Hostname: page-component-55f67697df-twqc4 Total loading time: 0 Render date: 2025-05-10T03:02:59.182Z Has data issue: false hasContentIssue false

Potassium homeostasis and signalling: from the whole plant to the subcellular level

Published online by Cambridge University Press:  08 May 2025

Lars H. Wegner
Affiliation:
International Research Center for Environmental Membrane Biology, School of Agriculture and Bioengineering, Foshan University, Foshan, China
Igor Pottosin
Affiliation:
Centro Universitario de Investigaciones Biomédicas, Universidad de Colima, Colima, México
Ingo Dreyer
Affiliation:
Electrical Signaling in Plants (ESP) Laboratory–Center of Bioinformatics, Simulation and Modeling (CBSM), Faculty of Engineering, Universidad de Talca, Talca, Chile
Sergey Shabala*
Affiliation:
International Research Center for Environmental Membrane Biology, School of Agriculture and Bioengineering, Foshan University, Foshan, China School of Biological Science, University of Western Australia, Crawley, WA, Australia
*
Corresponding author: Sergey Shabala; Email: [email protected]

Abstract

Potassium is an essential macronutrient required for plant growth and development. Over the recent decade, an important signalling role of K+ has emerged. Here, we discuss some aspects of such signalling at the various levels of plant functional organisation. The topic covered include: (1) mechanisms of long-distant K+ transport in the xylem and phloem and the molecular identity and regulation of K+ loading and unloading into plant vasculature; (2) essentiality and physiological roles of K+ cycling between shoots and roots; (3) plant sensing and signalling of low K+; (4) maintenance of K+ homeostasis at the cellular level; (5) stress-induced modulation of cytosolic K+ as a signal in plant adaptive responses to hostile environment; (6) stress-specific K+ “signatures” and mechanisms of their decoding by regulation of purine metabolism and H+-ATPase activity; (7) cytosolic K+ loss as a metabolic switch and a regulator of autophagy; and (8) vacuolar K+ transport and sensing.

Type
Review
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 (https://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

Potassium is the eighth most abundant element in the continental crust (2.1%) and also in seawater (Benito et al., Reference Benito, Haro, Amtmann, Cuin and Dreyer2014). Potassium is indispensable for all known living organisms. Despite its high abundance, however, the K+ food chain is not trivial. Animal organisms are not able to cover their K requirement from inorganic matter. Instead, they are dependent on plants that make this nutrient bioavailable in the first instance. This dependency may explain why potassium is likely the most investigated nutrient in plant science. In the model species Arabidopsis thaliana, 75 transporters may mediate K+ movements across cellular membranes; 35 of these are K+-selective (Very & Sentenac, Reference Very and Sentenac2003). Of these, high-affinity K+ transporter AtHAK5 and the inward rectifier K+ channel AKT1 are believed to be central to root K+ uptake (Lhamo et al., Reference Lhamo, Wang, Gao and Luan2021; Maierhofer et al., Reference Maierhofer, Scherzer, Carpaneto, Müller, Pardo, Hänelt, Geiger and Hedrich2024; Rubio et al., Reference Rubio, Nieves-Cordones, Horie and Shabala2020). The former transporter belongs to the HAK/KUP/KT K+ transporter family that operates at external K+ concentrations below 10 μM (Rubio et al., Reference Rubio, Nieves-Cordones, Horie and Shabala2020), while the latter belongs to Shaker-like group of K+ channels that mediate root K+ acquisition at higher concentrations. The same is true for many other species such as rice (Nieves-Cordones et al., Reference Nieves-Cordones, Martínez, Benito and Rubio2016), wheat (Zhang et al., Reference Zhang, Liu, Zhang, Yang, Guo and Wang2019), soybeans (Chao et al., Reference Chao, Li, Zhang, Huang, Ren, Xu and Huang2024), tomato (Rubio et al., Reference Rubio, Fon, Ródenas, Nieves-Cordones, Alemán, Rivero and Martínez2014) and maize (Ma et al., Reference Ma, Dong, Lu, Lu, Meng and Liu2020). Both types of transporters are regulated by complex pathways involving calcineurin B-like proteins (CBL) and CBL-interacting kinases (CIPK) and target of rapamycin (TOR) complexes that have been subjects of intensive studies (Cheong et al., Reference Cheong, Pandey, Grant, Batistic, Li, Kim, Lee, Kudla and Luan2007; Geiger et al., Reference Geiger, Becker, Vosloh, Gambale, Palme, Rehers, Anschuetz, Dreyer, Kudla and Hedrich2009; Li et al., Reference Li, Tang, Wang and Luan2023a,Reference Li, Xue, Tang and Luanb) and are not discussed here. Similarly, for other aspects, we refer to several excellent review articles (Anschütz et al., Reference Anschütz, Becker and Shabala2014; Britto et al., Reference Britto, Coskun and Kronzucker2021; Cakmak & Rengel, Reference Cakmak and Rengel2024; Ragel et al., Reference Ragel, Raddatz, Leidi, Quintero and Pardo2019; Zhang et al., Reference Zhang, Hu, Han, Chen, Lai and Wang2023; Zörb et al., Reference Zörb, Senbayram and Peiter2014) and apologize that we cannot cite all due to space limitations. Instead, we have decided to focus here on an aspect that has not received too much attention to date but is receiving increasing attention in research: the role of K+ as a signalling element.

2. K + homeostasis at the whole plant level

Nutrient allocation within the plant is mediated by the two major vascular systems; the xylem provides a pathway for water and nutrient transport from the roots to the shoot, and the main role of phloem is in translocating sugars from photosynthetically active source tissues, mainly the leaves, to the roots and other below-ground organs as well as to seeds and fruits (Rogiers et al., Reference Rogiers, Coetzee, Walker, Deloire and Tyerman2017; van Bel & Hafke, Reference van Bel and Hafke2005). The xylem is a microfluidic system formed by dead, interconnected tubes, whereas the phloem consists of living cells with flow occurring in the sieve elements. Via xylem and phloem, a circular transport of K+ is established (Figure 1).

Figure 1. A schematic diagram showing K+ circulation via xylem and phloem in higher plants; suc., sucrose. For more details, see the text.

Long-distance K+ transport, JK+.i, in the compartment i (either xylem or phloem) can be simply described by the following equation:

(1) $$\begin{align}{J}_{K+,i}={v}_i\ast {A}_i\ast {c}_{K+,i}\end{align}$$

with vi, Ai and cK+,i being the flow velocity, the cross-section of all xylem or phloem elements displaying flow and the xylem/phloem K+ concentration, respectively. An elegant way of assessing ${v}_i$ and Ai non-invasively in intact plants is by the magnetic resonance imaging (MRI) technique. Using this technique, Windt et al. (Reference Windt, Vergeldt, De Jager and Van As2006) monitored xylem and phloem flow in poplar, tomato, tobacco and Ricinus communis. Phloem flow velocity ranged from 0.25 to 0.4 mm/s in all species during daytime; this corresponded to a volume flow rate (vi* Ai) of 0.1 to 0.2 mm3/s in the latter three species and somewhat higher values (~1 mm3/s) in poplar. The xylem flow velocity was higher by a factor of at least 10, and the volume flow rate ranged from 1 mm3/s in tomato to 17 mm3/s in poplar. Similar volume flow rates in the xylem and phloem during nighttime indicated the cycling of water between both compartments. For intact transpiring plants, the xylem K+ concentration is directly accessible with the multifunctional xylem probe (Wegner & Zimmermann, Reference Wegner and Zimmermann2002), combining a pressure probe with a K+ selective microelectrode. The location of the probe tip in a vessel is verified by recording pressure values below atmospheric or even vacuum level. In transpiring maize seedlings, xylem K+ concentrations of about 1 mM were measured, varying slightly, among other things, with K+ in the media. Root-to-shoot K+ export can also be quantified by combining xylem K+ recordings with gravimetrical measurements of water uptake (Wegner & Zimmermann, Reference Wegner and Zimmermann2009). Measuring phloem K+ is also a technical challenge. By analysing the exudate from stylets of phloem-feeding aphids, Lohaus et al. (Reference Lohaus, Hussmann, Pennewiss, Schneider, Zhu and Sattelmacher2000) obtained values of about 50 mM K+ for maize leaves, indicating that the phloem sap is ~50 times more concentrated than the xylem sap with respect to K+. If the xylem volume flow was exceeding that in the phloem by a factor of 50, this would imply a 100% K+ circulation between compartments. Unfortunately, we lack MRI data for maize, but this factor was 70 in tomatoes and lower in other species. Apparently, a very large part of xylem K+ ends up in the phloem and returns to the root.

Current estimates of the extent to which K+ is circulating are still based on less sophisticated techniques which rely on isolating xylem and phloem sap with invasive methods and measuring leaf and root K+ content (Jeschke & Pate, Reference Jeschke and Pate1991; for possible technical limitations see Wegner, Reference Wegner2015). Values vary considerably depending on experimental conditions. For maize, Engels and Kirkby (Reference Engels and Kirkby2001), applying different temperature regimes to manipulate root and shoot growth, reported that between 34 and 93% of xylem K+ is recirculated via the phloem. Jeschke and Pate (Reference Jeschke and Pate1991) found values between 58.9 and 103 % for Lupin, Ricinus communis and barley at mild salinity. Touraine et al. (Reference Touraine, Grignon and Grignon1988) investigated nitrate-fed soybean and observed that 68% of K+ moving upward in the xylem returned back to the root via the phloem. In plants reducing N in the shoot, nitrate is the main counterion for K+ in the xylem sap, whereas in the phloem it is frequently malate (so-called Ben Zioni-Lips-model; Ben Zioni et al., Reference Ben Zioni, Vaadia and Lips1971). The extent to which K+ is recirculated depends on various factors including K+ shoot demand and the K+ nutritional status of the plant, suggesting that it serves as a shoot-to-root signal particularly under conditions of K+ deficiency.

The most compelling evidence in favour of this hypothesis comes from split-root experiments (Drew & Saker, Reference Drew and Saker1984). K+ export from a single seminal barley root supplied with full nutrient solution was monitored using 42K+ or 86Rb+ as a tracer for K+, while the rest of the root system that was placed in a separate container was either fed with the same solution or deprived of K+. The latter treatment leading to K+ deficiency in the shoot strongly stimulated K+ export from the single, well-fed root and significantly reduced phloem K+ concentration in that root (Drew et al., Reference Drew, Webb and Saker1990) compared to a situation of even K+ supply to all roots. Evidence for K+ re-circulation via the phloem also comes from blocking phloem transport, for example, in Arabidopsis mutants not expressing the NaKR1 (sodium potassium root defective 1) protein, leading to a K+ accumulation in leaves (Tian et al., Reference Tian, Baxter, Lahner, Reinders, Salt and Ward2010).

In order to understand how K+ allocation in the plant is organized, potential sites of control need to be identified. These are (i) xylem loading, that is, K+ release from stellar cells into adjacent dead xylem vessels in the root, (ii) K+ unloading from the xylem in leaf tissues and (iii) uptake of K+ by sieve tubes in assimilating leaf tissues. All three sites are associated with membrane transport steps that have been studied in some depth.

2.1. Xylem loading

Initial evidence for a key role of an outward rectifying K+ channel in root xylem loading was already presented by Wegner and Raschke (Reference Wegner and Raschke1994) and Wegner and de Boer (Reference Wegner and De Boer1997). These authors performed patch clamp experiments on protoplasts isolated from barley xylem parenchyma. Later Gaymard et al. (Reference Gaymard, Pilot, Lacombe, Bouchez, Bruneau, Boucherez, Michaux-Ferrière, Thibaud and Sentenac1998) reported that a Shaker-type channel, denoted as a stellar K+ outward rectifier (SKOR), served this function; in Arabidopsis SKOR knockout mutants the K+ transport to the shoot was reduced by 50%. Particularly under conditions of nitrate deficiency, SKOR is tightly controlled by the activity of the nitrate transporter NRT1.5, co-expressed in xylem parenchyma cells (Drechsler et al., Reference Drechsler, Zheng, Bohner, Nobmann, von Wirén, Kunze and Rausch2015), reminding us of the importance of a coordination of K+ loading with anion transport. Interestingly, gating of SKOR is tightly regulated by its substrate concentration at either side of the membrane (Figure 1). K+ transport capacity of SKOR increases with the cytosolic K+ activity by enhancing the channel open probability (Liu et al., Reference Liu, Li and Luan2006). Hence, K+ accumulation in xylem parenchyma cells, due to an increase in volume flow passing through these cells, would favour K+ release into the xylem (so-called concentration-polarisation effect). Consistently, K+ loading into the xylem was shown to strongly depend on transpirational water uptake by the roots (Wegner & Zimmermann, Reference Wegner and Zimmermann2009). Additionally, SKOR gating was also shown to depend on the external K+ concentration so that the shift of the equilibrium potential for K+, EK, was followed by almost equal shift of the voltage dependence of the channel open probability (Gaymard et al., Reference Gaymard, Pilot, Lacombe, Bouchez, Bruneau, Boucherez, Michaux-Ferrière, Thibaud and Sentenac1998; Wegner & De Boer, Reference Wegner and De Boer1997). Wegner and De Boer argued that K+ channel activity and, hence, K+ release into the xylem, will tend to decrease with the external K+ concentration if the K+ dependence of the cellular membrane potential (MP) of stellar cells is less pronounced than the shift in SKOR gating (which is likely, given the MP dependence on other conductances and the proton pump). This could be a way for K+ recirculation via the phloem affecting, by means of its potential impact on the stelar apoplastic K+ pool, SKOR gating. SKOR K+ dependence has been extensively studied using site directed mutagenesis; the pore of the channel was shown to be also responsible for sensing the external K+ concentration (Johansson et al., Reference Johansson, Wulfetange, Porée, Michard, Gajdanowicz, Lacombe, Sentenac, Thibaud, Mueller-Roeber, Blatt and Dreyer2006).

Xylem K+ concentration is also controlled along the way from the root to the leaves by cells bordering on the vessels. In Arabidopsis, AtHKT1.1 plays a key role in maintaining a favourable K+/Na+ ratio in plants suffering from salinity stress by mediating Na+ resorption by these cells and stimulating K+ release due to the related membrane depolarization (Hauser & Horie, Reference Hauser and Horie2010). Consistently, the K+/Na+ ratio was shown to be significantly lower in athkt1.1 knockout mutants than in the wildtype.

2.2. Xylem unloading in the leaf

For K+ unloading in the leaf, the bundle sheath cells (BSCs) enclosing the vascular tissue play an important role. Their effectiveness as a diffusion barrier towards the mesophyll is still under debate, particularly for C3 plants (Wigoda et al., Reference Wigoda, Moshelion and Moran2014). In the C4 plant maize, K+ is retrieved from the xylem sap via the BSCs, mainly in the smallest veins, as revealed by using 85Rb+ as a tracer for K+ (Keunecke et al., Reference Keunecke, Lindner, Seydel, Schulz and Hansen2001). Inward-rectifying K+ channels in Arabidopsis BSC protoplasts, which are likely to play a role in K+ resorption from the xylem, differed from mesophyll K+ channels with respect to their voltage dependence (Wigoda et al., Reference Wigoda, Pasmanik-Chor, Yang, Yu, Moshelion and Moran2017).

