Channel-mediated high-affinity K uptake into guard cells from Arabidopsis

Similar documents
Potassium Uptake Supporting Plant Growth in the Absence of AKT1 Channel Activity Inhibition by Ammonium and Stimulation by Sodium

Outer Pore Residues Control the H and K Sensitivity of the Arabidopsis Potassium Channel AKT3

Opening and closing of stomatal pores regulates gas exchange

Voltage-dependent gating characteristics of the K channel KAT1 depend on the N and C termini

Transport of glucose across epithelial cells: a. Gluc/Na cotransport; b. Gluc transporter Alberts

Internal Aluminum Block of Plant Inward K Channels

Date of birth: 19 August 1967 Marital status: married Children: 2 Sons Nationality: Dutch Place of birth: Assen, The Netherlands

Supplementary Figure 1

Regulation of Voltage Dependence of the KAT1 Channel by Intracellular Factors

Dipartimento di Produzione Vegetale, Università degli Studi di Milano, via Celoria 2, I Milano, Italy

Plant ion channels: from molecular structures to physiological functions Sabine Zimmermann* and Hervé Sentenac

The physiology of channel-mediated K + acquisition in roots of higher plants

Inhibition of S532C by MTSET at intracellular ph 6.8 indicates accessibility in the closed

Editorial. What is the true resting potential of small cells? Jean-Marc Dubois

Signal processing in nervous system - Hodgkin-Huxley model

Role of inward K + channel located at carrot plasma membrane in signal cross-talking of camp with Ca 2+ cascade

Identification and Characterization of the Inward K + Membrane of Brassica Pollen Protoplasts. Channel in the Plasma

Channels can be activated by ligand-binding (chemical), voltage change, or mechanical changes such as stretch.

Membrane Potentials, Action Potentials, and Synaptic Transmission. Membrane Potential

BRIEF COMMUNICATION 3,4-DIAMINOPYRIDINE A POTENT NEW POTASSIUM CHANNEL BLOCKER

Quantitative Electrophysiology

CELL BIOLOGY - CLUTCH CH. 9 - TRANSPORT ACROSS MEMBRANES.

Cellular Electrophysiology. Cardiac Electrophysiology

AtKUP1: An Arabidopsis Gene Encoding High-Affinity Potassium Transport Activity

Quantitative Electrophysiology

Salinity stress-tolerant and -sensitive rice (Oryza sativa L.) regulate AKT1-type potassium channel transcripts differently

Channelling auxin action: modulation of ion transport by indole-3-acetic acid

Edited by Maarten J. Chrispeels, University of California, San Diego, CA, and approved August 17, 2000 (received for review May 12, 2000) Methods

Supplemental Information. C. elegans AWA Olfactory Neurons Fire. Calcium-Mediated All-or-None Action Potentials

Differential Sodium and Potassium Transport Selectivities of the Rice OsHKT2;1 and OsHKT2;2 Transporters in Plant Cells 1[C][OA]

Site directed mutagenesis reduces the Na a nity of HKT1, an Na energized high a nity K transporter

Passive Membrane Properties

Differential polyamine sensitivity in inwardly rectifying Kir2 potassium channels

AD-" IONIC BASIS OF POTENTIAL REGULATION(U) BAYLOR COLLO / U U ijejmedicine L HOUSTON TX DEPT OF PHYSIOLOGY AND MOLECULAR 7 MEEE"..

Modeling action potential generation and propagation in NRK fibroblasts

Receptor-mediated activation of a plant Ca 2 -permeable ion channel involved in pathogen defense

Research review Research review

over K in Glial Cells of Bee Retina

Resting membrane potential,

BRIEF COMMUNICATION OF ASYMMETRY CURRENT SQUID AXON MEMBRANE FREQUENCY DOMAIN ANALYSIS

Different properties of SV channels in root vacuoles from near isogenic Al-tolerant and Al-sensitive wheat cultivars

Ca 2+ -Permeable, Outwardly-Rectifying K + Channels in Mesophyll Cells of Arabidopsis thaliana

