Skip to main content
Log in

A Structural Classification of Candidate Oscillatory and Multistationary Biochemical Systems

  • Original Article
  • Published:
Bulletin of Mathematical Biology Aims and scope Submit manuscript

Abstract

Molecular systems are uncertain: The variability of reaction parameters and the presence of unknown interactions can weaken the predictive capacity of solid mathematical models. However, strong conclusions on the admissible dynamic behaviors of a model can often be achieved without detailed knowledge of its specific parameters. In systems with a sign-definite Jacobian, for instance, cycle-based criteria related to the famous Thomas’ conjectures have been largely used to characterize oscillatory and multistationary dynamic outcomes. We build on the rich literature focused on the identification of potential oscillatory and multistationary behaviors using parameter-free criteria. We propose a classification for sign-definite non-autocatalytic biochemical networks, which summarizes several existing results in the literature. We call weak (strong) candidate oscillators systems which can possibly (exclusively) transition to instability due to the presence of a complex pair of eigenvalues, while we call weak (strong) candidate multistationary systems those which can possibly (exclusively) transition to instability due to the presence of a real eigenvalue. For each category, we provide a characterization based on the exclusive or simultaneous presence of positive and negative cycles in the associated sign graph. Most realistic examples of biochemical networks fall in the gray area of systems in which both positive and negative cycles are present: Therefore, both oscillatory and bistable behaviors are in principle possible. However, many canonical example circuits exhibiting oscillations or bistability fall in the categories of strong candidate oscillators/multistationary systems, in agreement with our results.

This is a preview of subscription content, log in via an institution to check access.

Access this article

Price excludes VAT (USA)
Tax calculation will be finalised during checkout.

Instant access to the full article PDF.

Fig. 1
Fig. 2
Fig. 3
Fig. 4
Fig. 5

Similar content being viewed by others

Notes

  1. We consider only the sign of each Jacobian entry as the weight of the corresponding arc in the graph.

  2. This definition holds for systems transitioning to instability from the right to the left neighborhood of \(\mu ^*\): In the above definition, it suffices to take \(\hat{\mu }= \mu ^* -\mu \) as the bifurcation parameter.

  3. A system is monotone with respect to some orthant cone if and only if its Jacobian is a Metzler matrix (namely its non-diagonal entries are nonnegative), or becomes a Metzler matrix after a proper change in the sign of some variables (Smith et al. 2008; Sontag 2007).

  4. Exploiting the Jacobian irreducibility assumption, it is actually possible to prove that at least two additional equilibria arise.

References

  • Alon U (2006) An introduction to systems biology: design principles of biological circuits. Chapman & Hall/CRC, Boca Raton

    MATH  Google Scholar 

  • Ambrosetti A, Prodi G (1995) A primer of nonlinear analysis. Cambridge studies in advanced mathematics, vol 34. Cambridge University Press, Cambridge

  • Angeli D, Sontag E (2003) Monotone control systems. IEEE Trans Autom Control 48(10):1684–1698

    Article  MathSciNet  Google Scholar 

  • Angeli D, Sontag E (2008) Oscillations in I/O monotone systems. IEEE Trans Circuits Syst 55:166–176

    Article  Google Scholar 

  • Angeli D, Ferrell JE, Sontag ED (2004) Detection of multistability, bifurcations, and hysteresis in a large class of biological positive-feedback systems. Proc Nat Acad Sci USA 101(7):1822–1827. doi:10.1073/pnas.0308265100

    Article  Google Scholar 

  • Atkinson MR, Savageau M, Myers J, Ninfa A (2003) Development of genetic circuitry exhibiting toggle switch or oscillatory behavior in Escherichia coli. Cell 113:597–607

    Article  Google Scholar 

  • Banaji M, Craciun G (2009) Graph-theoretic approaches to injectivity and multiple equilibria in systems of interacting elements. Commun Math Sci 7(4):867–900

    Article  MathSciNet  MATH  Google Scholar 

  • Blanchini F, Franco E (2011) Structurally robust biological networks. Bio Med Cent Syst Biol 5(1):74

    Google Scholar 

  • Blanchini F, Miani S (2008) Set-theoretic methods in control. Systems and control: foundations and Applications. Birkhäuser, Boston

