re. 7 3.5.! x. 3:15;. 2 - i... .. Fix; I. o 5.15:: an. .1. ,. .IT: 433...; . c9... :5... . i... 3.1:... 51... a v. I. (of. .k r an .53.; w 3 2 15.5? ’7... I is? .g... .15. . 25.. ,2. 1.. P 1 12.... 5).. 1 1 taut: 9.9.5.... .933 9.5.93...‘ {52.3, 13.! a I: ‘ 2 1:: V tyne In... ‘ 4... ~ 9.3.. «a... ’ v. r‘i‘pfiou‘fL-AIK .- .t i .- I . 18.». each... R 2:.- . . it. #28:..313. $33furjvflx \t . 3, ‘95“? 5.}... .- 1‘1 ‘7 a... .5... 1: 5’12... a. la: .1: ... . ....,,..?u.fi €312... , . .h4fi5‘r. u...) . I ,azlrfin 0.2, «£1.33: 5 . 1.. $30). . ‘ , v 1.321....) ._ .V:I9 1.12.3». u... 3 iv.\) .2 92”“ rats: maps goo l LIBRARY Michigan Sim: University #1 This is to certify that the dissertation entitled Light-induced short-term ion flux responses and associated transport mechanisms in leaf tissues of pea (Pisum sativum L. cv. Argenteum) presented by Bin Yan has been accepted towards fulfillment of the requirements for Ph.D. Horticulture degree in Jib; s: Z/JJID Major professor Date August 16, 2001 MS U i: an Affirmative Action/Equal Opportunity Institution 042771 PLACE IN RETURN BOX to remove this checkout from your record. TO AVOID FINES return on or before date due. MAY BE RECALLED with earlier due date if requested. DATE DUE DATE DUE DATE DUE 6/01 eJCIRC/DatoDuepGS-QIS Light-induced short-term ion flux responses and associated transport mechanisms in leaf tissues of pea (Pisum sativum L. cv. Argenteum) By Bin Yan AN ABSTRACT OF A DISSERTATION Submitted to Michigan State University in partial fulfillment of the requirements for the degree of Doctor of Philosophy Department of Horticulture 2001 Professor Irvin E. Widders ABSTRACT LIGHT -INDUCED SHORT-TERM ION FLUX RESPONSES AND ASSOCIATED TRANSPORT MECHANISMS IN LEAF TISSUES OF PEA (PISUM SA TIVUM L. CV. ARGENT E UM) By Bin Yan The study was conducted to investigate light-induced rapid flux responses of Ca2+, H+, K“ and Cl' and associated transport mechanisms in leaf tissues of pea (Pisum sativum L. cv. Argenteum). Light stimulation (450 nmol rn'2 3") resulted in a transient increase in net 1 in mesophyll tissues. The light-response lag time of Ca2+ influx of about 50 nmol In2 3' Ca2+ is the most fast (13 s), whereas net influxes of H+, Cl' and K+ were also rapidly induced to increase within approximately 29, 45, and 48 s by light exposure, respectively. After this initial response to light, a rapid Post-Initial Flux Response (PIFR) was measured for Ca2+ and Cl' within 2-3 min, due most likely to an increase in efflux. In contrast, PIFR of H”, K+ was slower. The light-induced alkalinization of mesophyll apoplast (0.4-0.5 pH increase) which occurred within initial 5-6 min could be explained by either C02 uptake into photosynthesizing tissues or increased K+ and Cl' uptake mediated by HJ'/K’r and H+/Cl' symports, respectively. LaCl3 (0.1 mM), a Ca2+ channel blocker, and or EGTA (l.0 mM), a carion/Ca2+ chelator inhibited the light-induced initial net Ca2+ influx, and lead to reduction of the light-induced PIFRs of Cl' and K+ around the mesophyll tissues over time. In contrast, the addition of Ca2+ ionophore A23187 (5 uM) or high Ca2+ concentration (2.0 mM) in the bathing solution produced an opposite effect. Therefore, Ca2+ influx resulting from increased Ca2+ channel activity may be an initial light response which could contribute to membrane depolarization. The plasma membrane H+-ATPase inhibitors, Na3VO3 (1.0 mM) and DCCD (0.1 mM), inhibited the light-induced initial response in net H+ influx by 50-60 % and eliminated the PIFR of H+, whereas fusicoccin (0.01 mM) significantly increased the light-induced PIFR of H+ and the initial response in net C1" influx. This suggests that H+-ATPase in the plasma membrane may not be directly involved in the initial light response in mesophyll tissues. In addition, the K+ outwardly-directed channels and the R-type anion channels are thought to play a role in the light-induced PIFR of K+ and Cl', respectively, whereas Cl' PIFR is Cay—dependent. Light-induced changes in net Ca2+, H+, K+ and Cl' flux from mesophyll tissues increased in a saturable manner with increasing fluence. However, high fluence (1800 and 2800 umol m’2 8") enhanced the temperature of the bathing solution and produced a possible heat effect on ion flux. White light (WL) of 450 umol m'2 s'1 stimulated significant changes in net flux of Ca2+, W, K" and Cl' from mesophyll tissues. Blue-light (BL) and red-light (RL) of similar fluence induced similar time course changes in net Ca2+, H+, K+ and Cl’ fluxes as WL. There were no significant differences in the magnitude of change in net Ca2+, H+, K+ and Cl’ fluxes between BL and RL. The result suggests that BL and RL play a similar role in rapid ionic responses to light. To my parents and my aunt who believe in me and whom I love. iv ACKNOWLEDGEMENTS I express sincere gratitude to my committee members, Dr. Irv Widders, Dr. Ken Poff, Dr. Eunice Foster and Dr. Eric Hanson, for the guidance and encouragement they have given me during my Ph.D. study. Their great helps made me finish the graduate program. First of all, I would like to thank my major professor, Dr. Irv Widders, for his didactic advise and tremendous contribution of his knowledge. His patience and supports are indispensable to my research. I extend my appreciation to Dr. Poff who enhanced my understanding of light signal transduction and gave me confidence. Drs. Foster and Hanson paid a lot attention and contribution to data analysis and experimental suggestion. Without their support my dissertation might be still an elementary step. I wish to give a special thanks to Drs. Ian Newman, Sergy Shabala and Olga Babourina from University of Tasmania, Australia. They provided great help with MIFE technique. A very warm thanks to Dr. Frank Dennis and Dr. Jose Saunders who edited my writing. I would like to thank Richard Wolthuis for his assistance in building wonderful chambers. Bob Forbes made a nice Faraday cage for the MIFE system. Also I express my appreciation to Dr. Erik Runkle for his friendly help of light measurements and English editing. Dr. Jose Masanbni gave me a lot of helps during my entire period in Horticulture. I would like to thank individuals in this Department — Pricilla Hockin, Joyce Jackson, Lowrie Kristi, Khanh Nguyan, Steve Brooks and HOGS people. Finally, I want to thank my parents, aunt and fi‘iends for their emotional support, friendship during my stay at MSU. TABLE OF CONTENTS Page ACKNOWLEDGEMENTS ...................... v LIST OF TABLES .......................... viii LIST OF FIGURES .......................... x LIST OF ABBREVIATIONS ..................... xiv INTRODUCTION .......................... 1 GENERAL INTRODUCTION ..................... 1 LITERATURE REVIEW ....................... 4 1. Generality of Signal Transduction ................ 4 2. Membrane Potential and Electrical Signalling ............ 6 3. Ion Fluxes and Signal Transduction ................ 9 3.1 Ca2+ fluxes and Ca2+ signalling ............... 10 3.2 H+ fluxes ........................ 15 3.3 K fluxes ........................ 19 3.4 CI‘ fluxes ........................ 22 4. Light Signal Transduction .................... 24 4.1 Photoreceptors ...................... 24 4.2 Light effects on membrane potentials and ion fluxes ..... . 26 Chapter 1. Characterization of Light-Induced Ca“, H”, K* and CI' Fluxes from Mesophyll and Epidermal Tissues of Pea Leaf Tissues ........... 30 ABSTRACT ........................... 31 INTRODUCTION ......................... 32 MATERIAL AND METHODS .................... 34 Plant material ......................... 34 Ion-selective microelectrodes ................... 35 MIF E system .......................... 36 Flux measurements ........................ 39 Experimental protocol ...................... 43 Data statistical analysis ...................... 46 RESULTS ............................ 51 DISCUSSION ........................... 69 LITERATURE CITED ....................... 81 vi Chapter 11. Possible Mechanism of Light-Induced Ca2+, HZ K+ and Cl‘ Fluxes from Pea Mesophyll Tissues ................... 85 ABSTRACT ........................... 86 INTRODUCTION ......................... 87 MATERIAL AND METHODS .................... 90 Plant materials ......................... 90 Flux measurements ....................... 90 Experimental protocol ...................... 92 Chemicals ........................... 93 Data statistical analysis ...................... 94 RESULTS ............................ 94 DISCUSSION .......................... 109 LITERATURE CITED ....................... 122 Chapter III. Effect of Light Intensity and Quality on Ion Fluxes from Pea Mesophyll Tissues ..................... 128 ABSTRACT ........................... 129 INTRODUCTION ......................... 129 MATERIAL AND METHODS ................... 132 Plant materials ......................... 132 Flux measurements ....................... 132 Light sources ......................... 133 Experimental protocol ...................... 136 Solution temperature ...................... 136 Data statistical analysis ..................... 137 RESULTS ............................ 137 DISCUSSION .......................... 150 LITERATURE CITED ....................... 155 SUMMARY ............................ 158 BIBLIOGRAPHY .......................... 164 vii LIST OF TABLES Page Chapter I. Table 1. Quantitative characteristics of the light-induced net fluxes and extracellular concentrations of HI, Ca2+, K+ and C1‘ in pea mesophyll tissues bathed in 0.5 mM CaSO4 and 1.0 mM KCl at pH 5.2 ........... 56 Table 2. Quantitative characteristics of the light-induced time responses of H“, Ca2+, K and C1' in pea mesophyll tissues bathed in 0.5 mM CaSO4 and 1.0 mM KCl at pH 5.2 ........................ 58 Table 3. Qualitative characteristics of light-induced ionic responses of pea mesophyll tissues bathed in 0.5 mM CaSO4 and 1.0 mM KCl at pH 5.2 . . . 59 Table 4. Quantitative characteristics of the light-induced induced net ion fluxes and extracellular ion concentrations in pea epidermal tissues bathed in 0.5 mM CaSO4 and 1.0 mM KCl at pH 5.2 ................ 63 Table 5. Effect of external pH in bathing solution on the light induced net fluxes of H”, Ca2+, K+ and Cl’ from pea mesophyll tissues ........... 64 Table 6. Effect of external pH in bathing solution on the light-induced extracellular concentrations of H‘”, Ca2+, K+ and Cl' and time responses in pea mesophyll tissues ....................... 65 Table 7. Effect of DCMU on the time responses of light-induced ion flux from pea mesophyll tissue ........................ 67 Chapter II. Table 1. Effects of H+-ATPase inhibitors and F C on light-induced net fluxes of H“, Ca2+, K+ and Cl‘ from pea mesophyll tissues ........ . . . . 97 Table 2. Effects of Ca2+ channel blocker and chelator on light-induced net fluxes of HI, Ca2+, K” and Cl' and time response in pea mesophyll tissues . . . 100 Table 3. Effects of K+ channel blocker on light-induced net fluxes of H‘”, Ca2+, viii K“ and Cl'from pea mesophyll tissues ................ 103 Table 4. Effects of Cl' channel blocker on light-induced net fluxes of H“, Ca2+, K+ and Cl“ and time response in pea mesophyll tissues .......... 105 Table 5. Effects of external K+ and Cl' concentrations on light-induced net fluxes of W, Ca”, K+ and Cl‘ from pea mesophyll tissues ......... 108 Chapter 111. Table 1. Effects of white-, blue- and red-light on net fluxes of H, Ca2+, K+ and Cl' from pea mesophyll tissues . . . . ............. 146 Table 2. Effects of white-, blue- and red- on net fluxes of H“, Ca2+, K and Cl' from pea mesophyll tissues at similar light fluences ...... . . . 149 ix LIST OF FIGURES Page Chapter I. Figure 1. Schematic illustration of the experimental set-up, the MIFE system and the position of the light source . . ............ Figure 2. The MIFE system and the position of the microelectrode and light source ............................ Figure 3. A segment of H+ flux data record from mesophyll to illustrate the voltage change from which the fluxes were estimated using the MIF E system . Figure 4. Time course of net ion fluxes and extracellular ion concentrations of W, Ca2+, Cl' and K+ at 1500-2000 pm from pea leaf surface using the MIF E system .......................... Figure 5. Calculation of light-induced change in net HI fluxes and extracelluler pH around pea mesophyll tissues. F0 or C0 is 5 min average of net fluxes or extracellular ion concentrations, respectively, before light-on. Fr is amplitude of light-induced Post-Initial Flux Response (PIFR). Fi is initial net ion flux or extracellular ion concentrations before PIFR. Ff or Cf is initial flux or concentration response, respectively, afier light-off ..... Figure 6. Time course of net fluxes and extracellular concentrations of H2 Ca2+, Cl' and K” around pea mesophyll tissues of one leaf sample under dark ............................. Figure 7. Li ght-induced time course in net fluxes and extracellular concentrations of H, Ca2+, Cl‘ and K+ around pea mesophyll tissues of one leaf sample .......................... Figure 8. Light-induced time course of net fluxes and extracellular concentrations of H, Ca2+, C1” and K around pea epidermal tissues of one leaf sample ......................... .44 ..47 .50 Fig Fi Figure 9. Effect of a photosynthetic inhibitor, DCMU, on light-induced initial response in net ion influx from pea mesophyll tissues. Data represents an average of 6 - 9 replicates. Vertical bars represent LSD values at p<0.05 .......... . . . . . .......... 67 Figure 10. Light-induced time course of net ion flux from pea mesophyll tissues. Data points represent a leaf sample. Net charge flux is a sum I-F + Ca” + K+ - Cl' fluxes ........................ 68 Figure 11. Light-induced time course of net ion flux from pea mesophyll and epidermal tissues. Data points represent an average of 7-12 leaf samples. Net charge flux is a sum of H” + Ca2+ + K+ - Cl’ fluxes ........... 68 Chapter 11. Figure 1. Effect of Hi-ATPase inhibitors, vanadate ('Na3VO4) and N, N- dicyclohexicarbo-diamide (DCCD), and the ATPase stimulator F usicocin (FC) on net H+ flux under low fluence (50 nmol m'2 5"). Data points represent an average of 6-8 replicates. Vertical bars represent LSD values at p<0.05 ............................ 95 Figure 2. Effect of HI-ATPase inhibitors, vanadate (Na3V04) and N, N- dicycloexicarbodiamide (DCCD), and the ATPase stimulator Fusicocin (F C) on the time course of light (450 nmol m’2 5“) induced change in net H+ flux from pea mesophyll tissues . . . . ............... 95 Figure 3. Effects of Ca2+ channel blocker (LaCl3), chelator (EGTA) and ionophore A23178 on net Ca2+ flux from pea mesophyll tissues under low fluence (50 nmol m‘2 3"). Data points represent an average of 6-8 replicates. Vertical bars represent LSD values at p<0.05 .......... 99 Figure 4. Effect of the Ca2+ chelator (EGTA) and channel blocker (LaCl3) on the time course of light (450 umol rn'2 s")-induced change in net Ca2+ flux from pea mesophyll tissues. Data points represent an average of 6-10 replicates. Vertical bars represent LSD values at p<0.05 .......... 99 Figure 5. Effect of K+ channel blockers tetraethylammonium (TEA) and Cesium chloride (CsCl) on the time course of light (450 umol rn‘2 s")- induced change in net Ki flux fiom pea mesophyll tissues. Data points represent an average of 6-10 replicates. Vertical bars represent LSD values at p<0.05 ............................ 102 Figure 6. Effect of Cl‘ channel blockers diisothiocyanatoslibene (DIDS) and anthracene—9-carboxylic (A-9—C) on the time course of light (450 umol m‘2 s' xi Ho Ch Fig Fig l)-induced change in net Cl' flux from pea mesophyll tissues. Data points represent an average of 6-10 replicates. Vertical bars represent LSD values at p<0.05 ........................... Figure 7. Effect of Ca2+ ionophore A23187 on the light-induced ion flux response of pea mesophyll tissues. Changes in net K+ and Cl' flux indicate the amplitude of light (450 nmol rn‘2 s")-induced Post-Initial Flux Response (PIFR), negative values. Data represents an average of 6-9 replicates. Vertical bars represent LSD values at p<0.05 ......... Figure 8. Effect of external Ca2+ concentration on the light-induced ion flux response of pea mesophyll tissues. Changes in net K+ and Cl‘ flux indicate the amplitude of light (450 umol m'2 s")-induced Post-Initial Flux Response (PIFR), negative values. Data represents an average of 6-9 replicates. Vertical bars represent LSD values at p<0.05 ......... Figure 9. A hypothetical model for light-induced ion flux response in pea mesophyll cells ......................... Chapter 111. Figure 1. Fluence of white, blue and red lights over a range of wavelengths . . . Figure 2. Effect of white-light intensity on flux responses HI, Ca2+, K” and Cl' around pea mesophyll tissues. Data points represent an average with 3-5 replicates. Vertical bars represent the standard error of the mean. Change in net ion influx indicates the light-induced initial response (a positive value). Amplitude of light-induced Post-Initial Flux Response (PIFR) indicates a decrease in influx or an increase in efflux or both (a negative value) ......................... Figure 3. Effect of increasing solution temperature on net fluxes H+, Ca2+, K+ and Cl‘ from pea mesophyll tissues. Data points represent an average of 2-5 replicates. Vertical bars represent standard error of the mean. Treatment temperature increases: T1 = 1-1.5 °C; T2 = 1.5-2.5 °C; T3 = 3-4 °C; T4 = 4-5°C. The base temperature was 23i1°C ......... Figure 4. Effect of white and red lights on temperature of the bathing solution. Data points represent an average of 3 replicates ............ Figure 5. Effects of light quality on time course of net flux of H*, Ca2+, K and Cl‘ fiom pea mesophyll tissues. Data points represent an average of 8- 10 replicates. Vertical bars represent LSD test result at p<0.05. Light intensity of white-light (WL), blue-light (BL) and red-light (RL) are 450, 32 and 225 nmol m‘2 s", respectively ................. xii . 102 . 107 . 107 . 122 135 . 138 . 140 143 Figure arou repi: in n Val; indi Va}! figure arc re; in \‘a in \‘E Figure 6. Effect of blue-light intensity on flux responses H+, Ca2+, K” and Cl' around pea mesophyll tissues. Data points represent an average with 3-5 replicates. Vertical bars represent the standard error of the mean. Change in net ion influx indicates the light-induced initial response (a positive value). Amplitude of light-induced Post-Initial Flux Response (PIF R) indicates a decrease in influx or an increase in efflux or both (a negative value) .............................. 144 Figure 7. Effect of red-light intensity on flux responses H”, Ca2+, K and Cl' around pea mesophyll tissues. Data points represent an average with 3-5 replicates. Vertical bars represent the standard error of the mean. Change in net ion influx indicates the light-induced initial response (a positive value). Amplitude of light-induced Post-Initial Flux Response (PIFR) indicates a decrease in influx or an increase in efflux or both (a negative value) .............................. 148 xiii i\ .4...A...filamklrddmvDmUrWMtL....UrLDnHumid. LrL.\. H\.D..RDnSS+llllrU. I. LIST OF ABBREVIATIONS A-9-C: Anthracene-9-carboxylic acid ANOVA: analysis of variance BL: blue light Ca2+: calcium ion Cl': chloride ion CsCl: Cesium chloride d: day DCCD: N-N’-dicyclohexil carbodiumide DCMU: 3-(3 ,4-dichlorophenyl)- l , 1 -dimethylurea Di: diffusion coefficient DIDS: Diisothiocyanatostilbene DSMO: dimethylsulfoxide En: Nemst potential EGTA: Ethyleneglycol-bis(B-aminoethylether)N, N, N’,N’-tetraacetic acid F: Faraday constant F C: F usicoccin H’: hydrogen ion HEPES: N-(2-Hydroxyrthyl)piperazine-N’-(2-thanesulfonic acid) K+z potassimn ion LaCl,: Lanthanum chloride LSD: Least Significant Difference MIFE: Microelectrode Ion Flux Estimation min: minus Na3VO4: sodium vanadate PIFR: Post-Initial Flux Response R: gas constant RL: red light 5: second SE: standard error T: temperature TEA: tetraethylammonium AV: electrochemical potential gradient VDI: valid data interval WL: white light 2: valency xiv INTRODUCTION GENERAL INTRODUCTION Plants are constantly exposed to fluctuations in various abiotic and biotic factors such as light, temperature, water, micro-organisms, and to mechanical and chemical stresses. Most of these signals evoke morphological, biochemical and physiological responses, indicating that there must be mechanisms in plants to perceive these external stimuli, and to elicit and transduce signals. In recent years, a great deal of progress has been made toward understanding signal transduction in plants (Felle, 1993; Leigh, 1993; Verhey and Lomax, 1993; Gilroy and Trewavas, 1994; Yang, 1996; Blumwald et al., 1998; Ebel and Mithofer, 1998; Grabov and Blatt, 1998; Genoud and Métraux, 1999). Most studies indicate that plants have developed multiple pathways to perceive and transduce signals. The inter-related molecular, cellular and tissue level events involved in triggering final responses have become important areas of research in plant science. However, the biochemical and biophysical mechanisms are not yet fully understood. External stimulation leads to changes in membrane potential and induce rapid electrical signal propagation either locally or systematically (Davies, 1987a; Wildon et al., 1992; Thain and Wayne, 1993; Rhodes et al., 1996). Most researchers agree that the electrical signalling involve ion fluxes across membrane systems (Beilby, 1984; Davies, 1987a; Felle, 1993; Wayne, 1993; Ward and Schroeder, 1997; Grabov and Blatt, 1998). r r93! lbs. +0 1.1. u WISE 8. arm Therefore, environmental stimulation most certainly induces changes of ion flux across cellular membranes. Calcium, a powerful second messenger, presumably plays a central role in signal transduction in plants. Recent investigations on stimuli-induced changes in cytosolic Ca2+ concentration have enhanced the understanding of the mode of Ca2+ action in plants (Bush, 1995; Webb et al., 1996; Blatt, 1999; Sanders et al, 1999; Trewavas, 1999). The identification of Ca2+- or Ca2+/calmodulin-dependent protein kinases in plant cells suggests that Ca2+ functions is a critical signalling ion. Further investigation is needed, however, to examine the role of Ca2+ in plants. Of the many environmental stimuli affecting plant growth and development, light may be the most important. Light has two principal functions in plants, energizing photosynthesis and photomorphogenesis. Plants have developed complex and sophisticated photosensory machinery that enable them to sense, transduce and respond to the direction, duration, quality, and quantity of light. Light-induced signalling systems in higher plants have been demonstrated by several authors (Chamovitz and Deng, 1996; Batschauer, 1998; Kurana et a1, 1998). The light responses of leaves involve at least two photoreceptors, a phytochrome and a blue/UV-A photoreceptor. An early response by leaf cells to light stimulation is a rapid change in the electrical potential across the plasma membrane (Bentrup, 1974; Cheeseman et al., 1982; Elzenga et al., 1995; Errnolayeva et al., 1996; Trebacz and Sievers, 1998). Changes in membrane potential indicate the involvement of ion fluxes in light signal transduction. Studies have attempted to determine if ion transporters, ion pumps, carriers and ion channels, in the pics: Grar 0' r__, I 101‘. USO L{) E}; 0f 1r {ECL' flue N831; ta. plasma membrane are involved in light-induced electrical signals (Mummert and Gradmann, 1991; Errnolayeva et al., 1996; Elzenga and van Volkenburgh, 1997; Szarek and Trebacz, 1999). Four inorganic ions were most highly correlated with light-induced bioelectrogensis, Ca2+, HI, KI and Cl'. Despite extensive research, it is unclear what role ion transport mechanisms play in the electrical response to light, which ions are important in electrical signalling, or whether Ca2+ is a key ionic messenger, and which photoreceptors may be linked to ion transporters and thus regulate their activity in cellular membrane. The interesting questions in need of answer include, how do leaves transduce light-elicited signals, what is the role of ion fluxes in the transduction of signals in photo-responses. In previous reports, pea leaves displayed rapid membrane depolarization, possibly involving H+, Cl' and Ca2+ (Elzenga and Van Volkenburgh, 1993; Elzenga et al., 1995; Stahlberg and Van Volkenburgh, 1999). However, there were no experimental evidence of ion fluxes for this conclusion. The Microelectrode Ion Flux Estimation (MIFE) technique has been introduced to investigate ion flux changes in root and leaf tissues (Lucas and Kochian, 1986; Newman et al., 1987; Shabala et al., 1997; Shabala and Newman, 1999). This noninvasive approach provides a direct tool to monitor ion dynamics around the plasma membrane of pea leaf cells, and to reveal possible ionic transport mechanisms. Ion-selective microelectrodes are utilized broadly in plant cell biology (F elle, 1992). This study proposes three hypotheses at the outset: (1) rapid light-induced electrical signalling results from ions fluxes driven by ion transporters localized in the plasma Klimt “1'..le .- uh AL ‘Av . LAI‘ H 055;: “E St LIT 1. Ge Lire . Mu Elk“; ~ one 0 men: Ecstn {Epic-a localig Jur 551mg membranes; (2) Ca2+ plays an initial role in affecting activity of other ion transporters within the plasma membrane; and (3) ion fluxes across the plasma membrane are influenced by light quantity and quality. Thus, the first goal of this research was to characterize light-elicited rapid changes in ion (Ca2+, K", H+ and Cl’) fluxes in leaf tissues, and investigated their relationships to early light electrical signals. Subsequently, we sought to identify the transport mechanisms involved in regulatory ion flux across the plasma membranes as well as the role of Ca2+ in light-induced ion flux responses. Finally, we measured the effects of light quantity and quality on the fluxes of Ca2+, K‘“, H+ and Cl'. This research should provide experimental evidence to elucidate photosensory mechanisms in leaves. LITERATURE REVIEW 1. Generality of Signal Transduction Like other biological organisms, plants can sense, transduce and respond to external signals via the built-in signalling systems. When stimulated, specific receptors induce one or several intracellular or intercellular sequences of biochemical and /or biophysical event leading to the modification of metabolic activities or the regulation of specific genes, which in turn produce a final response. Thus, sensory systems of plant cells typically involve three steps: perception, signal elicitation and transduction, and a localized or systemic plant response. Juniper (1976) thought that “perception” is an altered physical state, which is assumed to pass the first physiological phase. The term represents the initial step of " " . r. . ,_ vii-iii») . CCR’DOI PTCg‘ess [Bic-xi; Wm Peeing ilfiil “a signalling and involves interaction between the stimulus and the receptors in a plant cell. However, stimulus-perception may involved a complex of physical, biochemical and molecular processes. Genoud and Métraux (1999) imaged that perception is analogous to a group of interacting algorithms, including an input layer, an output layer, and an intermediate layer linking the output and the input. Thus, they suggested a Boolean network model. “Signal transduction” infers sequence from stimulus perception to final response, and may involve conversion of energy, alteration of metabolic pathways, and changes of pH and water potential. Ray (1999) thought that “signal transduction” is the common term used to define a diverse topic that encompasses a large body of knowledge about the biochemical mechanisms that regulate cellular physiology. “Response” involves downstream steps in signal transduction chains, including morphological, biochemical, or physical changes. Research on signal transduction in plants follows animal models. Some important progress in plants has been made, including the identification of the ethylene receptor (Bleecker and Schaller, 1996), phytochromes (Bowler, 1997), and detection of certain intermediate proteins, such as receptor-like protein kinases, GTP-binding proteins (G proteins), phospholipase C and Cazi/cahnodulin-dependent kinases (Gilory and Trewavas, 1994; Yang, 1996). To study plant sensory systems, molecular genetic, biochemical and biophysical approaches are used widely in plant research. Signal transduction mechanisms have been proposed in almost all environmental and hormonal signals, including gravity (Chen, 1999), water (Bray, 1997), pathogens (Blumwald et al., 1998, Ebel and Mithofer, 1998), light (Chamovitz and Deng, 1996; Khurana et al., 1998), mechanical stimuli (Verhey and Lomax, 1993), and abscisic acid (ABA) (Leung and 1‘4 I ‘i‘ldr A‘.“ 2. I] Giraudat, 1998). However, the mechanistic base of signal transduction in higher plants remains a “black box”. With the improvement of research technology, progress has been made toward understanding plant signal transduction processes. Some proposed models have been established to explain the biochemical pathways of plant response to external and internal signals (Gilory and Trewavas, 1994; Yang, 1996; Chang and Stewart, 1998). After perception by a specific receptor, the primary signal (first messenger) interacts and is transformed into transduction chains. G proteins serve a key regulatory role in initiating signal transduction. Next, the secondary messengers amplify the first messenger. Finally, the amplified signals drive and link cellular signalling pathways that result in final responses. Many secondary messengers/signalling elements are suggested in plants, including Ca2+, HI, lipids and cAMP (for reviews see Felle, 1993; Verhey and Lomax, 1993; Leckie et al., 1998). Among these second messengers, Ca2+ is best studied and accepted as a main signalling component. The Ca2+ messenger is thought to connect many signalling networks in plant cells. However, plant-signalling systems are being found to be highly complex, as new details are discovered. Interaction between signal transduction pathways may also occur when plants are stimulated concurrently by multiple factors. 2. Member potential and electrical signalling A voltage potential is defined as a voltage difference across a membrane, which is created by the net separation of positive charges fiom negative charges (Wayne, 1993). The voltage potential represents a quantitative expression of membrane potential. In plant cells. posit stirrui red item lit! Mint": Simulate tom the {Nine ‘3. . ~ ‘ 33mm. 'vl" ' , w A. r. A.” Per . . Lilia: cells, positive charges are carried by cations, e.g. KI, Ca2+, Na+ and HI, and negative charges are carried by anions, e.g. Cl', and organic acids. When a plant cell is not stimulated, the measured membrane potential is considered a resting potential with range from -100 to -200 mV. Membrane potential changes whenever the given membrane is perturbed or stimulated. When plasma membrane depolarization occurs in plant cells, there is a shift fi'om the normally negative resting potential to less negative potentials. Either a net positive charge movement into a cell or a net negative charge movement out of a cell is necessary to effect membrane depolarization. Conversely, membrane hyperpolarization is a process of either a net positive charge movement out of a cell or a net negative charge movement into a cell, which shifts membrane voltage potential to more negative. Many environmental stimuli can alter membrane potential resulting in depolarization or hyperpolarization (for reviews see Davies, 1987a; Thain and Wildon, 1993). Electrical signals involving membrane potential changes are thought to “control” many responses, for instant, plant cell growth (Pickard, 1971; Hush et al., 1991), metabolic changes associated defense to biotic factors (Wildon et al., 1989), and activation of ion channels, such as Ca2+, K+, and anion channels. Two classes of electrical signalling, action potentials and slow-wave potentials, potentially perform local or long- distance communication (Stankovic and Davies, 1998; Stankovic et al., 1998). Local electrical signalling occurs when electrical potential changes occur across membranes within a cell, whereas long-distance signalling involves the systemic transmission and Propagation of voltage potential changes along membrane systems between cells or organs. of sister induced Acri. earthy; Wfirnria 19:59; 1 cm; . and Fm organs. Many stimuli induce local membrane potential changes. However, the evidence of systemic transduction of voltage potential changes seems to be limited to wound- induced plants (W ildon et al., 1992; Rhodes et al., 1996). Action potentials in plants are characterized by a large, transient change in the membrane potential. Slow-wave potentials or variation potentials differ from action potentials by their amplitude and slower rate of propagation (Frachisse and Desbiez, 1989; Thain and Wildon, 1993; Stahlberg and Cosgrove, 1997). Action potentials are observed in a wide range of plant cells and thought to play a major role in intercellular and intracellular communication and in the regulation of ion transport and turgor at the cellular and organell levels (Retivin and Opritov, 1987; From, 1991). A remarkable example is Chara algae cell, which exhibits a large and transient action potential (Hope and Findlay, 1964). Many higher plant species generate action potentials in response to light (Trebacz and Zawadzki, 1985; Trebacz and Sievers, 1998), cold (Shiina and Tazawa, 1986; Pyatygin et al., 1992) and wounding (Wildon etal., 1992; Stankovic et al., 1998) Action potentials are considered to be the most rapid electrical phenomena in plants. The electrical signalling suggests that changes in ion flux across the plasma membrane occur. Wayne (1993) and Davies (1987a) proposed that effluxes of Cl’ and K+ and influx of Ca2+ give in part rise to the action potentials. Fromm and Spanswick ( 1993) provided an indirect evidence in willow (Salix viminalis) using ion transporter inhibitors. In light- induced liverwort (Conocephalum conicum), Cl' channels are involved in the depolarization event associated with action potentials (Trebacz et al., 1994). 3. Ion flux hbrolog}. in a cell it “r" m pidbilla mgr cmrels. p dents in lll\‘0i‘~ed membran Signals. cellular 5 Ph010rec< 1939). 5 SW11 Ce Mmbrar bi Ca3‘ ; lettiffin CZ'lol‘lasr Elan, 19. 3. Ion fluxes and signal transduction In biology, ion flux is a quantitative expression of the net movement of ions fi'om one site in a cell to another over a specific period of time and area. Ion transport across the plasma membrane is either by passive or active pathways. Passive transport, through ion channels, protein carriers and facilitated diffusion pathways, depends on electrochemical gradients of ions across the membrane. In contrast, active transport is an energy- dependent process. Membrane potential changes mediated by ion fluxes may be considered primary events in signal transduction. Ca2+, H", K+ and C1: are believed to be primary ions involved in effecting changes in membrane potential. Stimulus-induced changes in membrane potential could be coupled to perception and transduction of extracellular signals. Consequently, ion fluxes across the plasma membrane may be linked with cellular signalling pathways. Guard cells provide a valuable model system for study of photoreception, ion transport, and osmoregulation in plant cells (Serrano and Zeiger, 1989). Stomatal movement is operated through coordination of signalling elements in guard cell ion channels, and is regulated by ABA and auxin (Grabov and Blatt, 1998). Membrane depolarization in guard cells is evoked by ABA and auxin and is accompanied by Ca2+ influx and cytosolic alkalinization (McAinsh et al., 1997). In addition, outward- rectifying K+ channels are activated by membrane depolarization and a rise in Cytoplasmic Ca2+ concentration and pH (Plieth et al., 1998; Roelfsema and Prins, 1998; Blatt, 1999). Opening of anion channels leads to ion efflux and further depolarization I‘wl'll r1 .1111 and this _ and ach: model of The r membrr Danes ( 531.. 5" "T ram! 1 rommur that the secure 31.01" Calcium Men 1 UlfSSi‘ng mOdlficg ”1151010 Slé’nal rr (3.10.0135 *1. . 51E“ .«dilln: (Ward and Schroeder, 1997). The massive ion fluxes across the guard cells control turgor and thus stomatal aperture. This indicates that ion fluxes are involved in both signalling and achievement of the physiological response in guard cells. However, the similar model of stirnuli-response systems in other leaf tissues are still lacking. The ionic mechanism is based on the observation of ion influxes or effluxes across membranes. However, ion fluxes may be controlled by various transport mechanisms. Davies (1987b) has already underlined that “ions (...) are likely to play a major role, especially as local (intracellular) signals”, and that “electrical signals (...) need to attain respect from a greater number of plant researchers, as candidates for major intercellular communication devices”. Ion-distribution and ion-channel blocking experiments suggest that the fluxes of Ca2+, K+, H+, and Cl' across membrane systems may be involved in specific electrical signalling events in a transduction pathway. 3.1. Ca2+flwc and Ca“ signalling Calcium, as a second messenger, plays a central role in signal transduction in plants. When the primary stimuli perceived by receptors are transduced to intracellular messengers (e.g., Ca2+), the second messenger amplifies primary signals associated with modification of many enzymes and brings about changes in biochemical and physiological responses. Numerous reviewers have implicated Ca2+ fiinction in plant signal transduction (Bush, 1995; Blatt, 1999; Sanders et al., 1999; Trewavas, 1999). Cytoplasmic Ca2+ and Cay-dependent protein kinases have been studied to elucidate signalling mechanisms. 10 A lug cells. cu illiflfié ' and Ciel “? near“. . in lean! on SUIT Imznes 00:33:13 in. fiftiIOL (1r. J.‘ Ll m ~44 Ll I 33035 C CaE.‘ C32. pu. found ir am? (if: this is j C- 1061 The S‘to Cation W the third atli‘ralcc A huge Ca2+ concentration gradient exists across the plasma membrane. In resting cells, cytosolic Ca” concentrations are maintained within the range of 0.1-1.0 uM, whereas the level of apoplastic Ca2+ for land plants is approximately 1000 uM (Bjorkrnan and Cleland, 1991; Harker and Venis, 1991). This produces a typical concentration difference of about 10‘4 to 10'3 fold. Estimation of cytosolic Ca2+ concentration is based on some reliable techniques, including Ca2+-selective electrodes, fluorescent dyes and luminescent protein aequorin (Read et al., 1993; Roos , 2000). In general, the electrical potentials across the plasma membrane are -100 to -200 mV, and the resultant electrochemical difference in Ca2+, calculated from the Nemst equation, is approximately 50 kJ mol'1 which is comparable to the free energy for ATP hydrolysis, i.e. approximately -44 kJ mol", in living cells (Miller et al., 1990). The electrochemical gradients for Ca2+ across other cellular membranes, indicating those found the endoplasmic reticulum (ER) and tonoplast, are also large and rarely reversed (Bush, 1993). Calcium movement across membranes is facilitated by Cay-permeable channels and Ca2+ pumps. Passive influx of Ca2+ occurs mainly through Ca2+-permeable channels found in the plasma membrane. Considering ion selectivity, White (1998) suggested that there are at least twomajor classes of Ca2+ channels in the plasma membrane. The first class is relatively non-selective with respect to cations, and the other has a high single- channel conductance and is called a maxi-cation channel (White, 1993; White, 1994). The second class is referred to as voltage-dependent channels with greater selectivity for cation (White, 1994; Pineros and Tester, 1995; Pineros and Tester, 1997). In addition, the third Cay-channel has been reported, a stretch-activated channel, which can be activated by membrane tension (Cosgrove and Hedrich, 1991; Ding and Pickard, 1993). ll Among; cells. \of‘. 9 t ‘ sperm: stimuli. 1 a1. 