Slide 1. Slide 2. Membrane Transport Mechanisms II and the Nerve Action Potential. Epithelia

Gyrhofstr. 15, Köln, Germany. 3 Present address: University of Bern, Institute of Cell Biology,

Module Membrane Biogenesis and Transport Lecture 15 Ion Channels Dale Sanders

Supporting Information

Voltage dependence of K+ channels in guard-cell protoplasts

Voltage-clamp and Hodgkin-Huxley models

Inventory and Functional Characterization of the HAK Potassium Transporters of Rice 1

BIOELECTRICITY. Chapter 1. Electrical Potentials. Electrical Currents

ACTION POTENTIAL. Dr. Ayisha Qureshi Professor MBBS, MPhil

Membrane Biology. Quantitative Analysis of Outward Rectifying K + Channel Currents in Guard Cell Protoplasts from Vicia faba.

POTASSIUM PERMEABILITY IN

Voltage-Dependent K Channels as Targets of Osmosensing in Guard Cells

Rahaf Nasser mohammad khatatbeh

Chapter 1 subtitles Ion gradients

Voltage Clamp Limitations of Dual Whole-Cell Gap Junction Current and Voltage Recordings. I. Conductance Measurements

Single Channel Properties of P2X 2 Purinoceptors

Membrane Biology. Membrane Transport in Stomatai Guard Cells: The Importance of Voltage Control. G. Thiel*, E.A.C. MacRobbie, and M.R.

Mechanism of Ion Permeation in Skeletal Muscle Chloride Channels

Voltage-clamp and Hodgkin-Huxley models

Physiology Unit 2. MEMBRANE POTENTIALS and SYNAPSES

Neurons and the membrane potential. N500 John Beggs 23 Aug, 2016

CELLULAR NEUROPHYSIOLOGY CONSTANCE HAMMOND

DYNAMICS OF POTASSIUM ION CURRENTS IN

Differential Abscisic Acid Regulation of Guard Cell Slow Anion Channels in Arabidopsis Wild-Type and

Influence of permeating ions on potassium channel block by external tetraethylammonium

BIOELECTRIC PHENOMENA

Cold Transiently Activates Calcium-Permeable Channels in Arabidopsis Mesophyll Cells 1[W]

Comparative functional features of plant potassium HvHAK1 and

Neuroscience 201A Exam Key, October 7, 2014

Basic elements of neuroelectronics -- membranes -- ion channels -- wiring

Membrane-Pipette Interactions Underlie Delayed Voltage Activation of Mechanosensitive Channels in Xenopus Oocytes

Membrane Protein Channels

7 Membrane Potential. The Resting Membrane Potential Results From the Separation of Charges Across the Cell Membrane. Back.

Νευροφυσιολογία και Αισθήσεις

Three Components of Calcium Currents in Crayfish Skeletal Muscle Fibres

Decoding. How well can we learn what the stimulus is by looking at the neural responses?

Running head: ABA activation of anion channels in tobacco guard cells

Membrane Currents in Mammalian Ventricular Heart Muscle Fibers Using a Voltage-Clamp Technique

Actin Dynamics Regulates Voltage-Dependent Calcium-Permeable Channels of the Vicia faba Guard Cell Plasma Membrane

Plant KT/KUP/HAK Potassium Transporters: Single Family Multiple Functions

Neurophysiology. Danil Hammoudi.MD

Physiology Unit 2. MEMBRANE POTENTIALS and SYNAPSES

BME 5742 Biosystems Modeling and Control

Current- and Voltage-Clamp Recordings and Computer Simulations of Kenyon Cells in the Honeybee

Divalent cation gating of an ammonium permeable channel in the symbiotic membrane from soybean nodules

Effects of Polylinker uatgs on the Function of Grass HKT1 Transporters Expressed in Yeast Cells

S-type Anion Channels SLAC1 and SLAH3 Function as Essential Negative Regulators of Inward K + Channels and Stomatal Opening in Arabidopsis

Voltage-Dependent Membrane Capacitance in Rat Pituitary Nerve Terminals Due to Gating Currents

