Theoretical investigation on models of circadian rhythms based on dimerization and proteolysis of PER and TIM

  • Received: 19 May 2016 Revised: 15 January 2017 Published: 01 October 2017
  • MSC : Primary: 34Cxx; Secondary: 92Bxx

  • Circadian rhythms of physiology and behavior are widespread\break mechanisms in many organisms. The internal biological rhythms are driven by molecular clocks, which oscillate with a period nearly but not exactly 24 hours. Many classic models of circadian rhythms are based on a time-delayed negative feedback, suggested by the protein products inhibiting transcription of their own genes. In 1999, based on stabilization of PER upon dimerization, Tyson et al. [J. J. Tyson, C. I. Hong, C. D. Thron, B. Novak, Biophys. J. 77 (1999) 2411-2417] proposed a crucial positive feedback to the circadian oscillator. This idea was mathematically expressed in a three-dimensional model. By imposing assumptions that the dimerization reactions were fast and dimeric proteins were in rapid equilibrium, they reduced the model to a pair of nonlinear ordinary differential equations of mRNA and total protein concentrations. Then they used phase plane analysis tools to investigate circadian rhythms. In this paper, the original three-dimensional model is studied. We explore the existence of oscillations and their periods. Much attention is paid to investigate how the periods depend on model parameters. The numerical simulations are in good agreement with their reduced work.

    Citation: Jifa Jiang, Qiang Liu, Lei Niu. Theoretical investigation on models of circadian rhythms based on dimerization and proteolysis of PER and TIM[J]. Mathematical Biosciences and Engineering, 2017, 14(5&6): 1247-1259. doi: 10.3934/mbe.2017064

    Related Papers:

    [1] Changgui Gu, Ping Wang, Tongfeng Weng, Huijie Yang, Jos Rohling . Heterogeneity of neuronal properties determines the collective behavior of the neurons in the suprachiasmatic nucleus. Mathematical Biosciences and Engineering, 2019, 16(4): 1893-1913. doi: 10.3934/mbe.2019092
    [2] Miguel Lara-Aparicio, Carolina Barriga-Montoya, Pablo Padilla-Longoria, Beatriz Fuentes-Pardo . Modeling some properties of circadian rhythms. Mathematical Biosciences and Engineering, 2014, 11(2): 317-330. doi: 10.3934/mbe.2014.11.317
    [3] Ying Li, Zhao Zhao, Yuan-yuan Tan, Xue Wang . Dynamical analysis of the effects of circadian clock on the neurotransmitter dopamine. Mathematical Biosciences and Engineering, 2023, 20(9): 16663-16677. doi: 10.3934/mbe.2023742
    [4] Yanqin Wang, Xin Ni, Jie Yan, Ling Yang . Modeling transcriptional co-regulation of mammalian circadian clock. Mathematical Biosciences and Engineering, 2017, 14(5&6): 1447-1462. doi: 10.3934/mbe.2017075
    [5] Zhenzhen Shi, Huidong Cheng, Yu Liu, Yanhui Wang . Optimization of an integrated feedback control for a pest management predator-prey model. Mathematical Biosciences and Engineering, 2019, 16(6): 7963-7981. doi: 10.3934/mbe.2019401
    [6] Ya'nan Wang, Sen Liu, Haijun Jia, Xintao Deng, Chunpu Li, Aiguo Wang, Cuiwei Yang . A two-step method for paroxysmal atrial fibrillation event detection based on machine learning. Mathematical Biosciences and Engineering, 2022, 19(10): 9877-9894. doi: 10.3934/mbe.2022460
    [7] E.V. Presnov, Z. Agur . The Role Of Time Delays, Slow Processes And Chaos In Modulating The Cell-Cycle Clock. Mathematical Biosciences and Engineering, 2005, 2(3): 625-642. doi: 10.3934/mbe.2005.2.625
    [8] Jun Zhou . Bifurcation analysis of a diffusive plant-wrack model with tide effect on the wrack. Mathematical Biosciences and Engineering, 2016, 13(4): 857-885. doi: 10.3934/mbe.2016021
    [9] Hussein Obeid, Alan D. Rendall . The minimal model of Hahn for the Calvin cycle. Mathematical Biosciences and Engineering, 2019, 16(4): 2353-2370. doi: 10.3934/mbe.2019118
    [10] Ryotaro Tsuneki, Shinji Doi, Junko Inoue . Generation of slow phase-locked oscillation and variability of the interspike intervals in globally coupled neuronal oscillators. Mathematical Biosciences and Engineering, 2014, 11(1): 125-138. doi: 10.3934/mbe.2014.11.125
  • Circadian rhythms of physiology and behavior are widespread\break mechanisms in many organisms. The internal biological rhythms are driven by molecular clocks, which oscillate with a period nearly but not exactly 24 hours. Many classic models of circadian rhythms are based on a time-delayed negative feedback, suggested by the protein products inhibiting transcription of their own genes. In 1999, based on stabilization of PER upon dimerization, Tyson et al. [J. J. Tyson, C. I. Hong, C. D. Thron, B. Novak, Biophys. J. 77 (1999) 2411-2417] proposed a crucial positive feedback to the circadian oscillator. This idea was mathematically expressed in a three-dimensional model. By imposing assumptions that the dimerization reactions were fast and dimeric proteins were in rapid equilibrium, they reduced the model to a pair of nonlinear ordinary differential equations of mRNA and total protein concentrations. Then they used phase plane analysis tools to investigate circadian rhythms. In this paper, the original three-dimensional model is studied. We explore the existence of oscillations and their periods. Much attention is paid to investigate how the periods depend on model parameters. The numerical simulations are in good agreement with their reduced work.


