for the influence of varying concentrations of free enzyme. Our simplification on the other hand cannot account for enzyme saturation, but for the influence of varying concentrations of free enzyme. Moreover, for small values of the substrate concentration and for concentrations of free enzyme close to its total amount both simplifications behave similar. Hence we view our simplification as biochemically concurrent to the one based on the Michaelis-Menten approximation. Both simplifications fail to accommodate the complete behavior of the distributive and sequential double phosphorylation cycle of Figure 1a. But both cover different but equally important aspects of its behavior and hence are well worth studying.
Remark 6. If any of the networks (N1) – (N4) had a Hopf bifurcation giving rise to periodic solutions, then by [31], so would the full mechanism in 9. In particular, by [31] the same is true if any of the irreversible reactions is made reversible and in particular, if unbinding reactions are considered in the formation the complexes KS0, KS1, FS1 or FS2.
In this section we apply Theorem 1, using the discussion after it, to the networks (N1) – (N4). To this end we use Eq (8) to determine the Jacobian matrices J1(h,λ), …, J4(h,λ) of networks (N1) – (N4), correspondingly. The computations are done symbolically and can be found in the supporting file 'SupplMat1.mw'.
We first comment on some common features of the networks:
Remark 7. (ⅰ) Every network (N1) – (N4) consists of 7 species and 6 reactions.
(ⅱ) The stoichiometric matrix of every network has rank s=4, as every network has three conserved quantities (the total amount of substrate, kinase and phosphatase).
(ⅲ) For every network the cone (5) is spanned by two nonnegative vectors w0 and w1 such that λ1w0+λ2w1 is a positive vector if and only if λ1,λ2>0.
Observe that a scaling αλ of the vector λ=(λ1,λ2) with α>0 translates into a scaling ακ of the vector of rate constants κ under the correspondence of parameterization in Proposition 2. Further x∗ is a steady state for κ if and only if it is for ακ and αJ(κ,x∗)=J(ακ,x∗). The latter implies that any eigenvalue of J(ακ,x∗) is in fact an eigenvalue of J(κ,x∗) multiplied by α>0. Thus, J(ακ,x∗) has a pair of purely imaginary eigenvalues if and only if J(κ,x∗) has a pair of purely imaginary eigenvalues.
Hence, it is enough to take one of λ1,λ2 to be one (since both are positive), and we let the elements in the kernel of the stoichiometric matrices Ni be of the form
w0+λw1,λ>0. |
Therefore, we consider every Jacobian matrix Ji(h,λ) according to Eq (8) parametrized by 8 parameters: the parameter λ and h1,…,h7.
By Remark 7 (ⅰ) and (ⅱ) it follows that the characteristic polynomial of every Jacobian Ji(h,λ) is a degree 7 polynomial of the form
z3(a0(h,λ)z4+…+a3(h,λ)z+a4(h,λ)), |
where each ai depends on the 8 parameters λ,h1,…,h7 (cf. the file 'SupplMat1.mw'). Following the discussion after Theorem 1, for each network we compute a0(h,λ),…,a4(h,λ), detH1(h,λ), detH2(h,λ) and detH3(h,λ) (cf. Definition 1) and show the following proposition:
Proposition 3. With the notation above, we have
(i) Network (N1): detH1(h,λ) and detH2(h,λ) contain only positive monomials, a4(h,λ) and detH3(h,λ) contain monomials of both signs. But detH3(h,λ) is positive whenever a4(h,λ)>0 (Proposition 4 in Section 5.1).
(ii) Network (N2) and (N3): detH1(h,λ), detH2(h,λ) and detH3(h,λ) contain only positive monomials, a4(h,λ) contains monomials of both signs. Thus detHi(h,λ)>0, i=1,2,3 for positive h and λ and in particular, detH3(h,λ)≠0.
(iii) Network (N4): detH1(h,λ), detH2(h,λ) and detH3(h,λ) and a4(h,λ) contain only positive monomials. As in case of N2 and N3, one has detHi(h,λ)>0, i=1,2,3 for positive h and λ and in particular, detH3(h,λ)≠0.
