Abstract
Control of cell volume and osmolarity is central to cellular homeostasis in all eukaryotes. It lies at the heart of the century-old problem of how plants regulate turgor, mineral and water transport. Plants use strongly electrogenic H+-ATPases, and the substantial membrane voltages they foster, to drive solute accumulation and generate turgor pressure for cell expansion. Vesicle traffic adds membrane surface and contributes to wall remodelling as the cell grows. Although a balance between vesicle traffic and ion transport is essential for cell turgor and volume control, the mechanisms coordinating these processes have remained obscure. Recent discoveries have now uncovered interactions between conserved subsets of soluble N-ethylmaleimide-sensitive factor attachment protein receptor (SNARE) proteins that drive the final steps in secretory vesicle traffic and ion channels that mediate in inorganic solute uptake. These findings establish the core of molecular links, previously unanticipated, that coordinate cellular homeostasis and cell expansion.
Keywords: plant cell turgor, volume control, K+ channels, voltage-dependent, SNARE protein, Sec1-Munc18 protein, secretion
Trends
Vesicle trafficking (SNARE) proteins ‘commandeer’ the voltage sensor domains of Kv channels to confer a voltage dependence on secretory traffic for coordination with ion transport during cell expansion.
Sec1/Munc18 (SM) protein-mediated regulation of secretion is selective among plasma membrane SNAREs.
SM proteins and Kv channels bind the SNARE SYP121 at overlapping sites, implying a sequential interplay between these proteins to coordinate membrane traffic and transport.
How Do Plants Regulate Cell Turgor and Volume?
It is crucial for plants that osmolality, hydrostatic pressure, or turgor, and membrane tension are maintained in fine balance. The plant cell is encased by a cell wall, unable to migrate over the body of the organism, unlike the case in animal development, such that plant growth and development are entirely dependent on cell expansion. Cell expansion is also vital for plant movements, including those of stomatal guard cells, which support gas exchange for photosynthesis and control evaporative water loss. With few exceptions, expansion must be coupled with the deposition of new wall material as the cell wall is remodelled and, in all cases, the associated increase in cell volume must be coordinated with secretory traffic that adds new membrane surface as well as extracellular material to remodel the cell wall 1, 2.
Rates of cell expansion must also be matched by solute transport and its accumulation as the cell volume increases. Cell volume affects molecular crowding within the cytosol and intracellular ionic strength 3, 4. If uncompensated, osmotic turgor pressure, which is a primary driver for continued growth, will decline as the cell expands 5, 6. Plant cells synthesise organic solutes and take up substantial quantities of inorganic salts, especially of K+, to sustain osmotic pressures against the cell wall. Therefore, from a practical point of view, understanding how ion transport and secretory traffic are integrated in cell growth is bound to address the question that has puzzled plant biologists for over a century: how do plants regulate cell volume and turgor?
Much circumstantial evidence connects membrane traffic with cell volume, osmotic pressure, and ion transport. Decreasing the osmolality outside many plant cells stimulates exocytosis and increases cell volume, while increasing the osmolality outside stimulates endocytosis and a decrease in cell volume 7, 8, 9, 10, 11. The growth of filamentous fungi, root hairs, and pollen tubes are similarly affected by the osmotic gradient across the plasma membrane and affect ion transport activities 5, 6, 12, 13, 14. These and other studies support a model that links cell volume and turgor with membrane traffic and ion transport, but they offer few clues to the underlying mechanisms by which these processes are connected. Here, we revisit the question of how plants coordinate membrane traffic with ion transport, with a focus on recent discoveries that shed new light on these connections.
A Diversity of Functions for SNARE Proteins in Plants
The SNARE (see Glossary) superfamily of proteins were first characterised during the 1980s and later recognised as key mediators of secretory traffic in yeast, animals, and plants 15, 16, 17, 18. SNARE proteins incorporate conserved motifs that selectively assemble highly stable, coiled-coil structures to drive membrane fusion (Figure 1). At first glance, plant genomes appear to incorporate substantial redundancy: the arabidopsis (Arabidopsis thaliana) genome encodes 54 SNARE proteins compared with the 24 SNAREs of Saccharomyces cerevisiae and 36 SNAREs of humans 19, 20, 21, 22. Much gene duplication is thought to reflect a diversification in functions that were necessary for plants to colonise land 23, 24. Plant SNAREs align vesicle traffic with several cellular processes, including gravitropism, pathogen defence, symbiosis, abiotic stress responses, and, of course, cell expansion. Such diversification may relate to the unusual interactions that integrate SNARE activity within the physiological processes they support in plants.
Figure 1.
The SNARE Cycle. SNAREs assemble a tetrameric coiled-coil around a central core of three glutamine (Qa, Qb, and Qc) and one arginine (R) residues, in 1:1:1:1 stoichiometry, to draw vesicle and target membranes together for fusion. Each motif is defined by the amino acid at the core of the interacting hydrophobic layers between the respective peptide coils. In some cases, the Qb- and Qc-SNARE motifs reside on a single polypeptide, as is the case for the arabidopsis SNAP33 protein, but, more commonly, each motif in a SNARE complex is contributed by a separate protein 24, 100, 101, 102, 103. (A) Binding of the membrane- and vesicle-localised SNAREs in a cognate SNARE core complex involves priming of the Qa-SNARE at the target membrane. The Qa-SNARE transits from its ‘closed’ to ‘open’ conformation through the unfurling of the Habc domain, which exposes the Qa-SNARE domain for binding with the Qb- and Qc-SNAREs. The Qa-, Qb-, and Qc-SNAREs when bound together form the ‘acceptor’ complex, which is available for binding to the R-SNARE. (B) Formation of the trans-SNARE complex following interactions between the R-SNARE on the vesicle and the ‘acceptor’ complex at the target membrane draws the secretory vesicle to the target membrane for fusion. (C) Formation of a stabilised Qa-, Qb-, Qc-, and R-SNARE core complex overcomes the hydration energy barrier of the membrane surfaces and leads to fusion of the secretory vesicle with the target membrane and release of the vesicle cargo. (D) Following vesicle fusion, cis-SNARE complex disassembly allows for recycling of the SNARE proteins and sustains further secretory traffic. Abbreviations: SNARE, soluble N-ethylmaleimide-sensitive factor attachment protein receptor; TM, transmembrane.