2.3. K +  uptake and release by sieve tubes and the role of K +  in the phloem

K+ plays an important role in maintaining the source-to-sink translocation of sugars in the phloem, and, reciprocally, the phloem strongly contributes to K+ homeostasis in the plant. Evidence for an involvement of K+ in phloem sugar translocation came from the observation that even mild K+ deficiency affected assimilate export from leaves (Amir & Reinhold, Reference Amir and Reinhold1971). High phloem K+ concentrations and the existence of a source-to-sink K+ gradient (Vreugdenhil, Reference Vreugdenhil1985) prompted Lang (Reference Lang1983) to postulate a key role of K+ in maintaining a phloem turgor gradient from source to sink, which is driving volume flow according to Münch’s generally accepted ‘Druckstrom’ (pressure-flow) hypothesis. A K+ gradient would particularly come into play when sucrose gradients fail to drive a sufficient mass flow. Lang argued that phloem K+ gradients could be adjusted to optimize convective flow, which could thus even be uncoupled, to some extent, from processes of phloem sucrose loading and unloading at source and sink, respectively (see also van Bel & Hafke, Reference van Bel and Hafke2005). A deeper insight into K+ loading processes in the leaf came with the identification of the shaker-type K+ channel AKT2/3, playing a key role in this process (Deeken et al., Reference Deeken, Geiger, Fromm, Koroleva, Ache, Langenfeld-Heyser, Sauer, May and Hedrich2002). Interestingly, AKT2/3 can either be strongly voltage-dependent being activated by increasingly negative membrane potentials (but being inactive around the K+ Nernst potential), or it can operate in a voltage-independent mode with K+ channel activity being unaffected by the membrane potential (Dreyer et al., Reference Dreyer, Michard, Lacombe and Thibaud2001). Based on these two modes of operation, Gajdanowicz et al. (Reference Gajdanowicz, Michard, Sandmann, Rocha, Corrêa, Ramírez-Aguilar, Gomez-Porras, González, Thibaud, van Dongen and Dreyer2011) developed the K+ battery concept (see also Dreyer et al., Reference Dreyer, Gomez-Porras and Riedelsberger2017 and Figure 1): The normally inward-rectifying channel can be switched by post-translational modifications, that is, phosphorylation, into the non-rectifying mode, which is mechanistically achieved by lowering the activation threshold of the channel (Michard et al., Reference Michard, Lacombe, Porée, Mueller-Roeber, Sentenac, Thibaud and Dreyer2005). With these open channels, the transmembrane K+ gradient, priorly established by K+ accumulation in the sieve tube, then contributes to the electrical driving forces that energize other membrane transport processes, for example, sucrose loading into the sieve tube via a sucrose-H+ symporter of the SUT type. The importance of the K+ battery is particularly evident under energy-limiting growth conditions. When gating mode switching was disabled in transgenic Arabidopsis, the plants showed a retarded developmental phenotype at short days or reduced ambient O2 supply causing local ATP deficiency (Gajdanowicz et al., Reference Gajdanowicz, Michard, Sandmann, Rocha, Corrêa, Ramírez-Aguilar, Gomez-Porras, González, Thibaud, van Dongen and Dreyer2011). The hypothesis of the K+ battery was corroborated by a combination of experimental data and quantitative modelling. It should be noted that the mechanism relies on a sufficient H+ buffer capacity of the sieve tubes and comes with a progressive alleviation of the K+ gradient to maintain charge balance.

Phloem K+ unloading at the sink, which received much less attention so far, can either be symplastic via plasmodesmata connecting the phloem with adjacent tissues or apoplastic by K+ (and sucrose) release via the sieve tube plasma membrane. A shaker-type K+ channel from Vicia faba, VFK1, is likely to be involved in the latter process; it was shown to be predominantly expressed in stems and in leaves becoming sinks by keeping them in the dark and depriving them of CO2 (Ache et al., Reference Ache, Becker, Deeken, Dreyer, Weber, Fromm and Hedrich2001).

To sum up, the main switches directing long-distance K+ circulation in the plant may have been identified including some of the regulatory mechanisms, but we are still far from a truly systemic understanding of K+ cycling between the xylem and phloem. K+ allocation in the plant does not only rely on K+ availability in the soil and K+ demand in source and sink tissues; it is also strongly affected by the convective mass flow in xylem and phloem (Wegner, Reference Wegner2015). Transpirational flow has a strong impact on K+ delivery to the shoot both affecting the K+ concentration gradient at the plasma membrane of root parenchyma cells and JK+.x. On the other hand, mass flow in the phloem is dominated by assimilate synthesis in the source and demand in the sink. Moreover, K+ transport in the xylem is linked to that of nitrate in plants reducing N in the shoot. Hence, we are confronted with a process integrating many aspects of plant life. Only a modelling approach taking account of these potentially antagonistic regulatory factors will allow us to unravel fine-tuning of K+ allocation in the plant. This way, the molecular information on different K+ transport processes will help us to understand phenomena of long-distance K+ cycling known for several decades.

3. K +  homeostasis at the cellular level

Modelling approaches have been proven already as rather helpful in filling gaps that are hard to address in wet-lab experiments. The above-mentioned concept of the K+ battery in the phloem, for instance, was initially developed in computational simulations. Based on the experience from these dry-laboratory experiments pointed experiments with transgenic plants under energy-limiting conditions could be designed, in that case with low ambient O2 to reduce ATP formation by respiration. The successful wet-laboratory experiment was thus the proof of concept of the model (Gajdanowicz et al., Reference Gajdanowicz, Michard, Sandmann, Rocha, Corrêa, Ramírez-Aguilar, Gomez-Porras, González, Thibaud, van Dongen and Dreyer2011; Sandmann et al., Reference Sandmann, Skłodowski, Gajdanowicz, Michard, Rocha, Gomez-Porras, González, Corrêa, Ramírez-Aguilar, Cuin, van Dongen, Thibaud and Dreyer2011). Recent advances in modelling have also provided new insights into K+ homeostasis at the cellular level. The modelling of cellular steady-state conditions, initially thought to be ‘trivial’, revealed that an adjustable steady-state of the transmembrane K gradient can only be achieved if at least two differently energized K+ transporter types are present in the membrane (Dreyer, Reference Dreyer2021). If this condition is met, the transmembrane K+ gradient can be adjusted by regulating the transporter activities. However, if it is not fulfilled, the regulation of transporter activity has no influence on the steady-state gradient (Dreyer, Reference Dreyer2021; Dreyer et al., Reference Dreyer, Hernández-Rojas, Bolua-Hernández, Tapia-Castillo, Astola-Mariscal, Díaz-Pico, Mérida-Quesada, Vergara-Valladares, Arrey-Salas and Rubio-Meléndez2024). Remarkably, the gain of control over the transmembrane homeostatic condition has a metabolic cost. The steady state is inevitably accompanied by transmembrane H+ and K+ cycles driven by ATP-consuming proton pumping (Figure 2). Such nutrient cycling was reported experimentally, for instance, under salt stress (Munns et al., Reference Munns, Day, Fricke, Watt, Arsova, Barkla, Bose, Byrt, Chen, Foster, Gilliham, Henderson, Jenkins, Kronzucker, Miklavcic, Plett, Roy, Shabala, Shelden and Tyerman2020; Shabala et al., Reference Shabala, Chen, Chen and Pottosin2020) and considered “futile” cycles (Britto & Kronzucker, Reference Britto and Kronzucker2006). The model analysis indicated, however, that these cycles are not as futile as they initially appear, but an essential part of the homeostatic process (Dreyer, Reference Dreyer2021).

Figure 2. Homeostatic (steady state) condition of a transmembrane K+ gradient. In this example, the membrane voltage in a steady state is V ss = −130 mV. At this voltage, there is an efflux of H+ released by the proton pump. The charge transport is compensated by a K+ influx of the same magnitude via the K+ channel (KC). Without further compensation, the system would not be stable in steady state but mediate an H+/K+ antiport, which would at least change the transmembrane potassium gradient. E K would become increasingly negative. The system would only be stable if the red and green dots overlap with the light blue dot at −200 mV. Stability at V ss = −130 mV can be obtained with an additional H+/K+ antiporter. This transporter releases each accumulated K+ ion and in the same process retrieves the released H+ in an electroneutral manner. The system is in steady state, and V ss or E K do not change. However, there are ATP-consuming H+ and K+ cycles.

If the system is brought out of steady state by changing the potassium concentration on one side of the membrane, the transporter network (called homeostat) mediates a largely electroneutral H+/K+ antiport to restore the altered gradient to its original value. Such an H+/K+ exchange also takes place if there are no H+/K+ antiporters in the membrane via symporters and channels (Contador-Álvarez et al., Reference Contador-Álvarez, Rojas-Rocco, Rodríguez-Gómez, Rubio-Meléndez, Riedelsberger, Michard and Dreyer2025). The exchange of H+ for K+ upon a change in the K+ gradient has been known since long (Blatt & Slayman, Reference Blatt and Slayman1987; Conway & O’Malley, Reference Conway and O'Malley1944; Maathuis & Sanders, Reference Maathuis and Sanders1994; Rodriguez-Navarro et al., Reference Rodriguez-Navarro, Blatt and Slayman1986) and was also recently observed in guard cells and mesophyll cells from Nicotiana benthamiana or Nicotiana tabacum (Li et al., Reference Li, Grauschopf, Hedrich, Dreyer and Konrad2024). The homeostat concept now provides a detailed mechanistic explanation of these observations. Interestingly, H+/K+ antiport occurred in a coordinated manner across both the plasma membrane and the vacuolar membrane supporting further the model prediction that there is an electrochemical coupling to sequential membranes (Dreyer et al., Reference Dreyer, Vergara-Valladares, Mérida-Quesada, Rubio-Meléndez, Hernández-Rojas, Riedelsberger, Astola-Mariscal, Heitmüller, Yanez-Chávez, Arrey-Salas, San Martín-Davison, Navarro-Retamal and Michard2023). Remarkably, besides organizing transmembrane K+ transport the homeostat also generates a cytosolic pH signal that additionally depends on various cellular parameters (Contador-Álvarez et al., Reference Contador-Álvarez, Rojas-Rocco, Rodríguez-Gómez, Rubio-Meléndez, Riedelsberger, Michard and Dreyer2025). Thus, in addition to its role as an information carrier at the whole plant level, K+ can also transmit messages at the cellular level.

The homeostat concept, which considers transporter networks instead of individual transporters, has so far explained well several observations, especially on K+ homeostasis. However, it also provided new insights into the homeostasis of other ions and even hormones (Dreyer et al., Reference Dreyer, Li, Riedelsberger, Hedrich, Konrad and Michard2022; Dreyer et al., Reference Dreyer, Hernández-Rojas, Bolua-Hernández, Tapia-Castillo, Astola-Mariscal, Díaz-Pico, Mérida-Quesada, Vergara-Valladares, Arrey-Salas and Rubio-Meléndez2024; Geisler & Dreyer, Reference Geisler and Dreyer2024; Li et al., Reference Li, Grauschopf, Hedrich, Dreyer and Konrad2024). As just mentioned before, homeostats appear also to constitute important parts of signalling processes in plants (Contador-Álvarez et al., Reference Contador-Álvarez, Rojas-Rocco, Rodríguez-Gómez, Rubio-Meléndez, Riedelsberger, Michard and Dreyer2025). Furthermore, the homeostat concept applies not only to the plasma membrane but also to endomembrane (Li et al., Reference Li, Grauschopf, Hedrich, Dreyer and Konrad2024). It could therefore contribute in the future to a better understanding of the role of K+ at the subcellular level.

4. From the cellular to the subcellular level: stress-induced modulation of cytosolic K+ as a signal in plant adaptive responses to hostile environment

In recent years, a significant bulk of evidence has been accumulated demonstrating that, in addition to a well-known role of K+ as a nutrient, changes in intracellular K+ concentrations may act as important developmental and adaptive signals (reviewed in Anschütz et al., Reference Anschütz, Becker and Shabala2014; Shabala, Reference Shabala2017; Wu et al., Reference Wu, Zhang, Giraldo and Shabala2018), affecting key processes such as cell fate determination and autophagy. Stress-induced K+ loss has been demonstrated in response to a broad range of abiotic and biotic stresses (reviewed in Shabala & Pottosin, Reference Shabala and Pottosin2014). These responses are often very fast (<1 min) and comparable with reported changes in cytosolic Ca2+ signals that are traditionally considered being upstream of stress signalling (Dong et al., Reference Dong, Wallrad, Almutairi and Kudla2022). It was also shown that such stress-induced K+ signalling is highly tissue-specific (Ahmed et al., Reference Ahmed, Shabala and Shabala2021; Chakraborty et al., Reference Chakraborty, Bose, Shabala and Shabala2016; Shabala et al., Reference Shabala, Zhang, Pottosin, Bose, Zhu, Fuglsang, Velarde-Buendia, Massart, Hill, Roessner, Bacic, Wu, Azzarello, Pandolfi, Zhou, Poschenrieder, Mancuso and Shabala2016a), and that kinetics of stress-induced K+ loss differ drastically between plant species, leading to a new concept of the stress-induced K+ flux ‘signatures’ (Rubio et al., Reference Rubio, Nieves-Cordones, Horie and Shabala2020), similar to those reported for cytosolic Ca2+. The physiological role of stress-induced modulation of cytosolic K+ and the mechanisms by which these K+ “signatures” are decoded are understood much less and require more discussion.

4.1. Potassium as a “metabolic switch”

This concept, first put forward by Demidchik (Reference Demidchik2014), implies that a drop in the cytosolic K+ concentration, [K+]cyt, will inhibit energy-consuming anabolic reactions and allow plants to allocate more energy towards defence responses. Indeed, K+ is essential for operation of numerous enzymes, by either providing proper ionic strength or catalysing their activity as a cofactor. For example, K+ is critical for operation of the pyruvate kinase that catalyses the conversion of phosphoenolpyruvate (PEP) and adenosine diphosphate (ADP) to pyruvate and ATP in the glycolytic pathway, by enabling a proper electrostatic environment in the active site for reactants as well as for stabilization of the product (Cui & Tcherkez, Reference Cui and Tcherkez2021). Another example includes the succinyl-CoA thiokinase, an essential enzyme in tricarboxylic acid cycle catalysing the reversible conversion of succinyl-CoA to succinate, where K+ facilitates magnesium binding and promotes the reaction in the forward direction (Lynn & Guynn, Reference Lynn and Guynn1978). Potassium is also a key component of the ribosomes, being present at most crucial positions such as codon-decoding and peptidyl transferase domains (Rozov et al., Reference Rozov, Khusainov, El Omari, Duman, Mykhaylyk, Yusupov, Westhof, Wagner and Yusupova2019). This makes it essential for stabilization of mRNA binding and tRNA ligands, maintaining the correct frame position during elongation and reinforcing rRNA–protein interactions (Cui & Tcherkez, Reference Cui and Tcherkez2021). Most of these enzymes are designed to operate in a high mM K+ concentration range. For example, Vmax value for starch synthase operation is 100 mM, 20 mM for glutathione synthase, 50 mM for pyruvate kinase, 90 mM for PEP carboxylase, etc (Evans & Sorger, Reference Evans and Sorger1966). At the same time, under stress conditions, the cytosolic K+ level may drop several folds within minutes of stress exposure (e.g., down to 15 mM within 10 min of 50 mM NaCl treatment; Shabala et al., Reference Shabala, Demidchik, Shabala, Cuin, Smith, Miller, Davies and Newman2006). In lucerne plants, reduction in leaf K+ content from 3.8 to 1.3 mg g−1 DW resulted in a 3.5-fold reduction in Rubisco activity (Peoples & Koch, Reference Peoples and Koch1979), and starch synthase activity was reduced 6-fold by changing K+ levels from 50 to 1 mM (Nitsos & Evans, Reference Nitsos and Evans1969). Thus, a reduction in [K+]cyt is expected to significantly reduce both plant enzymatic activity and protein synthesis avoiding the competition for energy between metabolic and defence responses. Consistent with this notion are reports that overexpressing pyruvate kinase PK21 (with strict K+ dependency) has weakened soybean salt tolerance (Liu et al., Reference Liu, Wang, Zhang, Li, Wang, Xu, Dai and Zhang2024).

4.2. Potassium deprivation induces autophagy

Autophagy targets a wide variety of cytoplasmic cargo for degradation, including breaking down proteins into their constituent amino acids. Recently, Rangarajan et al. (Reference Rangarajan, Kapoor, Li, Drossopoulos, White, Madden and Dohlman2020) showed that K+ starvation promoted autophagy in yeasts through core autophagy-related kinases and the PI3-kinase complex. In their work, transferring yeast cells from 10 to 1 mM K+ media resulted in a 70% increase in the intensity of a Rosella signal, a fluorescent reporter of autophagy. Autophagy is also regulated by TORC1 (target of rapamycin complex 1). When [K+]cyt is high, TORC1 remains active and inhibits autophagy via phosphorylation (Li et al., Reference Li, Xue, Tang and Luan2023b), and a drop in the [K+]cyt deactivates TORC1 promoting assembly of autophagy-related proteins, with a 50-fold difference in the phosphorylation level of TORC-related genes reported for Arabidopsis plants after 2 h exposure to low-K+ (10 μM) or high-K+ (10 mM) liquid medium. TOR also interacts antagonistically with SnRK1 (sucrose nonfermenting 1 (SNF1)-related kinase 1), a heterotrimeric Ser/Thr protein kinase complex that is considered a key sensor of energy deficit and a central regulator of cellular energy homeostasis (Xiao et al., Reference Xiao, Zhou, Xie, Li, Su, He, Jiang, Zhu and Qu2024). SnRK1 is an upstream regulator of TOR and can induce autophagy by directly interacting with autophagy-related proteins in the case of ATP shortage (Xiao et al., Reference Xiao, Zhou, Xie, Li, Su, He, Jiang, Zhu and Qu2024).