19% martin: would bg- arrears ‘ "h I. ‘h‘ifli‘l: 51031 CO: °an pr: and Wa} 1W0 .iiPaSe 1 E0“ alfin 12mm Among the CaZI-permeable channels characterized in the plasma membrane of plant cells, voltage-activated Ca2+ channels may contribute to calcium signalling. Membrane depolarization causes activation of Cay-channels. Direct measurement using the patch clamp technique indicated that depolarization of the plasma membrane in carrot protoplasts involves activated Ca2+-permeable channels (Thuleau et al., 1994). Similar observation has been made in maize shoots (White, 1997). Such voltage gating which activates the channel may be critical during cellular signal transduction. Numerous stimuli, including blue light (Spalding and Cosgrove, 1992b), red light (Errnolayeva et al., 1996) and fungal elicitors (Kuchitsu et al., 1993), are known to induce rapid membrane depolarization. The resultant increase in Ca2+ channel activity and influx would be expected to enhance cytosolic Ca2+ concentration by influx. Concurrently, the increase in Ca2+ influx results in membrane depolarization. Therefore, voltage-gated Ca2+ channels provide a connection between stimuli and Ca2+ influx across the plasma membrane. However, compared with other ion channels (e.g. K“ channels), there is limited direct evidence that Ca2+ channels operate in the plasma membrane of leaves. Most conclusions are based on pharmacological and cell biological studies. These results only provide indirect evidence that voltage-dependent Ca2+ channels in the plasma membrane are important for the initiation of plant responses to external signals (Hepler and Wayne, 1985; Leonard and Helper, 1990; Thion et a1, 1996). Two energy-dependent Ca2+ transport systems have been shown to function. Ca2+- ATPase provides high affinity Ca2+ transport whereas proton-coupled antiporters provide low affinity Ca2+-transport (Evans and Williams, 1998). Calcium efflux from the plasma membrane is mediated by Cay-ATPase or ATP-hydrolyzing pumps and calcium 12 ethane: aall99i tantrum NKSdflCC 1937), Hg ”mm c t- u . pumam re requ? role in 1 19961 Central Wilma: cones“; mitoch. lacuoh exchangers, which catalyze an exchange of Ca2+ with H+ (Evans et al., 1991; De Michelis et al, 1992). The nH+/Ca2+ carriers have been reported in the plasma membrane (Rasi- Caldogno et al., 1987), tonoplast (Schumaker and Sze, 1986), and chloroplast thylakoid membrane (Ettinger et al., 1999). The plasma membrane CaZI-ATPase itself was considered to function as a ATP-dependent nH+/Ca2+ antiporter (Rasi-Caldogno et al., 1987). However, there is no further evidence supporting this. The Ca2+-ATPase is a primary efflux transporter functioning against a Ca2+ concentration gradient. Like the plasma membrane H+-ATPase, Ca2+-ATPase is also inhibited by vanadate (Williams et al., 1990; Camelli et al., 1992). The regulation of cytoplasmic Ca2+ involves a coordination of Ca2+ channels and the plasma membrane Ca2+-ATPase activity in plant cells. Low cytosolic Ca” concentrations are required for the stability and activity of many enzymes, and appear to play a critical role in protein synthesis in the secretory pathways (Rudolph et al., 1989; Gill et al., 1996). Evans et a1. (1991) suggested that the plasma membrane Ca2+-ATPase plays a central role in the regulation of Ca2+ homeostasis of the cell, and, in particular, the restoration of low cytosolic Ca2+ concentrations after a stimulus-induced increase in concentration. Some intracellular organelles, including vacuoles, endoplasmic reticulum, mitochondria, and even chloroplasts, are able to store and sequester Ca2+ effectively. The vacuole occupies 90% of the mature plant cell volume. It contains a high Ca2+ concentration (DuPont et al., 1990) and represents a major intracellular Ca2+ pool. The voltage-dependent or voltage—gated Ca2+ channels in tonoplasts contribute to the regulation of cytoplasmic Ca2+ concentrations (Sanders et al., 1999). These channels are sensitive to inositol-l,4,5-triphosphate (IP3) concentrations, which induce a release of 13 Ca" from 19%). A r 1~ {1121111515 | eff-3i .b:u. 15b iIIL'ZSCUC’. and 012i: plasma {Colbrx 511631; 13 C310: 0’” . ‘ Slim (high 1992)‘ Ca2+ from the vacuole (Schumaker and Sze, 1987; Alexandre et al., 1990; Gilroy et al., 1990). Afterward, cyclic ADP-ribose (cADPR) was found to activate tonoplast Ca2+ charmels (Allen et al., 1995), thereby indicating that tonoplast Ca2+ channels are ligand- regulated. IP3 is also proposed to be an important signalling messenger in cellular signal transduction (Gilory and Trewavas, 1994; Leckie et al., 1998). Calmodulin (CaM), upon binding to‘Ca”, is another unique protein which interacts with a number of key enzymes, and other structural proteins called CaM-binding proteins (Roberts et al., 1986). The plasma membrane Ca2+-ATPase, a CaM-binding protein, can be activated by CaM (Colbran and Soderling, 1990; Hsieh et al., 1991). The CaZI-binding-CaM enhances the capability of the plasma membrane Cay-ATPase to respond to an increase of cytoplasmic free Ca2+ concentrations. The spatial and temporal changes in cytosolic Ca2+ indicate that Ca2+ is involved in a wide range of cellular signalling networks, and Ca2+ oscillations are crucial to the specificity of Ca2+ signalling mechanisms (McAinsh and Hetherington, 1998). Increases in cytosolic free Ca2+ concentrations have been widely observed in response to a number of stimuli, including that of ABA (McAinsh et al., 1990), mechanical stress and cold (Knight et al., 1991; Knight et al., 1992), auxin (F elle, 1988a), red light (Shacklock et al., 1992), and during phototropism and gravitropism (Gehring et al., 1990). The role of Ca2+ in action potentials has been demonstrated (Beilby, 1984; Wayne, 1993). Calcium signalling is involved in cell grth and development (Brownlee et al., 1999). Ca2+- /CaM-dependent proteins and protein kinase C, driven by increased cytosolic Ca2+, regulate protein phosphrylation/dephosphorylation in plant cells (Poovaiah and Reddy, 14 {Ella C0551? rm mare Carl M“... «in-Qt ace; 1993; Trewavas and Malho, 1998), a primary mechanisms of signal transduction. Changes of in intra- and intercellular Ca2+ concentrations are considered to be a possible key component of cellular signalling pathways. Coordination of Ca2+ transporters in membranes is necessary to achieve the dramatic changes of cytosolic Ca” concentration, and time course in concentration oscillations. Trewavas (1999) suggested that the function of cytoplasmic Ca2+ oscillation may lie in constructing a cellular “intelligence”. Determination of changes in Ca2+ fluxes may be an important approach to the study of Ca2+ signalling mechanisms. Some common Ca2+- related signalling components, such as Cazidependent protein kinase (CDPK), and CaZI/CaM-dependent protein kinases, support the role of Ca2+ in cellular signal transduction. However, CaZI-independent pathways or other signalling systems are accepted as existing in plants (MacRobbie, 1992; Allan et al., 1994). 3. 2. H + fluxes Plant cells maintain a pH gradient across the plasma membrane of about 1.5 - 2.0 pH units, i.e., cytoplasm with pH 7.0-7.5 and apoplast with pH 5.0-5.5. The proton gradient is believed to energize many important transport systems through H+-coup1ed co- transporters, including (1) solute uptake into the cells against concentration gradients mediated by syrnporters, and (2) loading of organic compounds into the phloem for transport to sink tissues (Palrngren and Harper, 1998). Thus, this electrochemical gradient, known as the proton driving motive force (dmj), represents the sum of the membrane potential and the pH difference across a given membrane. The dmf provides 15 hipotiresr; star»; is: Proto' msmbrax membrrr lei H. e mimbm' L‘W energy for secondary active proton-driven solute transport across the plasma membrane, e.g. K+ (Marré et al., 1989) and sugars (Serrano, 1985). Mitchell’s chemiosmotic theory hypothesized that H+ gradients across biological membranes fimction as a source of energy for cellular work (Mitchell, 1961;, 1976;, 1985). Proton fluxes across the plasma membrane are chiefly catalyzed by the plasma membrane H+-ATPase (proton pump) and H+-coupled transporters. The plasma membrane H+-ATPases generate a chemiosmotic H+-gradient and are the main pathway for H+ efflux across the plasma membrane. Studies have demonstrated that the plasma membrane H+-ATPase is a proton pump and is electrogenic (i.e., it establishes a negative membrane potential that is positive on the outside, and negative on the inside) (Sze, 1985). The enzyme pumps outwardly H+ across the plasma membrane into the apoplast. The resulting extracellular acidification, i.e. efflux of positively charged H+, causes an initial repolarization of membrane potential. On the contrary, the inward movement of H” produces a membrane depolarization and cytoplasmic acidification. Biochemical and molecular research indicates that the enzyme is involved in many physiological roles (Palrngren and Harper, 1998; Sze et al., 1999). A major function of the plasma membrane H+-ATPase is the establishment of a pH gradient across the plasma membrane. Although the enzyme plays an indirect role for nutrient uptake into the cell, it actually serves as a main energy source in plant cells. For example, from the Nemst equation, En = 58.2 log [Ion]ou,/[Ion]m. A Nemst membrane potential (E,) of -175 mV (3x 58.2), [Ion].m,/[Ion]in would create a 1000-fold concentration gradient in a monovalent cation such K+; i.e., the cell can maintain 100 16 m“. K' :r me 5011 so Protor imports abut H‘ mM K+ inside by virtue of a channel protein even though K+ is present at only 100 uM in the soil solution (Stein, 1990). Proton influx into the cytoplasm is mediated by symporters, such as H+/K+ (Rodriguez-Navarro et al., 1986; Netting , 2000) and H+/Cl' (Sanders et al., 1985) in the plasma membrane. Both carriers also facilitate K+ and Cl' uptake by cells. However, the rate of transport through symporters is much slower than via K+ and anion channels (Palmgren and Harper, 1999). This is because solute fluxes through symporters are energetically coupled with H+ transport. The symport enzyme requires energy transformation for conformational changes. Clearly, both H+-ATPase and H“ coupled symports regulate proton fluxes across the plasma membrane. Experimental evidence about H+/K+, H+/Cl' or other symporters remain insufficient. A high cytoplasm pH is of utmost importance for maintaining cellular metabolic activities. A number of transporters capable of maintaining a stable cytoplasmic pH are well characterized (Sze et al., 1999). Except the plasma membrane H+-ATPase, the H‘- ATPase and H+-pyrophosphase in the tonoplast are also important pathways contributing to cytoplasmic ‘pH stat’ (Sze et al., 1992; Rea and Poole, 1993; Zhen et al., 1997). Moreover, the regulation of cytoplasmic pH regulation is thought to be linked to K+ channels (Blatt, 1992; Ilan et al., 1994), and anion channel activity (Johannes et al., 1998). In corn root tissue, Felle (1998) found that cation channels may contribute to apoplastic pH regulation. Several other mechanisms are involved in pH regulation, including the production and consumption of malate via the biochemical pH stat (Leigh, 1993). 17 -m- pli 1C membrrt. the arbor: lactic er rudamn ciropium and has more \ CZIlOSOilC $571333 5 and reduc extrusion fikag. Whollc c pH regulation in plant cells may reflect a dynamic equilibrium in ion fluxes across membranes. Protons are considered to be another second messenger in plants because of the above pH-dependent processes (Felle, 1989b; Felle, 1993; Grabov and Blatt, 1998; Leckie et al., 1998). The most remarkable observation is that ABA functions in the regulation of stomatal movement. The ABA action in stomata is strictly associated with cytoplasmic alkalinization by 0.2-0.4 pH (Iring et al., 1992; Blatt and Thiel, 1993). Blatt and Armstrong (1993) confirmed the key role of H+ as a messenger in regulating stomata aperture when they found that ABA-induced stomatal closure was blocked by low cytosolic pH. Another hormone, auxin, may also play a role in cytoplasmic pH signalling. Auxin at low concentrations can elicit stomata opening (Marten et al., 1991) and reduce cytosolic pH (Irving et al., 1992). This is because auxin stimulates proton extrusion by enhancing the activity of plasma membrane H+-ATPases (Felle, 1993). The resulting cytoplasm alkalinization activates a K+-inward rectifier channel, which leads to osmotic changes in guard cells and ultimate stomatal opening (Thiel and Wolf, 1997). In contrast, the key role of the H+ messenger for stomatal closure is achieved by activation of a K+-outward rectifier channel (Lemtiri-Chlieh and Macrobbie, 1994; Grabov and Blatt, 1997) and anion channels (Blatt and Grabov, 1997). The experimental evidence for the role of H+ as second messenger is limited to guard Cells so far. The cytoplasm pH regulation is known to link many ionic transport meChanisms. It thus is reasonable to assume that H+ messenger plays a similar role in Other tissues as guard cells. The regulation of plasma membrane ATPase may provide an connection for H“ fluxes and stimuli. Auxin has been demonstrated to activate gene 18 erpressio 1.7.70 (X: d'jr.‘ 0.5): i bhtllislu {limemc suzeestir. 3-3 A” ii; Pozusiur traction: of piam across I.‘ .9016me ci‘i’inas expression of H+-ATPase (Theolois, 1986). This enzyme was also phosphorylated in vitro (Xing et al., 1996). A high-affinity fusicoccin binding protein has also been identified in different plant tissues (Weiler et al., 1990). In addition, white light (Linnemeyer et al., 1990) and blue light (Shimazki et al., 1993) stimulate proton efflux, suggesting that H+-ATPase may be regulated by external factors. 3.3 K flux Potassium ion (K+) has been demonstrated to serve both biophysical and biochemical functions in plant cells. In the cytoplasm, K+ within the 50-500 mM range, is an indispensable cytosolic component in plant cells, a major contributor to osmotic potential of plant cells (Haschke and Liittge, 1975). Because of the high concentration gradient across the plasma membrane, K+ also exerts a dominant influence on the diffusion potential component of transmembrane electrical potential. Additionally, high cytoplasmic K” serves as a catalyst for a large number of enzymes (Evans and Wildes, 1971) and is a requirement for protein synthesis (Leigh and Wyn Jones, 1984; Memon et al., 1985). K+ fluxes across the plasma membrane are essential for many physiological fimctions, e.g., stomatal opening (I-Iumbel and Hsiao, 1970) and leaf movements (Lee, 1990). The regulation of K+ transport and cytosolic K+ homeostasis is accomplished by K+ channels and ion-coupled KI symport or antiport. High affinity uptake via a symport meChansim has been postulated in Neurospora (Blatt, 1987). The symport of HI/K+ results in K+ accumulation in the cytoplasm against its concentration gradient. The influxes of KI and W depolarize the membrane and render the apoplast alkaline (Netting , l9 den 5110'. hp: Char 2000). In contrast, a K+/H+ antiport facilitates H+ efflux coupled to K” influx. This mechanism has been referred to as a capable pump-independent pathway of short-tenn pH regulation (Marré et al., 1987; Felle, 1988b; F elle, 1989a). Two classes of K+ channels have been characterized; an outwardly-directed K+ channel facilitating K” efflux and an inwardly-directed K+ channel facilitating or K+ influx. The activity of both channels depend strongly upon their membrane potentials, various effectors, channel. blockers, as well as their respective gating mechanisms (Schroeder, 1992; Blatt and Thiel, 1993). Known K+ channels are highly permeable to K”, which is larger than for other monovalent cations, such as Pb+, NH+, and Na+ (Bentrup, 1990). Compared as to other cation channels, important progress has been made in understanding the molecular mechanisms and regulation of K+ channels (Czempinski et al., 1999). Two types of inwardly rectifying K+ channels (KTAl , AKTl) from Arabidopsis thaliana have been cloned (Anderson et al., 1992; Sentenac et al., 1992) Potassium inward rectifiers, or inwardly-directed K+ channels were originally demonstrated in guard cells (Schroeder et al., 1987; Blatt, 1992). These channels were shown to be activated in a voltage- and time-dependent manner when membranes are hyperpolarized fiom -50 mV to -200 mV (Blatt, 1992). Thus the inwardly-directed K" Channels are considered to represent a pathway for K+ uptake by plant cells. Another pathway for K+ influx is K“-H+ symport. However, it is not easy to distinguish SYn'Iporters from K" inward rectifiers (Hedrich and Schroeder, 1989). Another class of K+ channels, outward rectifiers, or outwardly-directed K+ channels, 20 hurls 193’; Ric chmds, ward: 1966.821? Sumner; 5 in gen dmmd finder Pins. 199; bill lilCiCa; represent a main mechanism for K+ efflux across the plasma membrane. This type of K” channel is activated by depolarization of the membrane potential (Iijima and Hagiwara, 1987; Moran et al., 1988; Schroeder, 1989). In guard cells, outwardly-directed K+ channels play a major role in stomatal closure (Thiel and Wolf, 1997). The same K+ outward currents were also observed in Arabidopsis thaliana mesophyll cells (Miedema et al. , 2000). The efflux of K+ mediated by outwardly rectifying K”r conductance may result in repolarization of action potentials in both algae and higher plant cells (Sikaoka, 1 966; Beilby and Coster, 1979; Simons, 1981) and takes part in motor cell movement of Samanea saman (Satter et al., 1988). In general, K+ fluxes mediated by K+ channels are thought to possibly be involved in signalling in guard cells, or participate in electrical signalling in excitable plants. Both classes of K+ channel are sensitive to changes in apoplastic pH and Ca2+ concentration (Herich et al., 1990; Blatt and Armstrong, 1993; Grabov and Blatt, 1997; Roelfsema and Prins, 1998). Cytoplasmic acidification reduces the conductance of the outward rectifier but increases that of the inward rectifier. Increased cytoplasmic Ca2+ is known to inactivate the K“ inward rectifier (Kelly et al., 1995) but to activate another one, i.e., the K+ outward rectifier (Schroeder and Hegiwara, 1989). The involvement of K” channels in Sensing numerous stimuli suggests that K+ fluxes or channels play a potential role in certain signalling pathways similar to that found in guard cells. Assmann (1996) hYpothesized that G-protein regulation of K+ channels may be a main ABA—signal transduction pathway. Further regulation of K+ channel gating may be accomplished by G proteins and protein phosphorylation (F airley and Assmann, 1991; Kelley et al., 1995). 21 i . me: 10: 3“ “ ' ryzumibl‘ y 3.4., U i ‘ ' 011030 “3.3630; W531i: Jm‘ mslan 9355i) 5 9 hr ,1 ”$007 Therefore, K+ fluxes or channels may be involved in many processes, including regulation of pH and Ca”, and hormone signalling. 3. 4. C1” fluxes Chloride ion (Cl') is a predominant inorganic anion in plant cells. A large electrochemical gradient for C1' exists across the plasma membrane. For example, a positive potential difference of 300-340 mV is typically present across the plasma membrane in Chara (Smith and Walker, 1976). The concentration gradient drives passive Cl’ efflux from the cytoplasm. Chloride efflux also play an important role in maintaining a negative membrane potential. It thus can be anticipated that anion channels are a primary route of Cl' efflux from the cytoplasm, while the influx or uptake for C1’ is mediated by a carrier mechanism (Assmann and Zeiger, 1987). Zeiger et a1. (1978) hypothesized that Cl‘ enters via a Cl'/H+ symport or a Cl'/OH‘ antiport. Chloride influx requires the input of a considerable amount of metabolic energy. The pH gradient across the plasma membrane, or proton motive driving force, is thought to provide this energy. Chloride import through a 2H*/Cl‘ symport against an opposing Cl' concentration gradient has been proposed and has been demonstrated experimentally in Chara (Smith and Walker, 1976; Sanders, 1980; Beilby and Walker, 1981; Sanders et al., 1985) and root-hair cell of Sinapis alba (F elle, 1994). However, experimental proof in higher plants has remained somewhat scant so far. Anion channels have higher selectivity for C1‘ than for malate (Hedrich and Martin, 1993; Schmidt and Schroeder, 1994). Many elements, such as nucleotides (Herdich et al., 1990), pH (Tyermen et al., 1986), malate (Hedrich and Marten, 1993), and 22 l phospnor am it: u . ‘1' ‘ Cildinfr: um i L Cr .1 . tot-*MA ' L :u AQ- (05.1.4. ..,.r.,. mluA ( (CiCih phosphorylation/dephosphorylation events (Terry et al, 1991; Hedrich and Marten, 1993), affect the activity of anion channels. Cytoplasmic Ca2+ has been shown to regulate anion channels (Hedrich et al, 1990; Schroeder and Hagiwara, 1990). A CaZI-activated Cl“ channel has been observed in an inside-out membrane patch of Chara plasma membrane (Okihara et al., 1991), whereas Cl' efflux in perfused Chara cells is sensitive to Ca2+ influx (Mimura and Shimmen, 1994). An anion-conducting chloride channel family (CIC) has been cloned (Lurin et al., 1996; J entsch and Giinther, 1997). Two modes of anion channel activity have been identified in guard cells: slow-type (S-type) and rapid-type (R-type) (Schroeder and Keller, 1992; Schroeder, 1995). The slow and sustained activation of S-type anion channels would produce a long-term depolarization which last a range of several minutes or longer (Schroeder and Hagiwara, 1 989; Linder and Raschke, 1992). ABA and Ca2+ cause long-term depolarization (Bisson and Gutlmecht, 1980; Findlay, 1982; Blatt, 1987), leading to the suggestion that the S- type anion channels may be activated by either ABA or Ca2+. On the other hand, the rapid anion channels (R-type), produce a short, transient depolarization of the plasma membrane (Hedrich et al., 1990). In light of voltage-dependent, anion channels can also be divided into at least two major classes: depolarization- (Tazawa et al., 1987; Keller et al., 1989; Schroeder and Hagiwara, 1989) and hyperpolarization-activated (Tyennan et 31-, 1986; Terry et al., 1991). Another class of anion channels can be activated by Stretching of membrane (Falke et al., 1988; Schroeder and Hedrich, 1989; Cosgrov and Hedrich, 1991). Tyerman (1992) and Schroeder (1995) suggested that anion channels play key roles in Controlling cellular functions, including turgor- and osmoregulation, stomatal movements 23 and anion transport. Most results in studies of guard cells indicate that anion channels in the plasma membrane may provide a control mechanism for stomatal closure. The proposed model suggests that anion channel activation and the resulting anion efflux from guard cells cause further membrane depolarization, thereby activating the non- selective Ca2+ channel which ultimately driving K+ efflux through outwardly-directed K+ channels required for stomatal closure (Schroeder and Hagiwara, 1989; Ward and Schroeder, 1997; Asamann and Shimazaki, 1999). Therefore, in guard cells, anion channels may be involved in ABA-induced stomatal movement by linking the Ca2+ messenger, Cl' and K+ effluxes. Furthermore, anion channels-controlling Cl' efflux constitutes one of main ionic components in leaf movement (Satter, 1988; Coté, 1995). Sensitivity of anion channels to a wide range of stimuli provides a possible role for involvement of Cl‘ flux in cellular signal transduction. Johannes et a1 (1998) proposed that anion channels play a central role in cytosolic pH regulation in plants, in addition to their established roles in turgor/volume regulation and signal transduction. 4 Light Signal Transduction 4. 1 Photoreceptors Light plays a primary role in regulating the growth and development of plants. Two important processes in plants, photomorphogenesis and photosynthesis are driven directly by light. Light also provides plants with many different kinds of information about their environment, such as quantity, quality, and direction. This information, known as light Signals, is perceived and integrated by a complex of sensory systems, and used to Optimize a variety of physiological and morphological responses. Recently, much 24 3051211011 pairs a} s in plan? photoreci pile-rose; Ame: sensier: present :r 199 LJJ ; CI. bOLL'ld to forms: a r Sex-m] m lBowler e “130111161“; messenger Prom b5 Until r Pom biUe ‘Kauiman. attention has focused on photosensory mechanisms in plants, and several light signalling pathways have been proposed (Chamovitz and Deng, 1996; Batschauer, 1998; Khurana et al., 1998). The most important progress has been identifying photoreceptors. Three classes of photoreceptor systems have been suggested for the photoperception in plants. These include phytochromes, blue/UV-A photoreceptors, and UV-B photoreceptors. The presence of multiple photoreceptors provides an approach to explore photosensory machinery. Among these photoreceptors, phytochromes are the best characterized with specific ' sensitivity for red/far red light in plants. Five phytochrome genes (phyA ~ phyE) are present in Arabidopsis (Sharrock and Quail, 1989; Quail et al., 1991; Neuhaus et al., 1993; Clark et al., 1994). Light absorption is mediated by a chromophore covalently bound to an apOprotein, and each phytochrome can exist in two photo-interconvertible forms: a red (R) absorbing Pr form and a far-red (FR) absorbing Rfr form (Bowler, 1997). Several models of phototransduction mediated by phytochromes have been hypothesized (Bowler et al., 1994b, Chamovitz and Deng, 1996). Biochemical and molecular research suggested that GTP binding proteins (G-proteins) are major components of signal machinery (Clark et al., 1993). Ca2+ and cyclic GMP (cGMP) are utilized as second messengers in phytochrome signalling (Neuhaus et al., 1993; Bowler et al., 1994a). However, the primary mechanism of phytochrome action remains to be determined. Progress in understanding photoregulation of plant processes activated by blue-light has until recently lagged far behind the phytochrome work (Briggs and Liscum, 1997). Four blue-light responses in higher plants have receiving concentrated attention (Kaufman, 1993). Two responses occur in the leaf (blue-light-induced increase in 25 stomatal aperture and blue-light-induced change in gene expression) and two responses occur in the stem (blue-light-induced suppression of stem elongation, blue-light-induced curvature of the stem, i.e. phototropism). However, blue-fUV-absorbing photosensory systems are still elusive. Higher plant blue-light photoreceptor have not been definitively isolated nor characterized (Short and Briggs, 1994). Based on their absorption properties, flavins, proteins or carotenoids have been proposed as blue-light receptors (cryptochromes) in higher plants (Galland and Senger, 1991; Horwitz and Berrocal, 1997). Phototropism is representative of blue-light-induced responses in plants. Briggs and co-workers have identified in maize coleptiles and etiolated pea seedlings a plasma- membrane—associated protein of about 120 kDa, which is connected to protein phosphorylation following exposure to blue light (Gallagher et al., 1988; Short and Briggs, 1990; Short et al., 1991; Palmer et al., 1993). Quarmby (1994) and Chamovitz and Deng (1996) have proposed that Ca2+ and cAMP have a role in blue-light signalling pathways. However, to our knowledge, the mechanism of blue-light signal transduction is still unclear. 4.2 Light effects on membrane potentials and ion fluxes An early response by leaf cells to light stimulation is a transient change in the membrane potential. Bentrup (1974) found that light induces changes in transmembrane electrical potential in green tissues. The transition from light to dark usually reduces the extracellular membrane potential prior to a subsequent increase, while the reverse transition has an opposite effect in T radescantia albbiflora leaves (Giroldini, 1988). A similar transient depolarization of membrane potential were found in light-induced 26 mesophyll cells of Dionaea muscipula Ellis (Trebacz and Sievers, 1998), pea (Elzenga et al., 1995), the intact leaves of Atriplex and Chenopodium (Liittge and Pallaghy, 1969), T radescantia and Commelina (Cheeseman et a1, 1982), Arabidopsis (Spalding et a1, 1992), as well as in several other aquatic and terrestrial plants (Novak and Ivankina, 1975). Most electrical responses to light involve membrane depolarization. However, light-induced membrane hyperpolarization has also been observed (Mimura and Tazawa, 1986; Takeshige et al., 1992). The changes in membrane potential indicate that light induces underlying changes in ion flux across membranes and infer the involvement of ion transporters. In the Characean algae, the action-potential-like depolarization results from activation of Ca2+ (Weisenseel and Ruppert, 1977) and Cl‘ channels (Lunevsky et a1, 1983). Chloride pumps as well as Ki and Cl' channels were involved in action potentials in Acetabularia (Mummert and Gradmann, 1991). In gametophytes of Asplenium trichomances, light- induced depolarization was suppressed by C1' channel blockers (Szarek and Trebacz, 1999). Blom-Zandstra et a1. (1997) reported transient light-induced changes in ion channel and proton pump activities in tobacco mesophyll protoplasts, and suggested that these ion transporters in the plasma membrane were involved in light-induced transient changes in the membrane potential. In Arabidopsis mesophyll, white light induced an increase in transient membrane depolarization and in K" channel activity within the plasma membrane (Spalding et al., 1992). A light-induced anion channel in the plasma membrane of pea mesophyll cells has been identified by Elzenga and Van Volkenburgh (1997). Shabala and Newman (1999) reported a light-induced flux response involving 27 Hi, Ca2+, Cl' and K” around bean mesophyll tissues. Light increased Ca2+ flux into xylem exudate from excised tomato petioles within less than 4 s after light-on or light-off treatments (Ries et al., 1995). Light also stimulated changes in the pH of the apoplasm of pea leaves (Van Volkenburgh and Cleland, 1980; Staal et al., 1994; Stahlberg and Van Volkenburgh, 1999). Similar result was reported in the cytoplasm of the unicellular green alga Ermosphaera viridis (Bethmann et al., 1998). Both red and blue light are thought to stimulate stomatal opening (Assmann and Shimazaki, 1999). Red light, matching the absorption spectrum of chlorophyll, stimulates an outward electrical current that may be carried by the plasma membrane H”- ATPase in Vicia guard cell protoplasts (Serrano et al., 1988). Blue light presented-as a single pulse stimulates stomatal opening (Zeiger et al., 1987; Zeiger, 1990) and induces proton extrusion from guard cell protoplasts (Assmann et al., 1985; Shimazaki et al., 1993). In addition, since an anion channel in Arabidopsis hypocotyl is activated by blue light (Cho and Spalding, 1996), blue light would probably affect anion channels in guard cells. Using calcium-sensitive fluorescent dyes, Shacklock et al. (1992) have shown that red light induces proplast swelling and causes a transient Ca2+ increase followed by a rapid reduction in cytoplasmic Ca2+ to below resting levels. The blue-light-induced membrane depolarization in etiolated cucumber hypocotyls involves inhibition of the proton pump and influx of calcium (Spalding and Cosgrove, 1992b), whereas K+ channels are activated by both phytochromes (Lew et al, 1990) and blue-light activated (Suh et al. , 2000) Clearly, light elicits transient electrical signals facilitated by ion fluxes across the 28 plasma membrane of plant cells. These changes may be one of rapid pathways of light signal transduction. Athough biochemical and molecular technology are becoming dominant approaches for the investigation of signalling intermediates and to elucidate the proposed signal transduction pathways, the ionic approach is still an important and powerful tool to understand photosensory mechanisms. 29 Chapter I. CHARACTRIZATION OF LIGHT-INDUCED CA”, H*, K+ AND CL' FLUXES FROM MESOPHYLL AND EPIDERMAL TISSUES OF PEA LEAF TISSUES. 30 ABSTRACT Peeled mesophyll and epidermal tissues fi'om excised leaflets of pea (Pisum sativum L. cv. Argenteum) bathed in 0.5 mM CaSO4 and 1.0 mM KCl at pH 5.2 were exposed to an approximately 30 min light (450 nmol m‘2 5") period. The net flux rates of Ca2+, Hi, K+ and Cl’ around mesophyll and epidermal tissues were measured simultaneously by noninvasive ion-selective microelectrodes. Light stimulation resulted in a transient net Ca2+ influx an increase in influx of about 50 nmol m'2 5'1 within 1 min in mesophyll tissue. The light-response lag time of Ca2+ is the most fast (13 5), whereas net influxes of H, Cl’ and 1C were also rapidly induced to increase within approximately 29, 45, and 48 s by light exposure, respectively. A light-induced extracellular alkalinization (pH increase of 0.4 - 0.5) was observed that lasted 5 - 6 min. Following the light-induced initial flux response, a rapid post-initial flux response occurred for Ca2+ and Cl‘ which led to a net efflux within 3 - 4 min. When light was turned off, net flux of all four ions decreased and then gradually exhibited a small change in magnitude. In contrast, in the epidermis, light only induced slight changes in the flux of Ca2+, H”, K+ and Cl'. Compared to pH 5.2, changing pH in the bathing solution to a more acidic conditions (pH 4.0) increased further the light-induced initial response in net K+ and Cl' influx, whereas alkalinization of external medium (pH 6.0) produced an opposite effect on K+ influx and enhanced the light-induced Post-Initial Flux Response of Cl' (i.e. a net efflux). Inhibition of ion flux by DCMU (0.05 mM) indicates that the light-induced change in ion fluxes is dependent upon photosynthesis. These results suggest that the light-induced increase in net Ca2+ influx is an initial response that may contribute to membrane depolarization. 31 Light-induced initial apoplastic alkalinization could be explained as either CO2 uptake by mesophyll tissues or increased uptake of K+ and Cl‘. INTRODUCTION Light-induced signalling systems have been found to exist in higher plants. One of the most remarkable light responses is a rapid change of membrane potential. Almost all green plant cells investigated react to sudden illumination by light-induced electrical signals. Because of the complexity of the effect and differences among plant species, either hyperpolarization (Mimura and Tazawa, 1986; Takeshige et al., 1992) or depolarization (Elzenga et al., 1995; Szarek and Trebacz, 1999) has been measured. Most results, however, showed that light induces a transient depolarization. The mechanism of light-induced electrical signalling is not yet entirely understood. Membrane potential represents a voltage difference between cations and anions across a given membrane. Membrane potential changes indicate that changes in ion flux across membrane systems occur. Therefore, it is possible that light also induces changes in ion fluxes across the plasma membrane of plant cells. Four ions, Ca2+, H“, K“ and Cl“, were shown to correlate with the light-induced bioelectrogensis. Some researchers tried to look for whether ion transporters in the plasma membrane are involved in light-induced electrical signalling (Mummert and Gradmann, 1991; Ermolayeva et al., 1996; Stahlberg and Van Volkenburgh, 1997 ; Szarek and Trebacz, 1999). Opinions in previous reports about the ionic basis involved in the light-induced electrical response are still controversial, and possibly there are somewhat different mechanisms in different species and even different tissues of the same species. In 32 ior. 66.: K01 New 1995 addition, experimental technology is often different in some reports. The characterization of the light-induced ionic responses, the magnitude of the responses, the specific ions involved and time responses depend strongly on the ionic compositions (Elzenga et al., 1995; Johannes et al., 1997) and pH (Prins et al., 1982; Remis et al., 1994; Shabala et al., 1997) of the medium environment. Pea leaves have been shown to display a rapid membrane depolarization followed by a slower membrane repolarization (Elzenga and Van Volkenburgh, 1993; Elzenga et al., 1995). These reports suggest that H, Ca2+ or Cl' fluxes across the plasma membrane may be involved in the light-induced electrical signals of pea mesophyll. However, there is an incomplete understanding about which ion is acting as the depolarizing agent in the initial phases of plasma membrane depolarization and how the ions are involved in the event. To study light-induced ionic responses, some research groups have used ion-selective microelectrodes to measure membrane potential changes. Ion-selective microelectrodes have been utilized broadly in plant cell biology. The Microelectrode Ion Flux Estimation (MIF E) technique was introduced to investigate ion fluxes around root (Lucas and Kochian, 1986; Newman et al., 1987; Shabala et al., 1997) and leaf tissues (Shabala and Newman, 1999). This non-invasive approach provides a direct tool to monitor ion dynamics around the plasma membrane of pea leaf cells, and to reveal rapid photosensory mechanisms. Shabala and Newman (1999) found that light induced a rapid increase in Ca2+ influx in bean mesophyll tissues and proposed that Ca2+ influx was involved in light- induced depolarization. However, Cl' efflux has also been proposed to be involved in the depolarization response to light stimulation (Spalding and Cosgrove, 1992; Elzenga et al., 1995) 33 Reports about the role of proton flux in plasma membrane electrical signalling are also contradictory. Plasma membrane H*-ATPase activity was reported to increase following light illumination (Marré et a1, 1989; Linnemeryer et al., 1990; Remis et al., 1994). The enzyme can mediate H+ efflux from leaf cells which leads to apoplastic acidification. However, it has been observed in many cases that light causes an initial extracellular alkalinization in leaf tissues (Prins et al., 1982; Shabala and Newman, 1999; Stahlberg and Van Volkenburgh, 1999). The reason for this contradiction could be that there has not been direct evidence to explain the experimental relationship between H+ fluxes and extracellular pH under light exposure. One can expect that external pH influences ion transports across the plasma membrane. Therefore, measurement of ion fluxes from pea mesophyll in the changing external pH environment may provide insights into pH and H” flux responses induced by light. The first investigation presented in this chapter deals with the characterization of light-induced ion fluxes from mesophyll and epidermal tissues of pea leaves. To explain the ionic basis of the light-induced transience of membrane depolarization of pea leaf tissues, the kinetics of Ca2+, K+, H“ and Cl‘ fluxes across the plasma membrane was measured using the non-invasive MIFE technique. External pH effects on light-induced ion flux changes are discussed. MATERIAL AND METHODS Plant Material Pea (Pisum sativum L. cv. Argenteum) seeds were soaked in distilled water for 6-8 hours, 34 and then germinated in an artificial clay shale (Turface manufacture) at room temperature (23i1°C). After 4-5 days of germination, the young seedlings were transplanted and transferred to a growth chamber with a 16 hour photoperiod (200 nmol m'2 s'1 at the top of the canopy), day/night temperatures of 20/25 °C and 75% relative humidity. The plants were cultured in a modified Hoagland solution (Johnson et al., 1957) following an increase in concentration from 0 —) ‘A —> V2 at 4-5 d intervals after transplanting. Ion-Selective Microelectrode Glass capillaries (catalog no. MIB 150-6; World Precision Instruments, Inc., Sarasota, FL, USA) were pulled using a two stage puller (PC-10, Narishige, Tokyo, Japan) to produce micropipettes with a tip diameter of about 2 pm. After pulling, the micropipettes were dried in the oven at 200°C for 4-5 h, and silanized with N- dimethyltrimethyldiylarnine (catalog no. 41716, F luka, Milwaukee, WI). Cooled microelectrodes were back-filled with 500 mM CaCl2 for Ca”, 500 mM KCI for K+ and Cl', and 15 mM NaCl plus 40 mM KHZPO, (adjusted to pH 6.0 with NaOH) for H. Electrode tips were then front-filled with a 50-80 rim-long column of the appropriate ion- selective sensor, including H+ (no. 95291), Ca2+ (no. 21048), K“r (no. 60031) and Cl' (no. 24902) cocktails (all from Fluka, Milwaukee, WI). Silanized micropipettes were used within one week of preparation, while the filled microelectrodes only were used within one-day period. Microelectrodes were calibrated for each ion using a known set of standard solutions before and after use. The standard solutions contained a pH 4.0 -7.0 for H, 0.1 — 2.0 mM 35 CaSO, for Ca”, or 0.1-5.0 mM KCl for K and CI. The average slope was 52-55 mV / pIon for monovalent ions and 26-29 mV / pCa2+ for Ca2+ electrodes. An Ag/AgCl reference electrode was prepared using a 4-5 cm section of 2 mm diameter polyethylene tubing filled with 2 % agar in 1 M KCl. A section of silver wire (diameter 0.38 mm) was coated in 0.2 M HCl. The coated silver wire was then inserted into one end of the tubing containing 2 % agar and wrapped with parafilm to fix in position. The reference electrode was stored in 1 M KCl solution and only used 1 (1. During measurements, the end of the reference electrode with silver wire was connected to the preamplifier, while the open end of the tubing was bathed into solution being measured. MIFE system Ion concentrations at the surface of the leaf mesophyll cells were measured using a non- invasive Microelectrode Ion Flux Estimation (MIFE) system (MIFE, Unitas Consulting, Hobart, Australia) (Shabala and Newman, 1997; Shabala et al., 1997). From time course data obtained by the MIFE system, the directionality and rates of ion flux could be estimated. Additional information on the MIFE is available at http://www.phys.utas.edu .au/physics /biophys. The set-up of the MIFE system is illustrated in Figure 1 and 2. An inverted microscope (Leitz Wetzlar, Germany) was placed on an anti-vibration table that was installed inside a Faraday cage to minimize electrical and magnetic noise. Three ion- selective electrodes were mounted on a coarse multi-microelectrode holder/manipulator 36 with three-dimensional positioning ability. After a tissue ample/solution chamber (8.0x 2.5 x 1.5 cm) was fastened to its holder, the electrodes were positioned at a 50 um distance parallel to the tissue surface. The distance between adjacent electrode tips was about 4-5 pm. For fine positioning of the microelectrodes, the holder supporting the tissue sample/solution chamber was controlled by a three way hydraulic micromanipulator (WR 88, Narishige, Tokyo, Japan) driven by a computer-controlled stepper motor (MO61-CE08, Superior Electric, Bristol, CT). Thus, the ion-selective electrode could be moved to any programmed position through computer control of the stepper motor. In this study the electrodes were moved vertically from 50 to 90 um above the leaf surface in a 20-s square-wave cycle with a frequency of 0.1 Hz, i.e. 40 pm for each movement (a half cycle) with 10 5 pass at each position. The three ion-selective electrodes and the reference electrode were connected to a pre- amplifier inside the Faraday cage. The pre-amplifier was then linked to a main amplifier outside the Faraday cage. A personal computer with an A/D card was used to monitor and record all measurements (Figure 1, 2). A small Plexi-glass curved mount in the shape of arc with a diameter 1.2 cm was positioned at the bottom of the measuring chamber. This mount caused a gentle bending of the leaf segment mounted in the chamber. Two braces were used in the chamber to prevent dislocation of the leaf segment from the mount. This set-up allowed for a clear view of leaf cells for electrode positioning through the inverted microscope. The microscope light position, tangential to leaf surface, provided background illumination of approximately 50 nmol m‘2 s'1 (Figure 1). The microscope lamp was turned off prior to 37 Main Amplifier 7W ¢ I Computer ['\ “V... m -_, w_ a- l,/ I Pre-ampllfier /,.’ A .// l N . .»’/ ‘— Mrcroelectrode holder .L/ 2 Microscope Iggy Light projector ’ ___.. Bathing solution / Chamber holder 5 I Microscope I L/ ‘ i. r I i /' ’ +—> ' /\ . .\ \ O Faraday Cage Hydraulic micromanipulator Stepper motor Figure 1. Schematic illustration of the experimental set-up, the MIFE system and the position of the light source 1: Micromanipulator; 2: Microelectrodes; 3: Ag/AgCl reference electrode; 4: leaf segment; 5: Measuring chamber. 38 iiiiliil.‘,'. concur calcui: Gilli COH'JO and 2 UN cont: mm b)‘ m. C1035 {Or iii: iOn ac initiating any experiments in order to allow the tissue to equilibrate the new environment. A fiber-optic light (450 nmol m2 s") was used as a light source (Figure 2). The 250 W lamp was from GS Edmumd Scientific Company (Barrington, NJ). A lens (no L-16- 16, World Precision Instruments, Inc., Sarasota, FL) was fixed to the top of a light guide to focus the beam. Light intensity was measured using a quantum sensor (model LI-189, LI-COR, Lincoln, NE). Flux measurements The MIFE system utilizes slowly vibrating-specific ion microelectrodes, enabling concurrent measurement of multiple ions. The MIF E data recording system and flux calculation protocol have been described by Shabala et al. (1997). In brief, a software CHART package (Unitas Consulting, Hobart, Australia) provides an automatic record, control all measurement processes in a real-time display on the computer screen (Figure 1 and 2). After measurements are completed, the recorded data files can be transferred by CHART to MS—Excel files. These programs were then used to calculate average ion concentrations and mean net fluxes at 10-s intervals (each half-cycle of manipulator movement). The concentration of each ion in the bathing solution was calculated for each position by measurement of voltage potential on the basis of Nemst equation. Ion concentrations close to the leaf were calculated from the calibration data and from the average mV value for the two positions. The magnitude and the directionality of the difference in estimated ion activity between the two positions, perpendicular to the leaf tissue surface, the 39 the microelectrode and light source Figure 2. The MIFE system and the position of the microelectrode and light source 40 concentration gradient, was the basis for estimating both the rate and directionality of the ion fluxes. The estimation methods for ion flux using ion-selective microelectrodes have been described by Newman et al. (1987), Kochian et al. (1992), and Shabala et al. (1997). Basically, ion fluxes were calculated using F ick’s first law of diffusion assuming a planer diffusion profile: D, (C, - C2) J.= —— (1) where J, is the net ion flux, D, is the self-diffusion coefficient for the ion in equation (in cm'2 s"); Cl and C2 are the ion activities at the two positions, A r is the amplitude of electrode movement. In this study, A r is 40 um. The net ion flux is affected by D,, temperature, leaf shape, Ar and electrochemical potential gradient. Temperature influences the Nemst equation according to E, = RT/zF In (C/Ce), where T is temperature, 2 is valency, F is Faraday constant (96,487 J mol" volt“), R is gas constant, and C/Ce is concentration difference between the membrane. The temperature was ignored due to a constant room temperature (23 °C). D, was used for Ca2+ (8.0 x 10'2 cm"2 s"), H+ (9.17 x 10'5 cm'2 s"), CI' (2.032 x 10-5 cm'2 s") and K+ (1.9 x 10’5 cm'2 3"). Leaf shape could be considered as a planar system. The general equation for estimation of planar diffusion based upon measurements using the MIF E system is: J = c u F (SS/Nemst slope) (dv/dx) (2) where c = ion concentration; u = mobility, and dv/dx is voltage gradient over distance x. 41 tier: elect} peter more 001621 that r The Nemst slope is obtained from a calibrated curve over a range of ionic concentration. The mobility u is specific for each ion and is related to the diffusion coefficient D,. Because D, equals uRT, one can appreciate that Di depends on temperature T (Noble, 1991). Therefore, the net ion flux is dependent on the voltage potential gradient (dv) when other conditions are constant. The ion flux values were expressed in terms of nmol m'2 s“. The calculation procedure for electrochemical potential gradient is illustrated in Figure 3. The electrometer’s AV is considered as the electrical voltage equivalent of the electrochemical potential difference over the distance A r. At 30 s the manipulator moved electrodes to position 1 (90 pm from leaf surface) within 10 s. The electrochemical potential measured by the electrometer was about 2.3 mV. At 40 s, the manipulator moved the electrodes to position 2 (40 pm from leaf surface). The electrochemical potential decreased to about 1.4 mV within 10 5. Position 3 had a higher mV (about 2.8) that when electrodes were positioned more distant from the leaf surface. The high-low potential was cycled each 20 3 period. Higher mV at position 1 more distance fiom leaf surface indicates a higher H+ concentration at this position. Hydrogen ion movement down its electrochemical potential gradient is toward the leaf cell, i.e. influx. An Opposite movement is an efflux, against the electrochemical gradient. During flux estimation, the first step in the calculation of ion flux was the determination of voltage potential at the extreme position of each manipulator movement. Since only ion activities at two positions perpendicular to the sample surface were measured, only net flux could be estimated. The net flux represents a balance between 42 influx“ scalar Ilux r calct talc: Perle estirr ll’lfi‘ cf Kim I) EXpe influx and efflux. The first 4 s (2 s for electrode movement and 2 s for electrochemical settling of the electrodes) of each half cycle (10 s) were not used for concentration and flux estimation. This was considered to be a delay time necessary to obtain a stable measurement following movement of the microelectrodes. Instability was thought to be due to solution imbalance and vibration noise following movement. The remaining 6 s at a position is defined as the valid data interval (VDI). Mean voltage potential during three consecutive VDIs, i.e. VDI,, VDI2 and VDI3 (Figure 3), at position 1, 2, and 3, respectively, were determined (2.31, 1.40 and 2.80 mV, respectively). The function for the line connecting the mean mV values for VDI, and VDI3 was established. The AV was the difference between the mean mV midtime (T2) value for VDI2 and also at T2 to this line firnction as shown in Figure 3. Ten 5 later, the next VDIs were repeated to create the next triplet, to be used to calculate AV again in similar manner. The electrochemical potential gradients (AV) were calculated by this process at 10 5 intervals which thus reflected running averages for 30 5 period. Finally, net flux was estimated from the AV, i.e. dv in equation (2) A positive estimated flux value indicates a net influx, while negative one is a net efflux. The above procedure makes reasonable considerations for the settling of voltage at the electrode tip after each movement. It also allows for drift of the signal, as seen Figure 3, which is to be expected when the tissue flux causes changes in the concentration of the ion being measured. Experimental Protocol 43 .Emgmthm wu:s_ 05. 0:6: uonEzmm one; moxaz or: :25; E0: omcmco omm=o> 2.: $962: 9. =anmoE m8 E9: Boom. Emu 5.: +1 ho EoEmom < .m 930E 89$ 9:: on om om 0v om ON — _ _ _ w _N».W r Amomtzm 9:60:81 ovvmm :oEmoa ‘ m_o> V r Amomtam 9: E0: E: o9 m Samoa A Aoomtaw e5 Eek E1 2: v c0368 9,5 329 cam—2 O 5225 San. u=m> H_n.._> _ o.o oé ON o.m o..v (Aw) uorrnros ureq am 10 96mm 44 Young expanding leaflets (leaf age about 2 weeks) were harvested from pea plants with a razor blade. For studies involving flux from mesophyll tissues, leaf sections of approxi- mately 15 x 10 mm (LxW) with limited veinal tissue were excised from the leaflet lamina. Using a fine forceps, a small strip of epidermal tissue was then detached from the abaxial surface of the lamina in such a manner so as to avoid major veins and to minimize mechanical damage. Isolated leaf segments were floated on buffer solution (0.5 mM CaSO4, 1.0 mM KCl, 2.5 mM HEPES-HCl, and 10 mM sucrose at pH 5.2) under 60 W m'2 illumination (Philips lamp). After 2.0-2.5 h, the peeled lamina segments were mounted within the sample chamber and bathed in 6-7 mL unbuffered solution (containing 0.5 mM CaSO,, 1.0 mM KCl, pH 5.2). The bathing solution was replaced 3 - 5 times. The sample tissue was then pretreated for 50-60 min in the dark. The dark treatment allowed for stabilization of ion flux across the plasma membrane as a result of equilibration of the tissue with the new environment. After dark pre-treatrnent, the leaf tissue was ready for measurement. Preliminary experiments showed that ion fluxes can vary significantly from one position to another over the surface of the lamina tissues. To minimize the time course variability of flux measurements, regions with stable flux were sought by moving manually the hydraulic manipulator stepper. After a suitable site was selected, ion fluxes were measured for a minimum 5-10 min before the treatment light was turned on. Data recording for flux estimation was continued for an average 30 min after the onset of illumination. The light source was then turned off, and data recording continued for an additional 10-1 5 min. The sample chamber was then removed and replaced with another 45 bairirt; dined until Iii. dunn, C011313 Signirt l 500‘: chamber with a flesh leaf sample. All experiments were conducted in unbuttered solution at room temperature (23i1°C). Arif et al. (1995) explained the reasons for not using buffers in the bath. This is because some of the HI crossing the tissue boundary diffuse as protoned buffer whose flux is not including in the flux calculated from the H+ electrochemical gradient (Arif et al., 1995). Heat emission from the light source was negligible (was discussed in Chapter III). The pH treatments were achieved by adding 0.1 mM HCl / NaOH to the unbuffered bathing solution. Photosynthetic inhibitor, DCMU (3-(3,4-dichlorophenyl)-1,1- dimethylurea, from SIGMA) was prepared in 0.5 % (v/v) DMSO and then dissolved in unbuffered solution adjusted to pH 5.2. DMSO of 0.5 % had no discernible effect on ion flux. To test effects of DCMU on ion flux, the control (an unbuffered solution) was replaced with DCMU solution (0.5 mM CaSO,, 1.0 mM KCl, 0.05 mM DCMU, pH 5.2) during the course on an experiment using a single leaf sample. Data statistical analysis By using ion-specific microelectrodes, net ion fluxes from leaf cells can be measured directly. One of the major concerns has been whether any artifacts were associated with ion flux measurement using the MIFE technique. No significant artifacts could be identified in this study. In most cases, changes in extracellular ion concentrations were consistent with changes in net ion flux. Secondly, light illumination did not cause significant changes in ion fluxes from leaf tissues when electrodes were lifted far (at least 1500-2000 pm) from leaf surface (Figure 4). Additionally, heat effects induced by light 46 pH 0|' (PM) K“ (MM) Ca2+ (11M) § 513 _ H Lighton . flux Light off 5.6 — -9— concentration I 5.4 — 5.2 — 5.0 — LL 640 - 032+ 620 >— F l 600 P— l— ’-1 , ":- A.) l ‘9' l‘.--.; rlrlrlrlrlrlrlrlrl 1400 l— CI- i l 1300 — l 1200 - 1100 1000 1400 1300 1200 1100 1000 900 l l l l 1 l l l l l l 0 2 4 6 8 10 12 14 16 18 Time (min) Nlrlrlrl 40 20 Net H+ flux (nmol m'2 5'1) 100 Net Ca2+ flux (nmol m'2 5'1) -100 -200 Net CI' flux (nmol rn'2 3'1) -300 N (A) 8 8 -2 s' ) S O -100 net K+ tlux (nmol m -200 ~300 Figure 4. Time course of net ion fluxes and extracellular ion concentrations of H+, Ca2+, CI' and K+ at 1500-2000 pm from pea leaf surface using the MIFE system. 47 treatment were considered to be negligible (see Chapter HI). Fluctuations in ion flux over time are thought to reflect normal ion transport processes and dynamics across the plasma membrane of leaf cells. Thus, the extracellular ion concentration changes measured by the MIF E system were considered to be results of real changes in net fluxes across the plasma membrane. Over 30 replicated leaf samples were conducted for the control. For analysis of ion flux response to light stimulation, there was concern about simply averaging replicate time-course measurements. The problem of averaging multiple replicates is that important time course changes in flux elicited by the light/dark treatments would be masked. For example, if two different replicate leaf samples exhibit a similar response in ion flux, but with a shift in time of several seconds or minutes, those shifts would result in a much difference response trend if the data were averaged. This, thus, would cause a loss of valuable information. For this reason, representative individual replicates were selected and shown in Figure 6 and 7 to best illustrate a typical response. In order to smooth the treatment data, and to reduce some of the noise for analysis of the responses, data (including concentration and flux) over 1 min interval were averaged. The data could be reproduced for replicates of leaf samples. One-way Analysis of variance (ANOVA) of the data was conducted using SAS 8 (SAS Institute Inc, Cary, NC). To determine if specific treatments (i.e. pH, DCMU) or comparisons among the four ions resulted in statistically significant differences in the light-responses for flux and extracellular ion concentration, LSD test (Fisher) at 0.05 was used. ANOVA results appear in the text and in the figures. Figure 5 illustrated how important features of the light-response for net H+ flux and 48 extracellular pH was calculated. The analysis for other ions (Ca2+, K+ and Cl') could also be made in a similar manner. The following terms were used in this study: F a or C0: Five min average net ion flux or extracellular ion concentration before light- on. Average change during the entire light period: the averaged net ion flux or extracellular ion concentration during the entire light period minus F0 or C0, respectively. Average change after 10 min light-off: the averaged net ion flux or extracellular ion concentration during the entire light period minus F, or C,, where Ff or Cf is 10 min average net fluxes or extracellular ion concentrations from light-on to light-off, respectively. Ion flux and extracellular ion concentration responses: Initial flux response (net influx): initial maximum net influx or minimum net efflux after light-on minus F0 (Figure 5). A positive value indicates a net influx. Initial concentration response (decrease): initial minimum extracellular ion concentration after light-on minus Co (Figure 5). However, pH response is an increase (positive), i.e. H+ concentration decrease. Maximum flux increase (net influx): maximum net influx or minimum net efflux during the entire light period minus F0. A positive value indicates increase in net influx. Maximum extracellular ion concentration decrease: minimum extracellular ion concentration during entire light period minus C0. Change during the first min: the average of net ion flux or extracellular ion concentration during the first min after light-on minus F0 or C0. Change during the first 49 ..co-Em: norm $622332 .mmcoamoc cosmbcoocoo 5 x3: .mEE m_ 5 .0 E NEE odor—on 20392850 :2 5.26355 .0 x3: co_ 6: .mEE m_ E .Euzn: 3:033. 55 figs-Boa 8635-59. .6 ou3__aEm m_ E .co-Em: 98mg 59:880. .mcozgcmocoo co_ cm_:__mombxo no mmxsc we: no mmmco>m 5E m m_ 00 .o o... doze-m: =>caomoE m8 959m Ia 222.0833 new mmxsc +16: 5 mmcmzo 3035-39. .0 co=m_:o_mo .m 059... 3:5 as: om mv ov mm on mm om 9 or m o . on- ~ g d _ — fl _ d a glfi a mi? — _ u _ _ — q _ d _ .— _ d d a — _ _ _ fi _ A _ 4 _ — _ _ _ a w _ _ _ O m a .. om-l ........... 00 .im m 000. r \O C. \u ,- . 9-W I mm + one. 2 {we so... 25. 8:88. 52. 3%: l u I r . hm .” . om . n . w m m x 09532 cozgceocooEEE I u r ) o .- . M - U ., w r - - - - - - . . l Vfi 0| 3:88. x3: REE lo M w e I . D 8 so: . 7v r ‘0 f0 . l m6 3. o -/ H Ir 0 a ( - ---- - -.....- .- . - 0. t 1 om l o o o o . . E . o o . O o . l . . . m m - A.._.£._.w..m_n._.._.n~c+-.......-.. .. . on I x3: IOI m Z J Em T In m to E 3 h. r ov 9m 50 Hd five min: the average value of net ion flux or extracellular ion concentration during the first 5 min after light-on minus F0 or C0. Amplitude of light-induced Post-Initial Flux Response (PIFR): the lowest influx or the highest efflux after Fi minus Fi (Figure 5), where F, is the net flux when the light- induced initial increase in net influx started to decline. PIF R is a negative value. Time response: Response lag time (second): the time between the light-on or light-off and change in net ion fluxes or extracellular ion concentrations. Initial light response time (min): the time to reach initial maximmn changes for ion fluxes or for extracellular ion concentrations after light-on (from dark to light) (Figure 5). T 50,, for initial change (min): the time to reach 50 % of initial response in net influx or in extracellular ion concentration after light-on. Time to maximum change (min): the time to reach maximum increase in net influx or maximum decrease in extracellular ion concentration during the entire light period. PIFR time: the time for whole PIFR process or the time from Fi to lowest influx or highest efflux after F i (Figure 5). RESULTS Ion flux change of mesophyll tissues in darkness In the dark, net fluxes (a balance between influx and efflux) of H", Ca2+, Cl' and K from pea mesophyll tissues in a bathing solution containing 0.5 mM CaSO4 plus 1.0 mM KCl, 51 5.8 5.6 5.0 540 520 C82+UM s g s § § 1400 1300 1200 Cl" uM 1100 1000 1600 1500 1400 K+uM 1300 1200 1100 -0- concentration -O— 1qu I l l l J l l l l 1 4i :- C32+ 4 :- rr/I "\’\\"\"lu’( ’./r r to r- -r r If r , ”a , (l \ ':*”‘\ “\‘I‘I’ ”(TACO q \l (M 1‘ \ \‘ \ V" “A ' ‘ f .‘ "r\\f,’ '(‘I/ Q ‘\ (N \s' ‘1, . I \\ M sf ‘ _‘ - Ul'" 2‘ I "q l I .. .i ‘ ...v : _. l I ‘ q i- ' l ‘ ' __1 -. ‘ j " l l J l l l L l J 1" -— C" —. l' ..— h— . ( ‘ | " V I ‘ ‘ I. i - |- . . 1 I l. O l _ ‘ ~ . l . I» l I , . l I ll , I ll ll '1 ‘ Ii ' ‘ P I ‘ ‘ tl ‘ ‘1 — \\ ‘ ,r If m . .. \5‘ r I s \.\ , _, L l l l J l l l l l l + ..- -_. K .- '- cl b ‘ I y - .. Ill . , . , l' ‘ ' H ‘1 ‘ ‘ ,l v I ' ., — r l I I ‘ I \. ‘- % ‘ d P r x,. x ‘ ; ' :_ ~ -l t " t -. 1 l l r l l r r r r- 0 5 10 15 20 25 30 35 40 45 50 Time (min) Figure 6. Time course of net fluxes and extracellular concentrations of H+, C32+, Cl' and K+ around pea mesophyll tissues of one leaf sample under dark. 52 40 O N O '8 Net H” flux (nmol m‘2 s") g g g 5; ° 3 s a e g Net Ca2+ tlux (nmol m'2 3'1) ° § § Net Cl' flux (nmol m'2 371) °§§§§ lb .3 8 8 Net K+ flux (nmol m'2 s'1) ‘é OW red be in( ex pet m at pH 5.2, exhibited small fluctuation over time (Figure 6). The extracellular concentrations of these ions were stable and fluctuated only slightly. Estimation of mean rates for the entire dark period for PF, K” and C1“ fluxes were -4, -51, -89 nmol m2 s", respectively. Thus, a net efflux of these ions occurred under dark conditions. In. contrast, a net Ca2+ influx (+9 nmol 111‘2 s") was observed during the dark period, indicating Ca2+ uptake by the mesophyll. Ion flux response of mesophyll tissues in light When pea mesophyll tissues were stimulated by exposure to light in the same bathing solution (0.5 mM CaSO4, 1.0 mM KCl, at pH 5.2), H”, Ca2+, K“ and cr exhibited different patterns of change in net fluxes and extracellular ion concentration over 30 min (Figure 7). In spite of some variability in the light-induced ion influx responses, especially of K+ and Cl', most replicate leaf samples showed similar light responses for each ion. In general, light induced a rapid increase in net flux (positive) and a decrease in extracellular ion concentration (negative). After a rapid large initial flux response, a reduced or slower net flux change was observed over time. The process, called the Post- Initial Flux Response (PIFR) in this study, typically involved a gradual return to flux rate closer to or lower than that measured prior to light stimulation. Since only net flux could be measured using the MIFE system, the PIFR could theoretically involve either independent changes or concurrent changes in both influx and efflux. Both Ca2+ and Cl‘ exhibited a rapid PIFR, whereas H‘” and K+ maintained a net influx for an extended period, which was then followed by a slow PIFR. Thus, the PIFR is a negative value. When the light was turned off, the net flux rate for each ion declined leading ultimately to 53 2+ (HM) l..A‘\ 40 A 3 N- E BEE x3: +162 20 0 A?» ~-E .055 x3.— +Nmo «oz mwmwmommm m m o :m N .5 .25; x2. .6 “oz mmm 40.8 r-» N m -E ass 3. +52 mmommm —ufi--_dd-——ddQHfld l _ . . q — J — a — q l l 45 40 35 15 10 Time (min) Figure 7. Light-induced time course in net fluxes and extracellular concentrations of H+, Ca2+, Cl' and K+ around pea mesophyll tissues of one leaf sample. 54 a net efflux. The flux change was small over time as compared to the response to light. The corresponding apoplastic ion concentrations increased, where pH decreased (Figure 7). The quantitative features for each ionic response are summarized in Table 1 and Table 2. A simple averaging of different periods (the entire period from dark to light or from light to dark) for each ion resulted in a smaller magnitude of change in either ion flux or extracellular ionic concentration (Table la, 1b). For H+, the average flux change over the entire light period was +11 nmol m'2 5", whereas the light-induced initial net flux change was +26 nmol m'2 s'l and maximum increase in net influx was +29 nmol m'2 s" (Table la). Thus, the average change over the entire light period was not used as a parameter to evaluate light-induced ionic responses. The light-induced initial responses in net flux and ion concentration were similar to the light-induced maximum changes for each ion (Table 1a and 1b). By comparison, Ca” flux exhibited the most rapid response to light. A transient increase in net Ca2+ influx of 48 nmol m”2 s'l occurred after light exposure (Figure 7 and Table la). This increase is considered a result of the increase in Ca2+ influx and not a decrease in Ca” efflux, because the initial Ca2+ flux before light exposure (F0) is a net influx. The net fluxes of H", K” and Cl' changed initially from a net efflux to a net influx with a magnitude of 26, 174, and 119 nmol m'2 s“, respectively, following exposure to light (Table 1a). Extracellular concentration for these ions surrounding mesophyll tissues also decreased, consistent with the development of their net influxes. The extracellular pH around mesophyll tissues increased by 0.4 - 0.5 during the 30 min (Table 1b). However, the average change in net Cl' flux during the first 5 min afier illumination was 55 12bit I film in 0.5 0] 20,13 'ifi '33.,“ I PIFR Table 1. Quantitative characteristics of the light-induced net fluxes and extracellular concentrations of H", Ca2+, K‘“ and CI' in pea mesophyll tissues bathed in 0.5 mM Case, and 1.0 mM KCl at pH 5.2. a: ion fluxes 1r Caz: K“ cr (nmol m:2 s“) F 0“ ~3 +8 ~5 1 ~61 Average change during +Ill +14 +84 +31 the entire light period Average change after -8 ~21 ~43 ~46 10 min after light-off Initial flux response (net influx) +26 +48 +174 +119 Change during the first min +6 +25 +20 +31 Change during the first five min +14 +31 +55 ~56 Maximum increase (net influx) +29 +50 +208 +190 Amplitude of PIFRt ~20 ~52 ~176 -l 71 b: extracellular ion concentrations H+ Ca” K“ Cl' pH uM uM uM C0§§ 5.02 562 1172 1126 Average change of +’0.36 ~37 ~45 ~30 the entire light period Average change after ~0.21 +31 +208 +213 ' 10 min after light-off Initial concentration response +0.43 ~49 ~109 ~108 (decrease) Change during the first min +0.04 -9 +11 ~15 Change during the first five min +0.24 ~27 ~50 +38 Maximum extracellular ion +0.47 -59 -128 ~132 concentration decrease All data are means with 30 replicates. 5 F0: five min average of net ion flux before light-on, where + or - indicate net ion influx or net ion efflux, respectively. ? - or + indicates decrease or increase of net ion flux and extracellular ion concentration, respectively. ’ PIFR: Light-induced Post-Initial Flux Response, a result of either a decrease in net influx or an increase in net efflux, or both. This value is negative. §§ Co: five min average of extracellular ion concentration before light-on. 56 near: | most time YES: Cont. lnstar also \ is St: negative (~56 nmol m'2 s", Table 1a), and extracellular Cl' concentration increased during the same period (Table 1b). This result indicates that a net Cl' efflux response occurred. Some ions responded more quickly to light than others. Compared to other ions (H2 1C and Cl'), Ca2+ flux was observed to have the smallest Tm, (1.5 min), the time for 50% change in initial net ion flux. The time to the maximum increase (influx) was also the most rapid for Ca” (3.5 min) (Table 2a). Chloride flux had a similar initial light response time (2.3 min) compared to Ca2+ (2.2 min), and both were significantly shorter than H” and K“. In considering light-response lag time, Ca2+ flux response was the most rapid within 13 s, then followed H+ (29 s) and by C1' (45 s) and K+ (48 3). Similar results were also observed in extracellular ion concentration (Table 2b). The PIF R time for Ca2+ (3.4 min) and Cl' (4.1 min) were significantly shorter than for H+ (8.4 min) and K+ (10.3 min) (Table 2a). When light was turned off, the response lag time for Ca” flux from light to dark was also significantly shorter (19 s) than other ions. The qualitative characteristics of the measured light-induced ionic responses are presented in Table 3. Changes in net flux and apoplastic ion concentration occurred concurrently following light stimulation when compared Table 1a with Table 1b. For instant, light exposure quickly stimulated Ca2+ uptake by the mesophyll within 13 s, and also caused a reduction in apoplastic Ca2+ concentration within 13 s. Thus, the evidence is strong that the Ca2+ flux response was not an artifact of some other factors. Light-induced ion fluxes from epidermal tissues 57 Table 2. Quantitative characteristics of the light-induced time responses of H“, Ca”, K+ and CI' in pea mesophyll tissues bathed in 0.5 mM CaSO4 and 1.0 mM KCI at pH 5.2. a: ion fluxes IF Caz; K+ Cl' Response lag time (s) 29 b 13 c 48 a 45 a (from dark to light) Response lag time (s) 26 b 19 c 50 a 45 a (from light to dark) Initial light response time 4.6 a 2.2 b 4.3 a 2.3 b (min) T50.) * 2.0 b 1.5 c 2.7 a 1.9 c (min) Time to maximum change 5.9 b 3.5 c 10.1 a 9.3 a (min) PIFR‘ time (min) 8.4 a 3.4 b 10.3 a 4.1 b b: extracellular ion concentrations H+ Ca2+ K+ Cl' Response lag time (s) 32 b 13 c 51 a 45 a (fi'om dark to light) Response lag time (s) 28 b 18 c 45 a 43 a (from light to dark) Initial light response time 5.3 a 3.3 b 5.9 a 3.4 b (min) Tm’ 2.6 a 1.6 b 2.7 a 2.6 a (min) Time to maximum change 15.1 a 8.7 b 7.0 b 8.1 b (min) All data are means with 30 replicates. Different letters indicate significant difference across rows according to LSD test at p < 0.05. ’ Tm: time to reach 50 % initial response of net ion flux or extracellular ion concentration, respectively. t PIFR: Light-induced Post—Initial Flux Response, a result of either a decrease in net influx or an increase in net efflux, or both. PIFR time: the time for whole PIFR process. 58 a in: ‘1 -.\.4 etc! Q . 1a.! envIS.V 2:. VIC :- 59822- Fun-v.7... __>.£QCV..:C 3.:— 50 nee-.aaa—tbh O‘CC- ~009=§=~|sn~u= $0 nuclshbsbflhfiflb O>-~====° .nr 4:2th 88:05 :50: 88:05 :60: 88:05 58: 88:00: m: 58: 20208200200 035:? 035:? 050:0. 030:0 88:05 5&0 :0: :0:: 88:05 5&0 :0: :0:: 88:05 5&0 :0: :0:: 88:05 5&0 :0: :0:: 88:00: 5&5 :0: 58:: 88:00: 5&5 :0: 58: 88:00: 5&5 :0: :50: 88:00: 5&5 :0: :50: 5% 5.0.2:»: :32 8:080: 30: 8:080: 308 8:080: 58: 8:080: :88 :5: 30:0 88:05 88:05 88:05 88:00: I: 20205200200 8:080: 58: 8:080: 308 8:080: :50: 8:088 30:0 88:05 5&0 :0: 0: 509 :0 88:05 5&0 :0: 50: :0 88:05 550 :0: £0: :0 88:05 0850 88:00: 5&5 :0: 50:: :0 88:00: 5&5 :0: :0 88:00: 5&5 :0: :0 88:00: 5&5 :0: §~\ 8:080: 555.80: 80.55-551— :5:8:: 805.: :0::05: w:0_ 5:05: 808 :0::5: :5 88:00: 58: 88:00: 30:: 88:00: 50:05:: 88:05 E: :30: 2028500200 5:05: :08 8:83: w:0: 55:5: :08 8:05: w:0: 5&5 :0: 0: x:&.:0 5&5 :0: 0: 5&0 5&5 :0: 0: 5&0 :0: 50:“: 8550 58: :0: 58.: 8550 58: 88:05 5&5 :0: 508:8: :0: 58.: 8550 :58 SQ 08888.3»: .55:— :0 .0: mac 5 .8 m: a 60: Es 0.: Es vOmaU 2:: m6 5 :0::0: 0.085: 593805 00: .:0 88:080.: 2:0: 80:65.3»: :0 8035:8355 258:8.0 .m 03:.H 59 An isolated strip of epidermal tissue (about 5 x 15 mm) was mounted in the same measuring chamber and bathed in the same solution as the mesophyll tissue. The experimental protocol was similar to that described previously. Unlike mesophyll tissues, light did not elicit a large flux change for any ion (Figure 8). No significant difference in PIFR were observed during about 25-30 min measurement. The data of Table 4 presented a possible characterization of the light-induced ion fluxes and extracellular ion concentrations around the epidermal tissues. After light stimulation, net Ca2+ influx increased slightly, net H“, K+ and Cl' effluxes decreased slowly (11+) or changed toward a net influx (K+ and Cl') during the first min or during the first 5 min after illumination (Table 4). Although net flux of each ion was estimated to have positively increased, the magnitude of increase in epidermal tissues (Table 4a) was much smaller than that of mesophyll tissues (Table la). In addition, light only induced a small initial increase in extracellular pH (Figure 8, Table 4b). External pH effects on light-responses of mesophyll tissues The mesophyll tissue was pretreated under dark for 2.0 ~ 2.5 h in a buffered solution with either pH 4.0 or 6.0. The net ion fluxes from mesophyll tissues were measured under light illumination in the unbuffered solution (containing 0.5 mM CaSO4 and 1.0 mM KCI) at the same pH as each buffered solution. Light period (30 min) was the same as at the control (pH 5 .2). Compared to the control, a higher pH (6.0) of the bathing solution resulted in a reduced light-induced initial response in net influx of both H+ and K+ by about 45 %, in particular, in a significant decrease in net flux change over 5 min 60 5.8 11* . ' -0- concentration Light off d 5.6 __ Light on . flux 5.4 _ l l , I Q b O o ' _ 5.2 .. .. ‘ (“wmummmw(11.111111111111111 (liillllll'll‘lllélllltllil‘t‘t‘ullllliiltill““UH“(“(H‘H‘U‘HH‘“H“. .\(l|lli _ 5.0 _ llllllluluullltulk * + l l J J l l l l *- .— 520 LCa2+ . i 1 fi 500 — : 3 480 — l .- 3 1' ‘ J . _ + ... - N 460 l . ' l ‘ . 4. ' 8 440 .. - I ' i . 1 1 420 F (\er 11(1 (L/“I “It” i p 400 ’_ l 4 l 1 I 1 J L _ 1400 Lcr _ - l . 1300 1— ‘ , M.— l- 1 - A 1 1 5 J 2 1200 l— 1 __ 3; L ' 4 '5 1100 _ _ L- . 111/ .fl 11111111011 i'\.\‘(\\“\ .1 "’HIHuan 1000 — ‘1 900 - -1, J i l l l 1 1 1 + .. 1400 .— K _ - l 1 Q .... 1300 — ‘ :‘ll‘ ..3 . 1» E P 1 l. 74 51"] ‘ 551200.. ' ma '...' ,. l by. 1. , _ + I" l’t ;" '0“ 7‘," .“ l'nu' 1(th r -“ 1. (.r‘mt s _ ‘5‘ A x b ,"1 ¢ V. l ('(1 H. u“ 'k‘ . “(”1 u 4 11m _ (‘.11\ I] . .— 1. —-l 1000 - d J— 1 - 1 1 1 1 1 1 J O 5 10 15 20 25 30 35 4 Time (min) 40 O N 0 1'0 0 Net H+ flux (nmol m'2 5'1) 63 g g o a s a s g Net Ca2+ llux (nmol m'2 s") I L o d N g 8 8 8 Net 01' flux (nmol m'2 s") §°§§§§ Net K+ flux (nmol rn'2 5'1) ‘é'é’ Figure 8. Light-induced time course of net fluxes and extracellular concentrations of H+, C32+, Cl' and K+around pea epidermal tissues of one leaf sample. 61 to {'11 “EYE c0116. mo S" l C01 111; illumination (Table 5). The alkaline medium also declined the light-induced PIFR of both ions (50 % for H and 36 % for K+). However, the increase in external pH did not alter the light-induced initial response in extracellular pH (Table 6), as well as in net influxes of Ca2+ and Cl' (Table 5). Acidification of the bathing solution (pH 4.0) produced a different response compared to the pH 5.2 treatment. Net H+ flux was always positive (net influx) while mesophyll were being illuminated. It started with a large influx (156 nmol m'2 5“) under dark conditions (Table 5). Light treatment initially induced a large increase in net influx (+44 nmol m‘2 s") which subsequently declined during the remainder of the light period (Table 5). Thus, the light-induced PIFR of H“ was a decrease in net influx (-108 nmol m'2 s") which was significantly larger than at both pH 5.2 and 6.0 (Table 5). Under pH 4.0, the light-induced initial increase in extracellular pH was much slower and reached to 0.35 within 19.5 min (Table 6). In contrast, the lower external pH resulted in a larger light- induced initial response in net influxes of K+ and Cl' as compared to pH 5.2 (Table 5). During 5 min after illumination, the net flux of Cl' still increased positively (54 nmol m'2 s") at pH 4.0. However, that of the pH 5.2 decreased negatively (-74 nmol m'2 5") compared with at pH 4.0 (Table 5). The similar difference was observed in change in extracellular Cl' concentration (Table 6). The light-induced PIFR of K+ was not influenced by lower pH (Table 5). However, the PIFR of Cl' was reduced by a large magnitudefrom 148 to 29 nmol rn‘2 s'l (Table 5). The pH 4.0 treatment did not affect the magnitude of light-induced Ca2+ initial response. However, the PIFR of Ca2+ and its initial light response time in either flux or 62 Table 4. Quantitative characteristics of the light-induced induced net ion fluxes and extracellular ion concentrations in pea epidermal tissues bathed in 0.5 mM CaSO4 (concentration decrease) and 1.0 mM KCI at pH 5.2. a: ion fluxes H+ Ca2+ Ki CI’ [11110] m-z s.1 F0§ -10 (9)‘ 2 (8) -9 (9) -1 (10) Initial flux response +l6 (9)+l 8 (8) +81 (8) (net influx) Change during the first min -5 (9) +5 (8) +23 (8) +36 (8) Change during the first five min -3 (9) +3 (8) +3 (8) +2 (8) Maximum increase (net influx) +9 (9) +23 (8) +121(8) +104 (8) b: extracellular ion concentrations H+ Ca2+ Ki" Cl' pH 11M 11M 11M C0§§ 5.04 (9) 529(8) 1126 (9) 1093 (8) Initial concentration response +“0.08 (9) ~28 (8) -75 (7) -36 (8) (decrease) Change during the first min +0.01 (9) O (8) +45 (7) +4 (8) Change during the first +0.03 (9) -7 (8) -ll (7) -1 l (8) five min Maximum extracellular ion +0.10 (9) -38 (8) -75 (7) -47 (8) 5 F0: five min average of net ion flux before light-on, where + or - indicate net ion influx or net ion efflux, respectively. t 3 The numbers in parentheses indicate the number of replicates. §§ Co: five min average of extracellular ion concentration before light-on. 63 - or + indicates decrease or increase of net ion flux and extracellular ion concentration, respectively. Table 5. Effect of external pH in bathing solution on the light induced net fluxes flux of H‘, Ca“, K+ and Cl‘ from pea mesophyll tissues. pH Hi Caz+ Ki Cl' inhathingiolutiom nmol m'2 s’1 F0§ 4.0 +156(8)§§ a +21(8) ~102 (8) a ~232 (7) a 5.2 ~0.4 (9) b 5 (12) ~81 (8) ab ~72 (8) ab 6.0 ~7 (8) b 0 (9) ns 24 (7) b ~26 (5) b Initial flux response (net influx) 4.0 +‘44 (8) a +50 (8) +272 (8) a +174 (7) a 5.2 +22 (9) ab +46 (12) +167 (8) b +101 (8) b 6.0 +12 (8) b +41 (9) ns +92 (5) b +131 (5) ab Change during the first min 4.0 ~15 (8) a +28 (8) +69 (8) +115 (7) a 5.2 +2 (9) b +25 (12) +39 (8) +17 (8) b 6.0 +1 (8) b +18 (9) us +28 (5) ns +31 (4) ab Change during the first five min 4.0 ~3 (8) a +15 (8) +89 (8) a +54 (7) a 5.2 +11(9) b +29 ( 12) +62 (8) a ~74 (8) b 6.0 +2 (8) a +12 (9) ns +1 (5) b ~10 (5) ab Amplitude of PIFR‘ 4.0 ~108 (8) a ~74 (8) a ~208 (8) a ~29 (7) b 5.2 ~22 (9) b ~59(12) ab ~203(8) a ~148 (8) a 6.0 ~11 (8)b ~42 (9) b ~131(5)b ~195 (5)a Different letters indicate significant difference between external pHs within columns according to LSD test at p < 0.05. “ns” indicates no significant difference. § F : five min avera e of net ion flux before 1i t-on, where + or — indicate net ion influx or net ion efflux, 0 respectively. §§ The numbers in parentheses indicate the number of replicates. * ~ or + indicates decrease or increase of net ion flux and extracellular ion concentration, respectively. t PIFR: Light-induced Post-Initial Flux Response, a result of either a decrease in net influx or an increase in net efflux, or both. This value is negative. Tab.