arXiv:2012.15437v3 [math.DS] 19 Feb 2021 hntemxmmnme fsal oiiesed ttsis states steady positive stable of number maximum the then h aiu ubro tbepstv taysae is states steady positive stable of number maximum the aiu ubris number maximum iiy tihoercsubspaces stoichiometric bility, https://sites.google.com/site/rootclassification/ the of sub-family [ a reversible) for and (possibly reactions species, two one most ad with network networks at a small if the equal for are [ numbers states betwe two steady sta these relation positive steady that the positive conjectured is and is problem states it a steady related that positive a Notice nondegenerate So, of is. states nondegenerate. steady be positive must and states steady positive 21 R (e.g., tools standard lenging. the because [ multistability [ criterion networks for reaction biochemical criterion of simpler list geom a algebraic to stead computational applied those cessfully on of based stability tw methods mu least the symbolic finding checking at Recently, first admits numerically is network then method given and typical a states), a that far, such of So parameters existence finding general. (i.e., the in positive problem Deciding hard stable a class. two is compatibility least stoichiometric at same correspon has the the kinetics) mass-action that under such (arising constants) (rate parameters positive h olwn eut.()I h aiu ubro positive of number maximum the If (1) results. following the oiiesed ttsi nodnumber odd an is states steady positive insses hc slne osic-iebhvo n decision-makin and behavior switch-like [ to signaling linked cellular is which systems, tion states? steady positive and states steady positive states? steady positive three least at admits open. widely are questions following the networks, tion , ∗ ‡ † M ujc classifications. subject AMS Abstract. e words. Key colo ahmtclSine,BiagUiest,Beij University, Beihang Sciences, Mathematical of School colo ahmtclSine,BiagUiest,Beij University, Beihang Sciences, Mathematical of School umte oteeiosDATE. editors the to Submitted Funding: 2 .Introduction. 1. ic hr a enals fnc rtrafrmlittoaiy(.. [ (e.g., multistationarity for criteria nice of list a been has there Since hs usin r oiae ytemlitblt rbe fbioche of problem multistability the by motivated are questions These Question Question IHOEDMNINLSOCIMTI SUBSPACES STOICHIOMETRIC ONE-DIMENSIONAL WITH , 13 , 11 3 , ,o lentvl ieadCiatciein[ Li´enard-Chipart criterion alternatively or ], 9 UTSAIIYO ECINNETWORKS REACTION OF MULTISTABILITY Twsfne yteNSFC12001029. the by funded was XT ) ti aua oakwa h eainbtentenmeso sta of numbers the between relation the what ask to natural is it ]), o h ecinntok ihoedmninlstoichiome one-dimensional with networks reaction the For 1.2. 1.1. iceia ecinntok,ms-cinkntc,mul kinetics, mass-action networks, reaction biochemical 10 N 2 +1 oegnrly hti h eainbtentenmeso s of numbers the between net relation the this is if what multistability generally, More admits network a that hold it Does , 1 IOINTANG XIAOXIAN otherwise. 12 , o h yaia ytm htaiefo iceia reac- biochemical from arise that systems dynamical the For 22 odgnrc ojcue,adti ojcuei proved is conjecture this and Conjecture], Nondegeneracy , , 7 .W a ewr dismlitblt fteeexist there if multistability admits network a say We ]. 24,92C45 92C40, N hnw rvd odto ntentoksc that such network the on condition a provide we then , † ). AND 1 12 HSU ZHANG ZHISHUO , N 18 2 − .I ssoni ueoscom- numerous in shown is It ]. 1 n,Cia( China ing, fti odto sstse,adthis and satisfied, is condition this if taysae sa vnnumber even an is states steady N 2 n,Cia( China ing, 2 ftemxmmnme of number maximum the If (2) . 8 )aecmuainlychal- computationally are ]) 20 .Hwvr esilneed still we However, ]. igdnmclsystem dynamical ding ‡ e.Frti problem, this For tes. [email protected] rcsbpcs eshow we subspaces, tric ittoaiy multista- tistationarity, [email protected] tts(.. [ (e.g., states y tbesed state steady stable isfiieymany finitely mits tyaeas suc- also are etry oiiesteady positive o taysae in states steady ntenumbers the en ewrswith networks ltistationarity outh-Hurwitz multistability rcs in process g ia reac- mical 6 , 17 work table ∗ 15 , ble N ]). 4 ). , , , 2 XIAOXIAN TANG, AND ZHISHUO ZHANG putational results that if a network has three nondegenerate positive steady states for a choice of parameters, then two of these steady states are stable [15, 20], which suggests the answer to Question 1.1 might be positive. But there is no answer to the more general problem Question 1.2. In this paper, for the reaction networks with one-dimensional stoichiometric sub- spaces, we answer both Question 1.1 Question 1.2 by the following three results. (1) There exists a network such that its maximum number of positive steady states is 3, and the maximum number of stable positive steady states is only 1 (see Example 3.8). So, the answer to Question 1.1 is negative. (2) If the maximum number of positive steady states is an even number N, then N the maximum number of stable positive steady states is 2 ( 3.4 (a)). (3) If the maximum number of positive steady states is an odd number N, then we provide a condition on the network such that the maximum number of N−1 stable positive steady states is 2 if this condition is satisfied, and this N+1 maximum number is 2 otherwise (Theorem 3.4 (b)). From the proof of Theorem 3.4, we also see that if a choice of parameters yields the maximum number of positive steady states, then it also yields the maximum number of stable positive steady states (Corollary 3.5). Besides the main results above, based on the lemmas for proving Theorem 3.4, we provide a self-contained proof for the nondegeneracy conjecture (Theorem 6.1). We also provide Corollary 3.6 for the networks with two reactions, which shows that multistability can be easily read off from the reaction coefficients. The idea of studying networks with one-dimensional stoichiometric subspaces is inspired by the fact that the small networks with two reactions (possibly reversible) studied in [12, 18, 19] have one-dimensional stoichiometric subspaces. Since it is easier to eliminate variables by using the conservation-law equations, these networks have a list of nice properties as follows. Their (nondegenerate) steady states are (simple) solutions to univariate (see Section 5.1 and Lemma 5.5). The stability of a steady state can be determined by checking the trace of a Jacobian (see Lemma 5.1). By these properties, one can first prove the nondegeneracy conjecture (Theorem 6.1) by perturbing a multiple solution of a univariate (see Lemma 6.3 and Lemma 6.4). It is worth mentioning that some similar but different ideas of perturbing parameters can be found in the study of (nondegenerate) multistationarity [18, 9]. Once the nondegeneracy conjecture is proved, it is not difficult to deduce the first part of the main result (Theorem 3.4 (a)) based on a “sign condition” (see Theorem 5.6). The challenging part of the main result (Theorem 3.4 (b)) is proved by computing the Brouwer degree of a univariate polynomial (5.9)ona bounded interval. It is remarkable that the Brouwer degree of steady-state equations is already known for the dissipative networks admitting no boundary steady states [3, Theorem 3]. However, for a large class of dissipative (or, conservative) networks (e.g., networks with exactly two irreversible reactions), if they admit multistability, then they definitely admit boundary steady states (see [19, Theorem 4.8]). So, the results in [3] do not apply here. The rest of this paper is organized as follows. In Section 2, we review the def- initions of multistationarity and multistability for the mass-action kinetics systems arising from reaction networks. In Section 3, we formally present the main result: Theorem 3.4. We also present two useful corollaries, and we illustrate how these re- sults work by two examples. In Section 4, we provide two useful lemmas on univariate polynomials. In Section 5, we prepare a list of nice properties for the networks with MULTISTABILITY OF ONE-DIMENSIONAL REACTION NETWORKS 3

one-dimensional stoichiometric subspaces. In Section 6, we prove the nondegeneracy conjecture for the networks with one-dimensional stoichiometric subspaces (Theorem 6.1). In Section 7, we prove Theorem 3.4 and its corollaries. Finally, we end this paper with some future directions inspired by Theorem 3.4, see Section 8. 2. Background. 2.1. Chemical reaction networks. In this section, we briefly recall the stan- dard notions and definitions on reaction networks, see [3, 12] for more details. A re- action network G (or network for short) consists of a set of s species {X1,X2,...,Xs} and a set of m reactions:

(2.1) α1j X1 + ··· + αsj Xs −→ β1j X1 + ··· + βsj Xs, for j =1, 2, . . . , m,

where all αij and βij are non-negative integers, and (α1j ,...,αsj ) 6= (β1j ,...,βsj ). We call the s × m matrix with (i, j)-entry equal to βij − αij the stoichiometric matrix of G, denoted by N . We call the image of N the stoichiometric subspace, denoted by S. We denote by x1, x2,...,xs the concentrations of the species X1,X2,...,Xs, re- spectively. Under the assumption of mass-action kinetics, we describe how these concentrations change in time by following system of ODEs:

α11 α21 αs1 κ1 x1 x2 ··· xs α12 α22 αs2 κ2 x x ··· x ⊤ 1 2 s (2.2) x˙ = f(κ; x) = (f1(κ; x),...,fs(κ; x)) := N·  .  , .   κ xα1m xα2m ··· xαsm   m 1 2 s    where x denotes the vector (x1, x2,...,xs), and each κj ∈ R>0 is called a rate constant corresponding to the j-th reaction in (2.1). By considering the rate constants as an unknown vector κ := (κ1,κ2,...,κm), we have polynomials fi(κ; x) ∈ Q[κ, x], for i =1, 2,...,s. Let d := s − rank(N ). A conservation-law matrix of G, denoted by W , is any row-reduced d × s-matrix whose rows form a basis of S⊥ (note here, rank(W ) = d). s Our system (2.2) satisfies W x˙ = 0, and both the positive orthant R>0 and its closure R≥0 are forward-invariant for the dynamics. Thus, any trajectory x(t) beginning 0 s at a non-negative vector x(0) = x ∈ R>0 remains, for all positive time, in the following stoichiometric compatibility class with respect to the total-constant vector c := W x0 ∈ Rd:

s Pc := {x ∈ R≥0 | W x = c}.

That means Pc is forward-invariant with respect to the dynamics (2.2). 2.2. Multistationarity and Multistability. A steady state of (2.2) is a con- ∗ s ∗ centration vector x ∈ R≥0 such that f(x ) = 0, where f(x) is on the right-hand side of the ODEs (2.2). If all coordinates of a steady state x∗ are strictly positive ∗ s ∗ ∗ (i.e., x ∈ R>0), then we call x a positive steady state. If a steady state x has zero ∗ s s ∗ coordinates (i.e., x ∈ R≥0\R>0), then we call x a boundary steady state. We say a ∗ ∗ ∗ steady state x is nondegenerate if im (Jacf (x )|S) = S, where Jacf (x ) denotes the Jacobian matrix of f, with respect to x, at x∗. A steady state x∗ is exponentially stable (or, simply stable in this paper) if it is nondegenerate, and all non-zero eigenval- ∗ ues of Jacf (x ) have negative real parts. Note that if a steady state is exponentially stable, then it is locally asymptotically stable [16]. 4 XIAOXIAN TANG, AND ZHISHUO ZHANG

Suppose N ∈ Z≥0. We say a network admits N (nondegenerate) positive steady states if for some rate-constant vector κ and for some total-constant vector c, it has N (nondegenerate) positive steady states in the stoichiometric compatibility class Pc. Similarly, we say a network admits N stable positive steady states if for some rate- constant vector κ and for some total-constant vector c, it has N stable positive steady states in Pc. The maximum number of positive steady states of a network G is

cappos(G) := max{N ∈ Z≥0 ∪ {+∞}|G admits N positive steady states}. The maximum number of nondegenerate positive steady states of a network G is

capnondeg(G) := max{N ∈ Z≥0 ∪ {+∞}|G admits N nondegenerate positive steady states}. The maximum number of stable positive steady states of a network G is

capstab(G) := max{N ∈ Z≥0 ∪ {+∞}|G admits N stable positive steady states}.