    1. Introduction

    Wide-type fruit flies, Drosophila melanogaster, might be the most extensively studied organism in circadian rhythm research. The researches of endogenous activity rhythm on Drosophila generally involve two different kinds of clock genes, called period (per, for short) [14,10] and timeless (tim, for short) [20,27]. Their encoded proteins, PER and TIM, bind to each other [5,0,27,29].

    PER protein and per mRNA cycle in a 24-hour period [7]. When PER protein is at a high level, per mRNA expression is repressed, suggesting that PER is an inhibitor of per mRNA accumulation [7]. The expression of per and tim genes is regulated by dCLOCK and CYC, and PER inhibits the transcription of per and tim by inactivating dCLOCK and CYC [1,3,18]. This negative feedback, introduced by PER inhibiting its own mRNA transcription, is the basis of many classic theoretical models of circadian rhythms [6,17,15,19].

    An alternative way to study circadian rhythms is based on a positive feedback, introduced by PER phosphorylation being an activator to PER [26]. Phosphorylation of PER is operated by a double-time gene encoded kinase, DOUBLE-TIME (DBT, for short) [13,16]. As suggested by the dbt mutants phenotypes, PER phosphorylation might be precluded to its degradation. PER and TIM stimulate transcription of per and tim genes by activating dClOCK [2]. Experimental results suggest that per mRNA is stabilized by PER/TIM dimers [24], and PER is stabilized by dimerization with TIM [13,16].

    The idea that PER phosphorylation introduces a positive feedback in PER accumulation can be expressed in a model of three-dimensional ordinary differential equations [26] (see (1) below). In [26], by imposing assumptions that the dimerization reactions were fast and dimeric proteins were in rapid equilibrium, they reduced the three-dimensional model to a pair of nonlinear ordinary differential equations of mRNA and total protein concentrations (see (2) below). Then they used the powerful phase plane portraiture to study the simplified two-dimensional model. In this paper, we explore the original three-dimensional model directly. It is shown that the circadian rhythms occur if the model possesses a unique equilibrium which is unstable. Furthermore, we deeply investigate how circadian rhythms are affected by several model parameters, including mRNA translation, mRNA degradation, monomer phosphorylation, protein proteolysis, association of PER/TIM protein and equilibrium constant for dimerization. The results help to explain some former-observed phenomena of circadian rhythms. In particular, our numerical results extremely agree with those given in [26], indicating that their reduction work is greatly reasonable.


    2. Mechanism and mathematical model

    In this section, we restate the model proposed by Tyson et al. [26]. The molecular mechanism for the circadian rhythm in Drosophila is summarized in Figure 1. Here the total PER (monomer + dimer) degradation rate does not increase proportionally with the total PER concentration's increasing.

