a" a an» a 5.41:5: . a , ..fi .35, .... ax!“ 5, in...“ «3:. it! 4 1.- li: A .4. 3')...- .H. . - unit. .a: (Jan. “-9... Ex. 2 fix: LT. . ‘Y a. . 5.. v I: . .3“... 9; r. "HESlS MICHIG l NWEHSITY LIBRARIES ll ill/IIll/lI/lI/II/l/I/II/I/lIll/Ill!!!” NIH/Ill”! This is to certify that the dissertation entitled GROWTH-DEPENDENT ALUMINUM UPTAKE AND DISCRETE LOCALIZATION IN CULTURED CELLS OF Nicotiana tabacum L. cv.! BY-Z: A POSSIBLE ROLE FOR MEMBRANE TRAFFIC AT THE CELL SURFACE presented by Victor Ao Vitorello has been accepted towards fulfillment of the requirements for PhoDo degree in Botany and Plant Pathology Palfggaw dot/«>14. t/Héag/ Major professor Date OQJOSXqu MS U i: an Affirmative Action/Equal Opportunity Institution 0-12771 LIBRARY Michigan State University PLACE N RETURN BOX to remove this checkout from your record. TO AVOID FINES rotum on or More data duo. DATE DUE DATE DUE DATE DUE GROWTH-DEPENDENT ALUMINUM UPTAKE AND DISCRETE LOCALIZATION IN CULTURED CELLS OF Nicotiana tabacum L. cv. BY-2: A POSSIBLE ROLE FOR MEMBRANE TRAFFIC AT THE CELL SURFACE By Victor A. Vitorello A DISSERTATION Submitted to Michigan State University in partial fiflfillment of the requirements for the degree of DOCTOR OF PHILOSOPHY Department of Botany and Plant Pathology 1 996 ABSTRACT GROWTH-DEPENDENT ALUMINUM UPTAKE AND DISCRETE LOCALIZATION IN CULTURED CELLS OF Nicotiana tabacum L. cv. BY-2: A POSSIBLE ROLE FOR MEMBRANE TRAFFIC AT THE CELL SURFACE By Victor A. Vitorello Short-term aluminum (Al) uptake and cellular localization of Al were studied in tobacco (Nicotiana tabacum L.) cv. BY-Z cultures, by using an in viva Al-morin fluorescence assay and graphite furnace atomic absorption spectrometry to detect Al. Results from these two methods were highly correlated (r2=0.94). Al uptake was dependent on cell growth. In growing cells, Al uptake was detectable within 5 min and had an initial rate of 16 nmol Al (10" cells) " h'l , but was negligible in stationary-phase cells. With respect to cellular localization, Al accumulated in discrete regions (2-30 pm long) at the cell periphery, usually one region per cell. Frequently, Al-morin fluorescence occurred along the interface (20-30 pm) of adjacent cells, as if along the plane of cellular division. The proportion of cells displaying localized Al varied from 0 to about 80%. Al-morin fluorescence was frequently associated with a protuberance of the cell surface and with vesicle-like or globular structures attached to the exterior surface of the cell and presumably formed in response to Al treatment. The plasma membrane and the cell wall tended to remain combined at regions of Al-morin fluorescence, following plasmolysis, but Al was found solely associated with the protoplast in rare cases of separation. The hypothesis that Al-accumulating regions correspond to regions of cell growth is raised. Increased external CaClz, low temperature (4°C), sorbitol (0.1 and 0.4 M), and pre-chelation of Al to citrate all largely reduced Al uptake (by 75-90%). Low pH (4.5) caused an increase in the percentage of cells permeable to trypan blue, in contrast to pH 5.6 where almost all cells were impermeable. This low pH- induced increase in membrane permeability was dependent on cell growth, it could be ameliorated by factors such as increased CaCl2 and it may play an important role in Al uptake. The inhibitors cremart (15-30 uM), brefeldin A (10 uM), caffeine (1-5 mM) and 2,6- dichlorobenzonitrile (20 uM) had no detectable effect on Al uptake. Growing cells lost their capacity for Al uptake when incubated in a minimal medium (0.5 to 3 h) prior to exposure to Al. However, Al uptake was sustained by treatment of cells with low concentrations (10 uM) of brefeldin A, an inhibitor of the Golgi secretory pathway. A physiological change in these cells, associated with vesicular traffic, is apparently necessary for cessation of Al uptake, but is reversibly blocked by brefeldin A. To Claudia, Tomas and Gabriela, To my parents and sisters, To those who love and truly dedicate themselves to mentoring. iv ACKNOWLEDGMENTS I would like to thank Dr. Alfied Haug, my research advisor, without whom this work would not have been possible. I am also grateful to Dr. Robert Bandurski, my academic advisor, and Dr. Patrick Hart and Dr. John Ohlrogge of my guidance committee. I wish to express my appreciation for the financial support from the Brazilian federal granting agency CNPq, grant # 200892/91-6 and by the University of 850 Paulo, Brazil. I also thank additional financial support from the Department of Botany and Plant Pathology and the Graduate School at MSU. My thanks to those who helped me in several ways. Dr. Robert Hausinger, Dr. Matthew Zabik, Dr. Alan Jones, Dr. Leah Bauer and Debbie Miller for use of equipment. Dr. Joanne Whallon and Dr. Shirley Owens for help on the confocal microscope. David Shintani and Mike Sullivan for providing me with the tobacco cell cultures. Kurt Stepnitz for the preparation of slides and photographs. Thanks to members of this department and to Dr. De Zoeten for their support and encouragement. I am also grateful to Dr. Holly Schaeffer, for support and critical readings. To my fiiends, I thank Ray Pacovsky, for fi'uitful discussions and encouragement, and Jorge Rodrigues and Maeli Melotto, for their years of friendship and for their always prompt help. I also thank my Brazilian fiiends in East Lansing and at CENA, University of 850 Paulo. Finally, special thanks to Dr. Ken Poff, who helped me change my perspective on science and life, who gave me encouragement, and who. through his example, gave me important lessons on mentoring. TABLE OF CONTENTS LIST OF TABLES ..................................................... ix LIST OF FIGURES .................................................... x ABBREVIATIONS ................................................... xiii CHAPTER 1 INTRODUCTION AND REVIEW OF LITERATURE ....................... 1 Chemistry of Aluminum ...................... y ...................... 2 General aspects of Al toxicity in plants ................................. 5 Aluminum toxicity in animals ....................................... 7 Low pH effects on roots ............................................. 9 A1 uptake in plants ................................................ 10 Sites of Al accumulation in roots ............................... 10 Al uptake at the cellular level: apoplastic or symplastic? ............ 1 1 Factors impacting Al uptake .................................. 13 Mechanisms of Al toxicity .......................................... 14 Al tolerance: mechanisms, genetics and Al-induced genes and proteins ....... 15 Processes involved in cell growth/secretion and Al toxicity ................ 16 REFERENCES .................................................. 21 CHAPTER 2 DEVELOPMENT AND VALIDATION OF AN IN VIVO ALUMINUM-MORIN FLUORESCENCE ASSAY FOR THE VISUALIZATION AND QUANTITATIVE DETECTION OF ALUMINUM IN TOBACCO BY-2 CELL SUSPENSIONS . . . 32 ABSTRACT ..................................................... 32 INTRODUCTION ................................................ 34 MATERIALS AND METHODS ..................................... 37 Cell cultures and Al uptake ................................... 37 Morin solutions and cell loading ............................... 38 Fluorescence intensity measurements ........................... 39 Graphite Furnace Atomic Absorption Spectrometry ................ 40 RESULTS ...................................................... 41 Morin uptake by cells ........................................ 41 Spectra ................................................... 41 Optimization of Al-morin fluorescence intensity .................. 45 Comparison of the Al-morin method to GFAAS ................... 49 vi DISCUSSION ................................................... 53 REFERENCES .................................................. 56 CHAPTER 3 SHORT-TERM ALUMINIUM UPTAKE BY TOBACCO CELLS: GROWTH DEPENDENCE AND EVIDENCE FOR INTERNALIZATION IN A DISCRETE PERIPHERAL REGION ............................................... 58 ABSTRACT ..................................................... 58 INTRODUCTION ................................................ 58 MATERIAL AND METHODS ...................................... 59 Chemicals and labware ...................................... 59 Cell cultures ............................................... 59 Aluminium uptake .......................................... 59 Analysis of Al by graphite furnace atomic absorption spectrometry . . . . 59 Morin fluorescence assay ..................................... 60 RESULTS ...................................................... 6O Dependence of Al uptake on cell growth ......................... 60 Time-course of Al uptake .................................... 60 Aluminium uptake depends on Ca *, temperature and pre-chelation to citrate ................................. 61 Distribution of morin in cells and cellular localization of Al ......... 63 Membrane permeability at low pH and Al uptake .................. 64 DISCUSSION ................................................... 64 REFERENCES .................................................. 66 CHAPTER 4 CESSATION OF ALUMINUM UPTAKE AND ALUMINUM-INDUCED ABNORMALITIES OF THE CELL SURFACE DEPENDS ON BREFELDIN A-SENSITIVE MEMBRANE TRAFFIC IN TOBACCO BY-2 CELLS ......... 67 ABSTRACT ..................................................... 67 INTRODUCTION ................................................ 68 MATERIALS AND METHODS ..................................... 70 Chemicals and labware ...................................... 70 Cell cultures and Al uptake procedure ........................... 70 Treatment of cells with inhibitors of growth-associated processes . . . . 71 Labelling of cells with morin and measurements of fluorescence intensity .................................. 72 Microscopy ................................................ 73 RESULTS ...................................................... 74 Confocal laser scanning microscopy of Al-accumulating regions in cells ....................................... 74 Cells lose their capacity for Al uptake when incubated in a minimal medium ............................................ 82 Effect of inhibitors of growth-associated processes on Al uptake ...... 82 Brefeldin A sustains Al uptake capacity by cells ................... 86 vii DISCUSSION ................................................... 93 REFERENCES ................................................. 100 CHAPTER 5 GENERAL DISCUSSION AND PERSPECTIVES ......................... 104 Future perspectives ........................................ 110 REFERENCES ................................................. 113 APPENDIX A GROWING TOBACCO CELLS RESPOND TO RAPID MEDIUM CHANGES WITH A TRANSIENT INCREASE IN LOCALIZED Ca2+ ACTIVITY ....... 116 APPENDIX B CELLULAR ASPECTS OF ALUMINUM TOXICITY IN PLANTS .......... 130 APPENDIX C ALUMINUM INTERACTION WITH PHOSPHOINOSITIDE-ASSOCIATED SIGNAL TRANSDUCTION ............................................ 137 APPENDIX D ALUMINIUM COORDINATION TO CALMODULIN: THERMODYNAMIC AND KINETIC ASPECTS ............................................. 144 viii LIST OF TABLES Chapter 3. Table 1. Effects of low temperature, pre-chelation of Al to citrate, and pre-wash pH on Al uptake in cultured tobacco cells, 3 days after transfer ....... 62 Table 2. Percentage of cells (3 days after transfer) permeable to trypan blue when incubated in medium at pH 4.5 or pH 5.6, and as affected by Ca2+ concentrations and presence of 20 uM Al. ............................... 64 Table 3. Percentage of cells permeable to trypan blue after 1h Al incubation as affected by growth stage of cells and pH of pre-wash medium .............. 64 Chapter 4. Table 1. Growth of tobacco BY-2 cells following transfer to fresh culture medium, incubation in minimal medium, or culture in the presence of brefeldin A (BF A). . 85 Table 2. The effect of cremart, dichlorobenzonitrile (DCB), caffeine and brefeldin A (BF A) on Al uptake in tobacco BY-2 cells. .................. 87 ix LIST OF FIGURES Chapter 2. Figure 1. Structure of the flavonoid morin (3, 5, 7, 2', 4' - pentahydroxyflavone) and presumed structure of its complex with Al ............................... 36 Figure 2. Uptake of morin into cells of tobacco BY-2 cultures and fluorescence intensity from Al-treated cells as a function of duration of incubation with morin ........................................................ 42 Figure 3. Excitation and emission spectra of free morin and Al-morin in tobacco BY-2 cells (A) and in an aqueous or ethanol/water/acetate solution (B). ........ 43 Figure 4. Dependence of cell Al-morin fluorescence intensity on the concentration of morin in the incubation medium and the cell titer during readings (insert). . . . . 46 Figure 5. Readings of Al-morin fluorescence intensity from tobacco BY-2 cells as affected by movement and sedimentation of cells and stability of the Al-morin fluorescence signal with time (insert). ........................ 48 Figure 6. Correlation between results of Al uptake by tobacco BY-2 cells measured by the Al-morin method or graphite furnace atomic absorption spectrometry (GFAAS). ............................................. 50 Figure 7. Aluminum uptake by tobacco BY—2 cells and by cellulose phophate, at pH 4.5 and 5.6, as reported by Al-morin fluorescence intensity or GFAAS. . . . 52 Chapter 3. Figure 1. Growth stage-dependent Al uptake in cultured tobacco cells .......... 61 Figure 2. Time course of Al uptake by cultured tobacco cells ................. 61 Figure 3. Photomicrographs of fluorescence from morin-labelled tobacco cells in the absence of added Al ........................................ 62 Figure 4. Photomicrographs of fluorescence from morin-labelled tobacco cells in the presence of added Al ....................................... 63 Chapter 4. Figure l. Photomicrographs of Al accumulation in deformities of the cell surface which originated after Al treatment. ........................ 75 Figure 2. Photomicrographs of Al accumulation in structures at the region of contact between two adjacent cells. ............................. 77 Figure 3. Photomicrographs of plasmolyzed, Al-treated cells. ................ 79 Figure 4. Aluminum uptake ceases in tobacco BY-2 cells after incubation in a minimal medium, but not in cells treated with brefeldin A. ............... 83 Figure 5. The effect of the duration of pre-incubation with EPA, prior to incubation in minimal medium, on blockage of the cessation of A1 uptake in tobacco BY—2 cells, by BFA. .................................. 88 Figure 6. Dose dependence curve for the suppression, by brefeldin A, of the cessation of Al uptake by tobacco BY-2 cells, induced by incubation in minimal medium. ................................................. 90 Figure 7. Reversibility of the effects of brefeldin A on loss of Al uptake capacity (A) and its effect after some loss of Al uptake capacity has already occurred (B) in tobacco BY-2 cells. .............................. 91 Chapter 5. Figure l. A schematic representation of a model for uptake of Al into regions of growing cells involved in the incorporation of new cell wall and plasma membrane. .................................................. 108 Appendix A Figure 1. A: Cellular Ca2+ activity, reported by chlorotetracycline fluorescence, as a function of the growth stage of tobacco cultures. ....................... 121 Figure 2. Changes in Ca2+ activity of growing tobacco cells, as reported by chlorotetracycline fluorescence, following incubation in a minimal medium, or maintenance in culture medium. ....................................... 122 Figure 3. Photomicrographs of growing tobacco cells (2 days after subculture) labelled with chlorotetracycline (50 MM, 15 min). ......................... 123 xi Figure 4. Growth of tobacco cells following incubation in different media, measured 24 h after beginning of treatment. .............................. 126 Appendix C Fig. 1 Simplified view of signal transduction along the Ca2‘/phosphoinositide pathway .......................................................... 138 xii ABBREVIATIONS ATP ......................... adenosine triphosphate BF A ......................... brefeldin A CCCP ........................ carbonyl cyanide m-chlorophenylhydrazone DCB ......................... 2,6-dichlorobenzonitrile DMSO ....................... dimethylsulfoxide DNP ......................... 2, 4 dinitrophenol EDTA ........................ ethylenediamine tetraacetic acid GFAAS ...................... graphite furnace atomic absorption spectrometry GTP ......................... guanosine triphosphate Mes .......................... 2-(N-morpholino) ethanesulfonic acid NTA ......................... nitrilotriacetate PCV ......................... packed cell volume PE ........................... phosphatidylethanolamine PS . . . . . . ..................... phosphatidylserine RGR ......................... relative growth rate (% per day) SE ........................... standard error xiii CHAPTER 1 INTRODUCTION AND REVIEW OF LITERATURE INTRODUCTION AND REVIEW OF LITERATURE Aluminum toxicity is a major soil constraint to food and biomass production throughout the world (F oy et al., 1978; Haug, 1984; Godbold et al., 1988). Its importance stems from the area covered, effects on crop productivity and often difficult correction. Estimates indicate that 40% of the arable soils of the world are acidic and may therefore present aluminum toxicity hazards (Osmond et al., 1980). Most of these areas are located in third world countries in South America, central Africa and southeast Asia. In Brazil, 580 million hectares are affected by acidic soils, covering 70% of its total territory. This is the largest area of acidic soils within a single country and corresponds almost to 75% of the continental USA. The problem is complicated by the difficulty in liming subsoil layers or forested areas (F oy et al., 1978), even when the farmer does have access to the necessary capital and machinery. Although Al toxicity has received considerable attention over the last 30 years, it is still poorly understood. There are numerous, but unsubstantiated, theories on the causes of Al toxicity in the literature (Kochian, 1995; Jones and Kochian, 1995). The cellular components and processes which have been reported to be affected by A1 are wide ranging, such as mitosis and cell division (Matsurnoto et al., 1976), physical properties and structure of the plasma membrane (V ierstra and Haug, 1978; Zhao et al., 1987), uptake of Ca2+ and other ions (Huang et al., 1992), phosphoinositide-mediated signal transduction (Haug et al., 1994 - Appendix B; Jones and Kochian, 1995), and actin and tubulin (MacDonald et al., 1 1987; Grabski and Schindler, 1995). Given the potential for Al to interact with numerous crucial cellular components such as ATP, GTP and phosphates, and thereby disrupt Mg-associated functions (Petterson and Bergman, 1989), the search for the site(s) of toxicity may reveal numerous targets. To illustrate, Crapper Mclachlan (1989) compiled a partial list of 46 processes and functions reportedly damaged by Al. Many of these processes involved phosphate metabolism and many were performed in vitro. That is, Al can be deterimental to many functions if it has access to them. Just as important, or maybe even more so, is understanding how Al is taken up by cells and reaches its targets. However, current understanding of most apects of Al uptake is also controversial and ambiguous (Lazof et al., 1994). Limitations for advances in this field are largely methodological (Taylor, 1988; Delhaize and Ryan, 1995). It was my objective in this study to characterize Al uptake, with respect to kinetics and cellular localization, and to examine factors affecting uptake. To achieve this goal, an in vivo fluorescence technique, based on the complex of morin with Al, was developed and optimized to detect Al in cells. This technique was then used to examine factors impacting Al uptake by Nicotiana tabacum L. cv BY-2 cells. Chemistry of Aluminum Aluminum is the most abundant metal in the earth’s crust (Hem, 1986). Mineral soils contain significant amounts of Al, most of which is locked in aluminosilicates or aluminum oxides (Sposito, 1989) of the clay fraction and does not pose a toxicity hazard. Upon acidification of a soil, a fraction of the total Al becomes soluble and potentially toxic to 3 plants. Thus, an acidic soil is almost synonymous with an Al-toxic soil. Acid soils cover up to 40 % of the arable soils of the world (Osmond et al., 1980) and are located mainly in tropical and subtropical areas, where soil acidity is a natural occurrence. The largest areas are in northern and central South America, central Africa and southeastem Asia. In temperate zones, soil acidity has become an increasing problem as a result of acid rain. In North America, problem areas are in the northeast, in the industrial regions of the USA and Canada, mainly because of acid rain, and also in the southeastern USA. With its high ionic charge and small ionic radius, Al has a ratio of charge-over-radius (z/r = 5.9) second only to beryllium (Orvig, 1993). Because of this, Al strongly polarizes the water molecules in its hydration shell (Baes and Mesmer, 1976). With a coordination number of six, Al is coordinated by six water molecules in an octahedral configuration. The high degree of polarization of the OH bond can result, depending on the pH of the medium, in the dissociation of a proton: Al(H20)63+ -‘= Al(H20)5(OH)2+ + H+ For simplification, these forms of A1 are represented without designating the associated water molecules, as in Al3+ or Al(OH)x’. Aluminum undergoes a pH-dependent hydrolysis series, as represented below: A1” a, ARCH)” ‘—-r A1(0H); *7 AI(OH)3° r AKOH)! I l Aln (0H),,‘3“"”+ precipitates polynuclear complexes 4 The neutral Al hydroxide species, Al(OH)3°, is insoluble and thus Al is largely insoluble in the pH range between 5 and 7. At pH values above 7.5, Al(OH)4' is formed and Al is again soluble. Under conditions with a high Al/OH' ratio, polynuclear species can form from Al(OH)2*, of which the “AID“ tridecameric polycation is the most important (Parker and Bertsch, 1992). In addition, hydrolysis and solubility of aluminum can be affected by chelation. In the presence of such agents, prediction of the fate of Al is difficult. Because of its high ratio of charge-to-radius, Al is a hard, highly reactive cation, and binds preferentially to hard negative donor groups (Da Silva and Williams, 1991). Fluoride, being the most electronegative of the anions, is the preferred inorganic monodentate ligand. However, since it is not incorporated into multidentate ligands, it is oxygen containing moieties of multidentate molecules which bind Al (Orvig, 1993). The most important ones are carboxyl (-COOH), hydroxyl (-OH), carbonyl (-CO) and phosphate (-PO3) groups. Sulfhydryl groups do not bind Al, even when part of a chelate ring (Toth et al., 1984). Amines bind Al only when part of multidentate ligands such as NTA and EDTA (Martin, 1992) Of the biological elements, the Al ion is closest in size (r=0.054 nm) to Mg2+ (r=0.072 nm) and the fen'ic ion F e3+ (r=0.065 nm), and it is size, not charge which is most important in metal ion substitution (Martin, 1988; Williams, 1992). Magnesium is closely associated with phosphate metabolism and use in cells (Williams, 1993). Al can bind to nucleoside triphosphates with an association constant 107 times that of Mg”, a cofactor for hydrolysis of nucleotides. Of special importance is the fact that Al has a slow rate of exchange in and out of its coordination sphere (Orvig, 1993). Ligand exchange rates for A1 are of the order of 1.3 s", 10’-fold slower than for Mi (Martin, 1992). These properties alone have the capacity to explain toxicity of Al. Fluoroaluminate complexes, most likely AlF4', can act as pseudophosphates, given structural similarities. By binding to nucleoside diphosphates, AlF4' can act as an analogue of the terminal y-phosphate of triphosphates. F luoroaluminates are not considered as the toxic species of Al, but, like GTPyS, have been used as a valuable experimental tool in studies of biochemical pathways involving GTP binding proteins. Since the fluoroaluminate acts as a non-hydrolyzable y-phosphate, these proteins are locked in an “activated” state (Chabre, 1990; Haug et al., 1994 - Appendix B). Complexation of A1 with fluoride can reduce Al toxicity and actually stimulate growth, as exemplified with pollen tubes (Konishi and Miyamoto, 1983). Some plants (e.g. tea) accumulate fluoride, and soils can be contaminated with fluoride as a result of industrial activity (Braen and Weinstein, 1985). However, the role of fluoride, in Al toxicity, from uncontaminated soils and from endogenous levels in plants, still needs to be well characterized. General aspects of Al toxicity in plants Aluminum toxicity is primarily root associated. Although symptoms of toxicity are manifested in the shoots, these are generally considered to be a consequence of the unhealthy root system. With the exception of Al accumulating plants (a classic example is tea), little Al is transported into the shoot (F oy et al., 1978). The initial and most dramatic symptom of Al toxicity is inhibition of root growth. Al-inj ured roots become short, stubby and thickened, with a brownish coloration. Fine branching and root hairs are reduced and the root system takes on a “corraloid” appearance (Taylor, 1988). Cells which are affected are the root cap, 6 meristem, elongating cells, root hairs and branch initials (Bennet et al., 1987; Jones and Kochian, 1995). Consequences can be very severe, with total loss of crop productivity. The site of Al-induced inhibition of root growth is localized in the root apex, and has been demonstrated by exposing only certain regions of the root to Al (Ryan et al., 1993). Root growth inhibition is rapid, and can be detected within 1-2 hours of Al exposure (Van et al., 1994; Ownby and Popham, 1989; Ryan et al., 1992). At least initially, this rapid inhibition is primarily the result of inhibited cell elongation, rather than reduced cell division (Kochian, 1995; Jones and Kochian, 1995). With the exception of some of the alterations in root morphology, overall toxicity symptoms can resemble Ca2+ deficiency. Since this was first recognized, a connection between A1 toxicity and Ca2+ metabolism has been sought and continues to be a major area of research in the field (see Hang and Vitorello, 1996 - Appendix D). Because of its pH—dependent solubility behavior, Al toxicity occurs only at soil pH values roughly below 5.0 or 5.5. Toxicity is most severe in cation-poor soils with low base saturation. This is frequently the case for highly weathered, acidic soils with low cation exchange capacity. Al toxicity can be totally eliminated by amending soils with lime, but because of the poor mobility of CaCO3 within the soil profile, this will be true only in those top layers that are tilled. The result will be a relatively shallow root system, particularly vulnerable to drought (Foy et al., 1978). In some cases, liming is not a feasible alternative for cash-strapped farmers in third world countries or remote locations. This practice is also not feasible in forests, such as those affected by acid rain in industrialized regions, and perennial crops. Although soil pH is not corrected, application of CaSO4 has recently gained wider use to ameliorate Al toxicity by increasing soil Ca2+ levels. CaSO4 is more mobile than 7 CaCO3 in the soil, reaching greater depths. and is inexpensive. Other major factors affecting severity of Al toxicity to roots are the concentrations of Ca2+ and other cations in the external solution (Kinraide and Parker, 1987), the ionic strength of solutions, temperature (Foy et al., 1978; Camargo, 1983), presence of chelators (Suhayda and Haug, 1986) and genotype of the plant (Foy et al., 1978; Wheeler et al.. 1993) The chemical species of Al which are toxic are presumably Al3+ and the mononuclear hydroxides, Al(OH)2* and Al(OH)2+ (Kinraide, 1991). High toxicity has also been attributed to the “Al,3“ tridecameric polycation (Parker and Bertsch, 1992), but this and its relevance to Al toxicity are questionable. These conclusions are based on correlations between root growth inhibition and changes in relative concentrations of Al species in solution, obtained predominantly by changing the pH of the solution. This co-linearity between pH and Al species is problematic, given the important effects pH has on the plant root (Kinraide, 1991). Aluminum toxicity in animals Although Al has generally been regarded as being of low toxicity to humans, in the last two decades, considerable evidence has implicated Al in various diseases, particularly neurological disorders, such as Alzheimer’s disease, amyotrophic lateral sclerosis and Parkinson-dementia (Ganrot, 1986; Kruck and McLachlan, 1988). That Al is neurotoxic has been shown directly by intracerebral injection of Al, causing progressive encephalopathy (Klatzo et al., 1965; Selkoe et al., 1979; Troncoso et al., 1985). It has also been clearly shown to be a causative agent of dialysis encephalopathy (Kerr et al., 1992). However, its role in 8 Alzheimer’s and other neurological disorders of natural cause are not clear. It is controversial whether brains afflicted by Alzheimer’s disease have an increased bulk Al content (Crapper et al., 1973; McDermott et al., 1979), but Al is claimed to be present in affected brain regions at concentrations sufficient to be toxic (McLachlan et al., 1992). It is associated with neurofibrillary tangles (Perl and Brody, 1980) and amyloid plaques (Candy et al., 1986) and Al was found to promote in vitro aggregation of amyloid p-protein, which accumulates in senile plaques (Mantyh et al., 1993; Kawahara et al., 1994). Because of the ubiquitous nature of Al, humans unintentionally ingest rather large amounts of Al (30-50 mg/day) in their diet (Ganrot, 1986). For the most part (99%), this Al is not absorbed by the intestines, and Al neurotoxicity is not expressed in normal, healthy individuals. Why normal exclusion of A] may fail is not clear. The accumulation of Al in the brain of patients afflicted with Alzheimer’s disease may be a consequence, rather than the cause, of nerve cell damage and other primary disfunctions, resulting in access to binding sites. A current hypothesis is that Al, although probably not a causative agent, is involved in the pathogenesis of Alzheimer’s disease (McLachlan et al., 1992). Because of the neutral internal pH values of animals, such as in the blood plasma, which would lead to precipitation, Al is transported in the chelated form. Citrate and transferrin are the major chelators of Al in the blood (Martin, 1988). In other animals, the toxic effects of A1 are most pronounced in fish, particularly the epithelium of gills (Driscoll et al., 1980; Youson and Neville, 1986). Until recently, Al concentrations in surface waters were generally below toxic levels (MacDonald and Martin, 1988). However, acidification due to acid rain has mobilized Al to toxic levels in many bodies of water. In addition to low pH, low Ca” concentrations (and probably other cations 9 as well) are also associated with Al toxicity to fish (Yousson and Neville, 1986) Low pH effects on roots As mentioned above, Al toxicity occurs in plants only at low pH, whether in soils or nutrient solutions. Low pH (<5.0) in itself is well known to be directly detrimental to root growth (Zsoldos and Erdei, 1981; Marschner, 1991; Yan et al., 1992), particularly in solutions of low ionic strength and low cation concentrations (Lund, 1970). As in A1 toxicity, increasing the concentration of Ca2+ and other cations in the external solution reduces or even abolishes the detrimental effects of acidity (F ageria et al., 1989). An evaluation of these low- pH effects is necessary for greater understanding and correct interpretation in studies of Al toxicity. Unfortunately, physiological effects related to low pH are usually not taken into consideration in Al toxicity studies and in some cases may have led to incorrect interpretation of results. At the cellular level, low pH has detrimental and distinct effects on the plasma membrane, particularly enhanced permeability (Zsoldos and Erdei, 1981; Yan et al., 1992). In addition, microviscosity of the plasma membrane is altered with decreasing pH (Berczi et al., 1981; Suhayda et al., 1988). These membrane effects are in part responsible for altered patterns of nutrient accumulation at low pH (Marschner, 1991). Although K" permeates the plasma membrane more readily at low pH, enhancement of efflux is greater, resulting in reduced net uptake (Zsoldos and Erdei, 1981). Low pH-induced membrane permeability can be alleviated by Ca2+ and other cations (Yan et al., 1992; Kinraide, 1993). Aluminum reportedly decreased H’-induced membrane permeability to K+ (Sasaki et al., 1994). It has 10 been suggested that Al-induced growth stimulation results from amelioration of proton toxicity (Llugany et al., 1995; Kinraide, 1993). At low pH, low concentrations of Al can be beneficial to root growth, particularly in Al-tolerant plants (Taylor, 1988). Recently, it was found that regions of the wheat root apex with high Al concentrations were correlated with regions in which a depletion in K* had occurred (Delhaize et al., 1993a). This was interpreted as being Al-induced loss of K’ from cells. However, since they did not present data clearly showing that this depletion of K” may have resulted regardless of the presence of Al, this depletion may well have been the result of the low pH of the medium, as has been shown to occur (Zsoldos and Erdei, 1981; Sasaki et al., 1994). If this is the case, would this mean that Al accumulates preferentially in cells that have low-pH-altered membranes? Al uptake in plants Sites of Al accumulation In intact roots, Al accumulates in the epidermis and in the cortex, although less in the inner layers of the cortex (Wagatsuma et al., 1987; Rengel, 1992; Jentschke et al., 1991, Ownby 1993; Delhaize et al, 1993a). The endoderrnis acts as a barrier (Rengel, 1992a) and transport to the shoot and leaves is generally small, although there are Al-accumulating species. Examples are the tea plant, pine trees (F oy et al., 1978) and some species native to the “cerrado” region of central Brazil (Haridasan, 1982; Haridasan and Araujo, 1988). Unfortunately, there is little in the literature as to mechanisms, cellular localization and chemical forms of the Al which accumulates in these plants. In tea, Al accumulates in older 11 leaves, particularly in the epidermis (Foy et al., 1978). With regard to the chemical form of Al in these leaves, it was concluded that most Al was chelated to the catechin group of polyphenols, and to a lesser extent to phenolic and organic acids and as Al-F complexes (Nagata et al., 1992). Whereas in some cases there is evidence that Al is chelated within the leaf cells (Haridasan et al., 1986), another report concluded that Al is deposited on the cell wall of the epidermis and palisade parenchyma (Memon et al., 1981). The Al species present after absorption from the soil is not yet known, but there is evidence that it may be transported as Al-F species (Nagata et al., 1993). Al uptake at the cellular level: apoplastic or symplastic? At the cellular level, a major unresolved issue is whether Al actually enters the cell or whether it accumulates solely in the cell wall or in the mucigel surrounding the root apex (Delhaize and Ryan, 1995; Kochian, 1995). Attempts to make this distinction have received considerable attention because of its implication to models of Al toxicity. While Al has been found in the nucleus, presumably bound to DNA (Matsumoto et al., 1976), it has been reported by others to be localized solely or predominantly in the cell wall (Clarkson, 1967; Ownby, 1993; Marienfel and Stelzer, 1993). This ambiguity results partially from differing experimental conditions, some of which may have been inadvertently flawed (Kochian, 1995). Earlier reports have employed long periods of exposure to Al, which raises questions as to the integrity of these stressed cells, and/or high concentrations of Al, which in turn raises questions as to precipitation of Al. Precipitation of Al phosphates in the cell wall can also occur. Maj or limitations to progress in Al uptake studies have been largely methodological. 