Hostname: page-component-5cf477f64f-rdph2 Total loading time: 0 Render date: 2025-03-31T06:47:07.542Z Has data issue: false hasContentIssue false

Dynamics of homeostats: the basis of electrical, chemical, hydraulic, pH and calcium signaling in plants

Published online by Cambridge University Press:  21 March 2025

Leslie Contador-Álvarez
Affiliation:
Programa de Doctorado en Ciencias mención Modelado de Sistemas Químicos y Biológicos, Universidad de Talca, Talca, Chile
Tamara Rojas-Rocco
Affiliation:
Programa de Doctorado en Ciencias mención Modelado de Sistemas Químicos y Biológicos, Universidad de Talca, Talca, Chile
Talía Rodríguez-Gómez
Affiliation:
Programa de Doctorado en Ciencias mención Modelado de Sistemas Químicos y Biológicos, Universidad de Talca, Talca, Chile
María Eugenia Rubio-Meléndez
Affiliation:
Electrical Signaling in Plants (ESP) Laboratory–Center of Bioinformatics, Simulation and Modeling (CBSM), Faculty of Engineering, Universidad de Talca, Talca, Chile
Janin Riedelsberger
Affiliation:
Electrical Signaling in Plants (ESP) Laboratory–Center of Bioinformatics, Simulation and Modeling (CBSM), Faculty of Engineering, Universidad de Talca, Talca, Chile
Erwan Michard
Affiliation:
Instituto de Ciencias Biológicas, Universidad de Talca, Talca, Chile
Ingo Dreyer*
Affiliation:
Electrical Signaling in Plants (ESP) Laboratory–Center of Bioinformatics, Simulation and Modeling (CBSM), Faculty of Engineering, Universidad de Talca, Talca, Chile
*
Corresponding author: Ingo Dreyer; Email: [email protected]

Abstract

Homeostats are important to control homeostatic conditions. Here, we have analyzed the theoretical basis of their dynamic properties by bringing the K homeostat out of steady state (i) by an electrical stimulus, (ii) by an external imbalance in the K+ or H+ gradient or (iii) by a readjustment of transporter activities. The reactions to such changes can be divided into (i) a short-term response (tens of milliseconds), where the membrane voltage changed along with the concentrations of ions that are not very abundant in the cytosol (H+ and Ca2+), and (ii) a long-term response (minutes and longer) caused by the slow changes in K+ concentrations. The mechanistic insights into its dynamics are not limited to the K homeostat but can be generalized, providing a new perspective on electrical, chemical, hydraulic, pH and Ca2+ signaling in plants. The results presented here also provide a theoretical background for optogenetic experiments in plants.

Type
Theories
Creative Commons
Creative Common License - CCCreative Common License - BY
This is an Open Access article, distributed under the terms of the Creative Commons Attribution licence (http://creativecommons.org/licenses/by/4.0), which permits unrestricted re-use, distribution and reproduction, provided the original article is properly cited.
Copyright
© The Author(s), 2025. Published by Cambridge University Press in association with John Innes Centre

1. Introduction

Homeostasis is the ability of organisms to maintain a stable internal state by compensating for changes in their environment through a regulated exchange of matter and energy with the exterior. An important example is ion homeostasis in plants (Serrano & Rodriguez-Navarro, Reference Serrano and Rodriguez-Navarro2001). Homeostatic ion balance is established through the coordinated activity of membrane transporters that act together in transporter networks. The networks display the inherent property of defining the transmembrane concentration gradient in a steady state depending on the relative activity of the individual transporters. We previously proposed to call these remarkable transporter network homeostats (Dreyer, Reference Dreyer2021). A homeostat comprises differently energized types of transporters that are permeable for a specific ion species, either alone or together with other ions, for example, H+. Unbiased, systemic modeling approaches combining thermodynamics with biophysics and mathematics showed that the homeostatic property of homeostats emerges from the coupling of the transporters’ activity through the membrane voltage and the electrochemical gradients of the permeating ions. In a previous report, we presented a hands-on tutorial for modeling a homeostat in (dynamic) equilibrium using the potassium (K) homeostat as an example (Dreyer et al., Reference Dreyer, Hernández-Rojas, Bolua-Hernández, Tapia-Castillo, Astola-Mariscal, Díaz-Pico, Mérida-Quesada, Vergara-Valladares, Arrey-Salas and Rubio-Meléndez2024). We derived analytical solutions and analyzed the K homeostat in computational simulations in its steady state. However, steady-state conditions are only part of the life of a plant cell.

Here, we now analyze the homeostat out of equilibrium and monitor its dynamic behavior. Using again the example of the K homeostat, we present the mathematical background that describes its deflection from a steady state and how the thermodynamic constraints determine its reactions to perturbations. In particular, we analyzed the responses to

  1. (i) An electrical stimulus such as occurs during an action potential, for instance,

  2. (ii) An imbalance in the K+ or H+ gradients caused, for example, by environmental stress,

  1. (iii) A change in transporter activities, for example, caused by posttranslational modifications such as (de)phosphorylation but also during hydraulic or chemical signaling.

The detour via mathematics and computer simulations reveals again far-reaching insights into fundamental physiological properties of homeostats in general (not only the K homeostat) that go beyond the mere establishment of homeostasis. The results strongly suggest that homeostats are deeply involved in electrical, chemical, hydraulic, pH and calcium signaling in plants.

The study of the K homeostat under nonequilibrium conditions could also provide important impetus for agriculture, as K+ plays a variety of important physiological roles (see the article on K+ homeostasis in this collection; Wegner et al., Reference Wegner, Pottosin, Dreyer and Shabala2025). Most of these functions are indirect effects of K+ due either to its concentration (osmolarity and electrostatic effects) or its fluxes (membrane voltage, indirect effect on pH and mass fluxes of ions). It is not always easy to understand the indirect effects; this is where mathematical modeling of homeostats can help. More fundamentally, the quantification of the activity of homeostats and its change through indirect physiological parameters such as voltage, concentration, pressure and pH requires numerical quantification, which is only possible through mathematical modeling. Understanding how transporter networks in plants respond to adverse conditions is important for developing strategies to improve crop resilience so that they can thrive in a changing environment.

2. Methods

The mathematical description of the K homeostat and its analysis in steady state have been presented in detail in Dreyer et al. (Reference Dreyer, Hernández-Rojas, Bolua-Hernández, Tapia-Castillo, Astola-Mariscal, Díaz-Pico, Mérida-Quesada, Vergara-Valladares, Arrey-Salas and Rubio-Meléndez2024). In short, the K homeostat (Supplementary Figure S1, Supplementary Table S1) is composed of K+ channels (KC), H+/K+ symporters (HKs), H+/K+ antiporters (HKa) and H+-ATPases, embedded in a membrane with membrane capacitance and membrane voltage (C, V). The membrane separates the internal and external compartments (volumes Volin , Volout ) with respective proton and potassium concentrations ([H+ ]in, [H+ ]out, [K+ ]in, [K+ ]out). The different K+ transporters are characterized by the parameters gKC , gHKs and gHKa , which indicate their activity relative to the maximal activity of the H+-ATPase. In steady state, the K+ and H+ fluxes through the transporters are represented by (Dreyer et al., Reference Dreyer, Hernández-Rojas, Bolua-Hernández, Tapia-Castillo, Astola-Mariscal, Díaz-Pico, Mérida-Quesada, Vergara-Valladares, Arrey-Salas and Rubio-Meléndez2024):

(1) $$\begin{align}{J}_{K, KC}^{ss}&=\frac{I_{Hmax}}{e_0}\cdotp {g}_{KC}\times \left({V}_{ss}-{E}_K^{ss}\right)\end{align}$$
(2) $$\begin{align}{J}_{K, HKs}^{ss}=\frac{I_{Hmax}}{e_0}\times {g}_{HKs}\times \left(2\times {V}_{ss}-{E}_H^{ss}-{E}_K^{ss}\right)\end{align}$$
(3) $$\begin{align}{J}_{H, HKs}^{ss}=\frac{I_{Hmax}}{e_0}\times {g}_{HKs}\times \left(2\times {V}_{ss}-{E}_H^{ss}-{E}_K^{ss}\right)\end{align}$$
(4) $$\begin{align}{J}_{K, HKa}^{ss}=-\frac{I_{Hmax}}{e_0}\times {g}_{HKa}\times \left({E}_K^{ss}-{E}_H^{ss}\right)\end{align}$$
(5) $$\begin{align}{J}_{H, HKa}^{ss}=\frac{I_{Hmax}}{e_0}\times {g}_{HKa}\times \left({E}_K^{ss}-{E}_H^{ss}\right)\end{align}$$
(6) $$\begin{align}{J}_{HATPase}^{ss}=\frac{I_{Hmax}}{e_0}\times {i}_p\left({V}_{ss}\right)\end{align}$$

with

(7) $$\begin{align}{i}_p\left({V}_{ss}\right)=\frac{1-{e}^{-\left(\frac{F}{RT}\times \left({V}_{ss}-{V}_{0, pump}\right)\right)}}{1+{e}^{-\left(\frac{F}{RT}\times {V}_{ss}+5.4\right)}+{e}^{-\left(0.1\times \frac{F}{RT}\times {V}_{ss}+2.5\right)}}\end{align}$$
(8) $$\begin{align}&{V}_{ss}-{E}_H^{ss}+{i}_p\left({V}_{ss}\right)\times \frac{g_{KC}+{g}_{HKs}+{g}_{HKa}}{4\times {g}_{HKs}\times {g}_{HKa}+{g}_{KC}\times {g}_{HKs}+{g}_{KC}\times {g}_{HKa}}=0\end{align}$$
(9) $$\begin{align}{E}_K^{ss}={E}_H^{ss}-{i}_p\left({V}_{ss}\right)\times \frac{g_{KC}+2\times {g}_{HKs}}{4\times {g}_{HKs}\times {g}_{HKa}+{g}_{KC}\times {g}_{HKs}+{g}_{KC}\times {g}_{HKa}}\end{align}$$

and

(10) $$\begin{align}{J}_K^{ss}&={J}_{K, KC}^{ss}+{J}_{K, HKs}^{ss}+{J}_{K, HKa}^{ss}=0\end{align}$$
(11) $$\begin{align}{J}_H^{ss}={J}_{H, HKs}^{ss}+{J}_{H, HKa}^{ss}+{J}_{HATPase}^{ss}=0\end{align}$$

