• No results found

Convergence Properties of Piecewise Power Approximations

N/A
N/A
Protected

Academic year: 2022

Share "Convergence Properties of Piecewise Power Approximations"

Copied!
16
0
0

Laster.... (Se fulltekst nå)

Fulltekst

(1)

http://dx.doi.org/10.4236/am.2016.713124

Convergence Properties of Piecewise Power Approximations

Arcady Ponosov1, Anna Machina2, Valeria Tafintseva1

1Department of Mathematical Sciences and Technology, Norwegian University of Life Sciences, Ås, Norway

2Department of Mathematics and Statistics, University of Victoria, Victoria, Canada Received 18 June 2016; accepted 8 August 2016; published 11 August 2016 Copyright © 2016 by authors and Scientific Research Publishing Inc.

This work is licensed under the Creative Commons Attribution International License (CC BY).

http://creativecommons.org/licenses/by/4.0/

Abstract

We address the problem of convergence of approximations obtained from two versions of the piecewise power-law representations arisen in Systems Biology. The most important cases of mean- square and uniform convergence are studied in detail. Advantages and drawbacks of the repre- sentations as well as properties of both kinds of convergence are discussed. Numerical approxi- mation algorithms related to piecewise power-law representations are described in Appendix.

Keywords

Power-Law Representations, Piecewise Nonlinear Approximations, Least-Squares Minimization, Mean-Square and Uniform Convergence

1. Introduction

For a given function v x

(

1,,xn

)

>0, defined in a domain Ω ∈n, let us calculate its partial derivatives in the logarithmic space:

( )

ln ,

ln

j j

j P j

v x v

f P

x v x

 ∂  ∂

=∂  =∂ (1) where P=

(

x1,,xn

)

is an arbitrary point. If fj

( )

P =const in the domain Ω for all j, then v clearly is a power function in Ω of the form

1 j

n g

j j

α x

= where α >0. In this paper we study piecewise constant ap- proximations of the quantities (1) or, in other words, nonlinear approximations of the function v by piecewise power functions.

(2)

This study is first of all motivated by applications in Systems Biology, where many networks can be de- scribed via compartment models

(

1, ,

) (

1, ,

)

, 1, , ,

i i n i n

x =V+ xxV xx i=  n (2) with the influx and efflux functions Vi+≥0 and Vi≥0, respectively.

For instance, in a typical metabolic network used in Biochemical Systems Theory the index i i

(

=1,,n

)

refers to the n internal metabolites xi ≥0. The influx Vi+

(

x1,,xn

)

≥0, resp. efflux Vi

(

x1,,xn

)

≥0 function accounts for the rate (velocity) of a production (synthesis), resp. degradation of the metabolite xi.

Another important example is gene regulatory networks which in many cases can be described as a system of nonlinear ordinary differential equations of the form

(

1

) (

1

)

d , , , , , 1, , ,

d

i

i n i n i

x F z z G z z x i n

t =  −  =  (3)

where x ti

( )

is the gene concentration (i = 1, ···, n) at time t, while the regulatory functions Fi and Gi depend on the response functions zk =zk

( )

xk , which control the activity of gene k and which are assumed to be sigmoid-type functions [1].

The derivatives fj

( )

P in the logarithmic space are very important local characteristics of biological net- works. In Biochemical Systems Theory these derivatives are known as the kinetic orders of the function v, while in Metabolic Control Analysis (see e.g. [2]) they are called elasticities. From the mathematical point of view, these quantities measure the local response of the function v to changes in the dependent variable (for instance, the local response of enzyme or other chemical reaction to changes in its environment). Thus, they describe local sensitivity of the function v, the terminology which is widespread in e.g. engineering sciences.

If all influx and efflux functions in (2) have constant kinetic orders, one obtains the so-called “synergetic system”, or briefly “S-system”:

1 1

, 1, , ,

ij ij

n n

g h

i i j i j

j j

x α x β x i n

= =

=

=

  (4)

where the exponents gij, hij represent all the (constant) kinetic orders associated with (4). The right-hand side of an S-system, thus, contains power functions, and analysis based on S-systems is, therefore, called “Power- Law (PL) Formalism”, see e.g. [3]-[7]).

The Power-Law Formalism has been successfully applied to a wide number of problems, for example, to metabolic systems [8], gene circuits [9], signalling networks [10]. Such systems are very advantageous in bio- logical applications, as the systems’ format considerably simplifies mathematical and numerical analysis such as steady state analysis, sensitivity, stability analysis, etc. For instance, calculation of steady states for the S- systems is a linear problem (see [7]). By these and other biological and mathematical reasons, it was suggested in [11] to classify such systems as “a canonical nonlinear form” in systems biology.