4.3. Cytosolic K +  levels can induce apoptosis

Apoptosis is characterized by a distinct series of morphological and biochemical changes that result in cell shrinkage, DNA breakdown, and, ultimately, cell death. Diverse external and internal stimuli trigger apoptosis, and enhanced K+ efflux has been shown to be an essential mediator of not only early apoptotic cell shrinkage but also of downstream caspase activation and DNA fragmentation in many living systems (Peters & Chin, Reference Peters and Chin2007; Remillard & Yuan, Reference Remillard and Yuan2004). The failure to maintain high [K+]cyt can also induce cell elimination via programmed cell death (PCD) by unblocking activities of caspase-like proteases and endonucleases in plants (Demidchik et al., Reference Demidchik, Cuin, Svistunenko, Smith, Miller, Shabala, Sokolik and Yurin2010; Shabala, Reference Shabala2009).

4.4. Plant sensing and signalling of low K +

In plant systems, the severity and specificity of stress is encoded by so-called cytosolic Ca2+ “signatures” that are then decoded by a sophisticated Ca2+ signalling network consisting of calcineurin B-like proteins (CBLs) and CBL-interacting kinases (CIPKs) that operate upstream of all key membrane transporters regulating plant ionic homeostasis (Dong et al., Reference Dong, Wallrad, Almutairi and Kudla2022; Li et al., Reference Li, Tang, Wang and Luan2023a). This is also true for low-K+ stress, where CBL1/9 (calcineurin B-like 1/9) and CIPK23 (CBL-interacting protein kinase 23) interact with Ca2+ signal triggered by low K+, to cause reorchestration of metabolism (Rodenas & Vert, Reference Rodenas and Vert2020). CBL1/9-CIPK23 complexes then phosphorylate and activate AKT1 (Behera et al., Reference Behera, Long, Schmitz-Thom, Wang, Zhang, Li, Steinhorst, Manishankar, Ren, Offenborn, Wu, Kudla and Wang2017) as well as regulate K+ release from the vacuole (Li et al., Reference Li, Tang, Wang and Luan2023a), to maintain [K+]cyt homeostasis. The latter authors showed that the K+ status regulates the protein abundance and phosphorylation of the CBL-CIPK-channel modules at the posttranslational level. CIPK9/23 kinases were responsible for phosphorylation of CBL1/9/2/3 in plant response to low-K+ stress and the HAB1/ABI1/ABI2/PP2CA phosphatases were responsible for CBL2/3-CIPK9 dephosphorylation upon K+ repletion, acting as a negative regulator (Li et al., Reference Li, Tang, Wang and Luan2023a). The above responses to K+ deprivation appear to be highly tissue-specific and rapid. Using Ca2+ reporter protein YC3.6, Behera et al. (Reference Behera, Long, Schmitz-Thom, Wang, Zhang, Li, Steinhorst, Manishankar, Ren, Offenborn, Wu, Kudla and Wang2017) reported an increase in cytosolic Ca2+ starting just after 60 sec of K+ deprivation and peaking at 140 sec. The response was most pronounced in the root elongation zone. Furthermore, Wang et al. (Reference Wang, Tan, Wallrad, Du, Eickelkamp, Wang, He, Rehms, Li, Han, Schmitz-Thom, Wu, Kudla and Wang2021) identified a so-called post-meristematic K+-sensing niche (KSN) where rapid K+ decline and Ca2+ signals coincided. The reported decline in K+ was very rapid (observed within 60 sec and peaked at 2 min) and then triggered phosphorylation of the NADPH oxidases RBOHC, RBOHD and RBOHF, via CIF peptide-activated SGN3-LKS4/SGN1 receptor complex (Wang et al., Reference Wang, Tan, Wallrad, Du, Eickelkamp, Wang, He, Rehms, Li, Han, Schmitz-Thom, Wu, Kudla and Wang2021). It should be noted that low-K+ signalling may also be intrinsically linked with Ca2+ signalling, via voltage coupling. Indeed, a hyperpolarization of the membrane occurs when K+ is removed which immediately leads to an increase in Ca2+ influx through hyperpolarization-activated Ca2+ channels (Gelli & Blumwald, Reference Gelli and Blumwald1997; Grabov & Blatt, Reference Grabov and Blatt1998; Lemtiri-Chlieh & Berkowitz, Reference Lemtiri-Chlieh and Berkowitz2004) and, hence, a rise in the cytosolic Ca2+.

Sensing of low-K environment also implies significant changes at the transcriptional level. Several K+ transporters in Arabidopsis are transcriptionally induced by K+-starvation, such as AtHAK5, AtKEA5, AtKUP3, AtCHX13 and AtCHX17 (Wang & Wu, Reference Wang and Wu2010). Among the genes encoding the channels and transporters expressed in root outer cell layers, AtHAK5 appears to be the most highly regulated (Chérel et al., Reference Chérel, Lefoulon, Boeglin and Sentenac2014). The relative abundance of OsHAK1 mRNA transcripts in rice seedlings was increased 2-fold 16 h after roots transfer to a K+-free medium for 16 h (Bañuelos et al., Reference Bañuelos, Garciadeblas, Cubero and Rodríguez-Navarro2002), and Nieves-Cordones et al. (Reference Nieves-Cordones, Alemán, Martínez and Rubio2010) reported a 600-fold increase in AtHAK5 transcript levels after 14 d of K+ starvation. Low-K stress-induced increase in ROS levels seems to be required for these transcriptional changes (Shin & Schachtman, Reference Shin and Schachtman2004) although the details remain scarce, as hundreds of other genes related to transcription, carbohydrate metabolism, hormonal signalling, protein phosphorylation, ROS metabolism and Ca2+ signal generation and transduction, are also up-regulated in response to K+-deficiency (Jin et al., Reference Jin, Yan, Li, Liu, Zhao, Zhang, Zhang, Zhu, Wang, Yu, Zhang, Yang and Tang2024; Ruan et al., Reference Ruan, Zhang, Xin, Zhang, Ma, Chen and Zhao2015; Schachtman & Shin, Reference Schachtman and Shin2007; Shankar et al., Reference Shankar, Singh, Kanwar, Srivastava, Pandey, Suprasanna, Kapoor and Pandey2013).

4.5. Decoding K +  “signatures”

In the Arabidopsis genome, 26 CIPK and 10 CBL are present, providing a plethora of possibilities for decoding stress specificity. Such decoding machinery for K+ “signatures” remains mostly unexplored. Here, two possible mechanisms are put forward and discussed.

4.5.1. Decoding by regulation of purine metabolism

One of the enzymes with a strong K+ dependency is the adenosine kinase (ADK) that catalyses the phosphorylation of adenosine to 5-AMP using ATP as a source of phosphate. It was shown that binding of K+ to Asp310 is essential for a transition to a catalytically productive structure and removing auto-inhibition of ADK (de Oliveira et al., Reference de Oliveira, Morales-Neto, Rocco, Sforca, Polo, Tonoli, Mercaldi, Cordeiro, Murakami and Franchini2018). Thus, a drop in [K+]cyt will reduce the amount of cAMP (3’,5’-cyclic adenosine monophosphate). The latter, in turn, is a second messenger for a wide range of stress responses (Blanco et al., Reference Blanco, Fortunato, Viggiano and de Pinto2020), with demonstrated ability to regulate plant ionic homeostasis by controlling activity of various ion channels and transporters (Kaplan et al., Reference Kaplan, Sherman and Fromm2007). One of them is SOS1 H+/Na+ exchanger that is considered as a critical component of the salt exclusion mechanism (Zhao et al., Reference Zhao, Zhang, Song, Zhu and Shabala2020). It was shown recently that SOS1 contains a cyclic nucleotide-binding domain (CNBD; Zhang et al., Reference Zhang, Hu, Han, Chen, Lai and Wang2023) between residues 732 and 883. Moreover, mutations at G777 and G784 sides affected folding and stability of CNBD thus significantly decreasing transport activity of SOS1 (Wang et al., Reference Wang, Pan, Chen, Xie, Gao, He, Li, Dong, Jiang and Zhao2023). These findings suggest a possible causal link between salt-induced reduction in cytosolic K+ content, ADK-mediated cAMP production and SOS1 activation. Also, in response to stress, small amounts of ATP (nanomolar concentrations; Xiao et al., Reference Xiao, Zhou, Xie, Li, Su, He, Jiang, Zhu and Qu2024) are released to apoplastic space to be sensed by purinoreceptors that activate secondary signalling cascades (Kim et al., Reference Kim, Yanders and Stacey2023).

4.5.2. Decoding by regulating H+-ATPase activity

Stress-induced depolarization of the membrane potential (MP) is arguably one of the fastest events detected at the cellular level in response to a broad range of abiotic stresses (reviewed in Shabala et al., Reference Shabala, Zhang, Pottosin, Bose, Zhu, Fuglsang, Velarde-Buendia, Massart, Hill, Roessner, Bacic, Wu, Azzarello, Pandolfi, Zhou, Poschenrieder, Mancuso and Shabala2016). This is also true in response to K+ deficiency, where the plasma membrane (PM) is hyperpolarized within a few minutes (Maathuis & Sanders, Reference Maathuis and Sanders1993; Nieves-Cordones et al., Reference Nieves-Cordones, Miller, Aleman, Martinez and Rubio2008), while other signals such as ROS or ethylene start to operate after at least several hours (Behera et al., Reference Behera, Long, Schmitz-Thom, Wang, Zhang, Li, Steinhorst, Manishankar, Ren, Offenborn, Wu, Kudla and Wang2017). Equally fast are changes in MP caused by hypoxia or salinity (Bose et al., Reference Bose, Shabala, Pottosin, Zeng, Velarde-Buendía, Massart, Poschenrieder, Hariadi and Shabala2014, Reference Bose, Rodrigo-Moreno, Lai, Xie, Shen and Shabala2015) (Figure 3). As many K+-permeable channels are voltage gated (Demidchik & Maathuis, Reference Demidchik and Maathuis2007; Jegla et al., Reference Jegla, Busey and Assmann2018), this comes with important implications for intracellular ion homeostasis. Recently, Maierhofer et al. (Reference Maierhofer, Scherzer, Carpaneto, Müller, Pardo, Hänelt, Geiger and Hedrich2024) have shown that AtHAK5 operates as a proton-driven K+ transporter that is activated by membrane hyperpolarization, emphasising the essentiality of voltage gating in regulation of K+ homeostasis in plant cells.

Figure 3. Difference in NaCl-induced K+ flux ‘signatures’ explain the difference in the strategy of adapting to salt stress between pea and barley. (a) Transient changes in MP of epidermal root cells in wheat and pea in response to acute NaCl treatment (based on Cuin et al., Reference Cuin, Betts, Chalmandrier and Shabala2008 and Bose et al., Reference Bose, Shabala, Pottosin, Zeng, Velarde-Buendía, Massart, Poschenrieder, Hariadi and Shabala2014, respectively). (b) Net K+ fluxes measured from plant roots by non-invasive ion flux measuring technique. (c) Predicted changes in cytosolic K+ concentration in pea and wheat root cytosol assuming no buffering from the vacuole. (d) Effect of vacuolar K+ buffering on cytosolic K+ kinetics in pea roots exposed to acute salinity treatment. Percentage values are the size of the vacuolar K+ available to be used to compensate cytosolic K+ loss.

In plants, the critical contributor to MP maintenance is the H+-ATPase pump. When the plasma membrane is depolarized under stress conditions (e.g., salinity stress), H+-ATPase activity is rapidly upregulated (Chen et al., Reference Chen, Pottosin, Cuin, Fuglsang, Tester, Jha, Zepeda-Jazo, Zhou, Palmgren, Newman and Shabala2007; Shabala et al., Reference Shabala, Zhang, Pottosin, Bose, Zhu, Fuglsang, Velarde-Buendia, Massart, Hill, Roessner, Bacic, Wu, Azzarello, Pandolfi, Zhou, Poschenrieder, Mancuso and Shabala2016), to restore MP. Such activation also provides an additional driving force for secondary active (H+-coupled) uptake of essential nutrients (i.e. PO4, NO3 or K+) and exclusion of toxins (such as Na+). K+ can bind to the proton pump at a site involving Asp617 in the cytoplasmic phosphorylation domain (Buch-Pedersen et al., Reference Buch-Pedersen, Rudashevskaya, Berner, Venema and Palmgren2006). Such binding can induce dephosphorylation of the E1P reaction cycle intermediate and, therefore, control the H+/ATP coupling ratio and uncouple ATP hydrolysis from transport. In other words, K+-induced dephosphorylation serves as a negative regulator of the transmembrane electrochemical gradient. We believe that this mechanism is ideally suited for decoding stress-induced K+ flux “signatures”, as illustrated in an example later.

Earlier, we have shown that pea and wheat plants have different strategies of responding to salt. Pea plants were able to quickly restore their MP within minutes upon acute salt treatment (Bose et al., Reference Bose, Shabala, Pottosin, Zeng, Velarde-Buendía, Massart, Poschenrieder, Hariadi and Shabala2014) (Figure 3a), while the MP in wheat root epidermal cells remained depolarized for much longer period (Cuin et al., Reference Cuin, Betts, Chalmandrier and Shabala2008) (Figure 3a). This difference was accompanied by a striking difference in the kinetics of the NaCl-induced K+ loss from their roots (Figure 3b). In pea, this K+ efflux was massive but transient while in wheat K+ efflux was an order of magnitude lower and remained more or less constant over the first 10 min. Can these differences in K+ flux “signatures” affect the operation of H+-ATPase pumps and explain the differences in MP kinetics?

To answer this question, some model calculations were done. Assuming the diameter of the root cell being 50 μm, then the cell volume is 6.54×10−14 m3 (65.4 pL) and the cell surface area is 7.85×10−9 m2. Accepting a basal cytosolic K+ concentration ([K+]cyt) as 100 mM, then the total amount of K+ in one cell is 6.54 pmol under control (unstressed conditions). In pea plants, the total K+ loss from root cells ranged from ∼ 520 nmol m−2s−1 1.5 min after stress onset to ∼ -40 nmol m−2s−1 20 min after (Figure 3b). The total K+ loss (in nmol) from the cell can be calculated as

(2) $$\begin{align}&\left[\mathrm{K}\;\mathrm{loss}\right]=\mathrm{Flux}\;\left(\mathrm{nmol}\;\mathrm{m}-2\mathrm{s}-1\right)\times \mathrm{surface}\kern0.17em \mathrm{area}\;\left(\mathrm{m}2\right)\nonumber\\&\qquad\qquad\times \mathrm{time}\;\left(\sec \right)\end{align}$$

Assuming that the efflux measured by non-invasive MIFE technique originates from the topmost (epidermal) cell layer; the cytosol representing 10% of the cell volume; no vacuolar buffering (e.g., no ability to replace lost cytosolic K+ from the vacuolar K+ pool); and feeding the flux data from Figure 3b into (Eq. 2), one can see that [K+]cyt in pea will drop by about 2-fold within 2 min after stress onset and will be reduced to zero within 6 min (Figure 3c). At 5% vacuolar buffering capacity, a 50% decline in [K+]cyt will occur by 6 min, and at 10% buffering – at 18 min (Figure 3d). Such rapid and drastic changes in [K+]cyt may explain rapid activation of H+-pumping (Bose et al., Reference Bose, Shabala, Pottosin, Zeng, Velarde-Buendía, Massart, Poschenrieder, Hariadi and Shabala2014) and repolarization of the MP (Figure 3a) following the removal of K+-induced dephosphorylation of H+-ATPase. In wheat, however, NaCl-induced reduction in [K+]cyt was much slower, and even without vacuolar buffering within 15 min [K+]cyt will drop only by 15 mM (Figure 1c). As a result, the H+-ATPase remained dephosphorylated for much longer and failed to rapidly restore the MP.

As shown in the above-mentioned examples, the extent of H+-ATPase dephosphorylation will be critically dependent not only on the magnitude of stress-induced K+ loss but also on the plant’s ability to replenish [K+]cyt loss from the vacuolar pool which may be highly specific for different cell types/tissues, as well as plant species. Thus, modulation of the coupling between H+-ATPase hydrolytic and transport activity by [K+]cyt “signatures” may represent a decoding mechanism and explain the difference in species’ strategies of handling salt stress.

The above modelling has been done without considering the fact that cells are connected by plasmodesmata, so cortical cells may also provide some additional buffering. However, giving the lack of the quantitative data on how rapid K+ from the neighbouring cells may be transported to root epidermis, modelling of this process is nearly impossible. It should be also kept in mind that these (cortical) cells will be also exposed to NaCl stress and will undergo the same membrane depolarisation and, hence, K+ loss. Thus, such additional “buffering capacity” is likely to play a relatively minor role in determining cytosolic K+ levels in root epidermal cells.