2.1. Perturbation of the system

To get an idea about the dynamics of the homeostat, we perturbed the system starting from the stable steady state and introduced perturbations in the electrochemical gradient, that is, in the membrane voltage ( $\Delta V=V-{V}_{ss}$ ), the potassium gradient ( $\Delta {E}_K={E}_K-{E}_K^{ss}$ ) and/or the proton gradient ( $\Delta {E}_H={E}_H-{E}_H^{ss}$ ). These perturbations change the driving forces of transport through each transporter and therefore affect the fluxes. In case of a linear dependence of the flux J on the driving force Δμ, the nonsteady-state fluxes are given by (Supplementary Text S1):

(12) $$\begin{align}{J}_{K, KC}&={J}_{K, KC}^{ss}+\frac{I_{Hmax}}{e_0}\times {g}_{KC}\times \left(\Delta V-\Delta {E}_K\right)\end{align}$$
(13) $$\begin{align}{J}_{K, HKs}={J}_{K, HKs}^{ss}+\frac{I_{Hmax}}{e_0}\times {g}_{HKs}\times \left(2\times \Delta V-\Delta {E}_H-\Delta {E}_K\right)\end{align}$$
(14) $$\begin{align}{J}_{H, HKs}={J}_{H, HKs}^{ss}+\frac{I_{Hmax}}{e_0}\times {g}_{HKs}\times \left(2\times \Delta V-\Delta {E}_H-\Delta {E}_K\right)\end{align}$$
(15) $$\begin{align}{J}_{K, HKa}={J}_{K, HKa}^{ss}-\frac{I_{Hmax}}{e_0}\times {g}_{HKa}\times \left(\Delta {E}_K-\Delta {E}_H\right)\end{align}$$
(16) $$\begin{align}{J}_{H, HKa}={J}_{H, HKa}^{ss}+\frac{I_{Hmax}}{e_0}\times {g}_{HKa}\times \left(\Delta {E}_K-\Delta {E}_H\right)\end{align}$$

The altered fluxes in turn cause changes in [K+ ] out , [K+ ] in , [H+ ] out , [H+ ] in , and the membrane voltage V. The changes in K+ concentrations are governed by the K+ flux:

(17) $$\begin{align}\frac{d{\left[{K}^{+}\right]}_{in}}{dt}&=\frac{-1}{Vol_{in}\cdot {N}_A}\times \left({J}_{K, KC}+{J}_{K, HKs}+{J}_{K, HKa}\right)\end{align}$$
(18) $$\begin{align}\frac{d{\left[{K}^{+}\right]}_{out}}{dt}=\frac{1}{Vol_{out}\cdot {N}_A}\times \left({J}_{K, KC}+{J}_{K, HKs}+{J}_{K, HKa}\right)\end{align}$$

Here, NA is the Avogadro constant, and Volin and Volout are the volumes of the compartments on the respective side of the membrane. Changes in proton concentrations are determined by the H+ flux and the buffer reactions:

(19) $$\begin{align}\frac{d{\left[{H}^{+}\right]}_{in}}{dt}&=\frac{-1}{Vol_{in}\cdot {N}_A} \times \left[{J}_{HATPase}+{J}_{H, HKs}+{J}_{H, HKa}-{k}_{v, in}\right.\nonumber\\ &\quad\left. \times \left({\left[{H}^{+}\right]}_{in}\times {\left[{Buf}^{-}\right]}_{in}-{K}_{s, in}\times {\left[ HBuf\right]}_{in}\right)\right]\\\end{align}$$
(20) $$\begin{align}\frac{d{\left[{H}^{+}\right]}_{out}}{dt}&=\frac{1}{Vol_{out}\cdot {N}_A}\times \left[{J}_{HATPase}+{J}_{H, HKs}+{J}_{H, HKa}-{k}_{v, out}\right.\nonumber\\ &\quad\left. \times \left({\left[{H}^{+}\right]}_{out}\times {\left[{Buf}^{-}\right]}_{out}-{K}_{s, out}\times {\left[ HBuf\right]}_{out}\right)\right]\end{align}$$

${K}_{s, in}={10}^{-{pK}_{s, in}}$ and ${K}_{s, out}={10}^{-{pK}_{s, out}}$ indicate the dissociation constants of the buffer reactions that buffer pHin and pHout to pKs,in and pKs,out , respectively. The buffer concentrations, in turn, change according to:

(21) $$\begin{align}\frac{d{\left[{Buf}^{-}\right]}_{in}}{dt}&=-\frac{d{\left[ HBuf\right]}_{in}}{dt}=\frac{-1}{Vol_{in}\cdot {N}_A}\nonumber\\ &\quad\times {k}_{v, in}\times \left({\left[{H}^{+}\right]}_{in}\times {\left[{Buf}^{-}\right]}_{in}-{K}_{s, in}\times {\left[ HBuf\right]}_{in}\right)\end{align}$$
(22) $$\begin{align}\frac{d{\left[{Buf}^{-}\right]}_{out}}{dt}&=-\frac{d{\left[ HBuf\right]}_{out}}{dt}=\frac{-1}{Vol_{out}\times {N}_A}\times {k}_{v, out}\nonumber\\ &\quad\times \left({\left[{H}^{+}\right]}_{out}\times {\left[{Buf}^{-}\right]}_{out}-{K}_{s, out}\times {\left[ HBuf\right]}_{out}\right)\end{align}$$

Charge transport from one side to the other is a net current

(23) $$\begin{align}I&={e}_0\times \left({J}_H+{J}_K\right)\nonumber\\ &={e}_0\times \left({J}_{HATPase}+{J}_{K, KC}+{J}_{K, HKs}+{J}_{H, HKs}+{J}_{K, HKa}+{J}_{H, HKa}\right)\end{align}$$

that changes the membrane voltage according to:

(24) $$\begin{align}\frac{dV}{dt}=\frac{-I}{C}=\frac{-{e}_0\times \left({J}_H+{J}_K\right)}{C}\end{align}$$

with the membrane capacitance C (unit F).

2.2. Kinetic changes

Changes in the parameters gX are usually not instantaneous and follow chemical reaction kinetics (Dreyer, Reference Dreyer2017; Dreyer et al., Reference Dreyer, Müller-Röber, Köhler and Blatt2004). To illustrate the effects, we modeled the time course of the change from ${g}_X^{start}$ to ${g}_X^{end}$ in the simplest way with mono-exponential relaxation kinetics:

(25) $$\begin{align}{g}_X(t)={g}_X^{end}+\left({g}_X^{start}-{g}_X^{end}\right)\times {e}^{-t/\tau }\end{align}$$

with the characteristic time constant $\tau$ that represents the particular features of the transporter, e.g. the relaxation of stretch-activated channels upon a change in the transmembrane pressure difference (Maksaev & Haswell, Reference Maksaev and Haswell2012).

The parameter gX often depends on other variable parameters, like the membrane voltage V, for example. In such a case, a change in V not only leads to the resulting adjustments of the homeostat but also to an additional change in gX with the subsequent consequences. Nevertheless, the change in gX and its effects are usually at least one order of magnitude slower (100 ms to 1 s) than the effects of changes in V (10 to 100 ms). To better understand the different contributions of the various parameters, we simulated the changes in the individual parameters in the model scenarios separately.

2.3. Computational simulation of the transporter system

To illustrate the dynamic behavior of the K homeostat using concrete examples, the system of differential equations was mathematically simulated using VCell Modeling and Analysis Software developed by the National Resource for Cell Analysis and Modeling, University of Connecticut Health Center (Schaff et al., Reference Schaff, Fink, Slepohenko, Carson and Loew1997). The different parameters are detailed in the text and the figures.

3. Results

The transporter network of the K homeostat establishes stable steady-state conditions for the membrane voltage V and the K+ gradient EK . Their values are determined by the activities of the different K+ transporters (gKC , gHKs and gHKa ) relative to the activity of the H+-ATPase [equations (8) and (9); (Dreyer et al., Reference Dreyer, Hernández-Rojas, Bolua-Hernández, Tapia-Castillo, Astola-Mariscal, Díaz-Pico, Mérida-Quesada, Vergara-Valladares, Arrey-Salas and Rubio-Meléndez2024)]. Any change in the gX values, an external voltage stimulus, ΔV, or a change in the transmembrane K+EK ) or H+ gradient (ΔEH ) is a disturbance of the steady state condition and causes a reaction of the transporter system. In the following, we will systematically analyze these disturbances and their consequences. It should be noted that changes in one parameter often result in changes in other parameters. For example, a change in membrane voltage also influences the activity of voltage-dependent transporters (Dreyer et al., Reference Dreyer, Sussmilch, Fukushima, Riadi, Becker, Schultz and Hedrich2021) or a change in the K+ concentration affects the activity of concentration-dependent transporters (Maierhofer et al., Reference Maierhofer, Scherzer, Carpaneto, Müller, Pardo, Hänelt, Geiger and Hedrich2024). Such connections and feedback loops make the interpretation of biological systems considerably more difficult, as it is no longer possible to clearly distinguish between cause and effect. The model scenarios presented in our theoretical analyses ultimately aim to convey a deeper understanding of the individual relationships between individual parameters and homeostat behavior. Against this background, it is sometimes a didactic necessity to initially disregard possible feedback effects.