In many models, however, the kinetic orders may vary considerably. A typical example is a model coming from Generalized Mass Action

(

1

)

1

, , , 1, , ,

p

i ir r n

r

x µ V x x i n

=

=

=

   (5)

where the power functions Vr

(

x1,,xn

)

describe the rates of the process no. r, while µir is a stoichiometric factor that stands for the number of molecules of xi produced, i.e. µ =ir 1, 2, or µ = − −ir 1, 2,. Collect- ing the processes in (5) in a net process of synthesis Vi+ (positive terms) and a net process of degradation Vi (negative terms) results in an aggregated system (2), which is not an S-system.

Another example of generic systems with non-constant kinetic orders stems from Saturable and Cooperativity Formalism [12] reflecting two essential features of biological systems, which gave the name to this formalism (see [13] for more details). In this case, the system (5) becomes

(

1

) (

1

)

1 1

,

j j

j j j j

n n

n m

j j

j j

i n n

n n m m

j j j j

j j

v x w x

x

K x L x

= =

= =

= −

+ +

∏ ∏

∏ ∏

 (6)

where i=1,,n and nj, mj, Kj and Lj are real numbers.

(3)

Another version of Saturable and Cooperativity Formalism, which is mentioned in [12], is defined as follows:

1 1

1 1

1 1

1 1

1 1, 1 1,

,

j j

j j j j

n n

n m

j j

j j

i n n n n n n

n n m n

j i j j i j

i i

j j j i j j j i

x x

x

x b x x c x

α β

α β

= =

= =

= = ≠ = = ≠

= −

+ +

∏ ∏

∑ ∑

∏ ∏ ∏ ∏

 (7)

where i=1,,n and nj, mj, bi, ci, α and β are real numbers.

In the case of gene regulatory networks (3) the sensitivities (1) are non-constant as well, even if one considers the functions Fi and Gi to be multilinear in zk. In addition, the usage of non-multilinear functions are also known in this theory [14].

Taking into account the importance of kinetic orders/elasticities/sensitivities (1) in Systems Biology, one one hand, and convenience of the well-developed analysis of S-systems (stability theory [7], parameter estimation routines [15], software packages) on the other, a new kind of generic representations of compartment systems (2) was suggested in [16] (see also [17] for further applications of this representation). According to this idea, the entire operating domain is divided to partition subsets where all kinetic orders can be viewed as constants. In other words, the system (2) is approximated by a set of S -systems, each being only active in its own partition subset. This way of representing (2) is called “Piecewise Power-Law Formalism” [18].

From the biological point of view, piecewise power-law representations are useful in many respects, when compared to other ways of approximations, as they take into account biologically relevant characteristics (kinetic orders) rather than the standard partial derivatives. Therefore, piecewise S-systems preserve important biological structures and, at the same time, do not destroy a relatively simple mathematical structure of plain S-systems. By this reason, approximations of a general target function by piecewise power approximations may be of a great importance for biological and other modelling. A rigorous mathematical justification of the idea of piecewise power-law approximations is the main purpose of the present paper. More precisely, we consider mean-square and uniform convergence of approximations by piecewise power functions to the target function provided that the associated partitions of the operating domain Ω satisfy some additional assumptions. One of the challenges is that partitions of the operating domain Ω may not be chosen freely in applications. For instance, the partitions may directly stem from biological properties of the model [17]. Other ways of con- structing partitions can be dictated by optimality-oriented algorithms. In Appendix (see also [18]) we describe such a method which goes back to the paper [19] and which is based on an automatical procedure, allowing to obtain simultaneously the best possible polyhedral partition and the respective best possible piecewise linear approximation in the logarithmic space.

The main results of the paper are presented in Section 3 (mean-square convergence of piecewise power approximations) and in Section 4 (uniform convergence of piecewise power approximations). Several auxiliary results are proved in Appendices A.1-A.3, while Appendix A.4 presents an approximation algorithm which provides an automated partition and the respective best possible approximation in the logarithmic space for a given number of subdomains. Finally, in Appendix A.5 we explane by example why a direct piecewise power- law fitting is ill-posed.

2. Preliminaries

Throughout the paper we use the following notations (see Table 1). Let v:n+\ 0

{ }

→+, and let

: n

ψ  → be given by ψ

( )

y =logv x

( )

, y=log .x

Let Ω be a domain in the space n+\ 0 ,

{ }

which we call Cartesian. We assume Ω to be closed and bounded (i.e. compact) subset of n. Let be its image in the logarithmic space n and

{ }

iN Ni=1 be a

measurable partition of ∆. This means that ∆Ni are all Borel measurable subsets of n, ∆ ∩ ∆ = ∅Ni Nj for every i j, =1,,N and

Ni=1∆ = ∆Ni for any natural N. In some results and algorithms ∆ will be a poly- hedron domain in the logarithmic space, and ∆iN will be a polyhedral partition.

