reaction extent or advancement of the reaction

75
arXiv:2111.05253v1 [physics.chem-ph] 8 Oct 2021 Reaction extent or advancement of the reaction: A new general definition Vilmos G´ asp´ ar and J´ anos T´ oth ,,§ Laboratory of Nonlinear Chemical Dynamics, Institute of Chemistry, ELTE E¨ otv¨ os Lor´ and University, Budapest, Hungary Budapest University of Technology and Economics, Department of Analysis, Budapest, Hungary §Chemical Kinetics Laboratory, Institute of Chemistry, ELTE E¨otv¨ os Lor´ and University, Budapest, Hungary E-mail: [email protected] Abstract Following the classical works by De Donder, Aris and Croce we extend the gen- eral investigation of the concept of reaction extent to the case of arbitrary number of species and (reversible or irreversible) reaction steps, not restricted to mass-action type kinetics. In typical cases the reaction extent tends to infinity. However, we have defined quantities tending to 1, as it is assumed many times. We have calculated the reaction extent also for exotic (oscillatory and chaotic) reactions to see their meaning. Our approach tries to adhere to the customs of the chemists and to be mathematically correct, and avoids some generally followed bad practice—mainly in the educational literature. Based on the talk given at the 2nd International Conference on Reaction Kinetics, Mechanisms and Catalysis. 20–22 May 2021, Budapest, Hungary 1

Upload: others

Post on 28-Dec-2021

2 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Reaction extent or advancement of the reaction

arX

iv:2

111.

0525

3v1

[ph

ysic

s.ch

em-p

h] 8

Oct

202

1

Reaction extent or advancement of the reaction:

A new general definition†

Vilmos Gaspar‡ and Janos Toth∗,¶,§

‡Laboratory of Nonlinear Chemical Dynamics, Institute of Chemistry, ELTE Eotvos

Lorand University, Budapest, Hungary

¶Budapest University of Technology and Economics, Department of Analysis, Budapest,

Hungary

§Chemical Kinetics Laboratory, Institute of Chemistry, ELTE Eotvos Lorand University,

Budapest, Hungary

E-mail: [email protected]

Abstract

Following the classical works by De Donder, Aris and Croce we extend the gen-

eral investigation of the concept of reaction extent to the case of arbitrary number

of species and (reversible or irreversible) reaction steps, not restricted to mass-action

type kinetics. In typical cases the reaction extent tends to infinity. However, we have

defined quantities tending to 1, as it is assumed many times. We have calculated the

reaction extent also for exotic (oscillatory and chaotic) reactions to see their meaning.

Our approach tries to adhere to the customs of the chemists and to be mathematically

correct, and avoids some generally followed bad practice—mainly in the educational

literature.

†Based on the talk given at the 2nd International Conference on Reaction Kinetics, Mechanisms andCatalysis. 20–22 May 2021, Budapest, Hungary

1

Page 2: Reaction extent or advancement of the reaction

Contents

1. Introduction 3

2. The concept of reaction extent 6

2.1. Typical errors . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6

2.2. Motivation and fundamental definitions . . . . . . . . . . . . . . . . . . . . . 7

2.2.1. Simple, recurring examples . . . . . . . . . . . . . . . . . . . . . . . . 7

2.2.2. The general case . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8

2.3. Properties of the reaction extent . . . . . . . . . . . . . . . . . . . . . . . . . 14

3. What is it that tends to 1? 19

3.1. Scaling by the initial concentration . . . . . . . . . . . . . . . . . . . . . . . . 20

3.1.1. One reaction step . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20

3.1.2. Stoichiometric initial condition. Excess, scarcity . . . . . . . . . . . . 23

3.2. Scaling by the ,,maximum” . . . . . . . . . . . . . . . . . . . . . . . . . . . . 27

3.3. Detailed balanced reactions . . . . . . . . . . . . . . . . . . . . . . . . . . . . 28

3.2.1. Ratio of two reaction extents . . . . . . . . . . . . . . . . . . . . . . . 28

3.2.2. Let us multiply the ratios . . . . . . . . . . . . . . . . . . . . . . . . . 32

3.4. Reactions with an attractive stationary point . . . . . . . . . . . . . . . . . . 35

3.5. The difference. Have we found the real reaction extent? . . . . . . . . . . . . 36

3.6. Reactant and product species . . . . . . . . . . . . . . . . . . . . . . . . . . . 40

4. And what if the conditions are not fulfilled? 41

4.1. Multistationarity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 41

4.2. Oscillation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 43

4.2.1. The Lotka–Volterra reaction . . . . . . . . . . . . . . . . . . . . . . . 43

4.2.1.1. Irreversible case . . . . . . . . . . . . . . . . . . . . . . . . . 43

4.2.1.2. Reversible case, detailed balanced . . . . . . . . . . . . . . . 44

2

Page 3: Reaction extent or advancement of the reaction

4.2.1.3. Reversible case, not detailed balanced . . . . . . . . . . . . . 45

4.2.2. The Rabai reaction of pH oscillation . . . . . . . . . . . . . . . . . . . 46

4.3. Chaos . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 48

5. Discussion, outlook 49

Acknowledgement 52

6. Notations 52

7. Appendix 52

7.1. On the form of the induced kinetic differential equation . . . . . . . . . . . . 52

7.2. Critical review of the literature . . . . . . . . . . . . . . . . . . . . . . . . . . 55

7.2.1. Neglectors . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 56

7.2.2. Refusers . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 56

7.2.3. Confused . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 57

References 59

1. Introduction

In some books on reaction kinetics one can find one of the expressions reaction extent

(extent of reaction), progress of reaction, the advancement of the reaction, con-

version, or reaction coordinate. Some other, not less important books do not even

mention these, or they may only mention without using them. (As a closure of our work we

shall present the results of a detailed literature search to prove these statements.) We have

found that it is possible and useful to give a general definition of reaction extent and system-

atically study its qualitative and quantitative properties in spite of the controversial views

in the literature. We also emphasize that such investigations should precede discussions of

possible teaching the concept.

3

Page 4: Reaction extent or advancement of the reaction

After a few decades of the introduction1 of the concept of reaction extent Aris2 has shown

a few examples.

The classical textbook on thermodynamic by Callen3 also uses the term degree of reac-

tion.

Our main goal is to fit the definitions and statements into the modern theory of reaction

kinetics, also called chemical reaction network theory, thus we cannot accept the formulation

by Carst4 according to which reaction extent is a unifying basis for stoichiometry.

A large collection of related concepts has been collected by Dumon et al.5 They also

introduce the reaction advancement ratio: ∆ξ∆ξmax

without giving a clear definition of ∆ξmax.

(It will turn out that in cases when something similar can be defined one should speak of

sup instead of max.) And what is one supposed to make of the expression ,,ξ = 1 when the

reaction has advanced one go.”?

Peckham6 proposes that IUPAC should use the normalized extent of reaction ξξmax

. The

problem is that the definition ξmax :=n0m

γmcan almost never be applied.

Croce7does know that ”one degree of advancement variable has to be defined for each

elementary step”, and also that ”the degree of advancement variable is not explained for the

general case of multistep reactions”. She ”reduces” the reaction steps (that we do not think

appropriate to call elementary) to

M∑

m=1

(βm,r − αm,r)X(m) = 0 (r = 1, 2, . . . , R)

the (bad) consequences of which we shall show in the paper. She proposes to divide the

variable we call reaction extent by V and calls the result degree of advancement. One must

admit that this trick would make some of the formulas simpler. When applying the concept

to multistep reactions her most important goal is to find a differential equation for a scalar

valued variable, either exactly, or approximately, using the Quasy Steady State Hypothesis—

in the usual naive way, i.e. without rigorously applying Tikhonov’s theorem. Our goal is to

4

Page 5: Reaction extent or advancement of the reaction

accept and extend some of the statements of the (most over the average) paper by Croce7

and correct other statements or offer them to the mathematician as a clearly defined open

problem.

Although Vandezande etal.8 uses the relationship n(t) = n(0)+γξ close to our definition,

they utter unacceptable statements like (p. 1179) ,,Extent of reaction can be thought of as

a one-dimensional vector...”. They also have the view that formulas do depend on how the

reaction is written meaning that one can multiply the reaction step. That is not an acceptable

view if one has in mind kinetics as tho goal. Percent completion is also introduced there.

Instead of maxima they speak about extrema—not defined more clearly than the maximum

by Dumon et al.5

Moretti9 has also vague or unacceptable sentences. He says that the x-axis must vary

from 0 to the theoretical maximum value ξmax, with the equilibrium value in the range

0 < ξeqn < ξmax. Also: Of course, ξmax = 1 is possible only when the initial conditions are

stoichiometric.

In the present paper we do not touch a few seemingly closely related topics, and these

are as follows: potential surfaces, transition states, thermodynamics. Atomic fluxes (see

e.g.10) will not be treated either, as these concepts are unrelated. Reaction coordinate as

a quantity characterizing the changes of the bonds in a transition state complex belongs to

another area. Also, we remind the reader that the expression flux of a reaction is also

used to mean the reaction rate of a step, or their sum, as used by Craciun et al.11 Of the

many possible names we shall use reaction extent as we have done above.

The structure of our paper is as follows. Section 2 introduces the concept for reaction

networks of arbitrary complexity (any number of reaction steps and species, mass action

kinetics not assumed). As it is a usual requirement that the reaction extent tends to 1

when ”the reaction tends to its end”, we try to find quantities derived from our reaction

extent having this property in Section 3. Exotic reactions (those showing oscillatory or

chaotic behaviour) are treated in Section 4. Discussions and a list of Notations follow next.

5

Page 6: Reaction extent or advancement of the reaction

An Appendix contains some mathematical background that would retard the expansion

of the thoughts of chemical relevance. Also, we enumerate citations from textbooks and

monographs neglecting the concept of reaction extent, or wrongly or too specifically using

it.

2. The concept of reaction extent

Before being constructive we enumerate a few general errors or useless approaches.

2.1. Typical errors

1. Most of the existing recent definitions cannot be applied to more complex systems; it is

mainly the case of a single reaction step that is treated, see e.g. the paper by Peckham.6

(This is not so with the original definition by De Donder and Van Rysselberghe.1)

2. It is quite common and absolutely unnecessary to use differentials (denoted by δ, d,∆,

etc.) when introducing the concept. (Cf.12 p. 61.) There is nothing infinitesimal here.

3. It is not exceptional to use the concept before defining it.

4. Several authors provide an implicit definition (that does not really define the given

quantity) and cannot get out from the trap built by themselves. Again, see e.g. the

paper by Peckham.6

5. Nothing is stated about the (mathematically or chemically) relevant properties of the

reaction extent.

6. Having introduced the concept this or that way, it is never ever used any more in many

works. (E.g. Pilling and Seakins13 mentions it in the Introduction and they do not

use the concept any more; this goes against the Chekhov principle: the gun shown at

the beginning should be used during the play.)

6

Page 7: Reaction extent or advancement of the reaction

Our aim is to present a treatment devoid of all these errors or failures and to study the

reaction extent more deeply as usual. We emphasize that this is not a paper on chemical ed-

ucation, the time for such a paper is only ripe when the corresponding concept is scientifically

sound.

2.2. Motivation and fundamental definitions

Before giving a general definition we take a simple example.

2.2.1. Simple, recurring examples

Let us start with an example that may be deemed chemically oversimplified but not too

trivial, still simple enough so as not to be lost in the details.