In particular, this proposition (which is proven below) tells us that in all four networks, detH3(h,λ)≠0 whenever a4(h,λ)>0. As a consequence we obtain the following theorem.
Theorem 2. For the networks (N1) – (N4) there do not exist rate constants κ and a corresponding positive steady state x∗ such that the Jacobian matrix Ji(κ,x∗) has a pair of purely imaginary eigenvalues. Thus, in particular, there do not exist κ and x∗ where a Hopf bifurcation occurs.
Proof. By Proposition 3 and the correspondence between the two parameterization of the Jacobian given in Proposition 2, there do not exist κ and a corresponding positive steady state x∗ such that the corresponding Hurwitz determinant detH3(κ,x∗) vanishes and the coefficient of lowest degree of the characteristic polynomial a4(κ,x∗) is positive. Further, for all networks detH1(κ,x∗)>0 and detH2(κ,x∗)>0 for all κ,x∗. Hence, by Theorem 3.1, the Jacobian matrix does not have a pair of purely imaginary eigenvalues.
Remark 8. By [33], the networks N1, N2 and N3 are multistationary, while N4 is not. Hence, by [32], the coefficient of lowest degree of the Jacobian must vanish for some κ and positive steady state x∗, and consistently a4(h,λ) must contain monomials of both signs.
Remark 9. We have focused on networks with two intermediates, as network ((N1)) provides the first non-trivial case and might shed light on how to approach the full network. All simplifications of the full network (N) obtained after removing three intermediates, that is, with only one intermediate left, satisfy that detHs−1(h,λ) is a sum of positive terms and hence does not vanish. Consequently, Hopf bifurcations do not arise.
All that remains is to show Proposition 3. This is done through mathematical reasoning aided by symbolic computations performed in Maple and Mathematica. In the following subsections we present for each network the stoichiometric matrix Ni, the kinetic order matrix Yi, the matrix Ei whose columns are vectors w0 and w1 that generate the cone (5) and the Jacobian matrix Ji(h,λ). For the networks ((N1)) – ((N4)) we found these vectors simply by finding a basis of the kernel of N, which has dimension 2, and deriving extreme vectors. For larger networks software like CellNetAnalyzer [34] can be used. Where appropriate we then discuss the coefficient a4(h,λ) and the determinants of the Hurwitz matrices detHi(h,λ). The computations are given in the supplementary file 'SupplMat1.mw'.
We denote by x1,x2,x3,x4,x5,x6,x7 the concentrations of K, F, S0, S1, S2, KS0, FS2 respectively. Then the stoichiometric matrix, the kinetic order matrix and the matrix E1 are
N1=[−110000000−110−10000101−101−1001−1001−100000001−10],Y1=[101000000101100000001001000100010000000010] and E1=[101001010110]. |
Using convex parameters, the Jacobian matrix in terms h1,…,h7,λ as given in (8) is
J1(h,λ)=[−h10−h300h600−h2λ00−h5λ0h7λ−h1h2−h3h4000−h1λ−h20h4(−1−λ)0h6h7λh1λ−h2λ0h4λ−h5λ00h10h300−h600h2λ00h5λ0−h7λ]. |
We observe that the coefficient a4(h,λ) of the characteristic polynomial contains monomials of both signs. We compute the associated Hurwitz determinants detH1(h,λ), detH2(h,λ) and detH3(h,λ) and obtain that detH1(h,λ) and detH2(h,λ) are sums of positive monomials and that detH3(h,λ) contains monomials of both signs as well. Hence both a4(h,λ) and detH3(h,λ) contain monomials of both signs and can potentially be zero.