The Qa-SNAREs, or syntaxins, are a case in point. In animals, these proteins lend combinatorial specificity to SNARE interactions for vesicle targeting and secretory membrane traffic, for example as dedicated to neurotransmitter release and insulin-induced GLUT4 traffic 25, 26. In arabidopsis, they have diversified further and engage in specialist roles that differentiate between abiotic and biotic stress responses, and even between responses to bacterial and fungal pathogens. Among the 19 arabidopsis Qa-SNAREs, the nine members of the SYP1 subgroup are evolutionarily related and associate with secretory traffic at the plasma membrane. SYP111 (=KNOLLE), is specialised for homotypic vesicle fusion late in cell division, which leads to the formation of the cell plate [20]; SYP121 (=SYR1/PEN1), SYP122, and SYP132 are expressed throughout the vegetative plant and contribute to abiotic and biotic stress responses 24, 27, 28; and SYP124 and SYP125 are expressed in pollen and are essential for pollen tube growth during fertilisation [29].
Of these Qa-SNAREs, SYP121 and SYP122 are close structural homologues, sharing 64% amino acid identity and 76% similarity across some 310 residues. Initially, these Qa-SNAREs were thought to be functionally redundant. Both are expressed throughout the plant; single mutation of either has little effect on plant growth under many conditions; and growth is greatly reduced only in the syp121/syp122 double mutant 30, 31. Both SYP121 and SYP122 assemble with the Qbc-SNARE SNAP33 and two near-identical R-SNAREs, VAMP721 and VAMP722, to form SNARE core complexes that drive vesicle fusion and secretion at the plasma membrane 18, 32, 33. However, SYP121 is uniquely important for abiotic stress resistance and stomatal movements; it contributes to the delivery and recycling of K+ channels and to long-term photosynthetic capacity under water stress 27, 34, 35; and it may be specialised for fungal pathogen defence [28]. SYP121, and its tobacco (Nicotiana tabacum) homologue NtSYR1, were first identified in association with the response to the water stress hormone abscisic acid (ABA) and its regulation of ion channels in the guard cells surrounding the stomatal pore of leaves 34, 36. It was recognised early on that the role of SYP121 in trafficking per se was not sufficient to explain much of the cellular physiology associated with abiotic stress 34, 35, 37. Similarly, VAMP721 and VAMP722 were previously assumed to be functionally redundant, the double vamp721/vamp722 mutant only exhibiting an embryo-lethal phenotype 29, 38. However, new evidence indicates a specialisation for VAMP721 that aligns with that of SYP121. As we note below, these SNAREs engage directly with a subset of K+ channels, and this specialisation is almost certainly a consequence of these interactions.
Solute Transport, Turgor, and Cell Volume
Plants energise the plasma membrane using H+-ATPases to pump H+ across the membrane, generating a substantial membrane voltage, inside negative, as well as a pH gradient, inside alkaline. Membrane voltage often comprises some 50% of the electrochemical gradient for H+, or ΔμH, across the plasma membrane. Voltage is used to drive the transport of other solutes, either together with the pH gradient (for transport coupled to H+ return across the membrane) or by itself [39]. K+ is the single, most prevalent, inorganic solute in the plant cell. Plants maintain K+ concentrations near 100 mM in the cytosol, even when soil K+ concentrations are 103- to 104-fold lower and often highly variable 40, 41. K+ uptake, along with that of NO3– and Cl–, is widely recognised to predominate in osmotic solute accumulation for turgor that powers cell expansion in plants. Indeed, in most agricultural soils, nutrient flux is dominated overwhelmingly by K+ and NO3− 40, 42, 43.
Electrophysiological studies of a variety of plant cells have shown that the dominant transporters – pumps, H+-coupled carriers, and channels – are generally present on the plasma membrane. They can be engaged to carry solute across the membrane, but are often kinetically limited by substrate (ion) concentration and, most importantly, by membrane voltage. Voltage controls the K+ channels that account for the bulk of K+ uptake under most physiological conditions, and it is also a key factor determining ion uptake through other transporters [40]. This situation contrasts with the common misconception that transporters are either present or absent from the membrane in response to various stimuli, including nutrient availability (Box 1). One inference to be drawn from these observations is that K+ flux is unlikely ever to be zero. Instead, K+ transport is expected to oscillate between uptake and loss over periods of tens of seconds to many minutes. Therefore, dynamic control of solute content is most likely achieved by ‘time-averaging’ net solute flux under control of the membrane voltage. The same applies to the other major osmotic solutes and the interpretation fits closely with observations of voltage and solute flux oscillations 14, 44, 45, 46, 47, with quantitative systems models of transport 44, 48, 49, 50, and with secretion and growth rate in root hairs and pollen, where these processes are easily quantified [14]. In short, membrane hyperpolarisation is essential to engage net solute uptake, its depolarisation facilitates net solute loss, and the balance between these two states of the membrane is connected to cell growth.
Box 1. Membrane Voltage Determines the Balance between Net Solute Uptake and Efflux.
