1 Introduction

Parameter identifiability in a system of ordinary differential equations mainly addresses the question of deciding whether the system parameters can be uniquely determined from data (see for instance Walter and Pronzato 1997; DiStefano 2014, Chapter 10). Since the pioneering paper (Bellman and Åström 1970), this problem has been broadly studied for general systems under different perspectives, including Taylor series and generating series approaches and differential algebra-based approaches. More details can be found in Pohjanpalo (1978), Ollivier (1990), Ljung and Glad (1994), Sedoglavic (2002), Xia and Moog (2003), Saccomani et al. (2003), Bellu et al. (2007), Meshkat et al. (2009), Chis et al. (2011a), Raue et al. (2014) and Hong et al. (2018a). Also, a variety of software tools for identifiability have been developed that work for general classes of models (e.g., polynomial or rational), such as DAISY (Bellu et al. 2007), COMBOS (Meshkat et al. 2014), GenSSI (Ligon et al. 2017) and SIAN (Hong et al. 2018b).

In this paper, we address the identifiability problem for a specific infinite class of models. Our aim is to obtain general statements about all the models in the class (see Walch and Eisenberg 2016; Brouwer et al. 2017 for prior results of this sort but for different classes of models). More precisely, we consider a particular class of systems of equations arising from biochemical reaction networks under mass-action kinetics, which induces polynomial autonomous systems of differential equations. In this framework, in Craciun and Pantea (2008), the authors describe necessary and sufficient conditions for the unique identifiability of the reaction rate constants (the parameters) of a chemical reaction network. Following their approach, we provide in this work sufficient conditions for uniquely identifying all the rate constants of a certain family of biochemical reaction networks from a reduced set of variables (see Definition 3). Unlike other authors (Anguelova et al. 2012), we do not consider all the possible minimal sets of variables allowing parameter identifiability, but we only focus on certain biologically relevant sets.

The family of networks we deal with is abundant in the literature. One example is the multisite phosphorylation system which describes the phosphorylation of a protein in L sites by a kinase(Y)/phosphatase(\({\tilde{Y}}\)) pair in a sequential and distributive mechanism (Deshaies and Ferrell 2001). The substrate \(S_i\) is the phosphoform obtained from the unphosphorylated substrate \(S_0\) by attaching i phosphate groups to it. Each phosphoform can accept (via an enzymatic reaction involving Y) or lose (via a reaction involving the phosphatase \({\tilde{Y}}\)) at most one phosphate (the mechanism is “distributive”), and there is a specific order to be followed for attaching and removing the phosphate groups (the phosphorylation is “sequential”).

Example 1

The reactions in the L-site sequential phosphorylation/dephosphorylation network are represented by the following labeled digraph:

$$\begin{aligned} \begin{array}{c} Y+S_0 \overset{a_1}{\underset{b_1}{\rightleftarrows }} U_1 \overset{c_1}{\rightarrow } Y+S_1 \overset{a_2}{\underset{b_2}{\rightleftarrows }} \dots \overset{a_L}{\underset{b_L}{\rightleftarrows }}U_L \overset{c_L}{\rightarrow } Y+S_L\\ {\tilde{Y}}+S_L \overset{{\tilde{a}}_L}{\underset{{\tilde{b}}_L}{\rightleftarrows }} V_L \overset{{\tilde{c}}_L}{\rightarrow } \dots \overset{{\tilde{c}}_2}{\rightarrow } {\tilde{Y}}+S_1 \overset{{\tilde{a}}_{1}}{\underset{{\tilde{b}}_1}{\rightleftarrows }} V_1 \overset{{\tilde{c}}_1}{\rightarrow } {\tilde{Y}}+S_0, \end{array}\end{aligned}$$

where \(U_1,\dots ,U_L,V_1,\dots ,V_L\) are intermediate enzyme-substrate species. The mass-action dynamical system for this network is [see identity (1) in Sect. 2.1]:

$$\begin{aligned}\begin{array}{rcl} {\dot{s}}_0 &{} = &{}-a_1ys_0+b_1u_1+{\tilde{c}}_1v_1,\\ {\dot{s}}_L &{} = &{}c_Lu_L-{\tilde{a}}_L{\tilde{y}}s_L+{\tilde{b}}_Lv_L,\\ \dot{u_i} &{} = &{} a_iys_{i-1}-(b_i+c_i)u_i, \ 1\le i \le L,\\ {\dot{v}}_i &{} = &{} {\tilde{a}}_i{\tilde{y}}s_i-({\tilde{b}}_i+{\tilde{c}}_i)v_i, \ 1\le i \le L,\\ {\dot{y}} &{} = &{}{i=1}{\overset{L}{\sum }}-a_iys_{i-1}+(b_i+c_i)u_i,\\ \dot{{\tilde{y}}} &{} = &{}{i=1}{\overset{L}{\sum }}-{\tilde{a}}_i{\tilde{y}}s_i+({\tilde{b}}_i+{\tilde{c}}_i)v_i,\\ {\dot{s}}_i&{} = &{}c_iu_i-a_{i+1}ys_i+b_{i+1}u_{i+1}+{\tilde{c}}_{i+1}v_{i+1}-{\tilde{a}}_i{\tilde{y}}s_i+{\tilde{b}}_iv_i, \; 1\le i\le L-1, \end{array}\end{aligned}$$

where lower-case letters represent the time-varying concentration of the corresponding chemical species. Here, the derivative with respect to time is represented with a dot over the corresponding variable.

As a consequence of Theorem 1 proved below, all the constants in the first connected component can be identified, in the sense of Definition 3, from the successive total derivatives of \(s_L\) up to order \(\max \{2,2L-1\}\) and all the constants in the second connected component can be identified from the successive total derivatives of \(s_0\) up to the same order. Moreover, as proved in Proposition 3, all the constants in the whole network can be identified from the successive total derivatives of \(s_L\) up to order \(\max \{2,2L-1\}\).

Another example of major biological importance is phosphorylation cascades, such as the mitogen-activated protein kinase (MAPK) cascade (Catozzi et al. 2016; Huang and Ferrell 1996; Kholodenko 2000; Shaul and Seger 2007). This cascade plays an essential role in signal transduction by modulating gene transcription in response to changes in the cellular environment. MAPK cascades participate in a number of diseases including chronic inflammation and cancer (Davis 2000; Kyriakis and Avruch 2001; Pearson et al. 2001; Schaeffer and Weber 1999; Zarubin and Han 2005) as they control key cellular functions (Hornberg et al. 2005; Pearson et al. 2001; Widmann et al. 1999). We depict in the following example the two-layer signaling cascade.

Example 2

Consider the graph associated with the two-layer simple phosphorylation cascade where the simplified diagram and the corresponding reactions are, respectively:

figure a

The corresponding mass-action dynamical system is [see (1)]:

$$\begin{aligned} {\dot{s}}_{1,0}= & {} -a_1es_{1,0}+b_1u_1+{\tilde{c}}_1v_1,\\ {\dot{s}}_{1,1}= & {} c_1u_1-{\tilde{a}}_1f_1s_{1,1}+{\tilde{b}}_1v_1-a_2s_{1,1}s_{2,0}+(b_2+c_2)u_2,\\ {\dot{e}}= -\dot{u_1}= & {} -a_1es_{1,0}+(b_1+c_1)u_1,\\ {\dot{f}}_1=-{\dot{v}}_1= & {} -{\tilde{a}}_1f_1s_{1,1}+({\tilde{b}}_1+{\tilde{c}}_1)v_1,\\ {\dot{s}}_{2,0}= & {} -a_2s_{1,1}s_{2,0}+b_2u_2+{\tilde{c}}_2v_2,\\ {\dot{s}}_{2,1}= & {} c_2u_2-{\tilde{a}}_2f_2s_{2,1}+{\tilde{b}}_2v_2,\\ {\dot{u}}_2= & {} a_2s_{1,1}s_{2,0}-(b_2+c_2)u_2,\\ {\dot{f}}_2=-{\dot{v}}_2= & {} -{\tilde{a}}_2f_2s_{2,1}+({\tilde{b}}_2+{\tilde{c}}_2)v_2. \end{aligned}$$

We prove in Theorem 2 that all the parameters in a signaling cascade system can be identified from a single variable: the last product of the last layer (\(S_{2,1}\) in the cascade presented in Example 2). This species is usually an output of interest for this type of cascades (Aoki et al. 2011; Chen et al. 2009; Hagen et al. 2013; Lin et al. 2009).

The organization of the paper is as follows. The next section provides introductory material on chemical reaction networks, mass-action kinetics equations and identifiability. Section 3 deals with the general assumptions required by the biochemical reaction networks we consider along the paper. In Sects. 4 and 5 we analyze the identifiability for sequential phosphorylation/dephosphorylation networks and phosphorylation cascades, respectively. We illustrate these results in Sect. 5.2 with a procedure to determine, from (noise-free) data, the 30 rate constants in the three-layer MAPK cascade, which relies on a heuristic to choose points to specialize the variables and solve for the rate constants. Finally, we include a section “Appendix” with the complete proofs of the results stated in the paper.

