Citation: Niclas Kruff, Sebastian Walcher. Coordinate-independent singular perturbation reduction for systems with three time scales[J]. Mathematical Biosciences and Engineering, 2019, 16(5): 5062-5091. doi: 10.3934/mbe.2019255
[1] | Linard Hoessly, Carsten Wiuf . Fast reactions with non-interacting species in stochastic reaction networks. Mathematical Biosciences and Engineering, 2022, 19(3): 2720-2749. doi: 10.3934/mbe.2022124 |
[2] | Peter Rashkov, Ezio Venturino, Maira Aguiar, Nico Stollenwerk, Bob W. Kooi . On the role of vector modeling in a minimalistic epidemic model. Mathematical Biosciences and Engineering, 2019, 16(5): 4314-4338. doi: 10.3934/mbe.2019215 |
[3] | Glenn Ledder . Forest defoliation scenarios. Mathematical Biosciences and Engineering, 2007, 4(1): 15-28. doi: 10.3934/mbe.2007.4.15 |
[4] | Andrei Korobeinikov, Aleksei Archibasov, Vladimir Sobolev . Order reduction for an RNA virus evolution model. Mathematical Biosciences and Engineering, 2015, 12(5): 1007-1016. doi: 10.3934/mbe.2015.12.1007 |
[5] | Kelum Gajamannage, Erik M. Bollt . Detecting phase transitions in collective behavior using manifold's curvature. Mathematical Biosciences and Engineering, 2017, 14(2): 437-453. doi: 10.3934/mbe.2017027 |
[6] | Qian Zhang, Haigang Li, Ming Li, Lei Ding . Feature extraction of face image based on LBP and 2-D Gabor wavelet transform. Mathematical Biosciences and Engineering, 2020, 17(2): 1578-1592. doi: 10.3934/mbe.2020082 |
[7] | Dan Yang, Shijun Li, Yuyu Zhao, Bin Xu, Wenxu Tian . An EIT image reconstruction method based on DenseNet with multi-scale convolution. Mathematical Biosciences and Engineering, 2023, 20(4): 7633-7660. doi: 10.3934/mbe.2023329 |
[8] | Hany Bauomy . Safety action over oscillations of a beam excited by moving load via a new active vibration controller. Mathematical Biosciences and Engineering, 2023, 20(3): 5135-5158. doi: 10.3934/mbe.2023238 |
[9] | Joseph D. Skufca, Erik M. Bollt . Communication and Synchronization in Disconnected Networks with Dynamic Topology: Moving Neighborhood Networks. Mathematical Biosciences and Engineering, 2004, 1(2): 347-359. doi: 10.3934/mbe.2004.1.347 |
[10] | OPhir Nave, Shlomo Hareli, Miriam Elbaz, Itzhak Hayim Iluz, Svetlana Bunimovich-Mendrazitsky . BCG and IL − 2 model for bladder cancer treatment with fast and slow dynamics based on SPVF method—stability analysis. Mathematical Biosciences and Engineering, 2019, 16(5): 5346-5379. doi: 10.3934/mbe.2019267 |
Ordinary differential equations involving a small parameter appear frequently in mathematics and in science. Their principal use in chemistry and biochemistry – which is the main topic of the present paper – is to find certain (attracting) invariant sets and to achieve reduction of dimension. The mathematical basis is singular perturbation theory, originally due to Tikhonov [1] and Fenichel [2], for systems with one small parameter ε (or, in other words, for systems with two time scales).
While Tikhonov's and Fenichel's theory is concerned with first order approximations in ε, there exist approaches to include higher order terms in ε, e.g. to improve accuracy in the approximation of invariant manifolds; see for instance the critical survey in Kaper and Kaper [3]. More recently, Noel et al. [4], Radulescu et al. [5], Samal et al. [6,7] developed an algorithmic method to compute slow-fast scenarios in chemical reaction networks, using tropical geometry. Concerning the existence (or persistence) of invariant sets obtained by such (a priori formal) calculations one may invoke hyperbolicity properties; for instance Theorem 4.1 in Chicone [8] is very useful in this respect. A direct method for chemical reaction networks involving different orders of a single small parameter, given certain properties of the system, is due to Cappeletti and Wiuf [9].
A different perspective is the consideration of systems with more than two time scales by introducing, cum grano salis, several small parameters ε1,ε2,…, and to obtain invariant manifolds and reduction on this basis. (One has the option to set all parameters equal in the end.)
Recently Cardin and Teixeira [10] generalized Fenichel's fundamental theorems, proving results on invariant sets and reductions of systems with more than two time scales. Here, the differential equation systems are assumed to have variables separated into blocks of fast, slow, "very slow" ones, and so on.
The present paper is based, on the one hand, on Cardin and Teixeira [10]. On the other hand, we extend earlier work [11,12] that is concerned with coordinate-independent reduction (not requiring an a priori separation of slow and fast variables), as well as with the basic question of finding – in arbitrary parameter dependent systems – critical parameter values from which singular perturbation reductions~emanate.
We will focus on the three time-scale setting, essentially to keep notation manageable, and will only briefly sketch extensions to more than three time scales. Furthermore we will mostly consider systems that satisfy not only the normal hyperbolicity conditions from [10] but have the stronger feature of exponential attractivity. One reason for this restriction lies in our interest in chemical reaction networks. But beyond this practical consideration, the algorithm to compute critical parameter values for singular perturbation scenarios indeed requires this additional property.
The paper is organized as follows. In Section 2 we review the work by Cardin and Teixeira [10]. Section 3 generalizes the coordinate-independent reduction algorithm from [11] to three-time scale systems. In Section 4 we start from a general parameter dependent system and extend the work from [12] on critical parameter values (Tikhonov-Fenichel parameter values) to three time scales (resp. two "small parameters"), and in Section 5 we discuss two classical examples (cooperativity with two complexes, competitive inhibition) from biochemistry in detail. Section 6 contains a few remarks about more than three time scales, and finally, for the reader's convenience, we prove some essentially known facts in an Appendix.
In this section we review and specialize results from Cardin and Teixeira [10] for a parameter dependent ordinary differential equation system
˙x1=dx1dt=ε1ε2f1(x,ε1,ε2)˙x2=dx2dt=ε1ε1f2(x,ε1,ε2)˙x3=dx3dt=ε1ε2f3(x,ε1,ε2);briefly ˙x=f(x,ε1,ε2). | (2.1) |
Here x=(x1,x2,x3)tr∈Rn with x1∈Rn1, x2∈Rn2, and x3∈Rn3, and f is smooth on an open neighborhood of U×[0,δ1)×[0,δ2), with U⊆Rn open and nonempty, and δ1>0, δ2>0.
We define
M1:={x∈U;f1(x,0,0)=0} | (2.2) |
and
M2:={x∈U;f1(x,0,0)=f2(x,0,0)=0}, | (2.3) |
and we will assume throughout that these sets are nonempty. Cardin and Teixeira require some hyperbolicity conditions, which we state here in slightly stronger versions, for the sake of simplicity:
● First hyperbolicity condition: For every x∈M1, all the eigenvalues of Dx1f1(x,0,0)* have nonzero real parts.
*For a smooth function g=g(x,y,…) we denote the partial derivatives by Dxg, Dyg etc.
For sufficiently small ε1,ε2 this condition implies local solvability of the implicit equation f1(x,ε1,ε2)=0 in the form x1=g(x2,x3,ε1,ε2), and one may furthermore write
˜f2(x2,x3,ε1,ε2):=f2(g(x2,x3,ε1,ε2),x2,x3,ε1,ε2). |
● Second hyperbolicity condition: For every x∈M2, all the eigenvalues of Dx2˜f2(x,0,0) have nonzero real parts.†
†In [10] the second hyperbolicity condition is erroneously written for f2 rather than ˜f2. The authors are aware of this and will publish a corrigendum.
By suitable choice of U, δ1 and δ2 we may assume that M1 and M2 are submanifolds.
Continuing to follow [10] we introduce the auxiliary system
0=dx1dτ2=ε2f1(x,0,ε2)˙x2=dx2dτ2=ε1f2(x,0,ε2)˙x3=dx3dτ2=ε2f3(x,0,ε2) | (2.4) |
on
Mε22:={x∈U;f1(x,0,ε2)=0}, |
and the intermediate reduced system
0=dx1dτ2=f1(x,0,0)˙x2=dx2dτ2=f2(x,0,0)˙x3=dx3dτ2=0 | (2.5) |
on M1. Thus in both equations (2.4) and (2.5) above the dot denotes differentiation with respect to τ2:=ε1t. By suitable choice of δ2 we may also assume that every Mε22 is a submanifold of Rn.
Finally we define the completely reduced system
0=dx1dτ3=f1(x,0,0)0=dx2dτ3=f2(x,0,0)˙x3=dx3dτ3=f3(x,0,0) | (2.6) |
on M2, hence the dot in (2.6) denotes differentiation with respect to τ3:=ε1ε2t. From here on we will follow [10] by just using dots for differentiations, with the appropriate time scale evident from the context.