12 Reasons for this are the complex chemistry of Al (Martin, 1988), interactions with the cell wall, the lack of a suitable Al radioisotope (Rengel, 1992; Schaedle et al., 1989; Taylor, 1988), and the limited sensitivity and resolution of microanalytical techniques (Delhaize et al., 1993a; Lazof and Lauchli, 1991; Lazof et a1, 1994). Because of the uncertain effectiveness of chelator washes in removing Al from the cell wall (Taylor, 1988), quantification of Al by graphite atomic absorption spectrometry (GFAAS), a sensitive and standard analytical technique, does not provide certainty as to cellular sites of accumulation. X-ray microanalysis, the major microanalytical technique, is limited by sensitivity and spatial resolution, the latter being relatively poor because of the interactions of the radiation with the specimen layer and distortions in the cell samples during preparation (Goldstein et al., 1981; Lazof and Lauchli, 1991; Lazof et al., 1994). Nevertheless, attempts have been made to surmount or circumvent these limitations and to address the problem of Al uptake and its cellular site of accumulation. The kinetics of short-term A1 uptake was found to be biphasic, and was first used in an attempt to distinguish cell wall A1 from uptake across the plasma membrane (Petterson and Strid, 1989; Zhang and Taylor, 1989; Zhang and Taylor, 1990). The rapid, initial, non-linear phase of uptake was interpreted as being accumulation of exchangeable A1 in the cell wall, and could be desorbed with citrate (Zhang and Taylor, 1990). The slower, second phase of uptake was linear and interpreted to be uptake across the plasma membrane (Pettersson and Strid, 1989) or metabolism-dependent accumulation in the cell wall (Zhang and Taylor, 1990). A more direct fiactionation procedure was also used (Tice et al., 1992). Following external washes aimed at removing apoplastic Al, soluble symplastic Al was washed out after cell membranes had been ruptured by freeze—thawing. The residual fraction was interpreted as being non- l3 soluble symplasmic Al and contained 50 to 70% of the total cell-associated Al. To improve upon microanalytical techniques for the detection of Al, secondary-ion mass spectrometry was used to detect Al in soybean root tips exposed to A1 for a 30 min period (Lazof et al., 1994). Although this method has a spatial resolution of about 2 pm, approximately one order of magnitude better than X-ray microanalysis, the association between A1 signal images and actual cell images was difficult (Figs. 4 and 6 of Lazof et al., 1994). Nevertheless, the authors concluded that Al was internalized by cells based on the fact that the Al signal was concentrated in circular areas similar in size (15 um in diameter) to individual cells. They argued that if present in the cell wall, the A1 signal image should show “rings” corresponding to the cell wall. In summary, there now seems to be evidence, although inconclusive, favoring the notion that Al is indeed rapidly internalized by cells (Kochian, 1995). Factors impacting Al uptake Whereas efforts have been made to determine the cellular sites of Al accumulation, much less attention has been given to factors impacting Al uptake. With regard to dependence of A1 uptake on metabolism and energy, results are again ambiguous. Depending on the criteria used, accumulation of Al has the characteristics of non-metabolic uptake (Clarkson, 1967; Wagatsuma, 1983; Foy et al., 1978). Metabolic inhibitors, such as DNP or CCCP, increase Al uptake (Wagatsuma, 1983; Rincon and Gonzales, 1992; Zhang and Taylor, 1991 ), rather than decrease it. Because this increase in uptake was more dramatic in Al-tolerant cultivars, an energy-dependent exclusion mechanism has been proposed (Zhang and Taylor, 1991; Rincon and Gonzales, 1992). In an apparent paradox, Al 14 accumulates in the root tip, which is physiologically very active (Bennet et al., 1985; Rincon and Gonzales, 1992) and a metabolism dependent-accumulation of Al in the cell wall has been proposed (Petterson and Strid, 1989; Zhang and Taylor, 1991). Low temperature has been shown to substantially decrease Al uptake (Pettersson and Strid, 1989; Zhang and Taylor, 1990). Calcium is known to ameliorate Al toxicity (Kinraide and Parker, 1987). Information on the effects of Ca2+ on Al uptake is scarce but in at least one report, has been shown to decrease Al uptake (Alva et al., 1986). A1 uptake has also been studied in non-plant systems lacking a cell wall, such as liposomes (Shi and Haug, 1988; Akeson and Munns, 1989) or mammalian cells (Shi and Haug, 1990). In neuroblastoma cells, Al uptake was suggested to depend on some sort of carrier, although it depended in part on the physical state of the membrane (Shi and Haug, 1990). Mechanisms of Al toxicity The search for the primary target of Al injury is probably the most active area in Al toxicity studies, and too many potential targets for Al toxicity have been suggested to adequately cover in this review. Several of them involve phosphate groups and their utilization by cells. A1 is capable of binding tightly to DNA, presumably to its phosphate backbone, or alternatively to associated histones (Dyrssen et al., 1987; Matsurnoto, 1991) and this led to one of the earlier hypotheses of toxicity, that cell division was impaired because of interactions of A1 with nuclear DNA. A1 also has a high binding affinity to free nucleotide triphosphates and a model for A1 toxicity based on its binding to ATP in the cytoplasm has 15 been proposed (Pettersson and Bergman, 1989). A1 has also been shown to alter the structure of the plasma membrane (Vierstra and Haug, 1978; Zhao et al., 1987) and has pronounced effects on ion fluxes across the membrane, particularly Ca2+ uptake (Huang et al., 1992). Al toxicity has frequently been linked to Ca2+ (Rengel, 1992) either because of Al-induced perturbations in cellular Caz’ metabolism or because of Ca2+ amelioration of A1 toxicity (Kinraide and Parker, 1987). Several investigations found profound Al-induced alterations in the structure of calmodulin, the chief mediator of intracellular Ca2+ signalling (Siegel and Haug, 1983; Siegel et al., 1983), and initiated considerable research on the role of calmodulin in Al toxicity (Haug and Vitorello, 1996 - Appendix C). More recently, phosphoinositide-mediated signal transduction, a pathway which also involves Caz” as an intracellular messenger, has been investigated as a primary site of Al toxicity (Fig. 1) in mammalian (Hang et al., 1994; Appendix B) and plant cells (Jones and Kochian, 1995). In both cases A1 treatment presumably inhibited the activity of phospholipase C or possibly the action of the trimeric G-protein (Shi and Haug, 1992; Shi et al., 1993). A1 tolerance: mechanisms, genetics and Al-induced genes and proteins Plant genotypes present wide variability with respect to their sensitivity to A1, and this has been known for some time. Traditional screening and breeding programs have resulted in some success. The mechanisms for tolerance, like those for toxicity, are not known (Delhaize and Ryan, 1995; Kochian, 1995). However, one mechanism, that of exclusion of Al from the root, has recently been revisited and has gained favor in this field 16 of research. In some cases, Al has been shown to accumulate more in the roots of A1- sensitive cultivars of wheat than in Al-tolerant ones (Delhaize et al., 1993b; Rincon and Gonzales, 1992). Recently, three independent groups have demonstrated that malate- secretion is enhanced in Al-tolerant cultivars compared to Al-sensitive ones (Basu et al., 1994; Delhaize et al., 1993b; Pellet et al., 1995). Tolerance would be the result of chelation of A1 to malate. From the few studies that have been performed, Al tolerance is a trait governed by only one or a few dominant genes (Aniol and Gustafson, 1984; Aniol, 1991; Berzonsky, 1992). In tobacco cell cultures selected for Al tolerance, a single dominant gene mutation was attributed to tolerance (Conner and Meredith, 1985). On the other hand, Al induces the synthesis of several proteins (Basu et al., 1994) and also the expression of several genes (Snowden et al., 1995). However, many of these proteins are induced in both Al-sensitive and Al-tolerant genotypes (Delhaize and Ryan, 1995). Some of these proteins have been identified and include phenylalanine ammonia-lyase, metallothionein-like proteins, proteinase inhibitors and asparagine synthetases (Snowden et al., 1995). Thus, to date, the proteins synthesized in response to A1 appear to be general stress-response proteins. Processes involved in cell growth/secretion and Al toxicity In plants, Al affects meristematic and elongating cells of the root apex (Ryan et al., 1993), root cap cells (Bennet and Breen, 1991) root hairs (Jones et al., 1995) and pollen tubes (Konishi and Miyamoto, 1983; Searcy and Mulcahy, 1990). In mammalian cells, neurological cells appear to be the most sensitive (Ganrot, 1986) and in fish A1 affects the 17 epithelium of the gill (Youson and Neville, 1986). These cells are engaged in growth, secretion, or synapsis, all of which involve similar processes at the cell surface, particularly membrane traffic and turnover. A1 may inhibit growth or disrupt secretory activity in these cells by direct toxicity to these regions. With regard to growth in plant cells, A1 inhibits tip growth in root hairs and pollen tubes, it reduces the incorporation of sucrose into cell walls (Huck, 1972) and induces callose synthesis (Schaeffer and Walton, 1990). Al was found to cause deformities in root hair tips, but only in those that were growing (Jones et al., 1995). For these reasons and because the relationship between cell growth and Al toxicity is a major theme of my dissertation, a brief review of processes associated with cell growth and their potential relevance to A1 toxicity is warranted. For a plant cell to expand and grow, new cell wall and plasma membrane must be incorporated to the cell surface. This involves many cellular components, ranging from the endomembrane system, the cytoskeleton and the plasma membrane. Cellulose microfibrils are synthesized on the outer side of the plasma membrane by a membrane-associated cellulose synthase complex (Delrner, 1987). Other cell wall components are synthesized in the Golgi dictyosomes and transported and secreted to the cell surface (Staehelin and Moore, 1995). The plasma membrane is under continual turnover in these regions. The number of exocytotic fusions in tip‘growing cells has been estimated to be SOOO/min (Heath, 1990). In plants, cell growth has been better studied in tip-growing cells (Steer and Steer, 1989). Regions of cellular growth may be involved in A1 toxicity for two reasons. First, access of Al to sites of toxicity may be facilitated in these regions, because of the dynamic nature of these membranes. Second, several components or processes involved with cell growth are potential targets for Al. 18 Membrane fusion, turnover and traffic are essential to growth and secretion. The mechanisms of membrane fusion are not well understood, with most knowledge deriving from information on liposomes rather than on actual fusion events in the cell. Nevertheless, from these liposome models, it is thought that non-bilayer lipid configurations are required as intermediate steps for fusion to occur. Lipid fUSIOl'I is promoted by lipids that do not form bilayers and by cations (Ca2+, H‘) and heat, which disrupt bilayers (Zirnmerberg, et al., 1993). Phosphatidylserine (PS) and phosphatidylethanolamine (PE) are considered important for fusion. In PE-deficient Escherichia coli cells, high concentrations of divalent cations were necessary for growth, and these results were interpreted to involve regulation of lipid polymorphism (Rietveld et al., 1994). Proteins are widely accepted as mediators of membrane fusion, however, with few exceptions (viral proteins), little information is available for such proteins (Zimmerberg et al., 1993). The plasma membrane probably plays an important role in Al toxicity, either as a permeability barrier or as a site of injury. At micromolar concentrations, Al reportedly induced rigidification and membrane firsion in unilamellar vesicles containing PS (Deleers et al., 1986). Binding of A1 to phosphatidylcholine was found to occur with an affinity 560 times higher than with Ca2+ (Akeson et al., 1989). A role of the plasma membrane in A1 toxicity was suggested by the fact that the average surface charge of protoplasts from Al- tolerant species was less negative than Al-sensitive species, and this property was used to separate and collect Al-tolerant and Al-sensitive protoplasts (Wagatsuma et al., 1995). Plasma membrane domains rich in negatively charged sites may constitute regions for binding and uptake of A1. Membrane instabilities, enhanced by low-pH, low Caz’ concentrations and possibly even by Al, may give rise to special conditions for uptake of A1. 19 Phosphatidylserine, a negatively charged, fusogenic lipid, may be present in such regions. Compared to the mechanisms of membrane fusion, other aspects of membrane trafficking are better understood, such as docking and the regulation of vesicular traffic. Both small and heterotrimeric GTP-binding proteins are involved in most phases, and GTP- binding proteins can be found in all cellular membranes (V errna et al., 1994; Nuoffer and Balch, 1994). There is experimental evidence that A] can inhibit the activity of GTP-binding proteins (Shi et al., 1992; Haug et al., 1994 - Appendix C). Perhaps Al has greater access to GTP-binding proteins of the plasma membrane in regions of cell growth. Calcium is necessary for membrane fusion to occur and is involved in the regulation of exocytosis and endocytosis (Steer, 1988). Modulation of fusion by Ca2+ can occur by interaction with phospholipids or by means of annexins, which are Ca2’-dependent phospholipid-binding proteins (Clark and Roux, 1995). In tip-growing cells such as pollen tubes and root hairs, Ca2+ gradients are established between the tip and basal regions (Miller . et al., 1992). These Ca2+ gradients are presumably generated and maintained by localized Ca2+ influx mediated through Ca2+ channels which accumulate in growing regions (Steer and Steer, 1989). Levels of calmodulin are high in the cytoplasm proximal to sites of growth (Hausser et al., 1984). Aluminum is thought to impact Ca2+ metabolism of cells and has been shown to interact with calmodulin, in vitro (Siegel and Haug, 1983; Haug and Vitorello, 1996- Appendix D) and Ca2+ channels (Huang et al., 1996). It is not known whether Al impacts annexins. Microtubules and microfilarnents are involved in directing and transporting secretory vesicles from the Golgi to the cell surface (Cole and Lippincott-Schwartz, 1995). Microtubules of the cortical array are involved in orienting cellulose microfibril deposition 2O (Delmer, 1987). Tubulin polymerization requires Mg-dependent hydrolysis of GTP. In vitro, A1 promoted microtubule assembly (MacDonald etal., 1987). However, Al has not yet been shown to interact with microtubules in vivo. The dynamics of actin are also dependent on GTP hydrolysis and Al reportedly induced an increase in the rigidity of transvacuolar strands of actin in cultured soybean cells (Grabski and Schindler, 1995). It has not yet been demonstrated whether Al actually reaches these actin strands. REFERENCES Akeson, M.A., Munns, D.N., Burau, R.G. 1989. Adsorption of Al to phosphatidylcholine vesicles. Biochim. Biophys. Acta 986:33-40. Akeson, M.A., Munns, ON. 1989. Lipid bilayer permeation by neutral aluminum citrate and by three alpha-hydroxy carboxylic acids. Biochim. Biophys. Acta 984:200-206. Alva, A.K., Asher, C.J., Edwards, DC. 1986. The role of calcium in alleviating aluminium toxicity. Aust. J. Agric. Res. 37:375-382. Aniol, A., Gustafson, J.P. 1984. Chromosome location of genes controlling aluminum tolerance in wheat, rye, and triticale. Plant Physiol. 104:701-705. Aniol, A. 1991. Genetics of acid tolerant plant. In Wright, R.J. (eds). Plant-soil interactions at low pH. Kluwer Academic, Dordrecht, The Netherlands, pp 1007-1017. Baes, C.F., Mesmer, RE. 1976. The hydrolysis of cations. J .Wiley & Sons, New York. pp 112-113. Basu, U., Godbold, D., Taylor, GJ. 1994. Aluminum resistance in T riticum aestivum associated with enhanced exudation of malate. J. Plant Physiol. 144:747-753. Bennet, R.J., Breen, C.M., Fey, M.V. 1985. The primary site of aluminium injury in the root of Zea mays L. S.Afi'. J. Plant Soil 228-17. Bennet, R.J., Breen, CM. 1991. The aluminium signal: new dimensions to mechanisms of aluminium tolerance. Plant Soil 134:153-166. Bennet, R.J., Breen, C.M., Fey, M.V. 1987. The effects of aluminium on root cap function and root development in Zea mays L. Environ. Exp. Bot. 27:91-104. Berczi, A., Erdei, L., Zsoldos, F. 1981. Membrane transport and ion transport properties in cereals under acidic and alkaline stress. Physiol. Plant. 53:475-478. Berzonsky, WA. 1992. The genomic inheritance of aluminum tolerance in Atlas 66 wheat. Genome 35:689-693. 21 22 Braen, S.N., Weinstein, L.H. 1985. Uptake of fluoride and aluminum by plants grown in contaminated soils. Water Air Soil Pollut. 24:215-223. Camargo, C.E. de 0. 1983. Efeito da temperatura da solucao nutritiva na tolerancia ao aluminio de cultivares de trigo. Bragantia 42: 51-63. Candy, J .M., Oakley, A.E., Klinowski, J. et al. 1986. Aluminosilicates and senile plaque formation in Alzheimer’s disease. Lancet 1:354-357. Chabre, M. 1990. Aluminofluoride and beryllofluoride complexes: new phosphate analogs in enzymology. Trends Biochem. Sci. 1526-10. Clark, G.B., Roux, SJ. 1995. Annexins of plant cells. Plant Physiol. 109:1133-1139. Clarkson, D.T. 1967. Interactions between aluminium and phosphorus on root surfaces and cell wall material. Plant Soil 27: 347-356. Cole, N.B., Lippincott-Schwartz, J. 1995. Organization of organelles and membrane traffic by microtubules. Current Opinion in Cell Biology 7: 55-64. Conner, A.J., Meredith, GP. 1985. Strategies for the selection and characterization of aluminum-resistant variants from cell cultures of Nicotiana plumbaginifolia. Planta 166: 466-473. Crapper, D.R., Krishnan, S.S., Dalton, A.J. 1973. Brain aluminum distribution in Alzheimer’s disease and experimental neurofibrillary degeneration. Science. 180:511- 513. Crapper Mclachlan, DR. 1989. Aluminum neurotoxicity: criteria for assigning a role in Alzheimer’s disease. In Lewis, T.E. (ed). Environmental Chemistry and Toxicology of Aluminum. Lewis Pub1., Boca Raton, FL. pp 299-315. Da Silva, J.J.R.F., Williams, R.J.P. 1991. The biological chemistry of the elements. The inorganic chemistry of life. Oxford University Press, Oxford. Deleers, M., Servais, J .P., Wiilfert, E. 1986. Biochim. Biophys. Acta 855: 271-276. Delhaize, E., Craig, S., Beaton, C.D., Bennet, R.J., Jagadish, V.C., Randall, P.J. 1993a. Aluminum tolerance in wheat (T riticum aestivum L.) I. Uptake and distribution of almninum in root apices. Plant Physiol. 103:685-693. Delhaize, B, Ryan, P.R., Randall, P.J. 1993b. Aluminum tolerance in wheat (Triticum aestivum L.). II. Aluminum-stimulated excretion of malic acid from root apices. Plant Physiol. 103:695-702. 23 Delhaize, E., Ryan, PR. 1995. Aluminum toxicity and tolerance in plants. Plant Physiol. 107: 315-321. Delrner, DP. 1987. Cellulose biosynthesis. Ann. Rev. Plant Physiol. 38: 259-290. Driscoll, C.T., Baker, J .P., Bisigni, J .J ., Schofield, CL. 1980. Effect of aluminium speciation on fish in dilute acidified waters. Nature 284: 161-164. Dyrssen, D., Haraldsson, C., Nyberg, E., Wedborg, M. 1987. Complexation of aluminum with DNA. J. Inorg. Biochem. 29:67-75. Foy, C.D., Chaney, R.L., White, MC. 1978. The physiology of metal toxicity in plants. Annu. Rev. Plant Physiol. 29: 511-566. Fageria, N.K., Baligar, V.C., Wright, R.J. 1989. Growth and nutrient concentrations of alfalfa and common bean as influenced by soil acidity. Plant Soil 119: 331-333. Ganrot, PD. 1986. Metabolism and possible health effects of aluminum. Environ. Health Perspect. 65: 363-441. Godbold, D.L., Fritz, E., Hiittermann, A. 1988. Aluminum toxicity and forest decline. Proc. Natl. Acad. Sci. USA. 85:3888-3892. Goldstein, J .I., Newbury, D.E., Echlin, P., Joy, D.C., Fiori, C., Lifshin, E. 1981. Practical techniques of x-ray analysis. In Goldstein, J .I., Newbury, Echlin, P., Joy, D.C., Fiori, C., Lifshin, E. (eds). Scanning electron microscopy and x-ray microanalysis. pp 393-446. Plenum Press, New York. Grabski, S., Schindler, M. 1995. Aluminum induces rigor within the actin network of soybean cells. Plant Physiol. 108:897-901. Haridasan, M. 1982. Aluminium accumulation by some cerrado native species of central Brazil. Plant Soil 65:265-273. Haridasan, M., Paviani, T.I., Schiavini, I. 1986. Localization of aluminimn in the leaves of some aluminium-accumulating species. Plant Soil 94:435-437. Haridasan, M., De Araujo, GM. 1988. Aluminimn-accumulating species in two forest communities in the Cerrado region of central Brazil. Forest Ecol. Manag. 24:15-26. Haug, A. 1984. Molecular aspects of aluminum toxicity. Crit. Rev. Plant Sci. 1: 345-373. Haug, A., Shi, B., Vitorello, V. 1994. Aluminum interaction with phosphoinositide- associated signal transduction. Arch Toxicol 68: 1-7 24 Haug, A., Vitorello, V. 1996. Aluminum coordination to calmodulin: Thermodynamic and kinetic aspects. Coordination Chemistry Reviews. (In press). Hausser, 1., Herth, W., Reiss, H.D. 1984. Calmodulin in tip growing plant cells, visualized by fluorescing calmodulin-binding phenothiazines. Planta 162:33-39. Heath, LB. 1990. Tip growth in plant and fungal cells. Academic Press, Orlando, Fla. Hem, J .D. 1986. Geochemistry and aqueous chemistry of aluminum. Kidney International, 29(18):s3-s7. Huang, J .W., Pellet, D.M., Papernik, L.A., Kochian, L.V. 1996. Aluminum interactions with voltage-dependent calcium transport in plasma membrane vesicles isolated from roots of aluminum-sensitive and -resistant wheat cultivars. Plant Physiol. 110: 561-569. Huang, J. W., Shaff, J.E., Grunes, D.L., Kochian, L.V. 1992. Aluminum effects on calcium fluxes at the root apex of aluminum-tolerant and aluminum-sensitive wheat cultivars. Plant Physiol. 98:230-237. Huck, MG. 1972. Impairment of sucrose utilization for cell wall formation in the roots of aluminum-damaged cotton seedlings. Plant Cell Physiol. 13:7-14. Jentschke, G., Schlegel, H., Godbold, D.L. 1991. The effect of aluminium on uptake and distribution of magnesium and calcium in roots of mycorrhizal Norway spruce seedlings. Physiol. Plant. 82:266-270. Jones, D.L., Kochian, L.V. 1995. Aluminum inhibition of the inositol 1,4,5-triphosphate signal transduction pathway in wheat roots: a role in aluminum toxicity? Plant Cell 7:1913-1922; Jones, D.L., Shaff, J .S., and Kochian, L.V. 1995. Role of calcium and other ions in directing root hair tip growth in Limnobium stoloniferum. 1. Inhibition of tip growth by aluminium. Planta 197:672-680. Kawahara, M., Muramoto, K., Kobayashi, K, Mori, H., Kuroda, Y. 1994. Aluminum promotes the aggregation of Alzheimer’s amyloid B-protein in vitro. Biochem. Biophys. Res. Comm. 198:531-535. Kerr, D.N.S. , Ward, M.K., Ellis, H.A., Simpson, W., Parkinson, LS. 1992. Aluminium intoxication in renal disease. In Chadwick, D.J., Whelan, J. (eds). Almninium in biology and medicine. pp 123-141. John Wiley & Sons, New York. Kinraide, T.B., Parker, DR. 1987. Cation amelioration of aluminum toxicity in wheat. Plant Physiol. 83: 546-551. 25 Kinraide, TE. 1991. Identity of the rhizotoxic aluminium species. Plant Soil 134: 167-178. Kinraide, TE. 1993. Aluminum enhancement of plant growth in acid rooting media. A case of reciprocal alleviation of toxicity by two toxic cations. Phys. Plant. 88:619-625. Klatzo, I., Wisniewki, H., Streicher, E. 1965. Experimental production of neurofibrillary degeneration. 1. Light microscopic observations. J. Neuropathol. Exp. Neurol. 24:187- 199. Kochian, L.V. 1995. Cellular mechanisms of aluminum toxicity and resistance in plants. Annu. Rev. Plant Physiol. Plant Mol. Biol. 46: 237-60. Konishi, S., Miyamoto, S. 1983. Alleviation of aluminum stress and stimulation of tea pollen tube growth by fluorine. Plant Cell Physiol. 24:857-862. Kruck, T.P.A., McLachlan, DR. 1988. Mechanisms of aluminum neurotoxicity - relevance to human disease. In: Sigel, H., Sigel, A. (eds). Metal ions in biological systems: Aluminum and its role in biology, vol. 24, pp 285-314. New York, Marcel Dekker. Lazof, D., Lauchli, A. 1991. Complementary analysis of freeze-dried and frozen-hydrated plant tissue by electron-probe X-ray microanalysis: Spectral resolution and analysis of calcium. Planta 184:327-333. Lazof, D.B., Goldsmith, J .G., Rufty, T.W., Linton, R. 1994. Rapid uptake of aluminum into cells of intact soybean root tips. A microanalytical study using secondary ion mass spectrometry. Plant Physiol. 106: 1107-1114. Llugany, M., Poschenrieder, C., Barcelo, J. 1995. Monitoring of aluminium-induced inhibition of root elongation in four maize cultivars differing in tolerance to aluminum and proton toxicity. Physiol. Plant. 93:265-271. Lund, Z.F. 1970. The effect of calcium and its relation to several cations in soybean root growth. Soil Sci. Soc. Am. Proc. 34:456-459. MacDonald, T.L., Humphreys, W.G., Martin, RB. 1987. Promotion of tubulin assembly by aluminum ions in vitro. Science 236: 1 83-1 86 Macdonald, T.L., Martin, RB. 1988. Aluminum ion in biological systems. Trends Biochem. Sci. 13:15-19. Mantyh, P.W., Ghilardi, J .R, Rogers, 8., DeMaster, E., Allen, C.J., Stimson, E.R., Maggio, J .E. 1993. Aluminum, iron, and zinc ions promote aggregation of physiological concentrations of B-amyloid peptide. J. Neurochem. 61:1171-1174. 26 Marienfeld, S., Stelzer, R. 1993. X-ray microanalyses in roots of Al-treated Avena sativa plants. J. Plant Physiol. 141 :569-573. Marschner, H. 1991. Mechanisms of adaptation of plants to acid soils. In Wright, R.J., Baligar, V.C., Murrmann, R.P. (eds). Plant-soil interactions at low pH. pp 683-702. Kluwer Academic Publishers. Netherlands. Martin, RB. 1988. Bioinorganic chemistry of aluminum. In: Sigel, H., Sigel, A. (eds). Metal ions in biological systems: Aluminum and its role in biology, vol. 24, pp 2-57. New York, Marcel Dekker. Martin, RB. 1992. Aluminium speciation in biology. In Chadwick, D.J., Whelan, J. (eds). Aluminium in biology and medicine. pp 5-25. John Wiley & Sons, New York. Matsumoto, H., Hirasawa, E., Torikai, H., Takahashi, E. 1976. Localization of absorbed aluminium in pea root and its binding to nucleic acids. Plant Cell Physiol. 17: 127-137. Matsumoto, H. 1991. Biochemical mechanism of the toxicity of aluminium and the _ sequestration of aluminium in plant cells. In Wright, R.J., Baligar, V.C. Murrrnann, R.P. (eds). Plant-soil interactions at low pH. pp 825-838. Dordrecht, Kluwer Acad. Publ. McDermott, J .R., Smith, A.I., Iqbal, K., Wisniewski, HM. 1979. Brain aluminum in aging and Alzheimer disease. Neurology 29:809-814. McLachlan, D.R., Fraser, P.E., Dalton, A]. 1992. Aluminium and the pathogenesis of Alzheimer’s disease: a summary of evidence. In Chadwick, D.J., Whelan, J. (eds). Aluminium in biology and medicine. pp 87-108. John Wiley & Sons, New York. Memon, A.R., Chino, M, Yatazawa, M. 1981. Microdistribution of aluminum and manganese in the tea leaf tissues as revealed by X-ray microanalyzer. Commun. Soil Sci. Plant Anal. 12:441-452. Miller, D.D., Callaham, D.A., Gross, D.J., Hepler, PK. 1992. Free Ca2+ gradient in growing pollen tubes of Lilium. J. Cell Sci. 101 :7-12. Nagata, T., Hayatsu, M, Kosuge, N. 1992. Identification of aluminium forms in tea leaves by 2"Al NMR. Phytochemistry 31:1215-1218. Nagata, T., Hayatsu, M, Kosuge, N. 1993. Aluminium kinetics in the tea plant using 27Al and "F NMR. Phytochemistry 32:771-775. Nuoffer, C., Balch, WE. 1994. GTPases: multifunctional molecular switches regulating vesicular traffic. Ann. Rev. Biochem. 63: 949-990. 27 Orvig, C. 1993. The aqueous coordination chemistry of aluminum. In Coordination chemistry of aluminum, (G.H. Robinson, ed), pp 85-121. VCH Publishers, New York, NY. ISBN 1-56081-059-9. Osmond, C.B., Bjorkman, 0., Anderson, DJ. 1980. Physiological processes in plant ecology, Springer-Verlag, p. 61. Ownby, J .D. 1993. Mechanisms of reaction of hematoxylin with aluminium-treated wheat roots. Phsysiol. Plant. 87:371-380. Ownby, J.D., Papham, HR. 1989. Citrate reverses the inhibition of wheat root growth caused by aluminum. J. Plant Physiol. 135:588-591. Parker, D.R., Bertsch, PM. 1992. Formation of the "All3" tridecameric polycation under diverse synthesis conditions. Environ. Sci. Technol. 26:914-921. Pellet, D.M., Grunes, D.L., Kochian, L.V. 1995. Organic acid exudation as an aluminum- tolerance mechanism in maize (Zea mays L.). Planta 196:788-795. Perl, D.P., Brody, AR. 1980. Alzheimer’s disease: x-ray spectrometric evidence of aluminum accumulation in neurofibrillary tangle-bearing neurons. Science 208:297-299. Petterson, A., Bergman, B. 1989. Effects of aluminium on ATP pools and utilization in the cyanobacterium Anabaena cylindrica: A model for the in viva toxicity. Physiol. Plant. 76:255-261. Pettersson, S., Strid, H. 1989. Initial uptake of aluminum in relation to temperature and phosphorus status of wheat (Triticum aestivum L.) roots. J. Plant Physiol. 134: 672-677. Rengel, Z. 1992. Role of calcium in aluminum toxicity. New Phytol. 121:499-513. Rietveld, A.G., Chupin, V.V., Koorengevel, M.C., Wienk, K.L.J., Dowhan, W., Kruijff, B. 1994. Regulation of lipid polymorphism is essential for the viability of phosphatidylethanolamine-deficient Escherichia cali cells. J. Biol. Chem. 269: 28670- 28675. Rincon, M., Gonzales, RA. 1992. Aluminum partitioning in intact roots of aluminum- tolerant and aluminum-sensitive wheat ( T riticum aestivum L.) cultivars. Plant Physiol. 99:1021-1028. Ryan, P.R., DiTomasso, J .M., Kochian, L.V. 1993. Aluminium toxicity in roots: an investigation of spatial sensitivity and the role of the root cap. J. Exp Bot 44:437-446. 28 Ryan, P.R., Shaif, J.E., Kochian, L.V. 1992. Aluminum toxicity in roots. Correlation among ionic currents, ion fluxes, and root elongation in aluminum-sensitive and aluminum- tolerant wheat cultivars. Plant Physiol. 99:1 193-1200. Sasaki, M., Kasai, M., Yamamoto, Y., Matsumoto, H. 1994. Comparison of the early response to aluminum stress between tolerant and sensitive wheat cultivars: root growth, aluminum content and efflux of K*. J. Plant Nutr. 17: 1275-1288. Searcy, K.B., Mulcahy, D.L. 1990. Comparison of the response to aluminum toxicity in gametophyte and sporophyte of four tomato (Lycapersican esculentum Mill.) cultivars. Theor. Appl. Genet. 80:289-295. Schaeffer, H.J., Walton, J .D. 1990. Aluminum ions induce oat protoplasts to produce an extracellular (1-3)B-D-glucan. Plant Physiol. 94: 13-19. Selkoe, D.J., Liem, R.K.H., Yen, S.H., Shelanski, ML. 1979. Biological and immunological characterization of neurofibrillary degeneration by aluminum. Brain Res. 163:235-252. Schaedle, M., Thornton, F.C., Raynal, D.J., Tepper, HE. 1989. Response of tree seedlings to aluminum. Tree Phsyiol. 52337-356. Shi, B., Haug, A. 1988. Uptake of aluminum by lipid vesicles. Tox. Environ. Chem. 17: 337- 349. Shi, B., Haug, A. 1990. Aluminum uptake by neuroblastoma cells. J. Neurochem. 55: 551- 558. Shi, B., Haug, A. 1992. Aluminum interferes with signal transduction in neuroblastoma cells. Pharmacol. Toxicol. 71 :308-313. Shi, B., Chou, K., Haug, A. 1993. Almninium impacts elements of the phosphoinositide signalling pathway in neuroblastoma cells. Mol. Cell. Biochem. 121:109-118. Siegel, N., Coughlin, R.T., Haug, A. 1983. A thermodynamic and electron paramagnetic resonance study of structural changes in calmodulin induced by aluminum binding. Biochem. Biophys. Res. Commun. 1 15:512-517. Siegel, N., Haug, A. 1983. Aluminum interaction with calmodulin: evidence for altered structure and function from optical and enzymatic studies. Biochim. Biophys. Acta 744:36-45. Snowden, K.C., Richards, K.D., Gardner, RC. 1995. Aluminum-induced genes. Induction by toxic metals, low calcium, and wounding and pattern of expression in root tips. Plant Physiol. 107: 341-348. 29 Staehelin, L.A., Moore, 1. 1995. The plant Golgi apparatus: structure, functional organization and trafficking mechanisms. Ann. Rev. Plant Physiol. Plant Mol. Biol. 46: 261-288. Steer, M.W. 1988. The role of calcium in exocytosis and endocytosis in plant cells. Physiol. Plant. 72: 213-220. Steer, M.W., Steer, J .M. 1989. Pollen tube tip growth. New Phytol. 111:323-358. Suhayda, C.G., Weis, C., Shi, B., Haug, A. 1988. pH-related changes in maize root plasma membranes. Commun. Soil Sci. Plant Anal., 19:1165-1175. Suhayda, C.G., Haug, A. 1986. Organic acids reduce aluminum toxicity in maize root membranes. Physiol. Plant. 682189-95. Sposito, G. 1989. The environmental chemistry of aluminum. CRC Press, Inc., Boca Raton, FL. 336 p. Taylor, G.J. 1988. The physiology of aluminium phytotoxicity. In: Sigel, H., Sigel, A. (eds). Metal ions in biological systems: Aluminum and its role in biology, vol. 24, pp 123-163. New York, Marcel Dekker. Tice, K.R., Parker, D.R., DeMason, DA 1992. Operationally defined apoplastic and symplastic aluminum fractions in root tips of aluminum-intoxicated wheat. Plant Physiol. 100: 309-318. Toth,1., Zekany, L., Brucher, E. 1984. Equilibrium study of the systems of Al(III), Ga(III), and 1n(III) with mercaptacetate, 3-mercaptopropionate and 2-mercaptobenzoate. Polyhedron 7:871-877. Troncoso, J.C., Hoffman, P.N., Griffin, J.W., Hess-Koslow, K.M., Price, D.L. 1985. Aluminum neurotoxicity: a disorder of neurofilament transport in motor neurons. Brain Res. 342:172-175. Van, H.L., Kuraishi, S., Sakurai, N. 1994. Aluminum-induced rapid root inhibition and changes in cell-wall components of squash seedlings. Plant Physiol. 106:971-976. Verrna, D.P.S., Cheon, C., Hong, Z. 1994. Small GTP-biding proteins and membrane biogenesis in plants. Plant Physiol. 106: 1-6. Viersna, R., Haug, A. 1978. The effect of Al” on the physical properties of membrane lipids in Thermaplasma acidaphilum. Biochem. Biophys. Res. Comm. 84: 138-143. Wagatsuma, T. 1983. Effect of non-metabolic conditions on the uptake of aluminum by plant roots. Soil Sci. Plant Nutr. 29: 323-333. 3O Wagatsuma, T., Minoru, K., Hayasaka, Y. 1987. Characteristics of upward translocation of aluminum in plants. Soil Sci. Plant Nutr. 30:345-358. Wagatsuma, T., Jujo, K., Ishikawa, S., Nakashima, T. 1995. Aluminum-tolerant protoplasts from roots can be collected with positively charged silica microbeads: a method based on differences in surface negativity. Plant Cell Physiol. 36: 1493-1 502. Wheeler, D.M.,. Edmeades, D.C., Christie, R.A., Gardner, R. 1993. Comparison of techniques for determining the effect of aluminium on the grth of, and the inheritance of aluminium tolerance in wheat. In Randall, P.J., Delhaize, E., Richards, R.A., Munns, R. (eds). Genetics aspects of plant mineral nutrition. pp 9-16. Kluwer Acad. Publ., Dordrecht. Williams, R.J.P. 1992. Aluminium in biology: an introduction. In Chadwick, D.J., Whelan, J. (eds). Aluminium in biology and medicine. pp 14 John Wiley & Sons, New York. Williams, R.J.P. 1993. Magnesium: An introduction to its biochemistry. In Birch, N.J. (ed). Magnesium and the cell. pp 15-30. Academic Press, London. Yan, F., Schubert, S., Mengel, K. 1992. Effect of low root medium pH on net proton release, root respiration, and root growth of corn (Zea mays L.) and broad bean (Viciafaba L.). Plant Physiol. 99: 415-421 . Youson, J ., Neville, C. 1986. Deposition of aluminum in the gill epithelium of rainbow trout (Salma gairdneri Richardson) subjected to sublethal concentrations of the metal. Can. J. Zool. 65:647-656. Zhang, G., Taylor, G.J. 1989. Kinetics of aluminum uptake by excised roots of aluminum tolerant and aluminum sensitive cultivars of T riticum aestivum L.. Plant Physiol. 91: 1094-1099. Zhang, G., Taylor, G.J. 1990. Kinetics of aluminum uptake in T riticum aestivum L.. Identity of the linear phase of aluminum uptake by excised roots of aluminum-tolerant and aluminum-sensitive cultivars. Plant Physiol. 94: 577-584. Zhang, G., Taylor, G.J. 1991. Effects of biological inhibitors on kinetics of aluminium uptake by excised roots and purified cell wall material of aluminium-tolerant and aluminium-sensitive cultivars of T riticum aestivum L. J. Plant Physiol. 138:533-539. Zhao, X.J., Sucoff, E., Stadelmann, E.J. 1987. Al3+ and c5: alteration of membrane permeability of Quercus rubra root cortex cells. 83:159-162 Zirnmerberg, J ., Vogel, S.S., Chemomordik, L.V. 1993. Annu. Rev. Biophys. Biomol. Str'uct. 22: 433-466. 31 Zsoldos, F ., Erdei, L. 1981. Membrane and ion transport properties in cereals under acidic and alkaline stress. I. Effects of pH on potassium uptake and grth of rice and wheat. Physiol. Plant. 53: 468-470. CHAPTER 2 DEVELOPMENT AND VALIDATION OF AN IN VIVO ALUMINUM-MORIN FLUORESCENCE ASSAY FOR THE VISUALIZATION AND QUANTITATIVE DETECTION OF ALUMINUM IN TOBACCO BY-2 CELL SUSPENSIONS ABSTRACT The development, optimization and validation of an in viva aluminum-morin fluorescence method for the detection of A1 in cultured Nicatiana tabacum L. cv. BY-2 is reported. Morin, a fluorogenic reagent for Al, was added to cells in an aqueous medium (pH 5.6) compatible with cellular integrity, and was found to accumulate rapidly (<30 min) in these tobacco cells. Accumulation occurred presumably via pH-dependent trapping in the cytoplasm, due to virtually complete dissociation (99%) of the 3-hydroxyl group at the higher cytoplasmic pH, as opposed to that (80%) in the loading medium (pH 5.6). The excitation (AW=42O-450 nm) and emission ()11m =500-520 nm) spectra of the Al-morin complex in cells and in aqueous solution were similar. This fact and the intensity and spectra of cell autofluorescence confirmed that fluorescence from cells treated with morin and/or Al originated almost entirely from morin or its complex with Al. Based on the emission spectra from cells, the Al-morin complex can be distinguished visually from free morin by intensity and coloration. Maximum fluorescence intensity was obtained by incubating cells with 100 MM morin. At higher concentrations, fluorescence intensity is decreased by quenching. The validity of this method for assessing Al in cells was tested by comparing results with graphite furnace atomic absorption spectrometry (GFAAS). Under various experimental conditions leading to wide-ranging levels of A1 uptake, results obtained fiom the Al-morin method were highly correlated (r2=0.94) with those obtained with GFAAS. The advantages of this method are several. It is quick, easy and inexpensive to perform, and permits quantitative 32 33 measurements as well as visualization under the microscope, with resolution comparable or better than other methods. This method is also apparently insensitive to precipitated Al. INTRODUCTION The toxic effects of aluminum on plants and its impact on crop productivity and forest decline in large areas throughout the world are well documented in the literature (Haug, 1984; Delhaize and Ryan, 1995; Kochian, 1995). Increases in Al contents of surface waters above previously innocuous levels, due to acid rain, have made it necessary to monitor Al in water and aquatic organisms. There is, however, a dearth of knowledge regarding mechanisms of Al uptake and toxicity in cells. Difficulties in Al toxicity studies result from problems such as the complex chemistry of Al (Martin, 1988; Delhaize and Ryan, 1995), particularly its pH-dependent hydrolysis. Precipitation of Al hydroxides, particularly within the cell wall, can result in false estimations of uptake and localization of A1 within the cell, when measured by graphite furnace atomic absorption spectrometry (GFAAS) or by microanalytical techniques. Also, the binding of Al to cell walls and its uncertain removal by chelating agents have made the distinction between uptake into the protoplast and apoplast difficult (Zhang and Taylor, 1990; Tice et a1, 1992; Rincon and Gonzales, 1992; Delhaize and Ryan, 1995; Kochian, 1995). Because progress in this field has in part been limited by methodology, further progress may require the use and further development of new techniques (Delhaize and Ryan, 1995) Ideally, a method for the detection of Al in cells should allow both the quantification and intracellular localization of A1 with high sensitivity and resolution. It should also 34 35 distinguish between precipitated and non-precipitated Al and should be easy, quick and inexpensive to perform. Most studies have used GFAAS, X-ray microanalysis and more recently, secondary-ion mass spectrometry for the detection and/or localization of cell- associated Al (Zhang and Taylor, 1990; Delhaize et al., 1993; Lazof, et a1. 1994). However, these methods do not meet several of the above mentioned requirements (e.g. see Lazof et al., 1994). Although application of secondary-ion mass spectrometry has improved sensitivity and resolution over X-ray microanalysis, it is still difficult to relate images of Al content with any cellular structure (Figs. 5 and 6 of Lazof et al., 1994) and requires extensive instrumentation and sample preparation. An alternative method involves the use of a flavonoid, morin (3, 5, 7, 2', 4' - pentahydroxyflavone). This highly sensitive fluorogenic reagent is known in analytical chemistry for the qualitative and quantitative determination of Al (Fig. l) in solutions (Katyal, 1968; Katyal and Prakash, 1977). It has also been used in histochemical applications as a dye for Al in microscope slide preparations (Eggert, 1970; Bani et al., 1974; Wen and Wisniewski, 1985) and has recently been used by Tice et a1. (1992) in the examination of apoplastic and symplastic Al fractions of wheat roots. Morin can bind Al even at low pH values of 4.5, thus avoiding precipitation of Al, but because it is poorly soluble in water at this pH, a water/alcohol/acetate mixture was generally employed. However, usage of morin in aqueous medium, at other pH values, was demonstrated in studies on the aqueous chemistry of morin (Browne et a1, 1990). The development, optimization and validation of an in viva Al-morin fluorescence method for the detection of Al in intact tobacco BY-2 cells is reported. Validation of this method was assessed by comparison of results with GFAAS. 36 0H + AI(OH),,<3'">+ o \AI(0H),<2-n)+ Figure 1. Structure of the flavonoid morin (3, 5, 7, 2', 4' - pentahydroxyflavone) and presumed structure of its complex with A1. MATERIALS AND METHODS Cell cultures and Al uptake Tobacco (Nicatiana tabacum L. cv BY-2) cell cultures were grown as described (Nagata et al., 1992), at 28°C, in 50 ml of liquid medium in 250 ml flasks and on a gyratory shaker set at 165 rpm. The medium consisted of the Murashige and Skoog salts (Nagata et al., 1992) supplemented with 30 g sucrose, 0.5 g MES, 100 mg myo-inositol, 1 mg thiamine- HCl, 180 mg KHZPO4, and 0.22 mg 2,4-D (2,4-dichlorophenoxyacetic acid) per liter of medium and the pH was adjusted to 5.7. Subcultures were made every 7 days with a 2.5 ml inoculum (5% v/v dilution). Cells at two or three days after subculture were used in the experiments. To prevent precipitation of Al, or formation of Al complexes, Al uptake experiments were performed in a medium consisting of 10 mM MES, 5 mM sucrose, 5 mM KCl and 2 mM CaCl2 (pH 4.5). Unless stated otherwise, the concentration of Al was 20 uM. Experiments were performed at room temperature (23 °C). Stock solutions of AlCl3 (lmM) were employed. Cells were first pre-washed for 10 min in uptake medium at a ratio of 40 ml/cm3 of packed cells on a gyratory shaker at 60 rpm. Cells were then distributed into 15 ml plastic tubes and, following removal of the pre-wash supernatant, 4 m1 of Al-containing (20 uM Al) uptake medium were added. The final cell titer was 4 x 10’ cells/ml. Cells were incubated with A1 during their allocated time on a shaker at 60 rpm. The pH of the cell suspension was monitored in both the pre-wash and incubation steps and was found to be stable during experiments. 