2 Preliminaries and Basic Notions

2.1 Chemical Reaction Systems

We briefly recall the basic setup of chemical reaction networks and how they give rise to autonomous dynamical systems under mass-action kinetics.

Given a set of s chemical species (denoted by capital letters), a chemical reaction network on this set of species is a finite directed graph whose vertices are indicated by complexes (non negative integer linear combinations of the species) and whose edges are labeled by parameters (positive reaction rate constants). The labeled digraph is denoted \(G = ({\mathcal {V}},{\mathcal {R}}, {\mathbf {k}})\), with vertex set \({\mathcal {V}}\), edge set \(\,{\mathcal {R}}\) and edge labels \({\mathbf {k}}\in {\mathbb {R}}_{>0}^{\#{\mathcal {R}}}\). If \((y,y')\in {\mathcal {R}}\), we note \(y\rightarrow y'\). The complexes determine vectors in \({\mathbb {Z}}_{\ge 0}^{s}\) (the coefficients of the linear combinations) according to the stoichiometry of the species they consist of. We identify each complex with its corresponding vector and also with the formal linear combination of species specified by its coordinates.

We present a basic example that illustrates how a chemical reaction network gives rise to a dynamical system. This example represents a classical mechanism of enzymatic reactions, usually known as the futile cycle (Huang and Ferrell 1996; Kholodenko 2000; Wang and Sontag 2008):

Example 3

Consider the following graph

$$\begin{aligned} E+S_0 \overset{a}{\underset{b}{\rightleftarrows }} U \overset{c}{\rightarrow } E+S_1 \qquad F+S_1 \overset{{\tilde{a}}}{\underset{{\tilde{b}}}{\rightleftarrows }} V \overset{{\tilde{c}}}{\rightarrow } F+S_0. \end{aligned}$$

The \(s=6\) variables U, V, \(S_0\), \(S_1\), E, F denote the chemical species. The source and the product of each reaction (i.e. the vertices) are the complexes (non negative linear combinations of the species). Finally, the edge labels in \({\mathbf {k}}=(a,b,c,{\tilde{a}},{\tilde{b}},{\tilde{c}})\) are the reaction rate constants describing how concentrations of the six species change in time as the reactions occur.

The first three complexes give rise to the vectors (0, 0, 1, 0, 1, 0), (1, 0, 0, 0, 0, 0) and (0, 0, 0, 1, 1, 0), while those in the second ones are (0, 0, 0, 1, 0, 1), \((0,1,0,0,0,0)\), and (0, 0, 1, 0, 0, 1).

A chemical reaction network G as above under the assumption of mass-action kinetics induces a polynomial dynamical system in the following way. Suppose that the species are \(X_1,\ldots ,X_s\) and their respective concentrations are denoted by \(x_1,\ldots ,x_s\) (denoted by small letters). We write \(k_{yy'}\) for the reaction rate of each reaction \(y\rightarrow y'\) in \({\mathcal {R}}\). We introduce the following chemical reaction dynamical system:

$$\begin{aligned} {{\dot{{\mathbf {x}}}}}~=~\left( \frac{\text {d}x_1}{\text {d}t} ,\frac{\text {d}x_2}{\text {d}t} ,\dots , \frac{\text {d}x_s}{\text {d}t} \right) ~=~ \underset{y\rightarrow y'}{\sum } k_{yy'} \, {\mathbf {x}}^y \, (y'-y), \end{aligned}$$
(1)

where \({\mathbf {x}}:=(x_1,\dots ,x_s)\) and \({\mathbf {x}}^y:=x_1^{y_1}\cdots x_s^{y_s}\) if \(y=(y_1,\ldots ,y_s)\). The right-hand side of each differential equation \({\dot{x}}_i\) is a polynomial \(f_i({\mathbf {x}},{\mathbf {k}})\), in the variables \(x_1,\dots , x_s\) with coefficients depending on the parameters \({\mathbf {k}}:=(k_{yy'})_{(y,y')\in {\mathcal {R}}}\).

For instance, in Example 3 this induced dynamical system is:

$$\begin{aligned} \begin{array}{rcll} {\dot{u}} &{} = &{} a es_0-(b+c)u,\\ {\dot{v}} &{} = &{}{\tilde{a}} fs_1-({\tilde{b}}+{\tilde{c}}) v,\\ {\dot{s}}_0 &{} = &{}-a es_0+b u+{\tilde{c}} v,\\ {\dot{s}}_1&{} = &{}-{\tilde{a}} fs_1+{\tilde{b}} v+c u,\\ {\dot{e}} &{} = &{}-a es_0+(b+c) u,\\ {\dot{f}} &{} = &{}-{\tilde{a}} fs_1+({\tilde{b}}+{\tilde{c}}) v. \end{array} \end{aligned}$$
(2)

2.2 Identifiability in Chemical Reaction Systems

Among all the different (not always equivalent) notions of identifiability in differential equations and control theory, we have chosen to work from the one introduced in Craciun and Pantea (2008) since it seems specially well suited to the dynamical biochemical systems we consider here (see, for instance, Chis et al. 2011a; Raue et al. 2014 for a survey on the state of the art).

One of the main differences in the various approaches to identifiability is an assumption on the number of experiments that can be conducted with the same parameter values but different initial conditions: a single-experiment approach assumes the experiment is performed only once with some (often generic) initial condition (see, for example, DiStefano 2014, Chapter 10), whereas the multi-experiment approach we adopt in this paper assumes that it is allowed to perform as many experiments as needed with the same parameter values but different initial conditions.

Definition 1

Let \(G = ({\mathcal {V}},{\mathcal {R}}, {\mathbf {k}})\) be a chemical reaction network with s species. Its associated reaction system (1) is called identifiable if the map \(\varPhi :{\mathbb {R}}_{>0}^{\# {\mathcal {R}}}\rightarrow {\mathbb {R}}[{\mathbf {x}}]^s\),

$$\begin{aligned} \varPhi ({\mathbf {k}})=\underset{y\rightarrow y'}{\sum } k_{yy'}{\mathbf {x}}^y(y'-y), \end{aligned}$$

is injective (here \({\mathbf {k}}=(k_{yy'})_{(y,y')\in {\mathcal {R}}}\) and \({\mathbb {R}}[{\mathbf {x}}]\) is the polynomial ring in the variables \(x_1,\ldots ,x_s\)).

Example 4

In Example 3 [see the corresponding differential equation system (2)], the domain of the map \(\varPhi \) is \({\mathbb {R}}_{>0}^6\), the target space is \({\mathbb {R}}[u,v,s_0,s_1,e,f]^6\) and the coordinate functions are the right-hand sides of the differential equations in (2). It is clear that \(\varPhi \) is injective and therefore, the reaction system is identifiable: the right-hand sides of \({\dot{s}}_0\) and \({\dot{s}}_1\) determine the six constants \({\mathbf {k}}=(a,b,c,{\tilde{a}},{\tilde{b}},{\tilde{c}})\).

Example 5

(see Craciun and Pantea 2008, Section 2, Fig. 1) Consider the following graph

Here \(s=2\), \(\# {\mathcal {R}}=6\) and the associated dynamical system is

$$\begin{aligned} \begin{array}{ccc} {\dot{x}}_1&{}=&{}(2k_1+k_4)x_2^2-(k_2+2k_6)x_1^2+(k_5-k_3)x_1x_2\\ {\dot{x}}_2&{}=&{}-(2k_1+k_4)x_2^2+(k_2+2k_6)x_1^2+(k_3-k_5)x_1x_2 \end{array}. \end{aligned}$$
(3)

Clearly, the map \(\varPhi \) is not injective: parameters \({\mathbf {k}}\in {\mathbb {R}}_{>0}^6\) define the same polynomials under \(\varPhi \) if and only if the linear forms \(2k_1+k_4\), \(k_2+2k_6\) and \(k_5-k_3\) take the same values when evaluated at \({\mathbf {k}}\). For instance, \(\varPhi (1,1,1,1,1,1)=\varPhi (1,1,2,1,2,1)=(3x_2^2-3x_1^2,-3x_2^2+3x_1^2)\). Therefore, the system (3) is not identifiable.

Definition 2

For a chemical reaction network G, we introduce the total derivative (or Lie derivative) associated to the induced differential equations system as follows: given a differentiable function \(\varphi :{\mathbb {R}}^s\rightarrow {\mathbb {R}}\), its total derivative \(\dot{\varphi }\) is defined as

$$\begin{aligned} \dot{\varphi }:=\displaystyle {\sum _{i=1}^s \dfrac{\partial \varphi }{\partial x_i} \dfrac{d x_i}{d t}}=\sum _{i=1}^s \dfrac{\partial \varphi }{\partial x_i} \sum _{y\rightarrow y'} k_{yy'}{\mathbf {x}}^y(y'_i-y_i), \end{aligned}$$

where each partial derivative \(\dfrac{\text {d} x_i}{\text {d} t}\) is replaced according to system (1). For an integer \(\ell \ge 1\), we denote by \(\varphi ^{(\ell )}\) the \(\ell \)th iteration of the total derivative of \(\varphi \) (in particular \(\varphi ^{(1)}=\dot{\varphi })\).