We say a network admits multistationarity if cappos(G) ≥ 2. We say a network admits multistability if capstab(G) ≥ 2. Conjecture 2.1 (Nondegeneracy Conjecture). [12, Conjecture 2.3] Given a net- work G, if cappos(G) < +∞, then capnondeg(G)= cappos(G). 3. Main Result. In this section, we first define a list of notions needed in the statement of the main result, and after that, we formally present the main result: Theorem 3.4, which reveals the relation between the maximum numbers of stable positive steady states and positive steady states. We also provide two useful corollar- ies. Corollary 3.5 shows that if we want to find the parameters such that a network (with a one-dimensional stoichiometric subspace) has the maximum number of stable positive steady states, then one method is to search the parameters such that the network has the maximum number of positive steady states. Corollary 3.6 shows that if a network contains two reactions, then we can determine if it admits multistability by looking at the reaction coefficients. Theorem 3.4 and the two corollaries will be proved in Section 7. We illustrate how these results work by two examples, where Example 3.8 answers Question 1.1. We remark that Theorem 3.4 generalizes [19, Theorem 3.16]. Assumption 3.1. For any network G with reactions defined in (2.1), by the def- inition of reaction network, we know (α11,...,αs1) 6= (β11,...,βs1). Without loss of generality, we will assume β11 − α11 6=0 throughout this paper. Suppose a network G (2.1) (with s species and m reactions) has a one-dimensional stoichiometric subspace. Note that the rank of its conservation-law matrix is s − 1. Given a total-constant vector c∗ ∈ Rs−1, under Assumption 3.1, we define

(3.1) A1 := 1, B1 := 0, ∗ βi1 − αi1 ci−1 (3.2) Ai := , Bi := − , for any i ∈{2,...,s}, β11 − α11 β11 − α11 and for every i ∈{1,...,s}, we define an equivalence class

Bk Bi {k ∈{1,...,s}|Ak 6=0, and A = A }, if Ai 6=0, (3.3) [i]c∗ := k i ({k ∈{1,...,s}|Ak =0}, if Ai =0.

∗ When c is clear from the context, we simply denote [i]c∗ by [i]. Denote by r the number of these equivalence classes, i.e.,

(3.4) r := |{ [i] | i ∈{1,...,s}}|≤ s. MULTISTABILITY OF ONE-DIMENSIONAL REACTION NETWORKS 5

Without loss of generality, we assume 1,...,r are the representatives of these equiva- lence classes. Recall m is the number of reactions in G (2.1). For every k ∈{1,...,r}, we define

(3.5) ϕk := min { αij }, where αij ’s are the reaction coefficients, see (2.1). 1≤j≤m iX∈[k] And for every j ∈{1,...,m}, we define

(3.6) γkj := αij − ϕk. iX∈[k] Define two sets of indices

(3.7) J := {i ∈{1,...,s}|Ai 6=0}, and

(3.8) H := {k ∈{1,...,r} ∩ J |γk1,...,γkm are not all the same}.

Assumption 3.2. Given a network G, suppose the stoichiometric subspace of G is one-dimensional. For a total-constant vector c∗, let H be the set of indices defined in (3.8). If there exists a rate-constant vector κ∗ such that G has N (0

(− Bi , +∞) if A > 0 Ai i (3.9) Ii := (0, +∞) if Ai =0 . Bi (0, − ) if Ai < 0  Ai Under Assumption 3.2, define an open interval

(3.10) I := Ik, where Ik is defined in (3.9). k \∈H

Remark 3.3. Under Assumption 3.2, we have 1 ∈ H. So, by (3.10), I ⊂ I1 = (0, +∞). Therefore, the left endpoint of I (3.10) is no less than 0. Let

Bτ (3.11) τ be the index in H such that Aτ > 0 and − is the left endpoint of I. Aτ

Note that by (3.5) and (3.6), there exists j ∈{1,...,m} such that γτj = 0. We define

(3.12) L := {j ∈{1,...,m}|γτj = 0}. Now, we are prepared to define two important sets of parameters, which are depending on the given network G: (3.13) ∗ ∗ ∗ W := {(κ ,c )|for the rate-constant vector κ ,G has cappos(G) positive steady states in Pc∗ }, and (3.14) α B A B ∗ ∗ ∗ αij ij k k τ γkj B := {(κ ,c )|X(β1j − α1j )κj Y |Ai| Y Bi Y ( − ) > 0} |A | |A | Aτ j∈L i∈J i∈{1,...,s}\J k∈H,k6=τ k k 6 XIAOXIAN TANG, AND ZHISHUO ZHANG

(here, we remark that in (3.14), the values of Ai, Bi, γkj , J , H, τ, and L are depending on c∗ (recall (3.1)–(3.12))). We explain the meaning of the above two sets as follows. (i) The set W collects all “witnesses” (κ∗,c∗) such that the number of positive steady states reaches its maximum. (ii) The set B collects all choices of parameters (κ∗,c∗) such that the “Brouwer degree” is positive (see more discussion in Section 8, and also see the meaning of Brouwer degree in Remark 7.5, or in [3, Theorem 1]). Theorem 3.4. Given a network G (2.1) with a one-dimensional stoichiometric subspace, suppose cappos(G) < +∞. cappos(G) (a) If cappos(G) is even, then capstab(G) = 2 . (b) If cappos(G) is odd, then

cappos(G)−1 2 if W∩B = ∅ capstab(G) = cappos(G)+1 , ( 2 if W ∩ B= 6 ∅ where W and B are defined in (3.13)–(3.14). Corollary 3.5. Given a network G (2.1) with a one-dimensional stoichiometric subspace, suppose cappos(G) < +∞. Then, there exist a rate-constant vector κ and a total-constant vector c such that the network G has cappos(G) positive steady states in the stoichiometric compatibility class Pc, where capstab(G) of these positive steady states are stable. Corollary 3.6. Given a network G (2.1), suppose G contains two reactions and cappos(G) < +∞. cappos(G) (a) If cappos(G) is even, then capstab(G) = 2 . (b) If cappos(G) is odd, then

cappos(G)−1 2 if W∩B2−react = ∅ capstab(G) = cappos(G)+1 , ( 2 if W∩B2−react 6= ∅ where W is defined in (3.13), and

∗ ∗ (3.15) B2−react := {(κ ,c )|(γτ2 − γτ1)(βτ1 − ατ1) > 0}

(here, recall τ (3.11) and γτj (j ∈ {1, 2}) (3.6) are depending on the total constant- vector c∗). Especially, if τ is the only index in [τ] (recall (3.3)), then the inequality stated in (3.15) can be replaced with

(3.16) (ατ2 − ατ1)(βτ1 − ατ1) > 0.

Example 3.7. This example illustrates how to compute capstab(G) using Theorem 3.4. Consider the following network G

κ1 κ2 κ3 X2 −→ X1, X1 −→ X2, 2X1 + X2 −→ 3X1.

By (2.1), for this network, we have s =2. Note that there is only one total constant c1 since s − 1=1. By [19, Theorem 2.6], we know cappos(G)=3. The key idea of computing capstab(G) is to check if we have W ∩ B 6= ∅ for the two sets W (3.13) and B (3.14). In fact, we have 1 3 (3.17) (κ∗ = ,κ∗ = 16,κ∗ = ,c∗ = −9) ∈W∩B 1 2 2 3 2 1 MULTISTABILITY OF ONE-DIMENSIONAL REACTION NETWORKS 7

(we list the details in the next paragraph). Therefore, by Theorem 3.4 (b), we have cappos(G)+1 capstab(G)= 2 =2. Below, we show that (3.17) is true. By (2.1), for this network, we have β11−α11 = ∗ 1 1 and β21 − α21 = −1. So, for the total constant c1 = −9 (∈ R ), by (3.1), we have A1 = 1 and B1 = 0, and by (3.2), we have A2 = −1 and B2 = 9. Hence, it is straightforward to verify that there are two equivalence classes defined in (3.3):

[1] = {1}, and [2] = {2}, and the set of indices defined in (3.7) is J = {1, 2}. By (3.5), for each equivalence class, we have

ϕ1 = min {α1j } =0, and ϕ2 = min {α2j } =0, 1≤j≤3 1≤j≤3 and so, by (3.6), we have γij = αij for every i ∈ {1, 2} and for every j ∈ {1, 2, 3}. Hence, the set of indices defined in (3.8) is H = {1, 2}, and the open interval defined in (3.10) is I = I1 ∩I2 = (0, 9), where I1 = (0, +∞) and I2 = (0, 9). Note that the left endpoint of I is 0. So, the index τ defined in (3.11) is τ = 1, and the set of indices defined in (3.12) is L = {1} (the set of indices j such that γ1j =0). So, for the rate ∗ 1 ∗ ∗ 3 constants κ1 = 2 ,κ2 = 16, and κ3 = 2 , the left-hand side of the inequality in (3.14) is

∗ αij αij Bk Ak Bτ γkj (βj1 − αj1)κj |Ai| Bi ( − ) |Ak| |Ak| Aτ j i k ,k τ X∈L Y∈J i∈{1,...,sY }\J ∈HY6=

2 γ21 B2 A2 B1 9 = (β − α )κ∗ |A |αi1 − = > 0. 11 11 1 i |A | |A | A 2 i=1 2 2 1 Y   That means, we have 1 3 (κ∗ = ,κ∗ = 16,κ∗ = ,c∗ = −9) ∈B 1 2 2 3 2 1 Note that by [19, Theorem 2.6], for these parameters, G has three (the maximum number) nondegenerate positive steady states. So, the above choice of parameters is also in the set of witnesses W. Hence, we indeed have (3.17). Example 3.8. Consider the network G

κ1 κ2 2X1 +2X2 + X3 −→ 3X1 + X2, X1 +2X3 −→ X2 +3X3.

We show by Corollary 3.6 that for this network, the answer to Question 1.1 is negative. First, we point out that cappos(G)=3. In fact, from the total degree of the steady-state system h shown later in Example 5.2, we see that the number of positive solutions of h = 0 is upper bounded by 3. And it is straightforward to check that for ∗ ∗ 8 the rate constants κ1 = 1, and κ2 = 3 , G has three nondegenerate positive steady states in the stoichiometric compatibility class P ∗ 11 . c =(−3,− 4 ) Next, we will prove that capstab(G)=1. By Corollary 3.6 (b), it is sufficient to show that W∩B2−react = ∅ (here, B2−react is defined in (3.15)). By (2.1), for this network, we have s =3, m =2, β11 − α11 =1, β21 − α21 = −1, and β31 − α31 = −1. Notice that the values of Ai’s defined in (3.1)–(3.2) have nothing to do with the choice ∗ of the total-constant vector c (Ai’s only depend on the values of αij and βij ). Indeed, for this network, by (3.1) and (3.2), we have A1 =1, A2 = −1, and A3 = −1 for any 8 XIAOXIAN TANG, AND ZHISHUO ZHANG

choice of total-constant vector c∗ ∈ R2. So, by (3.9) and (3.10), the left endpoint of the interval I defined in (3.10) is always 0. Hence, the index τ defined in (3.11) is τ =1. Note also, it is straightforward to check by the conservation laws (see Example ∗ ∗ ∗ ∗ 5.2) that if (κ ,c ) ∈ W, then c1 6= 0 and c2 6= 0. So, by (3.1)–(3.3), there is only one index τ in [τ]. Therefore, by Corollary 3.6, for any (κ∗,c∗) ∈ W, the left-hand side of the inequality (3.16) is always

(ατ2 − ατ1)(βτ1 − ατ1) = (α12 − α11)(β11 − α11)= −1 < 0.