    MATH  Google Scholar 

  • Craciun G, Feinberg M (2006a) Multiple equilibria in complex chemical reaction networks: I. The injectivity property. SIAM J Appl Math 65(5):1526–1546

  • Craciun G, Feinberg M (2006b) Multiple equilibria in complex chemical reaction networks: II. The species-reaction graph. SIAM J Appl Math 66(4):1321–1338

  • Craciun G, Tang Y, Feinberg M (2006) Understanding bistability in complex enzyme-driven reaction networks. Proc Natl Acad Sci 103(23):8697–8702

    Article  MATH  Google Scholar 

  • Davidson EH, Rast JP, Oliveri P, Ransick A, Calestani C et al (2002) A genomic regulatory network for development. Science 295(5560):1669–1678

    Article  Google Scholar 

  • Domijan M, Kirkilionis M (2009) Bistability and oscillations in chemical reaction networks. J Math Biol 59(4):467–501

  • Domijan M, Pécou E (2011) The interaction graph structure of mass-action reaction networks. J Math Biol 51(8):1–28

    MATH  Google Scholar 

  • Edelstein-Keshet L (2005) Mathematical models in biology. Classics in applied mathematics, vol 46. SIAM, Philadelphia

  • Elowitz MB, Leibler S (2000) A synthetic oscillatory network of transcriptional regulators. Nature 403(6767):335–338

    Article  Google Scholar 

  • Feinberg M (1987) Chemical reaction network structure and the stability of complex isothermal reactors I. The deficiency zero and deficiency one theorems. Chem Eng Sci 42:2229–2268

    Article  Google Scholar 

  • Franco E, Blanchini F (2012) Structural properties of the MAPK pathway topologies in PC12 cells. J Math Biol 67(6–7):1633–1668

  • Franco E, Friedrichs E, Kim J, Jungmann R, Murray R, Winfree E, Simmel FC (2011) Timing molecular motion and production with a synthetic transcriptional clock. Proc Natl Acad Sci 108(40):E784–E793

    Article  Google Scholar 

  • Gardner TS, Cantor CR, Collins JJ (2000) Construction of a genetic toggle switch in Escherichia coli. Nature 403(6767):339–342

    Article  Google Scholar 

  • Goldbeter A, Gérard C, Gonze D, Leloup JC, Dupont G (2012) Systems biology of cellular rhythms. FEBS Lett 586:2955–2965

    Article  Google Scholar 

  • Golnaraghi F, Kuo BC (2009) Automatic control systems, 9th edn. Wiley, New York

    Google Scholar 

  • Gouze JL (1998) Positive and negative circuits in dynamical systems. J Biol Syst 6:11–15

    Article  MATH  Google Scholar 

  • Haddad WM, Chellaboina V (2004) Stability theory for nonnegative and compartmental dynamical systems with time delay. Syst Control Lett 51(5):355–361

    Article  MathSciNet  MATH  Google Scholar 

  • Hirsch MW (1988) Stability and convergence in strongly monotone dynamical systems. J Reine Angew Math 383:1–53

    MathSciNet  MATH  Google Scholar 

  • Hofbauer J (1990) An index theorem for dissipative semiflows. Rocky Mt J Math 20(4):1017–1031

    Article  MathSciNet  MATH  Google Scholar 

  • Hori Y, Hara S (2011) Time delay effects on oscillation profiles in cyclic gene regulatory networks: harmonic balance approach. In: American Control Conference (ACC) 2011, pp 2891–2896

  • Jiang JF (1991) On the global stability of cooperative systems. Bull Lond Math Soc 26:455–458

    MathSciNet  MATH  Google Scholar 

  • Khalil HK (2002) Nonlinear systems. Prentice Hall, Englewood Cliffs

    MATH  Google Scholar 

  • Kaufman M, Soule C, Thomas R (2007) A new necessary condition on interaction graphs for multistationarity. J Theor Biol 248(4):675–685

    Article  MathSciNet  Google Scholar 

  • Kim J, Winfree E (2011) Synthetic in vitro transcriptional oscillators. Mol Syst Biol 7:465

    Article  Google Scholar 

  • Kim J, White KS, Winfree E (2006) Construction of an in vitro bistable circuit from synthetic transcriptional switches. Mol Syst Biol 1:68