    Figure 1. A simple molecular mechanism for the circadian clock in Drosophila. Redrawn from [26]. PER and TIM proteins are synthesized in the cytoplasm, where they may be destroyed by proteolysis or they may combine to form relatively stable heterodimers. Heteromeric complexes are transported into the nucleus, where they inhibit transcription of per and tim mRNA. Here it is assumed that PER monomers are rapidly phosphorylated by DBT and then degraded. Dimers are assumed to be poorer substrates for DBT.

    The mechanism in Figure 1 could be translated into a set of six differential equations, for per and tim mRNAs, PER and TIM monomers, and PER/TIM dimers in the cytoplasm and nucleus. Such a complicated set of equations could not efficiently illustrate the importance of positive feedback in the reaction mechanism. So by noticing that PER and TIM messages and proteins followed roughly similar time courses in vivo, Tyson et al. [26] lumped them into a single pool of clock proteins. In addition, they assumed that the cytoplasmic and nuclear pools of dimeric protein were in rapid equilibrium. Then they established the following differential equations for [mRNA]=M, [monomer]=P1, and [dimer]=P2:

    {dMdt=vm1+(P2/Pcrit)2kmM,dP1dt=vpMkp1P1JP+P1+rP2kp3P12kaP21+2kdP2,dP2dt=kaP21kdP2kp2P2JP+P1+rP2kp3P2. (1)

    Here monomer was assumed to be phosphorylated more quickly than dimer, i.e., kp1kp2. The parameter r determined the inhibition of dimer to monomer phosphorylation. For convenience, in this paper we follow them by taking r=2.

    In their work, it was further assumed that the dimerization reactions were fast (ka and kd are large) such that monomers and dimers were always in equilibrium with each other. Then, by equilibrium conditions: P2=KeqP21, Keq=ka/kd, they obtained the reduced two-dimensional system:

    {dMdt=vm1+(Pt(1q)/(2Pcrit))2kmM,dPtdt=vpMkp1Ptq+kp2PtJP+Ptkp3Pt, (2)

    where Pt=P1+2P2=[total protein], kp1=kp1kp2kp1, and

    q=q(Pt)=21+1+8KeqPt.

    Two widely concerned points of circadian rhythms are whether the endogenous rhythms exist and how long the periods are. Since the mechanism has already been translated into mathematical models, attentions are drawn to examine the existence of periodic orbits and calculate the periods. In their work, system (2) has been thoroughly analyzed. In this paper, we try to study system (1). A typical oscillating solution of system (1) is illustrated in Figure 2, where the corresponding parameter values are chosen from Table 1.

    Figure 2. Numerical solution of (1). Parameter values are chosen as in Table 1. We take ka=106 and kd=ka/Keq.
    Table 1. Parameter values suitable for circadian rhythm of wild-type fruit flies.
    NameValueUnits Ea/RTDescription
    vm1 Cmh6Maximum rate of synthesis of mRNA
    km0.1 h14First-order rate constant for mRNA degradation
    vp0.5 CpCmh6Rate constant for translation of mRNA
    kp110 Cph6 Vmax for monomer phosphorylation
    kp20.03 Cph6 Vmax for dimer phosphorylation
    kp30.1 h16First-order rate constant for proteolysis
    Keq200 C1p-12Equilibrium constant for dimerization
    Pcrit0.1 Cp6Dimer concen at the half-maximum transcription rate
    JP0.05 Cp-16Michaelis constant for protein kinase (DBT)
    This table is adapted from Tyson et al. [26]. Parameters Cm and Cp represent characteristic concentrations for mRNA and protein, respectively. Ea is the activation energy of each rate constant (necessarily positive) or the standard enthalpy change for each equilibrium binding constant (may be positive or negative). The parameter values are chosen to ensure temperature compensation of the wild-type oscillator.
     | Show Table
    DownLoad: CSV