That means W∩B2−react = ∅ for this example. 4. Univariate Polynomials. Definition 4.1. Suppose p(z) is a univariate polynomial in R[z], we say z∗ ∈ ∗ dp ∗ ∗ R is a simple solution of p(z)=0 if p(z )=0 and dz (z ) 6= 0. We say z ∈ ∗ dp ∗ R is a multiplicity-N (N ≥ 2) solution of p(z)=0 if p(z )=0 , dz (z )=0, N N d −1p ∗ d p ∗ ..., dzN−1 (z )=0 and dzN (z ) 6=0. n Lemma 4.2. Let p(z) := anz + ··· + a1z + a0 be a univariate polynomial in R[z]. If z1 and z2 are two distinct real solutions of the equation p(z)=0, and any real solution of p(z)=0 in the open interval (z1,z2) has an even multiplicity, then we dp dp have dz (z1) dz (z2) ≤ 0. Proof. We write p(z) = (z − z1)(z − z2)h(z), where h(z) ∈ R[z]. Note that dp dh (z) = (z − z )h(z) + (z − z )h(z) + (z − z )(z − z ) (z). dz 2 1 1 2 dz So, dp dp (4.1) (z ) (z ) = −(z − z )2h(z )h(z ). dz 1 dz 2 1 2 1 2

˜ mb Let V := {b ∈ (z1,z2)|p(b)=0}. Then, we can write h(z) = h(z) b∈V (z − b) , where m is an even number, and h˜(z) = 0 has no real solutions in (z ,z ) (which b Q 1 2 implies h˜(z1)h˜(z2) ≥ 0). So, by (4.1), we have dp dp (z ) (z ) = −(z − z )2h˜(z )h˜(z ) (z − b)mb (z − b)mb ≤ 0. dz 1 dz 2 1 2 1 2 1 2 b Y∈V We are done. Lemma 4.3. Let p(z) be a univariate polynomial in R[z]. Suppose the equation p(z)=0 has exactly N (N > 0) real solutions z(1) < ··· < z(N) in an open interval (a,M) (a ∈ R, and M ∈ R ∪{+∞}), where z(1) is simple, and any multiple solution dp (1) has an odd multiplicity. If p(a) 6=0, then sign(p(a)) = −sign( dz (z )). Proof. We write

N (4.2) p(z) = C(z) (z − z(i))mi , i=1 Y where C(z) is a non-zero polynomial in R[z], m1 = 1, and for any i ≥ 2, mi is odd. Then, we have

N N dp dC (z) = (z) (z − z(i))mi + C(z) m (z − z(i))mi−1 (z − z(k))mk . dz dz i i=1 i=1 k i Y X Y6= MULTISTABILITY OF ONE-DIMENSIONAL REACTION NETWORKS 9

Thus, we have N dp (4.3) (z(1)) = C(z(1)) (z(1) − z(k))mk . dz k Y=2 Note that C(a) 6= 0 since p(a) 6= 0. Also, C(z) does not change its sign over (a,M) since p(z) = 0 has exactly N solutions in (a,M). Hence, for any z ∈ (a,M), C(z) has the same sign with C(z(1)). So, sign(C(a)) = sign(C(z(1))) since C(z) is continuous. Therefore, by (4.2) and (4.3), the sign of p(a) is equal to

(1) dp (1) −sign(C(a)) = −sign(C(z )) = −sign( dz (z )), if N is odd, (1) dp (1) (sign(C(a)) = sign(C(z )) = −sign( dz (z )), if N is even. We complete the proof. 5. Networks with one-dimensional stoichiometric subspaces. In this sec- tion, we focus on the networks with one-dimensional stoichiometric subspaces. We present a list of nice results on these networks. In Section 5.1, we show the steady-state system (augmented with the conservation laws) can be reduced to a univariate poly- nomial (see (5.3)–(5.5)). In Section 5.2, we present a criterion for stability (Lemma 5.1). In Section 5.3, we show a nondegenerate steady state corresponds to a simple real root of a univariate polynomial (Lemma 5.5). In Section 5.4, we present a sign condition (Theorem 5.6), which implies an upper bound of the number of stable pos- itive steady states (Corollary 5.7). In Section 5.5, we prepare a list of lemmas for the networks admitting at least one positive steady state. We remark that in the first author’s previous work [19], Lemma 5.1, Lemma 5.3, Lemma 5.4, and Lemma 5.5 are proved for networks with two reactions (possibly re- versible). Those proofs can be directly to extend to any networks with one-dimensional stoichiometric subspaces. Also, Lemmas 5.3–5.5 are implied by the results [5, Propo- sitions 9.1–9.2] on the “reduced Jacobian matrix” [5, Definition 9.8]. So, we omit the proofs here. Note also Theorem 5.6 is a straightforward generalization of [19, Theorem 3.5]. 5.1. Steady States. If the stoichiometric subspace of a network G (2.1) is one- dimensional, then under Assumption 3.1, for every j ∈{1,...,m}, there exists λj ∈ R such that

β1j − α1j β11 − α11 . . (5.1)  .  = λj  .  . β − α β − α  sj sj   s1 s1      Note here, we have λ1 = 1, and by the definition of chemical reaction network (which requires (α1j ,...,αsj ) 6= (β1j ,...,βsj ) in (2.1)), we have λj 6= 0 for all j ∈{2,...,m}. We substitute (5.1) into f(x) in (2.2), and we have m s αkj (5.2) fi = (βi1 − αi1) λj κj xk , i =1, . . . , s. j=1 k X Y=1 We define the steady-state system augmented with the conservation laws: m s αkj (5.3) h1 := f1 = (β11 − α11) λj κj xk j=1 k X Y=1 (5.4) hi := (βi1 − αi1)x1 − (β11 − α11)xi − ci−1 (here, ci−1 ∈ R), 2 ≤ i ≤ s. 10 XIAOXIAN TANG, AND ZHISHUO ZHANG

We first solve xi from hi = 0 for i ∈ {2,...,s}, and then we substitute the symbolic solution into h1. We get a polynomial

g(κ; c; x1) := h1(x1,...,xs)| β21−α21 c1 βs1−αs1 cs−1 x2= x1− , ..., xs= x1− β11−α11 β11−α11 β11−α11 β11−α11 m s α βk1 − αk1 ck−1 (5.5) = (β − α ) λ κ x 1j ( x − )αkj . 11 11 j j 1 β − α 1 β − α j=1 k 11 11 11 11 X Y=2 ∗ ∗ ∗ Clearly, if for a rate-constant κ , x is a steady state in Pc∗ , then x is a common ∗ ∗ ∗ solution to the equations h1(x ) = ... = hs(x ) = 0, and the first coordinate of x , ∗ ∗ ∗ denoted by x1, is a solution to the univariate equation g(κ ; c ; x1) = 0. 5.2. Stability. Lemma 5.1. [19, Lemma 3.4] If the stoichiometric subspace of a network G (2.1) is one-dimensional, then a nondegenerate steady state x∗ is stable if and only if s ∂fi |x x∗ < 0, where fi’s are the polynomials defined in (5.2). i=1 ∂xi = Example 5.2. Consider the network in Example 3.8. It is straightforward to P 8 check that the equality (5.1) holds for λ1 = 1 and λ2 = −1. Let κ1 = 1, κ2 = 3 , 11 c1 = −3, and c2 = − 4 . Then we have the steady-state system h: s s 8 h = (β − α ) λ κ xαk1 + λ κ xαk2 = x x (x x2 − x ), 1 11 11 1 1 k 2 2 k 1 3 1 2 3 3 k k ! Y=1 Y=1 h2 = (β21 − α21)x1 − (β11 − α11)x2 − c1 = −x1 − x2 +3, and 11 h = (β − α )x − (β − α )x − c = −x − x + . 3 31 31 1 11 11 3 2 1 3 4

By solving the equations h1(x) = h2(x) = h3(x)=0, we find three nondegenerate positive steady states:

3 x(1) ≈ (2.577, 0.423, 0.173), x(2) = (2, 1, ), x(3) ≈ (1.423, 1.577, 1.327). 4 It is straightforward to check by Lemma 5.1 that only x(2) is stable. Here, we complete the computation by Maple2020 [14], see “Example.mw” in Table 1. 5.3. Nondegenerate Steady States. Lemma 5.3. [19, Lemma 3.7] Let f and h be the systems defined in (5.2)–(5.4). Denote by |Jach| the determinant of the Jacobian matrix of h with respect to x. Then, we have s s−1 ∂fi |Jach| = (α11 − β11) . ∂xi i=1 X Lemma 5.4. [19, Lemma 3.8] For the system h defined in (5.3)–(5.4) and the polynomial g defined in (5.5), if for a rate-constant vector κ∗ and a total-constant ∗ ∗ ∗ ∗ vector c , x is a solution to h1(x )= ... = hs(x )=0, then

s−1 ∂g ∗ ∗ ∗ ∗ (α11 − β11) (κ ; c ; x1) = |Jach(x )|. ∂x1 Lemma 5.5. [19, Lemma 3.11] Given a network G (2.1), suppose the stoichiomet- ric subspace of G is one-dimensional. Let h and g be defined as in (5.3)–(5.5). If for a rate-constant vector κ∗, there is a steady state x∗ in the stoichiometric compatibility class Pc∗ , then the following statements are equivalent. MULTISTABILITY OF ONE-DIMENSIONAL REACTION NETWORKS 11

(i) The steady state x∗ is nondegenerate. ∗ (ii) |Jach(x )|6=0. (iii) ∂g (κ∗; c∗; x∗) 6=0. ∂x1 1 5.4. Sign condition. Theorem 5.6. Given a network G, suppose the stoichiometric subspace of G is one-dimensional. Let h be the system defined in (5.3)–(5.4). If for a rate-constant vec- ∗ (1) (N) tor κ , G has N distinct positive steady states x ,...,x in Pc∗ , where these steady (1) (N) states are ordered according to their first coordinates (i.e., x1 <...

P s ∗ s−1 ∂fi ∗ |Jach(x )|= (α11 − β11) (x ). ∂xi i=1 X So, the conclusion follows from Theorem 5.6. (b) The conclusion directly follows from (a). 5.5. Networks Admitting Positive Steady States. In this section, we prove a list of lemmas for the networks admitting at least one positive steady state. First, we ∗ will factor g(κ; c; x1) (see (5.8)) for a given total-constant vector c , and we deduce a polynomial q(κ; x1) (see (5.9)). In Lemmas 5.9–5.10, we show that the first coordinate of a (nondegenerate) positive steady state is a (simple) real solution of the equation q(κ; x1) = 0. Following from Lemma 5.9, Lemma 5.11 explains why Assumption 3.2 is reasonable (also see Remark 5.12 and Example 5.13). From the equation q(κ; x1) = 0, we can symbolically solve a carefully-chosen (see (5.10)) parameter κℓ, and we deduce a function φ (see (5.13)). In Lemma 5.15, we show a list of properties of this function, which will be useful in the next two sections. Suppose a network G (2.1) (with s species and m reactions) has one-dimensional stoichiometric subspace. Recall Section 5.1. By the steady-state system h augmented with conservation laws defined in (5.3)-(5.4), we can deduce a polynomial g(κ; c; x1) ∗ in (5.5). Suppose for a total constant-vector c , the notions [i], Ai, Bi, r, ϕk, γkj , J , H, Ii, I, τ, and L are defined in (3.3)–(3.12). For every i ∈{1,...,s}, we define

1 (Aix + Bi), i ∈ J (i.e., Ai 6= 0) |Ai| 1 (5.6) Yi(x1) := (1, i 6∈ J (i.e., Ai = 0).