5. Vacuolar K+-permeable channels sense vacuolar K+ content

The vacuole is the main intracellular K+ store in plant cells and, except extreme K+-starving conditions, the electrochemical K+ gradient across the tonoplast is cytosol-directed. Thus, K+ accumulation in the vacuole is mediated by an active K+/H+ antiport, whereas passive vacuolar K+ release is channel-mediated (Li et al., Reference Li, Grauschopf, Hedrich, Dreyer and Konrad2024). The vacuolar K+ concentration tends to be more variable as compared to cytosolic K+. During stomata closure there is a massive decrease of intracellular (presumably, mainly vacuolar) K+ content, from 450 to 95 mM in Commelina communis (Leigh, Reference Leigh1997). For several plant species, collectively, upon stomatal closure the cytosolic K+ concentration decreases from 150–247 mM to 55–93 mM, whereas the vacuolar one from 181–454 mM to 38–92 mM, (Jezek & Blatt, Reference Jezek and Blatt2017). In open stomata after the midday, K+ is substituted by 50% with sugars (Talbott & Zeiger, Reference Talbott and Zeiger1996). In salinized barley leaves K+ dropped (substituted by Na+) both in the vacuole and the cytosol; these changes were dramatic (about 5-fold) in epidermal cells, and less prominent in mesophyll cells, where cytosolic K+ tends to be maintained at the expense of the vacuolar K+ pool (Cuin et al., Reference Cuin, Miller, Laurie and Leigh2003), thus improving cytosolic K+/Na+ ratio, underlying a higher photosynthetic activity and salt tolerance (James et al., Reference James, Munns, Von Caemmerer, Trejo, Miller and Condon2006). The refilling of the cytosol by vacuolar K+ is most pronounced in the case of K+ deficiency: the vacuolar K+ activity can drop 10-fold, whereas the cytosolic K+ activity remains almost constant (Walker et al., Reference Walker, Leigh and Miller1996).

There are three types of ion channels, which mediate passive K+ fluxes across the tonoplast: (i) VK (K+-selective) channel, mainly encoded by TPK1, (ii) slow vacuolar (SV) channels; non-selective cation channels encoded by TPC1 and (iii) fast vacuolar (FV) channels (also non-selective; with unknown molecular identity). Only VK/TPK channels, but not SV/TPC1 or FV, can discriminate between K+ and Na+ (Brüggemann et al., Reference Brüggemann, Pottosin and Schönknecht1999a; Gobert et al., Reference Gobert, Isayenkov, Voelker, Czempinski and Maatthuis2007; Pottosin et al., Reference Pottosin, Martínez-Estévez, Dobrovinskaya and Muñiz2003; Pottosin & Dobrovinskaya, Reference Pottosin and Dobrovinskaya2014; Ward & Schroeder, Reference Ward and Schroeder1994). FV and SV channel activity and/or conductance is modulated by a variety of cytosolic and vacuolar factors, including Ca2+ (and Mg2+), voltage and polyamines. FV activity is suppressed by cytosolic Ca2+, Mg2+ and polyamines in the micromolar–submillimolar range (Brüggemann et al., Reference Brüggemann, Pottosin and Schönknecht1998, Reference Brüggemann, Pottosin and Schönknecht1999b; Dobrovinskaya et al., Reference Dobrovinskaya, Muñiz and Pottosin1999; Tikhonova et al., Reference Tikhonova, Pottosin, Dietz and Schönknecht1997). In the virtual absence of these cations at both tonoplast sides and symmetric K+ in a physiological range, it is activated at an increased voltage difference of either sign (Allen et al., Reference Allen, Amtmann and Sanders1998; Bonales-Alatorre et al., Reference Bonales-Alatorre, Shabala, Chen and Pottosin2013; Pottosin et al., Reference Pottosin, Martínez-Estévez, Dobrovinskaya and Muñiz2003; Tikhonova et al., Reference Tikhonova, Pottosin, Dietz and Schönknecht1997), with a minimum open probability at about resting tonoplast potential of −20 to −30 mV (Miller et al., Reference Miller, Cookson, Smith and Wells2001; Walker et al., Reference Walker, Leigh and Miller1996; Wang et al., Reference Wang, Dindas, Rienmuller, Krebs, Schumacher, Wu, Hedrich and Roelfsema2015). Vacuolar polycations at millimolar concentrations convert FV into an outward rectifier, suppressing the cytosol-directed K+ flux (Tikhonova et al., Reference Tikhonova, Pottosin, Dietz and Schönknecht1997). The SV/TPC1 channel is activated at cytosol-positive voltage and its opening requires high (ten micromolar) cytosolic Ca2+, whereas vacuolar Ca2+ (and Mg2+) locks the channel in the resting state, thus increasing the threshold for its voltage activation (reviewed in Pottosin & Dobrovinskaya, Reference Pottosin and Dobrovinskaya2022). VK/TPK channels are mainly expressed in guard cells and to a lesser extent in root and mesophyll, yielding 10 to 200 VK copies per vacuole (Pottosin et al., Reference Pottosin, Martínez-Estévez, Dobrovinskaya and Muñiz2003; Tang et al., Reference Tang, Zhao, Yang, Wang, Li, Kleist, Lemaux and Luan2020; Ward & Schroeder, Reference Ward and Schroeder1994), whereas a typical vacuole contains thousands of active copies of SV/TPC1 and FV channels. Yet, due to the fact that VK/TPK channels are voltage-independent, insensitive to Mg2+ and polyamines and only require elevated (>1 μM) cytosolic Ca2+, which can activate the channel directly or via the CBL-CIPK pathway (Allen et al., Reference Allen, Amtmann and Sanders1998; Pottosin et al., Reference Pottosin, Martínez-Estévez, Dobrovinskaya and Muñiz2003; Tang et al., Reference Tang, Zhao, Yang, Wang, Li, Kleist, Lemaux and Luan2020; Ward & Schroeder, Reference Ward and Schroeder1994), it is conceivable that they can dominate K+ flux from vacuole at moderate cytosolic Ca2+ elevations. Strict negative control of SV/TPC1 channels by physiologically relevant factors can question its direct contribution to tonoplast K+ and Ca2+ fluxes. However, it has been shown that in vivo SV/TPC1 activation by a depolarization is transmitted in some way to the VK/TPK activation, mediating vacuolar K+ release by the latter, as if the activity of these two tonoplast channels is functionally coupled (Jaslan et al., Reference Jaslan, Dreyer, Lu, O'Malley, Dindas, Marten and Hedrich2019). It is conceivable that such a transduction occurs in a direct mechanical way so that the movement of the voltage sensor of the SV/TPC1 channels favours the activation of associated VK/TPK ones; for such mechanism full activation (opening) of SV/TPC1 channels, which is only observed at rather high cytosolic Ca2+, is not necessary so that the channel protein may operate in a non-canonical ion flux-independent way (Pottosin & Dobrovinskaya, Reference Pottosin and Dobrovinskaya2022).

VK/TPK channels were proven to play key roles in vacuolar K+ release upon ABA-induced stomatal closure (Gobert et al., Reference Gobert, Isayenkov, Voelker, Czempinski and Maatthuis2007) and vacuolar K+ remobilization upon K+ starvation (Gu et al., Reference Gu, Lu, Ren and Lu2024; Tang et al., Reference Tang, Zhao, Yang, Wang, Li, Kleist, Lemaux and Luan2020). VK/TPK activity is postulated to be predominant over the activity of non-selective SV/TPC1 and FV channels under salt stress, not only for the means of refilling the cytosolic K+ pool but also for a proper Na+ compartmentalization, avoiding a futile energy-consuming vacuolar Na+ cycling (Shabala et al., Reference Shabala, Chen, Chen and Pottosin2020; Wu et al., Reference Wu, Zhang, Giraldo and Shabala2018).

Vacuolar K+ dynamics appears to be controlled by the vacuolar K+ content. Firstly, upon the increase of the K+ concentration in the growth medium, K+ cellular (principally, vacuolar) content tends to approach the upper limit of 100–250 mM when external K+ reaches 24–95 μM (species-dependent) and does not increase upon further increase of external K+ to 1 mM (Leigh, Reference Leigh2001). On the other hand, under stomatal closure, induced by low and optimal ABA concentrations, the release of the vacuolar K+ analogue 86Rb, albeit substantially slower at low ABA dosage, ceased when the same loss of vacuolar tracer resulted as if “different numbers of channels are activated at high and low amounts of ABA, but the behaviour of individual ion channels is sensitive to ion content (volume)” (MacRobbie, Reference MacRobbie1995, Reference MacRobbie1998). MacRobbie opted for a volume/ stretch sensitive K+ channel and osmo-/stretch sensitivity was indeed demonstrated for TPK1 channels from diverse plant species (Maathuis, Reference Maathuis2011). In contrast to FV and SV/TPC1 channels, which are equally permeable for Rb+ and K+ (Brüggemann et al., Reference Brüggemann, Pottosin and Schönknecht1999a; Pottosin & Dobrovinskaya, Reference Pottosin and Dobrovinskaya2014), VK/TPK channels hardly conduct Rb+ (Gobert et al., Reference Gobert, Isayenkov, Voelker, Czempinski and Maatthuis2007; Ward & Schroeder, Reference Ward and Schroeder1994). Yet, a significant role of SV/TPC1 channels in ABA-induced stomata closure seems to be ruled out (Islam et al., Reference Islam, Munemasa, Hossain, Nakamura, Mori and Murata2010; Peiter et al., Reference Peiter, Maathuis, Mills, Knight, Pelloux, Hetherington and Sanders2005), but what about FV channels? For the sake of completeness, one needs to take also into account the coupling between the fluxes of K+ and small anions (Cl), governing stomatal movements, not only for the same membrane, but also between plasma membrane and tonoplast. Indeed, plasma membrane and tonoplast communicate across the common space, cytosol. Thus, the mutation of the main anion pathway in the plasma membrane, formed by SLAC1 channel, predictably slows down the global transport across both membranes in series (Horaruang et al., Reference Horaruang, Hills and Blatt2020). Voltage dependence of the outwardly rectifying K+ channel GORK in the guard cell plasma membrane senses external K+, which changes manifold upon stomatal close-open transitions (Jezek & Blatt, Reference Jezek and Blatt2017). This K+ modulation ensures that at any external K+ GORK channels will mediate K+ efflux (Dreyer & Blatt, Reference Dreyer and Blatt2009). However, such a mechanism defines the direction of K+ flux but not the condition of its cessation. A more compatible K+-sensing mechanism, which is the voltage-independent control of K+ channels activity by cytosolic K+ was described for SKOR, another outward-rectifying K+ channel, with a high degree of homology to GORK (Liu et al., Reference Liu, Li and Luan2006). But to the best of our knowledge, yet such a mechanism has not been confirmed for GORK. Thus, next we consider K+ sensing mechanisms of the major tonoplast cation channels.

A focused study demonstrated that FV channels sense both vacuolar and cytosolic K+ changes (Pottosin & Martínez-Estévez, Reference Pottosin and Martínez-Estévez2003). An increase in the cytosolic K+ resulted in a shift of the whole voltage dependence in parallel with the EK shift. In contrast, increase of vacuolar K+ only affected the voltage-dependent process at cytosol negative potentials, decreasing the respective voltage threshold and increasing channel’s open probability, hence potentially inducing vacuolar K+ release. Overall, this analysis demonstrated that at tonoplast attainable potentials the FV activity was almost independent of cytosolic and strongly potentiated by vacuolar K+ (Pottosin & Martínez-Estévez, Reference Pottosin and Martínez-Estévez2003). Meanwhile, SV/TPC1 was also sensitive to vacuolar K+: in the virtual absence of vacuolar divalent cations an increase of vacuolar K+ resulted in the increase of the threshold for voltage activation in parallel with the right-hand EK shift (Ivashikina & Hedrich, Reference Ivashikina and Hedrich2005; Pottosin et al., Reference Pottosin, Martínez-Estévez, Dobrovinskaya and Muñiz2005). Substitution of K+ with NMDG+ demonstrated that this shift was caused by shielding of the negative membrane surface charge. Yet, the effect was beyond the totally unspecific cation binding, so that vacuolar Na+ caused a larger shift as compared to K+ (Ivashikina & Hedrich, Reference Ivashikina and Hedrich2005; Pottosin & Dobrovinskaya, Reference Pottosin and Dobrovinskaya2014). It was speculated then that an increase of the threshold for voltage activation by high vacuolar Na+ can be a mechanism for the prevention of a Na+ leak from the vacuole at salt stress (Ivashikina & Hedrich, Reference Ivashikina and Hedrich2005). Yet, in the presence of submillimolar Ca2+ at the luminal side the direction of the K+ (Na+)-induced shift of the voltage dependence is reversed, so that the channel tends to open at more cytosol negative potentials. Luminal Ca2+ is a principle negative regulator of the SV/TPC1 activity, and shielding the negative surface charge by K+, and, more strongly, by Na+, decreases the local Ca2+ concentration in the vicinity of the membrane surface/luminal Ca2+ binding site and alleviates the Ca2+ inhibitory effect (Pérez et al., Reference Pérez, Wherrett, Shabala, Muñiz, Dobrovinskaya and Pottosin2008; Pottosin et al., Reference Pottosin, Martínez-Estévez, Dobrovinskaya and Muñiz2005; Pottosin & Dobrovinskaya, Reference Pottosin and Dobrovinskaya2014). So that, at physiological vacuolar Ca2+, an increased content of Na+ in the lumen would promote vacuolar Na+ release via SV/TPC1 channels. The same appears to be true for FV channels (Bonales-Alatorre et al., Reference Bonales-Alatorre, Shabala, Chen and Pottosin2013). To summarize, at non-salinized conditions, when K+ is a dominant cellular cation, both FV and SV/TPC1 channels act as vacuolar K+ sensors, so that their open probability increases with the increase of vacuolar K+ and decreases when vacuolar K+ dropped. The latter can be suitable for a condition of extreme K+ deficiency, when the electrochemical K+ gradient is vacuole-directed (Walker et al., Reference Walker, Leigh and Miller1996), to diminish channel-mediated K+ re-absorption by the vacuole (Figure 4). Yet, this sensing mechanism is not selective for K+ vs Na+, so that at salt stress additional regulatory factors need to be invoked to suppress the non-selective channels´ activity and reduce the vacuolar Na+ leak (Pottosin et al., Reference Pottosin, Olivas-Aguirre, Dobrovinskaya, Zepeda-Jazo and Shabala2021; Shabala et al., Reference Shabala, Chen, Chen and Pottosin2020).

Figure 4. Potassium permeable channels of FV and SV types act as safety valves, preventing K+ overload at excessive K+; vacuolar K+ filling is mediated by K+ (Na+)/ H+ antiporter of NHX type (left). Suppression of the activity of FV and SV channels by low vacuolar K+ prevents channel-mediated vacuolar K+ re-uptake and allows the maintenance of cytosolic K+ at the expense of the vacuolar K+ pool. Cytosol refilling by vacuolar K+ can be mediated by a hypothetical K+: H+ symporter (right). Note that in the K+ replete conditions a significant ‘futile’ H+ and K+ cycling results, whereas under K+ deficiency (upon the change in the direction of transtonoplast K+ gradient) this cycling is reduced. Above transport system acts as K+ sensor/effector at non-saline conditions but is unable per se to discriminate between K+ and Na+ under high salinity (see text).

6. Conclusion and outlook

Potassium is the most abundant and probably also the most versatile cation in plants. Besides its rather static role as an osmotic agent and charge balancer, there is more and more evidence that it also plays significant roles in intra-plant and intra-organelle communication. In addition, potassium gradients can also serve as an important supporting energy source, for example, for phloem loading. Although our knowledge of K in plants seems to be by far the most advanced of all relevant ions, there are still many open questions that need to be answered in future research. Status of current knowledge provides the basis for a comprehensive quantitative modelling approach including properties of membrane transport proteins, volume flow and regulatory features of K+. Recent computational modelling revealed the role of apparently futile K+ and H+ cycling for intracellular K+ homeostasis and steady state membrane potential maintenance. These conclusions received an experimental confirmation, but more wet-lab work is needed to support them. The loss of K+ is a common denominator of abiotic stress responses and there is an emerging view for its physiological importance as an energy switch from metabolism to defence needs. There is contrasting behaviour of different plant species and tissues in handling of intracellular K+ and membrane potential in a response to stress, which raises the question of the specificity of K+ sensing, decoding and downstream signalling. Some plausible K+ sensing and decoding mechanisms are addressed in this review. Several K+ and K+-conducting channels and transporters such as SKOR, HAK5 and vacuolar SV/TPC1 and FV may be plausible candidates to act as K+ sensors and effectors regulating the activity of target proteins. Therefore, the puzzle of K+ homeostasis and signalling in plants is gaining shape. Obviously, to complete this job, a systemic modelling approach with focused experimentation are required.

Data availability statement

Data used for quantitative analysis are available on request.