لجنة الطب البشري رؤية تنير دروب تميزكم

Main idea of this lecture:

me239 mechanics of the cell - syllabus me239 mechanics of the cell me239 mechanics of the cell - grading me239 mechanics of the cell - overview

Parameters for Minimal Model of Cardiac Cell from Two Different Methods: Voltage-Clamp and MSE Method

Synaptic dynamics. John D. Murray. Synaptic currents. Simple model of the synaptic gating variable. First-order kinetics

Cytokinin. Fig Cytokinin needed for growth of shoot apical meristem. F Cytokinin stimulates chloroplast development in the dark

Action Potential (AP) NEUROEXCITABILITY II-III. Na + and K + Voltage-Gated Channels. Voltage-Gated Channels. Voltage-Gated Channels

Separation and Characterization of Currents through Store-operated CRAC Channels and Mg 2 -inhibited Cation (MIC) Channels

Lecture 04, 04 Sept 2003 Chapters 4 and 5. Vertebrate Physiology ECOL 437 University of Arizona Fall instr: Kevin Bonine t.a.

Potassium channels in plant cells

Transcription:

Proc. Natl. Acad. Sci. USA Vol. 96, pp. 3298 3302, March 1999 Plant Biology Channel-mediated high-affinity K uptake into guard cells from Arabidopsis LIOUBOV BRÜGGEMANN*, PETRA DIETRICH*, DIRK BECKER*, INGO DREYER*, KLAUS PALME, AND RAINER HEDRICH* *Julius-von-Sachs-Insititut für Biowissenschaften, Lehrstuhl für Molekulare Pflanzenphysiologie und Biophysik, Universität Würzburg, Julius-von-Sachs-Platz 2, 97082 Würzburg, Germany; and Max-Delbrück-Laboratorium in der Max-Planck Gesellschaft, Carl-von-Linné Weg 10, 50829 Cologne, Germany Communicated by Jozef S. Schell, Max Planck Institute for Breeding Research, Cologne, Germany, December 24, 1998 (received for review November 30, 1998) ABSTRACT Potassium uptake by higher plants is the result of high- or low-affinity transport accomplished by different sets of transporters. Although K channels were thought to mediate low-affinity uptake only, the molecular mechanism of the high-affinity, proton-dependent K uptake system is still scant. Taking advantage of the high-current resolution of the patch-clamp technique when applied to the small Arabidopsis thaliana guard cells densely packed with voltage-dependent K channels, we could directly record channels working in the concentration range of high-affinity K uptake systems. Here we show that the K channel KAT1 expressed in Arabidopsis guard cells and yeast is capable of mediating potassium uptake from media containing as little as 10 M of external K. Upon reduction of the external K content to the micromolar level the voltage dependence of the channel remained unaffected, indicating that this channel type represents a voltage sensor rather than a K -sensing valve. This behavior results in K release through K uptake channels whenever the Nernst potential is negative to the activation threshold of the channel. In contrast to the H - coupled K symport shown to account for high-affinity K uptake in roots, ph-dependent K uptake into guard cells is a result of a shift in the voltage dependence of the K channel. We conclude that plant K channels activated by acid ph may play an essential role in K uptake even from dilute solutions. Since the initial observation of Epstein that K uptake into plant cells can be decomposed into the activity of high- and low-affinity transport systems (1), the molecular structure of three different K transporters, K channels, and two distinct carrier types, has been identified. In 1992, AKT1 and KAT1 (2, 3) were shown to represent the first members of a large family of inward-rectifying K channels (4), channel subtypes that had been recognized from patch-clamp studies in almost all plant cell types studied so far (5). In the presence of millimolar K concentrations and upon hyperpolarization of the plasma membrane, these inward rectifiers mediate K uptake. Based on studies