For every j ∈{1,...,m}, we define

αij αij (5.7) Cj := |Ai| Bi . i Y∈J i∈{1,...,sY }\J 12 XIAOXIAN TANG, AND ZHISHUO ZHANG

∗ ∗ When c = c , we can write g(κ; c ; x1) in (5.5) as

r ∗ ϕk (5.8) g(κ; c ; x1) = q(κ; x1) Yk(x1) , k Y=1 where m r αij αij γkj q(κ; x1) := (β11 − α11) λj κj |Ai| Bi Yk(x1) , j=1 i k X Y∈J i∈{1,...,sY }\J Y=1 m γkj = (β11 − α11) λj κj Cj Yk(x1) j=1 X k∈{1Y,...r}∩J m γkj (5.9) = (β11 − α11) λj κj Cj Yk(x1) . j=1 k X Y∈H In (5.9), the second equality follows from the definition of Cj (5.7) and the fact that when k 6∈ J , Yk(x1) = 1 (see (5.6)). The last equality holds because when k 6∈ H, γkj = 0. In fact, if k 6∈ H, then by (3.8), γk1,...,γkm are all the same. So, by (3.5) and (3.6), for any j ∈{1,...,m}, γkj = 0. Now, for the index τ defined in (3.11), we define

(5.10) ℓ := argmin {γτj}. j∈{1,...,m} and for every j ∈{1,...,m}, define

γkj (5.11) Pj (x1) := Cj Yk(x1) k Y∈H Then q(κ; x1) in (5.9) can be written as

m

(5.12) q(κ; x1) = (β11 − α11) λj κj Pj (x1). j=1 X

We solve κℓ from q(κ; x1) = 0, and we define the symbolic solution as a function

m λj κjPj (x1) j=1,j6=ℓ (5.13) φ(ˆκ; x1) := − , PλℓPℓ(x1)

whereκ ˆ := (κ1,...,κℓ−1,κℓ+1,...,κm). Here, we recall that λℓ 6= 0 by (5.1). For any z ∈ R, we say φ(ˆκ; x1) is well-defined at z if Pℓ(z) 6= 0. Lemma 5.8. Given a network G, suppose the stoichiometric subspace of G is one- ∗ dimensional. For a total-constant vector c , let Ii be the interval defined in (3.9), and ∗ ∗ let Yi(x1) be the function defined in (5.6). If x1 ∈ Ii, then Yi(x1) > 0. ∗ Proof. Clearly, if Ai = 0, then by (5.6), we have Yi(x1)=1 > 0. Suppose Ai 6= 0. By (5.6), we have Y (x )= 1 (A x + B ). If x∗ ∈ I , then by (3.9), i 1 |Ai| i 1 i 1 i

∗ Bi Aix1 + Bi > Ai(− )+ Bi = 0. Ai ∗ So, we have Yi(x1) > 0. MULTISTABILITY OF ONE-DIMENSIONAL REACTION NETWORKS 13

Lemma 5.9. Given a network G, suppose the stoichiometric subspace of G is one- ∗ s−1 dimensional. For a total-constant vector c ∈ R , let q(κ; x1) be the polynomial ∗ s defined in (5.9). Then, for a rate-constant vector κ ∈ R>0, G has a positive steady ∗ state in the stoichiometric compatibility class Pc∗ if and only if q(κ ; x1)=0 has a s solution in Ii, where Ii is the open interval defined in (3.9). i=1 ∗ ∗ ∗ Proof. “T⇒:” Suppose G has a positive steady state x := (x1,...,xs) in Pc∗ . By ∗ ∗ ∗ (5.3), (5.4) and (5.5), we have x1 is a solution of g(κ ; c ; x1)=0. By(3.1), (3.2), ∗ ∗ ∗ (5.3), and (5.4), for any i (1 ≤ i ≤ s), xi = Aix1 + Bi > 0. So x1 is contained in the s ∗ ∗ interval Ii. By(5.8) and Lemma 5.8, we know x1 is a solution of q(κ ; x1) = 0. i=1 s T ∗ ∗ ∗ “⇐:” Let x1 be a solution of q(κ ; x1) = 0 in Ii. Then, by (5.8), x1 is a i=1 ∗ ∗ ∗ ∗ solution of g(κ ; c ; x1) = 0. For every i (2 ≤ i ≤ s),T let xi = Aix1 + Bi. Then, by ∗ ∗ (3.9), we have xi > 0 for every i. By(5.3), (5.4) and (5.5), x is a positive solution ∗ ∗ of h1 = ··· = hs = 0 for κ = κ and for c = c , i.e., it is a positive steady state in the stoichiometric compatibility class Pc∗ Lemma 5.10. Given a network G, suppose the stoichiometric subspace of G is one-dimensional. Let g(κ; c; x1) be the polynomial defined in (5.5). For a total- ∗ s−1 constant vector c ∈ R , let q(κ; x1) be the polynomial defined in (5.9). If for a ∗ s ∗ rate-constant vector κ ∈ R>0, G has a positive steady state x in the stoichiometric ∂g ∗ ∗ ∗ ∂q ∗ ∗ compatibility class Pc∗ , then (κ ; c ; x ) has the same sign with (κ ; x ), and ∂x1 1 ∂x1 1 additionally, the steady state x∗ is stable if and only if ∂q (κ∗; x∗) < 0. ∂x1 1 r ∂g ∗ ∗ ∗ ∂q ∗ ∗ ∗ ϕ Proof. By (5.8), we have (κ ; c ; x )= (κ ; x ) Yk(x ) k . By the proof ∂x1 1 ∂x1 1 1 k=1 s ∗ Q of Lemma 5.9, x1 ∈ Ii, where Ii is defined in (3.9). By Lemma 5.8, we have i=1 ∗ ∂g ∗ ∗ ∗ ∂q ∗ ∗ Yk(x ) > 0 for any kT. Therefore, sign( (κ ; c ; x )) = sign( (κ ; x )). By 1 ∂x1 1 ∂x1 1 Lemma 5.1, Lemma 5.3 and Lemma 5.4, the steady state x∗ is stable if and only if ∂g (κ∗; c∗; x∗) < 0, and thus, we have the conclusion. ∂x1 1 Lemma 5.11. Given a network G, suppose the stoichiometric subspace of G is one-dimensional. If for a total-constant vector c∗ ∈ Rs−1, the set of indices H defined (3.8) is empty, then for any rate-constant vector κ ∈ Rs, G has either no positive steady states, or infinitely many positive steady states in the stoichiometric compati- bility class Pc∗ . Proof. Suppose for a rate-constant vector κ∗ ∈ Rs, x∗ is a positive steady state in the stoichiometric compatibility class Pc∗ . By the proof of Lemma 5.9, the first s ∗ ∗ ∗ coordinate of x , say x1, is a solution of the equation q(κ ; x1) = 0 in Ii, where Ii i=1 is the open interval defined in (3.9). If H = ∅, then by (5.9), we have T m ∗ ∗ ∗ q(κ ; x1) = (β11 − α11) λj κj Cj . j=1 X m ∗ By Assumption (3.1), β11 − α11 6= 0. So, we must have j=1 λj κj Cj = 0. Note also s s ∗ Ii is nonempty since it contains x1. Therefore, any pointP in Ii is a solution of i=1 i=1 ∗ qT(κ ; x1) = 0, and so, the conclusion follows from Lemma 5.9. T Remark 5.12. In this remark, we explain why Assumption 3.2 is reasonable. If for ∗ a rate-constant vector κ , G has N positive steady states in Pc∗ , where 0

then Lemma 5.11 guarantees that there exists one index in H, say k. If1 6∈ H, then we can relabel the species X1,...,Xs as X˜1,..., X˜s such that X˜1 = Xk, X˜k = X1 and for any i 6∈ {1, k}, Xi = X˜i. We illustrate this relabelling by Example 5.13. Example 5.13. Consider the following network G

κ1 κ2 X1 +2X2 + X3 +2X4 −→ 2X1 + X2 +3X4, X1 +2X3 + X4 −→ X2 +3X3.

The conservation-law equations are

(5.14) −x1 − x2 − c1 =0, −x1 − x3 − c2 =0, x1 − x4 − c3 =0.

∗ ∗ ∗ ∗ 3 We choose a total-constant vector c := (c1,c2,c3) = (−3, 1, 2) (∈ R ). It is straight- forward to check that the set of indices defined in (3.7) is J = {1, 2, 3, 4}, and there are 4 equivalence classes defined in (3.3):

[1] = {1}, [2] = {2}, [3] = {3}, [4] = {4}.

By (3.5), for each equivalence class, we have

ϕ1 = min {α1j} =1, ϕ2 = min {α2j } =0, 1≤j≤2 1≤j≤2 ϕ3 = min {α3j} =1, ϕ4 = min {α4j } =1, 1≤j≤2 1≤j≤2

and so, by (3.6), we have

γ11 =0, γ21 =2, γ31 =0, γ41 =1, γ12 =0, γ22 =0, γ32 =1, γ42 =0.

Hence, the set of indices defined in (3.8) is H = {2, 3, 4}, and 1 6∈ H. However, we can relabel X1 and X4 as what is suggested in Remark 5.12. Consider the network after relabelling:

κ1 κ2 X˜4 +2X˜2 + X˜3 +2X˜1 −→ 2X˜4 + X˜2 +3X˜1, X˜4 +2X˜3 + X˜1 −→ X˜2 +3X˜3.

For this network, the conservation-law equations are

(5.15) −x˜1 − x˜2 − c˜1 =0, −x˜1 − x˜3 − c˜2 =0, x˜1 − x˜4 − c˜3 =0.

Comparing (5.14) and (5.15), and noting that x1 =x ˜4, x4 =x ˜1, x2 =x ˜2, and x3 = ∗ x˜3, we have c˜1 = c1 +c3, c˜2 = c2 +c3, and c˜3 = −c3. So, for c = (−3, 1, 2), the total- constant vector becomes c˜∗ = (−1, 3, −2) after the relabelling. It is straightforward to check that the set of indices defined in (3.8) is H˜ = {1, 2, 3} for c˜∗. Now, we do have 1 ∈ H˜. Lemma 5.14. Given a network G, suppose the stoichiometric subspace of G is ∗ one-dimensional. For a total-constant vector c , let Cj be the number defined in (5.7). If for a rate-constant vector κ∗, G has at least one positive steady state x∗ = ∗ ∗ (x1,...,xs ) in the stoichiometric compatibility class Pc∗ , then for any j ∈{1,...,m}, the number Cj is positive. ∗ ∗ Proof. By (3.1), (3.2), (5.3), and (5.4), for any i (1 ≤ i ≤ s), xi = Aix1 + Bi. If ∗ i ∈ {1,...,s}\J , then we have Ai = 0 by (3.7), and so, we have Bi = Aix1 + Bi = ∗ xi > 0. MULTISTABILITY OF ONE-DIMENSIONAL REACTION NETWORKS 15

Lemma 5.15. Given a network G, suppose the stoichiometric subspace of G is one-dimensional. For a total-constant vector c∗, let I be the open interval defined in (3.10), and let φ(ˆκ; x1) be the function defined in (5.13). If for a rate-constant vector ∗ ∗ κ , G has a positive steady state x in the stoichiometric compatibility class Pc∗ , then we have the following results. (1) For any z ∈ I, φ(ˆκ; x1) is well-defined at z. (2) For any k ∈ H\{τ}, we have Y (− Bτ ) > 0, where H, τ and Y are defined k Aτ k in (3.8), (3.11) and (5.6), and recall by (3.11) that − Bτ is the left endpoint Aτ of I. (3) The function φ(ˆκ; x1) is well-defined at the left endpoint of I. γkℓ Proof. (1) By (5.11), we have Pℓ(x1)= Cℓ Yk(x1) . Recall that I = Ik k∈H k∈H (3.10), where Ik is defined in (3.9). So, by LemmaQ5.8 and Lemma 5.14, for any zT∈ I, we have Pℓ(z) > 0, and thus, φ(ˆκ; x1) is well-defined at z. (2) Since G has at least one positive steady state x∗ in the stoichiometric compat- s ibility class Pc∗ , by Lemma 5.9, the interval Ii is nonempty. Note that by (3.10), i=1 s T we have Ii ⊂ I. Thus, I is also nonempty. For any k ∈ H\{τ}, recall Yk(x1) = i=1 1 Bτ 1 Bτ (Akx T+ Bk) (5.6). We show below that Yk(− ) = (Bk − Ak ) > 0. In |Ak| 1 Aτ |Ak| Aτ fact, we have the following cases. Bk Bτ • If Ak > 0, then by (3.9), Ik = (− , +∞). Note that a = − is the left Ak Aτ Bτ Bk endpoint of I and I ⊂ Ik by (3.10). So, we have − > − , and hence, Aτ Ak