We replace the hyperbolicity conditions from [10] by stronger requirements, since in our applications we focus on attracting invariant manifolds.
Definition 1. We say that system (2.1) satisfies the hyperbolic attractivity condition (HA) if Dx1f1(x,0,0) has only eigenvalues with negative real part on M1 and if furthermore Dx2˜f2(x,0,0) has only eigenvalues with negative real part on M2.
Our starting point is the following theorem, specialized from Cardin and Teixeira [10], Theorems A, B and Corollary A. Some of our statements are informal; for rigorous statements and pertinent definitions we refer to [10].
Theorem 1. Let system (2.1) be given, with (HA) satisfied.
(a) Let N⊆M2 be a compact submanifold (with nonempty interior in the relative topology, and possibly with boundary). Then for all sufficiently small ε1,ε2 there exists a locally invariant manifold Nε1,ε2 for system (2.1) which is O(ε1+ε2) close to N, diffeomorphic to N and locally exponentially attracting. Given the appropriate time scales, solutions of (2.1) on Nε1,ε2 converge to solutions of (2.6) on N.
(b) Let ε2 be sufficiently small and let L⊆Mε22 be a compact submanifold (with nonempty interior in the relative topology, and possibly with boundary). Then for all sufficiently small ε1 there exists a locally invariant manifold Lε1,ε2 for system (2.1) which is O(ε1+ε2) close to L, diffeomorphic to L and locally exponentially attracting. Given the appropriate time scales, solutions of (2.1) on Lε1,ε2 converge to solutions of (2.5) on L.
Note that the passage from (2.1) to (2.6) may be seen as a singular perturbation reduction for two time scales. From this perspective, system (2.5) indeed represents an intermediate step. For systems with two time scales this leads to the question whether intermediates exist.
As given, the part regarding ˜f2 in condition (HA) is not ready to use in applications. We provide two equivalent versions.
Proposition 1. Condition (HA) is equivalent to either of the following conditions.
(i) Dx1f1(x,0,0) has only eigenvalues with negative real parts on M1, and
B1(x):=−Dx1f2(x,0,0)Dx1f1(x,0,0)−1Dx2f1(x,0,0)+Dx2f2(x,0,0) |
has only eigenvalues with negative real parts on M2.
(ii) Dx1f1(x,0,0) has only eigenvalues with negative real parts on M1, and for all sufficiently small ε>0 the matrix
B2(x,ε):=(Dx1f1(x,0,0)Dx2f1(x,0,0)εDx1f2(x,0,0)εDx2f2(x,0,0)) |
has only eigenvalues with negative real parts on M2.
Proof. We use the notions introduced with the hyperbolicity condition (H) and Definition 1. From
f1(g(x2,x3,ε1,ε2),x2,x3,ε1,ε2)=0 |
one gets by the chain rule
Dx2g(x2,x3)=−Dx1f1(x,,ε1,ε2)−1Dx2f1(x,,ε1,ε2) |
when f1(g(x2,x3,ε1,ε2),x2,x3,ε1,ε2)=0, and a further application of the chain rule shows the equivalence of (HA) and (ⅰ). The equivalence of (ⅰ) and (ⅱ) follows from Lemma 3 in the Appendix.
Remark 1. (a) One may rewrite systems (2.1) through (2.6) to some extent, with no effect on the reductions. Using Hadamard's lemma, one may restate (2.1) as
˙x1=ˆf1(x,ε2)+ε1ˆf1,1(x,ε1)+ε1ε2ˆf1,2(x,ε1,ε2)˙x2=+ε1ˆf2(x,ε1)+ε1ε2ˆf2,2(x,ε1,ε2)˙x3=+ε1ε2ˆf3(x,ε1,ε2) |
with only the ˆfi remaining in the subsequent reductions. Thus the auxiliary system becomes
0=ε2ˆf1(x,ε2)˙x2=ε1ˆf2(x,0)˙x3=ε2f3(x,0,ε2) |
and there are analogous modifications for the intermediate and the fully reduced system.
(b) The passage from (2.1) to the completely reduced system (2.6) can evidently be obtained in the following manner: Fix ε1>0 and reduce (2.1) with respect to the small parameter ε2 (in time scale ε2t). Then let ε1→0, rescaling time once more to τ3. We will use this observation later on.
In the present section we generalize the coordinate-independent reduction procedure from [11,13] to the three-time scale setting. The first task is to intrinsically characterize those systems which admit a transformation to "standard form" (2.1). Reversing matters, applying a (local) smooth coordinate transformation to Eq (2.1) yields a smooth system
˙x=g(0,0)(x,ε1,ε2)+ε1(g(1,0)(x,ε1,ε2)+ε2g(1,1)(x,ε1,ε2)) | (3.1) |
on an open neighborhood of ˜U×[0,δ1)×[0,δ2)⊆Rn×R×R (˜U⊆Rn open), evidently satisfying the following conditions:
(ⅰ) For all sufficiently small ε1≥0,ε2≥0, the zeros of g(0,0)(x,ε1,ε2) form a submanifold ˜M1⊆˜U, of codimension n1, 1≤n1<n. Given any compact submanifold P1⊆˜M1, there exists θ1>0 such that at every y∈P1 the derivative Dxg(y,ε1,ε2) admits the eigenvalue zero with algebraic and geometric multiplicity n−n1, and the remaining eigenvalues have real parts ≤−θ1.
(ⅱ) For all sufficiently small ε1>0,ε2≥0 the zeros of
g(0,0)(x,ε1,ε2)+ε1g(1,0)(x,ε1,ε2) |
form a submanifold ~M2⊆˜U, of codimension n1+n2, 1≤n2<n−n1. Moreover, for any compact submanifold P2⊆˜M2 there exists a θ2>0 with the following property: At every y∈P2 the derivative Dxg(0,0)(y,ε1,ε2)+ε1Dxg(1,0)(y,ε1,ε2) admits the eigenvalue zero with algebraic and geometric multiplicity n−n1−n2, and the remaining eigenvalues have real parts ≤−θ2ε1.
By Remark 1 one may assume that system (3.1) is in the special form
˙x=g(0,0)(x,ε2)+ε1(g(1,0)(x,ε1)+ε2g(1,1)(x))+O(ε2(ε1+ε2)), | (3.2) |
adjusting conditions (ⅰ) and (ⅱ) accordingly. Conditions (ⅰ) and (ⅱ) are certainly necessary for (3.1) or (3.2) to be a transformed version of (2.1). The first part of the next lemma shows sufficiency.
Lemma 1. (a) There exists a local diffeomorphism transforming system (3.2) to a system of type (2.1) with condition (HA) if and only if conditions (i) and (ii) above hold.
(b) Condition (i) for system (3.2) is equivalent to the following: For any y∈~M1 there exist a neighborhood U1,y, a smooth map P1:U1,y→Rn×n1 such that P1(y,ε2) has rank n1, and a smooth map μ1:U1,y→Rn1 such that Dxμ1(y,ε2) has rank n1, yielding a decomposition
g(0,0)(x,ε2)=P1(x,ε2)μ1(x,ε2), |
and moreover there is a θ1>0 such that
A1(x,ε2):=Dμ1(x,ε2)P1(x,ε2) |
has only eigenvalues with real part ≤−θ1, for all x∈U1,y.
(c) In presence of condition (i), condition (ii) for system (3.2) is equivalent to the following: For every (sufficiently small) ε1>0 and any y∈~M2 there exist a neighborhood U2,y, a smooth map P2:U2,y→Rn×n2 such that (P1(y,ε2),ε1P2(y,ε1)) has rank n1+n2, and a smooth map μ2:U2,y→Rn2 such that (Dxμ1(y,ε2),Dxμ2(y,ε2))tr has rank n1+n2, yielding a decomposition
g(0,0)(x,ε2)+ε1g(1,0)(x,ε1)=P1(x,ε2)μ1(x,ε2)+ε1P2(x,ε1)μ2(x,ε1), |
and moreover there is a θ2>0 such that
A2(x,ε1,ε2):=(Dμ1(x,ε2)Dμ2(x,ε1))(P1(x,ε2)ε1P2(x,ε1)) |
has only eigenvalues with real part ≤−θ2ε1, for all x∈U2,y.
Proof The nontrivial assertion of part (a) follows from the existence of n−n1 independent first integrals of g(0,0) in a neighborhood of y, which was noted by Fenichel [2], Lemma 5.3 for smooth vector fields, and shown in [13], Proposition 2.2 for the analytic setting, and likewise from the existence of n−n1−n2 independent first integrals of g(0,0)+ε1g(1,0) in a neighborhood of y. These first integrals determine slow and "very slow" variables. Parts (b) and (c) are straightforward applications of [11], Theorem 1, Remark 4 and Remark 2.
Remark 2. The existence of the decomposition g(0,0)=P1μ1 in part (b) (as well as the decomposition in part (c)) is a consequence of the implicit function theorem in the smooth or analytic case. For polynomial or rational vector fields there exists a decomposition with rational functions as entries of P1 and μ1, and there is an algorithmic approach to its computation. See [11] for details.