In the remainder of this section we prove the following:
Proposition 4. Consider the coefficient a4(h,λ) and the Hurwitz determinant detH3(h,λ) of network (N1) and given in the file 'SupplMat1.mw'. Then
ifa4(h,λ)>0,thendetH3(h,λ)>0. |
Proof. The coefficient a4(k,λ) of the Jacobian is
−λ2(h1h2h3h4+h1h2h3h5+h1h2h4h5+h1h3h4h5+h1h3h4h7−h1h4h5h7+h2h3h4h5−h2h3h4h6+h2h4h5h6−h3h4h5h6−h3h4h5h7−h3h4h6h7−h3h5h6h7−h4h5h6h7). |
Since λ2 factors out and it does not affect the sign, we consider
c0:=−h1h2h3h4−h1h2h3h5−h1h2h4h5−h1h3h4h5−h1h3h4h7+h1h4h5h7−h2h3h4h5+h2h3h4h6−h2h4h5h6+h3h4h5h6+h3h4h5h7+h3h4h6h7+h3h5h6h7+h4h5h6h7. |
We show that c0>0 implies detH3>0, omitting the argument of H3. The computations are performed in Maple, but we explain here the computational procedure for the proof.
We start by noting that c0 can be written as:
b0=(h6−h1)(h2+h5+h7)h3h4+(h7−h2)(h1+h3+h6)h4h5+(h6h7−h1h2)h3h5. |
We see immediately that if h1>h6 and h2>h7, then c0<0. We do not need to study this case.
We consider the case h6≥h1 and h7≥h2, such that one of the two inequalities is strict, otherwise c0=0. In this case c0≥0. We introduce new nonnegative parameters v1,v2 and substitute h6=h1+v1 and h7=h2+v2. This encodes the inequalities. We perform this substitution into detH3 using Maple, expand the new polynomial, and check the sign of the coefficients. All coefficients in detH3 are positive, meaning that detH3 will be positive in this case. This holds if v1,v2>0 or if one of the two parameters is set equal to zero.
The latter means that we need to study the case h1≥h6 and h2≤h7. We now perform the substitution h1=h6+v1 and h7=h2+v2, again with one of v1 or v2 nonzero. We observe that if h5≥h3, then again detH3 is positive. So we restrict to h3>h5 and perform the substitution h3=h5+v3 into detH3.
When we do that, detH3 has coefficients of both signs, and therefore the sign is not clear. We have still to impose c0>0 for this scenario. We perform the substitutions into c0 and obtain:
c0=(−2h2h4h5−2h2h4v3−h2h25−h5h2v3−h4h25−h4h5v3−h4v2v3)v1+h4h25v2+2h4h5h6v2+h4h5v2v3+h25h6v2+h5h6v2v3 |
For c0 to be positive, we need v1 to be smaller than the root of c0 seen as a polynomial in v1, which is:
z1:=h5v2(h4h5+2h4h6+h4v3+h6h5+h6v3)2h2h4h5+2h2h4v3+h2h25+h2h5v3+h4h25+h4h5v3+h4v2v3. |
Now, we need to check whether detH3 can be negative when v1 is smaller than z1. To check that, we make the substitution
v1=μμ+1z1. |
Then any number in the interval [0,z1) is of this form for some nonnegative μ. We perform this substitution in Maple, gather the numerator of the resulting detH3, and confirm that all signs are positive. Further detH3 is positive even if some of μ,v2,v3 are zero.
The other case h2≥h7 and h1≤h6 is analogous by the symmetry of the system. This finishes the argument, since we have explored all possibilities for c0>0, and they all give that detH3 is positive.
We use the following ordering: x1,x2,x3,x4,x5,x6,x7 for the concentration of K,F,S0,S1,S2,KS0,FS1 respectively. Under this ordering the stoichiometric matrix, the kinetic order matrix and the matrix E2 are
N2=[−1100000000−11−10000101−11−10001−1001−1000000001−1],Y2=[101000000110100000001010000100010000000001] and E2=[101001011010]. |
With this parametrization, the Jacobian of the system evaluated at a steady state defined by (h1,…,h7,λ) is:
J2(h,λ)=[−h10−h300h600−h20−h400h7−h10−h3000h7−h1λh2(−1+λ)0h4(−1−λ)h5λh60h1λ−h2λ0h4λ−h5λ00h10h300−h600h20h400−h7]. |
detH3(h,λ) contains only positive monomials and in particular it does not vanish for any positive h,λ.