Bτ Bk Bk − Ak > Bk − Ak = 0. Aτ Ak

Bk Bτ Bk • If Ak < 0, then by (3.9), Ik = (0, − ). So, we have − < − , and hence, Ak Aτ Ak

Bτ Bk Bk − Ak > Bk − Ak = 0. Aτ Ak

(3) By (3.5), (3.6) and (5.10), we have γτℓ = 0. So, by (5.11), we have

γkℓ Pℓ(x1) = Cℓ Yk(x1) . k∈H\{Y τ}

So, by Lemma 5.14 and (2), we have Pℓ(a) > 0. Thus, φ(ˆκ; x1) is well-defined at a. 6. Nondegeneracy Conjecture. In this section, the goal is to prove Conjecture 2.1 (Nondegeneracy Conjecture) for the networks with one-dimensional stoichiometric subspaces, see Theorem 6.1. Theorem 6.1. Given a network G, if the stoichiometric subspace of G is one- dimensional, and if cappos(G) < +∞, then capnondeg(G)= cappos(G). Below, we first present a list of lemmas (with examples). The lemmas for proving Theorem 6.1 are also very useful in Section 7 (e.g., see the proof of Lemma 7.4). After the proof, we provide Corollary 6.6, which will play a key role in the next section when we prove the part (b) of Theorem 3.4 (the main result). Lemma 6.2. Given a network G, suppose the stoichiometric subspace of G is ∗ one-dimensional. For a total-constant vector c , let q(κ; x1) be the polynomial de- fined in (5.9), let ℓ be the index defined in (5.10) and let φ(ˆκ; x1) be the rational 16 XIAOXIAN TANG, AND ZHISHUO ZHANG

∗ ∗ function defined in (5.13). If for a rate-constant vector κ , x1 is a multiplicity-N ∗ ∗ (N ≥ 2) solution of q(κ ; x1)=0, and if φ(ˆκ; x1) is well-defined at x1, then we have N N ∗ ∗ ∗ ∂φ ∗ ∗ ∂ −1φ ∗ ∗ ∂ φ ∗ ∗ φ(ˆκ ; x ) = κ , (ˆκ ; x )=0, ..., N−1 (ˆκ ; x )=0 and N (ˆκ ; x ) 6= 0, where 1 ℓ ∂x1 1 1 ∂x 1 ∂x1 1 ∗ ∗ ∗ ∗ ∗ κˆ = (κ1,...,κℓ−1,κℓ+1,...,κm). ∗ ∗ ∗ ∗ ∗ ∗ Proof. Define µ(κℓ, x1) := q(κ1,...,κℓ−1,κℓ,κℓ+1,...,κm, x1). If for κ = κ , x1 ∗ is a multiplicity-N solution of q(κ ; x1) = 0, then ∗ ∗ (I) µ(κℓ , x1) = 0, N ∂µ ∗ ∗ ∂ −1µ ∗ ∗ (II) (κ , x ) = 0, . . . , N−1 (κ , x )= 0, and ∂x1 ℓ 1 ℓ 1 ∂x1 N ∂ µ ∗ ∗ (III) N (κℓ , x1) 6= 0. ∂x1 For any x1 ∈ R such that φ(ˆκ; x1) is well-defined, by (5.12) and (5.13), we have ∗ ∗ ∗ φ(ˆκ ; x1)= κℓ and

∗ (6.1) µ(φ(ˆκ ; x1), x1) = 0.

Take the derivative of equation (6.1), via the chain rule:

∂µ ∗ ∂φ ∗ ∂µ ∗ (6.2) (φ(ˆκ ; x1), x1) (ˆκ ; x1)+ (φ(ˆκ ; x1), x1) = 0 . ∂κℓ ∂x1 ∂x1

∗ ∗ ∗ ∗ Evaluating this equation at x1 = x1, and recalling that φ(ˆκ ; x1)= κl , we obtain:

∂µ ∗ ∗ ∂φ ∗ ∗ ∂µ ∗ ∗ (6.3) (κl , x1) (ˆκ ; x1)+ (κl , x1) = 0 . ∂κl ∂x1 ∂x1 Note that by (5.12),

∂µ ∗ ∗ ∂q ∗ ∗ ∗ (κl , x1)= (κ ; x1) = (β11 − α11)λℓPℓ(x1) 6=0. ∂κl ∂κl

∗ (By Assumption 3.1, β11 − α11 6= 0. Recall λℓ 6= 0 by (5.1). And Pℓ(x1) 6= 0 because ∗ ∂µ ∗ ∗ φ(ˆκ; x1) is well-defined at x .) Note also (κ , x ) = 0, by (II). Thus, the equality 1 ∂x1 l 1 (6.3) implies ∂φ (ˆκ∗; x∗) = 0. Next, take another derivative, applying the chain rule ∂x1 1 ∗ to equation (6.2), and then evaluate at x1 = x1:

2 2 2 2 2 ∂µ ∗ ∗ ∂ φ ∗ ∗ ∂ µ ∗ ∗ ∂φ ∗ ∗ ∂ µ ∗ ∗ ∂φ ∗ ∗ ∂ µ ∗ ∗ (κl ,x1 ) (ˆκ ; x1 ) + (κl ,x1 ) (ˆκ ; x1 ) + 2 (κl ,x1 ) (ˆκ ; x1 ) + (κl ,x1 ) = 0. ∂κ ∂x2 ∂κ2 ∂x ! ∂x ∂κ ∂x ∂x2 l 1 l 1 1 l 1 1

∂φ ∗ ∗ ∂µ ∗ ∗ ∂2φ ∗ ∗ ∂2µ ∗ ∗ Thus, we deduce from (ˆκ ; x1) = 0 that (κl , x1) 2 (ˆκ ; x1)+ 2 (κl , x1) = 0. ∂x1 ∂κl ∂x1 ∂x1 i i ∂µ ∗ ∗ ∂ φ ∗ ∗ ∂ µ ∗ ∗ Similarly, for any 3 ≤ i ≤ N, we can deduce (κl , x1) i (ˆκ ; x1)+ i (κl , x1) = 0. ∂κl ∂x1 ∂x1 By (II) and (III), we have the conclusion. Lemma 6.3. Given a network G, suppose the stoichiometric subspace of G is one- ∗ dimensional. For a total-constant vector c , let q(κ; x1) be the polynomial defined in (5.9), let ℓ be the index defined in (5.10) and let φ(ˆκ; x1) be the rational function ∗ ∗ defined in (5.13). If for a rate-constant vector κ , x1 is a multiplicity-2N (N ∈ Z>0) ∗ ∗ solution of q(κ ; x1)=0, and if φ(ˆκ; x1) is well-defined at x1, then for every ǫ > 0, there exists δ ∈ R such that for all δ′ ∈ (0, |δ|), ∗ ∗ ∗ ′ ∗ ∗ (1) for κ = (κ1,...,κℓ−1,κℓ + sign(δ)δ ,κℓ+1,...,κm), the equation q(κ; x1)=0 (1) (2) (1) ∗ has two simple real solutions x1 and x1 , for which 0 < x1 − x1 < ǫ and (2) ∗ −ǫ < x1 − x1 < 0, and MULTISTABILITY OF ONE-DIMENSIONAL REACTION NETWORKS 17

∗ ∗ ∗ ′ ∗ ∗ (2) for κ = (κ1,...,κℓ−1,κℓ − sign(δ)δ ,κℓ+1,...,κm), the equation q(κ; x1)=0 ∗ ∗ has no real solution in the interval (x1 − ǫ, x1 + ǫ). Proof. By Lemma 6.2, we have 2N−1 2N ∗ ∗ ∗ ∂φ ∗ ∗ ∂ φ ∗ ∗ ∂ φ ∗ ∗ φ(ˆκ ; x1)= κl , (ˆκ ; x1)=0,..., 2N−1 (ˆκ ; x1)=0, and 2N (ˆκ ; x1) 6=0. ∂x1 ∂x1 ∂x1 ∗ ∗ ∗ ∗ ∗ Here, we recall thatκ ˆ = (κ1,...,κℓ−1,κℓ+1,...,κm). For any ǫ> 0, without loss of N ∂2 φ ∗ ∗ ∗ generality, we assume 2N (ˆκ ; x1) 6= 0 for any x1 ∈ (x1 − ǫ, x1 + ǫ). So, the Taylor ∂x1 ∗ ∗ ∗ expansion of φ(ˆκ ; x1) over (x1 − ǫ, x1 + ǫ) is 2N−1 k 2N ∗ ∗ ∗ 1 ∂ φ ∗ ∗ ∗ k 1 ∂ φ ∗ ∗ 2N φ(ˆκ ; x1) = φ(ˆκ ; x1)+ X (ˆκ ; x1)(x1 − x1) + (ˆκ ; ξ)(x1 − x1) k! ∂xk (2N)! ∂x2N k=1 1 1 1 ∂2N φ κ∗ κ∗ ξ x x∗ 2N , ξ x∗ ǫ,x∗ ǫ . (6.4) = l + 2N (ˆ ; )( 1 − 1) where ∈ ( 1 − 1 + ) (2N)! ∂x1 1 N N ∂2 φ ∗ ∂2 φ ∗ 2N Choose δ ∈ R such that sign(δ) = sign( 2N (ˆκ ; ξ)) and δ(2N)! / 2N (ˆκ ; ξ) <ǫ. ∂x1 ∂x1 For any δ′ ∈ (0, |δ|), let   δ˜′ := sign(δ)δ′. ∗ ∗ ˜′ By (6.4), we solve the equation φ(ˆκ ; x1) = κl + δ , and we get two distinct real solutions: 1 1 ∂2N φ 2N d2N φ 2N x(1) = x∗ + δ˜′(2N)! / (ˆκ∗; ξ) , and x(2) = x∗ − δ˜′(2N)! / (ˆκ∗; ξ) . 1 1 ∂x2N 1 1 dx2N  1   1  (1) (2) ∗ By (5.12) and (5.13), x1 and x1 are real solutions of the equation q(˜κ ; x1) = 0, whereκ ˜∗ := (κ∗,...,κ∗ ,κ∗ + δ˜′,κ∗ ,...,κ∗ ). By (6.4), we have ∂φ (ˆκ∗; x(i)) 6=0 1 l−1 l l+1 m ∂x1 1 (1) (2) ∗ for i ∈{1, 2}. So, by Lemma 6.2, both x1 and x1 are simple solutions of q(˜κ ; x1)= ∗ ∗ ˜′ 0. Clearly, by (6.3), the equation φ(ˆκ ; x1)= κl − δ has no real solutions for x1, and ∗ ∗ ∗ ˜′ ∗ ∗ so, for κ = (κ1,...,κl−1,κl − δ ,κl+1,...,κm), the equation q(κ; x1)=0 has no real solutions. Lemma 6.4. Given a network G, suppose the stoichiometric subspace of G is one- ∗ dimensional. For a total-constant vector c , let q(κ; x1) be the polynomial defined in (5.9), let ℓ be the index defined in (5.10) and let φ(ˆκ; x1) be the rational function ∗ ∗ defined in (5.13). If for a rate-constant vector κ , x1 is a simple solution, or a ∗ multiplicity-(2N + 1) (N ∈ Z>0) solution of q(κ ; x1)=0, and if φ(ˆκ; x1) is well- ∗ ′ defined at x1, then for every ǫ> 0, there exists δ ∈ R such that for all δ ∈ (0, |δ|), ∗ ∗ ∗ ′ ∗ ∗ (1) for κ = (κ1,...,κℓ−1,κℓ + sign(δ)δ ,κℓ+1,...,κm), the equation q(κ; x1)=0 (1) (1) ∗ has a simple real solution x1 , for which 0 < x1 − x1 <ǫ. ∗ ∗ ∗ ′ ∗ ∗ (2) for κ = (κ1,...,κℓ−1,κℓ − sign(δ)δ ,κℓ+1,...,κm), the equation q(κ; x1)=0 (2) (2) ∗ has a simple real solution x1 , for which −ǫ < x1 − x1 < 0. Proof. Similar to the proof of Lemma 6.3, for any ǫ> 0, without loss of generality, N ∂2 +1φ ∗ ∗ ∗ we assume 2N+1 (ˆκ ; x1) 6= 0 for any x1 ∈ (x1 −ǫ, x1 +ǫ). By Lemma 6.2, the Taylor ∂x1 ∗ ∗ ∗ expansion of φ(ˆκ ; x1) over (x1 − ǫ, x1 + ǫ) is 2N k 2N+1 ∗ ∗ ∗ 1 ∂ φ ∗ ∗ ∗ k 1 ∂ φ ∗ ∗ 2N+1 φ(ˆκ ; x1) = φ(ˆκ ; x1)+ X (ˆκ ; x1)(x1 − x1) + (ˆκ ; ξ)(x1 − x1) k! ∂xk (2N + 1)! 2N+1 k=1 1 ∂x1 1 ∂2N+1φ κ∗ κ∗ ξ x x∗ 2N+1, ξ x∗ ǫ,x∗ ǫ . (6.5) = l + 2N+1 (ˆ ; )( 1 − 1) where ∈ ( 1 − 1 + ) (2N + 1)! ∂x1 18 XIAOXIAN TANG, AND ZHISHUO ZHANG