Next we use the decompositions to compute reductions.
Proposition 2. (a) In arbitrary coordinates the reduction corresponding to the passage from system (2.1) to the auxiliary system may be obtained as follows:
Given ε2≥0, determine the projection matrix
Q1(x,ε2):=In−P1(x,ε2)A1(x,ε2)−1Dxμ1(x,ε2). |
The auxiliary system (2.4) for ε2>0 then corresponds to
˙x=Q1(x,ε2)(g(1,0)(x,0)+ε2g(1,1)(x)) |
on the local invariant manifold defined by μ1(x,ε2)=0. The equation corresponding to the intermediate reduced system (2.5) is obtained by setting ε2=0.
(b) In arbitrary coordinates the reduction corresponding to the passage from system (2.1) to the completely reduced system (2.6) may be obtained as follows:
Given ε1>0, determine the projection matrix
˜Q2(x,ε1):=In−(P1(x,0),ε1P2(x,0))A2(x,ε1,0)−1(Dxμ1(x,0)Dxμ2(x,0)). |
Then ˜Q2(x,ε1) extends smoothly to a matrix valued function Q2(x) at ε1=0. The equation corresponding to the completely reduced system in arbitrary coordinates is given by
˙x=Q2(x)g(1,1)(x) |
on the local invariant manifold defined by μ1(x,0)=μ2(x,0)=0.
Proof. Part (a) is a direct application of [11], Theorem 1. For part (b) this theorem is also applicable, but there is a technical problem involving ˜Q2 as ε1→0, since A2(x,0) is non-invertible. To resolve this difficulty, recall that ˜Q2(x,ε1) is the projection map onto the kernel of
Dxg(0,0)(x,0)+ε1Dxg(1,0)(x,ε1) |
along the image, for x∈˜M2 (see [11], Remark 1). With the conditions given in Lemma 1 (c) the image is equal to the column space of (P1,ε1P2), which in turn equals the column space W1 of (P1,P2). The latter matrix has full rank at ε1=0, and its entries depend smoothly on ε1 and x. Moreover the kernel is equal to the kernel of (Dxμ1,Dxμ2)tr, and we may assume w.l.o.g. that
(Dxμ1Dxμ2)=(M1M2) |
with invertible M1, whence the kernel is equal to the column space W2 of the matrix
(−M−11M2I) |
with entries depending smoothly on ε1. Thus there remains to verify that the matrix of the projection onto W2 along W1 depends smoothly on ε1. For the sake of completeness we give a proof of this fact in Lemma 4, Appendix.
We note that the reduction also works, including convergence properties, under the weaker assumption corresponding to (H) rather than (AH) for A1 and A2 in Lemma 1.
Remark 3. While Proposition 2 provides the reduced equations, one also needs initial values for these, which may be obtained from an initial value y of system 3.2 with the help of the first integrals noted in the proof of Lemma 1(a); see [11], Proposition 2:
● Assuming that y is sufficiently close to ˜M1, the corresponding initial value (up to an error of order ε1+ε2) for the auxiliary system and for the intermediate reduced system is the (locally unique) intersection of ˜M1 and the level sets of n−n1 independent first integrals of ˙x=g(0,0)(x,0) which contain y.
● Assuming that y is sufficiently close to ˜M2, the corresponding initial value (up to an error of order ε1+ε2) for the auxiliary system and for the intermediate reduced system is the (locally unique) intersection of ˜M2 and the level sets of n−n1−n2 independent common first integrals of ˙x=g(0,0)(x,0) and ˙x=g(1,0)(x,0,0) which contain y. (A direct application of Proposition 2 in [11] would lead to simultaneous first integrals of ˙x=g(0,0)(x,0)+ε1g(1,0)(x,ε1,0) for all ε1. This is equivalent to the condition stated.)
To illustrate the procedure with an example, we recall the competitive inhibition network with substrate S, enzyme E, inhibitor I and two complexes C1,C2; see for instance Keener and Sneyd [14]. The reaction scheme is given by
E+Sk1⇌k−1C1k2⇀E+P,E+Ik3⇌k−3C2 |
which leads (with the usual assumptions of mass action kinetics, spatial homogeneity and constant thermodynamical parameters) to the differential equation system
˙s=k−1c1−k1s(e0−c1−c2)˙c1=k1s(e0−c1−c2)−(k−1+k2)c1˙c2=k3(e0−c1−c2)(i0−c2)−k−3c2 | (3.3) |
for the concentrations. (The original system is five dimensional; the two linear first integrals e+c1+c2 and i+c2 yield reduction to dimension three.) There is a familiar two time scale reduction for this system, viz. the quasi-steady state (QSS) reduction for both complexes; see Eqs (1.51–1.52) in Keener and Sneyd [14]. As was shown in [15], section 3.2 this QSS reduction is (up to irrelevant higher order terms) identical to the singular perturbation reduction with small parameter e0=εe∗0, hence Tikhonov guarantees its validity. (It is not generally true that QSS reductions amount to Tikhonov-Fenichel reductions. A sufficient condition is stated in [16], Prop. 5, and a list of possible QSS reductions for competitive inhibition - including a characterization which correspond to singular perturbation reductions - is given in section 5.2 of that paper.)
Example 1. In system (3.3) set x=(s,c1,c2)tr and assume k2=ε1ε2k∗2, k3=ε1k∗3 and k−3=ε1k∗−3. (Colloquially speaking, binding to the inhibitor and degradation from the inhibitor complex are slow, while degradation from the substrate complex to enzyme and product is very slow. The related two time scale reduction to a one dimensional equation corresponds to slow formation of enzyme and complex.) This is of the type (3.2), with
g(0,0)(x)=(k−1c1−k1s(e0−c1−c2)k1s(e0−c1−c2)−k−1c10),g(1,0)(x,ε1)=(00k∗3(e0−c1−c2)(i0−c2)−k∗−3c2),g(1,1)(x,ε1,ε2)=(0−k∗2c10). |
Moreover ˜M2 is contained in the common zero set of
μ1=k−1c1−k1s(e0−c1−c2) and μ2=k∗3(e0−c1−c2)(i0−c2)−k∗−3c2, |
˜M1 is contained in the zero set of μ1, and we have
P1(x,ε2)=(1−10),P2(x,ε1)=(001). |
We determine the auxiliary system and the intermediate reduced system. With
Dμ1=(−k1(e0−c1−c2,k1s+k−1,k1s) |
one has
Dμ1P1=−k1(e0−c1−c2−(k1s+k−1)=:−ν1 |
and furthermore
Q1=I3+1ν1(∗k1s+k−1k1s∗−(k1s+k−1)−k1s000)=1ν1(∗k1s+k−1k1s∗k1(e0−c1−c2)−k1s00ν1). |
Application to
g(1,0)+ε2g(1,1)=μ2(001)−ε2k∗2c1(001) |
yields the auxiliary system (in time scale ε1t) on ˜M1:
(˙s˙c1˙c2)=μ2ν1(k1s−k1sν1)−ε2k∗2c1ν1(k1s+k−1k1(e0−c1−c2)0). |
Setting ε2=0 one obtains the intermediate reduced system.
When the initial values for system (3.3) are given by (s0,c1,0,c2,0), to obtain the approximate initial values (s∗0,c∗1,0,c∗2,0) on ~M1 one uses (according to Remark 3) the two first integrals s+c1 and c2 of g(0,0) and the defining equation for ~M1, thus the system
s+c1=s0+c1,0c2=c2,0k−1c1−k1s(e0−c1−c2)=0 |
which leads to quadratic equations for s and c1.
To find the fully reduced system one first computes
Dμ2=(0,−k∗3(i0−c2),−k∗3(e0+i0−c1−2c2)−k∗−3) |
and
A2=(Dμ1Dμ2)(P1,ε1P2)=(−k1(e0−c1−c2)−k1s−k−1ε1⋅k1sk∗3(i0−c2)−ε1⋅(k∗3(e0+i0−c1−2c2)+k∗−3)). |
The computation of the projection matrix is straightforward (although a software system is helpful) but the output is sizeable. We just record the fully reduced system (in time scale ε1ε2t). It is given by
˙x=1ν2⋅(ξ1ξ2ξ3) |
with
ν2=sc1k1k∗3+sc2k1k∗3−se0k1k∗3−c21k1k∗3−3c1c2k1k∗3+2c1e0k1k∗3+c1i0k1k∗3−2c22k1k∗3+3c2e0k1k∗3+c2i0k1k∗3−e20k1k∗3−e0i0k1k∗3−sk∗−3k1+c1k∗−3k1+c1k−1k∗3+c2k∗−3k1+2c2k−1k∗3−e0k∗−3k1−e0k−1k∗3−i0k−1k∗3−k∗−3k−1 |
and
ξ1=k∗2(se0k∗−3k1+c1e0k−1k∗3−c1i0k−1k∗3+c22k−1k∗3−c2e0k−1k∗3−c2i0k−1k∗3+e0i0k−1k∗3−c2k∗−3k−1),ξ2=k1k∗2k∗3(c31(k∗3)2−2c21e0(k∗3)2+2c21i0(k∗3)2+c1e20(k∗3)2−2c1e0i0(k∗3)2+c1i20(k∗3)2+c32(k∗3)2−c22e0(k∗3)2−2c22i0(k∗3)2+2c2e0i0(k∗3)2+c2i20(k∗3)2−e0i20(k∗3)2−c21k∗−3k∗3+c1e0k∗−3k∗3+2c1i0k∗−3k∗3−3c22k∗−3k∗3+2c2e0k∗−3k∗3+3c2i0k∗−3k∗3−2e0i0k∗−3k∗3+2c2(k∗−3)2),ξ3=−k∗−3k1k∗2k∗3(c1i0k∗3−c22k∗3+c2e0k∗3+c2i0k∗3−e0i0k∗3+c2k∗−3) |
restricted to the invariant curve ˜M2.