Below

{ }

Ni i=N1 is the measurable partition of Ω which is the image of the partition

{ }

iN i=N1 under the inverse logarithmic transformation.

(4)

Table 1. Overview of the basic terminology and notation used in the paper (LS-least-squares).

Cartesian space Logarithmic space

Space n+\ 0{ } n

Target function v x( ) ψ( )y =logv x( )

Domain

Measurable partition

{ }

1

N N i i=

{ }

1

N N i i=

LS power approximation vN( )x ψN( )y =log

(

vN( )x

)

LS linear approximation VN( )x =exp

(

ΨN( )y

)

ΨN( )y We also put δN =1max diam≤ ≤i N

( ( )

iN

)

.

Let

( )

1 ijN

n f

N N

i i j

j

v x γ x

=

=

be a least-squares (LS) power-law fitting to the function v on ΩNi ,i=1,,N. For x∈ Ω we consider the piecewise power function vN

( )

x =viN

( )

x , whenever x∈ ΩiN. We put

( )

log

( ( ) )

.

N N

y v x

ψ =

Let

( )

1 n

N N N

i ij j

j

y c g y

=

Ψ = +

be a LS linear approximation to the function ψ on ∆Ni , i=1,,N. For y∈ ∆ we consider the piecewise linear function ΨN

( )

y = ΨiN

( )

y , whenever y∈ ∆iN. We put also

( )

exp

( ( ) )

.

N N

V x = Ψ y

We remind that the parameters cN and gijN of the linear functions ΨiN are uniquely obtained from the following minimization criterion in the logarithmic space:

(

1

) (

1

)

2 1

1

, , , , d d min .

N i N

N

n i n n

i

y y y y y y

ψ

= ∆

 − Ψ  →

 

∑ ∫

(8)

Alternatively, one can define approximations of the target function v by power functions minimizing the distance in the space Ω:

(

1

) (

1

)

2 1

1

, , , , d d min .

iN N

N

n i n n

i

v x x v x x x x

= Ω

 −  →

 

∑ ∫

(9)

Our last minimization criterion looks similar to (6), but is, in fact, very different

(

1

) (

1

)

2 1

1

, , , , d d min,

iN N

N

n i n n

i

y y y y y y

ψ

= ∆

 − Ψ  →

 

∑ ∫

(10)

as the minimum here is taken over all polyhedral partitions

{ }

Ni i=N1 of the polyhedral domain ∆, and all corresponding linear functions ΨNi (i=1,,N ).

The main advantage of the criteria (8) and (10) is their linearity that provides the uniqueness of the solution and also makes the process of finding the solution computationally cheap, as it is based on explicit matrix formulas. On the other hand the use of the logarithmic transformation requires caution. The influences of the data values will change, as will the error structure of the model. Yet, the criterion (8) only requires a standard linear regression, while the criterion (10) requires a special regression algorithm, still linear, but much more involved (see Appendix A.4 for details).

The criterion (9) gives best possible approximation in terms of the LS error in the Cartesian space. However, a nonlinear regression algorithm should be used in this case, which is less advantageous, especially when the number of the estimated parameters is big. In addition, the nonlinear regression may have other drawbacks, one of which is ill-posedness (see Appendix A.5).

(5)

3. Mean-Square Convergence of Piecewise Power Approximations

The results of this section provide the mean-square convergence (L2-convergence) of piecewise approximations by power functions. The involved parameters may be e.g. obtained according to one of the minimization criteria (8) or (9).

The main technical challenges stemming from the nature of these minimization algorithms can be sum- marized as follows: 1) the L2-convergence of the approximations in the logarithmic space may not imply the L2-convergence of their images in the Cartesian space (and vice versa); 2) it is not evident that automatic dissections of the operating domain, as e.g. in the algorithms based on the minimization criterion (10), make the diameters of the partition subsets go to zero even if the number of partition subsets tends to ∞.

Three propositions below deal with L2-convergence in the logarithmic domain.

Proposition 1. Let the target function v>0 be measurable and bounded on and ψ =logv. Suppose that the measurable partitions

{ }

iN Ni=1 satisfy the property δ →N 0 (N→ ∞). Then for the corresponding LS approximations ΨN in and vN in we have Ψ →N ψ , vNv in the respective L2-norms, if

N→ ∞.

To prove this proposition we need the following lemma, the proof of which can be found in Appendix A.1:

Lemma 1. Let v be measurable and 0< ≤m v y

( )

M on for some constants m M . Let , ψ =logv and the measurable partitions

{ }

Ni i=N1 satisfy the property δ →N 0 (N→ ∞). Then there exists a sequence

{ }

θN of continuous on functions satisfying the properties θN

( )

y =const on any iN, θN

( )

y ψ

( )

y

for all y∈ ∆ and

{ }

θN (resp.

{ }

ΘN =

{

exp

( )

θN

}

) L2-converges to ψ (resp. v) on

(resp.

