Your new experience awaits. Try the new design now and help us make it even better

HYPOTHESIS AND THEORY article

Front. Syst. Biol., 21 October 2025

Sec. Multiscale Mechanistic Modeling

Volume 5 - 2025 | https://doi.org/10.3389/fsysb.2025.1693064

Structural properties and asymptotic behavior of bacterial two-component systems

  • 1Department of Information Engineering, Università degli Studi di Padova, Padova, Italy
  • 2Department of Molecular Medicine, Università degli Studi di Padova, Padova, Italy

Bacteria rely on two-component signaling systems (TCSs) to detect environmental cues and orchestrate adaptive responses. Despite their apparent simplicity, TCSs exhibit a rich spectrum of dynamic behaviors arising from network architectures, such as bifunctional enzymes, multi-step phosphorelays, transcriptional feedback loops, and auxiliary interactions. This study develops a generalized mathematical model of a TCS that integrates these various elements. Using systems-level analysis, we elucidate how network architecture and biochemical parameters shape key properties such as stability, monotonicity, and signal amplification. Analytical conditions are derived for when the steady-state levels of phosphorylated proteins exhibit robustness to variations in protein abundance. The model characterizes how equilibrium phosphorylation levels depend on the absolute and relative abundances of the two components. Specific scenarios are explored, including the MprAB system from Mycobacterium tuberculosis and the EnvZ/OmpR system from textit Escherichia coli, to describe the potential role of reverse phosphotransfer reactions. By combining mechanistic modeling with system-level techniques, such as nullcline analysis, this study offers a unified perspective on the design principles underlying the versatility of bacterial signal transduction. The generalized modeling framework lays a theoretical foundation for interpreting experimental dynamics and rationally engineering synthetic TCS circuits with prescribed response dynamics.

1 Introduction

Bacteria rely on two-component systems (TCSs) as their primary signaling modules to detect environmental cues and orchestrate adaptive responses. A canonical TCS consists of a membrane-bound sensor histidine kinase (SHK) and a cytoplasmic response regulator (RR). Upon stimulation, the SHK autophosphorylates on a conserved histidine and transfers the phosphoryl group to an aspartate on the RR, generating the active form (RR-P) that typically regulates gene expression. This minimal architecture is remarkably versatile, underpinning processes such as chemotaxis, nutrient sensing, antibiotic resistance, and virulence regulation (Tierney and Rather, 2019; Tiwari et al., 2017; Kirby, 2009; Ramos et al., 2022; Alvarez and Georgellis, 2023).

Despite their apparent simplicity, TCSs display a rich spectrum of topologies and dynamic behaviors (Zschiedrich et al., 2016; Groisman, 2016; Stock et al., 2000). In some systems, exemplified by CheA in bacterial chemotaxis, SHK functions exclusively as a kinase, phosphorylating the RR. However, in many TCSs, SHK is bifunctional, participating in both phosphorylation and dephosphorylation of its cognate RR. In such cases, the input signal can modulate either one or both of these enzymatic activities, effectively tuning the rates of kinase and/or phosphatase reactions. TCSs may implement single-step phosphotransfers or multi-step phosphorelays, adding regulatory complexity and potentially delaying signal propagation.

At the transcriptional level, many TCSs feature autoregulation: the phosphorylated RR activates transcription of both its own gene and the gene encoding its partner SHK, thereby forming a positive feedback loop (Goulian, 2010). This feedback can alter steady-state behavior, activation, and inactivation kinetics and generate transient overshoot or “memory” effects, whereby the system responds faster to repeated stimuli. Although less common, negative autoregulation—or even mixed positive and negative feedback—has been observed in specific systems, providing an additional layer of response modulation. Auxiliary proteins can further diversify TCS behaviors, either by directly interacting with SHKs or RRs or by mediating cross-talk between otherwise independent TCS pathways (Rao et al., 2021; Groisman, 2016).

Mathematical modeling has been pivotal in elucidating the emergent properties of TCSs (summarized in Table 1). Batchelor and Goulian (2003) demonstrated that the steady-state level of RR-P can be robust to protein abundance fluctuations when SHK is limiting, a property supported by experimental data. Shinar et al. (2007) formalized the conditions for input-output robustness, showing that robustness is compromised when multiple independent phosphorylation or dephosphorylation routes exist. Igoshin et al. (2008) identified conditions for bistability, particularly when unphosphorylated SHK and RR form “dead-end” complexes or when alternative phosphatases modulate RR-P turnover. Ray and Igoshin (2010), Mitrophanov et al. (2010), and Zorzan et al. (2021) explored the role of transcriptional feedback, showing that autoregulation can alter response speed, overshoot amplitude, and even affect the effective sign of feedback, enabling TCSs to switch between positive and negative regulatory modes depending on signal strength. These studies collectively highlight how bifunctionality, phosphorelays, and feedback loops produce rich dynamic behaviors—including robustness, bistability, and adaptive memory—that are now central themes in systems-level analyses of TCSs.

Table 1
www.frontiersin.org

Table 1. Comparison of previous findings on bacterial TCSs with results from this study’s model.

In this study, we develop a systems-level model of a generalized TCS model focusing on the MprAB system from Mycobacterium tuberculosis that integrates canonical phosphorylation cycles, bifunctional enzymatic activity, transcriptional feedback, and potential auxiliary interactions. Our modeling framework seeks to (i) dissect how network architecture and parameter regimes shape dynamic properties and provide robustness, to be adopted as a building block to implement overshoots, oscillations, and bistability, and (ii) provide a predictive foundation for interpreting experimental dynamics and guiding synthetic circuit design in bacterial signal transduction.

By combining mechanistic modeling with systems-level analysis, this study elucidates how bifunctionality, phosphorelays, and feedback loops shape the dynamic behavior of TCSs, providing insights into bacterial adaptation and a framework for the rational engineering of synthetic signaling circuits (Mukherji and van Oudenaarden, 2009; Pasotti et al., 2017; Müller et al., 2025).

2 Two-component system: mathematical model

The model we consider is a general version of the model proposed in Tiwari et al. (2010) to describe the functioning of the two-component system MprA/MprB in M. tuberculosis in its active state.