37 38 Incubation of cells with Al was terminated by adding 0.1 M citrate (pH 4.5) to a final concentration of 5 mM, followed by removal of the Al-containing supernatant (operationally defined as the first wash). For samples destined to GFAAS, cells were washed twice more with 4 ml of uptake medium containing 5 mM citrate (pH 4.5) to further reduce the presence of extracellular and cell wall A1. For some of the optimization experiments, different levels of A1 uptake were desired and generated by using cells at two different growth stages (2 or 3 days after subculture). In these cases, results were normalized and expressed as percentage of response. Ion exchange cellulose (pre-swollen high capacity carboxymethyl ion exchange cellulose) from Pierce Chemical Co. (Rockford, IL, USA) was used as a control for A1 precipitation. Cellulose treatment was identical to that of cells, with 5 mg dry weight of cellulose substituting for the 0.2 cm3 of packed cells. Morin solutions and cell loading Morin (3, 5, 7, 2', 4' - pentahydroxyflavone, C,5HIOO7, F.W. 302.2 g) was from Sigma (St. Louis, MO). Stock solutions fi'om two different lots of morin differed in intensity of coloration, but did not produce different results in the fluorescence assays. Morin was prepared as a 0.1 M solution in DMSO. This concentrated stock solution was stable and resulted in only a small carry-over of DMSO, which is commonly used in experimental cell biology. In aqueous solution, morin was poorly soluble at pH values between 4 and 5. Upon addition of morin to water, the pH of the solution slowly dropped from neutral to about 4.5 39 with resultant precipitation. To prevent this, aqueous solutions were always prepared in solutions buffered with 5 or 10 mM MES. Solubility and the development of a yellow coloration increased with increasing pH. Consistent with this pH-dependent behavior, the absorbance peak of morin in aqueous solution was found to shift from 360 to 385 nm, and to increase in intensity by 64% in solutions of pH 4.8 to 6.5 (data not shown), as previously reported in aqueous solutions (Browne et al., 1990). For comparisons, morin was also prepared in an alcoholic solution of 0.2 % morin in 85% ethanol and 0.5% acetic acid, pH 4.5. Following exposure to Al, cells were incubated with morin in fresh uptake medium (pH 5.6) at a titer of 4 x 105 cells/m1. Morin concentrations ranging from 10 to 200 ,uM and lengths of incubation ranging from 5 to 60 min were employed. The composition of this medium was the same as that used for Al uptake except for a pH of 5.6. Although more soluble at higher pH, morin was sufficiently soluble at pH 5.6 and presented a light yellow coloration. After incubation with morin, the supernatant was removed and cells were resuspended with 3.6 ml of uptake medium (pH 5.6) plus 0.2 ml of 0.1 M EDTA (pH 5.6). The purpose of the presence of EDTA was to quench extracellular Al-morin fluorescence. Final cell titer was 4 x 10’ cells/ml. Fluorescence intensity measurements Measurements of fluorescence intensity were performed on a Perkin Elmer (N orwalk, CT, USA) LS-5 spectrofluorometer coupled to a microcomputer. Spectra were scanned at 120 nm/min, with a step interval of 0.5 nm. For fluorescence intensity measurements, at 40 fixed wavelengths, readings were monitored for 5 or 15 s at 0.5 5 intervals, for solutions and cell suspensions, respectively. In all measurements slit widths of 10 and 5 nm for excitation and emission, respectively, were employed. Graphite Furnace Atomic Absorption Spectrometry Following Al uptake and citrate washes to remove extracellular Al and Al bound to the cell wall, cell pellets or cellulose material were digested with 250 [21 of 8 N HNO3 at 135 °C in Teflon screw-cap vials. After excess acid was dried off, samples were redissolved in 1 or 2 ml of 25 mM HNO3 and analyzed for A1 with a Varian (Sunnyvale, CA, USA) S pectrAA 400 graphite furnace atomic absorption spectrometer with Zeeman background correction. The furnace program consisted of seven temperature steps set at 85, 95, 120, 1000, 1400, 2500 and 2600 °C for 5, 40, 10, 5, 6, 2.8, and 2 s, respectively. Pyrolytically partitioned graphite tubes were flushed with N2 gas. Sample volume was 20 a1 and readings were performed during atomization at 2500°C, at 396.2 nm, with a slit width of 0.5 nm. Absorbance values were based on peak height and standard solutions of AlCl3 from Sigma (St. Louis, M0) were used for calibration. RESULTS Morin uptake by cells Uptake of morin into cells (incubation at pH 5.6) occurred largely within the first 30 min of incubation, as judged by the development of fluorescence from the cells (Fig. 2). Based on the decrease in absorbance at 380 nm of the external solution following a 20 min period of uptake, it was estimated that approximately 14 % and 10% of the morin applied was taken up by the cells at pH 5.6 and 6.5, respectively, at three different concentrations of morin (20, 50 and 100 aM). Morin accumulation was found to be homogeneous throughout the cytoplasm, with little or no accumulation of morin within the vacuole or cell wall (V itorello and Haug, 1996). Incubation of cells with morin or with concentrations of DMSO equivalent to those carried over with morin, was found to have no affect on cell viability, as judged by trypan blue. Spectra The excitation and emission spectra of morin and Al-morin in cells were similar to those recorded in aqueous solution and in a water/ethanol/acetate mixture (Fig. 3). The excitation peak ranged from 420-450 nm and the emission peak ranged from 500-520nm, similar to results previously reported in the literature. Excitation maxirna have been reported to range from 410-440 nm and emission from 510-525 nm (Ueno et al., 1992; Saarl and Seitz, 1983; Browne et al., 1990). The excitation spectrum of Al-morin in cells was 41 140 - 100 120 - .L A cells + Al .3 + morin c in— E‘ v o g 80 A g '2 - 60 g 3 m o 3 :: S : ° — 4o 8 8 cells - Al 3 3 + morin a a B I - 20 = = = ———— + Al autofluorescence - Al O I I I I I I - O o 10 ' 20 30 4o 50 60 70 time of incubation with morin (min) Figure 2. Uptake of morin into cells of tobacco BY-2 cultures and fluorescence intensity from Al-treated cells as a function of duration of incubation with morin. Cells (2 and 3 days after subculture) were pro-washed in uptake medium pH 4.5, then incubated for 1 h with 0 or 20 aM A1 (pH 4.5). Following this, cells were incubated with 100 11M morin in uptake medium, pH 5.6, then resuspended in the same medium, free of morin and containing 5 mM EDTA. To verify whether the concentration of Al in the cells determined the time needed for morin incubation, differing levels of Al uptake were obtained by using cells of different ages, and the resulting Al-morin fluorescence intensities were normalized (ordinate on the right). Uptake medium consisted of 10 mM MES, 5 mM sucrose, 5 mM KCl and 2 mM CaCl,, pH 4.5 or 5 .6. Bars represent SE (n=3). 43 Figure 3. Excitation and emission spectra of free morin and Al-morin in tobacco BY-2 cells (A) and in an aqueous or ethanol/water/acetate solution (B). Scanning was performed at 120 nm/min. The slit widths were 10 and 5 nm for excitation and emission, respectively. A. Aluminum uptake by cells (2 days afier subculture) and morin loading were performed as described in Fig. 2. B. The morin concentration was 10 and 50 “M in the aqueous solution (uptake medium, pH 5.6) and 6.6 mM in the 85% ethanol/0.5 % acetic acid solution. The Al concentration was 10 MM. fluorescence (arbitrary units) fluorescence (arbitrary units) 200 excitation emission 150 - + Al, + morin + A' I monn 100 — 5O - " Al, "’ ”0'1“ - Al, + morin f“? A ‘~ g = = _ -- "-=. g 0 autofluorescence autofluoresce-r: “- B excitation emission 200 — 10.3.32?" CQUQOUS '0' AI 150 -' ethanol! acetate 100 '7 aqueous. aqueous 50 Pr£°"“ 5O - 50 uM morin 4' Al O " — = = "" = = = = __ __ ._ 380 400 420 440 460 480 500 520 540 560 580 wavelength (nm) Figure 3. 45 closer to that found in aqueous medium than to that obtained from a water/ethanol/acetate mixture whereas the opposite occurred for the emission spectrum (Fig. 3). These small differences in the excitation and emission spectra of the Al-morin complex in cells compared to those in solution are expected given the more diverse physico-chemical environments within cells as opposed to that in simple solution. Shifts in the spectra of Al-morin depend on the chemical environment (Escriche et al., 1983; Saarl and Seitz, 1983). Optimization of Al-morin fluorescence intensity As a method for detection of Al in these tobacco cells, the use of Al-morin fluorescence was optimized, aiming for maximum fluorescence intensity and sensitivity. At a fixed duration of incubation with morin (10 min), the fluorescence intensity from cells treated with Al depended on the concentration of morin in the labelling medium (Fig. 4), increasing with concentration up to about 100 “M, after which it decreased as a result of quenching of fluorescence intensity, presumably by free morin. This quenching pattern was also observed in solution albeit at concentrations of morin one order of magnitude lower. Quenching probably explains why the Al-morin fluorescence signal was fully developed within the first 5-10 min of incubation with morin (Fig. 2), although morin uptake continued to increase and was completed only after about 30 to 60 min of incubation (Fig. 2). Thus, the greater accumulation of morin within the cell was counter-balanced by quenching. Since quenching did occur, optimization of cell titer for fluorescence intensity readings was also performed (insert Fig. 4). At the cell densities used, no decrease in fluorescence at higher cell densities was observed, but an inflection in the fluorescence 46 120 100 - $ 2' 80 - 0 g 250 a A 2 days In 60 — 0 3 200 - +N 2 g 'E E 3 3 150 — 8 3 E 3 days a 40 "‘ 8 g 100 _ '0' AI e a e h 5 50 - cells-Al 3‘2? .— a— -- -" ‘— 20 _ 0 ”re 2 days I l l I o 200 400 600 800 cell titer (1o3 cells/ml) o l l l l O 50 100 150 200 Figure 4. Dependence of cell Al-morin fluorescence intensity on the concentration of morin in the incubation medium and the cell titer during readings (insert). Al uptake by cells (1 h, 20 uM A1, pH 4.5) and morin loading were performed as described in Fig. 2. The effect of morin concentration was tested with different levels of A1 uptake, obtained by using cells of different ages (2 and 3 days after subculture). The resulting Al-morin fluorescence intensities were normalized and expressed as percentage of maximum response. Bars represent SE (n=3). morin concentration (uM) 250 47 intensity curve was observed between 3 and 5 x 105 cells/ml. Although less pronounced, this presumably resulted from quenching by free morin and/or absorption of light by the cells themselves. Because fluorescence readings fiom these cells are performed on a suspension rather than in solution, readings fluctuated because of the random movement of the cells within the cuvette and as a result of sedimentation of cells in the cuvette. From time course readings of fluorescence intensity (Fig. 5), reductions in fluorescence intensity due to settling of cells were detectable after about one minute. Despite the problem of sedimentation of cells, the signal was sufficiently stable for enough time so that measurements could be made. To compensate for fluctuations in readings during this time, measurements were made at 0.5 5 intervals and averaged over 10 to 30 5. When cells were allowed to settle to the bottom of the cuvette afier a few minutes (Fig. 5), the fluorescence signal decreased and approached zero, reaching even lower values if the cells were centrifuged. Thus, fluorescence originating from the solution bathing the cells was minimal, but was higher in cells treated with morin in an alcohol/acetate solution, which compromises cell integrity. Additionally, EDTA was added at a concentration of 5 mM to the cell suspensions following incubation with Al and morin and was always present during fluorescence intensity readings. In solution, 5 mM EDTA (A1 binding affinity constant of log K=16.5; Orvig, 1993) was capable of quenching all fluorescence (99% reduction) from the Al-morin complex (binding affinity constant estimated to be log K= 4.2-6.5; Katyal and Prakash, 1977; Saarl and Seitz, 1983) and can thus by utilized to further reduce the contribution of fluorescence emanating from the bathing solution. Citrate, with an Al binding affinity constant (log K=7.8) closer to that of morin, was less effective at quenching Al-morin fluorescence (37% reduction). b 00 100 I a V100 _ m 8 r: 80 - 8 e"? 80 - a, 50 _ a: 2 C a 40 - 3 5 E‘ 2 20 - (U ‘ or E 60 _ D. 0 I 1 fl 1 1 fi- 0 20 40 so 80100 3' time following morin u incubation (min) 5 40 - o 0 in o I. 3 -Al : 20 - + morin o I . I ' . O 50 100 150 200 250 time (seconds) Figure 5. Readings of Al-morin fluorescence intensity from tobacco BY-2 cells as affected by movement and sedimentation of cells and stability of the Al-morin fluorescence signal with time (insert). For readings during sedimentation, measurements were made at l 5 intervals and collected with a microcomputer. Cells (3 days after subculture) had almost completely sedimented within the time frame involved. To test the stability and reproducibility of the Al-morin fluorescence signal, readings were made on the same cell samples at different times following morin incubation. This was tested in cells with different concentrations of A1, by using cells at 2 and 3 days after subculture and the resulting Al- morin fluorescence intensities were normalized and expressed as percentage of maximum response. Cells were incubated with Al (1 h, 20 uM A1, pH 4.5) and morin (100 uM, 10 min) as described in Fig. 2. Bars represent SE (n=3). 49 The Al-morin fluorescence signal from cells was found to be repeatable overtime (insert Fig. 5). Essentially the same readings could be obtained from the same sample over a period of about an hour, with small decreases thereafter. In a spectrophotometric application for the determination of Al by morin (Ahmed and Hossan, 1995), absorbance values were found to be stable for 48 hours. Comparison of the Al-morin method to GFAAS From the optimization experiments described above, a desirable compromise of conditions was established, taking into consideration maximum fluorescence intensity for greater sensitivity, quenching, time spent, stability and reproducibility of results. This was achieved by loading cells with morin at pH 5.6, 100 uM morin, for 10 min. To ultimately validate this Al-morin technique, a comparison was made between results obtained from this technique and graphite furnace atomic absorption spectrometry, a standard analytical technique for quantification of Al, and the predominant analytical method employed in Al uptake studies. Good agreement (r2=0.94) was found between results from these two techniques (Fig. 6) when both were used to measure cell-associated A1 in parallel samples. This comparison covered wide-ranging levels of Al content, which had been obtained under various experimental conditions (V itorello and Haug, 1996). The above comparisons were evaluated at a pH of 4.5, at which Al precipitation does not occur at [A1] 3 20 uM. Control experiments with cellulose provided supporting evidence that precipitation did not occur. This control experiment was undertaken because GFAAS will also quantify precipitated Al present in the cell pellets. In contrast, the Al-morin 50 25 20 _ y=x*0.13+0.71 O O a; r2=0.94 2 '5 15 _ 2o” < E (“5 a .0 10- E r: 5.. 0 l i I l l I i l 0 20 40 60 80 100 120 140 160 AI-morin fluorescence (arbitrary units) Figure 6. Correlation between results of A1 uptake by tobacco BY-2 cells measured by the Al-morin method or graphite furnace atomic absorption spectrometry (GFAAS). Al uptake was performed under different experimental conditions and resulted in wide-ranging levels of Al uptake and concentrations of cellular A1. Factors which were varied were: age of cells, time of incubation with Al, concentrations of CaCl2 in the uptake medium, concentrations of sorbitol in uptake medium, temperature, pre-chelation of A1 to citrate and pH of the pre- wash solution (V itorello and Haug, 1996 - chapter 3). In all cases, cells were incubated with 100 uM morin for 10 min, following Al uptake. 51 technique appeared to be insensitive to precipitated Al (Fig. 7). When Al (20 uM) was incubated with cells in a solution at pH 5.6, a pH value which will knowingly lead to at least some precipitation of unchelated Al (Kragten, 1978), uptake results differed between the A1- morin technique and GFAAS. Compared to pH 4.5, uptake at pH 5.6 measured with GFAAS indicated increased levels of Al associated with the cells, whereas the Al-morin method reported less Al associated with the cells (Fig. 7). This discrepancy in results can be attributed to precipitation of Al, which can cause overestirnations of Al uptake. That precipitation of Al did indeed occur was evidenced by control experiments with ion exchange cellulose (Fig. 7). At pH 4.5, but not at pH 5.6, Al was quantitatively removed from the cellulose after citrate washes, as determined by GFAAS. At pH 5.6, even after the citrate washes, about 60 % of the applied Al was found to be associated with the cellulose, with nearly identical results in cells, indicating precipitation of Al. Intemalization of Al by the cell should be expected to decrease at pH 5.6 since Al is not toxic at this pH and a correlation between A1 toxicity and Al uptake has been demonstrated in tobacco cell cultures (Yamamoto et al., 1994). Thus, GFAAS can overestimate Al uptake when precipitation occurs whereas the Al-morin method probably reports results closer to genuine uptake. It should be noted that precipitation can also occur at lower pH values if rather high Al concentrations are used. In addition, Al solutions are most often unbuffered. Thus, even though Al may be added in a solution of sufficiently low pH, high A1 levels, close to the limit of solubility, associated with fluctuations in pH might cause precipitation of Al and overestimation of Al uptake. 52 150 (55 125 - — 50 ‘6 E _ — 45 a E. 100 J < I- x < E in" I 5, 75 - [15 : at < O In a a, a % o _ o g 50 g g ~10 E g o 2 a: g g. 25 ' o 3 — 5 3 O E O o — o a b c a b c pH 4.5 pH 5.6 Figure 7. Aluminum uptake by tobacco BY-2 cells and by cellulose phophate, at pH 4.5 and 5.6, as reported by Al-morin fluorescence intensity or GFAAS. Cells were incubated with Al (1 h, 20 “M Al) and morin (100 uM, 10 min) as described in Fig. 2, with the exception of the pH of the medium.Cellulose phosphate was treated identically to that used for the cells, with 5 mg substituting for each 0.2 cm3 of packed cells. Bars represent SE (n=3). DISCUSSION The major findings of this report are that morin can be successfully loaded into intact cells of tobacco BY-2 cells in an aqueous medium, and that morin can be used to reliably detect Al in these cells following uptake. The conditions for use of this technique were optimized for maximum fluorescence intensity. As discussed below, this technique has several advantages over previously used methods. Loading of intact cells with morin, in an aqueous medium, was possible presumably because of its pH-dependent solubility behavior (Fig. 2). Compared to the extracellular solution employed (pH 5.6), the higher pH value of the cytoplasm allows for more extensive dissociation of the 3-hydroxyl group, higher solubility of morin and thus presumably results in a pH-dependent trapping of the dye. The pK., for the first hydroxyl group (3-hydroxyl group) to undergo dissociation has been estimated to be 5.0 (Browne et al, 1990). At pH 5.6, about 80% of morin will be dissociated in contrast to 99% at pH 7.0. Thus, at the lower pH, there is still a reasonable quantity of undissociated, neutral morin to traverse the plasma membrane whereas there is almost none once the compound reaches the interior of the cell. That the dye is contained within the protoplast has been demonstrated by plasmolysis experiments (V itorello and Haug, 1996). The same principle of pH-dependent ion trapping, which can explain entry of morin into the cell can also explain why the dye was not detected to accumulate within the vacuole. Loading of cells with naringin (4', 5, 7 - trihydroxyflavanone 7- rhamnoglucoside) a glycosylated flavonoid, with higher solubility in water than morin, was also attempted but was unsuccessful. From the excitation and emission spectrum of Al and morin loaded cells and from 53 54 relative intensity values (Fig. 3.), it is safe to conclude that the fluorescence measured from these cells originated largely from morin and its complex with Al, and that the fluorescence behavior of this reagent is not significantly altered by the cellular medium. Additionally, as confirmed by the emission spectrum of morin and Al-morin in cells, a distinction between complexed and non-complexed morin can be achieved based on coloration and intensity, which is of importance for the observation of cells under the microscope. Complexed morin is bluish-green, in agreement with its emission maximum, while free morin, because of its low and broad emission spectrum, appears olive-green or yellowish green. Actual color tones may, however, vary depending on the filter set used in the microscope. In addition to determining conditions for maximizing fluorescence intensities, the fact that quenching of the morin signal occurs, may be advantageous for reproducibility of results. Employing morin concentrations in the plateau area probably helps stabilize the fluorescence signal against small fluctuations in morin uptake resulting from variances in duration of incubation or possible heterogeneities in cell populations (Figs. 2 and 4). The advantages of this Al-morin method for the detection of Al in cells are manifold. It is easy, quick, and inexpensive to perform. It allows both for quantitative measurements and for visualization and localization under an epifluorescence microscope, using filter sets similar to those used for the common fluorescent dye fluorescein. In comparison to other methods for the cellular localization of Al, it has higher resolution, which theoretically should approach the limit of optical microscopy. Additionally, superior imaging of cellular Al could probably be obtained if confocal microscopy were utilized. Another advantage of this method is that it is apparently not sensitive to precipitated Al. Another possible advantage of this technique is that it probably distinguishes between 55 apoplastic and symplastic Al. Evidence for this and a discussion are presented by Vitorello and Haug, 1996. Whether or not Al can be totally removed from the cell wall, from sites with unknown binding affinity, by use of chelating agents has been disputed. If contained within the cell wall, fluorescence of the Al-morin complex is achieved either because morin strips Al off these sites or (more likely) because of the possibility of multiple ligand binding. In either case it should still be possible, in principle, to quench Al-morin fluorescence with EDTA, because the relative binding affinities of morin and EDTA are known, regardless of whether Al is actually removed from its site on the cell wall. Whereas the application of this technique to cell suspensions is straightforward, its application to other plant model systems, such as roots, may require some adaptations, but should nevertheless prove to be advantageous. As for observations with the microscope, tissues made up of several layers of cells could be observed by confocal microscopy. Despite its numerous advantages, this method reports only relative amounts of Al, not absolute ones. Thus it is imperative that when using this method in a new cell or tissue system, that results obtained with Al-morin be initially checked or calibrated with some other standard method of Al quantification. This may also be important since the chemical environment, particularly whether hydrophobic or hydrophilic, of the Al-morin complex can alter fluorescence levels. Al-morin is an advantageous technique to be employed when numerous samples are being studied, such as in screening programs. If not used as the primary method for detection of Al, it would be advantageous as a preliminary method, such as in exploratory experiments, which could be later confirmed with other methods of analysis such as GFAAS. REFERENCES Ahmed, M.J., Hossan, J. 1995. Spectrophotometric determination of aluminium by morin. Talanta 42:1135-1142. Boni, U. De, Scott, J.W., Crapper, DR. 1974. Intracellular aluminum binding: a histochemical study. Histochemistry 40:31-37. Browne, B.A., McColl, J.G., Driscoll, CT. 1990. Aluminium speciation using morin: I. Morin and its complexes with aluminium. J. Environ. Qual. 19: 65-72. Delhaize, B, Ryan, RR. 1995. Aluminum toxicity and tolerance in plants. Plant Physiol. 107: 315-321. Delhaize, E., Craig, S., Beaton, C.D., Bennet, R.J., Jagadish, V.C., Randall, P.J. 1993. Aluminum tolerance in wheat (T riticum aestivum L.) I. Uptake and distribution of aluminum in root apices. Plant Physiol. 103: 685-693. Eggert, DA. 1970. The use of morin for fluorescent localization of aluminum in plant tissues. Stain Technol. 45: 301-303. Escriche, J .M., De la Guardia Cirugeda, M., Hernandez, F .H. 1983. Increase in the sensitivity of the fluorescent reaction of the complexing of aluminium with morin using surfactant agents. Analyst. 108: 1386-1391. Haug, A. 1984. Molecular aspects of aluminum toxicity. Crit. Rev. Plant Sci. 1: 345-3 73. Katyal, M. 1968. Flavones as analytical reagents - a review. Talanta 15: 95-106. Katyal, M., Prakash, S. 1977. Analytical reactions of hydroxiflavones. Talanta 24: 367-375. Kochian, L.V. 1995. Cellular mechanisms of aluminum toxicity and resistance in plants. - Annu. Rev. Plant Physiol. Plant Mol. Biol. 46: 237-60. Kragten, J. 1978. Atlas of metal-ligand equilibria in aqueous solution. John Wiley & Sons, New York, pp 16-39, 52-68. 56 57 Lazof. D.B., Goldsmith, J .G., Rufty, T.W., Linton, R. 1994. Rapid uptake of aluminum into cells of intact soybean root tips. A microanalytical study using secondary ion mass spectrometry. Plant Physiol. 106: 1107-1114. Martin, RB. 1988. Bioinorganic chemistry of aluminum. In: Sigel, H., Sigel, A. (eds). Metal ions in biological systems: Aluminum and its role in biology, vol. 24, pp 2-57. New York, Marcel Dekker. Nagata, T., Nemoto, Y., Hasezawa, S. 1992. Tobacco BY-2 cell line as the “HeLa” cell in the cell biology of higher plants. Int. Rev. Cytol. 132:1-30. Orvig, C. 1993. The aqueous coordination chemistry of aluminum. In Robinson GH. (ed). Coordination chemistry of aluminum . pp 85-121. VCH Publishers. New York. Rincon, M., Gonzales, RA. 1992. Aluminum partitioning in intact roots of aluminum- tolerant and aluminum-sensitive wheat (T riticum aestivum L.) cultivars. Plant Physiol. 99:1021-1028. Saarl, L.A., Seitz, W.R. 1983. Immobilized morin as fluorescence sensor for determination of aluminum (1H). Anal. Chem. 55: 667-670. Tice, K.R., Parker, D.R., DeMason, D.A. 1992. Operationally defined apoplastic and symplastic aluminum fractions in root tips of aluminum-intoxicated wheat. Plant Physiol. 100: 309-318. Uemura, E., Minachi, M. and Lartius, R. 1992. Enhanced neurite growth in cultured neuroblastoma cells exposed to aluminum. Neuroscience Letters 142: 171-174. Ueno, K., Irnamura, T., Cheng, KL. 1992. Handbook of Organic Analytical Reagents. CRC Press, Boca Raton, pp 25-29, 158. Vitorello, VA. and Haug, A. 1996. Short-term aluminium uptake by tobacco cells: Growth dependence and evidence for internalization in a discrete peripheral region. Physiol. Plant. (in press). Wen, G.Y. and Wisniewski, HM. 1985. Histochernical localization of aluminum in the rabbit CNS. Acta Neuropathologica 68: 175-184. Zhang, G., Taylor, G.J. 1990. Kinetics of aluminum uptake in Triticum aestivum L.. Identity of the linear phase of aluminum uptake by excised roots of aluminum-tolerant and aluminum-sensitive cultivars. Plant Physiol. 94: 577-584. Yamamoto, Y., Rikiishi, 8., Chang, Y., Ono, K., Kasai, M., Matsumoto, H. 1994. Quantitative estimation of aluminium toxicity in cultured tobacco cells: correlation between aluminium uptake and growth inhibition. Plant Cell Physiol. 35: 575-583. CHAPTER 3 SHORT-TERM ALUMINIUM UPTAKE BY TOBACCO CELLS: GROWTH DEPENDENCE AND EVIDENCE FOR INTERNALIZATION IN A DISCRETE PERIPHERAL REGION Vitorello, V.A., Haug, A. 1996. Physiologia Plantarum (in press) 58 PHYSIOLOGA MW 96: xxx-an. 1996 mum-aw“ womb-Int!" must-an Short-term aluminium uptake by tobacco cells: Growth dependence and evidence for internalization in a discrete peripheral region VidorLVltorelloandAlh'edflang V.A.aud A.l996.Shut-unmaluminimunakebytohaeeoeens: “mile. m . . il I wuemtammham 96: ii 2 5 E m- 222 222 22 22 =1 2 5% 2" 232 2,21% . f 2 .222- ,3 :2 2 E 3 2 E F i 2 22 E E E 2 mm-Wmmmwmmnw www.mmmmammm uh. W. SA:CaInde£ns iaNadeerna Bo “Plant? M 'mSmUuEmEaser- W m W mascara». ..Brusil (W “1.1. Hang (eon!- MW; Dept alum um. Sum um. East mm. MI 43324. . lntruduetion Aluminium toxicity is a major factor limiting plant pounhmacidicsoilseovefingvastregionsoftheworld (Foyetal. 1978.ng l984).andhasalsoreeeived rec- ognition asapotential probleminhumaus and animals (Ganrot 1986). Despite considerable research endeav- ans.meehanismsofAltoxieityarestillpoorlymder- modpuucularlywithrespeetwcellularsitesofwxie- ity(l(oehian l995).ltisstilldisprnedwhetherinitial micefl’eetsofAlnemanifestedintheeellwalla withiutheplanteellLStudiesouAluptakeanditseellu- Received 180mb.l995:uvisad6m1996 “~91!” h localiztim are central to the elucidation of such Ryan 1995.Kochim 1995).Diacrepaneies.inpart.result from methodological dificuhies. such as the complex aolutionchenusuyofAHOrvig l993).thelaekofasuit- able Al radioisotope. the potential fa ”face binding 6 AlandformatimofAlruecipitatesintheeellwallmin- laidel991.'liceetal.l992).aswellastheratherlinuted resolution of microanalytical techniques. such as elec- uoupmbex-raynucmaualyais (Lazofetal. 1994). Concerning Al uptake studies. Zhang md Tayla' (1989.1990)analyaedthekinetiesofAluptakewiththe 59 purposeofdistinguishingbetweenAlresidingintheap- oplastarrdthatinthesymplast'l'tceetal.(1992)deviscd afractionationprocedureinroots.andmorerecently Lard et al. (1994) utilized secondary-ion mass spec- nometrytoexaminethedistributionot'Alinrootapices. Onbalance.evidenceappearstofavourtheideathatAl israpidlytakenupintotherootapicalsymplamal- thoughconsensusisapparentlylackingandpoblems esistineachoftheseapproaches(lcocbianl995). Apmfiomhausfikewmpuuuremwnmd Strid 1989. Zhang and Taylor 1990). genotype (Zhang and Taylor 1989). and metabolic inhibitors (Wagatsuma 1983). thereeaistsadearthofinformationregardingpa- rameters impacting A1 uptake. To illustrate. althouflr Ca"iswellknowntoameliorateAltoxicity(Kinraide andParker1987.DelhaiuandRyan1995).thereis scantinformation(Alvaetal.1986)astotheactualup- takeofAlbyplantcellsanditsdepardencementerual Ca” concentrations. ForthisAluptakesmdy.weae1ectedamluuedto- baccocelllinecharactetisedbyhighgrowthratesand high homogeneity (NagataetaL1992). Guidedinpart byornprevious experimentsonmammaliancellsfihi andl-laug1990).Aluptakewasassessedbypaphitefur- nace atomic absorption spectromeu'y (GFAAS).Inaddi- tion. a novel fluorescent dye assay was developed. em- ploying morin in aqueous solution. Our results indicate thatAlisrapidlytakenupbyactivelygrowingtobacco cellswhereAlisapparentlyloulizfiathastinitially. indisceteaorresneuthecellularpaiphery. Abbreviations-HM graphitel'tanaceauuulcabsuptiou mam.mmm Materialsand methods Chm-Id hbwue MunshigeandSkoogbasalsaltmixuueandmmin (2'.3'.4',S.7-pentahydroxyflavone) were purchased from Sigma (St. Louis. MO. USA). All other chemical re- agemsusedwereofanalyticalgnde.61ssswarewasem- ployed only as the culture vessel and for some pipetting. andwuwashedm'th 10%lm03htordertodinu'nish backgrormd levels of Al. Otherwise polypropylene or polyethylene labware was utilized. KOl-l uHCl (0.1 toS M)wasusedforadjustingpl-lofallsolutions.1-lighqual- itywater. obtaineduaptoductafterreverseosmosis. deioninticnanddistillation.wasuscd'mallertperimems. Cellular. CensuspensionatlnuesofNiearimmbemLev. Bright Yellow2(BY-2)weregrownatpl-15.6underpe- viously described conditions (Nagataet al. 1992). Cul- ttu'esweregrown.at_28'C.in50mlofliquidmediumin 250ml£rlenmeyerflaskashakingatl65rpmmagyrr tory shaker. Subculttues were prepared every 7 days witha25m1-inoculum. 2 Cellgrowthwasassessedbymeastuingpackedcell vohmeG—SmhtcennifugationatZOOgingaduatcd tubes). Eachcm’ ofpackedcellscontained7.9x 10‘ cellsandcorrespondedtoo.6goffreshweigbt.Cells werecotmtedafterdisaw'egationwithsfio-Ofiaque- ous).ar70C.fcr30min.Proteincontentwasl.3tol.S mgptoteinperlo‘cells.meastuedbythe3radfud (1976)microassay . Cellmembrane "tywasassmdwithtrypan b1ue.Cellswereincubatedfor$mininal:1(v/v)mix- uueofcellstrspensionarrdo.4%(wlv)u'ypanblue.Cells inculturewerevirtuallyimpermeabletouypmblueo 98%ofcells). Ankh-uptake Uptake medium consisted of 10 mM 2-(N-mor- pholino)etlnnesulfonicacid(MES).5mMsucrose.and SmMKQpH45.Calciumconcentmimswereeither 0.2.2.0,a’mmMCa03.'l'heAlconcenuationwas20 uMinallenperiments.AlthoughMEShasapK.of6.1. itwasusedasbufierbecauscorganicacids ' em- ployedintlu‘spHrangecanchelateAl(Orvigl993).ln 1ieuofamtuesuitab1ebuffer.MBSwillat1eastresistin- creasesinpl-Lamaja'probleminAltmticitystudies. Allertperimentswereperformedatroomwmperature (23C)un1essstatedotherwise.A1uminiumstoeksolu— tims(1mM)wereprepared.gena-allythesamedayof anexperimartbyadditioncfAlClfiI-Ifitointenaely stirredwaterpeviouslyadjustedtoapl-lofSSto3A. likewiseAlChstockwassimilarlyaddedtouptakeme- diumimmedmelypriortoincubatiatofcells. Cellswetelne-washedinuptakemeditmrataratioof 40mlcm‘ofpackedcellsandincubatedfa10minma shakerat60rpn.8tandardpe-washsohuionshadapll d45andcootained20mMCaCh. Subscqueotly.cells waedisuibutedirno15mlplasticurbes,eachutbereceiv- ingOlcm’ dpaekedcellal-‘ollowmgremovalotthepe- wuhsupanatmtcellswaeincubatedwith4mlofthe Al-containinguptakemediquH45).onashakerat60 mafiadhrgafinaleelltiued’4xto’cellsmr'. Total elapsedtimewasappoahnatelywminfrunthebcgin- uingofpe-washsteptothebegimingofAlinatbatiut. StanthrduptakemedimncontainedOlmMCaCla. The pl-lofthecellsuspensionwasfoundtobestableduring experimentsAluptahewashahedbyadditionofOJmI ofO.1Mciune.pH45. followedbyremovaloftheAl- calmingsuperana’sunplesdcsdnedeAAS. cellswerewashedtwicemorewith4mloftmtakeme- dimnphrsSmMcitrate(pl-l4.5.2mMCaCl,J.Elapaed tinnwasZSminfromtheeodofAlwonfiuitialad- ditionofciuate)bcompleticnofwashes. A-lylselAlbygrapflschrr-esasc-Ieabam-pflco m Sampleswereoven-driedat115Cinlooselycappcd.5- mlehnsaew-capfiahthendigestedwichSOulof “‘97.!“ 60 8Mlm0,intighdycappedvialsat135C.Exces'sacid was dried off and digested samples were redissolved in lor2mlof25li-lNO,. Sampleswereanalyacdfa’ AlwithaSpectrAAItOOgraphitefinnaceatomicab— sorption spectrometer (Varian. Stmnyvale. CA. USA) wicheemanbackgrormdcorrection. Readingswerere- corded in duplicate. at 396.2 nm. Pyrolytically coated pardtionedgraphitembeswereflushedwithNggasA seven-step furnace program was applied as recom- mended by the manufacturer. Amintion temperature wassettoZSOOCwithapmdingchartemper-anueof IOOO’C. RecoveryofappliedAlfromthecellpelletsandwash supernatantsofseveralsampleswasdetermincdtobe 102:3.4%(se).simi1artofindingsonAlrecoverycm- ductedonnetu'oblastomacells(8hiandl-laug1990). However. cell pellets yielded more remoducible results and werethereforeused the ' 'tm. BackgrotmdlevelsofaboutZumolAl(10‘cells)“ftom controlcells(noaddedAl)weresubtractedfromthere- sultsofAl-treatedsamples. Mailman-y ln=paralle1toGPAAS.theAlconsentincellswasalso evaluatcdinvivowithmain.afluorescentdyewith highsensitivityforAHinviuodetectabilitylimitof about 0.04 uM). commonly used in analytical applica- donsmltyllmdPrakashlm).Morinisuaditionally employed in a methanol and/or acetic acid based solu- nonmltydandPrakashlm.‘l’tceetal.1992).1-low- ever.wehavesucceededinloadingthisdyeintointl:t cells (see Results. Fig. 3).inabuffered aqueous solution (pl-1 5.6).freecfmethanol and acetate'l'beaqueoussol- ubihtyofmainaweakpmnprndcaddislowbutin- creaseswithincreasingpl-lmenoetal. 1992).inaccord with the pit-dependent dissociation of its hydroxyl poupaTheB'hydmylpwphuapnofSDandis theonlyoneofthefivetomtdergomajorprotondissoci- ationatapl-llessthan7.0(Browneetal. 1990). Thus. the dye was insoluble and precipitated in aqueous solu- tionsun'thapllbetween4t05.butwhenbuffetedatpll 5.6 yielded a yellow. translucent solution. Stock solutions of morin (0.1 M in dimethyl sulfox- ide)werestoredinthedarkandwerepepsredfreshev- erytSdastmmediatelyafterAluptakeandciMe wash.cellswereresuspendedin4rnlof100uMmorin inuptake meditnn (pl-l 5.6.2mMCaC13) and incubated for 10 min.at60rpm. Following removal of the super- natanLcellswereresuspeodedwith3.6mlofuptakeme- diurn(pl-1'.a.n.2mMCaC1;)p1usO.2mlofO.1Methyl- enediaminetetraacetate (EDTA) (pH 5 6.) affording a fi- naltitreof4x10’ce11sml". Tominimizedyeself- quenching. these conditions were selected following ex- perimentswheremorinconcenuation (10—500 uM).in- cubau'on time (5-60 min). and cell titre (1.5-8 x 10’ cellsml“)weretestedandoptimiaed.inthepresence andabsenceofAl(20|rM.1hincubationofce11s). “M9119” CellswereobsetvedwithaWLstandardresearchmi- eroscope equipped withanepifluorescence condenser NFL (Carl Zeiss. Oberkochen, Germany). The fine set usedconsistedofabandpassfilterml’wO-MOnm) andalong-wavepassfilterMHO nm).filmutilized fuphotomiaographywaskoyalGolleOOCEssunan KodakCo..Rochester.NY.USA). Fluorescence intensities were derivcdfromsuspen- sionswithals-SlauninesceoceSpcctromcteerin- Elmer. Norwalk. CT. USA) coupled to a microcomputer and chart recorder. The excitation and emission wave- lengthsweresettoMO:5nmand515:2.5nm.re- spectively.Cellsweresuspendedinacuvetteandthe fluorescence intensity was subsequently monitored at 0.5sintervalsEachsamplereadingwastakenastheav- erageofthcftrst30sofmonitoting.llesultingfromset- dmgofcenstedttcdonsinfluorescenceintensitycould bedetecsedonlyafteratleastlmin.Whencellswereal- lowcdtosettlecompletelleuorewencereadingsofthe super-unantapproachedsero.dernonsu-atingtheabsence dmhuitscompleawithALinthemedimFluo- rescenceintensityofcellsampleswasstableforatleast 1h. Ranks minty-keno!” AltuninitunupnkeaotrMJhincubatiomwasassesaed atdifferentstagesofthecellculunecyclem. 1A).and wasfoundtobecriticallydepeodentontheageofthe culuue. Under standard conditions (pH 4.5. 0.2 mM CaCl.).Aluptakepeahcdat2daysaftersubcultmebut wasnegligibleatSand6days(Frg.1A).whichcorre- spondcdrespeetivelynoperiodsofhighandlowgrowth activity(l’tg.18).Relativegrowthrate(RGR)isanindi- rectindicatorofthepercentageofcellsengagedin growthanddivisionatagiventimemditsmaximum valuewasformdtobe1129day"at2daysaftertransfer (fig.18).lndeed.tbemitoticindexwasfoundtopeakin dtistobaccocelllineatZ-Bdaysaftersubculun'eanda similarRGRwasreportchilagataetal. 1992).Asignif- icantcorrelationbetweenAluptakemdrelativegrowth ratewasfouud(insert.fig. 1A). Therefore. activecell growth is required fa A1 uptake under conditions of short-tam expostu'etoAl. Regardinglong-termexpo- stne. Aluptake(100uMAl.pH4.0)occurredaftera10 hperiodinlogarithmic-phasecellsofawbaccocellcul- mbmwunotobsu'vableataninstationmy-phaae cells('Yamamotoeta1. 1994). Allsubscquentuptakeesperimentswereperformed withceusathayspost-uansferratherthmwithcellsat 2dayspost-transfer.becauseofthegreaterquantityof cellsavailableandthebetterreproducibilityofdata. The-coursedAluptake Aluminiumuptakecouldbeatectedjustminutesafter initialincubation(pl-l45,0.2mMCaCl,)byboththe 3 €200 ' g .3 €150 r- i .- '5 100 - =— 2 E so . E g 0 1‘ nus-u stunt were) “i: 8‘ :9 125 g i g 1m 5 e ’5 l 2' mi 5 25 i '0 and inmeditunthatcontainedSliTA Acelltitreof xlo’ cellsmr'wasusedintbpre-washstep. nd4x10’cellsmr' gcmu'clforfluo- 10mMMBS.5mMsucrcse.5mMKCl. and2.0(me-wash)cr0.2fl(Alincubatiat)mHCaC1..Valnsle themeanofatleastd forAl-morinllutaesaence-d anhcanonsquFAAS. anrerxesentsalnsutCorrelation betweenAluptake.asaessedbybotbGFAASandAl-morinfiu- aeacence.andrelativegrowthrle(RGR. Fig. lB)of¢lls.L'm- earmpession onsandthecerrespondiugcoelftcierasof determimtiar. .werey-0.13RGR+1..9 t’nOJJfor OFAAS.andy-1.25RGR+8..8.r’-O79forAl-mminfluoru- cence. B. Growthoftobaccocellculunesandrelativegrowth rate(RGR).AWeibull-typemotblwasusedsofitagrowth eurvetothecelluuedultelativegrowtbrasewucalculaed bydseWRGR-(lltineflduen Al-mtxinfluuescencemetltodandGFAAsmg.2).Up- takeratewashighestduringthefirstwminfmitialrue ca 16 umolAl [10‘ cells)" h").afterwhichitdeaaased (ca4 nmol A1 [10‘ cellsl’| h"). but saturation was not observable within the experimental period. OurvaluesonAluptakeintobaccocellsngMe- semble those observed in bean cell suspensions (Me- 4 61 nrnolAl (to'cotrar‘ is a 8 0&0 Montreal-rum unit» 5 o883 Fig.2.‘l'lrnc daysafter timand mean . 2110.0 “pli‘dgx'ath of 032 (O). 2.0 (I). and 20 mM tn rescue CaC1.(A).respecti y.Dnerminedideraieally.butintheab- senceofaddchLconu'olvalues forfluaescence'mtenslty 2 E: fr .3 E ii (2hangand'l'hylor1989.1990).