).

Proof of Proposition 1. We use the sequences

{ }

θN and

{ }

ΘN =

{

exp

( )

θN

}

from the lemma 1, which both converge in the L2-sense in the respective domains.

Since ΨN

( )

y is the LS piecewise linear approximation in ∆, we have

( ) ( )

2 2 0

N N

L L

ψ θ ψ

Ψ − ≤ − →

as N→ ∞.  In the next proposition we do not assume that δ →N 0.

Proposition 2. Let be a polyhedral domain in n, the function ψ be square integrable in and

{ }

iN Ni=1 be the optimal polyhedral partition of obtained by the algorithm described in Appendix A.4. Then for the corresponding LS approximations ΨN in we have Ψ →N ψ in the L2-norm, if N→ ∞.

Proof. Evidently, for the L2-function ψ there exists a sequence of polyhedral partitions

{ }

i i=N1 of ∆ such that max diam

( ( )

i

)

0

i

δ = ∆ → as N→ ∞ and a sequence of piecewise constant functions

{

ΨN

( )

y

}

given

by ψN

( )

y =CiN, whenever y∈ ∆iN for which Ψ →N ψ in the L2-norm if N→ ∞. For the optimal polyhedral approximation ΨN

( )

y we obtain

( ) ( )

2 2 0

N N

L L

ψ ψ

Ψ − ≤ Ψ − →

as N→ ∞.  In particular, the assumption on ψ is fulfilled if the target function v is measurable and bounded on Ω. The case of the L2-convergence of the approximations VN, given as VN =exp

( )

ΨN , is more involved. The reason for that is that the L2-convergence of the sequence

{ }

ΨN does not necessarily imply the L2- convergence of the sequence

{

exp

( )

ΨN

}

.

We introduce the following notation. Given a partition subset ∆Ni of ∆ we put

( )

0, 1,

( )

,

( ) (

1, ,

)

,

N N N

i i k i

e y = e k y = yy k k=  n (11)

(6)

where the point

(

yiN

( )

k

)

nk=1∈ ∆iN is the center of mass of the convex set ∆Ni given by

( )

1 d , 1, , .

mes N

i N

i N k

i

y k y y k n

= =

(12)

Let AiN =

(

aiN

( )

k l,

)

be the symmetric n n× -matrix with the entries defined as

( )

, 1

(

,

) ( )

, d , , 1, , ,

mes N

i

N N N

i N i i

i

a k l e k y e l y y k l n

= =

(13)

Below we fix a matrix norm . . All matrix norms are equivalent. One of the norms is Euclidean, which is defined via the maximal eigenvalues: A = λmax

( )

A AT . In the case of symmetric, positive definite matrices (like AiN above) we can write that Amax

( )

A .

We say that the sequence of partitions

{ }

Ni i=N1 (N=1, 2,) of satisfies the condition (∆) if there exists a constant c0 >0 such that

( )

1 0

diam∆iN AiN c.

If the chosen norm is Euclidean, then the latter estimate can be rewritten as

0

diam ,

N i N i

λ c

∆ ≤

where λiN is the least (positive) eigenvalue of the matrix AiN (i=1,,N ; N=1, 2,).

Informally speaking, this property means that the partition subsets cannot be too different from each other in the shape. Assume, for instance, that the partition sets are enclosed in rectangular boxes. The result below says that if the ratio of the longest and the shortest edges of the boxes is bounded above, i.e. boxes are not “too thin”, then the sequence of such boxes satisfies the property (∆).

Proposition 3. A sequence of rectangular boxes

{ }

PN satisfies the property (∆) if and only if

1,2,

sup

N N N

a b

=

< ∞

, where aN (resp. bN) is the length of the smallest (resp. biggest) edge of the box PN. Proof. We calculate the matrix (13).

We fix N and the Nth rectangular box P=PN given by

1l 1 1r, 2l 2 2r, , nl n nr. yyy yyyyyy

Let ,

2

l r

i i

i

y y

y = + i=1,, ,n be the center of mass and e i y

( )

, = yiyi. The substitution

[ ] ( )

, where , 0,1 , 1 2 ,

l r l

i i i i i i i i i i i i

y = yz δ =yy zyyz − yields

(

,

) ( )

, d 1 n k l

(

k 1 2

)(

l 1 2 d

) (

1 ,

)

P

e k y e l y y δ δ δ δ z z z k l n

Σ

= − − ≤ ≤

∫ ∫

where Σ =

[ ] [ ]

0,1 × 0,1 × ×

[ ]

0,1 n. Since

(

1 2

)(

1 2 d

)

0

( ) (

, 1 2

)

2d 1 ,

k l k 12

z z z k l z z

Σ Σ

− − = ≠ − =

∫ ∫

and mes∆ =iN δ1δn, the matrix (13) becomes

2 2

diag 1 , ,

12 12

δn

δ

 

 

  . The least eigenvalue of the matrix is equal

(7)

to 1 12 min

( {

δ12,,δn2

} )

, i.e. to 1 12

( )

aN 2. The diameter of the box can be estimated above by the constant n bN, which also dominates the asymptotics of the diameter. Therefore the condition (∆) is fulfilled for the given sequence of rectangular boxes if and only if the sequence

{

bN aN

}

is bounded above. 