For the sake of generality, we refer to “response regulator” (RR) and “sensor histidine kinase” (SHK) rather than to MprA and MprB, respectively. Denoting by r (r*) and s (s*), the concentration of RR (phosphorylated RR) and SHK (phosphorylated SHK), respectively, the dynamic evolution of the two-component system is described by the following set of ODEs (see Supplemental Information of Tiwari et al. (2010), Equations (S39)–(S42)):

ṙ=kpKPr*sktKTrs*+kexdr*kexpr+νrkpdegr(1)
ṙ*=kpKPr*s+ktKTrs*kexdr*+kexprkpdegr*(2)
ṡ=kads*kaps+ktKTrs*+νskpdegs(3)
ṡ*=kads*+kapsktKTrs*kpdegs*(4)

—where

νr and νs are the production rate constants of RR and SHK, respectively;1

kp is the rate constant for the SHK-dependent dephosphorylation of RR*;

KP is the Michaelis–Menten constant for RR* dephosphorylation by SHK;

kt is the rate constant for the SHK*-dependent phosphorylation of RR;

KT is the Michaelis–Menten constant for RRSHK* phosphotransfer;

kexp and kexd are the exogenous phosphorylation and dephosphorylation rate constants, respectively;

kap and kad are the autophopshorylation and autodephosphorylation rate constants, respectively;

kpdeg is the protein degradation rate (assumed equal for RR and SHK).

One additional assumption worth highlighting is that the system is always considered to be in the active state. This is biologically reasonable as external stimuli often saturate the sensing capacity of the TCS. As a result, the transition of the sensor s from the inactive to the active state upon binding external stimuli can be neglected in the model, as well as the availability of ATP inside the cell to provide phosphate groups for the phosphorylation steps.

The overall system can be represented as in Figure 1.

Figure 1
Diagram illustrating a two-component system involving sensor histidine kinase (SHK) and response regulator (RR). In panel a, an input signal activates SHK, leading to phosphorylation and gene expression changes resulting in cellular responses. In panel b, the processes of phosphatase and phosphotransferase, regulation kinetics, and feedback autoregulation are depicted, showing dynamic interactions between SHK and RR, influencing gene expression.

Figure 1. Schema of the generalized TCS. Binding of the signal molecule and general activation of genes are reported in panel (a), while in panel (b) the part of the system described by Equations 14 is reported.

We define the total amount of RR and SHK as RT=r+r* and ST=s+s*, respectively, and rewrite the previous model presented in Equations 14 in the form shown in Equations 58:

ṘT=kpdegurRT(5)
ṙ*=kpKPr*STs*+ktKTRTr*s*kexdr*+kexpRTr*kpdegr*(6)
ṠT=kpdegusST(7)
ṡ*=kads*+kapSTs*ktKTRTr*s*kpdegs*(8)

—where urνr/kpdeg (usνs/kpdeg) is the net production rate of RR (SHK). Due to the separation of timescales between protein accumulation and phosphorylation/dephosphorylation events, we can assume that total concentrations of RR and SHK are preserved—namely, that RT and ST are constant. Under this assumption, we can normalize all state variables and consider the phosphorylated portion of RR and SHK

r*r*RTands*s*ST,

the dynamics of which are described by

ṙ*=kexd+kexp+kpdegr*kpKPSTr*1s*ktKTSTr*s*+ktKTSTs*+kexpṡ*=kad+kap+kpdegs*ktKTRTs*1r*+kap

Since we aim to provide a model describing the functioning of general two-component systems (TCSs) and unveiling its structural and asymptotic properties, from now on we will consider the following general formulation:

ṙ*=α1+α2r*α3STr*1s*α4STr*s*+α4STs*+α2f1r*,s*(9)
ṡ*=β1+β2s*β3RTs*1r*β4RTr*s*+β4RTr*+β2f2r*,s*(10)

Differential Equations 9, 10 describe the dynamics of the phosphorylated portions of RR and SHK—that is, ratio phosphorylated-RR (phosphorylated-SHK) over total RR (SHK)—under the assumption that total concentrations RT and ST are constant. Notice that in Equation 10, the terms β4RTr*s* and β4RTr* have been included for reasons of symmetry. Of course, this general formulation can be tailored to the specific two-component system under investigation. For instance, we immediately verify that, upon defining

α1=kexd+kpdeg,α2=kexp,α3=kpKP,α4=ktKTβ1=kad+kpdeg,β2=kap,β3=ktKT,β4=0,

Equations 9, 10 reduce to the MprA-MprB system proposed in Tiwari et al. (2010).

2.1 Structural properties

We note that, by the way that r* has been defined, it is dimensionless, and such that for every t0 it holds 0r*(t)1, r*=0 means that all RR are unphosphorylated, while r*=1 represents the situation with all RR phosphorylated. Clearly, the same holds for s*, and hence every state trajectory of the bidimensional system Equations 9, 10 belongs to the feasibility set C(r*,s*):0r*1,0s*1.

Proposition 1. The TCS model Equations 9, 10 exhibits a unique equilibrium point (req*,seq*) within the feasibility set C.

Proof. First, notice that the set C is positively invariant with respect to systems Equations 9, 10, so that if the state trajectory starts in C, then it stays in C for any t0. Positive invariance of the convex and compact set C ensures that there exists at least one equilibrium point in C—that is, a limit cycle or at least one stable equilibrium point (Blanchini and Miani, 2015—Theorem 4.21).

We now resort to Bendixon’s theorem to rule out the existence of closed orbits.2 Note that

df1dr*r*,s*=α1+α2α3ST1s*α4STs*<0,r*,s*Cdf2ds*r*,s*=β1+β2β3RT1r*β4RTr*<0,r*,s*C

Hence, div(f)df1dr*+df2ds* is not identically zero in any sub-region of the simply connected region C and does not change sign in C. Then, by Bendixon’s theorem (Sastry, 1999—Theorem 2.7), the set C contains no closed orbits of system Equations 9, 10.

Finally, we resort to nullcline analysis to prove the uniqueness of steady states. Setting dr*/dt=0 and ds*/dt=0 yields the following expressions for r* and s* nullclines:

r*=α4STs*+α2α1+α2+α3ST1s*+α4STs*gs*(11)
s*=β4RTr*+β2β1+β2+β3RT1r*+β4RTr*hr*(12)

A typical figure of RR and SHK nullclines is reported in Figure 2.