3.1. Voltage changes

At first, we considered in a concrete scenario an electrical signal that significantly changed the membrane voltage (by $\Delta V(0)=V(0)-{V}_{ss}$ ) but not the K+ gradient ( $\Delta {E}_K=0$ ). Such voltage changes have their physiological equivalent in action potentials (Brownlee, Reference Brownlee2022; Cuin et al., Reference Cuin, Dreyer and Michard2018; Hedrich et al., Reference Hedrich, Salvador-Recatalà and Dreyer2016; Scherzer et al., Reference Scherzer, Böhm, Huang, Iosip, Kreuzer, Becker, Heckmann, Al-Rasheid, Dreyer and Hedrich2022) but are also induced in optogenetic experiments (Huang et al., Reference Huang, Ding, Roelfsema, Dreyer, Scherzer, Al-Rasheid, Gao, Nagel, Hedrich and Konrad2021; Papanatsiou et al., Reference Papanatsiou, Petersen, Henderson, Wang, Christie and Blatt2019; Reyer et al., Reference Reyer, Häßler, Scherzer, Huang, Pedersen, KAS, Bamberg, Palmgren, Dreyer and Nagel2020). For illustration, we chose a physiological model scenario, in which a steady state has been established at ${V}_{ss}=-150\; mV$ , ${E}_K^{ss}=-115\; mV$ ([K+]in = 150 mM, [K+]out = 1 mM), and ${E}_H^{ss}=92.1\; mV$ (pHin = 7.2, pHout = 5.6). This could be achieved with the parameter set gKC = 8 V-1, gHKs = 0.02 V-1, and gHKa = 1.378 V-1. Together with an IHmax = 50 pA and a membrane surface of 1000 μm2 (C = 10 pF), the respective H+ and K+ current densities were similar to those observed in guard cells (Dietrich et al., Reference Dietrich, Dreyer, Wiesner and Hedrich1998; Lohse & Hedrich, Reference Lohse and Hedrich1992). This system was allowed to equilibrate in its steady state and then challenged with a very short (1 ms) voltage stimulus in the range of $\Delta V=\pm 50\; mV$ (Figure 1a,e). Such a change in the membrane voltage altered the driving forces for K+ and H+, which caused a net current (Figure 1b,f), a net K+ flux (Figure 1c,g) and a net H+ flux (Figure 1d,h; Supplementary Text S2) via the different transporters. A hyperpolarization (ΔV < 0; Figure 1ad) provoked a negative net current and a K+ and H+ influx, while a depolarization (ΔV > 0; Figure 1eh) provoked a positive current and effluxes. In this scenario, we have excluded any dependence of the transporter activities on the membrane voltage to clearly work out the effect of the voltage change on the driving forces. The effects of changes in transporter activity that would superimpose the consequences presented here are discussed below.

Figure 1. Response of the K homeostat to a very rapid voltage pulse (ΔV). The K homeostat in steady state is challenged at t = 0 s by a rapid (1 ms) hyperpolarization (a–d) or depolarization (e–h) and is then left to its own devices again. The induced net currents (b, f) cause the membrane voltage to reset over time (a, e). The net currents are carried by net fluxes of K+ (JK ; c, g) and H+ (JH ; d, h) across the membrane.

The net charge transport had a rapid effect on the membrane voltage, with a tendency to restore the original voltage within tens of milliseconds, that is, $\Delta V(t)$ and the ion fluxes varied with time. The time until the steady state (dV/dt = 0) was reached, varied with the K-channel activity. With higher activity, the reset time was shorter than with low activity. For comparison, we chose a second scenario, in which the K-channel activity was 10-fold lower than in the first scenario but with the same steady state condition ( ${V}_{ss}=-150\; mV$ , ${E}_K^{ss}=-115\; mV$ ). This was achieved with gKC = 0.8 V-1, gHKs = 0.474 V-1, and gHKa = 0.77 V-1 (Supplementary Figure S2). The slightly longer relaxation time in the second scenario was mainly due to the lower net current, which restored the membrane voltage. As expected for a system with fewer K channels, the K+ flux (JK ; Supplementary Figure S2c and g) in particular was reduced and was only partially compensated by a larger H+ flux (JH ; Supplementary Figure S2d and h).

Irrespective of the scenario, the K homeostat showed a very fundamental short-term behavior of homeostats in general: a voltage change ΔV induced a net current that was proportional to the voltage change (Figure 2) and which counteracted the deflection from a steady state. The positive current induced by a depolarization (ΔV > 0) drove the membrane voltage in the negative direction while the negative current induced by a hyperpolarization (ΔV < 0) caused a positive shift of the voltage. Thus, the dynamics of the homeostat stabilized the membrane voltage at the steady state value. The more K channels were active in the K homeostat, the stronger was this autoregulative feedback. Current–voltage (I/V) curves of the type shown in Figure 2 are not limited to K homeostats, but a characteristic feature of all types of homeostats. Actually, they are the fundamental basis for setting the membrane voltage in the steady state.

Figure 2. Fundamental response of a homeostat to a voltage stimulus. A voltage stimulus (ΔV) induces a net current, which in turn causes a negative restoring force for the membrane voltage. The system is stable at ΔV = 0.

The K+ and H+ fluxes induced by the voltage stimulus were in the range of hundreds of attomols (10−18 mols) per second. Over the time course of the relaxation process, they summed up to a transmembrane net transfer of up to 1–4 attomols (Supplementary Figure S3). To estimate the potential consequences of these at first sight extremely tiny fluxes, it needed to be put in relation to the cellular volume. Considering in our model scenarios a spherical cell with C = 10 pF (cell surface = 1000 μm2), the internal volume would be ~3 pL (3000 μm3). Considering further that the vacuole could occupy about 90% of the internal cellular space, we calculated with a cytosolic volume of 0.3 – 3 pL. A mass transfer of 3 attomols into or out of such a volume would change the concentration by 1 μM (3 pL) or 10 μM (0.3 pL). For K+ with a concentration of [K+]in = 100 mM, such a change was insignificant. In contrast, for H+ with a resting concentration of [H+]in = 0.063 μM (pH 7.2) such a change would have implied an alteration of 1500%–15,000% in the absence of any pH buffer system. So far, the cellular simulations were carried out with a potent pH buffer ([Buf-] = [HBuf] ≥ 10 mM). Under this condition, the H+ fluxes did not change pHin in the considered short time interval. However, when we repeated the simulations with 100-fold smaller buffer concentrations ([Buf-] = [HBuf] ~ 0.1 mM), the cytosolic pH was strongly affected (Figure 3). Although this low buffer capacity might not be physiological (Felle, Reference Felle2001), this example is intended to illustrate the consequences of small buffering capacities. A depolarization alkalinized the cytosol, while a hyperpolarization caused an acidification. The magnitude of the pH changes was different in the two homeostat scenarios. It was more pronounced in scenario 2 with fewer K-channel activity than in scenario 1 because the smaller K+ fluxes were partially compensated by larger H+ fluxes in the second scenario. Thus, a homeostat has a resetting/stabilizing effect on a disturbance of the membrane voltage. In addition, it could in principle convert a voltage signal into a pHcyt signal. How sensitive this conversion is, depends on the buffer capacity and on the composition of the homeostat. Interestingly, also the feature of converting an electrical into a chemical signal could be generalized to other homeostats. For example, a homeostat with Ca2+-permeable transporters could convert a voltage signal into a cytosolic Ca2+ signal. Resting [Ca2+]cyt – like [H+]cyt – is very low and can therefore be quickly changed by fluxes in the range of tens to hundreds of attomoles per second. The time course of the conversion of the electrical signal into a chemical signal depends strongly on the respective chemical buffer capacity of the cell. The lower the buffer capacity, the more sensitive is the conversion process.

Figure 3. Buffer-dependent changes of the internal proton concentration in response to a voltage stimulus. Hyperpolarization (a) or depolarization (b) provoke H+ fluxes that may significantly modify the internal pH (pHin). The magnitude of this pH change depends on the buffer capacity of the cell and the composition of the homeostat. Compared are cells with a strong (≥10 mM; black curves) and a very weak (0.1 mM) buffer capacity, as well as a K homeostat with high () and one with low () K-channel activity.

3.2. Change in K + concentration

Next, we considered in our model scenarios a sudden change in the concentrations of the permeating ions, that is, in the case of the K homeostat K+ (this section) and H+ (next section). Also these changes affected the driving forces of transport. Although a slower change is more in line with physiological reality, it is by analyzing abrupt changes that one can learn the most about the dynamics of the system. First of all, we changed [K+]out, which affected the K+ gradient ([K+]out increase → $\Delta {E}_K>0$ ; [K+]out decrease → $\Delta {E}_K<0$ ). When we started with the K homeostat in steady state conditions and then doubled or halved [K+]out we induced a $\Delta {E}_K=\pm 17.3 mV$ . The cellular system responded to such a disturbance in two temporally separated phases (Supplementary Text S3). In a first phase of a few tens of milliseconds, the membrane voltage quickly reached a new, transient peak (i.e. dV/dt ≈ 0; Figure 4a), while in the second phase, a stable, long-lasting H+/K+ antiport occurred (Figure 4c,f), which could theoretically last for several tens of minutes (up to hours) with a correspondingly strong pH buffer. To illustrate this process free from secondary pH-effects, we clamped in our simulations [H+]in and [H+]out. The effects of a limited buffer capacity are illustrated further below. The condition dV/dt = 0 implied JK = −JH , i.e. an electroneutral H+/K+ antiport [equation (24)]. Thus, in the second phase, after equalization of the membrane voltage, the K homeostat mediated an exchange of K+ for H+, which slowly changed the potassium and proton concentrations (in our simulations clamped) until also here a new steady was established. Exactly, such a H+/K+ antiport has been observed experimentally in guard cells and mesophyll cells (Li et al., Reference Li, Grauschopf, Hedrich, Dreyer and Konrad2024), which underpins the correctness of the theoretical considerations. It should be noted that the H+/K+ antiport occurred regardless of whether the K homeostat contained antiporters or not. Further inspection revealed that the H+/K+ antiport was not exactly electroneutral, which is illustrated for a positive ΔEK in the following. After the first depolarization caused by a decaying negative net current with an amplitude in the pA-range (Figure 4b), in the following, the net current was not exactly zero but slightly positive with an amplitude in the sub-fA-range (Figure 4e, inset), that is, the H+-efflux was slightly larger in amplitude than the K+-influx. This positive net current led to a very slow return of the membrane voltage to the stable steady state at ${V}_{ss}=-150\; mV$ (Figure 4d). The negative K+ flux (JK , Figure 4c,f) caused an increase of [K+]in with the tendency of $\Delta {E}_K\to 0$ and restauration of the original ${E}_K^{ss}$ , that is, the larger [K+]out was compensated by a larger [K+]in. When we repeated the simulations with the K homeostat of the second scenario (fewer K channels), we observed essentially a similar behavior (Supplementary Figure S4a–f). Just the amplitudes of the transient voltage change, the net currents and the H+ and K+-fluxes were smaller indicating that in this case the K homeostat of scenario 1 was more sensitive than that of scenario 2. Qualitatively identical results, but reversed in sign, were obtained when a $\Delta {E}_K=-17.3\; mV$ (reduction of [K+]out) was applied (Supplementary Figure S4g–l).