For instance, for the network given in Example 3, its associated dynamical system (2) and the function \(\varphi =u^4+v\), we have

$$\begin{aligned} \dot{\varphi }=4u^3(a es_0-(b+c)u)+{\tilde{a}} fs_1-({\tilde{b}}+{\tilde{c}})v. \end{aligned}$$

Note that for a differentiable function \(\varphi :{\mathbb {R}}^s\rightarrow {\mathbb {R}}\), the total derivative \(\varphi ^{(\ell )}\) can be regarded as a function depending on the \((s+\#{\mathcal {R}})\)-variables \({\mathbf {x}},{\mathbf {k}}\).

Definition 3

Let \(G = ({\mathcal {V}},{\mathcal {R}}, {\mathbf {k}})\) be a chemical reaction network with s species. We say its associated reaction system (1) is identifiable from the variables\(x_{i_1},\dots ,x_{i_t}\) if there exists a positive integer D such that the following injectivity condition holds: if \({\mathbf {k}}^{*},{\mathbf {k}}^{**}\in {\mathbb {R}}_{>0}^{\# {\mathcal {R}}}\) verify

$$\begin{aligned} x^{(\ell )}_{i_j}({\mathbf {x}},{\mathbf {k}}^{*})=x^{(\ell )}_{i_j}({\mathbf {x}},{\mathbf {k}}^{**}), \end{aligned}$$

for all \(1\le \ell \le D\), \(1\le j\le t\), then \({\mathbf {k}}^{*}={\mathbf {k}}^{**}\).

The introduction of the Lie derivative in identifiability is a usual and quite natural approach suitable adapted to our purposes (see, for instance, Chis et al. 2011a). Among other works following this approach, Sedoglavic (2002), Chiş et al. (2011b) and Anguelova et al. (2012) also include a discussion about the number of derivatives needed for the proposed identifiability analysis.

Definitions 1 and 3 are related in the obvious way:

Proposition 1

A chemical reaction system in the variables \({\mathbf {x}}=x_1,\ldots , x_s\) is identifiable in the sense of Definition 1 if and only it is identifiable from the variables \(x_1,\ldots , x_s\) in the sense of Definition 3.

Proof

First we observe that the identity \(\varPhi ={\dot{x}}_1\times {\dot{x}}_2\times \cdots \times {\dot{x}}_s\) holds as functions of the argument \({\mathbf {k}}\). Thus, if \(\varPhi \) is injective, the condition of Definition 3 is satisfied for the variables \(x_1,\ldots , x_s\) and the integer \(D=1\). Conversely, suppose that the chemical reaction system is identifiable from the variables \(x_1,\ldots , x_s\) using a certain number D of successive total derivatives. Then the function \(\varPhi \) is necessarily injective in the arguments \({\mathbf {k}}\): if it is not the case, there exist \({\mathbf {k}}^{*}\ne {\mathbf {k}}^{**}\) such that \({\dot{x}}_i({\mathbf {x}},{\mathbf {k}}^{*})={\dot{x}}_i({\mathbf {x}},{\mathbf {k}}^{**})\) as functions of the variables \({\mathbf {x}}\) for all \(i=1,\ldots ,s\). Since the values of \({\mathbf {k}}^{*},{\mathbf {k}}^{**}\) are constants with respect to the total derivative we conclude that \(x_i^{(\ell )}({\mathbf {x}},{\mathbf {k}}^{*})=x_i^{(\ell )}({\mathbf {x}},{\mathbf {k}}^{**})\) for all \(\ell \in {\mathbb {N}}\) and all \(1\le i \le s\), arriving at a contradiction. \(\square \)

Example 6

Consider the graph

$$\begin{aligned} X_1+X_2 \overset{k_1}{\longrightarrow } X_3 \overset{k_2}{\longrightarrow } X_4. \end{aligned}$$

and its associated system

$$\begin{aligned} {\dot{x}}_1=-k_1x_1x_2,\quad {\dot{x}}_2=-k_1x_1x_2,\quad {\dot{x}}_3=k_1x_1x_2-k_2x_3,\quad {\dot{x}}_4=k_2x_3. \end{aligned}$$

The system is identifiable in the sense of Definition 1. Following Definition 3, the system is identifiable from the single variable \(x_3\) with one derivative (i.e. in this case \(D=1\) in Definition 3). It is also identifiable from the variable \(x_4\), but its total derivative of second order is needed in order to determine all the parameters (i.e. \(D=2\) for this variable). On the other hand, the system is not identifiable from the set of variables \(\{x_1,x_2\}\), since the constant \(k_2\) does not appear in any of the successive total derivatives of \(x_1\) nor \(x_2\).

For technical reasons, we need to slightly generalize the notion of identifiability introduced in Definition 3. The following definition is related to the notion of identifiability of parameter combinations (Boulier 2007; Meshkat et al. 2009):

Definition 4

Let \(G = ({\mathcal {V}},{\mathcal {R}}, {\mathbf {k}})\) be a chemical reaction network. Let \(p\in {\mathbb {N}}\) and \(\psi : {\mathbb {R}}_{>0}^{\# {\mathcal {R}}}\rightarrow {\mathbb {R}}^p\) be a map from the space of parameters in an affine space \({\mathbb {R}}^p\). We say that the map \(\psi \) is identifiable from the variables\(x_{i_1},\dots ,x_{i_t}\) if there exists a positive integer D such that the following injectivity condition holds: if \({\mathbf {k}}^{*},{\mathbf {k}}^{**}\in {\mathbb {R}}_{>0}^{\# {\mathcal {R}}}\) verify

$$\begin{aligned} x^{(\ell )}_{i_j}({\mathbf {x}},{\mathbf {k}}^{*})=x^{(\ell )}_{i_j}({\mathbf {x}},{\mathbf {k}}^{**}), \end{aligned}$$

for all \(1\le \ell \le D\), \(1\le j\le t\), then \(\psi ({\mathbf {k}}^{*})=\psi ({\mathbf {k}}^{**})\).

Roughly speaking, Definition 4 says that the value of the function \(\psi \) is uniquely determined by the values of the successive derivatives \(x^{(\ell )}_{i_j}\).

Observe that the notion of identifiability of a system from the variables \(x_{i_1},\ldots ,x_{i_t}\) as it is defined in Definition 3 can be translated in the sense of Definition 4 as the identifiability of the function \(\psi :{\mathbb {R}}_{>0}^{\# {\mathcal {R}}}\rightarrow {\mathbb {R}}^{\# {\mathcal {R}}}\), \(\psi ({\mathbf {k}})={\mathbf {k}}\).

For instance, in the (non identifiable) Example 5, the function \(\psi : {\mathbb {R}}_{>0}^6\rightarrow {\mathbb {R}}^3\), defined as \(\psi ({\mathbf {k}}):=(2k_1+k_4,k_2+2k_6,k_5-k_3)\), is identifiable from \(x_1\) (or \(x_2\), or both variables). In this case we say simply that the constants \(2k_1+k_4,\ k_2+2k_6,\ k_5-k_3\) can be identified from \(x_1\).

This notion will be useful along the paper. We will typically consider very simple functions \(\psi \) whose coordinates are either the rate constants or the sum of all the rate constants leaving from one complex.

3 Assumptions on the Biochemical Reaction Networks

We will analyze the identifiability problem for a specific kind of chemical reaction networks. We start by describing the assumptions on the networks we will consider in the sequel.

First, we assume that the “building blocks” of the network have the following shape:

$$\begin{aligned} X_{1}+X_{2} \overset{a}{\underset{b}{\rightleftarrows }} U \overset{c}{\rightarrow } X_{1}+X_{3}, \end{aligned}$$

where U is a species that only participates in those three reactions along all the network. We call U an intermediate species, and we say that species \(X_1\) acts as an enzyme, species \(X_2\) acts as a substrate and species \(X_3\) acts as a product.

Definition 5

We say an intermediate species Ureacts to the non-intermediate species \(X_1\) if there exists another non-intermediate species \(X_2\) such that the reaction \(U\rightarrow X_1+X_2\) exists. We say the non-intermediate species \(X_1\)reacts with the non-intermediate species \(X_2\) if there exists an intermediate species U such that the reaction \(X_1+X_2\rightarrow U\) exists.

Example 7

(Example 2 continued) Species \(U_1,V_1,U_2,V_2\) are the intermediate species. E and F act as enzymes. \(S_{1,0}\) acts as a substrate in the first connected component and as a product in the second one. Species \(S_{2,0}\) and \(S_{2,1}\) also act as both substrates and products (in the third and fourth connected components). Finally, \(S_{1,1}\) acts as a product in the first connected component, as a substrate in the second one, and as an enzyme in the third one.