in the whole-cell configuration of the patch-clamp technique applied to Vicia faba guard cell protoplasts, Schroeder and Fang (6) calculated an affinity constant of 3.5 mm for these guard cell K channels. Analysis of the current voltage relation obtained from guard cells and root cells exposed to varying external K concentrations revealed that the activation potential of the inward currents shifted to 20- to 30-mV more negative values with decreasing K concentrations (6 8), a behavior from which it was concluded that K channels can sense the external potassium. Thus, channel opening was supposed to occur only when the K gradient is inward, restricting these transporters to function as K -sensing K The publication costs of this article were defrayed in part by page charge payment. This article must therefore be hereby marked advertisement in accordance with 18 U.S.C. 1734 solely to indicate this fact. PNAS is available online at www.pnas.org. uptake valves (9). Supporting this hypothesis, these channels were reported to not open in the absence of extracellular K (9). Because of this effect, K uptake channels were not supposed to form shunt pathways through which K, accumulated by high-affinity carriers, can leak out of the cells. To date, two high-affinity plant K carriers have been identified. In 1994, Schachtman and Schroeder reported on the discovery of HKT1, a proton-driven K symporter from Hordeum vulgare (10). Later, however, detailed studies revealed that HKT1 mediates Na -driven rather than H -driven K uptake (11). In search for the molecular entity generating the H K phenomenon, Escherichia coli-like K transporters recently were identified in barley (12) and Arabidopsis (13). Later studies showed that members of this family mediate both high- and low-affinity K transport (14, 15). It is, however, still unclear which members of the channel or carrier families account for acid-induced K uptake observed in vivo (16). Here we report that hyperpolarization-activated K channels in guard cells are capable of mediating high-affinity K uptake, the activity of which strongly depends on the external proton concentration. Supporting this finding we show that K uptake-deficient yeast, when complemented with this guard cell K channel, regained the ability to grow in micromolar K media. Therefore, acid-activated K uptake channels might represent a molecular equivalent contributing to the H K phenomenon. MATERIALS AND METHODS Yeast Genetics. Potassium-dependent growth assays were performed by using the Saccharomyces cerevisiae strain JR Y339 (17). KAT1 was subcloned into the yeast expression vector pgk (18) and transformed into JR Y339 using the lithium acetate method (19). Growth assays were performed on SDAP medium (20) containing 2% purified Agar (Sigma A7921) supplemented with potassium at concentrations indicated in the figure legends. Patch-Clamp Experiments. Arabidopsis thaliana seedlings (L. cv. C24; Arabidopsis Stock Center, Columbus, Ohio) were grown in a growth chamber, and guard cell protoplasts were isolated as described before (21). Current measurements were performed by using an EPC-9 patch-clamp amplifier (HEKA Electronics, Lambrecht, Germany) and low-pass-filtered with an eight-pole Bessel filter. Whole-cell data were low-passfiltered with a cut-off frequency of 2 khz. Data were sampled at 2.5 times the filter frequency (for single-channel recording this factor was 5), digitized (ITC-16; Instrutech, Mineola, NY), stored on hard disk, and analyzed with the software PULSE and PULSEFIT from Instrutech on a Gravis TT200. Patch pipettes were prepared from Kimax-51 glass (Kimble Glass, Vineland, NJ) and coated with silicone (Sylgard 184 silicone elastomer kit; Dow-Corning). To determine membrane potentials, the To whom reprint requests should be addressed. 3298