Figure 4. Response of the K homeostat to a rapid change in the K+ gradient (ΔEK ). The K homeostat in steady state is challenged at t = 0 s by a doubling of [K+]outEK > 0). The induced net currents (b, e) cause the membrane voltage to reset over time (a, d). The net currents are carried by net fluxes of K+ (JK , green) and H+ (JH , blue; c, f) across the membrane. Please note the different answers in the millisecond (short-term response, a–c) and minutes time scale (long-term response, d–f).

To illustrate the effect of the pH buffer, we repeated the experiment shown in Figure 4 (clamped pHin, i.e., infinite buffer capacity), where [K+]out has been doubled at t = 0, with finite buffer capacities (100 mM, 50 mM, and 10 mM; Kurkdjian & Guern, Reference Kurkdjian and Guern1989) (Supplementary Figure S5). When the buffer capacity had no limit (Figure 4), the H+/K+ exchange proceeded until also [K+]in had doubled, the K+ gradient was as before (ΔEK = 0), and the original membrane voltage was restored (ΔV = 0, i.e., V = −150 mV; Supplementary Figure S5, black curves). However, when the buffer capacity reached its limits, the efflux of protons could only be partially compensated and the internal pH value increased (Supplementary Figure S5e; gray curves). This in turn changed the transmembrane proton gradient (EH ) and caused the H+/K+ exchange to terminate earlier (Supplementary Figure S5b and d), so that neither the K+ gradient (Supplementary Figure S5c) nor the membrane voltage (Supplementary Figure S5a) was fully restored. Instead, a new pHin-dependent steady-state condition was established. Thus, a homeostat can convert a concentration gradient signal into a voltage and a pHin signal. The sensitivity of this conversion depends on the pH buffer capacity and the composition of the homeostat.

3.3. Change in H + concentrations

Our considerations have shown so far that a K homeostat could process (i) a voltage signal into a pHin signal and (ii) a change in the K+ gradient into a voltage and a pHin signal. As next we considered a change in the proton gradient (ΔEH ). We could have simulated similar scenarios as shown before for ΔV (Figures 1– 3 and Supplementary Figures S2 and S3) or ΔEK (Figure 4 and Supplementary Figures S4 and S5). However, closer mathematical inspection revealed that the scenario of ΔEH could be explained by a combination of the previously considered perturbations (ΔV and ΔEK ). Because the steady state of the K homeostat ( ${V}_{ss},{E}_K^{ss}$ ) depends on EH [equations (8, 9)], a change in this value sets a new steady-state condition of the homeostat ( ${V}_{ss, new}$ , ${E}_K^{ss, new}$ ). Consequently, the homeostat, which was originally in dynamic equilibrium ( ${V}_{ss, old}$ , ${E}_K^{ss, old}$ ), was set by ΔEH at t = 0 to a state corresponding to a deflection from its new dynamic equilibrium. After the change, the situation of the homeostat could therefore also be explained by $\Delta V={V}_{ss, old}-{V}_{ss, new}$ and ${\Delta E}_K={E}_K^{ss, old}-{E}_K^{ss, new}$ , causing responses such as those described in the previous two sections. In a rapid reaction, the membrane voltage and, under very weak buffer conditions, also pHin were adjusted within tens of milliseconds to a temporary steady state. In the following, the net K+ and H+ fluxes as well as the tiny, but non-zero net current caused by the concentration disequilibrium changed then slowly [K+]in, [H+]in (depending on the buffer capacity) and V until a new steady state was reached (compare Figures 1– 4 and Supplementary Figures S2–S5). Thus, from a mechanistic point of view, the cases of ΔEH and ΔEK did not differ.

3.4. Changes in transporter activities (parameters g X )

Finally, we considered what happened if internal or external stimuli modified the activities of the membrane transporters of the homeostat. Physiologically, this happens very frequently, for example, when expression levels change or when posttranslational modifications occur. However, physical parameters such as voltage or hydraulic pressure can also be regulators of the transporter proteins.

As in the previous case, we were able to gain important insights by taking a closer look at the mathematical representation of the homeostat. When being in a steady state, the values ( ${V}_{ss},$ ${E}_K^{ss}$ ) were determined according to equations (8, 9) by the activity of the proton ATPase, the K+ channel, the H+/K+ symporter and the H+/K+ antiporter, represented by (gKC , gHKs , gHKa ). A change in any of the gX implied a new steady state condition, which was determined by the altered parameter set. The current condition ( ${V}_{ss, old}$ , ${E}_K^{ss, old}$ ) could thus be interpreted, as in the case of ΔEH , as a deviation by ΔV and ΔEK from this new steady state ( ${V}_{ss, new}$ , ${E}_K^{ss, new}$ , Supplementary Text S4).

Changes in gX might have been caused by an altered expression level, protein turnover, or posttranslational modifications, for instance. These processes are usually rather slow (minutes to hours) suggesting that in these cases also ( ${V}_{ss, new}$ , ${E}_K^{ss, new}$ ) are slowly adjusted. But they could also be faster as in the case of wounding when a hydraulic wave is assumed to propagate quickly within the tissues and modifies the activity of tension-regulated channels/transporters (Moe-Lange et al., Reference Moe-Lange, Gappel, Machado, Wudick, C. S. A., Schott-Verdugo, Bonus, Mishra, Hartwig and Bezrutczyk2021). To classify the consequences of such quicker processes, we exemplarily altered the parameter of the K channel, gKC , in scenarios 1 and 2 considered before. In scenario 1, we decreased gKC by a factor of 10 to the value of gKC of scenario 2, and in scenario 2, we increased gKC by a factor of 10 to the gKC value of scenario 1. These alterations changed the steady-state values from ( ${V}_{ss, old}=-150\; mV$ , ${E}_K^{ss, old}=-115\; mV$ ) to ( ${V}_{ss, new1}=-168.3\; mV$ , ${E}_K^{ss, new1}=-7.3\; mV$ ) and ( ${V}_{ss, new2}=-149.2\; mV$ , ${E}_K^{ss, new2}=-141.5\; mV$ ), respectively. Instead of an abrupt change of gKC , we chose an exponential decay kinetics characterized by a time constant τ. We did this because ensembles of ion channels or transporters, which individually are stochastically acting objects, adapt their activity to external stimuli with exponential kinetics (Dreyer, Reference Dreyer2017; Dreyer et al., Reference Dreyer, Müller-Röber, Köhler and Blatt2004).

To understand the short-term effects of the gKC changes on the membrane voltage, we first considered the consequences of the equivalent changes in ΔV and ΔEK (Supplementary Text S3). A decrease of gKC caused a ${V}_{ss, new1}=-168.3\; mV$ more negative than the original ${V}_{ss, old}=-150\; mV$ , that is, $\Delta V={V}_{ss, old}-{V}_{ss, new1}>0$ , and a ${E}_K^{ss, new1}=-7.3\; mV$ more positive than the original ${E}_K^{ss, old}=-115\; mV$ , that is, $\Delta {E}_K={E}_K^{ss, old}-{E}_K^{ss, new1}<0$ . Both, $\Delta V>0$ and $\Delta {E}_K<0$ provoked a positive net current (Figures 1f, 2 and Supplementary Figure S4h and k) that hyperpolarized the membrane toward a temporary minimum. The time course of this hyperpolarization depended on τ, the reaction time constant of the channel ensemble (Figure 5a). With τ = 100 ms, the first rapid adjustment of the voltage took less than a second, while with a nonunrealistic τ = 1s (Brüggemann et al., Reference Brüggemann, Dietrich, Dreyer and Hedrich1999; Dietrich et al., Reference Dietrich, Dreyer, Wiesner and Hedrich1998; Maksaev & Haswell, Reference Maksaev and Haswell2012) it could take around 10 s. Please note that the change in gKC provoked a ${\Delta E}_K<0$ , which caused a temporary membrane voltage peak (−182 mV, Figure 5a) that was negative of the new final steady state ( ${V}_{ss, new1}=-168.3\; mV$ ), a phenomenon observed already before (Supplementary Figure S4g and j). This imbalance would equilibrate in the slow phase of the adjustment (tens of minutes) that is dominated by a change of the concentrations due to the net fluxes of K+ and H+ (compare Supplementary Figure S4g–l).

Figure 5. Changes in the membrane voltage as a response to changes in gKC of the K homeostat. Shown is the first phase of the biphasic adjustment of the voltage. (a) In scenario 1, gKC started at t = 0 to decay exponentially with time constant τ toward the gKC value of scenario 2 (decay by factor 10). (b) In scenario 2, gKC started at t = 0 to increase with exponential saturation with time constant τ toward the gKC value of scenario 1 (increase by factor 10).

Similar conclusions as those illustrated for a decrease of gKC could also be obtained when increasing gKC , albeit with inverse signs. Compared to the starting conditions ( ${V}_{ss, old}=-150\; mV$ , ${E}_K^{ss, old}=-115\; mV$ ) the new steady state ${V}_{ss, new2}$ was more positive i.e. $\Delta V={V}_{ss, old}-{V}_{ss, new2}<0$ , and ${E}_K^{ss, new2}$ more negative than the starting value ${E}_K^{ss, old}$ , i.e. $\Delta {E}_K={E}_K^{ss, old}-{E}_K^{ss, new2}>0$ . Both deviations from the steady state caused a negative current (Figures 1b and 2, S4b and e) that triggered a temporary depolarization of the membrane voltage (Figure 5b). In the long term (tens of minutes), the membrane voltage would repolarize to the new steady state ( ${V}_{ss, new2}=-149.2\; mV$ ) alongside with an increase of [K+]in ( ${E}_K\to {E}_K^{ss, new2}$ ) due to the net K+ influx (Supplementary Figure S4c and f). The long-term fluxes were strongly influenced by the buffer capacity of the cell (compare Supplementary Figure S5), as a K+ influx was accompanied by an H+ efflux and a K+ efflux by an H+ influx. In summary, these considerations indicated that a change in the activity of an ion channel type involved in a homeostat is a strong signal that affects the membrane voltage, the pH conditions and in the long-term the ionic conditions of the cell.