The next lemma is proved in Appendix A.2.

Lemma 2. Assume that the target function v>0 is measurable and bounded on and ψ =logv. Assume further that the sequence of partitions

{ }

iN i=N1 (N =1, 2,) of satisfies the condition (∆). Then the corresponding LS approximations VN

( )

x and ΨN

( )

y are uniformly bounded on and ∆, re- spectively, i.e. there exist constant C1>0 and C2>0 such that

( )

1sup

( )

N

x

V x C v x

∈∆

for all N=1, 2, and all x∈ ∆,

( )

2sup

( )

N

y

y C ψ y

∈∆

Ψ ≤

for all N=1, 2, and all y∈ ∆.

The main result of this section is the following theorem:

Theorem 4. Let the target function v be measurable and bounded on .

1) If the measurable partitions

{ }

iN i=N1 have the property δ →N 0 (N→ ∞), then vNv and VNv in the L2

( )

-norm as N→ ∞.

2) Assume that ∆ is a polyhedral domain in n, while a sequence of polyhedral partitions

{ }

Ni i=N1 of

and associated LS piecewise linear approximations ΨN satisfy the criterion (10) for each N=1, 2,. Assume further that the partitions

{ }

Ni i=N1 (N=1, 2,) satisfy the condition (∆). Then VNv in the L2

( )

-

norm as N→ ∞.

Proof. To prove the first part of the theorem, we apply Lemma 1 and obtain

( ) ( )

2 2 0

N N

L L

v v ψ

− ≤ Θ − →

as N→ ∞, since vN

( )

x is the LS piecewise power approximation in Ω.

In the second part of the theorem, we use either Proposition 1 or Proposition 2, which yields the L2- convergence of the LS approximations ΨN to the function ψ =logv. Applying Lemma 2 we obtain the uniform boundedness of the approximations: ψ

( )

y ,ΨN

( )

y M for some M and any N =1, 2,. Then we have

( )

( ) ( )

( ( ) ) ( ( ) ) ( )

( )

2

2 2

2 1

d

exp exp exp d

,

N N

L

N

n

N L

V v V x v x x

y y y y y

C

ψ ψ

− = −

= Ψ − + +

≤ Ψ −

where exp

( )

max exp

(

1 n

)

.

y

C M y y

= ∈∆ ++ The latter estimate is due to the uniform Lipschitz continuity of the function exp

( )

u on the interval

[

0,M

]

: exp

( )

u −exp

( )

v ≤exp

( )

M uv.

This estimate proves the L2-convergence of the LS approximations VN to the target function v. 

4. Uniform Convergence of Approximations

In the previous section we studied convergence of LS approximations in the L2-norm. In many applications, however, it is desirable to consider their uniform convergence. This may be, for instance, of interest if we include the obtained approximations into the models based on differential equations, as it is well-known that convergence of (approximations of) solutions is only guaranteed by the uniform convergence of (approximations

(8)

of) the right-hand sides.

The main result of this section is formulated in terms of kinetic orders fj

( )

x and fjN

( )

x of the target function v x

(

1,,xn

)

and its piecewise power approximations wN.

Theorem 5. Let the target function v>0 be a C1-function (i.e. differentiable with the continuous partial derivatives). Let the sequence of partitions

{ }

iN Ni=1 (N=1, 2,) of have the following two properties:

1) The closure of each ΩNi coincides with the closure of its interior intNi . 2) δ →N 0 (N→ ∞).

Assume, in addition, that for any i=1,,N , N=1, 2, there exist points PiN∈ ΩiN, QiNintiN such that piecewise power approximations wN (=wiN on iN) satisfy

( )

iN N

( )

iN , j

( )

iN jN

( )

iN .

v P =w P f Q = f Q (14) Then wNv uniformly on as N→ ∞.

Proof. We fix N and consider the corresponding partition ∆Ni of the domain ∆. Let ψ =logv,

N log N

Φ = w , yiN =logPiN, ciN =logQiN. Clearly, ψ

( )

ciN =

(

f1

( )

PiN ,,fn

( )

PiN

)

. By assumption, for

N

y∈ ∆i we have ΦiN

( )

y = ∇ψ

( )

ciN

(

yy0

)

+ψ

( )

yiN , where yiN∈ ∆Ni ,ciNintiN. On the other hand, the mean value theorem yields ψ

( )

y ψ

( )

yiN = ∇ψ

( )

c

(

yyiN

)

where c∈ ∆iN depends on y. Therefore

( ) ( ) ( ) ( )

.