As a strong simplification of reality, assume that water formation follows the simple

reaction step 2H2 + O2=2H2O. Let the forward step be represented in a more abstract

way: 2X+Y −−→ 2 Z. This means that we do not take into consideration either the atomic

structure of the species, or the realistic details of water formation. The number of species

(that will be denoted below as M) is 3, the number of (irreversible) reaction steps (to be

denoted below as R) is 1. The 3 × 1 stoichiometric matrix γ consisting of stoichiometric

numbers (differences of the coefficients of the species on the right and on the left side of

the reaction arrow) is:

−2

−1

2

. In case this step occurs five times the vector of the number

of individual species will change as follows:

NX −N0X

NY −N0Y

NZ −N0Z

= 5

−2

−1

2

,

where N0X is the number of molecules of species X at the beginning, and NX is the number

7

Page 8: Reaction extent or advancement of the reaction

of pieces of species X after five steps, and so on. If one has the reversible reaction step

2X + Y −−⇀↽−− 2 Z, and the backward reaction step takes place three times then the total

change is

NX

NY

NZ

N0X

N0Y

N0Z

= 5

−2

−1

2

+ 3

2

1

−2

. (1)

Note that the number of molecules and the occurrence of reaction steps are positive integers.

Here we follow more or less Kurtz14 and Toth et al.15 We cannot rely on a discrete state

deterministic model of reaction kinetics—that would be desirable—because such a model

does not exists as far as we know.

2.2.2. The general case

Before providing general definitions we mention that Dumon et el.5 formulated a series of

requirements that should be obeyed by a well-defined reaction extent. We are unable to

accept most of these requirements. Let us mention only one: the reaction extent should be

independent of the choice of stoichiometric coefficients (invariant under multiplication), i.e.

it should have the same value for the reaction 2H2 + O2 −−→ 2H2O and for the reaction

H2 +12O2 −−→ H2O. This is not an error of the mentioned authors. Our (different) point

of view is that the reaction extent is strongly connected to kinetics, and it is not a tool to

describe stoichiometry as some further authors4 also think.

Following the books by Feinberg16 and by Toth et al.15 we consider a complex chemical

reaction, simply reaction, or reaction network as a set consisting of reaction steps as

follows:{

M∑

m=1

αm,rX(m) −−→M∑

m=1

βm,rX(m) (r = 1, 2, . . . , R)

}

; (2)

where

1. the chemical species are X(1),X(2),. . . ,X(M);

8

Page 9: Reaction extent or advancement of the reaction

2. the reaction steps are numbered from 1 to R (M and R being positive integers);

3. α := [αm,r] and β := [βm,r] are M × R matrices of non-negative integer components

called stoichiometric coefficients, with the properties that ∀m∃r : βm,r 6= αm,r,

meaning that all the species take part in at least one reaction step, and ∀r∃m : βm,r 6=

αm,r, or: all the reaction steps do have some effect, and finally

4. γ := β −α is the stoichiometric matrix of stoichiometric numbers.

Instead of (2) some authors prefer writing this:

M∑

m=1

γm,rX(m) = 0 (r = 1, 2, . . . , R). (3)

This formulation immediately excludes reaction steps like X + Y −−→ 2Y (used e.g. to

describe a step in the Lotka–Volterra reaction), or reduces it to X −−→ Y. Similarly, an

autocatalytic step that may be worth being studied, see e.g. pp. 63 and 66,2 like X −−→ 2X

appears oversimplified as 0 −−→ X.

Another possibility is to exclude the empty complex, meaning that we get rid of the

possibility to simply represent in- and outflow with 0 −−→ X and X −−→ 0, respectively.

These last two examples evidently mean mass-creation and mass-destruction. If you

do not like these you should explicitly say so, because sometimes mass-creation and mass-

destruction is not so obvious as above, see the reaction

X −−→ Y + U, Z −−→ X+ U, Z −−→ U.

It may happen that one would like to exclude reaction steps with more than two (three)

molecules on the left side, such as

2MnO −4 + 6H+ + 5H2C2O4 = 2Mn +

2 + 8H2O+ 10CO2.

9

Page 10: Reaction extent or advancement of the reaction

But such steps do occur e.g. on page 1236 of Kovacs et al.17 when dealing with overall

reactions. The theory and applications of decomposition of overall reactions into elementary

steps17,18 would have been impossible without this framework.

Someone may be interested in complex chemical reactions only consisting of reversible

steps. Then, (s)he will write down all the reaction steps and the corresponding ,,anti-

reactions steps.

Taking into consideration restrictions of the above kinds usually does not make the math-

ematical treatment easier. Sometimes it needs hard work to figure out what they mean, as it

is in the case of mass conservation of models containing species without atomic structure,15,19

or in the case of the existence of oscillatory reactions.20

To sum up: an author has the right to make any restriction (s)he finds to be chemically

important, but (s)he should declare these restrictions at the outset (and certainly not in the

middle of a discussion).

Finally, we mention our restriction: all the steps in (2) are really present, i.e. they

proceed with positive rate whenever the species on their left side are present.

We assume throughout that pressure, temperature and volume (V ) are constant. With

these one can write the generalized form of (1) as

N1

N2

. . .

NM

N01

N02

. . .

N0M

=

R∑

r=1

γ1,r

γ2,r

. . .

γM,r

Wr,

or shortly

N−N0 = γW,

where component Wr of the vector W =

[

W1 W2 . . . WR

]⊤

gives the number of occur-

rences of the rth reaction step. Note that we do not speak about infinitesimal changes.

10

Page 11: Reaction extent or advancement of the reaction

(NX(m) and Nm will be used interchangeably.)

With a slight abuse of notation let W(t) be the number of occurrences of reaction steps

in the interval [0, t], a step function. Then:

N(t)−N0 = γW(t),

or turning to moles

n(t)− n0 =N(t)−N0

L= γ

W(t)

L= γξ(t),

where L is the Avogadro constant having the unit mol−1, and

n(t) :=N(t)

L,n0 :=

N0

L, ξ(t) :=

W(t)

L.

(We had to choose the less often used notation L to avoid mixing up with other notations.)

The relationship in concentrations can be expressed as

c(t)− c0 =n(t)− n0

V=

γ

Vξ(t), (4)

where V ∈ R+ is the volume of the reaction vessel, assumed as said above to be constant,

and c(t) := n(t)V

, c0 := n0

V. (Note that the component cX(m) or cm of c is traditionally denoted

in chemical textbooks as [X(m)], see e.g. Section 1.2 of13).

The concentration in Eq. (4) is again a step function; however, if the number of particles is

large, as very often it is, it may be considered to be a continuous (what is more: differentiable)

function.

Remember that the components of ξ(t) have the dimension amount of substance: mol.

Let us give a general formal explicit definition valid for any number of species and

reaction steps, and not restricted to mass action type kinetics. (However, a few

qualitative restrictions are usually made on the function rate, see15,16,21 of which we now only

11

Page 12: Reaction extent or advancement of the reaction

mention continuity.) In order to do so we rewrite the induced kinetic differential equation

c(t) = γrate(c(t)) (5)

together with the initial condition c(0) = c0 into an (equivalent) integral equation:

c(t)− c0 = γ

∫ t

0

rate(c(t)) dt. (6)

and for its physical counterpart defined for a scalar variable measured in seconds and having

coordinate values measured in moldm3 .)

The component rater of the vector rate provides the reaction rate of the rth reaction

step. Note that in the mass action case (5) specializes into

c = γk⊙ cα (7)

or, in coordinates

cm(t) =

R∑

r=1

γmrkr

M∏

p=1

cαp,r

p (m = 1, 2, · · · ,M),

where k is the vector of (positive) reaction rate coefficients kr, and in (7) we used the usual

vectorial operations, see e.g. Section 13.2 in Toth et al.15 Their use in formal reaction

kinetics has been initiated by Horn and Jackson.22

In accordance with what has been said up to now we introduce the definition.

Definition 1. The reaction extent or conversion of the complex chemical reaction or

reaction network (2) is the scalar variable, vector valued function defined by the formula

ξ(t) := V

∫ t

0

rate(c(s)) ds ; (8)

its derivative ξ(t) = V rate(c(t)) is usually called the rate of conversion.

12

Page 13: Reaction extent or advancement of the reaction

This reaction extent has been derived from the number of reaction steps in order to

reveal its connection to the concentrations. (In what follows we shall not use the expression

progress of reaction, and even less the reaction coordinate, but we accept the proposal

by Croce7) to call 1Vξ the degree of advancement.

One can formulate the following trivial consequences of the definition:

n = γξ, c =1

Vγξ c(t) =

1

Vγξ (9)

mentioned also by Laidler,23 sometimes as definitions, sometimes as statements (!). Note

that neither the rate of the reaction rate(c(t)) = ξ

V, nor the reaction extent ξ (fulfilling

thus one of the requirements formulated by5), nor the rate of conversion ξ depends on the

stoichiometric matrix γ.

What is wrong with the almost ubiquitous implicit ”definition” (4)? We show an example

to enlighten this.

Example 1. Consider the reaction

Xk1−−→ Y, X+ Y

k2−−→ 2Y

(expressing the fact that the same step occurs parallelly, with and without autocatalysis)

with

γ =

−1 −1

1 1

.

Then (6) specializes into

cX(t)− cX(0) = −∫ t

0

k1cX(s) ds−∫ t

0

k2cX(s)cY(s) ds = −ξ1(t) + ξ2(t)

V

cY(t)− cY(0) =

∫ t

0

k1cX(s) ds +

∫ t

0

k2cX(s)cY(s) ds =ξ1(t) + ξ2(t)

V,

and these two relations do not determine ξ1 and ξ2 individually, only their sum (even if one

13

Page 14: Reaction extent or advancement of the reaction

utilizes cY(t) = cX(0) + cY(0)− cX(t)). The problem originates in the fact that the reaction

steps are not linearly independent as reflected in the singularity of the matrix γ.

Example 2. If the reaction steps of a complex chemical reaction are independent the situ-

ation is better.

In some special cases there is a way of making the ”definition” (4) into a real, explicit

definition. Assume that R ≤ M , and that the stoichiometric matrix γ is of the full rank, i.e.

the reaction steps are independent. Then, one can rewrite (4) in two steps as follows:

γ⊤(c(t)− c0) =1

Vγ⊤γξ(t)

ξ(t) = V (γ⊤γ)−1γ⊤(c(t)− c0).

Now one can accept as a definition ξ(t) := V (γ⊤γ)−1γ⊤(c(t) − c0). Nevertheless, in this

special case (6) implies

γ⊤(c(t)− c0) = γ⊤γ

∫ t

0

rate(c(s)) ds

and

(γ⊤γ)−1γ⊤(c(t)− c0) =1

Vξ(t) =

∫ t

0

rate(c(s)) ds,

thus this definition is the same as the one in (8). This derivation can always be done if

R = 1—a not-so-interesting case. Also, R ≤ M does not happen very often. In case of

combustion reactions Law’s law24 (see page 11) states that R ≈ 5M.

2.3. Properties of the reaction extent

The usual assumptions on the vector valued function rate are as follows, see.15,21

1. All of its components are continuously differentiable functions defined on RM taking

only non-negative values. This is usual, e.g. in the case of mass action kinetics, but—

14

Page 15: Reaction extent or advancement of the reaction

with some restrictions—also in the case when the reaction rates are rational functions

as in the case of Michaelis–Menten or Holling type kinetics, see e.g.23,25–27

2. The value of rater(c) is zero if and only if some of the species needed for the rth

reaction step to take place is missing, i.e. for some m : αm,r > 0 and cm = 0 (see p.

613, Condition 1 in21). We shall say in this case that reaction step r cannot start

from the concentration vector c.

The second assumption implies—even in the general case, i.e. without restriction to the mass

action type kinetics—that rater(c) > 0 if all the necessary (reactant, see below) species are

present initially: αm,r > 0 =⇒ cm > 0.

Let us sum up the relevant qualitative characteristics of the reaction extent. In case

there is no misunderstanding, we shall use the same notation for a physical

quantity (expressed by its numerical value and unit) and for its numerical value,

as it is customary.

Theorem 1.

1. The domain of ξ is the same as that of c.

2. Both c and ξ ∈ C2(J,RR); with some open interval J ⊂ R such that 0 ∈ J.

3. ξ obeys the following initial value problem:

ξ(t) = V rate(c0 +1

Vγξ(t)), ξ(0) = 0. (10)

4. The velocity vector of the reaction extent (also called the rate of conversion) points

into the interior of the first orthant at the beginning, and this property is kept for all

times in the domain of ξ.