Finally, given initial values (s0,c1,0,c2,0) for system (3.3), approximate initial values for the fully reduced system may be determined by solving the algebraic equations s+c1=s0+c1,0, μ1=0 and μ2=0.
Typically in applications one starts with a general parameter dependent system, rather than a system of type (2.1) or (3.1) with pre-assigned "small parameters". Therefore the first task is to determine critical parameter values, for which small perturbations lead to singular perturbation scenarios. Thus we consider Tikhonov-Fenichel parameter values, as defined in [12] for two time scales, and extend the notion to the three time scale setting.
Tikhonov-Fenichel parameter values (TFPV) were introduced in [12] for polynomial (or rational) parameter dependent systems
˙x=h(x,π),x∈Rn,π∈Π⊆Rm. | (4.1) |
A TFPV ˆπ is characterized by the property that small perturbations π=ˆπ+ερ+⋯ along a smooth curve in parameter space Π give rise to a singular perturbation reduction for
˙x=h(x,ˆπ+ερ+⋯)=h(x,ˆπ)+εDπh(x,ˆπ)ρ+⋯ |
with locally exponentially attracting critical manifold. (The definition extends easily to smooth systems but the algorithmic approach relies on the stronger assumption.) There exists an intrinsic characterization of TFPV's, see [12] Lemmas 1 and 2, for which the characteristic polynomial
χ(τ,x,π)=τn+σn−1(x,π)τn−1+⋯+σ1(x,π)τ+σ0(x,π) | (4.2) |
of the Jacobian Dxh(x,π) is relevant. We recall:
Lemma 2. Given 0<s<n, a parameter value ˆπ is a TFPV with locally exponentially attracting critical manifold Zs (depending on ˆπ) of dimension s, and x0∈Zs, if and only if the following hold:
● h(x0,ˆπ)=0.
● The characteristic polynomial χ(τ,x,π) from (4.2)) satisfies
(i) σ0(x0,ˆπ)=⋯=σs−1(x0,ˆπ)=0;
(ii) all roots of χ(τ,x0,ˆπ)/τs have negative real parts.
● The system ˙x=h(x,ˆπ) admits s independent local analytic first integrals at x0.
All the conditions in the lemma can be represented by polynomial equations and inequalities. The condition on the roots of χ(τ,x0,ˆπ)/τs is characterized by inequalities: There exist n−s Hurwitz determinants (see e.g. Gantmacher [17], Ch. V, §6, Thm. 4 ff.) which must attain values >0. Moreover, the existence requirement for s independent first integrals leads to a series of polynomial equations via degree by degree evaluation of Taylor expansions. While there are notable exceptions for chemical reaction networks, with possible first integrals from stoichiometry, this condition is the most problematic to verify in general. One may proceed as follows: For every d>0 there is an induced action of Dxh(x,π) on the space S1+⋯+Sd of polynomials in n variables with zero constant term and of degree ≤d. Extending condition (ⅰ), the characteristic polynomial of this action (which coincides with (4.2) for d=1) must have vanishing coefficients for all sufficiently small powers of the indeterminate. No further inequalities appear, due to the structure of the eigenvalues for this action, and thanks to Hilbert's Basissatz, finitely many of these equations suffice. With increasing dimension, one has to deal with feasibility problems. In applications the common procedure is step by step, looking for simplifications at every stage. A full account is given in [12].
For the remainder of this section we assume that Π⊆Rm+ is a semi-algebraic set, and that system (4.1) admits the positively invariant subset Rn+. Then, as was shown in [12], the Tikhonov-Fenichel parameter values for dimension s, 1≤s<n form a semi-algebraic subset Πs⊆Rm. We will denote the Zariski closure of Πs by Ws. Thus the elements of Ws satisfy all defining equations for Πs but not necessarily the defining inequalities.
Generalizing the approach to TFPV in [12]), and taking into account the special form of (3.1), it seems reasonable to consider surfaces in parameter space. Thus consider a smooth surface of the special form
γ(ε1,ε2)=ˆπ+ε1(ρ1(ε1)+ε2ρ2(ε1,ε2)) |
defined in some nighborhood of (0,0). Substitute γ(ε1,ε2) for π in (4.1) to get
h(x,γ(ε1,ε2))=h(x,ˆπ)⏟=:g(0,0)+h(x,γ(ε1,0))−h(x,ˆπ)⏟=:ε1⋅g(1,0)+(h(x,γ(ε1,ε2))−h(x,ˆπ))−(h(x,γ(ε1,0))−h(x,ˆπ))⏟=:ε1ε2g(1,1) | (4.3) |
with the g(i,j) smooth by Hadamard's lemma. In order to obtain a system (3.1) that also satisfies the conditions (ⅰ) and (ⅱ) preceding Lemma 1, the following is necessary: There exist s>0 and k>0 such that ˆπ∈Πs+k, and ˆπ+ε1⋅ρ1(ε1)∈Πs for all sufficiently small ε1>0. (Note that ε2 plays no role in these conditions.) This observation gives rise to:
Definition 2. Given system (4.1) and s,k>0 with s+k<n, let δ>0 and let
β:(−δ,δ)⟶Π, ε1↦β(ε1) |
be a smooth curve such that
(i) β(ε1)∈Πs for all ε1>0,
(ii) ˆπ:=β(0)∈Πs+k.
Then we call ˆπ a Tikhonov-parameter value (for dimension s+k) nested in ¯Πs.
We note some properties of nested TFPV.
Proposition 3. (a) Any TFPV ˆπ∈Πs+k which is nested in ¯Πs lies in the boundary of Πs relative to its Zariski closure Ws.
(b) Let β as in Definition 2, and for ε1>0 consider the decomposition
h(x,β(ε1))=P∗(x,ε1)μ∗(x,ε1) |
according to [11], Theorem 1. Then
detDμ∗(x,0)P∗(x,0))=0 |
on the critical manifold.
Proof. Part (a) is a direct consequence of the definition. As for part (b), at ε1=0, with ˆπ∈Πs+k and x0∈Zs+k (using notation from Lemma 2), the coefficient σs(x0,ˆπ) of the characteristic polynomial (4.2) of
Dxh(x0,ˆπ)=P∗(x0,0)Dμ∗(x0,0) |
must vanish. This is equivalent to non-invertibility of Dμ∗(x,0)P∗(x,0)); see e.g. [11], Remark 4.
Remark 4. Proposition 3 opens a starting point for the computation of nested TFPV: Start with system (4.1) corresponding to "generic" parameter values in Πs, i.e. parameter values in the intersection of Πs with an irreducible component of the Zariski closure Ws. In order to find nested parameters for higher dimension one only needs to look at the boundary of ¯Πs, and one can use part (b) in order to obtain necessary conditions. Practically this may be realized by determining the decomposition P⋅μ for generic π∈Πs and then looking at zeros of Dμ⋅P, with parameters in the boundary. (The boundary may also contain further parameter values in Πs.)
For chemical reaction networks (CRN) the parameter region is usually given by Π=Rm+, thus
π=(π1⋮πm)∈Rm+, |
and for many such systems and given s, the irreducible components of Ws are just determined by the vanishing of certain of the πi; see e.g. [12,16,18]. (The underlying reason for this fact is the subject of forthcoming work.) Thus we have, for π in a given irreducible component:
(ⅰ) Upon relabelling, there is an ℓ, 0<ℓ<m such that πi=0 for all i∈{ℓ+1,⋯m};
(ⅱ) the remaining parameters are nonnegative.
In other words, the intersection of Πs with the given irreducible component of Ws corresponds to some subset of ¯Rℓ+, with boundary ¯Rℓ+∖Rℓ+. This leads to an obvious case-by-case analysis. Note that boundary points may or may not be contained in Πs, but there is no loss in starting with "generic" parameter values in the interior Rℓ>0. We look at a particular example.