We make the following assumption concerning the structure of the network:

Assumption 1

  1. 1.

    Each connected component of the graph is of the following form:

    $$\begin{aligned}Y+S_0\overset{a_1}{\underset{b_1}{\rightleftarrows }} U_1 \overset{c_1}{\rightarrow } Y+S_1\overset{a_2}{\underset{b_2}{\rightleftarrows }} U_2 \overset{c_2}{\rightarrow } \dots Y+S_{L-1}\overset{a_L}{\underset{b_L}{\rightleftarrows }} U_L \overset{c_{L}}{\rightarrow }Y+S_L, \end{aligned}$$

    where there is a unique enzyme Y acting on all the reactions of the connected component.

  2. 2.

    The intermediate species \(U_j\) appearing in the entire network are all different.

  3. 3.

    The non-intermediate species \(S_j\) in each connected component are all different, but they may also appear in other connected components.

  4. 4.

    Each complex lies in a unique connected component of the network.

Although the above assumption seems restrictive, it is satisfied by many networks such as the multisite phosphorylation system described in Example 1, the phosphorylation cascades as the one described in Example 2 and also the network in Example 3. As we observed before, in Examples 1 and 3 each species plays a unique role but in Example 2 the species \(S_{1,1}\) acts alternatively as a product (in the first connected component), as a substrate (in the second one) and as an enzyme (in the third one).

For an intermediate species U, we call

$$\begin{aligned} {\mathscr {S}}_U=\{S: S \text { acts as a substrate or a product in the }\\ \text {connected component determined by } U \}. \end{aligned}$$

For instance, in Example 2 we have \({\mathscr {S}}_{U_1}={\mathscr {S}}_{V_1}=\{S_{1,0},S_{1,1}\}\) and \({\mathscr {S}}_{U_2}={\mathscr {S}}_{V_2}=\{S_{2,0},S_{2,1}\}\).

We finish our assumptions on the kind of graphs we consider with a slightly technical condition.

Assumption 2

There is a partition of the species of the graph, that is, a decomposition into nonempty disjoint subsets:

$$\begin{aligned} {\mathscr {S}}={\mathscr {S}}^{(0)} \bigsqcup {\mathscr {S}}^{(1)} \bigsqcup \dots \bigsqcup {\mathscr {S}}^{(M)}, \end{aligned}$$

where \(M \ge 2\), \(\sqcup \) denotes the disjoint union, \({\mathscr {S}}^{(0)}\) is the set of intermediate species and given an intermediate species U with Y acting as an enzyme in the corresponding connected component, there exists \(\alpha \ge 1\) with \({\mathscr {S}}_U\subseteq {\mathscr {S}}^{(\alpha )}\) and \(Y\notin {\mathscr {S}}^{(\alpha )}\).

Remark 1

Under Assumption 1, the new condition imposed on the graph by Assumption 2 implies the following fact: if \(X_1\) reacts with \(X_2\), then there exists \(\alpha \ne \beta \) such that \(X_1\in {\mathscr {S}}^{(\alpha )}\) and \(X_2\in {\mathscr {S}}^{(\beta )}\). In particular, if \(S_i\) and \(S_j\) are two substrates or products in the same connected component, the complex \(S_i + S_j\) is not present in the network.

Example 8

In Example 2 we can consider the following partition \({\mathscr {S}}^{(0)}=\{U_1,V_1,U_2, V_2\}\), \({\mathscr {S}}^{(1)}=\{S_{1,0},S_{1,1}\}\), \({\mathscr {S}}^{(2)}=\{S_{2,0},S_{2,1}\}\), \({\mathscr {S}}^{(3)}=\{E\}\), \({\mathscr {S}}^{(4)}=\{F_1\}\), \({\mathscr {S}}^{(5)}=\{F_2\}\).

However it is not the unique possible partition: for instance, another choice could be \({\mathscr {S}}^{(0)}, {\mathscr {S}}^{(1)}\) and \({\mathscr {S}}^{(2)}\) as before, but \({\mathscr {S}}^{(3)}\), \({\mathscr {S}}^{(4)}\) and \({\mathscr {S}}^{(5)}\) are replaced by the single set \(\{E,F_1,F_2\}\).

4 Identifiability in Connected Components

This section is devoted to dealing with the identifiability problem for chemical reaction networks satisfying the assumptions stated in Sect. 3. Our aim is to show that all reaction constants of the network can be identified from the successive derivatives of the variables in a certain family of non-intermediates.

In order to do this, we choose a suitable subset of variables and estimate the maximum number of successive derivatives of them that we need to identify all the reaction constants. Namely, we choose variables \(x_{i_1},\dots , x_{i_t}\) and determine a number \(D_j\) of successive derivatives of \(x_{i_j}\), for \(1\le j\le t\), so that the injectivity condition in Definition 3 holds for \(D = \max \{D_j\}\).

Since the derivatives \(x_{i_j}^{(\ell )}({\mathbf {x}}, {\mathbf {k}})\) are polynomials in the variables \({\mathbf {x}}\) with coefficients that are polynomials in the reaction rate constants \({\mathbf {k}}\), showing that the parameters \({\mathbf {k}}\) are identifiable from \(x_{i_j}^{(\ell )}({\mathbf {x}}, {\mathbf {k}})\) for \(1\le \ell \le D_j\), \(1\le j \le t\), is the same as showing that they are uniquely determined by the coefficients of the polynomials \(x_{i_j}^{(\ell )}({\mathbf {x}}, {\mathbf {k}})\). Thus, our strategy to proving identifiability will be to locate suitable subsets of monomials in the derivatives \(x_{i_j}^{(\ell )}\) that enable us to prove that the values of all the reaction constants can be uniquely determined from their corresponding coefficients.

4.1 Identifying the Constants in One Connected Component from One Variable

The aim of this section is to show that all the reaction constants in a connected component

$$\begin{aligned} Y+S_0\overset{a_1}{\underset{b_1}{\rightleftarrows }} U_1 \overset{c_1}{\rightarrow } Y+S_1\overset{a_2}{\underset{b_2}{\rightleftarrows }} U_2 \overset{c_2}{\rightarrow } \dots Y+S_{L-1}\overset{a_L}{\underset{b_L}{\rightleftarrows }} U_L \overset{c_{L}}{\rightarrow }Y+S_L \end{aligned}$$
(4)

of a network satisfying the assumptions stated in Sect. 3 are identifiable from a limited number of successive derivatives of the variable \(s_L\) representing the concentration of the last product.

We start by showing that all the constants \(c_L, a_L, b_L\), and, for \(1\le j \le L-1\), \(a_j\) and \(b_j+c_j\) can be identified (in the sense of Definition 4) simply from the first three derivatives of this variable. Then, we proceed to identify recursively all the constants \(c_{j}\) (and consequently, also the constants \(b_j\)) for \(j=L-1,\dots , 1\), from higher-order derivatives of \(s_L\). The main result of this section is the following:

Proposition 2

All the constants in a connected component

$$\begin{aligned} Y+S_0\overset{a_1}{\underset{b_1}{\rightleftarrows }} U_1 \overset{c_1}{\rightarrow } Y+S_1\overset{a_2}{\underset{b_2}{\rightleftarrows }} U_2 \overset{c_2}{\rightarrow } \dots Y+S_{L-1}\overset{a_L}{\underset{b_L}{\rightleftarrows }} U_L \overset{c_{L}}{\rightarrow }Y+S_{L} \end{aligned}$$

of a network satisfying the assumptions in Sect. 3 can be identified from \(s^{(\ell )}_{L}\) with \(1\le \ell \le \mathrm {max}\{2,2L-1\}\).

The strategy in the proof of this result consists in the exact computation of the coefficients of certain distinguished monomials in the successive derivatives of \(s_L\). This explicit computation enables us to achieve the identifiability of all the constants of the connected component by means of a recursive procedure that we summarize in Table 1. For a complete proof, see Proposition 4 in “Appendix A.”

Table 1 The constants in the connected component (4) can be identified from \(s_L\)

We illustrate the procedure underlying the proof of the previous statement with a simple example.

Example 9

Consider the network

$$\begin{aligned}\begin{array}{c} Y+S_0\overset{a_1}{\underset{b_1}{\rightleftarrows }} U_1 \overset{c_1}{\rightarrow } Y+S_1\overset{a_2}{\underset{b_2}{\rightleftarrows }} U_2 \overset{c_2}{\rightarrow } Y+S_2\\ Z+S_2\overset{{\tilde{a}}_1}{\underset{{\tilde{b}}_1}{\rightleftarrows }} W \overset{{\tilde{c}}_1}{\rightarrow }Z+S_3 \end{array} \end{aligned}$$

According to Proposition 2, all the constants in the first connected component can be identified from \(s^{(\ell )}_2\) with \(1\le \ell \le 3\). In fact, if we call \(K_1=b_1+c_1\), \(K_2=b_2+c_2\) and \({\tilde{K}}_1={\tilde{b}}_1+{\tilde{c}}_1\),

where the constants \(c_2,a_2,K_2\) (thus, also \(b_2= K_2-c_2\)), \(a_1,K_1\) and \(c_1\) (thus, also \(b_1= K_1 - c_1\)) are identified in Table 1.