Author contributions

S.S. conceived the concept. All authors wrote the article.

Funding statement

L.H.W. acknowledges a support by the National Science Foundation of China (Grant No. 32070277), I.D. 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) and by Fondo Nacional de Desarrollo Científico, Tecnológico y de Innovación Tecnológica (FONDECYT/Chile, Grant No. 1220504).

Competing interest

The authors declare no competing interests.

Open peer review

To view the open peer review materials for this article, please visit http://doi.org/10.1017/qpb.2025.10.

Footnotes

Associate Editor: Dale Sanders

References

Ache, P., Becker, D., Deeken, R., Dreyer, I., Weber, H., Fromm, J., & Hedrich, R. (2001). VFK1, a Vicia faba K+ channel involved in phloem unloading. The Plant Journal, 27(6), 571580. https://doi.org/10.1046/j.1365-313X.2001.t01-1-01116.x CrossRefGoogle ScholarPubMed
Ahmed, H. A. I., Shabala, L., & Shabala, S. (2021). Tissue-specificity of ROS-induced K+ and Ca2+ fluxes in succulent stems of the perennial halophyte Sarcocornia quinqueflora in the context of salinity stress tolerance. Plant Physiology and Biochemistry, 166, 10221031. https://doi.org/10.1016/j.plaphy.2021.07.006 CrossRefGoogle ScholarPubMed
Allen, G. H., Amtmann, A., & Sanders, D. (1998). Calcium-dependent and calcium-independent K+ mobilization channels in Vicia faba guard cell vacuoles. Journal of Experimental Botany, 49(Special Issue), 305318. https://doi.org/10.1093/jxb/49.Special_Issue.305 CrossRefGoogle Scholar
Amir, S., & Reinhold, L. (1971). Interaction between K-deficiency and light in 14 C-sucrose translocation in bean plants. Physiologia Plantarum, 24(2), 226231. https://doi.org/10.1111/j.1399-3054.1971.tb03483.x CrossRefGoogle Scholar
Anschütz, U., Becker, D., & Shabala, S. (2014). Going beyond nutrition: Regulation of potassium homoeostasis as a common denominator of plant adaptive responses to environment. Journal of Plant Physiology, 171(9), 670687. https://doi.org/10.1016/j.jplph.2014.01.009 CrossRefGoogle ScholarPubMed
Bañuelos, M. A., Garciadeblas, B., Cubero, B., & Rodríguez-Navarro, A. (2002). Inventory and functional characterization of the HAK potassium transporters of rice. Plant Physiology, 130, 784795. https://doi.org/10.1104/pp.007781 CrossRefGoogle ScholarPubMed
Behera, S., Long, Y., Schmitz-Thom, I., Wang, X. P., Zhang, C. X., Li, H., Steinhorst, L., Manishankar, P., Ren, X. L., Offenborn, J. N., Wu, W. H., Kudla, J., & Wang, Y. (2017). Two spatially and temporally distinct Ca2+ signals convey Arabidopsis thaliana responses to K+ deficiency. New Phytologist, 213(2), 739750. https://doi.org/10.1111/nph.14145 CrossRefGoogle ScholarPubMed
Ben Zioni, A. B., Vaadia, Y., & Lips, S. H. (1971). Nitrate uptake by roots as regulated by nitrate reduction products of the shoot. Physiologia Plantarum, 24(2), 288290. https://doi.org/10.1111/j.1399-3054.1971.tb03493.x CrossRefGoogle Scholar
Benito, B., Haro, R., Amtmann, A., Cuin, T. A., & Dreyer, I. (2014). The twins K+ and Na+ in plants. Journal of Plant Physiology, 171(9), 723731. https://doi.org/10.1016/j.jplph.2013.10.014 CrossRefGoogle ScholarPubMed
Blanco, E., Fortunato, S., Viggiano, L., & de Pinto, M. C. (2020). Cyclic AMP: A polyhedral signalling molecule in plants. International Journal of Molecular Sciences, 21(14), Article 4862. https://doi.org/486210.3390/ijms21144862 CrossRefGoogle ScholarPubMed
Blatt, M. R., & Slayman, C. L. (1987). Role of active potassium-transport in the regulation of cytoplasmic pH by nonanimal cells. Proceedings of the National Academy of Sciences USA, 84, 27372741. https://doi.org/10.1073/pnas.84.9.273 CrossRefGoogle ScholarPubMed
Bonales-Alatorre, E., Shabala, S., Chen, Z. H., & Pottosin, I. (2013). Reduced tonoplast fast-activating and slow-activating channel activity is essential for conferring salinity tolerance in a facultative halophyte, quinoa. Plant Physiology, 162(2), 940952. https://doi.org/10.1104/pp.113.216572 CrossRefGoogle Scholar
Bose, J., Rodrigo-Moreno, A., Lai, D., Xie, Y., Shen, W., & Shabala, S. (2015). Rapid regulation of the plasma membrane H+-ATPase activity is essential to salinity tolerance in two halophyte species, Atriplex lentiformis and Chenopodium quinoa . Annals of Botany, 115(3), 481494. https://doi.org/10.1093/aob/mcu219 CrossRefGoogle ScholarPubMed
Bose, J., Shabala, L., Pottosin, I., Zeng, F., Velarde-Buendía, A. M., Massart, A., Poschenrieder, C., Hariadi, Y., & Shabala, S. (2014). Kinetics of xylem loading, membrane potential maintenance, and sensitivity of K+-permeable channels to reactive oxygen species: Physiological traits that differentiate salinity tolerance between pea and barley. Plant, Cell & Environment, 37(3), 589600. https://doi.org/10.1111/pce.12180 CrossRefGoogle ScholarPubMed
Britto, D. T., & Kronzucker, H. J. (2006). Futile cycling at the plasma membrane: A hallmark of low-affinity nutrient transport. Trends in Plant Science, 11(11), 529534. https://doi.org/10.1016/j.tplants.2006.09.011 CrossRefGoogle ScholarPubMed
Britto, D. T., Coskun, D., & Kronzucker, H. J. (2021). Potassium physiology from Archean to Holocene: A higher-plant perspective. Journal of Plant Physiology, 262, Article 153432. https://doi.org/10.1016/j.jplph.2021.153432 CrossRefGoogle ScholarPubMed
Brüggemann, L. I., Pottosin, I. I., & Schönknecht, G. (1998). Cytoplasmic polyamines block the fast-activating vacuolar cation channel. The Plant Journal, 16(1) 101105. https://doi.org/10.1046/j.1365-313x.1998.00274.x CrossRefGoogle Scholar
Brüggemann, L. I., Pottosin, I. I., & Schönknecht, G. (1999a). Selectivity of the fast-activating vacuolar cation channel. Journal of Experimental Botany, 50(335), 873876. https://doi.org/10.1093/jxb/50.335.873 Google Scholar
Brüggemann, L. I., Pottosin, I. I., & Schönknecht, G. (1999b). Cytoplasmic magnesium regulates the fast activating vacuolar cation channel. Journal of Experimental Botany, 50(339), 15471552. https://doi.org/10.1093/jxb/50.339.1547 CrossRefGoogle Scholar
Buch-Pedersen, M. J., Rudashevskaya, E. L., Berner, T. S., Venema, K., & Palmgren, M. G. (2006). Potassium as an intrinsic uncoupler of the plasma membrane H+-ATPase. Journal of Biological Chemistry, 281(50), 3828538292. https://doi.org/10.1074/jbc.M604781200 CrossRefGoogle Scholar
Cakmak, I., & Rengel, Z. (2024). Potassium may mitigate drought stress by increasing stem carbohydrates and their mobilization into grains. Journal of Plant Physiology, 303, Article 154325. https://doi.org/10.1016/j.jplph.2024.154325 CrossRefGoogle ScholarPubMed
Chakraborty, K., Bose, J., Shabala, L., & Shabala, S. (2016). Difference in root K+ retention ability and reduced sensitivity of K+-permeable channels to reactive oxygen species confer differential salt tolerance in three Brassica species. Journal of Experimental Botany, 67(15), 46114625. https://doi.org/10.1093/jxb/erw236 CrossRefGoogle Scholar
Chao, M. N., Li, L. B., Zhang, J. Y., Huang, L., Ren, R., Xu, X. J., & Huang, Z. W. (2024). Structural features and expression regulation analysis of potassium transporter gene GmHAK5 in soybean (Glycine max). Plant Growth Regulation, 102, 471483. https://doi.org/10.1007/s10725-023-01079-w CrossRefGoogle Scholar
Chen, Z., Pottosin, I. I., Cuin, T. A., Fuglsang, A. T., Tester, M., Jha, D., Zepeda-Jazo, I., Zhou, M., Palmgren, M. G., Newman, I. A., & Shabala, S. (2007). Root plasma membrane transporters controlling K+/Na+ homeostasis in salt-stressed barley. Plant Physiology, 145(4) 17141725. https://doi.org/10.1104/pp.107.110262 CrossRefGoogle ScholarPubMed
Cheong, Y. H., Pandey, G. K., Grant, J. J., Batistic, O., Li, L., Kim, B. G., Lee, S. C., Kudla, J., & Luan, S. (2007). Two calcineurin B-like calcium sensors, interacting with protein kinase CIPK23, regulate leaf transpiration and root potassium uptake in Arabidopsis. Plant Journal, 52, 223239. https://doi.org/10.1111/j.1365-313X.2007.03236.x CrossRefGoogle ScholarPubMed
Chérel, I., Lefoulon, C., Boeglin, M., & Sentenac, H. (2014). Molecular mechanisms involved in plant adaptation to low K+ availability. Journal of Experimental Botany, 65, 833848. https://doi.org/10.1093/jxb/ert402 CrossRefGoogle Scholar
Contador-Álvarez, L., Rojas-Rocco, T., Rodríguez-Gómez, T., Rubio-Meléndez, M. E., Riedelsberger, J., Michard, E., Dreyer, I. (2025). Dynamics of homeostats: The basis of electrical, chemical, hydraulic, pH and calcium signaling in plants. Quantitative Plant Biology, 6, eX, 110. https://doi.org/10.1017/qpb.2025.6 CrossRefGoogle ScholarPubMed
Conway, E. J., & O'Malley, E. (1944). Nature of the cation exchanges during short-period yeast fermentation. Nature, 153, 555556. https://doi.org/10.1038/153555b0 CrossRefGoogle Scholar
Cui, J., & Tcherkez, G. (2021). Potassium dependency of enzymes in plant primary metabolism. Plant Physiology and Biochemistry, 166, 522530. https://doi.org/10.1016/j.plaphy.2021.06.017 CrossRefGoogle ScholarPubMed
Cuin, T. A., Betts, S. A., Chalmandrier, R., & Shabala, S. (2008). A root’s ability to retain K+ correlates with salt tolerance in wheat. Journal of Experimental Botany, 59(10), 26972706. https://doi.org/10.1093/jxb/ern128 CrossRefGoogle ScholarPubMed
Cuin, T. A., Miller, A. J., Laurie, S. A., & Leigh, R. (2003). Potassium activities in cell compartments of salt-grown barley leaves. Journal of Experimental Botany, 54(383), 657661. https://doi.org/10.1093/jxb/erg072 CrossRefGoogle ScholarPubMed
de Oliveira, R. R., Morales-Neto, R., Rocco, S. A., Sforca, M. L., Polo, C. C., Tonoli, C. C. C., Mercaldi, G. F., Cordeiro, A. T., Murakami, M. T., & Franchini, K. G. (2018). Adenosine Kinase couples sensing of cellular potassium depletion to purine metabolism. Scientific Reports, 8(1), Article 11988. https://doi.org/10.1038/s41598-018-30418-5.CrossRefGoogle ScholarPubMed
Deeken, R., Geiger, D., Fromm, J., Koroleva, O., Ache, P., Langenfeld-Heyser, R., Sauer, N., May, S., & Hedrich, R. (2002). Loss of the AKT2/3 potassium channel affects sugar loading into the phloem of Arabidopsis. Planta, 216(2), 334344. https://doi.org/10.1007/s00425-002-0895-1 CrossRefGoogle ScholarPubMed
Demidchik, V. (2014). Mechanisms and physiological roles of K+ efflux from root cells. Journal of Plant Physiology, 171(9), 696707. https://doi.org/10.1016/j.jplph.2014.01.015 CrossRefGoogle ScholarPubMed
Demidchik, V., & Maathuis, F. J. M. (2007). Physiological roles of nonselective cation channels in plants: From salt stress to signalling and development. New Phytologist, 175(3), 387404. https://doi.org/10.1111/j.1469-8137.2007.02128.x CrossRefGoogle ScholarPubMed
Demidchik, V., Cuin, T. A., Svistunenko, D., Smith, S. J., Miller, A. J., Shabala, S., Sokolik, A., & Yurin, V. (2010). Arabidopsis root K+-efflux conductance activated by hydroxyl radicals: Single-channel properties, genetic basis and involvement in stress-induced cell death. Journal of Cell Science, 123(Pt 9), 14681479. https://doi.org/10.1242/jcs.064352 CrossRefGoogle ScholarPubMed
Dobrovinskaya, O. R., Muñiz, J., & Pottosin, I. I. (1999). Inhibition of vacuolar ion channels by polyamines. Journal of Membrane Biology, 167(2), 127140. https://doi.org/10.1007/s002329900477 CrossRefGoogle ScholarPubMed
Dong, Q. Y., Wallrad, L., Almutairi, B. O., & Kudla, J. (2022). Ca2+ signaling in plant responses to abiotic stresses. Journal of Integrative Plant Biology, 64(2), 287300. https://doi.org/10.1111/jipb.13228 CrossRefGoogle Scholar
Drechsler, N., Zheng, Y., Bohner, A., Nobmann, B., von Wirén, N., Kunze, R., & Rausch, C. (2015). Nitrate-dependent control of shoot K homeostasis by the nitrate transporter1/peptide transporter family member NPF7. 3/NRT1. 5 and the stelar K+ outward rectifier SKOR in Arabidopsis. Plant Physiology, 169(4), 28322847. https://doi.org/10.1104/pp.15.01152 Google ScholarPubMed
Drew, M. C., & Saker, L. R. (1984). Uptake and long-distance transport of phosphate, potassium and chloride in relation to internal ion concentrations in barley: Evidence of non-allosteric regulation. Planta, 160(6), 500507. https://doi.org/10.1007/BF00411137 CrossRefGoogle ScholarPubMed
Drew, M. C., Webb, J., & Saker, L. R. (1990). Regulation of K+ uptake and transport to the xylem in barley roots; K+ distribution determined by electron probe X-ray microanalysis of frozen-hydrated cells. Journal of Experimental Botany, 41(7), 815825. https://doi.org/10.1093/jxb/41.7.815 CrossRefGoogle Scholar
Dreyer, I. (2021). Nutrient cycling is an important mechanism for homeostasis in plant cells. Plant Physiology, 187(4), 22462261. https://doi.org/10.1093/plphys/kiab217 CrossRefGoogle Scholar
Dreyer, I., Michard, E., Lacombe, B., Thibaud, J.-B. (2001). A plant Shaker-like K+ channel switches between two distinct gating modes resulting in either inward-rectifying or ‘leak’ current. FEBS Letters, 505(2), 233239. https://doi.org/10.1016/j.tplants.2009.04.001 CrossRefGoogle ScholarPubMed
Dreyer, I., & Blatt, M. R. (2009). What makes a gate? The ins and outs of Kv-like K+ channels in plants. Trends in Plant Science, 14, 383390.CrossRefGoogle ScholarPubMed
Dreyer, I., Gomez-Porras, J. L., & Riedelsberger, J. (2017). The potassium battery: A mobile energy source for transport processes in plant vascular tissues. New Phytologist, 216(4), 10491053. https://doi.org/10.1111/nph.14667 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(4), Article 104078. https://doi.org/10.1016/J.ISCI.2022.104078 CrossRefGoogle ScholarPubMed
Dreyer, I., Vergara-Valladares, F., Mérida-Quesada, F., Rubio-Meléndez, M. E., Hernández-Rojas, N., Riedelsberger, J., Astola-Mariscal, S. Z., Heitmüller, C., Yanez-Chávez, M., Arrey-Salas, O., San Martín-Davison, A., Navarro-Retamal, C., & Michard, E. (2023). The surprising dynamics of electrochemical coupling at membrane sandwiches in plants. Plants, 12(1), 204. https://doi.org/10.3390/PLANTS12010204 CrossRefGoogle ScholarPubMed
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
Engels, C., & Kirkby, E. A. (2001). Cycling of nitrogen and potassium between shoot and roots in maize as affected by shoot and root growth. Journal of Plant Nutrition and Soil Science, 164(2), 183191.3.0.CO;2-B>CrossRefGoogle Scholar
Evans, H. J., & Sorger, G. J. (1966). Role of mineral elements with emphasis on univalent cations. Annual Review of Plant Physiology, 17, 4776 https://doi.org/10.1146/annurev.pp.17.060166.000403 CrossRefGoogle Scholar