    Google Scholar 

  • Leite MCA, Wang Y (2010) Multistability, oscillations and bifurcations in feedback loops. Math Biosci Eng 7(1):83–97

  • Lewis J (2003) Autoinhibition with transcriptional delay: a simple mechanism for the zebrafish somitogenesis oscillator. Curr Biol 13(16):1398–1408

    Article  Google Scholar 

  • Liu X, Yu W, Wang L (2010) Stability analysis for continuous-time positive systems with time-varying delays. IEEE Trans Autom Control 55(4):1024–1028

    Article  MathSciNet  Google Scholar 

  • Maybee J, Quirk J (1969) Qualitative problems in matrix theory. SIAM Rev 11:30–51

    Article  MathSciNet  MATH  Google Scholar 

  • Meiss JD (2007) Differential dynamical systems. Mathematical modeling and computation, vol 14. SIAM, Philadelphia

  • Mincheva M (2011) Oscillations in biochemical reaction networks arising from pairs of subnetworks. Bull Math Biol 73:2277–2304

    Article  MathSciNet  Google Scholar 

  • Mincheva M, Craciun G (2008) Multigraph conditions for multistability, oscillations and pattern formation in biochemical reaction networks. Proc IEEE 96(8):1281–1291

    Article  Google Scholar 

  • Montagne K, Plasson R, Sakai Y, Fujii T, Rondelez Y (2011) Programming an in vitro DNA oscillator using a molecular networking strategy. Mol Syst Biol 7:466

  • Novak B, Tyson JJ (2008) Design principles of biochemical oscillators. Nat Rev Mol Cell Biol 9(12):981–991

    Article  Google Scholar 

  • Ortega R, Campos J (1995) Some applications of the topological degree to stability theory. Topological methods in differential equations and inclusions. Kluwer Academic Publishing, Dordrecht, pp 377–409

    Chapter  Google Scholar 

  • Padirac A, Fujii T, Rondelez Y (2012) Bottom-up construction of in vitro switchable memories. Proc Natl Acad Sci 109(47):E3212–E3220

    Article  Google Scholar 

  • Qiao L, Nachbar RB, Kevrekidis IG, Shvartsman SY (2007) Bistability and oscillations in the Huang-Ferrell model of MAPK signaling. Public Libr Sci Comput Biol 3(9):e184

    MathSciNet  Google Scholar 

  • Richard A, Comet JP (2011) Stable periodicity and negative circuits in differential systems. J Math Biol 63(3):593–600

    Article  MathSciNet  MATH  Google Scholar 

  • Richeson D, Wiseman J (2002) A fixed point theorem for bounded dynamical systems. Ill J Math 46(2):491–495

    MathSciNet  MATH  Google Scholar 

  • Richeson D, Wiseman J (2004) Addendum to: “a fixed point theorem for bounded dynamical systems” [Illinois J. Math. 46(2), 2002, pp. 491–495]. Ill J Math 48(3):1079–1080

  • Savageau MA, Voit EO (1987) Recasting nonlinear differential equations as S-systems: a canonical nonlinear form. Math Biosci 87:83–115

    Article  MathSciNet  MATH  Google Scholar 

  • Smith HL (2008) Monotone dynamical systems: an introduction to the theory of competitive and cooperative systems. American Mathematical Society, Providence

    Book  Google Scholar 

  • Snoussi E (1998) Necessary conditions for multistationarity and stable periodicity. J Biol Syst 6:3–9

    Article  MATH  Google Scholar 

  • Sontag E (2007) Monotone and near-monotone biochemical networks. Syst Synth Biol 1:59–87

    Article  Google Scholar 

  • Soulé C (2004) Graphic requirements for multistationarity. ComPlexUs 1(3):123–133

    Google Scholar 

  • Srzednicki R (1985) On rest points of dynamical systems. Fund Math 126(1):69–81

    MathSciNet  MATH  Google Scholar 

  • Stricker J, Cookson S, Bennett MR, Mather WH, Tsimring LS, Hasty J (2008) A fast, robust and tunable synthetic gene oscillator. Nature 456(7221):516–519

    Article  Google Scholar 

  • Strogatz SH (1994) Nonlinear dynamics and chaos. Perseus Book Publishing, Reading