4. Discussion and conclusion

Homeostats are networks of differentially energized transporters that transfer the same sort of ions/molecules across a membrane. A previous study argued on proper thermodynamic grounds that homeostats are the hidden rulers of homeostasis in plant cells (Dreyer et al., Reference Dreyer, Hernández-Rojas, Bolua-Hernández, Tapia-Castillo, Astola-Mariscal, Díaz-Pico, Mérida-Quesada, Vergara-Valladares, Arrey-Salas and Rubio-Meléndez2024). We have now analyzed their particular dynamic properties and propose that homeostats also play a key role in electrical, hydraulic and chemical signaling, such as pH and calcium signaling.

Exemplarily for the K homeostat we could show that homeostats equip the membrane with a self-regulatory, voltage-stabilizing property (Figure 2), in which a depolarization causes a positive net current that drives the membrane voltage in the negative direction, while a hyperpolarization provokes the opposite. Such a feature is not limited to K homeostats. In fact, it can be generalized. Such a characteristic curve has already been determined experimentally for a Ca2+ homeostat in vacuoles (Dindas et al., Reference Dindas, Dreyer, Huang, Hedrich and Roelfsema2021), indicating that also there several differentially energized transporter types involved in Ca2+ transport form a new functional unit. Also other types of homeostats, for example, anion homeostats (Dreyer, Reference Dreyer2021; Dreyer et al., Reference Dreyer, Li, Riedelsberger, Hedrich, Konrad and Michard2022; Geisler & Dreyer, Reference Geisler and Dreyer2024; Li et al., Reference Li, Grauschopf, Hedrich, Dreyer and Konrad2024), would show the same current–voltage characteristic. Related to this general feature is the rapid stabilization of the membrane voltage that takes tens of milliseconds (Figure 1). In the same time frame also abrupt changes in the K+ ( $\Delta {E}_K$ ) or the H+ gradient ( $\Delta {E}_H$ ) or changes in transporter activities ( $\Delta {g}_X$ ) affect the voltage (Figure 6a). While rapid voltage changes ( $\Delta V$ ) are an important feature in electrical signaling, e.g. by action potentials (Brownlee, Reference Brownlee2022; Cuin et al., Reference Cuin, Dreyer and Michard2018; Hedrich et al., Reference Hedrich, Salvador-Recatalà and Dreyer2016), abrupt changes of the other parameters ( $\Delta {E}_K$ , $\Delta {E}_H$ or $\Delta {g}_X$ ) are not necessarily realistic. Usually, these alterations take place more slowly. We nevertheless examined exemplarily instantaneous settings of $\Delta {E}_K$ (Figure 4 and Supplementary Figures S4 and S5) to illustrate the fast dynamics of the system and show that this case explains in principle also the features of rapid changes of $\Delta {E}_H$ or $\Delta {g}_X$ (Supplementary Text S2 and S4). In particular the latter (rapid $\Delta {g}_X$ ) may occur during the optogenic control of ion channels artificially introduced in plants (Christie & Zurbriggen, Reference Christie and Zurbriggen2021; Ding et al., Reference Ding, Zhou, Becker, Yang, Krischke, Scherzer, Yu-Strzelczyk, Mueller, Hedrich and Nagel2024; Huang et al., Reference Huang, Ding, Roelfsema, Dreyer, Scherzer, Al-Rasheid, Gao, Nagel, Hedrich and Konrad2021; Huang et al., Reference Huang, Shen, Roelfsema, Becker and Hedrich2023; Konrad et al., Reference Konrad, Gao, Zurbriggen and Nagel2023; Papanatsiou et al., Reference Papanatsiou, Petersen, Henderson, Wang, Christie and Blatt2019; Reyer et al., Reference Reyer, Häßler, Scherzer, Huang, Pedersen, KAS, Bamberg, Palmgren, Dreyer and Nagel2020). Our analyses may therefore form a theoretical basis for these experiments.

Figure 6. Schematic overview of the dynamic properties of the K homeostat. Changes in the membrane voltage ( $\Delta V$ ), the potassium gradient ( $\Delta {E}_K$ ), the proton gradient ( $\Delta {E}_H$ ), and the transporter activities ( $\Delta {g}_X$ ) influence each other. The effects manifest rapidly within tens of milliseconds (a) and slowly on a time scale of tens of minutes (b). The interactions shown with dashed arrows in (a) only occur if the buffer capacity is absent or low. In addition to the shown interactions, there might also be feedbacks on an intermediate time scale from $\Delta V$ , $\Delta {E}_K$ , and/or $\Delta {E}_H$ on $\Delta {g}_X$ that are not indicated; for example, in the case of voltage-gated ion channels, where $\Delta V$ feeds back on $\Delta {g}_X$ (Dreyer et al., Reference Dreyer, Müller-Röber, Köhler and Blatt2004).

In addition to the ad hoc changes, we also chose for $\Delta {g}_X$ a physiological scenario with an exponential time course. Membrane transporters such as ion channels behave stochastically in response to regulatory parameters. This manifests itself in exponential activation and deactivation time courses of the channel ensemble (Dietrich et al., Reference Dietrich, Dreyer, Wiesner and Hedrich1998; Dreyer, Reference Dreyer2017; Dreyer et al., Reference Dreyer, Müller-Röber, Köhler and Blatt2004). In order to exploit the detailed knowledge from the analysis of abrupt changes, these time courses can be considered to be subdivided into small time intervals with stepwise changes of ${g}_X$ . The membrane voltage adjusts within a few milliseconds to the condition determined by the actual ${g}_X$ value and thus follows the delayed time course of the $\Delta {g}_X$ change (Figure 5).

The scenario shown in Figure 5 is not limited to K homeostats and could explain in general terms the principle of basic sensing mechanisms in plants allowing them to react to a diverse set of physical and chemical stimuli, such as a hydraulic stimulus, or hormones and second messengers. The corresponding homeostat only needs to comprise one type of transporter that reacts sensitively to the respective stimulus. The stimulus alters the activity of the transporter and thus deflects the homeostat from its steady state. This in turn causes ΔV as the first rapid reaction, possibly followed also by pH (and Ca2+) signals.

The ion fluxes that modify the membrane voltage in the short term are generally not large enough to significantly change the cellular concentrations. With two exceptions: H+ and Ca2+. The cytosolic concentration of these two ion species is in the tens to hundreds of nanomolar range (Brownlee & Wheeler, Reference Brownlee and Wheeler2025; Felle, Reference Felle2001; Kader & Lindberg, Reference Kader and Lindberg2010; Köster et al., Reference Köster, DeFalco and Zipfel2022; Kudla et al., Reference Kudla, Becker, Grill, Hedrich, Hippler, Kummer, Parniske, Romeis and Schumacher2018; Luan & Wang, Reference Luan and Wang2021) and could therefore fundamentally change by the ion fluxes quantitatively estimated in this study (Supplementary Figure S3). These fluxes might be fully or partially absorbed by the cellular buffer systems (Brownlee & Wheeler, Reference Brownlee and Wheeler2025; Feng et al., Reference Feng, Fan, Miller and Xu2020; Kosegarten et al., Reference Kosegarten, Grolig, Wieneke, Wilson and Hoffmann1997; Niñoles et al., Reference Niñoles, Rubio, García-Sánchez, Fernández, Bueso, Alejandro and Serrano2013; Oja et al., Reference Oja, Savchenko, Jakob and Heber1999; Pfanz & Heber, Reference Pfanz and Heber1986; Wegner et al., Reference Wegner, Li, Zhang, Yu, Shabala and Hao2021; Zhou et al., Reference Zhou, Hao and Yang2021). Nevertheless, the buffer capacities for both ions are likely to be different, with the capacity for H+ presumably being higher than for Ca2+. Thus, we hypothesize that in homeostats involving calcium the Ca2+-response accompanied to a rapid voltage change is faster than the suppressed or delayed H+-response shown in this study (Figure 3). Although we have not analyzed Ca2+ homeostats here (but initial analyses were presented in Dreyer et al., Reference Dreyer, Li, Riedelsberger, Hedrich, Konrad and Michard2022), we can still state that fast signals (ΔV) are differentially converted to Ca2+ and H+ signals, as the larger pH buffer capacity of the cell can attenuate and delay the H+ signal. Thus, a similar translation mechanism (ΔV→pHcyt / ΔV→Ca2+ cyt) can nevertheless give rise to a multitude of finely tuned cellular signals.

In addition to the short-term responses, there is also a mutual long-term interplay between $\Delta V$ , $\Delta {E}_K$ , $\Delta {E}_H$ and $\Delta {g}_X$ that take place within tens of minutes (Figure 6b). These changes are mainly due to the slow change of concentrations caused by the net K+ and H+ fluxes (Figures 4f and Supplementary Figure S4f, 1 and S5). Interestingly, in these long-term fluxes the K homeostat mediates an almost electroneutral K+/H+ exchange even in the absence of functional H+/K+ antiporters. This feature of K homeostats has been observed experimentally already (Li et al., Reference Li, Grauschopf, Hedrich, Dreyer and Konrad2024). Thus, long-term uptake of potassium is always accompanied by an efflux of protons, which would quickly overload the chemical pH buffers of the cells if it is not compensated for in some other way. Here, anion (A) homeostats step in, which mediate an electroneutral A-/H+ symport (Li et al., Reference Li, Grauschopf, Hedrich, Dreyer and Konrad2024). Together, K and A homeostats form a flexible functional super-unit that not only manages the uptake of these ions but also participates in the pH-buffering process (see Graphical Abstract). Thus, the common view that Cl (or another anion like NO3) is the counterion of K+ guaranteeing electroneutrality is just half of the truth and does not acknowledge the flexibility in pH control by the two involved homeostats.