The membrane of plant cells typically exhibits two quasi-stable states, one hyperpolarised often to values negative of –100 mV and one depolarised and positive of the equilibrium voltage for K+ (EK). Membrane voltage when free running – not voltage clamped – is determined by the balance of current circulation through all charge-carrying transporters, such that the sum of all currents is always zero. Electrophysiologists commonly use current-voltage (IV) curves, such as the two shown in Figure I in the central panel (black, hyperpolarised; grey, depolarised), to describe current across the membrane. The IV curves shown were generated using OnGuard 49, 50, 104 and are typical of IV curves of arabidopsis guard cells and root epidermal cells. Experimentally, such IV curves are derived from measurements in which the membrane is clamped across a range of voltages and clamp current at each voltage is recorded [105]. Each curve is the sum of current through all transporters in the membrane. The point at which a curve crosses the voltage axis marks the free-running voltage; away from this point, there is either a net negative (inward) or positive (outward) current. Typical voltage ranges of the two quasi-stable states of the membrane are indicated by the green shading along the voltage axis.
The total membrane IV curve can be separated into the component IV curves for each transporter. Experimentally, separating transporter currents often relies on challenge with (ant-)agonists and, for voltage-gated channels, on the comparatively long timescales needed for channel gating to respond to changes in clamp voltage [105]. Each of the total membrane IV curves in Figure I is broken down to show the major components, namely the H+-ATPase (long-broken line), the inward- (dotted line) and outward-rectifying (broken-dotted line) K+ channel, and the Cl– channel (short-broken line) currents. Current amplitudes and directions at the free-running voltage (where total membrane IV curve crosses the voltage axis) are indicated by the coloured arrows to the side in each case. Such analysis shows that several transporters are kinetically limited by voltage, notably the two K+ channels. It also shows that these transporters remain functional, even when their activity is constrained by voltage. K+ current through inward- (for uptake, primarily KAT1 in guard cells) and outward-rectifying (for efflux, primarily GORK) K+ channels is evident under voltage clamp, even when K+ is taken up during growth 55, 60, 106. Here, the outward-rectifying channels are evident as the substantial positive current near and positive of 0 mV in both of the total membrane IV curves. The main difference between the black and grey IV curves is a greater H+-ATPase and smaller Cl– channel current in the hyperpolarised state (black IV curve). Also evident in this state is a larger inward-rectifying K+ channel current. These variations in transport capacity arise from post-translational regulation by cytosolic-free [Ca2+]. Thus, the dynamic range of voltages between the two states is dictated primarily by the two K+ channels and the voltage dependence of their activities.
Of course, translational and post-translational controls introduce further regulatory controls on solute transport. The voltage range over which many ion channels activate, often referred to as ‘gating’, may be affected, for example, by Ca2+ binding, nitric oxide, and protein (de-)phosphorylation 48, 51, 52, 53. Voltage-dependent gating is best described in the whole cell by its midpoint voltage (V1/2) at which half-maximal activity (conductance) is achieved. The actions of these several signals may displace V1/2 and, hence, the apparent threshold for channel activation, by 20-40 mV or more. Given that the plant cell membrane typically varies across a relatively narrow range of voltages, approximately between –180 mV and –50 mV, such changes in V1/2 can have significant effects on the time-averaged net rates of ion flux. Channel activity also depends on the population of K+ channels present at the membrane, although, in the first instance, such changes will affect the maximum conductance of the population rather than V1/2 8, 27, 35, 37. Only over longer timescales may translational control alter the distribution between channel subpopulations 54, 55. Even so, these additional controls leave unresolved the most fundamental questions about the mechanisms by which solute uptake is regulated dynamically with membrane traffic as the cell expands.
SNARE Binding Affects the Voltage Sensitivity of Kv Channels
The first clues to a mechanism coordinating vesicle traffic and ion transport appeared in 2009 with the publication by Honsbein et al. [56] of a direct, physical interaction between the Qa-SNARE SYP121 and the Kv channel KC1. Subsequent work showed that SYP121 also interacts with a second Kv channel, KAT1, a dominant inward-rectifying K+ channel in guard cells and several other cell types [57]. Channel binding in each case was specific for SYP121, but not for SYP122, and binding occurred at the plasma membrane independent of any traffic of the channel to the membrane. KC1 is a close homologue of KAT1, and assembles with another inward-rectifying K+ channel subunit, AKT1, as part of a major pathway for K+ uptake at the root epidermis [54]. By virtue of its heteromeric assembly, these channels activate only near –200 mV, the voltage extreme of the plant cell, but on interacting with SYP121, the channels activated at voltages approximately 70–100 mV more positive 56, 58. Thus, SYP121 interaction with KC1 is essential to displace V1/2 for the KC1-AKT1 channel into the physiological range of voltages normally observed in the root epidermis to facilitate K+ uptake. In support of this interpretation, the mutant syp121, but not syp122, phenocopied the akt1 and kc1 mutants, showing little measurable inward-rectifying K+ current and failed to grow at submillimolar K+ concentrations when channel-mediated K+ uptake was limiting.
Subsequent studies [58] resolved the binding site on SYP121 as the linear motif F9xRF close to the cytosolic N terminus of the Qa-SNARE and confirmed the ability of SYP121 to complement the syp121 mutant. Intriguingly, the cognate R-SNARE VAMP721, which is localised to the vesicle membrane before fusion, also binds and modulates the activity of the K+ channels. Zhang et al. [59] found that VAMP721 binds both the KC1 and KAT1 channel subunits via a site centred around Tyr57 within the longin domain of the R-SNARE and separate from the core α-helix that assembles in the SNARE complex. Binding displaced the V1/2 for gating to more negative voltages and reduced the maximum conductance, both in K+ channels assembled of KC1 and AKT1 subunits and in the KAT1 channels when expressed heterologously. Overexpressing VAMP721 in arabidopsis also suppressed the inward-rectifying K+ current of the root epidermis. From the viewpoint of the capacity for K+ uptake, these actions effectively counter those of the Qa-SNARE: it is as though channel activity were coordinated in concert with the respective SNARE binding partner, and it begs the question ‘Could K+ flux be temporally coupled with a binding exchange between cognate SNAREs during the process of vesicle fusion?’.