Figure 2
Graph showing the relationship between \( s^* \) and \( r^* \). A blue curve represents \( s^* = h(r^*) \), and a red curve represents \( r^* = g(s^*) \). Two points are marked: a blue star at \( (\alpha, \bar{\alpha}) \) and a red star at \( (\beta, \bar{\beta}) \). The gray shaded area denotes \( C_{eq} \). The axes range from 0 to 1.

Figure 2. Nullclines for α1=0.5, α2=1, α3=1, α4=7, β1=0.3, β2=1, β3=1, β4=4, RT=1, and ST=1. The Ceq region corresponds to the subregion where the equilibrium point is located, as detailed in Proposition 2.

From expression 11, it is easy to obtain s*=g1(r*):

g1r*=α1+α2+α3STr*α2α4ST1r*+α3STr*(13)

We define the function Δ(r*)h(r*)g1(r*) and note that, by the way Δ(r*) has been defined, if req*,seq* is an equilibrium point, then Δ(req*)=0; vice versa, if Δ(r̄*)=0 then r̄*,h(r̄*)=req*,seq* is an equilibrium point. It is a matter of computation to verify that Δ(r*) is a rational function—Δ(r*)=n(r*)d(r*)—and that both the numerator and denominator are polynomials of order 2:

nr*=β4RTr*+β2α4ST1r*+α3STr*+β1+β2+β3RT1r*+β4RTr*×α1+α2+α3STr*α2
dr*=β1+β2+β3RT1r*+β4RTr*α4ST1r*+α3STr*

Note that d(r*)>0 for every r*[0,1], and hence Δ(r*)=0 for some r*[0,1] if and only if n(r*)=0 for some r*[0,1]. Since n(0)>0 and n(1)<0, there certainly exists req*[0,1] such that n(req*)=0, and hence Δ(req*)=0—as already demonstrated, the system admits at least one equilibrium point in C. On the other hand, since n(r*) is a second-order polynomial, such an req* belonging to the interval [0,1] is unique—the system admits a unique equilibrium point C.

Remark 1. Remark 1. A closed-form expression for the equilibrium point of the TCS can be computed as the unique root in interval [0,1] of the second-order polynomial n(r*).

req*=α3β3RTSTα4β4RTST±A2α3β3RTSTα4β4RTST+α1β3RT+α2β3RTα1β4RTα2β4RT

with. A=(α3β3RTST+α4β4RTSTα1β3RT2α2β3RT+α2β4RTα3β1STα4β2STα1β1α2β1α1β2α2β2)24(α2β3RT+α4β2ST+α2β1+α2β2)(α3β3RTSTα4β4RTST+α1β3RT+α2β3RTα1β4RTα2β4RT)+α1β3RT+2α2β3RTα2β4RT+α3β1ST+α4β2ST+α1β1+α2β1+α1β2+α2β2

Proposition 1 states that all trajectories with initial conditions in C converge to a unique equilibrium point req*,seq*C. This means that, independently of the initial relative amounts of phosphorylated and unphosphorylated proteins, the proportion of phosphorylated to total RR will asymptotically equal req*, while the proportion of phosphorylated to total SHK will asymptotically tend to seq*. The following proposition identifies a subregion CeqC where the equilibrium point is located and hence provides upper and lower bounds to the phosphorylation levels req* and seq* asymptotically reached by the TCS.

Proposition 2. Consider the TCS described by models Equations 9, 10. The unique equilibrium point of the system, denoted by req*,seq*, belongs to the subregion

Ceqr*,s*:rmin*r*rmax*,smin*s*smax*C,

where

rmin*:=α2α1+α2+α3ST,rmax*:=α4ST+α2α1+α2+α4ST,smin*:=β2β1+β2+β3RT,smax*:=β4RT+β2β1+β2+β4RT

Proof. Consider the expression for RR nullcline Equation 11 and note that

gs*s*=STα1α4+α4α3ST+α2α3α1+α2+α3ST1s*+α4STs*2>0for everys*0,1,

and hence r* is strictly monotonically increasing in s*. The bounds on req* then follow from

g0=α2α1+α2+α3STrmin*,andg1=α4ST+α2α1+α2+α4STrmax*,

Analogous computations on SHK nullcline Equation 12 lead to upper and lower bounds on seq*.

The set Ceq is reported in Figure 2 for the set of parameters considered. We conclude this section with the following Lemma, which will be useful for subsequent derivations (see again Figure 2).

Lemma 1. Consider the TCS described by models Equations 9, 10, and define

ᾱ:=α2α3α2α3+α1α4α̂:=α2α1+α2β̄:=β2β3β2β3+β1β4β̂:=β2β1+β2

Then, RR nullcline Equation 11 always passes through (ᾱ,α̂)g(ᾱ)=α̂while SHK nullcline Equation 12 always passes through (β̄,β̂)h(β̄)=β̂.

This behavior can also be observed in Figure 3, where the dotted lines indicate the nullclines associated with higher values of RT and ST, while the dashed lines are the nullclines obtained with lower values of RT and ST, as described in the caption.

Figure 3
Graph showing two functions, \(s^* = h(r^*)\) in blue and \(r^* = g(s^*)\) in red, with dashed and dotted lines representing different behaviors. Both curves intersect around \(r^* = 0.5\) and \(s^* = 0.6\).

Figure 3. Nullclines for α1=0.5, α2=1, α3=1, α4=7, β1=0.3, β2=1, β3=1, and β4=4. The solid lines have RT=1 and ST=1 as in Figure 2; the dashed lines are obtained with RT=0.4 and ST=0.4; the dotted lines with RT=2 and ST=2.

Since verifying that g(ᾱ)=α̂ and h(β̄)=β̂ is just a matter of computation, the proof of Lemma 1 is omitted.

At this point, two observations are in order. First, the dimensionless values req* and seq* depend on the total amounts of RR and SHK proteins present within the system (recall that, due to time scale separation, so far we have assumed that the quantities RT and ST are constant). In other words, req* and seq* are continuous functions of RT and STreq*=req*(RT,ST) and seq*=seq*(RT,ST). The second observation is that uniform monotonicity of req*(RT,ST) and seq*(RT,ST) with respect to their arguments is not guaranteed. Depending on the values taken by the system parameters, equilibrium req* might decrease with RT when RT belongs to a specific interval, and increase with RT when it belongs to a different interval.

3 Relative concentrations

3.1 Low vs high RT concentration

In this section, we assume that RT and ST are independent.