N N N

i y ψ y ψ c ψ ci y yi

Φ − ≤ ∇ − ∇ −

The uniform continuity of the continuous vector function ψ

( )

y on and the property that

( )

1

maxdiam 0

N N

i N i N

δ = ≤ ≤ ∆ → → ∞

imply that, given an ε>0, the estimate

( ) ( )

.

N

i y ψ y ε

Φ − < (15) is fulfilled for sufficiently large N.

Since (15) holds for any i

{

1, 2,,N

}

, we also obtain that for sufficiently large N ΦN

( )

y −ψ

( )

y, i.e. Φ →N ψ uniformly on ∆ as N→ ∞.

As the uniform convergence of the sequence

{ }

ΦN implies its uniform boundedness, there is M such that

( )

,

( )

N y ψ y M

Φ ≤ for all y∈ ∆. Therefore,

( ) ( )

exp

( )

exp

( ) ( ) ( )

,

N N N

w xv x = Φ y − ψ yCΦ y −ψ y

where C=expM. This gives the uniform convergence of wN to v as N→ ∞.  Our last result shows that the LS approximations converge uniformly in the scalar case. This is due to the fact that in the scalar case the equalities (14) are always fulfilled.

Corollary 1. Let the target function v be continuous on Ω =

[

A B,

]

(A> >B 0) and ψ =logv. Assume

that the sequence of partitions

{ }

iN Ni=1 (N=1, 2,) of has the property δ →N 0 (N→ ∞).

Then for the corresponding LS power approximations VN and vN we have VNv, vNv (N→ ∞) uniformly on Ω.

The proof of the theorem follows directly from the previous theorem and the following lemma, the proof of which is given in Appendix A.3.

Lemma 3. Let a linear function l:

[ ]

a b, be the LS approximation of a C1 function ψ :

[ ]

a b, on the entire interval

[ ]

a b . Then there exist , y0

[ ]

a b, and c

( )

a b, such that l y

( )

0

( )

y0 ,

( ) ( )

.

l c′ =ψ′ c

5. Discussion and Conclusions

Piecewise power-law representations may be very useful as practical approximations to target functions which are defined analytically or numerically. However, a strict mathematical justification of these approximations is not always paid attention to. Unfortunately, such an analysis is not always straightforward, especially if one puts additional a priori assumptions on the approximations, which is quite common in many applications.

(9)

We showed in the present paper that under additional assumptions power approximations do converge to the target function. We studied least-squares and uniform convergence, both of which are widely used (explicitly or implicitly) in applications.

Our analysis dealt with two types of regression: linear regression in the logarithmic space and power-law regression in the Cartesian space. The first procedure has all the advantages of the linear regression, but the transformation back to the Cartesian space distorts the error structure of the problem; the least squares error for the resulting piecewise power-law fitting is in general less accurate than the corresponding error for a power-law regression of the original data. As a partial remedy, it may be advantageous to apply power-law regression to the original data over each of the partition subsets back in the Cartesian space. Yet, being nonlinear regression this procedure is essentially ill-posed. Thus, both kinds of regression have their strong and weak sides, so that the choice between them must be undertaken by modeling consideration.

In many cases, it may also be advantageous to use the classical linear regression in combination with optimal partitions of the operating domain. In the logarithmic space this procedure is again linear and can be auto- matized, but this may also cause several technical problems when proving the convergence of the corresponding approximations.

In the present paper, we offered a partial mathematical justification of the analysis based on piecewise power approximations, stemming from both kinds of regression, by verifying their convergence in the mean-square (L2) and uniform sense. Uniform convergence is e.g. important if target functions are included in differential eq- uations, as it is the uniform, and not L2-convergence, which is inherited by the solutions of the equations. How- ever, a comprehensive analysis of convergence of solutions of differential equations, approximated by piecewise S-systems, is beyond the scope of this paper and will be discussed in a separate publication.

Acknowledgements

The work of the first author was partially supported by a EEA grant coordinated by Universidad Complutense de Madrid, Spain, and by the grant #239070 of the Norwegian Research Council.

References

[1] de Jong, H. (2002) Modeling and Simulation of Genetic Regulatory Systems: A Literature Review. Journal of Compu- tational Biology, 9, 67-104. http://dx.doi.org/10.1089/10665270252833208

[2] Kascer, H. and Burns, J. (1973) The Control of Flux. Symposia of the Society for Experimental Biology, 27, 65-104.