    3. Method and result

    It is well-known that for higher dimensional ordinary differential equations, there is no so-called Poincaré-Bendixson theory: any limit set is a limit cycle if it contains no steady state. So, in order to use the powerful phase plane analysis tools, Tyson et al. [26] reduced (1) into (2) by imposing some assumptions. Fortunately, we observe that (1) is a three-dimensional competitive system in some sense [8,0,22,23]. For n-dimensional competitive ordinary differential equations, the dynamics is co-dimensional one. Every limit set lies on a Lipschitz manifold with one dimension lower, and this manifold is homeomorphic to an (n1)-dimensional Euclidean space [28]. As for a three-dimensional competitive system, though there is no phase plane, one has a two-dimensional Lipschitz manifold (qualitatively exists but is unknown), where recurrent motions of system lie in. As a result, any limit set for such a system consists of either limit cycle, or steady state, or steady states connected with homoclinic or heteroclinic orbits. Suppose that all forward orbits for a three-dimensional competitive system are bounded and the system has a unique steady state. Then by the Perron-Frobenius theory, the linearized matrix at the steady state has a negative eigenvalue. The system has a one-dimensional stable manifold which is a strictly monotone curve [21], so it also rules out the third choice for limit set. For more details, please see [23] or the Appendix. We summarize the above discussion into the following theorem which can be found in [30]:

    Theorem 3.1. Suppose (1) has a unique steady state E. If the linearized matrix of (1) at E has one negative eigenvalue and two positive real part eigenvalues, then (1) has at least one stable limit cycle.

    Numerical calculation suggests that (1) has a unique equilibrium in a large region of parameter values. However, limit cycles do not exist all the time. According to Theorem 3.1, when either Keq or ka is small, (1) has no limit cycles but a unique equilibrium (see Table 2), and the equilibrium appears to be a global stable steady state.

    Table 2. Equilibrium of (1) and corresponding eigenvalues of its Jacobian matrix vary with Keq and ka.
    KeqkaEquilibrium1Eigenvalues
    200106 (10.00,0.05,0) {50.20,0.40,0.1}
    103 (10.00,0.05,6×106) {50.19,0.40,0.1}
    1 (8.62,0.10,0.04){25.96,0.01±0.11i}
    103 (1.38,0.04,0.24) {164.97,0.11±0.41i}
    106 (1.36,0.04,0.25) {1.47×105,0.12±0.42i}
    15106 (10.00,0.05,0) {50.20,0.40,0.1}
    103 (10.00,0.05,6×106) {50.19,0.40,0.1}
    1 (9.60,0.08,0.10) {30.77,0.03±0.08i}
    102 (5.09,0.08,0.10) {63.98,0.66±0.28i}
    103 (5.03,0.08,0.10) {417.94,1.43,0.54}
    106 (5.02,0.08,0.10) {3.9×105,1.57,0.52}
    1106 (10.00,0.05,0) {50.20,0.40,0.1}
    103 (10.00,0.05,6×106) {50.19,0.40,0.1}
    1 (10.00,0.05,2×103) {46.81,1.13,0.10}
    103 (10.00,0.05,3×103) {1240,28.12,0.10}
    106 (10.00,0.05,3×103) {1.2×106,28.49,0.10}
    1 Those zeros in equilibrium terms are actually very small positive numbers. Other parameter values are as given in Table 1.
     | Show Table
    DownLoad: CSV

    Comparing with the two-dimensional system (2), there are two more parameters ka and kd in system (1). Considering the equilibrium condition, Keq=ka/kd, one only needs to detect how periods of (1) are influenced by ka. As shown in Table 3 and Figure 3E, if we take Keq=200, periodic orbits occur when ka is larger than a critical value ka=0.9. In that region, as ka goes up, the period starts with a rapid decline, and then becomes quite insensitive. At first, we guess that the period is decreasing when ka is sufficiently large, but numerical calculations tell that it is not the case. In fact, the period even has a tendency to increase when ka is larger than 2.9×106 (see Table 3). The similar situations are observed with Keq=15 (see Table 4).