N ∂2 +1φ ∗ Choose δ ∈ R such that sign(δ) = sign( 2N+1 (ˆκ ; ξ)) and ∂x1

1 2N+1 2N+1 ∂ φ ∗ δ(2N + 1)! / 2N+1 (ˆκ ; ξ) < ǫ.  ∂x1  For any δ′ ∈ (0, |δ|), let δ˜′ := sign(δ)δ′.

∗ ∗ ∗ ∗ ˜′ ∗ ∗ ∗ Then, forκ ˜ = (κ1,...,κl−1,κl + δ ,κl+1,...,κm), the equation q(˜κ ; x1) = 0 has 1 N (1) ∗ ˜′ ∂2 +1φ ∗ 2N+1 a real solution x1 = x1 + δ (2N + 1)! / 2N+1 (ˆκ ; ξ) . By(6.5), we have ∂x1 ∂φ (ˆκ∗; x(1)) 6= 0. So, by Lemma 6.2, we know ∂q (ˆκ∗; x(1) ) 6= 0, and hence, x(1) is a ∂x1 1 ∂x1 1 1 ∗ ∗ ∗ ˜′ ∗ ∗ simple solution. Similarly, for κ = (κ1,...,κl−1,κl − δ ,κl+1,...,κm), the equation q(κ; x1) = 0 has a simple real solution

1 ∂2N+1φ 2N+1 x(2) = x∗ − δ˜′(2N + 1)! / (ˆκ∗; ξ) . 1 1 ∂x2N+1  1 

Example 6.5. We illustrate by Figure 1 how Lemma 6.3 and Lemma 6.4 work. ∗ Consider the network G in Example 3.7. Choose c1 = −9. Then, we can compute ∗ ∗ that the polynomial defined in (5.9) for c = (c1) is 2 q(κ; x1) = κ1(−x1 + 9) − κ2x1 + κ3x1(−x1 + 9). By the computation in Example 3.7, we know the index τ defined in (3.11) is τ =1. Also, we know γ11 = 0, γ12 =1, and γ13 =2. So, the index ℓ defined in (5.10) is 1. So, the function defined in (5.13) is 2 κ2x1 − κ3x1(−x1 + 9) φ(ˆκ; x1) = . −x1 +9 ∗ ∗ 3 We substitute κˆ = (κ2,κ3) = (16, 2 ) into q(κ; x1), and we get a bivariate polynomial 3 µ(κ , x ) := κ (−x + 9) − 16x + x2(−x + 9). 1 1 1 1 1 2 1 1 ∗ ∗ ∗ ∗ 1 3 ∗ (a) For κ := (κ1,κ2,κ3) = ( 2 , 16, 2 ), the equation q(κ ; x1)=0 has a multiplic- ∗ ∗ ity-2 solution, which is x1 ≈ 0.6961103652. Also, x1 is a multiplicity-2 so- ∗ ∗ lution of µ(κ1, x1)=0. Clearly, φ(ˆκ; x1) is well-defined at x1. We plot the ∗ ∗ point (κ1, x1) and µ(κ1, x1)=0 for 0 ≤ κ1 ≤ 0.8 and 0 ≤ x1 ≤ 1.5 in Figure 1 ′ 1 (a). Choose δ = − 2 . It is seen from Figure 1 (a) that for any δ ∈ (0, |δ|), ∗ ′ the equation µ(κ1 − δ , x1)=0 has two distinct simple real solutions (e.g., see the two solid boxes for κ1 =0.3 in Figure 1 (a)), which are sufficiently close ∗ ∗ ′ to x1, and the equation µ(κ1 + δ , x1)=0 has no real solution. ∗ ∗ ∗ ∗ 3 3 ∗ (b) For κ := (κ1,κ2,κ3) = ( 10 , 16, 2 ), the equation q(κ ; x1)=0 has a simple ∗ ∗ solution, which is x1 ≈ 1.201246094. Clearly, φ(ˆκ; x1) is well-defined at x1. ∗ ∗ We plot the point (κ1, x1) and µ(κ1, x1)=0 for 0 ≤ κ1 ≤ 0.8 and 0.8 ≤ x1 ≤ 1 1.5 in Figure 1 (b). Choose δ = 4 . It is seen from Figure 1 (b) that for any ′ ∗ ′ δ ∈ (0,δ), the equation µ(κ1 + δ , x1)=0 has a simple real solution (e.g., see the solid box at the intersection of the blue line κ1 = 0.5 and the red curve ∗ in Figure 1 (b)), which is sufficiently close to x1, and similarly, the equation ∗ ′ µ(κ1 − δ , x1)=0 has a simple real solution too. MULTISTABILITY OF ONE-DIMENSIONAL REACTION NETWORKS 19

(a) (b)

Fig. 1. The two figures (a) and (b) illustrate Lemma 6.3 and Lemma 6.4, respectively. See the corresponding examples in Example 6.5. These figures are generated by Maple2020, see “Plot.mw” in Table 1.

Proof of Theorem 6.1. Proof. We only need to show cappos(G) ≤ capnondeg(G). Suppose cappos(G)= N, i.e., there exist a rate-constant vector κ∗ and a total-constant vector c∗ such that G (1) (N) has N distinct positive steady states x ,...,x in Pc∗ . For the total-constant ∗ vector c , let q(κ; x1) be the polynomial defined in (5.9), let ℓ be the index defined in (5.10) and let φ(ˆκ; x1) be the rational function defined in (5.13). By the proof (1) (N) of Lemma 5.9, the first coordinates of those positive steady states x1 ,...,x1 are s ∗ solutions of q(κ ; x1) = 0 in the interval Ii, where Ii is defined in (3.9). Note that i=1 s T Ii ⊂ I, where I is defined in (3.10). So, by Lemma 5.15 (1), the function φ(ˆκ; x1) i=1 T (i) is well-defined at x1 for any i ∈ {1,...,N}. Without loss of generality, assume the (1) (v) multiplicity of any solution in {x1 ,...,x1 } (v ≤ N) is even, and any solution in s (v+1) (N) {x1 ,...,x1 } is simple or has an odd multiplicity. Denote the interval Ii by i=1 (b, E). Here, we only prove the case when E 6= +∞. If E = +∞, the argumentT is similar. Let (i) (i) min {x1 }− b (i) (j) E − max {x1 } 1≤i≤N |x − x | 1≤i≤N ǫ = min{ , min{ 1 1 |1 ≤ i < j ≤ N}, }. 2 2 2

By Lemma 6.3, for ǫ> 0, and for every k ∈{1,...,v}, there exists δk ∈ R such that ′ for any δ ∈ (0, |δk|), q(κ; x1) = 0 has two distinct simple solutions in the interval (k) (k) (x1 − ǫ, x1 + ǫ) for ∗ ∗ ∗ ′ ∗ ∗ κ = κ1,...,κℓ−1,κℓ + sign(δk)δ ,κℓ+1,...,κm , (k) (k) and q(κ; x1) = 0 has no real solutions in the interval (x1 − ǫ, x1 + ǫ) for ∗ ∗ ∗ ′ ∗ ∗ κ = κ1,...,κℓ−1,κℓ − sign(δk)δ ,κℓ+1,...,κm .  20 XIAOXIAN TANG, AND ZHISHUO ZHANG

v v Note that there exist ⌈ ⌉ indices k1,...,k⌈ ⌉ in {1,...,v} such that δk1 ,...,δk⌈ v ⌉ 2 2 2

have the same sign. Without loss of generality, assume δk1 > 0. By Lemma 6.4, for ǫ > 0, and for every j ∈ {v +1,...,N}, there exists δj > 0 such that for any ′ (j) (j) δ ∈ (0,δj), q(κ; x1) = 0 has a simple solution in the interval (x1 − ǫ, x1 + ǫ) for ∗ ∗ ∗ ′ ∗ ∗ κ = κ1,...,κℓ−1,κℓ ± δ ,κℓ+1,...,κm . Let δ = min{|δ1|,..., |δv|,δv+1,...,δN }. ′ ∗ ∗ ∗ ′ ∗ ∗ Therefore, for any δ ∈ (0,δ), for κ = κ1,...,κℓ−1,κℓ + δ ,κℓ+1,...,κm , q(κ; x1)= v  0 has at least 2⌈ 2 ⌉ + N − v distinct simple solutions in (b, E). By Lemma 5.9, the v  network G admits at least 2⌈ 2 ⌉ + N − v positive steady states. By Lemma 5.10 v and Lemma 5.5, these 2⌈ 2 ⌉ + N − v steady states are nondegenerate. So, we have v cappos(G)= N ≤ 2⌈ 2 ⌉ + N − v ≤ capnondeg(G). Corollary 6.6. Given a network G (2.1) with a one-dimensional stoichiometric subspace, suppose cappos(G) < +∞. Assume W and B are the two sets of parameters defined in (3.13) and (3.14). Define a subset Wnondeg of W:

∗ ∗ ∗ ∗ (6.6) Wnondeg := {(κ ,c ) ∈ W|for (κ ,c ), all positive steady states of G are nondegenerate}.