- con 4“ .. - mu. C. 11111; In. hill .0 A 31: D (ii p .. 1. m C t a _ .u .1. r1; Table 6. Effect of external pH in bathing solution on the light-induced extracellular concentrations of H“, Ca“, K“ and Cl’ and time responses in pea mesophyll tissues. pH Hi Ca2+ K+ Cl‘ Mammalian (pH) (AM) (.uM) CHM) 4.0 4.46(8) “c 608 (8) 1159 (8) ab 1168 (8) 5.2 5.00 (9) b 558 (12) 1200 (8) a 1065 (8) 6.0 5.40 (8) a 541 (9)ns 1015 (7) b 1175 (5) ns Initial concentration response (decrease) 4.0 +1035 (8) ~46 (8) ~192 (8) ~192(6) a 5.2 +0.42 (9) ~46(10) ~128(8) ~11 1(8) ab 6.0 +0.30 (8) ns ~44 (9) ns ~92 (5) ns ~66 (5) b Change on the first five min 4.0 +0.07 (8) b ~24 (8) ~82 (8) ~135 (8) a 5.2 +0.21 (9) a ~37 (12) ~63 (8) ~29 (8) b 6.0 +0.16 (8) ab ~18 (9) ns ~30 (5) ns ~21 (5) b Initial light response time (min) 4.0 19.5 (8) a 3.1 (8) b 8.3 (8) a 4.4 (8) a 5.2 5.3 (9) b 2.6(12) b 4.6 (8) b 2.3 (8) b 6.0 6.3 (8) b 5.4 (9) a 4.0 (5) b 3.4 (5) ab Different letters indicate significant difference between external pHs within columns according to LSD test at p < 0.05. “ns” indicates no significant difference. i Co: five min average of extracellular ion concentration before light-on. §§ The numbers in parentheses indicate the number of replicates. ' ~ or + indicates decrease or increase of net ion flux and extracellular ion concentration, respectively. 65 6X17 Tn: COT“. E111 Phc “35 Sum 111.172 net 2‘. ions DIE. extracellular concentration were significantly larger at pH 4.0 than pH 6.0 (Table 5, 6). The lower pH medium prolonged the initial light response time of extracellular K+ and Cl' concentration compared with pH 5.2 (Table 6). Effect of photosynthetic inhibitor on light-induced ion fluxes Photosynthetic inhibitor, DCMU can block electron transfer. This inhibitor at 0.05 mM was added to the bathing solution (0.5 mM CaSO4 and 1.0 mM KCI) and the pH adjusted to 5 .2 with HCl. The addition of DCMU significantly inhibited the light-induced initial response in net influxes of H+, Ca2+ and 1C around mesophyll tissues (Figure 9), and elonged the light response lag time of H“ and Ca2+ fluxes (Table 7). However, the light- induced initial response in net Cl‘ influx were not significantly influenced by DCMU application (Figure 9). Sum of light-induced net ion fluxes When the magnitude and the directionality of net charge flux, based upon the combined net fluxes of W, Ca2+, K+ and Cl' (Figure 7), the light-induced net charge flux of the four ions increased transiently, lasting for about 5 min, and then declined (Figure 10 and 11). The light-induced initial increase in net influx (about 300 nmol m'2 3") indicates that a positive net charge move to the cytoplasm and would be expected to contribute to a membrane depolarization. Compared with mesophyll tissues, however, net charge flux rates in epidermal tissues did not exhibit any response to light (Figure 11). 66 Table 7. Effect of DCMU on the light-induced the time responses of H“, Ca“, K+ and CI' in pea mesophyll tissue. Hi Ci” K+ Cl‘ (8) Light response lag time ~DCMU (Control) 28 (5)"b 13 (6) b 46(6) 42(5) + 50 11M DCMU 57(6) a 32 (6) a 47 (5) ns 46 (5) ns Different letters indicate significance between the control and DCMU treatment according to LSD test at p< 0.05. “ns” indicates no significance. . Response lag time is the time between the light-on or light-off and change in net ion fluxes. .. The numbers in parentheses indicate the number of replicates. ‘5; 160 ~ 1:1 control - : +0.05 mM DCMU E '7" (I) 30.1 120 F - E 9 5 1 s 2 g ‘x’ w l— l —1 g 3 3 5 LSD(0.05) E 8 z “ 4° " I j .C .9 1 , _, . 0 WA , // H“ (2.2+ K” of Figure 9. Effect of a photosynthetic inhibitor, DCMU, on light-induced initial response in net ion influx from pea mesophyll tissues. Data represents an average of 6-9 replicates. Vertical bars represent LSD test result at p<0.05. 67 400 + Mesophyll Light off A 300 1— — '7.» Lighten N IS 200 _ — ‘6 S v 100 — _ m a) x 5.3. c O — _ .9 ‘6 Z '100 F- fi .200 l l l l l J J l l O 5 10 15 20 25 3O 35 4O 45 50 Time (min) Figure 10. Light-induced time course of net ion flux from pea mesophyll tissues. Data points represent a leaf sample. Net charge flux is a sum of H” + Ca2+ + K” - cr fluxes 250 + Mescphyll 200 _ '6- Epidemis _ ~70, Lighton Light off 0" 150 -- —l E E 100 — - r: m a 50 - ~ :1 a .5 o _ _ 8 Z -50 _ ...1 -100 1 1 1 1 1 1 1 1 1 1 0 5 10 15 20 25 30 35 40 45 50 Time (min) Figure 11. Light-induced time course of net ion flux from pea mesophyll and epidermal tissues. Data points represent an average of 7~12 leaf samples. Net charge flux is a sum of H+ + Ca2+ + K+ - Cl' fluxes. 68 Exp: (1pr- al.. 1r 11 ' hon 61C. 1111 mt”? dtp st :7 Sir: Pl;- D11, DISCUSSION Which ion is the initial depolarizing agent? Exposure of leaf mesophyll cells to light has been shown to result in a transient depolarization of the membrane potential, followed by a slow repolarization (Elzenga et al., 1995). Similar observations have been reported in other species including in Arabidopsis (Lewis et al., 1997), Phaseslus vulgaris (Nishizaki, 1992), Asplenium trichomanes (Szarek and Trebacz, 1999), etc. The membrane depolarization responses, however, have varied in their magnitude of membrane potential and the time responses, etc. In general, changes in membrane potential are considered to be the result of ion fluxes across the plasma membrane. Thus, the ionic component of light-induced membrane depolarization seems to be a foundation for understanding electrical signalling. Many authors suggested that Ca2+, H+, K+ and Cl' play a major role in stimuli- induced electrical signalling. Cation influx or anion efflux from leaf cells results in membrane depolarization. However, the identify of the ions involved in transient plasma membrane depolarization is still controversial in the literature. Elzenga et al. (1995) demonstrated that the light-induced plasma membrane depolarization (about 25 mV) in pea mesophyll tissues occurred at about 1-2 min after stimulation, which was followed by a slow repolarization. Assuming that four ions (Ca2+, Hi, Ki and Cl‘) are primarily involved in light-induced membrane potential changes, the summed net charge flux rates (Figure 10) measured in the present study exhibited a similar time course trend as the light-induced membrane potential response reported previously (Elzenga et al., 1995; Stahlberg and Van Volkenburgh, 1999). Therefore, we propose that the change in the net charge flux involving Ca2+, H“, K“ and Cl' following 69 13‘». ct. 0C TC light exposure could potentially be responsible for the reported light-induced membrane potential response. A light-induced transient increase in the net charge flux could be expected to depolarize the plasma membrane. Thus, the findings in this study provide indirect experimental evidence for the involvement of Ca2+, Hi, K+ and Cl' fluxes in light- induced electrical signalling. Shabala and Newman (1999) suggested that Ca2+ is a depolarizing ion based upon direct measurement of light-induced ion fluxes in bean mesophyll tissues. Their conclusion was based on time-course data following illumination, i.e., Ca2+ flux exhibited a more rapid (5 3) change than H”, K+ and Cl’ fluxes. The result in pea mesophyll (Figure 7 and Table 1, 2) agrees with their results in bean leaf tissues. However, they did not evaluate the transition from light to dark. Our results indicate that the light-induced flux changes of Ca2+, Hi“, K+ and Cl' declined and returned gradually to the rate which existed before light treatment when the light was turned off (Figure 7). Considering the time of the response to light, net Ca2+ influx increased rapidly to the maximum magnitude within 3.5 min afier light exposure. The maximum increase in net H+ and K+ influx, however, occurred after a 5.9-min and a 10.1-min delay, respectively (Table 2a). The initial light response time, T50.) and light response lag time of Ca2+ were also relatively short (Table 2a). Among the four different ions measured, Ca2+ should be considered as a main candidate for membrane depolarization in pea mesophyll. After about 1 min of illumination, net Ca2+ influx increased by approximately 25 nmol m'2 5'1 near mesophyll tissues. This result should be explained as an increase in Ca2+ influx, and not a decrease in Ca2+ efflux. According to the calculation of Nobel (1991), charge transfer (Q) = capacitance of 70 membrane (C) x voltage generated (AE), i.e. Q = CAE. The C of most biological membranes are approximately are luF cm'2 (Liittge and Pitman 1976). If membrane potential change (AB) is 25 mV, the Q required is 25 x 10'9 Coulombs cm‘z. One mole of monovalent ions has charge of 96500 Coulombs, or a mole of divalent ions has approximately 2 x 105 Coulombs. Therefore, a net transfer (Q) of 0.125 pmol of Ca2+ can realize 25 mV membrane change. In this study, an increase in_ Ca2+ influx of 25 nmol m'2 s" or 2.5 pmol cm”2 s'1 in pea mesophyll was observed within 1 min after light exposure. Actual measured membrane depolarization resulting from light exposure has been reported to rang between 25 and 35 mV (Elzenga et al., 1995; Stahlberg and Van Volkenburgh, 1999.). In addition, the initial response time for transient Ca2+ influx (about 2 min) measured in this study, thus, is also consistent with light-induced depolarization (2-3 min) reported by Elzenga et al. (1995). Therefore, Ca2+ can be considered to be an initial depolarizing ion. However, one should not exclude other ions from possible involvement in the light- induced membrane depolarization. A light-induced Arabidopsis plasma membrane conductance change for C1' has been suggested by Spalding et al. (1992). In pea mesophyll, Elzenga et al. (1995) found that a transient depolarization depended upon external Cl' concentrations and was unaffected by changing external Ca” or K concentration. On the other hand, light-induced membrane depolarization of pea epidermal cells was enhanced by increasing Ca2+ concentration in the bathing solution. Thus, these investigators concluded that the ionic mechanism for light-induced depolarization differs between epidermis and mesophyll. The membrane potential 71 op. P1: 1, pi 1L 168 10: changes in the epidermis were explained as being due to changes in Ca2+ flux and activity of the plasma membrane H+~ATPase (Elzenga at al., 1995). However, their conclusion was based on experiments in which the ionic concentrations of Ca2+ and Cl‘ were modified. Also ion flux was not measured directly. Because of numerous interactive and feedback regulatory mechanisms for ion transporters in the plasma membrane, the measurement of only membrane potential is not sufficient to reveal involvement of the specific ion fluxes. Only a few groups have directly measured light-induced ion flux changes in leaf tissues. Johannes et al. (1997) observed that Ca2+, K+ and anion-permeable channels were open at the peak of light-induced membrane depolarization in red-light-induced moss Phycomitrella patens. While Ca2+ influx and anion efflux coincided with the depolarizing phase (2-15 5), K+ influx occurred only during the first 30 s after light stimulation. Their results indicated that a transient K” efflux appeared later and induced the plasma membrane repolarization. However, their conclusions on ion flux dynamics are questionable because they measured ion fluxes using 30 s or 1 min intervals which is longer than the time necessary to attain the peak of membrane depolarization (2-15 5). Shabala and Newman ( 1999) rejected Cl' efflux and 1C influx as depolarizing agents afier measuring ion fluxes using the MIFE system. They found both large Cl' and K influxes after a 2 min delay between the onset of light stimulation and the onset of flux changes in bean mesophyll. In this study, changes in net Cl' and K+ influx did not occur until approximately 50 s (response lag time) afier illumination (Table 2a). The magnitudes of increase in net influx from pea mesophyll were smaller than that from bean mesophyll (about 1000 nmol m'2 3"). After 2.3 min, the net Cl’ influx declined to a net 72 6111 be. aml 11111 exp 1m de' )11 C01 me Ira m . ill= efflux below the F0 (Table 2a). Thus, this process, i.e. the PIFR, could be considered to be a result of an increase in C1' efflux not a decrease in C1' influx. This is because the amplitude of PIFR (~171nmol m'2 s") was greater than the magnitude of light-induced initial increase in net Cl' influx (119 nmol m'2 3") (Table 1a). The anion efflux could be expected to be involved in membrane depolarization. Elzenga and Van Volkenburgh (1997) tested anion channels by using patch-clamp techniques and found high anion channel activity in pea mesophyll cells after a 30 3 light treatment. They concluded that Cl' efflux is involved in light-induced membrane depolarization. In this study, however, light induced a rapid initial response in net Cl‘ influx. This net Cl' influx could be involved in the repolarization process and might counteract the light-induced depolarization. However, summation of net charge fluxes (including Ca2+, H", K+ and Cl' ) in Figure 10 and Figure 11 suggests that light still caused a net influx increase that contributes to a transient depolarization. Thus, Cl' efflux is involved in light-induced membrane depolarization but not in the initial transient depolarization event. The K“ exerts a dominating influence on the diffusion potential component of transmembrane electrical potential differences by virtue of its relatively high permeability in biological membranes. Even under dark conditions, K+ maintained a variable net flux rate around £100 nmol m2 s'1 which is larger than H and Ca2+ (Figure 6). Among Ca2+, K+ and Cl‘, which the electro-chemical potential for K+ is the lowest at the unperturbed membrane voltage of approximately ~150 mV. This is because the equilibrium voltage for K+ (Ex) is probably near ~100 mV (Leigh and Wyn Jones, 1984). Spalding and Cosgrove (1992) put forward that K+ is not a major depolarizing ion in blue-light-induced cucumber hypocotyls because TEA, an inhibitor of K+ channels, did not affect the onset 73 of 1 P351 mer efll C11 ICE [it I in: 0.5 of depolarization. In most cases, K+ is considered as a charge balancing ion, moving passively to compensate for light—induced ion movement of C1' or Hi. However, membrane depolarization of Desmodium motorium appeared to be associated with K+ efflux (Antkowiak et al., 1991). This conclusion was not based on direct flux measurement. In the present study and in bean mesophyll (Shabala and Newman, 1999), direct K+ measurement using the MIFE system indicated that K+ does not fimction in initial light-induced electrical signalling. The data also suggested that H+ is not an initial depolarizing ion. Net H” influx and apoplast pH exhibited only relatively small increases within 1 min after light stimulation (Table la). The time-course response of H+ was slower than that of Ca2+ (Table 2). The delay between light application and the beginning of the H”, K+ and Cl' flux responses could be explained as the time required to elevate cytosolic free Ca2+ concentrations to a level that can elicit a cellular signalling cascade response. This observation is in agreement with the hypothesis that Ca2+ is a major ion initiating the plasma membrane depolarization in pea mesophyll. Light-induced alkalinization and photosynthetic effects Initial extracellular alkalinization in response to light was observed in leaf tissues of many species including sunflower (Hoffmann and Kosegarten, 1995), pea (Stahlberg and Van Volkenburgh, 1999), and bean (Muhling et al., 1995; Shabala and Newman, 1999). The same result was observed in Figure 7. A light-induced increase in H+ influx occurred after about a 30 s lag time (Table 2a), whereas Shabala and Newman (1999) found a lag time of l~2 min in bean mesophyll. The magnitude of the increase in apoplastic pH, 0.4 ~ 0.5, is similar to the finding of Stahlberg and Van Volkenburgh (1999). In addition, they 74 113; Off pun re P 1 ”“1 an 1101. 11:31 dec tho bel ape also reported an extracellular acidification after 15 min of light exposure with the parallel occurrence of a gradual hyperpolarization of membrane potential. They concluded that growth stimulation of pea leaves is due to H+ efflux mediated by a plasma membrane Hi- pump. Unfortunately, however, H+ fluxes were not measured in their study. In the present experiment, direct H” flux measurement does not support their speculation. Light treatment caused a rapid increase in net H" influx, not an efflux. Light-induced alkalinization was slowed down and extracellular pH started to decline after approximately 5-6 min, whereas net H+ influx was decreased or reversed to net efflux or both, i.e. PIFR. These results seem to correspond to the “acidification” response described by Stahlberg and Van Volkenburgh (1999). Their observation, however, occurred as a result of a 1.0-1.5 h light treatment and the “acidification” was due to a decrease in apoplastic pH from the maximum initial light-induced pH. It should be noted though that the reported reduction in pH was not lower than the original level measured before light was turned on. Termination of light exposure resulted in a decrease of apoplastic pH (Figure 7) in agreement with Stahlberg and Van Volkenburgh (1999). In bean mesophyll tissues, Shabala and Newman (1999) showed a light-induced initial net H+ efflux but not a net influx, even though extracellular pH increased. They explained that this apparent inconsistency could be a result of rapid CO2 uptake by photosynthesizing tissues. Another possible consideration is the effect of other ion transporters which regulate extracellular pH. Several mechanisms can be involved in apoplastic pH regulation (Leigh, 1993), including production and consumption of malate via the biochemical pH ‘stat’, transport of protons to and from the vacuole, and fluxes of 75 811' [1.11 H+ at the plasma membrane. Other ions, such as K“, Ca2+ and anions may be involved in cytoplasmic acidification or extracellular alkalinization which contribute to an initial depolarization of membrane potential (influx of positively charged H“) followed by a repolarization as the H+ pump is activated (Franchisse and Desbiez, 1989). Gating of the K+~inward and Kl~outward rectifiers in guard cells is sensitive to external pH (Thiel and Wolf, 1997). Therefore, it is possible that the activity of other ion transporters in the plasma membrane also influence Hi flux and extracellular pH. The photosynthetic inhibitor DCMU is often used to test the effects of photosynthesis on light-induced electrical signalling in leaf tissues (Nishizaki, 1992; Elzenga et al., 1995; Stahlberg and van Volkenburgh, 1999; Szarek and Trebacz, 1999). In pea mesophyll, 0.05mM DCMU inhibited to a large extent membrane depolarization (Stahlberg and van Volkenburgh, 1999). The presence of DCMU at the same concentration inhibited the light-induced initial response in not H, Ca2+ and K+ influxes (Figure 9) and delayed the initial light response (Table 7). This result experimentally demonstrates the correlation between photosynthesis and light-induced ion flux change. Carbon dioxide (C02) uptake was thought to be one reason for light-induced apoplast alkalinization (Gotow et al., 1985; Shimazaki and Zeiger, 1987; Shabala and Newman, 1999). In lower plants, light-induced pH increase is due to the uptake of HCO‘ ions in (Merrett et al., 1996). Formation of HCO’ ions and CO2 is expected from the following equilibrium: CO2 + H20 <:> HZCO3 ¢:> H” + HCO‘3 <=> CO'3 + 2H“. Reduction in either CO2 or HCO‘3 results in a decline H+ concentration or pH increase in solution. In the bathing solution, CO2 reacts with H20 to form HCO‘3 with a combined pK of 6.3 76 1N: by 11111“ 1‘1‘ 611: 1101 (Neumann and Levine, 1971). In this study, when light was turned on, the uptake of CO2 by mesophyll cells shifted the equilibrium switch toward production of H2C03. This resulted in a decrease in H+ concentration and thus alkalinization. Therefore, DCMU inhibition of photosynthesis which lowers CO2 uptake may significantly inhibit light- induced H+ influx as suggested by this study (Figure 9, Table 7). Additional evidence for the light-induced alkalinization comes from external pH effects. High extracellular pH (7.0) initiated further alkalinization (0.6-0.7 pH unit, data not shown) compared with pH 5.2 (0.4 ~ 0.5 pH unit) of Table 1b. Enhanced pH shifis the equilibrium toward H2C03 formation (Yin et al., 1996) resulting in a further increase in pH. On the other hand, a shift to a more acidic environment (pH 4.0) would be expected to reduce HZCO3 formation and inhibit the rapid pH rise illustrated in Table 6. After 5 min of illumination, only a small pH increase (0.07) was observed in the acidic bathing solution (Table 6), whereas at pH 7.0, extracellular pHiwas transiently increased by 0.4 ~ 0.5 during the same period (data not shown). Similar observations were made in light-induced Cyanidium cadarium (Kura-Hotta and Enami, 1981) and Dunaliella acidophia (Remis et al., 1994). Therefore, it seems reasonable to suggest that CO2 uptake might cause apoplastic alkalinization in the light-induced pea mesophyll tissues. Moreover, the results using different external pH treatment indicate that the regulation of extracellular pH is not solely dependent on H+ flux across the plasma membrane. The high net H+ flux under pH 4.0 was similar to the increase in extracellular pH as under pH 6.0 and pH 5.2 (Table 6). 77 13+ Furthermore, another possible explanation may be that ion uptake is driven by the proton motive potential. H" influx into cytoplasm is mediated by symporters within the plasma membrane, e.g. H+/K+ (Hedrich and Schroeder, 1989; Netting, 2000) and H+/C1' (Sanders et al., 1985). The uptake of H+ by plant cells is energetically favorable due to the negative membrane potential and a low cytosolic H+ concentration in the cytoplasm compared to the apoplast. Thus, H+ fluxes are believed to serve as the energy source for membrane conformational changes that accompany solute transport involving symporters. The influx or uptake of Cl' is mediated by active transports (Assmann and Zeiger, 1987), and hypothesized to occur via a Cl'/H" symport or a Cl'/OH' antiport (Zeiger et al., 1978). The C1“ import requires the input of a considerable amount of metabolic energy. To meet this requirement, 2Hl/Cl' co-transport against an opposing Cl' gradient has been postulated which has indeed been demonstrated and kinetically modeled for Chara (Sanders et al., 1985) and root-hair cell of Sinapis alba (F elle, 1994). The experimental evidence in higher plants, however, has remained somewhat scant to date. In pea mesophyll, a light-induced initial response in net Cl‘ influx was observed within 2-3 min (Table 1a, 2a). Extracellular Cl' concentration declined following the net Cl” influx (Table 1b). With decreasing pH or increasing H+ concentration in the bathing solution (4.0), the light-induced initial response in net Cl' influx was significantly enhanced as compared to that at pH 5.2 and pH 6.0 (Table 5). However, high external pH (6.0) resulted in a higher light-induced PIFR of Cl' than at pH 4.0 (Table 5). Similar results were observed for the Kl response to external pHs (Table 5). Therefore, Cl'/H+ and H+/K+ symport might be hypothesized to mediate light-induced apoplastic 78 C1 "(J {/J alkalinization of pea mesophyll. More experimental evidence is required to clarify this issue though. No direct attempt has been made to experimentally determine the influence of extracellular pH on Ca2+ flux fiom leaf cells so far. The Hl/Ca2+ antiporter is a likely candidate for H+ ~ Ca2+ exchange which in principle mediates transport of Ca2+ against its concentration gradient and requires a H+/Ca2* stochiometry of at least three (Schumaker and Sze, 1987). However, activity of Hl/Ca” activity frequently was associated with membranes of mitochondria and chloroplast and only rarely with the plasma membrane (Bush, 1995). In this experiment, external pH did not influence the light-induced initial Ca2+ flux change (Table 5). Compared with pH 4.0, the alkalinizing environment (pH 6.0) inhibited the light-induced PIFR of Ca2+ (Table 5). Presumably changes in external pH alter the pH gradient across the plasma membrane or affect feedback regulation by cytoplasmic pH, thus influencing cytosolic Ca“ homeostasis. The light-induced PIFR is evidenced in the current study. Shabala and Newman (1999) did not discuss this event. The PIFR may reflect a combined response of Ca2+, H“, K” and Cl' flux in leading to membrane repolarization which could be a cellular mechanism to regulate external pH. Another possible function of the PIFR could be to regulate a cytoplasmic Ca2+ messenger which is involved in light-induced signal transduction. Further investigation on the interaction between ion transporters in the plasma membrane is needed. Light-induced ion fluxes from epidermal cells. Shabala and Newman (1999) only measured a relatively small change of net ion fluxes 79 from bean epidermal strip when exposed to light. They suggested that the epidermis is an effective barrier to ion flux but not for CO2 diffusion into the leaf. Elzenga et al. (1995) proposed that the ionic mechanisms of light-induced depolarization for mesophyll and epidermis of pea are different. In the current research, the light-induced initial ion flux changes in epidermal tissues were also smaller than in mesophyll (Figure 8, Table 4). Light initially induced a small decrease in net Hl efflux resulting in apoplastic alkalinization (0.08 pH unit), and a reduction in extracellular Ca2+ concentration (28 11M) as well. The summation of light-induced net flux Ca2+, Hl, K+ and C1' in epidermal tissues was smaller than in mesophyll tissues (Figure 11). This result was similar to the relatively small light-induced membrane potential change reported by Elzenga et al. (1995). Although light-induced ion flux changes were observed, it is impossible to speculate on the ionic mechanism for electrical signalling in epidermal tissue. The results in this study explain in part the ionic basis for light-induced bioelectrogenesis. However, more evidence is needed on the specific ion transporters involved in electrical signal transduction. Further investigations should also focus on the role of ion channels and proton pumps in light responses. 80 LITERATURE CITED Antkowiak S, Mayer WE, Engelmann W. 1991. Oscillation of the membrane potential of pulvinar motor cells in situ in relation to leaflet movements of Deodium motorium. J Exp Bot 240:901-910. Arif I, Newman IA, Keenlyside N. 1995. Proton flux measurements fiom tissues in buffered solution. Plant Cell Environ 1 8: 1 3 1 9-1324. Assmann SM, Zeiger E. 1987. Guard cell bioenergetics. In E Zeiger, E Farquhar, I Cowan, eds, Stomatal Function. Stanford University Press, Stanford, CA, ppl63-l94. Bush DS. 1995. Calcium regulation in plant cells and its role in signaling. Annu Rev Plant Physiol Plant Mol Biol 45:95-122. Elzenga JTM, Prins HBA, Van Volkenburgh E. 1995. Light-induced membrane potential changes of epidermal and mesophyll cells in growing leaves of Pisum sativum. Planta l97:127~134. Elzenga JTM, Van Volkenburgh E. 1993. Ion channels in the plasma membrane of epidermal and mesophyll cells of growing pea leaves. Plant Physiol 102:S106. Elzenga JTM, Van Volkenburgh E. 1997. Characterization of a light-controlled anion channel in the plasma membrane of mesophyll cells of pea. Plant Physiol 113:14l9~ 1426. Ermolayeva E, Hohmeyer H, Johannes E, Sanders D. 1996. Calcium-depend membrane depolarization activated by phytochrome in the moss Physcomitrella patens. Planta 199:352-358. Felle HH. 1994. The Hi/Cl' symporter in root-hair cells in Spinais alba. Plant Physiol 106:1131-1136. Frachisse JM, Desbiez MO. 1989. Investigation of the wave of electrogenic active and passive membrane elements in Riccafluitans. Biochem Biophys Acta 646: 1 51-160. Gotow K, Sakaki T, Tondo N, Kobayashi K, Syono K. 1985. Light-induced alkalinization of the suspending medium of guard cell protoplasts from Vicia fava L. Plant Physiol 79:825-828. Hedrich R, Schroeder JI. 1989. The physiology of ion channels and electrogenic pump in higher plants. Annu Rev Plant Physiol Plant Mol Biol 40:539-569. Hoffmann B, Kosegarten H. 1995. FITC-dextran form measuring apoplast pH and apoplastic pH gradient between various cell types in sunflower leaves. Physiol Plant 81 95:327-335. Johannes E, Ermolayeva E, Sanders D. 1997. Red-light-induced membrane potential transients in the moss Physcomitrella patens: ion channels interaction in phytochrome signalling. J Exp Bot 48:599-608. Johnson CM, Stout PR, Broyer TC, Carton AB. 1957. Comparative chlorine requirements of different plant species. Plant and Soil 8:337-353. Kochian LV, Shatt JE, Kuhtreiber WM, J affe LF, Lucas WJ. 1992. Use of extracellular, ion selective, vibrating microelectrode system for the quantification of K, H, and Ca2+ fluxes in maize roots and maize suspension cells. Planta 188:601—610. Kura-Hotta M, Enarni I. 1981. Light-induced H+ efflux from intact cells of Cyanidium caldarium. Plant Cell Physiol 22:1175-1183. Leigh RA, Wyn Jones RG. 1984. A hypothesis relating critical potassium concentrations for growth to the distribution and functions of this ions in the plant cell. New Phytol 97:1-13. Leigh RA. 1993. Perception and transduction of stress by plant cells. In L Fowden, T Mansfield, J Stoddart, eds, Plant adaptation to environmental stress. Chapmem and Hall, pp223~237. Lewis BD, Karlin-Neumann G, Davis RW and Spalding EP. 1997. Cay-activated anion channels and membrane depolarization induced by blue light and cold in Arabidopsis seedlings. Plant Physiol 114:1327-1334. Linnemeyer PA, Van Volkenburgh 1E, Cleland RE. 1990. Characterization and effect of light on the plasma membrane H-ATPase of bean leaves. Plant Physiol 94: 1671-1676. Long JM, Widders IE. 1990. Quantification of apoplastic potassium content by elution analysis of leaf lamina tissue from pea (Pisum sativum L. cv. Argenteum). Plant Physiol 94:1040-1047. Lucas WJ, Kochian LV. 1986. Ion transport processes in corn roots: an approach utilizing microelectrode techniques. In WD Gensler, ed, Advanced agricultural instrumentation, design and use. Martinus Nijhoff, Dordrecht, The Netherlands, pp 402. Lfittge U, Pitman MG. 1976. Transport in Plant 11, Part A, Cell; Part B, Tissues and Organs. Encycl Plant Physiol, New Series, V012, Springer-Verlag, Berlin. Marré MT, Albergoni FG, Moroni A, Marré E. 1989. Light-induced activation of electrogenic Hi extrusion and K+ uptake in Elodea densa depends on photosynthesis and is mediated by the plasma membrane H+-ATPase. J Exp Bot 40:343-352. 82 Merrett MJ, Nimer NA, Dong LF. 1996. The utilization of bicarbonate ions by the marine microalga Nanochlorepsis oculate (Droop) Hibberd. Plant Cell Environ 19:478- 484. Mimura T, Tazawa M. 1986. Light-induced membrane hyperpolarization and adenine nucleotide levels in perfirsed Characean cells. Plant Cell Physiol 27:319-330 Muhling KH, Plieth C, Hansen UP, Sattehnacher B. 1995. Apoplastic pH of intact leaves of Viciafava as influenced by light. J Exp Bot 46:377-382. Mummert H, Gradmann D. 1991. Action potentials in Acetabularia: measurement and stimulation of voltage-gated fluxes. J Membr Biol 124:266-273. Netting AG. 2000. pH, abscisic acid and the integration of metabolism in plants under stressed and non-stressed conditions: cellular response to stress and their implication for plant water relations. J Exp Bot 51:147-158. Neumann J, Levine RP. 1971. Reversible pH changes in cells of Chlamydomonas reinhardtii resulting from CO2 fixation in the light and its evolution in the dark. Plant Physiol 47:700-704. Newman IA, Kochian LV, Grusak MA, Lucas WJ. 1987. Fluxes of H+ and K” in corn roots. Plant Physiol 84:1177-1184. Nishizaki Y. 1992. Effects of inhibitors on the light-induced changes in electric potential in pulvinar motor cells of Phaseolus vulgaria L. Plant Cell Physiol 33:1073- 1078. Nobel PS. 1991. Physiochemical and Environmental Plant Physiology. Academic Press, Inc. pp110-l39. Prins HBA, Snel JFH, Zanstra PE, Helder RJ. 1982. The mechanism of bicarbonate assimilation by the polar leaves of Potamogeton and Elodea: CO2 concentrations at the leaf surface. Plant Cell Environ 52207-214. Remis D, Treffny B, Girnmler H. 1994. Light-induced Hl transport across the plasma membrane of the acid-resistant green alga Dunaliella acidophila. Plant Physiol Biochem 32:75-84. Sanders D, Smith FA, Walker NA. 1985. Proton/chloride cotransport in Chara: mechanism of enhanced influx after rapid external acidification. Planta 163:411-418. Schmnaker KS, Sze H. 1987. Inositol 1,4,5-triphosphate releases Ca2+ from vacuolar membrane vesicles of oat roots. J Biol Chem 262:3944~3946. 83 Shabala NS, Newman IA. 1997. Proton and calcium flux oscillations in the elongation region correlate with root mutation. Physiol Plant 100:917-926. Shabala NS, Newman IA, Morris J. 1997. Oscillations in H+ and Ca2+ fluxes around the elongation region of corn and effects of external pH. Plant Physiol 113:111-118. Shabala NS, Newman IA. 1999. Light-induced changes in hydrogen, calcium, potassium, and chloride ion fluxes and concentration from the mesophyll and epidermal tissues of bean leaves. Understanding the ionic basis of light-induced bioelectrogensis. Plant Physiol 119:1115-1124. Shimazaki K, Zeiger E. 1987. Red light-dependent CO2 uptake and oxygen evolution in guard cell protoplasts of Viciafava L.: evidence for photosynthetic CO2 fixation. Plant Physiol 84:7-9. Spalding EP, Cosgrove DJ. 1992. Mechanism of blue-light-induced plasma membrane depolarization in etiolated cucumber hypocotyls. Planta 188:199-205. Spalding EP, Slayman CL, Goldsmith MHM, Gradmann D, Bertl A. 1992. Ion channels in Arabidopsis plasma membrane . Transport characteristics and involvement in light- induced voltage changes. Plant Physiol 99296-102. Staal M, Elzenga JTM, Van Elk AG, Prins HBA, Van Volkenburgh E. 1994. Red and blue light-stimulated proton efflux by epidermal leaf cells of the argenteum mutant of Pisum sativum. J Exp Bot 45:1213-1218. Stahlberg R, Van Volkenburgh E. 1999. The effect of light on membrane potential, apoplastic pH and cell expansion in leaves of Pisum sativum L. var. Argenteum. Planta 208:188-195. Szarak I, Trebacz K. 1999. The role of light-induced membrane potential changes in guttation in gametophytes of Asplem'um trichomanes. Plant Cell Physiol 42:1280-1286. Thiel G, Wolf A H. 1997. Operation of K+~channels in stomatal movement. Trends Plant Sci 2:339-343. Yin ZH, Huve K, Heber U. 1996. Light-dependent proton transport into mesophyll vacuoles of leaves of C3 plants as revealed by pH-indicating fluorescent dyes: a reappraisal. Planta 19929-17. Zeiger E, Bloom AJ, Helper PK. 1978. Ion transport in stomatal guard cells, A chemiosmotic hypothesis: What’s new. Plant Physiol 77:29-32. 84 Chapter II. POSSIBLE MECHANISM OF LIGHT-INDUCED CA”, H‘, K‘ AND CL' FLUXES FROM MESOPHYLL TISSUES. 