Proposition 3. (Low RT concentration.) Consider the TCS described by models Equations 9, 10 and let the total SHK concentration ST be arbitrary but fixed. When the total RR concentration is extremely low—that is, for RT0—the equilibrium point asymptotically reached by the system is given by req*,seq*=(g(β̂),β̂).

Proof. By taking the limit for RT0 of the function h(r*) defined in Equation 12 and representing SHK nullcline,3 it can be seen that seq*=β̂. The result then follows by plugging seq* into RR-nullcine Equation 11.

Proposition 4. (High RT concentration.) Consider the TCS described by models Equations 9, 10 and let the total SHK concentration ST be arbitrary but fixed. When the total RR concentration is extremely high—that is, for RT+—the equilibrium point asymptotically reached by the system is rhR*,h(rhR*), with rhR* being the (unique) solution in the interval [0,1] of the quadratic equation A(r*)2+Br*+C=0, where

A:=α1+α2+α3STβ3α2+α4STβ4α1β4B:=α1+α2+α3STβ3+α2+α4STβ4α2β3C:=α2β3

More specifically, rhR*=BB24AC/2A.

Proof. Note that when RT+, the upper and lower bounds on seq* are given by smin*=0 and smax*=1, respectively, and hence do not provide any useful information. Taking the limit for RT+ of RR and SHK nullclines Equations 13, 12 yields

limRT+g1r*=α1+α2+α3STr*α2α4ST1r*+α3STr*limRT+hr*=β4r*β4r*+β31r*

Solving for limRT+g1(r*)=limRT+h(r*) leads to the quadratic equation A(r*)2+Br*+C=0. The result now follows upon noting that if α1+α2+α3STβ3>α1+α2+α4STβ4, then A>0 and B<0, otherwise A<0; by Descartes’ rule of signs, the quadratic equation has a unique positive solution.

Corollary 1. Consider the TCS described by models Equations 9, 10 and let the total SHK concentration ST be arbitrary but fixed. Assuming the total RR concentration to be very high—that is, RT+—then if β30 and β4=0, the equilibrium point asymptotically reached by the system is rmin*,0; if β3=0 and β40, the equilibrium point is rmax*,1.

Proof. Consider the scenario with β30 and β4=0 and note that in this case, smax*=β2β1+β2. Taking the limit for RT+ of SHK nullcline (12) yields

seq*=limRT+β2β1+β2+β3RT1r*=0,smin*=0.

Then, from RR nullcline Equation 11, we have req*=g(seq*)=rmin*4. The proof for the case β3=0 and β40 follows the same line and is hence omitted.

Figure 4 reports, for an illustrative set of parameters, equilibrium values req* and seq* as a function of RT.

Figure 4
Chart showing two curves on a grid. The red curve represents \(r^*_{eq}\) and stabilizes around 0.9. The blue curve, \(s^*_{eq}\), increases sharply from about 0.75, then levels off around 0.95. Key points marked with colored symbols: red circle for \(g(\hat{\beta})\), blue circle for \(\hat{\beta}\), red square for \(r^*_{hR}\), and blue square for \(h(r^*_{hR})\). The horizontal axis is labeled \(R_T\).

Figure 4. Equilibrium values req* and seq* continuously depend on the total amount of RR protein RT. Parameter values: α1=0.5, α2=1, α3=1, α4=7, β1=0.3, β2=1, β3=1, β4=4, and ST=1.

By symmetry, analogous results on the equilibrium point hold when the SHK total amount is extremely low or extremely high—ST0 or ST+.

3.2 Uniform monotonicity of the equilibrium with respect to RT and ST

We now consider small perturbations of RT and ST concentrations and investigate their effects on the equilibrium point req*,seq*.

We assume first that ST is constant and consider small perturbations of RT. The equilibrium values continuously depend on RT—that is, req*,seq*=gseq*,RT,h(req*,RT)—and this dependence is quantitatively described by

req*RT=gs*seq*RT(14)
seq*RT=hr*req*RT+hRT(15)

Conversely, if we assume that total concentration RT is constant while ST slowly varies, we have

req*ST=gs*seq*ST+gST(16)
seq*ST=hr*req*ST(17)

Putting together Equations 1417 and solving for the variation of equilibria with respect to RT and ST, we obtain

req*RT=hRTgs*1hr*gs*seq*RT=hRT1hr*gs*(18)
req*ST=gST1hr*gs*seq*ST=gSThr*1hr*gs*(19)

Proposition 5. Consider the TCS described by model Equations 9, 10, and let req*,seq* denote the (unique) equilibrium point of the system. The equilibrium values req*=req*(RT,ST) and seq*=seq*(RT,ST) are:

i) monotonically increasing in their arguments if α̂>β̄ and β̂>ᾱ;

ii) monotonically decreasing in their arguments if α̂<β̄ and β̂<ᾱ.

Proof. Observe that

hr*=RTβ1β4+β4β3RT+β3β2β1+β2+β3RT1r*+β4RTr*20for every r*0,1,

and by symmetry, also gs*0 for every s*[0,1]. Moreover, recall that the function Δ(r*)h(r*)g1(r*) is such that Δ(0)>0 and Δ(1)<0 (see proof of Theorem 1), and hence at the equilibrium Δr*=hr*g1r*<0hr*<g1r*. This, in turn, implies that

0<hr*g1r*=hr*gs*<1

Then, the sign of the partial derivatives Equations 18, 19 are solely determined by hRT and gST since all other terms are always non-negative. It is a matter of computation to verify that

hRT=β1β4+β2β3r*β2β3β1+β2+β3RT1r*+β4RTr*2,

and hence at equilibrium signhRT=sign(req*β̄). Exploiting again the symmetry of the system, we can claim that signgST=sign(seq*ᾱ). Hence, provided that req* (seq*) is greater than β̄ (respectively, ᾱ), both req* and seq* are monotonically increasing functions of RT (respectively, ST). Similarly, provided that req* (seq*) is smaller than β̄ (respectively, ᾱ), both req* and seq* are monotonically decreasing functions of RT (respectively, ST). It is clear from Figure 2 that when α̂>β̄ and β̂>ᾱ, the equilibrium values necessarily satisfy the inequalities req*>β̄ and seq*>ᾱ, and the thesis follows.