    Google Scholar 

  • Thomas R (1981) On the relation between the logical structure of systems and their ability to generate multiple steady states or sustained oscillations. In: Dora J, Demongeot J, Lacolle B (eds) Numerical methods in the study of critical phenomena, Springer Series in Synergetics, vol 9. Springer, Berlin, pp 180–193

    Chapter  Google Scholar 

  • Thomas R (1994) The role of feedback circuits: positive feedback circuits are a necessary condition for positive real eigenvalues of the Jacobian matrix. Berichte der Bunsengesellschaft für physikalische Chemie 98(9):1148–1151

    Article  Google Scholar 

  • Tigges M, Marquez-Lago TT, Stelling J, Fussenegger M (2009) A tunable synthetic mammalian oscillator. Nature 457(7227):309–312

    Article  Google Scholar 

  • Tsai TYC, Choi YS, Ma W, Pomerening JR, Tang C, Ferrell JEJ (2008) Robust, tunable biological oscillations from interlinked positive and negative feedback loops. Science 321(5885):126–129

    Article  Google Scholar 

  • Wang R, Li C, Chen L, Aihara K (2008) Modeling and analyzing biological oscillations in molecular networks. Proc IEEE 96(8):1361–1385

    Article  Google Scholar 

  • Watkins DS (2007) The matrix eigenvalue problem: GR and Krylov subspace methods. Society for Industrial and Applied Mathematics, Philadelphia

    Book  MATH  Google Scholar 

Download references

Acknowledgments

The authors would like to thank Professor F. Zanolin for his extremely valuable suggestions, J. Kim for his feedback on the manuscript and the anonymous reviewers for their constructive comments. Elisa Franco acknowledges financial support from NSF grant CMMI-1266402 and from the Bourns College of Engineering at the University of California at Riverside.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Franco Blanchini.

Appendices

Appendix 1: Non-dimensionalization and Boundedness of the Two-Gene Network

We will carry out the non-dimensionalization procedure for the toggle switch network, leaving the derivation for the other cases to the reader. We follow non-dimensionalization steps similar to those proposed in Elowitz and Leibler (2000); Franco et al. (2011) and Kim and Winfree (2011). Consider the simple (dimensional) model:

$$\begin{aligned}&\tau \,\dot{R}_1= c_{1}+a_1\frac{1}{K_{M1}^{n} + P_2^{n}} - R_1, \quad \dot{P}_1=k_p R_1 - k_d P_1, \end{aligned}$$
(13a)
$$\begin{aligned}&\tau \,\dot{R}_2 =c_{2}+a_2\frac{1}{K_{M2}^{n} + P_1^{n}} - R_2, \quad \dot{P}_2 =k_p R_2 - k_d P_2. \end{aligned}$$
(13b)

Here \(c_{i}\) is the “leak” transcription of RNA. For simplicity, we assume that the translation and degradation rates for the proteins are the same. Constant \(\tau \) is the mRNA half-life in the system. Constants \(K_{Mi}\) represent the number of proteins necessary to half-maximally repress \(R_i\). Finally, assume the translation efficiency of each RNA species is given by \(\bar{p}_i\), which corresponds to the average number of proteins produced by a single RNA molecule. We define the non-dimensional variables: \( r_i={R_i}/{\bar{p}_i}\), \(p_i={P_i}/{K_{Mj}}\), \((i,j)\,\in \,\{(1,2),(2,1)\}\). We rescale time as \( \tilde{t} = t / \tau \), and also define the non-dimensional parameters:

$$\begin{aligned}&\gamma _i=\frac{c_i}{\bar{p}_i},\quad \alpha _i=\frac{a_i}{\bar{p}_i\,K_{Mi}^n},\quad \beta _i=\frac{k_p\,\bar{p}_i}{k_d\,K_{Mj}},\quad T=\frac{1}{\tau k_d}. \end{aligned}$$

The resulting non-dimensional equations are:

$$\begin{aligned}&\dot{r}_1= \gamma _1+\alpha _1\frac{1}{1+p_2^{n}} - r_1, \quad T \dot{p}_1=\beta _1 r_1 - p_1, \end{aligned}$$
(14a)
$$\begin{aligned}&\dot{r}_2 =\gamma _2+\alpha _2\frac{1}{1+p_1^{n}} - r_2, \quad T \dot{p}_2 =\beta _2 r_2 - p_2, \end{aligned}$$
(14b)

Finally, if we assume \(T\approx 1\), we get a system in the same form as Eq. (1).