5. The components of ξ are either positive, strictly monotonously increasing functions,

or constant zero. The second case implies that some of the reaction steps cannot start

15

Page 16: Reaction extent or advancement of the reaction

at the beginning.

Proof.

1. This follows from (8).

2. As c is the solution of a differential equation with a continuously differentiable right-

hand-side, it is twice continuously differentiable. The definition (8) implies ξ ∈

C2(J,RR).

3. Take the derivative of (8) and use (6). The initial condition also comes from (8).

4. The derivative of ξr is non-negative.

5. If the derivative of ξr is positive, then ξr is strictly monotonously increasing.

If it is zero at some time t0, then for some m : αm,r > 0 and cm(t0) = 0. However,

cm(t0) = 0 can only hold if cm(0) = 0 held at the beginning, because an initially

positive concentration cannot turn to zero, see Theorem 1 on p. 617 in the book by

Volpert and Hudyaev.21 But then cm(t) = 0 for all t ∈ R+0 . Thus in this case for all

t ∈ R+0 : ξr(t) = 0, this, however, together with the initial condition ξr(0) = 0 implies

that for all t ∈ R+0 : ξr(t) = 0; therefore ξr is (not strictly) monotone, it is constant

zero in this case.

Let us make a few remarks.

The last property mentioned in the Theorem can be realized with limt→+∞ ξ(t) = +∞

(the simplest example for this being 0 −−→ X), or with a finite positive value of limt→+∞ ξ(t),

see the example X −−→ Y −−→ Z below.

Eq. (10) shows that we would have simpler formulas if we used ξ(t)V

as proposed by Aris2

on p. 44, but we keep the more complicated form, as is is more often used.

In the mass action case both c and ξ are infinitely many times differentiable.

16

Page 17: Reaction extent or advancement of the reaction

If one uses a kinetics different from the mass action type not fulfilling Conditions 1 and

2, or—as Pota28 has shown—if one applies an approximation, then it may happen that some

of the initially positive concentrations turn to zero.

The condition in the last sentence of Theorem 1 is only necessary but not sufficient as the

example below shows. But before that we need a technical remark. We label the first axis

(usually: horizontal) in the figures with t/s, to emphasize that the figure shows a relation

between pure numbers and not between physical quantities. Other axis labels are formed in

a similar way.

And now, let us start the consecutive reaction Xk1−−→ Y

k2−−→ Z from the initial vector of

concentrations

[

c0X 0 0

]⊤

, and suppose k1 6= k2. Then the second step cannot start at the

beginning, yet the second reaction extent is positive for all positive times as the solution of

the equations

ξ1 = V k1

(

c0X − ξ1V

)

, ξ2 = V k2

(

0 +ξ1 − ξ2

V

)

(11)

being

ξ1(t) = V (c0X − e−k1t),

ξ2(t) = V(k1(c

0X − e−k2t) + k2(−c0X + e−k1t))

k1 − k2.

shows.

Positivity also follows from the fact that the velocity vector of the differential equation

(11) points inward, into the interior of the first quadrant.

Note that ξ2/mol in Fig. 1 has an inflection point, because its second derivative is zero at

some positive time T/s for all choices of the reaction rate coefficients, and the third derivative

is not zero at T/s. ξ1/mol in Fig. 1 is concave from below no matter what the reaction rate

coefficients are.

One should take into consideration that although in the practically interesting cases

the number of equations in (10) is larger than those in (5), the equations are of a simpler

17

Page 18: Reaction extent or advancement of the reaction

2 4 6 8 10t /s

0.2

0.4

0.6

0.8

1.0

reaction

extents

/mol

ξ1(t)

ξ2(t)

Figure 1: Reaction extents in the simple consecutive reaction Xk1−−→ Y

k2−−→ Z.k1 = 1 1

s, k2 = 2 1

s, c0X = 1 mol

dm3 , c0Y = c0Z = 0 moldm3 , V = 1 dm3.

structure.

Monotonicity mentioned in the Theorem implies that all the components of the reaction

extent do have a finite or infinite limit as t tends to sup(Dom(c)) =: t∗; a finite or infinite

time. It is an interesting open question: when does a coordinate of the reaction extent vector

tend to infinity?

As the emphatic closure of this series of remarks we mention that strict monotonicity of

the reaction steps shows that the reaction steps do not cease to proceed during equilibrium,

if they have started at all. This is the meaning of dynamic equilibrium taught also in

high schools. This important fact

• is independent on the form of kinetics,

• is a property of the deterministic model of reaction kinetics,

• and no reference to thermodynamics or statistical physics is needed to prove

it.

Example 3. In case when the domain of c is a proper subset of the non-negative real

numbers, ξ has the same property. Let us consider the induced kinetic differential equation

18

Page 19: Reaction extent or advancement of the reaction

c = kc2 of the quadratic autocatalytic reaction 2Xk−−→ 3X with the initial condition

c(0) = c0 > 0. Then, c(t) = c01−kc0t

,(

t ∈[

0, 1kc0

[

( [0,+∞[)

. Now (10) specializes into

ξ = V k(c0 + ξ/V )2, ξ(0) = 0 having the solution ξ(t) = V c0kt

1−kc0t,(

t ∈[

0, 1kc0

[)

, thus ξ

blows up at the same time (t∗ := 1kc0

) when c.

Definitions and a few statements about blow-up in kinetic differential equations are given

in the works by Csikja et al.29,30

As the right hand side of (10) is a polynomial (vector-vector) function, the following

problem naturally arises.

Problem 1. What are the necessary or sufficient conditions that (10) is Hungarian,31

or kinetic (to use the original name)? (For the definition and characterization of kinetic

differential equations in the class of polynomial equations see the Appendix, Lemma 1, for

related results see also32,33)

In most of the examples below the differential equations for the reaction extent are

Hungarian, thus it will suffice to show a case when it is not.

Example 4. Consider the equations for the reaction extents in the case of the reaction

Xk1−−→ 0, X

k2−−→ 2X:

ξ1 = V k1(c0 +ξ2 − ξ1

V), ξ2 = V k2(c0 +

ξ2 − ξ1V

) = V k2c0 −k2ξ1 + k2ξ2. (12)

The boxed term shows that (12) is not Hungarian.

3. What is it that tends to 1?

Our interest up to this point was the number of occurrences of reaction steps. However,

many authors think it to be useful and visually attractive that the ”reaction extent tends to

1 when the reaction tends to its end,” see e.g. Fig. 1.34 Peckham6 also argues for [0,1].

19

Page 20: Reaction extent or advancement of the reaction

Another approach is given by Moretti,9 Dumon et al.5 and others to introduce the re-

action advancement ratio ξξmax

and state that this ratio is always between 0 and 1. It is

really instructive that Atkins35 pp. 272–276 shows a figure ofG versus ∆ξ, where the first

axis is labeled from zero to one. The next edition36 however on pp. 216–217 have changed

their graphs to leave the first axis without 1 as an upper bound.

Being loyal to the usual belief,6,37 we are looking for quantities (pure numbers) tending

to 1 as e.g. t → +∞. Scaling might help find such quantities.

3.1. Scaling by the initial concentration

3.1.1. One reaction step

Now we are descending from the height of generality by considering a single irreversible step

(R = 1), assuming that the kinetics is of the mass action type. Thus, the reaction is

M∑

m=1

αmX(m)k−−→

M∑

m=1

βmX(m) (13)

therefore one has for the reaction extent

ξ(t) = V k

c01

c02

. . .

c0M

+

γ1

γ2

. . .

γM

ξ(t)

V

α

= V k

M∏

m=1

(

c0m + γmξ(t)

V

)αm

, ξ(0) = 0; (14)

with γm := βm − αm.

Theorem 2.

1. If the reaction step (13) cannot start then ξ(t) = 0 for all non-negative real times t:

∃m : (αm 6= 0 & c0m = 0) =⇒ ∀t ∈ R : ξ(t) = 0. (15)

20

Page 21: Reaction extent or advancement of the reaction

2. If the reaction step (13) can start and all the species are produced, then ξ(t) tends to

infinity (blow-up included):

∀m : γm := βm − αm > 0 =⇒ limt→t∗

ξ(t) = +∞, (16)

where t∗ := sup(Dom(ξ)).

3. If some of the species decreases: ∃m : γm < 0, then

limt→+∞

ξ(t) = min

{

−V c0mγm

; γm < 0

}

. (17)

Proof.

1. The condition in (15) implies that ∀t ∈ R+0 and for all such m ∈ M for which αm 6= 0

one has cm(t) = 0, thus the right-hand-side of (14) is zero, and this, together with the

initial condition ξ(0) = 0 implies the statement.

2. The condition in (16) implies that the right-hand-side of (14) is positive, and

∀t ∈ Dom(ξ) : ξ(t) = k

M∑

m=1

αmγmc0m + γmξ/V

M∏

p=1

(c0p + γpξ(t)/V )αp > 0,

thus ξ is not only monotonously increasing but also strictly convex from below implying

the statement.

3. First of all, if one can show that

sup(Dom(c)) = +∞, (18)

then sup(Dom(ξ)) = sup(Dom(c)) = +∞ follows. If all the species decrease then (18)

21

Page 22: Reaction extent or advancement of the reaction

is true. If X(m) decreases and X(p) increases then − cmγm

+ cpγp

= 0 implies

−cm(t)

γm+

cp(t)

γp= −cm(0)

γm+

cp(0)

γp,

which—together with the non-negativity of the concentrations see pp. 613–615 in21—

proves this part of the statement.

The last case of the Theorem includes the situation when the number of species is 1. The

simplest example to illustrate this is X −−→ 0.

The example X −−→ Y with c0X = 0, c0Y > 0 shows that a species can have positive

concentration for all positive times in a reaction where ”none of the steps” can start.

To have a dimensionless quantity, Aris2 proposes the division by the sum of the initial

concentrations in a special case on p. 47.

The table below shows a series of examples illustrating the different possibilities.

Table 1: Different cases of reaction extent

Step c0 ξ = ξ(t) = Case

X1−−→ 2X 0 V ξ 0 (15)

X1−−→ 2X 1 V (1 + ξ) V (et − 1) (16)

2X1−−→ 3X 1 V (1 + ξ/V )2 V t

1−t(16)

2X + Y1−−→ 2 Z V (c0X − 2ξ/V )2(c0Y − ξ/V ) (17)

Example 5. Here we give the details of the last example. In the case of the reaction

2X +Y1−−→ 2 Z mimicking water formation one has the following quantities: R := 1,M :=

3,X := H2,Y := O2,Z := H2O. Furthermore,

α =

2

1

0

, β =

0

0

2

, γ =

−2

−1

2

.

22

Page 23: Reaction extent or advancement of the reaction

The initial value problem to describe the time evolution of reaction extent is

ξ = V k

c0X

c0Y

c0Z

+

−2

−1

2

ξ

V

2

1

0

= V k(c0X − 2ξ

V)2(c0Y − ξ

V), ξ(0) = 0. (19)

At the moment we can only provide the inverse of the solution to (19). Different initial

conditions lead to different results, see Figs. 2–4.

Obviously, the last part of the theorem is of main practical use. For this case one has

the following statement.

Corollary 1. In the third case of the Theorem above dividing ξ by the initial concentration

scaled by the quantity −γm/V we obtain a (pure) number tending to 1, as t tends to infinity:

limt→+∞ξ(t)

V c0m/(−γm)= 1.

3.1.2. Stoichiometric initial condition. Excess, scarcity

Alas! Before studying the mentioned figures, however, we need some definitions in order to

avoid the sin of using a concept without having defined it. The concepts of stoichiometric

initial condition and initial stoichiometric excess are elsewhere often used and never

defined.