[3] Savageau, M.A. (1969) Biochemical Systems Analysis. I. Some Mathematical Properties of the Rate Law for the Component Enzymatic Reactions. Journal of Theoretical Biology, 25, 365-369.

http://dx.doi.org/10.1016/S0022-5193(69)80026-3

[4] Savageau, M.A. (1969) Biochemical Systems Analysis. II. The Steady-State Solutions for an n-Pool System Using a Power-Law Approximation. Journal of Theoretical Biology, 25, 370-379.

http://dx.doi.org/10.1016/S0022-5193(69)80027-5

[5] Savageau, M.A. (1970) Biochemical Systems Analysis. III. Dynamic Solutions Using a Power-Law Approximation.

Journal of Theoretical Biology, 26, 215-226. http://dx.doi.org/10.1016/S0022-5193(70)80013-3

[6] Voit, E.O., et al. (1991) Canonical Nonlinear Modeling. S-System Approach to Understanding Complexity. Van No- strand Reinhold, NY.

[7] Voit, E.O. (2000) Computational Analysis of Biochemical Systems: A Practical Guide for Biochemists and Molecular Biologists. Cambridge University Press, New York.

[8] Alvarez-Vasquez, F., Sims, K.J., Cowart, L.A., Okamoto, Y., Voit, E.O. and Hannun, Y.A. (2005) Simulation and Va- lidation of Modeled Sphingolipid Metabolism in Saccharomyces cerevisiae. Nature, 433, 425-430.

http://dx.doi.org/10.1038/nature03232

[9] Atkinson, M.R., Savageau, M.A., Myers, J.T. and Ninfa, A.J. (2003) Development of Genetic Circuitry Exhibiting Toggle Switch or Oscillatory Behavior in Escherichia coli. Cell, 113, 597-607.

http://dx.doi.org/10.1016/S0092-8674(03)00346-5

[10] Vera, J., Rath, O., Balsa-Canto, E., Banga, J.R., Kolch, W. and Wolkenhauer, O. (2010) Investigating Dynamics of In- hibitory and Feedback Loops in ERK Signalling Using Power-Law Models. Molecular BioSystems, 6, 2174-2191.

http://dx.doi.org/10.1039/c0mb00018c

[11] Savageau, M.A. and Voit, E. (1987) Recasting Nonlinear Differential Equations as S-Systems: A Canonical Nonlinear

(10)

Form. Mathematical Biosciences, 87, 83-115. http://dx.doi.org/10.1016/0025-5564(87)90035-6

[12] Hernandez-Bermejo, B., Fairen, V. and Sorribas, A. (2000) Power-Law Modeling Based on Least-Squares Criteria:

Consequences for System Analysis and Simulation. Mathematical Biosciences, 167, 87-107.

http://dx.doi.org/10.1016/S0025-5564(00)00039-0

[13] Sorribas, A., Hernández-Bermejo, B., Vilaprinyo, E. and Alves, R. (2007) Cooperativity and Saturation in Biochemical Networks: A Saturable Formalism Using Taylor Series Approximations. Biotechnology and Bioengineering, 97, 1259- 1277. http://dx.doi.org/10.1002/bit.21316

[14] Tafinsteva, V., Machina, A. and Ponosov, A. (2013) Polynomial Representations of Piecewise-Linear Differential Eq- uations Arising from Gene Regulatory Networks. Nonlinear Analysis: Real World Applications, 14, 1732-1754.

http://dx.doi.org/10.1016/j.nonrwa.2012.11.008

[15] Chou, I.C. and Voit, E.O. (2009) Recent Developments in Parameter Estimation and Structure Identification of Bio- chemical and Genomic Systems. Mathematical Biosciences, 219, 57-83. http://dx.doi.org/10.1016/j.mbs.2009.03.002 [16] Savageau, M.A. (2002) Alternative Designs for a Genetic Switch: Analysis of Switching Times Using the Piecewise

Power-Law Representation. Mathematical Biosciences, 180, 237-253.

http://dx.doi.org/10.1016/S0025-5564(02)00113-X

[17] Coelho, P.M.B.M., Salvador, A. and Savageau, M.A. (2010) Relating Mutant Genotype to Phenotype via Quantitative Behavior of the NADPH Redox Cycle in Human Erythrocytes. PLoS ONE, 5, e13031.

http://dx.doi.org/10.1371/journal.pone.0013031

[18] Machina, A., Ponosov, A. and Voit, E.O. (2010) Automated Piecewise Power-Law Modeling of Biological Systems.

Journal of Biotechnology, 149, 154-165. http://dx.doi.org/10.1016/j.jbiotec.2009.12.016

[19] Ferrari-Trecate, G. and Muselli, M. (2002) A New Learning Method for Piecewise Linear Regression. Lecture Notes in Computer Science, 2415, 444-449. http://dx.doi.org/10.1007/3-540-46084-5_72

(11)

Appendix

A.1. Proof of Lemma 1, Section 1

Let us prove the lemma for the domain ∆. Notice that the function ψ =logv is measurable on the compact set ∆ and satisfies the estimate logmψ

( )

y logM on . By Lusin’s theorem, for any N, there is a uniformly continuous function ϕN on ∆ such that logm≤ϕN ≤logM for all N and the Lebesgue measure of the set y∈ ∆:ϕN

( )

y ≠ψ

( )

y is less than 1N. Now, there exists a δ >0 for which yy<δ, y y′∈ ∆, implies ϕN

( )

y −ϕN

( )

y′ <1N. We define θN

( )

y =ϕN

( )

yiN for some yiN∈ ∆iN.