Gajdanowicz, P., Michard, E., Sandmann, M., Rocha, M., Corrêa, L. G. G., Ramírez-Aguilar, S. J., Gomez-Porras, J. L., González, W., Thibaud, J.-B., van Dongen, J. T., & Dreyer, I. (2011). Potassium (K+) gradients serve as a mobile energy source in plant vascular tissues. Proceedings of the National Academy of Sciences of the United States of America, 108(2), 864869. https://doi.org/10.1073/pnas.1009777108 CrossRefGoogle ScholarPubMed
Gaymard, F., Pilot, G., Lacombe, B., Bouchez, D., Bruneau, D., Boucherez, J., Michaux-Ferrière, N., Thibaud, J. B., & Sentenac, H. (1998). Identification and disruption of a plant shaker-like outward channel involved in K+ release into the xylem sap. Cell, 94(5), 647655. https://doi.org/10.1016/s0092-8674(00)81606-2 CrossRefGoogle Scholar
Geiger, D., Becker, D., Vosloh, D., Gambale, F., Palme, K., Rehers, M., Anschuetz, U., Dreyer, I., Kudla, J., & Hedrich, R. (2009). Heteromeric AtKC1.AKT1 channels in Arabidopsis roots facilitate growth under K+-limiting conditions. Journal of Biological Chemistry, 284, 2128821295. https://doi.org/10.1074/jbc.M109.017574 CrossRefGoogle Scholar
Geisler, M., & Dreyer, I. (2024). An auxin homeostat allows plant cells to establish and control defined transmembrane auxin gradients. New Phytologist, 244, 14221436. https://doi.org/10.1111/nph.20120 CrossRefGoogle ScholarPubMed
Gelli, A., & Blumwald, E. (1997). Hyperpolarization-activated Ca2+-permeable channels in the plasma membrane of tomato cells. Journal of Membrane Biology, 155, 3545. https://doi.org/10.1007/s002329900156 CrossRefGoogle ScholarPubMed
Gobert, A., Isayenkov, S., Voelker, C., Czempinski, K., & Maatthuis, F. J. M. (2007). The two-pore channel TPK1 gene encodes the vacuolar K+ conductance and plays a role in K+ homeostasis. Proceedings of the National Academy of Sciences USA, 104(25), 1072610731. https://doi.org/10.1073/pnas.0702595104 CrossRefGoogle Scholar
Grabov, A., & Blatt, M. R. (1998). Membrane voltage initiates Ca2+ waves and potentiates ca2+ increases with abscisic acid in stomatal guard cells. Proceedings of the National Academy of Sciences USA, 95, 47784783. https://doi.org/10.1073/pnas.95.8.4778 CrossRefGoogle ScholarPubMed
Gu, H., Lu, Z., Ren, T., & Lu, J. (2024). From hidden to visual K+ deficiency in plants: Integration of subcellular K+ distribution and photosynthetic performance. Crop and Environment, 3(2), 8490. https://doi.org/10.1016/j.crope.2024.02.001 CrossRefGoogle Scholar
Hauser, F., & Horie, T. (2010). A conserved primary salt tolerance mechanism mediated by HKT transporters: A mechanism for sodium exclusion and maintenance of high K+/Na+ ratio in leaves during salinity stress. Plant, Cell & Environment, 33(4), 552565. https://doi.org/10.1111/j.1365-3040.2009.02056.x CrossRefGoogle Scholar
Horaruang, W., Hills, A., & Blatt, M. R. (2020). Communication between the plasma membrane and tonoplast is an emergent property of ion transport. Plant Physiology, 182, 18331835. https://doi.org/10.1104/pp.19.01485 CrossRefGoogle Scholar
Islam, M. M., Munemasa, S., Hossain, M. A., Nakamura, Y., Mori, I. C., & Murata, Y. (2010). Roles of AtTPC1, vacuolar Two Pore Channel 1, in Arabidopsis stomatal closure. Plant & Cell Physiology, 51(2), 302311. https://doi.org/10.1093/pcp/pcq001 CrossRefGoogle ScholarPubMed
Ivashikina, N., & Hedrich, R. (2005). K+ currents through SV-type vacuolar channels are sensitive to elevated luminal sodium levels. The Plant Journal, 41(4), 606614. https://doi.org/10.1111/j.1365-313X.2004.02324.x CrossRefGoogle ScholarPubMed
James, R. A., Munns, R., Von Caemmerer, S., Trejo, C., Miller, C., & Condon, C. A. G. (2006). Photosynthetic capacity is related to the cellular and subcellular partitioning of Na+, K+ and Cl- in salt-affected barley and durum wheat. Plant Cell and Environment, 29(12), 21855197. https://doi.org/10.1111/j.1365-3040.2006.01592.x CrossRefGoogle Scholar
Jaslan, D., Dreyer, I., Lu, J. P., O'Malley, R., Dindas, J., Marten, I., & Hedrich, R. (2019). Voltage-dependent gating of SV channel TPC1 confers vacuole excitability. Nature Communications, 10(1), 2659. https://doi.org/10.1038/s41467-019-10599-x CrossRefGoogle ScholarPubMed
Jegla, T., Busey, G., & Assmann, S. M. (2018). Evolution and structural characteristics of plant voltage-gated K+ channels. Plant Cell, 30(12). 28982909. https://doi.org/10.1105/tpc.18.00523 CrossRefGoogle ScholarPubMed
Jeschke, W. D., & Pate, J. S. (1991). Cation and chloride partitioning through xylem and phloem within the whole plant of Ricinus communis L. under conditions of salt stress. Journal of Experimental Botany, 42(9), 11051116. https://doi.org/10.1093/jxb/42.9.1105 CrossRefGoogle Scholar
Jezek, M., & Blatt, M. R. (2017). The membrane transport system of the guard cell and its integration for stomatal dynamics. Plant Physiology, 174, 487519. https://doi.org/10.1104/pp.16.01949 CrossRefGoogle ScholarPubMed
Jin, R., Yan, M. X., Li, G. H., Liu, M., Zhao, P., Zhang, Z., Zhang, Q. Q., Zhu, X. Y., Wang, J., Yu, Y. C., Zhang, A. J., Yang, J., & Tang, Z. H. (2024). Comparative physiological and transcriptome analysis between potassium-deficiency tolerant and sensitive sweetpotato genotypes in response to potassium-deficiency stress. BMC Genomics, 25, 61. https://doi.org/10.1186/s12864-023-09939-5 CrossRefGoogle ScholarPubMed
Johansson, I., Wulfetange, K., Porée, F., Michard, E., Gajdanowicz, P., Lacombe, B., Sentenac, H., Thibaud, J., Mueller-Roeber, B., Blatt, M. R., & Dreyer, I. (2006). External K+ modulates the activity of the Arabidopsis potassium channel SKOR via an unusual mechanism. The Plant Journal, 46(2), 269281. https://doi.org/10.1111/j.1365-313X.2006.02690.x CrossRefGoogle ScholarPubMed
Kaplan, B., Sherman, T., & Fromm, H. (2007). Cyclic nucleotide-gated channels in plants. FEBS Letters, 581(12), 22372246. https://doi.org/10.1016/j.febslet.2007.02.017 CrossRefGoogle ScholarPubMed
Keunecke, M., Lindner, B., Seydel, U., Schulz, A., & Hansen, U. P. (2001). Bundle sheath cells of small veins in maize leaves are the location of uptake from the xylem. Journal of Experimental Botany, 52(357), 709714. https://doi.org/10.1093/jexbot/52.357.70 CrossRefGoogle ScholarPubMed
Kim, D., Yanders, S., & Stacey, G. (2023). Salt stress releases extracellular ATP to activate purinergic signaling and inhibit plant growth. Plant Physiology, 193(3) 17531757. https://doi.org/10.1093/plphys/kiad429 CrossRefGoogle ScholarPubMed
Lang, A. (1983). Turgor-regulated translocation. Plant, Cell & Environment, 6(9), 683689. https://doi.org/10.1111/1365-3040.ep11589312 Google Scholar
Leigh, R. A. (1997). Solute composition of vacuoles. Advances in Botanical Research, 25, 171194. https://doi.org/10.1016/S0065-2296(08)60152-4 CrossRefGoogle Scholar
Leigh, R. A. (2001). Potassium homeostasis and membrane transport. Journal of Plant Nutrition and Soil Science, 164(2), 193198.3.0.CO;2-7>CrossRefGoogle Scholar
Lemtiri-Chlieh, F., & Berkowitz, G. A. (2004). Cyclic adenosine monophosphate regulates calcium channels in the plasma membrane of Arabidopsis leaf guard and mesophyll cells. Journal of Biological Chemistry, 279, 3530635312. https://doi.org/10.1074/jbc.M400311200 CrossRefGoogle ScholarPubMed
Lhamo, D., Wang, C., Gao, Q. F., & Luan, S. (2021). Recent advances in genome-wide analyses of plant potassium transporter families. Current Genomics, 22, 164180. https://doi.org/10.2174/1389202922666210225083634 CrossRefGoogle ScholarPubMed
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(4), 15251542. https://doi.org/10.1111/NPH.19436 CrossRefGoogle Scholar
Li, K. L., Tang, R. J., Wang, C., & Luan, S. (2023a). Potassium nutrient status drives posttranslational regulation of a low-K response network in Arabidopsis. Nature Communications, 14. https://doi.org/10.1038/s41467-023-35906-5.Google Scholar
Li, K. L., Xue, H., Tang, R. J., & Luan, S. (2023b). TORC pathway intersects with a calcium sensor kinase network to regulate potassium sensing in Arabidopsis. Proceedings of the National Academy of Sciences USA, 120(47), e2316011120. https://doi.org/10.1073/pnas.2316011120 CrossRefGoogle Scholar
Liu, K., Li, L., & Luan, S. (2006). Intracellular K+ sensing of SKOR, a Shaker-type K+ channel from Arabidopsis. The Plant Journal, 46(2), 260268. https://doi.org/10.1111/j.1365-313X.2006.02689.x CrossRefGoogle Scholar
Liu, W., Wang, Y. B., Zhang, Y. W., Li, W., Wang, C. J., Xu, R., Dai, H. Y., & Zhang, L. F. (2024). Characterization of the pyruvate kinase gene family in soybean and identification of a putative salt responsive gene GmPK21. BMC Genomics, 25(1), Article 88. https://doi.org/8810.1186/s12864-023-09929-7.CrossRefGoogle ScholarPubMed
Lohaus, G., Hussmann, M., Pennewiss, K., Schneider, H., Zhu, J.-J., & Sattelmacher, B. (2000). Solute balance of a maize (Zea mays L.) source leaf as affected by salt treatment with special emphasis on phloem retranslocation and ion leaching. Journal of Experimental Botany, 51(351), 17211732. https://doi.org/10.1093/jexbot/51.351.1721 CrossRefGoogle ScholarPubMed
Lynn, R., & Guynn, R. W. (1978). Equilibrium-constants under physiological conditions for reactions of succinyl coenzyme-A synthetase and hydrolysis of succinyl coenzyme-A to coenzyme-A and succinate. Journal of Biological Chemistry, 253, 25462553.CrossRefGoogle ScholarPubMed
Ma, N. N., Dong, L. N., Lu, W., Lu, J., Meng, Q. W., & Liu, P. (2020). Transcriptome analysis of maize seedling roots in response to nitrogen-, phosphorus-, and potassium deficiency. Plant Soil, 447, 637658. https://doi.org/10.1007/s11104-019-04385-3 CrossRefGoogle Scholar
Maathuis, F. J. M. (2011). Vacuolar two-pore K+ channels act as vacuolar osmosensors. New Phytologist, 191(1), 8491. https://doi.org/10.1111/j.1469-8137.2011.03664.x CrossRefGoogle ScholarPubMed
Maathuis, F. J. M., & Sanders, D. (1994). Mechanism of high-affinity potassium uptake in roots of Arabidopsis thaliana . Proceedings of the National Academy of Sciences USA, 91, 92729276. https://doi.org/10.1073/pnas.91.20.9272 CrossRefGoogle ScholarPubMed
Maathuis, F. J. M., & Sanders, D. (1993). Energization of potassium uptake in Arabidopsis thaliana . Planta, 191, 302307. https://doi.org/10.1007/BF00195686 CrossRefGoogle Scholar
MacRobbie, E. A. C. (1995). Effect of ABA on 86Rb+ fluxes at plasmalemma and tonoplast of stomatal guard cells. The Plant Journal, 7(5), 835843.CrossRefGoogle Scholar
MacRobbie, E. A. C. (1998). Signal transduction and ion channels in guard cells. Philosophical transactions of the Royal Society of London. Series B, Biological sciences, 353(1374), 14751488. https://doi.org/10.1098/rstb.1998.0303 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. https://doi.org/10.1038/s41467-024-52963-6 CrossRefGoogle Scholar
Michard, E., Lacombe, B., Porée, F., Mueller-Roeber, B., Sentenac, H., Thibaud, J. B., & Dreyer, I. (2005). A unique voltage sensor sensitizes the potassium channel AKT2 to phosphoregulation. Journal of General. Physiology, 126, 605617. https://doi.org/10.1085/jgp.200509413 CrossRefGoogle ScholarPubMed
Miller, A. J., Cookson, S. J., Smith, S. J., & Wells, D. M. (2001). The use of microelectrodes to investigate compartmentation and the transport of metabolized inorganic ions in plants. Journal of Experimental Botany, 52(356), 541549. https://doi.org/10.1093/jxb/52.356.541 CrossRefGoogle ScholarPubMed
Munns, R., Day, D. A., Fricke, W., Watt, M., Arsova, B., Barkla, B. J., Bose, J., Byrt, C. S., Chen, Z. H., Foster, K. J., Gilliham, M., Henderson, S. W., Jenkins, C. L. D., Kronzucker, H. J., Miklavcic, S. J., Plett, D., Roy, S. J., Shabala, S., Shelden, M. C., … Tyerman, S. D. (2020). Energy costs of salt tolerance in crop plants. New Phytologist, 225(3), 10721090.CrossRefGoogle ScholarPubMed
Nieves-Cordones, M., Alemán, F., Martínez, V., & Rubio, F. (2010). The Arabidopsis thaliana HAK5 K+ transporter is required for plant growth and K+ acquisition from low K+ solutions under saline conditions. Molecular Plant, 3, 326333. https://doi.org/10.1093/mp/ssp102 CrossRefGoogle Scholar
Nieves-Cordones, M., Martínez, V., Benito, B., & Rubio, F. (2016). Comparison between Arabidopsis and rice for main pathways of K+ and Na+ uptake by roots. Frontiers in Plant Science, 7(992). https://doi.org/10.3389/fpls.2016.00992 CrossRefGoogle ScholarPubMed
Nieves-Cordones, M., Miller, A. J., Aleman, F., Martinez, V., & Rubio, F. (2008). A putative role for the plasma membrane potential in the control of the expression of the gene encoding the tomato high-affinity potassium transporter HAK5. Plant Molecular Biology, 68(6) 521532. https://doi.org/10.1007/s11103-008-9388-3 CrossRefGoogle ScholarPubMed
Nitsos, R. E., & Evans, H. J. (1969). Effects of univalent cations on activity of particulate starch synthetase. Plant Physiology, 44, 12601266+https://doi.org/10.1104/pp.44.9.1260 CrossRefGoogle ScholarPubMed
Peiter, E., Maathuis, F. J. M., Mills, L. N., Knight, H., Pelloux, J., Hetherington, A. M., & Sanders, D. (2005). The vacuolar Ca2+-activated channel TPC1 regulates germination and stomatal movement. Nature, 434(7031), 404408. https://doi.org/10.1038/nature03381 CrossRefGoogle ScholarPubMed
Peoples, T. R., & Koch, D. W. (1979). Role of potassium in carbon-dioxide assimilation in Medicago sativa L. Plant Physiology, 63, 878881. https://doi.org/10.1104/pp.63.5.878 CrossRefGoogle ScholarPubMed
Pérez, V., Wherrett, T., Shabala, S., Muñiz, J., Dobrovinskaya, O., & Pottosin, I. (2008). Homeostatic control of slow vacuolar channels by luminal cations and evaluation of the channel-mediated tonoplast Ca2+ fluxes in situ. Journal of Experimental Botany, 59(14), 38453855. https://doi.org/10.1093/jxb/ern225 CrossRefGoogle ScholarPubMed
Peters, J., & Chin, C. K. (2007) Potassium loss is involved in tobacco cell death induced by palmitoleic acid and ceramide. Archives of Biochemistry and Biophysics, 465(1), 180186. https://doi.org/10.1016/j.abb.2007.05.025 CrossRefGoogle ScholarPubMed
Pottosin, I., & Dobrovinskaya, O. (2014). Non-selective cation channels in plasma and vacuolar membranes and their contribution to K+ transport. Journal of Plant Physiology, 171(9), 723742. https://doi.org/10.1016/j.jplph.2013.11.013 CrossRefGoogle ScholarPubMed
Pottosin, I., & Dobrovinskaya, O. (2022). Major vacuolar TPC1 channel in stress signaling: What matters, K+, Ca2+ conductance or an ion-flux independent mechanism? Stress Biology, 2, Article 31. https://doi.org/10.1007/s44154-022-00055-0 CrossRefGoogle ScholarPubMed