Commandeering Kv Channel Voltage Sensors for Voltage-Sensitive Secretory Traffic
If these findings show that SYP121 and VAMP721 bind in multimeric complexes to control channel-mediated K+ uptake, they leave unresolved the consequences for membrane traffic itself. How does SYP121 bind the channels and does channel binding also promote secretory traffic, as it does channel activity? Given the cytosolic location of the F9xRF motif of SYP121, it comes as no surprise that the complementary binding site on the Kv channels resides on the cytosolic side of the membrane. However, what was unexpected was the isolation of the critical binding motif RYxxWE to the VSD [60], well removed from the channel pore and adjacent the base of the S1 α-helix. Voltage-gated K+ channels of plants (including KC1, AKT1, and KAT1) belong to the superfamily of Kv channels that are found across phyla. Kv channels typically comprise six membrane-spanning α-helices. The first four α-helices (S1–S4) form a semiautonomous voltage-sensor domain (VSD), and the remaining two α-helices (S5 and S6) line the channel pore 61, 62. The VSDs of these channels incorporate a series of fixed positive charges that transduce changes in voltage across the membrane, driving highly conserved conformational changes that move the VSD part-way across the membrane and open the channel pore 61, 63, 64, 65.
The consequences for secretory traffic were equally surprising. Overexpressing the KC1 channel or the VSD alone was sufficient to rescue traffic in the presence of the dominant-negative Qa-SNARE peptides SYP121ΔC and SYP122ΔC (which normally block traffic to the plasma membrane), but only in arabidopsis expressing SYP121. These findings showed that the critical interaction occurs at the plasma membrane with the native Qa-SNARE. Furthermore, VSD expression on its own promoted secretory traffic, but only if the VSD achieved the open channel conformation or could be driven to this conformation by voltage: thus, a mutant VSD that effectively ‘locked’ the voltage sensor in a conformation corresponding to an open channel promoted traffic, whereas a mutant that locked the VSD in the conformation corresponding to the closed channel did not [63]. The findings demonstrate that channel binding facilitates secretory traffic in a voltage-dependent manner. In effect, the Qa-SNARE commandeers the channel VSD to ‘sense’ the voltage bias for transport and adjust the rates of secretory traffic in parallel with solute uptake.
Such a close juxtaposition of transport with vesicle traffic also underpins the complementary actions of the dominant-negative peptides of the Qa-SNARE and K+ channel. Expressing the cytosolic SYP121ΔC peptide was observed in early experiments to block secretory traffic, most likely because the peptide retains the ability to bind its cognate SNARE partners, but is unable to drive vesicle fusion without the C-terminal transmembrane domain 33, 66, 67. Unexpectedly, SYP121ΔC was found also to lead to a substantial rise in cellular K+ and osmotic content 59, 67, 68. As noted above, expressing the channel VSD promotes secretory traffic; it also suppresses K+ uptake and osmotic content [59]. A simple and overarching explanation lies in the coordinate stimulation of vesicle traffic and ion transport mediated by the interacting partners. SYP121ΔC retains the capacity for K+ channel binding [59] and, thereby, for activation of the K+ channels; therefore, SYP121ΔC can be expected both to compete with the native SYP121 for cognate SNARE partners, thus blocking traffic, and to promote K+ uptake through its binding with the K+ channels at the plasma membrane. Similarly, the VSD retains the ability to bind SYP121, but it does not conduct K+ across the membrane. So, VSD binding can be expected both to promote vesicle traffic through its interaction with SYP121 and, by competing with the native K+ channels for Qa-SNARE binding, to suppress SNARE-mediated stimulation of K+ uptake (Figure 2). In short, these experiments elegantly demonstrate that the protein fragments, which retain the corresponding partner-binding domains but only partial functionality, are each able to uncouple secretory traffic from solute transport with predictable and complementary effects.
Figure 2.
Complementarity of SYP121ΔC and the Kv Channel Voltage-Sensor Domain Is A Predictable Consequence of Their Dual Functionalities. K+ channels are shown with orange pore domains and purple voltage-sensor domains (VSDs). SYP121 is shown in red with membrane anchor, longer Qa-SNARE domain, and shorter Habc domain; the dominant-negative SYP121ΔC is shown lacking the membrane anchor; VAMP721 is shown in blue with a membrane anchor, longer R-SNARE domain and a shorter longin domain. The FxRF motif of SYP121 with the VSD is shown as a circle at the SYP121 N terminus. Other SNARE proteins have been omitted for clarity. Comparison of the relative rates for vesicle traffic and solute influx are represented by bar graphs (right). (A) In the absence of either dominant-negative fragment, SNARE binding with the K+ channels facilitates the activation and balance of both processes. (B) Overexpressing the channel VSD facilitates SNARE-mediated traffic, but competes with the native channels for SNARE binding, thereby suppressing K+ influx and its accumulation. (C) Overexpressing the SYP121ΔC fragment facilitates K+ flux by binding and activating the K+ channels, but the SYP121ΔC fragment competes with the native SNARE at the plasma membrane for assembly in the SNARE complex needed to drive vesicle fusion. Thus, in each case, the dominant-negative fragment uncouples vesicle traffic and K+ uptake with complementary effects on the two processes. Abbreviation: SNARE, soluble N-ethylmaleimide-sensitive factor attachment protein receptor.