Remark 2. The conditions on the system parameters provided by proposition 5 are sufficient (but not necessary) for uniform monotonicity of the equilibrium concerning total concentrations RT and ST. It is worth noticing that such a result is extremely powerful; its strength resides in the fact that it does not depend on the specific form of the functions RT and ST (provided they are monotone). More specifically, let RT=fR(uext) and ST=fS(uext), where uext is an external signal and fR and fS are monotone functions. Then, uext=fR1(RT), and the relationship between ST and RT is given by ST=fSfR1(RT) (note that the composite function fSfR is itself monotone). Proposition 5 states that if α̂>β̄ and β̂>ᾱ, monotonicity of the equilibrium with respect to RT and ST is ensured independently on the specific form of the monotone functions fR and fS. If the previous conditions are not satisfied, uniform monotonicity is not guaranteed.

We now focus on the case where a proportionality relationship among RT and ST can be assumed: ST=fSfR(RT)=λRT. Note that this is a perfectly reasonable assumption when phosphorylated RR activates the transcription of both its gene and the gene encoding its partner SHK—see, for example, the mathematical description of the MprA/MprB two-component system adopted in (Tiwari et al., 2010).

Theorem 1. Consider the TCS described by models Equations 9, 10, and assume that total RR and SHK concentrations are related by ST=λRT, where λ>0 is a fixed (not necessarily known) proportionality coefficient. When the total RR concentration is extremely high—that is, for RT+—the (unique) equilibrium point asymptotically reached by the system is

req*,seq*=0,0,ifα3β3α4β4>11,1,ifα3β3α4β4<1

Proof. Compute the limit for RT+ of RR and SHK nullclines Equations 11, 12:

r*=limRT+gs*,λRT=α4s*α3+α4α3s*,for every λ>0(20)
s*=limRT+hr*,RT=β4r*β3+β4β3r*(21)

From expression Equation 21, it is easy to obtain

r*=h1s*=β3s*β4+β3β4s*

Substituting the previous expression into Equation 20 and solving for s* yields the following quadratic equation:

s*2α4β4α3β3s*+α3β3α4β4=0

Then, the only two possible equilibrium points are (req*,seq*)=(0,0) and (req*,seq*)=(1,1). To determine which is the right solution, we need to resort to the intersection condition hr*gs*<1 (see the proof of Proposition 5). Indeed, it is straightforward to verify that

limRT+hr*=β3β4β3+β4β3r*2limRT+gs*=α3α4α3+α4α3s*2,

and hence

hr*gs*|(r*,s*)=(0,0)=β4β3α4α3hr*gs*|(r*,s*)=(1,1)=β3β4α3α4,

which uniquely determines the limiting equilibrium pair once the quantity β4β3α4α3 is known.

Remark 3. The previous result does not require knowledge of the value assumed by the proportionality coefficient λ; we just need to know that a proportionality coefficient continuously relates RT and ST

4 Absolute concentrations

We have thus far analyzed the properties (asymptotic behavior and monotonicity) of relative concentrations: of the ratio between phosphorylated and unphosphorylated protein concentrations. A fundamental and crucial point is that these properties do not necessarily hold for absolute concentrations too: the fact that the relative concentration req* tending to 0 does not imply that absolute concentration req* tends to 0; similarly, uniform monotonicity of req* for RT does not imply uniform monotonicity of r* to RT. To understand this point, note that the relative concentration req* tends to 0 when total RR concentration asymptotically grows to infinity (i.e., RT+) and r* asymptotically approaches a given saturation level req*0. Regarding monotonicity, since r*=r*RT, it holds that

r*RT=r*RTRT+r*

It is clear that if req* is a monotonically increasing function of RT (namely, r*RT>0), so is req*. On the contrary, if req* is a monotonically decreasing function of RT, and hence r*RT<0; monotonicity of req* with respect to RT is not guaranteed.

In the following, we analyze the asymptotic behavior of absolute concentrations req* and seq* when RT grows to infinity, under the assumption that RR and SHK total concentrations are linearly related with the proportionality coefficient λST=λRT.

Theorem 2. Consider the TCS described by models Equations 9, 10 and assume that the total RR and SHK concentrations are linearly related by ST=λRT, where λ>0 is a fixed proportionality coefficient. When total RR concentration is sufficiently high—that is, for RT+, RR and SHK—then absolute concentrations asymptotically approach the equilibrium values:

req*=α2β3+λα4β2λα3β3α4β4seq*=α2β4+λα3β2α3β3α4β4,

respectively.

Proof. We claim that for a sufficiently high RT, absolute equilibrium concentrations req* and seq* asymptotically approach saturation levels ρ and σ:

limRT+req*RT,λRTRT=ρ,limRT+seq*RT,λRTλRT=σ

We now seek to determine the values ρ and σ. First, we note that

ρ=limRT+gseq*RT,λRT,λRTRT=limRT+α4λRTseq*RT,λRT+α2α1+α2+α3λRT1seq*RT,λRT+α4λRTseq*RT,λRTRT=limRT+α4σ+α2α1+α2+α3λRT+α4α3σRT=α4σ+α2λα3

Analogously, the limit of seq* for RT+ can be computed as

σ=limRT+hreq*RT,λRT,λRTλRT=limRT+β4RTreq*RT,λRT+β2β1+β2+β3RT1req*RT,λRT+β4RTreq*RT,λRTλRT=limRT+β4ρ+β2β1+β2+β3RT+β4β3ρλRT=β4ρ+β2λβ3

Therefore, we need to solve the linear system:

ρλα3=α4σ+α2σβ3=β4ρ+β2λ

Solving for ρ and σ yields

ρ=α2β3+λα4β2λα3β3α4β4,σ=α2β4+λα3β2α3β3α4β4,

Thus, the proof is concluded.

It follows from Theorem 2 that for sufficiently high RT, while the amount of phosphorylated SHK increases with λ, the amount of phosphorylated RR is a decreasing function of λ, such that

req*=1λα2β3+α4β2α3β3α4β4(22)

5 Discussion

A distinguishing feature of the proposed TCS mathematical model is that it accounts for a variety of reactions, including RR phosphorylation and dephosphorylation through external (exogenous) pathways, SHK autophosphorylation and autodephosphorylation, RR phosphorylation via phosphotransfer from SHK, and RR dephosphorylation via SHK. Of course, by setting 0 for one or more parameters, the model can be tailored to specific two-component systems (TCSs) and/or situations in which some of the previous reactions are negligible.