As indicated, the coupling of different homeostats represents a further level of dynamic complexity. Changes in the setting of one homeostat, that are limited to this homeostat at first glance (e.g. $\Delta {E}_K$ for the K homeostat) also affect all other connected homeostats, with far-reaching physiological implications. As an example, recently a coupling between the auxin and anion homeostats has been proposed as the reason why plants use two different auxin exporters, one of which even requires the energy from ATP hydrolysis for a downstream transport (Geisler & Dreyer, Reference Geisler and Dreyer2024). The different homeostats “communicate” with each other predominantly via the membrane voltage, the cytosolic pH, but also by Ca2+-regulated protein kinases (Dong et al., Reference Dong, Bai, Almutairi and Kudla2021; Hamel et al., Reference Hamel, Sheen and Séguin2014; Kudla et al., Reference Kudla, Becker, Grill, Hedrich, Hippler, Kummer, Parniske, Romeis and Schumacher2018; Wang et al., Reference Wang, Liu, Han, Liu, Sadeghnezhad and Liu2023). We are only just beginning to understand the grammar of this sophisticated language. The long-term simulations presented in this study should therefore not be overinterpreted as we have considered a homeostat in isolation. In real life, cellular countermeasures involving other homeostats and regulatory feedbacks will certainly avoid extreme values for [K+]in and pHin. In this context, it should also be noted that non or poorly selective transporters can link different types of homeostats. For example, if we consider the K homeostat from this work and add e.g. K+- (but also Na+-) permeable channels of the HKT type (Riedelsberger et al., Reference Riedelsberger, Miller, Valdebenito-Maturana, Piñeros, González and Dreyer2021) then the K homeostat is linked to the Na homeostat (HKT & SOS1/NHX; Gámez-Arjona et al., Reference Gámez-Arjona, Park, García, Aman, Villalta, Raddatz, Carranco, Ali, Ali and Zareen2024) and forms another new super-unit. All these new units act according to the basic rules shown in Figures 2 and 6. However, their particular characteristics and the question of how one ion can influence the homeostasis of the other still need to be explored in more detail in the future.

In summary, our analyses strongly suggest that it is not individual transporters but the dynamic properties of homeostats that underlie electrical, chemical, hydraulic, pH and calcium signaling in plants. Homeostats are universal building bricks (like LEGO) that then group into larger units to accomplish the complicated tasks of physiology.

Supplementary material

The supplementary material for this article can be found at http://doi.org/10.1017/qpb.2025.6.

Data availability statement

The authors confirm that the data supporting the findings of this study are available within the article and its supplementary materials. Raw data sets are available from the corresponding author, ID, upon reasonable request.

Author contributions

I.D. conceived the study. All authors contributed to the design of the study. I.D. conducted the data analyses. L.C.-A., T.R.-R., T.R.-G., and M.E.R.-M. participated in data visualization. All authors contributed to writing and editing of the manuscript and approved its final version.

Funding statement

This work was supported by the Agencia Nacional de Investigación y Desarrollo de Chile (ANID), grant No. Anillo-ANID ATE220043 (the multidisciplinary center for biotechnology and molecular biology for climate change adaptative in forest resources; CeBioClif) to I.D. and by Fondo Nacional de Desarrollo Científico, Tecnológico y de Innovación Tecnológica (FONDECYT/Chile), grant no. 11251254 to M.E.R.-M., no. 1210920 to E.M, no. 1220504 to I.D., no. 1240167 to J.R.

Competing interest

The authors declare none.

Open peer review

To view the open peer review materials for this article, please visit http://doi.org/10.1017/qpb.2025.6.

Footnotes

Leslie Contador-Álvarez, Tamara Rojas-Rocco, and Talía Rodríguez-Gómez these authors contributed equally.