    Table 3. Period of endogenous rhythms of wild-type flies varies as ka (Keq=200) varies.
    ka0.0010.10.80.9110100
    Periodnonenonenone72.4463.1050.8932.51
    ka500100050001045×1041055×105
    Period28.6126.9024.8624.5424.2724.2424.21
    ka1062×1062.5×1062.9×1063×106
    Period24.2124.2124.2124.3024.44
    Periodic oscillations happen when ka is larger than the bifurcation value ka=0.9. Other parameter values are as given in Table 1.
     | Show Table
    DownLoad: CSV
    Figure 3. Relation between the oscillator period of (1) and some parameter values. In each diagram, other parameter values are chosen as in Table 1 and ka=106, and periodic oscillations occur only when the correlate parameter is in the interval [a,b]. In case A, a=0.2 and b=1.4; in case B, a=0.02 and b=0.44; in case C, a=7 and b=46; in case D, a=0 and b=0.4; in case E, a=0.9 and b=; in case F, a1=a2=4, b1=570 and b2=588. For the convenience of numerical integration, curve (1) is shown only with Keq40 in case F. As for 4Keq40, a decreasing period is suggested by curve (2) with increasing Keq. Particularly, on curve (1) the period maintains 24.2-25.2 when the parameter Keq varies in the interval [c,d]=[50,460].
    Table 4. Period of endogenous rhythms of perL mutant varies as ka (Keq=15) varies.
    ka0.0010.11.11.2210100500
    Periodnonenonenone57.1955.6741.3430.9829.21
    ka1000200050001041057×1057×1057×105
    Period28.9428.8028.7128.6728.6528.6529.2030.37
    Periodic oscillations occur when ka is beyond the bifurcation value ka=1.2. Other parameter values are as in Table 1.
     | Show Table
    DownLoad: CSV

    Based on Tables 3 and 4, one can choose a suitable value of ka to calculate the rhythms for wild-type and mutant flies. The numerical results are presented in Table 5, where temperature compensation is found in wild-type flies but not in perL mutant flies.

    Table 5. Period of the endogenous rhythms of wild-type and mutant flies based on (1).
    Genotype KeqTempPeriodGenotype kp1 kp2Period
    Wild type2452024.2 dbt+(1×)100.0324.2
    2002524.2 dbt+(2×)150.0624.3
    1643024.2 dbt+(3×)200.0925.7
    perL18.42026.5 dbtS100.317.6
    15.02528.7 dbt+100.0324.2
    12.33030.4 dbtL100.00325.1
    To simplify the integration, we take ka=106 for wild-type flies and ka=5000 for mutant flies. Other conditions are as in Table 6.
     | Show Table
    DownLoad: CSV

    Table 5 is due to the original three-dimensional system (1). As a comparison, we state Table 6, which is cited from [26] and based on the reduced two-dimensional system (2). Clearly, one can see that Table 5 and Table 6 are almost the same, which indicates that the reduction in [26] is greatly reasonable from this perspective.

    Table 6. Period of the endogenous rhythms of wild-type and mutant flies based on (2).
    Genotype KeqTempPeriodGenotype kp1 kp2Period
    Wild type2452024.2 dbt+(1×)100.0324.2
    2002524.2 dbt+(2×)150.0624.4
    1643024.2 dbt+(3×)200.0925.7
    perL18.42026.5 dbtS100.317.6
    15.02528.7 dbt+100.0324.2
    12.33030.5 dbtL100.00325.2
    This table is copied out of Tyson et al. [26]. It is assumed that each parameter k varies with temperature according to k(T)=k(298)exp{εa(1298/T)}, with values for k(298) and εa=Ea/(0.592kcalmol1) given in Table 1. The dbt+(n×) means n copies of the wild-type allele.
     | Show Table
    DownLoad: CSV

    In the next section, we will see more about the relation between circadian rhythms and parameters of (1).


    4. Discussion

    In the actual experiment, parameters of the circadian rhythms models are hard to be measured, or even unmeasurable. Parameter values in Table 1 have been chosen to yield a period close to 24-hours and ensure temperature compensation of the wild-type oscillator. The parameter values are arbitrary. Other combinations of parameter values may also yield circadian oscillations with possibly different periods.

    It is significant to study how parameters of (1) affect its periodic oscillations. The numerical results are given in Figure 3, where the following parameters are considered: mRNA translation, mRNA degradation, monomer phosphorylation, protein proteolysis, association of PER/TIM protein and equilibrium constant for dimerization.

    As shown in Figure 3A, periodic oscillation disappears when the protein synthesis rate vp is below a critical value. That coincides with the truth that the inhibiting effect of protein synthesis may eventually suppress the circadian rhythmicity [11,4,25]. Moreover, the period decreases when the protein synthesis rate is greater than a certain value. That matches the observations of anisomycin in the mollusk Bulla [12].