One of the best characterized examples of TCS is the EnvZ/OmpR system in Escherichia coli, which responds to changes in environmental osmolality by regulating the expression of the outer membrane porins OmpF and OmpC. As in many TCSs, EnvZ is a bifunctional sensor histidine kinase, meaning that it phosphorylates and dephosphorylates the response regulator OmpR. Batchelor and Goulian (2003) proposed a mathematical model of the EnvZ/OmpR TCS and experimentally tested the model’s predictions. Their main finding was that for sufficiently high amounts of OmpR, when total EnvZ in the cell is much less abundant than total OmpR5, the steady-state level of phosphorylated OmpR is robust (insensitive) to fluctuations in EnvZ and OmpR concentrations. This model accounts for the autokinase, phosphotransfer, and phosphatase activities of EnvZ and neglects the exogenous phosphorylation and dephosphorylation of OmpR. Casting such a scenario into our mathematical framework means setting α2 and β4 to 0. Theorem 2 then implies that the equilibrium absolute concentration for OmpR is given by req*=α4β2α3β3, and hence, consistent with Batchelor and Goulian (2003), does not depend on EnvZ total concentration. However, our model shows that if an exogenous RR phosphorylation flux is present (α20), the previous result fails; when an external pathway for OmpR phosphorylation is present, the steady-state concentration of phosphorylated OmpR is (higher and) decreasing with λ (see Equation 22). Notably, Batchelor and Goulian (2003) predicted, via theoretical analysis and experimental verification with fluorescent reporter strains, that when condition STRT does not hold, the steady-state value of OmpR-P decreases with increasing total EnvZ concentration. This is consistent with our theoretical results, which also shed light on the role of an EnvZ-independent mechanism for OmpR phosphorylation.

Furthermore, our analysis allows the characterization of the steady-state concentration of the histidine kinase: seq*=λβ2β3 (recall that β4=0). As expected, our model predicts that the amount of phosphorylated EnvZ increases with more vigorous autokinase activity (β2) and decreases with stronger phosphotransfer activity of the histidine kinase (β3).

Finally, while our analysis demonstrates the existence of a single robust equilibrium of the system (Theorem 1), it is instructive to consider the possibility of using such a building block as part of a closed-loop system with positive retroactivity, which could lead to oscillatory or bistable behaviors (Igoshin et al., 2008; Zorzan et al., 2021; Tiwari et al., 2010).

5.1 Phosphotransfer and reverse phosphotransfer reactions

Bifunctional sensor histidine kinase exerts both positive and negative control through SHK phosphotransfer and phosphatase activity, respectively. While the biochemical reactions underlying SHK kinase activity are reasonably well understood, the mechanisms of phosphatase activity represent a long-standing question, the investigation of which has led to the formulation of multiple hypotheses (see Huynh and Stewart, 2011 for an overview). An early hypothesis, first proposed by Dutta and Inouye (1996), identified reverse transfer of the phosphoryl group from phosphorylated RR to SHK as a potential RR dephosphorylation mechanism. Such a hypothesis was prompted by experimental results conducted on EnvZ/OmpR system in E. coli (Dutta and Inouye, 1996; Zhu et al., 2000), showing that reverse transfer of the phosphoryl group from OmpR-P to EnvZ was detected in the early period of the phosphatase reaction with domain A of EnvZ—specifically with the EnvZ kinase phosphatase+ mutant (EnvZ.N347D), and, under certain conditions, with the wild-type EnvZ.

Even if later experiments invalidated the reverse phosphotransfer model (Hsing and Silhavy, 1997), it is universally recognized that reverse phosphotransfer can occur under certain conditions. As pointed out by Gao and Stock (2009), multiple mechanisms may have evolved for phosphatase activities, and individual histidine kinases may utilize different regulatory strategies. We now aim to theoretically investigate a scenario in which both direct and reverse phosphotransfer reactions occur, and a distinct phosphatase activity of the sensor histidine is present.

Since the kinase activity of SHK takes the form of a phosphotransfer reaction (by which a phosphoryl group is transferred from phosphorylated SHK to RR), reaction rates α4 and β3 are actually equal—α4=β3. We first assume that only SHK exhibits phosphotransfer activity (β4=0), and we rename α3 as α3p, where superscript p stands for “phosphatase activity” (of the SHK). It follows from Theorem 2 that when total RR concentration is sufficiently high, steady-state absolute concentrations are given by req*=1λα2+β2α3p and seq*=λβ2β3.

When reverse phosphotransfer from phosphorylated RR to SHK occurs, the reaction rate β4 is non-zero and α3=α3p+α3rt, with α3rt=β4 (where superscript rt stands for “reverse phosphotransfer”). Then, recalling that α4=β3, Theorem 2 yields

req*=1λα2+β2α3p and seq*=λβ2β3+α2β4+λα3rtβ2α3pβ3=λβ2β3+α3rtα2+λβ2α3pβ3

This indicates that, even if reverse phosphotransfer occurs, the absolute concentration of phosphorylated RR remains unchanged. While this may seem contradictory at first, it is easily explained by noting that reverse phosphotransfer from phosphorylated RR to SHK is exactly compensated by the increased direct phosphotransfer from phosphorylated SHK to RR. On the contrary, when the reverse phosphotransfer reaction occurs, our analysis shows that the absolute concentration of SHK increases and that such an increase is larger for higher values of the reverse phosphotransfer rate (bigger α3rt) and/or for larger amounts of total SHK concentration (bigger λ).

This study’s main findings are summarized here in comparison with the literature.

6 Conclusion

We here developed a generalized mathematical model for bacterial two-component signaling systems that integrates canonical phosphorylation cycles, bifunctional enzymatic activities, transcriptional feedback, and potential auxiliary interactions. Through systems-level analysis, we elucidated how network architecture and parameter regimes shape key dynamic properties and robustness.

Our modeling framework provides a predictive foundation for interpreting experimental dynamics, as illustrated for the EnvZ/OmpR system, and for guiding the rational design of synthetic signaling circuits. We demonstrated that the bifunctionality of the sensor histidine kinase, multi-step phosphorelays, and transcriptional feedback, which are incorporated into the model, enable rich behaviors that allow TCSs to precisely tune cellular responses to diverse environmental stimuli.