Definition 2. Consider the induced kinetic differential equation (5) of the reaction (2) with

the initial condition c(0) = c0 6= 0. This initial condition is said to be a stoichiometric

initial condition (and c0 is a stoichiometric initial concentration), if for all such m ∈

M, r ∈ R for which γm,r < 0 the ratios c0m−γm,r

are independent from m and r. If the ratios are

independent of r, but for some p ∈ M the ratioc0p

−γp,ris larger than the others, then X(p)

is said to be in initial stoichiometric excess, or it is in excess initially. If the ratios are

independent of r, but for some p ∈ M the ratioc0p

−γp,ris smaller than the others, then X(p) is

23

Page 24: Reaction extent or advancement of the reaction

said to be in initial stoichiometric scarcity, or it is in scarcity initially. The last notion

is mathematically valid, but in such cases one prefers saying that all the other species are in

excess.

Note that in combustion theory the expressions stoichiometric, fuel lean and fuel rich

are used in the same sense, see page 115 of.10

Although the above classes of initial conditions are disjoint, they do not cover all the

possible cases. A series of examples will shed some light on the fine points of the definition.

Example 6.

1. 2X + Y −−→ 0: here γX = −2, γY = −1, thus the stoichiometric initial concentrations

are of the form

[

2c0Y c0Y

]⊤

with arbitrary positive concentration c0Y.

2. 2X+Y −−→ 2 Z: here γX = −2, γY = −1, thus the stoichiometric initial concentrations

are of the form

[

2c0Y c0Y c0Z

]⊤

with arbitrary non-negative concentrations c0Y > 0, c0Z ≥

0.

3. 2X −−→ X or X −−→ 0: here c0X > 0 is always a stoichiometric initial concentration,

because in this case the number of species is 1, therefore the constancy criterion is

automatically fulfilled. (We admit that this case is trivial or uninteresting from the

chemist’s point of view.)

4. X −−→ 2X, X+Y −−→ 2Y, Y −−→ 0: any initial concentration would be stochiometric

for the second step, and also for the third one, considered separately. However, for the

first step the definition cannot be applied.

5. 5X+Y −−→ 3X+ 4Y: as γY > 0, the definition only applies to X, and c0X > 0 can be

arbitrary to be stoichiometric, if c0Y > 0 also holds.

6. 2X + 3Y −−→ X + Y:

[

c0X 2c0X

]⊤

are the stoichiometric initial concentrations with

arbitrary c0X > 0.

24

Page 25: Reaction extent or advancement of the reaction

7. X + 2Y −−→ Z, 2X + Y −−→ Z: the conditionsc0X

−1=

c0Y

−2and

c0X

−2=

c0Y

−1can be

fulfilled with no positive initial concentration vector, therefore no stoichiometric initial

condition can be found. Funny to realize, ifc0Y

2< c0X < 2c0Y, then X is present in

both stoichiometric excess and scarcity at the same time. As it would be desirable

to eliminate such cases, the problem arises: determine those reactions for which this

cannot happen.

8.∑M

m=1 αmX(m) −−⇀↽−−∑M

m=1 βmX(m) : If the species are ordered in such a way that

γ1, γ2, . . . , γK < 0; γK+1, γK+2, . . . , γM ≥ 0,

(for some 1 ≤ K ≤ M) then

[

−γ1d0 . . . −γKd0 c0K+1 . . . c0M

]⊤

is a stoichiomet-

ric initial concentration vector for any numbers d0 > 0; c0K+1 ≥ 0, . . . , c0M ≥ 0.

9. It does not seem to be easy to give a similar general representation of the stoichiometric

initial concentrations in the general case (2).

0 20 40 60t /s

0.2

0.4

0.6

0.8

1.0

ξ(t)/mol

Figure 2: k = 1.0 dm6

mol2s, c0X/2 = 0.7 mol

dm3 < c0Y = 1.2 moldm3 , c0Z = 3 mol

dm3 , V = 1 dm3. Y isin stoichiometric excess.

25

Page 26: Reaction extent or advancement of the reaction

0 100 200 300 400t /s

0.2

0.4

0.6

0.8

1.0

ξ(t)/mol

Figure 3: k = 1.0 dm6

mol2s, c0X/2 = 0.6 mol

dm3 = c0Y, c0Z = 3 mol

dm3 , V = 1 dm3. Stoichiometricinitial condition; slow convergence.

0 20 40 60t /s

0.2

0.4

0.6

0.8

1.0

ξ(t)/mol

Figure 4: k = 1.0 dm6

mol2s, c0X/2 = 0.6 mol

dm3 > c0Y = 0.5 moldm3 , c0Z = 3 mol

dm3 , V = 1 dm3. X is instoichiometric excess.

At this point it may not be obvious how to generalize Corollary 1 for more complicated

cases. We shall see this below.

Peckham6 also argues for [0,1], and under stoichiometric initial conditions he gives a

definition of ξmax and proposes to use ξ(t)ξmax

—in extremely special cases.

Note that instead of saying that the scaling factor is some initial concentration in the

26

Page 27: Reaction extent or advancement of the reaction

0 20 40 60t /s

0.2

0.4

0.6

0.8

1.0

ξ (t)

VcX0 2

0 20 40 60t /s

0.2

0.4

0.6

0.8

1.0

ξ (t)

VcX0 2

0 20 40 60t /s

0.2

0.4

0.6

0.8

1.0

ξ (t)

VcY0

Figure 5: Scaled reaction extents tend to 1. The reaction rate coefficient and the initial dataare the same as in Figs. 2–4.

third case of Theorem 2 one can equally well say that the divisor is the stationary value of the

reaction extent, as it is in the cases in the figures: V c0X/2, V c0X/2 = V c0Y, V c0Y, respectively.

This result will come handy below.

3.2. Scaling by the ,,maximum”

In cases when ξ∗ := limt→+∞ ξ(t) is finite, then ξ∗ = sup{ξ(t); t ∈ R}, thus ξ∗ may be

identified (mathematically incorrectly) with ξmax, and surely limt→+∞ξ(t)ξ∗

= 1. That is the

27

Page 28: Reaction extent or advancement of the reaction

procedure applied by most authors.5,8,9

3.3. Detailed balanced reactions

Definition 3. The complex chemical reaction

M∑

m=1

αm,rX(m)kr−−⇀↽−−k−r

M∑

m=1

βm,rX(m) (r = 1, 2, . . . , R); (20)

endowed with mass action kinetics is said to be conditionally detailed balanced at the

positive stationary point c∗ if

kr(c∗)α.,r = k−r(c

∗)β.,r (21)

holds. It is (unconditionally) detailed balanced if (21) holds for any choice of (positive)

reaction rate coefficients.

Note that all the steps in (20) are reversible. Furthermore, in such cases the reaction steps

are indexed by r and −r, and it is our choice in which order the forward and backward steps

are written, expressing the fact that ”backward” and ”forward” has no physical meaning.

3.2.1. Ratio of two reaction extents

Suppose we have a single reversible reaction step:

M∑

m=1

αmX(m) −−⇀↽−−M∑

m=1

βmX(m). (22)

(This reaction is unconditionally detailed balanced because the number of the forward and

backward pairs is 1.) Then the initial value problem for the reaction extents is as follows.

ξ1 = V k1

M∏

m=1

(c0m + γm(ξ1 − ξ−1)/V )αm , ξ1(0) = 0,

ξ−1 = V k−1

M∏

m=1

(c0m + γm(ξ1 − ξ−1)/V )βm, ξ−1(0) = 0,

28

Page 29: Reaction extent or advancement of the reaction

where γm := βm − αm.

Proposition 1. Under the above conditions one has limt→+∞ξ1(t)ξ−1(t)

= 1.

Proof. First of all, let us note that the concentrations (and therefore the reaction extents)

do not blow up,38 i.e. sup(Dom(c)) = +∞.

Next, as limt→+∞ ξ1(t) = +∞ and limt→+∞ ξ−1(t) = +∞ and limt→+∞ ξ1(t) = V k1(c∗)α

and limt→+∞ ξ−1(t) = V k−1(c∗)β = V k1(c

∗)α one can apply l’Hospital’s Rule to get the

desired result.

Note that initially one only knows that it is the derivatives of the reaction extents that

have the same value at equilibrium.

Example 7. Consider the reversible reaction 2X + Y −−⇀↽−− 2 Z for ”water formation” with

the data

k1 = 1dm6

mol2s, k−1 = 1

dm3

mol s, c0Y = 1

mol

dm3 , c0Z = 1

mol

dm3 . (23)

• If c0X = 3 moldm3 , then X is in excess initially;

• if c0X = 2 moldm3 , then one has a stoichiometric initial condition;

• if c0X = 1 moldm3 , or c0X = 1

2moldm3 , then Y is in excess initially.

Note that it is not the excess or scarcity that is relevant, see Conjecture 1 below. The initial

rates of the forward and backward reactions are as follows:

1 · 9 · 1 > 1 · 1,

1 · 4 · 1 > 1 · 1,

1 · 1 · 1 = 1 · 1,

1 · 1/4 · 1 < 1 · 1.

The results are in accordance with Conjecture 1 below and can be seen in Figs. 6 and 7.

29

Page 30: Reaction extent or advancement of the reaction

0.0 0.2 0.4 0.6 0.8t /s

0.5

1.0

1.5

2.0

2.5

ξ1(t)/ξ-1(�)

l����

ratio

0.0 0.2 0.4 0.6 0.8t /s

0.5

1.0

1.5

2.0

2.5

ξ1(t)/ξ-1(t)

limit

ratio

Figure 6: The ratio ξ1(t)ξ−1(t)

tending to 1 from above: X is in excess at the top and the initialcondition is stoichiometric at bottom. Data are given in the text.

30

Page 31: Reaction extent or advancement of the reaction

0.0 0.2 0.4 0.6 0.8t /s

0.2

0.4

0.6

0.8

1.0

1.2

ξ1(t)/ξ-1(t)

limit

ratio

0.0 0.2 0.4 0.6 0.8t /s

0.2

0.4

0.6

0.8

1.0

1.2

ξ1(t)/ξ-1(t)

limit

ratio

Figure 7: The ratio ξ1(t)ξ−1(t)

being constant and tending to 1 from below. Y is in excess in bothcases. Data are given in the text.

31

Page 32: Reaction extent or advancement of the reaction

Now we formulate our experience collected on several models as a conjecture. Consider

reaction (22).

Conjecture 1. If k1(c0)α > k−1(c

0)β (or k1(c0)α < k−1(c

0)β) then the ratio ξ1(t)ξ−1(t)

tends to

its limit in strictly monotonously decreasing (increasing) way. If k1(c0)α = k−1(c

0)β, then

the ratio is constant 1, and c0 = c∗.

Convergence has been proved above. The ratio at t = 0 is not defined, but the limit of

the ratio when t → +0 can be calculated using the l’Hospital Rule as

limt→+0

ξ1(t)

ξ−1(t)= lim

t→+0

ξ1(t)

ξ−1(t)=

k1k−1

(c0)α−β,

and k1k−1

(c0)α−β < 1 is equivalent to saying that k1(c0)α < k−1(c

0)β, and the remaining cases

can be similarly treated.

What is left to prove is that k1(c0)α < k−1(c

0)β implies that ξ1ξ−1

is strictly monotonously

increasing.

The meaning of the above conjecture is quite obvious: if the forward reaction proceeds

slower at the beginning than the backward reaction, then the limit 1 of the ratio is approached

from below etc. The concept of stoichiometric initial condition seem to play no role here.

Instead of studying other simple reactions we generalize the above result.

3.2.2. Let us multiply the ratios

Theorem 3. If the complex chemical reaction (20) is detailed balanced, i.e. (21) is fulfilled,

then limt→+∞∏R

r=1ξr(t)ξ−r(t)

= 1.

Proof. It is similar to that of Proposition 1: all the separate factors tend to 1 as t → +∞.

Example 8. Consider the example in Fig. 8 that is not unconditionally detailed balanced.