A Molecular Governor for the Control of Cell Expansion
In principle, vesicle traffic to and from the plasma membrane may affect the population of ion channels and other transporters, thereby altering the capacity for ion transport. So, why should coordinating vesicle traffic and ion transport directly, independent of such traffic, be so important? During any one day, more than twice the total water content of a plant, along with a substantial quantity of solutes, passes from root to shoot, driven by evaporative demand 69, 70, 71. Variations in water and solute flux over the diurnal cycle may vary rapidly, they affect cellular water potentials, and, consequently, turgor, with direct consequences for cell growth; for example, cell expansion rates in leaves vary in inverse proportion with transpiration and in direct proportion with leaf hydration 70, 72. Clearly, close and dynamic control of membrane transport, solute accumulation, and water flux is essential at the cellular level to integrate these activities with the variations in water and solute availability throughout the hydraulic network of the plant. Most importantly, this control needs to be coupled to vesicle traffic to balance solute transport with the rates at which new membrane surface is added as the cell volume increases.
Direct, mechanical coupling between key components of the vesicle trafficking machinery and a dominant subset of solute transporters is one answer to this need. Such a basal level of control may also be seen as an energy-efficient solution. It does not require a costly overhead in regulatory feedback networks or additional transcription, translation, or protein turnover in SNARE components and membrane transporters. Indeed, in plants, for which voltage comprises a significant fraction of the energetic charge on the cell membrane, simply modulating the population of transporters, such as K+ channels, through vesicle traffic and channel turnover is an inefficient way to control the rates of solute uptake. Both experimental evidence and quantitative models have shown that the benefits of altering the number of functional channels at the plasma membrane are offset by compensatory changes in voltage itself 60, 73, 74. Again, these and other studies point to voltage itself as the most efficient point of control for traffic, as it is for ion transport (see the section ‘Solute Transport, Turgor, and Cell Volume’ and Box 1).
In effect, membrane voltage is both a driver for, and is responsive to, all charge-carrying ion transport across the plasma membrane. Membrane hyperpolarisation is usually driven by activation of H+-ATPases and is generally important to accelerate solute uptake. Therefore, it makes good biological sense for the plant cell to use voltage as a proxy for transport activity. The voltage proxy is especially important when transport is expected to oscillate with voltage and dynamic control of solute content is achieved by ‘time-averaging’ net flux. Given that channel VSDs respond to transmembrane voltage by undergoing substantial conformational changes, the conformational status of the VSD provides an excellent running measure of voltage with a resolution timeframe of milliseconds. Similar to James Watt's application of the mechanical governor 75, 76, SNARE-VSD binding serves as a molecular governor, a coupling that enables mutual coordination of the two processes so that each facilitates and tempers the activity of the other.
How VSD conformation is coupled to the Qa-SNARE remains a big question. Conformational changes of Kv and other channel VSDs within membranes are known to affect the exposure of significant portions of VSD structure to the cytosolic side of the membrane 61, 62. Thus, with the RYxxWE motif of the channel positioned close to the cytosolic surface of the membrane [59], its availability for SNARE binding may depend on VSD conformation (Figure 3). Conceivably, then, if binding and debinding are sufficiently rapid, SNARE binding may simply report the time-averaged conformation of the VSD. Equally plausible, however, SNARE binding may be retained, VSD conformation being transduced through the bound SNARE to affect vesicle traffic. In either of these scenarios, binding may also prolong the open state of the (inward-rectifying) channel or reduce the time it remains in the closed state, thereby promoting K+ influx. Almost certainly, together with these processes, we also need to consider an exchange of channel binding between cognate SNAREs [59] and how this might be coordinated during vesicle fusion.
Figure 3.
Voltage Alters the Exposure of the Kv Channel Voltage Sensor to the Cytosol. Side views of the voltage sensor domain (VSD) of the KAT1K+ channel are shown in the ‘up’ (A) and ‘down’ (B) conformations. For clarity, the membrane bilayer is not shown and only water molecules (light blue) on either side (inside and outside) of the membrane are visible. Transmembrane α-helices are colour coded in green (S1), black (S2), red (S3), and yellow (S4). Residues RYxxWE that form the binding site for SYP121 are included in blue as stick representations. The molecular dynamics of each conformation were modelled using as a template the entire α-subunit of Kv1.2 obtained by the Rosetta method 63, 107. Each model was analysed after embedding within a lipid bilayer in a periodic boundary condition box with water molecules, K+ and Cl– ions, and was optimised using energy minimisation and equilibration at 298 K for 5 ns with a harmonic restraint of 0.5 kcal/mol Å2 applied to the backbone atom using NAMD [108]. As the membrane is hyperpolarised, movement of the S4 domain within the transmembrane electric field drives the transition to the ‘down’ conformation and is coupled to rotation of the S5 and S6 α-helices to open the channel pore (not shown). Here, the ‘down’ conformation corresponds to the open channel. On the basis of simulations [63], it appears also to promote the aqueous exposure of the RYxxWE motif at the cytosolic surface of the membrane. Images created with VMD [109].
Temporal Coordination of Vesicle Traffic with Ion Transport: A Role for Sec1/Munc18 Proteins?
One clue to a possible exchange in channel binding between SNAREs comes from the basic structure of SYP121. Qa-SNAREs, including SYP121, incorporate a C-terminal membrane anchor preceded by a cytosolic α-helical (H3) domain that comprises the Qa-SNARE motif and assembles with cognate SNAREs in the core complex. At the N terminus of the H3 domain are a set of three antiparallel α-helices, the Habc domain, and an unstructured N-terminal sequence of 20–40 residues. The Habc domain folds back on the H3 α-helix to prevent access to that α-helix, rendering the SNARE inactive in a so-called ‘closed’ conformation 21, 24, 77. Priming of the Qa-SNARE involves unfurling of the Habc domain to expose the Qa-SNARE motif in a fusion competent conformation (Figure 1).