We use the following ordering: x1,x2,x3,x4,x5,x6,x7 for the concentration of K,F,S0,S1,S2,KS0,KS1 respectively. Under this ordering the stoichiometric matrix, the kinetic order matrix and the matrix E3 are
N3=[−11−1100000000−10000101−101−10001−101−10000001−100],Y3=[101000000011100000001001000010010000000100] and E3=[101001010110]. |
With this parametrization, the Jacobian of the system evaluated at a steady state defined by (h1,…,h7,λ) is:
J3(h,λ)=[h1(−1−λ)0−h3−h4λ0h6h7λ0000000−h1h2−h3h4000−h1λh2(−1+λ)0h4(−1−λ)h5λh600−h2λ00−h5λ0h7λh10h300−h60h1λ00h4λ00−h7λ]. |
detH3(h,λ) contains only positive monomials and in particular it does not vanish for any positive h,λ.
We use the following ordering: x1,x2,x3,x4,x5,x6,x7 for the concentration of K,F,S0,S1,S2,KS1,FS1 respectively. Under this ordering the stoichiometric matrix, the kinetic order matrix and the matrix E4 are
N4=[0−110000000−11−1000011−101−10001−10001−100000001−1],Y4=[110000000110100000010010000100001000000001] and E4=[100101011010]. |
With this parametrization, the Jacobian of the system evaluated at a steady state defined by (h1,…,h7,λ) is:
J4(h,λ)=[−h1λ00−h4λ0h6λ00−h20−h400h7−h10−h3000h7h1(1−λ)h2(−1+λ)h3h4(−1−λ)h5λ000−h2λ00−h5λh6λ0h1λ00h4λ0−h6λ00h20h400−h7]. |
detH3(h,λ) contains only positive monomials and in particular it does not vanish for any positive h,λ.
As discussed in the Introduction, in [16,Section 5.36] a simplification of the mass action model of Figure 1a is examined and the authors provide steady state concentration values and rate constants of a candidate Hopf bifurcation point. Here we want to explain how this point fails to give a pair of purely imaginary eigenvalues (Proposition 1).
In [16], the authors consider the following irreversible version of the network (N):
S0+Kκ1→KS0κ3→S1+Kκ4→KS1κ6→S2+KS2+Fκ7→FS2κ9→S1+Fκ10→FS1κ12→S0+F. | (Nf) |
In network Nf the constants κ2, κ5, κ8 and κ11 describing the 'backward' reactions in the full network (N) have been assigned the value zero. Hence the matrix Ef whose columns generate the cone (5) is
ETf=[101000000101000111111000]. |
Using x1,x2,x3,x4,x5,x6,x7,x8,x9 for the concentrations of S0,K,KS0,S1,KS1,S2,F,FS2,FS1 respectively one obtains the Jacobian matrix
Jf(h,λ)=[−h1λ1−h2λ1000000h9λ1−h1λ1h2(−λ1−λ2)h3λ1−h4λ2h5λ20000h1λ1h2λ1−h3λ10000000−h2λ2h3λ1h4(−λ1−λ2)00−h7λ1h8λ200h2λ20h4λ2−h5λ200000000h5λ2−h6λ2−h7λ200000−h4λ10−h6λ2h7(−λ1−λ2)h8λ2h9λ100000h6λ2h7λ2−h8λ20000h4λ100h7λ10−h9λ1]. |
The authors of [16] provide the candidate point where detH5(h∗,λ∗)=0:
h∗1=9.15394021721585⋅10−6h∗2=8.438690345203897⋅10−6h∗3=9.15394021721585⋅10−6h∗4=9.15394021721585⋅10−6h∗5=0.0000234589h∗6=0.00391442h∗7=0.0625077h∗8=0.00391442h∗9=1λ∗1=1λ∗2=1. |
We find that
detH1(h∗,λ∗)=−1.13292detH2(h∗,λ∗)=0.0803339detH3(h∗,λ∗)=−1.7440236556291417⋅10−6detH4(h∗,λ∗)=2.5421805536611004⋅10−15detH5(h∗,λ∗)=−1.3900880766102185⋅10−42a6(h∗,λ∗)=−1.682281311658486⋅10−18. |
While detH4(h∗,λ∗), detH5(h∗,λ∗) and a6(h∗,λ∗) are all very small numbers, detH5(h∗,λ∗) is by orders of magnitude smaller. And if 10−42≈0, then detH4(h∗,λ∗) and a6(h∗,λ∗) are nonzero. In particular a6(h∗,λ∗)<0 and hence this point gives rise to a pair of symmetric real eigenvalues. We have additionally verified our claim by considering the set of corresponding eigenvalues (computed with Mathematica):
{−1.06643,−0.0661828,−0.000208571,−0.0000990197,0.0000339721,−0.0000339721,−4.701752723492088⋅10−16,1.2668321503904372⋅10−17,1.495648385287835⋅10−18}. |
All eigenvalues are real numbers, three eigenvalues are ≈0 (as expected as the system has three conservation relations) and the real values 0.0000339721 and −0.0000339721 give a pair of symmetric eigenvalues (up to 10 digits).