In this case the condition of detailed balancing to hold is

k−1k−2k−3 = k1k2k3, k3k5 = k−3k−5, k3k4 = k−3k−4

32

Page 33: Reaction extent or advancement of the reaction

k1

k-3

k-2

k4

k

k5

k-5

X +Y

2 ZY +Z

2 X

3 XX +Z

Figure 8: A triangle coupled with a Wegscheider-type reaction

33

Page 34: Reaction extent or advancement of the reaction

as applying either the circuit conditions and the spanning forest conditions39 or use

the algebraic condition coming from the Fredholm alternative theorem (see p. 133 of15)

gives. (The tedious calculations providing the above equalities can be carried out using the

function DetailedBalancedQ of the package ReactionKinetics, supplement to the book

by Toth et al.15)

0 1 2 3 4t /s

0.5

1.0

1.5

2.0

r=1

R

r (t)/ -r (t)

Figure 9: Convergence of the product of ratios from above in case of reaction in Fig. 8 withdata implying detailed balancing, as follows: k1 = 1 dm6

mol2s, k−1 = 2 dm3

mol·s , k2 = 3 dm3

mol·s ,

k−2 = 1 dm3

mol·s , k3 = 13

dm3

mol·s , k−3 = 12

dm6

mol2s, k4 = 3 dm3

mol·s , k−4 = 2 1s, k5 = 3 dm6

mol2s,

k−5 = 2 dm3

mol·s , c0X = c0Y = c0Z = 1 mol

dm3 .

Even if the reaction is not detailed balanced, it will not blow up,38 the stationary point

exists and is unique, and the product of the ratios will converge; although the limit will be

different from 1, see Fig. 10. Let us use l’Hospital’s Rule to show this in the case of a single

factor:

limt→+∞

ξr(t)

ξ−r(t)= lim

t→+∞

ξr(t)

ξ−r(t)=

ξr(+∞)

ξ−r(+∞)=

krk−r

cα(.,r)∗

cβ(.,r)∗

.

One can form quite different and complicated expressions that tend to 1, e.g. 1+ ξ2−ξ3ξ1

is

also a possible choice in case of R = 3. We may as well start from the pseudo-Helmholtz

34

Page 35: Reaction extent or advancement of the reaction

0 50 100 150 200t /s

0.5

1.0

1.5

2.0

r=1

R

r (t)/ -r (t)

Figure 10: Convergence of the product of ratios in case of reaction in Fig. 8 with datanot fulfilling detailed balance, as follows: k1 = 1 dm6

mol2s, k−1 = 2 dm3

mol·s , k2 = 3 dm3

mol·s ,

k−2 = 1 dm3

mol·s , k3 = 1 dm3

mol·s , k−3 = 1 dm6

mol2s, k4 = 3 dm3

mol·s , k−4 = 2 1s, k5 = 3 dm6

mol2s,

k−5 = 2 dm3

mol·s , c0X = c0Y = c0Z = 1 mol

dm3 .

function introduced on p. 98 by Horn and Jackson22

H(c) := c⊤ · (ln(c)− ln(c∗)− 1)

(see also Liptak et al.40) and rewrite it in terms of the reaction extent H(ξ) := H(c0+γξ/V ),

finally divide it with the stationary value: Ψ(ξ(t)) := H(ξ(t))

H(ξ∗). The function Ψ has been defined

with a non-zero denominator and is tending to 1 as t → +∞ (at least in the case of weakly

reversible zero deficiency reactions endowed with mass action kinetics). As this function uses

all the reaction extents we shall often use it below.

3.4. Reactions with an attractive stationary point

All our observations can be summarized in the trivial proposition below. Before stating it

we need a definition for general ordinary differential equations.

35

Page 36: Reaction extent or advancement of the reaction

Let M ∈ N, f ∈ C1(RM ,RM) and consider the initial value problem

x(t) = f(x(t)),x(0) = x0 (∈ RM). (24)

Definition 4. The stationary point x∗ of (24) is said to be attractive, if all the solutions

starting from a neighbourhood of x∗ are defined for all positive time, and tend to it as

t → +∞.

Note that attractiveness is less then being asymptotically stable and different from being

stable. The neighbourhood mentioned in the definition is the domain of attraction.

As a trivial reformulation of the definitions we arrive at our general statement.

Proposition 2. Let x∗ an attractive stationary point of (24), and assume that x0 is located

in the attracting domain of x∗. Assume furthermore, that for some g ∈ C(RM ,R) : g(x∗) 6= 0,

then limt→+∞g(x(t))g(x∗)

= 1.

We applied this statement when mentioning the expression g(ξ) := 1 + ξ2−ξ3ξ1

, and also

when calculating the ratios of reaction extents.

3.5. The difference. Have we found the real reaction extent?

Example 9. Let us try to fit the example on p. 204 of the book by Atkins and De Paula41

N2(g) + 3H2(g) −−→ 2NH3(g)

into the framework presented here (using abstract notations, aligning with the book). The

reaction is of the form

A + 3Bk−−→ 2C,

36

Page 37: Reaction extent or advancement of the reaction

thus the induced kinetic differential equation for the concentrations of species (assuming

mass action kinetics) is as follows.

cA = −kcAc3B, cB = −3kcAc

3B, cC = 2kcAc

3B,

therefore for the amounts we have

nA = − k

V 3nAn

3B, nB = −3

k

V 3nAn

3B, nC = 2

k

V 3nAn

3B.

The authors provide nA(0) = 10 mol, but they do not specify either nB(0) or nC(0). A tacit

assumption may be nB(0) = −γBnA(0) = 30 mol (the stoichiometric initial condition), but

our arguments work for any

nB(0) ≥ 3nA(0), (25)

and for any nC(0). (If, however, nB(0) < 3nA(0), then limt→+∞ ξ(t) = nB(0)3

< nA(0), thus

one can in no sense speak about the fact that ξ(T ) = 10 for some T ∈ R+.) As

nA(t)− nA(0) = −ξ(t),

for some T ∈ R+ : ξ(T ) = 1mol implies nA(T ) = 9mol, and we also have nB(T )− nB(0) =

−3mol, nC(T ) − nC(0) = 2mol. For no finite T ∈ R+ we can have ξ(T ) = 10mol; this

can only be understood in such a way that limt→+∞ ξ(t) = 10mol (or, to use the short

notation ξ(+∞) = 10mol) and, certainly, limt→+∞ nA(t) = 0mol (if one has enough H2 at

the beginning as assumed in (25)). Note that the above results are independent on the value

of k and V , and—as far as (25) holds—on the value of nB(0) = nH2(0).

Example 10. Let us consider the reversible version of the previous example, Problem 7.25

of41 with a rich moral of the fable. The problem set by the authors is as follows. ”Express

the equilibrium constant of a gas-phase reaction A+3B −−⇀↽−− 2C in terms of the equilibrium

value of the extent of reaction, ξ, given that initially A and B were present in stoichiometric

37

Page 38: Reaction extent or advancement of the reaction

proportions.”

A well-prepared student in classical physical chemistry is supposed to proceed in the

following way.

Initially, one had 1, 3 and 0 mols of A, B, and C. At equilibrium the changes of the species

(in mols) are as follows: −ξ∗,−3ξ∗,+2ξ∗; (where ξ∗ is the reaction extent at equilibrium),

thus at equilibrium we have the following amounts of the species A, B, and C: 1− ξ∗, 3(1−

ξ∗), 2ξ∗; therefore the total amount of the species at equilibrium is ntotal = 1 − ξ∗ + 3(1 −

ξ∗) + 2ξ∗ = 2(2− ξ∗). The equilibrium constant (having the unit mol−2) expressed in terms

of mols is

Kn=

M∏

m=1

nγmX(m) =

(2ξ∗)2

(1− ξ∗)(3(1− ξ∗)3)=

4(ξ∗)2

27(1− ξ∗)4. (26)

The molar fractions in equilibrium are

1− ξ∗

ntotal

=1− ξ∗

2(2− ξ∗),

3(1− ξ∗)

ntotal

=3(1− ξ∗)

2(2− ξ∗),

2ξ∗

ntotal

=2ξ∗

2(2− ξ∗);

thus the equilibrium constant (being a pure number) expressed in terms of molar ratios is

Kx=

M∏

m=1

xγmX(m) =

( 2ξ∗

2(2−ξ∗))2

1−ξ∗

2(2−ξ∗)(3(1−ξ∗)2(2−ξ∗)

)3=

2(ξ∗)2(2− ξ∗)2

27(1− ξ∗)4.

Now let us try to fit the solution of the problem into the general framework presented

here, and en route let us discern the tacit assumptions in the classical derivation. The

differential equations for the reaction extents are

ξ1 = V k1

(

n0A − (ξ1 − ξ−1)

V

)(

n0B − 3(ξ1 − ξ−1)

V

)3

,

ξ−1 = V k−1

(

n0C + 2(ξ1 − ξ−1)

V

)2

,

38

Page 39: Reaction extent or advancement of the reaction

Assuming V = 1dm3, n0A = 1mol, n0

B = 3mol, n0C = 0mol we get

ξ1 = k1(

n0A − (ξ1 − ξ−1)

) (

n0B − 3(ξ1 − ξ−1)

)3,

ξ−1 = k−1

(

n0C + 2(ξ1 − ξ−1)

)2,

Let us introduce ζ := ξ1 − ξ−1 (that will turn out to be the advancement of the reaction

in this special case), then one can write

cA(t) = 1− ζ(t), cB(t) = 3− 3ζ(t), cC(t) = 2ζ(t),

because of (9). If t → +∞, then limt→+∞ c(t) exists and is positive, thus limt→+∞ ζ(t) =: ζ∗

also exists and 0 < ζ∗ < 1.

The evolution equation for ζ is

ζ = k1 (1− ζ) (3− 3ζ)3 − k−1 (2ζ)2 .

If t → +∞, then the limit of the right-hand-side exists, therefore limt→+∞ ζ(t) also exists

and is zero, therefore 0 = k1 (1− ζ) (3− 3ζ)3 − k−1 (2ζ)2 , implying

k1k−1

=(2ζ∗)2

(1− ζ∗) (3− 3ζ∗)3=

4(ζ∗)2

27(1− ζ∗)4, (27)

cf. (26).

Be careful, (27) does not mean that the equilibrium constant depends on the stationary

(equilibrium) concentrations. It is the equilibrium value of ζ∗ that depend on them so that

the equilibrium constant is really constant.

It would be comforting to see that ζ∗ is a(n asymptotically) stable stationary point.

We have seen that the cited author gave no definition whatsoever for the advancement

of the reaction. It turned out that he uses the difference of the two reaction extents useful

39

Page 40: Reaction extent or advancement of the reaction

only for a single pair of reactions. Our definition works for any initial concentration vectors,

whereas he used one possible stoichiometric initial concentration for the irreversible

step A+ 3B −−→ 2C.

Let us mention that Aris2 also uses the trick turning to the difference of two reaction

extents on p. 86.

3.6. Reactant and product species

Favorite expressions in the chemists’ vocabulary are reactant species and product species.

In order to give a general definition of these concepts we can say that a species is a reactant

species if it occurs at the left side of a reaction arrow (or reaction step), and it is a product

species if it occurs at the right side of a reaction arrow.

Table 2 below shows that their use is much more limited than thought.

Table 2: Reactants and products

Reaction Reactant Product Bothstep(s) species species

only onlyX −−→ Y X Y —X+ 2Y −−→ 3 Z X, Y Z —X+Y −−→ 2Y X — YX+ 2Y −−⇀↽−− 3 Z — — X, Y, ZX −−→ Y −−→ Z X Z YX −−→ Y — —2Y −−→ 2X — — X, YX −−→ Y −−→ Z — —2Z −−→ 2Y X — Y, Z

One should be very careful, however. Although the above definition is in no contradiction

with the common use of these words, still we cannot provide a formal refinement of the

classification above although it would be useful.