    The PER/TIM complex formation plays a key role in the model. Circadian rhythm is markedly affected by the dimerization reaction, precisely in the model, by the association rate constant ka and dissociation rate constant kd. In Figure 3E the period decreases as ka increases, which coincides with the suggestion that the heterodimeric dimerization is attenuated in the long-period perL mutant [5]. Here the attenuation is probably due to the competition of PER homodimeric complexes [0]. Note again that it always has kd=ka/Keq in this paper. The results of Figure 3F imply that circadian rhythm occurs only when kd is in a bounded range, and the period can be recognized as an increasing function of the dimer disassociation rate.

    In Figure 3B we show how the oscillation is affected by mRNA synthesis. Periodic rhythm requires the mRNA synthesis rate km to be bounded, implying that the oscillation may be destroyed if mRNA synthesizes either too slow or too fast. Moreover, the period of oscillation becomes shorter as the mRNA synthesis rate goes up. The qualitatively similar results (Figure 3D) are detected when we consider the proteins proteolysis rate kp3, except that periodic oscillation happens even if there is no proteins proteolysis.

    According to Theorem 3.1, in Figure 4 we inspect the dependence of oscillations on parameters Keq and kp1. A U-shape region is found, whose boundary is almost the same as the locus of Hopf bifurcation in [26] (see Figure 4 in [26]). Within that region the system exhibits periodic oscillations, and in the outside area the system exhibits no limit cycle but a stable steady state. Figure 3C and 3F help to investigate the variation of period in this U-shape region. In Figure 3C, periodic oscillation requests that kp1 is neither too small nor too large, which means that protein monomers are sufficient but not too unstable. In Figure 3F, periodic oscillation vanishes when Keq is smaller than a critical value, which implies that the proteins tend to dimerize. Meanwhile, when Keq varies within a quite large region beyond a certain value, the period of (1) remains virtually unchanged, suggesting that the wild-type oscillation has temperature compensation (see also Table 5). Furthermore, when Keq decreases in a large region the period increases, which agrees with the consensus that perL mutant introduces a longer period for the perL-encoded protein to reduce its tendency to form dimers [5,10]. By Table 5, one can also tell that perL mutant loses temperature compensation.

    Figure 4. Two-parameter (Keq and kp1) bifurcation diagram for system (???). Here Keq and kp1 are allowed to vary, and other parameter values are fixed as in Table 1. We take ka=106. Periodic oscillations happen only within the U-shape region bounded by the two curves. Outside this region the system evolves toward a stable steady state. We note that for any Keq one can find a kp1 such that oscillations happen, which differs from the boundedness requirement of Keq as in Figure 3F.

    Appendix. The concentrations of mRNA, monomers and dimers are naturally nonnegative. We therefore focus on the first orthant R3+={(M,P1,P2):M0,P10,P20}. It is easy to see that R3+ is a positively invariant set of system (1), i.e., any solution ((M(t),P1(t),P2(t)) of system (1) through a point in the first orthant lies in it when t0.

    Let a>vm/km,  b>vpa/kp3,  c=Keqb2 and B(a,b,c)={(M,P1,P2):0Ma, 0P1b, 0P2c}. Denote (f1,f2,f3) the vector field of (1). By estimating the sign of the vector field at vertexes on the boundary of B(a,b,c), one has

    {f1(a,P1,P2)=vm1+(P2/Pcrit)2kma<vmkma<0,f2(M,b,P2)=vpMkp1bJP+b+rP2kp3b2kab2+2kdP2<vpakp3b<0,f3(M,P1,c)=kaP21kdckp2cJP+P1+rckp3c<kp2cJP+P1+rckp3c<0.

    The vector field for (1) on the boundary of B(a,b,c) is shown in Figure 5, which implies that B(a,b,c) is positively invariant. Note that for any point in R3+, one can find such (a,b,c) satisfying that box B(a,b,c) contains the point. It follows immediately that any forward solution of system (1) is bounded. We summarize the above discussion into the following proposition:

    Figure 5. The vector field for (1) on the boundary of B(a,b,c).

    Proposition 1. For any a>vm/km,  b>vpa/kp3 and c=Keqb2, B(a,b,c) is positively invariant for (1), that is, all forward solutions for (1) are bounded.

    From Proposition 1, there are at least one steady state in B(a,b,c). Suppose that the steady state E is unique and there is no zero real part eigenvalue for the linearized matrix at E. Then the equilibrium E is either locally asymptotically stable, or has a two-dimensional unstable manifold. The following arguments show that the latter case provides the existence of limit cycles.