85 ABSTRACT Mechanisms of light-induced rapid changes in net Ca2+, H“, K+ and C1” flux around mesophyll tissues from excised leaflet of pea (Pisum sativum L. cv. Argenteum) were analyzed. Calcium channel blocker LaCl3 (0.1 mM) or chelator EGTA (1.0 mM) inhibited 42-44 % of the light-induced initial response of net Ca2+ influx and the light~ induced Post-Initial Flux Response (PIFR) of Cl’ and K“ over time. In contrast, the Ca2+ ionophore A23187 (5 11M) and increased exogenous Ca2+ (2.0 mM) significantly enhanced the light-induced (PIFR) of Cl', whereas low external Ca” (0.1 mM) produced an opposite effect. The evidence suggests that Ca2+ influx, resulting from increased Ca2+ channel activity, might be an initial step in the rapid response by leaf mesophyll to light. The plasma membrane H+~ATPase inhibitors, Na3VO4 (1.0 mM) and DCCD (0.1 mM), inhibited the light-induced initial response in net H+ influx by 50 - 60 % and eliminated the PIFR of H“. The presence of Na3VO4 in the bathing solution also significantly suppressed the light-induced increase in net K+ influx. However, these inhibitors did not affect the light-induced response in net Ca2+ influx. Furthermore, fusicoccin (0.01 mM) significantly increased the light-induced PIFR of Hi and the initial response in net Cl' influx. These results suggest that H+~ATPase in the plasma membrane is not involved directly in the initial light response, whereas H“-K+ and H+~Cl' symports may contribute to the light-induced initial extracellular alkalinization and the uptake of K+ and C1' by mesophyll tissues. In addition, TEA (5.0 mM) significantly blocked most of the light- induced PIFR of K. The R-type anion channel blocker DIDS (0.1 mM) completely suppressed light-induced net Cl' efflux within the first 5 min of illumination, but the S- type blocker A~9~C (0.1 mM) did not. These observations indicate that the outwardly- 86 directed K“ channels and the R-type anion channels may play a role in light-induced PIFR of K+ and Cl', respectively, whereas that of Cl' is Cay-dependent. INTRODUCTION Rapid light-induced changes in membrane potential are ubiquitous in higher green plants. This bioelectrogenesis can be explained as light-induced changes in ionic fluxes across the plasma membrane. Light has been shown to elicit rapid changes in ion flux in leaves of Dunalliella acidophila (Remis et al., 1994), Physcomitrella patens (Johannes et al., 1997), bean (Vicia fava) (Shabala and Newman, 1999). These reports propose that ion transport across the plasma membrane is highly regulated, involving specific transport mechanisms during periods of light exposure. However, few studies have related ion transport activity and regulatory mechanisms to responses to light. In general, plant physiological responses to light depend on cellular signalling systems that link light perception to terminal responses. Therefore, 1i ght-induced ion flux changes may link one or several signalling pathways. Ion fluxes across cell membrane systems are controlled by many ion channels, carrier proteins, and electrogenic ion pumps. These ion transporters regulate transport of ions across the plasma membranes and maintain ion homeostasis, and large chemical potential gradients. Light stimulation may activate or up/down regulate certain mechanisms that mediate changes of ion fluxes across membranes of plant cells. Light-induced leaf movement has been shown to involve K+, Cl‘ and H‘“ fluxes mediated by K+ channels and proton pumps in extensor and flexor tissues (Coté, 1995; Mayer et al., 1997). Other studies have focused on stomatal guard cell changes in turgidity controlled by ion fluxes across the plasma membrane of guard 87 cells (Grabov and Blatt, 1998; Assmann and Shimazaki, 1999). The dominant ions involved in signal transduction in guard cells include Hi, Ca2+, K+ and Cl‘. Ion channels (such as Ca2+~, inward K+ rectifier and outward K+ rectifier and anion-channels) and H1 ATPase are considered to play important roles in light-responses of guard cells. However, research is lacking on the regulation and responsiveness of ion transport mechanisms in other leaf tissues to environmental stimuli. Data on the ionic basis of light-induced membrane potential changes are also controversial and vary with plant species and experimental conditions. To date, it is not fully understood which ion transporters are associated with light- induced ion flux responses in leaf tissues. Many possibilities exist because multiple ions are involved in the regulation of plasma membrane electrical potential. Shabala and Newman (1999) measured directly ion fluxes from bean mesophyll tissues and suggested that light-induced Ca2+ influx is a main depolarizing agent. They did not, however, explain the relationship between Ca2+ channel activity and the light-induced transient net Ca2+ influx. The results of Johannes et al. (1997) support a role for Ca“ in the red-light- induced flux change in Physcomitrella patens. Blom-Zandstra et al. (1997) reported transient light-induced changes in ion channel and proton pump activities in the plasma membrane of tobacco mesophyll protoplasts, and suggested that ionic transporters in the plasma membrane might be involved in the light-induced transient changes in membrane potential. In Arabidopsis mesophyll, white light induced an increase in K“ channel activity in the plasma membrane and a transient membrane depolarization (Spalding et al., 1992). In pea mesophyll cells, the light-controlled anion channels in the plasma membrane were proposed as a mechanism for membrane depolarization (Elzenga and 88 Van Volkenburgh, 1997). Further investigation indicated that the plasma membrane Hi- ATPase is not directly involved in rapid light-induced membrane depolarization (Stahlberg and Van Volkenburgh, 1999). They also suggested that several different mechanisms participate in light-induced membrane depolarization in the mesophyll and the epidermis (Elzenga at al., 1995). However, ion flux in leaf cells was not directly measured in the above studies. In Chapter I, the light-induced rapid changes in the net flux of H", Ca2+, K+ and Cl' were demonstrated in pea mesophyll tissues. The transient net Ca2+ influx and the concomitant apoplastic alkalinization were thought to involve Caz’ channels, or Hi-K+ or Hl-Cl' symporter, as well as or H+~ATPase in the plasma membrane. The light-induced Post-Initial Flux Response (PIFR) of Cl‘ from a net influx to a net efflux suggests that anion channels may also play a role whereas the function of K“ fluxes in light response has still not been confirmed in pea mesophyll. Differential changes in ionic flux reveal that there are several ion transporters involved in the cellular mechanisms regulating light signal transduction. Therefore, firrther research is necessary to determine which transport mechanisms are associated with the light-induced ion flux changes. Specific chemicals that inhibit or activate ion channels or ion pumps are used widely to determine the effect of environmental stimuli on specific ion transporters in the plasma membrane. However, most previous studies have only measured light-induced membrane potential change. The role of ion transporters cannot be elucidated from only membrane potential data alone. To test our hypothesis, two different approaches were used. The first consisted of treating pea mesophyll tissues with selective inhibitors or activators of H“, Ca2+, K" and Cl’ transporters in the plasma membrane before light 89 illumination. Light-induced dynamic ion flux changes of H, Ca2+, K+ and Cl' were then monitored using the Microelectrode Ion Flux Estimation (MIFE) system. The second approach was to observe ionic flux responses to light in different external ionic solution environments. In this way, one is able to study possible mechanisms involved in the ionic response to light. MATERIALS AND METHODS Plant Material Seeds of pea (Pisum sativum L. cv. Argenteum) were germinated 4-5 (1 and then transferred to a growth chamber with a 16 h photoperiod, day/night temperatures of 25/25 °C, 7 5% relative humidity, and a radiance of 200 pmol m’2 s" at the top of the canopy. The plants were cultured using modified Hoagland’s solution (Johnson et al., 1957), gradually increasing the concentration from 0 to 1A to I/2 strength at 4-5 (1 intervals. Flux measurements Ion fluxes and concentrations at the surface of the leaf mesophyll were measured using the MIFE system. Technical details were generally as described in Chapter I and Shabala et a1. (1997). Additional information on the MIFE system is available at http://www.phys. utas.edu.au/physics/biophys. In brief, micropipettes with a tip diameter of 2-3 um were pulled fiom 1.5-mm-diameter borosilicate glass capillaries (catalog no. MIB 150-6; World Precision Instruments, Inc., Sarasota, FL) with a horizontal electrode puller (PC-10, Narishige, Tokyo, Japan). After silanization, cooled microelectrodes were back-filled with 500 mM CaCl2 for Ca“, 500 mM KCI for K+ and Cl', and 15 mM NaCl 90 plus 40 mM KH,,PO4 (adjusted to pH 6.0 with NaOH) for H”. Electrode tips were then front-filled with a liquid-membrane-type of the appropriate ion-selective sensor, Hi (no. 95291), Ca2+ (no. 21048), K+ (no. 60031) and Cl‘ (no. 24902) cocktails (all from F luka, Milwaukee, WI), until a 50- to 80- um long column of ionophore was produced. Microelectrodes were calibrated for each ion using a known set of standards before and after use (see Chapter I). The Ag/AgCl reference electrode was made using a plastic tube containing 1000 mM KCl in 2% agar. Calibrated ion-selective electrodes were positioned over the desired cells in a leaf segment using a micromanipulator with three-dimensional control. The set-up was mounted on the stage of an inverted microscope (Leitz Wetzlar, Germany) which provides a clear view for electrode positioning. The sample chamber with mounted leaf segment was placed on a three-way hydraulic micromanipulator (WR 88, Narishige, Tokyo, Japan) driven by a computer-controlled stepper motor (MO61-CE08, Superior Electric, Bristol, CT). The MIFE system allows for measurement of three ions simultaneously. Three microelectrodes can be slowly vibrated at any angle in a two- dimensional plane by separately controlling the amplitude of the vibration of the micromanipulator. I The MIFE system measured the voltage differences at two extreme positions of a 50 um vibration above the leaf tissue surface. The magnitude and the directionality of the difference in estimated ion activity between the two positions was the basis for estimating both the rate and directionality of the ion fluxes. In this research the electrodes were moved in a 20—s square wave cycle between 50 to 90 um above the leaf surface at a low frequency (0.1 Hz). The concentrations of ions were calculated from their 91 electrochemical potentials at each position. The flux for each specific ion was calculated from the measurements of the difference in the electrochemical potential between these positions (Shabala et al., 1997, or see Chapter I). During analysis, the first 4 s of each half cycle (10 s duration) were not used for concentration and flux estimation. This 4 3 period accounted for the time for electrode movement to a new position and for 'voltage potential stability. Experimental protocol Young expanding leaflets (leaf age about 2 weeks) were harvested and a leaf section of approximately 15 x 10 mm (LxW) with limited veinal tissue was cut surgically from the lamina. A small strip of epidermal tissue was peeled from the abaxial surface of the lamina of each section so as to prevent mechanical damage to the underlying veinal and mesophyll tissue (Long and Widders, 1990). Isolated leaf segments were floated on the buffered solution (0.5 mM CaSO4, 1.0 mM KCI, 2.5 mM HEPES-HCI, and 10 mM sucrose at pH 5.2) under 60 W m‘2 illumination (Philips lamp). After 2.0-2.5 h, the cut segment was mounted in a sample chamber. The sample chamber contained 6-7 mL unbuffered solution or bathing solution (0.5 mM CaSO,, 1.0 mM KCl, pH 5.2), and the segments then were exposed to 50—60 min of darkness. After 5-10 min of measurement, the light was turned on. Flux measurement was continued for an average 30 min after the start of illumination. After light was turned off, the measurement was conducted for another lO-15 min. The sample chamber was then removed and a new leaf sample was installed. 92 A fiber-optic light source (delivering 450 pmol m'2 s") as measured by a quantum sensor, model LI-189, LI-COR, Lincoln, NE) was used with a 250 W halogen lamp (GS Edmumd Scientific Company, Barrington, NJ). A lens (no L~16~16, World Precision Instruments, Inc., Sarasota, FL) was fixed on the top of light guide to focus the beam. Chemicals Chemical treatments were selected according to their effects on ion transporters. These treatments included a Ca2+ channel blocker (LaC13), a Ca“ chelator (EGTA), a Ca2+ ionophore A23187, H+~ATPase inhibitor (vanadate and DCCD), a Hl-ATPase activator fusicoccin (FC), K+ channel blockers (TEA and CsCl), and anion channel blockers (A-9- C and DIDS). Vanadate (Fisher, Lawn, NJ), CsCl, EGTA, and LaCl3 (SIGMA, St. Louis, M0) were dissolved directly in an unbuffered solution (0.5 mM CaSO,,, 1.0 mM KCI, pH 5.2). Others including FC, DCCD, A-9-C, DIDS, Ca2+ ionophore A23187 and TEA (from SIGMA, St. Louis, M0) were prepared in DMSO as stock solutions. To test the effects of these chemicals (V anadate, DCCD, F C, EGTA, LaC13, Ca2+- ionophore A23187, A-9-C, DIDS, TEA, CsCl) on ion fluxes under light stimulation (450umol m'2 s“), all chemicals from their stock were dissolved with the unbuffered solution adjusted to pH 5.2. The final DSMO concentration was below 0.5 % (v/v) in the bathing medium. The leaf segment was incubated in the unbuffered bathing solution (containing one of these chemicals) under darkness for 50-60 min. Following the procedure of experimental protocol described above, ion flux was measured over an average 30 min period of light exposure. Vanadate, DCCD, FC, EGTA, LaC13, Ca2+ ionophore A23187, A-9-C, DIDS, TEA, 93 CsCl were added to bathing solution, respectively, to test then individual effects on a specific ion flux under low fluence (50umol m'2 5"). Different external concentrations of Ca2+ (0.1 and 2.0 mM), K” (0.2, 5.0 and 10 mM) and Cl’ (0.2, 2.0, and 10 mM) were achieved by changing the salt concentrations in the bathing solution. Data statistical analysis Data analysis has been described in Chapter 1. Analysis of variance (ANOVA) of the data was conduced using SAS 8 (SAS Institute Inc, Cary, NC). RESULTS Effects of H+-ATPase inhibitor and FC. To test the possible involvement of a plasma membrane proton pump in the light-induced flux change, two plasma membrane H+~ATPase inhibitors, DCCD (at 0.1 mM) and Na3VO, (at 1.0 mM), as well as a stimulator of H“-ATPase, Fusicoccin (FC) at 0.01 mM, were applied to the bathing solution (0.5 mM CaSO,, 1.0 mM KCI, adjusted to pH 5.2). The PC is a powerfirl effector that stimulates the plasma membrane IP-ATPase and is used as an experimental tool to investigate the mechanism of physiological modulation of ' this enzyme (Marré, 1979). Under low background light (50 pmol m‘2 s"), the net H“ flux from mesophyll tissues of the control (Figure 1) exhibited a small change around zero. However, the addition of DCCD and Na3VO4 inhibited H+ efflux and resulted in a high net H+ influx of 16-20 nmol m'2 s“. In contrast, FC stimulated H+ efflux and resulted in decrease in net H+ flux of 18 nmol m'2 s'l after approximately 30 min (Figure l). 94 1' -0- control ‘ + 0.1 mM DCCD A 30 _ + 1mMNa v04 -* ...— + 0.01 mM FC to " .l N 'E 20 — ‘_._____________. — ‘5 Add N E I - II 5 x 10 - 1 -~ — 3 t: ' v - s 0 - r C .E ‘ " ‘ a) 3» ~10 — ~- _ ‘0 _- 5 - LSD(0,05) \‘ q ~20 — _ 1 1 1“ 1 1 1 1 1 0 5 10 15 20 25 30 35 40 Time (min) Figure 1. Effect of H+-ATPase inhibitors. vanadate (Na3VO )and N, N-dicyclohexi cabodiamide (DCCD), and the ATPase stimulator Fusicocin (FC) on net H flux under fluence (50 pmol 111‘2 s'1 ). Data points represent an average of 4-6 replicates. Vertical bars represent LSD values at p<0.05. 20 <>-ammx + 0.1 mM DCCD . r, -O— 1 mM Na V04 Light-off .. ,5" + 0.01 mM F0 .m l ‘1‘ 10 l— — E Light-on o e - 1 w 25 O — _. +¢ A I E . . c .E 81 ~10 — - c cu .1: O 1 . -20 1 1 1 1 1 1 1 1 1 1 0 5 10 15 20 25 30 35 40 45 Time (min) Figure 2. Effect of H+-ATPase inhibitors, vanadate (N33VO4) and N. N-dicyclohexicabodiamide (DCCD), and the ATPase stimulator Fusicocin (FC) on the time course of light (450 pmol rn'2 s‘1) -induced change in net H+ flux from pea mesophyll tissues. Data points represent an average of 6~1O replicates. 95 After DCCD, NagVO, and FC were added (at the same concentrations as in low light) to the bathing solution (0.5 mM CaSO,, 1.0 mM KCl, adjusted to pH 5.2), the leaf tissue was incubated for 50-60 min in the dark. Net H+ flux and extracellular pH around the pea mesophyll tissues were then measured following light stimulation (450 pmol rn‘2 3"). Light induced a rapid increase in both net H+ influx and extracellular pH. After 5-10 min this influx (the control) decreased or changed to a net efflux, which is known as the light- induced PIFR (Figure 2, Table l). The PIFR could involve a gradual return to flux rate closer to or lower than that measured prior to light stimulation. Either decrease in influx or increase in efflux (or both) caused the PIFR. Addition of DCCD and Na,VO4 resulted in a significant inhibition of the light-induced initial response (61% and 54% of the control respectively) in net H+ influx (Table 1) and. extracellular pH (data not shown). A 50-60 % reduction in H+ PIFR was observed in the presence of DCCD or Na.,VO4 (Table l). The F C also reduced the light—induced initial response in net H+ influx within 5 min of exposure to light and suppressed the light-induced extracellular alkalinization. As compared to the control (~22 nmol m'2 s"), F C caused a higher amplitude of PIFR (~46 nmol 111‘2 s") from net influx to net efflux (Table 1), indicating that F C had stimulated H+ efflux from the plasma membrane. As discussed previously in Chapter I, light induced a transient increase in net Ca2+ influx. Both DCCD and Na,VO4 treatments significantly reduced the light-induced PIF R of Ca2+ (38 % of the control), whereas Na,VO4 significantly inhibited the light-induced initial response (28 %) in net K+ influx (Table 1). FC increased light-induced initial response in net Cl' influx (240 %) but inhibited that of net Ca2+ influx (Table 1). During the first 5 min of light treatment, the light-induced net Cl' flux change was 27 nmol m'2 s'1 96 Table 1. Effects of H+-ATPase inhibitors and FC on light-induced net fluxes of 11“, Ca“, K" and Cl' from pea mesophyll tissues. 11* Ca” K: cr Qmol m'2 8") Initial flux response (influx) control +126(10) §a +48(10) a +171 (8) a +101 (8) b 1.0 mM Na3VO, +12 (10) b +57 (7) ab +132 (8) b +72 (7) b 0.1 mM DCCD +12 (8) b +52 (8) ab +139 (7) ab +67 (7) b 0.01 mM FC +24 (7) a +30 (6) b +200 (6) a +242 (6) a Flux change during the first min control +4 (10) +26 (10) a +24 (8) a +13 (8) ab 1.0 mM Na:,VO4 +2 (10) +35 (9) a +19 (8) a -9 (7) b 0.1 mM DCCD +4 (8) +7 (10) ab +27 (7) a ~20 (7) b 0.01 mM FC ~1 (8) ns +1 (8) b ~62 (6) b +69 (6) a Flux change during the first five min control +13 (10) +29 (10) a +67 (8) ~60 (8) 1.0 mM Na3VO, +4 (10) +41 (8) a +16 (7) ~55 (7) 0.1 mM DCCD +6 (8) +17 (10) ab +17 (7) -36 (7) 0.01 mM FC +4 (8) ns ~13 (6) b +9 (6) ns +27 (6) ns Amplitude of PIFRi control ~22 (10)a -63 (10) a ~188 (8) ab -151 (8) 1.0 mM Na3VO4 ~9 (8) c ~40 (8) b ~138 (8) a -213 (5) 0.1 mM DCCD ~10 (8) c -49 (8) b ~200 (7) b ~l73 (7) 0.01 mM FC ~46 (6) b ~53 (6) ab ~181 (6) ab -207 (6) ns Different letters indicate significant difference between treatments within columns at p < 0.05 according to LSD test. “ns” indicates no significant difference. § The numbers in parentheses indicate the number of replicates. 1 ~ or + indicates decrease or increase of net ion flux and extracellular ion concentration, respectively. ‘ PIFR: Light-induced Post-Initial Flux Response, a result of either a decrease in net influx or an increase in net efflux, or both. This value is negative. 97 under F C, indicating a net influx. However, in other treatments, net Cl' flux changes were negative, indicating a net efflux. Additionally, the light-induced PIFR of Cl' (-207 nmol rn’2 5") occurred after 5.3 min, whereas the control only needed 2.1 min. FC did not affect the light-response of K” flux. These results indicated that the inhibition of H1 ATPases caused a decrease in the light induced initial influxes of H+ and KL, and in the PIFRs of IF and Ca2+. On the other hand, an increase in IF-ATPase activity enhanced the light induced initial influxes of Cl' and PIFR of H”. Effects of Ca2+ channel blocker and chelator The Ca“ channel blocker (LaC13) or chelator EGTA were used to evaluate their effects on Ca2+ channel activity. The Ca2+~ionophore A23187 is thought to mediate Cazl/ZH+ exchange across the membrane (Pressman, 1976). When at the same low light background (50 pmol In2 8"), net Ca2+ flux from pea mesophyll tissues in the control only displayed a small change (Figure 3). However, the net Ca2+ influx was inhibited by both LaCl3 (0.1 mM) and EGTA (1.0 mM) but increased by Cay-ionophore A23187 of 5 11M (Figure 3). Light (450 pmol In2 S") elicited a transient increase in net Ca2+ influx followed by a PIFR (Figure 4, or see Chapter 1). Compared with the control, both LaCl3 and EGTA significantly reduced the light-induced initial response in net Ca2+ influx (by 42-44% of the control) and corresponding response in extracellular Ca” concentration (64% and 39%, data not shown), as well as the light-induced PIFR of Ca2+ around pea mesophyll (27 % and 37 %, respectively). Its initial light response time also was delayed by both 98 60 T -O-1mMEGTA .. + 0.1mM ‘_-~ j + SuM Calfic‘ignophore - (:0 40 _ + 5 Ca-ionOphore _ 'E . . ‘o E 20 ~ _ 5 x 1 . .2 + 0 l" - N (U 0 h 1 2 ~20 — - .E O F . g -40 _ _ 6 . . ‘60 " 1 1 1 1 1 1 l 1 T 0 5 10 15 20 25 30 35 40 Time (min) Figure 3. Effects of Ca 2+ channel blocker (LaCl3). chelator (EGTA) and ionophore A23178 on net Ca2+ flux from pea mesophyll tissues under low fluence (50 pmol rn'2 3'1). Data points represent an average of 6-8 replicates. Vertical bars represent LSD values at p<0.05. 40 b) O N O ..L o Change in net C32+ flux (nmol m'2 s'1) O -10 '20 _ Lsoom. _ -30 _ .. 1 1 1 1 1" 1 1 I 1 1 0 5 10 15 20 25 3O 35 40 45 Time (min) Figure 4. Effect of the C82+ chelator (EGTA) and channel blocker (LaCl3) on the time course of light (450 pmol m”2 3'1)-induced change in net 032+ flux from pea mesophyll tissues. Data points represent an average of 6~10 replicates. Vertical bars represent LSD values at p<0.05. 99 Table 2. Effects of Ca“ channel blocker and chelator on light-induced fluxes of H“, Ca“, K1 and Cl' and time response in pea mesophyll tissues. H1 Ca2+ K+ Cl‘ Initial flux response (influx) (nmol m'2 5") control +126(10)§ a +45 (8) a +169 (8) a +96 (8) 1.0 mM EGTA +7 (8) b +25 (9) b +121 (8) b +138 (8) 0.1 mM LaCl, +13 (8) b +26 (10)b +75 (8) 0 +85 (9) ns Flux change during the first min (nmol m"2 3") control +3( 10) +24 (8) a +22 (8) +10 (8) 1.0 mM EGTA +1 (8) +13 (9) b +52 (8) +71 (7) 0.1 mM LaCl, +3 (8) ns +4 (10) b +29 (8) us +12 (9) ns Flux change during the first five min (nmol m’2 s") control +13 (10) a +32(8) a +63 (8) ~57 (8) b 1.0 mM EGTA +1 (8) b +9 (9) b +30 (8) +61 (7) a 0.1 mM LaCl, +8 (8) a +100) 0 +13 (8) ns -4 (9) c Amplitude of PIFR‘ (nmol rn‘2 8“) control ~26 (10) a ~60 (8) a ~201 (8) a ~143 (8) a 1.0 mM EGTA ~18 (8) ab -38 (9) b -121 (8) b -34 (4) b 0.1 mM LaCl3 -1l (7) b -44(10) b -99 (8) b -50 (9) b Initial light response time (min) control 3100) 2.1 (8) b 3.9 (8) 2.3 (8) b 1.0 mM EGTA 4.5 (8) 3.7 (9) a 6.6 (8) 4.8 (8) b 0.1 mM LaCl, 3.4 (8) ns 3.8(10) a 4.4 (8) ns 5.6 (9) a Different letters indicate significant difference between treatments within columns at p < 0.05 according to LSD test. “ns” indicates no significant difference. § The numbers in parentheses indicate the number of replicates. I ~ or + indicates decrease or increase of net ion flux and extracellular ion concentration, respectively. ‘ PIFR: Light-induced Post-Initial Flux Response, 8 result of either a decrease in net influx or an increase in net efflux, or both. This value is negative. 100 (Table 2). This indicates that Ca2+ channel blocker and chelator inhibit the light response in net Ca2+ influx. Simultaneously, the light-induced initial responses of net H+ and K“ influx were also significantly suppressed in the presence of LaCl3 and EGTA (Table 2). Neither Ca2+ influx inhibitor affected the light-induced initial response of net Cl' influx. However, the light-induced PIFR of Cl‘ was greatly inhibited by LaCl3 and EGTA (Table 2). During the first 5 min after illumination, a negative change in net Cl' flux, a net efflux, had been observed in the presence of EGTA. The change was significantly higher than that of LaCl3 and EGTA. Like Cl', LaCl3 and EGTA also produced a decrease in the light- induced PIFR of K+ and H+ (only LaC13) (Table 2). Effects of K“ channel blocker Tetraethylammonium (TEA) is a powerful blocker for both classes of K+ channels, inward— and outward-rectifiers (Bentrup, 1990). CsCl blocks inward K+ current (Moran et al., 1990). To identify whether K” channels are involved in the light-induced Kl flux change, TEA-Cl' (5.0 mM) and CsCl (2.0 mM) were added to the bathing solution (0.5 mM C380,, 1.0 mM KCl, adjusted to pH 5.2). In general, light stimulation produced a rapid increase in net Kl influx followed by a PIFR which may be due to a decrease in net K+ influx or an increase in net K or both (Figure 5 or see Chapter I). Compared to the control, TEA significantly blocked the light-induced initial response (35 % of the control) in net K+ influx and the PIFR of K” (66 %) (Table 3). The effects of CsCl on the light- induced K+ flux change were similar but smaller in the magnitudes as compared to that of TBA. 10] 120 -0- control + 2 mM CsCl . - A 100 + 5mM TEA L'QM‘O" _ f...” l- -l ,. l E 80 — _. '5 g ” . :g 60 - - a t l + x 40 _ _ 8 C - . .E G C . .1 2 o 0 — _ r- J- -l _20 1 1 1 1 1 l 1 1 1 1 0 5 1O 15 20 25 30 35 40 45 Time (min) Figure 5. Effect of K+ channel blockers tetraethylammonium (TEA) and Cesium chloride (CsCl) on the time course of light (450 pmol rn'2 s'1Hnduced change in net K+ flux from pea mesophyll tissues. Data points represent an average of 6-10 replicates. Vertical bars represent LSD values at p<0.05. 200 i 1mMA-9-C -O- " -O- 0. + 0.1mMDlDS _ 150 - - LSD(0.05) 4 10° ‘- Licht- - . Light-off . o Change m net 01' flux (nmol m'2 s'1) C 1 O c o 1 -100 r- , -150 i 1 1 1 1 1 1 1 1 1 0 5 10 15 20 25 30 35 4O 45 Time (min) Figure 6. Effect of Cl' channel blockers diisothiocyanatoslibene (DIDS) and anthracene~9~ carboxylic (A-9-C) on the time course of light (450 pmol rn'2 5'1)-induced change in net Cl' flux from pea mesophyll tissues. Data points represent an average of 6-10 replicates. Vertical bar represents LSD values at p<0.05. ‘ _ 102 Table 3. Effects of K+ channel blocker on light-induced net fluxes of H2 Ca“, K+ and Cl‘ from pea mesophyll tissues. H+ Ca2+ K” CL (nmol m'2 8") Initial flux response (influx) control +125 (10)§a +47 (8) +183 (10)a +91 (8) b 5.0 mM TEA +16 (8) b +42 (6) ns +120 (9) b +246 (7) a 2.0 mM CsCl +12 (5) b n.m. +142 (8) ab n.m. Flux change during the first min control +4 (10) +21 (8) +41 (10) +15 (10) b 5.0 mM TEA +5 (8) +30 (6) ns +18 (9) +148 (7) a 2.0 mM CsCl +3 (5) ns n.m. +32 (8) ns n.m. Flux change during the first five min control +11 (10) a +21 (10) +77 (10) ~34 (8) b 5.0 mM TEA +9 (8) ab +10 (6) ns +24 (9) +93 (7) a 2.0 mM CsCl +4 (5) b n.m. +53 (8) ns n.m. Amplitude of PIFR‘ control ~18 (10) ~57 (8) ~189(10) a -l49 (8) a 5.0 mM TEA -20 (8) ~65 (6) ns ~64 (9) b ~39 (3) b 2.0 mM CsCl ~11 (5) ns n.m. ~95 (8) b n.m. Different letters indicate significant difference between treatments within columns at p < 0.05 according to LSD test. “ns” indicates no significant difference. § The numbers in parentheses indicate the number of replicates. 1 ~ or + indicates decrease or increase of net ion flux and extracellular ion concentration, respectively. ‘ PIFR: Light-induced Post-Initial Flux Response, a result of either a decrease in net influx or an increase in net efflux, or both. This value is negative. 103 Both TEA and CsCl also significantly reduced the light-induced initial response (36 and 52 %, respectively) (Table 3) and maximum increase (38 and 50 %, respectively, data not shown) in not H influx. The blocker TEA fluther increased the light-induced initial response of net Cl' influx, possibly because of high Cl’ concentrations in the bathing solution. However, it did not affect the light responses of net Ca2+ flux (Table 3). Effects of Cl' Channel blocker Two anion channel blockers, A~9~C (for the Slow-type channels) and DIDS (for the Rapid-type channels), were used to determine whether anion channel activity was correlated with the light-induced Cl' flux change. Light elicited a rapid and short response in net Cl‘ influx and followed by a PIFR (Figure 6, Table 4, or see Chapter I). The A-9-C treatment did not significantly affect the light-induced initial response of net Cl' influx (Table 4). In contrast to A—9-C and the control, DIDS did not cause a net Cl' efflux but a net influx (48nmol m‘2 5") during the first 5 min of illumination (Table 4). A PIFR of Cl' occurred after 6.3 min in the presence of DIDS, whereas the control only needed 2.3 min of light exposure. When the light was turned off, the net Cl' flux changed to a net efflux for the control whereas A-9-C and DIDS increased the net Cl‘ influx (Figure 6). Both A-9-C and DIDS, when applied at the same concentration (0.1 mM), did not significantly influence the light-induced initial response of net influx of IF, Ca” and K” (K+ data for DIDS not measured experimentally). By comparison to the control, DIDS significantly reduced the light-induced response of H“ and Ca2+ influxes during the first 5 min of illumination and extended the initial light response time of Ca2+ (Table 4). 104 Table 4. Effects of Cl' channel blocker on light-induced net fluxes of H”, Ca“, K+ and CI’ and time response in pea mesophyll tissues. H+ Ca2+ K” C1” Initial flux response (influx) (nmol m2 8") control +127 (8)§ +48 (6) +161 (8) +110 (8) 0.1 mM A-9-C +23 (11) +59 (7) +171 (8) ns +164 (7) 0.1 mM DIDS +18 (8) ns +45 (7) ns n.m. +201 (5) ns Flux change during the first min (nmol m'2 5") control +8 (8) +22 (6) a +35 (8) a +64 (8) 0.1 mM A-9-C +7 (1 1) +21 (7) a ~43 (8) b +139 (7) 0.1 mM DIDS +4 (8) ns -7 (7) b n.m. +30 (7) ns Flux change during the first five min (nmol m'2 s") control +13 (8) a +22 (6) a +68 (8) a ~30 (8) ab 0.1 mM A~9~C +12 (1 1) a +22 (7) b +25 (8) b ~99 (5) b 0.1 mM DIDS +6 (8) b +2 (7) b n.m. +48 (7) a Initial light response time (min) control 4.1 (10) 2.2 (6) b 4.3 (8) b 2.3 (8) b 0.1 mM A-9-C 3.2(11) 3.0 (7) b 8.0 (8) a 1.7 (7) b 0.1 mM DIDS 3.5 (8)ns 7.7 (7) a n.m. 6.3 (7) a Different letters indicate significant difference between treatments within columns at p < 0.05 according to LSD test. “ns” indicates no significant difference. “n.m.”. not experimentally measured. § The numbers in parentheses indicate the number of replicates. I - or + indicates decrease or increase of net ion flux and extracellular ion concentration, respectively. 105 Effects of Ca“ ionophore and external Ca“, K“ and Cl' concentrations Treatment with the Ca2+~ionophore A23187 (5 uM) significantly stimulated the light- induced initial response in net Ca2+ influx from 47 to 67 nmol m‘2 s'l (Figure 7). Compared to the control, the light-induced initial response of not H’“, K+ and Cl' influx were not significantly influenced by the 'Caztionophore (K+ and CT data not shown) (Figure 7). In the same treatment, the light-induced PIFR of Cl' increased significantly from ~144 to -241 nmol m'2 8", whereas the PIFR of K+ was not significantly affected (Figure 7). Therefore, the Ca2+~ionophore and LaCl3 or EGTA produce different effects on light-induced H“, Ca2+, Cl' and K+ flux change in pea mesophyll tissues. When the Ca2+ concentration in the external bathing solution was increased to 2.0 mM (2.0 mM CaSO,, 1.0 mM KCI, pH 5.2), a significant increase in net Ca“ influx response to light was observed (Figure 8). The light-induced PIFR of Cl' was also significantly enhanced by the higher Ca2+ concentration within the bathing solution (Figure 8). Compared with the control (0.5 mM Ca2+), the light-induced initial response in net Hi, K+ and Cl' influxes were not significantly affected (K+ and CT data not shown). However, when the external Ca“ concentration was experimentally set at 0.1 mM, the PIFR of K+ and Cl' declined significantly. The light-induced initial response in net Kl influx increased at higher external K+ concentration (5.0 mM) in the bathing solution and decreased at the lower concentration (0.2 mM K”) (Table 5a). Thus, the initial light response of K+ flux appears to be concentration-dependent. However, changing the K+ concentration did not significantly affect the light-induced initial response nor the PIFR of Hi, Ca2+ and Cl' (Table 5a). 106 300 a E f: 150 - / _/ _ . - / / - .5: 100 — % é _ g1 . 1150905) % / ‘ o 50 +- . ¢ ¢ / _ 0 i % / / A ‘ WWW ..... mama" .... we new Figure 7. Effect of 032+ ionophore A23187 on the light-induced ion flux response of pea mesophyll tissues. Changes in net K+ and CI' flux indicate the amplitude of light (450 pmol m'2 s'1)-induced Post-Initial Flux Response (PIFR), negative values. Data represents an average of 6-9 replicates. Vertical bars represent LSD values at p<0.05. 300 r 1:) 0.1 men? .1 :21 0.5 mM of” (control) ._"‘ 250 L- - 2 mM ea “ _ 'm ‘1‘ l . E '5 200 — _ E 5 ~ . X a 150 r- —l C ‘5 L ‘ O C .s 100 ~ — 8, _ LSD(0.05) . 5 . l. . 0 ra + , initial 11* hflux Initial Cruz" hflux Amplitude Amplitude of response response PFIR for K PFIR for 0' Figure 8. Effect of external Ca2+ concentration on the light-induced ion flux response of pea mesophyll tissues. Changes in net K“ and or flux indicate the amplitude of light (450 pmol m'2 3'1)~induced Post-Initial Flux Response (PIFR), negative values. Data represents an average of 6—9 replicates. Vertical bars represent LSD values at p<0.05. 107 Table 5. Effects of external Kr and Cl’ concentration on light-induced net fluxes of H”, Ca2+, K+ and Cl' from pea mesophyll tissues. a: external K+ concentrations Hi Ca2+ K+ Cl' (nmol mas") ' Initial flux response (influx) control(l.0 mM K”) +l21(7)§ +48 (8) +153 (6) b +91 (7) 0.2 mM K+ +19 (7) +43 (9) +49 (7) c +106 (7) 5.0mM K+ +25 (6) ns +38 (6) ns +255 (7)-a +153 (6) ns Amplitude of PIFR‘ control(1.0 mM K‘) ~16 (7) ~51 (8) ~18] (6) a -143 (7) 0.2 mM K+ ~19 (7) -52 (9) ~82 (7) b ~119 (6) 5.0 mM K+ ~11 (6) ns ~51 (6) ns ~262 (7) ab ~111 (4) ns b: external Cl concentrations H+ Q24» K+ Cl- (InnolnlL2 8") Initial flux response (influx) Control(1.0mM Cl') +27 (8) a +48 (8) a +176 (6) b +103 (8) b 0.2 mM Cl' +17 (5) a +28 (6) b +107 (8) b +101 (6) b 2.0 mM Cl’ +20 (7) a +43 (6) a n.m. +160 (5) b 10 mMKCl +8 (6) b +37 (7) ab +739 (8) a +312 (7) a Amplitude of PIFR Control(1.0mM Cl') ~24 (8) a -59 (8) -193 (6) b ~184 (8) 0.2 mM Cl' ~17 (5) a ~48 (5) ~155 (8) b ~112 (6) 2.0 mM Cl' -23 (7) a ~45 (6) n.m. ~126 (5) 10 mMKCI -8 (5) b ~51 (7) ns ~509 (8) a ~157 (6) ns Different letters indicate significant difference between treatments within columns at p < 0.05 according to LSD test. “ns” indicates no significant difference. “n.m.” not experimentally measured. § The numbers in parentheses indicate the number of replicates. ’ ~ or + indicates decrease or increase of net ion flux and extracellular ion concentration, respectively. 3 PIFR: Light-induced Post-Initial Flux Response, a result of either a decrease in net influx or an increase in net efflux, or both. This value is negative. 108 If the Cl' concentration in the bathing solution was lowered to 0.2 mM (did not alter Ca2+ and K” concentration), the light-induced initial response in net Ca2+ influx was significantly inhibited (Table 5b). Doubling the K” and Cl' concentrations in the external bathing medium to 10 mM KCI dramatically reduced the light-induced initial response in net H+ influx and significantly enhanced net K+ and Cl' influx in response to light (Table 5b). DISCUSSION Ca2+ fluxes and Ca2+ transporters Increase in cytosolic free Ca2+ concentrations is considered to be necessary for cellular signal transduction in response to external stimuli. Passive influx of Ca2+, facilitated by the opening of Ca2+-permeable channels in the plasma membrane, is one of the primary means of increasing cytosolic free Ca2+. Numerous physiological, pharmacological and Ca2+ flux studies have provided indirect evidence that Ca2+ channels in the plasma membrane of higher plant cells play a central role in the initiation of various signal transduction sequences, including, plant-pathogen interactions (Bach et al., 1993), light- induced leaf movement (Mayer et al., 1997), and honnone- and light-regulated stomatal responses (Ward and Schroeder, 1997; Grabov and Blatt, 1998). However, direct evidence for the light-induced Ca2+ channels in the plasma membrane of mesophyll cells has not been obtained. Accurate measurement of Ca2+ influx across the plasma membrane is quite difficult. In the current research, net Ca2+ flux was measured by a noninvasive, ion-selective, slowly vibrating microelectrode. A 1i ght-induced transient increase in net Ca2+ influx was identified and proposed as an initial depolarizing agent in 109 Chapter 1, suggesting that an increase in Ca2+ influx not a decrease in Ca2+ efflux is one major reason for the initial Ca2+ influx response (see Chapter 1). Although H“~Ca2+ exchanges occurred in the cell wall of Chara australis cells (Ryan et al., 1992), the result of net Ca2+ flux mainly represents free Ca2+ movement within the apoplast of the tissues. Reid and Smith (1992) did not find significant effects on Ca2+ influx following changes in external pH or following changes in internal pH by separating the cell wall from cell contents of Chara corallina, indicating that exchanges between H+~Ca2+ are relatively constant under a certain range of pH in the bathing solution. Therefore, it is reasonable to believe that the light-induced net Ca2+ influx is the result of Ca2+ uptake facilitated by Ca2+~channels in the plasma membrane. Further evidence for the role of putative Ca2+ channels is obtained by examining the effects of a channel blocker and a Ca2+ chelator. Both LaCl3 and EGTA significantly suppressed the light-induced transient increase in net Ca2+ influx (Table 2), demonstrating the relationship between Ca2+ influx and Ca2+ channel activity. At least two major classes of Ca2+ channels reside in the plasma membrane (White, 1998). The first class includes those that are relatively non-selective with respect to cations and possess a high single-channel conductance; these are known as maxi-cation channels (White, 1993; White, 1994). The second class is relatively more selective for cations, and exhibits a smaller single-channel conductance. Since the addition of LaCl3 and EGTA also resulted in a significant reduction of the light-induced initial response of net K+ influx (Table 2), the influx of both Ca2+ and K+ is assumed to occur by diffusion through non—selective cation channels. The same finding was reported in Samanea saman protoplasts (Moran and Satter, 1990) and during light-induced swelling of 110 Phaseolus coccineus protoplasts (Mayer et al., 1997). The Ca2+ or K+ influx across the putative cation-channels was dependent upon the extemal concentration of Ca2+ or K+ but not upon each other (Table 2, 5). However, an organic Ca2+ channel blocker, verapamil (0.1 mM), did not affect light-induced Ca2+ influx or other ion fluxes (data not shown). Reid and Smith (1992) also made similar observations in Chara. In animal cells, La“ can block at least two types (L~ and N-type) of voltage-activated Ca2+ channels, whereas verapamil only blocks the L~type (Tsien et al., 1987). Apparently, not all types of cation- channel are responsive to light in pea leaf tissues. The effect of Ca2+ on the light-induced PIFR of Cl‘ provides more evidence for Ca2+ influx as an initial response to light. Clearly, increased cytoplasmic Ca2+ has been shown to trigger anion channels and to precede Cl' efflux (Schroeder, 1995). The C1' efflux across the plasma membrane was activated by intracellular free Ca2+ in Chara cells (Mimura and Shimmen, 1994; Johannes et al., 1998). A rapid light-induced PIFR of Cl' after 2-3 min of illumination was detected and explained as a result of an increase in C1‘ efflux (See Chapter I ). This time allows for Ca” activation of anion channels and a change in the directionality of the anion flux. Three facts support the role of Ca” in regulating Cl‘ efflux: (a) LaCl, and EGTA significantly blocked the light-induced PIFR of Cl’ (Table 2); (b) the addition of Cay-ionophore A23187 (Figure 7) or (0) increasing Ca2+ concentration in the bathing medium resulted in a significant increase in the light-induced PIF R of Cl‘ (Figure 8). Nevertheless, the light-induced Ca2+ influx was not influenced by C1” channel blockers (Table 4). Both LaCl3 and EGTA dramatically inhibited the light-induced initial response in net H” influx (Table 2). F elle (1998) hypothesized that cation channels may contribute to the 111 regulation of the apoplastic pH in roots. He also observed that extracellular Ca2+ was apparently less effective in influencing apoplastic pH. Using the Ca2+~ionophore A23187 as a Cay/H+ exchanger did not significantly affect the light-induced Hi influx (Figure 7). This result is different from that of LaCl3 and EGTA treatments. A possible explanation is that proton pumps are deactivated when cation channels are blocked. Kinoshita et al. (1995) reported that the activation of IF-ATPase required elevated cytoplasmic Ca2+ concentration in guard cells. Under the light, perhaps the interaction of Ca2+ and H+ (and other ions, such as K, Cl’) with Hi-ATPase in the plasma membrane stimulates its phosphorylation/dephosphorylation, as occurs under stresses (Netting, 2000). No work seems to have been done on the possible mechanisms for such a light-induced response. The present result might constitute a linkage between Call-channels or Ca2+ messenger and Hi flux across the plasma membrane during light signal transduction. All considerations taken together, the simplest interpretation is that light-induced transient Ca2+ influx facilitated by Ca2+ channels could initiate other ionic responses in pea mesophyll. It is interesting to note that external Ca2+ concentrations influenced ion flux responses to light. The first effect of apoplastic Ca2+ is the light-induced initial response in net Ca2+ influx which was significantly increased by higher Ca2+ concentration (2.0 mM), whereas the initial flux responses in other ions were not significantly influenced by changing extracellular Ca2+ concentrations (Figure 8) (K+ and Cl' data not shown). However, the light-induced PIFR of K and CT was significantly influenced by apoplastic Ca2+ concentrations (Figure 8). These results suggest that apoplastic K+ also play an important role in regulating ionic responses to light. 