For many Qa-SNAREs, passage through this conformational switch depends on binding and debinding of one of the Sec1/Munc18 (SM) family of proteins 33, 78. SM proteins form a clothes-peg-shaped structure with a major cleft that binds Qa-SNAREs in the closed (inactive) conformation, stabilising the protein against promiscuous interactions. They also show at least two other modes of binding. One of these modes engages the Qa-SNARE N terminus, binding it within a minor cleft on the SM surface, and the other involves ‘clamping’ of the SM major cleft over the assembled SNARE complex to stabilise it 79, 80. Binding of the Qa-SNARE N terminus within the SM minor cleft is thought to spatially coordinate and facilitate the unfurling of the Qa-SNARE, and it affects the speed of fusion 80, 81, 82. Even so, what drives the conformational switch between binding modes is unclear and is almost certain to depend on ancillary interactions. Recent studies highlighted the binding of Munc13 with the binary Qa-SNARE-SM pair of Syn1A and Munc18 to accelerate SNARE complex assembly and neurotransmitter release [81]. Munc13, and its homolog Unc13 of Caenorhabditis elegans [83], are large, Ca2+- and phospholipid-binding proteins that have some similarities to proteins of the exocyst complex involved in vesicle tethering before fusion 84, 85, 86.
The arabidopsis SM protein SEC11 (=KEULE) was originally identified as a factor supporting Qa-SNARE SYP111 activity during cell plate formation [87]. SEC11 also binds SYP121 at the plasma membrane and affects its assembly with VAMP721 and SNAP33 as well as secretory traffic [88]. Arabidopsis harbours six proteins with weak homologies to Munc13, but only PATROL1 functions in vesicle traffic in vegetative tissues [89]. The other five gene products are pollen specific 90, 91 or are isolated to chloroplasts and mitochondria [92]. Significantly, PATROL1 affects H+-ATPase traffic to the plasma membrane, which is independent of SYP121 and SYP122 [37] and does not affect their associated cargos 24, 37, 89, 93. Thus, PATROL1 targets a trafficking pathway that is distinct from that of SYP121 and SEC11.
How, then, does SEC11 disengage from SYP121 to activate the Qa-SNARE? Similar to other SM proteins, SEC11 binding depends on the N terminus of the Qa-SNARE [94], and especially on a highly -conserved Phe9 residue present in SYP121 to promote secretory traffic [33]. Here, the Qa-SNARE interacts with the minor cleft and Leu128 near the N-terminal surface of SEC11 33, 88. Thus, the N-terminal site for SEC11 binding overlaps with the F9xRF motif that is essential for Kv channel binding [58]. Furthermore, only SYP121 and the SNARE complexes that it assembles will bind SEC11 at the plasma membrane 33, 88. It has been found that the channel-binding domain peptide KC1Δ92-677, which includes the RYxxWE motif, affects SEC11 binding to SYP121 as well as that of the cognate SNAREs in complex (S. Waghmare, R. Karnik and M.R. Blatt, unpublished, 2016), suggesting both a role for the Kv channel in SEC11 binding as well as a possible sequence involving the channel in a ‘handover’ between the Qa- and R-SNARE (Figure 4). In short, these findings are consistent with the idea that the Kv channel VSD is key to supporting secretory traffic in vivo through its action in promoting SYP121-mediated SNARE complex assembly.
Figure 4.
A Model for Kv Channel-Binding Exchange between VAMP721 and SYP121. K+ channels are shown with orange pore domains and purple voltage-sensor domains (VSDs). SYP121 is shown in red with membrane anchor, longer Qa-SNARE domain, and shorter Habc domain with the FxRF motif shown as a circle and Phe9 highlighted; VAMP721 is shown in blue with membrane anchor, longer R-SNARE domain, and shorter longin domain with Tyr57 indicated. SEC11 is depicted in green with the major cleft formed by the black line and residue Leu128, important for SEC11 interactions with SYP121 33, 88, is indicated near the N-terminal surface. Other SNARE proteins have been omitted for clarity. Voltage is denoted by charges (+ and –), with membrane hyperpolarisation indicated with increased negative charge inside the cell. (A) SYP121 in its ‘closed’ conformation is stabilised by SEC11 bound to the SYP121 N terminus with the interaction with Leu128 in the SEC11 minor cleft essential for unfurling of the SYP121 Habc domain. (B,C) VAMP721 on the secretory vesicle binds the closed K+ channel to help recruit the vesicle to the plasma membrane. (D) VAMP721 and the open conformation of SYP121 exchange channel binding, facilitated by membrane hyperpolarisation and channel opening with VSD in the ‘down’ state, assembling the trans-SNARE complex, which is stabilised by SEC11. (E) SEC11 facilitates release from channel and disassembly of the cis-SNARE complex. (F) SNARE complex disassembles to recycle the cognate SNAREs for subsequent fusion events. Abbreviation: SNARE, soluble N-ethylmaleimide-sensitive factor attachment protein receptor.
Concluding Remarks: A Century-Old Problem
Rates of cell expansion must be matched by solute transport and its accumulation as the cell volume increases. For the plant cell, it is vital that transport of inorganic solutes is maintained to provide osmotic pressure against the cell wall and, concurrently, that this process is controlled together with that of secretory traffic that delivers new membrane and cell wall material as the cell expands. Such coordination is intimately connected to mineral and water use by plants; its failure is associated with drought and salt sensitivities, with physiopathologies and diseases that reflect a loss in the control of cell volume and osmotic homeostasis, and with plant water relations and vegetative growth 67, 95, 96.