We also report a brief evidence of the global boundedness of the two-gene system trajectories. The system expressed by Eq. (1) can be rewritten as

$$\begin{aligned} \left[ \begin{array}{c} \dot{r}_1 \\ \dot{p}_1 \\ \dot{r}_2 \\ \dot{p}_2 \end{array} \right]&= \left[ \begin{array}{cccc} ~-1~ &{} ~0~ &{} ~0~ &{} ~0~ \\ \beta _1 &{} -1 &{} 0 &{} 0 \\ 0 &{} 0 &{} -1 &{} 0 \\ 0 &{} 0 &{} \beta _2 &{} -1 \end{array} \right] \, \left[ \begin{array}{c} r_1 \\ p_1 \\ r_2 \\ p_2 \end{array} \right] + \left[ \begin{array}{c} \gamma _1 + H_1(p_2) \\ 0 \\ \gamma _2 + H_2(p_1) \\ 0 \end{array} \right] \\&= \tilde{M} \left[ \begin{array}{c} r_1 \\ p_1 \\ r_2 \\ p_2 \end{array} \right] + \tilde{H}(p_1, p_2). \end{aligned}$$

\(\tilde{M}\) is an asymptotically stable matrix and \(\tilde{H}(p_1, p_2)\) is a bounded quantity (\(\Vert \tilde{H}(p_1, p_2) \Vert \le \eta \)), because \(\gamma _i\) are constants and \(H_i(p_j)\) are Hill functions. The right-hand side of the equation is given by the sum of a linear term and a globally bounded nonlinear term. If we neglect the nonlinear part, we achieve a linear system which is asymptotically stable and admits a global quadratic Lyapunov function \(V(x) = x^\top P x\), where \(\tilde{M}^\top P + P \tilde{M} = - Q\), with \(Q\) positive definite. Hence, all the system solutions will be globally ultimately bounded in a set of the form \(x^\top P x \le k\), for some \(k>0\) which depends on \(P\), \(Q\) and \(\eta \) (see Khalil 2002 Section 5.2 and Blanchini and Miani 2008 Section 4.4.1).

Appendix 2: Critical Systems

In the main text, we have considered non-critical systems in order to provide unified results. Here we discuss our results and examine the validity of our characterization for possibly critical systems. Note that Propositions 3 and 4, concerning candidate multistationary systems, hold true for critical systems as well. The assumption comes into play when we consider the characterization of candidate oscillatory systems.

Validity of Proposition 1

For critical networks, the sufficiency part of Proposition 1 does not hold: The existence of negative cycles of order two does not assure that the system can oscillate. For example, consider the following system:

$$\begin{aligned} \dot{x}_1&=- b_1x_1 + H_{12}(x_2),\nonumber \\ \dot{x}_2&=- b_2x_2 + H_{21}(x_1), \end{aligned}$$
(15)

where \(b_1\) and \(b_2\) are positive constants, and \(H_{ij}\) are positive, bounded, and monotone functions. To ensure the existence of a negative cycle, we assume that (for example) \(H_{12}\) is non-increasing, while \(H_{21}\) is non-decreasing. (We can take \(H_{ij}\) as Hill functions.) For any choice of \(b_i\) and \(H_{ij}\), the system has a single, asymptotically stable equilibrium point. However, oscillations are possible if we take \(b_1=b_2=0\), i.e., if we remove our assumption that negative self-loops must exist at each node.

The necessity part of Proposition 1 still holds, because if there are no negative cycles the system cannot have equilibria with oscillatory instability.

Validity of Proposition 2

In the presence of negative cycles only, instability must be oscillatory. Thus, Proposition 2 still holds. However, some critical systems cannot be destabilized at all, as just shown in example (15).

It is legitimate to ask whether a critical system with only negative cycles (all of order two) is necessarily stable. This would extend Proposition 1 as follows:

Conjecture

A system is a candidate oscillator in the weak sense if and only if it has at least a negative cycle of order greater than two.