112 It is remarkable that light immediately stimulated a rapid PIFR of Ca2+ following a transient Ca2+ influx (see Chapter 1,). Perhaps the PIFR is associated with Ca” efflux and reflects light-regulation of cytosolic Ca2+ homeostasis. When cytosolic Ca2+ concentration increases, active Ca2+~transport systems work to restore the resting cytoplasmic Ca2+ to submicromolar levels (Trewavas, 1999). Vanadate is not completely specific for the plasma membrane Hl-ATPase inhibition. The plasma membrane Ca2+- ATPase also is inhibited by vanadate (Palmgren and Harper, 1999). Thus, it is not surprising that vanadate and even DCCD significantly reduced the light—induced PIF R of Ca2+ (Table l) and enhanced the light-induced maximum increase in net Ca2+ influx (data not shown). On the other hand, Ca2+ channel blocker and EGTA also reduced the light- induced PIFR of Ca2+ (Table 2). Therefore, the decrease in Ca2+ influx is not a contributor to this process. Ca2+ efflux across the plasma membrane mediated by Ca”- ATPase in the plasma membrane would be expected to be a main pathway for the light- induced PIF R of Ca2+. Cl' fluxes and Cl‘ transporters Elzenga and Van Volkenburgh (1997) suggested that light-induced transient depolarization across the plasma membrane of pea mesophyll cells is at least due in part to an increased plasma membrane conductance to anions. They found that one of the anion channels identified by the patch clamp technique is characterized with a single- channel conductance of 32 picasiemens as a light-induced candidate. What is the role of Cl' efflux and anion channels? Anion channels have higher selectivity for C1' than for malate (Hedrich, 1994; Schroeder, 1995). In plant cells, anion channels regulate Cl' 113 efflux from cytoplasm into the extracellular space, driven by the anion-gradient across the plasma membrane and negative membrane potentials. In two types of anion channels, the S-type anion channels would produce a long-term and sustained depolarization, whereas the R-type anion channels would be involved in short and transient depolarization. The voltage-dependence and kinetics of R-type anion channels are also reminiscent of voltage-dependent Ca2+ and K+ channels in excitable membranes (Bisson, 1984). The data presented showed that the two anion channel inhibitors displayed different effects on the light-induced change in net Cl' flux. The S-type inhibitor, A-9-C, did not affect the light-induced PIFR of Cl' (an increase in net Cl' efflux), which was eliminated by the R- type blocker DIDS (Figure 8). A-9-C did not affect the light response in net influx of other ions (H1, Ca2+ and Kl). By comparison, DIDS significantly reduced the influx of H+ and Ca2+ influx during the first 5 min of illumination (Table 4). Thus, R-type, and not S- type anion channels, appear to mediate Cl' efflux during the light response. Most previous evidence supports the model that Cl’ efflux catalyzed by the opening of anion channels is a component of light-induced depolarization (Spalding and Cosgrove, 1992; Trebacz et al., 1994; Elzenga and Van Volkenburgh, 1997). However, Cl' flux was not measured in their studies. The CI' ion can move inwardly or outwardly across the plasma membrane. Indeed, light yielded an initial and short increase in net Cl' influx; in other words, Cl‘ influx exceeds Cl' efflux during the initial light response (see Chapter I). On the other hand, this result provides some evidence that‘ H+/Cl' symport might be a pathway for the light-induced influx of both Cl' and H1 as discussed in Chapter I. Fusicoccin (FC) significantly increased the light-induced initial response in net Cl' influx (Table l) which also was enhanced in acidic bathing medium (see Chapter 1). Thus, Cl' 114 efflux mediated by anion channels might be involved, but is probably not the initial depolarizing component. Calcium-dependent activation of anion channels seems to be one of its main pathways in cellular signal transduction. Conversely, Cl' efflux in response to activation of anion channels produces further depolarization, which, in turn, leads to rapid signal amplification as additional anion channels respond to increased depolarization (Ward and Schroeder, 1997). Ward and Schroeder (1997) hypothesized that anion efflux induced by cellular stimulation results in membrane depolarization which in turn triggers the activation of voltage-dependent Ca2+~channels that mediate Ca2+ influx. In this study, low Cl' concentration (0.2 mM) in the bathing solution significantly limited the light-induced response in net Ca2+ influx. At the same low concentration of exogenous Cl’, the light- induced initial response in net Cl' influx significantly decreased (Table 5b), suggesting a possible relationship between Ca2+ influx and Cl' fluxes or concentrations during the light response. Elzenga et a1. (1995) showed that light-induced depolarization of pea mesophyll was eliminated by high Cl‘ concentrations (about 10 mM, but they did not test Cl' concentration below 1.0 mM). High external KCl concentrations, 10 mM, greatly increased the light-induced response in net Cl‘ and K+ influx but affected Hi influx much less (Table 5b). This provides an explanation for their results. However, the possibility that high ion concentrations in the bathing medium may generate large variable ion flux changes cannot be excluded. K+ fluxes and K’r transporters As shown previously, the first K+ response to light is a rapid increase in net K” influx 115 across the plasma membrane of pea mesophyll tissues (Figure 5). Uptake of K+ across the plasma membrane is dominated by three possible pathways. One of these is K+ influx through a K+ inward rectifier. When a K1 inward channel blocker (CsCl) was added to the bathing solution, the light-induced initial response in net K+ influx was reduced slightly, but the maximum increase in net K+ influx was inhibited later. Seemingly, whether the K+ inwardly rectifier is involved in the initial light response in net K“ influx is still unclear. Another candidate is the non-selective cation channels which are permeable for both Ca2+ and K“. Both LaCl3 and EGTA resulted in significant reductions in the light-induced net K+ influx response (Table 2), which were also dependent on external K+ concentration (Table 5a). The final possibility for K“ influx is active K“~H+ symport. The proton motive force (pH gradient) generated by the H“ pump can energetically drive K+ uptake through a carrier (Hedrich and Schroeder, 1989; Netting , 2000). Two facts support the view that light increases K‘i-H+ symport across the plasma membrane of pea mesophyll. First, changing apoplastic pH caused a significant change in the light—induced initial response in net K+ influx (See Chapter I). Acidic external bathing solution increased the K‘ influx. Second, the plasma membrane Hl-ATPase inhibitor, Na,VO,, decreased the light-induced response in net K+ influx (Table l). Roelfsema and Prins (1998) reported that light stimulated proton extrusion and resulted in apoplastic acidification enhancing the conductance of K+ inward rectifying channels. Hedrich and Schroeder (1989) tried to compare the H+~K+ symporter with the Kl~inward rectifier. However, the flux measurement of the present experiment could not distinguish between these alternative pathways. To complicate things further, both K+ channel blockers also significantly reduced the light-induced initial response of not H* influx 116 (Table 3), seemingly supporting the existence of a H+~K+ symporter. As discussed previously, the light-induced K+ influx is not a main depolarizing ionic event. Considering light-induced transient apoplastic alkalinization and the CsCl effect'on the initial response of net K+ influx, the light-induced H+~K+ symport is proposed as an initial pathway for K+ influx. Unlike Ca2+ and Cl', light only elicited a slow PIFR of K+ following the initial response in net K+ influx (see Chapter I). This PIFR may be due to a decrease of K+ influx or an increase of K+ efflux mediated by K“ outward rectifiers. In pea mesophyll, the K+ channel blocker TEA eliminated most the light-induced PIFR of K” (Table 3), indicating that outwardly-directed K+ channels are a main pathway for K+ efflux and contribute to the PIFR. However, the PIFR of K+ probably counteracts Cl’ efflux and thus results in a membrane repolarization. Cay-activated K+ channels were found in higher plants or green algae (Elzenga and Van Volkenburgh, 1993; Johannes et al., 1997; Bauer etal., 1998). Exogenous LaCl3 and EGTA or low external Ca2+ concentration (0.1 mM) reduced the light-induced PIFR of K+ (Table 2, Figure 8). However, high external Ca2+ concentration (2.0 mM) or addition of Ca2+~ionophore did not affect the PIP R of K (Figure 7, 8). A possible reason is related to the decrease in net K+ influx when inward K+ rectifiers were deactivated by elevated cytosolic Ca” concentrations (Lemtiri-Chlieh and MacRobbie, 1994; Blatting, 1999). In general, the Ki outward rectifiers are activated by membrane depolarization and up-regulated by cytoplasmic pH increase, but are virtually insensitive to cytosolic Ca2+ (Lemtiri-Chlieh and MacRobbie, 1994; Grabov and Blatting, 1997). Keunecke and Hansen (2000) found a different pH-dependence of K+ channel activity between bundle sheath and mesophyll cells of maize leaves. 117 Alternatively, as in guard cells (Allan et al., 1994), Cay-independent pathways are assumed to also exist during the light-induced PIFR of K in pea mesophyll. This process may be associated with the regulation of H+ homeostasis and membrane potentials, which also involve other ions. The regulation of K1 channel gating in guard cells may be accomplished by G proteins and protein phosphorylation (Thiel and Wolf, 1997). K+ fluxes in mesophyll do not have the clearly defined specialized functions that are obvious in guard cells (Assmann, 1996). The regulation of K” channel activation in mesophyll cells still needs to be investigated. The plasma membrane H+-ATPase is not directly involved in light-induced H+ influx and apoplastic alkalinization Transient apoplastic alkalinization in mesophyll tissues has been demonstrated (Elzenga et al., 1995; Shabala and Newman, 1999; or see Chapter I). Stahlberg and Van Volkenburgh (1999) proposed that the plasma membrane Hl-ATPase does not directly take part in the light-induced membrane depolarization and alkalinization of pea mesophyll cells. However, previous results indicated that the plasma membrane H‘- ATPase activity increased following illumination (Marré et al., 1989; Linnemeryer et al., 1990; Remis et al., 1994). The resulting apoplastic acidification is thought to be coupled with H” efflux from the leaf cells. As proposed previously, CO2 uptake by mesophyll cells and H+~K+ or Hl-Cl' symport are possible reasons for the alkalinization (see Chapter I). In the present study, light induced an initial increase in extracellular pH (data not shown) and in net H+ influx both of which were inhibited by H+~ATPase inhibitors (Figure 2). Why did the H+~ATPase inhibitor also reduce alkalinization? Possibly 118 because the inhibition of the H+ pump by vanadate or DCCD decreased H+ extrusion from the cytoplasm thereby reducing the proton motive force, the H+ gradient across the plasma membrane. The resulting decrease in the light-induced net K+ influx could be related to the blockage of H+-ATPase (Table 1). Considering the consistent change of H+ flux and extracellular pH, it is assumed that the plasma membrane H+~ATPase may contribute to the light-controlled apoplastic alkalinization. If light activates the plasma membrane H+~ATPase and Hi extrusion, or if H+~ATPase is a primary channel for H1 transport, the net flux rate should be shifted to efflux. In fact, however, a transient increase in H+ influx occurred after the onset of light stimulation, indicating that more H+ moves inwardly or H+ efflux was inhibited at least. Even though in the presence of FC, which hyperpolarizes the plasma membrane by activating Hi extrusion and simultaneously enhancing K+ influx (Ullrich and Guem, 1990), light still elicited a rapid increase in net H+ influx (Figure 2, Table 1). Remarkably, H+ influx is larger than its efflux; in other words, at least Hl-ATPase does not function as a direct agent in the light- induced initial response. Therefore, the present results support the conclusion of Stahlberg and Van Volkenburgh (1999). The fimction of Hl-ATPase may involve feedback regulation of ion flux. Cytoplasmic pH homeostasis is attributed to the regulation of H+ pumps and other ionic transporters, including K+ channels and anion channels such as in guard cells (Grabov and Blatting, 1998). Stahlberg and Van Volkenburgh (1999) observed a long period of apoplastic acidification immediately following light-induced transient alkalinization of pea mesophyll, whereas vanadate (0.5 mM) eliminated this acidification but did not influence the light-induced transient depolarization. Light induced a PIFR of Hi after a 119 rapid initial increase in net H+ influx (Figure 2). An increase in H+ efflux or a decrease in H“ influx resulted in this process, which may contribute possibly to extracellular acidification. Treatment with F C significantly increased the PIFR of H, which was eliminated by vanadate or DCCD (Table 1), suggesting that the H+ pump could be involved in this response by facilitating H+ efflux. Proton signalling has been proposed as a second messenger in guard cells (Leckie et al., 1998). Alternatively, it is possible that H+ flux or pH, another candidates as signals, may also be involved in light responses in pea mesophyll through a Ca2+~independent pathway. Dynamic cytoplasmic H+ changes may link Caz‘ldependent pathways, or share some common signalling components in cellular signal transduction. The role of the H* messenger remains to be investigated. A model: ionic pathways for light response Figure 9 presents a schematic model integrating the observed and hypothetical events in light-induced ionic signal transduction in pea mesophyll. Upon activation of putative photoreceptors, Ca2+ influx through non-selective cation channels would play an initial role in transmitting the light signal and depolarizing the membrane. K“ and Cl' uptake coupled with H+ influx contribute to apoplastic akalinization. Light-induced initiation of Ca2+ channels drives several differential secondary signalling elements, e.g., intracellular Ca2+ and pH, K”~ and anion-channels. A common feature for all ions measured is that light induced PIFR of ion fluxes from pea mesophyll. The processes are associated with the activities of ionic transporters (e.g., Cay-ATPase, H+~ATPase, K+ channels and anion Channels) in the plasma membrane or intracellular membrane systems, which contribute to membrane repolarization. 120 However, this research does not exclude the possibility of Ca2+~independent pathways. Proton signalling coupled with ion flux changes in cellular signal transduction challenges the hypothesis above. Identification of some specific ion transporters, e.g., inward or outward K+ rectifiers, anion channels, Ca2+~ATPase, can allow discrimination among specific pathways that are presumably involved in the rapid light response. Direct measurement of dynamic changes of cytosolic ion concentration will provide more valuable evidence. Further investigation will need to link the observed early events to downstream steps, such as protein phosphorylation/dephosphorylation, Ca2+lcahnodulin dependent—kinases and gene expression involved in photoreception. 121 apoplast plasma membrane ADP + Pi H-t- « H+ cytoplasm \ H+~ATPase ATP ' ’ + '1' repolarization K ——-.—-—- K /[pH] .mm' 4. anion channel ‘1' (332+ Ca2+ v Ca -- ’ase - depolarization K+ K+ 111+ H: [Ca2+] ) c1- ‘ Cl' alkalinization V mm mm \ (332+ arm; C a” Ca" channell j [intracellular stores] ,. o 0'- ,. a a '0 ,0," '- Light": — >C photorecepto ' H+< H’~ATPase Figure 9. A hypothetical model for light-induced ion flux response in pea mesophyll cells 122 LITERAURE CITED Allan AC, Ficker MD, Ward JL, Beale MH, Trewavas. 1994. Two transduction pathways mediate rapid effects of abscisic acid in Commelina guard cells. Plant cell 6:1319-1328. Assmann SM. 1996. G—protein regulation of plant K+ channels. In DPS Verama, ed, Signal transduction in plant grth and development. Springer Wien New York, pp39~ 61. Assmann SM, Shmazaki K. 1999. The multisensory guard cell: Stomatal response to blue light and abscisic acid. Plant Physiol 119:809-815. Bach M, Schnitzler JP, Seitz HU. 1993. Elicitor~induced changes in Ca2+ influx, K“ efflux, and 4-hydroxybenzoic acid synthesis in protoplasts of Dcucus carats L. Plant Physiol 103:407-412 Bauer CS, Plieth C, Hansen UP, Simonis W, Schdnknecht G. 1998. A steep Ca2+- dependence of a K+ channel in a unicellular green alga. J Exp Bot 49:1761-1765. Bentrup F W. 1990. Potassium ion channels in the plasmalemma. Physiol Plant 79:705- 711. Bisson MA. 1984. Calcium effects on electrogenic pump and passive permeability of the plasma membrane of Chara corallina. J Membr Biol 81:59-67. Blatt MR. 1999. Reassessing roles for Ca2+ in guard cell signalling. J Exp Bot 50:989- 999. Blom-Zandstra M, Koot HTM, Van Hattum J, Van Vogalzang SA. 1997. Transient light-induced changes in ion channel and proton pump activities in the plasma membrane of tobacco mesophyll protoplasts. J Exp Bot 48:1623-1630. Coté GG. 1995. Signal transduction in leaf movement. Plant Physiol 1092729~734 Elzenga JTM, Prins HBA, Van Volkenburgh E. 1995. Light-induced membrane potential changes of epidermal and mesophyll cells in growing leaves of Pisum sativum. Planta l97:127~134. Elzenga JTM, Van Volkenburgh E. 1993. Ion channels in the plasma membrane of epidermal and mesophyll cells of growing pea leaves. Plant Physiol 102:S106. Elzenga JTM, Van Volkenburgh E. 1997. Characterization of a light-controlled anion channel in the plasma membrane of mesophyll cells of pea. Plant Physiol 113:1419- 1426. 123 Felle H. 1998. The apoplasic pH of the Zea mays root cortex as measured with pH- sensitive microelectrodes: aspects of regulation. J Exp Bot 49:987-995. Grabov A, Blatt MR. 1997. Parallel control of the inward-rectifier Kl channel by cytosolic free Ca2+ and pH in Vicia guard cells. Planta 201 :84-95. Grabov A, Blatt MR. 1998. Co-ordination of signalling elements in guard cell ion channels control. J Exp Bot 49:351-360. Hedrich R. 1994. Voltage-dependent chloride channels in plant cells: identification, characterization, and regulation of a guard cell anion channel. Curr Topics in Membrane 42:1-33. Hedrich R, Schroeder JI. 1989. The physiology of ion channels and electrogenic pump in higher plants. Annu Rev Plant Physiol Plant Mol Biol 40:539-569. Johannes E, Ermolayeva E, Sanders D. 1997. Red-light-induced membrane potential transients in the moss Physcomitrella patens: ion channels interaction in phytochrome signalling. J Exp Bot 48:599-608. Johannes E, Crofts A, Sanders D. 1998. Control of Cl‘ efflux in Chara corallins by cytosolic pH, free Ca2+, and phosphorylation indicates a role of plasma membrane anion channels in cytoplasmic pH regulation. Plant Physiol 118:173-181 Johnson CM, Stout PR, Broyer TC, Carton AB. 1957. Comparative chlorine requirements of different plant species. Plant and Soil 8:337-353. Keunecke M, Hansen UP. 2000. Different pH-dependences of K+ channel activity in bundle sheath and mesophyll cells of maize leaves. Planta 210:792-800. Kinoshita T, Nishimura M, Shimazaki K. 1995. Cytosolic concentration of Ca2+ regulates the plasma membrane H+~ATPase in guard cell of fava bean. Plant Cell 7:1333- 1342. Leckie CP, McAinsh MR, Montgomery L, Priestley AJ, Staxen I, Webb AAR, Hetherington AM. 1998. Second messengers in guard cells. J Exp Bot 49:339-349. Lemliri-Chlieh F, MacRobbie EAC. 1994. Role of calcium in the modulation of Vicia guard cell potassium channels by abscisic acid: a patch clamp study. J Membr Biol 137 299-107. Linnemeyer PA, Van Volkenburgh E, Cleland RE. 1990. Characterization and effect of light on the plasma membrane H-ATPase of bean leaves. Plant Physiol 94:1671-1676. Long JM, Widders IE. 1990. Quantification of apoplastic potassium content by elution 124 analysis of leaf lamina tissue from pea (Pisum sativum L. cv. Argenteum). Plant Physiol 94:1040-1047. Marré E. 1979. Fusicoccin: a tool in plant physiology. Annu Rev Plant Physiol 30:273- 288. Marré MT, Albergoni F G, Moroni A, Marré E. 1989. Light-induced activation of electrogenic H” extrusion and K“ uptake in Elodea densa depends on photosynthesis and is mediated by the plasma membrane Hl-ATPase. J Exp Bot 40:343-3 52 Mayer WE, Hohloch C, Kalkuhl A. 1997. Extensor protoplasts of the Phaseolus pulvinus: light-induced swelling may require extracellular Ca2+ influx, dark-induced shrinking inositol 1,4,5-t1iphosphate-induced Ca2+ mobilization. J Exp Bot 307:219-228. Mimura T, Shimmen T. 1994. Characterization of the Ca2+-dependent Cl' efflux in perfused Chara cells. Plant Cell Physiol 35:793-800. Moran N, Fox D, Satter RL. 1990. Interaction of the depolarization-activated K+ channel of Samanea saman with inorganic ions: a patch-clamp study. Plant Physiol 94:424-431. Netting AG. 2000. pH, abscisic acid and the integration of metabolism in plants under stressed and non-stressed conditions: cellular response to stress and their implication for plant water relations. J Exp Bot 51:147-158. Pahngren MG. 1998. Proton gradients and plant growth: role of the plasma membrane H+~ATPase. Adv Bot Res 28:2-70. Palrngren MG, Harper MG. 1999. Pumping with plant P-type ATPases. J Exp Bot 50:883-893. Pressman BC. 1976. Biological applications of ionophores. Annu Rev Biochem. 45:501-530. Reid R], Smith FA. 1992. Regulation of calcium influx in Chara: effects of K1, pH, metabolic inhibition, and calcium channel blockers. Plant Physiol 100:637~643. Remis D, Treffny B, Girnmler H. 1994. Light-induced H+ transport across the plasma membrane of the acid-resistant green alga Dunaliella acidophila. Plant Physiol Biochem 32:75-84. Roelfsema MRG, Prins HRA. 1998. The membrane potential of Arabidopsis thaliana guard cells: depolarizations induced by apoplastic acidification. Planta 205: 100-1 12. Ryan PR, Newman IA, Arif I. 1992. Rapid calcium exchange for proton and potassium in cell walls of Chara. Plant Cell Environ 15:675-683. 125 Schroeder II. 1995. Anion channels as central mechanisms for signal transduction in guard cells and putative functions in roots for plant-soil interactions. Plant Mol Biol 28:353-361. Schroeder JI, Hedrich R. 1989. Involvement of ion channels and active transport in osmoregulation and signalling of higher plant cells. Trends Biochem Sci 14:187-192. Shabala NS, Newman IA, Morris J. 1997. Oscillations in H“ and Ca2+ fluxes around the elongation region of corn and effects of external pH. Plant Physiol 113:111-118. Shabala NS, Newman IA. 1999. Light-induced changes in hydrogen, calcium, potassium, and chloride ion fluxes and concentration from the mesophyll and epidermal tissues of bean leaves. Understanding the ionic basis of light-induced bioelectrogensis. Plant Physiol 119:1115-1124. Spalding EP, Cosgrove DJ. 1992. Mechanism of blue-1ight-induced plasma membrane depolarization in etiolated cucumber hypocotyls. Planta 188: l99~205. Spalding EP, Slayman CL, Goldsmith MHM, Gradmann D, Bertl A. 1992. Ion channels in Arabidopsis plasma membrane. Transport characteristics and involvement in light- induced voltage changes. Plant Physiol 99:96-102. Stahlberg R, Van Volkenburgh E. 1999. The effect of light on membrane potential, apoplastic pH and cell expansion in leaves of Pisum sativum L. var. Argenteum. Planta 208:188-195. Thiel G, Wolf A H. 1997. Operation of K+~channels in stomatal movement. Trends Plant Sci 21339-343. Trebacz K, Simonis W, Schonknecht G. 1994. Cytoplasmic Ca2+ Kl, Cl', and NO,’ activities in the liverwort Conocephalum conicum L. at rest and during action potential. Plant Physiol 106:1073-1084. Trewavas AJ. 1999. Le calcium, C’est la vie: calcium makes waves. Plant Physiol 120:1-6. Tsien RW, Hess E, McCleskey EW, Rosenberg RL. 1987. Calcium channels: mechanism of selectivity, permeation and block. Annu Rev Biophys Chem 16:265-290. Ulhich CI, Guem J. 1990. Ion fluxes and pH changes induced by trans-plasmalemma electron transfer and fusicoccin in Lemna gibba G1. Planta l80:390~399. Ward JM, Schroeder JI. 1997. Roles of ion channels in initiation of signal transduction in higher plants. In Signal transduction in plants. P Aducci, ed, Brikhauser Verlag Basel, 126 Switzerland, pp 1-22. White PJ. 1993. Characterization of a high-conductance, voltage-dependent cation channel from the plasma membrane of rye roots in planer lipid bilayers. Planta 1912541- 551. White PJ. 1994. Characterization of a voltage-dependent cation channel from the plasma membrane of rye (Secale carele L.) roots in planer lipid bilayers. Planta 191 :186-193. White PJ. 1998. Calcium channels in the plasma membrane of root cells. Ann Bot 81:173-183. 127 Chapter III Effects of light intensity and quality on ion flux from pea mesophyll tissues. 128 ABSTRACT Mesophyll tissues from excised leaflet of pea (Pisum sativum L. cv. Argenteum) were exposed to different light intensity treatments. Five fluence rates (225, 450, 900, 1800 and 2800 pmol m"2 s") were used to test fluence-response of ion flux. Light-induced changes in net Ca2+, H, K“ and Cl' flux increased in a saturable manner with increasing fluence. However, high fluence increased the temperature of the bathing solution, which may have caues a heat effect on ion flux. White light (WL) of 450 pmol m'2 s'1 stimulated significant changes in net flux of net Ca2+, Hi, K+ and Cl' from mesophyll tissues. When leaf tissues were exposed to the blue-light (BL) and red-light (RL) of similar intensity, the change in net Ca2+, IT, K+ and Cl‘ flux exhibited a similar pattern to WL. At an intensity of 225 nmol m'2 s", the RL-induced initial increase in net Kl influx was more rapid than under BL. However, there were no significant differences in the magnitude of change in net Ca2+, Hi, Ki and Cl' fluxes between BL and RL. The BL~ and RL-induced flux responses of all ions were saturated at intensities higher than 16 pmol m‘ 2 s‘1 (BL) and 75 pmol m'2 5’1 (RL), respectively. The result suggests that BL and RL play a similar role in rapid ionic responses to light. INTRODUCTION Light provides plants with many different kinds of information, as a result of variations in light fluence, quantity, direction and duration. Plants sense light signals through a complex sensory system consisting of different photoreceptors, which detect different light qualities over a wide spectral range, but with overlapping action spectra (Batschauer, 1998). Light energy is then used either directly or indirectly to optimize a 129 variety of physiological and morphological processes. Recently much attention has focused on photosensory mechanisms in plants and several light signalling pathways have been proposed (Chamovitz and Deng, 1996; Batschauer, 1998; Khurana et al., 1998). The most important aspect of understanding the regulation of photosensory mechanisms is to elucidate the function of photoreceptors. At least three classes of photosensory pigments (photoreceptors) have been implicated in the perception of light signals in plants, including phytochromes, blue/UV-A photoreceptors, and UV-B photoreceptors. The multiple photoreceptor classes imply the presence of a highly responsive, intricate, and yet finely tuned, cellular signal transduction machinery (Batschauer, 1998). However, the chemical structure and function of receptors are not fully understood. Early events in the transduction of light-mediated signals in plants are often associated with transient changes of membrane potential. This bioelectrogensis reflects changes in ion fluxes across the plasma membrane. Light-induced rapid ion flux changes have been demonstrated in many plant species, such as Dunalliella acidophila (Remis et al., 1994), Physcomitrella patens (Johannes et al., 1997), and Vicia fava (Shabala and Newman, 1999). Both red and blue lights have been reported to elicit electrical signalling or ionic flux changes, especially for guard cells (Serrano and Zeiger, 1989; Assmann and Shimazaki, 1999). Red light matching the absorption spectrum of chlorophyll, stimulates an outward electrical current that may be canied by the plasma membrane Hi-ATPase (Serrano et al., 1988). Blue light given in a single pulse stimulates stomatal opening (Zeiger et al., 1987) and induces proton extrusion from guard cell protoplasts (Assmann et al., 1985; Shimazaki et al., 1993). In addition, since an anion channel in Arabidopsis hypocotyl tissue was shown to be activated by BL (Cho and Spalding, 1996), it is likely 130 that BL would affect similar anion channels in guard cells. Shacklock et al. (1992) demonstrated that RL induces protoplast swelling and causes a transient Ca2+ increase followed by a rapid reduction in cytoplasmic Ca2+ concentrations to below resting levels. The BL-induced depolarization in etiolated cucumber hypocotyls involves the inhibition of a proton-pumping ATPase and an influx of Ca2+ (Spalding and Cosgrove, 1992). However, the reports on pulvinar movements are complicated by the fact that phytochromes, blue-light photoreceptors, and underlying circadian rhythms all play roles in regulating turgor changes (Briggs and Liscum, 1997). Recently, Suh et al. (2000) characterized BL activated K“ efflux channels in flexor cells fiom Samanea saman. Lew et a1. (1990) also had reported phytochrome-activated K+ channels in Mougeotis protoplasts. Clearly, light elicits transient electrical signalling which is mediated by ion fluxes across the plasma membranes of plant cells. However, the identity of the light quality which plays a main role in regulating ion flux remains to be investigated. In previous reports, the light-induced rapid flux changes of Ca2+, IT, K+ and Cl' had been demonstrated in pea mesophyll. Some specific ion transporters were proposed to be involved in the light~induced ionic flux responses (See Chapter II). Based upon the differential reports above, we hypothesize that either BL or RL, or both, may be involved in the regulation of light—induced ion flux responses. In addition, light intensity would be expected to influence ion flux changes in pea mesophyll cells. To test our hypotheses, BL and RL, were used to evaluate the effects of light quality on ion fluxes. Different fluence treatments were also imposed to determine ionic responses. 131 MATERIALS AND METHODS Plant materials Seeds of pea (Pisum sativum L. cv. Argenteum) were germinated for 4-5 d. Plants then were transferred to a growth chamber with a 16 h photoperiod, day/night temperatures of 20/25 °C, 75% relative humidity, and a fluence of 200 pmol m'2 s" at the top of the canopy. The pea seedlings were cultured using modified Hoagland’s solution (Johnson et al., 1957), gradually increasing the concentration from 0 to 1A to ‘/2 strength at 4-5 d intervals. Flux measurements The ion fluxes and extracellular ion concentrations at the surface of the leaf mesophyll were measured using the MIFE system. Technical details of this system are similar to that generally as described in Chapter I and by Shabala et al. (1997). From time course data obtained by the MIFE system, the directionality and rates of ion flux could be estimated. Additional information on the MIFE is available at http://www.phys.utas.edu.au/ physics/biophys. In brief, micropipettes with a tip diameter of 2 - 3 urn were pulled fi'om 1.5 mm-diameter borosilicate glass capillaries (catalog no. MIB ISO-6; World Precision Instruments, Inc., Sarasota, FL) with a horizontal electrode puller (PC-10, Narishige, Tokyo, Japan). After silanization, cooled microelectrodes were back-filled with 500 mM CaCl2 for Ca“, 500 mM KCI for K” and Cl‘, and 15 mM NaCl plus 40 mM KHZPO, (adjusted to pH 6.0 with NaOH) for H“. Electrode tips were then front-filled with a liquid-membrane-type of the appropriate ion-selective sensor, Hi (no. 95291), Ca2+ (no. 21048), K” (no. 60031) and Cl' (no. 24902) cocktails (all from Fluka, 132 Milwaukee, WI), until a 50- to 80- um long column of ionophore was produced. Microelectrodes were calibrated for each ion using a known set of standards before and after use. The Ag/AgCl reference electrode was prepared using a plastic tube containing 1000 mM KCl in 2% agar. After calibration, the ion-selective electrodes were slowly vibrated by a micromanipulator which provided three-dimensional positioning. The microelectrode/ micromanipulator set-up was mounted on the stage of an inverted microscope (Leitz, Wetzlar, Germany). A measuring chamber with leaf sample was placed on a holder that was connected to a three-way hydraulic micromanipulator (WR 88, Narishige, Tokyo, Japan) driven by a computer-controlled stepper motor (MO6l-CE08, Superior Electric, Bristol, CT). The MIFE system allows measurement of three ions simultaneously. Three microelectrodes can be vibrated at any angle in a two-dimensional plane by separately controlling the amplitude of the slow vibration of the micromanipulator. The MIFE system measured the voltage differences at the two extreme positions of vibration, which provided the basis for estimation of both the rate and directionality of net ion fluxes. In this study the electrodes were moved in a 20-s square wave cycle between 50 to 90 um above the leaf surface at a low fiequency (0.1 Hz). Ion concentration of ions was calculated from the voltage potentials at each position using the calculation curve for the specific microelectrode. The net flux for each specific ion was estimated fiom the measurements of the difference in the electrochemical potential between these positions (Shabala et al., 1997). The first 4 s of each half cycle (10 s) were not used for concentration and flux estimation. This was considered to be the minimum delay time necessary to obtain a stable measurement following movement of the 133 microelectrodes. Instability was thought to be caused by solution turbulence and physical vibrational noise following movement. Light sources Light sources were provided by a fiber-optic light projector (CUDA Products Corporation, Jacksonville, FL) with a 250 W halogen bulb (GS Edmumd Scientific Company, Barrington, NJ). A lens (no L~16-16, World Precision Instruments, Inc., Sarasota, FL) was fastened on the top of a light guide to focus the beam. The blue light (BL) and red light (RL) treatments were achieved by a blue glass filter (5-60, thickness 4.5 mm, Kopp Glass Inc., Pittsburgh, PA) and a red glass filter (2-61, thickness 4.5 mm, Kopp Glass Inc., Pittsburgh, PA), respectively, which were placed in front of the optics in the lens. The spectra of white-, blue- and red-light sources were determined using Li- 1800 spectroradiometer (LI-COR, Lincoln, NE). The white light delivered a broad spectra between 400-1000 nm (Figure 1A). However, most of light lies between 500-800 nm. The long wavelength (above 800 nm) fluence was decreased by 45-47 % after filtering with distilled water (Figure 1). The blue source emitted light of wavelengths between 375-525 nm with peak emission at around 450 nm (Figure l). The output of the red source included light of wavelengths between 610-1000 nm. The peak of light emission was between 650-740 nm (Figure 1). The blue— and red-light fluence rates were measured using a quantum sensor (model LI-189, LI-COR, Lincoln, NE). In this study, White-light (WL) of 450 pmol m’2 s“ was used as light treatment. When the WL was filtered by either the blue- or red-glass, the BL of 32 pmol m'2 s'1 and the RL of 225 pmol m'2 s'1 were generated, respectively. 134 —L N I + - water -0- + water —L O O) ITITITI Fluence (pmol m'2 5'1) A White~light . 2 0 ’1 4 1 1 1 1 L l 1 . 1 . 1 . 1 . 1‘ 0.8 — Blue-light — +7, - 1 0'1 1 E .1 '5 _ E 3 8 _ c a) 2 u. _ “L 1 1 . 1 . 1 . 1 . 1 . 1 . 1 . 1‘ i Red-light ‘ FA 3 F- — 'm (y .- E 5 2 — _ E i . 3 8 1 — - u. 0 IIIIIIIIIIOOOJ _ . 1 . 1 . 1 . 1 . 1 A 1 . 1 n 1 300 400 500 500 700 800 900 1 000 1 1 00 Figure 1. Fluence of white, blue and red lights over a range of wavelengths. Wavelength (nm) 135 Experimental protocol Young expanding leaflets were harvested and a leaf section with limited veinal tissue of approximately 15 x 10 mm (LxW) was surgically cut from the lamina. A small strip of epidermal tissue was peeled from the abaxial surface of the lamina of each section so as to avoid incurring mechanical damage to the underlying veinal and mesophyll tissue (Long and Widders, 1990). Isolated leaf segments were incubated in a buffered solution (0.5 mM CaSO,, 1.0 mM KCI, 2.5 mM HEPES-HCl, and 10 mM sucrose at pH 5.2) under 60 W rn‘2 illumination (Philips lamp). After 2.0-2.5 h, the cut segment was mounted in a measuring chamber. The measuring chamber contained about 6-7 mL of unbuffered solution or bathing solution (0.5 mM CaSO,, 1.0 mM KCI, pH 5.2), and the segments then were exposed to 50 - 60 min of darlmess before light treatment. All experiments were conducted in unbuffered solution at room temperature (23:1°C). Flux measurement was continued for an average of 30 min from the onset of illumination and then following the onset of darkness for 10-15 min. Solution temperature The temperatures of the solutions during illumination were measured using a thermocouple thermometer (Model 600-2050, Barnant Company, Barrington, IL). To increase the solution temperature, a warmer bathing solution was expelled into the bathing solution (23i1°C) and then repeatedly sucked and expelled with a mL pipette for 5 - 6 times. The period of time during which warmer solution had been added, mixed and equilibrated was ignored when analyzing ionic concentration and flux data. This 136 represented about 3-4 min total. Change of temperature of the bathing solution could be maintained within a range of 1-2 °C during 15-20 min measurement. All solutions were unbuffered (0.5 mM CaSO,, 1.0 mM KCl, pH 5.2). Data statistical analysis Data analysis has been described in Chapter 1. Analysis of variance (ANOVA) of the data was conducted using SAS 8 (SAS Institute Inc, Cary, NC). RESULTS Effects of light intensity on ion fluxes Since not flux is a balance between influx and efflux, a positive net influx response is the result of enhanced influx or reduced efflux of the ion in question. In this study, five fluence rates of White-light (WL) (225, 450, 900, 1800 and 2800 umol m‘2 s") was used to test the effects of light intensity. Under different light intensities, the net fluxes of Ca2+, H, K and Cl' around pea mesophyll exhibited a similar response trend, i.e. a light- induced initial response in net ion influx (more positive net flux) followed by a Post- Initial Flux Response (PIF R) (see Chapter I). The PIF R could theoretically involve either independent changes or concurrent changes in both influx and efflux. Therefore, either a decrease in influx or an increase in efflux or both can elicit a negative PIFR. In general, the light-induced initial responses in net influx of all ions were the smallest at the lowest intensity (225 pmol m'2 s"). Further increase in light intensity did not produce a significant difference in the responses of net H+ and Cl’ influx between intensities from 450 to 2800 umol m'zs" (Figure 2). Net Ca2+ and K+ fluxes exhibited saturation kinetics 137 .wmzwmz €58me mma 959m .6 new +v_ . ( t_s z_w Iowu) adld £0 apmudwv ooom com. com. 00v. com. com. cow- ow- mm- mv. 0v. mm- om- .Am:_m> gamma: 3 Eon ..o xan 5 3855 cm .0 x35 5 .3383 m mmumoGE Eur: mmcoammm xsi _m_=c_-fiom 8039.33: do 333:2 .629 928a 3 $539 $55 8085-29. 9: $286.: x35 :2 E: 5 $220 .58.: 9: do .6th Emucsm 9: E8959 mama _8_tm> .mmfigam: m-m 53> omega cm “command $58 98 comm ooom a wl. H oomw _ ooow _ +N rum «-8 .06.: 853E com _ 1 1 l [— d d d l I +Nm0 4 o ow om omr cow com om 0v ow om ooom cow. ovw- omr- 00?. cm. ow- 0v. 0N- mw. comm ooom com? a _ ii a 000? _ o r T 1 < 11 A! d com a H _ . _ . 4ll . 4‘ E53 833% .«u Axacccomcane .55 .w. J i 4 a a 1 om 00F omr com com com 0 ‘— O N on no .+I do momcoamm. x3: :0 b_mc$c_ £9792; Co 89cm .N 9:9“. ( t-S Z_LU lowu) xnuul )au u! efiueqo 138 in the initial influx response at light intensities above 900 pmol m'2 s". The light-induced PIFRs of W, Ca2+ and K+ were saturated above 450 pmol rn’2 s". In contrast, high light intensities (1800 - 2800 pmol m’2 5") enhanced dramatically the amplitude of PIFR of Cl' (Figure 2). Effects of solution temperature on ion fluxes When the temperature of the bathing solution increased by 1.0-1.5 °C or 1.5-2.5 °C, the net fluxes of Ca2+, Hi, K1 and Cl‘ from pea mesophyll did not change significantly in darkness during 15-20 min of measurement (Figure 3). A 30 min illumination of 225 and 450 pmol m2 5'1 generated an increase in solution temperature of 0.6 °C and 1.3 °C, respectively (Figure 4). This indicates that the heat effect of these two light intensities is not consequential to ion flux. However, when light intensity was further increased to 900, 1800 and 2800 pmol m'2 s", the temperature of the bathing solution was enhanced by 2.6, 3.8 and 5.0 °C with 20 - 30 min, respectively (Figure 4). At 2800 umol m'2 s", in particular, the increase of temperature was quite rapid. Red-light at intensities of 450, 1170 and 1490 pmol rn'2 s'l also heated up the bathing solution by 1.1, 2.1 and 3.2 °C within 30 min, respectively (Figure 4). The highest fluence of red-light (1490 pmol m'2 s’ I) produced a rapid increase in solution temperature. A large flux change for Ca”, Hi, K+ and Cl' was observed when the temperature was increased by 3 - 4 or 4 - 5 °C in darkness within 15-20 min (Figure 3). Therefore, high fluence is expected to cause heat effects on ion fluxes during experiments of 30 min. 139 380%.: men Eat -_0 can +x . 82 9: 0.3 u 0.. E8. 9:2. ...N a or _ —HD 0 .5552... r———O :1“: film: 3*. firm. TTTTT l I :nmu 1 1L1 14 1 i J 1 com. 00 T 00.. com (vs z.m pwu) mg _0 ran u; 96qu (L-s z_m pum) xnu +z2:) ran at 315mm 2:5 9:: .0c 3mm mm; 939an2 E. ”0. Ym u m._. ”on n.m-me u E. ”0.. m. T? u E. ”$3905 mcamcmaEQ E958; down. 05 .0 Soto Emucflm E393. mama .mo_to> $06239 Wm do 3996 cm E893. £58 Ema .woammz 2958..» c..mu L ow m0 .