‘l'heinitialrateofup- takeinorntobaccocellfiafterconvertingornrmitsmd assruningSfidryweiut)wasl.8ugAl(gdryweidrt)" min".1..anofetal.(1994)estimatedatissueconcenua- tiarof70nmolAl(gfreshweight)"insoybeanroottips aher30nrinofincubatimavaluecomparablet064 nmolA1(gfreshweight)" forutdinotn'cellsafter30 minofincubation. ”WWQCR’J-pmn-dpu- madam lnpnttoexaminefactorscthnthancellMandin puttoftlthersu'engthenthegoodcorrelaimtbetween ““911” Tab.l.Effectsoflow 'onofAltocin'ate. andpe-washpl-lonAluptakeinculnrredtobwcocellstays sf- tertransfer C.ellswerepre-washed(pl-l4.5ta'5620mMC followcdbyAlC13inarbation(20pM,l;npna;15.0.2mMCaClz) fal h.at23C.unlessind_icatedodrawise. 5 mM citric sc- id.adjustedtopl-l4.5 wi KOl-l. Allotherexpaimentalcondi' and solutions was as described (Fig. 1A). Conuol values for fluo- intens (mean equalto fluorescence units). rescence ity identicallybutinthesbsenceofALwaembtrmdfromdretotal “ ‘1,“ "Bheplications. Aluptake-lhincuban'on Al-mtrin GFAAS (nmol Al Uptake conditions (arbiuury units) [10‘ cells]") Pre-wash at pH 4.5 Standardincubation 825:3.1 9.11:1.3 4Cpre-washandincubanon 20.7:55 2.3103 Prechelati ntocitrate ll..7:4l 1.6304 Pro-wash at pH 5.6 Standard incubation 39.0 t 4.4 5.4 t 0.6 GFAAS and Al-morin fluorescence. we examined the in- fluence of Ca”. low temperature. and pre-chelation to citrate on Al uptake. Aluminium uptake was only re- duced by about 10% (Fig. 2) when the level of Ca“ in A 62 the uptake solution was increased from 0.2 to 2.0 mM Ca". However. the presence of 20 mM Ca 2' in both the pre-wash and uptake solutions drastically reduced Al up- take by 90% (Fig. 2). even after 4 h of incubation. Post- application of 20 mM Ca”. however. failed to afiect cel- lular Al levels (data not shown). It is noteworthy that Ca" reportedly can ameliorate Al toxicity (Kinraide and Parker 1987). This rs one of the first reports to examine efiects of Ca“ on Al uptake in intact plant cells. Prior to this, Alva et al. (1986) found that Ca” concenu'ations varying from 0.5 to 15 mM reduced water-extractable and acid-enn'actable (0.1 M HNO,) Al in soybean and subterranean clover roots. While the water-exnactablc Al can be expected to be associated with the cell wall, it is. however. not clear how much of the acid-extractable Al was previously associated with the cell wall or with the symplasm. Pro-chelation cfAl to 5 mM citrate also greatly re- duced Al uptake by 85% (Tab. 1). Upon lowering the temperattne of the standard pre-wash and uptake solu- tionsfrom23Cto4C,Aluptakewasdiminishedby about 75% (Tab. 1). in accord with observations on Triti- cum aestivum L. roots (Pettersson and Strid 1989. Zrang and Taylor 1990). The expression of Al toxicity has also been shown to be lass pronounced at lower temperatures (Camargo 1983). Fig. 3. Photomicrographs of flutxesoence from morin-labelled tobacco cells (3 days aha transfer)i III the absence of added Al. and autofluoresce ence fran control cells (not treated with morin). ment (atretofluoresce nansmt tted images of the same cells lsbe morin Experimental manipulation of cells was identical to that tnA ms(Fig. 1A).)brg with Al absent. A. fluorescence and B luptake geseofth samecellswithout morintreat- lit-tranrmrgh .Flutrescence' Images of cellsl labesllcgsrthggnn (0.1 mM 10 min); E. Fluorescence and F. light- ssm wt 5 M sorbitol. showin the ofmorin flo- rescence in the cell protoplast but undeteéltsble in the cell wall: The scyale 3 presence I: ebumpmeuufimmdaumiaographsareonthesame scale. A camera shutter release speed of 0.5 s was utilized in gene-sting all fluorescence' images nmmnlm Distribution of morin in cells and cellular localization of Al As judged by comparison with autofluorescence of con- trol cells (Fig. 3A. B). the distribution of morin fluores- cence in cells not exposed to Al was homogeneous throughout the cytoplasm and nucleus. and morin did not detectably accumulate in the vacuole (Fig. 3C. D) or in the cell wall of plasmolyzed cells (Figs. 3E. F). Typi- cal fluorescence intensities from morin-loaded cells in the absence of Al were 18 to 26 arbitrary units (e.g. see Fig. 1A) compared to 4—8 units from autofluorescence of control cells. Following exposure to Al and subsequent chelate wash. brightly fluorescent Al-mon'n zones could be vi- sualized at the periphery of cells (Figs. 4A, B). in addi- tion to apparently non complexed morin distributed throughout the cytoplasm. In cells not treated with Al. morin fluorescence appeared pale green. as opposed to a darker (and much more intense) green fluorescence from these regions where the dye was seemingly complexed to Al. The proportion of cells showing these discrete zones (ranging from almost none to about 70—80%) de- pended on the extent of Al uptake. When present. one or two fluorescent zones per cell were visible. Often. these fluorescent zones appeared at the region of contact be- 63 tween two adjacent cells (Fig. 4C). as it along the plane of cellular division; these fluorescent zones were gener— ally larger (about 20—30 pm) than those located in other regions (about 2—18 pm) of the cell periphery (Fig. 4A. B vs C). Visualization of the fluorescent zones became possible shortly ( 5 min) after A1 incubation. In cells 5 and 6 days after subculture. in which Al uptake was neg- lible (Fig. 1A). none or very few discrete fluorescent re- gions were perceptible and the fluorescence pattern of these cells was the same as that of cells not treated with A] (Fig. 3C, D). This discrete localization of Al does not appear to re- sult from artefacts such as unequal distribution of morin in cells (above), or from morin being unable to compete with cellular ligands. At longer AI exposure times (>1 h). some cells (less than 5% of the population) displayed Al-morin fluorescence throughout the cytoplasm. with an even brighter region presumably originating from the nucleus (Fig. 4D). demonstrating that morin can com- pete with intracellular ligands. Virtually the same pattern of Al-morin distribution was rapidly generated in almost all cells when incubated with Al in the presence of 5 mM acetate. which drastically affects membrane penneabil- ity at pH 4.5 (Jackson and St. John 1980). Interestingly. post-application of acetate, following Al incubation. did N. ‘. . ‘2. - n . ' ,./’ - Fig. 4. Photomicrpgraphs of fluorescence from morin-labelled tobacco cells (3 days afte- transfer. except D) in the presence of added AI. Aluminium uptake conditions were as described in Fig. 1A. A and B. Cells incubated with AI (20 M. 30 min) under stan- dard conditions and labelled With morin (0.1 mM. 10 min). AI-morin fluorescence is localized in adiscrete region of the cell periph- cry (arrow). Fluorescence from remainderof cell is from non complexed morin. similar to control cells labelled with morin in ab- scence of applied Al (Figs. 3C. D). C. Similar to A and B. but demonsu’ating Al-mon'n fluorescence located at the region of contact between two adjacent cells (arrow 1. D. Cells (2 days after u-ansfcr) incubated with Al for 90 min. showing localization of Al throughout cell. but parti.t:ularly in nucleus (arrow). E. Fluorescence and F. light-transmitted images of the same morin and trypan blue-labelled cells. showmg the presence of an Al-morin region in a cell (arrow in E) im cant to trypan blue (F). as opposed to one that is permeable (arrow in F). Scale bars in A and C represent 20 and 25 um. respectively. B and D are on same scale as A. E and F the same as C. A camera shutter release speed of 0.2.5 5 was utilized in generating all fluorescence images. 6 Physiol. Plant. 97, I996 nor ageueialiaeddisuibudonofAanrdidit detectablyalterthediseietepatternofAldisuibutiuiat thecell periphery (data not shown). Specialmentionismadeoftheobservatimthatthe smespatialpattemofAldisuibutionocciuredirrespee- tiveoftheorderofAlandrnorinloading(i.e.ineells firstexposedtomorinmdthentoAlorviceversa);evai whencellswerestainedwithniorininau-aditionalrneth- anal/acetic acid solution. the same Al distribution in dis- ereteperiphaalregionswasobtained(dannotshowu). However. in cells preloaded with morin. the fluores- ceneeintensitywaslessintense.pesinnahlyuiginling ficmmorinleakagebyeellsduingpe—washandineu- bationwithAl. TodistinguishwbetherthedisueteAl-inorinflues- centiegioriwasassociatedwiththesymplamcellwall. orboth.plasinolysisofcellsneuedwithAlwasat- tempted. similarlytothatperforinedinnon-Al-ueated cells (Figs 3E.P').However.theaecellsalmostalways plasmolyzed poorly. shrivelling instell. with little disso- ciadonofthesymplaunn'omthecenwanlntheiare occasions in which plasmolysis was more favorable. thesyrnplaststillrernainedasaociatedwiththeeellwall alongaseetionhubotuingmeeeAl-mainfluaeseent regions. ltisnotewonhythatthefluaeseeueeofAl-m’uiwas notquenchedbythepreaenceofMAflmMfinatleast aSO—foldexeessoverrna'in(appliedato.lrnMandex- eesswashedout).ED‘l‘Ahasabindingaflinityemstant for Al (log K = 16.5; Orvig 1993) about 10" times higherthanthatofniorinaogK1-65;Xatyaland Prakashl977)andeanbeutiliaedtoqueuehAl-inuin {lg-gm in analytical applications (Ueno et a1. 1 ). H.DI'IIC pain-hilly I IowpfludAluptaha Althoughcellsinculuueandthoseineubatedlnuptake mediumatpHS.6werevirtuallyinipei-ineableu>u'ypui blue. lowering media pl! to4.5 enhanced dye permeabil- ity.partieularlyafter4hofineubation('lhh. 2). 'l'histype of H‘-induwd membrane permeability could be amelio- medbyseveralfactors. Indeed. cellage(i..e yowth- stage). coneenn'ationofCa". orpreseueeofAlinthe medium.andternpentureallimpaetedmernbnneper- nieabiliryatpll450‘abs2and3).asjudgedbyuypan bhieuptakeintobaeeocellsAlthoughlessinform is availdile for cell ailtures. increased membrane per- meabilityofrootsreportedlyresultedfiurndeaeasedex- terrnIleYanetal. 1992).pat‘ticularlyasillustrandby enhanced efflux of K’ from cells (Zsoldos and Erdei 1981. Sasakietal. 1994). Thus.pH-uiggeredinernbraneperinediiliryniaybea factorinflueneingAluptakeHowevermellperuieahility touypanblueitselfisapparentlynotarequirenientfu Aluptake. Fineduallabellingofcellswithinorinand u'ypanbluerevealedfluoreseentAl-rnorinsouesincells thatwerenOtperuiediletouypanblueflsmBJee- “M.W.!” 1a». 2Wof¢lhadaysmmpmbleo ny- pnbluewluiinaihatedinrnediumatpl-HSa'pHSLandufl- head? Ca’enncuiuauonsandthepnaeneeoi‘zoiiMALCells (2x1 cellsinr‘)waehiaibuadinrnediaidentialineunposi- umbthatutilizadinAluptakeespawoeelegendl-‘rgJM mmmaummmm anirdrhl lOnunhnibaucupammeuednuuwasreplacedbyanequal vohneotheahnndhundaunemhiean3sehun3 septic-5on3 Mun fiCellspaindleuiuypnbhie lbw 4h“ prise 01ml)” 1830.4 2230.8 pins Minna" 19.4328 73333.0 2.0inMCa” 5.5313 29.1315 MEMO“ 1.7306 3331.4 Madly-039”“ 1230.2 8.1313 md.eellspe-washedatp85.6.ratherthanatpll4.5. renainediinperineablemuypanbluewitbin l hofAJ uptakeahb.3).butAluptakewasstillhalfofthatob- tainedunderstandarduptakeconditionst'l‘ab. l).'l'hird. thepercentageofcellsthatshoweddisereteregionswith Al-niorinfluoreaceucewas higherthanthepercentage permeableuiu'ypanhlue.'l'hatis.wefoundthatabout mwMofMay-oldeelhdemonsuatedAl-minau- mtregimwhileaboutMoftheeellswerealso permeabletouypanblueaab.3)afteral-hAlineuba- tionperiod. Lastly.thediseretecellularlocalinriouof Atasmonimiedbymoiin.peisistedregaidhssofeellu- -hrperniediilitytouypanblue. M Ou'resultssupportthenotiouthatrapidAluptakeiseiit- icaflydepadentmceflMInmwingtobaccoeells aeonsiderableunarntofAlisapparentlylocaliaed.at leastiiiitially.iriadiscieteregionofthecell.podhly comprisingpeiipheraleytosolandloraplasinarneinbrme aeeum.’ oralteinauvely' .inthedyouung‘” cellwallregion. 1&3. MdalhpumlenuypnhheamrlhAlh- mam» deellsand d nedthellswerepe—mwashed pH M bwdbymwhhawhhanmmuumfl4fi. 02ml! cmfclhltllother conditions andsohiuoi- mnmwmlALMen3sam3replicanoas. §Cellsperrnnhletotmlnblue Dania-tam +Al -Al mill-l4: 2m 3532.1 46.0345 Bay 14.9322 28333.8 Shy 23305 1.7305 mum 3m 2.830.? 15.3333 Some or possibly most of the Al within this discrete region was apparently sequestered into at least one com- partment which was not directly accessible to the exter- nal solution. i.e. suggesting that Al was indeed internal- iaed by the cell. In view of uncertainties as to whether chelate washes are capable of removing all Al from the cell wall (Zhang and Taylor. 1990). supposedly because of high affinity binding sites within the wall. evidence for Al internalization was derived from several aspects of our studies. Whereas a considerable portion of the q)- plied Al was still associated with the cell after several ci~ trate washes following Al uptake. prechelatiorr of Al to citrate reduced Al uptake by about 85% (Tab. 1). This large difference is unexpected ifAl is binding to a site in the cell wall. Similarly. CaCl, (20 mM) was ineffectual in reversing a preceding Al accumulation by the cell. but application of these levels of CaCl, during incubation with Al greatly reduced Al uptake (Fig. 2). Furthermore. even if it is assumed that the observed fluorescence orig- inaes from Al tightly bound in the cell wall. application of EDTA. while possibly not removing the Al. should still be capable of displacing morin and quenching fluo- rescence. because of its much higher binding affinity. However. fluorescence of Al-morin was not quenched by externally applied high concentrations of EDTA. ln- ternalization of Al is further suggested by the observa- tion that morin accumulated in the cytoplasm (Figs 3C- F) and that membrane permeability appears to play a role in Al uptake. Apart from the possibility that Al may be localized in a discrete region of the cell wall. which in turn com- prises a mosaic of wall domains with different composi- tion and functions (Roberts 1990). our studies indicated the absence of detectable amounts of Al in the remaining larger portion of the cell wall. These findings were in part further documented by the virtually negligible up- take of AI by older cells. In this context it is worth men- tioning that Al was not associated with pectin uronic ac- idsashasbeenshowninthreedifferentpearootsections treated with 1 mM AlCl,. pH 4.5 (Matsumoto et al. 1977). Taken together. our evidence for Al internalization supports the emerging notion that Al mainly accumu- lates in the symplasm rather than the apoplasm (Kochian 1995). In animal cells. which lack a cell wall. it was demonstrated that Al internalization is required to inhibit formation of inositol phosphates. key products of phos- phoinositide-related signal transduction (Shi et al. 1993). It is a reasonable assumption that Al-triggered charges in these types of second messengers are early indicators of Al toxicity in cells. The plasma membrane of growing cells. possibly a discrete region. therefore plays a role in Al uptake. Given the pronormced dependence of Al uptake on cell growth and its absence in stationary-phase cells. we can only speculate a to the underlying mechanisms respon- sible for the results assented here. A certain region of theplasmamembranethatpmticipatesingrowthpro— 65 cesses. such as membrane turnover and cell wall synthe- sis. might be less effective as a permeability barrier and could serve as an entry port for Al uptake. This nright explain the discrete localization of a major amount of Al at the cell periphery. In addition. the physical state of this membrane region could be modulated by the exter- nal low pH (Lindblom and Rilfors l992). which in turn might favour Al-tn'ggered membrane phase changes (Vi- erstra and Haug 1978) and facilitate Al entry. as was demonsu'ated on phosphatidylserine-containing lipid vesicles (Shi and Haug 1988). pl-l-related changes in the plasma membrane of some tobacco cells were indicated by enhanced permeability to trypan blue (Tabs 2 and 3). especially upon longer exposure to low pH. These u'y- pan blue-leaky cells were probably alive because cell death occurred at an external pH < 2 according to "P nu- clear magnetic resonance studies on suspension cultures (Fox and Ratclifie 1990). and cellular ATP levels were most likely maintained under our atlture conditions. Un- der high external Ca” concentrations (20 mM. Fig. 2). or at low temper-mire (4'C. Tab. 1). greatly diminished Al uptakemightresultfromalterationsofthediscretere- gioninacellsubjecttothistypeoftreatment. OurfindingsthatAluptakeiscriticallydependeuton cellulm'growthareinaccm'dwith wholeplantobserva- nonsManifestationsofAltoxicityinrootsoccurin meristematiccells.elongatingcellsandroothairs(Mat- sumoto et al. 1976. Ryan et al. 1993). which are indeed theactivelygiowingcellsofaroot. Tothebestofourknowledge. this isthefirstreport where morin in a solely aqueous medium has been ap- plied to cells. i.e. free of commonly used methanol and acetic acid. Since cellular integrity is better maintained upon application of aqueous solutions. a methanol/ace- tare-free procedure was preferable because Al redistribu- tion is less probable. Additionally. extracellular morin was reduced. allowing for cell-derived fluorescence measurements. Given the simplicity and resolution of light microscopy (about 0.2 pm). and the ease of speci- men preparation. the morin fluorescence method is thoughttobesuperiortothesecondaryion mass spec- u'omeu'y technique recently applied to ascertain Al dep- osition in plant cells (Lazof et al. 1994). Thus. a morin- based technique should prove advantageous in Al uptake studies. particularly when many data points are desired. or when Al incorporation by genetically different culti- vars is evalutated. AW-VA.V.wmsupputedinpartbyafellow- shiny-rt! 2013921916. from theConselho Nacional deDe~ senvolvimento Cientffrco e Tecuoldgico (CNPq). Brasil. This work was also partially funded by the nt of Botany and PM! Pamology at MSU. We are grateful to Drs R. Hausinger. M. Zahik. A. lures. and 1.. Bauer for use of the GFAAS. thior'raneter. cpl-fluorescent miaoscope. and water shaker. respectively. We also thank R. Pacovsky for useful dis- cussions and D. Shiatsu and M. Sullivan fr: sumlyiug us with use cell culture. met-391.1006 66 References Alva. A. K.. Asher. C. J. 8: Edwards. D. G. 1986. The role of calcium in alleviating aluminium mxicity. - Aust. J. Apic. Res. 37:375-382. Bradford. M. 1976. A rapid and sensitive method in the quarti- tation of microgram quantities of protein utilizing the princi- ples of protein-dye binding. - Anal. Biochem. 72: 248-254. Browne. B. A., McColl. J. G. 8: Driscoll. C. T. 1990. Alumino ium speciation using morin: l. Morin and its complexes with aluminium. - J. Environ Qual. 19: 65-72. Carnargo. C. E. de O.1983.E.feito da unmet-sun da soluclo nutritiva na tolerlncia so aluminio dc cultivates dc trigo. — Bragantia 42: 51-63. Delhaize. E. 8:. Ryan. P. R. 1995. Alumirnim toxicity lid tolero ancc in plains. - Plant Physiol. 107 : 315—321. Fox. G. G. at Rstcliffe. R. G. 1990. ”P-NMR obsu'vation on the effects of the external pH on use inuacellular pH values in plant cell suspension culnires. - Plain Physiol. 93: 512-321. Foy. C. D.. Chaney. R. l... a White. M. C. 1978. The physiology of metal toxicity in plants. - Annu. Rev. Plant Physiol. 29: 511-566. Ganrot. P. O. 1986. Metabolism and possible health effects of alumiiaim. - Envirar. Health Perspect. 65: 363-441. . Haug. A. 1984. Molecular aspects of alum'utum toxicity. - Crit. Rev. Plant Sci. 1: 345-373. Jackson. P. C. at St. John. J. B. 1980. Charges in memu'ane lip- idsofrootsassociatcd withchanesirrpermesbil'uy.-Plant Physiol. 66: 801-804. Katyal. M. at Prakash. S. 1977. Analytical raciicnsofhymy- .. flavones. - Talarsa 24: 367-375. Kinraide, T. B. 1991.1dcntity ofthe thiamnxic alt-aim spe- cies. - Plant Soil 134: 167-178. - &Parker.D.R.1987.Cationainelicatirsrdahmiimimeox- icity in wheat - Pill! Physiol. 83: 546-551. Kochian. L. V. 1995. Cellular meclnnisms of aluminum msicity and resistance in plants. - Annu. Rev. Plait Physiol. Plant M01. Biol. 46: 237-60. Lazof. D. B.. Goldsmith. J. G.. Rufty. T. W. & Liam R. 1994. Rapid upmke of shimiraim into cells of imact soybem root tips. A micromalytical snidy using in mass spectrometry. - Plant Physiol. 106: 1107-1114. Lindblom. G. at Rilfors. 1... 1992 Cubic lid hexagonal lipid phases -lnStructureandDynamic PropertieeofLipidsand Membranes (P. J. Quinn and R. J. Cherry. eds). FP- 51-76. Portland Press. Lonbn. ISBN 1-855-78u014-3. Matsurnoto. l-l.. Hirasawa. E.. Torikai. H. dt Takahashi. E. 1976. Localization ofabscrbed aluminiuminpearootanditsln'nd- ing to nucleic mids. - Plant Cell Physiol. 17: 127-137. - . Mortmura S. at Takahashi. E. 1977. Less involvernern of pectin in the precipitation of aluminium in pea root. - Plant Cell Physiol. 18: 325-335. McDonald-Stephens. J. 1... Jr Taylor. G. J. 1995. Kinetics of alu- minum uptake by cell suspensions of Phased» vulgaris l... — J. Plant Physiol. 145: 327—334. Nagatlal. gheNemoto. II. at Hasmws. S. 1992 ‘lbbaeco BY-2 cc ' asthe“e1.a"cellintleccllbrology' ofhi plats-1m. Rev. Cytol. 132: 1-30. 8"" Why?! “The”. 1996 Orvig, C. 1993. The aqueous coadination chemistry of alumi- num. - In Coordination Chemistry of Alumiiaim (G. 11. Robinson. ed). pp. 85-121. VCl-l Publishers. New York NY. lsBN 1-56081-059-9. Peuersson. S. a Su-id. 1-1. 1989. Initial uptake of aluminum in re- lxion to temperature and phosphorus status of wheat (Triu- eiim antiwar 1..) roots. - J. Plant Physiol. 134: 672-677. Robats. K. 1990. Structures at the plant surface. - Curr. Opirt Cell Biol. 2: 920-928. Ryan. P. R.. Ditomaso. J. M. a Kochian. L. V. 1993. Aluminium toxicity in roots: an investigation of spatial sensitivity and the role of the root cap. - J. Exp. Bot 44:437-446. Sasaki. M., Kasai. M.. Yamamoto, Y. 6r Matsumoto, H. 1994. 'stsr of the ecly response to aluminum sums be- tween tolerant and sensitive wheat cultivars: root growth. alumirmm cement and efl'lux of K‘. - J. Pill! Nun. 17: 1275-1288. Shi.B.&Heug. A. 1988.Uptakeofaluminumby lipidvesicles. -Tox. Environ. Chan. 17: 337-349. -&1-lnig.A. 1990.A1uminumuptahbyneuroblastoun .-J. Neurochem. 55: 551-558. - Chou. K & Haug. A. 1993. Aluminium impacts elements of the phosphoinositide sipslling pathway in neuroblas- tans cells - Mol. Cell. Biochem. 121:109-118. Tice. K. R.. Parka. D. R. a DeMason. D. A. 1992. Opermion- ally defined apoplastic and symplastic aluminum fractions in root tips of aluminum-intoxicmed wheat. - Plant Physiol. 1(1): 309-318. Ueuo. L1mamura.T.&ChengK.L. 1992.1-llrdbookof0r- ganicAnalyticalReagents.-CRCPiess.BocaRatou.FL.pp. 2.5-29. 158. ISBN 0-8493-4287-2 Vita-sue R. a Bang. A. 1978. The effect ofAl"on the physical matte} ofmanbranelip'dsinmmplom acrdoplrr' ' .-Biochem. Bligpbys. Res. Comm. 84: 138-143. Wagmuma. T. 1983. act of non-metabolic conditions on do uptakcofaluminumbyplantroots.-Soil Sci. Plant Nutr. 29: 153-333. Yamamom. Y.. Rikiishi. 8.. Grang. Y.. Ono. K.. Kasai. M. a Matsuinom. 1-1. 1994. Qumtitative estimation of aluminium toxicity in culuired mbacco cells: correlaion between alu- minium upmke mid growth inhibition. - Plant Cell Physiol. 35: $75—$83. VII. F.. Schubert. S. & Mengel. K. 1992. Effect of low root me- d'annplionnctprotonreleasc.roarespiration.androot ofcorniZeasinysLMndbroadbethciafabo ). - Plain Physiol. 9% 415-421. Zhang. G. & Taylor. G. J. 1989. Kinetics of aluminum uptake by excised roots of aluminum tolaant and aluminrnn sensitive cgigvms of Tritieisn oern‘vum 1.... - Plant Physiol. 91: 1094- - R Taylor. G. J. 19%. Kinetics of aluminum uptake in Thri- cisn oestr'vum 1.... ldemity of the linear phase of aluminum spake by excised roots of aluminum-tolerant and alumi- niirnoaensitive cultivars. - Plant Physiol. 94: $77-$84. Zsoldos, F. & Erdei. L. 1981. Membrane and ion transport prop- ertiesincereals underacidicmrdalkalinestrees.1.£ffectsd onpotassitmtuptakeandgrowthofriceandwhea.- $291.35:sz E 70. CHAPTER 4 CESSATION OF ALUMINUM UPTAKE AND ALUMINUM-INDUCED ABNORMALITIES OF THE CELL SURFACE DEPENDS ON BREFELDIN A- SENSITIVE MEMBRANE TRAFFIC IN TOBACCO BY-Z CELLS ABSTRACT Continuing previous work [Vitorello and Haug, 1996], the relationship between cell growth and Al uptake in tobacco BY-Z cells was further examined. As observed by confocal laser scanning microscopy of morin-labelled cells, new aspects of Al-accumulating regions emerged. Al-morin fluorescence was frequently associated with a protuberance of the cell surface and with vesicle-like or globular structures attached to the exterior surface of the cell and presumably produced in response to Al treatment. Similar structures were also frequently observed along the plane of contact between adjacent cells. These events occurred only in growing cells, consistent with the hypothesis that Al accumulates in regions of the cell where new cell wall and plasma membrane are being incorporated. Treatment of cells with cremart (IO-30 uM), brefeldin A (10 uM), caffeine (1-5 mM) and 2,6-dichlorobenzonitrile (20 uM), inhibitors of microtubule polymerization, Golgi-mediated secretion, cell plate formation and cellulose synthesis, respectively, had no discernible effect on Al uptake. Brefeldin A also had no effect on Al-induced release of material by the growing cells. However, cells almost completely lost there capacity for Al uptake when incubated in a minimal medium, at pH 5.6, for 1 to 3 h prior to exposure to Al. Incubation of cells in minimal medium reduced subsequent growth in culture medium. Brefeldin A (10 11M) suppressed loss of Al uptake capacity, upon incubation in minimal medium, in a dose-dependent and reversible manner. Once cells lost their capacity for Al uptake, brefeldin A did not restore it. 67 INTRODUCTION An understanding of Al uptake processes is central to the elucidation of mechanisms of Al toxicity. Al toxicity was shown to be correlated with long-term uptake in tobacco cell cultures (Yamamoto et al., 1994). In mammalian cells, Al internalization was required for inhibition of phosphoinositide-mediated signal transduction (Shi et al., 1993). Short-term aluminum uptake has recently been shown to be dependent on cell growth in Nicotiana tabacum L. cv BY-2 cells (V itorello and Haug, 1996), consistent with the fact that Al is detrimental to and predominantly accumulates in growing cells of the root apex (Ryan et al., 1993; Rincon and Gonzales, 1992). Al accumulated predominantly in a discrete region at the cell periphery, ofien between adjacent cells, as if along the plane of cell division. It is hypothesized that these regions might coincide with regions of growth in cell surface and volume (V itorello and Haug, 1996). Access of Al to target sites may be facilitated in such regions. Recently, it was shown that Al is detrimental (within 30 min) to tip growth in root hairs, demonstrating that Al can rapidly inhibit cell elongation in non-dividing cells (Jones et al., 1995). This provided additional evidence that inhibition of cell elongation, rather than cell division, is the primary cause of rapid, Al-induced root growth inhibition (Kochian, 1995). The tips of these root hairs were swollen and deformed and afier prolonged exposure to Al, sometimes burst (Jones et al., 1995). Moreover, A1 (0.3 mM) has been found to cause rapid changes in the ultrastructure of the Golgi apparatus and to reduce secretory activity of 68 69 peripheral root cap cells (Bennet and Breen, 1991). In addition, Al reportedly inhibited utilization of sucrose for cell wall synthesis (Huck, 1972) and induced callose production (Schaeffer and Walton, 1990; Schreiner et al., 1994). Taken together, circumstantial evidence supports the notion that Al may accumulate in and disturb regions of the cell that are involved in processes of cell growth and/or secretion, although it is not clear how this occurs. In this chapter, the relationship between cell growth and Al uptake and accumulation is further examined. By using confocal laser scanning microscopy and low morin concentrations for the detection of Al, improved images are generated showing Al-induced abnormalities of the cell surface of growing cells. Subsequent studies, using inhibitors of processes involved in cell growth, led to the discovery that a functional, brefeldin A-sensitive secretory pathway is required for cells to lose their capacity for Al uptake, upon prior incubation in a minimal medium and concomitant reduction in growth. MATERIALS AND METHODS Chemicals and labware Brefeldin A, morin and Murashige and Skoog basal salts were from Sigma (St. Louis, MO) and caffeine was from Aldrich (Milwaukee, WI). Cremait (butamiphos) was a gifi from Dr. K. Saito of Sumitomo Chemical Co. (Hyogo, Japan), and 2,6-dichlorobenzonitn'le (DCB) was a reference standard from the Environmental Protection Agency (Research Triangle Park, NC). All other chemical reagents were of analytical grade. Acid-washed glassware or, preferably, plastic labware was employed. The water used in all experiments was sequentially purified by reverse osmosis, deionization and distillation. Cell cultures and Al uptake procedure Cells of tobacco (Nicotiana tabacum L.) cv BY-2 were grown as described previously (V itorello and Haug, 1996). Growing cells, at 2 or 3 days after subculture, were used in all experiments. Procedures used for Al uptake were also as described previously (V itorello and Haug, 1996) with the exception of Al uptake prior to plasmolysis experiments. Al uptake was performed in a minimal medium composed of 10 mM Mes, 5 mM sucrose, 5 mM KCl, 2 mM CaClz, pH 4.5 (or pH 5.6). In some experiments, cells were plasmolyzed with 0.5 M sorbitol, following Al treatment, to allow separation of the plasma membrane from the cell wall and thus distinguish whether Al-morin fluorescence was in the cell wall or in the protoplast. However, 70 71 following incubation with A1 at low pH, cells plasmolyzed poorly, tending to shrivel, with incomplete separation of the plasma membrane from the cell wall. Attempting to improve this, cells were also briefly pre-washed at pH 5.6 and incubated with Al for shorter periods of time (<30 min), also in minimal medium at pH 5.6. Treatment of cells with inhibitors of growth-associated processes Cremart, brefeldin A, caffeine and 2,6-dichlorobenzonitn'le are inhibitors of microtubule polymerization (Doonan et al., 1988), the Golgi secretory pathway (Driouich et al., 1993), cell plate formation (Hepler and Bonsignore, 1990), and cellulose synthesis (Shedletzky et al., 1992), respectively. Stock solutions of these inhibitors of growth- associated processes were prepared as follows. A 10 mM stock solution of brefeldin A (BF A; MW= 280.4) was prepared in methanol (Yasuhara et al., 1995) and stored refrigerated with desiccant. A stock solution of cremart [O-ethyl-O-(3-methyl-6-nitrophenyl) N-sec- butylphosphorothioarnidate; MW = 332] was prepared by adding 10 til of cremart per ml of DMSO (ca. 30 mM). Caffeine (MW=194.2) was prepared as an aqueous 0.1 M stock solution, and 2,6-dichlorobenzonitrile (DCB; MW=172) was prepared as a 20 mM stock solution in DMSO. To test the effects of these inhibitors on Al uptake, these compounds were individually added to 10 ml of cell culture which had been transferred to 50 ml Erlenmeyer flasks, and were incubated for 30-60 min on a gyratory shaker 160-180 rpm, at 28°C. Following this treatment, Al uptake (1 h, pH 4.5) was performed as described above. The final concentrations of these compounds used were 15 and 30 1.1M cremart, 1 and 5 mM caffeine, 10 iiM BFA and 20 iiM DCB. In the experiments regarding the effects of BFA on the cessation of Al uptake, cells were pre-incubated with BF A prior to other experimental treatments by transferring a volume of suspension culture (generally 10 to 20 ml) to 50 ml Brlemneyer flasks, followed by incubation with BFA on a gyratory shaker at 160-180 rpm, 28 °C. Concentrations of BFA employed ranged from 1 to 50 uM, but was generally 10 uM. Unless otherwise stated, pre- incubation with BFA was for 30 min. Control cells were treated identically, with an equivalent amount of methanol, but no BF A. Labelling of cells with morin and measurements of fluorescence intensity For the detection of Al, cells were labelled with the Al-binding fluorescent dye morin (100 uM, 10 min) and measurements of fluorescence intensity were made with a LS-S luminescence spectrometer (Perkin-Elmer, Norwalk, CT) coupled to a microcomputer, as described previously (V itorello and Haug, 1996). However, for observations under the microscope, cells were treated with lower concentrations of morin (5 or 10 uM) to reduce glare and saturation of Al-fluorescent regions, thus improving the quality of the images. 73 Microscopy Aluminum and morin-treated cells were observed with a Zeiss 210 laser scanning confocal microscope (Carl Zeiss, Inc., Thomwood, NY). Images were collected in laser scanning transmitted brightfield and confocal fluorescence modes using the blue 488 nm line from a 10 mW dual-line argon ion laser. For confocal fluorescence, a long pass 520 nm barrier filter was used. A 40 x Plan-Neofluor oil immersion (numerical aperture of 1.3) objective was used. For black and white photography, Kodak Tmax 100 film (Eastman Kodak Co., Rochester, NY, USA) was used on a Matrix Multicolor computerized camera unit (Agfa-Matrix, Orangeburg, NY), directly coupled to the Zeiss microscope. Conventional, epifluorescence microscopy was also employed to examine cells. A band pass 400-440 nm and a long pass 470 nm filter set were used with a Zeiss WL standard research microscope equipped with an epifluorescence condenser IV FL (Carl Zeiss, Oberkochen, Germany). RESULTS Confocal laser scanning microscopy of Al-accumulating regions in cells Improved images of Al accumulating regions were obtained (Figs. 1-3), compared with our previous report (Vitorello and Haug, 1996), by using confocal microscopy and lower concentrations of morin for detection of Al. This resulted in reduced glare and saturation of fluorescence. As reported previously, Al accumulated in a discrete region at the cell periphery, frequently along the boundary between two cells, as if along the plane of cell division. A bulge or protuberance of the cell surface was often observed at sites of Al-morin fluorescence (Fig. 1A). Adjacent to sites of Al accumulation, but exterior to the cell, Al- morin fluorescence was also frequently observed in what appeared to be large vesicle-like or globular structures (Fig. 1B, C). These structures were attached to the cell, as evidenced when cells rotated and moved freely in a newly placed drop of cell suspension on a microscope slide. Al-morin fluorescence was not uniform along the plane of contact between adjacent cells, where Al frequently accumulated, but rather was associated with distinct structures (Fig. 2), somewhat similar to the vesicle and globular-like structures described above (Fig. 1). Another new observation was that Al-morin fluorescence occurred at the margins of the plane of contact between cells, as judged by focussing the confocal images up and down the specimen. These events could be observed after short-term exposure of the cells to Al. However, these structures were more pronounced following longer periods of exposure to Al (60 min) than shorter periods (15 min). These structures also became more pronounced with time, 74 75 Figure l. Photomicrographs of Al accumulation in deformities of the cell surface which originated afier A] treatment. Cells (2 days after subculture) were pre-washed (pH 4.5) then incubated with 10 or 20 uM AlCl3 (pH 4.5) for 15 or 30 min. Uptake was halted by addition of 5 mM citrate. For detection of Al, cells were incubated with 5 pM morin for 10 min (pH 5.6), then resuspended in fresh medium (pH 5.6) containing 5 mM EDTA. Images were collected with a confocal laser scanning microscope. Images of each specimen (A-C) were collected in both transmitted laser scanning brightfield and confocal fluorescence modes. The left half side of each image is a transmitted laser scanning brightfield image showing a general view of the cell. The right half side is a confocal fluorescence image of the same cell showing only Al-morin fluorescence. A) Protuberance of the cell surface (arrow in left), containing A1 (image on right), which formed in response to Al treatment. The fluorescence from the remainder of the cell is weak and does not show up under the conditions used for observation. B) A globular structure attached to the cell surface (arrow in B), also containing A1 (image on right). C) F orrnation of a vesicle-like structure (arrow in C) at the cell surface and associated Al-morin fluorescence (image on right). 76 Figure 1. 77 Figure 2. Photomicrographs of Al accumulation in structures at the region of contact between two adjacent cells. Cells (2 days after subculture) were treated with Al and morin as described in Fig. 1. From top to bottom, images are confocal fluorescence, transmitted laser scanning brightfield and an overlay of both modes. Images were collected with a confocal laser scanning microscope. A, B and C) A view of Al accumulation between two cells. D, E, and F) A higher magnification View of Al accumulation between two cells, showing more detailed structure. 78 Figure 2. 79 Figure 3. Photomicrographs of plasmolyzed, Al-treated cells. Cells were treated with Al (10 uM, 15 min) as described in Fig. 1. Afier labelling with morin, cells were plasmolyzed with 0.25 M sorbitol, followed 5-10 min later with a final concentration of 0.5 M sorbitol. Images are overlays of confocal fluorescence images over transmitted laser scanning brightfield images. A and C) Images illustrating tendency for the cell wall and plasma membrane to remain attached in the region containing A]. B) Plasmolyzed cell with Al-morin fluorescent region associated with the protoplast rather than with the cell wall. However, separation of the protoplast and cell wall was rarely achieved in the region containing Al. 80 Figure 3. 81 after treatment with Al had terminated. None of these events were observed to occur in growing control cells treated identically but without Al, or in non-growing cells (stationary- phase cells) treated with Al. These observations are similar to Al-induced deformities in root hair tips observed by Jones et al. (1995), although longer periods of exposure to Al were required in root hairs. After 2 to 3 hours of exposure to Al (20 uM), the root hair tips of a pondweed, Limnobium stoloniferum, were deformed and swollen and, after prolonged exposure, were sometimes observed to burst. Al-treated cells were plasmolyzed to visually distinguish Al possibly in the cell wall and in the protoplast. This distinction has been difficult, because of shrivelling and incomplete separation of the plasma membrane from the cell wall. Modest improvements were obtained when Al exposure was performed in a medium of pH 5.6 and was decreased to 15 or 30 min. From repeated plasmolysis experiments, at pH 4.5 or 5.6, it became apparent that there was a tendency for the protoplast to remain associated with the cell wall at these regions of Al-morin fluorescence (Fig. 3A, B). Occasionally, however, it was possible to observe apparent dissociation of the plasma membrane from the cell wall, with Al-morin fluorescence associated with the protoplast (Fig . BB). Plasmolysis of cells prior to Al treatment could be accomplished successfully. However, Al uptake was reduced by 76 to 84% when cells were treated with 0.1 or 0.3 M sorbitol prior to and during exposure to Al, as determined by both Al-morin fluorescence and graphite furnace atomic absorption spectrometry. Al-accumulating regions could not be observed in cells plasmolyzed prior to Al treatment. 