While the “if” part of this conjecture is true, unfortunately the the “only if” part is false. This can be seen by counterexample. We modify example (15) by introducing a new variable \(x_3\) and a positive cycle:

$$\begin{aligned} \dot{x}_1&= - b_{1}x_1 + H_{12}(x_2) \\ \dot{x}_2&= - b_{2}x_2 + H_{21}(x_1) + \mu x_3 \\ \dot{x}_3&= - b_{3}x_3 + {\nu x_2.} \end{aligned}$$

The system Jacobian is:

$$\begin{aligned} J = \left[ \begin{array}{ccc} -b_{1} &{} -\delta &{} 0 \\ \gamma &{} - b_{2} &{} \mu \\ 0 &{} \nu &{} - b_{3} \end{array} \right] \end{aligned}$$

This matrix may well have unstable complex eigenvalues. Let us assume \(b_{1}= b_{2} = 0\). The characteristic polynomial is:

$$\begin{aligned} s^3 + b_{3}s^2 + (\gamma \delta -\mu \nu )s + \gamma \delta b_{3} \end{aligned}$$

Building the Routh–Hurwitz table for this polynomial, one can check that for \(\mu ,\nu >0\) two roots are complex with positive real parts. For \(b_{1}, b_{2} >0\) small enough, the property is preserved.

A property relevant to our discussion on critical cases is the following known result (see for instance Edelstein-Keshet 2005, Chapter 6.5):

Proposition 5

If all the self-loops are negative, and all the other cycles are negative and of order two, then any equilibrium of the system is stable.

Appendix 3: Systems Affected by Delays

Many molecular systems have been successfully modeled using delayed differential equations. A notable example is given by gene networks, where the transport of RNA and proteins across cellular membranes are well captured by explicit delays (Hori , Hara 2011; Lewis 2003). Delay differential equations are infinite dimensional systems and a formal, exhaustive treatment of this case would require a more sophisticated setup. However, we can show that local, cycle-based sufficient conditions for OTIs and RTIs can be stated in a wide class of systems with delays. Consider the delay differential equation model

$$\begin{aligned} \dot{x}(t) = f(x(t),x(t-\tau _1),x(t-\tau _2), \dots , x(t-\tau _M)). \end{aligned}$$
(16)

Under standard differentiability assumptions, the corresponding linearized system around an equilibrium point is

$$\begin{aligned} \dot{\xi }(t) = A_0 \xi (t) + \sum _{k=1}^M~A_k \xi (t-\tau _k). \end{aligned}$$
(17)

For simplicity assume that there are no delayed self-loops, i.e., that the matrices \(A_k\) have zero diagonal entries for \(k \ge 1\).

It is well known that the stability of the system above can be established by inspecting the roots of this equation:

$$\begin{aligned} \det \left[ A_0 + \sum _{k=1}^M~A_k e^{-\tau _k s}-sI\right] =0. \end{aligned}$$
(18)

Stability is ensured if the roots have negative real parts. We now consider the following auxiliary system, in which all delays are (fictitiously) set to \(0\):

$$\begin{aligned} \dot{\zeta }(t) = \left[ A_0+\sum _{k=1}^M~A_k\right] \zeta (t) = \bar{A} \zeta (t). \end{aligned}$$
(19)

Matrix \(\bar{A} \) is the same Jacobian we would obtain by setting the delays \(\tau _i =0\) in system (16):

$$\begin{aligned} \dot{x}(t) = f(x(t),x(t),x(t), \dots , x(t)). \end{aligned}$$

Note that equilibria are delay independent, so the delay-free system above has the same equilibria as the delayed system (16).

As previously done in this paper, we can associate graphs with systems (17) and (19), where nodes correspond to species concentrations and signed, directed arcs correspond to the dynamic interactions among species (delayed or not), defined by matrices \(A_k\), \(k=0,\ldots ,M\). Delays do not change the sign of the cycles. Therefore, any positive/negative cycle of (17) corresponds to a positive/negative cycle of (19).

Proposition 6

Assume that system (17) has only negative cycles. Then, the system admits solely OTIs.

Proof

Ab absurdo, assume that the system admits a real transition to instability, and thus, one root of Eq. (18) crosses the imaginary axis with value \(s=0\). This is equivalent to writing Eq. (18) as

$$\begin{aligned} \det \left[ A_0+\sum _{k=1}^M~A_k \right] =0. \end{aligned}$$

If this is true, then also the auxiliary system (19) must admit a zero eigenvalue. But this is impossible if the system has only negative cycles, according to Proposition 2. \(\square \)

Proposition 7

Assume that system (17) has only positive cycles. Then, the system admits solely RTIs.