Notably, we derived analytical conditions in Propositions 3, Propositions 4, Propositions 5 and Theorem 1 under which the steady-state levels of phosphorylated proteins exhibit input–output robustness, overshoot, or bistability. We also characterized in Sections 34 how the equilibrium phosphorylation levels depend on the absolute and relative abundances of the two components. These insights are critical for understanding natural mechanisms of bacterial adaptation and for forward-engineering synthetic gene circuits with prescribed dynamics.

By combining the mechanistic modeling framework with systems analysis techniques, such as nullcline analysis, this study provides a unified perspective on the structural design principles that underlie the remarkable versatility of two-component signal transduction. The proposed generalized model lays a theoretical foundation for further experimental investigations, such as exploring reverse phosphotransfer mechanisms, and establishes a framework for rationally harnessing two-component systems in synthetic biology applications.

Data availability statement

The original contributions presented in the study are included in the article/supplementary material; further inquiries can be directed to the corresponding author.

Author contributions

IZ: Writing – original draft, Methodology, Formal Analysis, Investigation, Conceptualization, Visualization. CC: Methodology, Writing – review and editing, Formal Analysis, Visualization. LS: Funding acquisition, Resources, Supervision, Methodology, Writing – original draft, Conceptualization. MB: Visualization, Project administration, Investigation, Supervision, Writing – review and editing.

Funding

The author(s) declare that no financial support was received for the research and/or publication of this article.

Conflict of interest

The authors declare that the research was conducted in the absence of any commercial or financial relationships that could be construed as a potential conflict of interest.

Generative AI statement

The author(s) declare that no Generative AI was used in the creation of this manuscript.

Any alternative text (alt text) provided alongside figures in this article has been generated by Frontiers with the support of artificial intelligence and reasonable efforts have been made to ensure accuracy, including review by the authors wherever possible. If you identify any issues, please contact us.

Publisher’s note

All claims expressed in this article are solely those of the authors and do not necessarily represent those of their affiliated organizations, or those of the publisher, the editors and the reviewers. Any product that may be evaluated in this article, or claim that may be made by its manufacturer, is not guaranteed or endorsed by the publisher.

Footnotes

1Actually, in (Tiwari et al., 2010) production of RR and SHK is described by the summation of two activating Hill functions. As explained later, since we are focusing on the functioning of the Two-Component System, the separation of time scales allows us to assume constant production rates.

2Since every limit cycle is a closed orbit, ruling out the existence of closed orbits automatically excludes the existence of limit cycles.

3An equivalent way to see that seq*=β̂ when RT0 is noticing that in this case both smin* and smax* tend to β2/β1+β2=β̂, and hence the subregion Ceq reduces to a line.

4Alternatively, the result directly follows from Proposition 4 with β3=0.

5As reported, for instance, in (Hsing and Silhavy, 1997), in vivo OmpR is nearly 100-fold more abundant than EnvZ.

References

Alvarez, A. F., and Georgellis, D. (2023). Environmental adaptation and diversification of bacterial two-component systems. Curr. Opin. Microbiol. 76, 102399. doi:10.1016/j.mib.2023.102399

PubMed Abstract | CrossRef Full Text | Google Scholar

Batchelor, E., and Goulian, M. (2003). “Robustness and the cycle of phosphorylation and dephosphorylation in a two-component regulatory system,”, 100. Proceedings of the National Academy of Sciences, (Washington, D.C., USA: National Academy of Sciences), 691–696. doi:10.1073/pnas.0234782100

CrossRef Full Text | Google Scholar

Blanchini, F., and Miani, S. (2015). Set-theoretic methods in control systems and control: Foundations and applications. Basel, Switzerland: Birkhäuser Basel.

Google Scholar

Dutta, R., and Inouye, M. (1996). Reverse phosphotransfer from OmpR to EnvZ in a kinase-/phosphatase+ mutant of EnvZ (EnvZ.N347D), a bifunctional signal transducer of Escherichia coli. J. Biol. Chem. 271, 1424–1429. doi:10.1074/jbc.271.3.1424

PubMed Abstract | CrossRef Full Text | Google Scholar

Gao, R., and Stock, A. (2009). Biological insights from structures of two-component proteins. Annu. Rev. Microbiol. 63, 133–154. doi:10.1146/annurev.micro.091208.073214

PubMed Abstract | CrossRef Full Text | Google Scholar

Goulian, M. (2010). Two-component signaling circuit structure and properties. Curr. Opin. Microbiol. 13, 184–189. doi:10.1016/j.mib.2010.01.009

PubMed Abstract | CrossRef Full Text | Google Scholar

Groisman, E. (2016). Feedback control of two-component regulatory systems. Annu. Rev. Microbiol. 70, 103–124. doi:10.1146/annurev-micro-102215-095331

PubMed Abstract | CrossRef Full Text | Google Scholar

Hsing, W., and Silhavy, T. (1997). Function of conserved histidine-243 in phosphatase activity of EnvZ, the sensor for porin osmoregulation in Escherichia coli. J. Bacteriol. 179, 3729–3735. doi:10.1128/jb.179.11.3729-3735.1997

PubMed Abstract | CrossRef Full Text | Google Scholar

Huynh, T., and Stewart, V. (2011). Negative control in two-component signal transduction by transmitter phosphatase activity. Mol. Microbiol. 82, 275–286. doi:10.1111/j.1365-2958.2011.07829.x

PubMed Abstract | CrossRef Full Text | Google Scholar

Igoshin, O., Alves, R., and Savageau, M. (2008). Hysteretic and graded responses in bacterial two-component signal transduction. Mol. Microbiol. 68, 1196–1215. doi:10.1111/j.1365-2958.2008.06221.x

PubMed Abstract | CrossRef Full Text | Google Scholar

Kirby, J. R. (2009). Chemotaxis-like regulatory systems: unique roles in diverse bacteria. Annu. Rev. Microbiol. 63, 45–59. doi:10.1146/annurev.micro.091208.073221

PubMed Abstract | CrossRef Full Text | Google Scholar

Mitrophanov, A., Hadley, T., and Groisman, E. (2010). Positive autoregulation shapes response timing and intensity in two-component signal transduction systems. J. Mol. Biol. 401, 671–680. doi:10.1016/j.jmb.2010.06.051

PubMed Abstract | CrossRef Full Text | Google Scholar

Mukherji, S., and van Oudenaarden, A. (2009). Synthetic biology: understanding biological design from synthetic circuits. Nat. Rev. Genet. 10, 859–871. doi:10.1038/nrg2697