Associate Editor: Richard Morris

References

Brownlee, C. (2022). Plant physiology: Anatomy of a plant action potential. Current Biology, 32, R1000R1002.CrossRefGoogle ScholarPubMed
Brownlee, C., & Wheeler, G. (2025). Cellular calcium homeostasis and regulation of its dynamic perturbation. Quantitative Plant Biology, 6, e5.CrossRefGoogle ScholarPubMed
Brüggemann, L., Dietrich, P., Dreyer, I., & Hedrich, R. (1999). Pronounced differences between the native K+ channels and KAT1 and KST1 alpha-subunit homomers of guard cells. Planta, 207, 370376.Google ScholarPubMed
Christie, J. M., & Zurbriggen, M. D. (2021). Optogenetics in plants. New Phytologist, 229, 31083115.CrossRefGoogle ScholarPubMed
Cuin, T. A., Dreyer, I., & Michard, E. (2018). The role of potassium channels in Arabidopsis thaliana long distance electrical signalling: AKT2 modulates tissue excitability while GORK shapes action potentials. International Journal of Molecular Sciences, 19, 926.CrossRefGoogle ScholarPubMed
Dietrich, P., Dreyer, I., Wiesner, P., & Hedrich, R. (1998). Cation sensitivity and kinetics of guard-cell potassium channels differ among species. Planta, 205, 277287.CrossRefGoogle Scholar
Dindas, J., Dreyer, I., Huang, S., Hedrich, R., & Roelfsema, M. R. G. (2021). A voltage-dependent Ca2+-homeostat operates in the plant vacuolar membrane. New Phytologist, 230, 14491460.CrossRefGoogle ScholarPubMed
Ding, M., Zhou, Y., Becker, D., Yang, S., Krischke, M., Scherzer, S., Yu-Strzelczyk, J., Mueller, M. J., Hedrich, R., Nagel, G., et al. (2024). Probing plant signal processing optogenetically by two channelrhodopsins. Nature, 633, 872877.CrossRefGoogle ScholarPubMed
Dong, Q., Bai, B., Almutairi, B. O., & Kudla, J. (2021). Emerging roles of the CBL-CIPK calcium signaling network as key regulatory hub in plant nutrition. Journal of Plant Physiology, 257, 153335.CrossRefGoogle ScholarPubMed
Dreyer, I. (2017). Plant potassium channels are in general dual affinity uptake systems. AIMS Biophysics, 4, 90106.CrossRefGoogle Scholar
Dreyer, I. (2021). Nutrient cycling is an important mechanism for homeostasis in plant cells. Plant Physiology, 187, 22462261.CrossRefGoogle Scholar
Dreyer, I., Hernández-Rojas, N., Bolua-Hernández, Y., Tapia-Castillo, V., Astola-Mariscal, S. Z., Díaz-Pico, E., Mérida-Quesada, F., Vergara-Valladares, F., Arrey-Salas, O., Rubio-Meléndez, M. E., et al. (2024). Homeostats: The hidden rulers of ion homeostasis in plants. Quantitative Plant Biology, 5, e8.CrossRefGoogle ScholarPubMed
Dreyer, I., Li, K., Riedelsberger, J., Hedrich, R., Konrad, K. R., & Michard, E. (2022). Transporter networks can serve plant cells as nutrient sensors and mimic transceptor-like behavior. iScience, 25, 104078.CrossRefGoogle ScholarPubMed
Dreyer, I., Müller-Röber, B., & Köhler, B. (2004). Voltage gated ion channels. In Blatt, M. R. (Ed.), Annual plant reviews, (Vol. 15, Membrane transport in Plants, pp. 150192). Blackwell Publishing.Google Scholar
Dreyer, I., Sussmilch, F. C., Fukushima, K., Riadi, G., Becker, D., Schultz, J., & Hedrich, R. (2021). How to grow a tree: Plant voltage-dependent cation channels in the spotlight of evolution. Trends in Plant Science, 26, 4152.CrossRefGoogle Scholar
Felle, H. H. (2001). pH: Signal and messenger in plant cells. Plant Biology, 3, 577591.CrossRefGoogle Scholar
Feng, H., Fan, X., Miller, A. J., & Xu, G. (2020). Plant nitrogen uptake and assimilation: Regulation of cellular pH homeostasis. Journal of Experimental Botany, 71, 43804392.CrossRefGoogle ScholarPubMed
Gámez-Arjona, F., Park, H. J., García, E., Aman, R., Villalta, I., Raddatz, N., Carranco, R., Ali, A., Ali, Z., Zareen, S., et al. (2024). Inverse regulation of SOS1 and HKT1 protein localization and stability by SOS3/CBL4 in Arabidopsis thaliana . Proceedings of the National Academy of Sciences of the United States of America, 121, e2320657121.CrossRefGoogle ScholarPubMed
Geisler, M., & Dreyer, I. (2024). An auxin homeostat allows plant cells to establish and control defined transmembrane auxin gradients. New Phytologist, 4, 14221436.CrossRefGoogle Scholar
Hamel, L. P., Sheen, J., & Séguin, A. (2014). Ancient signals: Comparative genomics of green plant CDPKs. Trends in Plant Science, 19, 7989.CrossRefGoogle ScholarPubMed
Hedrich, R., Salvador-Recatalà, V., & Dreyer, I. (2016). Electrical wiring and long-distance plant communication. Trends in Plant Science, 21, 376387.CrossRefGoogle ScholarPubMed
Huang, S., Ding, M., Roelfsema, M. R. G., Dreyer, I., Scherzer, S., Al-Rasheid, K. A. S., Gao, S., Nagel, G., Hedrich, R., & Konrad, K. R. (2021). Optogenetic control of the guard cell membrane potential and stomatal movement by the light-gated anion channel GtACR1. Science Advances, 7, eabg4619.CrossRefGoogle ScholarPubMed
Huang, S., Shen, L., Roelfsema, R. M. G., Becker, D., & Hedrich, R. (2023). Light-gated channelrhodopsin sparks proton-induced calcium release in guard cells. Science, 382, 13141319.CrossRefGoogle ScholarPubMed
Kader, M. A., & Lindberg, S. (2010). Cytosolic calcium and pH signaling in plants under salinity stress. Plant Signaling & Behavior, 5, 233.CrossRefGoogle ScholarPubMed
Konrad, K. R., Gao, S., Zurbriggen, M. D., & Nagel, G. (2023). Optogenetic methods in plant biology. Annual Review of Plant Biology, 74, 313339.CrossRefGoogle ScholarPubMed
Kosegarten, H., Grolig, F., Wieneke, J., Wilson, G., & Hoffmann, B. (1997). Differential ammonia-elicited changes of cytosolic pH in root hair cells of rice and maize as monitored by 2[prime],7[prime]-bis-(2-Carboxyethyl)-5 (and -6)-Carboxyfluorescein-fluorescence ratio. Plant Physiology, 113, 451461.CrossRefGoogle Scholar
Köster, P., DeFalco, T. A., & Zipfel, C. (2022). Ca2+ signals in plant immunity. EMBO Journal, 41, e110741.CrossRefGoogle ScholarPubMed
Kudla, J., Becker, D., Grill, E., Hedrich, R., Hippler, M., Kummer, U., Parniske, M., Romeis, T., & Schumacher, K. (2018). Advances and current challenges in calcium signaling. New Phytologist, 218, 414431.CrossRefGoogle ScholarPubMed
Kurkdjian, A., & Guern, J. (1989). Intracellular pH: Measurement and importance in cell activity. Annual Review of Plant Biology, 40, 271303.CrossRefGoogle Scholar
Li, K., Grauschopf, C., Hedrich, R., Dreyer, I., & Konrad, K. R. (2024). K+ and pH homeostasis in plant cells is controlled by a synchronized K+/H+ antiport at the plasma and vacuolar membrane. New Phytologist, 241, 15251542.CrossRefGoogle Scholar
Lohse, G., & Hedrich, R. (1992). Characterization of the plasma-membrane H+-ATPase from Vicia faba guard cells - modulation by extracellular factors and seasonal changes. Planta, 188, 206214.CrossRefGoogle ScholarPubMed
Luan, S., & Wang, C. (2021). Calcium signaling mechanisms across kingdoms. Annual Review of Cell and Developmental Biology, 37, 311340.CrossRefGoogle ScholarPubMed
Maierhofer, T., Scherzer, S., Carpaneto, A., Müller, T. D., Pardo, J. M., Hänelt, I., Geiger, D., & Hedrich, R. (2024). Arabidopsis HAK5 under low K+ availability operates as PMF powered high-affinity K+ transporter. Nature Communications, 15, 8558.CrossRefGoogle Scholar
Maksaev, G., & Haswell, E. S. (2012). MscS-like10 is a stretch-activated ion channel from Arabidopsis thaliana with a preference for anions. Proceedings of the National Academy of Sciences of the United States of America, 109, 1901519020.CrossRefGoogle ScholarPubMed
Moe-Lange, J., Gappel, N. M., Machado, M., Wudick, M. M., C. S. A., Sies, Schott-Verdugo, S. N., Bonus, M., Mishra, S., Hartwig, T., Bezrutczyk, M., et al (2021) Interdependence of a mechanosensitive anion channel and glutamate receptors in distal wound signaling. Science Advances, 7, eabg4298.CrossRefGoogle ScholarPubMed
Niñoles, R., Rubio, L., García-Sánchez, M. J., Fernández, J. A., Bueso, E., Alejandro, S., & Serrano, R. (2013). A dominant-negative form of Arabidopsis AP-3 β-adaptin improves intracellular pH homeostasis. The Plant Journal, 74, 557568.CrossRefGoogle ScholarPubMed
Oja, V., Savchenko, G., Jakob, B., & Heber, U. (1999). pH and buffer capacities of apoplastic and cytoplasmic cell compartments in leaves. Planta, 209, 239249.CrossRefGoogle ScholarPubMed
Papanatsiou, M., Petersen, J., Henderson, L., Wang, Y., Christie, J. M., & Blatt, M. R. (2019). Optogenetic manipulation of stomatal kinetics improves carbon assimilation, water use, and growth. Science, 363, 14561459.CrossRefGoogle ScholarPubMed
Pfanz, H., & Heber, U. (1986). Buffer capacities of leaves, leaf cells, and leaf cell organelles in relation to fluxes of potentially acidic gases. Plant Physiology, 81, 597602.CrossRefGoogle ScholarPubMed
Reyer, A., Häßler, M., Scherzer, S., Huang, S., Pedersen, J. T., KAS, A.-R., Bamberg, E., Palmgren, M., Dreyer, I., Nagel, G., et al. (2020). Channelrhodopsin-mediated optogenetics highlights a central role of depolarization-dependent plant proton pumps. Proceedings of the National Academy of Sciences of the United States of America, 117, 2092020925.CrossRefGoogle ScholarPubMed
Riedelsberger, J., Miller, J. K., Valdebenito-Maturana, B., Piñeros, M. A., González, W., & Dreyer, I. (2021). Plant HKT channels: An updated view on structure, function and gene regulation. International Journal of Molecular Sciences, 22, 1892.CrossRefGoogle ScholarPubMed
Schaff, J., Fink, C. C., Slepohenko, B., Carson, J. H., & Loew, L. M. (1997). A general computational framework for modeling cellular structure and function. Biophysical Journal, 73, 11351146.CrossRefGoogle ScholarPubMed
Scherzer, S., Böhm, J., Huang, S., Iosip, A. L., Kreuzer, I., Becker, D., Heckmann, M., Al-Rasheid, K. A. S., Dreyer, I., & Hedrich, R. (2022). A unique inventory of ion transporters poises the Venus flytrap to fast-propagating action potentials and calcium waves. Current Biology, 32, 42554263, e5.CrossRefGoogle ScholarPubMed
Serrano, R., & Rodriguez-Navarro, A. (2001). Ion homeostasis during salt stress in plants. Current Opinion in Cell Biology, 13, 399404.CrossRefGoogle ScholarPubMed
Wang, L., Liu, Z., Han, S., Liu, P., Sadeghnezhad, E., & Liu, M. (2023). Growth or survival: What is the role of calmodulin-like proteins in plant? International Journal of Biological Macromolecules, 242, 124733.CrossRefGoogle ScholarPubMed
Wegner, L. H., Li, X., Zhang, J., Yu, M., Shabala, S., & Hao, Z. (2021). Biochemical and biophysical pH clamp controlling net H+ efflux across the plasma membrane of plant cells. New Phytologist, 230, 408415.CrossRefGoogle ScholarPubMed
Wegner, L. H., Pottosin, I., Dreyer, I., Shabala, S. (2025) Potassium homeostasis and signalling: From the whole plant to the subcellular level. Quantitative Plant Biology, 6, eX.Google Scholar
Zhou, J. Y., Hao, D. L., & Yang, G. Z. (2021). Regulation of Cytosolic pH: The contributions of plant plasma membrane H+-ATPases and multiple transporters. International Journal of Molecular Sciences, 22, 12998.CrossRefGoogle Scholar
Figure 0

Figure 1. Response of the K homeostat to a very rapid voltage pulse (ΔV). The K homeostat in steady state is challenged at t = 0 s by a rapid (1 ms) hyperpolarization (a–d) or depolarization (e–h) and is then left to its own devices again. The induced net currents (b, f) cause the membrane voltage to reset over time (a, e). The net currents are carried by net fluxes of K+ (JK; c, g) and H+ (JH; d, h) across the membrane.

Figure 1

Figure 2. Fundamental response of a homeostat to a voltage stimulus. A voltage stimulus (ΔV) induces a net current, which in turn causes a negative restoring force for the membrane voltage. The system is stable at ΔV = 0.

Figure 2

Figure 3. Buffer-dependent changes of the internal proton concentration in response to a voltage stimulus. Hyperpolarization (a) or depolarization (b) provoke H+ fluxes that may significantly modify the internal pH (pHin). The magnitude of this pH change depends on the buffer capacity of the cell and the composition of the homeostat. Compared are cells with a strong (≥10 mM; black curves) and a very weak (0.1 mM) buffer capacity, as well as a K homeostat with high () and one with low () K-channel activity.

Figure 3

Figure 4. Response of the K homeostat to a rapid change in the K+ gradient (ΔEK). The K homeostat in steady state is challenged at t = 0 s by a doubling of [K+]outEK > 0). The induced net currents (b, e) cause the membrane voltage to reset over time (a, d). The net currents are carried by net fluxes of K+ (JK, green) and H+ (JH, blue; c, f) across the membrane. Please note the different answers in the millisecond (short-term response, a–c) and minutes time scale (long-term response, d–f).

Figure 4

Figure 5. Changes in the membrane voltage as a response to changes in gKC of the K homeostat. Shown is the first phase of the biphasic adjustment of the voltage. (a) In scenario 1, gKC started at t = 0 to decay exponentially with time constant τ toward the gKC value of scenario 2 (decay by factor 10). (b) In scenario 2, gKC started at t = 0 to increase with exponential saturation with time constant τ toward the gKC value of scenario 1 (increase by factor 10).

Figure 5

Figure 6. Schematic overview of the dynamic properties of the K homeostat. Changes in the membrane voltage ($\Delta V$), the potassium gradient ($\Delta {E}_K$), the proton gradient ($\Delta {E}_H$), and the transporter activities ($\Delta {g}_X$) influence each other. The effects manifest rapidly within tens of milliseconds (a) and slowly on a time scale of tens of minutes (b). The interactions shown with dashed arrows in (a) only occur if the buffer capacity is absent or low. In addition to the shown interactions, there might also be feedbacks on an intermediate time scale from $\Delta V$, $\Delta {E}_K$, and/or $\Delta {E}_H$ on $\Delta {g}_X$ that are not indicated; for example, in the case of voltage-gated ion channels, where $\Delta V$ feeds back on $\Delta {g}_X$ (Dreyer et al., 2004).