A direct consequence of Proposition 2 is the following theorem:

Theorem 1

If a chemical reaction network satisfying the assumptions in Sect. 3 consists of N connected components

$$\begin{aligned}&Y_1+S_{1,0}\overset{a_{1,1}}{\underset{b_{1,1}}{\rightleftarrows }} U_{1,1} \overset{c_{1,1}}{\rightarrow } Y_1 +S_{1,1}\overset{a_{1,2}}{\underset{b_{1,2}}{\rightleftarrows }} U_{1,2} \overset{c_{1,2}}{\rightarrow } \dots \\&\dots Y_1+S_{1,L_1-1}\overset{a_{1,L_1}}{\underset{b_{1,L_1}}{\rightleftarrows }} U_{1,L_1} \overset{c_{1,L_1}}{\rightarrow }Y_1+S_{1,L_1}\\&\vdots \\&Y_N+S_{N,0}\overset{a_{N,1}}{\underset{b_{N,1}}{\rightleftarrows }} U_{N,1} \overset{c_{N,1}}{\rightarrow } Y_N+S_{N,1}\overset{a_{N,2}}{\underset{b_{N,2}}{\rightleftarrows }} U_{N,2} \overset{c_{N,2}}{\rightarrow } \dots \\&\dots Y_N+S_{N,L_N-1}\overset{a_{N,L_N}}{\underset{b_{N,L_N}}{\rightleftarrows }} U_{N,L_N}\overset{c_{N,L_N}}{\rightarrow }Y_N+S_{N,L_N}, \end{aligned}$$

then the associated system is identifiable from the variables \(s_{1,L_1},\dots ,s_{N,L_N}\) corresponding to the last products of each connected component of the network. Moreover, for every \(1\le i \le N\), the order of derivation needed for the variable \(s_{i,L_i}\) is at most \(\max \{2,2L_i-1\}\).

4.2 Identifying the Constants in Two Connected Components from One Variable

In this subsection, we analyze the identifiability problem for a subclass of the networks we have been considering. More precisely, we consider networks containing pairs of connected components of the following type:

$$\begin{aligned} \begin{array}{c} Y+S_0\overset{a_1}{\underset{b_1}{\rightleftarrows }} U_1 \overset{c_1}{\rightarrow } Y+S_1\overset{a_2}{\underset{b_2}{\rightleftarrows }} U_2 \overset{c_2}{\rightarrow } \dots Y+S_{L-1}\overset{a_L}{\underset{b_L}{\rightleftarrows }} U_L \overset{c_{L}}{\rightarrow }Y+S_L,\\ {{\widetilde{Y}}} +S_L\overset{{\tilde{a}}_L}{\underset{{\tilde{b}}_L}{\rightleftarrows }} V_L \overset{{\tilde{c}}_L}{\rightarrow } {{\widetilde{Y}}}+S_{L-1}\overset{{\tilde{a}}_{L-1}}{\underset{{\tilde{b}}_{L-1}}{\rightleftarrows }} V_{L-1} \overset{{\tilde{c}}_{L-1}}{\rightarrow } \dots {{\widetilde{Y}}}+S_1\overset{{\tilde{a}}_1}{\underset{{\tilde{b}}_1}{\rightleftarrows }} V_1 \overset{{\tilde{c}}_1}{\rightarrow }{{\widetilde{Y}}}+S_0. \end{array} \end{aligned}$$
(5)

As before, we work under the assumptions made in Sect. 3.

By Proposition 2, we know that all the constants in the first connected component in (5) can be identified from a certain number of successive derivatives of \(s_L\). Using the specific structure of the second component, we can prove that the same derivatives also enable the identification of the reaction rate constants of that component.

We first prove that the constants \({\tilde{a}}_L, {\tilde{b}}_L, {\tilde{c}}_L\), and, for \(1\le j \le L-1\), \({\tilde{a}}_j\) and \({\tilde{b}}_j + {\tilde{c}}_j\) can be identified from \({\dot{s}}_L\) and \(\ddot{s}_L\) and, then, by means of a recursive explicit computation of coefficients of a family of distinguished monomials in higher-order derivatives of \(s_L\), we show how to successively identify the constants \({\tilde{b}}_j\) for \(j={L-1},\dots , 1\), and, consequently, also the constants \({\tilde{c}}_j\). In this way, we deduce:

Proposition 3

Given a chemical reaction network satisfying the assumptions in Sect. 3, all the constants in two connected components of the type

$$\begin{aligned} \begin{array}{c} Y+S_0\overset{a_1}{\underset{b_1}{\rightleftarrows }} U_1 \overset{c_1}{\rightarrow } Y+S_1\overset{a_2}{\underset{b_2}{\rightleftarrows }} U_2 \overset{c_2}{\rightarrow } \dots Y+S_{L-1}\overset{a_L}{\underset{b_L}{\rightleftarrows }} U_L \overset{c_{L}}{\rightarrow }Y+S_{L},\\ {\widetilde{Y}}+S_{L}\overset{{\tilde{a}}_L}{\underset{{\tilde{b}}_L}{\rightleftarrows }} V_L \overset{{\tilde{c}}_L}{\rightarrow } {\widetilde{Y}}+S_{L-1}\overset{{\tilde{a}}_{L-1}}{\underset{{\tilde{b}}_{L-1}}{\rightleftarrows }} V_{L-1} \overset{{\tilde{c}}_{L-1}}{\rightarrow } \dots {\widetilde{Y}}+S_1\overset{{\tilde{a}}_1}{\underset{{\tilde{b}}_1}{\rightleftarrows }} V_1 \overset{{\tilde{c}}_1}{\rightarrow }{\widetilde{Y}}+S_0 \end{array} \end{aligned}$$

can be identified from \(s^{(\ell )}_{L}\) with \(1\le \ell \le \mathrm {max}\{2,2L-1\}\).

We summarize the identifiability procedure underlying the proof of the previous proposition in Table 2, and we also illustrate the result in Example 10. For a complete proof, see Proposition 5 in “Appendix A.”

Table 2 The constants in the two connected components in (5) can be identified from \(s_L\)

Example 10

Consider the network

$$\begin{aligned} \begin{array}{c} Y+S_0\overset{a_1}{\underset{b_1}{\rightleftarrows }} U_1 \overset{c_1}{\rightarrow } Y+S_1\overset{a_2}{\underset{b_2}{\rightleftarrows }} U_2 \overset{c_2}{\rightarrow } Y+S_2 \\ {\tilde{Y}}+S_2\overset{{\tilde{a}}_2}{\underset{{\tilde{b}}_2}{\rightleftarrows }} V_2 \overset{{\tilde{c}}_2}{\rightarrow } {\tilde{Y}}+S_1\overset{{\tilde{a}}_1}{\underset{{\tilde{b}}_1}{\rightleftarrows }} V_1 \overset{{\tilde{c}}_1}{\rightarrow } {\tilde{Y}}+S_0 \end{array} \end{aligned}$$

According to Proposition 3, all the constants in the two connected components can be identified from \(s^{(\ell )}_2\) with \(1\le \ell \le 3\). In fact, if we call \(K_1=b_1+c_1\), \(K_2=b_2+c_2\), \({\tilde{K}}_1={\tilde{b}}_1+{\tilde{c}}_1\) and \({\tilde{K}}_2={\tilde{b}}_2+{\tilde{c}}_2\):

Here, the constants \(c_2\), \({\tilde{a}}_2, {\tilde{b}}_2\), \(a_2, K_2\) (then, \(b_2\)), \({\tilde{c}}_2 \), \({\tilde{a}}_1, {\tilde{K}}_1\), \(a_1\), \(K_1\), \(c_1\) (then, \(b_1\)) and \({\tilde{b}}_1\) (then, \({\tilde{c}}_1\)) are identified in Table 2.