Example 2. We again consider competitive inhibition; see Eq (3.3). Here the parameters are of the form
π=(e0k1k−1k2i0k3k−3)∈R7+. |
From [12], Proposition 8 (see also Goeke's dissertation [18], section 9.3) we have the necessary condition e0k1k2k−3=0 for a TFPV in Π1, with each of the four cases (e.g. e0=0 and all other parameter values ≥0) yielding a singular perturbation reduction with attracting one dimensional critical manifold. Hence W1 has four irreducible components. For the cases with some ki=0, which amount to a distinction of slow and fast reactions, the two time scale reductions may be obtained using a computational procedure outlined in Heinrich and Schauer [19]; see also Stiefenhofer [22]. In order to find nested TFPV's for dimension 2 we perform a case-by-case investigation. We only consider one case here; see Section 5 for the remaining ones.
For the case k2=0 the system is given by
˙s=k−1c1−k1se˙c1=k1se−k−1c1˙c2=k3ei−k−3c2, |
where we have used the abbreviations e=e0−c1−c2 and i=i0−c2; note that e≥0 and i≥0 by design of (3.3). By Remark 4, nested TFPV's for dimension two and corresponding points in the critical manifold necesarily satisfy det(Dμ⋅P)=0, with
μ=(k−1c1−k1sek3ei−k−3c2), P=(10−1001), |
Dμ=(−k1ek1s+k−1k1s0−k3i−(k3i+k3e+k−3)). |
Proceeding according to Remark 4, we determine the vanishing set of
det(−(k1e+k1s+k−1)k1sk3i−(k3i+k3e+k−3))=k1k3ie+k1k3e2+k1k−3e+k1k3se+k1k−3s+k−1k3i+k−1k3e+k−1k−3. |
Since all the variables and parameters are nonnegative, this sum equals zero if and only if every summand vanishes. In particular, k−1⋅k−3 has to vanish for any nested TFPV. We look at the two ensuing cases.
(ⅰ) k−1=0: Then the remaining conditions are
k1k3ie=k1k3e2=k1k−3e=k1k3se=k1k−3s=0. |
If k1=0 or k3=k−3=0 we obtain a two dimensional variety of stationary points. Checking the attractivity conditions (HA), one finds that these cases yield nested TFPV. If e=s=0 holds then we get e0−c1−c2=0=s which corresponds to a one dimensional variety. In case e=k−3=0 we get c1=0 while c2 and s are arbitrary, thus we have a two dimensional (attracting) variety of stationary points.
(ⅱ) k−3=0: In this case there remains
k1k3ie=k1k3e2=k1k3se=k−1k3i=k−1k3e=0. |
In view of case (ⅰ) we only have to check k3=0 or e=i=0. In both cases we get a variety of dimension two.
The case k3=k−3=0 leads to system (3.3) with k3=ε1k∗3, k−3=ε1k∗−3 and k2=ε1ε2k∗2, the reduction of which was discussed in Example 1.
In this section we continue the discussion of the competitive inhibitor network, to some extent, and furthermore present a fairly complete investigation of a cooperative system with two complexes, following the strategy outlined in Remark 4. The mathematical analysis yields a rather large number of possible reductions, with parameter conditions that allow an interpretation in a biochemical context. It seems that most of these reductions have not been discussed before in the literature. Here one should note that the principal interest in biochemistry (as in the monograph by Keener and Sneyd) lies in identifying and discussing reaction velocities for given parameter constellations; thus the focus is different from the one in the present paper. (Noel et al. [4], Radulescu et al. [5], Samal et al. [6,7] investigate questions related to those in the present paper.) Missing from a complete analysis are some cases concerned with boundary points in Π1⊆W1 which themselves belong to Π1, as well as certain degenerate cases for Π2. Moreover we will not generally record routine calculations to verify conditions such as (HA), and for ease of notation we will frequently use the term "critical manifold" for the Zariski closure of this object, without mentioning the inequalities to be satisfied.
We continue to investigate the competitive inhibition network; see Eq (3.3), Examples 1 and 2. The analysis of TFPV which was started in Example 2 will be finished here. For Π1 there are three remaining cases, viz. e0=0, k1=0 and k−3=0.
(a) For e0=0, the system is given by
˙s=(k1s+k−1)c1+k1sc2˙c1=−(k1s+k−1+k2)c1−k1sc2˙c2=−k3ic1−(k3i+k−3)c2 | (5.1) |
We only consider the generic case for Π1, thus all the remaining parameters are >0. Then the (only possible) decomposition P⋅μ for the right hand side is given by
μ=(c1c2),P=(k1s+k−1k1s−(k1s+k−1+k2)−k1s−k3i−(k3i+k−3)). |
For nested TFPV, a simple computation yields the necessary condition
0=det(Dμ⋅P)=ik−1k3+ik2k3+sk−3k1+k−3k−1+k−3k2, |
with all terms positive; thus every summand must vanish, and in particular
det(Dμ⋅P)=0 ⇒ (k−1+k2)k−3=0 ⇒ k−1=k2=0 or k−3=0. |
In case k−1=k2=0 system (5.1) admits a two dimensional variety of stationary points, with k1s≠0 only if c1+c2=0. The intersection of this variety with the positive orthant is only one dimensional, thus we do not obtain a two dimensional critical manifold. The cases with k1s=0 translate to k1=εk∗1 for the system with small parameters. (Otherwise the critical manifold would be given by s=0, which does not contain the line given by c1=c2=0.) Moreover we have e0=ε1ε2e∗0, hence every term k1e0s is of the form ε21ε2⋅(⋯), and the completely reduced system is necessarily trivial. Likewise, the case k−3=0 in system (5.1) leads to k1s=0.
To summarize, for the case e0=0, which leads to the familiar QSS reduction in the two time scale setting, there are no interesting reductions for the three time scale scenario: No parameter value leads to an intermediate reduction to dimension two with nontrivial complete reduction.
(b) Next we consider the system with k1=0, i.e.
˙s=k−1c1˙c1=−(k−1+k2)c1˙c2=k3ei−k−3c2. |
Because of Example 2 we may assume that k2≠0, which yields c1=0 for stationary points.
In case k−3=k3=0 we indeed have a two dimensional critical manifold. Turning to small parameters we have k1=ε1ε2k∗1, k3=ε1k∗3 and k−3=ε1k∗−3, and (3.3) becomes
(˙s˙c1˙c2)=(k−1c1−(k−1+k2)c10)+ε1(00k∗3ei−k∗−3c2)+ε1ε2k∗1es(−110) | (5.2) |
We compute the reductions for this case. For the auxiliary system (on the critical variety defined by c1=0) we obtain the decomposition
(k−1c1−(k−1+k2)c10)=(k−1−(k−1+k2)0)⏟P1⋅c1⏟μ1, |
and a straightforward computation yields the projection matrix
Q1=(1k−1/(k−1+k2)0000001) |
and the auxiliary system
(˙s˙c1˙c2)=(00k∗3ei−k−3∗c2)+ε2k∗1k2esk−1+k2(−100) |
on the variety defined by c1=0. (This is a QSS reduction for complex C1, according to [16], Proposition 5.) The intermediate reduced system is obtained setting ε2=0.
Turning to the complete reduction, the decomposition of the "fast part" of (5.2) is given by
(k−10−(k−1+k2)00ε1)(c1μ2),with μ2=k∗3ei−k∗−3c2. |
One obtains
A2=(−(k−1+k2)0(k−1+k2)k3i−ε1(k3(e+i)+k−3c2)) |
and may continue as prescribed by Proposition 2. There are shortcuts, though: First note that the critical manifold is given by c1=0, and c2 constant and equal to the smaller solution ˜c2 of the quadratic equation
0=μ2(0,c2)=k∗3(e0−c2)(i0−c2)−k∗−3c2. |
The completely reduced system will automatically yield ˙c1=0 and ˙c2=0, hence only the first row of the projection matrix needs to be computed. As the final result of the reduction procedure we get the equation
˙s=−k∗1k2k−1+k2(e0−˜c2)s, |
with the dot denoting differentiation with respect to ε1ε2t.