Plant Biology: Brüggemann et al. Proc. Natl. Acad. Sci. USA 96 (1999) 3299 command voltages were corrected off-line for series resistances and liquid junction potentials according to Neher (22). The standard pipette solution (cytoplasm) contained 300 mm potassium gluconate 2 mm MgCl 2 3 mm CaCl 2 5 mm EGTA 2 mmmgatp 10 mm Hepes-Tris, ph 7.5. The bathing medium contained 10 mm citrate-tris, ph 4.5 or 6.0 or 10 mm Mes-Tris, ph 6.0, in addition to 0.26 mm MgCl 2 or 2.6 mm MgCl 2, respectively (0.1 mm free Mg 2 ). Potassium was adjusted by using potassium gluconate as indicated in the figure legends. The K content of all solutions was verified by atomic-absorption spectroscopy. Single-channel recordings were performed in 150 mm symmetrical potassium gluconate using standard pipette solution and a bathing medium containing 1 mm CaCl 2 10 mm citrate-tris, ph 5.0. All solutions were adjusted to 540 mosmol kg by using D-sorbitol. Chemicals were obtained from Sigma. Biophysical Analysis. Relative open probabilities were deduced from a double voltage-step protocol. Time- and voltagedependent K currents were elicited in response to hyperpolarization. During the second voltage step, K currents relaxed in a time-dependent manner. The instantaneous current voltage relationship, obtained from extrapolating the relaxation time course of the second pulse to t 0 with an exponential function, is proportional to the relative open probability, p O (V), at the end of the activation pulse. Relative open probabilities were fitted with a Boltzmann function: P O 1 1 V V 1 2 V S Here, V 1/2 denotes the voltage at which 50% of the channels are active, and V S is the slope factor that is correlated to the charge of the voltage sensor. RESULTS AND DISCUSSION When we complemented the yeast mutant lacking the K transporters trk1 and trk2 with KAT1, a K channel expressed in A. thaliana guard cells (23), growth was rescued in low-k medium (Fig. 1), a behavior that has been used previously for cloning the first plant K uptake channels, KAT1 and AKT1 (2, 3). Although the mutant containing the vector alone did not grow at K concentrations below 10 mm, KAT1 supported yeast growth in as low as 10 M K in the culture medium, indicating that K channels are able to catalyze high-affinity K uptake. FIG. 1. Potassium concentration-dependent growth of the K uptake-deficient yeast mutant JR Y339 complemented with the A. thaliana guard cell K channel KAT1 or the empty vector.. In the presence of 150 mm K, inward-rectifying K channels in guard cells as well as the gene products of KAT1 and KST1 [guard cell K channel of Solanum tuberosum (24)] expressed in Xenopus oocytes are characterized by a unitary conductance in the order of 5 8 ps (Fig. 2A; ref. 21). This conductance is the result of, e.g., 1.1-pA current driven by a voltage drop across the membrane of about 180 mv (Fig. 2A, Nernst potential for K,E K 0). According to the substrate dependency of K currents (see Fig. 3A), in 100 M K or even less, the same potential difference (E K E M ) would elicit single-channel K currents in the order of only a few femtoampere. This is far below the resolution limit of the patchclamp technique [100 fa (25)]. To perform high-resolution current recordings at K concentrations representing the high-affinity K uptake range, we used A. thaliana guard cell protoplasts characterized by a membrane patch-like wholecell surface area (capacitance 1.4 0.7 pf, n 253) but high channel density [ 5 8 channels m 2 corresponding to around 500 1,000 channels per cell (21)]. In the presence of basically 300 mm K in the pipette and 1.14 mm K in the bath, inward currents were elicited upon hyperpolarization of the plasma membrane negative to around 130 mv (Fig. 2 B and C). At potentials more positive than, e.g., 60 mv, outward-rectifying K channels were gated open (not shown here). Because this channel type is not involved in K uptake but, rather, K release, it will not be taken into account in this study. When we reduced the external K concentration to, e.g., 60 M (Fig. 2B Right), a concentration within the working range of high-affinity K transporters (1), inward currents required membrane potentials negative of 200 mv because of the change in the driving force. At voltages positive to the Nernst potential for K, channel activity resulted in outward currents, which decreased with further depolarization. Detailed analysis of the current voltage curves revealed that channel activation is K -independent from 30 mm down to the micromolar range (Fig. 2D; see also ref. 26). An identical behavior was seen with oocytes expressing KAT1 (down to 500 MK, not shown), proving high-affinity uptake to represent an intrinsic property of the K channel. Thus, in contrast to the current literature (see ref. 9 for review), this channel type conducts outward K currents positive from E K, whereas negative to the equilibrium potential it mediates K uptake. Additional evidence for this hyperpolarizationactivated K channel to function as voltage sensor rather than K sensor was obtained from the analysis of the open probability in the presence of 20, 60, 130, 330, and 1140 M K in the bath. As shown in Fig. 2D, the activation curves as well as half-activation potentials (V 1/2 ) superimposed for K levels representing the high- and low-affinity range. Therefore, the voltage range at which K efflux occurs is a function of the difference between the voltage threshold for K channel activation and the Nernst potential for potassium. The wholecell conductance of the inward current through the K channel plotted as a function of the K concentration displayed a Michaelis Menten type of behavior characterized by an affinity constant of about 2 mm (Fig. 3A), a K M similar to that of potassium channels in V. faba guard cells (6). This channel, therefore, represents a low-affinity transport system that, because of its K -independent gating, also operates in the high-affinity range. Because this channel type is capable of conducting K currents of the same order of magnitude as described for ph-dependent K uptake in plants (9), we also explored the ph sensitivity of the high-affinity K uptake component of the K channel. After activating inward K currents through membrane hyperpolarization to 220 mv in the presence of 60 M K and ph 6.0, we applied pulses of higher proton concentration to the guard cell protoplasts (Fig. 3B, ph 4.5). Upon this acidification the current amplitude increased dramatically (Fig. 3B), a feature reminiscent of H K symport-