A direct consequence of Proposition 3 is the following corollary:

Corollary 1

If a chemical reaction network satisfying the assumptions in Sect. 3 consists of 2N connected components of the shape

$$\begin{aligned}&Y_1+S_{1,0}\overset{a_{1,1}}{\underset{b_{1,1}}{\rightleftarrows }} U_{1,1} \overset{c_{1,1}}{\rightarrow } Y_1+S_{1,1}\overset{a_{1,2}}{\underset{b_{1,2}}{\rightleftarrows }} U_{1,2} \overset{c_{1,2}}{\rightarrow } \dots \\&\dots Y_1+S_{1,L_1-1}\overset{a_{1,L_1}}{\underset{b_{1,L_1}}{\rightleftarrows }} U_{1,L_1} \overset{c_{1,L_1}}{\rightarrow }Y_1+S_{1,L_1}\\&{\widetilde{Y}}_1+S_{1,L_1}\overset{{\tilde{a}}_{1,L_1}}{\underset{{\tilde{b}}_{1,L_1}}{\rightleftarrows }} V_{1,L_1} \overset{{\tilde{c}}_{1,L_1}}{\rightarrow } {\widetilde{Y}}_1+S_{1,L_1-1}\overset{{\tilde{a}}_{1,L_1-1}}{\underset{{\tilde{b}}_{1,L_1-1}}{\rightleftarrows }} V_{1,L_1-1} \overset{{\tilde{c}}_{1,L_1-1}}{\rightarrow } \dots \\&\dots {\widetilde{Y}}_1+S_{1,1}\overset{{\tilde{a}}_{1,1}}{\underset{{\tilde{b}}_{1,1}}{\rightleftarrows }} V_{1,1} \overset{{\tilde{c}}_{1,1}}{\rightarrow }{\widetilde{Y}}_1+S_{1,0}\\&\vdots \\&Y_N+S_{N,0}\overset{a_{N,1}}{\underset{b_{N,1}}{\rightleftarrows }} U_{N,1} \overset{c_{N,1}}{\rightarrow } Y_N +S_{N,1}\overset{a_{N,2}}{\underset{b_{N,2}}{\rightleftarrows }} U_{N,2} \overset{c_{N,2}}{\rightarrow } \dots \\&\quad \dots Y_N+S_{N,L_N-1}\overset{a_{N,L_N}}{\underset{b_{N,L_N}}{\rightleftarrows }} U_{N,L_N} \overset{c_{N,L_N}}{\rightarrow }Y_N+S_{N,L_N}\\&{\widetilde{Y}}_N+S_{N,L_N}\overset{{\tilde{a}}_{N,L_N}}{\underset{{\tilde{b}}_{N,L_N}}{\rightleftarrows }} V_{N,L_N} \overset{{\tilde{c}}_{N,L_N}}{\rightarrow } {\widetilde{Y}}_N+S_{N,L_N-1}\overset{{\tilde{a}}_{N,L_N-1}}{\underset{{\tilde{b}}_{N,L_N-1}}{\rightleftarrows }} V_{N,L_N-1} \overset{{\tilde{c}}_{N,L_N-1}}{\rightarrow } \dots \\&\dots {\widetilde{Y}}_N +S_{N,1}\overset{{\tilde{a}}_{N,1}}{\underset{{\tilde{b}}_{N,1}}{\rightleftarrows }} V_{N,1} \overset{{\tilde{c}}_{N,1}}{\rightarrow }{\widetilde{Y}}_N+S_{N,0} \end{aligned}$$

then the associated system is identifiable from the variables \(s_{1,L_1},\dots ,s_{N,L_N}\). Moreover, for every \(1\le i \le N\), the order of derivation needed for the variable \(s_{i,L_i}\) is at most \(\max \{2,2L_i-1\}\).

5 Identifying the Cascade

We will consider in this section networks that are called cascades. Signaling cascades are biochemical networks of major biological importance as they participate in a number of several diseases and also control key cellular functions (Davis 2000; Kyriakis and Avruch 2001; Pearson et al. 2001; Schaeffer and Weber 1999; Widmann et al. 1999; Zarubin and Han 2005). The mitogen-activated protein kinase (MAPK) cascade is a network present in all eukaryotic cells and one of the most extensively modeled signaling systems (Hornberg et al. 2005; Huang and Ferrell 1996; Qiao et al. 2007). A schematic representation of the network is the following

figure b

where \(S_{1,0}\) represents the kinase MAPKKK, and \(S_{1,1}\) represents the activated form MAPKKK\(^*\). \(S_{2,0}\), \(S_{2,1}\) and \(S_{2,2}\) stand for MAPKK, MAPKK-P and MAPKK-PP, respectively. And finally, \(S_{3,0}\), \(S_{3,1}\) and \(S_{3,2}\) stand for MAPK, MAPK-P and MAPK-PP, respectively. \(F_1\) represents the enzyme that deactivates MAPKKK\(^*\), and \(F_2\) and \(F_3\) represent the corresponding phosphatase of each layer.

More generally, cascades consist of \(N\ge 1\) layers and are represented by the following scheme:

figure c

One important feature of cascades is that the enzyme on the first connected component of a certain layer is the last product of the first component of the previous layer. For instance, \(S_{1,L_1}\) is the enzyme on the second layer and so on. The corresponding reaction network for the N-layer cascade is the following

$$\begin{aligned}&E+S_{1,0} \; {\overset{a_{1,1}}{\underset{b_{1,1}}{\rightleftarrows }}} \; U_{1,1} \, {\overset{c_{1,1}}{\rightarrow }} \; E+S_{1,1}\; {\overset{a_{1,2}}{\underset{b_{1,2}}{\rightleftarrows }}} \; U_{1,2} \, {\overset{c_{1,2}}{\rightarrow }} \dots \nonumber \\&\dots E+S_{1,L_1-1}{\overset{a_{1,L_1}}{\underset{b_{1,L_1}}{\rightleftarrows }}} U_{1,L_1} {\overset{c_{1,L_1}}{\rightarrow }}E+S_{1,L_1} \nonumber \\&F_1+S_{1,L_1}{\overset{{\tilde{a}}_{1,L_1}}{\underset{{\tilde{b}}_{1,L_1}}{\rightleftarrows }}} V_{1,L_1} {\overset{{\tilde{c}}_{1,L_1}}{\rightarrow }} F_1+ S_{1, L_1-1} \; {\overset{{\tilde{a}}_{1,L_1-1}}{\underset{{\tilde{b}}_{1,L_1-1}}{\rightleftarrows }}} \; V_{1,L_1-1} {\overset{{\tilde{c}}_{1,L_1-1}}{\rightarrow }}\dots \nonumber \\&\dots F_1+S_{1,1} {\overset{{\tilde{a}}_{1,1}}{\underset{{\tilde{b}}_{1,1}}{\rightleftarrows }}} V_{1,1} {\overset{{\tilde{c}}_{1,1}}{\rightarrow }} F_1+S_{1,0} \nonumber \\&S_{1,L_1}+S_{2,0}\; {\overset{a_{2,1}}{\underset{b_{2,1}}{\rightleftarrows }}} \; U_{2,1} \, {\overset{c_{2,1}}{\rightarrow }} \; S_{1,L_1}+S_{2,1}{\overset{a_{2,2}}{\underset{b_{2,2}}{\rightleftarrows }}} \; U_{2,2} \, {\overset{c_{2,2}}{\rightarrow }} \dots \nonumber \\&\dots S_{1,L_1} +S_{2,L_2-1}{\overset{a_{2,L_2}}{\underset{b_{2,L_2}}{\rightleftarrows }}} U_{2,L_2} {\overset{c_{2,L_2}}{\rightarrow }} S_{1,L_1}+S_{2,L_2} \nonumber \\&F_2+S_{2,L_2}{\overset{{\tilde{a}}_{2,L_2}}{\underset{{\tilde{b}}_{2,L_2}}{\rightleftarrows }}} V_{2,L_2} {\overset{{\tilde{c}}_{2,L_2}}{\rightarrow }} F_2+ S_{2, L_2-1} \; {\overset{{\tilde{a}}_{2,L_2-1}}{\underset{{\tilde{b}}_{2,L_2-1}}{\rightleftarrows }}} \; V_{2,L_2-1} {\overset{{\tilde{c}}_{2,L_2-1}}{\rightarrow }} \dots \nonumber \\&\dots F_2+S_{2,1} {\overset{{\tilde{a}}_{2,1}}{\underset{{\tilde{b}}_{2,1}}{\rightleftarrows }}} V_{2,1} {\overset{{\tilde{c}}_{2,1}}{\rightarrow }}F_2+S_{2,0} \nonumber \\&\vdots \nonumber \\&S_{N-1,L_{N-1}}+S_{N,0}{\overset{a_{N,1}}{\underset{b_{N,1}}{\rightleftarrows }}} U_{N,1} {\overset{c_{N,1}}{\rightarrow }} \dots \nonumber \\&\dots S_{N-1,L_{N-1}} +S_{N,L_N-1} {\overset{a_{N,L_N}}{\underset{b_{N,L_N}}{\rightleftarrows }}} U_{N,L_N} {\overset{c_{n,L_n}}{\rightarrow }} S_{N-1,L_{N-1}}+S_{N,L_N}\nonumber \\&F_N+S_{N,L_N}{\overset{{\tilde{a}}_{N,L_N}}{\underset{{\tilde{b}}_{N,L_N}}{\rightleftarrows }}} V_{N,L_N} {\overset{{\tilde{c}}_{N,L_N}}{\rightarrow }} F_N+ S_{N, L_N-1} \; {\overset{{\tilde{a}}_{N,L_N-1}}{\underset{{\tilde{b}}_{N,L_N-1}}{\rightleftarrows }}} \; V_{N,L_N-1} {\overset{{\tilde{c}}_{N,L_N-1}}{\rightarrow }}\dots \nonumber \\&\dots F_N +S_{N,1}{\overset{{\tilde{a}}_{N,1}}{\underset{{\tilde{b}}_{N,1}}{\rightleftarrows }}} V_{N,1} {\overset{{\tilde{c}}_{N,1}}{\rightarrow }} F_N+S_{N,0}. \end{aligned}$$
(6)

We will assume \(F_i\ne F_j\) if \(i\ne j\) and consider the following partition of the non-intermediate species, which satisfies Assumption 2:

$$\begin{aligned} {\mathscr {S}}={\mathscr {S}}^{(1)} \bigsqcup {\mathscr {S}}^{(2)} \bigsqcup \dots \bigsqcup {\mathscr {S}}^{(2N+1)}, \end{aligned}$$

with \({\mathscr {S}}^{(m)}=\{S_{m,0}, \dots ,S_{m,L_m} \}\) and \({\mathscr {S}}^{(N+m)}=\{F_m\}\), for \(1\le m \le N\), and \({\mathscr {S}}^{(2N+1)}=\{E\}\).