Let kN be chosen in such a way that δkN <δ . If y∈ ∆iN, then we have

( )

N

( ) ( ) ( )

N 1 ,

N y y N y N yi N

ϕ −θ =ϕ −ϕ <

as y y, iN∈ ∆Ni and diam∆ <iN δkN <δ. Therefore, the Lebesgue measure of the set

( ) ( )

: N 1

y∈ ∆ θ y −ψ y > N is less than 1 N, so that the sequence

{ }

θN converges to ψ in measure, and due to its uniform bondedness, also in the L2-sense on ∆.

A similar argument applies to the sequence

{ }

ΘN on , where we use the sequence

{

exp

( )

θN

}

instead

of

{ }

θN .

A.2. Proof of Lemma 2, Section 1

Clearly, ψ

( )

y is measurable and bounded on ∆. Let sup

( )

y

c ψ y

= ∈∆ .

Let us fix a partition subset ∆iN. Our aim now is to find estimates for the norms of orthonormal basis functions

{ }

ek , k=0,,n in the linear subspace of the space L2

( )

Ni consisting of all linear functions and equipped with the scalar product

( ) ( )

1 d .

mes N

i N i

y y y

ξ η ξ η

⋅ = ∆

One basis is given by the set (11). However, this set is not necessarily orthogonal.

First of all, we choose e0=1 and observe that its norm is equal to 1. Using the description (11) of the basis functions eiN

(

k y,

)

defined via the center of mass we directly deduce from (12) that e0 =1 is orthogonal to any linear combination of the other basis functions. The challenge is therefore to estimate the norms of linear combinations ek =

ln=1αl ieN

( )

l y, , where αl are real numbers.

In the proof below we often omit one of the variables in eiN

( )

l y, , that is either l, or y, depending on a particular interpretation of this basis. Writing eiN

( )

y means that we regard it as a vector for each particular y, i.e. eiN

( )

y =

(

eiN

( )

1,y ,,eiN

(

n y,

) )

(the component eiN

( )

0,y is excluded in further considerations).

Omitting y (eiN

( )

l ) means that we treat eiN

( )

l y, as a function of y for a given l, i.e. as an element of the space

( )

2 N

Li .

As ek =1, we require the following constraints on the coefficients:

( )

2

( ) ( )

, =1

1 1

1 d , , d ,

mes N mes N

i i

n

N N T N

k i i l k i

N N

i k l i

e y y e k y e l y α α y α A α

= = =

where

α = ( α

l

)

nl=1. Therefore,

( ) ( )

max max

N N

i i

N

k i

y y

e y α e y

∈∆  = ∈∆ ⋅ (16) (where . is the Euclidean norm in n and a b is the scalar product of two vectors) with the constraint

TAiN 1 α α = .

Diagonalization of the symmetric, positive definite matrix AiN with the help of an orthogonal matrix Q gives the matrix containing the eigenvalues λ >k 0 of AiN on the diagonal. Putting β =Qα and using Qy = y , we obtain from (16) that

( ) ( ) ( )

1 2 2

1

max max max diam

N N N

i i i

n

N N N

k i i i k

y y y k

e y β Qe y β e y β

∈∆ ∈∆ ∈∆ =

 

= ⋅ ≤ ≤ ∆  

Referanser

RELATERTE DOKUMENTER

In the given paper a segmented regression approach and a greedy algorithm have been proposed and tested to simplify the piecewise linear generation function used to model

authentication of user traffic across networks. The purpose of the analysis is to show that there exist several use cases where such authentication is needed. The analysis

However, in comparison to the complex hydrological models (SMAR, HBV and SWAT), the simple single input-output models of the artificial neural network (ANN) and the

build a signal-specialized parameterization in a mutigrid hierarchy and minimize the signal approximation error, the error between the original surface signal and its re-

In this paper we present an algorithm that uses these properties to generate a piecewise linear approximation of implicit curves and surfaces, that is isotopic to the curve or

Given a point cloud, in the form of unorganized points, the problem of auto- matically connecting the dots to obtain an aesthetically pleasing and piecewise-linear closed

Surface-reconstructing growing neural gas (sgng) presented in this dissertation constitutes an artificial neural network that takes a set of sample points lying on an object’s

We present an algorithm for extracting these lines from piecewise linear tensor fields by finding and connecting all intersections with the cell faces of a data set.. The core of