3300 Plant Biology: Brüggemann et al. Proc. Natl. Acad. Sci. USA 96 (1999) FIG. 2. Voltage-, time-, and K -dependent properties of the K uptake channel in A. thaliana guard cells. (A) Activation of single K uptake channels in an outside-out patch in response to hyperpolarizing voltage pulses to values indicated starting from a holding potential of 67 mv. For better resolution of current amplitudes, the K concentration was 150 mm K on both sides. The current voltage relation of single channels from four to six individual measurements as a function of the membrane voltage corresponded to a slope conductance of 8.6 0.3 ps (not shown). (B) Macroscopic outward and inward K currents (whole-cell configuration) through the inward rectifier elicited by 1.5-s, too pulses to voltages indicated in the presence of 1140 MK (Left) and 60 MK (Right). Because of the experimental conditions K out channels present in the plasma membrane, too, are not active in the voltage range shown. (C) Current voltage curves of steady-state K currents in the presence of different external K concentrations. Pulse protocols are as in B. The data shown in B and C are representative of three to six independent experiments. (D) Voltage-dependent open probabilities (Upper) and half-activation potentials (V 1/2, Lower) as a function of the K concentration. Open probabilities were recorded in the presence of 30 mm (Upper, E) and 20 M K (Upper, F) and 20, 60, 130, 330, and 1,140 M K (Lower, n 3 9). ers. When comparing the current activation curves under both conditions, we could relate this increase in K current to an acid-induced shift in the voltage dependence by 65 mv to less-negative membrane potentials (Fig. 3C). This facilitation of K channel activation by protons thus is qualitatively similar to that recognized for guard cell channels from other plants as well as KAT1 and KST1 expressed in Xenopus oocytes in the presence of millimolar K concentrations (21, 24, 27 30). Plotting the reversal potentials of the macroscopic currents against the external K concentration showed that they perfectly followed the Nernst equation (Fig. 3D, 58.9 mv decade). This indicates that acid activation of the channel is not superimposed by H K symporter activity, and, consequently, proton gradients do not energize K channeldependent K uptake into guard cells. From the experimental evidence presented here we conclude further that (i) K uptake channels are capable of mediating high-affinity K transport that is stimulated upon acidification, (ii) K uptake channels are equipped with a channel-intrinsic voltage sensor rather than a K sensor, (iii)