+I do moxac 8: :0 939382 5:28 mEmmmbE _o 90th .m Sign (ks z.m pmu) xnu +3331; ug 3509K) 140 (vs 8.!” mm) xnu +H tau ug afiueqo 29 28 :27 A0 0‘3 ‘99 0 g? 26 5% 32 25 0 85 O 24 23 28 c27 Ac 0'3 “v? 2 26 g? 5% (1.0 Eo25 :25 “5 24 23 -e— 225 pmol m'g s': -O- 450 pmol m' s' White-light 4 "‘ + 900 pmol m'zf'11 ‘ _ +1800pmolm'1s'1 ‘ + 2800 pmol m' s' Light-on ‘ ‘/A/o/"/A r l ‘ ~ 4". o . . . . l l l l J J l “ -e— 225 pmol mi 8': Red-light ‘ L_ -O- 450pmol m' 2'. 1 _ + 1170 pmol m'2 s'1 _ + 1490 pmol m' s' . — -l Light-on W—A/G/a _. l /A _ L— WOW—4 . o . — + l l l A l l l O 5 10 15 20 25 30 Time (min) Figure 4. Effect of white and red lights on temperature of the bathing solution. Data points represent an average of 3 replicates. 141 Effects of blue-light on ion fluxes Like WL, a BL exposure of 32 pmol m‘2 8" induced an initial response in net Hi, Ca2+, K+ and Cl' influxes (Figure 5). This was followed by Post-Initial Flux Responses (PIFRs) of Hi, Ca2+, K+ and C1. The net fluxes of all ions decreased to a net efflux (negative value) or returned to the original value before the BL treatment when BL was turned off (Figure 5). Compared with the WL of 450 pmol m'2 s", the magnitude of BL-induced initial response in net influx of H+ and Ca2+ was significantly smaller (Table 1). The BL- induced changes in net H+ flux during the first min declined as compared to WL. In particular, the BL-induced change in net K+ flux was much smaller than WL-induced during the first 5 min (Figure 5). The BL-induced PIFRs for Ca”, K” and Cl’ decreased significantly, especially for K” and Cl' which were 58 % and 74 % lower than the WL- induced PIFR, respectively. During the first 5 min, the flux change of Cl' was positive, i.e., BL caused a significant increase in net Cl' flux. In contrast, WL caused a decrease in net Cl' flux (Figure 5). The BL-induced initial responses of net K+ and Cl‘ flux were approximately 3-4 min and 2-3 min slower, respectively, than the WL-induced initial response (Table 1). To examine the BL-fluence-response relationships of ion flux, five BL intensities were used: 16, 32, 70, 160, and 225 pmol m'2 3". These BL treatments elicited a similar ion flux response as shown in Figure 5. The initial response of net ion influx was the smallest under the low BL intensity (16 pmol m'2 s" ) for all ions (Ca2+, Hi, K+ and Cl') (Figure 6). Above 32 pmol m“2 s", the BL-induced initial response in net 11‘, K+ and Cl‘ influx increased and exhibited light saturation kinetics. For Ca2+, higher BL intensities (above 70 pmol m‘2 5") generated larger initial response in net influx (Figure 6). With 142 22:88.2 .3 «-5 .25 mum can an .8... 2m 3% 59.69 8m :8 59.83 .35 3972;; do £292 £94 .modva um :32 «m2 own “comma? 9mg .mo_tm> .3623. 07m .6 mmmcmzw cm “commune. Egon Ema .moammz =>caommE mma E0: .0 new v. . m0 .+I do xac 5c ho 09.38 oE: :0 3:90 Em: do «ootw .m 939m - + +N 2:5 as: :55 95 mv ov mm on mm om or or m 0 me ov mm on ma cm or or m o 4 _ a _ _ _ _ _ _ a our- 3 J _ q _ a _ _ A _ 1 cu. m u w 1 o m r U o. . m I 3. U W I w I 8 ..c r 8.22.. s... I 1 l -.0 :35: l on ( +x :35: on _ _ _ _ a a n _ _ _ _ a _ _ _ _ _ _ _J or or- w u . m w .1 m. o m. .. w - o or a. .. M 1 m \XI i U cm W. I or m. I a 8.23 i z on sir .I 9 dmw ( :3: ._ L. +1 ON (1'5 z_tu pwu) xnu 3| tau ug afiueuo 143 (vs 2-1» pwu) xnu +H iau ug afiueuo .Am:_m> m>=mmmc mv £8 .0 xaEm E mmmmbE cm .0 x35 c_ mmmobmn m 3.865 Amnzn: omcoamom x2... _m_._c_-.m0n_ umoauEéQ. .0 83:22. .329 o>£moa 3 3:83. BEE 0832739. m5 8.865 x35 c2 .9. E omcmzo .538 on. .o .o..m n.mucmfi 9: E393. m.mn _wo_to> 898:3. Wm 53> 2652... cm Emmmao. 958 Sun. .833: $383... mom 9.305 .._0 van +v. . mo .+I .0 $288. x3: :0 £9.25 22.2.3 .6 .omtm .o 0.39”. +N ATm N.E _oE3 853E omm com on? on: on o omm com o? co? om o a a 4 _| 4 ~ _ # q C? OON- . a a a . — 1 _ _ 0V 8- i 1 o8- I 1 fil .. om? I l om 1 I. ow a 1 - H r 1 om? l H 1 V w 1 om. 02- T 1 om. m. m. ow- l 1 e W. om- T m D. T 1 1 1 9 a I. O . oo. 1 v. u M om- I -_o 1 8. or 1 + 1 8. m... 13 q a i a q a q . i a I. a o . E a . 1 1 a _ J 4 ma ) mm: In 1 w R. 1 m; 0 l .I w . om 1 9 w ~./u ON- T. W1 S I. 1 I 1 . ( .3- I 1 s. r l on 1 mw (Ir 1 2- 1 - om- I a a 1- 1 on . oo 3 Em 1T 1 1 cc m .I max: u : m no r I Queccomcoamefiac. 1%: +N + 144 increasing intensity, the BL-induced PIFR of the four ions increased but was saturated above 70 pmol m'2 s‘1 (for H+ and K) or above 160 pmol m‘2 s" (for Ca2+ and Cl') (Figure 6). Effects of red-light on ion fluxes Red-light of 225 pmol m'2 3'1 produced similar net flux responses for H, Ca2+, K+ and Cl‘ as observed following WL or BL exposure. A return to dark conditions caused a decrease in net ion flux (negative) similar to WL and BL (Figure 5). Compared with the WL of 450 nmol m'2 s", the initial net influx responses of H", Ca2+, K” and Cl’ were not significantly influenced by RL of 225 nmol m’2 5’1 (Table 1). However, the RL-induced PIFRs of K and Cl’ were significantly lower than WL (Table 1). In most cases, there were no significant differences between RL- and BL—effects. The significant difference was that RL induced a rapid net K” influx (75 and 38 nmol rn’2 s", respectively) which was larger than BL (2 and 8 nmol m‘2 s", respectively) during the first min and during the first 5 min after illumination (Figure 5). In addition, RL caused a decrease in net Cl' flux (~15 nmol m'2 5") whereas BL still increased net Cl’ flux (+26 nmol m’2 8") during the first 5 min of illumination (Figure 5). The initial RL response time for C1’ (3.0 min) was significantly shorter than BL (5.5 min) (Table 1). Five RL intensities, 75, 225, 450, 1170 and 1490 pmol m'2 s", were tested to evaluate the RL fluence-response on ion flux. The highest two intensities caused an increase in solution temperature of approximately 3-4 and 4-5 °C with in 20 - 30 min (Figure 4). Therefore, these RL exposures would be expected to generate heat effects on ion flux. However, low fluence (below 450 pmol m'2 8") did not cause significant heating of the 145 Table 1. Effects of white-, blue- and red-light on net fluxes of H+, Cu“, K" and Cl' from pea mesophyll tissues. Hi Ca2+ K+ Cl’ Initial flux response (influx) (nmol m'2 s") White light +*22 (8)5a +44 (8)a +142 (10) +124 (8) Blue-light +14(8) b +35 (8)b +118 (8) +93 (8) Red-light +19(8) ab +41 (8) ab +145 (8)ns +80 (8) ns Amplitude of PIFRI (nmol m’2 s") White light -17 (8) -59 (8) a -l96 (10)a -161 (8) a Blue-light -11 (8) ~41 (8) b -82 (8) b -39 (8) b Red-light -19 (8) ns -45 (8) ab -111 (8) b -85 (8) c Initial light response time (min) White light 4.1 (8) 2.4 (8) 4.3 (10) b 2.8 (8) b Blue-light 4.3 (8) 3.6 (8) 8.4 (8) a 5.5 (8) a Red-light 3.0(8)ns 4.0(8)ns 2.5 (8) b 3.0(8)b Different letters indicate significant difference between treatments within columns at p < 0.05 according to LSD test. “ns” indicates no significant difference. White-light fluence is 450umol m’2 5". Blue— and red-light fluences are 32 and 225 pmol m'2 s", respectively, as a result of filtering the white light. 5 The numbers in parentheses indicate the number of replicates for each treatment. ' - or + indicates decrease or increase of net ion flux and extracellular ion concentration, respectively. 3 PIFR: Light-induced Post-Initial Flux Response, a result of either a decrease in net influx or an increase in net efflux, or both. This value is negative. 146 solution. Like WL, all RL intensities exhibited a similar response in net ion fluxes as shown in Figure 5. There were no significant differences of the initial response of net Hi, Ca2+, and Cl' influx among all RL intensities except at the lowest (75 pmol m‘2 s“) (Figure 7). The RL-induced initial response of net K+ influx was intensity-dependent below 450 pmol m‘2 s‘1 but became saturated when RL intensity was further increased. Higher RL fluence rates (1170 and 1490 pmol m‘2 s") also enhanced the initial H+ influx response (Figure 7). This again was possibly due to heat effects. The RL-induced PIF R by the four ions increased in magnitude but became saturated above 225 pmol 111'2 s" for H”, Ca2+ and K”. However, the higher fluence (1490 pmol m'2 5") caused a large increase in the PIFR of Cl' which is also thought to be due to the heating effects (Figure 7). Comparison between blue- and red-light effects on ion fluxes Comparison was conduced when all lights were a similar fluence of 225 pmol m'2 s". Blue-light produced a smaller change in net Ca2+ flux than WL did during the first min (Table 2a). The PFIR of H+ induced by both BL and RL were significantly larger than the WL-induced PIFR. However, no significant differences in the initial response of net H” influx were observed between the three lights. In the same manner, the WL, BL and RL did also not produce a significant difference in either the light-induced initial flux response or the light-induced PIFR of Cl' and K (Table 2a). When both BL and RL intensities were reduced to 70-75 pmol m'2 s", no significant differences of ionic light response were observed between BL and RL. Both BL and RL 147 (Vs 3-111 Iowu) Hdld 10 epnmdwv .829 m>zmmmc 3 £8 .0 xafim £ 030.05 cm .0 x35 E 33.26 m 3.865 Euznc mmcoamom xai EzEEmon nmoanc_-Em__ .o mungaE< .329 o>Emoa 8 3:83. _mz_c_ 302051.50: or: 3629: x23 :2 6: E @9ch .59.. m5 .0 .o..m n.mucmfi m5 Emmeam. 2mg .mo_tm> .mmfiozqo. m1m 2:2, omm.m>m cm Emmmam. $58 Ema .mmzwmz ingoon moa 9.39m 15 new +v_ . 009 com? com oov +N «.0 .+I .0 39.83. x2. :0 36:95 £978. .0 SEE o 0mm- . a com. I 09.. I cor- I on. I 1 l OYI cm. I ON- ow on our on ov on :1» N15 .083 853E 000 F cow- om? I ONT I 3.1 oomr _ com 09.» _ _ mm- 1. E53 82.3.5 1Hu 93.2.: 09.83. 325 1w: .... 939“. o o 1 on 1 OS 1 on. 1 0 w 1 8m .w 1 3 m. 1 08 u 1 w w. nu 1 or m 1U) m m. w. Z on s. on 148 Table 2. Effects of white-, blue- and red-light on ion flux rate in pea mesophyll tissues at similar light fluences. a: at 225 pmol m" s" If Ca: K+ Cl‘ (nmol m“2 5") Initial flux response (influx) White light +ll4(5)§ +31 (5) +71 (3) +27 (3) Blue-light +17 (5) +26 (3) +89 (4) +95 (4) Red-light +21 (3) us +41 (4) ns +113 (4) ns +106 (4) ns Flux change during the first min White light +4 (5) +24 (5) a +42 (3) +3 (3) Blue-light +11 (5) -8 (3) b +5 (4) +12 (4) Red-light +2 (5) ns +5 (4) ab +92 (4) ns +14 (5) ns Amplitude of PIFRt White light -6 (5) b -33 (5) -74 (3) -73 (3) Blue-light -22 (3) a -45 (3) -124 (4) -61 (2) Red-light -16 (5) a -43 (4) ns -1 l4 (4) ns -56 (3) ns 1): at 70-75 pmol m'2 s" H Ca2+ K+ Cl' (nmol m‘2 3") Initial flux response (influx) Blue-light +14 (4) +21(4) +115 (4) +83 (3) Red-light +16 (4) ns +31(5) ns +84 (4) ns +42 (4) ns Amplitude of PIFR Blue-light -18 (4) -34 (4) -96 (4) -25 (2) Red-light -11 (4) ns -31 (5) ns -63 (4) ns -43 (3) ns Different letters indicate significant difference between treatments within columns at p < 0.05 according to LSD test. “ns” indicates no significant difference. § The numbers in parentheses indicate the number of replicates. I - or + indicates decrease or increase of net ion flux and extracellular ion concentration, respectively. : PIFR: Light-induced Post-Initial Flux Response, a result of either a decrease in net influx or an increase in net efflux, or both. This value is negative. 149 produced similar initial flux responses or amplitudes of PIFR of W, Ca2+, K+ and Cl (Table 2b). DISCUSSION Effects of light intensity Previous studies on light-induced ionic or electrical responses in plants have used an intermediate fluence within the range of 200 - 400 pmol In2 5". Only a few studies have evaluated light responses to different light intensities. Elzenga et al. (1995) found that light-induced depolarization in pea mesophyll tissues depended, in a saturable way, on light intensities between 50-1000 pmol m'2 5". Lower intensities were observed to result in less depolarization of the plasma membrane and a delay in the light response. The light-induced polar pH reaction in Elodea canadensis was also found to be reduced at lower light intensities (Elzenga and Prins, 1989). In the current study, the light responses in the net fluxes of Ca2+, H”, K+ and Cl' were limited at the lowest WL fluence (225 pmol rn’2 s" ) (Figure 2). Although higher fluence (450 - 1800 pmol m‘2 5") produced a higher magnitude of net influx changes for all ions, that of net H” and Cl' influx appeared to saturate. Under 450 pmol m'2 s", the light-induced PIFR of Ca2+, H“ and K+ also were saturated (Figure 2). Elzenga et al. (1995) also reported that light-induced depolarization of pea mesophyll was saturated at approximately 400 pmol m'2 5". However, the increase in solution temperature during light exposure constitutes a potential problem in measurement of ion flux at high fluence. Both factors, light intensity and solution temperature, should be considered when one evaluates the effect of 150 light on ion flux. Most researchers ignore or do not consider the li ght-induced heat effects on leaf tissues at lower radiance levels (Stahlberg and Van Volkenburgh, 1999, Shabala and Newman, 1999). High fluence was found to increase bathing solution temperature by approximately 4 - 5 °C within 20 min (Figure 4). Therefore, temperature effects caused by high fluence on ion flux must be considered and calculated. Bathing solution temperature increase of 3 - 5 °C clearly increase net flux kinetics of Ca“, H, K* and Cl’ under darkness. By contrast, smaller increases in temperature (1.0 - 2.5 °C) did not affect net ion flux in a measurable manner (Figure 3). Therefore, high fluence- induced changes of ion fluxes could be expected to result fiom two factors, light intensity and heat. In other words, ion flux changes induced by high intensities (1800 and 2800 pmol m‘2 3") should not be only explained on the basis of light fluence, but also include possible heat effects. Temperature directly affects the kinetics of ion transport across the membranes in biological systems. For example, Lee and Satter (1988) found that temperature had a large influence on H+ release fi'om extensor cells of Samanea saman pulvini. In this study, it is difficult to differentiate between light- and temperature-effects at high fluence. The present work indicates that radiance of 450 of nmol m'2 s'1 is enough to stimulate ionic responses to light. Effects of light quality Light (WL) induced a rapid increase in net H+ influx from pea mesophyll tissues (see Chapter 1). Both RL and BL also elicited a similar time-course change of net Hi (Figure 5). At the same fluence, no significant differences in the initial response in net H+ influx 151 was observed among the three light treatments (Table 2). However, the BL- and RL- induced PIFR for H were significantly larger than that of WL- at 225 pmol m'2 5’1 (Table 2a). As discussed in Chapter H, the plasma membrane Hi-ATPase may contribute to the light-induced PIFR of H+ causing apoplastic acidification due to increased proton efflux from mesophyll cells. Both lights stimulate proton efflux (Assmann et al., 1985; Serrano et al., 1988). However, several contradictory opinions exist in the literature. Nishizaki (1992) found that BL and RL elicited membrane depolarization and hyperpolarization, respectively. He proposed that a pulse of BL inhibited the activity of the proton-ATPase in pulvinar motor cells. Spalding and Cosgrove (1992), on the other hand, suggested that the mechanism of BL-induced membrane depolarization of cucumber hypocotyls initially involves inhibition of a HI-ATPase. This study provides indirect evidence for the role of both BL and RL in stimulating proton efflux. Additionally, the fluence-responses to both BL and RL were similar for the initial response of net HI influx (Figure 6, 7). It is well established that light induces a transient increase in net Ca2+ influx. Like Hi, the effects of both BL and RL on net Ca2+ flux were similar to that of WL (Figure 5). Elevated concentrations of cytosolic Ca2+ were found in RL- (Shacklock et al., 1992; Neuhaus et al., 1993; Ermolayeva et al., 1996) and BL-treated cells (Christie and Jenkins, 1996; Sharma et al., 1997). The role of Ca2+ as a cellular messenger, has been suggested to play a role in both BL and RL signal transduction. However, evidence for a direct link between Ca2+ transport and light quality is still lacking. The present results support the hypothesis that both light qualities might initially increase net Ca2+ influx. The most likely pathway for Ca2+ influx into cytoplasm would be through Ca2+ channels in the plasma membrane (see Chapter II) resulting in a transient increase in cytosolic free Ca” 152 concentration. Both BL and RL stimulated K+ uptake by guard cells (MacRobbie, 1988) and by bean mesophyll (Blum et al., 1992). In pea mesophyll, RL induced a rapid and large increase in net K+ influx from pea mesophyll (Figure 5). In comparison, the BL-induced change in net K+ influx was slower and smaller during the first min of illumination. However, such result was not observed at low RL fluence of 75 pmol rn‘2 5'1 (Table 2b). Similarly, no difference between both BL and RL existed in the magnitude of the initial response in net K+ influx (Table 2). These differences are mainly reflected in the initial light response time. Furthermore, the BL- and RL-induced PIFRs of K+ were also similar (Table 2). As proposed in Chapter II, the light-induced PIFR of K+ flux change may be associated with a enhanced net KI efflux mediated by K+ channels. Both light qualities were found to stimulate KI channel activity. For example, BL activated K” efflux channels in flexor cells of Samanea saman (Suh et al. 2000), whereas phytochrome activated K+ channels in Mougeotis protoplasts (Lew et al., 1990). Subsequent research demonstrated that one or two Ca2I-activated K+ channels were involved (Lew et al., 1992). According to the present results in pea mesophyll, BL and RL should be considered to play a role in the light-induced K+ flux response in a similar manner. Blue-light-activated anion channels have been identified in leaves of Arabidopsis (Cho and Spalding, 1996; Lewis et al., 1997). These authors speculated that Cl‘ efflux, mediated by anion channels, is responsible for BL-induced membrane depolarization. However, the activity of anion channels does not explain fully the change of Cl' flux across the plasma membrane. Except efflux through anion channels, Cl' can move 153 inwardly into cytoplasm through a Hi/Cl‘ symport mechanism (Sanders et al., 1985). The net flux reflects a balance between influx and efflux across the plasma membrane in pea mesophyll. In the present study, both BL and RL stimulated similar changes in net Cl' influx and the PIFR of Cl' at the same fluence (Table 2). The light-induced PIFR of Cl“ has been considered to be the result of Cl‘ efflux catalyzed by anion channels within the plasma membrane (see Chapter 11). Taken together the present data suggest that both BL and RL all induced flux responses of H’“, Ca2+, K+ and CI' in a similar manner. The same phenomenon was also found in effects of BL and RL on the activity of ion transport across the plasma membrane of guard cells (Assmann et al., 1985; Zeiger et al., 1987; Serranno et al., 1988; Shimazaki et al., 1993). This indicates that BL + RL regions of the light spectrum can elicit ionic flux responses in pea mesophyll tissues. The similarity in ionic responses to BL and RL are consistent with the conclusion in Chapter I that ion flux responses in pea mesophyll are photosynthesis dependent. This is supported by the finding that DCMU, a photosynthestic inhibitor, also inhibited the light-induced initial responses in net ion flux (see Chapter I) and removed light-induced membrane depolarization (Stahlberg and Van Volkenburgh, 1999) in pea mesophyll. The current study provides a starting point for further investigations on potential involvement of photoreceptor. 154 LITERATURE CITED Assmann SM, Shmazaki K. 1999. The multisensory guard cell: Stomatal response to blue light and abscisic acid. Plant Physiol 119:809-815. Assmann SM, Simoncini L, Schroeder JI. 1985. Blue light activates electrogenic ion pumping in guard cell protoplasts of Viciafava. Nature 318:285-287. Batchauer A. 1998. Photoreceptors of higher plants. Planta 206:479-492. Blum DE, Elzenga JTM, Linnemeyer PA, Van Volkenburgh E. 1992. Stimulation of grth and ion uptake in bean leaves by red and blue light. Plant Physiol 100:1968- 1975. Briggs WR, Liscum E. 1997. Blue light-activated signal transduction in higher plants. In P Aducci, ed, Signal transduction in plants. Brikhauser Verlag Basel/Switzerland, plO7-135. Chamovitz DA, Deng XW. 1996. Light signaling in plants. Crit Rev Plant Sci 15:455- 478. Cho MH, Spalding EP. 1996. An action potential in Arabidopsis hypocotyls activated by blue light. Proc Natl Acad Sci USA 93:8134—8134. Elzenga JTM, Prins HBA. 1989. Light-induced polar pH changes in leaves of Elodea canadensis. 1. effects of carbon concentration and light intensity. Plant Physiol 91 :62-67 Elzenga JTM, Prins HBA, Van Volkenburgh E. 1995. Light-induced membrane potential changes of epidermal and mesophyll cells in growing leaves of Pisum sativum. Planta 197:127-134. Ermolayeva IE, Hohmeyer H, Johannes E, Sanders D. 1996. Calcium—depend membrane depolarization activated by phytochrome in the moss Physcomitrella patens. Planta 199:352-358. Johannes E, Ermolayeva E, Sanders D. 1997. Red-light-induced membrane potential transients in the moss Physcomitrella patens: ion channels interaction in phytochrome signalling. J Exp Bot 48:599-608. Johnson CM, Stout PR, Broyer TC, Carton AB. 1957. Comparative chlorine requirements of different plant species. Plant and Soil 8:337-353. Lee Y, Satter RL. 1988. Effects of temperature on H+ uptake and release during circadian rhythmic movements of excised Samanea motor organs. Plant Physiol 86:352- 354 155 Lew RL, Serlin BS, Schauf CL, Stockton ME. 1990. Calcium activation of Mougeotia potassium channels. Plant Physiol 92:831-836. Lewis BD, Karlin-Neumann G, Davis RW and Spalding EP. 1997. Cay-activated anion channels and membrane depolarization induced by blue light and cold in Arabidopsis seedlings. Plant Physiol 114:1327-1334. Long JM, Widders IE. 1990. Quantification of apoplastic potassium content by elution analysis of leaf lamina tissue from pea (Pisum sativum L. cv. Argenteum). Plant Physiol 94:1040-1047. MacRobbie EAC. 1988. Control ion flux in stomatal guard cells. Bot Acta 101:140- 148. Neuhaus G, Bowler C, Kern R, Chua NH. 1993. Calcium/calmodulin-dependent and - independent phytochrome signal transduction pathways. Cell 73:937-952. Nishizaki Y. 1992. Effects of inhibitors on the light-induced changes in electric potential in pulvinar motor cells of Phaseolus vulgaria L. Plant Cell Physiol 33:1073- 1078. Remis D, Treffny B, Gimmler H. 1994. Light-induced H+ transport across the plasma membrane of the acid-resistant green alga Dunaliella acidophila. Plant Physiol Biochem 32:75-84 Sanders D, Smith FA, Walker NA. 1985. Proton/chloride cotransport in Chara: mechanism of enhanced influx after rapid external acidification. Planta 163:411-418. Serrano EE, Zeiger E, Hagiwara S. 1988. Red light stimulates an electrogenic proton pump in Vicia guard cell protoplasts. Proc Natl Acad Sci USA 85:436-440. Serrano EE, Zeiger E. 1989. Sensory transduction and electrical signaling in guard cells. Plant Physiol 91 :795-799. Shabala NS, Newman IA, Morris J. 1997. Oscillations in H and Ca2+ fluxes around the elongation region of corn and effects of external pH. Plant Physiol 113:111-118. Shabala NS, Newman IA. 1999. Light-induced changes in hydrogen, calcium, potassium, and chloride ion fluxes and concentration from the mesophyll and epidermal tissues of bean leaves. Understanding the ionic basis of light-induced bioelectrogensis. Plant Physiol 119:1115-1124. Shacklock PS, Read ND, Trewavas AJ. 1992. Cytosolic free calcium mediates red light- induced photomorphologensis. Nature 358:753-755. 156 Shanna VK, Jain PK, Malik MK, Maheshwari SC, Khurana JP. 1997. Light- and calcium-modulated phosphorylation of proteins from wheat seedlings. Phytochemistry 44:781-786. Shimazaki K, Omasa K, Kinoshita T, Nishimura M. 1993. Properties of the signal transduction pathway in the blue light response of stomatal guard cells of Vicia fava and Commelina benghalensis. Plant Cell Physiol 34:1321-1327 Spalding EP, Cosgrove DJ. 1992. Mechanism of blue-light-induced plasma membrane depolarization in etiolated cucumber hypocotyls. Planta 188:199-205. Suh S, Moran, Lee Y. 2000. Blue light activates potassium-efflux channels in flexor cells from Samanea saman motor organs via two mechanisms. Plant Physiol 1232833- 843. Zeiger E, Lino M, Shimazaki K, Ogawa T. 1987. The blue-light response of stomata: mechanism and function. In E Zeiger E, GD Farquhar, IR Cowan, eds, Stomatal function. Stanford University Press, Stanford, pp 209-227. 157 SUMMARY Light signalling is becoming a main area of research in plant science. Plant scientists try to elucidate photosensory mechanisms through genetic, biochemical and biophysical approaches. Membrane potential changes are commonly observed responses to many stimuli. However, the ionic basis for such stimulus-elcited membrane potential changes remains to be determined. Ion transport mechanisms and kinetics are frequently not understood. It is known that ion activity gradients and fluxes exist across cellular membranes, and this fact provides a potential target for studying electrical signalling. Because electrical signals are short-lived, it is difficult to use biochemical and molecular techniques for investigation of such responses. In contrast, ion-microelectrode technology can measure rapid ion kinetics in a time-course manner around plant cells without tissue damage. Microelectrode Ion Flux Estimation (MIFE) is a slowly vibrating ion-selective microelectrode system developed by Shabala and Newman (Shabala et al., 1997; Shabala and Newman, 1999). The MIFE has three advantages. First, it can measure the simultaneous flux of three ions, thus one can observe multiple ion movements across mesophyll plasma membrane. Second, it provides a stable and anti-noise environment during an experiment. In most cases, changes in extracellular ion concentration with are consistent with changes in net ion flux, the enabling the identification and exclusion of artifacts. Third, this system contains a sofiware package to automatically record all 158 measurements and to estimate flux. Therefore, the MIFE allows direct, dynamic and continuous quantitative estimation of ion kinetics around cells. To focus on mesophyll cells, a small strip of epidermal tissue was gently removed from pea leaflets with little or no wounding. One cannot fully discount the possibility, however, that peeling may elicit some type of long-term wound response which may pre- dispose the tissue to responding to light in a certain manner and thus give rise to artifacts. On the other hand, extensive research with peeled leaf lamina tissue indicates that ion flux kinetics is similar to intact tissue. The peeled mesophyll cells can easily be viewed with an inverted microsc0pe. This configuration ensures that measured ion fluxes are predominately from mesophyll cells. Although Elzenga et al. (1995) observed a rapid light-induced membrane depolarization (25 mV) in pea mesophyll cells, they did not provide further evidence for the ionic basis of bioelectrogenesis. This dissertation demonstrates a significant, short- term, light-induced ion flux response of pea mesophyll tissues. Rapid initial increases in net Ca2+, H, K” and Cl' flux from mesophyll cells were measured afier onset of illumination. The response of Ca2+ flux was the most rapid (13 5), whereas HI, Cl‘ and K+ flux responses started within approximately 29, 45, and 48 s, respectively, after stimulation. Following the initial response, a light-induced Post-Initial Flux Response (PIFR) occurred, which led to a decrease in net flux. The PIFR time of Ca2+ and Cl' was shorter (3 -4 min). Ion transport across the plasma membrane can alter membrane potentials. Assuming that Ca2+, HI, K+ and C1' are primary ions involved in light-induced changes in membrane potential, this study showed that the time course of summed net charge flux was similar 159 to that of the light-induced membrane potential reported previously (Elzenga et al., 1995; Stahlberg and Van Volkenburgh, 1999). A light-induced initial increase in the net charge flux would be expected to depolarize the plasma membrane. Therefore, this dissertation presents a effective and usefiil method to estimate the involvement of ion fluxes in electrical signalling. Calcium fimctioning as a second messenger is thought to play a critical role in cellular signal transduction in plant cells. An increase in free Ca2+ concentrations in the cytoplasm occurs in most stimulus-responses. The present results support Ca2+ influx as an initial depolarizing agent in the light-induced membrane depolarization (Spalding et al., 1992; Shabala and Newman, 1999). However, other researchers have not yet provided experimental evidence for a Ca2+ influx mechanism. This study showed that the light-induced initial net Ca2+ influx was significantly inhibited by 42-44 % in the presence of the Ca2+ channel blocker LaCl3 and the chelator EGTA. Apparently, the light-induced initial response in net Ca2+ flux is the result of an increase in influx mediated by Ca2+ channels, rather than a decrease in efflux. The increased Ca2+ influx would be expected to elevate cytoplasmic Ca2+ concentrations, thus initiating cellular signal transduction. The effects of Ca2+ influx on other ion flux responses also provide additional evidence for this hypothesis. Both LaCl3 and EGTA significantly reduced the light-induced initial H+ and K response and PIFR of C1‘. The light-induced PIFR of Cl' were also significantly enhanced by both Ca2+ ionophore A23187 and high exogenous Ca2+ concentration. Therefore, this dissertation may be the first to experimentally demonstrate that Ca2+ influx through Ca2+ channels represents an initial step in light- induced membrane depolarization. 160 P. 1 ~11... A light-induced extracellular alkalinization (pH increase of 0.4 - 0.5) was observed in the current study. The results were anticipated due to CO2 uptake and HI-mediated K+ or Cl‘ uptake by mesophyll cells. Shabala and Newman (1999) also concluded that CO2 uptake of bean mesophyll cells was a reason of light-induced increase in apoplast pH. However, the plasma membrane HI-ATPase inhibitors, Na3VO4 and DCCD, inhibited the light-induced initial response in net H+ influx by 50-60 % and eliminated the PIFR of H“. The presence of Na,VO4 in the bathing solution also significantly suppressed the light- induced increase in net K+ influx. In contrast, the ATPase activator, F usicoccin significantly increased the light-induced PIFR of H+ and the initial response in net Cl‘ influx. In addition, external pH in the bathing solution affected fitrther the light-induced response in net KI and Cl' influx. These results suggest that HI-ATPase in the plasma membrane is not directly involved in the initial light response. The enzyme is involved in mediating Hi/K+ and HI/Cl' symports, which contribute to light-induced initial extracellular alkalinization. Anion channels may also be involved in membrane potential changes. Elzenga and Van Volkenburgh (1997) also reported increased activity of an anion channel in pea mesophyll cells in response to light. However, they did not measure Cl' flux, which was a main reason for the light-induced PIFR of C1' in my study. Testing with anion channel blockers indicated that rapid-type anion channels play a role in increasing Cl' efflux, but anion channels are not involved in light-induced initial Cl‘ flux response. This result is different from that of Elzenga and Van Volkenburgh (1997). Another contribution of this dissertation is the demonstration that the light-induced PIFR is associated with ion transporters, including Ca2"-ATPase, HI-ATPase, K+ channels 161 and anion channels in the plasma membrane. Differential alteration in ion transporter activity suggests that multiple ion transport pathways fimction during short-term light responses. To date, such ion flux evidence has not been reported. Most research has just focused on the effects of ion transporter inhibitors or activators on membrane potentials or on ion concentration change in bulk solution surrounding plant tissues. In this dissertation, I proposed a model for short-term light responses, based on experimental evidence, and provided a possible explanation for bioelectrogenesis. Ion flux responses represent an important component of cellular signalling systems in plants to external stimuli. This model also provides a potential approach for future research in signal transduction of plants. Both blue light and red light display a similar action in controlling stomotal movement (Serranno and Zeiger, 1989; Assmann and Shimazaki et al., 1999). Similar effects of both wavelengths on ion fluxes in pea mesophyll cells were found in the present study. The fact that DCMU treatment inhibited the light-induced H+ and Ca2+ flux responses suggests that the ionic responses to light are photosynthesis-dependent. The wide action spectra of the ion flux response introduces a further question; namely whether photoreceptors are ever involved in the light-induced ion flux response? Combining MIFE technology with specific mutants will enhance the possibility of ascertaining the involvement of specific photoreceptors in the future research. Although the MIFE system can estimate net ion flux dynamics in the apoplast, the technique does not directly measure intracellular ion concentration change. Some attempts have been made to detect cytosolic ion concentrations and pH with improved ion-selective microelectrode techniques, such as the double-barreled electrode and the 162 pH-electrode (Felle, 1993). In further studies, the MIFE system might be used to test Ca2+ concentration and pH in the cytoplasm with improved microelectrodes. The present study does not exclude the possibility of CaZI-independent mechanisms. Hydrogen signalling, coupled with ion flux changes may constitute an alternate mechanism for cellular signal transduction. Further research is needed to test such a hypothesis. A combination of both MIFE and patch-clamp techniques would provide further discrimination among specific ion transporters, for instance, inward or outward K” rectifiers and anion channels. My data in this dissertation and studies by others (Shabala et al., 1997; Shabala and Newman, 1999) indicate that, among the ion microelectrode techniques, the MIFE system is a most powerfirl tool to measure ion flux in plant cells. The fiirther improvement of the MIF E system will open many new interesting avenues of research, including the elucidation of signal transduction pathways in response to pathogen infection, salinity, cold, wounding, etc. It would also be interesting to investigate localized ion flux changes in remote tissues of whole plants when stimuli have been shown to elicit systemic response in plants, thus indicating the occurrence of long-distance signal systems. 163 BIBLIOGRAPHY Antkowiak S, Mayer WE, Engelrnann W. 1991. Oscillation of the membrane potential of pulvinar motor cells in situ in relation to leaflet movements of Deodium motorium. J Exp Bot 240:901-910. Arif I, Newman IA, Keenlyside N. 1995. Proton flux measurements from tissues in buffered solution. Plant Cell Environ 18:1319-1324. Alexandre L, Lassalles JP, Kado RT. 1990. Opening of Ca2+ channels in isolated red beet root vacuole membrane by inositol-1,4,5-triphosphat. Nature 343:567-570. Allan AC, Ficker MD, Ward JL, Beale NIH, Treanas. 1994. Two transduction pathways mediate rapid effects of abscisic acid in Commelina guard cells. Plant cell 621319-1328. Allen GJ, Muir SR, Sanders D. 1995. Release of Ca2+ individual plant vacuoles by both IP3 and cyclic ADP-ribose. Science 268:735-737. Anderson JA, Hupriker SS, Kochian LV, Lucas WJ, Gaber RF. 1992. Functional expression of a probable Arabidopsis thaliana potassium channel in Saccaromyces cerevisias. Proc Natl Acad Sci USA 92:9520—9524. Assmann SM. 1993. Signal transduction in guard cells. Annu Rev Cell Biol 92345-375. Assmann SM. 1996. G-protein regulation of plant K channels. In DPS Verama, ed, Signal transduction in plant growth and development. Springer Wien New York, pp39- 61. Assmann SM, Shmazaki K. 1999. The multisensory guard cell: Stomatal response to blue light and abscisic acid. Plant Physiol 119:809-815. Assmann SM, Simoncini L, Schroeder H. 1985. Blue light activates electrogenic ion pumping in guard cell protoplasts of Viciafava. Nature 318:285-287. Assmann SM, Zeiger E. 1987. Guard cell bioenergetics. In E Zeiger, E Farquhar, I Cowan, eds, Stomatal Function. Stanford University Press, Stanford, CA, pp163-l94. Batchauer A. 1998. Photoreceptors of higher plants. Planta 206:479-492. Bach M, Schnitzler JP, Seitz HU. 1993. Elicitor-induced changes in Ca2+ influx, K" efflux, and 4-hydroxybenzoic acid synthesis in protoplasts of Dcucus carats L. Plant 164 Physiol 103:407-412 Bauer CS, Plieth C, Hansen UP, Simonis W, Schonknecht G.‘ 1998. A steep Caz“- dependence of a K” channel in a unicellular green alga. J Exp Bot 49: 1 761-1765. Beilby MJ, Coster HGL. 1979. The action potential in Chara corallina. 11 two activation transient in voltage and excitation in nerve. Aust J Plant Physiol 6:323-335. Beilby MJ, Walker NA. 1981. Chloride transport in Chara. 1. Kinetics and current- voltage curves for a probable proton symport. J Exp Bot 32:43-54. Beilby M. 1984. Calcium and plant action potentials. Plant Cell Environ 7:415-421. Bentrup F W. 1974. Lichtabh'angige membrane potential bei pflanzen. Ber Detsch Botany Ges 87:515-528. Bentrup FW. 1990. Potassium ion channels in the plasmalemma. Physiol Plant 79:705- 71 1. Bethmann B, Simonis W, Schonknecht H. 1998. Light-induced changes of cytosolic pH in Eremosphaera viridis: recordings and kinetic analysis. J Exp Bot 49:1129-1137. Bisson MA. 1984. Calcium effects on electrogenic pump and passive permeability of the plasma membrane of Chara corallina. J Membr Biol 81:59-67. Bisson MA, Gutknecht J. 1980. Osmotic regulation in algae. In RM Spanswick, WJ Lucas, J Dainty, eds, Plant Membrane Transport: Current Conceptual Issues. Amsterdam: Elsevier, North Holland, pp131-142. Bjorkman T, Cleland RE. 1991. The role of extracellular fi'ee-calcium gradients in gravitropic signalling in maize roots. Planta 185:379-384. Blatt MR. 1987. Electrical characteristics of stomatal guard cells: the ionic basis of the membrane potential and the consequence of potassium chloride leakage fiom microelectrodes. Planta 170:272-287. Blatt MR. 1992. K+ channels of stomatal guard cells: characteristics of the inward rectifier and its control by pH. J Gen Physiol 99:615-644. Blatt MR. 1999. Reassessing roles for Ca” in guard cell signalling. J Exp Bot 50:989- 999. Blatt MR, Grabov A. 1997. Signal rebundacy, gates and integration in the control of ion channels for stomatal movement. J Exp Bot 48:529-537. 165 Blatt MR, Armstrong F. 1993. K+ channels of stomatal guard cells: abscisic acid- evoked control of the outward rectifier mediated by cytoplasmic pH. Planta 191:330-341. Blatt MR, Rodriguez-Navaro A, Slayman CL. 1987. Potassium-proton symport in Neurospora: kinetic control by pH and membrane potential. J Membr Biol 98: 169-195. Blatt MR, Thiel G. 1993. Hormonal control of ion channel gating. Annu Rev Plant Physiol Plant Mol Biol 44:543-567. Bleecker AB, Schaller CE. 1996. The mechanism of ethylene perception. Plant Physiol 111:653-660. Blom-Zandstra M, Koot HTM, Van Hattum J, Vogalzang SA. 1997. Transient light- induced changes in ion channel and proton pump activities in the plasma membrane of tobacco mesophyll protoplasts. J Exp Bot 48:1623-1630. Blum DE, Elzenga JT M, Linnemeyer PA, Van Volkenburgh E. 1992. Stimulation of growth and ion uptake in bean leaves by red and blue light. Plant Physiol 100:1968- 1975. Blumwald E, Aharon QS, Lam BCH. 1998. Early signal transduction pathways in plant- pathogen interactions. Trends Plant Sci 3:342-346. Bowler C. 1997. The transduction of light signals by photochrome. In P Aducci, ed, Signal transduction in plants. Brikhauser Verlag Basel, Switzerland, pp137-152. Bowler C, Neuhaus G, Yamagata H, Chua NH. 1994a. Cyclic GMP and calcium mediate phytochrome phototransduction. Cell 77:7 3-8 1. Bowler C, Yamagata H, Neuhaus G, Chua NH. 1994b. Phytochrome signal transduction pathways are regulated by recipotcal control mechanisms. Genes Dev 8:2188-2202. Bray E. 1997. Plant responses to water deficit. Trends Plant Sci 2:48-53. Briggs WR, Liscrun E. 1997. Blue light-activated signal transduction in higher plants. In P Aducci, ed, Signal transduction in plants. Brikhauser Verlag Basel/Switzerland, p107-135. Brownlee C, Goddard H, Hetherington AM, Peake LA. 1999. Specificity and integration of responses: Ca2+ as a signal in polarity and osmotic regulation. J Exp Bot 50:1001-1011 Bush DS. 1993. Regulation of cytosolic calcium in plants. Plant Physiol 103:7-13. Bush DS. 1995. Calcium regulation in plant cells and its role in signaling. Annu Rev Plant Physiol Plant Mol Biol 45:95-122. 