As our running example for this section, we will consider the two-layer cascade with 18 reactions.

Example 11

$$\begin{aligned}&E+S_{1,0} \overset{a_{1,1}}{\underset{b_{1,1}}{\rightleftarrows }} U_{1,1} \overset{c_{1,1}}{\rightarrow } E+S_{1,1} \\&F_1+S_{1,1} \overset{{\tilde{a}}_{1,1}}{\underset{{\tilde{b}}_{1,1}}{\rightleftarrows }} V_{1,1} \overset{{\tilde{c}}_{1,1}}{\rightarrow } F_1+S_{1,0} \\&S_{1,1}+S_{2,0} \overset{a_{2,1}}{\underset{b_{2,1}}{\rightleftarrows }} U_{2,1} \overset{c_{2,1}}{\rightarrow } S_{1,1}+S_{2,1} \overset{a_{2,2}}{\underset{b_{2,2}}{\rightleftarrows }} U_{2,2} \overset{c_{2,2}}{\rightarrow } S_{1,1}+S_{2,2} \\&F_2+S_{2,2} \overset{{\tilde{a}}_{2,2}}{\underset{{\tilde{b}}_{2,2}}{\rightleftarrows }} V_{2,2} \overset{{\tilde{c}}_{2,2}}{\rightarrow } F_2+S_{2,1} \overset{{\tilde{a}}_{2,1}}{\underset{{\tilde{b}}_{2,1}}{\rightleftarrows }} V_{2,1} \overset{{\tilde{c}}_{2,1}}{\rightarrow } F_2+S_{2,0}. \end{aligned}$$

The first layer consists of two connected components. The first component consists of one modification performed by the enzyme E on the substrate \(S_{1,0}\), which is transformed into the product \(S_{1,1}\). On the second connected component, the enzyme \(F_1\) performs the reverse modification on the substrate \(S_{1,1}\). The second layer is similar. For this network we have \({\mathscr {S}}^{(1)}=\{S_{1,0},S_{1,1}\}\), \({\mathscr {S}}^{(2)}=\{S_{2,0},S_{2,1},S_{2,2}\}\), \({\mathscr {S}}^{(3)}=\{F_1\}\), \({\mathscr {S}}^{(4)}=\{F_2\}\) and \({\mathscr {S}}^{(5)}=\{E\}\).

5.1 Identifiability of Constants in a General Cascade

The aim of this section is to show that all the constants in the cascades introduced in (6) can be identified from successive derivatives of the variable corresponding to the last product of the last layer, \(S_{N,L_N}\). In order to prove this, we relate the derivatives of the last product of a given layer of the cascade with the derivatives of the last product of the layer immediately above.

To shorten notation, we will denote \(K_{m, j} = b_{m,j}+ c_{m,j}\) and \({\tilde{K}}_{m, j} = {\tilde{b}}_{m,j}+ {\tilde{c}}_{m,j}\) for every \(1\le m\le N\), \(1\le j \le L_m\). Also, for unifying purposes, we set \(S_{0,L_{0}}:= E\).

For \(1\le n\le N\), consider the variable \(s_{n, L_n}\) corresponding to the last product of the nth layer of the cascade. We have that

$$\begin{aligned} {\dot{s}}_{n,L_n}= & {} c_{n,L_n} u_{n,L_n} -{\tilde{a}}_{n,L_n} s_{n,L_n} f_n+{\tilde{b}}_{n,L_n}v_{n,L_n}\\&- \sum _{j=1}^{L_{n+1}} a_{n+1,j} s_{n, L_n} s_{n+1, j-1}+ \sum _{j=1}^{L_{n+1}} K_{n+1,j} u_{n+1,j} \end{aligned}$$

and, for \(n=N\), only the three first terms appear in the derivative, i.e.\(a_{N+1,j}=0\), \(K_{N+1,j}=0\) for all j. The second derivative of \(s_{n,L_n}\) is

We can see that the variable \(s_{n-1, L_{n-1}}\) corresponding to the last product of the \((n-1)\)th layer appears in the second derivative of \(s_{n, L_n}\). More precisely, from the above expression, it follows easily that it only appears in the term \(c_{n, L_n} a_{n, L_n} s_{n-1, L_{n-1}} s_{n, L_n-1}\), since \(S_{n-1, L_{n-1}}\) does not react with or to \(F_n\) or \(S_{n+1, j}\) for any j. Thus, two differentiation steps enable us to “jump” from one layer of the cascade to the layer immediately above. Inductively, the idea is that, for \(m<n\), by taking \(2(n-m)\) derivatives of \(s_{n, L_n}\) we will reach the mth layer; that is, the variable \(s_{m, L_m}\) will appear and so, the successive derivatives of \(s_{m, L_m}\) will appear in higher-order derivatives of \(s_{n, L_n}\).

Now, by the results in Sect. 4.2 for the case of two connected components of the form (5), we can identify all constants in the mth layer of the cascade by looking at the coefficients of certain monomials of the derivatives of \(s_{m, L_m}\). Then, our previous considerations will imply that those constants can be identified from successive derivatives of \(s_{n, L_n}\) as well. In order to ensure that this can be achieved, we prove that certain monomials effectively appear in the derivatives of \(s_{n, L_n}\) and compute their coefficients (see Proposition 6 in “Appendix A” for a precise statement and its proof).

When considering the last product of the last layer of the cascade, we obtain our main result:

Theorem 2

All the constants in the network (6) can be identified from \(s^{(\ell )}_{N,L_N}\) with \(1\le \ell \le \max \{2N; 2(N-m+L_m)-1, \ 1\le m\le N\}\).

We now summarize the identifiability procedure which proves the previous theorem. The procedure obtains recursively, for \(m= N, N-1, \dots , 1\), the values of the constants \(a_{m, j}\), \({\tilde{a}}_{m, j}\), \(b_{m, j}\), \({\tilde{b}}_{m, j}\), \(c_{m, j}\) and \({\tilde{c}}_{m, j}\), for \(1\le j \le L_m\), from the successive derivatives of \(s_{N, L_N}\), according to Table 3.

In order to shorten notation, let \({\mathcal {P}}_N:=1\), \({\mathcal {C}}_N:= 1\), \({\mathcal {K}}_N:=0\) and, for \(1\le m \le N-1\), \({\mathcal {P}}_m:=\prod \nolimits _{i=m+1}^N s_{i, L_i-1}\), \({\mathcal {C}}_m :=\prod \nolimits _{i=m+1}^N c_{i, L_i}a_{i, L_i}\) and \({\mathcal {K}}_m := \sum \nolimits _{i=m+1}^N K_{i, L_i}\).

Table 3 The constants in the cascade can be identified from \(s_{N,L_N}\)

Example 12

(Example 11 continued) Here we find the monomials relevant for identifiability in the two-layer cascade. We highlight with blue boxes the constants that we are identifying in each derivative. We moreover highlight with green boxes the monomials that we used to identify \(b_{1,1}\) and \({\tilde{c}}_{1,1}\) from \(K_{1,1}\) and \({\tilde{K}}_{1,1}\), respectively (see rows 5 and 6 in Table 3).

5.2 An Example of How to Obtain the Rate Constants from Data

Here, we will illustrate our previous theoretical identifiability results in a specific example, showing how they can be used as a guidance in experimental design for practical parameter identification from observable data.