Pottosin, I., & Martínez-Estévez, M. (2003) Regulation of the fast vacuolar channel by cytosolic and vacuolar potassium. Biophysical Journal, 84(21), 977986. https://doi.org/10.1016/S0006-3495(03)74914-5 CrossRefGoogle ScholarPubMed
Pottosin, I., Olivas-Aguirre, M., Dobrovinskaya, O., Zepeda-Jazo, I., & Shabala, S. (2021). Modulation of ion transport across plant membranes by polyamines: Understanding specific modes of action under stress. Frontiers in Plant Science, 11, Article 616077. https://doi.org/10.3389/fpls.2020.616077 CrossRefGoogle ScholarPubMed
Pottosin, I. I., Martínez-Estévez, M., Dobrovinskaya, O. R., & Muñiz, J. (2003). Potassium-selective channel in the red beet vacuolar membrane. Journal of Experimental Botany, 54(383), 663667. https://doi.org/10.1093/jxb/erg067 CrossRefGoogle ScholarPubMed
Pottosin, I. I., Martínez-Estévez, M., Dobrovinskaya, O. R., & Muñiz, J. (2005). Regulation of the slow vacuolar channel by luminal potassium: Role of surface charge. Journal of Membrane Biology, 205, 103111. https://doi.org/10.1007/s00232-005-0766-3 CrossRefGoogle ScholarPubMed
Ragel, P., Raddatz, N., Leidi, E. O., Quintero, F. J., & Pardo, J. M. (2019). Regulation of K+ nutrition in plants. Frontiers in Plant Science, 10, Article 281. https://doi.org/10.3389/fpls.2019.00281 CrossRefGoogle ScholarPubMed
Rangarajan, N., Kapoor, I., Li, S., Drossopoulos, P., White, K. K., Madden, V. J., & Dohlman, H. G. (2020). Potassium starvation induces autophagy in yeast. Journal of Biological Chemistry, 295(41), 1418914202. https://doi.org/10.1074/jbc.RA120.014687 CrossRefGoogle ScholarPubMed
Remillard, C. V., & Yuan, J. X. J. (2004). Activation of K+ channels: An essential pathway in programmed cell death. American Journal of Physiology-Lung Cellular and Molecular Physiology, 286(1), L49L67. https://doi.org/10.1152/ajplung.00041.2003 CrossRefGoogle ScholarPubMed
Rodenas, R., & Vert, G. (2020). Regulation of root nutrient transporters by CIPK23: ‘One kinase to rule them all’. Plant and Cell Physiology, 62(4), 553563. https://doi.org/10.1093/pcp/pcaa156 CrossRefGoogle Scholar
Rodriguez-Navarro, A., Blatt, M. R., & Slayman, C. L. (1986). A potassium-proton symport in Neurospora crassa . Journal of General Physiology, 87, 649674. https://doi.org/10.1085/jgp.87.5.649 CrossRefGoogle ScholarPubMed
Rogiers, S. Y., Coetzee, Z. A., Walker, R. R., Deloire, A., & Tyerman, S. D. (2017). Potassium in the grape (Vitis vinifera L.) berry: Transport and function. Frontiers in Plant Science, 8, Article 1629. https://doi.org/10.3389/fpls.2017.01629 CrossRefGoogle Scholar
Rozov, A., Khusainov, I., El Omari, K., Duman, R., Mykhaylyk, V., Yusupov, M., Westhof, E., Wagner, A., & Yusupova, G. (2019). Importance of potassium ions for ribosome structure and function revealed by long-wavelength X-ray diffraction. Nature Communications, 10(1), Article 2519. https://doi.org/251910.1038/s41467-019-10409-4.CrossRefGoogle ScholarPubMed
Ruan, L., Zhang, J. B., Xin, X. L., Zhang, C. Z., Ma, D. H., Chen, L., & Zhao, B. Z. (2015). Comparative analysis of potassium deficiency-responsive transcriptomes in low potassium susceptible and tolerant wheat (Triticum aestivum L.). Scientific Reports, 5(1), 10090. https://doi.org/10.1038/srep10090 CrossRefGoogle ScholarPubMed
Rubio, F., Fon, M., Ródenas, R., Nieves-Cordones, M., Alemán, F., Rivero, R. M., & Martínez, V. (2014) A low K+ signal is required for functional high-affinity K+ uptake through HAK5 transporters. Physiologia Plantarum, 152, 558570. https://doi.org/10.1111/ppl.12205 CrossRefGoogle Scholar
Rubio, F., Nieves-Cordones, M., Horie, T., & Shabala, S. (2020). Doing ‘business as usual’ comes with a cost: Evaluating energy cost of maintaining plant intracellular K+ homeostasis under saline conditions. New Phytologist, 225(3), 10971104. https://doi.org/10.1111/nph.15852 CrossRefGoogle ScholarPubMed
Sandmann, M., Skłodowski, K., Gajdanowicz, P., Michard, E., Rocha, M., Gomez-Porras, J. L., González, W., Corrêa, L. G. G., Ramírez-Aguilar, S. J., Cuin, T. A., van Dongen, J. T., Thibaud, J.-B., & Dreyer, I. (2011). The K+ battery-regulating Arabidopsis K+ channel AKT2 is under the control of multiple post-translational steps. Plant Signaling and Behavior, 6(4), 558562. https://doi.org/10.4161/psb.6.4.14908 CrossRefGoogle Scholar
Schachtman, D. P., & Shin, R. (2007). Nutrient sensing and signaling: NPKS. Annual Reviews of Plant Biology, 58, 4769. https://doi.org/10.1146/annurev.arplant.58.032806.103750 CrossRefGoogle ScholarPubMed
Shabala, L., Zhang, J., Pottosin, I., Bose, J., Zhu, M., Fuglsang, A. T., Velarde-Buendia, A., Massart, A., Hill, C. B., Roessner, U., Bacic, A., Wu, H., Azzarello, E., Pandolfi, C., Zhou, M., Poschenrieder, C., Mancuso, S., & Shabala, S. (2016a). Cell-type-specific H+-ATPase activity in root tissues enables K+ retention and mediates acclimation of barley (Hordeum vulgare) to salinity stress. Plant Physiology, 172(4), 24452458. https://doi.org/10.1104/pp.16.01347 CrossRefGoogle Scholar
Shabala, S. (2009). Salinity and programmed cell death: Unravelling mechanisms for ion specific signalling. Journal of Experimental Botany, 60(3), 709711. https://doi.org/10.1093/jxb/erp013 CrossRefGoogle ScholarPubMed
Shabala, S. (2017). Signalling by potassium: Another second messenger to add to the list? Journal of Experimental Botany, 68(15), 40034007. https://doi.org/10.1093/jxb/erx238 CrossRefGoogle ScholarPubMed
Shabala, S., & Pottosin, I. (2014). Regulation of potassium transport in plants under hostile conditions: Implications for abiotic and biotic stress tolerance. Physiologia Plantarum, 151(3), 257279. https://doi.org/10.1111/ppl.12165 CrossRefGoogle ScholarPubMed
Shabala, S., Chen, G., Chen, Z., & Pottosin, I. (2020). The energy cost of the tonoplast futile sodium leak. New Phytologist, 225(3), 11051110. https://doi.org/10.1111/nph.15758 CrossRefGoogle ScholarPubMed
Shabala, S., Demidchik, V., Shabala, L., Cuin, T. A., Smith, S. J., Miller, A. J., Davies, J. M., & Newman, I. A. (2006). Extracellular Ca2+ ameliorates NaCl-induced K+ loss from Arabidopsis root and leaf cells by controlling plasma membrane K+-permeable channels. Plant Physiology, 141, 16531665. https://doi.org/10.1104/pp.106.082388 CrossRefGoogle ScholarPubMed
Shankar, A., Singh, A., Kanwar, P., Srivastava, A. K., Pandey, A., Suprasanna, P., Kapoor, S., & Pandey, G. K. (2013). Gene expression analysis of rice seedling under potassium deprivation reveals major changes in metabolism and signaling components. PLoS One, 8(7), e70321. https://doi.org/10.1371/journal.pone.0070321 CrossRefGoogle ScholarPubMed
Shin, R., & Schachtman, D. P. (2004). Hydrogen peroxide mediates plant root cell response to nutrient deprivation. Proceedings of the National Academy of Science USA, 101, 88278832. https://doi.org/10.1073/pnas.0401707101 CrossRefGoogle ScholarPubMed
Talbott, L. D., & Zeiger, E. (1996). Central roles for potassium and sucrose in guard cell osmoregulation. Plant Physiology, 111(4), 10511057. https://doi.org/10.1104/pp.111.4.1051 CrossRefGoogle ScholarPubMed
Tang, R. J., Zhao, F. G., Yang, Y., Wang, C., Li, K., Kleist, T. J., Lemaux, P. G., & Luan, S. (2020). A calcium signaling network activates vacuolar K+ remobilization to enable plant adaptation to low-K environments. Nature Plants, 6(4), 384393. https://doi.org/10.1038/s41477-020-0621-7 CrossRefGoogle ScholarPubMed
Tian, H., Baxter, I. R., Lahner, B., Reinders, A., Salt, D. E., & Ward, J. M. (2010). Arabidopsis NPCC6/NaKR1 is a phloem mobile metal binding protein necessary for phloem function and root meristem maintenance. The Plant Cell, 22(12), 39633979. https://doi.org/10.1105/tpc.110.080010 CrossRefGoogle ScholarPubMed
Tikhonova, L. I., Pottosin, I. I., Dietz, K. J., & Schönknecht, G. (1997). Fast-activating cation channel in barley mesophyll vacuoles Inhibition by calcium. The Plant Journal, 11(5), 10591070. https://doi.org/10.1046/J.1365-313X.1997.11051059.X CrossRefGoogle Scholar
Touraine, B., Grignon, N., & Grignon, C. (1988). Charge balance in NO3−fed soybean: Estimation of K+ and carboxylate recirculation. Plant Physiology, 88(3), 605612. https://doi.org/10.1104/pp.88.3.605 CrossRefGoogle ScholarPubMed
van Bel, A. J., & Hafke, J. B. (2005). Physiochemical determinants of phloem transport. In Vascular transport in plants (pp. 1944). Elsevier. https://www.sciencedirect.com/science/article/pii/B9780120884575500046 CrossRefGoogle Scholar
Very, A. A., & Sentenac, H. (2003). Molecular mechanisms and regulation of K+ transport in higher plants. Annual Review of Plant Biology, 54, 575603. https://doi.org/10.1146/annurev.arplant.54.031902.134831 CrossRefGoogle ScholarPubMed
Vreugdenhil, D. (1985). Source-to-sink gradient of potassium in the phloem. Planta, 163 ( 2), 238240. https://doi.org/10.1007/BF00393513 CrossRefGoogle ScholarPubMed
Walker, D. J., Leigh, R. A., & Miller, A. J. (1996). Potassium homeostasis in vacuolate plant cells. Proceedings of the National Academy of Sciences USA, 93(19), 1051010514. https://doi.org/10.1073/pnas.93.19.10510 CrossRefGoogle ScholarPubMed
Wang, F. L., Tan, Y. L., Wallrad, L., Du, X. Q., Eickelkamp, A., Wang, Z. F., He, G. F., Rehms, F., Li, Z., Han, J. P., Schmitz-Thom, I., Wu, W. H., Kudla, J., & Wang, Y. (2021). A potassium-sensing niche in Arabidopsis roots orchestrates signaling and adaptation responses to maintain nutrient homeostasis. Developmental Cell, 56(6), 781794. https://doi.org/10.1016/j.devcel.2021.02.027 CrossRefGoogle ScholarPubMed
Wang, Y., & Wu, W. H. (2010). Plant sensing and signaling in response to K+-deficiency. Molecular Plant, 3, 280287. https://doi.org/10.1093/mp/ssq006 CrossRefGoogle ScholarPubMed
Wang, Y., Dindas, J., Rienmuller, F., Krebs, W., Schumacher, K., Wu, W. H., Hedrich, R., & Roelfsema, M. R. G. (2015). Cytosolic Ca2+ signals enhance the vacuolar ion conductivity of bulging Arabidopsis root hair cells. Molecular Plant, 8(11), 16651674. https://doi.org/10.1016/j.molp.2015.07.009 CrossRefGoogle ScholarPubMed
Wang, Y. H., Pan, C. C., Chen, Q. H., Xie, Q., Gao, Y. W., He, L. L., Li, Y., Dong, Y. L., Jiang, X. Y., & Zhao, Y. (2023). Architecture and autoinhibitory mechanism of the plasma membrane Na+/H+ antiporter SOS1 in Arabidopsis . Nature Communications, 14(1), Article 4487. https://doi.org/448710.1038/s41467-023-40215-y.CrossRefGoogle ScholarPubMed
Ward, J. M., & Schroeder, J. I. (1994). Calcium-activated K+ channels and calcium-induced Ca2+ release by slow vacuolar ion channels in guard cell vacuoles implicated in the control of stomatal closure. The Plant Cell, 6(5), 669683. https://doi.org/10.1105/tpc.6.5.669 CrossRefGoogle ScholarPubMed
Wegner, L. H. (2015). Interplay of water and nutrient transport: A whole-plant perspective. In Progress in botany (pp. 109141). Springer. http://link.springer.com/chapter/10.1007/978-3-319-08807-5_5 CrossRefGoogle Scholar
Wegner, L. H., & De Boer, A. H. (1997). Properties of two outward-rectifying channels in root xylem parenchyma cells suggest a role in K+ homeostasis and long-distance signaling. Plant Physiology, 115(4), 17071719. https://doi.org/10.1104/pp.115.4.1707 CrossRefGoogle Scholar
Wegner, L. H., & Raschke, K. (1994). Ion channels in the xylem parenchyma of barley roots. A procedure to isolate protoplasts from this tissue and a patch-clamp exploration of salt passageways into xylem vessels. Plant Physiology, 105(3), 799813. https://doi.org/10.1104/pp.105.3.799 CrossRefGoogle Scholar
Wegner, L. H., & Zimmermann, U. (2002). On-line measurements of K+ activity in the tensile water of the xylem conduit of higher plants. The Plant Journal, 32(3), 409417. https://doi.org/10.1046/j.1365-313x.2002.01426 CrossRefGoogle ScholarPubMed
Wegner, L. H., & Zimmermann, U. (2009). Hydraulic conductance and K+ transport into the xylem depend on radial volume flow, rather than on xylem pressure, in roots of intact, transpiring maize seedlings. New Phytologist, 181(2), 361373. https://doi.org/10.1111/j.1469-8137.2008.02662.x CrossRefGoogle Scholar
Wigoda, N., Moshelion, M., & Moran, N. (2014). Is the leaf bundle sheath a “smart flux valve” for K+ nutrition? Journal of Plant Physiology, 171(9), 715722. https://doi.org/10.1016/j.jplph.2013.12.017 CrossRefGoogle Scholar
Wigoda, N., Pasmanik-Chor, M., Yang, T., Yu, L., Moshelion, M., & Moran, N. (2017). Differential gene expression and transport functionality in the bundle sheath versus mesophyll–a potential role in leaf mineral homeostasis. Journal of Experimental Botany, 68(12), 31793190. https://doi.org/10.1093/jxb/erx067 CrossRefGoogle ScholarPubMed
Windt, C. W., Vergeldt, F. J., De Jager, P. A., & Van As, H. (2006). MRI of long-distance water transport: A comparison of the phloem and xylem flow characteristics and dynamics in poplar, castor bean, tomato and tobacco. Plant, Cell & Environment, 29(9), 17151729. https://doi.org/10.1111/j.1365-3040.2006.01544.x CrossRefGoogle ScholarPubMed
Wu, H., Zhang, X., Giraldo, J. P., & Shabala, S. (2018). It is not all about sodium: Revealing tissue specificity and signalling roles of potassium in plant responses to salt stress. Plant and Soil, 431 (1–2), 117. https://doi.org/10.1007/s11104-018-3770-y CrossRefGoogle Scholar
Xiao, J. Q., Zhou, Y. J., Xie, Y. Y., Li, T. T., Su, X. G., He, J. X., Jiang, Y. M., Zhu, H., & Qu, H. X. (2024). ATP homeostasis and signaling in plants. Plant Communications, 5(4), Article 100834. https://doi.org/10083410.1016/j.xplc.2024.100834.CrossRefGoogle ScholarPubMed
Zhang, H. N., Liu, Z. H., Zhang, Y. M., Yang, R. J., Guo, X. L., & Wang, G. Y. (2019). Salt stress changes the ion homeostasis via differential K+ uptake ways in salt-sensitive/tolerant wheat (Triticum aestivum) varieties. International Journal of Agriculture and Biology, 21, 733742. https://doi.org/10.17957/IJAB/15.0951 Google Scholar
Zhang, M., Hu, Y., Han, W., Chen, J., Lai, J., & Wang, Y. (2023). Potassium nutrition of maize: Uptake, transport, utilization, and role in stress tolerance. The Crop Journal, 11(1), 10481058. https://doi.org/10.1016/j.cj.2023.02.009 CrossRefGoogle Scholar
Zhao, C. Z., Zhang, H., Song, C. P., Zhu, J. K., & Shabala, S. (2020). Mechanisms of plant responses and adaptation to soil salinity. The Innovation, 1(1), Article 100017. https://doi.org/10001710.1016/j.xinn.2020.100017.CrossRefGoogle ScholarPubMed
Zörb, C., Senbayram, M., & Peiter, E. (2014). Potassium in agriculture–status and perspectives. Journal of Plant Physiology, 171(9), 656669. https://doi.org/10.1016/j.jplph.2013.08.008 CrossRefGoogle Scholar
Figure 0