Then, W ∩ B= 6 ∅ if and only if Wnondeg ∩ B 6= ∅. Proof. Since Wnondeg ⊂ W, we only need to prove that if W ∩ B= 6 ∅, then ∗ ∗ Wnondeg ∩ B 6= ∅. Suppose (κ ,c ) ∈W∩B. By the definition of W (see (3.13)), ∗ ∗ ∗ for κ , G has cappos(G) positive steady states in Pc∗ . If(κ ,c ) 6∈ Wnondeg, then by the proof of Theorem 6.1, there exist an index ℓ ∈ {1,...,s} and a positive number ′ ∗ ∗ ∗ ′ ∗ ∗ δ such that for any δ ∈ (0,δ), for κ = κ1,...,κℓ−1,κℓ + δ ,κℓ+1,...,κm or κ = κ∗,...,κ∗ ,κ∗ − δ′,κ∗ ,...,κ∗ , G has cap (G) nondegenerate positive steady 1 ℓ−1 ℓ ℓ+1 m pos  states in Pc∗ . Notice that the left-hand side of the inequality in (3.14) is continuous  with respect to κ∗. Thus, there exists δ′′ ∈ (0,δ) such that (κ∗∗,c∗) ∈ B, where ∗∗ ∗ ∗ ∗ ′′ ∗ ∗ κ := κ1,...,κℓ−1,κℓ ± δ ,κℓ+1,...,κm . Therefore, we have Wnondeg ∩ B 6= ∅.

7. Multistability. The goal of this section is to prove Theorem 3.4 (the main result), Corollary 3.5, and Corollary 3.6. From the proof presented later, we will see that the part (a) of Theorem 3.4 follows from the results in the previous two sections. So, the main task is to prove the part (b). The most crucial step is to show q(κ∗; a) 6= 0 for some (κ∗,c∗) (see Lemma 7.4), where the polynomial q is defined in (5.9), and the number a is the left endpoint of the interval I defined in (3.10). After this step, we can determine the number of stable positive steady states by the sign of q(κ∗; a) (see Lemma 7.7 and Corollary 7.8). From the proof, we will see the left-hand side of the inequality in (3.14) is an explicit expression of q(κ∗; a) in terms of (κ∗,c∗). We first prepare a list of lemmas. Lemma 7.1. Given a network G, suppose the stoichiometric subspace of G is one- dimensional. For a total-constant vector c∗, let I be the open interval defined in (3.10), ∗ and let q(κ; x1) be the polynomial defined in (5.9). If for a rate-constant vector κ , ∗ q(κ ; x1)=0 has N different solutions in I, then there exists another total-constant ∗ ∗ ∗ vector c˜ such that the equation q˜(κ ; x1)=0 corresponding to c˜ has at least N s ˜ ˜ different solutions in the interval ∩j=1Ij , where Ij is the interval defined in (3.9) for the total-constant vector c˜∗. ∗ (1) (N) Proof. Assume that q(κ ; x1)=0 has N different solutions x1 ,...,x1 in I for ∗ some rate-constant vector κ . For any j ∈{1,...,s}, let Aj and Bj be the numbers defined in (3.1) and (3.2) for c∗. Let H be the set of indices defined in (3.8) for c∗. Since H is finite, we can choose a positive number ρ such that for any k ∈ H, ρ> Bk . Ak For any j ∈{2,...,s}, suppose j ∈ [k]c∗ , where k ∈{1,...,r} (recall the definition of [k]c∗ in (3.3), and recall that r denotes the number of equivalence classes). We define MULTISTABILITY OF ONE-DIMENSIONAL REACTION NETWORKS 21 the (j − 1)-coordinate of a new total-constant vectorc ˜∗:

∗ cj−1, if k ∈ H, or Aj =0 ∗ −ρ(βj − αj ), if k 6∈ H, and Aj > 0 (7.1) c˜j−1 :=  1 1  (i) (βj1 − αj1)( max {x1 } + ρ), if k 6∈ H, and Aj < 0.  1≤i≤(N+1)  Now, forc ˜∗, letH˜ be the set of indices defined in (3.8). By (7.1), it is easy to verify that (I) H˜ = H, and ∗ ∗ (II)q ˜(κ ; x1)= q(κ ; x1). (1) (N) For any k ∈ {1,...,s}, So, by (II), we know x1 ,...,x1 are also solutions to ∗ q˜(κ ; x1) = 0. s (i) ˜ Below, we show that for every i ∈{1,...,N}, we have x1 ∈ Ij . In fact, note j=1 (i) T that {1,...,s} = ( [k]c˜∗ )∪( [k]c˜∗ ). We know x1 ∈ I = Ik. For every k ∈ H, k∈H˜ k6∈H˜ k∈H S ˜ S ˜ (i) T˜ by (7.1), we have Ik = Ik. So, by (I), for every k ∈ H, x1 ∈ Ik, and hence, for any (i) ˜ ˜ ˜ ˜ j ∈ [k]c˜∗ , x1 ∈ Ij since by (3.3) and (3.9), Ik = Ij . On the other hand, for any k 6∈ H, ∗ and for every j ∈ [k]c˜∗ , let A˜j and B˜j be the numbers defined in (3.1) and (3.2) forc ˜ . ˜ By (3.1), we see that A = A˜ . By(7.1), if A > 0, then we have x(i) > − Bj = −ρ j j j 1 Aj (i) (i) B˜j (i) (since x1 > 0), and if Aj < 0, then we have x1 < − = max {x1 } + ρ. Notice Aj 1≤i≤N (i) that the left endpoint of I is no less than 0 (recall Remark 3.3). So, we have x1 > 0. (i) ˜ ∗ So, by (3.9), we always have x1 ∈ Ij . Therefore, we know thatq ˜(κ ; x1)=0 has at s least N different solutions in the interval I˜j . j=1 Lemma 7.2. Given a network G, supposeT the stoichiometric subspace of G is one- dimensional. For a total-constant vector c∗, let I be the open interval defined in (3.10), and let q(κ; x1) be the polynomial defined in (5.9). If cappos(G)= N (0 ≤ N < +∞), ∗ ∗ then for any rate-constant vector κ , q(κ ; x1)=0 has at most N different solutions in I. ∗ Proof. If q(κ ; x1) = 0 has at least N + 1 different solutions in I, then by Lemma ∗ ∗ 7.1, there exists a total-constant vectorc ˜ such that the equationq ˜(κ ; x1) = 0 cor- ∗ s ˜ responding toc ˜ has at least N + 1 different solutions in the interval ∩j=1Ij , where ∗ I˜j is the interval defined in (3.9) for the total-constant vectorc ˜ . Thus, by the proof of Lemma 5.9, for κ = κ∗, the network G has at least N + 1 positive steady states in Pc˜∗ , which is a contradiction to the hypothesis that cappos(G)= N. The following definition for a degenerate steady state with a multiplicity is moti- vated by Lemma 5.5, Lemma 5.9 and Lemma 5.10. Definition 7.3. Given a network G (2.1), suppose the stoichiometric subspace ∗ of G is one-dimensional. For a total-constant vector c , let q(κ; x1) be the polynomial defined in (5.9). Suppose for a rate-constant vector κ∗, the network G has a positive ∗ ∗ ∗ steady state x in Pc∗ . If x1 is a multiplicity-N (N ≥ 2) solution of q(κ ; x1)=0, then we say x∗ is degenerate with multiplicity-N. Lemma 7.4. Given a network G, suppose the stoichiometric subspace of G is one- dimensional. For a total-constant vector c∗, let I := (a,M) be the open interval defined in (3.10), and let q(κ; x1) be the polynomial defined in (5.9). If cappos(G)= N (0 ≤ N < +∞), and if for a rate-constant vector κ∗, G has N positive steady states 22 XIAOXIAN TANG, AND ZHISHUO ZHANG

in the stoichiometric compatibility class Pc∗ , and any degenerate positive steady state has an odd multiplicity, then we have q(κ∗; a) 6=0. Proof. Suppose for the total-constant vector c∗ and for the rate-constant vector ∗ (1) (N) (1) (N) κ , the N positive steady states are x ,...,x such that x1 <...

1 (1) (i) (i+1) (N) ǫ = min { |a − x1 |, min {|x1 − x1 |}, |x1 − M|} . 2 1≤i≤N−1 ∗ For the total-constant vector c , let ℓ be the index defined in (5.10), and let φ(ˆκ; x1) be the function defined in (5.13). By Lemma 5.15 (1) and (3), φ(ˆκ; x1) is well-defined (i) (i) at x1 for any i ∈ {1,...,N}, and at a. By Lemma 5.10 and Lemma 5.5, x1 is a ∗ simple, or a multiplicity-(2K + 1) solution of q(κ ; x1) = 0 (here, K ∈ Z>0). So, by Lemma 6.4, for every i ∈ {1,...,N}, for ǫ > 0, there exists δi > 0 such that for all ′ ∗ ∗ ∗ ′ ∗ ∗ δ ∈ (0,δi), for κ = (κ1,...,κℓ−1,κℓ ± δ ,κℓ+1,...,κm), the equation q(κ; x1)=0 has (i) (i) (i) a simple real solution y1 , for which |y1 − x1 |<ǫ. ∗ Either if a is a multiplicity-2K (K ∈ Z>0) solution of q(κ ; x1) = 0, or if a is a ∗ simple, or a multiplicity-(2K +1) (K ∈ Z>0) solution of q(κ ; x1) = 0, then by Lemma ′ 6.3 (1) or Lemma 6.4 (1), for ǫ> 0, there exists δ0 ∈ R such that for all δ ∈ (0, |δ0|), ∗ ∗ ∗ ′ ∗ ∗ for κ = (κ1,...,κℓ−1,κℓ + sign(δ0)δ ,κℓ+1,...,κm), the equation q(κ; x1)=0 hasa simple solutiona ˜ such that 0 < a˜ − a<ǫ (i.e., a< a 0. ∂x1 ∂x1 MULTISTABILITY OF ONE-DIMENSIONAL REACTION NETWORKS 23

Thus, by Lemma 4.2, for any i = 1,...,N, if x(2i) is degenerate, then it has an odd multiplicity. Then, the conclusion follows from (a). Lemma 7.7. Given a network G (2.1) with a one-dimensional stoichiometric sub- ∗ space, suppose cappos(G)=2N +1 (0 ≤ N < +∞). If for a rate-constant vector κ and a total-constant vector c∗, G has 2N +1 positive steady states in the stoichio- metric compatibility class Pc∗ , and if N +1 of these positive steady states are stable, then

∗ αij αij Bk Ak Bτ γkj (7.2) (β1j − α1j )κj |Ai| Bi ( − ) > 0. |Ak| |Ak| Aτ j i k ,k τ X∈L Y∈J i∈{1,...,sY }\J ∈HY6= ∗ where for the total constant-vector c , the notions Ai, Bi, γkj , J , H, τ, and L are defined in (3.1)–(3.12). (1) (2N+1) (1) Proof. Suppose the positive steady states are x ,...,x such that x1 < (2N+1) (1) ...