Hence, this justifies our claim that the question of whether or not a Hopf bifurcation can occur in the full network (N) is still open. For the full network, detH1,…,detHs−2 are all positive, and hence the problem is reduced to making detHs−1 vanish while having as positive. Network (N1) is the smallest simplified network where both detHs−1 and as attain both signs and hence poses the same challenge.
For networks admitting pairs of purely imaginary eigenvalues, a typical approach in this situation is to find values for which detHs−1 vanishes, and then verify that as is positive when evaluated at these values. But if the two conditions happen to be incompatible, as it might be the case for the full network (N), then it is unclear how to proceed. Automated approaches [16] cannot handle the analysis of the signs of these two large polynomials at the same time. Here, by combining guided heavy symbolic computations with manual intervention based on the visual inspection of as, we have managed to prove that the two sign conditions are incompatible for network (N1).
The idea is to reduce the problem into checking the sign of the coefficients of a polynomial. We first break the condition as>0 into subcases involving simple inequalities between the parameters. Then, one uses the inequalities to substitute one parameter in Hs−1 by a new parameter that is allowed to take any positive value (e.g. k1>k2 leads to the substitution k1=k2+a, such that the new constraints are k2,a>0). Finally, Hs−1 is transformed into a polynomial (or rational function) that only needs to be evaluated at positive values to guarantee as>0. If all coefficients of Hs−1 are positive, then trivially the polynomial is positive. If some coefficients are negative, then further exploration is required.
This strategy has worked nicely for network (N1). However, substitutions of the type k1=k2+a typically increase dramatically the number of terms as a polynomial of the form kd1p(κ,x) becomes (k2+a)dp(κ,x). We also performed substitutions of the form k1=μμ+1p(κ,x)q(κ,x) with μ>0, which introduced large denominators.
All this illustrates the difficulty in the analysis of the full network: the computation of the Hurwitz determinants is very demanding, but a posterior analysis, in par with the analysis of network (N1), is prohibitive. Nevertheless, we think ideas introduced here in the analysis of network (N1) might be applied to other networks posing similar challenges, and maybe similar ideas end up being successful for the full network either after increasing computational power or by depicting new strategies to simplify computations.
As explained in Remark 6, the existence of periodic orbits in simplified models can be lifted to more complex models under certain network modifications [31]. However, the non-existence of periodic orbits or Hopf bifurcations does not, in principle, tell us anything about complex models including the simplified models in some form. Considering the demanding computational cost involved in describing the dynamics of a reaction network for arbitrary parameter values, it would be of great help to have a better understanding of what type of network operations or parameter regimes guarantee that the non-existence of a behavior in a simplified model is preserved in the more complex model. For example, we are not aware of any result nor reasoning we could employ to ensure that our conclusions on networks (N1)-(N4) are preserved after introducing unbinding reactions with a small rate constant.
CC acknowledges funding from Deutsche Forschungsgemeinschaft, 284057449. EF acknowledges funding from the Danish Research Council for Independent Research.
The authors thank Angélica Torres for providing source code to compute Hurwitz determinants and Alan Rendall for helpful discussions about (simple) Hopf bifurcations. The authors thank the hospitality Erwin Schrödinger International Institute for Mathematics and Physics in Vienna, where this project was started during the workshop 'Advances in Chemical Reaction Network Theory'.
The authors declare there is no conflicts of interest.