82 Cells lose their capacity for A1 uptake when incubated in a minimal medium Al uptake in the tobacco cells is dependent on growth and is highest in cells 1 to 3 days after subculture (V itorello and Haug, 1996). When 2-or 3-day cells were removed from their culture medium, and first incubated in a minimal medium at pH 5.6 prior to exposure to Al, subsequent Al uptake (1 h, pH 4.5) decreased significantly (Fig. 4). After 30 min of incubation, Al uptake was reduced by 30-50%, and most of the Al uptake capacity of these cells was abolished after 1 or 2 h. Al-morin fluorescent regions could not be observed, under the microscope, in cells which had undergone this loss of Al uptake capacity and, in this respect, resembled stationary-phase cells. Incubation of cells in minimal medium for a defined period of time (1 or 3 h), followed by fresh culture medium, significantly reduced cell growth when compared to uninterrupted culture or to direct transfer into fresh culture medium (Tab. 1). Thus, the cessation of Al uptake, induced by incubation in minimal medium, was presumably caused by reduced or halted cell growth. Effect of inhibitors of growth-associated processes on Al uptake Cells were incubated for one hour, prior to treatment with A1, with cremart (15 and 30 pM), BFA (10 uM), 2,6—dichlorobenzonitrile (ZOuM) and caffeine (1 and 5 mM), which are inhibitors of microtubule polymerization, Golgi-mediated secretion, cell wall synthesis and cell plate formation, respectively (Doonan et al., 1988; Driouich et al., 1993; Hepler and Bonsignore, 1990; Shedletzky et al., 1992). These processes are intimately involved in cell 83 Figure 4. Aluminum uptake ceases in tobacco BY-2 cells after incubation in a minimal medium, but not in cells treated with brefeldin A. Cells at 2 and 3 days after subculture (A and B, respectively) were used. Prior to any other treatment, cells were first pre-incubated with O or 10 uM brefeldin A for 30 min, then removed from their culture medium and incubated (4 x 105 cells/ml) in a minimal medium (pH 5.6) also containing 0 or 10 uM brefeldin A, for 0.5 to 3 h. Cells (2 or 3 days after subculture) were pre-washed (pH 4.5) then incubated with 20 uM AlCl3 (pH 4.5) for l h. Uptake was halted by addition of 5 mM citrate. For detection of Al, cells were incubated with 100 iiM morin for 10 min (pH 5.6), then resuspended in fi'esh medium (pH 5.6) containing 5 mM EDTA. Control cells were manipulated identically but were maintained in culture medium prior to Al uptake. Bars represent SE (n=3). AI uptake (1 hr) fluorescence (arbitrary units) Al uptake (1 hr) fluorescence (arbitrary units) Figure 4. .3 .3 N uh O _a. O O l O 84 ' J r A 2 day cells .. ________ —o culture medium (control) j minimal medium 6° ‘ (+ 10 “M BFA) 40 -* 20 a minimal medium (- BFA) 0 B 3 day cells minimal medium 30 ‘ (+ 10 "M BFA:E O) .5 O l N O I O culture medium —. (control) minimal medium (- BFA) fi. 1 I I 0 1 2 3 duration of incubation (hours) 85 Table 1. Growth of tobacco BY-2 cells following transfer to fresh culture medium, incubation in minimal medium, or culture in the presence of brefeldin A (BF A). Growth was measured 24 h after the beginning of each treatment, as the increment in cell titer (106 cells/ml) of the suspension. Cells were utilized 2 days after subculture. a) Cells were left undisturbed in the culture medium; b) Cells were transferred to a fresh culture medium; c) and (1) Cells were incubated for 1 and 3 h, respectively, in a minimal medium, followed by incubation in fresh culture medium; e) and f) Cells in culture medium were treated with 10 uM brefeldin A (BF A) continuously, or for 1 h, followed by incubation in fresh culture medium, respectively. Growth of cells - increment in cell titer Treatment (10° cells/ml) a) continuous culture 1.05 i 0.05 b) fresh medium 0.69 i 0.04 c) l h minimal medium 0.40 i 0.06 d) 3 h minimal medium 0.33 i 0.04 e) 10 uM BF A - continuous culture 0.03 d: 0.03 f) 1 h 10 uM BFA ' 0.56 :t 0.04 86 growth. However, treatment of growing cells in culture medium with these inhibitors prior to exposure to Al had no significant effect on Al uptake (Tab. 2). In the specific case of BFA, an inhibitor of the Golgi apparatus and associated secretory pathway (Klausner et al., 1992; Driouich et al., 1993), the effects of this compound on Al-induced production of extracellular material, as described above (Fig. 1 and 2), was also tested. As judged by observations with the microscope, treatment of cells with 20 uM BF A bad no effect on Al-induced formation of extracellular material. Cells were treated with BF A one hour before application of Al, but BFA was also maintained during Al and morin incubations, as a precaution, because the effects of BF A are known to be highly reversible (Driouich et al., 1993). Brefeldin A sustains Al uptake capacity by cells Although application of BF A to continually growing cells did not cause any significant changes in Al uptake, it suppressed the loss of cell Al uptake capacity induced by incubation in minimal medium (Fig. 4). This suppression was complete in 3 day cells (Fig. 4B), but not in 2 day cells (Fig. 4A). Simply co-incubating cells with BFA and minimal medium was sufficient to inhibit most of the reduction in Al uptake capacity, but the effects of BFA were more pronounced if growing cells, in culture medium, were pre-incubated with BF A for 15-30 min prior to exposure to minimal medium (Fig. 5). In preliminary experiments, cremart (30 11M) or caffeine (5 mM) did not suppress this dr0p in Al uptake capacity (data not shown). Tests were performed to verify whether these were pleiotropic effects of BFA. A 87 Table 2. The effect of cremart, dichlorobenzonitrile GJCB), caffeine and brefeldin A (BF A) on Al uptake in tobacco BY-2 cells. Growing cells (2 and 3 days) in culture medium were incubated with these inhibitors for 1 h prior to Al uptake, with the exception of BF A which was applied 0.5 hr before. Al uptake (1 h) was performed as described in Fig. 4. Al uptake was determined by Al-morin fluorescence. Results were pooled (n=3), and are expressed as percentage of Al uptake in control cells not treated with these inhibitors. Al uptake treatment percentage of control (%) (mean :1: SE) Cremart (15 and 30 uM) 100.2 i 11.3 DCB (20 uM) 98.2 :t 2.9 Caffeine (1 and 5 mM) 118.4 :t 13.8 BFA(10 11M) 104.8 s 5.41 88 120 Al uptake _ Zcontrol 100 - A E ZOpMBFA S a 1:: 3‘ 8° ‘ i g SpMBFA 27% 83 60 g o < 3 4° ‘ \ AI uptake '5 minimal medium a (no BFA) 20 — 0 l l I l l I O 5 10 15 20 25 30 35 time (min), pre-incubation BFA Figure 5. The effect of the duration of pre-incubation with BF A, prior to incubation in minimal medium, on blockage of the cessation of Al uptake in tobacco BY-2 cells, by BFA. Cells 2 days after subculture were used. Prior to any other treatment, cells were pre-incubated with 5 and 20 uM BFA for 0 to 30 min, then incubated in minimal medium + BFA for 1 h. Al uptake (1 h) and morin treatment were performed as described in Fig. 4. The lower and upper dashed lines represent control cells not treated with BF A and control cells not treated with both BFA and minimal medium, respectively. 89 dose-dependence curve, for BFA inhibition of Al uptake reduction, was performed (Fig. 6). Effects were noticeable at concentrations as low as 1 uM. Although complete saturation of the response was not observed, 5 to 10 uM BFA produced a response close to maximum. Further increments, with use of up to 50 uM of BFA, were small (Fig. 6). These concentrations and duration of treatment with BFA are low, consistent with a primary and specific effect upon the Golgi apparatus. In plant cells, concentrations of BFA which have been employed range from about 0.5 to 700 11M (Satiat-Jeunemaitre et al., 1996). In mammalian systems, concentrations vary from 0.2 to 20 11M depending on whether intact cells or a cell-free system is employed (Robinson, 1993). A hallmark of BFA effects upon the Golgi apparatus is its rapid reversibility (Driouich et al., 1993). Once BF A was removed from the minimal medium, the capability of cells for Al uptake dropped similarly to those cells treated only with minimal medium (Fig. 7 A). Finally, addition of BFA to cells that had already experienced some loss in their A] uptake capacity did not result in recovery of Al uptake, rather it had the effect of maintaining the same levels of Al uptake and avoiding further drops in Al uptake capacity (Fig. 7 B). 90 120 _ AI uptake g 100— C 3 ‘i‘t‘ .1: _ 3g 80 CE $3 60 ‘58 38 <01 2 40 O .. 3 C 20— 0 l l l l T o 10 20 30 40 so 60 BFA concentration (11M) Figure 6. Dose dependence curve for the suppression, by brefeldin A, of the cessation of Al uptake by tobacco BY-2 cells, induced by incubation in minimal medium. Cells after 2 days of subculture were used. Cells were incubated in minimal medium for 1 h, followed by incubation with Al (1 h), and treatment with morin, as described in Fig. 4. The dashed line represents Al uptake in control cells neither incubated in minimal medium nor with BFA prior to uptake. Bars represent SE (n=3). 91 Figure 7. Reversibility of the effects of brefeldin A on loss of Al uptake capacity (A) and its effect after some loss of A1 uptake capacity has already occurred (B) in tobacco BY-2 cells. Prior to any other treatment, cells were pre-incubated with O or 10 uM BFA for 30 min, then incubated in minimal medium (i BFA) for various times. Cells were treated with Al for 1 h and with morin as described in Fig. 4. Cells 3 days after subculture were used. A) After 1 h incubation in minimal medium and BF A, some cells were reincubated in minimal medium free of BFA (dashed line). B) BFA (10 uM) was added, at two different times, to cells incubated in minimal medium, after some loss of Al uptake capacity had occurred.Bars represent SE (n=3). AI uptake (1 hr) Al uptake (1 hr) fluorescence (arbitrary units) 3 l 100 92 A / .75 _ 'E 80 — 3 . . . Z‘ ' . minimal medium E 41 (+10 11M BFA) s: o _ '9 6 release g - 3 from BFA -. o A . 0 ~, minimal 5 40 A medium 9 in E g 20 A a: minimal medium (-BFA) O B 50 1. addition of 10 11M BFA addition of 10 11M BFA minimal medium (- BFA) 0 Figure 7. l H I l 1 2 3 4 duration of incubation (hours) DISCUSSION Three major new findings emerge from this work. First, treatment of growing cells with Al promotes deformities in the cell surface, presumably in regions where new cell wall and plasma membrane are being added. Second, upon reduction of cell growth following incubation in minimal medium, Al uptake is also reduced. Third, Al uptake is sustained when cells are treated with BF A, despite incubation in minimal medium and reductions in growth. It is suggested that alterations in the plasma membrane or cell wall occur concomitant with cessation of growth, and result in reduced Al uptake. Cells treated with BF A are incapable of these alterations and therefore continue to take up Al. Based on observations with the microscope and using Al-morin fluorescence, it is clear that Al is associated with deformities (protuberances, vesicle-like and globular structures) of the cell surface (Fig. 1). It is concluded that Al induced these abnormalities of the cell surface because: 1) These structures were not observable in control cells treated identically but without Al; 2) These structures were more pronounced after longer periods of incubation with Al; 3) Presumably because of these abnormal structures, there was a tendency for the cell wall and protoplast to remain attached at these sites following plasmolysis. This was observable only in Al-treated cells. With respect to the nature of these regions, it is important to try to distinguish whether these events are occurring because of interactions of A1 with the plasma membrane, with the cell wall, or both. In at least some cells plasmolyzed after Al treatment, Al was apparently completely associated with the protoplast (Fig. 3B). Evidence favoring the notion that Al uptake involves passage across the plasma membrane was provided previously 93 94 (Vitorello and Haug, 1996 - chapter 3). Involvement of the plasma membrane is again suggested by the fact that Al uptake was almost completely eliminated in growing cells treated with sorbitol prior to exposure to Al. These regions of Al accumulation possibly coincide with regions of cell growth, where new cell wall and plasma membrane are being added (V itorello and Haug, 1996). This is suggested because of the dependence of Al uptake on cell growth and because of the patterns of Al accumulation in the cell. The fluorescence observed along the region of contact between adjacent cells (Fig. 2) was suggestive of recent observations made on cytokinesis in this same BY-2 cell line, which were in the process of fusing their cell plate with the parent plasma membrane and cell wall (Samuels et al., 1995). Fusion occurs at the edges of the plane separating the daughter cells, which seems to colocalize with Al-morin fluorescence. Furthermore, fusion occurs through finger-like projections at the margins of the cell plate. If Al accumulates at these sites of firsion, this would explain the presence of several discrete fluorescing regions observed along the plane of separation (Fig. 2). Alternatively, this pattern of Al accumulation may coincide with maturation of the cell plate, which involves closing of plate fenestrae (also towards the margins) and cellulose synthesis (Samuels et al., 1995). The hypothesis that aluminum accumulates in regions of cell wall and plasma membrane synthesis is consistent with the fact that Al toxicity is manifested in growing cells of the root apex (Ryan et al., 1993; Kochian, 1995), that it inhibits tip growth in root hairs and pollen tubes (Konishi and Miyamoto, 1983; Searcy and Mulcahy, 1990; Jones et al., 1995), that it disrupts secretory activity and causes derangements of the Golgi in root cap cells (Bennet et al., 1991), that incorporation of sucrose into cell walls is impeded (Huck, 95 1972) and that callose synthesis is induced (Schaeffer and Walton, 1990). Most compelling are recent observations that Al induces deformities only in the tips of growing root hairs, following a few hours of exposure to Al (Jones et al., 1995). It seems reasonable that the observed abnormalities occurred because of derangements caused by Al in the incorporation of cell wall and plasma membrane. Such derangements can explain the tendency for the cell wall and plasma membrane to remain attached in Al-treated, plasmolyzed cells. It may also explain some previously reported observations in the literature. In wheat roots, Al uptake was interpreted to depend on an intact protoplast and to involve metabolism dependent uptake into the cell wall (Zhang and Taylor, 1990). In one of the few previous applications of morin to studies of Al in plant cells, accumulation of Al in the cell walls of elongating cells may have been wrongly interpreted as being a consequence of Al precipitation (Fig. 4 of Tice et al., 1992). Although it is not yet clear how and why Al accumulates in these discrete regions, intriguing information on the cessation of Al uptake in cells, and its suppression by BFA, has been obtained. This may contribute to a further understanding of Al uptake. On the other hand, inhibitors of cell growth processes had almost no effect on Al uptake (Tab. 2), whereas cells largely lost their capacity for Al uptake following incubation of cells in a minimal medium. This occurred presumably because of reduced growth upon incubation in minimal medium. The manner in which growth is reduced may determine whether Al uptake is affected or not, since BF A halted cell grth (Tab. 1) but did not reduce Al uptake. An intriguing and important finding was that BF A sustained the capacity of cells to incorporate Al, following incubation in minimal medium and concomitant reduction in growth (Fig. 4). Two points should be stressed. Brefeldin A did not reverse the loss in Al 96 uptake capacity of cells (Fig. 7b) but rather prevented its occurrence (Fig. 4). Brefeldin A itself had no significant effect on Al uptake, except possibly afier prolonged exposure of cells to BFA (Tab. 2, Figs. 4 and 7). The antibiotic brefeldin A, a fungus-derived macrocyclic lactone, is an inhibitor of vesicular traffic between the endoplasmic reticulum (ER) and the cistemae of the Golgi apparatus. Its application to cells results in disassembly of the Golgi apparatus and inhibition of the secretory pathway, particularly as assessed by protein secretion (Misumi et al, 1986; Driouich et al., 1993). Brefeldin A is widely used as an experimental tool, particularly in mammalian cells, for the study of Golgi-associated events (Klausner et al., 1992; Staehelin and Moore, 1995). However, in plant cells the effects of BF A on the Golgi are not as well defined (Satiat-Jeunemaitre et al., 1996). It does not always result in disassembly of the Golgi, and its effects have been observed to be more pronounced on the trans-Golgi network (Driouich et al., 1993). Brefeldin A is thought to inhibit binding of GTP to the ADP ribosylation factor (ARF), a small GTP-binding protein. This prevents the association of ARF and coatomer proteins with the membranes of the ER and Golgi, and consequently budding of vesicles is impeded (Torii et al., 1995). Prior application of GTPyS and AlF,‘ can inhibit the effects of BFA (Klausner et al., 1992). It is concluded that the effects of BFA on maintaining Al uptake in these tobacco cells resulted from the known interference of BF A on vesicular traffic, rather than secondary effects because: 1) The concentration of BFA employed was low (10uM), similar to those used in mammalian cells and is among the lowest reportedly used for plant cells (Robinson, 1993; Satiat—Jeunemaitre et al., 1996). In these tobacco BY-2 cells, 20 1.1M BFA was sufficient to promote disassembly of the Golgi (Yasuhara et al., 1995); 2) The effect of BF A 97 was highly reversible (Fig. 7A), a hallmark of the effect of this compound on the Golgi apparatus and its associated secretory pathway (Driouich et al., 1993); and 3) Addition of BFA to the cells only during Al incubation had no effect on Al uptake, i.e. there is no evidence that it somehow facilitated uptake of Al. Thus, by apparently inhibiting the Golgi secretory pathway, BFA maintains the cells capable of Al uptake upon incubation in minimal medium. In growing cells, the bulk of vesicular traffic is expected to occur between the trans Golgi apparatus and regions of cell growth and expansion. Indeed, in tobacco BY-2 cells, BF A inhibited the formation of cell plates (Yasuhara et al., 1995) and it inhibited auxin-mediated growth and secretion of cell- wall proteins in maize coleoptiles (Schindler et al., 1994). Aluminum accumulates in regions at or close to the cell surface (F igs.1, 2 and 3; Vitorello and Haug, 1996). This putative role of the Golgi secretory pathway in Al uptake lends support to the hypothesis that Al accumulates in regions of cell growth. The mechanism by which BFA inhibits the loss of A1 uptake capacity in cells is not clear. The effects of BFA on vesicular traffic do not seem to be limited to the Golgi apparatus, per se. There is evidence that it affects vesicular traffic along the exocytotic and endocytotic pathway also (Pelham, 1991; Driouich et al., 1993; Torii et al., 1995; Satiat- J eunemaitre et al., 1996). Thus, it is likely that in addition to reduced secretion, plasma membrane turnover is inhibited by BFA, caused either by inhibition of the Golgi secretory pathway or by inhibition of exocytosis and endocytosis. It is also likely that there are important changes in the composition of the cell surface when a region ceases growth (or vice versa). These changes require, in part, plasma membrane turnover and the transport of proteins, precursors etc. to the cell surface. Indeed, BFA reportedly inhibited the secretion 98 of auxin-binding protein in maize suspension cultures (Jones and Herman, 1993) and of a “wall-loosening factor”in maize coleoptiles (Schindler et al., 1994). Taken together, it is hypothesized that alterations in the composition of the cell surface are responsible for cessation of Al uptake. Maintenance of Al uptake by BFA results from the inhibition of these changes at the cell surface. What are these changes? Proteins or lipids could be involved. The transit of glycolipids and sphingomyelin through the Golgi to the plasma membrane was inhibited by BF A in BHK21 cells (Kallen et al., 1993). However, the transport of phosphatidylethanolamine, in hepatocytes, and cholesterol, in chinese hamster ovary cells, to the plasma membrane were not, even though protein secretion was greatly reduced (Urbani and Sirnoni, 1990; Vance et al., 1991). With respect to proteins, BF A differentially inhibited secretion of proteins in tobacco cell cultures cv. Havanna (Kunze et al., 1995). A clue to this puzzle may be the fact that low pH induces enhanced membrane permeability (Yan et al., 1992). This was found to occur in growing cells, but not in non- growing cells of these tobacco BY-2 cells, and this low-pH induced increase in membrane permeability may play a role in Al uptake (V itorello and Haug, 1996). Protonation of proteins and lipids involved in membrane fusion could be responsible for the destabilization of the plasma membrane. For the purpose of speculation and to illustrate a possible scenario, one possibility may be annexins, which are Ca21-dependent proteins which bind to acidic phospholipids and can be regulated by phosphorylation. These proteins can be transported to the cell surface by the Golgi secretory pathway and play a role in a wide variety of essential processes, among which are exocytosis, membrane fusion and vesicle trafficking. GTP-binding proteins are 99 sometimes associated in these processes (Clark and Roux, 1995). Annexins have been immunolocalized to tip-growing regions of pollen tubes and fern rhizoids (Blackboum et al., 1992; Clark et al., 1995) and also to secretory cells such as outer cells of the root cap (Clark et al., 1992;1995). In addition to involvement in exocytosis and secretory activity, annexins have been implicated with cellulose and callose synthases (Andrawis et al., 1993; Clark and Roux, 1995), and also with interactions of microfilaments, particularly actin, with the plasma membrane (Clark and Roux, 1995). These processes have all been reported to be affected by Al (Grabski et al., 1995; Schaeffer and Walton, 1990), and it is conceivable that a member of the annexin family is involved with Al accumulation in cells, either by generating membrane instability or perhaps by direct binding of Al. BF A treatment may cause this protein to remain at the cell surface upon incubation of cells in minimal medium and cessation of growth. REFERENCES Andrawis, A., Solomon, M., Delrner, DP. 1993. Cotton fiber annexins: a potential role in the regulation of callose synthase. Plant J. 3: 763-772. Bennet, R.J., Breen, CM. 1991. The aluminium signal: new dimensions to mechanisms of aluminium tolerance. Plant Soil 134:153-166. Blackboum, H.D., Barker, P.J., Huskisson, N.S., Battey, NH. 1992. Properties and partial protein sequence of plant annexins. Plant Physiol. 99:864-871. Clark, G.B., Dauwalder, M., Roux, SJ. 1992. Purification and immunolocalization of annexin-like protein in pea seedlings. Planta 187: 1-9. Clark, G.B., Tumwald, S. Tirlapur, U.K., von der Mark, K., Roux, S.J., Scheuerlein, R. 1995. Induction and polar distribution of annexin-like proteins during phytochrome- mediated rhizoid initiation and growth in spores of the ferns Diyopteris and Anemia. Planta 197:376-3 84. Clark, G.B., Roux, SJ. 1995. Annexins of plant cells. Plant Physiol. 109:1133-1139. Doonan, J .H., Cove, D.J., Lloyd, CW. 1988. Microtubules and microfilaments in tip growth: evidence that microtubules impose polarity on protonemal growth in Physcomitrella patens. J. Cell Sci. 89:533-540. Driouich, A., Zhang, G.F., Staehelin, LA. 1993. Effect of brefeldin A on the structure of the Golgi apparatus and on the synthesis and secretion of proteins and polysaccharides in sycamore maple (Acer pseudoplatanus) suspension-cultured cells. Plant Physiol. 101:1363-1373. Grabski, S., Schindler, M. 1995. Aluminum induces rigor within the actin network of soybean cells. Plant Physiol. 108:897-901. Hepler, P.K., Bonsignore, CL. 1990. Caffeine inhibition of cytokinesis: ultrastructure of cell plate/degradation. Protoplasma 157: 182-192. Huck MG (1972) Impairment of sucrose utilization for cell wall formation in the roots of aluminum-darnaged cotton seedlings. Plant Cell Physiol 13:7-14 100 101 Jones, AM. and Herman, EM. 1993. KDEL-containing auxin-binding protein is secreted to the plasma membrane and cell wall. Plant Physiol. 101:595-606. Jones, D.L., Shaff, J .S., and Kochian, L.V. 1995. Role of calcium and other ions in directing root hair tip growth in Limnobium stoloniferum. 1. Inhibition of tip growth by aluminium. Planta 197:672-680. Kallen, K.J., Quinn, P., Allan, D. 1993. Effects of brefeldin A on sphingomyelin transport and lipid biosynthesis in BHK21 cells. Biochem. J. 289:307-312. Klausner, R.D., Donaldson, J .G., Lippincott-Schwartz, J. 1992. Brefeldin A: insights into the control of membrane traffic and organelle structure. J. Cell. Biol. 116: 1071-1080. Kochian, L.V. 1995. Cellular mechanisms of aluminum toxicity and resistance in plants. Annu. Rev. Plant Physiol. Plant Mol. Biol. 46: 237-60. Konishi, S., Miyamoto, S. 1983. Alleviation of aluminum stress and stimulation of tea pollen tube growth by fluorine. Plant Cell Physiol. 24:857-862. Kunze, I., Hillmer, S., Kunze, G., Muntz, K. 1995. Brefeldin A differentially affects protein secretion from suspension-cultured tobacco cells (Nicotiana tabacum L.). Misumi, Y., Miki, A., Takatsuki, A., Tamura, G., Ikehara, Y. 1986. Novel blockade by brefeldin A of intracellular transport of secretory proteins in cultured rat hepatocytes. J. Biol. Chem. 261 :1 1398-1 1403. Pelham, HR. 1991. Multiple targets for Brefeldin A. Cell 67:449-451. Rincon, M., Gonzales, RA. 1992. Aluminum partitioning in intact roots of aluminum- tolerant and aluminum-sensitive wheat (Triticum aestivum L.) cultivars. Plant Physiol. 99:1021-1028. Robison, D.G. 1993. Brefeldin A: a tool for plant cell biologists? Bot. Acta 106: 107-109. Ryan, P.R., DiTomasso, J.M., Kochian, L.V. 1993. Aluminium toxicity in roots: an investigation of spatial sensitivity and the role of the root cap. J. Exp Bot 44:437-446. Samuels, A.L., Giddings, T.H., Staehelin, LA. 1995. Cytokinesis in tobacco BY-2 and root tip cells: a new model of cell plate formation in higher plants. J. Cell Biol. 130: 1345- 1357. Satiat-Jeunemaitre, B., Hawes, C., 1992. Reversible dissociation of the plant Golgi apparatus by brefeldin A. Biol. Cell 74:325-328. 102 Satiat-Jeunemaitre, B., Cole, L., Bourett, T., Howard, R., Hawes, C. 1996. Brefeldin A effects in plant and fungal cells: something new about vesicle trafficking. J. Microscopy 181:162-177. Schaeffer, H.J., Walton, J .D. 1990. Aluminum ions induce oat protoplasts to produce an extracellular (1-3)B-D—glucan. Plant Physiol. 94:13-19. Shi, B., Chou, K., Haug, A. 1993. Aluminium impacts elements of the phosphoinositide signalling pathway in neuroblastoma cells. Mol. Cell. Biochem. 121:109-118. Schindler, T., Bergfeld, M., Hohl, M., Schopfer, P. 1994. Inhibition of Golgi-apparatus function by brefeldin A in maize coleoptiles and its consequences on auxin-mediated growth, cell-wall extensibility and secretion of cell-wall proteins. Planta 192: 404-413. Schreiner, K.A., Hoddinott, J ., Taylor, G.J. 1994. Aluminum-induced deposition of (1,3)- beta-glucans (callose) in Triticum aestivum L. Plant Soil 162:273-280. Searcy, K.B., Mulcahy, D.L. 1990. Comparison of the response to aluminum toxicity in gametophyte and sporophyte of four tomato (Lycapersican esculentum Mill.) cultivars. Theor. Appl. Genet. 80:289-295. ' Shedletzky, B., Shrnuel, M., Trainin, T., Kalman, S., Delmer, D. 1992. Cell wall structure in cells adapted to growth on the cellulose-synthesis inhibitor 2,6-dichlorobenzonitrile. Plant Physiol. 100: 120-130. Staehelin, L.A., Moore, 1. 1995. The plant Golgi apparatus: structure, functional organization and trafficking mechanisms. Ann. Rev. Plant Physiol. Plant Mol. Biol. 46: 261-288. Tice, K.R., Parker, D.R., DeMason, D.A. 1992. Operationally defined apoplastic and symplastic aluminum fractions in root tips of aluminum-intoxicated wheat. Plant Physiol. 100: 309-318. Torii, S., Banno, T., Watanabe, T., Ikehara, Y., Murakami, K., Nakayama, K. 1995. Cytotoxicity of Brefeldin A correlates with its inhibitory effect on membrane binding of COP coat proteins. J. Biol. Chem. 270: 11575-1580. Urbani, L., Sirnoni, RD. 1990. Cholesterol and vesicular stomatitis virus G protein take separate routes from the endoplasmic reticulum to the plasma membrane. J. Biol. Chem. 265:1919-1923. Vance, J .E., Aasman, E.J., Szarka, R. 1991. Brefeldin A does not inhibit the movement of phosphatidylethanolamine from its sites of synthesis to the cell surface. J. Biol. Chem. 266:8241-8247. 103 Vitorello, VA. and Haug, A. 1996. Short-term aluminium uptake by tobacco cells: Growth dependence and evidence for internalization in a discrete peripheral region. Physiol. Plant. (in press). Yasuhara, H., Sonobe, S., Shibaoka, H. 1995. Effects of brefeldin A on the formation of the cell plate in tobacco BY-2 cells. Eur. J. Cell Biol. 66:274-281. Yamamoto, Y., Rikiishi, S., Chang, Y., Ono, K., Kasai, M., Matsumoto, H. 1994. Quantitative estimation of aluminium toxicity in cultured tobacco cells: correlation between aluminium uptake and growth inhibition. Plant Cell Physiol. 35: 575-583. Yan, F., Schubert, S., Mengel, K. 1992. Effect of low root medium pH on net proton release, root respiration, and root growth of corn (Zea mays L.) and broad bean (Viciafaba L.). Plant Physiol. 99: 415—421. Zhang, G., Taylor, G.J. 1990. Kinetics of aluminum uptake in T riticum aestivum L.. Identity of the linear phase of aluminum uptake by excised roots of aluminum-tolerant and aluminum-sensitive cultivars. Plant Physiol. 94: 577-584. CHAPTER 5 GENERAL DISCUSSION AND PERSPECTIVES GENERAL DISCUSSION AND PERSPECTIVES In this study, some novel findings on Al uptake by cultured tobacco cells are reported. A major theme to emerge from this research is that cell growth is somehow implicated with Al uptake. Non-growing cells do not take up Al, neither do they appear to be sensitive to it, as evidenced from the absence of observable symptoms of toxicity in non-elongating root cells (Ryan et al., 1993). Another major novelty is the discovery of discrete Al-accumulating regions at the periphery of the cell and Al-induced abnormalities at the cell surface. An intriguing finding is that brefeldin A-sensitive membrane/vesicular traffic plays a role in whether a cell becomes incapable of Al uptake. What is the relationship between cell growth and Al uptake? What role do these Al-accumulating regions play? How does membrane/vesicular traffic affect a cell’s capability to take up Al? These are some of the questions that are posed. The objective of this chapter is to summarize and discuss the findings on Al uptake, to present a model for uptake of A1 and to discuss future research perspectives. Using the fluorescence assay developed for the detection of Al in these tobacco BY-2 cells (chapters 2 and 3) and also GFAAS, Al uptake was found to be correlated with and dependent on cell growth (Chapter 3). When growing cells were incubated in minimal medium, leading to reduced or arrested growth, Al uptake almost ceased to occur (Chapter 4). The underlying causes for the association between A1 uptake and cell growth are not yet known. However, greater availability of metabolic energy for Al uptake, during cell growth, 104 105 does not appear to be the cause. In intact and excised roots, metabolic inhibitors were demonstrated to increase Al uptake (Wagatsuma, 1983; Zhang and Taylor, 1989; Zhang and Taylor, 1991) rather than decrease it. However, metabolic inhibitors (DNP) had no effect on Al uptake in Neuroblastoma cells (Shi and Haug, 1990). Some additional findings are summarized. The kinetics of Al uptake in these cultured cells was found to be similar to those found in roots, which is important if use of these cells are desired as a plant cell model for Al uptake. Other than age, factors which were found to affect A1 uptake were Ca2+ concentration, low temperature, increased osmoticum, and pre- chelation of Al to citrate (Chapters 3 and 4). The low pH of 4.5, required in Al toxicity studies, was itself detrimental to cells, causing increased membrane permeability. Some evidence for a correlation between A1 uptake and this low-pH induced membrane permeability was found (Chapter 3). The ameliorative effects of factors such as cations (ionic strength) and low temperature, upon Al uptake, may result from their modulating effect on low-pH induced alterations in membrane structure. The physiological effects of the low pH of solutions which are necessary for Al toxicity to occur, must obviously be taken into consideration when examining the effects of Al. As judged with this Al-morin fluorescence technique, Al accumulated in discrete peripheral regions of the cell, in particular in protuberances and in material presumably originating therefrom (Chapters 3 and 4). These structures were induced by treatment with Al. Generalized binding of Al to the cell wall was not detected. The cell wall and the plasma membrane appeared to become closely associated in Al-accumulating regions, as evidenced by plasmolysis of Al- and morin-labelled cells, in which the plasma membrane tended to remain in contact with the cell wall in regions where Al accumulated. Plasmolysis of cells 106 prior to exposure to Al resulted in very low levels of Al uptake and virtually no localized regions of Al-morin fluorescence. It appears that any accumulation of Al in the cell wall, in these discrete Al-accumulating regions, is the result of the release of cellular material, induced by Al itself, into the cell wall. First, however, Al is taken up by the protoplast. This is supported by the statements made above, by the observation of Al associated with the symplast of plasmolyzed cells, and by arguments in Chapter 3, which support Al internalization. It is suggested that the regions of Al accumulation are regions where the cells are growing, i.e., where new membrane and cell wall are being added. This is suggested because 1.) A1 uptake occurs only in growing cells; 2) Al accumulation is localized to a discrete region close to the cell surface, and the pattern of Al localization is suggestive of events occurring during cytokinesis in these tobacco BY-2 cells (Samuels et al., 1995); 3) Al induces deformities of the cell surface, suggestive of deformities induced by Al in growing root hair tips (Jones et al., 1995); 4) BFA-sensitive membrane/vesicular traffic affects the cell’s capability of Al uptake. Two major experimental tools were instrumental for generating the new information. A fluorescence assay for the detection of A1 and use of a cell culture as a plant cell model. The fluorescence method was adapted, optimized and a comparison of results obtained with this technique and with GFAAS validated this method as a technique for the detection of Al (Chapter 2). This method has several advantages over other methods of detection previously employed, such as GFAAS itself, X-ray microanalysis and the recently employed method of secondary-ion mass spectrometry (Lazof et al., 1994). It is sensitive and allows both for quantitative measurements and intracellular localization, with higher resolution than 107 previously possible. It is performed on intact cells, without sample preparation and is easy, rapid and inexpensive to perform. It is also apparently insensitive to Al precipitates. Because of these advantages, this method should be useful in future studies on Al toxicity, a field which has been limited by appropriate methodology (Taylor, 1988; Delhaize and Ryan, 1995) and is potentially applicable to other systems such as root tips. A tentative model for uptake of Al by plant cells is proposed (Fig. 1). Al has access to potential targets because of events occurring in regions of cell growth, during membrane turnover and cell wall deposition. Membrane instabilities which occur during membrane fusion events serve as a port of entry for Al, or expose binding sites of membrane-bound proteins. The observed increases in membrane permeability at low pH (Chapter 3) presumably results from enhanced membrane instability of these regions. Low pH has been shown to promote non-bilayer configurations in some lipids (Lindblom and Rilfors, 1992). Alternatively, membrane-bound proteins become protonated at low pH and low cation concentration, and could be responsible for membrane instabilities. Thus, any treatment which has an effect on membrane stability, could affect Al uptake. This would be consistent with the effects of low temperature, high Ca2+ and cation concentration, reduced cell turgor and metabolic inhibitors on Al uptake. Such a model based on membrane properties could explain why non-metabolic conditions provided by cold treatment and metabolic inhibitors produce Opposing effects on Al uptake (Wagatsuma, 1983; Zhang and Taylor, 1989, 1990). In this model, the role of BF A-sensitive membrane/vesicular traffic in determining a cell’s capability for Al uptake rests on its ability to transport material (proteins, lipids, polysaccharides etc.) to and from the cell surface. Thus, the arrival or departure of key components at the cell surface may greatly determine whether AI will accumulate in this 108 growing PM CW ER cell BFA i vesicles % A|3+ membrane instability at low pH and low > ionic strength Figure l. A schematic representation of a model for uptake of Al into regions of growing cells involved in the incorporation of new cell wall and plasma membrane. Membrane instabilities in these regions resulting from fusion events, and enhanced by low-pH and low ionic strength, facilitate access of Al to the interior of the cell or to Al-binding proteins of the plasma membrane. Accumulation of Al in these regions causes derangements in processes involved in growth, and eventually in deformations of the cell surface. Upon cessation of growth, changes in the cell surface impede Al uptake. These changes depend on vesicular traffic, which is inhibited by action of BFA. ER-endoplasmic reticultun, PM- plasma membrane, CW-cell wall, BFA-brefeldin A. 109 region or not. The question is what are these components? Following Al uptake, Al-induced derangements in this cell region could cause the observed release of cellular material (Chapter 4). Since this material was found to be exterior to the rest of the cell, this might help explain why there has been much controversy and ambiguity over whether Al accumulates in the apoplast or the symplast. Tice et a1. (1992), in one of few previous applications of morin on plant tissue, observed localized Al in the cell wall, which was interpreted to be Al precipitation, but is perhaps better explained by the release of Al-containing material. Furthermore, this can also explain Zhang and Taylor’s (1990, 1991) “metabolism dependent” accumulation of Al within the cell wall. Yarnamoto et al. (1994) have recently demonstrated a correlation between A1 uptake and Al toxicity in tobacco cell cultures. Exclusion of A] may be a mechanism for avoiding Al toxicity. Increased cation concentration and chelation of Al to citrate in the solution both reduced Al uptake (Chapter 3) and both are well known to ameliorate Al toxicity in roots (Kinraide and Parker, 1987; Suhayda and Haug, 1986). Recently, it has been reported that tolerance to Al may be associated with enhanced malate exudation (Delhaize et al., 1993, Basu et al., 1994; Pellet et al, 1995) which would fimction to keep Al outside of the cell. That A1 is excluded from non-growing cells is clearly demonstrated in Chapter 3. The root apex, specifically the meristem and elongating cells, has recently been shown to be the primary site of inhibition of root grth by Al (Ryan et al., 1993). In addition to these cells, Al has also been shown to cause injuries to root cap and root hair cells (Bennet and Breen, 1991). Our results and suggested model point to a common denominator between these cells, namely the fact that they are growing or actively involved in secretory processes. 