40

Page 41: Reaction extent or advancement of the reaction

4. And what if the conditions are not fulfilled?

What happens if one takes an exotic reaction, such as one having multiple stationary points,

oscillation or chaos?

4.1. Multistationarity

1

ε

1

ε

� X

� + 2 Y � �

+ �

Figure 11: Reaction having three stationary points with 0 < ε < 1/6

Horn and Jackson,22 (see p. 110) has shown that the complex chemical reaction in Fig. 11

has three (positive) stationary points in every stoichiometric compatibility class if 0 < ε < 16.

To be more specific, let us choose ε = 110, and let c0X = c0Y = 1 mol

dm3 . Then, easy calculation

shows that in the stoichiometric compatibility class defined by {(x, y); x+ y = 2}

c∗1 :=

[

1−√33

1 +√33

]

41

Page 42: Reaction extent or advancement of the reaction

is an attracting stationary point with the attracting domain

{(x, y); 0 < x < 1−√3

3, 1 +

√3

3< y < 1), x+ y = 2},

and

c∗3 :=

[

1 +√33

1−√33

]

is an attracting stationary point with the attracting domain

{(x, y); 1 +√3

3< x < 1, 0 < y < 1−

√3

3, x+ y = 2},

and c∗2 := (1, 1) (a special case of the stoichiometric initial concentration

[

c0X c0X

]

with

c0X > 0) is a non-attracting stationary point. Let us start the reaction from the attracting

domain of the first stationary point and let us calculate the value of the normed pseudo-

Helmholtz function Ψ as a function of time. The result can be seen in Fig. 12. This example

0.5 1.0 1.5t /s

0.9980

0.9985

0.9990

0.9995

Ψ

Figure 12: Convergence to 1 in reaction of Fig. 11.

fits into the validity region of Proposition 2, it is only interesting because of the presence of

multiple stationary points.

Let us mention that criteria of multistationarity have been collected by Joshi and Shiu.42

42

Page 43: Reaction extent or advancement of the reaction

4.2. Oscillation

We mention here two oscillatory reactions: the often used theoretical Lotka–Volterra reaction

and the experimentally based Rabai reaction43 aimed at describing pH oscillations. (One

may say that Brusselator would be a more realistic choice, as it leads to a limit cycle, but

on the other hand it has a third order step. The calculations to be carried out here would

be almost the same.)

4.2.1. The Lotka–Volterra reaction

4.2.1.1. Irreversible case It is known20,44 that under some mild conditions the only two-

species reaction to show oscillation is the (irreversible) Lotka–Volterra reaction Xk1−−→ 2X,

X +Yk2−−→ 2Y, Y

k3−−→ 0. (Cf. also the paper by Toth and Hars.45) It has a single positive

stationary point that is stable but not attractive, therefore one cannot apply Proposition 2.

2 4 6 8t /s

0.55

0� �

����

����

����

0.80

Ψ (t)

Figure 13: Time evolution of the function Ψ in the irreversible Lotka–Volterra reaction withk1 = 3 1

s, k2 = 4 dm3

mols, k3 = 5 1

sand c0X = 1 mol

dm3 , c0Y = 2 mol

dm3 .

What we see in Fig. 13 is that the pseudo-Helmholtz function is oscillating. The question

is if it can be used for some purposes. Note, however, that the individual reaction extents

do not oscillate, they are ”pulsating” (and tending to infinity), they have an oscillatory

derivative, and the zeros of their second derivative clearly show the endpoints of the periods,

43

Page 44: Reaction extent or advancement of the reaction

see Fig. 14. It may be a good idea to calculate any kind of reaction extent or the pseudo-

Helmholtz function for a period in case of oscillatory reactions. We shall return to this

point in a following paper.

2 4 6 8t /s

10

20

30

ξ (t)/m���

ξ1

ξ2

ξ3

2 4 6 8t /s

2

4

6

8

10

12

ξ ' (t)/�� !/s

ξ1'

ξ2'

ξ3'

2 4 6 8t /s

-40

-20

20

40

ξ '' (t)/"#$%/s^2

ξ1''

ξ2''

ξ3''

Figure 14: The individual reaction extents and their first and second derivatives of theirreversible Lotka–Volterra reaction with k1 = 3 1

s, k2 = 4 dm3

mols, k3 = 51

sand

c0X = 1 moldm3 , c0Y = 2 mol

dm3 .

4.2.1.2. Reversible case, detailed balanced The reversible Lotka–Volterra reaction

Xk1−−⇀↽−−k−1

2X, X + Yk2−−⇀↽−−k−2

2Y, Yk3−−⇀↽−−k−3

0 is also worth studying. First, let us note that for all

values of the reaction rate coefficients it has a single, positive stationary point, because the

reaction is reversible, therefore it is permanent,46,47 the trajectories remain in a compact set.

If the trajectories remain in a compact set, then they are either tending to a limit cycle, or

the stationary point is asymptotically stable. The first possibility is excluded by the above

mentioned theorem by Pota, thus, it is only the second possibility that remains.

Let us note that both existence and uniqueness of the stationary state also follow from

the Deficiency One Theorem (p. 106 in Feinberg,16 or p. 176 in Toth et al.15)

44

Page 45: Reaction extent or advancement of the reaction

Next, if the reaction is detailed balanced holding if and only if

k1k2k3 = k−3k−2k−1 (28)

is true, then our Proposition 2 implies that limt→+∞Ψ(ξ(t)) = 1. Based on Theorem 3) we

0.5 1.0 1.5t /s

-1.5

-1.0

-0.5

0.5

1.0

Ψ (t)

Figure 15: Time evolution of the function Ψ in the reversible detailed balanced Lotka–Volterra reaction with k1 = 1 1

s, k2 = 2 dm3

mols, k3 = 3 1

sk−1 = 4 dm3

mols, k−2 = 5 dm3

mols,

k−3 = 158

1s, and c0X = 1 mol

dm3 , c0Y = 2 moldm3 .

can state the same about the product of the ratios of the reaction extents.

4.2.1.3. Reversible case, not detailed balanced If (28) does not hold, the reaction

still has an attracting stationary point, what is more, it has an asymptotically stable

stationary point as we have seen above. Fig. 17 shows the behavior of the individual reaction

extents.

One would expect that the behaviour of the pseudo-Helmholtz function is similar to the

irreversible case if the backward rate coefficients are small. However, this situation cannot

be realized, because Condition (28) shows that it is impossible to decrease the backward

rates without changing the forward ones.

45

Page 46: Reaction extent or advancement of the reaction

2 4 6 8t /s

5

10

15

ξ (t)/mol

ξ1

ξ-1

ξ2

ξ-2

ξ3

ξ-3

0.5 1.0 1.5t /s

5

10

15

ξ ' (t)/mol/s

ξ1′

ξ-1′

ξ2′

ξ-2′

ξ3′

ξ-3′

0.5 1.0 1.5t /s

-15

-10

-5

5

10

ξ '' (t)/mol/s^2

ξ1′′

ξ-1′′

ξ2′′

ξ-2′′

ξ0&′

ξ-3′′

Figure 16: The individual reaction extents and their first and second derivatives of thereversible, detailed balanced Lotka–Volterra reaction with k1 = 1 1

s, k2 = 2 dm3

mol s,

k3 = 3 1sk−1 = 4 dm3

mol s, k−2 = 5 dm3

mol s, k−3 = 15

81s, and c0X = 1 mol

dm3 , c0Y = 2 moldm3 .

4.2.2. The Rabai reaction of pH oscillation

Here we include a reaction proposed by Rabai43 to describe pH oscillations. This reaction has

a much more direct contact to chemical kinetic experiments and is much more challenging

from the point of view of numerical mathematics than the celebrated Lotka–Volterra reaction.

Rabai43 starts with the steps

A− +H+ k1−−⇀↽−−k−1

AH,

AH+ H+ + {B} k2−−→ 2H+ + P−,

where {B} is an external species, i.e. one considered to have constant concentration. This

46

Page 47: Reaction extent or advancement of the reaction

0.2 0.4 0.6 0.8t /s

0.0

0.2

0.4

0.6

0.8

1.0

Ψ (t)

Figure 17: Time evolution of the function Ψ in the reversible, not detailed balanced Lotka–Volterra reaction with k1 = 1 1

s, k2 = 2 dm3

mol s, k3 = 3 1

sk−1 = 4 dm3

mol s, k−2 = 5 dm3

mol s,

k−3 = 158

1s, and c0X = 1 mol

dm3 , c0Y = 2 mol

dm3 .

reaction has a single stationary point

c∗A− = 0, c∗H+ = c0H+ + c0AH, c∗AH = 0, c∗P = c0A− + c0AH + c0P

specializing into c∗A−

= 0, c∗H+ = c0

H+, c∗AH = 0, c∗P = c0A−

with the natural restriction on the

initial condition c0AH = 0, c0P = 0.

Putting the reaction into a CSTR (continuously stirred tank reactor) means in the terms

of formal reaction kinetics that all the species can flow out and some of the species flow in,

i.e. in the present case the following steps are added

A− k0−−→ 0, (29)

0k0c0A−−−−→ A−, (30)

H+ k0−−→ 0, (31)

0k0c0H+−−−→ H+, (32)

AHk0−−→ 0. (33)

47

Page 48: Reaction extent or advancement of the reaction

0.5 1.0 1.5t /s

2

468

10

ξ (t)/mol

ξ1

ξ-1

ξ2

ξ-2

ξ3

ξ-3

0.5 1.0 1.5t /s

2

&'8

10

ξ ' (t)/mol/s

ξ1′

ξ-1′

ξ2′

ξ-2′

ξ3′

ξ-3′

0.5 1.0 1.5t /s

-15

-10

-5

5

10

ξ '' (t)/mol/s^2

ξ1′′

ξ-1′′

ξ2′′

ξ-2′′

ξ0&′

ξ-3′′

Figure 18: The individual reaction extents and their first and second derivatives of thereversible, not detailed balanced Lotka–Volterra reaction with k1 = 1 1

s, k2 = 2 dm3

mol s,

k3 = 3 1sk−1 = 4 dm3

mol s, k−2 = 5 dm3

mol s, k−3 = 15

81s, and c0X = 1 mol

dm3 , c0Y = 2 moldm3 .

As a result of adding these steps, multistability (existence of more than one stationary points)

may occur with appropriately chosen values of the parameters.

When the reaction step

H+ + {C−} k3−−→ CH (34)

is also added, one may obtain periodic solutions having appropriate parameter values, see

Fig. 19.

Let us remark that neither the Lotka–Volterra reaction nor the Rabai reaction is mass-

conserving. However, we could have chosen a mass-conserving oscillatory reactions, as e.g.

the Ivanova reaction: X + Y −−→ 2Y,Y + Z −−→ 2 Z,Z + X −−→ 2X. Finally, we remark

that the connection between multistationarity and oscillation is far from being as simple as

part of the literature asserts, see.48

4.3. Chaos

Here we use a version of the Rabai reaction that can numerically be shown to exhibit

chaotic behaviour, see Fig. 20 and is a good model for experimental pH oscillators also show-

48

Page 49: Reaction extent or advancement of the reaction

400 600 800t /s

()*

5.0

5.5

+,-

./5

789

pH(t)

Figure 19: Time evolution of the pH and the projection of the first three coordinates ofthe trajectory in the oscillating Rabai reaction with k1 = 1010 dm3

mol s, k−1 = 103 1

s,

k2 = 106 dm3

mol s, k3 = 1 dm3

mol s, k0 = 10−3 1

s, and c0

A−= 5

1000moldm3 , c0H+ = 1

1000moldm3 ,

c0AH = 0 moldm3 , c

0P = 0 mol

dm3 .

ing chaotic behaviour. (Let us mention that rigorously verified kinetic differential equation

is not known to exist.)

Including the removal of CH

CHk11−−→ 0 (35)

and using appropriate parameters and favorable input concentrations chaotic solutions are

obtained. Fig. 20 illustrates (does not prove!) this fact.

