• Keine Ergebnisse gefunden

5. Thermokinetic Modeling 89

5.4. Transformation and Translation

The thermokinetic modeling formalism is a convenient reformulation of the thermodynamic modeling formalism. Both formalisms are basically equivalent and the transformation and translation methods for thermodynamic models (see Section 4.1.1, p. 51) can be applied.

Here we give the transformation and translation rules for thermokinetic models. Since we re-strict ourselves to diagonal resistance matrices, we can only adapt methods where the occurrence of off-diagonal elements in the resistance matrix can be avoided.

Further, this section introduces an additional transformation method, namely the translation of stoichiometric coefficients that is specific for thermokinetic models.

5.4.1. Linear Transformation

Due to the restriction to diagonal resistance matrices, linear transformations of fluxes are not possible. Such transformations lead to non-diagonal resistance matrices. Thus, the following corollary considers only a transformation of concentrations.

Corollary 5.32 (Transformation of a thermokinetic model). Let M be a thermokinetic model (see Definition 5.29, p. 100), and let Tc, Tc,e andTJ,e be square matrices of full rank. Then, the solution of the system Mˆ with

E =TcNE, NˆP =TcNP, SˆE =Tc,eSE, SˆP =Tc,eSP, ˆ

c( ˆξ,ξˆe) =Tcc(TcT# ˆξ, Tc,eT # ˆξe), R( ˆˆ ξ,ξˆe) =R(TcT# ˆξ, Tc,eT # ˆξe), Nˆe =TcNeTJ,e, ξˆ0 =TcT ,−10.

is equivalent to that of M with ˆ

c=Tcc, TcT# ˆξ =ξ, Tc,eT # ˆξee, Jˆ=J, Fˆ =F, TJ,ee=Je.

5. Thermokinetic Modeling

Proof. This transformation rule follows from the Corollary 4.2 (p. 52).

§ 5.33 (Negative stoichiometric coefficients). In natural coordinates the entries of the matrices NE, NP, SE and SP are non-negative. After a transformation of concentrations, the entries of the corresponding transformed matrices may be negative.

5.4.2. Translation of the Variables

Section 4.1.2 (p. 54) introduced rules for the translation of the variablesc,µandJ in thermody-namic models. Analog rules can be formulated for thermokinetic models. This section discusses only the translation of chemical potentials because this operation is needed for an appropriate scaling of the state variables of the simulation equations. It is equivalent to a change of the reference state of the Gibbs formation energies.

Corollary 5.34 (Scaling of a TK model). Let M be a thermokinetic model (see Definition 5.29, p. 100), and let δξ ∈ Ri+0, δξe ∈ Ri+0,e be vectors with NT log(δξ) +ST log(δξe) = 0. Then, the solution of the system Mˆ with

E/P =NE/P, SˆE/P =SE/P, Nˆe=Ne, ˆ

c( ˆξ,ξˆe) =c

ξˆ◦δξ(−1),ξˆe◦δξe(−1)

, ξˆ00◦δξ(−1) and

R( ˆˆ ξ,ξˆe) = NET#δξ◦SET#δξ◦R

ξˆ◦δξ(−1),ξˆe◦δξe(−1)

=NPT#δξ◦SPT#δξ◦R

ξˆ◦δξ(−1),ξˆe◦δξe(−1) is equivalent to that of M with

ˆ

c=c, ξˆ=ξ◦δξ, ξˆee◦δξe, Jˆ=J, Fˆ =F, Jˆe =Je.

Proof. The proof follows the lines of the proof of Corollary 4.10 (p. 54). Write down the model equations ofM. Now replaceJ = ˆJ,F = ˆF,ξ = ˆξ◦δξ(−1)andξe= ˆξe◦δξe(−1). This immediately leads to the functions ˆc( ˆξ,ξˆe) and the matrices NˆE/P, SˆE/P and Nˆe. To prove the expression for Rˆ we start with the equation Rˆ◦Jˆ= ˆF and prove that it is equivalent to R◦J =F. The equationRˆ◦Jˆ= ˆF can be expanded to

Rˆ

z }| { NET#δξ◦SET#δξ

| {z }

δFE

◦R◦Jˆ= ˆF = ˆNET# ˆξ◦SˆET# ˆξe−NˆPT# ˆξ◦SˆPT# ˆξe

=NET#ξ◦SETe◦NET#δξ◦SET#δξe

| {z }

δFE

−NPT#ξ◦SPTe◦NPT#δξ◦SPT#δξe

| {z }

δFP

. (5.5)

5. Thermokinetic Modeling

Due to the specific choice of δξ and δξe, we have NET log(δξ) +SET log(δξe) = NPT log(δξ) + SPT log(δξe) and thus δFE = δFP. For this reason these terms cancel out and the equation R◦J =F remains.

Example 5.35 (Scaling of capacities and resistances). The above corollary is often neces-sary to numerically scale the simulation equations. Consider the single reaction of glucose-6-phosphate to fructose-6-glucose-6-phosphate (g6p f6p) that is catalyzed by the phosphoglucose iso-merase. Alberty [1] gives transformed Gibbs formation energies∆fG0◦i atpH = 7,T = 298.15 K, I = 0.25 M and c = 1 M:

fG0◦g6p =−1318.92 kJ mol−1, ∆fG0◦f6p =−1315.74 kJ mol−1.

These values may be used as chemical standard potentials µi and lead to the capacities (Ci = cmexp(−µi/(RT))):

Cg6p = 1.160×10231 M, Cf6p = 3.217×10230 M.

Clearly, these values are not suited for direct use in numerical simulations. The tables from Alberty [1] are based on the convention that chemical elements in the standard state in their most stable form have ∆fG0◦i = 0. Most relevant (bio)chemical species have a rather low, negative ∆fGi because the respective compounds are energetically more favorable than an analogous mixture of pure elements.