Plant Biology: Brüggemann et al. Proc. Natl. Acad. Sci. USA 96 (1999) 3301 FIG. 3. High-affinity inward currents through guard cell plasma membrane K channels is not driven by the proton gradient. (A) Potassium-dependent saturation of the K channel conductance recorded in bath medium of ph 4.5 and K concentrations ranging from 20 M to 30 mm. Relative conductance was determined according to ref. 6 (n 2 6). (B) Proton-dependent stimulation of inward K currents during a solute pulse of ph 4.5 within 60 M K and ph 6.0. The membrane potential was clamped to 220 mv. The bar illustrates the duration of the acid pulse generated by a fast perfusion system. (C) Shift in voltage-dependent open probabilities of the K channels in response to a ph change from 6.0 to 4.5. Relative open probabilities were normalized with respect to the maximum at ph 4.5. Lines represent Boltzmann fits to the voltage-dependent open probabilities at the two ph values. The data represent mean SD of seven measurements. (D) Ideal Nernstian behavior of the K current reversal potentials. Reversal potentials (V rev ) were deduced from steady-state current voltage relations as shown in Fig. 2C in the presence of 20 1,140 M external K at ph 4.5. The line represents a fit according to the Nernst equation, revealing a shift in V rev of 58.9 mv per 10-fold change in the K concentration. the channel-intrinsic ph sensor (29, 31) is active under highand low-affinity transport conditions, and (iv) K uptake channels, upon potassium starvation, could represent a K - efflux pathway. In line with our high-resolution K current measurements on A. thaliana guard cells, the K uptakedeficient yeast mutant trk1/2, when complemented with KAT1, took advantage of the voltage-dependent properties of this ion channel. Activated by hyperpolarization and acidic ph, KAT1 does not inactivate, allowing long-term K accumulation, which is required for yeast growth (Fig. 1). To drive channel-mediated K influx with 10 M K in the culture medium, yeast cells containing, e.g., 100 mm cytoplasmic K have to pump their membrane potential negative to 232 mv. Resting potentials even more negative have been recorded in guard cells as well as in root apex cells (32, 33). Our results also have shown that at micromolar extracellular potassium concentrations, K uptake channels could represent a significant K shunt if the membrane potential is positive to E K.K symporters that are driven by gradients other than or in addition to the electrical potential are able to operate in the voltage window between E K and 100 mv or even more positive (10 13, 15, 34). In A. thaliana root protoplasts, net K uptake currents carried by these symporters could at least balance channel-mediated K efflux (35). Furthermore, conditions such as K depletion or K starvation seem to induce the expression of K symporters probably to limit channel-mediated K loss (12, 15). In this context, upregulation of nonchannel K transporters may explain why, in 10 M K, plants lacking AKT1 still reach 50% of the fresh weight compared with the wild type (32). In summary, we conclude that acid-activated, high-affinity K uptake displays an intrinsic biophysical property of plant K uptake channels extending the dynamic range of K channel action. Because K channels operate on the basis of ak concentration-independent voltage sensor, K uptake at potentials positive to E K might represent the working range of nonchannel K transporters. We are grateful to J. Vanderleyden (Leuven, Belgium) and his group for stimulating discussion as well as to D. Sanders for comments on the manuscript. For atomic absorption measurements of K concentrations, we thank W. Kaiser (Würzburg, Germany). We also thank J. D.