Proof

To prove this proposition, we need two observations:

  1. (a)

    In the absence of negative cycles (with the exception of self-loops), system (17) is a linear positive system with delay. Therefore, matrix \(\bar{A}\) in system (19) is a Metzler matrix.

  2. (b)

    We invoke a well-known property of positive linear systems with delay (with no delayed self-loops). System (18) is stable if and only if its delay-free, auxiliary counterpart (19) is stable (Haddad and Chellaboina 2004; Liu et al. 2010) (see also Wang et al. 2008, Theorem 6.5).

If all its cycles are positive, the auxiliary system (19) can transition to instability only by means of a pole in \(s=0\). Again ab absurdo let us assume that the delay system (18) admits instability with a pair of dominant imaginary eigenvalues and no real eigenvalue crossing the origin. Then, the auxiliary system (19) would also be unstable. However, the dominant eigenvalue of the Metzler matrix \(\bar{A}\) is real; therefore, the auxiliary system (19) would transition to instability with a pole at \(s=0\). But this would also imply that Eq. (18) is satisfied with \(s=0\), which is a contradiction. Hence, we conclude that if all its cycles are positive, a transition to instability with a pair of imaginary eigenvalues is impossible for system (17). \(\square \)

As a comment to the above proposition, we stress that equilibria are delay independent; therefore, if the auxiliary subsystem presents multiple equilibria, so does the delayed system.

We conclude this appendix by noting that Proposition 5 is not valid in the presence of delays. Indeed, the presence of a delay in a “second order negative cycle” may compromise stability, as it is well known in control theory.

Appendix 4: Structural Cross-Constraints Among Functions

We have not considered models presenting cross-linked dynamic terms in several equations. Cross-linked terms appear typically in models built using the mass-action kinetics formalism. For instance, consider the reactions

where \(\emptyset \) indicates elimination of a species from the system (degradation or out-flow). Indicating the concentration of a species with the corresponding small letter, the differential equation model is:

$$\begin{aligned} \dot{x}&= x_0 - g_X(x) -pg_{XZ}(x,z) \end{aligned}$$
(20)
$$\begin{aligned} \dot{y}&= g_X(x) -g_{Y}(y) \end{aligned}$$
(21)
$$\begin{aligned} \dot{z}&= g_Y(y) -g_{XZ}(x,z). \end{aligned}$$
(22)

Here, all the reaction rates are increasing functions. Identical dynamic terms appear in the three equations, so the Jacobian has dependent entries:

$$\begin{aligned} J = \left[ \begin{array}{ccc} -(\alpha + p\mu ) &{} 0 &{} -p\nu \\ ~\alpha &{} -\beta &{} ~0 \\ -\mu &{} ~\beta &{} - \nu \end{array} \right] , \end{aligned}$$

where, for a fixed equilibrium, all Greek letters represent positive constants. This structure presents both negative and positive cycles, but it may only undergo oscillatory transitions to instability. Thus, it is a candidate oscillator in the strong sense. This fact can be seen by writing the characteristic polynomial:

$$\begin{aligned} p(s) = (s+\alpha ) (s+\beta ) (s+\nu ) + \mu p (s+\beta ) s +p\nu \alpha \beta , \end{aligned}$$

where all coefficients are positive, and thus, there cannot be real positive roots.

This would seem in contradiction with the presence of the positive cycle \(1 \rightarrow 3 \rightarrow 1\). However, our results are not invalidated because we do not consider cross-constraints among entries in the Jacobian, such as the fact that \(J_{22} = - J_{32}\). Clearly, if we could change all entries independently (without changing sign), this system could present a real transition to instability.

For systems with cross-constrained dynamics, we believe that algorithmic/numerical methods are the best approach to discriminate admissible transitions to instability.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Blanchini, F., Franco, E. & Giordano, G. A Structural Classification of Candidate Oscillatory and Multistationary Biochemical Systems. Bull Math Biol 76, 2542–2569 (2014). https://doi.org/10.1007/s11538-014-0023-y

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s11538-014-0023-y

Keywords

Navigation