(c) Finally, we deal with the case k−3=0, which does not automatically yield a one dimensional variety of stationary points. System (3.3) becomes
˙s=k−1c1−k1se˙c1=k1se−(k−1+k2)c1˙c2=k3ei. |
We may assume that k1≠0 and k2≠0, otherwise one would arrive at (non-generic) subcases of previously discussed systems. From this we obtain c1=0 and es=0 as necessary conditions. Now e=0 and nonnegativity of varaibles imply e0=0; a previously discussed case, therefore every stationary point satisfies s=0. If k3≠0 then i=0 forces c2=i0; the corresponding parameter values are not in Π1. So the only case remaining is k3=k−3=0 (very slow binding to the inhibitor, very slow degradation of the inhibitor complex), with system
˙s=k−1c1−k1se˙c1=k1se−(k−1+k2)c1˙c2=0. |
To obtain a two dimensional variety of stationary points one has to check the boundary of Π1 for nested TFPV, which splits into four cases. We only discuss the case k1=0 here, thus (3.3) with small parameters becomes
(˙s˙c1˙c2)=(k−1c1−(k−1+k2)c10)+ε1k∗1es(−110)+ε1ε2(00k∗3ei−k∗−3c2) | (5.3) |
The computation of the auxiliary system runs similar to the reduction of (5.2) and yields
(˙s˙c1˙c2)=k∗1k2esk−1+k2(−100)+ε2(00k∗3ei−k−3∗c2), |
on the invariant variety given by c1=0. Finally, the completely reduced system lives on the variety defined by c1=s=0 (a coordinate subspace), and therefore by [16], Proposition 5 the reduced system may be directly obtained via "classical" QSS reduction; yielding
˙c2=k∗3(e0−c2)(i0−c2)−k∗−3c2. |
In this subsection we study the standard cooperative system involving substrate S, two complexes C1,C2, enzyme E and product P. The reaction scheme
S+Ek1⇌k−1C1k2⇀E+PS+C1k3⇌k−3C2k4⇀C1+P |
yields, with the usual assumptions and stoichiometry, the differential equation
˙s=−k1e0s+(k−1+k1s−k3s)c1+(k1s+k−3)c2˙c1=k1e0s−(k−1+k2+k1s+k3s)c1+(k−3+k4−k1s)c2˙c2=k3sc1−(k−3+k4)c2 | (5.4) |
where all appearing constants are non-negative; see Keener and Sneyd [14], equations (1.62–64), where the QSS reduction for complexes C1 and C2 is discussed. A direct proof of validity via singular perturbation theory - for two time scales - is given in 21ray [21], Section 6.5 when e0=εe∗0; alternatively [16], Proposition 5 is applicable. In [16], subsection 5.4 a list of parameter conditions for various types of QSS is given, and their relation to Tikhonov-Fenichel reductions is discussed. Keener and Sneyd also mention the (partial) equilibrium approximation, by which one generally determines the product forming velocity given equilibrium in a certain subnetwork. In the present context [14], section 1.7, Exercise 10 this means that k2=εk∗2 and k4=εk∗4 are small. The discussion in 5.2.4 shows, among other things, that these conditions are not compatible with singular perturbation reductions. (One needs to add further "small parameters", and the resulting reduction formulas are more complicated.) Cooperative systems with three complexes are mentioned in [14], and the singular perturbation reduction in the case of arbitrary many complexes, with small e0=εe∗0, is derived in [16], section 5.3. Now we turn to the three time scale setting.
According to Goeke [18], Kap. 9.4, necessary conditions for TFPV are given by
e0k1k2(k−3+k4)=0. |
When we substitute k1=0 in Eq (5.4) we obtain
˙s=(k−1−k3s)c1+k−3c2˙c1=−(k−1+k2+k3s)c1+(k−3+k4)c2˙c2=k3sc1−(k−3+k4)c2. |
Hence considering the generic case (all remaining parameters >0) we obtain an irreducible component of W1 given by k1=0, and the critical manifold is given by c1=c2=0. We get a decomposition with
P=(−sk3+k−1k−3−(sk3+k−1+k2)k−3+k4sk3−(k−3+k4)),μ=(c1c2), |
and necessary conditions for nested TFPV from
0=detDμ⋅P=(k−1+k2)(k−3+k4)⇒k−1=k2=0 or k−3=k4=0. |
Thie first set of conditions does not, by itself, yield a two dimensional critical manifold, and we will not pursue it further here. The second set, i.e. k−3=k4=0, yields the two dimensional variety given by c1=0.
Considering this setting, we introduce the small parameters in our original system by substituting k1=ε1ε2k∗1, k−3=ε1k∗−3, k4=ε1k∗4. Ordering the parameters as e0,k1,k−1,k2,,k3,k−3,k4, we thus consider the surface in parameter space given by
γ(ε1,ε2)=(e00k−1k2k300)+ε1⋅((00000k∗−3k∗4)+ε2(0k∗100000)), |
and with x=(s,c1,c2)tr we get
h(x,ε1,ε2)=g(0,0)(x)+ε1⋅(g(1,0)(x,ε1)+ε2⋅g(1,1)(x,ε1,ε2)) | (5.5) |
with
g(0,0)(x)=((−sk3+k−1)c1−(sk3+k−1+k2)c1k3sc1)g(1,0)(x,ε1)=(k∗−3c2(k∗−3+k∗4)c2−(k∗−3+k∗4)c2)g(1,1)(x,ε1,ε2)=(sk∗1c1+sk∗1c2−k∗1e0s−(sk∗1c1+sk∗1c2−k∗1e0s)0). |
For this system we compute the complete reduction on c1=c2=0 and the intermediate reduction on c1=0. In order to compute the completely reduced system, a factorization of g(0,0)+ε1g(1,0) is given by
(P1,ε1P2)⋅(μ1μ2) |
with μ1=c1, μ2=c2, and
P1=(−sk3+k−1−(sk3+k−1+k2)k3s),P2=(k∗−3k∗−3+k∗4−(k∗−3+k∗4)). |
The projection matrix is
Q2=(1−−sk3k∗4−k∗−3k−1−k−1k∗4k∗−3k−1+k2k∗−3+k−1k∗4+k2k∗4−−sk3k∗4−2k∗−3k−1−k2k∗−3−k−1k∗4k∗−3k−1+k2k∗−3+k−1k∗4+k2k∗4000000), |
and the fully reduced system in very slow time on the invariant manifold c1=c2=0 is given by the equation
˙s=−k3k∗4s+k∗−3k2+k∗4k2k∗−3k−1+k2k∗−3+k−1k∗4+k2k∗4⋅k∗1e0s. |
Similarly one computes the intermediate system on the two dimensional variety given by c1=0 from the decomposition P1⋅μ1:
(˙s˙c1˙c2)=1sk∗3+k−1+k2(−(sc2k3k∗4+2k−1c2k∗−3+k2k∗−3c2+c2k∗4k−1)0−(k−1c2k∗−3+k2k∗−3c2+c2k∗4k−1+c2k2k∗4)). |
In this context, a brief digression may be appropriate: In their discussion of reaction velocities, Keener and Sneyd [14] also derive a Hill function in Eq (1.69)) with arguments that could a priori be interpreted as involving three time scales: They consider "k1→0 and k3→∞, with constant product k1k3". Rescaling time, one might think of this as a case of very small k1, with k3 of order one. One system satisfying these conditions is (5.5), but the completely reduced equation manifestly does not follow Hill kinetics. One might also consider the case when k3 is of order one, k1 is very small and all the other ki are small. This corresponds to a degenerate (but a priori legitimate) subcase of "k−3=k4=0" discussed above, with the surface in parameter space now given by
γ(ε1,ε2)=(e0000k300)+ε1⋅((00k∗−1k∗20k∗−3k∗4)+ε2(0k∗100000)). |
A straightforward analysis shows, however, that the resulting system does not satisfy the rank (or eigenvalue) conditions from Lemma 1(c), hence Tikhonov-Fenichel reduction is not applicable. To summarize: For the given cooperative network, a reaction velocity of Hill type cannot be obtained from time scale arguments.
From e0=0 one also obtains a component of W1, and system (5.4) specializes to
˙s=(k−1+k1s−k3s)c1+(k1s+k−3)c2˙c1=−(k−1+k2+k1s+k3s)c1+(k−3+k4−k1s)c2˙c2=k3sc1−(k−3+k4)c2. |
The right hand side has a factorization P⋅μ with
μ=(c1c2), P=(sk1−sk3+k−1sk1+k−3−sk1−sk3−k−1−k2−sk1+k−3+k4sk3−k−3−k4), |
and in order to obtain nested TFPV we examine all variable-parameter configurations that satisfy
0=det(Dμ⋅P)=sk1⋅(sk3+k−3+k4)+(k−1+k2)⋅(k−3+k4). |
The plane given by s=0 is not a viable candidate for a two dimensional critical manifold since it does not contain the line c1=c2=0. This leaves the cases k1=k−1=k2=0, k1=k−3=k4=0 and k3=k−3=k4=0.
The first of these yields a two dimensional variety (defined by k3sc1−k−3c2=0) only under the additional condition k4=0. The second case, whenever k1≠0, yields a variety whose intersection with the positive orthant has dimension one, hence is of no relevance. For the third case we obtain a two dimensional variety only if k1=0 or k2=0.