How widespread, then, are SNARE–channel interactions that might couple through membrane voltage in plants? Surveys of extant plant genomes 57, 60, 97 show that the F9xRF motif is closely conserved within a subset of plasma membrane Qa-SNAREs in vascular land plants, but not in bryophytes; neither is it found in animal Qa-SNAREs. The RYxxWE motif is similarly conserved within a subset of plant Kv channel VSDs in all vascular plant genomes [60], implying its early coevolution coincident with the expansion of SNAREs when plants colonised land [23]. Functional evidence to date is restricted to arabidopsis and tobacco, and there is much of SNARE protein function in plants that is still to be explored. Of interest, the same SNAREs also interact with a subset of plasma membrane-localised aquaporins 93, 98, suggesting functional controls that also extend to water flux rates across this membrane. It is likely that similar associations will surface in other vascular plants. What sets SNARE–-channel interactions in plants apart from the animal models [99] is the importance of regulating of membrane traffic and ion transport in parallel, with a basal mechanism of control through shared coupling to VSD conformation. Voltage is intimately connected with the osmotic solute uptake in plants. Its connection to membrane traffic and cell volume leaves little doubt now that this direct, mechanistic link lies at the heart of the century-old problem of how plants regulate turgor and cellular volume. From a practical point of view alone, understanding the mechanisms behind such coordination is bound to inform strategies for crop improvement in the face of global environmental change (see Outstanding Questions).
Outstanding Questions.
What is the sequence of binding and debinding events that leads to secretory vesicle fusion in association with K+ channel activation?
Does channel binding recruit SM protein binding or compete with it during SNARE complex assembly?
Do pathogens or symbionts take direct advantage of SNARE-channel interactions during their association with the plant host?
How might SNARE-channel interactions be manipulated to enhance growth, water, and mineral use efficiency by the plant?
Figure I.
Total Membrane IV Curves and Their Decomposition.
Acknowledgements
The authors acknowledge funding from the UK Biotechnology and Biological Sciences Research Council, including grants BB/I024496/1, BB/K015893/1, BB/L001276/1, BB/M01133X/1, BB/M001601/1, and BB/L019205/1. Past work by the authors was supported by grants from the Chinese Scholarship Council, the Royal Society, the Gatsby Charitable Foundation, and the John Simon Guggenheim Memorial Foundation.
Glossary
- Gating
a process of ion channel opening and closing, often responding to membrane voltage, ligand, or protein binding and debinding.
- Kv channel
members of the gene superfamily of voltage-gated K+ channels that facilitate K+ flux across biological membranes; Kv channels assemble from four subunits, each with six transmembrane-spanning α-helices.
- Sec1/Munc18 (SM) proteins
soluble proteins that bind SNAREs, regulating their assembly and stabilising the SNARE complex.
- Soluble N-ethylmaleimide-sensitive factor attachment protein receptor (SNARE) proteins
membrane-anchored proteins that drive the final steps of membrane fusion.
- Trans- and cis-SNARE complex
refers to the SNARE complex before fusion with proteins anchored in two membranes and after vesicle fusion with all SNARE proteins anchored in a single membrane, respectively.
- V1/2
the voltage at which a population of voltage-gated ion channels achieves half-maximal activity; also known as the midpoint voltage, this parameter is a key descriptor of the gating process.
References
- 1.Thiel G., Battey N. Exocytosis in plants. Plant Mol. Biol. 1998;38:111–125. [PubMed] [Google Scholar]
- 2.Zonia L., Munnik T. Life under pressure: hydrostatic pressure in cell growth and function. Trends Plant Sci. 2007;12:90–97. doi: 10.1016/j.tplants.2007.01.006. [DOI] [PubMed] [Google Scholar]
- 3.Biemans-Oldehinkel E. A sensor for intracellular ionic strength. Proc. Nat. Acad. Sci. U.S.A. 2006;103:10624–10629. doi: 10.1073/pnas.0603871103. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 4.Spitzer J.J., Poolman B. Electrochemical structure of the crowded cytoplasm. Trends Biochem. Sci. 2005;30:536–541. doi: 10.1016/j.tibs.2005.08.002. [DOI] [PubMed] [Google Scholar]
- 5.Zonia L. Hydrodynamics and cell volume oscillations in the pollen tube apical region are integral components of the biomechanics of Nicotiana tabacum pollen tube growth. Cell Biochem. Biophys. 2006;46:209–232. doi: 10.1385/CBB:46:3:209. [DOI] [PubMed] [Google Scholar]
- 6.Mathur J. Local interactions shape plant cells. Curr. Opin. Cell Biol. 2006;18:40–46. doi: 10.1016/j.ceb.2005.12.002. [DOI] [PubMed] [Google Scholar]
- 7.Meckel T. Guard cells undergo constitutive and pressure-driven membrane turnover. Protoplasma. 2005;226:23–29. doi: 10.1007/s00709-005-0106-6. [DOI] [PubMed] [Google Scholar]
- 8.Hurst A.C. Trafficking of the plant potassium inward rectifier KAT1 in guard cell protoplasts of Vicia faba. Plant J. 2004;37:391–397. doi: 10.1046/j.1365-313x.2003.01972.x. [DOI] [PubMed] [Google Scholar]
- 9.Homann U., Thiel G. The number of K + channels in the plasma membrane of guard cell protoplasts changes in parallel with the surface area. Proc. Nat. Acad. Sci. U.S.A. 2002;99:10215–10220. doi: 10.1073/pnas.152324399. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 10.Shope J.C., Mott K.A. Membrane trafficking and osmotically induced volume changes in guard cells. J. Exp. Bot. 2006;57:4123–4131. doi: 10.1093/jxb/erl187. [DOI] [PubMed] [Google Scholar]