The reaction extents tend to +∞ in such a way that their derivative is chaotic, as

expected, see Fig. 21.

5. Discussion, outlook

We have introduced the concept of reaction extent for reaction networks of arbitrary com-

plexity (any number of reaction steps and species) without assuming mass action kinetics.

This reaction extent gives the advancement of each individual irreversible reaction step; in

case of reversible reactions we have a pair of reaction extents. In all the practically important

cases the fact that the reaction extent is strictly monotonous, implies that the reaction steps

do not cease to proceed even during equilibrium. Thus, we provided a meaning of dynamic

49

Page 50: Reaction extent or advancement of the reaction

2500 3000 3500t /s

:;<

=>?

@AB

6.0

CDE

7.0

pH(t)

FGH IJK LMN 6.0 OPQ 7.0pAH

4.0

RST

UVW

Z[\

6.0

]^_

7.0

pH

Figure 20: Time evolution of the pH and the projection of the logarithm of the first twocoordinates of the trajectory in the oscillating Rabai reaction with the same parameters asin Fig. 19 and k11 = 5

1001s.

equilibrium usually taught also in high schools. Note that we made no allusion to either

thermodynamics or statistical mechanics.

After a few statement on the qualitative behaviour of the reaction extent we made a

few effort to connect the notion with the traditional ones. Thus, we have shown that if the

number of reaction steps is one, as in all the practically important cases, the reaction extent

in the long run (as t → +∞) tends to 1 if appropriately scaled.

Then, for an arbitrary number of reversible detailed balanced reaction steps the product

of the ratios of the reaction extents are shown to tend to 1 as t → +∞.

Our most general statement follows for arbitrary reactions having an attracting stationary

point and with a function not vanishing on the stationary point: in this case the value of the

chosen function along the time dependent concentrations divided by the value of the given

function at the stationary concentration tends to 1. Thus, this statement is true not only

for the reaction extent but also for any appropriate functions.

Recalculations of some examples by Atkins and DePaula41 make clear the vague meaning

of some further concepts and statements including the concept of reactant, intermediate and

product species.

50

Page 51: Reaction extent or advancement of the reaction

400 600 800 1000 1200t /s

`abcde

fghijk

nopqrs

ξ11(t)

Figure 21: Time evolution of the reaction extent of the step (34) in the chaotic Rabai reactionwith the same parameters as in Fig. 20.

We have numerically studied a few examples of reactions where our statements cannot

be applied: multistationary, oscillatory and chaotic reactions.

One should take into consideration that although in the practically interesting cases the

number of equations in (10) is larger than those in (5), the equations for the reaction extents

are of a much simpler structure, as the right hand side of the differential equations describing

them only consist of a single term.

As a by-product we have given an exact definition of stoichiometric initial concentration

and initial concentration in excess. We have also shown that reactant and product can be

defined in full generality—contrary to the literature.

In the Appendix we cited a few examples proving that peer review of textbooks are not

less important than that of papers.

One can say that the concept of reaction extent can be usefully applied to a larger class

of reactions than usually, but in some (exotic) cases its use need further investigations.

However, it is the reader to decide if we succeeded in avoiding all the traps mentioned at the

beginning.

Quite a few authors treats the methodology of teaching the concept; we think this ap-

proach will only have its raison d’etre when the scientific background will have been clarified.

51

Page 52: Reaction extent or advancement of the reaction

Acknowledgement

The present work has been supported by the National Office for Research and Development

(2018-2.1.11-TET-SI-2018-00007 and SNN 125739). JT is grateful to Dr. J. Karsai (Bolyai

Institute, Szeged University) and for Daniel Lichtblau (Wolfram Research) for their help.

Members of the Working Party for Reaction Kinetics and Photochemistry, especially Prof.

T. Turanyi and Dr. Gy. Pota, made a number of useful critical remarks.

6. Notations

Some of the readers may appreciate that we have collected the used notations.

7. Appendix

7.1. On the form of the induced kinetic differential equation

It is relatively easy to characterize those polynomial differential equations that can emerge

as the mass action type induced kinetic differential equation of a complex chemical reaction.

Lemma 1 (32). Suppose the right hand side f of a differential equation c(t) = f(c(t)) is a

polynomial. It is the induced kinetic differential equation of a reaction endowed with mass

action type kinetic if and only if its mth coordinate function is of the form

fm(c) = gm(cm)− cmhm(c), (36)

where all the coefficients of the polynomials gm and hm are non-negative, and gm does not

depend on the scalar cm : cm :=

[

c1 c2 . . . cm−1 cm+1 . . . cM

]⊤

.

A right-hand-side of the form (36) is sometimes said to be Hungarian. One would think

52

Page 53: Reaction extent or advancement of the reaction

Table 3: Notations I

Notation Meaning Unit Typical value=⇒ implies∈ belongs to∀ universal quantor ”for all”∃ existential quantor ”there is”⊙ coordinate-wise product of vectors

cm, cX concentration of X(m) and X moldm3

c vector of concentrationsc0m initial concentration of X(m) mol

dm3

c∗m stationary concentration of X(m) moldm3

Ci(A,B) i times continuouslydifferentiable functions

Dom(u) the domain of the function uJ ⊂ R an open interval

k, kr, k−r reaction rate coefficient 1s( moldm3 )1−

∑Mm=1 αm,r

M the number of chemical species ∈ N

nm the quantity of species X(m) mol ∈ N

N the set of positive integersN0 the set of

non-negative integersn0m the initial quantity mol ∈ N

of species X(m)n the vector of

the quantity of speciesn0 the vector of

the initial quantity of species

rater rate of the rth reaction step moldm3s

rate vector of reaction ratesR the set of real numbersR+ the set of

positive real numbersR+

0 the set ofnon-negative real numbers

t time ∈ R

53

Page 54: Reaction extent or advancement of the reaction

Table 4: Notations II

Notation Meaning Unit Typical value

V volume dm3

Wr number of occurrences ∈ N0

of the rth stepW vector of

number of occurrencesxm mth dependent variable in a

differential equationx vector of variables in a

differential equationX,Y,X(m) chemical speciesαm,r, αm stoichiometric coefficient 1 0, 1, 2, 3

in the reactant complexα matrix of

reactant complex vectorsβm,r, βm stoichiometric coefficient 1 0, 1, 2, 3

in the product complexβ matrix of

product complex vectorsγm,r, γm stoichiometric number 1 −3,−2, . . . , 2, 3

in the reactant complexγ stoichiometric matrixξr reaction extent mol

of the rth stepξ vector of

reaction extents

54

Page 55: Reaction extent or advancement of the reaction

that this property is equivalent to

fm(c1, c2, · · · , cm−1, 0, cm+1, · · · , cM) ≥ 0, (37)

but it is not true, as the following example shows.

c1 = c22 −2c2c3 + c23 + · · · , c2 = · · · , c3 = · · · .

The first equation fulfils (37), but is not Hungarian because of the boxed term.

Note that any polynomial equation can be transformed into a Hungarian one (as far as

the trajectories in the open first order are concerned) by multiplying the right-hand-side

with the product of the variables, an idea of Samardzija49 generalized by Gh. Craciun.

7.2. Critical review of the literature

The authors dealing with reaction kinetics seem to show three different kinds of behaviour.

1. Totally or practically neglect the concept, some of them do not even mention any of the

names reaction extent (extent of reaction), advancement/progress of reaction, reaction

flux. (It is not clear what the reaction coordinate means but it is surely different from

the previous concept(s).)

2. Strongly refuse the use of the concept.

3. Utter scientifically unacceptable confused, ambiguous sentences.

Let us starts with remarks on two classics. The concept comes from De Donder and Van

Rysselberghe,1 although almost always it is only the first author who is mentioned. Here

the general case is considered, the definition is implicit and the properties are not studied.

The definition of the degree of reaction on page 169 of the classic textbook by Callen3

commits most of the faults enumerated above.

Now let us see the point of view of more recent authors.

55

Page 56: Reaction extent or advancement of the reaction

7.2.1. Neglectors

It is an acceptable standpoint at the present status of things.

• Fig. 17.15 of the popular textbook50 by Dillard and Goldberg is typical (first, horizon-

tal coordinate: reaction coordinate; second, vertical coordinate: energy). Fig. 17.17

pretends to show the effect of a catalyst. The figure actually shows a potential energy

surface, and this concept is not related to the one treated here.

• Feinberg16 is not interested.

• Pilling and Seakins13 on p. 14 show the typical figure in the Introduction (!), and never

return to it.

• On page 90 Toth et al.15 provided an initiative of the general investigations. In a later

chapter, see p. 268, they show a similar approach for the stochastic model. Historically

(for us) it was Kurtz14 the first to give a general clear definition, but that author was

not interested in the details of the topic.

• Turanyi and Tomlin10 are not interested.

• The introduction of the degree of reaction in the classic textbook3 (see p. 169)also

commits some of the faults but one may say that the author is not really interested in

the use of the concept.

7.2.2. Refusers

• The opinion of Lente51 as expressed on p. 96 is: The x axis is often called ”reaction

coordinate” but typically lacks any meaningful definition.

• Lente in a paper with Schmitz52 emphasizes the restricted use of the concept: ”At

this point, it should be emphasized that the extent of reaction only makes sense if the

studied reaction is a reasonably simple one that does not have any intermediates in

significant concentrations.”

56

Page 57: Reaction extent or advancement of the reaction

7.2.3. Confused

Our standpoint is that if an author seems to be mistaken then it is not enough to express

our disapprovement, we have to find out their unrevealed errors and tacit assumptions, etc.

• The view shared by most of the chemists is summarized in a few key words by Laidler23

We make a few remarks on those key words using our notations.

Extent of Reaction (or Extent of Conversion) Under the usual assumptions the

extent of reaction at any time is defined by ξ(t) := nm(t)−n0m

γmwhere n0

m is the initial

amount of the reactant or product entity X(m), n0m is the amount at time t (both

measured in moles), and γm is the stoichiometric number for that species in the

reaction equation as written. The extent of reaction defined in this way

depends on how the reaction equation is written, but it is independent of

which entity in the reaction equation is used in the definition.

For the general reaction (13) the extent of reaction is

ξ :=∆n1

γ1=

∆n2

γ2= · · · = ∆nM

γM.

As the author (without mentioning this fact) excludes steps of the form X+Y −−→

X + Z, we had to essentially change his equation to be meaningful, also avoiding

to speak about reactant, product and signs of stoichiometric coefficients. Note

that general here means a single, irreversible step.

Rate of Conversion The rate of conversion is defined as the time derivative of the

extent of reaction, ξ = nm

γm.

Rate of Reaction The rate of reaction is the rate of conversion per unit volume:

rater := ξV, thus it is a derived quantity (?!). For the species X(m) the rate of

reaction is thus given by 1V γm

nm. If the volume does not change during the course

of reaction the rate of reaction may be expressed in terms of concentrations as

57

Page 58: Reaction extent or advancement of the reaction

rater =1γm

cX(m).

• Let us scrutinize the ”Bible” of Physical Chemistry,41 published in eleven editions up

to the writing of the present paper. (We do not have access to all the editions.)

1. On page 201 they give a definition for the reaction A −−⇀↽−− B.

2. On page 201 they provide a general (?) definition: ”The general case of a

reaction.—Consider the reaction 2A + B −−→ 3C + D. A more sophisticated

way of expressing the chemical equation is to write it in the symbolic form 0 =

3C + D − 2A − B by subtracting the reactants from both sides (and replacing

the arrow by an equals sign).” This is a less sophisticated way as one loses part

of the information in case a species occurs on both sides.

3. Let us see p. 233 of the cited work: ”Checklist of key ideas—The extent of reaction

ξ is defined such that, when the extent of reaction changes by a finite amount ∆ξ,

the amount of A present changes from nA,0 to nA,0 −∆ξ.” This sentence should

be taught what a definition is not.

4. On p. 233, after the nebulous quasi-definitions the rate of reaction—that is a

clearly understood, measurable quantity—it is now defined as a derived quantity

using the extent of reaction.