Using the above corollary, we can scale the capacities to values more suited for numerical computations. The stoichiometric matrix is NT = (−1,1) and thus the condition for the scaling factors is−log(δξg6p) + log(δξf6p) = 0 or equivalently δξg6p =δξf6p.

Assume that g6p has a typical concentration of 1 mM. From a numerical perspective, an optimal value of its capacity is1 mMbecause then the thermokinetic potentialξg6pthat is a state variable of the model varies around the typical value1. We chooseδξg6p =δξf6p = 1.160×10234 and get the new relation

cg6p =Cg6pδξg6p−1

| {z }

Cˆg6p

ξˆg6p cf6p =Cf6pδξf−16p

| {z }

Cˆf6p

ξˆf6p

with Cˆg6p = 1 mM and Cˆf6p = 0.2773 mM. These values are moderate and thus can be used for numerical computations.

§ 5.36. The above example shows that tables of Gibbs formation energies can be used to determine the capacities. The standardized and widely accepted definition of the reference state that underlies the tables of Gibbs formation energies allows for the direct exchange of model parameters between different models. These tables should be used if they are available.

This often leads to very large capacities because most Gibbs formation energies are rather low. To avoid numerical problems, the models can be scaled before simulation and analysis by applying the above corollary. This scaling of the model can be performed automatically by the modeling or simulation tool.

5. Thermokinetic Modeling

5.4.3. Translation of the Stoichiometric Coefficients

The thermokinetic modeling formalism explicitly distinguishes the stoichiometric matrices of reactants and products. In the thermodynamic modeling formalism, this distinction is not necessary because the information is coded in the expression for the thermodynamic resistances.

For example, the ideal thermodynamic resistance function for mass-action kinetics depends on the thermodynamic forces exerted by reactants and products∆µE/P,j =P

i∈E/PjνE/P,ijµi,j that depend on the stoichiometric coefficients of reactants and products (see Definition 3.22, p. 47).

The explicit introduction of NE/P and SE/P in TKM allows for a further translation operation.

Two reactions with the same overall stoichiometry may differ in their reactant and product stoichiometry. Consider the reaction from A to B catalyzed by X. Its stoichiometry can be described by two equivalent variants: (1) The reaction equation A+X B+X shows that X is a catalyst of the reaction. (2) The formulationA B has the same overall stoichiometry but does not refer to the catalyst X. Assume mass-action kinetics and small concentrations of A and B such that saturation effects of the catalyst X can be neglected. Then, the kinetics of variant (1) can be described by a constant thermokinetic resistance R1 = const, whereas the resistance of the variant (2) is ξX-dependent in order to model the catalyzing activity of X, i. e.R2 is proportional toξX−1. From a formal point of view, both model variants are equivalent.

Depending on the context, the modeler will prefer variant (1) or (2).

In the following corollary we establish the general rules that allow transforming such different representations into each other. These rules are based on a translation of the stoichiometric coefficients.

Corollary 5.37 (Translation of stoichiometric coefficients). Let M be a thermokinetic model (see Definition 5.29, p. 100) and let δN ∈ Ri0×j0 and δS ∈ Ri0,e×j0. Then, the solution of the system Mˆ with

E/P =NE/P +δN, SˆE/P =SE/P +δS, Nˆe=Ne, ˆ

c( ˆξ,ξˆe) = c( ˆξ,ξˆe), R( ˆˆ ξ,ξˆe) = R( ˆξ,ξˆe)◦δN#ξ◦δS#ξe, ξˆ00 is equivalent to that of M with

ˆ

c=c, ξˆ=ξ, ξˆee,

Jˆ=J, Fˆ =F ◦δN#ξ◦δS#ξe, Jˆe=Je.

Proof. We compare the model equations of M and Mˆ (see Definition 5.29, p. 100) under appli-cation of the above transformation rules.

Since N =NP −NE = ˆN andS =SP−SE = ˆS, the mole balance equations are equal c˙ˆ= ˙c. Since cˆ= c, ξˆ= ξ and ξˆe = ξe, the equations ˆc= ˆc( ˆξ,ξˆe) and c = c(ξ, ξe) are also equivalent.

Further, we have thatFˆ = ˆNE# ˆξ◦SˆE# ˆξe−NˆP# ˆξ◦SˆP# ˆξe=NE#ξ◦δN#ξ◦SEe◦δS#ξe− NP#ξ ◦ δN#ξ ◦ SPe ◦ δS#ξe = δN#ξ ◦ δS#ξ ◦ (NE#ξ ◦ SEE −NP#ξ ◦SPP) = δN#ξ◦δS#ξ◦F. Thus, the forces follow the given transformation rule. With this result it

5. Thermokinetic Modeling

follows that RˆJˆ= ˆF is equivalent to R J = F. Thus, all four parts of the model equations given in Definition 5.29 are equivalent.

Example 5.38(Catalyzed reaction).Let the reactionA+X B+Xhave a constant resistance R. The species X has an overall stoichiometric coefficient of zero (νXX,P −νX,E = 0) and is therefore a catalyst of the reaction. With the ordering(A, B, X), we haveNET = (1,0,1)and NPT = (0,1,1). By using δNT = (0,0,−1) we get the reaction A B with NˆET = (1,0,0), NˆPT = (0,1,0)and a resistanceRˆ=R ξX−1. Thus, the effect of the catalyst could be moved from the stoichiometric submodel to the kinetic submodel. Observe that in the case of a constant CX the model Mˆ can be reduced, since the concentration and the thermokinetic potential of the catalyst are constant: ξX =Cx−1cx = const. In the translated model ξX, can be treated as a model parameter and not as a state variable.