    By computing the Jacobian matrix of (1), one has

    Df=(0++0+),

    where '''' represents that the entry is strictly negative and ''+'' means strict positivity. According to [23], the system is competitive with respect to the cone K={(M,P1,P2)R3:M0,P10,P20}. By applying the theory on competitive systems in [23], we have

    Theorem A. Suppose (1) has a unique steady state E. If the linearized matrix of (1) at E has one negative eigenvalue and two positive real part eigenvalues, then (1) has at least one stable limit cycle.

    Therefore, in order to study the oscillations for (1), one only needs to discuss its steady state and the local stability of the steady state.


    [1] [ R. Allada,N. E. White,W. V. So,J. C. Hall,M. Rosbash, A mutant Drosophila homolog of mammalian clock disrupts circadian rhythms and transcription of period and timeless, Cell, 93 (1998): 791-804.
    [2] [ K. Bae,C. Lee,D. Sidote,K-Y. Chuang,I. Edery, Circadian regulation of a Drosophila homolog of the mammalian clock gene: PER and TIM function as positive regulators, Mol. Cell. Biol., 18 (1998): 6142-6151.
    [3] [ T. K. Darlington,K. Wager-Smith,M. F. Ceriani,D. Staknis,N. Gekakis,T. D. L. Steeves,C. J. Weitz,J. S. Takahashi,S. A. Kay, Closing the circadian loop: CLOCK-induced transcription of its own inhibitors per and tim, Science, 280 (1998): 1599-1603.
    [4] [ A. Eskin,S. J. Yeung,M. R. Klass, Requirement for protein synthesis in the regulation of a circadian rhythm by serotonin, Proc. Natl. Acad. Sci. USA, 81 (1984): 7637-7641.
    [5] [ N. Gekakis,L. Saez,A.-M. Delahaye-Brown,M. P. Myers,A. Sehgal,M. W. Young,C. J. Weitz, Isolation of timeless by PER protein interaction: Defective interaction between timeless protein and long-period mutant PERL, Science, 270 (1995): 811-815.
    [6] [ A. Goldbeter, A model for circadian oscillations in the Drosophila period protein (PER), Proc. R. Soc. Lond. B, 261 (1995): 319-324.
    [7] [ P. E. Hardin,J. C. Hall,M. Rosbash, Feedback of the Drosophila Period gene product on circadian cycling of its messenger RNA levels, Nature, 343 (1990): 536-540.
    [8] [ M. W. Hirsch, Systems of differential equations which are competitive or cooperative. Ⅰ: Limit sets, SIAM J. Math. Anal., 13 (1982): 167-179.
    [9] [ M. W. Hirsch, Systems of differential equations that are competitive and cooperative. Ⅳ: Structural stability in three dimensional systems, SIAM J. Math. Anal., 21 (1990): 1225-1234.
    [10] [ Z. J. Huang,K. D. Curtin,M. Rosbash, PER protein interactions and temperature compensation of a circadian clock in Drosophila, Science, 267 (1995): 1169-1172.
    [11] [ M. W. Karakashian,J. W. Hastings, The effects of inhibitors of macromolecular biosynthesis upon the persistent rhythm of luminescence in Gonyaulax, J. Gen. Physiol., 47 (1963): 1-12.
    [12] [ S. B. S. Khalsa,D. Whitmore,G. D. Block, Stopping the circadian pacemaker with inhibitors of protein synthesis, Proc. Natl. Acad. Sci. USA, 89 (1992): 10862-10866.
    [13] [ B. Kloss,J. L. Price,L. Saez,J. Blau,A. Rothenfluh,C. S. Wesley,M. W. Young, The Drosophila clock gene double-time encodes a protein closely related to human casein kinase lϵ, Cell, 94 (1998): 97-107.
    [14] [ R. J. Konopka,S. Benzer, Clock mutants of Drosophila melanogaster, Proc. Natl. Acad. Sci. USA, 68 (1971): 2112-2116.
    [15] [ J.-C. Leloup,A. Goldbeter, A model for circadian rhythms in Drosophila incorporating the formation of a complex between PER and TIM proteins, J. Biol. Rhythms, 13 (1998): 70-87.