With the exception of this very last case, the completely reduced system will always be trivial, due to k1=ε1k∗1 and e0=ε1ε2e∗0, which implies k1e0=O(ε21ε2). We consider one special case, viz. the intermediate reduction coresponding to the nested TFPV with k1=k3=k−3=k4=0; here c1=0 defines the two dimensional critical manifold. Considering
γ(ε1,ε2)=(00k−1k2000)+ε1⋅((0k∗100k∗3k∗−3k∗4)+ε2(e∗0000000)), |
we compute:
g(0,0)=(c1k−1−(k−1+k2)c10)g(1,0)=((sk∗1−sk∗3)c1+(sk∗1+k∗−3)c2−(sk∗1+sk∗3)c1+(−sk∗1+k∗−3+k∗4)c2k∗3sc1−(k∗−3+k∗4)c2)g(1,1)=(−ε1k∗1e∗0sε1k∗1e∗0s0). |
The intermediate reduced system on the invariant variety c1=0 is then:
(˙s˙c1˙c2)=((sc2k∗1k2+2c2k∗−3k−1+c2k∗−3k2+c2k−1k∗4)(k−1+k2)0−(k∗−3+k∗4)c2). |
These conditions define a component of W1, and generically the critical manifold is given by s=c1=0. System (5.4) is given by
˙s=−k1e0s+(k−1+k1s−k3s)c1+k1sc2˙c1=k1e0s−(k−1+k2+k1s+k3s)c1−k1sc2˙c2=k3sc1 |
and the product decomposition (which we do not write down here) yields
0=detDμ⋅P=k1(e0−c2)(k2+2k3s). |
as necessary conditions for nested TFPV. One possible case is k1=0 with critical manifold c1=0. The remaining cases are:
(ⅰ) k2=k−1=0 with variety s=0;
(ⅱ) k2=k3=0 with variety k1(e0−c1−c2)s−k−1c1=0.
Note that the condition e0−c2=0 does not yield a two dimensional critical variety.
In this situation system (5.4) simplifies to
˙s=−k1e0s+(k−1+k1s−k3s)c1+(k1s+k−3)c2˙c1=k1e0s−(k−1+k1s+k3s)c1+(k−3+k4−k1s)c2˙c2=k3sc1−(k−3+k4)c2. |
The condition k2=0 by itself does not define an irreducible component of W1; in other words it does not guarantee the existence of a one dimensional variety of stationary points. Therefore we first investigate sufficient conditions, using the observation ˙s+˙c1+2˙c2=−k4c2.
(a) For k4≠0 this observation implies that any stationary point satisfies c2=0, and the remaining condition is k3sc1=0.
(ⅰ) In case k3≠0 we have either s=0, with the variety of stationary points given by s=c2=0; in turn this yields the parameter configuration
k−1=k2=0. |
(ⅱ) Alternatively we have c1=0, the variety is given by c1=c2=0, and one must have k1e0=0. We obtain the possible parameter configurations
k1=k2=0 or e0=k2=0. |
(b) In case k4=0 the remaining system is
˙s=−k1es+k−1c1−k3sc1+k−3c2˙c1=k1es−k−1c1+k3sc1+k−3c2˙c2=k3sc1−k−3c2. |
Adding the first two equations for stationary points shows that k−3c2=0, and combining this with the third equation yields k3sc1=0; in addition one has k1es−k−1c1=0. Thus there are further conditions for the existence of a one dimensional critical variety.
(ⅰ) Given that k3≠0 and k−3≠0, the variety is given either by c2=s=0, which yields the parameter conditions
k2=k−1=k4=0; |
or the variety is given by c1=c2=0, with parameter conditions
k2=k4=k1=0 or k2=k4=e0=0; |
all of these are special cases from (a).
(ⅱ) In case k3=0 we obtain the one dimensional variety given by c2=0 and k1(e0−c1)s−k−1c1=0; thus we have the parameter condition
k2=k3=k4=0 |
which defines a component of W1.
(ⅲ) In case k−3=0 one obtains the variety s=c1=0, with parameter conditions
k2=k4=k−3=0. |
For all these parameters the next task is to discuss conditions for embedded TFPV. We will only do so for two cases.
1. In the case k−1=k2=0 one has a decomposition
(−k1e−k3c1k−3k1e−k3c1k−3+k4k3c1−(k−3+k4))⋅(sc2) |
which yields
detDμ⋅P=k1(k−3+k4)e+k3k4c1. |
We take a closer look at the case k1=k3=0, with critical variety c2=0. The surface in parameter space
γ(ε1,ε2)=(e00000k−3k4)+ε1⋅((0k∗100k∗300)+ε2(00k∗−1k∗2000)) |
yields
g(0,0)=(c2k−3(k−3+k4)c2−(k−3+k4)c2)g(1,0)=(−k∗1e0s+(sk∗1−sk∗3)c1+sk∗1c2k∗1e0s−(k∗1s+k∗3s)c1−k∗1sc2k∗3sc1)g(1,1)=(k∗−1c1−(k∗−1+k∗2)c10). |
The intermediate reduced system is as follows:
˙s=sc1k−3k∗1+k4k∗1c1s−sc1k∗3k4−k∗1e0sk−3−k∗1e0sk4k−3+k4˙c1=−sc1k∗1+k∗1e0s˙c2=0, |
and the completely reduced system (on s=c2=0) is given by:
˙c1=−c1(c1k−3k∗1k∗2−c1k∗−1k∗3k4+c1k∗1k∗2k4−c1k∗2k∗3k4−e0k−3k∗1k∗2−e0k∗1k∗2k4)c1k−3k∗1+c1k∗1k4−c1k∗3k4−e0k∗1k−3−e0k∗1k4 |
2. In the case k2=k3=k4=0 we have the one dimensional critical manifold
k1s⋅(e0−c1)−k−1c1=0, c2=0 |
and the right hand side of the system at k2=k3=k4=0 can be decomposed into P⋅μ, with
P=(1sk1+k−3−1−sk1+k−30−k−3)μ=(k1s⋅(e0−c1)−k−1c1c2). |
This yields
det(Dμ⋅P)=(k1(e0−c1)+k1s+k−4)⋅k−3. |
We investigate the case k−3=0. Additionally setting k−3=0 we obtain the two dimensional critical manifold defined by
μ2:=−k1e0s+(k−1+k1s)c1+k1sc2=0. |
Following the usual procedure we consider the surface
γ(ε1,ε2)=(e0k1k−10000)+ε1⋅((00000k∗−30)+ε2(000k∗2k∗30k∗4)) |
in parameter space, and thus
g(0,0)=(−k1e0s+(sk1+k−1)c1+k1sc2k1e0s−(sk1+k−1)c1+−k1sc20)g(1,0)=(k∗−3c2k∗−3c2−k∗−3c2)g(1,1)=(−sk∗3c1−(sk∗3+k∗2)c1+k∗4c2sk∗3c1−k∗4c2). |
Here the intermediate reduced system is given by
˙s=1sk1+k1(e0−c1−c2)+k−1⋅(sc2k∗−3k1+2c2k∗−3k−1)˙c1=1sk1+k1(e0−c1−c2)+k−1⋅(sc2k∗−3k1−2k1c2k∗−3c1−2k1k∗−3c22+2e0k1k∗−3c2)˙c2=−c2k∗−3 |
on μ2=0, and the fully reduced system is given by
˙s=1sk1+k1(e0−c1)+k−1⋅(−se0k1k∗2)˙c1=1sk1+k1(e0−c1)+k−1⋅(k1k∗2c21−k∗2e0k1c1)˙c2=0. |
In this section we give a brief outline on extending the coordinate-free approach to more than three time scales. Thus let N≥3 and first consider a system with N−1 small parameters of the form
˙xi=(∏1≤j<iεj)⋅fi(x,ε1,…,εN−1),1≤i≤N; briefly ˙x=f(x,ε) | (6.1) |
with separated variables. By a smooth coordinate transformation this becomes
˙x=g(0,…,0)+ε1(g(1,0,…0)+ε2(g(1,1,0,…,0)+ε3(⋯))) | (6.2) |
with the very last term in the embedded brackets being εN−1g(1,…,1). Here all g(i1,…,iN−1) are functions of (x,ε1,…,εN−1). Moreover conditions (ⅰ), (ⅱ) preceding Lemma 1 generalize in an obvious manner to the vanishing sets of
g(0,…,0)g(0,…,0)+ε1g(1,0,…0)g(0,…,0)+ε1(g(1,0,…0)+ε2g(1,1,0,…,0))etc. |
and as in Proposition 2 one obtains decompositions
g(0,…,0)=P1μ1g(0,…,0)+ε1g(1,0,…0)=(P1,ε1P2)(μ1μ2)g(0,…,0)+ε1(g(1,0,…0)+ε2g(1,1,0,…,0))=(P1,ε1P2,ε1ε2P3)(μ1μ2μ3)etc. |
Likewise, one generalizes the definitions of A1,A2 and the constructions of the projection matrices Qj, the latter extending smoothly to ε1=⋯εN−1=0. This yields the various (intermediate) reductions.
Given a general parameter dependent system (4.1), nested Tikhonov-Fenichel parameter values may be found via the ansatz
γ:(ε1,…,εN−1)↦ˆπ+ε1(ρ1(x,ε1)+ε2(ρ2(x,ε1,ε2)+ε3(⋯))) |
and the ensuing decomposition of h(x,γ(ε1,…,εN−1)) analogous to the one in (4.3). Thus the problem is to find s>0, 0<k1<⋯<kN−1 with s+kN−1<n and a smooth map
β:(ε1,…,εN−2)→Π, |
defined in some neighborhood of 0, such that
β(ε1,…,εN−2)∈Πsβ(ε1,…,εN−3,0)∈Πs+k1⋮β(0,…,0)∈Πs+kN−1 |
whenever all εj>0. Rather obvious generalizations of Proposition 3 hold, and the strategy outlined in Remark 4 remains applicable.