Figure 1. A schematic diagram showing K+ circulation via xylem and phloem in higher plants; suc., sucrose. For more details, see the text.

Figure 1

Figure 2. Homeostatic (steady state) condition of a transmembrane K+ gradient. In this example, the membrane voltage in a steady state is Vss = −130 mV. At this voltage, there is an efflux of H+ released by the proton pump. The charge transport is compensated by a K+ influx of the same magnitude via the K+ channel (KC). Without further compensation, the system would not be stable in steady state but mediate an H+/K+ antiport, which would at least change the transmembrane potassium gradient. EK would become increasingly negative. The system would only be stable if the red and green dots overlap with the light blue dot at −200 mV. Stability at Vss = −130 mV can be obtained with an additional H+/K+ antiporter. This transporter releases each accumulated K+ ion and in the same process retrieves the released H+ in an electroneutral manner. The system is in steady state, and Vss or EK do not change. However, there are ATP-consuming H+ and K+ cycles.

Figure 2

Figure 3. Difference in NaCl-induced K+ flux ‘signatures’ explain the difference in the strategy of adapting to salt stress between pea and barley. (a) Transient changes in MP of epidermal root cells in wheat and pea in response to acute NaCl treatment (based on Cuin et al., 2008 and Bose et al., 2014, respectively). (b) Net K+ fluxes measured from plant roots by non-invasive ion flux measuring technique. (c) Predicted changes in cytosolic K+ concentration in pea and wheat root cytosol assuming no buffering from the vacuole. (d) Effect of vacuolar K+ buffering on cytosolic K+ kinetics in pea roots exposed to acute salinity treatment. Percentage values are the size of the vacuolar K+ available to be used to compensate cytosolic K+ loss.

Figure 3

Figure 4. Potassium permeable channels of FV and SV types act as safety valves, preventing K+ overload at excessive K+; vacuolar K+ filling is mediated by K+ (Na+)/ H+ antiporter of NHX type (left). Suppression of the activity of FV and SV channels by low vacuolar K+ prevents channel-mediated vacuolar K+ re-uptake and allows the maintenance of cytosolic K+ at the expense of the vacuolar K+ pool. Cytosol refilling by vacuolar K+ can be mediated by a hypothetical K+: H+ symporter (right). Note that in the K+ replete conditions a significant ‘futile’ H+ and K+ cycling results, whereas under K+ deficiency (upon the change in the direction of transtonoplast K+ gradient) this cycling is reduced. Above transport system acts as K+ sensor/effector at non-saline conditions but is unable per se to discriminate between K+ and Na+ under high salinity (see text).

Author comment: Potassium homeostasis and signalling: from the whole plant to the subcellular level — R0/PR1

Comments

Dear Dale,

this review paper is submitted following your earlier invitation. We hope it fulfills your expectations.

Best regards, Sergey

Review: Potassium homeostasis and signalling: from the whole plant to the subcellular level — R0/PR2

Conflict of interest statement

N/A As per prior communication with the journal, this manuscript was co-reviewed with John Ward. Both read the review, wrote and exchanged comments.

Comments

This is a review of the roles of K+ at the subcellular, cellular, and up to the whole-plant level. It reviews some less discussed but material topics. It also introduces some hypotheses that will be useful to the field. I have some suggestions for improving the manuscript.

1. In the section on K+ signaling (pages 7 and 8), it would be good to note in the beginning (lines 242-244) that changes in K+ concentration could be a signal, as you said, but that that K+ deficiency signaling could lead to additional changes discussed here such as differences in cell fate or autophagy. K+ deficiency signaling is discussed next so it fits together.

2. In the section on long-distance transport, the role of HKT1 is not mentioned. Although HKT1 is mainly studied due to its strong effect on sodium tolerance, there are examples of HKT1 importance for potassium homeostasis. For example, by removing Na+ from the xylem, data has shown that the K+ concentration in the xylem sap is larger in wildtype controls than in hkt1 mutants in Arabidopsis, presumably due to HKT1-dependent Na+ mediated depolarisation and the ensuing K+ efflux from xylem parenchyma cells.

3. In the discussion around Fig 3, something appears to be incorrect. For Fig. 3B, line 361 “wheat K+ efflux was an order of magnitude lower and gradually increased with time”. For wheat, there appears to be a small rapid efflux that is constant over time. That would be consistent with a linear decrease in K+ concentration seen in Fig. 3C.

4. Also, Figure 3 could be improved, red and green colors are not good for two reasons, color blind readers can’t distinguish them and when printed in black and white they are not distinguishable. Change to two different shapes such as open and closed circles and squares. And some of the labeling, especially the axes and legend, appears to be low resolution.

5. The high-affinity HAK/KUP/KT K+ transporter family should be introduced and described. Only HAK5 is briefly mentioned in the Conclusions. Papers investigating the phenotypes for K+ flux and accumulation in roots of hak5 single and hak5 akt1 double mutants should be discussed.

6. lines 293 -314: With respect to the low K+ sensing and ensuing Ca2+ increase, the connection to hyperpolarisation-activated Ca2+ permeable channels should be described.

7. Studies showing effects of low K+ induced transcriptome changes should be discussed when reviewing K+ sensing.

Minor points

line 68 “with ambient K+” is not clear enough, maybe change to “with the concentration of K+ of available in the soil” or “with soil K+ ” or “with K+ in the media” depending on how the experiment was performed.

In the discussion about long-distance transport of K+ in the phloem and evidence that K+ is recycled from the leaves to the roots via the phloem, it would be useful to cite genetic evidence from Tian et al., 2010 Plant Cell. 2010 22:3963-79, who showed that when NaKR1 in Arabidopsis is knocked out, phloem is defective, and sodium and potassium accumulate in leaves. Maybe lines 91-97.

The subtitle “Cytosolic K+ levels determine cell fate” could be misinterpreted and should be more specific to the proposed hypothesis.

line 109 “SKOR is under tight control of nitrate transporter NTR1.5” it is unclear which is in control, I think this should be “”SKOR is tightly controlled by the activity of nitrate transporter NTR1.5".

line 112 “either site” should be “either side”

lines 113-116 “Hence, K+ retention at the plasma membrane, induced by an increase in volume flow passing through the cells, would favour K+ release into the xylem (so-called concentration-polarisation effect).” the meaning is not clear, maybe “Potassium uptake into xylem parenchyma cells favors K+ release into the xylem (so-called concentration-polarisation effect).”

line 221 “but only symporters and channels” not clear change to “via symporters and channels”

line 356 “withing” change to “within”

line 392 (Cuin et al. 200) should be Cuin et al., 2008.

line 439 “proved” should be “proven”, but “shown” would be better.

line 450 “slower at low ABA dose” change to “ slower at low ABA dosage”

line 467 “independent on cytosolic,” change to “independent of cytosolic K+,”

line 493 “act as safe vales” change to “act as safety valves”

line 511 “dry-lab analysis” is confusing. I believe computational modeling is meant here and would be more direct.

Line 519 the reference “Maierhofer et al” is missing.

line 522 “and hand-to-hand going focused experimentation are required.” could be stated better, maybe “ with focused experimentation are required” is better.

Review: Potassium homeostasis and signalling: from the whole plant to the subcellular level — R0/PR3

Conflict of interest statement

Reviewer declares none.

Comments

Wegner et al ‘Potassium homeostasis...’

Wegner et al set out here to cover a range of topics around K homeostasis and signalling that they perceive to have been less covered in recent years. Although the abstract lists eight separate topics, the contents can be condensed to three, overarching themes, (i) long-distance K circulation, (ii) cellular K homeostasis and K sensing, and (iii) the role of tonoplast K transport in homeostasis and sensing. Some effort to condense and focus the initial descriptions might therefore be a help for the reader, as these are interwoven.

General

Overal, the issues of K circulation, and of xylem and phloem loading are explained well and the arguments for the concept of a ‘K battery’ are easily accessed. The model has been promoted by Dreyer in the past and is an intriguing one, but the presentation would benefit from some more detail to highlight the experimental evidence to support this model (l. 147-63). Equally important, readers would find useful some effort to highlight what unique predictions come from the model and how they might be tested.

Elsewhere, the presentation is strained and occasionally confused. In general, it is not obvious how the distinction is made between what determines true sensor/receptor processes and what are determined by the intrinsic kinetic features of the dynamic interactions that arise through shared intermediates, whether of the membrane voltage or of common pools of ions in each compartment. Some clarity around these distinctions is vital, if only to point out the difficulties in establishing what constitutes an ion sensory process. It is also unclear at times as to what tissues/cells the text refers to. This matter is equally important, as many of the processes are tissue-specific. Several of the detailed points below relate to these concerns.

Specific points (in order of appearance)

l.218, ff The text here and elsewhere confuses ‘steady state’ with ‘equilibria’. The two phenomena are not the same. What is described here is not a system that is ‘brought out of equilibrium’ but simply one that is displaced from its steady state. The result is a compensating change in flux that tends to return the system to the original steady state. Nothing more and no equilibrium.

Figure 2 The authors should reconsider the description of the H flux and its equilbrium here. The H-ATPase pumps one H per cycle (ATP), so the energetics imply an equilibrium voltage around -380 mV under physiological conditions, ie. pHi of 7.5 and pHo of 5.5

l.222-4 There is a long history connecting K and H flux that the authors should consider citing here. Among others, K starvation in fungi leads to a 1:1 exchange of H for K (cf. Conway & O’Malley 1944 Nature; Conway & O’Malley 1946 Biochem J) that was shown in Neurospora to arise through a H-K symport operating with the H-ATPase (cf. Rodgriguez-Navarro et al 1986 J Gen Physiol; Blatt & Slayman 1987 PNAS). The same processes and mechanisms were later shown in Arabidopsis roots (Matthuis & Sanders 1994 PNAS). There is also clear evidence in guard cells for similar mechanisms of H and K exhange (cf. Clint & Blatt 1989 Planta; Blatt & Clint 1989; Wong et al 2021 Plant Phys; to name a few).

l.227 “It former generates...” missing subject?

l. 255, ff Potassium is well-known as a factor important for many enzyme functions, if only in charge balance and protein stabilization. These are broad effects and for many, concentrations of only a few millimolar are sufficient, so well below the concentrations found in the cytosol even in K-starved cells. For the specific examples here - and generally for the benefit of readers - it would be helpful to compare and assess the concentrations of K needed for these activities and how they compare with the concentrations of K found under stress, such as evident under K starvation. If possible, it would be useful to compare the values of Ki with respect to [K] or its depletion. Otherwise, much of the proposed roles for K as a ‘switch’ are not particularly meaningful. For example, the reader must ask ‘How low must [K] go before it deactivates TORC1?’, ‘What is the Ki for TORC1 inhibition by K?’ and so on.

l. 306, ff Some clarity and detail around these experiments is called for. The Behera study cited here is easily misunderstood, principally because the experiments did not include measurements of membrane voltage. A simple explanation for the ‘K signature’ is that hyperpolarization of the membrane occurs when K is removed which leads immediately to an increase in Ca influx through hyperpolarization-activated Ca channels (and hence a rise in cytosolic Ca). Such phenomena are well-known and proven to connect through membrane voltage in many tissues (cf. Grabov & Blatt 1998 PNAS; Gelli & Blumwald 1997 J Memb Biol; Lemtiri-Chlieh & Berkowitz 2004 JBC). It is therefore unnecessary to invoke additional regulatory controls or to suggest a ‘K signature’ beyond the connection through membrane voltage.

l. 342, ff. please clarify that the reference here of voltage gating is to ion channels, not transporters in general. An additional point is missed altogether here is that coupled transporters, such as the H-Cl and H-K (HAK) symporters as well as the H-ATPase, are also under kinetic control by membrane voltage and this kinetic control is generally non-linear. The authors' own Fig. 2 makes this point clear for the H-ATPase. So even without voltage gating (of channels), membrane voltage will have very important impacts on ion homeostasis.

l.345-6 Again, the authors need to consider - and make the distinction between - regulation such as by changes in the number of active transporters (here the H-ATPase) vs the effect of voltage on the kinetics of transport. For the H-ATPase, simply depolarizing the membrane will stimulate H flux instantaneously by relaxing the voltage limitation on H-ATPase kinetics, as shown in the authors' Fig. 2.

l. 363-96 The calculations here do not take account of plasmodesmatal connections between cells and the extensive ‘buffer capacity’ that these connections confer on the root. It would be a good idea to bring considerations of cellular coupling into these calculations. Otherwise, the estimates appear wildly outside reality.

l. 367 symbols missing here

l. 402-5 The numbers here are somewhat out of date and, in all events, are highly variable. The authors will find a more recent assessment in the review of Jezek & Blatt 2017 Plant Phys.

l. 413, ff The literature around vacuolar K and Ca flux is certainly convoluted. It would be helpful for the reader if some of the details presented in this paragraph are omitted/condensed and the presentation simplified. In particular, as there is still no functional evidence for TPC1, it might be an idea to state this up front and move on.

l. 445-55 The authors would do well here to include some discussion of the OnGuard models that have proven successful in predicting the mechanics of how stomata work. The analysis that Chen et al (2012) and Jezek et al (2021) present supports the idea of the connection between content and tonoplast flux, only the mechanistic interpretation does not require stretch activated channels, just the underlying kinetic constraints of the different channels. A discussion of these connections is also to be found in Horaruang et al 2020 Plant Phys.

l. 460, ff The opening sentences are completely devoid of citation, yet the information is presented as fact. Please clarify. This entire paragraph would benefit from a rewriting. It’s almost impossible to follow, and the directionalities of the various fluxes are lost on the reader. There is much speculation also around surface charge, yet the publications cited offer little basis to make the claims here.

l.520 what is a K ‘effector’? Please clarify

Minor

l. 143 Please use English term (’Pressure-Flow') rather than the German or include it for the non-German speakers

l.174-4, 227, 416, “does not only interact”, “but hardly” and elsewhere, please consider rephrasing to avoid germanic constructions

Recommendation: Potassium homeostasis and signalling: from the whole plant to the subcellular level — R0/PR4

Comments

Thank you for submitting this review article to QPB.

Copies of two independent reviews of your manuscript are attached. As you will note, the reviews are authoritative and detailed. The reviewers propose a number of suggestions that would enhance the quality of the manuscript. Major revision is recommended in the hope you will utilise the constructive comments of the reviewers to generate what I think will comprise a landmark review in the field of understanding K homeostasis in plants.

When resubmitting the manuscript, please detail your responses to the reviewers' points.

Decision: Potassium homeostasis and signalling: from the whole plant to the subcellular level — R0/PR5

Comments

No accompanying comment.

Author comment: Potassium homeostasis and signalling: from the whole plant to the subcellular level — R1/PR6

Comments

No accompanying comment.

Recommendation: Potassium homeostasis and signalling: from the whole plant to the subcellular level — R1/PR7

Comments

No accompanying comment.

Decision: Potassium homeostasis and signalling: from the whole plant to the subcellular level — R1/PR8

Comments

No accompanying comment.