    [16] [ J. L. Price,J. Blau,A. Rothenfluh,M. Abodeely,B. Kloss,M. W. Young, double-time is a novel Drosophila clock gene that regulates PERIOD protein accumulation, Cell, 94 (1998): 83-95.
    [17] [ P. Ruoff,L. Rensing, The temperature-compensated Goodwin model simulates many circadian clock properties, J. Theor. Biol., 179 (1996): 275-285.
    [18] [ J. E. Rutila,V. Suri,M. Le,M. V. So,M. Rosbash,J. C. Hall, CYCLE is a second bHLH-PAS clock protein essential for circadian rhythmicity and transcription of Drosophila period and timeless, Cell, 93 (1998): 805-814.
    [19] [ T. Scheper,D. Klinkenberg,C. Pennartz,J. van Pelt, A mathematical model for the intracellular circadian rhythm generator, J. Neurosci., 19 (1999): 40-47.
    [20] [ A. Sehgal,J. L. Price,B. Man,M. W. Young, Loss of circadian behavioral rhythms and per RNA oscillations in the Drosophila mutant timeless, Science, 263 (1994): 1603-1606.
    [21] [ J. F. Selgrade, Asymptotic behavior of solutions to single loop positive feedback systems, J. Diff. Eqns., 38 (1980): 80-103.
    [22] [ H. L. Smith, Periodic orbits of competitive and cooperative systems, J. Diff. Eqns., 65 (1986): 361-373.
    [23] [ H. L. Smith, Monotone Dynamical Systems: An Introduction to the Theory of Competitive and Cooperative Systems, Amer. Math. Soc. Providence, Rhode Island, 1995.
    [24] [ V. Suri,A. Lanjuin,M. Rosbash, TIMELESS-dependent positive and negative autoregulation in the Drosophila circadian clock, EMBO J., 18 (1999): 501-791.
    [25] [ W. R. Taylor,J. C. Dunlap,J. W. Hastings, Inhibitors of protein synthesis on 80s ribosomes phase shift the Gonyaulax clock, J. Exp. Biol., 97 (1982): 121-136.
    [26] [ J. J. Tyson,C. I. Hong,C. D. Thron,B. Novak, A simple model of circadian rhythms based on dimerization and proteolysis of PER and TIM, Biophys. J., 77 (1999): 2411-2417.
    [27] [ L. B. Vosshall,J. L. Price,A. Sehgal,L. Saez,M. W. Young, Block in nuclear localization of period protein by a second clock mutation, timeless, Science, 263 (1994): 1606-1609.
    [28] [ Y. Wang,J. Jiang, The general properties of discrete-time competitive dynamical systems, J. Diff. Eqns., 176 (2001): 470-493.
    [29] [ H. Zeng,Z. Qian,M. P. Myers,M. Rosbash, A light-entrainment mechanism for the Drosophila circadian clock, Nature, 380 (1996): 129-135.
    [30] [ H.-R. Zhu,H. L. Smith, Stable periodic orbits for a class of three dimensional competitive systems, J. Diff. Eqns., 110 (1994): 143-156.
  • This article has been cited by:

    1. Shuang Chen, Jinqiao Duan, Ji Li, Dynamics of the Tyson–Hong–Thron–Novak circadian oscillator model, 2021, 420, 01672789, 132869, 10.1016/j.physd.2021.132869
    2. Alessio Franci, Marco Arieli Herrera-Valdez, Miguel Lara-Aparicio, Pablo Padilla-Longoria, Synchronization, Oscillator Death, and Frequency Modulation in a Class of Biologically Inspired Coupled Oscillators, 2018, 4, 2297-4687, 10.3389/fams.2018.00051
  • Reader Comments
  • © 2017 the Author(s), licensee AIMS Press. This is an open access article distributed under the terms of the Creative Commons Attribution License (http://creativecommons.org/licenses/by/4.0)
通讯作者: 陈斌, bchen63@163.com
  • 1. 

    沈阳化工大学材料科学与工程学院 沈阳 110142

  1. 本站搜索
  2. 百度学术搜索
  3. 万方数据库搜索
  4. CNKI搜索

Metrics

Article views(3561) PDF downloads(449) Cited by(2)

Article outline

Figures and Tables

Figures(5)  /  Tables(6)

Other Articles By Authors

/

DownLoad:  Full-Size Img  PowerPoint
Return
Return

Catalog