110 Future perspectives The Al-morin fluorescence assay is potentially very useful in Al toxicity studies and also in plant screening programs. In a very recent application (Larsen et al., 1996), morin was utilized to examine Al in Al-sensitive mutants of Arabidopsis. Imaging of Al in intact root tissues could be accomplished with confocal microscopy and the Al-morin technique. Hopefully, the work described here will foster continued investigations into the matter of how and where Al accumulates in cells. Further supporting evidence for the hypothesis that Al accumulates in cell regions involved in new plasma membrane and cell wall synthesis might be obtained by examining Al-morin fluorescence in tip-growing cells such as pollen tubes and root hairs. The growing region of these cells is well defined. Similar examinations could be performed in budding yeast. Use of markers of regions of cell growth would be ideal to determine if Al accumulation coincides with these regions. Annexins, Cay-dependent phospholipid-binding proteins involved in membrane fusion, are possible candidates for use as markers. Immunolocalization studies have determined that annexins are concentrated in the tip region of polarly growing cells such as pollen tubes and fern rhizoids (Blackbourn et al., 1992; Clark et al., 1995) and also in secretory cells such as the outer cells of the root cap (Clark et al., 1992). In a recent study, Al induced rigidification in transvacuolar strands of actin in soybean (Glycine max) cell suspensions (Grabski and Schindler, 1995). It has not been shown whether Al reaches this actin. IfAl accumulates in regions at or near the cell surface, how does it interact with actin? Membrane-associated annexins have been shown to interact with and modulate actin bundling (Clark and Roux, 1995; Glenney et al., 1987). Thus, III studies relating the cellular distribution of Al to immunolocalization of annexins and actin would be informative. In such studies, it would also be of interest to examine if actin and annexins are present in Al-induced extracellular material (chapter 4.). In preliminary studies (data not shown) these external structures were found to be detergent-insoluble, as are cytoskeleton fractions. The feasibility of such studies would depend, in part, on the accessibility to needed antibodies. Another approach to examine the nature of these Al-accumulating regions would be the use of yeast as a model system. Processes involved in cell growth and in the secretory pathway are better understood in yeast than in plants, and a larger selection of experimental tools is available. In yeast there are many mutants of growth-associated processes, such as the Golgi-mediated secretory process. The genetic dissection of these processes is an active area of research (Chant, 1994). Such an approach could prove very useful. However, little is known about Al toxicity in yeast and their susceptibility to A1. The possible relationship between low pH, plasma membrane permeability to ions, Al uptake and cellular regions of growth should be examined. One possibility would be to perform patch-clamp studies on growing cells. Tip-growing cells, such as pollen tubes or root hairs, could be used to examine whether changes in membrane permeability are localized to regions of cell growth, and whether there is a correlation with accumulation of Al. An obvious obstacle to this approach is the presence of the cell wall. In at least one study, however, laser microsurgery was used to expose the plasma membrane of the growing apex of F ucus spiralisn rhizoid cell so that conventional patch clamp studies could be performed (Taylor and Brownlee, 1992). A better alternative to patch clamping is the use of extracellular vibrating ion-selective microelectrodes, which permits non-invasive 112 measurements of ion fluxes (Smith et al., 1994).This technique has recently been used to determine the magnitude and spatial localization of Ca”, K’ and H" fluxes in root hairs of Limnobium stolom'ferum (Jones et al., 1995). This technique could be used in conjunction with the localization of Al by morin fluorescence to answer important questions. Is Al accumulation co-localized with low-pH increases in plasma membrane permeability to ions? ls Al uptake modulated by factors which similarly modulate low-pH induced plasma membrane ion permeability? What is the role of Ca2+ in stabilizing the membrane and in modulating Al uptake? A question which arises from these considerations is whether a correlation can be found between tolerance to low pH and tolerance to Al in plants. Finally, to understand what processes are involved with Al uptake in growing cells, studies comparing the physiology of cells capable of Al uptake and those which were induced to lose their capacity for Al uptake could be performed. For example, are there changes in plasma membrane composition (both in terms of lipids and proteins) between these cells? Does BFA block these changes? An obvious problem with this approach is that changes in a relatively small region of the plasma membrane may not be detectable if whole cells or even plasma membrane preparations are used for analysis. Furthermore, finding such differences would not necessarily establish a causal link. REFERENCES Basu, U., Godbold, D., Taylor, G.J. 1994. Aluminum resistance in Triticum aestivum associated with enhanced exudation of malate. J. Plant Physiol. 1442747-753. Bennet, R.J., Breen, CM. 1991. The aluminium signal: new dimensions to mechanisms of aluminium tolerance. Plant Soil 134:153-166. Bennet, R.J., Breen, C.M., Fey, M.V. 1987. The effects of aluminium on root cap function and root development in Zea mays L. Environ. Exp. Bot. 27:91-104. Blackboum, H.D., Barker, P.J., Huskisson, N.S., Battey, NH. 1992. Properties and partial protein sequence of plant annexins. Plant Physiol. 99:864-871. Chant, J. 1994. Cell polarity in yeast. Trends in Genetics 10: 328-333. Clark, G.B., Dauwalder, M., Roux, SJ. 1992. Purification and immunolocalization of annexin-like protein in pea seedlings. Planta 187: 1-9. Clark, G.B., Roux, SJ. 1995. Annexins of plant cells. Plant Physiol. 109:1133-1139. Clark, G.B., Tumwald, S. Tirlapur, U.K., von der Mark, K., Roux, S.J., Scheuerlein, R. 1995. Induction and polar distribution of annexin-like proteins during phytochrome- mediated rhizoid initiation and growth in spores of the ferns Dryopteris and Anemia. Planta 197:376—384. Delhaize, B., Ryan, P.R., Randall, P.J. 1993. Aluminum tolerance in wheat (Triticum aestivum L.). 11. Aluminum-stimulated excretion of malic acid from root apices. Plant Physiol. 103:695-702. Delhaize, B., Ryan, PR. 1995. Aluminum toxicity and tolerance in plants. Plant Physiol. 107: 315-321. Glenney, J.R., Tack, B., Powell, MA. 1987. Calpactins: two distinct Ca“ -regulated phospholipid- and actin-binding proteins isolated from lung and placenta. J. Cell. Biol. 104: 503-511. Grabski, S., Schindler, M. 1995. Aluminum induces rigor within the actin network of soybean cells. Plant Physiol. 108:897-901. 113 114 Jones, D.L., Shaff, J .S., and Kochian, L.V. 1995. Role of calcium and other ions in directing root hair tip growth in Limnobium stoloniferum. 1. Inhibition of tip growth by aluminium. Planta 197:672-680. Kinraide, T.B., Parker, DR. 1987. Cation amelioration of aluminum toxicity in wheat. Plant Physiol. 83: 546-551. Larsen, P.B., Tai, C.Y., Kochian, L.V., Howell, SH. 1996. Arabidopsis mutants with increased sensitivity to aluminum. Plant Physiol. 110:743-751. Lazof, D.B., Goldsmith, J .G., Rufty, T.W., Linton, R. 1994. Rapid uptake of aluminum into cells of intact soybean root tips. A microanalytical study using secondary ion mass spectrometry. Plant Physiol. 106: 1107-1114. Lindblom, G., Rilfors, L. 1992. Cubic and hexagonal lipid phases. In Structure and dynamic properties of lipids and membranes, (P.J. Quinn & R.J. Cherry, eds), pp 51-76. Portland Press, London. ISBN 1-855-78-014-3. Pellet, D.M., Grunes, D.L., Kochian, L.V. 1995. Organic acid exudation as an aluminum- tolerance mechanism in maize (Zea mays L.). Planta 196:788-795. Ryan, P.R., DiTomasso, J .M., Kochian, L.V. 1993. Aluminium toxicity in roots: an investigation of spatial sensitivity and the role of the root cap. J. Exp Bot 44:437-446. Samuels, A.L., Giddings, T.H., Staehelin, LA. 1995. Cytokinesis in tobacco BY-2 and root tip cells: a new model of cell plate formation in higher plants. J. Cell Biol. 130: 1345- 1357. Shi, B., Haug, A. 1990. Aluminum uptake by neuroblastoma cells. J. Neurochem. 55: 551- 558. Smith, J .S., Sanger, R.H., Jaffe, LR 1994. The vibrating Caz’ electrode: a new technique for detecting plasma membrane regions of Ca2+ influx and efflux. Meth. Cell Biol. 40: 115- 134. Suhayda, C.G., Haug, A. 1986. Organic acids reduce aluminum toxicity in maize root membranes. Physiol. Plant. 68:189-195 Taylor, G.J. 1988. The physiology of aluminium phytotoxicity. In: Sigel, H., Sigel, A. (eds). Metal ions in biological systems: Aluminum and its role in biology, vol. 24, pp 123-163. New York, Marcel Dekker. Taylor, A.R., Brownlee, C. 1992. Localized patch clamping of plasma membrane of a polarized plant cell. Plant Physiol. 99:1686-1688. 115 Tice, K.R., Parker, D.R., DeMason, D.A. 1992. Operationally defined apoplastic and symplastic aluminum fiactions in root tips of aluminum-intoxicated wheat. Plant Physiol. 100: 309-318. Wagatsuma, T. 1983. Effect of non-metabolic conditions on the uptake of aluminum by plant roots. Soil Sci. Plant Nutr. 29: 323-333. Yamamoto, Y., Rikiishi, S., Chang, Y., Ono, K., Kasai, M., Matsumoto, H. 1994. Quantitative estimation of aluminium toxicity in cultured tobacco cells: correlation between aluminium uptake and growth inhibition. Plant Cell Physiol. 35: 575-583. Zhang, G., Taylor, G.J. 1989. Kinetics of aliuninum uptake by excised roots of aluminum tolerant and aluminum sensitive cultivars of Triticum aestivum L.. Plant Physiol. 91: 1094-1099. Zhang, G., Taylor, G.J. 1990. Kinetics of aluminum uptake in Triticum aestivum L.. Identity of the linear phase of aluminum uptake by excised roots of aluminum-tolerant and aluminum-sensitive cultivars. Plant Physiol. 94: 577-584. Zhang, G., Taylor, G.J. 1991. Effects of biological inhibitors on kinetics of aluminium uptake by excised roots and purified cell wall material of aluminium-tolerant and aluminium-sensitive cultivars of Triticum aestivum L. J. Plant Physiol. 138:533-539. APPENDIX A GROWING TOBACCO CELLS RESPOND TO RAPID MEDIUM CHANGES WITH A TRANSIENT INCREASE IN LOCALIZED Cit2+ ACTIVITY Vitorello, V.A., Haug, A. 1996. Plant Cell Reports (in press). ABSTRACT A suspension of tobacco cells, Nicotiana tabacum L. BY-2, was subjected to a rapid medium change, resulting in a disturbance of growth. A subpopulation of growing cells responded to such a nutritional signal by establishing a transient, localized Caz’ accumulation, as judged by chlorotetracycline fluorescence. Residing near or at the plasma membrane, this initial Ca2+ signal began to relax after I h to a value presumably corresponding to an equilibrium Ca2+ level. This response was susceptible to treatment with brefeldin A, an agent impacting vesicular traffic, as indicated by a fiirther fluorescence increase. By contrast, undisturbed growing and non-growing cells did not display a Ca2+ response, regardless of the presence of brefeldin A. 116 INTRODUCTION Plant cells are confronted with the task of continually monitoring their microenvironment and of adjusting intracellular parameters in response to varying external conditions. Given the significance of nutrient availability in the plant's environment, signalling pathways must exist coupling growth promotion or arrest to nutrition-linked signals. How plants sense their environment, and how signals are recognized and transduced are poorly understood mechanisms. A potential participant in this type of signal transmission is the Ca2+ signal, which in turn is triggered or modulated via upstream pathways involving components like phosphoinositides, GTPases, and protein kinases (Benidge 1995). In plant cells, Ca2+ is a key player in diverse cellular processes such as tip growth during pollen germination and elongation (Herth et al. 1990; Zonia and Tupy 1995). Evidence also indicates that Ca2+ exerts a profound influence upon the transport and distribution of Golgi-synthesized, noncellulosic material, destined for the cellular surface, via specialized carrier vesicles (Schmid and Damke 1995). For example, in animal cells (Robinson et al. 1996), in maize, and growing pollen tubes (Clark and Roux 1995), annexin proteins have been identified which interact in a Cay-dependent manner with secretory vesicles mediating exocytotic processes. The Golgi apparatus of growing plant cells, as opposed to that of mammalian cells (Rothman and Orci 1990; Cole and Lippincott-Schwartz 1995) is intensively engaged in synthesizing complex cell wall polysaccharides (Staehelin and Moore 1995). To decipher mechanisms of vesicular traffic, in part Ca2’-dependent, 117 118 usage of yeast mutants, defective in certain steps of vesicular pathways (Chant 1994), and treatment of mammalian cells (Klausner et al. 1992; Jackson and Kepes 1994), and plant cells (Staehelin and Moore 1995), with a macrocyclic lactone, brefeldin A (Merck 1989), have provided considerable insight into vesicular traffic. Given the paucity of information on pathways mediating signals in plant cells, we examined the intracellular Ca2+ activity in growing tobacco cells, stressed by a rapid change in medium. In response to such a signal, a transient localized Ca2+ accumulation is observed, based on chlorotetracycline fluorescence. Furthermore, the transient increase in Ca2+ activity responds to brefeldin A-treatrnent. MATERIALS AND METHODS Cell culture and treatments Cell suspension cultures of Nicotiana tabacum L. BY-2 were grown in a culture medium (Nagata et al. 1992), at pH 5.6, 28°C, and an extracellular CaCl2 concentration of 3 mM. Cell growth was determined by measuring packed cell volume (3-5 min centrifugation at 200 g in graduated tubes); each cm3 of packed cell volume corresponded to 7.9x10" cells. Cellular grth was disturbed by rapidly (about 5 min) transferring cells from the culture medium into a minimal medium composed of 10 mM MES, 5 mM sucrose, 5 mM KCl, 2 mM CaClz, pH 5.6, 23°C. Brefeldin A (Sigma, St. Louis, MO) was disssolved in methanol (10 mM), applied to cells at a final concentration of 10uM, and incubated for 30 min prior 119 to and during the duration of the treatment. Fluorescence assay Employing a titer of 4x1 05 cells/ml, intracellular Caz’ was evaluated with chlorotetracycline (Sigma, St. Louis, MO). After removal of the supernatant, and resuspension of cells in minimal medium devoid of the dye, but with 5 mM EDTA, fluorescence measurements were performed on a Perkin-Elmer LS-S luminescence spectrometer coupled to a microcomputer. Wavelength settings were 395:5 nm for excitation and 520:2.5 nm for emission. Each sample reading was taken as the average of the first 15 s of monitoring. Reductions in fluorescence intensity, attributable to cell settling, were detectable only after at least 1 min. Loading of cells with chlorotetracycline presented no difficulties in contrast to problems encountered in loading plant cells with Ca2+ indicators like Fluo-3 (Timmers et al. 1991), because of cell wall-associated esterases (Callaham and Hepler 1991). After entering the cell, this fluorescent antibiotic reports Ca” in the solution adjacent to hydrophobic sites like membranes, with which the Ca2+ is in equilibrium (Tsien 1989). Microscopy Cells were observed with a Zeiss (Carl Zeiss, Oberkochen, Germany) WL research microscope equipped with an epifluorescence condenser IV FL. The filter set used consisted of a band pass filter (BP 400-440 nm) and a long-wave pass filter (LP 470 nm). For photomicroscopy a Kodak Tmax p-3200 (Eastman Kodak Co., Rochester, NY) film was 120 employed. RESULTS AND DISCUSSION When tobacco cells were rapidly transferred from the culture medium into a minimal medium, then labelled with chlorotetracycline (50 uM, 30 min), a pronounced fluorescence intensity was observed. This fluorescence intensity was dependent on the growth stage of the culture, and their relative growth rate (figs. 1A and B). Actively growing cells, at about 1 to 3 days after subculture, displayed the highest fluorescence levels, while cells close to or at stationary phase (5 and 6 days after subculture), manifested fluorescence levels close to that of autofluorescence. If maintained in culture medium, rather than incubated in minimal ‘ medium, actively growing cells (2 days after subculture), labelled with chlorotetracycline, also displayed a weak fluorescence signal close to that of autofluorescence. As to cells transferred to minimal medium, the fluorescence intensity increase was detected after a rather short period (15 min) of incubation in this type of medium, peaked after about 1 h, and then gradually declined towards the initial level (fig. 2). To follow this type of response, chlorotetracycline labelling of cells was reduced to 15 min, and was performed at the end of each incubation period. As observed by epifluorescence microscopy on growing cells, exposed to minimal medium, the enhanced Ca2+ activity was found to be initially localized in a discrete region near or at the plasma membrane (fig. 3A). However, with longer exposures to minimal medium (2 1 h), fluorescence was often observed to occur over the entire cellular periphery. 121 1e 0 i- A or». _ €1512 ”i 5 _ 85’ (I 534.. c 5 i: A g ~125§ E ~100§ = fl 8 45% o 3 ~50 e 3 or c. c "25 > a: \ 3 0 Fa r T l '1‘ ‘0 2 0 1 2 3 4 5 6 days after subculture Fig. 1 A: Cellular Ca2+ activity, reported by chlorotetracycline fluorescence, as a function of the growth stage of tobacco cultures. Cells in minimal medium were labelled with 50 12M dye for 30 min. Values are reported after subtraction of autofluorescence. Bars represent SE (n=3). B: Growth curve and relative growth rates (RGR) of the cultures. The growth curve can be described by PCV=67.6 - 64.4exp(- 0.00599), 1'2 = 0.98, where PCV is the packed cell volume (%), and t is the time (days). The relative growth rate was calculated from the expression RGR =d(PCV)/dt(1/(PCV). 122 0 minimal medium I continuous culture _ _ 1g m)brefeldin A ...\ x§\\\§ -. removal of hnhMMA ..a ...a O N I I co 1 .5 l \ \ chlorotetracycline fluorescence (arbitrary units) or 1 O I l l l I 0 30 60 90 120150180 time (min) Fig. 2. Changes in Ca2+ activity of growing tobacco cells, as reported by chlorotetracycline fluorescence, following incubation in a minimal medium, or maintenance in culture medium. Cells (2 days after subculture) were labelled with the dye (50 HM) during the last 15 min of the incubation period. Some cells were also treated with 10 11M brefeldin A (30 min before and throughout medium treatment). To test the reversibility of brefeldin A effects, the treatment solution was replaced after 1 h with a fresh minimal medium devoid of the antibiotic. The mean chlorotetracycline fluorescence intensity of cells close to the stationary phase (5 days after subculture) was about two fluorescence units, regardless of incubation in minimal medium. Values are reported afier subtraction of autofluorescence. Bars represent SE (n=3). 123 Fig. 3 Photomicrographs of growing tobacco cells (2 days after subculture) labelled with chlorotetracycline (50 MM, 15 min). A: Growing cells maintained in culture medium, displaying lack of localized fluorescence; B: growing cells transferred to and incubated in minimal medium for 1 h, displaying localized fluorescence. The camera's shutter release speed (0.5 s) was‘the same for both images. 124 20 um Figure 3. 125 By contrast, continuously growing and stationary phase cells lacked any localized fluorescence (such as in fig. 3B). In minimal medium, only a subpopulation of the cells, ranging from 7 to 16 % of the total cell population, displayed a transient, localized Ca2+ response. The observed changes in fluorescence intensity (fig. 2) can in part be attributed to variations in the percentage of cells displaying localized fluorescence. At 15, 60, and 180 min of exposure to minimal medium, 7, 13, and 9 % of the cells displayed a response, respectively. In this subpopulation, the Ca2+ signal relaxed within a few hours to a value presumably corresponding to a global equilibrium Ca2+ level, representative of the stationary phase. Since the cell population is not synchronized, and only a subpopulation displayed a transient Ca”, the response is expected to be sharper and more intense in a synchronized culture. In this respect, the tobacco cultivar employed may serve as a good plant model because these cells can be highly synchronized (Nagata et al 1992). When cells were growing (2 days after subculture) in culture medium, cell growth decreased in response to medium changes (fig. 4). This reduction in growth was presumably temporary, indicative of a new lag phase. Alterations in cellular growth can be expected to be accompanied by changes in the Golgi-mediated secretory pathway, wherein Caz’ is known to participate in the regulation of vesicular traffic (Schmid and Damke 1995). Application of 10 “M brefeldin A to growing cells resulted in further enhancement of cellular Ca2+ activity, following transfer of cells to a minimal medium, but not when these cells were kept in culture medium (fig. 2). The effects of brefeldin A were found to be highly reversible (fig. 2). With 10 1.1M brefeldin present, 12, 14, and 16% of the cells responded, respectively. The lack of detectable local Ca2+ accumulations in tobacco cells growing in culture medium, both in the presence or absence of brefeldin A (figs. 2 and 3A), seems to be in accord with 126 16 14 — s ,, ° 12 — \ ° :5 a E \ E ,, u y i R 8 E 3 2 31° ‘ R a 8 8 4:2 3 \ ._ __ .5 Z/sz 3 a Rs " s “f a: o - \‘:$ n w? 32 4 ‘ RR Z632 “3 E a”: R.. 1 W Rm W} 2 — R33? 0 {/3 J R / 0 ;\\\\ :::::':: 7% abcdef (D Fig. 4. Growth of tobacco cells following incubation in different media, measured 24 h after beginning of treatment. Cells were utilized 2 days after subculture. Treatment were (a) cells lefi undisturbed in the culture medium; (b) the culture medium was replaced by a fresh one; (c) and (d) cells were incubated for 1, and 3 h, respectively, in a minimal medium, followed by incubation in fresh culture medium; (e) and (f) cells in culture medium were treated with 10 11M brefeldin A (BFA) continuously, or for 1 h, followed by incubation in fresh culture medium, respectively; (g) cells in culture medium were grown continuously in the presence of 0.1 % methanol, as a control for methanol introduced as a solvent for brefeldin A. Values are reported after subtraction of the initial percentage of packed cell volume (8.1%). Bars represent SE (n=3)_ 127 recent observations on growing green algae where cytosolic [Cay] determinations and 2* gradients injection of Ca2+ buffers also failed to provide evidence for detectable Ca (Holzinger et al. 1995). In growing cells, the Ca” response to medium change was enhanced by brefeldin A treatment (fig. 2). This did not appear to be an artifact since chlorotetracycline fluorescence was not affected by brefeldin A in stationary phase or undisturbed growing cells (fig. 2). Furthermore, the high degree of reversibility of this effect is typical of brefeldin A action on the Golgi apparatus (Klausner et al. 1992), where considerable vesicular traffic, in part under Ca2+ control, originates from. V Although poorly understood, Cay-mediated signalling pathways are assumed to be operative in plant cells (Bush 1995). Progress has been largely hampered by methodology and lack of good model systems. As reported here, the tobacco cell line, BY-2, seems to be a good model system for studies on signalling pathways in plant biology (Nagata et al. 1992). REFERENCES Berridge, M.J. 1995. Calcium signalling and cell proliferation. Bioessays 17: 491-500 Bush, D.S. 1995. Calcium regulation in plant cells and its role in signaling. Annu. Rev. Plant Physiol. Plant Mol. Biol. 46: 95-122 Callaham, D.A., Hepler, PK. 1991. Measurement of free calcium in plant cells. In: McCormack, J .G., Cobbold, P.H. (eds) Cellular Calcium, A Practical Approach. IRL Press, Oxford, pp 383-410 Chant, J. 1994. Cell polarity in yeast.Trends Genetics 10:328-333 Clark, G.B., Roux, SJ. 1995. Annexins of plant cells. Plant Physiol. 109 :1133-1139 Cole, N.B., Lippincott-Schwartz, J. 1995. Organization of organelles and membrane traffic by microtubules. Curr. Opin. Cell Biol. 7:55-64 Herth, W., Reiss, H.D., Hartrnann, E. 1990. Role of calcium ions in tip growth of pollen tubes and moss protonema cells. In: Heath IB (ed) Tip Growth in Plant and Fungal Cells. Academic Press, San Diego CA, pp 91-118 Holzinger, A., Callaham, D.A., Hepler, P.K., Meindl, U. 1995. Free calcium in Micrasterias: local gradients are not detected in growing lobes. Eur. J. Cell Biol. 67:363- 371 Jackson, C.L., Kepes, F. 1994. BFRI, a multicopy suppressor of brefeldin A-induced lethality, is implicated in secretion and nuclear segregation in Saaccharomyces cerevisae. Genetics 137: 423-437 Klausner, R.D., Donaldson, J .G., Lippincott-Schwartz, J. 1992. Brefeldin A: Insights into the control of membrane traffic and organelle structure. J. Cell Biol. 11621071-1080 Merck Index , 11-th edition 1989. Merck Inc, Rahway NJ, Nr 1363 Nagata, T., Nemoto, Y., Hasezawa, S. 1992. Tobacco BY-2 cell line as the “HeLa” cell in the cell biology of higher plants. Int. Rev. Cytol. 132:1-30 128 129 Robinson, M.S., Watts, C., Zerial, M. 1996. Membrane dynamics in endocytosis. Cell 84:13-21 Rothman, J .E., Orci, L. 1990. Movement of proteins through the Golgi stack: a molecular dissection of vesicular transport. F ASEB J. 4: 1460-1468 Schmid, S.L., Damke, H. 1995. Coated vesicles: a diversity of form and function. FASEB J. 921445-1453 '3 Staehelin, L.A., Moore, 1. 1995. The plant golgi apparatus: structure, functional organization and trafficking mechanisms. Annu. Rev. Plant Physiol. Plant Mol. Biol. 46:261-288 Timmers, A.C.J., Reiss, H.D., Schel, J .H.N. 1991. Digitonin-aided loading of fluo-3 into embryogenic plant cells. Cell Calcium 12:515-521 Tsien, KY. 1989. Fluorescent indicators of ion concentrations. Meth. Cell Biol. 30:127-156 Zonia, L., Tupy, J. 1995. Lithium-sensitive calcium activity in the germination of apple (Malus x domestica Borkh.), tobacco (Nicotiana tabacum L.), and potato (Solanum tuberosum L.). J. Exptl. Bot. 46:973-979 APPENDIX B CELLULAR ASPECTS OF ALUMINUM TOXICITY IN PLANTS Haug, A., Vitorello, V. 1996. In: Yasui, M., Ota, K., Strong, M.J., Verity, M.A. (eds). Mineral and metal neurotoxicology. CRC Press, Boca Raton, F L (in press). 130 Chapter 4 Cellular Aspects of Aluminum Toxicity in Plants CONTENTS 4.1 Introduction........... .... - ................................................................... 4.2 Al(III) Chernisuy ............................................................................................. 4.3 Al(III) and Plant Growth ..................................................................... - 4.4 Genetics and Aluminum Tolerance ............................................................................................ 4.5 Molecular Features - .. ........................................................................ 4.5.1 Al(III) Uptake ............................................................................................. 4.5.2 Interiorized Al(III) ......................................................................................................... 4.6 Concluding Remarks .................................................................................................................. Acknowledgment .................................................................................................................................. References .......................... ........................................................................................ 4.1 INTRODUCTION Soil acidity is a major factor limiting plant growth and biomass production. In the heterogeneous soil system the pH of the aqueous phase plays a fundamental role in properties like the solubility and hydrolysis of ions. the electric charges of soil colloids, and the uptake of nutrients by plant roots. As the pH decreases, the solubility of ions like Al(III), Mn(II). and Cu(II) generally increases, eventually reaching levels toxic for many crap plants. Regarding metal toxicity. that caused by aluminum is by far the major factor for plant growth limitation on naturally acidic soils (pH <5.0). For example, 70% of the soils in tropical America are affected by aluminum toxicity. Industrial activities and fertilizer usage have also contributed to the acidification and thus to the levels of free aluminum in terrestrial and aquatic environments previously not afflicted with aluminum toxicity."3 Despite the long history of studies on aluminum-related limitation of plant growth. there still exists a dearth of knowledge concerning the molecular basis for aluminum toxicity and the impact of low pH on plant root biochemistry."3 In this chapter we are therefore focusing our attention primarily on molecular aspects of aluminum toxicity in plant cells rather than on phenomenological descriptions of the anatomy/physiology of plants under aluminum stress. Regarding mechanisms of aluminum toxicity in plant cells. information will also be provided, and conclusions drawn, from appropriate experiments conducted on mammalian cells. We hope that such a strategy will stimulate interest in studies directed towards elucidating primary mechanisms of aluminum toxicity mani- fested in both plant and mammalian cells. 08493-7664-5/96601XkSJO c 1996 by crtc Press Inc. 35 Al(III) IN PLANTS 13 l 36 42 Al(III) CHEMISTRY In soils. aluminum is predominantly found as aluminosilicate and clay minerals. In the environment, chemical reactions of aluminum are mostly occurring in water. For these reasons, and also because the water content of a typical biological cell is approximately 80% of the cellular weight. it is necessary to understand reactions of aluminum in aqueous solution. Concerning details about aluminum properties (thermodynamic data, kinetics. chelation) in minerals and in solution, infor- mative articles are available.” In short, in aqueous solution the charged Al(III) exists as a hexahydrated species. Al(H20)63‘, at low pH. As the pH increases, in particular in the range from 5 to 6.2, the hydrated metal ion has a strong tendency to form various hydrolytic species. A series of polymeric aluminum species is also generated, particularly at higher pH values. Besides the pH value and Al(III) concentration. the extent of hydrolysis is strongly dependent on the presence of chelating ligands. The hydrated Al(III) ion has a considerable aflinity to oxygen-containing negatively charged ligands such as carboxyl, hydroxyl, carbonyl, and phosphate groups associated with organic molecules. Apart from binding to hydroxyl, strong inorganic complexes are formed with fluoride, sulfate. and phosphate ligands. The mode of metal uptake across the plant plasma membrane may depend upon a chelate formed by Al(III). Following interiorization, the movement of Al(III) to biochemically critical targets in the cytosol and/or membranes is presumably also dependent upon the types of ligands available. Participation of Al(III) in chemical reactions is in part governed by the slow exchange rate of hydration water. a consequence of the small ionic radius (0.053 nm) of Al(III). 4.3 Al(III) AND PLANT GROWTH A pH value of 5.0 is the approximate threshold below which aluminum toxicity becomes manifest in whole plants. This threshold is rather ill-defined since it also depends on factors such as the presence of ions Ca(II) and phosphate, chelators, and the plant genotype.L3 In a “typical" acidic soil, the pH is around 4.5 and the concentration of mobile aluminum may vary between 10 and 40 M (1 ppm Al(III) equals 37 1.1M). However, in acidic forest ecosystems displaying forest decline. pH values around 3.8 and mobile aluminum levels of about 1 mM have been reported.. Toxicity has been tentatively ascribed to the hexahydrated, trivalent Al(III) species, rather than to its hydrolysis products with lower electric charge.“3 Al(III)-stressed plants show inhibition of growth. mainly that of the roots; the root apex is apparently a critical tissue.9 Al(III)-related derangements of root metabolism reportedly impact transport of essential nutrients between roots and plant tissues above ground. thus reducing plant biomass accumulation."3 Symptoms on roots include root thickening and discoloration. At the cellular level, perturbations of plasma membrane structure. vacuolation of root cap cells, and changes in nuclear structure are observable.3~9 Since primary target(s) of Al(III) action on cells have not been clearly identified, some or all of these symptoms may reflect secondary injuries to the cell. Regarding growth of Al(III)-stressed plants. considerable attention has been paid to interactions between Al(III) and Ca(II). Not attributable to ionic strength effects, elevated Ca(II) or Mg(II) levels in the root medium have been found to ameliorate the effects of toxic Al(III) on plant growth.1 An atn'active feature of these observations is that an increase in soil acidity is generally accompanied by a loss of exchangeable cations like Ca(II) and Mg(II).‘ Findings derived from these types of amelioration studies led to the hypothesis that toxic Al(III) ions are displacing Ca(Il) from the membrane surface, thus impairing the functioning of surface ligands and transport channels critical for proper root growth."3 Recent experiments on wheat discredited this type of Ca(II)-displacement hypothesis for explaining Al(III)-related rhizotoxicity.lo Rather than unspecifically residing at the cell surface, the initial Al(III)-induced lesion seems to be associated with regulatory processes which are in part dependent upon signal-transducing elements of the plasma membrane, at least in mammalian cells.“-12 Al(III) IN PLANTS 132 37 4.4 GENETICS AND ALUMINUM TOLERANCE Among plant species and genotypes of a given species there exist wide variations regarding tolerance to Al(IIl).‘ In the acidic cerrado region of central Brazil, for example, Al(III)-accumulating tree species have been found which apparently even require Al(III) in the growth medium for normal growth.13 However, the biochemical and genetic control mechanisms underlying tolerance to Al(III) are poorly understood. At this time, hard evidence is lacking whether tolerance to Al(III) is associated with processes outside the cell (exclusion) and/or inside the plant cell (internal detoxi- fication). Given this dearth of knowledge, and findings that rhizotoxicity seems to be a pronounced feature of toxic Al(III) on plants,1 parameters like root length or root elongation of Al(III)-stressed plants are generally selected as indicators when studying the inheritance of Al(III) tolerance in crop plants. Based on these selection techniques, there is evidence that a single incomplete dominant gene is responsible for the inheritance of aluminum tolerance in wheat." Al(III) resistance of a tobacco variant was also attributed to a single dominant gene mutation when tobacco cells were selected for Al(III) resistance in cell cultures.” Furthermore. by selecting biochemical parameters as indicators for Al(III) sensitivity, experiments on two wheat varieties suggested that the effects of Al(III) on K(I) uptake are not primary factors associated with sensitivity towards toxic Al(III)." Despite the paucity of information on the molecular basis for Al(III) resistance, significant progress has been achieved. solely relying on conventional plant breeding programs. in selecting and releasing Al(III)-tolerant plants for crop production on acidic soils.‘ 4.5 MOLECULAR FEATURES 4.5.1 Al(III) Uptake Across The plasma membrane Numerous studies have been conducted concerning Al(III) uptake and distribution in intact plants or in plant tissue!“ However. considerable ambiguity remains concerning the relationships between the measured uptake kinetics and putative Al(III)-binding compartments. Indeed, few data are available on the effects of Al(III) at the cellular and molecular levels, such as mode(s) of Al(III) uptake across the plasma membrane, role of metal speciation in membrane traversal, and initial targets critical for cellular physiology. Outside the plant cell, in the Donnan free space of the root cell, major aluminum-binding sites are negatively charged carboxyl groups of pectic substances."3 However, we believe that these surface-exposed. unspecific binding sites can be dismissed as primary aluminum targets. Rather, likely primary targets for aluminum-triggered rhizotoxicity are expected to reside in the cytosol and/or in the plasma membrane of the plant cell,“ i.e., Al(III) must be internalized by the cell. This notion is also supported by findings on mammalian cells, where key elements of phosphoinositide-associated signal u'ansduction pathways. coupled with Ca(II) regulation, were implicated as potential primary targets for toxic Al(III) ions.“-‘2 In oak root cortex cells, application of Al(III) (pH 4; 37 11M to 3.7 mM) enhanced membrane permeability to nonelectrolytes (methyl urea. urea), and lowered permeability to water. '9 The plasma membrane was also implicated as a putative Al(III) target in studies on Amaranthus protoplasts, where “Ca“ uptake was inhibited within minutes by Al(III) application (pH 4.5; 10 and 50 1.11%)” These kinds of permeability changes are presumably related to pronounced Al(III)-triggered mem- brane lipid phase changes, in vivo and in membrane vesicles isolated from Thermoplasma acido- philum (pH 2).2 Moreover. compared with nonstressed mycelia, the proportion of less-ordered membrane domains was decreased when spin-labeled fungal mycelia were exposed to Al(III) stress (1 to 10 mM).21 Al(III)-induced changes in lipid packing were also observed in response to Al(III)- binding (20 M) to the erythrocyte membrane.22 The phosphodiester group was found to be the primary binding site of Al(III) with lipids.23 Furthermore, based on findings that Al(III) neutralizes the surface charge of phosphatidyl choline vesicles. this type of interaction was implicated in the nonspecific. Al(III)-related inhibition of cation uptake by root cells.“ In the presence of Al(III) (20 to 50 1.1M), changes in the membrane surface charge are generated by displacing virtually all bivalent as 133 cations, such as Ca(II), present at bulk concentrations as high as 1 mM.15 To reiterate, Al(III)- related Ca(II) displacement at the membrane surface does not seem to be the initial lesion for rhizotoxicity of plants.” Nevertheless, it is reasonable to assume that nonspecific Al(III) binding to the membrane surface. followed by traversal of the membrane, are sequential events required for Al(III) interiorization by the cell. Membrane traversal, perhaps involving some sort of carrier?‘5 may in part depend on the physical state of the membrane. 4.5.2 Interiorized Al(III) Once interioriwd, Al(III) may interact with membrane-bound, cytosolic, and nuclear constituents of the cell. Given the slow solvent exchange rate of Al(III)] it must be kept in mind, however, that initially established interactions with certain sites may gradually change until a final equilibrium has been reached where Al(III) is irreversibly coordinated to a ligand, perhaps different from the initial one. Forming the interface between the cell and its microenvironment. the plasma membrane detects external stimuli at the surface, processes these signals across the membrane, then converts them into a cascade of reactions involving second messengers like Ca(II). In animal cells, chief among cellular communication channels is the phosphoinositide signaling pathway. A major product of this pathway is a second messenger, inositol-1.4,S-triphosphate (1P3) which, in turn, plays a crucial role in mobilizing intracellular Ca(II).” In plants, the elements of this pathway are present but evidence is lacking as to the Operation of this system in response to a specific stimulus.” Regarding aluminum toxicity, short-terrn studies on neuroblastoma cells demonstrated that application of Al(III) (pH 6.8, 2 W to 1 mM) inhibited intracellular Ca(II) mobilization which, in turn. coincided with a decrease in inositol phosphate production. Being crucial for 1?, formation. phosphatidyli- nositol-4,5-diphosphate-specific phospholipase C probably harbors an Al(III) target. A second Al(III) target seems to reside on molecular switching elements of the phosphoinositide signaling pathway, viz.. on Mg(II)/guanine nucleotide binding (Gp) proteins. It was suggested that Al(III) is occupying a site at or near the GTP binding center on the Gp protein."-12 Besides direct interactions with proteins, Al(III) may perturb the protein's lipid milieu, especially at higher.Al(III) For example. the hydrolyzing efficacy of phospholipases is in part dependent on the physical state of their lipid environment.‘2 Evidently further studies are necessary to identify Al(III) targets on elements of the signal transduction pathway in eukaryotic cells. Concerning Al(III) interactions with membrane-bound enzymes. studies on barley root plasma membranes indicated that Al(III) ions are capable of inhibiting a membrane-associated Ca(II)- ATPase activity. The Al(III)-induced lower activity could in part be restored by applying massive amounts (50 mM) of monovalent cations. As a result of increased ionic strength. Al(III) was presumably displaced from lipids in the boundary lipid region of the enzyme, thus restoring the motional requirements of the intact protein.29 There is also the possibility that ATPase activity is diminished resulting from the formation of an Al(III)-ATP chelate when millimolar AlCl, levels are employed.30 Furthermore. applying Al(III) to vesicles enriched in right-side-out plasma mem- branes isolated from maize roots, Al(III) (pH 6.0; 40 11M) inhibited NADH-dependent electron transfer processes," i.e., reactions generally associated with the uptake of iron by the plant root.32 Following internalization by the wheat root cell, Al(III) was distributed in the cytosolic fraction (generally 55% of total Al[III]), nuclear fraction (around 30%), and the mitochondrial fraction (around 13%).33 Interestingly, a roughly similar distribution of Al(III) was found in neuroblastoma cells."5 In the cytosol there exists a multitude of ligands capable of binding Al(III), e.g., ATP,30 citric acid,“-7 various organic acids.’ and acidic polypeptides.2 However, evidence is lacking whether intracellular chelators are instrumental in intrinsic Al(III) tolerance mechanisms of plant cells. On the other hand, like in mammalian cells."5 this large chelator pool may contribute to the high and rapid accumulation of Al(III) by plant cells. By maintaining a low intracellular-free [Al(III)] the equilibrium is shifted in favor of Al(III) influx. Al(III) 1N PLANTS 134 39 Al(III) ions reportedly also interact with mitochondrial membranes. Experiments on isolated mitochondria from untreated wheat roots indicated that Al(III) ions (pH 7; 12 to 75 11M) interacted with the mitochondrial respiratory pathway, presumably by interfering with the oxidation of sub- strates that donate electrons to Complexes I and 11.34 Also involved in energy-generating processes are NADP-dependent dehydrogenases. which were strongly inhibited by Al(III) (pH 6.85; 2 11M A1C13) with respect to the substrate. isocitrate.” In the cytosol, interiorized Al(III) may bind to calmodulin.2 a highly conserved protein (about 17 kDa) crucial for Ca(II)-mediated regulation in eukaryotic cells. Responding to altered intracel- lular Ca(II) levels. e.g., triggered via signaling pathways, this protein is activated by undergoing conformational changes generated by specifically binding four Ca(II) per protein. The activated protein can then associate with and stimulate certain calmodulin-dependent partner enzymes." Concerning Al(III)-related alterations of calmodulin, only in vitro experiments are available. Pre- sumably binding only weakly to calmodulin,” the protein undergoes helix-coil transitions in the presence of Al(III), thus changing the protein's internal dynamics and its efficacy in correctly docking with partner proteins.3“‘° Al(III)-induced changes on calmodulin are instrumental in malfunctions of calmodulin-dependent reactions, e.g., that of 3’.5’-cyclic nucleotide phosphodi- esterase, certain plant AT'Pases,2 and NAD kinase activity in wheat root tips.“ The latter region harbors high levels of calmodulin.‘2 and has been implicated as being a site for Al(III) injury? At this time evidence is lacking whether Al(III) inactivates calmodulin in a cell. Abundantly present, cytosolic chelators are possibly protecting this important regulatory protein from Al(III) injury.‘3 The nucleus may be an important sink for interiorized Al(III), especially in cells under long- term Al(III) stress. For example, when onion root tips were treated with AlCl, (pH 5.5, 1 mM, several hours), cell division was inhibited.“ Various findings indicate that Al(III) accumulates on DNA,“ where the phosphodiester group is apparently a binding Site.45 Under cytosolic conditions, Al(III) binding to the DNA is presumably weak.“5 Given the condensed state of DNA in the nucleus of a cell, interstrand DNA cross-linking via Al(III) bridges is highly probable."I Such cross-linking would be consistent with observations that Al(III)-treated pea chromatin is more heat stable.“ Moreover, Al(III)-induced changes in DNA structure might be expected to disrupt the physiolog- ical processes of gene expression. It is noteworthy that the mRNA levels of calmodulin and tubulin were considerably diminished in Al(III)-treated rabbit brains.“ Compared with Al(III)-induced derangements of membrane structure and signaling processes (detectable within minutes),”-‘3 Al(III) interactions at the DNA, and numerous toxicity responses (growth, cell secretion, callose formation, etc.) are long-term effects.‘-3~3-‘9 Whether some of these long-term responses are directly caused by Al(III) is questionable. 4.6 CONCLUDING REMARKS Only rudimentary information is currently available on primary targets and the initial phases in the temporal progression of aluminum toxicity, both in plant and animal cells. What. at present. appears to be a multitude of Al(III) effects in cells, may eventually originate from a single (or a few) primary target(s) critical for the temporal development of the aluminum toxicity syndrome. Given the significance of an intact plasma membrane and of functional communication pathways for cellular survival, any Al(III)-uiggered change therein is expected to have severe repercussions on the cell's ability to respond to extemal/internal signals. By focusing on identifying initial key interaction sites for toxic Al(III), we may also contribute to. and in fact simplify, our endeavors of deve10ping Al(III)-resistant varieties through biotechnological means. ACKNOWLEDGMENT V.V. was supported in part by a fellowship, grant Nr. 200892/91-6, from the Conselho Nacional de Desenvolvimento Cientifico e Tecnologico (Brasil). 40 BS REFERENCES I. la) sev- IO. 11. 12. 13. 14. 16. I7. 18. 19. 20. 21. 22. 23. 24. 25. 26. 27. 28. 29. 30. 31. Foy CD. Chaney RL. and White MC: The physiology of metal toxicity in plants. Annu. Rev. Plant Physiol.. 1978; 29: 511-566. . Haug A: Molecular aspects of aluminum toxicity. CRC Crit. Rev. Plant Sci.. 1984; 1: 345-373. . Taylor GJ: The physiology of aluminum phytotoxicity. In Sigel I-I (ed): Metal Ions in Biological Systems. Vol. 24. New York, Marcel Dekker. 1988. pp 123-163. . Sposito G (ed): The Environmental Chemistry of Aluminum. Boca Raton, CRC Press. 1989. Akitt JW: Multinuclcar studies of aluminum compounds. Prog. Nucl. Magn. Reson. Spectrosc. 1989: 21: 1-149. Martin RB: The chemistry of aluminum as related to biology and medicine. Clin. Chem. 1986: 32: 1797-1806. Orvig C: The aqueous coordination chemistry of aluminum. In Robinson GH (ed): Coordination Chemistry of Aluminum. New York. VCH Publishers. 1993. pp 85-121. . Godbold Dl.. Fritz E. and Hilttermann A: Aluminum toxicity and forest decline. Proc. Natl. Acad Sci. U.S.A.. I988; 85: 3888-3892. . Bennet RJ and Breen CM: Aluminium toxicity: Towards an understanding of how plant roots react to the physical environment. In Randall PJ. Delhaize E. Richards RA. and Munns R (eds): Genetic Aspects of Plant Mineral Nutrition. Dordrecht. Kluwer Academic. 1993. pp 103-116. Kinraide TB. Ryan PR. and Kochian LV: AIP-Ca" interactions in aluminum rhizotoxicity. 11. Evaluating the Ca"- displacement hypothesis. Planta. I994; 192: 104-109. ' Shi B, Chou K. and Haug A: Aluminium impacts elements of the phosphoinositide signalling pathway in neuroblas- toma cells. Mol. Cell. Biochem.. 1993: 121: 109-118. Haug A. Shi B. and Vitorello V: Aluminum interaction with phosphoinositide-associated signal transduction. Arch. Toxicol. 1994; 68: 1-7. Haridasan M. and de Araujo G M: Alurninium-accurnulating species in two forest communities in the cerrado region of central Brazil. Forest Ecol. Manage. 1988; 24: 15-26. Wheeler DM. Edrneades DC, Christie RA. and Gardner R: Comparison of techniques for determining the effect of aluminium on the growth of. and the inheritance of aluminium tolerance in wheat. In Randall PJ. Delhaize E. Richards RA. and Munns R (eds): Generic Aspects of Plant Mineral Nutrition. Dordrecht. Kluwer Academic. 1993. pp 9-16. . Conner AJ. and Meredith CP: Strategies for the selection and characterization of aluminum-resistant variants from cell cultures of Nicotiana plumbaginifolia. Planta, 1985; 166: 466-473. Pettersson S, and Strid H: Effects of aluminium on growth and kinetics of K’('°Rb’) uptake in two cultivars of wheat with different sensitivity to aluminium. Physiol. Plant. 1989; 76: 255-261. Zhang G. and Taylor GJ: Effects of biological inhibitors on kinetics of aluminium uptake by excised roots and purified cell wall material of aluminium-tolerant and aluminium-sensitive cultivars of Triticum aesri'vum L. J. Plant Physiol.. 1991: 138: 533-539. Haug A. and Shi B: Biochemical basis of aluminium tolerance in plant cells. In Wright RJ. Baligar VC. and Murrmann RP (eds): Plant-soil interactions at low pH. Dordrecht. Kluwer Academic. 1991. pp 839-850. Zhao XJ. Sucoff E. and Stadelmann El: Al" and Ca” alteration of membrane permeability of Quercus rubra root cortical cells. Plant Physiol., 1987; 83: 159-162. Haug and Vitorellols Rengel Z. and Elliott DC: Aluminium inhibits "Ca” uptake by Amaranthus protOpIasts. Biochem. Physiol. Pflanz. 1992: 188: 177-186 Zel J. Svetek J. Cine H. and Schara M: Effects of aluminum on membrane fluidity of the mycorrhizal fungus Amanita mascaria. Physiol. Plant. 1992: 89: 172-176. Weis C. and Haug A: Aluminum-altered membrane dynamics in human red blood cell white ghosts. Thromb. Res. 1989: 54: 141-149. Hauser H. and Phillips MC: Interactions of the polar groups of phospholipid bilayer membranes. In Cadenhead DA. and Danielli JF (eds): Progress in Surface and Membrane Science. Vol. 13. New York. Academic Press. 1979. pp 297-413. Akeson MA. Munns DN. and Burau RG: Adsorption of A13’to phosphatidylcholine vesicles. Biochim. Biophys. Acta. 1989: 986: 33—40. Deleers M: Cationic atmosphere and cation competition binding at negatively charged membranes: Pathological implications of aluminum. Res. Commun. Chem. Pathol. Pharmacol. 1985; 49: 277-294. Shi B. and Haug A: Aluminum uptake by neuroblastoma cells. J. Neurochem.. 1990: 55: 551-558. Berridge MJ: Inositol triphosphate and calcium signalling. Nature, 1993: 361: 315-325. Cote' GG. and Crain RC: Biochemistry of phosphoinositidcs. Annu. Rev. Plant Physiol.. 1993: 44: 333-356. Caldwell CR. and Haug A: Divalent cation inhibition of barley root plasma membrane-bound Gait-ATPase activity and its reversal by monovalent cations. Physiol. Plant. 1982: 54: 112-118. Panchalingam K. Sachedina S. Pettegrew JW. and Glonek T: Al-ATP as an intracellular carrier of Al(III) ion. Int. J. Biochem.. 1991; 23: 1453-1469. Loper M. Brauer D. Patterson D. and Tu 81: Aluminum inhibition of NADI-l-linkcd electron transfer by com root plasma membranes. J. Plant Nutr.. 1993: 16: 507-514. Al(III) IN PLANTS 136 41 32 33. 34. 35. 36. 37. 38. 39. 40. 41. 42. 43. 44. 45. 46. 47. 48. 49. Holden Ml, Luster DO. and Chaney RL: Enzymatic iron reduction at the root plasma membrane: Partial purification of the NADI-l-Fe chelate reductase. 1n Manthey JA. Crowley DE. and Luster DG (eds): Biochemistry of Metal Micronutrients in the Rhizosphene. Boca Raton, Lewis Publishers. 1994. pp 285-294. Nieda'ela G. and Aniol A: Subcellular distribution of aluminium in wheat roots. Acta Biochim. Pol.. 1983: 30: 99-105. de Lima ML. and Copeland L: The efiect of aluminium on respiration of wheat roots. Physiol. Plant. 1994; 90: 51-58. Yoshino M. Yamada Y. and Murakami K: Inhibition by aluminum ion of NAD- and NADP-dependent isocitrate dehydrogenases from yeast. Int. J. Biochem. 1992: 24: 1615—1618. Cohen P. and Klee CB (eds): Calmodulin. Amsterdam. Elsevier. 1988. Martin RB: Bioinorganic chemistry of aluminum. In Sigel H. and Sigel A (eds): Metal Ions in Biological Systems. Vol 24. New York. Marcel Dekker. 1988. pp 1-57. Yuan S. and Haug A: Frictional resistance to motions of bimane-labelled spinach calmodulin in response to ligand binding. FEBS Lett.. 1988: 234: 218-223. Weis C. and Haug A: Aluminum-altered conformational changes in calmodulin alter the dynamics of interaction with rnelittin. Arch. Biochem. Biophys.. 1987: 254: 304-312. l-Iaug A. and Vitorello V: Aluminium coordination to calmodulin. thermodynamic and kinetic aspects. Coord Chem. Rev.. in press. Slaski II: Effect of aluminium on calmodulin-dependent and calmodulin-independent NAD kinase activity in wheat root tips. J. Plant Physiol.. 1989: 133: 696-701. Allan E. and Trewavas A: Quantitative changes in calmodulin and NAD kinase during early cell develoth in the root apex of Pisum sativum L Planta. 1985: 165: 493-501. Suhayda CG. and I-Iaug A: Citrate chelation as a potential mechanism against aluminum toxicity in cells: the role of calmodulin. Can J. Biochem. Cell Biol. 1985: 63: 1167-1175. Matsurnoto 1-1: Biochemical mechanism of the toxicity of aluminium and the sequestration of aluminium in plant cells. In Wright RJ. Baligar VC. and Murrrnann RP (eds): PlantoSoil Interactions at Low pH. Dordrecht. Kluwer Academic. 1991. pp 825-838. ' Dyrssen D. Haraldsson C. Nyberg E. and Wedborg M: Complexation of aluminum with DNA. J. Inorg. Biochem.. " 1987; 29: 67-75. Martin RB: Aluminium speciation in biology. 1n Aluminium in Biology and Medicine (Ciba Foundation Symposium no. 169). Chichester. John Wiley 8: Sons. 1992. pp 5-25. . Karlik SJ. and Eichhom GL: Polynucleotide crosslinking by aluminum. J. Inorg. Biochem.. 1989: 37: 259—269. Crapper Mclachlan DR: Aluminum neurotoxicity: Criteria for ”signing a role in Alzheimer's disease. In Lewis TE (ed): Environmental Chemistry and Toxicology of Aluminum. Boca Raton, Lewis Publishers. 1989. pp 299-315. Basu A. Basu U. and Taylor GI: Induction of microsornal membrane proteins in roors of an alurninumoresistant cultivar of Triticum aestivum 1.. under conditions of aluminum stress. Plant Physiol.. 1994: 104: 1007-1013. APPENDIX C ALUMINUM INTERACTION WITH PHOSPHOINOSITIDE-ASSOCIATED SIGNAL TRANSDUCTION Haug, A., Shi, B., Vitorello, V. 1994. Arch. Toxicol. 68:1-7. Arch Toxicol (1994) 68: l-7 Review article 137 fihcology © Springer-Verlag 1994 Aluminum interaction with phosphoinositide-associated signal transduction Alfred Haug, Biao Shi, and Victor Vitorello Department of Microbiology and Pesticide Research Center. Michigan State University. East Lansing. MI 48824. USA Received 14 June l993/Accepted 15 September 1993 Abstract. Concerning molecular and cellular mechanisms of aluminum toxicity. recent studies support the hypothesis that interactions of aluminum ions with elements of signal transduction pathways are apparently primary events in cells. In the case of the phosphoinositide-associated sig- nalling pathway of neuroblastoma cells. guanine nucleo- tide-binding proteins (G proteins) and a phosphatidylino- sitol-4.5-diphosphate (PIP2)-specific phospholipase C are probable interaction sites for inhibitory actions of alumi- num ions. Following interiorization of aluminum by the cell. metal interactions decrease the accumulation of ino- sitol phosphates. especially that of inositol-1.4.5-triphos- phate (1P3). concomitant with derangements of intracellular Ca1+ homeostasis. In the presence of high concentrations of Calt. formation of 1?: is also diminished in aluminum- pretreated cells. presumably involving a process not re- quiring Mgzt-dependent G proteins. At higher aluminum doses. metal-induced changes in the lipid milieu of the membrane-bound phospholipase may play a role. These types of primary interactions of aluminum ions with ele- ments of cellular communication channels are probably crucial in the manifestation of the multifacetted aluminum toxicity syndrome. If present as a phosphate-like fluoro- aluminate. a stimulatory role of aluminum ions is displayed in G protein-coupled transmembrane signalling. Key words: Aluminum toxicity - Calcium regulation - G proteins - Neuroblastoma - Phosphoinositides - Phos- pholipase C - Signal transduction Introduction Over the past decade. concern has intensified regarding the toxic effects of aluminum on man (Ganrot 1986; Marquis 1989; Schmidt et al. 1991). In the environment. aluminum is recognized as a serious global problem since vast areas of Correspondence to: A. Haug the world suffer from natural soil acidity which in turn mobilizes aluminum for uptake by crop plants. Moreover. industrial development and usage of fertilizers have con- . tributed to the acidification of agricultural areas and aquatic habitats. Aluminum can therefore enter the food chain (Haug 1984; Lewis 1989; Taylor 1991). Although considerable evidence has emerged regarding clinical aspects (Wills and Savory 1985: Nebelter and Co- burn 1986). only scant information is available on specific cellular and molecular mechanisms associated with alu- minum toxicity. Nevertheless. there are indications that aluminum interferes with signalling pathways in cells (Woods et al. 1990: Miller et al. 1989; Schofl et al. 1990: Shi and l-Iaug 1992). Indeed. the hypothesis can be formulated that elements of signalling pathways. like that associated with phos- phoinositides, harbor a primary target for aluminum inter- action(s) with a cell. Depending on conditions. these in- teractions may either stimulate or inhibit the signalling pathway. It is therefore our objective to provide a critical review of cellular and molecular aspects pertinent to the above hypothesis. To set the stage for such a focussed discussion. we first summarize some background infon'na- tion on the biochemistry of the interrelated Ca2+lphos- phoinositide signalling pathway. the best-known commu— nication channel. Signal transduction along the Gaze/phosphoinositide pathway Cells require communication with their environment. By carrying messages along signalling pathways. input signals are transformed into specific biological responses. such as changes in intracellular Ca2+ concentrations (Rasmussen 1990). In eukaryoes. chief among cellular communication channels is the phosphoinositide signalling pathway (Fig. 1). Following binding of a primary messenger (hor- mone. neurotransmitter) to a receptor at the cell surface. second messengers. viz. inositol-1.4.5otriphosphate (1P3) and diacylglycerol. are produced through hydrolysis of 138 L) r. 0 W extracellular "W matrix lrocetitt'ir "-r ~ ' ". ~r wk w t“ ._. a "2 DAG - " ’ . ' w L. f‘ . 1 — Q‘ a- "1'9 I a i ... ~ ‘ Osman ac: ‘ g . (’3’ ... _ -- an . .. :25: i ' .W -..Jxa. '. J .....I-a «L... ---...1. ' I I :23: . 3%?1 ; El ...=~ @ al,—.e- £25: . v' a i ’3 'P P 751’.- l ;~::": r 2. f‘ 2:- \ “m, I“ .25.“. I retictm 9‘3: : cellular .-- . Fig. 1. Simplified view of signal transduction along the Cab/phos- phoinositide pathway. The shaded area indicates signal transduction elements which are presumably impacted by aluminum ions. Signal input occurs by binding of a primary messenger to a receptor at the plasma membrane surface. thus triggering transmission of a signal to a heterotrimeric G, protein. The Ca subunit dissociates from the Gay subunits in exchange of GDP for GTP. The Go subunit activates a phosphoinositide (PIP2)-specific phospholipase C (PLC). thus produ- membrane-associated phosphatidylinositol-4.5-diphosphate (PIPz) by a specific phospholipase C (Fig. l). 11’: plays a key role in mobilizing intracellular Ca2+ levels (Berridge 1993). while diacylglycerol activates protein kinase C which in turn is involved in protein phosphorylation pro- cesses. Following an array of reactions. 11’: and dia- cylglycerol are metabolized and PIP: is resynthesized. thus closing the phosphatidylinositol cycle (Berridge 1987; Rana and Hokin I990: Majerus l992). Heterotrimeric Gp proteins (M around 100 kDa) serve as intermediaries (Fig. 1) that link cell surface receptors to phospholipase C (Bimbaumer et al. 1990: Hall 1992). In response to the activation of an appropriate receptor. the Gp protein undergoes a conformational change which facil- itates the Mgzt-dependent exchange of bound GDP for GTP. GTP binding to the alpha subunit reduces the affinity to the receptor. Following dissociation from the beta/ gamma subunits. the Mgzhactivated alpha subunit (Fig. I) interacts with the specific effector. phospholipase C. Signal transduction is terminated by the hydrolysis of bound GTP through the GTPase activity of the alpha subunit. Sub- sequent reassociation of this subunit with the beta/gamma subunits leads to reconstitution of the G, protein (Freiss- muth et a1. 1989). Fluoroaluminate-Induced phosphoinositide metabolism and Cu2t mobilization Concerning signal transduction. fluoroaluminates have be- come valuable tools in elucidating functional aspects of signalling pathways. For example. in vitro experiments cing two second messengers as signal output. viz. 11’: and diacylglyo cerol (DAG). The off-phase of signal transmission is initiated by GTP hydrolysis allowing reassembly of G, subunits. Through a series of reactions. e.g. dephosphorylations leading to IP: and 1P. [Pi and diacylglycerol are metabolized and PIP: is resynmesized via PI and PIP (Berridge I987). PKC. protein kinase C: protein-P. phosphorylated protein demonstrated that application of millimolar fluoride. in the form of NaF. activated regulatory uimeric G proteins. which in turn are known to participate in stimulating G protein-coupled transmembrane signalling. This activation process has been shown to be dependent on the presence of trace amounts of aluminum. leading to the formation of fluoroaluminates (Sternweis and Gilman 1982). Indeed. under ordinary experimental conditions. aluminum con- tamination from chemicals and glassware was sufficient for this fluoride-related activation. Even when precautions against contamination were taken. aluminum levels close to 1 M were found in neuroblastoma tissue culture cells (Shi and Haug 1990). The fluoroaluminate species which activates G proteins is not known unambiguously. With regard to thermo- dynamic equilibria. aluminum apparently formed soluble complexes A1133“. where x = 1-6. in aqueous solutions (Savchenko and Tananaev 1959). G proteins were maxi- mally activated in the presence of millimolar NaF con- centrations. Under these conditions. the resulting major species. AIFI. was proposed to act as a pseudophosphate by binding to the G protein adjacent to GDP. in a tetrahedral configuration. However. unlike the gamma-phosphate of GTP. this fluoroaluminate bound to GDP does not undergo hydrolysis and therefore the catalytic center of the protein is locked into the activated configuration (Chabre 1990). In support of the notion that certain fluoroaluminate species behave as a pseudophosphate are studies showing that fluoroaluminates mimicked phosphate binding to the Cazr-free conformation of sarcoplasmic reticulum ATPase. In the presence of Mg“. phosphorylation by phosphate (Pi) was inhibited as a result of AlFI binding. Fluoroaluminate also acted as an analogue of the gamma-phosphate of ATP in an ADP/AIFI complex of the Ca2t-bound form of the enzyme. Bound on the external high affinity sites. the Ca3+ ions became occluded upon binding of an ADP/fluoro- aluminate complex (T roullier et al. 1992). Observed cellular responses to fluoroaluminates are in accord with the action that these compounds are interacting with G proteins. Treatment of hepatocytes with NaF mi- micked the effects of Cab-mobilizing hormones. Cytosolic free Ca2+ levels rose as a result of G protein-mediated PIP: hydrolysis and release of 1P3 (Blackmore et al. 1985). Following treatment of rat parotid acinarcells with milli- molar concentrations of NaF. the observed amylase release was mediated by phosphoinositide degradation and in- tracellular Ca2+ mobilization. even in the absence of ex- tracellular Cazt Sustained increases in intracellular Ca1+ levels probably resulted from enhanced entry of extra- cellular Cazt. Trace amounts of aluminum probably suf- ficed to form the active chelate AIFI (Tojyo et al. 1991). Moreover. by stimulating human umbilical vein endothelial cells with fluoroaluminates. inositol phosphates were gen- erated in a dose- and time-dependent manner. arachidonic acid was released. and prostacyclin was produced. The arachidonic acid release was dependent on intracellular Ca?t stores. while prostacyclin formation was in part re- lated to protein kinase C activity (Magnusson et al. 1989). In rat cortical slices. NaF (2-20 mM) application stimu- lated phosphoinositide hydrolysis. rapidly producing IP as the major product within a few min (lope I988). Em- ploying murine keratinocytes. fluoroaluminates (rnicro- molar AlClalmillimolar NaF) stimulated total inositol phosphate formation. in the presence of 0.05 mM Ca” in the medium. The enhanced phosphoinositide metabolism was. however. not sufficient to induce cellular differentia- tion (Lee and Yuspa 1991). On the other hand. in osteo- blasts low concentrations of aluminum (10-25 M Alz(SOa)3. or AlCls. or Al(N03)3) stimulated DNA syn- thesis and alkaline phosphatase activity through a me- chanism apparently not involving fluoride (SO—200 itM NaF) (Lau et al. 1991). At these fluoride concentrations one would not expect formau'on of phosphate-like fluoro- alurninates. Various cells have been reported to exhibit oscillations in cytosolic Ca2+ concentrations periodically released from intracellular Ca“ reservoirs (Berridge 1993). For example. in single rat hepatocytes such Cazt oscillations could be induced by cell surface receptor agonists. However. when generated by direct activation of G proteins with fluoro- aluminates (1 uM AlClsll mM NaF). these oscillations became less regular in both frequency and temporal de- pendence compared with agonist-induced ones. The am- plitudes of the oscillations ranged from a resting value of about 200 nM to maximally 600 nM of intracellular Ca2+ levels. which were monitored with rnicroinjected aequorin. The irregularity in the fluoroaluminate-induced. receptor- independent oscillations was attributed to a modulatory role in the transient generation by the agonist-receptor complex (Woods et al. 1990). The notion that all effects of fluoroaluminates can be ascribed to the formation of a pseudo-phosphate in the 139 3 GDP-G protein complex was thrown into doubt by ex- periments on the bacterial peptide chain elongation factor EF-Tu (Kraal et al. 1990). a member of the family of G proteins (Bosch et al. 1989). However. taking into account the three-dimensional location of the three consensus ele- ments involved in guanine nucleotide binding (La Cour et al. 1985: Dever et al. 1987) and data from NMR experi- ments (Hazlett et al. 1990). it appears that the guanine binding sites of ET-Tu and other G proteins are structurally non-equivalent. At higher concentrations. the presence of aluminum in cells generally leads to inhibition rather than stimulation of biochemical processes. Using fluoride-stimulated (5—10 mM NaF) murine neuroblastoma cells. A103 (50 uM to 1 mM) reduced inositol phosphate formation in a dose- dependent manner. Aluminum apparently did not interfere with subsequent kinase or phosphatase reactions. down- stream of 1?; production. The primary aluminum action rather occurred at an upstream site(s) responsible for 1?; production. At low aluminum doses (about 0.2 uM to 50 uM). the fluoride-stimulated inositol phosphate forma- tion was virtually unaffected (Shi et al. 1993). As opposed to heterotrimeric G proteins discussed above. there also exists a second family of smaller (20—25 kDa). monomeric G proteins (Downward 1990). These smaller G proteins cannot be activated by aluminum. fluoride. and Mg2+ (Kahn 1991). In the preceding paragraphs fluoroaluminates were in- troduced as valuable tools in studying signal transduction. To generate biochemically active. i.e. phosphate-like. spe- cies of fluoroaluminates. typically millimolar NaF is ap- plied-to cellular suspensions or enzyme solutions. Given these evidently high doses of NaF. the question has to be raised regarding the toxicological significance of fluoride in aluminum toxicology. The normal levels of fluoride in soft tissues varied be- tween25and53uM(l ppm=53uMF).whereasinteeth and bones concentrations typically ranged from 5 to about 16 mM (Clarkson 1991). In the blood plasma of healthy patients. baseline aluminum levels varied from about 0.2 to 0.4 uM; levels above 3 M were considered indicative of increased aluminum burden (Van der Voet et al. 1991). Employing laser microprobe mass analysis. in- tracytOplasmic aluminum concentrations of about 1 mM were found in tangle-bearing neurons of patients afflicted with amyotrophic sclerosis (Perl and Good 1987). Given the above micromolar fluoride levels found in soft tissues. it is thus fairly improbable that phosphate-like fluoro- aluminates are being formed. This conclusion is based on known dissociation constants of complex fluoroaluminates in aqueous solution (Savchenko and Tananaev 1959) as- suming the existence of thermodynamic equilibria. To the best of our knowledge. there exists no information re- garding fluoroaluminate speciation taking into account ki- netic equilibria. i.e. conditions prevalent in a biological microenvironment. When millimolar levels of NaF are applied to cells. there is the possibility that NaF elicits responses in- dependent of those triggered by the simultaneously formed. phosphate-like AIFI. This was demonstrated on rabbit fe- moral arteries where aluminum fiuoride stimulated phos- 140 4 pholipase C whereas NaF (10 mM NaF plus deferoxamine to chelate contaminating aluminum) apparently stimulated calcium channels (Ruth and Blackmore 1990). A dual effect of fluoride was also observed on phosphoinositide meta- bolism in rat brain cortex (Claro et al. 1990). Interactions of aluminum with elements of the phosphoinositide-associated signalling pathway Concerning direct effects of aluminum ions on elements of the phosphoinositide-linked signalling pathway. doses of AlCl3 (l—lOO uM) inhibited PIP: hydrolysis by bovine heart phospholipase C activity (McDonald and Mamrack 1988). and also the release of total inositol phosphates in rat cortical slices stimulated by carbachol (Johnson and lope 1986). or fluoride (lope 1988). Applying patch-clamp techniques to single internally perfused pancreatic acinar cells. pulsatile Cab-dependent Cl- currents. normally evoked by acetylcholine. disappeared when 1 mM AlCh was present in the perfusion solution (Wakui et al. 1990). Furthermore. in hormone-stimulated hepatocytes. A103 (10 M) reportedly perturbed oscillatory phosphoinositide- mediated Ca2+ signalling (Schofl et al. 1990). In intact neuroblastoma cells. the bradykinin-triggered 1P3 release and the concomitant change in intracellular Ca2+ levels were appreciably less in alurninum-pretreated cells. compared with corresponding values in the control group (Shi and Haug l992). Employing digitonin-permeabilized murine neuroblastoma cells. stimulated with nonhydrolyz- able guanosine-5'-(gammathio) triphosphate (GTPISl). ap plication of various aluminum doses (2 11M to 1 mM) considerably interfered with the formation of IE. IPz. and IP. Competition experiments (5 mM GTP. 5 mM Mg“. 50 uM AlCls) showed that the aluminum binding site on the Gp protein was probably different from that for GTP (Shi and Haug 1992). Furthermore. at enhanced in- tracellular free CaZt concentrations a Gilt-mediated path- way seemed to become Operative. bypassing the Mgzthp- mediated pathway. In the Cal's-mediated pathway. appli- cation of 50 itM aluminum mainly diminished the release of 1P3. Following interiorization (Shi and Haug 1990). aluminum therefore seems to target two elements crucial for PIP: hydrolysis. viz. the Gp protein and a phosphoino- sitide-specific phospholipase C (Fig. l) (Shi et al. 1993). Rather than binding directly to proteins. e.g. phospho- lipase C. the possibility cannot be excluded that aluminum perturbed the enzyme‘s lipid microenvironment (Pan- chalingam et al. 1991) because the hydrolyzing efficacy of phospholipases was reportedly dependent on the physico- chemical pr0perties of their lipid milieu (Dennis 1983). Furthermore. these types of lipid-mediated effects may be also transmitted to G proteins since certain specific lipids. e.g. is0prenoids. were found to be covalently attached to G proteins (Casey 1992). To illustrate. on measuring GTP- induced inositol phosphate formation in digitonin-permea- bilized neuroblaStoma cells exposed to increasing AlC13 doses (about 0 1.1M to 1000 uM). a biphasic inositol pro- duction was observed. The data may in part be interpreted in terms of aluminum-related changes in the phospholipase lipid environment at aluminum doses higher than 50 M (Shi et al. 1993). This notion is consistent with observations that fluorescent hydrophobic membrane probes exhibited constraints in motional parameters when human eryth- rocyte white ghosts were exposed to 20 uM AlC13 (Weis and Haug 1989). Since in vitro experiments had indicated that aluminum induces structural changes in calmodulin (Yuan and Haug 1988). aluminum may also impair Calt- calmodulin-dependent phospholipase C activity (Levine et al. 1990), and other calmodulin-dependent enzymes. Sub- units of certain G proteins are known to inhibit calmodulin- sensitive adenylate cyclase (Mangels et al. 1992). In this context it is noteworthy that aluminum ions. administered as AlCls. reportedly bound to calpains. i.e. non-lysosomal intracellular cysteine proteinases which harbor calmodulin- like CaZt-binding domains and are activated by Celt (Zhang and Johnson 1992). Moreover. calpain-mediated proteolysis of low molecular weight subunits of human neurofilaments was inhibited by AlCl3 (ICso = 200 uM). mainly by affecting the substrate directly (Nixon et al. 1990). In the absence of experimental data. theoretical ar- guments suggested an avid binding of aluminum to P1P: and/or 1?: (Birchall and Chappell 1988). which in turn may adversely affect the phosphoinositidemediated Ca2+ sig- nalling pathway. Studies on receptor-generated Ca2+ tran- sients in hepatocytes indicated. however. that aluminum does not interact with 1P3 by perturbing its metabolism or its Cazt-releasing properties. It was concluded that alumi- num might rather interact with either phospholipase C or G proteins (Shears et al. 1990). consistent with recent findings on neuroblastoma cells (Shi and Hung 1992: Shi et al. 1993). Clearly more studies are necessary to further de- lineate aluminum target(s) with elements of the phospho— inositide signalling pathway. The aluminum concentrations adminisrered in the above experiments on cells generally varied from 0 to about 1000 M. thus ranging from intracellular background level. around 1 RM. to pathophysiological. and finally to very high concentrations. Given this range of aluminum doses applied. data derived from isolated cells appear to be re- levant regarding the impact of aluminum on the signalling pathway in cells associated with tissue. in particular that of the brain. Regarding pathophysiological levels in experi- mental toxicology. total brain aluminum concentrations average about 100 uM, and in some neuronal populations aluminum levels may reach 400 M as a result of non- uniform accumulation (Nixon et al. 1990). even as high as 1 mM in tangle-bearing neurons of patients afflicted with amyotrophic sclerosis (Perl and Good 1987). As opposed to the extracellular aluminum dose. the intracellular free Al3+ concentration is difficult to estimate. because of un- certainties surrounding the kinetics of aluminum. binding to various intracellular ligands. and the lack of information regarding metal speciation in a cellular environment. Nevertheless. model calculations indicate that the intra- cellular free Al3+ level is in the subnanomolar range. at physiological pH values (Martin 1986; Shi et al. 1993). As indicated above. the majority of experimental studies were conducted by exposing cells to aluminum salts. mostly AlCla. However. when dealing with the bioavail- ability of toxic aluminum in tissues like the brain. the nature of the metal’s coordination sphere (speciation) seems to play an important role. This is illustrated in ex- periments on aluminum-related changes in the permeability of the blood-brain barrier to sucrose of rats. Employing compounds like aluminum lactate and aluminum maltolate. the permeability characteristics were dependent on metal speciation (Favarato et al. 1992). Similar observations were made in studies on the aluminum distribution into brain and liver of rats and rabbits in response to the intravenous ad- ministration of aluminum lactate or citrate (Yokel et al. 1991). Aluminum binding to G proteins other than G, proteins At subnanomolar free Al3+ concentrations. the trivalent cation could compete with Mgzr. at 1 mM. for the ex- changeable nucleOtide (GTP or GDP) binding site. the so- called E site of tubulin (Macdonald et al. 1987). a member of the G protein family (Bosch et al. 1989). All“ inhibited GTPase activity once GTP was bound to the protein. Given the slow ligand exchange rate of A13‘ (Martin 1986). dis- sociation of GTP was inhibited on Apr-bound GTP-tubu- lin, leading to aberrant rrricrotubule assembly and dis- assembly in a pure tubulin system (Macdonald et al. 1987). On the other hand. employing a more realistic system. viz. in vitro assays with tubulin/microtubule associated pro- teins. no structural peculiarities of microtubules have been found in response to various AlC13 doses (0.2-l mM) (Schmidt et al. 1991). Additionally. in studies of mice fed high dietary aluminum (up to l mg/g diet. several weeks). impaired brain microtubule function did not appear to be an early biochemical lesion in aluminum toxicosis (Oteiza et al. 1989). Fluoroaluminates reportedly did not bind to soluble tu- bulin even in the presence of GDP. However. fluoro- aluminates bound with rrricromolar affinity to tubulin in- corporated in microtubules. and inhibited their depoly- merization (Carlier et al. 1988). The fluoride complex ap- parently mimicked the terminal phosphate only after its dissociation from the nucleotide but before its release from the protein (Chabre 1990). Aluminum ions also inhibited the activation of the G protein-related transducin (Miller et al. 1989), which cou- ples rhodopsin to the stimulation of a guanosine-3’-5'- cyclic monophosphate (cGMP)-specific phosphodiesterase in retinal rod outer segments (Stryer 1986). By competing with Mg2+ for a high affinity binding site on transducin. subnanomolar concentrations of intracellular free Ali“. following addition of A103. markedly inhibited GTP binding to the protein. rather than inhibiting GTPase ac- tivity once GTP was bound. as in the case of Al3+ inter- actions with tubulin. At these ion aetivities. the rod cGMP phosphodiesterase was only slightly inhibited (Miller et al. 1989). In the presence of A13». cells are therefore impaired in processing light-induced signals in the vision mecha- msm. Fluoroaluminates also impacted transducin by binding to the Mg2+-GDP-G protein complex. Behaving like a pseudOphosphate. binding of fluoroaluminate to u'ansducin led to the formation of a GTP-alpha subunit complex of the G protein. This latter complex could activate cGMP 141 5 phosphodiesterase in response to rhodopsin photoexcitation (Kanaho et al. 1985). As opposed to the effect of free Al?" (see above). fluoride-chelated aluminum thus enhanced signal transduction from a photoexcited receptor. Another cellular communication channel that operates via G proteins is the receptor-regulated adenylate cyclase system which produces a second messenger. cyclic AMP (Freissmuth et al. 1989). As to aluminum interactions with this communication channel. experiments on rat brains demonstrated that intracerebroventricularly administered aluminum citrate (1 M. pH 7.2) caused pronounced. long- lasting changes in cyclic AMP synthesis in the cortex (Johnson 1988). Moreover. dual effects of aluminum as activator and inhibitor were observed on adenylate cyclase in a parasite. the liver fluke. Micromolar amounts of MCI; augmented the enzyme activation by fluoride. Ale in- hibited the GTP-stimulated cyclase at Mgzt concentrations ranging from 1 to 4 mM (Mansour et al. 1983). Since aluminum has been implicated in contributing to neuronal dysfunction in Alzheimer‘s disease (Crapper McLachlan I989). experiments were performed to de- termine whether the level of alpha subunits of five G pro- teins was altered in Alzheimer patients. Employing im- munoblotting techniques. no significant differences in the level of G protein subunits between Alzheimer‘s diseased and age-matched postmortem brains were detectable (McLaughlin et al. 1991). Conclusions In the past few years. our understanding of primary inter- actions of aluminum with cells has appreciably increased. Current findings provide evidence that aluminum interacts with key elements of cellular communication channels (Fig. 1). At the very least. some of the features of the multifacetted aluminum toxicity syndrome presumably originate from these types of interactions. As a result. aluminum-triggered derangements occur on the levels of second messengers and the temporal progression of signal events which in turn are fairly rapidly reflected in meta- bolic dysfunctions. Regarding long-term changes. alumi- num-related modifications of signalling pathways may lead to altered gene expression and thus to impaired cellular malfunctions since signal transduction pathways are in- volved in regulating gene expression by modulating the activity of nuclear transcription factors (Karin and Smeal l992). Evidently more work is needed to clearly identify direct target(s) of aluminum with wellopurified G proteins. and subunits thereof. What aluminum-triggered structural and functional changes do take place following binding to the protein? Application of recombinant DNA technology tools is expected to enhance our understanding of aluminum- sensitive regions of these proteins. A decision must be reached whether or not. and at which doses. aluminum indirectly interacts with these proteins via metal-related changes in the lipid matrix of the membrane. Additionally. the impact of aluminum on phospholipase C and its lipid microenvironment deserves particular attention. Given ap- parent differences in tissue sensitivity towards aluminum. 