The three-layer cascade with \(L_1=1, L_2=L_3=2\) represents the well-known MAPK signaling cascade with \(s_{3,2}\) representing the concentration of the doubly phosphorylated kinase MAPK-PP (Catozzi et al. 2016; Huang and Ferrell 1996; Kholodenko 2000; Shaul and Seger 2007). Consider, then, the cascade (6) for \(N=3\) and \(L_1=1\), \(L_2=L_3=2\), whose schematic representation is introduced at the beginning of Sect. 5. In this case, we have 22 species concentrations \({\mathbf {x}}\) and 30 rate constants \({\mathbf {k}}\) which can be identified from \(s^{(\ell )}_{3,2}\), \(1\le \ell \le 6\), by Theorem 2. According to Definition 3, this means that if we consider the polynomial system

$$\begin{aligned} s_{3,2}^{(\ell )}({\mathbf {x}},{\mathbf {k}})=p_{\ell }({\mathbf {x}},{\mathbf {k}}) \end{aligned}$$
(7)

for the corresponding polynomials \(p_{\ell }\) obtained from (1) by computing the successive total derivatives of \(s_{3,2}\), the function that maps the vector of rate constants \({\mathbf {k}}\) to the coefficients of the polynomials \(p_{\ell }\)’s (considered as polynomials in the species concentrations \({\mathbf {x}}\)) is injective. This means that all the rate constants can be recovered from noise-free data by a suitable interpolation procedure: if we evaluate these polynomials at “sufficiently many” points \({\mathbf {x}}\in {\mathbb {R}}^{22}\), we may reconstruct the coefficients and, consequently, determine uniquely the values of the rate constants.

However, it is not clear which \({\mathbf {x}}\in {\mathbb {R}}^{22}\) are suitable for identifying the parameters of the system, nor how many of them are enough for this purpose. We give here a heuristic to choose a list of \({\mathbf {x}}\in {\mathbb {R}}^{22}\) based on the monomials in the second column of Table 4, which is the adapted version of Table 3 for this particular case. This heuristic can be used as an aid to design experiments to obtain the rate constants values. Each initial state \({\mathbf {x}}\in {\mathbb {R}}^{22}\) is in correspondence with a different experiment.

In order to recover the value of the 30 rate constants in this case, we propose the following algorithm:

  1. Step 1.

    Consider \({\mathbf {x}}_1,{\mathbf {x}}_2,\dots ,{\mathbf {x}}_{30}\in {\mathbb {R}}^{22}\) defined as follows: for the ith monomial in Table 4, consider \({\mathbf {x}}_i\in {\mathbb {R}}^{22}\) where all the coordinates are 0 except for those coordinates corresponding to variables that divide the monomial, which are equal to 1. For example, for the monomial \(u_{1,1}s_{2,1}s_{3,1}\), all the coordinates of the associated point are equal to 0, except for the three coordinates corresponding to \(u_{1,1}\), \(s_{2,1}\) and \(s_{3,1}\) that are equal to 1.

  2. Step 2.

    For each \(i\in \{1,\dots ,30\}\), obtain the value \(s_{3,2}^{(\ell )}({\mathbf {x}}_i,{\mathbf {k}})\) for the order \(\ell \) that corresponds to the ith monomial in Table 4. Ideally, these values should be obtained experimentally, for instance considering \({\mathbf {x}}_i\) the initial state at time \(t=0\).

  3. Step 3.

    Construct a (nonlinear) polynomial equation system from (7), of 30 equations in the 30 unknowns \({\mathbf {k}}\), by evaluating the right-hand sides at \({\mathbf {x}}_1,\dots ,{\mathbf {x}}_{30}\) and replacing the left-hand sides with the values obtained in the previous step.

  4. Step 4.

    Solve the polynomial system in the unknowns \({\mathbf {k}}\).

A vague explanation of why this heuristic works is that each monomial in Table 4 incorporates a new variable that comes paired with the new rate constant to be identified. Further research is needed to find a rigorous proof for this conjecture.

We implemented the algorithm above by reconstructing the values of the left-hand sides of (7) with the rate constants in the third column of Table S2 in the Supporting Information of Qiao et al. (2007). We used Maple (2014) to solve the system of equations and successfully obtained the following values (in a few seconds using a standard desktop computer).

$$\begin{aligned}&a_{1,1} = 337.2299998,\quad a_{2,1} = 1226.000001,\quad a_{2,2} = 3383.7,\\&a_{3,1}\! = 229.5699981,\quad a_{3,2} = 3388.7,\quad {\tilde{a}}_{1,1}\! = 1841.000002,\quad {\tilde{a}}_{2,1} \!=\! 2960.300016, \\&{\tilde{a}}_{2,2} = 1956.8,\quad {\tilde{a}}_{3,1} = 297.0,\quad {\tilde{a}}_{3,2} = 974.7,\\&b_{1,1} = 261.1000013,\quad b_{2,1} = 623.1700002,\quad b_{2,2} = 605.3100002,\\&b_{3,1} =694.13,\quad b_{3,2} = 485.3499999,\quad {\tilde{b}}_{1,1} = 198.47,\quad {\tilde{b}}_{2,1} = 163.0,\\&{\tilde{b}}_{2,2} = 48.804,\quad {\tilde{b}}_{3,1} = 301.09,\quad {\tilde{b}}_{3,2} = 587.45,\\&c_{1,1} = 146.07,\quad c_{2,1} = 420.0000001,\quad c_{2,2} = 214.65,\quad c_{3,1} = 43.658, \\&c_{3,2} = 65.732,\quad {\tilde{c}}_{1,1} = 338.4400021,\quad {\tilde{c}}_{2,1} = 668.2000111,\\&{\tilde{c}}_{2,2} = 67.97000003,\quad {\tilde{c}}_{3,1} = 31.743,\quad {\tilde{c}}_{3,2} = 175.91. \end{aligned}$$

The same three-layer cascade may be completely identified also by means of the result stated in Theorem 1: in this case the rate constants in each connected component can be identified from \(s^{(\ell _1)}_{1,1}\), \(s^{(\ell _2)}_{1,0}\), \(s^{(\ell _3)}_{2,2}\), \(s^{(\ell _4)}_{2,0}\), \(s^{(\ell _5)}_{3,2}\), and \(s^{(\ell _6)}_{3,0}\), respectively, for \(1\le \ell _1,\ell _2\le 2\) and \(1\le \ell _3,\ell _4,\ell _5,\ell _6\le 3\). By Corollary 1 we can also identify the constants from \(s^{(\ell _1)}_{1,1}\), \(s^{(\ell _3)}_{2,2}\) and \(s^{(\ell _5)}_{3,2}\), for \(1\le \ell _1\le 2\) and \(1\le \ell _3,\ell _5\le 3\). We adapted the procedure above and implemented it in Maple, and we obtained the same rate constants as before.

Throughout the article, we assume that one can use noise-free data in order to recover the rate constants values. Nevertheless, there are certain numerical errors that appear at Step 4, when the polynomial system in the unknowns \({\mathbf {k}}\) is solved. If we moreover implement the algorithm with numerical approximations of the total derivatives, more numerical errors are bound to occur. The major drawback of considering the last two approaches, based on Theorem 1 or Corollary 1, is that more species have to be measured. However, the value that has to be numerically estimated corresponds to a derivative of order at most three, which can be approximated more accurately and with fewer time measurements than those values of higher-order derivatives.

The Maple code for both procedures can be found at http://cms.dm.uba.ar/Members/mpmillan/identifiability.

Table 4 The constants in the three-layer cascade with 30 constants can be identified from \(s_{3,2}\)

6 Discussion and Further Work

The main contribution of this paper has been to prove that all the rate constants in several well-known chemical reaction networks that are abundant in the literature can be identified from a reduced set of kinetic variables. The work here extends previous results by Craciun and Pantea (2008) and avoids computationally expensive procedures such as differential elimination and Gröbner basis (Bellu et al. 2007; Boulier 2007; Meshkat et al. 2009).

We should point out that we assumed that there is a special partition of the set of chemical species and that every connected component of the chemical reaction network has a particular shape (see Sect. 3). Both assumptions are natural when modeling multisite phosphorylation systems and signaling cascades (Wang and Sontag 2008; Huang and Ferrell 1996). We have then shown, in Sect. 4, how to identify the rate constants in every connected component, or two related connected components, from a single species. In Sect. 5 we have moreover proved that all the rate constants in signaling cascade networks can be identified from only one species: the last product of the first component of the last layer. Additionally, we have presented in Sect. 5.2 an example showing how to compute the values of the rate constants from noise-free data according to our theoretical results in the previous sections. The procedure is based on a heuristic to choose the right input data; it would be of great interest to find a formal proof for establishing a good set of sufficient data for any network of the class considered in this paper.

We expect that the techniques used in this paper could be applied for identifiability from a few variables to a number of modifications of the networks we have considered here. For instance, it would be interesting to introduce more intermediate complexes within different reactions. Another potential adaptation is relaxing the assumption \(F_i\ne F_j\) for \(i\ne j\) in the cascade network, and allowing for repetition of these enzymes. Both modifications are natural extensions of the networks we have analyzed, and we conjecture that similar results can be obtained. We moreover would like to apply our techniques to more general but hence well structured networks such as MESSI networks (Pérez Millán and Dickenstein 2018). Another future research direction is to characterize which other variables can be considered to identify the rate constants of either a whole connected component or the entire biochemical network.