PubMed Abstract | CrossRef Full Text | Google Scholar

Müller, M. M., Arndt, K. M., and Hoffmann, S. A. (2025). Genetic circuits in synthetic biology: broadening the toolbox of regulatory devices. Front. Synthetic Biol. 3, 1548572. doi:10.3389/fsybi.2025.1548572

CrossRef Full Text | Google Scholar

Pasotti, L., Bellato, M., Casanova, M., Zucca, S., Cusella De Angelis, M. G., and Magni, P. (2017). Re-using biological devices: a model-aided analysis of interconnected transcriptional cascades designed from the bottom-up. J. Biol. Eng. 11, 50. doi:10.1186/s13036-017-0090-3

PubMed Abstract | CrossRef Full Text | Google Scholar

Ramos, A. L., Aquino, M., García, G., Gaspar, M., de la Cruz, C., Saavedra-Flores, A., et al. (2022). Rpus/r is a novel two-component signal transduction system that regulates the expression of the pyruvate symporter mctp in Sinorhizobium fredii ngr234. Front. Microbiol. 13, 871077. doi:10.3389/fmicb.2022.871077

PubMed Abstract | CrossRef Full Text | Google Scholar

Rao, S., and Igoshin, O. (2021). Overlaid positive and negative feedback loops shape dynamical properties of phopq two-component system. PLoS Comput. Biol. 17, e1008130. doi:10.1371/journal.pcbi.1008130

PubMed Abstract | CrossRef Full Text | Google Scholar

Ray, J., and Igoshin, O. (2010). Adaptable functionality of transcriptional feedback in bacterial two-component systems. PLOS Comput. Biol. 6, e1000676–10. doi:10.1371/journal.pcbi.1000676

PubMed Abstract | CrossRef Full Text | Google Scholar

Sastry, S. (1999). Nonlinear systems. Analysis, stability, and control. Interdiscip. Appl. Math. 10. doi:10.1007/978-1-4757-3108-8

CrossRef Full Text | Google Scholar

Shinar, G., Milo, R., Martínez, M., and Alon, U. (2007). “Input-output robustness in simple bacterial signaling systems,”, 104. Proceedings of the National Academy of Sciences, (Washington, D.C., USA: National Academy of Sciences), 19931–19935. doi:10.1073/pnas.0706792104

CrossRef Full Text | Google Scholar

Stock, A., Robinson, V., and Goudreau, P. (2000). Two-component signal transduction. Annu. Rev. Biochem. 69, 183–215. doi:10.1146/annurev.biochem.69.1.183

PubMed Abstract | CrossRef Full Text | Google Scholar

Tierney, A. R., and Rather, P. N. (2019). Roles of two-component regulatory systems in antibiotic resistance. Future Microbiol. 14, 533–552. doi:10.2217/fmb-2019-0002

PubMed Abstract | CrossRef Full Text | Google Scholar

Tiwari, A., Balazsi, G., Gennaro, M., and Igoshin, O. (2010). The interplay of multiple feedback loops with post-translational kinetics results in bistability of mycobacterial stress response. Phys. Biol. 7, 036005. doi:10.1088/1478-3975/7/3/036005

PubMed Abstract | CrossRef Full Text | Google Scholar

Tiwari, S., Jamal, S. B., Hassan, S. S., Carvalho, PVSD, Almeida, S., Barh, D., et al. (2017). Two-component signal transduction systems of pathogenic bacteria as targets for antimicrobial therapy: an overview. Front. Microbiol. 8, 1878. doi:10.3389/fmicb.2017.01878

PubMed Abstract | CrossRef Full Text | Google Scholar

Zhu, Y., Qin, L., Yoshida, T., and Inouye, M. (2000). Phosphatase activity of histidine kinase EnvZ without kinase catalytic domain. Proc. Natl. Acad. Sci. 97, 7808–7813. doi:10.1073/pnas.97.14.7808

PubMed Abstract | CrossRef Full Text | Google Scholar

Zorzan, I., Del Favero, S., Giaretta, A., Manganelli, R., Di Camillo, B., and Schenato, L. (2021). Mathematical modelling of sige regulatory network reveals new insights into bistability of mycobacterial stress response. BMC Bioinforma. 22, 558. doi:10.1186/s12859-021-04372-5

PubMed Abstract | CrossRef Full Text | Google Scholar

Zschiedrich, C., Keidel, V., and Szurmant, H. (2016). Molecular mechanisms of two-component signal transduction. J. Mol. Biol. 428, 3752–3775. doi:10.1016/j.jmb.2016.08.003

PubMed Abstract | CrossRef Full Text | Google Scholar

Keywords: two-component systems, MprAB Mycobacterium, EnvZ, OmpR, synthetic biology, sensor histidine kinase, response regulator, odes

Citation: Zorzan I, Cimolato C, Schenato L and Bellato M (2025) Structural properties and asymptotic behavior of bacterial two-component systems. Front. Syst. Biol. 5:1693064. doi: 10.3389/fsysb.2025.1693064

Received: 26 August 2025; Accepted: 19 September 2025;
Published: 21 October 2025.

Edited by:

Luis Diambra, National University of La Plata, Argentina

Reviewed by:

Alan Givré, National Scientific and Technical Research Council (CONICET), Argentina
Juan Ignacio Marrone, National Scientific and Technical Research Council (CONICET), Argentina

Copyright © 2025 Zorzan, Cimolato, Schenato and Bellato. This is an open-access article distributed under the terms of the Creative Commons Attribution License (CC BY). The use, distribution or reproduction in other forums is permitted, provided the original author(s) and the copyright owner(s) are credited and that the original publication in this journal is cited, in accordance with accepted academic practice. No use, distribution or reproduction is permitted which does not comply with these terms.

*Correspondence: Irene Zorzan, em9yemFuLmlyZW5lQGdtYWlsLmNvbQ==; Massimo Bellato, bWFzc2ltby5iZWxsYXRvQHVuaXBkLml0

These authors have contributed equally to this work and share last authorship

Disclaimer: All claims expressed in this article are solely those of the authors and do not necessarily represent those of their affiliated organizations, or those of the publisher, the editors and the reviewers. Any product that may be evaluated in this article or claim that may be made by its manufacturer is not guaranteed or endorsed by the publisher.