The Maple computations for the proof of Theorem 5.1 are provided in the file 'SupplMat1.mw'. For the convenience of the readers without access to Maple, we provide as well as pdf version of the file.
[1] | C. Conradi and A. Shiu, Dynamics of post-translational modification systems: recent progress and future directions, Biophys. J., 114 (2018), 507-515. |
[2] | T. Suwanmajo and J. Krishnan, Exploring the intrinsic behaviour of multisite phosphorylation systems as part of signalling pathways, J. R. Soc. Interface, 15 (2018), 20180109. |
[3] | J. Gunawardena, Multisite protein phosphorylation makes a good threshold but can be a poor switch, Proc. Natl. Acad. Sci. U.S.A., 102 (2005), 14617-14622. |
[4] | C. Salazar and T. Hofer, Multisite protein phosphorylation-from molecular mechanisms to kinetic models, FEBS J., 276 (2009), 3177-3198. |
[5] | M. Thomson and J. Gunawardena, Unlimited multistability in multisite phosphorylation systems, Nature, 460 (2009), 274-277. |
[6] | C. Conradi, D. Flockerzi and J. Raisch, Multistationarity in the activation of an MAPK: parametrizing the relevant region in parameter space, Math. Biosci., 211 (2008), 105-131. |
[7] | C. Conradi and M. Mincheva, Catalytic constants enable the emergence of bistability in dual phosphorylation, J. R. S. Interface, 11. |
[8] | J. Hell and A. D. Rendall, A proof of bistability for the dual futile cycle, Nonlinear Anal. Real World Appl., 24 (2015), 175-189. |
[9] | D. Flockerzi, K. Holstein and C. Conradi, N-site Phosphorylation Systems with 2N-1 Steady States, Bull. Math. Biol., 76 (2014), 1892-1916. |
[10] | L. Wang and E. D. Sontag, On the number of steady states in a multiple futile cycle, J. Math. Biol., 57 (2008), 29-52. |
[11] | C. Conradi and A. Shiu, A global convergence result for processive multisite phosphorylation systems, Bull. Math. Biol., 77 (2015), 126-155. |
[12] | C. Conradi, M. Mincheva and A. Shiu, Emergence of oscillations in a mixed-mechanism phosphorylation system, Bull. Math. Biol., 81 (2019), 1829-1852. |
[13] | T. Suwanmajo and J. Krishnan, Mixed mechanisms of multi-site phosphorylation, J. R. Soc. Interface, 12 (2015), 20141405. |
[14] | N. Obatake, A. Shiu, X. Tang and A. Torres, Oscillations and bistability in a model of ERK regulation, J. Math. Biol., (2019), https://doi.org/10.1007/s00285-019-01402-y. |
[15] | Y. A. Kuznetsov, Elements of Applied Bifurcation Theory, Springer-Verlag, 1995. |
[16] | H. Errami, M. Eiswirth, D. Grigoriev, et al., Detection of hopf bifurcations in chemical reaction networks using convex coordinates, J. Comput. Phys., 291 (2015), 279-302. |
[17] | M. El Kahoui and A. Weber, Deciding Hopf bifurcations by quantifier elimination in a softwarecomponent architecture, J. Symbolic Comput., 30 (2000), 161-179. |
[18] | W. M. Liu, Criterion of hopf bifurcations without using eigenvalues, Journal of Mathematical Analysis and Applications, 182 (1994), 250-256. |
[19] | X. Yang, Generalized form of Hurwitz-Routh criterion and Hopf bifurcation of higher order, Appl. Math. Lett., 15 (2002), 615-621. |
[20] | B. L. Clarke, Stability of Complex Reaction Networks, 1-215, John Wiley & Sons, Ltd, 2007. Available from: https://onlinelibrary.wiley.com/doi/abs/10.1002/9780470142622.ch1. |
[21] | B. L. Clarke, Stoichiometric network analysis, Cell Biochem. Biophys., 12 (1988), 237-253. |
[22] | R. T. Rockafellar, Convex Analysis, Princeton University Press, 1970. |