142 6 do these variations in part result from differential sensi- tivities of phospholipase isozymes found in the respective tissue? In aluminum-tolerant tissues. have cells adapted to continual aluminum stress by employing communication channels less susceptible to aluminum interference? A hint of such a possibility is given by observations that in the presence of higher Cab concentrations second messengers can be generated by bypassing G proteins. Elucidation of primary aluminum interactions with key elements of cel- lular communication pathways will ultimately yield insight into initial reactions associated with the pathophysiology of aluminum toxicity. References Berridge Ml (I987) lnositol triphosphate and diacylglycerol: two in- teracting second messengers. Annu Rev Biochem 56: 159—193 Berridge Ml (19931 lnositol u'iphosphate and calcium signalling. Nature 361: 315-325 Birchall JD. Chappell JS (1988) Aluminium. chemical physiology. and Alzheimer's disease. Lancet 11: 1008-1010 Bimbaumer L. Abramowitz l. Yatani A. Okabe K. Mattera R. Graf R. Sanford J. Codrna J. Brown AM (1990) Roles of G proteins in coupling of receptors to ionic channels and other effector systems. CRC Crit Rev Biochem Mol Biol 25: 225-244 Blackmore PF. Bocckino SB. Wayniek LE. Exton JH (1985) Role of guanine nucleotide-binding regulatory protein in the hydrolysis of hepatocyte phosphatidylinositol-4.5-biphosphate by calcium-mo- bilizing hormones and the control of cell calcium. J Biol Chem 27: 14477-14483 Bosch L. Kraal B. Parmeggiani A (eds) ( I989) The guanine-nucleotide binding proteins. Common structural and functional properties. Plenum Press. New York Carlier MF. Didry D. Melki R. Chabre M. Pantaloni D (1988) Stabi- lization of microtubules by inorganic phosphate and its skeletal analogues. the fluoride complexes of aluminum and beryllium. Biochemisu'y 27: 3555-3559 Casey P (1992) Visual differences: G proteins. Nature 359: 671-672 Chabre M (1990) Aluminofluoride and beryllofluoride complexes: new phosphate analogs in enzymology. Trends Biochem Sci 15: 6—10 Clarkson TW (1991) Inorganic and organometal pesticides. ln: Hayes WJ. Laws ER (eds) Handbook of Pesticide Toxicology. vol 2. Academic Press. New York. pp 559—565 Claro E. Wallace MA. Fain IN (1990) Dual effect of fluoride on phosphoinositide metabolism in rat brain cortex. Biochem J 268: 733-737 Crapper Mclachlan DR (1989) Aluminum toxicity: criteria for as- signing a role in Alzheimer's disease. In: Lewis TE (ed) En- vironmental chemistry and toxicology of aluminum. Lewis. Boca Raton. Ha. pp 299-315 Dennis EA (1983) Phospholipases. In: Boyer PD (ed) The Enzymes. vol 16. Academic Press. New York. pp 307-353 Dever TE. Glynias Ml. Merrick WC (1987) GTP-binding domain: three consensus sequence elements with distinct spacing. Proc Natl Acad Sci USA 84: 1814-1818 Downward J (1990) The ras superfarnily of small GTP-binding pro- teins. Trends Biochem Sci 15: 469-472 Favarato M. Zatta P. Perazzolo M. Fontana I... Nicolini M (1992) Aluminum (Ill) influences the permeability of the blood-an barrier to ["CJsucrose in rats. Brain Res 569: 330-335 Freissmutlr M. Casey Pl. Gilman AG (1989) G proteins control diverse pathways of transmembrane signaling. FASEB J 3: 2125-2131 Ganrot P0 (1986) Metabolism and possible health effects of alumi- num. Environ Health Perspect 65: 363-441 Hall A (1992) Signal transduction through small GTPases - a tale of two GAPs. Cell 69: 389-391 Haug A (1984) Molecular aspects of aluminum toxicity. CRC Crit Rev Plant Sci 1: 345-373 Hazlett TL. Higashijima T. Jameson DM (1990) An aluminum fluoride binding site was not detected on elongation factor Tu-GDP. Bio- phys] 57: 289a Johnson GVW. lope RS (1986) Aluminum impairs glucose utilization and cholinergic activity in rat brain in vitro. Toxicology 40: 93-l02 Johnson GVW (1988) The effects of aluminum on agonist-induced alterations in cyclic AMP and cyclic GMP concentrations in rat brain regions in vivo. Toxicology 51: 299-308 lope RS (1988) Modulation of phosphoinositide hydrolysis by NaF and aluminum in rat conical slices. J Neurochem 5|: 1731-1736 Kahn RA ( I991) Fluoride is not an activator of the smaller (20—25 kDa) GTP-binding proteins. J Biol Chem 266: 15595-15597 Kanaho Y. Moss 1. Vaughn M (1985) Mechanism of inhibition of transducin GTPase activity by fluoride and aluminum. J Biol Chem 260: “493-11497 Karin M. Smeal T (1992) Control of transcription factors by signal transduction pathways: the beginning of the end. Trends Biochem Sci 172418-422 _ Kraal B. de Graaf JM. Mesters JR. van Hoof PJM. laquet E. Par- meggiani A (1990) Fluoroaluminates do n0t affect the guanine nucleotide binding centre of the peptide chain elongation factor EF-Tu. Eurl Biochem 192: 305-309 La Cour TFM. Nyborg J. Thin 5. Clark BFC (1985) Structural details of the binding of guanosine diphosphate to elongation factor Tu from E. coli as studied by X-ray crystallography. EMBO J 4: 2385-2388 . Lau KHW. Yoo A. Wang SP (1991) Aluminum stimulates the pro- liferation and differentiation of osteoblasts in vitro by a mechanism that is different from fluoride. Mol Cell Biochem 105: 93-105 Lee E. Yuspa SH (1991) Aluminum fluoride stimulates inositol phosphate metabolism and inhibits expression of differentiation markers in mouse keratinocytes. J Cell Physiol 148: 106-115 Levine SN. Sonnier GB. Abroe K (1990) Effects of diabetes mellitus and aluminum toxicity on myocardial calcium transport. Toxicol- ogy 65: 137-147 Lewis TE (ed) (1989) Environmental chemistry and toxicology of aluminum. Lewis. Boca Raton. Fla. Macdonald TL. Humphreys WG. Martin RB (1987) Promotion of tu- bulin assembly by aluminum ions in vitro. Science 236: 183-186 Magnusson MK. Halldorsson H. Kjeld M. Thorgeirsson G (1989) Endothelial inositol phosphate generation and prostacyclin pro- duction in response to G protein activation by AlFI. Biochem J 264: 703-711 Majerus PW (1992) lnositol phosphate biochemistry. Annu Rev Bio- chem 61: 225-250 Mangels LA. Neubig RR. Hamm HE. Gregy ME (1992) Calmodulin binding distinguishes between beta gamma subunits of activated G proteins and transducin. Biochem J 283: 683—690 Mansour JM. Ehrlich A. Mansour TE (1983) The dual effects of alu- minum as activator and inhibitor of adenylate cyclase in the liver fluke Fasciola hepatica. Biochem Biophys Res Commun 112: 911-918 Marquis lK (1989) Neurotoxicity of aluminum. In: Lewis TE (ed) Environmental chemistry and toxicology of aluminum. Lewis. Boca Raton. Fla. pp 289-298 Martin RB (1986) The chemistry of aluminum as related to biology and medicine. Clin Chem 32: 1797-1806 McDonald Ll. Manuack MD (1988) Aluminum affects phosphoino- sitide hydrolysis by phosphoinositidase C. Biochem Biophys Res Comrnun 155: 203-208 McLaughlin M. Ross BM. Milligan G. McCulloch J. Knowler IT (1991) Robustness of G proteins in Alzheimer‘s disease: an im- munoblot study. J Neurochem 57: 9-14 Miller lL. Hubbard CM. Litrnan Bl. Macdonald TL (1989) Inhibition of transducin activation and guanosine uiphosphatase activity by aluminum ion. J Biol Chem 264: 243—250 Nebelter HG. Coburn JW (1986) Aluminum and renal osteodystrophy. Annu Rev Med 37: 79-95 Nixon RA. Clark JF. Logvinenko KB. Tan MKH. Hoult M. Grynspan F (1990) Aluminium inhibits calpain-mediated proteolysis and in- duces human neurofilament proteins to form proteaseoresistant high molecular weight complexes. J Neurochem 55: 1950—1959 Oteiza Pl. Golub MS. Gershwin ME. Donald JM. Keen CL (1989) The influence of high dietary aluminum on brain microtubule poly merization in mice. Toxicol Lett 47: 279-285 Panchalingam K. Sachedina S. Pettegrew lW. Glonek T (1991) Al- ATP as an intracellular carrier of Al(lll) ion. Int J Biochem 23: 1453-1469 . Perl DP. Good PF ( I987) The association of aluminium. Alzheimer's disease and neurofibrillary tangles. J Neural Transm 24: 205-211 Rana RS. Hokin LE (1990) Role of phosphoinositides in transmem- brane signaling. Physiol Rev 70: 115-164 Rasmussen H (1990) The complexities of intracellular Ca“ signalling. Biol Chem Hoppe-Seyler 371: 191-206 Rath HP. Blackmore PF (1990) Differential activation of rabbit fe- moral arteries by aluminum fluonde and sodium fluoride. l Phar- macol Exp Ther 254: 514—520 Savchenko GS. Tananaev IV (1959) Types of complex fluoro- aluminates in aqueous solution. Soviet research in fluorine cheni- istry. 1949-1956. part 11. English edition. Consultants Bureau. New York 203-212 Schmidt R. BOhm K. Vater W. Unger E (1991) Aluminium induced osteomalacia and encephalopathy - an aberration of the tubulin assembly into microtubules by Al"? Progr Histo- Cytochem 23: 355-364 Schdfl C. Sanchez~Bueno A. Dixon Cl. Woods NM. Lee JAC. Cuth- bertson KSR. Cobbold PH. Birchall JD (1990) Aluminium perturbs oscillatory phosphoinositide-mediated calcium signalling in hor- mone-stimulated hepatocytes. Biochem J 269: 547-550 Shears SB. Dawson AP. Loomis-Husselbee lW. Cullen Pl ( 1990) The perturbation. by aluminium. of receptor-generated calcium tran- sients in hepatocytes is not due to effects of 1ns(1.4.5)P3-stimulated Ca" release or lns(1.4.5)P3 metabolism by the 5—phosphatase and 3okinase. Biochem J 270: 837 Shi B. Haug A (1990) Aluminum uptake by neuroblastoma cells. J Neurochem 55: 551-558 Shi B. Haug A (1992) Aluminium interferes with signal transduction in neuroblastoma cells. Pharmacol Toxicol 71: 308-313 143 7 Shi B. Chou K. Haug A (1993) Aluminium impacts elements of the phosphoinositide signalling pathway in neuroblastoma cells. Mol Cell Biochem 121: 109-118 Stemweis PC. Gilman AG (1982) Aluminum: a requirement for acti- vation of the regulatory component of adenylate cyclase by fluoride. Proc Natl Acad Sci USA 79: 4888-4891 Stryer L (1986) Cyclic GMP cascade of vision. Annu Rev Neurosci 9: 87-119 Taylor GJ (1991) Current views of the aluminum stress response; the physiological basis of tolerance. Curr Top Plant Biochem Physiol 10: 57-93 Tojyo Y. Tanimura A. Matsui S. Matsumoto Y. Sugiya H. Furuyama S (1991) NaF-induced amylase release from rat parotid cells is mediated by P1 breakdown leading to Ca3° mobilization. Am J Physiol 260: C194-C200 Troullier A. Girardet lL. Dupont Y (1992) Fluoroaluminate complexes are bifunctional analogues of phosphate in sarcoplasmic reticulum Cale-ATPase. J Biol Chem 267: 22821—22829 Van der Voet GB. Marani E. de Wolff FA (1991) Aluminium neuro- toxicity. Progr Histo- Cytochem 23: 235-242 Wakui M. ltaya K. Nirchall D. Peterson H (1990) Intracellular alu- minium inhibits acetylcholine- and caffeine-evoked Calt mobili- zation. FEBS Len-267: 301—304 Weis C. Haug A (1989) Aluminum-altered membrane dynamics in human red blood cell white ghosts. Thromb Res 54: 141-149 Wills MR. Savory l (1985) Water content of aluminum. dialysis de- mentia. and osteomalacia. Environ Health Perspect 63: 141-147 Woods NM. Dixon Cl. Cuthbertson KSR. Cobbold PH (1990) Fluoroaluminate mimics agonist application in single rat hepato- cytes. Biochem J 265: 613-615 Yokel RA. Lidums V. McNamara Pl. Ungerstedt U (1991) Aluminum distribution into brain and liver of rats and rabbits following in- travenous aluminum lactate or citrate: a microdialysis study. Toxicol Appl Pharrnacol 107: 153-163 Yuan S. Haug A (1988) Frictional resistance to morions of bimane- labelled spinach calmodulin in response to ligand binding. FEBS Lett 234: 218-223 Zhang H. Johnson P (1992) Differential effects of aluminum ion on smooth muscle calpain land calpain ll activities. Int J Biochem 24: 1773-1778 APPENDIX D ALUMINUM COORDINATION TO CALMODULIN: THERMODYNAMIC AND KINETIC ASPECTS Haug, A., Vitorello, V. 1996. Coord. Chem. Rev. (in press) 144 lilllllllllNllllllN CHEMISTRY ' . ' Coordination Chemistry Reviews 0 (1995) REVIEWS ELS EVI R 000-000 Aluminium coordination to calmodulin: Thermodynamic and kinetic aspects Alfred Haug ', Victor Vitorello ”'1 ‘ Department ofMiaobiology. Michigan State University, East Busing, MI 48824. USA " Department of Boraw. Michigan State University. East Lansing. MI 48824. USA Received 23 January 1995; revised 3 April 1995 Contents Abstract ...................................................... .000 1. Introduction ................................................ 000 2. Calmodulin ................................................. 000 3. Al( III) in biology ............................................. 000 3.1. A1( III) in aqueous solution ..................................... 000 3.2. Al(III) in cells ............................................ 000 4. Al(III) impacts calmodulin ........................................ 000 4.1. Thermodynamic aspects of Al(III) binding to calmodulin ................... 000 4.2. Kinetics of Al(III) removal by dtric acid from calmodan .................. 000 4.3. Al(III) impairs recomition of calmodan by target proteins ................. 000 4.4. Al(III)-related changes in calmodulin's internal dynamics ................... 000 Summary .................................................... 000 Acknowledgement ............................................... 000 References .................................................... 000 Abstract Calmodulin is an acidic, Ca(II)-binding protein (about 17000 d) which plays a key role in regulating a broad spectrum of target enzymes in eukaryotic cells. Within each lobe of the dumbbell-shaped molecule there are two adjacent Ca(II)-binding domains. which form a single cooperative unit. Hydrophobic forces are crucial for interfacing the Ca(II)-activated regulatory protein with its target enzyme. Presumably not involving the {our specific Ca(II)- binding domains on calmodulin, binding of Al(III) to calmodan generates considerable dehydration entropies. Al(III )-triggered changes in the a-helix content of calmodulin are involved in the mismatch between calmodulin and its target protein. By altering calmodulin's ‘ Permanent address: Centro de Energia Nuclear na Agrintltura/USP. Caixa Postal 96. Piracicaba. 13400 Sao Paulo. Brasil. 0010-8545/95/309.50 c 1995 ElseVier Science S.A. All rights reserved 5501 0010~8545(95)01168-4 145 2 internal dynamies. Al(III) apparently interferes with the proteins capacity to search out conformational substates suitable for proper docking with a specific target protein. Keywords: Al(III) binding; Al(III)-related dehydration; Al(III) removal; Calmodulin; Calmodulin partners; Protein dynamics 1. Introduction In eukaryotic cells a small protein, calmodulin, plays a pivotal role in regulating a broad spectrum of target enzymes which in turn are involved in vital cellular processes, such as protein phosphorylation, cyclic nucleotide metabolism, motility, cell division, and sensory perception [1-3]. The regulatory action is accomplished by binding Ca(II) to calmodulin, thereby modulating its conformation in such a manner as to promote specific interactions between the small protein and the respective target. As a cytosolic protein, calmodulin senses changes in intracellular free Ca(II) concentrations, [Ca(II)], ranging from about 10 nM to 1 itM. Changes in [Ca(II)] are brought about by release of Ca(II) from intracellular stores, or from Ca(II) signals generated in response to the transduction of external input signals across the plasma membrane. Chief among signal transduction pathways is that associated with membrane-bound phosphoinositides [4-6]. Calmodulin is therefore an essential communication link necessary for transforming extracellular information into specific biological responses. 2. Calmodan The key to calmodulin’s ability to function as a Ca(II)-dependent transducer resides in its primary protein structure. This acidic, globular protein (about 17000 d, 148 amino acid residues, isoelectric point about 4.1) has been highly conserved during the evolution of eukaryotic cells. suggesting the existence of an ancestral gene E 7,8]. X-ray diffraction studies demonstrated that calmodulin resembles a dumbbell (length about 6.5 nm) with the two lobes interconnected by an eight-turn a-helix. This central helix (about 2 nm long, amino acid residues 65-92) carries an appreciable net dipole which may play a role in the sequential binding of Ca(II) to the apoprotein [9.10]. Each lobe (diameter about 2 nm) harbours two Ca(II)-binding domains whose helix-loop—helix arrangement (11:12:11 residues reSpectively) is similar to the structure of Ca(II)—binding units in other Ca(II)-binding proteins [11]. In - calmodulin, only the loop regions contain antiparallel B-sheet strands. Coupling antiparallel strands of adjacent Ca(II)-binding loops, hydrogen bonds assist in stabilizing the macromolecule [9,10]. Counting from the amino terminus of calmo- dulin, the Ca(II)-binding domains are designated as I, II, III and IV respectively. Thermodynamic studies on domain organization in calmodulin indicated that within a lobe the two adjacent Ca(II)-binding domains form a single cooperative block, especially between domains I and II. The two lobes seem to repel each other [12]. 146 3 The four domains in calmodulin permit specific binding of 4 mol Ca(II) per mol of protein. At approximate vertices of an octahedron, the liganding residues harbour oxygen-containing side chains, such as those of aspartic acid, and the peptidyl carbonyl oxygen. Hydration water is probably part of the ligand sphere [13]. Despite considerable similarities in the primary sequence of the domains’ loops, the dissoci- ation constants for Ca(II) vary between 1 uM and 10 MM, at physiological conditions. The Ca(II)-binding properties and the sequence of Ca(II)-binding to the protein are critically dependent on the presence of monovalent ions, pH, and divalent cations like Mg(II). Current models indicate that domains III and IV (residing on the C-terminus lobe) are c00peratively binding Ca(II) with high amnity, as compared with the low afinity cooperative binding in domains I and II [2,14]. NMR experi- ments on calmodulin in solution further corroborate the existence of high and low affinity Ca(II)-binding domains [14]. For the high affinity domains, the rates of Ca(II) dissociation range from 3 to about 10 3”, while those from the low afinity domains are in the 300 to 550 5" range, at room temperature [15,16]. Ca(II) association constants are expected to range from 107 to 10‘I s“, since these values are related to the rates of water exchange of hydrated Ca(II) ions [17]. Upon binding of 1 to 4 mol Ca(II) per mol of apoprotein, pronounced structural transformations take place in the protein such as increase of the a-helix content. and enhanwd formation of hydrophobic regions on the protein surface. Ca(II)-binding generates a readjustment/formation of the fi-sheet within the binding 100ps of the domains [18]. These helical rearrangements may in part result from coupled move- ments of a pair of helices [l9]. Elicited by coordination reactions [20], especially by those involved in the binding of two Ca(II) in the high affinity domains of the C-terminus lobe, local perturbations in structure are also transmitted to the N-terminus lobe of the flexible macromolecule [21]. Ca(II) binding also efl'ects the protein to elongate by about 7% [22]. Ca(II)-triggered short and long range conformational changes cause an enhanced exposure of hydrophobic regions on the protein surface [23], thus preparing calmoduhn for docking with and the activation of the target protein. The molecular basis for calmodulin’s ability to activate numerous target proteins has not been elucidated. Nevertheless, experiments on calmodulin-binding oligopep- tides. like the amphiphilic melittin (2845 d, about 1.5 nm long, 26 amino acid residues), demonstrated that hydrophobic forces are crucial for interfacing the Ca(II)-calmodulin with its target [24]. Regarding calmodulin interactions with its partner (molar ratio 1:1), binding constants vary from 10 itM to 100 pM [25]. The Ca(II)-calmodulin structure apparently collapses around its partner [26], involving hydrophobic and salt bridges [27]. 3. Al(III) in biology 3.1. Al(III ) in aqueous solution Since the water content of a typical cell comprises about 80% of the cellular weight, it is essential to understand the behaviour of Al(III) in aqueous solution. I47 4 Regarding details of the aqueous chemistry of Al(III) we are referring to informative review articles [28—30]. Briefly, the coordination of the trivalent cation, Al(III), in aqueous solution is generally either tetrahedral, pH > 6.2, like in Al(OH): , or octahe- dral, pH<5, like in Al(H,O)2‘. The ionic radius is 0.054 nm; the hydrated radius has a value of 0.475 nm [31]. As the pH increases, the hexahydrated species hydro- lyzes, especially in the range from 5]=6:1), the removal of Al(III) was accelerated, as indicated by k3- 2551-05 min" and k“ n0.5i0.04 min". As judged by the rates of Al(III) removal of the second and third Al(III), the binding sites are heterogeneous and probably noninteracting. The diflerences in the rates, i.e. k, vs. k“, presumably arise from the accessibility of the chelator to the respective Al(III) binding sites in the flexible apocalmodulin [56], or the more compact Ca(II)-saturated calmodulin [10,57]. Given these observations, one can speculate that the first Al(III) is coordinated to ligands in the central helix region which is readily accessible in the absence or presence of Ca(II). The second and third Al(III) are coordinated to ligands in the vicinity of, or at, the Ca(II)-binding domains, as hypothesized above, where chelator access might be restricted in the more compact Ca(II)-saturated protein compared with the more open apocalmo— dulin. These types of restriction and the profound denaturation initially produced by Al(III) in the protein [41] may account for the lack of ready reversibility of the Al(III)—calmodulin complex upon applican'on of citrate. It is noteworthy that the rate constants, k, and k,, correspond to half-lives of about 10 and 70 s respectively. 150 7 Similar time constants, ranging from 10 to about 100 s, have been observed in thermal unfolding reactions of ribonuclease A [58], whose molecular weight (M :13 700) is close to that of calmodulin. Citrate-related removal of Al(III) is a considerably slower process than dissociation of Ca(II) from calmodulin. Employing Ca(II)- sensitive fluorescence chelators, the latter rate constants, kg“, 550 s" and about 5 s" (19‘C), correspond to the fast and slow release of two Ca(II) each from the protein [44]. Al(III) application inhibited the Ca(II)- and calmodulin-dependent 3',5'-cyclic nucleotide phosphodiesterase activity by impairing calmodufin rather than the basal activity of the enzyme, examined in the absence of calmodulin [41]. Following exhaustive citrate removal by dialysis, the Al(III)-depleted calmodulin stimulated the Ca(II)—calmodulin-dependent phosphodiesterase activity in a manner identical to a calmodulin which had never seen Al(III) and citrate. Despite such a complete restoration of biological activity, restoration of the a-helix content of calmodulin remained incomplete, irrespective of the presence or absence of Ca(II) [55]. This lack of complete restoration may result from local rearrangements of ligands in the region previously occupied by Al(III). Clearly, more experiments are necessary to understand mechanisms of chelator- triggered dissociation of Al(III) from calmodulin. Al(III)-sensitive, fluorescent chela- tors may be used to observe processes related to the removal of the first Al(III). The Al(III)-related pathway of calmodulin refolding should be analysed relative to that operative upon Ca(II) dissociation. 4.3. Al(III ) impairs recognition of calmodulin by target proteins Al(III)-triggered structural changes in calmodulin reportedly led to the inhibition of Ca(II)—calmodulin-dependent enzymes like 3',5'-cyclic nucleotide phosphodiester- ase [41], a plant membrane-bound ATPase activity [59], and calpains [60.61]. To get insight into the tOpography of the enzyme’s binding region with Ca(II )-activated calmodulin, a high-amnity, amphiphilic, 26-residue peptide [24,25] from bee venom, melittin, was used [43]. The steady state and time-dependent fluorescence character- istics of melittin’s single tryptOphanyl residue are sensitive parameters to derive information on the peptide’s interface with calmodulin, with Al(III) absent or present. With Al( III) present. the tryptophanyl residue experienced a more polar microenvi- ronment at the interface of melittin with Ca(II)-calmodulin. These findings are probably a consequence of the Al(III)-triggered breakage of a-helices in Ca(II)—calmodulin [41] which in turn would lead to a more open Ca(II)—calmodulin structure compared with the compact Ca(H)-calmodulin.structure [62], observed in the absence of Al(III). Al(III)-induced structural changes are coupled with a rearrangement of salvation water. which plays a key role in protein dynamics [63,64] and as a hydrogen bond sink [65]. An open protein structure is also indicated by: (a) a decreased immobilization of a spin probe covalently attached to Al(III )-treated Ca(II)—calmodulin [42]; (b) an improved access of diffusion-controlled, collisional quenching molecules to the fluorophore of melittin, associated ( 1 : l) with calmodulin; (c) an enhanwd rotational correlation time of melittin’s fluorophore. This faster 151 3 correlation time (3.5 as) is indicative of segmental motions in the vicinity of the tryptophanyl residue compared with restricted motions (8.1 ns) observed in the hydrophobic-hydrogen bonding [65,66] interplay of the compact melittin-calmo- dulin complex, existing in the absence of Al(III) [43]. Since the basic moiety of melittin apparently associates with acidic amino acid residues [26,67] of the central helix in biochemically active Ca(II)-calmodulin [9.10], it is noteworthy that forma- tion of time-dependent hydrogen bonding patterns by the residues Arg 74, Arg 86,‘ and Arg 90 are crucial in the compaction process during complexation [68]. The central helix has also been suggested to harbour a Mg(II)-binding site [14] which in turn is a putative site for Al(III), as discussed previously. Arg 86 is flanked by three glutamate residues [24] (Glu 82, Glu 83, Glu 84) potentially capable of forming chelate rings through side chain donor atoms [47]. Given the proximity of key residues for the formation of hydrogen bonding patterns and the putative Al(III)— binding site on the central helix, Al(III)-triggered breakage of hydrogen bonds would have profound repercussions on the compaction process during calmodulin’s inter- action with melittin. Taken together, the presence of Al(III) apparently induces a mismatch between calmodulin and its partner. Such a mismatch may be instrumental in Al(III)-related malfunctions of Ca(II)—calmodulin dependent enzymatic reactions. Concerning the Al(III)-related lack of recognition between calmodulin and its specific partner, experiments assessing solely the binding constants for the two proteins are obviously insuficient. Detailed structural and dynamic studies are required to ascertain Al(III)-related changes in the complementary surface, crucial for recognizing calmodulin’s specific partner protein. Particular attention should be paid to the role of hydration water presumably involved in the Al(III)-disturbed compaction process upon complexation. 4.4. Al(III )-related changes in calmodulin's internal dynamics Like other globular proteins, calmodufin has a distinct equilibrium structure [10]. Afl'ording calmodulin the capacity to associate with specific partners, other conforma- tions. separated by low energy barriers, and different from the average equilibrium structure, must exist [63,64,69]. Although a clear picture has not yet emerged, the space-time dynamics in calmo- dulin was illustrated by nanosecond motions of a specific site [70.71], by thermally induced fluctuations modulating dipole—dipole interactions [72], by nanosecond responses elicited via ligand-triggered conformational perturbations [21], and by internal and side chain mobilities in NMR experiments [15,44]. Whether low fre- quency motions (wave numbers less than 50 cm") are important for calmodulin’s biological activity has not been elucidated [73]. For the purpose of diagnosing Al(III)-related anomalies in the dynamics of calmo- dulin, relative to that in the presence of Ca(II), the frictional resistance to local motions of a protein-bound fluorophore was evaluated [74]. The fluorescence anisot- ropy and lifetime of protein-associated fluorophore are sensitive to global motions of the protein and to local rotations in the fluorophore's microenvironment From the temperature dependence of these fluorescence parameters the thermal coefiicient 152 9 of the frictional resistance to the fluorophore motions can be extracted [75,76]. As to calmodulin, Al(III)-generated changes in fluorescence parameters were measured (0.1 M KCl, 10 mM Mops bufl‘er, pH 6.5, 5 mol Ca(II) and 3 mol Al(III) per mol calmodulin, range 10 to 40 °C) through a fluorophore covalently anchored at site 26 of spinach calmodulin where a cysteinyl residue is residing. Both in the presence or absence of Al(III), calmodulin underwent a substate change at a critical temperature of 25 °C [74]. This temperature is appreciably lower than that of the major unfolding of the protein: at 55 °C for apocalmodulin and above 90°C in the presence of Ca(II) [77]. In the presence of Al(III), the enthalpy, AH=8.2 kcal mol", was close to that of apocalmodufin (8.0 kcal mol“), as opposed to biochemically active Ca(II)-calmodulin (5.3 kcal mol"), measured in the absence of Al(III). The lower AH value in biochemically active calmodan apparently reflects the sensitive dynamic aspects essential for the transfer of information within the protein in response to regulatory stimuli. As indicated by a AH =8.2 kcal mol"1 and the thermal resistance coeflicients, the fluorOphore motions appeared to be less disposed to temperature changes when Al(III) was associated (3: 1) with calmodulin. Since A1( III) application to calmoduhn caused hydrogen bond breakage [41], some of the temperature- dependent changes in AH and motional characteristics described above may be attributed to alterations in protein hydration (solvent fluctuations) relative to that seen by biochemically active Ca(II)—calmodulin. The observed restricted fluorophore mobility at site 26, with Al(III) present, may in part result from the increased viscosity in the fluorophore‘s microenvironment [74], because solvent atoms exert a damping efl'ect on the atoms at the surface of the protein [64]. Taken together, by altering calmodulin’s conformational substates, Al(III) apparently interferes with the protein’s capacity to search out substates suitable for proper docking with a specific partner protein harbouring melittin-like amino acid sequences. 5.Summary To fully understand the regulatory properties of multifunctional Ca(II)-calmodulin, the systematic features of the protein's space-time dynamics must be clarified. Links must be established between the critical conformational substate at the onset of ligand binding (e.g. Ca(II) coordination) and the final conformational substate(s) responsible for correct docking between calmodulin and its respective target. By the same token, the temporal and structural gap must be bridged between the initial substate of the Al(III)-coordinated protein and the dysfunctional substate, nonOperative for docking, to understand molecular details of the Al(III)-induwd lesion on calmodulin. By employing bioengineered calmodulin with a single intrinsic fluorescence emitter (tryptOphan), located at selectively defined sites along the protein backbone, fluctuations of the local state can be monitored at various regions of calmodulin subjected to Al(III) coordination. Recent studies had indicated that Al(III) is primarily bound to membrane- associated elements of the Ca(II)-phosphoinositide signalling pathway [38,39], rather than to cytosolic calmodulin. However, the possibility cannot be excluded 10 153 that Al(III) is also binding to calmodulin-dependent enzymes associated with the membrane [78], or to calmodulin involved in membrane channel functions [79]. Given the considerable amount of information available on structural details (X-ray structure, amino acid sequence), and the availability of genetic tools [53] to design calmodulins differing in amino acid composition, we believe that calmodulin is a unique instrument to derive a microscopic picture on the impact of Al(III) coordination on an important macromolecule. Acknowledgement One of us (V.V.) was supported in part by a fellowship, grant no. 200892/91-6, from the Conselho Nacional de Desenvolvimento Cientifico e Tecnologico (Brasil). References [1] CB. Klee and TC Vanaman. Adv. Protein Chem. 35 (1982) 213. [2] C.L.A. Wang and PC Leavis, in M.P. Thompson (Ed), Calcium-Binding Proteins. VoL 1., CRC Press, Boca Raton. FL. 1988, p. 97. [3] DC. Wylie and T.C. Vanaman. in P. Cohen and CB. Klee (eds), Calmodulin. Elsevier. Amsterdam, 1988, p. 1. [4] MJ. Berridge, Nature. 361 (1993) 315. [5] P.W. Majerus, Annu. Rev. Biochem., 61 (1992) 225. [6] H. Rasmussen, Biol. Chem. Happe-Seyler, 371 (1990) 191. [7] M. Goodman, Progr. Biophys. Melee. Biol, 37 (1981) 105. [8] A.R. Means. LA. Putkey and P. Epstein. in P. Cohen and CB. Klee (Eds), Calmodulin. Elsevier, Amsterdam, 1988, p. 17. [9] Y5. Babu. 1.5. Sack. TJ. Greenhough. CE Bugg, A.R. Means and WJ. Cook, Nature. 315 (1985) 37. [10] Y.S. Babu. CE. Bugg and WJ. Cook. in P. Cohen and CB. Klee (Eds), Calmodulin, Elsevier. Amsterdam. 1988, p. 83. [11] H. Kawasaki and R.1-l. Kretsinger, in P. Shelterline (Ed), Protein Profile, Vol. 1, Academic Press, . London, 1994, p. 343. [12] TN. Tsalkova and PL Privalov, .1. MOL Biol, 181 (1985) 533. [13] N.CJ. Strynadka and M.N.G. James. Annu. Rev. Biochem., 58 (1989) 951. [14] GB. Klee, in P. Cohen and CB. Klee (Eds), Calmodulin, Elsevier, Amsterdam. 1988. p. 35. [15] IS. Evans, BA. Levine. R.J.P. Williams and MR. Womald, in P. Cohen and CB. Klee (Eds), Calmodulin. Elsevier. Amsterdam, 1988. p. 57. [16] P. Bayley. P. Ahlstrom. S.R. Martin and S. Forsen. Biochem. Biophys. Res. Commun., 120 ( 1984) 185. [17] D.W. Margerum. G.R. Cayley, D.G. Weatherburn and GK Pagenkopf, in AB. Martel! (Ed), Coordination Chemistry. VoL 2, American Chemical Society, Washington. 1978. p. l. [18] C.L.A. Wang, Q. Zhan. T. Tao and J. Gergeley, Biophys. J., 22 (1986) 48a. [19] O. Herzberg. 1. Moult and M.N.G. James, 1. BioL Chem. 261 (1986) 2638. [20] 6.1... Eichhorn. Coord Chem. Rev. 128 (1993) 167. [21] S. Yuan and A. Haug. Eur. J. Biochem.. 175 (1988) 119. [22] B.A. Seatom J.P. Had. D.M. Engelman and EM. Richards, Biochemistry, 24 (1985) 6740. [23] BE. Finn. T. Drakenberg and S. Forsen. FEBS Lett. 336 (1993) 368. [24] LA. Cox. M. Comte, A. Mamar-Bachi, M. Miles and JJ. Schaer. in C. Gerday, L. Bolis and R Gilles (Eds), Caldum and Calcium Binding Proteins. Springer-Veda; Berlin. 1988. p. 141. 154 u [25] SR. Anderson and DA. Malencik. in W.Y. Cheung (Ed), Calcium and Cell Function, Vol. V1, Academic Press. Orlando. FL. 1986. p. 1. [26] P. James. T. Vorherr and E Carafoli. Trends Biochem. Sci, 20 (1995) 38. [27] WE Meador, A.R. Means and EA. Quiocho. Science, 262 (1993) 1718. [28] CF. Baes and 11.5. Mesmer, The Hydrolysis of Cations. Wiley, New York, 1976, p. 112. [29] A.P. Martel] and RJ. Motekaitis. in T.E. Lewis (Ed), Environmental Chemistry and Toxicology of Aluminum. Lewis. Boa Raton, FL. 1989, p. 3. [30] R.B. Martin. in H. Sigel (Ed), Metal Ions in Biological Systems. Vol. 24, Marcel Dekker. New. York. 1988. p. 1. [31] ER. Nightingale. J. Phys. Chem. 63 (1959) 1381. [32] C. Orvig. in OH. Robinson (Ed), Coordination Chemistry of Aluminum. VCH. New York. 1993. p. 85. [33] J.W. Akin. Progr. NMR Spectr. 21 (1989) 1. [34] RM. Bertsch. in G. Sposito (Ed), The Environmental Chemistry of Aluminum. CRC Press, Boca Raton. FL. 1989, p. 81. [35] R.B. Martin. Clin Chem. 32 (1986) 1797. [36] K. Panchalingam. S Sachedina. J.W. Pettegrew and T. Glonek. Int. J. Biochem.. 23 (1991) 1453. [37] SJ. Karlik. GA. Egavish and GI. Eichhorn, J. Am. Chem. Soc.. 105 (1983) 602. [38] B. Shi. K. Chou and A. Haug. Molec. Cell. Biochem.. 121 (1993) 109. [39] A. Haug, B. Shi and V. Vitorello, Arch. ToxicoL. 68 (1994) 1. [40] D.A. Buckingham. in GI... Eichhorn (Ed), Inorganic Biochemistry, Elsevier. Amsterdam. 1975. p. 3. [41] N. Siegel and A. Haug. Biochim. Biophys. Acta. 744 (1983) 36. [42] N. Siegel R. Coughlin and A. Haug. Biochem. Biophys. Res. Commun.. 115 (1983) 512. [43] C. Weis and A. Haug. Arch. Biochem. Biophys., 254 (1987) 304. [44] S. Forsen. HJ. Vogel and T. Drakenberg. in W.Y. Cheung (Ed), Calcium and Cell Function. Vol. VI, Academic Plus. FL. 1986, p. 113. [45] GA. MePhalen. N.C.J. Strynadka and M.N.G. James. Adv. Protein Chem. 42 (1991) 77. [46] MT Record. CF. Anderson and T.M. Lohman. Quart. Rev. Biophys.. 11 (1978) 103. [47] '1'. Kiss. in K. Burger (Ed), Biocoordination Chemistry. Ellis Horwood. New York. 1990, p. 56. [48] N. Siegel and A. Haug. Inorgan. Chim. Acta. 79 (1983) 230. [49] J.P. Glusker. Adv. Protein Chem, 42 (1991) 1. [50] J.P. McManus. CW. Hague, BJ. Manden. M. Sikorska and AG. Szabo, J. BioL Chem. 265 (1990) 10358. [51] J. Trewhella. WK. Liddle, D.B. Heidorn and N. Strynadka. Biochemistry, 28 (1989) 1294. [52] G. Lindeberg. H. Bennich and A. Engstroem. Int. J. Peptide Protein Res. 38 (1991) 253. [53] G.K. Ackers and ER. Smith. Annu. Rev. Biochem.. 54 (1985) 597. [54] 1.1.. Miller. CM. Hubbard. BJ. Litman and T.L Macdonald, J. Biol. Chem. 264 (1989) 243. [55] CG. Suhaya and A. Haug. Can. J. Biochem. Cell BioL. 63 (1985) 1167. [56] P.K. Lambooy. Rf. Steiner and H. Sternberg. Arch. Biochem. Biophys.. 217 (1982) 517. [57] DC. Laporte, B.M. Wierman and D.R. Strom, Biochemistry. 19 (1980) 3814. [58] CR. Cantor and PR Schimmel. Biophysical Chemistry. Freeman. San Francisco. 1980. p. 1095. [59] N. Siegel and A. Haug. Physiol. Plant. 59 (1983) 285. [60] H. Zhang and P. Johnson. Int. J. Biochem.. 24 (1992) 1773. [61] R.A. Nixon. 1.1". Clarke. K.B. Logvinenko, M.K.H. Tan. M. Hoult and F. Grynspan. J. Neurochem.. 55 (1990) 1950. [62] H. Weinstein and EL Mehler. Annu. Rev. Physiol., 56 (1994) 213. [63] G. Careri. Order and Disorder in Matter. Benjamin—Cummings. Menlo Parks. CA. 1984. [64] W17. van Gunsteren and M. Karplus. Biochemistry. 21 (1982) 2259. [65] IL Finney. J. Phys (Paris) Colloq. 45 (C7) (1984) 197. [66] A.R. Persht Trends Biochem. Sci, 12 (1987) 301. [67] RC. Weber. TJ. Lukas, 12A. Craig, E. Wilson. M.M. King, A.P. Wiatkowski and D.M. Watterson. Proteins”. Struct. Funct. Genet. 6 (1989) 70. [68] LL Pascual-Ahuir. EL Mehler and H. Weinstein. Mol. Eng. 1 (1991) 231. ,2 155 [69] J.A. McCammon and SC. Harvey, Dynamics of Proteins and Nucleic Acids, Cambridge Univiversity Press London, 1987. [70] S. Pundak and RS. Roche. in B. de Bernard, G.l... Sottocasa G. Sandri, E Carafoli, A.N. Taylor. T.C. Vanaman and R.J.P. Williams (Eds). Calcium Binding Proteins, Elsevier, Amsterdam, 1983. p. 171. [71] M. Chabberl. M.C. Kilhofler. D.M. Watterson. J. Haiech and H. Lami, Biochemistry. 28 (1989) 6093. [72] PB. O'Hara. KM. Gorski and M.A. Rosen. Biophys. J., 53 (1988) 1007. [73] KC. Chou. BiOphys. Chem. 30 (1988) 3. [74] S. Yuan and A. Haug. FEBS Lett.. 234 (1988) 218. [75] G. Weber. S. Smrlata and M. Rholam. Biochemistry. 23 (1984) 6785. [76] S. Scarlau. M. Rholam and G. Weber. Biochemistry, 23 ( 1984) 6789. [77] S.R. Martin and RM. Bayley, Biochem. J. 238 (1986) 485. [78] S.N. Levine. G.B. Sonnier and K. Abroe, Toxicol.. 65 (1990) 137. [79] H.C. Yang. M.M. Reedy, C.L. Burke and GM. Strasburg. Biochemistry. 33 (1994) 518. - "Illllllllllllllll