∂q ∗ (1) ∗ sign( (κ ; x1 )) = −sign(q(κ ; a)). ∂x1 By Lemma 4.2, for every i ∈{2,..., 2N +1},

∂q ∗ (i) ∂q ∗ (i−1) (7.3) sign( (κ ; x1 ))sign( (κ ; x1 )) ≤ 0. ∂x1 ∂x1 So, by Lemma 5.10, if there exist N +1 stable positive steady states, then q(κ∗; a) > 0. Therefore, the condition (7.2) follows from the fact that

∗ ∗ Bτ ∗ Bk Ak Bτ γkj q(κ ; a) = q(κ ; − ) = (β11 − α11) λj κj Cj ( − ) , Aτ |Ak| |Ak| Aτ j k ,k τ X∈L ∈HY6=

and the equality (β11 − α11)λj = β1j − α1j (see (5.1)). Corollary 7.8. Given a network G (2.1) with a one-dimensional stoichiometric subspace, suppose cappos(G)=2N +1 (0 ≤ N < +∞). If for a rate-constant vector κ∗ and a total-constant vector c∗, G has 2N +1 nondegenerate positive steady states in Pc∗ , then N +1 of these positive steady states are stable if the condition (7.2) holds, and N of these positive steady states are stable if the condition (7.2) does not hold. Proof. Notice that if for (κ∗,c∗), all positive steady states of G are nondegenerate, then the equality (7.3) in the proof of Lemma 7.7 can be replaced with

∂q ∗ (i) ∂q ∗ (i−1) sign( (κ ; x1 ))sign( (κ ; x1 )) < 0. ∂x1 ∂x1 Then, the conclusion follows from a similar argument with the proof of Lemma 7.7. Proof of Theorem 3.4. Proof. (a) If cappos(G)=2N (0 ≤ N < +∞), then by Theorem 6.1, we have ∗ ∗ capnondeg(G)=2N. Thus, there exists a choice of (κ ,c ) such that G has exactly 2N nondegenerate positive steady states in Pc∗ (note here, there is no degenerate positive 24 XIAOXIAN TANG, AND ZHISHUO ZHANG

steady states since cappos(G)=2N). By Lemma 5.1, Lemma 5.3, and Theorem 5.6, N of these 2N positive steady states are stable. Note that by Corollary 5.7 (b) and the fact that cappos(G)=2N, we have capstab(G) ≤ N. Therefore, we conclude capstab(G)= N. (b) Assume cappos(G)=2N +1 (0 ≤ N < +∞). Notice that by Corollary 5.7 (b) and by the fact that cappos(G)=2N + 1, we have

(7.4) capstab(G) ≤ N +1.

We have the following two cases. (Case 1.) If W ∩ B= 6 ∅, then by Corollary 6.6, Wnondeg ∩ B= 6 ∅ (recall the definition of Wnondeg in (6.6)). That means there exists a choice of parameters (κ∗,c∗) such that G has 2N + 1 nondegenerate positive steady states and the condition (7.2) holds (recall that the inequality stated in the definition of B (3.14) is exactly the condition (7.2)). By Corollary 7.8, there are N +1 stable positive steady states. So, by (7.4), we have capstab(G)= N + 1. (Case 2.) Suppose W∩B = ∅. By(7.4), it is sufficient to prove that ∗ ∗ ∗ ∗ capstab(G) ≤ N. For any (κ ,c ), if (κ ,c ) ∈/ W, then G has at most 2N positive steady states. By Corollary 5.7 (b), at most N positive steady states are stable. If (κ∗,c∗) ∈ W, then G has 2N + 1 positive steady states. Note that (κ∗,c∗) ∈/ B since W∩B = ∅. By Corollary 5.7 (b) and Lemma 7.7, at most N positive steady states are stable. Therefore, we conclude that capstab(G) ≤ N. Proof of Corollary 3.5. Proof. If cappos(G)=2N (0 ≤ N < +∞), then the conclusion follows from the proof of the part (a) in Theorem 3.4. If cappos(G)=2N +1 (0 ≤ N < +∞), we have the following two cases. (Case 1.) If W ∩ B= 6 ∅, then the conclusion follows from Case 1 in the proof of Theorem 3.4. (Case 2.) If W∩B = ∅, then by Theorem 3.4 (b), capstab(G) = N. Note that by Theorem 6.1, capnondeg(G)=2N +1, and hence, there exists a choice of parameters (κ∗,c∗) such that G has exactly 2N + 1 nondegenerate positive steady states in Pc∗ . Since W∩B = ∅, we know the condition (7.2) does not hold for (κ∗,c∗) (again, recall the the inequality stated in the definition of B (3.14) is the condition (7.2)). By Corollary 7.8, N of these positive steady states are stable. Lemma 7.9. Given a network G (2.1) with two reactions, suppose cappos(G) = 2N +1 (0 ≤ N < +∞). If for a rate-constant vector κ∗ and a total-constant vector ∗ c , G has 2N +1 positive steady states in Pc∗ , and if N +1 of these steady states are stable, then

(7.5) (γτ1 − γτ2)(βτ1 − ατ1) < 0,

where τ and γτj (j ∈ {1, 2}) are defined in (3.11) and (3.6) for the total constant- vector c∗. Especially, if τ is the only index in [τ], then the condition (7.5) becomes (3.16), i.e., (ατ1 − ατ2)(βτ1 − ατ1) < 0. Proof. By [12, Lemma 4.1], if G contains only two reactions (i.e., m = 2), and if cappos(G) ≥ 1, then the stoichiometric subspace is one-dimensional. Since G contains two reactions, the set of indices L defined in (3.12) has only one element. So, the MULTISTABILITY OF ONE-DIMENSIONAL REACTION NETWORKS 25 left-hand side of the inequality (7.2) is

∗ Bk Ak Bτ γ (β − α ) κ C ( − ) k1 , if γτ <γτ 11 11 1 1 |Ak| |Ak| Aτ 1 2 (7.6) k∈H,k6=τ .  ∗ Bk Ak Bτ γk (β − α ) κ C Q ( − ) 2 , if γτ >γτ 12 12 2 2 |Ak| |Ak| Aτ 1 2  k∈H,k6=τ Q  Bτ 1 Bτ By Lemma 5.15 (2), for any k ∈ H\{τ}, we have Yk(− )= (Bk −Ak ) > 0. By Aτ |Ak| Aτ [12, Lemma 4.1], we have (β11 − α11)(β12 − α12) < 0. By Lemma 5.14, we have Cj > 0 for any j ∈{1, 2}. Therefore, the sign of (7.6) is equal to −sign(γτ1 −γτ2)(β11 −α11). βτ1−ατ1 Note also, by (3.1) and (3.11), we know Aτ = > 0. So, the inequality (7.2) β11−α11 stated in Lemma 7.7 becomes (γτ1 − γτ2)(βτ1 − ατ1) < 0 (i.e., the condition (7.5)). If there is only one index in the equivalence class [τ], then by (3.6), we have (γτ1 − γτ2) = (ατ1 − ατ2). So, the inequality in (3.15) (i.e., the condition (7.5)) becomes (3.16). Proof of Corollary 3.6. Proof. By lemma 7.9, the argument is similar to the proof of Theorem 3.4. 8. Discussion. Given a network G with a one-dimensional stoichiometric sub- ∗ space, suppose cappos(G) = 3. We assume for a rate-constant vector κ , G has 3 nondegenerate positive steady states in a stoichiometric compatibility class Pc∗ . Let ∗ q(κ; x1) be the function defined in (5.9) for the total-constant vector c . From Remark 7.5 and the proof of Theorem 3.4, we see that the sign of the left-hand side of the ∗ inequality in (3.14) is the Brouwer degree of the function q(κ ; x1) over a bounded subset of the interval I defined in (3.10). By (3.14), we see this Brouwer degree de- pends on the choice of the total-constant vector c and the rate-constant vector κ. We remark that it is proved in [3, Theorem 3], if a dissipative network admits no boundary steady states, then the Brouwer degree does not depend on the choice of c and κ. So, a natural question is for a general network G, if there exist two choices of parameters c and κ such that their corresponding Brouwer degrees have different signs but the network has three nondegenerate positive steady states for both choices of parameters. To put it simply, we wonder if there exist two choices of parameters yielding three nondegenerate positive steady states such that one yields multistability but the other one does not.

REFERENCES

[1] Christoph Bagowski, and James Ferrell Jr. Bistability in the JNK cascade. Curr. Biol., 11(15):1176–82, 2001. [2] Murad Banaji, and Casian Pantea. Some results on injectivity and multistationarity in chemical reaction networks. SIAM J. Appl. Dyn. Syst., 15(2):807–869, 2016. [3] Carsten Conradi, Elisenda Feliu, Maya Mincheva, and Carsten Wiuf. Identifying parameter regions for multistationarity. PLoS Comput. Biol., 13(10):e1005751, 2017. [4] Carsten Conradi, and Dietrich Flockerzi. Switching in Mass Action Networks Based on Linear Inequalities. SIAM J. Appl. Dyn. Syst., 11(1):110-134, 2012. [5] Carsten Conradi, and Casian Pantea. Multistationarity in biochemical networks: Results, analysis, and examples. Algebraic and Combinatorial Computational Biology, Academic Press, 279-317, 2019. [6] Gheorghe Craciun, and Martin Feinberg. Multiple equilibria in complex chemical reaction networks: I. the injectivity property. SIAM J. Appl. Math., 65:1526–1546, 2005. [7] Gheorghe Craciun, Yangzhong Tang, and Martin Feinberg. Understanding bistability in com- plex enzyme-driven reaction networks. PNAS, 103(23):8697–8702, 2006. [8] Biswa Nath Datta. An elementary proof of the stability criterion of Li´enard and Chipart. Linear Appl., 22:89–96, 1978. 26 XIAOXIAN TANG, AND ZHISHUO ZHANG

[9] Alicia Dickenstein, Mercedes Perez Millan, Anne Shiu, and Xiaoxian Tang. Multistationarity in Structured Reaction Networks. Bull. Math. Biol.. 81(5):1527–1581, 2019. [10] James Ferrell Jr., and Eric Machleder. The biochemical basis of an all-or-none cell fate switch in Xenopus oocytes. Science, 280(5365):895–898, 1998 [11] Hoon Hong, Xiaoxian Tang, and Bican Xia. Special algorithm for stability analysis of multi- stable biological regulatory systems. J. Symbolic Comput., 70:112–135, 2015. [12] Badal Joshi, and Anne Shiu. Which small reaction networks are multistationary? SIAM J. Appl. Dyn. Syst., 16(2):802–833, 2017. [13] Stefan M¨uller, Elisenda Feliu, Georg Regensburger, Carsten Conradi, Anne Shiu, and Alicia Dickenstein. Sign conditions for injectivity of generalized polynomial maps with applica- tions to chemical reaction networks and real algebraic . Found. Comput. Math., 16(1):69–97, 2016. [14] Maple (2020) Maplesoft, a division of Waterloo Maple Inc., Waterloo, Ontario. [15] Nida Obatake, Anne Shiu, Xiaoxian Tang, and Angelica Torres. Oscillations and bistability in a model of ERK regulation. J. Math. Biol., 79:1515–1549, 2019. [16] Lawrence Perko. Differential equations and dynamical systems, Volume 7 of Texts in Applied Mathematics. Springer-Verlag, New York, third edition, 2001. [17] Guy Shinar, and Martin Feinberg. Concordant chemical reaction networks. Math. Biosci., 240(2):92–113, 2012. [18] Anne Shiu, and Timo de Wolff. Nondegenerate multistationarity in small reaction networks. Discrete Contin. Dyn. Syst. B, 24(6):2683–2700, 2019. [19] Xiaoxian Tang, and Hao Xu. Multistability of small reaction networks. Accepted by SIAM J. Appl. Dyn. Syst., arXiv:2008.03846. [20] Ang´elica Torres, and Elisenda Feliu. Detecting parameter regions for bistability in reaction networks. SIAM J. Appl. Dyn. Syst., 20(1):1–37, 2021. [21] Carsten Wiuf, and Elisenda Feliu. Power-law kinetics and determinant criteria for the preclusion of multistationarity in networks of interacting species. SIAM J. Appl. Dyn. Syst., 12:1685– 1721, 2013. [22] Wen Xiong, and James Ferrell Jr. A positive-feedback-based bistable ‘memory module’ that governs a cell fate decision. Nature, 426:460–465, 2003. SUPPLEMENTARY MATERIAL Table 1 lists all files at the online repository: https://github.com/ZhishuoCode/ small-network

Table 1 Supporting Information Files

Name File Type Results Example.mw/.pdf Maple/PDF Example 3.8/Example 5.2 Plot.mw/.pdf Maple/PDF Example 6.5/Figure 1