[23] | B. D. Aguda and B. L. Clarke, Bistability in chemical reaction networks: Theory and application to the peroxidase-oxidase reaction, J. Chem. Phys., 87 (1987), 3461-3470. |
[24] | K. Gatermann, M. Eiswirth and A. Sensse, Toric ideals and graph theory to analyze hopf bifurcations in mass action systems, J. Symbol. Comput., 40 (2005), 1361-1382. |
[25] | C. Conradi and D. Flockerzi, Switching in mass action networks based on linear inequalities, SIAM J. Appl. Dyn. Syst., 11 (2012), 110-134. |
[26] | E. Feliu, C. Lax, S. Walcher, et al., Quasi-steady state and singular perturbation reduction for reaction networks with non-interacting species, arXiv, 1908.11270. |
[27] | A. Cornish-Bowden, Fundamentals of Enzyme Kinetics, 3rd edition, Portland Press, London, 2004. |
[28] | A. Goeke, S. Walcher and E. Zerz, Determining "small parameters" for quasi-steady state, J. Differ. Equations, 259 (2015), 1149-1180. |
[29] | A. Goeke, S. Walcher and E. Zerz, Classical quasi-steady state reduction-a mathematical characterization, Phys. D, 345 (2017), 11-26. |
[30] | H.-R. Tung, Precluding oscillations in Michaelis-Menten approximations of dual-site phosphorylation systems, Math. Biosci., 306 (2018), 56-59. |
[31] | M. Banaji, Inheritance of oscillation in chemical reaction networks, Appl. Math. Comp., 325 (2018), 191-209. |
[32] | C. Conradi, E. Feliu, M. Mincheva, et al., Identifying parameter regions for multistationarity, PLoS Comput. Biol., 13 (2017), e1005751. |
[33] | A. Sadeghimanesh and E. Feliu, The multistationarity structure of networks with intermediates and a binomial core network, Bull. Math. Biol., 81 (2019), 2428-2462. |
[34] | A. von Kamp, S. Thiele, O. Hädicke, et al., Use of CellNetAnalyzer in biotechnology and metabolic engineering, J. Biotechnol., 261 (2017), 221-228. |
1. | Elisenda Feliu, Nidhi Kaihnsa, Timo de Wolff, Oğuzhan Yürük, The Kinetic Space of Multistationarity in Dual Phosphorylation, 2020, 1040-7294, 10.1007/s10884-020-09889-6 | |
2. | Carsten Conradi, Nida Obatake, Anne Shiu, Xiaoxian Tang, Dynamics of ERK regulation in the processive limit, 2021, 82, 0303-6812, 10.1007/s00285-021-01574-6 | |
3. | Thapanar Suwanmajo, Vaidhiswaran Ramesh, J. Krishnan, Exploring cyclic networks of multisite modification reveals origins of information processing characteristics, 2020, 10, 2045-2322, 10.1038/s41598-020-73045-9 | |
4. | Beatriz Pascual‐Escudero, Elisenda Feliu, Local and global robustness at steady state, 2022, 45, 0170-4214, 359, 10.1002/mma.7780 | |
5. | Máté László Telek, Elisenda Feliu, Jason M. Haugh, Topological descriptors of the parameter region of multistationarity: Deciding upon connectivity, 2023, 19, 1553-7358, e1010970, 10.1371/journal.pcbi.1010970 | |
6. | Carsten Conradi, Maya Mincheva, In distributive phosphorylation catalytic constants enable non-trivial dynamics, 2024, 89, 0303-6812, 10.1007/s00285-024-02114-8 | |
7. | Nicola Vassena, Symbolic hunt of instabilities and bifurcations in reaction networks, 2023, 0, 1531-3492, 0, 10.3934/dcdsb.2023190 | |
8. | Nidhi Kaihnsa, Máté L. Telek, Connectivity of Parameter Regions of Multistationarity for Multisite Phosphorylation Networks, 2024, 86, 0092-8240, 10.1007/s11538-024-01368-z | |
9. | Xiaoxian Tang, Kaizhang Wang, Hopf Bifurcations of Reaction Networks with Zero-One Stoichiometric Coefficients, 2023, 22, 1536-0040, 2459, 10.1137/22M1519754 |