166 Camelli A, De Michelis MI, Rasi-caldogno FR. 1992. Plasma membrane Ca-ATPase of radish seedlings. 1. Biochemical characterization using ITP as a substrate. Plant Physiol 98:1196-1201. Chamovitz DA, Deng XW. 1996. Light signaling in plants. Crit Rev Plant Sci 15:455- 478. Chang C, Stewart R. 1998. The two-component systems. Plant Physiol 117:723-731. Cheeseman JM, Edwards M, Meidner H. 1982. Cell potentials and turgor pressures in epidermal cells of T radescantia and Commelina. J Exp Bot 33:761-770. Chen R, Rosen E, Masson PH. 1999. Gravitropsim in higher plant. Plant Physiol 120:343-350. Cho MH, Spalding EP. 1996. An action potential in Arabidopsis hypocotyls activated by blue light. Proc Natl Acad Sci USA 93:8134-8134. Clark T, Mathews S, Sharrock PA. 1994. The phytochrome apoprotein family in Arabidopsis is encoded by five genes: the sequences and expression of PHYD and PHYE. Plant Mol Biol 25:413-427. Clark GB, Menon AR, Tong CG, Thompson GA, Roux SJ. 1993. Phytochrome regulates GTP-binding protein activity in the envelop of pea nuclei. Plant J 4:399-402. Colbran RJ, Soderling TR. 1990. Calcium/calmodulin-dependent protein kinase II. Curr Topics Cell Regul 31 :181-200. Coleman HA. 1986. Chloride currents in Chara — a patch clamp study. J Membr Biol. 83:109-118. Cosgrove DJ, Hedrich R. 1991. Stretch-activated chloride, potassium, and calcium channels coexisting in plasma membranes of guard cells of Viciafava L. Planta 186:143- 153. Coté GG. 1995. Signal transduction in leaf movement. Plant Physiol 1092729-734 Czempinski K, Gaedeke N, Zimmerrnann S, Miiller-Rdber B. 1999. Molecular mechanism and regulation of plant ion channels. J Exp Bot 50:955-966. Davies E. 1987a. Action potential as multifunctional signals in plant: a unifying hypothesis to explain apparently disparate wound response Plant Cell and Environment 10:623-631. 167 Davis E. 1987b. Plant response to wounding. In PK Stumpf, EE Conn, eds, Biochemistry of Plants, Vol. 12, New York, Academic Press, pp 243-264. Ding JP, Pichard BG. 1993. Modulation of mechanosensitive calcium-selective cation channels by temperature. Plant J 3:713-720. DuPont FM, Bush DS, Windle J], Jones RL. Calcium and proton transport in membrane vesicles from barley roots. Plant Physiol 94:179-188 Ebel J, Mithofer A. 1998. Early events in the elicitation of plant defence. Planta 206:335-348. Elzenga JTM, Prins HBA. 1989. Light-induced polar pH changes in leaves of Elodea canadensis. 1. effects of carbon concentration and light intensity. Plant Physiol 91:62-67 Elzenga JTM, Prins HBA, Van Volkenburgh E. 1995. Light-induced membrane potential changes of epidermal and mesophyll cells in growing leaves of Pisum sativum. Planta 197:127-134. Elzenga JTM, Van Volkenburgh E. 1993. Ion channels in the plasma membrane of epidermal and mesophyll cells of growing pea leaves. Plant Physiol 102:SIO6. Elzenga JTM, Van Volkenburgh E. 1997. Characterization of a light-controlled anion channel in the plasma membrane of mesophyll cells of pea. Plant Physiol 113:1419- 1426. Ermolayeva E, Hohmeyer H, Johannes E, Sanders D. 1996. Calcium-depend membrane depolarization activated by phytochrome in the moss Physcomitrella patens. Planta 199:352-358. Ettinger WF, Clear AM, Fanning KJ, Peck ML. 1999. Identification of a Cay/HI antiport in the plant chloroplast and thylakoid membrane. Plant Physiol 119:1379-1385. Evans DE, Briars SA, Williams LE. 1991. Active calcium transport by plant cell membranes. J Exp Bot 42:285-323. Evans HJ, Wildes RA. 1971. Potassium and its role in enzyme activation: potassium in biochemistry and physiology. Proceeding of the 8th colloquium. International Potash Institute, Bern, Switzerland, pp 13-39. Evans DE, Williams LE. 1998. P-type calcium ATPases in higher plants — biochemical, molecular and functional properties. Biochem Biophys Acta 1376:1-25. Fairley KA, Assmann SM. 1991. Evidence for G-protein regulation of inward potassium channel current in guard cells of fava bean. Plant Cell 3:1037-1044. 168 Falke LC, Edwards KL, Pickard BG, Misler S. 1988. A stretch-activated anion channel in tobacco protoplasts. FEBS Lett 237:141-144. Felle HH. 1988a. Auxin causes oscillations of cytosolic free calcium and pH in Zea mays L.: interaction of Ca2+ and pH? Plants 176:248-255. F elle HH. 1988b. Short-term pH regulation in plants. Physiol Plant 74:583-591. Felle HH. 1989a. KI/HI-antiport in Riccia fluitans: an alternative to the plasma membrane H+ pump for short-term regulation? Plant Sci (lett.) 61 :9-15. Felle HH. 1989b. pH as a second messenger in plants. In WF Boss, DJ Marré, eds, Second messengers in plant growth and development. New York, Alan R. Liss Inc. pp145-166. Felle HH. 1992. Ion-selective microelectrodes: their use and important in modern plant biology. Bot Acta 106:5-12. Felle HH. 1993. Developmental physiology: signal transduction. Progress in Bot 54:254-267. F elle HH. 1994. The H+/Cl' symporter in root-hair cells in Spinais alba. Plant Physiol 106:1131-1136. Felle HH. 1998. The apoplasic pH of the Zea mays root cortex as measured with pH- sensitive microelectrodes: aspects of regulation. J Exp Bot 49:987-995. Frachisse JM, Desbiez MO. 1989. Investigation of the wave of electrogenic active and passive membrane elements in Riccafluitans. Biochem Biophys Acta 646:151-160. Fromm J. 1991. Control of phloem unloading by action potential in Mimosa. Physiol Plant 83:529-533. Fromm J, Spanswick R. 1993. Electric signals released from roots of willow (Salix vimz'nalis L.) change transpiration and photosynthesis. J Plant Physiology, 141:673-680. Furuya M. 1993. Phytochromes: their molecular species, gene families, and functions. Annu Rev Plant Physiol Plant Mol Biol 44:617-645. Gallagher S, Short TW, Ray PM, Pratt LH, Briggs WR. 1988. Light-mediate changes in two proteins found associated with plasma membrane fractions from pea stem sections. Proc Natl Acad Sci USA 85:8003-8007. Galland P, Senger H. 1991. Flavins as possible blue-light photoreceptor. In MG 169 Holmes, ed, Photoreceptor evolution and function. Academic Press, London, p 65-124. Gehring CA, Williams DA, Cody SH, Parish RW. 1990. Phototropism and geotropism in maize coleopitles are spatially correlated with increases in cytosolic free calcium. Nature 345:528-530. Genoud T, Métraux JP. 1999. Crosstalk in plant cell signaling: structure and function of the genetic network. Trends Plant Sci 4:503-507. Gill DL, Waldron RT, Rys-Sikora KE, Ufret-Vincenty CA, Graber MN, Favre CJ, Alfonso A. 1996. Calcium pools, calcium entry, and cell growth. Bioscience Reports 16:139-157. Gilory S, Read ND, Trewavas T. 1990. Elevation of cytoplasmic calcium by caged calcium or caged inositol triphosphate initiates stomatal closure. Nature 346:769-771. Gilory S, Trewavas T. 1994. A decade of plant signals. Bioessays. 16:677-682. Giroldini W. 1988. Extracellular light-evoked potentials in T radescantia albbiflora leaves. Plant Physiol Biochem 26: 133-138. Gotow K, Sakaki T, Tondo N, Kobayashi K, Syono K. 1985. Light-induced alkalinization of the suspending medium of guard cell protoplasts from Vicia fava L. Plant Physiol 79:825-828. Grabov A, Blatt MR. 1997. Parallel control of the inward-rectifier K” channel by cytosolic free Ca2+ and pH in Vicia guard cells. Planta 201:84-95. Grabov A, Blatt MR. 1998. Co-ordination of signalling elements in guard cell ion channels control. J Exp Bot 49:351-360. Harker FR, Venis MA. 1991. Measurement of intracellular and extracellular free calcium in apple fi'uit cells using calcium selective microelectrodes. Plant Cell Environ 14:525-530. Haschke KP, Lfittge K. 1975. Interaction between 1AA, potassium and malate accumulation and growth in Avena coleoptile segments. Z pflanzenphysiol 76:450-455. Hedrich R. 1994. Voltage-dependent chloride channels in plant cells: identification, characterization, and regulation of a guard cell anion channel. Curr Topics in Membrane 42:1-33. Hedrich R, Busch H, Raschke K. 1990. Ca2+ and nucleotide dependent regulation of voltage dependent anion channels in the plasma membrane of guard cells. EMBO J 9:3889-3892. 170 Hedrich R, Martin I. 1993. Malate induced feedback regulation of plasma membrane anion channels could provide a C02 sensor to guard cells. EMBO J. 12:897-901. Hedrich R, Schroeder II. 1989. The physiology of ion channels and electrogenic pump in higher plants. Annu Rev Plant Physiol Plant Mol Biol 40:539-569. Helper PK, Wayne R0. 1985. Calcium and plant development. Annu Rev Plant Physiol 36:397-439. Hoffrnann B, Kosegarten H. 1995. FITC-dextran form measuring apoplast pH and apoplastic pH gradient between various cell types in sunflower leaves. Physiol Plant 95:327-335. HOpe AB, Findlay GP. 1964. The action potential in Chara. Nature 191:811-812. Horwitz BA, Berrocal GM. 1997. A spectroscopic view of some recent advances in the study of blue light photoreception. Bot Acta 110:360-368. Hosoi S, Lino M, Shimozaki K. 1988. Outward-rectifying Kl” channels in stomatal guard cell protoplasts. Plant Cell Physiol 209:207-211. Hsieh WL, Pierce WS, Sze H. 1991. Calcium-pumping ATPases in vesicles from carrot cells. Plant Physiol 1535-1544. Humble CD, Hsiao TC. 1970. Light-dependent influx and efflux of potassium of guard cells during stomatal opening and closing. Plant Physiol 46:483-487. Hush JM, Overall RL, Newman IA. 1991. A calcium influx precedes organogenesis in Graptopetalum. Plant Cell Environ 14:657-665. Iijima T, Hagiwara S. 1987. Voltage dependent K+ channels in protoplasts of trap-lobe cells of Dionea muscipula. J Membr Biol 100:73-81. Ilan N, Schwartz A, Moran N. 1994. External pH effects on the depolarisation-activated K channels in guard cell protoplasts of Viciafava. J Gen Physiol 103:807-831. Irving HR, Gehring CA, Parish RW. 1992. Changes in cytosolic pH and calcium of guard cells precede stomatal movements. Proc Natl Acad Sci USA 89: 1790-1794. Jentsch TJ, Giinther W. 1997. Chloride channels: an emerging molecular picture. BioEssays 19:117-126. Johannes E, Ermolayeva E, Sanders D. 1997. Red-light~induced membrane potential 171 transients in the moss Physcomitrella patens: ion channels interaction in phytochrome signalling. J Exp Bot 48:599-608. Johannes E, Crofts A, Sanders D. 1998. Control of Cl' efflux in Chara corallins by cytosolic pH, fiee Ca”, and phosphorylation indicates a role of plasma membrane anion channels in cytoplasmic pH regulation. Plant Physiol 118:173-181 Juniper BE. 1976. Geotropisurn. Annu Rev Plant Physiol 27:385-406. Julien JL, Desbies MO, Jaepgher G, Frachisse LM. 1991. J Exp Bot 42:131-137. Kaufman LS. 1993. Transduction of blue-light signals. Plant Physiol 1022333-337. Keller BU, Hedrich R, Raschke K. 1989. Voltage-dependent anion channels in the plasma membrane of guard cells. Nature 341 :450-453. Kelly WB, Esser JE, Schroeder JI. 1995. Effects of cytosolic calcium and limited, possible dual, effects of G protein modulators on guard cell inward potassium channels. Plant J 8:470-489. Keunecke M, Hansen UP. 2000. Different pH-dependences of K+ channel activity in bundle sheath and mesophyll cells of maize leaves. Planta 210:792-800. Khurana JP, Kochhar A, Tyagi AK. 1998. Photosensory perception and signal transduction in higher plants - molecular genetic analysis. Crit Rev Plant Sci 17:465-539. Kinoshita T, Nishimura M, Shimazaki K. 1995. Cytosolic concentration of Ca2+ regulates the plasma membrane HI-ATPase in guard cell of fava bean. Plant Cell 7:1333- 1342. Knight MR, Compbell AK, Smith SM, Trewavas AJ. 1991. Transgenic plant aequorin reports the effects of touch and cold-shock and elicitors on cytoplasmic calcium. Nature 352:524-526. Knight MR, Compbell AK, Smith SM, Trewavas AJ. 1992. Wind-induced plant motion immediately increases cytosolic calcium. Proc Natl Acad Sci USA 89:4467-4471. Knight H, Trewavas AJ, Knight MA. 1997. Calcium signalling in Arabidopsis thaliana responding to drought and salinity. Plant J 114:1327-1334. Kochian LV, Shatt JE, Kuhtreiber WM, Jaffe LF, Lucas WJ. 1992. Use of extracellular, ion selective, vibrating microelectrode system for the quantification of K, H*, and Ca2+ fluxes in maize roots and maize suspension cells. Planta 188:601-610. Kuchitsu K, Kikuyama M, Shibuya N. 1993. N-acetylchitooligosaccharides, biotic 172 elicitor for phytoalexin production induce transient membrane depolarization in suspension-cultured rice cells. Protoplasma 174:79-81. Kura-Hotta M, Enami I. 1981. Light-induced HI efflux from intact cells of Cyanidium caldarium. Plant Cell Physiol 22:1175-1183. Leckie CP, McAinsh MR, Montgomery L, Priestley AJ, Staxen I, Webb AAR, Hetherington AM. 1998. Second messengersinguard cells. J Exp Bot 49:339-349. Lee Y. 1990. Ion movements that control pulvinar curvature in nyctinastic legumes. In RL Satter, HL Gorton, TC Vogelmann, eds, The pulvinus: motor organ for leaf movement. American Society of Plant Physiologists, Rockville, MD, pp 130-141. Leigh RA, Wyn Jones RG. 1984. A hypothesis relating critical potassium concentrations for growth to the distribution and functions of this ions in the plant cell. New Phytol 97:1-13. Leigh RA. 1993. Perception and transduction of stress by plant cells. In L Fowden, T Mansfield, J Stoddart, eds, Plant adaptation to environmental stress. Chapmem and Hall, pp223-237. Lemliri-Chlieh F, MacRobbie EAC. 1994. Role of calcium in the modulation of Vicia guard cell potassium channels by abscisic acid: a patch clamp study. J Membr Biol 137:99-107. Leonard RT, Hepler PK, eds. 1990. Calcium in plant growth and development. American Society of Plant Physiologists, Rockville, MD. Leung J, Giraudat J. 1998. Abscisic acid signal transduction. Annu Rev Plant Physiol Plant Mol Biol 49:199-222. Lew RL, Serlin BS, Schauf CL, Stockton ME. 1990. Calcium activation of Mougeotia potassium channels. Plant Physiol 92:831-836. Lewis BD, Karlin-Neumann G, Davis RW and Spalding EP. 1997. Cay-activated anion channels and membrane depolarization induced by blue light and cold in Arabidopsis seedlings. Plant Physiol 114:1327-1334. Linder B, Raschke K. 1992. A slow anion channel in guard cells, activating at large hyperpolarization, may be principal for stomatal closing. FEBS Lett 313:27-30 Linnemeyer PA, Van Volkenburgh E, Cleland RE. 1990. Characterization and effect of light on the plasma membrane H-ATPase of bean leaves. Plant Physiol 94:1671-1676. Lino B, Baizabal-Aguirre VM, de la Vara LEG. 1998. The plasma membrane H-ATPase 173 from beet root is inhibited by a calcium-dependent phosphorylation. Planta 204:352- 359. Long JM, Widders IE. 1990. Quantification of apoplastic potassium content by elution analysis of leaf lamina tissue from pea (Pisum sativum L. cv. Argenteum). Plant Physiol 94: 1040- l 047. Lucas WJ, Kochian LV. 1986. Ion transport processes in corn roots: an approach utilizing microelectrode techniques. In WD Gensler, ed, Advanced agricultural instrumentation, design and use. Martinus Nijhoff, Dordrecht, The Netherlands, pp402. Lunevsky VZ, Zherelova OM, Vostrikov IY, Berestovsky GN. 1983. Excitation of Characean cell membranes as a result of activation of calcium and chloride channels. J Membr Biol 72:43-58. . Lurin C, Geelen D, Barbier—Brygoo H, Guem J, Maurel C. 1996. Cloning and functional expression of a plant voltage-dependent chloride channel. Plant Cell 8:701-711 Lfittge U, Pitrnan MG. 1976. Transport in Plant H, Part A, Cell; Part B, Tissues and Organs. Encycl Plant Physiol, New Series, Vol 2, Springer-Verlag, Berlin. MacRobbie EAC. 1988. Control ion flux in stomatal guard cells. Bot Acta 101:140- 148. MacRobbie EAC. 1992. Calcium and ABA-induced stomatal closure. Phil Trans R. Soc Lond B 338:5-18 Marré E. 1979. Fusicoccin: a tool in plant physiology. Annu Rev Plant Physiol 30:273- 288. Marré E, Beffagna N, Romani G. 1987. Potassium transport and regulation of intracellular pH in Elodea densa leaves. Bot Acta 1:24-31. Marré MT, Albergoni F G, Moroni A, Marré E. 1989. Light-induced activation of electrogenic H+ extrusion and K uptake in Elodea densa depends on photosynthesis and is mediated by the plasma membrane HI-ATPase. J Exp Bot 40:343-352 Marten I, Lohse G, Herich R. 1991. Plant growth hormones control voltage-dependent activity of anion channels in the plasma membrane of guard cells. Nature 353:758-762. Mayer WE, Hohloch C, Kalkuhl A. 1997. Extensor protoplasts of the Phaseolus pulvinus: light-induced swelling may require extracellular Ca” influx, dark-induced shrinking inositol 1,4,5-triphosphate-induced Ca2+ mobilization. J Exp Bot 307:219-228. McAinsh MR, Brownlee C, Hetherington AM. 1990. Abscisic acid-induced elevation of 174 guard cell cytosolic Ca2+ precedes stomatal closure. Nature 343:186-188. McAinsh MR, Brownlee C, Hetherington AM. 1997. Calcitun ions as second messengers in guard cell signal transduction. Physiol Plant 100:16-29. McAinsh MR, Hetherington AM. 1998. Encoding specificity in Ca2+ signaling systems. Trends Plant Sci 3:32-36. Memon AR, Saccmani M, Glass ADM. 1985. Efficiency of K+ utilization by barley varieties: the role of subcellular compartrnentation. J Exp Bot 36:1860—1876. Merrett MJ, Nimer NA, Dong LF. 1996. The utilization of bicarbonate ions by the marine microalga Nanochlorepsis oculate (Droop) Hibberd. Plant Cell Environ 19:478- 484. De Michelis MI, Rasi-Caldogne, Pugliarello MC. 1992. The plasma membrane Ca2+ pump: potential role in Ca2+ homeostasis. In CM Karssen, LC van Loon, D Vreugdenhil, eds, Progress in plant grth regulation. Kluwer Acad, Netherlands, pp 675-685. Miedema H, Romano LA, Assmann SM. 2000. Kinetic analysis of the KI-selective outward rectifier in Arabidopsis mesophyll cells: a comparison with other plant species. Plant Cell Physiol 41 :209-217. Miller AJ, Vogg G, Sanders D. 1990. Cytosolic calcium homeostasis in fungi: roles of plasma membrane transport and intracellular sequestration of calcium. Proc Natl Acad Sci USA 87:9348-9352. Mimura T, Tazawa M. 1986. Light-induced membrane hyperpolarization and adenine nucleotide levels in perfused Characean cells. Plant Cell Physiol 27:319-330. Mimura T, Shimmen T. 1994. Characterization of the Cay-dependent Cl' efflux in perfused Chara cells. Plant Cell Physiol 35:793-800. Mitchell P. 1961. Coupling of phosphorylation to electron and hydrogen transfer by a chemi-osmatic type of mechanism. Nature 191:144-148. Mitchell P. 1976. Vectical chemistry and the molecular mechanics of chemosmatic coupling: power transmission by proticity. Biochemical Society Transaction 4:399-430. Mitchell P. 1985. The correlation of chemistry and osmatic forces in biochemistry. J Biochem 97:1-18. Moran N, Ehrenstein G, Iwasa K, Bare C, Satter RL. 1988. Potassium channels in motor cells of Samanea saman with inorganic ions: a patch-clamp study. Plant Physiol 88:643- 648. 175 Moran N, Fox D, Satter RL. 1990. Interaction of the depolarization-activated KI channel of Samanea saman with inorganic ions: 3 patch-clamp study. Plant Physiol 94:424-431. Muhling KH, Plieth C, Hansen UP, Sattelrnacher B. 1995. Apoplastic pH of intact leaves of Viciafava as influenced by light. J Exp Bot 46:377-382. Mummert H, Gradmann D. 1991. Action potentials in Acetabularia: measurement and stimulation of voltage-gated fluxes. J Membr Biol 124:266-273. Netting AG. 2000. pH, abscisic acid and the integration of metabolism in plants under stressed and non-stressed conditions: cellular response to stress and their implication for plant water relations. J Exp Bot 51:147-158. Neuhaus G, Bowler C, Kern R, Chua NH. 1993. Calcium/cahnodulin-dependent and - independent phytochrome signal transduction pathways. Cell 73:937-952. Neumann J, Levine RP. 1971. Reversible pH changes in cells of Chlamydomonas reinhardtii resulting from CO2 fixation in the light and its evolution in the dark. Plant Physiol 47:700-704. Newman IA, Kochian LV, Grusak MA, Lucas WJ. 1987. Fluxes of H+ and K+ in corn roots. Plant Physiol 84:1177-1184. Nishizaki Y. 1992. Effects of inhibitors on the light-induced changes in electric potential in pulvinar motor cells of Phaseolus vulgaria L. Plant Cell Physiol 33:1073- 1078. Nobel PS. 1991. Physiochemical and Environmental Plant Physiology. Academic Press, Inc. pp110-139. Novak VA, Ivankina NG. 1975. Comparative study of light-induced changes of electric potentials in plants. F iziol Rast 22:36-40 Okihara K, Ohkawa T, Tsutsui I, Kasai M. 1991. A Ca2+- and voltage dependent-Cl sensitive anion channel in the Chara plasmalemma: a patch clamp study. Plant Cell Physiol 32:593-601. Palmgren MG. 1998. Proton gradients and plant growth: role of the plasma membrane H"-ATPase. Adv Bot Res 28:2-70. Palmgren MG, Harper MG. 1999. Pumping with plant P-type ATPases. J Exp Bot 50:883-893. Palmer JM, Short TW, Gallagher S, Briggs WR. 1996. Blue light-induced 176 phosphorylation of a plasma membrane-associated protein in Zea mays L. Plant Physiol 102:1211-1218. Pickard BG. 1971. Action potentials resulting from the mechanical stimulation of pea epicotyls. Planta 97:106-115. Pickard BG. 1973. Action potentials in higher plants. Bot Rev 39:172-201. Plieth C, Sattehnacher B, Hansen UP. 1998. Light-induced cytosolic calcium transients in green plant cells. H. the effect on the K+ channel as studied by a kinetic analysis in Chara corallina. Planta 207:52-59. Poovaiah BW, Reddy ASN. 1993. Calcium and signal transduction in plants. Crit Rev Plant Sci 12:185-211. Pineros M, Tester M. 1995. Characterization of a voltage-dependent Cay-selective channel from wheat roots. Planta 195:478-488. Pineros M, Tester M. 1997. Calcium channels in high plant cells: selectivity, regulation and pharmacology. J Exp Bot 48: 551-557 Pressman BC. 1976. Biological applications of ionophores. Annu Rev Biochem. 45:501-530. Prins HBA, Snel JFH, Zanstra PE, Helder RJ. 1982. The mechanism of bicarbonate assimilation by the polar leaves of Potamogeton and Elodea: CO2 concentrations at the leaf surface. Plant Cell Environ 52207-214. Pyatygin SS, Opritov VA, Khudyakov VA. 1992. Subthreshold changes in excitable membranes of Cucurbis pepo L. stem cells during cooling-induced action potential generation. Planta 186:161-165. Quail PH, Boylan MT, Parks BM, Short TW, Yu Y, Wagner D. 1995. Phytochromes: Photosensory perception and signal transduction. Science 268:675-680. Quail PH, Hershey HP, Idler KB, Sharrock RA, Charistensen AH. 1991. Phy-gene structure, evolution, and expression. In B Thomas, CB Johnson, eds, Phytochrome properties and biological action. NATO ASI Series, H50, Berlin:Springer-Verlag, pp 13- 38. Quarmby LM. 1994. Signal transduction in sexual life of Chlarnydomonas. Plant Mol Biol 26:1271-1287. Rasi-Caldogno F, Pugliarello MC, De Michelis MI. 1987. The Cay-transport ATPase of plat plasma membrane catalyzes an nH+/Ca2+ exchange. Plant Physiol 83:994-1000. 177 Ray LB. 1999. Signal transduction. Science 284:755-756. Rea PA, Poole RJ. 1993. Vacuolar HI-translocation pyrophosphate. Annu Rev Plant Physiol Plant Mol Biol 44:157-180. Read ND, Shacklock PS, Knight MR, Trewavas AJ. 1993. Imaging calcium dynamics in living plant cells and tissues. Cell Biol Int Rep 17:111-125. Reid RJ, Smith FA. 1992. Regulation of calcium influx in Chara: effects of K‘“, pH, metabolic inhibition, and calcium channel blockers. Plant Physiol 100:637-643. Reid RJ, Tester MA, Smith FA. 1997. Voltage control of calcium influx in intact cells. Aust J Plant Physiol 24:805-810. Remis D, Treffny B, Gimmler H. 1994. Light-induced H+ transport across the plasma membrane of the acid-resistant green alga Dunaliella acidophila. Plant Physiol Biochem 32:75-84. Retivin VG, Opritov VA. 1987. Cable properties of the stem of higher plants. Fiziol Rast 34:5-12. Rhodes JD, Thain JF, Wildon DC. 1996. The pathway for systemic electrical signal conduction in the wounded tomato plant. Planta 200:50-57. Ries SK, Widders I, Baughan R. 1995. Light affects the flux of Ca2+ from excised tomato leaves within four seconds. J Exp Bot 46: 1 867-1 876. Roberts DM, Lukas TJ, Watterson DM. 1986. Structure, function, and mechanism of action of calmodulin. Crit Rev Plant Sci 4:311-335. Rodriguez-Navarro A, Blatt MR, Slayman CL. 1986. A potassium-proton symport in Neurospora crassa. J Gen Physiol 87:649-674. Roelfsema MRG, Prins HRA. 1998. The membrane potential of Arabidopsis thaliana guard cells: depolarizations induced by apoplastic acidification. Planta 205:100-112. Roos W. 2000. Ion mapping in plant cells — methods and applications in signal transduction research. Planta 347-370. Rudolph HK, Antebi A, Fink GR, Buckley CM, Dorman TE, LeVitre J, Davidow LS, Mao JI, Moir DT. 1989. The yeast secretory pathways is perturbed by matantions in PMRI, a member of Ca2+ ATPase family. Cell 58:133-145. Ryan CA, farmer EE. 1991. Oligosaccharide signals in plants: A current assessment. 178 Annu Rev Plant Physiol Plant Mol Biol 42:651-674. Ryan PR, Newman IA, Arif I. 1992. Rapid calcium exchange for proton and potassium in cell walls of Chara. Plant Cell Environ 15:675-683. Sanders D. 1980. The mechanism of Cl' transport at the plasma membrane of Chara corallina. I. cotransport with Cl'. J Membr Biol 53:129-141. Sanders D, Brownlee C, Harper JF. 1999. Communicating with calcium. Plant Cell 11:691-706. Sanders D, Smith FA, Walker NA. 1985. Proton/chloride cotransport in Chara: mechanism of enhanced influx after rapid external acidification. Planta 163:411-418. Satter RL, Morse MJ, Lee Y, Carian RC, Cote G, Moran N. 1988. Light and clock- controlled leaflet movements in Samanea saman: a physiological, biophysical and biochemical analysis. Bot Acta 101:205-213. Schmidt C, Schroeder II. 1994. Anion-selectivity of slow anion channels in Vicia fava guard cells: large nitrate permeability. Plant Physiol 106:383-391. Schroeder JI. 1988. K+ transport properties K+ channels in the plasma membrane of Viciafava guard cells. J Gen Physiol 92:667-683. Schroeder JI. 1989. A quantitative analysis of outward rectifying K” channels in guard cell protoplasts from Viciafava. J Membr Biol 107:229-235. Schroeder H. 1992. Plasma membrane ion channel regulation during abscisic acid- induced closing of stomata. Philos Trans R Soc London B Ser 338:83-89. Schroeder JI. 1995. Anion channels as central mechanisms for signal transduction in guard cells and putative functions in roots for plant-soil interactions. Plant Mol Biol 28:353-361. Schroeder JI, Hagiwara S. 1989. Cytosolic calcium regulates ion channels in the plasma 7 membrane of Viciafava guard cells. Nature 338:427-430. Schroeder JI, Hagiwara S. 1990. Repetitive increase in cytosolic Ca2+ of guard cells by abscisic acid activation of nonselective CaZI-permeable channels. Proc Natl Acad Sci USA 87:935-9309. Schroeder JI, Hedrich R. 1989. Involvement of ion channels and active transport in osmoregulation and signalling of higher plant cells. Trends Biochem Sci 14:187-192. Schroeder JI, Keller BU. 1992. Two types of anion channels currents in guard cells with 179 distinct voltage regulation. Proc Natl Acad Sci USA 89:5025-5029. Schroeder JI, Raschke K, Neker E. 1987. Voltage dependence of K” channels in guard- cell protoplasts. Proc Natl Acad Sci USA 84:4108-4112. Schumaker KS, Sze H. 1986. Calcium transport into the vacuole of oat roots. Characterization of H+/Ca2+ exchange activity. J Biol Chem 261 12172-12178. Schumaker KS, Sze H. 1987. Inositol 1,4,5-triphosphate releases Ca2+ from vacuolar membrane vesicles of oat roots. J Biol Chem 262:3944~3946. Sentenac H, Bonneaud N, Minet M, Lacroute F, Salmon J-M, Gaymard F, Grignon C. 1992. Cloning and expression in yeast of a plant potassium ion transport system. Science 256:663-665. Serrano R. 1985. Plasma membrane ATPase of plants & fimgi. CRC press. Boca Raton, 174 pp. Serrano EE, Zeiger E, Hagiwara S. 1988. Red light stimulates an electrogenic proton pump in Vicia guard cell protoplasts. Proc Natl Acad Sci USA 85:436-440. Serrano EE, Zeiger E. 1989. Sensory transduction and electrical signaling in guard cells. Plant Physiol 91:795-799. Shabala NS, Newman IA. 1997. Proton and calcium flux oscillations in the elongation region correlate with root mutation. Physiol Plant 100:917-926. Shabala NS, Newman IA, Morris J. 1997. Oscillations in H+ and Ca2+ fluxes around the elongation region of corn and effects of external pH. Plant Physiol 113:111-118. Shabala NS, Newman IA. 1999. Light-induced changes in hydrogen, calcium, potassium, and chloride ion fluxes and concentration from the mesophyll and epidermal tissues of bean leaves. Understanding the ionic basis of light-induced bioelectrogensis. Plant Physiol 119:1115-1124. Shacklock PS, Read ND, Trewavas AJ. 1992. Cytosolic free calcium mediates red light- induced photomorphologensis. Nature 358:753-755. Sharma VK, Jain PK, Malik MK, Maheshwari SC, Khurana JP. 1997. Light- and calcium-modulated phosphorylation of proteins from wheat seedlings. Phytochemistry 44:781-786. Sharrock RA, Quail PH. 1989. Novel phytochrome sequences in Arabidopsis thaliana: structure, evolution, and differential expression of a plant regulatory photoreceptor family. Genes Dev 3:1745-1757. 180 Shiina T, Tazawa M. 1986. Action potential in Lufla cylindlica and its effects on elongation growth. Plant Cell Physiol 27: 1081-1089. Shimazaki K, Omasa K, Kinoshita T, Nishimura M. 1993. Properties of the signal transduction pathway in the blue light response of stomatal guard cells of Vicia fava and Commelina benghalensis. Plant Cell Physiol 34:1321-1327 Shimazaki K, Zeiger E. 1987. Red light-dependent CO, uptake and oxygen evolution in guard cell protOplasts of Viciafava L.: evidence for photosynthetic CO2 fixation. Plant Physiol 84:7-9. Short TW, Briggs WR. 1990. Characterization of a rapid, blue light-mediated change in detectable phosphorylation of a plasma membrane protein from etiolated pea (Pisum sativum L.) seedlings. Proc Natl Acad Sci USA 93:8129-8133. Short TW, Porst M, Briggs WR. 1991. On the trail of the photoreceptor for phototropism in higher plants. In F Lenci, G Columbetti, DP Hader, PS Song, eds, Biophysics of Photoreceptors and Photomovement in Microorganisms. New York, Plenum, pp 173-180. Short TW, Briggs WR. 1994. The transduction of blue light signals in higher plants. Annu Rev Plant Physiol Plant Mol Biol 45:143-171. Sibaoka T. 1966. Action potentials in plant organs. Symp Soc Exp Biol 20:49-74. Simons PJ. 1981. The role of electricity in plant movements in higher plants. Annu Rev Plant Physiol 87:11-37. Smith FA, Walker NA. 1976. Chloride transport in Chara corallina during acid loads. J Exp Bot 27:451-459. Spalding EP, Cosgrove DJ. 1992a. Rapid alterations in growth rate and electrical potentials upon stem excision in pea seedlings. Planta 187:523-531. Spalding EP, Cosgrove DJ. 1992b. Mechanism of blue-light-induced plasma membrane depolarization in etiolated cucumber hypocotyls. Planta 188:199-205. Spalding EP, Slayman CL, Goldsmith MHM, Gradmann D, Bertl A. 1992. Ion channels in Arabidopsis plasma membrane . Transport characteristics and involvement in light- induced voltage changes. Plant Physiol 99:96-102. Staal M, Elzenga JTM, van Elk AG, Prins HBA, Van Volkenburgh E. 1994. Red and blue light-stimulated proton efflux by epidermal leaf cells of the argenteum mutant of Pisum sativum. J Exp Bot 45:1213-1218. I81 Stahlberg R, Cosgrove DJ. 1997. The propagation of slow wave potential in pea epicotyls. Plant Physiol 113:209-217. Stahlberg R, Van Volkenburgh E. 1999. The effect of light on membrane potential, apoplastic pH and cell expansion in leaves of Pisum sativum L. var. Argenteum. Planta 208:188-195. Stankovic B, Davies E. 1998. The wound response in tomato involves rapid growth and electrical responses, systemically up-regulatory transcription of proteinase inhibitor and cahnodulin and down-regulated translation. Plant Cell Physiol 39:268-274. Stankovic B, Witters DL, Zawadzki T, Davies E. 1998. Action potential and variation potential in sunflower: an analysis of their relationships and distributing characteristics. Physiol Plant 103:51-58. Stein WD. 1990. Channels, carriers and pumps: an introduction to membrane transport. San Diego, CA, Academic Press. Stevenson JM, Perera IY, Heihnann I, Persson S, Boss WF. 2000. Inotitol signalling and plant growth. Trends Plant Sci 5:252-258. Suh S, Moran, Lee Y. 2000. Blue light activates potassium-efflux channels in flexor cells from Samanea saman motor organs via two mechanisms. Plant Physiol 123:833- 843. Szarak I, Trebacz K. 1999. The role of light-induced membrane potential changes in guttation in gametophytes of Asplenium trichomanes. Plant Cell Physiol 42:1280-1286. Sze H. 1985. HI-translocating ATPases: Advances using membrane vesicles. Annu Rev Plant Physiol 36:175-208. Sze H. Ward JM, Lai S. 1992. Vacuolar H+-ATPases fi'om plants. J Bioenerg Biomembr 24:371-381. Sze H, Li X, Pahngren MG. 1999. Energization of plant cell membranes by Hi-pumping ATPases: regulation and biosynthesis. Plant Cell 11:677-690. Takagi S, Nagai R. 1988. Light-affected Ca2+ fluxes in protoplasts from Vallisneria mesophyll cells. Plant Physiol 88:228-323. Takeshige K, Mitsumori F, Tazawa M, Mimura T. 1992. Role of cytoplasmic inorganic phosphate in light-induced activation of H‘-pumps in the plasma membrane and tonoplast of Chara corallina. Planta 186:466-472. 182 Tazawa M, Shimmen T, Timura T. 1987. Membrane control in the Characea. Annu Rev Plant Physiol 38:95-116. Terry BR, Tyerman SD, Findlay GP. 1991. Ion channels in the plasma membrane of Amaranthus protoplasts: one cation and one anion channel dominate the conductance. J Membr Biol 121:223-236. Thain JF Wildon JF. 1993. Electrical signaling in Plants. In J Schultz, I Raskin, eds, Plant signals in interaction with other organisms. Rochkville, MD, American Society of Plant Physiologists, pp 102-115. Theologis A. 1986. Rapid gene expression by auxin. Annu Rev Plant Physiol 37:407- 438. Thiel G, Wolf A H. 1997. Operation of KI-channels in stomatal movement. Trends Plant Sci 2:339-343. Thion L, Mazars C, Thuleau P, Graziana A, Rossignol M, Moreau M, Ranjeva R. 1996. Activation of plasma membrane voltage-dependent calcium-channels by disruption of microtubles in carrot cells. FEBS Lett 340:45-50. Thompson WF, White MJ. 1991. Physiological and molecular studies of light-regulated nuclear genes in higher plants. Annu Rev Plant Physiol Plant Mol Biol 42:423-466. Thuleau P, Ward JM, Ranjeva R, Schroeder II. 1994. Voltage-dependent calcium- perrneable channels in the plasma membrane of a higher plant cell. The EMBO J 13:2970-2975. Trebacz K, Sievers A. 1998. Action potentials evoked by light in traps of Dionaea muscipula Ellis. Plant Cell Physiol 39:369-372. Trebacz K, Simonis W, Schonknecht G. 1994. Cytoplasmic Ca2+ KI, Cl‘, and NO,‘ activities in the liverwort Conocephalum conicum L. at rest and during action potential. Plant Physiol 106: 1073- 1 084. Trebacz K, Zawadzki T. 1985. Light-triggered action potentials in liverwort Conocephalum conicum. Physiol Plant 75:24-30. Trewavas AJ, Gilroy S. 1991. Signal transduction in plant cells. Review 7:354—361. Trewavas AJ, Knight M. 1994. Mechanical signaling, calcium and plant form. Plant Mol Biol 26:1329-1341. Trewavas AJ, Malho R. 1998. Ca2+ signaling in plant cells: the big network! Curr Opinion Plant Biol 1:428-433. 183 Trewavas AJ. 1999. Le calcium, C’est la vie: calcium makes waves. Plant Physiol 120:1-6. Tsien RW, Hess E, McCleskey EW, Rosenberg RL. 1987. Calcium channels: mechanism of selectivity, permeation and block. Annu Rev Biophys Chem 162265-290. T yerrnan SD, 1992 Anion channels in plants. Annu Rev Plant Physiol plant Mol Biol, 43:351-373. Tyerman SD, Findley GP, Paterson GJ. 1986. Inward membrane current Chara inflata. 11. Effects of pH, Cl‘ channel blockers and NH}, and significance for hyperpolarised sate. J Membr Biol 89:153-161. Ulhich CI, Guem J. 1990. Ion fluxes and pH changes induced by trans-plasmalemma electron transfer and firsicoccin in Lemna gibba G1. Planta 180:390-399. Van Volkenburgh E, Cleland RE. 1980. Proton excretion and cell expansion in bean leaves. Planta 1482273-278. Van Volkenburgh E, Cleland RE, Watanabe M. 1990. Light-stirnulated cell expansion in bean (Phaseolus vulgaris L.) leaves, 11 quantity and quality of light required. Planta 182:77-80. Verhey SD, Lomax TL. 1993. Signal transduction in vascular plants. J Plant Growth Regul 12:179-195. Ward JM, Schroeder JI. 1997. Roles of ion channels in initiation of signal transduction in higher plants. In Signal transduction in plants. P Aducci, ed, Brikhéiuser Verlag Basel, Switzerland, pp 1-22. Wayne R. 1993. Excitability in plant cells. America Scientist 81:140-151. Wayne R. 1994. The excitability of plant cells: with special emphasis on Characean intemodal cells. Bot Rev 60:265-367. Webb AAR, Taylor JE, McAinsh MR, Hetherington AM. 1996. Calcium ion as intracellular second messengers in plants. Adv Bot Res 22:45-96. Weiler EW, Meyer C, Oecking C, Feyerabend M, Mithofer A. 1990. The fusicoccin receptor of higher plant. In CL Lamb, RN Beachy, eds, Plant gene transfer. Wiley-Liss, New York, pp 153-164. Weisenseel MH, Ruppert I-IK. Phytochrome and calcium ions are involved in light- induced membrane depolarization in Nitella. Planta 137:225-229. 184 White PJ. 1993. Characterization of a high-conductance, voltage-dependent cation channel from the plasma membrane of rye roots in planer lipid bilayers. Planta 191:541- 551. White PJ. 1994. Characterization of a voltage-dependent cation channel from the plasma membrane of rye (Secale carele L.) roots in planer lipid bilayers. Planta 191:186-193. White PJ. 1997. Voltage-dependent Ca2+ uptake by right-side-out plasma membrane vesicles derived from maize shoots. J Plant Physiol 152:17-24. White PJ. 1998. Calcium channels in the plasma membrane of root cells. Ann Bot 81:173-183. Wildon DC, Doherty HM, Eagles G, Bowles DJ, Thain JF. 1989. Systemic responses arising from localized heat stimuli in tomato plants. Ann Bot 64:691-695. Wildon DC, Thain JF, Minchin PEH, Gubb IR, Reilly AJ, Skipper YD, Doherty HM, O’Donnell PJ, Bowles DJ. 1992. Electrical signaling and systemic proteinase inhibitor induction in the wounded plant. Nature 360:62-65. Williams LE, Schueler SB, Briskin DP. 1990. Further characterization of the red beet plasma membrane Ca2+-ATPase using GTP as an alternative substrate. Plant Physiol 92:747-754. Xing T, Higgins VJ, Blumwald E. 1996. Regulation of plant defense response to fimgal pathogens: two types of protein kinase in the reversible phosphorylation of the host plasma membrane H +-ATPase. Plant Cell 82555-564. Yang Z. 1996. Signal transduction proteins in plants: an overview. In DPS Verama, ed, Signal transduction in plant grth and development. Springer Wien New York, ppl-31. Zeiger E. 1990. Light perception in guard cells. Plant Cell Environ 13:739-747. Zeiger E. 1994. The photobiology of stomatal movements. In Kendrick RE, Kronenberg GI-IM, eds, Photomorphogenesis in Plant. Kluwer, Dordrecht, p 683-706. Zeiger E, Bloom AJ, Helper PK. 1978. Ion transport in stomatal guard cells, A chemiosmotic hypothesis: What’s new. Plant Physiol 77:29-32. Zeiger E, Lino M, Shimazaki K, Ogawa T. 1987. The blue-light response of stomata: mechanism and function. In E Zeiger E, GD Farquhar, IR Cowan, eds, Stomatal function. Stanford University Press, Stanford, pp 209-227. 185 Zhen RG, Kim B, Rea PA. 1997. The molecular and biochemical basis of pyrophosphate-energized ion translocation at the vacuole membrane. Adv Bot Res 27:297-337. ‘ 186