5. P. 809 clearly illustrates why it is an even heavier task to incorporate the concept

reaction coordinate into the area of rational science: ”The reaction coordinate

is the collection of motions, such as changes in inter-atomic distances and bond

angles, that are directly involved in the formation of products from reactants.

(The reaction coordinate is essentially a geometrical concept and quite distinct

from the extent of reaction.)”

6. The concept is applied on p. 238: ”Express the equilibrium constant of the

gas phase reaction A + 3B −−⇀↽−− 2C in terms of the equilibrium value of the

58

Page 59: Reaction extent or advancement of the reaction

extent of reaction ξ, given that initially A and B were present in stoichiometric

proportions.” This example has been treated in detail above.

References

(1) De Donder, T.; Van Rysselberghe, P. Thermodynamic Theory of Affinity: A Book of

Principles ; Stanford University Press: California, 1936.

(2) Aris, R. Introduction to the Analysis of Chemical Reactors ; Prentice-Hall: Englewood

Cliffs, N. J., 1965.

(3) Callen, H. B. Thermodynamics and an Introduction to Thermostatistics, 2nd ed.; John

Wiley & Sons: New York, Chichester, Brisbane, Toronto, Singapore, 1985.

(4) Garst, J. F. The extent of reaction as a unifying basis for stoichiometry in elementary

chemistry. J. Chem. Educ. 1974, 51, 194.

(5) Dumon, A.; Lichanot, A.; Poquet, E. Describing chemical transformations: From the

extent of reaction ξ to the reaction advancement ratio χ. J. Chem. Educ. 1993, 70,

29–30.

(6) Peckham, G. D. The extent of reaction, ∆ξ—Some nuts and bolts. J. Chem. Educ.

2001, 78, 508.

(7) Croce, A. E. The application of the concept of extent of reaction. Journal of chemical

education 2002, 79, 506.

(8) Vandezande, J. E.; Vander Griend, D. A.; DeKock, R. L. Reaction extrema: Extent of

reaction in general chemistry. Journal of Chemical Education 2013, 90, 1177–1179.

(9) Moretti, G. The “extent of reaction”: a powerful concept to study chemical transfor-

mations at the first-year general chemistry courses. Foundations of Chemistry 2015,

17, 107–115.

59

Page 60: Reaction extent or advancement of the reaction

(10) Turanyi, T.; Tomlin, A. S. Analysis of Kinetic Reaction Mechanisms ; Springer, 2014.

(11) Craciun, G.; Jin, J.; Yu, P. Y. Uniqueness of weakly reversible and deficiency zero

realization. arXiv preprint arXiv:2010.04316 2020,

(12) Truesdell, C. A. Rational Thermodynamics, second edition ed.; Springer-Verlag: New

York, 1984.

(13) Pilling, M. J.; Seakins, P. W. Reaction Kinetics ; Oxford University Press: Oxford,

1995.

(14) Kurtz, T. G. The relationship between stochastic and deterministic models for chemical

reactions. J. Chem. Phys. 1972, 57, 2976–2978.

(15) Toth, J.; Nagy, A. L.; Papp, D. Reaction Kinetics: Exercises, Programs and Theorems.

Mathematica for Deterministic and Stochastic Kinetics ; Springer Nature: Berlin, Hei-

delberg, New York, 2018.

(16) Feinberg, M. Foundations of Chemical Reaction Network Theory ; Springer, 2019.

(17) Kovacs, K.; Vizvari, B.; Riedel, M.; Toth, J. Decomposition of the permanganate/oxalic

acid overall reaction to elementary steps based on integer programming theory. Physical

Chemistry Chemical Physics 2004, 6, 1236–1242.

(18) Papp, D.; Vizvari, B. Effective solution of linear Diophantine equation systems with an

application in chemistry. J. Math. Chem. 2006, 39, 15–31.

(19) Deak, J.; Toth, J.; Vizvari, B. Mass conservation in complex chemical reactions. Alk.

Mat. Lapok 1992, 16, 73–97, In Hungarian.

(20) Pota, G. Two-component bimolecular systems cannot have limit cycles: A complete

proof. The Journal of Chemical Physics 1983, 78, 1621–1622.

60

Page 61: Reaction extent or advancement of the reaction

(21) Volpert, A.; Hudyaev, S. Analyses in Classes of Discontinuous Functions and Equations

of Mathematical Physics ; Martinus Nijhoff: Dordrecht, 1985.

(22) Horn, F.; Jackson, R. General mass action kinetics. Archive for Rational Mechanics

and Analysis 1972, 47, 81–116.

(23) Laidler, K. J. A glossary of terms used in chemical kinetics, including reaction dynamics

(IUPAC Recommendations 1996). Pure and applied chemistry 1996, 68, 149–192.

(24) Law, C. K. Combustion at a crossroads: Status and prospects. Proceedings of the

Combustion Institute 2007, 31, 1–29.

(25) Polcz, P.; Kulcsar, B. A.; Szederkenyi, G. Computation of rational parameter dependent

Lyapunov functions for LPV systems. EasyChair Preprint 2018, 175 .

(26) Polcz, P.; Peni, T.; Szederkenyi, G. Reduced linear fractional representation of nonlinear

systems for stability analysis. IFAC-PapersOnLine 2018, 51, 37–42.

(27) Kiss, K.; Toth, J. n-Dimensional ratio-dependent predator-prey systems with memory.

Differential Equations and Dynamical Systems 2009, 17, 17–35.

(28) Pota, G. Solutions defined on finite time intervals in a model of a real kinetic system.

Journal of Mathematical Chemistry 2016, 54, 1879–1883.

(29) Csikja, R.; Papp, D.; Toth, J. Blow-up and semidiefnite programming. 2021, In prepa-

ration.

(30) Csikja, R.; Toth, J. Blow up in polynomial differential equations. Enformatika. Int. J.

Appl. Math. Comput. Sci 2007, 4, 728–733.

(31) Cardelli, L.; Tribastone, M.; Tschaikowski, M. From electric circuits to chemical net-

works. Natural Computing 2020, 19, 237–248.

61

Page 62: Reaction extent or advancement of the reaction

(32) Hars, V.; Toth, J. On the inverse problem of reaction kinetics. Colloquia Mathematica

Societatis Janos Bolyai. 1979; pp 363–379.

(33) Craciun, G.; Johnston, M.; Szederkenyi, G.; Tonello, E.; Toth, J.; Yu, P. Y. Realizations

of kinetic differential equations. Mathematical Biosciences and Engineering 2019, 17,

862–892.

(34) Glasser, L. Correct use of Helmholtz and Gibbs function differences, ∆A and ∆G: The

van’t Hoff reaction box. Journal of Chemical Education 2016, 93, 978–980.

(35) Atkins, P. W. Physical Chemistry (5th edition); Oxford University Press: Oxford, 1994.

(36) Atkins, P. W. Physical Chemistry (6th edition); Oxford University Press: Oxford, 1998.

(37) Treptow, R. S. Free energy versus extent of reaction: Understanding the difference

between ∆G and ∂G/∂ξ. Journal of chemical education 1996, 73, 51.

(38) Boros, B.; Hofbauer, J. Permanence of weakly reversible mass-action systems with a

single linkage class. SIAM Journal on Applied Dynamical Systems 2020, 19, 352–365.

(39) Feinberg, M. Necessary and sufficient conditions for detailed balancing in mass action

systems of arbitrary complexity. Chemical Engineering Science 1989, 44, 1819–1827.

(40) Liptak, G.; Pereira, M.; Kulcsar, B.; Kovacs, M.; Szederkenyi, G. Traffic Reaction

Model. arXiv preprint arXiv:2101.10190 2021,

(41) Atkins, P.; De Paula, J. Atkins’ Physical Chemistry (8th edition); Oxford University

Press: Oxford, 2006.

(42) Joshi, B.; Shiu, A. A survey of methods for deciding whether a reaction network is

multistationary. Mathematical Modelling of Natural Phenomena 2015, 10, 47–67.

(43) Rabai, G. Modeling and designing of pH-controlled bistability, oscillations, and chaos

in a continuous-flow stirred tank reactor. 1998, 135, 381–392.

62

Page 63: Reaction extent or advancement of the reaction

(44) Schuman, B.; Toth, J. No limit cycle in two species second order kinetics. Bulletin des

sciences mathematiques 2003, 127, 222–230.

(45) Toth, J.; Hars, V. Specification of oscillating chemical models starting from a given

linearized form. Theoretica Chimica Acta 1986, 70, 143–150.

(46) Simon, P. Globally attractive domains in two-dimensional reversible chemical, dynam-

ical systems. Ann. Univ. Sci. Budapest, Sect. Comp. 1995, 15, 179–200.

(47) Boros, B. Existence of positive steady states for weakly reversible mass-action systems.

SIAM Journal on Mathematical Analysis 2019, 51, 435–449.

(48) Toth, J. Multistationarity is neither necessary nor sufficient to oscillations. Journal of

mathematical chemistry 1999, 25, 393–397.

(49) Samardzija, N.; Greller, L. D.; Wasserman, E. Nonlinear chemical kinetic schemes

derived from mechanical and electrical dynamical systems. J. Chem. Phys. 1989, 90,

2296–2304.

(50) Dillard, C. R.; Goldberg, D. E. Chemistry: Reactions, Structure, and Properties ;

Macmillan New York: New York City, NY, 1978.

(51) Lente, G. Deterministic Kinetics in Chemistry and Systems biology: The Dynamics of

Complex Reaction Networks ; Springer, 2015.

(52) Schmitz, G.; Lente, G. Fundamental concepts in chemical kinetics. ChemTexts 2020,

6, 1–7.

63

Page 64: Reaction extent or advancement of the reaction

2 4 6 8t

10

20

30

ξ (t)

ξ1

ξ2

ξ3

Page 65: Reaction extent or advancement of the reaction

2 4 6 8t

10

20

30

ξ (t)

ξ1

ξ2

ξ3

Page 66: Reaction extent or advancement of the reaction

2 4 6 8t

2

4

6

8

10

12

ξ ' (t)

ξ1'

ξ2'

ξ3'

Page 67: Reaction extent or advancement of the reaction

2 4 6 8t

-40

-20

20

40

ξ '' (t)

ξ1''

ξ2''

ξ3''

Page 68: Reaction extent or advancement of the reaction

2 4 6 8t

1.0

uvw

xyz

{|}

~��

���

ξ ' (t)

ξ1′

ξ-1′

ξ2′

ξ-2′

ξ3′

ξ-3′

Page 69: Reaction extent or advancement of the reaction

1 2 3 4t

-0.6

-0.4

-0.2

0.2

ξ '' (t)

ξ1′′

ξ-1′′

ξ2′′

ξ-2′′

ξ0&′

ξ-3′′

Page 70: Reaction extent or advancement of the reaction

2 4 6 8t

-���

-���

-0.5

0.0

Ψ (t)

Page 71: Reaction extent or advancement of the reaction

0.5 1.0 1.5t

4

6

10

ξ

ξ1

ξ-1

ξ�

ξ-�

ξ�

ξ-�

Page 72: Reaction extent or advancement of the reaction

2 4 6 8t

0.0

0.2

0.4

0.6

0.8

1.0

Ψ (t)

Page 73: Reaction extent or advancement of the reaction

0.5 1.0 1.5t

4

6

10

ξ ' (t)

ξ1′

ξ-1′

ξ2′

ξ-2′

ξ3′

ξ-3′

Page 74: Reaction extent or advancement of the reaction

0.5 1.0 1.5t

-15

-10

-5

5

10

ξ '' (t)

ξ1′′

ξ-1′′

ξ2′′

ξ-2′′

ξ0&′

ξ-3′′

Page 75: Reaction extent or advancement of the reaction

0.5 1.0 1.5t

1

4

ξ

ξ1

ξ-1

ξ�

ξ-�

ξ�

ξ-