Supplementary material: File

Contador-Álvarez et al. supplementary material

Contador-Álvarez et al. supplementary material
Download Contador-Álvarez et al. supplementary material(File)
File 925.5 KB

Author comment: Dynamics of homeostats: the basis of electrical, chemical, hydraulic, pH and calcium signaling in plants — R0/PR1

Comments

Dear Editor,

enclosed please find the second part of our analyses of homeostats in plants. In the first part (DOI: 10.1017/qpb.2024.8), we provided a hands-on tutorial for modeling homeostats and anlyzed homeostats in steady state. Now, we went one step further and investigated the reaction of homeostats on disturbances of this steady state. In the first part, we concluded that homeostats are the hidden rulers of ion homeostasis in plants, and now in the second part, we have to conclude that homeostats provide also the basis of electrical, hydraulic, pH, and Ca signaling in plants. We consider these as important insights that we would like to share with the scientific community.

Looking forward for a positive response

Best regards

Ingo Dreyer

Review: Dynamics of homeostats: the basis of electrical, chemical, hydraulic, pH and calcium signaling in plants — R0/PR2

Conflict of interest statement

Reviewer declares none.

Comments

A logical continuation of the formulation of the model of a K+ homeostat in their previous studies is the testing of its behavior upon perturbations of membrane voltage, sudden changes of external K+ and H+ concentrations, and the turnover rate of individual transporters (here,as an example, ensemble conduction of a population of K+ channels, which could be, for instance, artificially evoked in an optogenetic experiment). Results and conclusions have a fundamental character and represent an important step in a construction of a “superhomeostat”, which combines diverse homeostat circuits.

I have few queries:

-An important conclusion is a “voltage-stabilizing property” of the K+ homeostat. No doubts, it should be true for the circuit which includes H+ pumps, H+/K+ symporters and antiporters, and K+-selective channels. But, what about nonselective cation channels, which hardly differentiate between K+ and Na+? Or, worse, with a low cation-anion preference? Experimentally, there are proofs for long (tens of minutes-days) lasting membrane depolarizations upon strong salt or oxidative stress. Apparently, K+ homeostat fails in this case. Should it be meaningful to specify these limitations, when a disturbance is beyond a capacity of a K+ homeostat?

- Lines 232-233, this very low level for pH buffer capacity seems unrealistic.

-Line 433. I would disagree. Intracellular buffer for Ca2+ (a plethora of Ca2+ binding molecules including proteins/ Ca2+ sensors), of course, exists, although in plants, up to my knowledge, there were no systematic studies to evaluate its capacity. In my opinion, its approximation may be a submillimolar value for the cytoplasm.

- Please, check, is it correct to use the term “equilibrium” in lines 313, 314, 420, when it is apparently speaking about steady state condition.

Review: Dynamics of homeostats: the basis of electrical, chemical, hydraulic, pH and calcium signaling in plants — R0/PR3

Conflict of interest statement

Reviewer declares none.

Comments

This theoretical study by Contador -Álvarez et al. is a follow-up work to a previous publication in the same journal dealing with the K+ homeostat (and elaborating on previous papers on this issue by the corresponding author, Ingo Dreyer). In that previous work the homeostat was introduced as a tool to analyse K+ and H+ concentrations and fluxes under steady state conditions. K+ and H+ homeostasis were shown to result from the interplay of a H+ ATPase, a K+ conductance and H+/K+ symport and antiport, also including H+ buffering. Here, the response to a perturbation of the system is analysed with the same computational tool, analysing the response to sudden changes in (i) membrane voltage, (ii) K+ concentration, and (iii) K+ membrane conductance. This is indeed a logical step forward. The simplicity of this approach proved to be a great advantage for extending our basic knowledge and understanding of membrane transport in plants, but it also comes with a severe limitation: The scenarios are rather artificial, since they refer to a cell with only K+ and H+ as permeable cations, and membrane-impermeable anions required for charge balance, a situation we won’t find in nature. The authors are aware of this limitation, since they state at the end of the discussion that “The long-term simulations presented in this study should therefore not been overinterpreted as we have considered a homeostat in isolation. In real life, cellular countermeasures involving other homeostats and regulatory feedbacks”. However, it is conceivable to design model systems with exactly these properties, e.g. oocytes co-expressing this set of transporters (and maybe inhibiting native anion channels); maybe this should still be pointed out explicitly.

I am sure the authors will explain to us in the near future how K+, H+ and Ca2+ signatures come about, and how they can be decoded by the cell.

I have a number of critical points to raise:

l. 62 …as an example

l. 172 some K+ is transported, so better say: effects on [K+] are negligible..

l. 170 onwards: The homeostat is used in this scenario to analyse the response of the membrane to a stepwise change in voltage at a timescale up to 100 ms. However, this leads us to a fundamental problem: It was not designed for that purpose, but rather, to describe a steady state. Hence, although it does include the kinetics of the buffer reactions adjusting pH, but it doesn’t include the kinetics of the K+ channel and the transporters which will respond with a delay to a stepwise change in voltage (and. likewise, K+). For an ion channel this delay will be in the range of maybe 10 to 100 ms, as numerous patch clamp experiments and analysis in oocytes after heterologous expression have shown, and for a H+ K+ cotransporter such as HAK5 I expect similar kinetics (though I didn’t find any data in the literature). As for a response to a change in K+ concentration, a recent publication by Maierhofer et al. (2024; Nature Communications 15.1; 8558) revealed that it takes even seconds until a new steady state is reached (refers to the next experiment). So the curves shown here are of little practical significance and rather show the properties of the algorithm used here; in the model applied to this scenario all transporters respond instantaneously to a change in parameters. The authors are in principal aware of that problem later saying that “Changes in the parameters gx are usually not instantaneous and follow chemical reaction kinetics” (l. 145), but they treat this issue separately instead of including it when considering changes in voltage and K+ concentration.

Insets to figures: In addition to membrane transport, circular fluxes are shown. The arrows are superimposed but don’t match, which makes the figure look messy and over-complicated.

l. 401: “…are not necessarily realistic…”

l. 411: …manifest themselves in group behaviours, leading to exponential activation and deactivation time courses…

l. 413 ff: This paragraph is a little awkward and needs rephrasing: e.g. we can subdivide the gp timecourse to infinitesimal stepwise changes with delta V following delta gp…

The scenario described here does not only apply to pressure sensing, but to any other (chemical) stimulus, such as hormone, second messengers…

l. 435 There is Ca2+ buffering as well! Bound Ca2+ is exceeding free Ca2+ by orders of magnitude.

l. 450 …long-term interplay between…

l. 455 It seems that H+ can serve as a balance between K+ and anion homeostat?!

Supplement:

Fig S5E: I really have a problem to understand the pH profiles. I would expect an s-shaped curve, with slow pH change at the beginning when the buffer capacity is at maximum and an increase in slope until the buffer is ‘broken’ and a rapid transition to a new steady state when the H+ flux ends.

Text S1, figure S7: mx and sx correspond to chord conductance and slope conductance. The latter has little significance, whereas the former is, from a thermodynamic standpoint, the relevant parameter. Hence, approximating the curve with a Taylor series is not of much practical use either.

Recommendation: Dynamics of homeostats: the basis of electrical, chemical, hydraulic, pH and calcium signaling in plants — R0/PR4

Comments

Dear Ingo et al,

Thank you for submitting your manuscript on your exciting work on homeostat dynamics to QPB. The reviewers and I are positive about this submission but feel the work could be strengthened by clarifying a number of points (see full reviews for some excellent comments and questions).

I look forward to receiving your revised manuscript.

With best wishes

Richard

Decision: Dynamics of homeostats: the basis of electrical, chemical, hydraulic, pH and calcium signaling in plants — R0/PR5

Comments

No accompanying comment.

Author comment: Dynamics of homeostats: the basis of electrical, chemical, hydraulic, pH and calcium signaling in plants — R1/PR6

Comments

Dear Richard and Olivier,

thank you very much for the assessment of our manuscript. All comments were very helpful to identify those parts of the MS that needed clarification and further explanation. Based on the comments we could elaborate a revised version of the MS. Please find attached:

*) the point-by-point answer to the comments of the reviewers/editors

*) the MS with tracked changes

Best regards

Ingo

P.S.: Confidential comment: There are two new references to manuscripts/papers that are still under review/revision in QPB. These MS were submitted within the frame of the Special Research Topic that Dale and I organize:

*) Brownlee & Wheeler: an excellent MS about Ca2+ in plants, for which I serve as editor. This citation fits perfectly to the issue of Ca-buffering that was addressed by both reviewers. For the MS “Brwonlee&Wheeler” only a very minor revision is needed. As soon as the authors submit their revision, it is expected to be accepted (without further review).

*) Wegner et al: the contribution about potassium in the RT.

Reason: Both references would fit and they would create links between some contribrutions of the RT.

Nevertheless, to be on the safe side, it might be necessary to wait until both are accepted for publication. If QPB wants to proceed faster, one or both references could be easily taken out to avoid this obstacle.

Review: Dynamics of homeostats: the basis of electrical, chemical, hydraulic, pH and calcium signaling in plants — R1/PR7

Conflict of interest statement

Reviewer declares none.

Comments

The authors have addressed my specific queries.

Review: Dynamics of homeostats: the basis of electrical, chemical, hydraulic, pH and calcium signaling in plants — R1/PR8

Conflict of interest statement

Reviewer declares none.

Comments

The authors have addressed my previous concerns properly and convincingly. I wish to congratulate them on this work!

My only objection refers to the use of the homeostat for modelling short-term changes in voltage and ion concentrations. For modelling K+/H+/Ca2+ signatures, which is the ultimate goal of this work, the authors will have to amend their homeostat such that the complex, short-term interplay of the parameters (voltage, ion concentrations…) is reflected in a still more realistic way.

Recommendation: Dynamics of homeostats: the basis of electrical, chemical, hydraulic, pH and calcium signaling in plants — R1/PR9

Comments

Hi Ingo,

Please accept my congratulations on this excellent contribution. We are delighted you chose QPB for this work.

There is a very helpful comment from one of the reviewers regarding the use of your model for short time-scales that you may wish to consider and address before uploading the final version of your manuscript.

With best wishes

Richard

Decision: Dynamics of homeostats: the basis of electrical, chemical, hydraulic, pH and calcium signaling in plants — R1/PR10

Comments

No accompanying comment.