The work of both authors has been supported by the bilateral project ANR-17-CE40-0036 and DFG-391322026 SYMBIONT. The authors thank two anonymous reviewers for helpful comments and suggestions.
Both authors declare no conflicts of interest in this paper.
For the reader's convenience we state and prove here two lemmas.
Lemma 3. Let r1,r2 be positive integers and
A∈Rr1×r1,B∈Rr1×r2,C∈Rr2×r1,D∈Rr2×r2. |
Moreover assume that all eigenvalues of A have real part <0. Then the following are equivalent.
(i) All eigenvalues of −CA−1B+D have real part <0.
(ii) There exists δ>0 such that, for every ε∈(0,δ), all eigenvalues of
(ABεCεD) |
have real part <0.
Proof. Consider the singularly perturbed linear differential equation
˙x=Ax+By˙y=εCx+εDy |
Introducing z:=x+A−1By one can rewrite this as
˙z=Az+ε(⋯)˙y=ε((−CA−1B+D)y+Cz) |
Here the fast system is just given by ˙z=Az, and the slow system (on the critical manifold defined by z=0) is given by
˙y=ε(−CA−1B+D)y. |
Using Tikhonov's theorem (in the form given e.g. in Verhulst [23], Ch. 8), one sees that both conditions (ⅰ), (ⅱ) are equivalent to exponential attractivity of the stationary point 0 for the linear system.
Lemma 4. Let V⊆Rn be open and nonempty, 0<r<n, δ>0 and
B1:V×[0,δ)→Rn×r,(x,ε)↦B1(x,ε)B2:V×[0,δ)→Rn×(n−r),(x,ε)↦B2(x,ε) |
be smooth functions (defined in some neighborhood of V×[0,δ)) such that Rn is the sum of the image W1 of B1 and the image W2 of B2, for every (x,ε). Then the entries of the matrix Q(x,ε)∈Rn×n which sends v∈Rn to its W2-component with respect to the direct sum decomposition W1⊕W2 depend smoothly on (x,ε).
Proof. We suppress the arguments (x,ε) in the notation. By assumption C:=(B1,B2) is invertible, and the entries of C−1 depend smoothly on (x,ε). With the projection matrix given by
Q=(0B2)C−1, |
the assertion is obvious.
[1] | A. N. Tikhonov, Systems of differential equations containing a small parameter multiplying the derivative (in Russian), Math. Sb, 31 (1952), 575–586. |
[2] | N. Fenichel, Geometric singular perturbation theory for ordinary differential equations, J. Differ. Equations, 31 (1979), 53–98. |
[3] | H. G. Kaper and T. J. Kaper, Asymptotic analysis of two reduction methods for systems of chemical reactions, Physica D, 165 (2002), 66–93. |
[4] | V. Noel, D. Grigoriev, S. Vakulenko, et al., Tropicalization and tropical equilibrium of chemical reactions, In: G.L. Litvinov, and S.N. Sergeev (eds): Tropical and idempotent mathematics and applications, Contemporary Math., 616, 261–275. Amer. Math. Soc., Providence, 2014. |
[5] | O. Radulescu, S. Vakulenko and D. Grigoriev, Model reduction of biochemical reactions networks by tropical analysis methods, Math. Model. Nat. Phenom., 10 (2015), 124–138. |
[6] | S. S. Samal, D. Grigoriev, H. Fröhlich, et al., Analysis of reaction network systems using tropical geometry, In: V.P. Gerdt, W. Koepf, W.M. Seiler and E.V. Vorozhtsov (eds.): Computer Algebra in Scientific Computing. 17Mth International Workshop, CASC 2015., 424–439, Lecture Notes in Computer Science, 9301, Springer-Verlag, Cham, 2015. |
[7] | S. S. Samal, D. Grigoriev, H. Fröhlich, et al., A geometric method for model reduction of biochemical networks with polynomial rate functions, Bull. Math. Biol., 77 (2015), 2180–2211. |
[8] | C. Chicone, Ordinary differential equations with applications, 2nd edition, Springer-Verlag, New York, 2006. |
[9] | D. Capelletti and C. Wiuf, Uniform approximation of solutions by elimination of intermediate species in deterministic reaction networks, SIAM J. Appl. Dyn. Syst., 16 (2017), 2259–2286. |
[10] | P. T. Cardin and M. A. Texeira, Fenichel theory for multiple time scale singular perturbation problems, SIAM J. Appl. Dyn. Syst., 16 (2017), 1425–1452. |
[11] | A. Goeke and S. Walcher, A constructive approach to quasi-steady state reduction, J. Math. Chem., 52 (2014), 2596–2626. |
[12] | A. Goeke, S. Walcher and E. Zerz, Determining "small parameters" for quasi-steady state, J. Differ. Equations, 259 (2015), 1149–1180. |
[13] | L. Noethen and S. Walcher, Tikhonov's theorem and quasi-steady state, Discrete Contin. Dyn. Syst. Ser. B, 16 (2011), 945–961. |
[14] | J. Keener and J. Sneyd, Mathematical physiology I: Cellular physiology, 2nd edition, Springer- Verlag, New York, 2009. |
[15] | A. Goeke, C. Schilli, S. Walcher, et al., Computing quasi-steady state reductions, J. Math. Chem., 50 (2012), 1495–1513. |
[16] | A. Goeke, S. Walcher and E. Zerz, Classical quasi-steady state reduction – A mathematical characterization, Physica D, 345 (2017), 11–26. |
[17] | F. R. Gantmacher, Applications of the theory of matrices, Dover, Mineola, 2005. |
[18] | A. Goeke, Reduktion und asymptotische Reduktion von Reaktionsgleichungen., Ph.D thesis, RWTH Aachen, 2013. Available from: http://darwin.bth.rwth-aachen.de/opus3/ volltexte/2013/4814/pdf/4814.pdf. |
[19] | R. Heinrich and M. Schauer, Quasi-steady-state approximation in the mathematical modeling of biochemical networks, Math. Biosci., 65 (1983), 155–170. |
[20] | C. Lax and S. Walcher, Singular perturbations and scaling, Discrete Contin. Dyn. Syst. Ser. B, to appear. arXiv:1807.03107 |
[21] | J. D. Murray, Mathematical biology. I. An introduction, 3rd edition, Springer, New York, 2002. |
[22] | M. Stiefenhofer, Quasi-steady-state approximation for chemical reaction networks, J. Math. Biol., 36 (1998), 593–609. |
[23] | F. Verhulst, Methods and applications of singular perturbations. Boundary layers and multiple time scale dynamics, Springer-Verlag, New York, 2005. |
1. | Ian Lizarraga, Martin Wechselberger, Computational Singular Perturbation Method for Nonstandard Slow-Fast Systems, 2020, 19, 1536-0040, 994, 10.1137/19M1242677 | |
2. | Samuel Jelbart, Nathan Pages, Vivien Kirk, James Sneyd, Martin Wechselberger, Process-Oriented Geometric Singular Perturbation Theory and Calcium Dynamics, 2022, 21, 1536-0040, 982, 10.1137/21M1412402 | |
3. | Samuel Jelbart, Christian Kuehn, Discrete geometric singular perturbation theory, 2023, 43, 1078-0947, 57, 10.3934/dcds.2022142 | |
4. | Ian Lizarraga, Bob Rink, Martin Wechselberger, Multiple timescales and the parametrisation method in geometric singular perturbation theory, 2021, 34, 0951-7715, 4163, 10.1088/1361-6544/ac04bf | |
5. | Niclas Kruff, Christoph Lüders, Ovidiu Radulescu, Thomas Sturm, Sebastian Walcher, Algorithmic Reduction of Biological Networks with Multiple Time Scales, 2021, 15, 1661-8270, 499, 10.1007/s11786-021-00515-2 | |
6. | G. A. Kurina, M. A. Kalashnikova, Singularly Perturbed Problems with Multi-Tempo Fast Variables, 2022, 83, 0005-1179, 1679, 10.1134/S00051179220110017 | |
7. | S Jelbart, C Kuehn, Extending discrete geometric singular perturbation theory to non-hyperbolic points, 2024, 37, 0951-7715, 105006, 10.1088/1361-6544/ad72c5 | |
8. | Seho Park, Herschel C. Pangborn, 2024, Multi-Timescale System Separation via Data-Driven Identification Within a Singular Perturbation Framework, 979-8-3503-8265-5, 2721, 10.23919/ACC60939.2024.10644768 | |
9. | S. Jelbart, C. Kuehn, S.-V. Kuntz, Geometric Blow-Up for Folded Limit Cycle Manifolds in Three Time-Scale Systems, 2024, 34, 0938-8974, 10.1007/s00332-023-09987-x | |
10. | Justin Eilertsen, Santiago Schnell, Sebastian Walcher, Natural Parameter Conditions for Singular Perturbations of Chemical and Biochemical Reaction Networks, 2023, 85, 0092-8240, 10.1007/s11538-023-01150-7 |