3302 Plant Biology: Brüggemann et al. Proc. Natl. Acad. Sci. USA 96 (1999) Reid for providing us with the yeast mutant trk 1/2. This work was funded by Deutsche Forschungsgemeinschaft grants to R.H. 1. Epstein, E. (1976) in Encyclopedia of Plant Physiology, eds. Lüttge, U. & Pitman, N. G. (Springer, Berlin), Vol. 2, pp. 70 94. 2. Anderson, J. A., Huprikar, S. S., Kochian, L. V., Lucas, W. J. & Gaber, R. F. (1992) Proc. Natl. Acad. Sci. USA 89, 3736 3740. 3. Sentenac, H., Bonneaud, N., Minet, M., Lacroute, F., Salmon, J. M., Gaymard, F. & Grignon, C. (1992) Science 256, 663 665. 4. Hedrich, R., Hoth, S., Becker, D., Dreyer, I. & Dietrich, P. (1998) in Cellular Integration of Signalling Pathways in Plant Development, NATO ASI Series, eds. LoSchiavo, R., Last, R. L., Morelli, G. & Raiknel, N. V. (Springer, Berlin), Vol. H104, pp. 35 45. 5. Hedrich, R. & Dietrich, P. (1996) Bot. Acta 109, 94 101. 6. Schroeder, J. I. & Fang, H. H. (1991) Proc. Natl. Acad. Sci. USA 88, 11583 11587. 7. Maathuis, F. J. & Sanders, D. (1995) Planta 197, 456 464. 8. Schroeder, J. I., Ward, J. M. & Gassman, W. (1994) Annu. Rev. Biophys. Biomol. Struct. 23, 411 471. 9. Maathuis, J. M., Ichida, A. M., Sanders, D. & Schroeder, J. I. (1997) Plant Physiol. 114, 1141 1149. 10. Schachtman, D. P. & Schroeder, J. I. (1994) Nature (London) 370, 655 658. 11. Rubio, F., Gassmann, W. & Schroeder, J. I. (1995) Science 270, 1660 1663. 12. Santa-Maria, G. E., Rubio, F., Dubcovsky, J. & Rodriguez- Navarro, A. (1997) Plant Cell 9, 2281 2289. 13. Quintero, F. J. & Blatt, M. R. (1997) FEBS Lett. 415, 206 211. 14. Fu, H. H. & Luan, S. (1998) Plant Cell 10, 63 73. 15. Kim, E. J., Kwak, J. M., Uozumi, N. & Schroeder, J. I. (1998) Plant Cell 10, 51 62. 16. Walker, N. A., Sanders, D. & Maathuis, F. J. (1996) Science 273, 977 979. 17. Becker, D., Dreyer, I., Hoth, S., Reid, J. D., Busch, H., Lehnen, M., Palme, K. & Hedrich, R. (1996) Proc. Natl. Acad. Sci. USA 93, 8123 8128. 18. Kang, Y. S., Kane, J., Kurjan, J., Stadel, J. M. & Tipper, D. J. (1990) Mol. Cell. Biol. 10, 2582 2590. 19. Gietz, R. D. & Schiestl, R. H. (1991) Yeast 7, 253 263. 20. Rodriguez-Navarro, A. & Ramos, J. (1984) J. Bacteriol. 159, 940 945. 21. Brüggemann, L., Dietrich, P., Dreyer, I. & Hedrich, R. (1999) Planta 207, 370 376. 22. Neher, E. (1992) Methods Enzymol. 207, 123 131. 23. Nakamura, R. L., McKendree, W. L. J., Hirsch, R. E., Sedbrook, J. C., Gaber, R. F. & Sussman, M. R. (1995) Plant Physiol. 109, 371 374. 24. Mueller-Roeber, B., Ellenberg, J., Provart, N., Willmitzer, L., Busch, H., Becker, D., Dietrich, P., Hoth, S. & Hedrich, R. (1995) EMBO J. 14, 2409 2416. 25. Sakmann, B. & Neher, E. (1995) Single Channel Recording (Plenum, New York). 26. Blatt, M. R. (1991) J. Membr. Biol. 124, 95 112. 27. Blatt, M. R. (1992) J Gen. Physiol. 99, 615 644. 28. Dietrich, P., Dreyer, I., Wiesner, P. & Hedrich, R. (1998) Planta 205, 287. 29. Hoth, S., Dreyer, I., Dietrich, P., Becker, D., Mueller-Roeber, B. & Hedrich, R. (1997) Proc. Natl. Acad. Sci. USA 94, 4806 4810. 30. Ilan, N., Schwartz, A. & Moran, N. (1996) J. Membr. Biol. 154, 169 181. 31. Hedrich, R., Moran, O., Conti, F., Busch, H., Becker, D., Gambale, F., Dreyer, I., Kuech, A., Neuwinger, K. & Palme, K. (1995) Eur. Biophys. J. 24, 107 115. 32. Hirsch, R. E., Lewis, B. D., Spalding, E. P. & Sussman, M. R. (1998) Science 280, 918 921. 33. Lohse, G. & Hedrich, R. (1992) Planta 188, 206 214. 34. Maathuis, F. J. & Sanders, D. (1994) Proc. Natl. Acad. Sci. USA 91, 9272 9276. 35. Maathuis, F. J., Sanders, D. & Gradmann, D. (1997) Planta 203, 229 236.