- 11.Shope J.C. Changes in surface area of intact guard cells are correlated with membrane internalization. Plant Physiol. 2003;133:1314–1321. doi: 10.1104/pp.103.027698. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 12.Zonia L., Munnik T. Osmotically induced cell swelling versus cell shrinking elicits specific changes in phospholipid signals in tobacco pollen tubes. Plant Physiol. 2004;134:813–823. doi: 10.1104/pp.103.029454. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 13.Kroeger J.H. Regulator or driving force? The role of turgor pressure in oscillatory plant cell growth. PLoS One. 2011;6:e18549. doi: 10.1371/journal.pone.0018549. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 14.Campanoni P., Blatt M.R. Membrane trafficking and polar growth in root hairs and pollen tubes. J. Exp. Bot. 2007;58:65–74. doi: 10.1093/jxb/erl059. [DOI] [PubMed] [Google Scholar]
- 15.Sollner T. SNAP receptors implicated in vesicle targeting and fusion. Nature. 1993;362:318–324. doi: 10.1038/362318a0. [DOI] [PubMed] [Google Scholar]
- 16.Sollner T. A protein assembly-disassembly pathway in-vitro that may correspond to sequential steps of synaptic vesicle docking, activation, and fusion. Cell (1993) 1993;75:409–418. doi: 10.1016/0092-8674(93)90376-2. [DOI] [PubMed] [Google Scholar]
- 17.Duman J.G., Forte J.G. What is the role of SNARE proteins in membrane fusion? Am. J. Physiol. Cell Physiol. 2003;285:C237–C249. doi: 10.1152/ajpcell.00091.2003. [DOI] [PubMed] [Google Scholar]
- 18.Sanderfoot A.A. The Arabidopsis genome. An abundance of soluble N- ethylmaleimide-sensitive factor adaptor protein receptors. Plant Physiol. 2000;124:1558–1569. doi: 10.1104/pp.124.4.1558. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 19.Sutter J.U. Setting SNAREs in a different wood. Traffic. 2006;7:627–638. doi: 10.1111/j.1600-0854.2006.00414.x. [DOI] [PubMed] [Google Scholar]
- 20.Lipka V. SNARE-Ware: the role of SNARE-Domain proteins in plant biology. Annu. Rev. Cell Dev. Biol. 2007;23:147–174. doi: 10.1146/annurev.cellbio.23.090506.123529. [DOI] [PubMed] [Google Scholar]
- 21.Jahn R., Scheller R.H. SNAREs - engines for membrane fusion. Nat. Rev. Mol. Cell Biol. 2006;7:631–643. doi: 10.1038/nrm2002. [DOI] [PubMed] [Google Scholar]
- 22.Hong W. SNAREs and traffic. Biochem. Biophys. Acta. 2005;1744:120–144. doi: 10.1016/j.bbamcr.2005.03.014. [DOI] [PubMed] [Google Scholar]
- 23.Sanderfoot A. Increases in the number of SNARE genes parallels the rise of multicellularity among the green plants. Plant Physiol. 2007;144:6–17. doi: 10.1104/pp.106.092973. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 24.Bassham D.C., Blatt M.R. SNAREs - cogs or coordinators in signalling and development? Plant Physiol. 2008;147:1504–1515. doi: 10.1104/pp.108.121129. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 25.Varlamov O. i-SNAREs: inhibitory SNAREs that fine-tune the specificity of membrane fusion. J. Cell Biol. 2004;164:79–88. doi: 10.1083/jcb.200307066. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 26.Paumet F. The specificity of SNARE-dependent fusion is encoded in the SNARE motif. Proc. Nat. Acad. Sci. U.S.A. 2004;101:3376–3380. doi: 10.1073/pnas.0400271101. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 27.Eisenach C. The trafficking protein SYP121 of Arabidopsis connects programmed stomatal closure and K + channel acivity with vegetative growth. Plant J. 2012;69:241–251. doi: 10.1111/j.1365-313X.2011.04786.x. [DOI] [PubMed] [Google Scholar]
- 28.Collins N.C. SNARE-protein-mediated disease resistance at the plant cell wall. Nature. 2003;425:973–977. doi: 10.1038/nature02076. [DOI] [PubMed] [Google Scholar]
- 29.El Kasmi F. SNARE complexes of different composition jointly mediate membrane fusion in Arabidopsis cytokinesis. Mol. Biol. Cell. 2013;24:1593–1601. doi: 10.1091/mbc.E13-02-0074. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 30.Assaad F.F. The PEN1 syntaxin defines a novel cellular compartment upon fungal attack and is required for the timely assembly of papillae. Mol. Biol. Cell. 2004;15:5118–5129. doi: 10.1091/mbc.E04-02-0140. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 31.Zhang Z.G. A SNARE-protein has opposing functions in penetration resistance and defence signalling pathways. Plant J. 2007;49:302–312. doi: 10.1111/j.1365-313X.2006.02961.x. [DOI] [PubMed] [Google Scholar]
- 32.Enami K. Differential expression control and polarized distribution of plasma membrane-resident SYP1 SNAREs in Arabidopsis thaliana. Plant Cell Physiol. 2009;50:280–289. doi: 10.1093/pcp/pcn197. [DOI] [PubMed] [Google Scholar]
- 33.Karnik R. Arabidopsis Sec1/Munc18 Protein SEC11 is a competitive and dynamic modulator of SNARE binding and SYP121-dependent vesicle traffic. Plant Cell. 2013;25:1368–1382. doi: 10.1105/tpc.112.108506. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 34.Leyman B. A tobacco syntaxin with a role in hormonal control of guard cell ion channels. Science. 1999;283:537–540. doi: 10.1126/science.283.5401.537. [DOI] [PubMed] [Google Scholar]
- 35.Sutter J.U. Abscisic acid triggers the endocytosis of the Arabidopsis KAT1 K+ channel and its recycling to the plasma membrane. Curr. Biol. 2007;17:1396–1402. doi: 10.1016/j.cub.2007.07.020. [DOI] [PubMed] [Google Scholar]
- 36.Leyman B. Localization and control of expression of Nt-Syr1, a tobacco SNARE protein. Plant J. 2000;24:369–381. doi: 10.1046/j.1365-313x.2000.00886.x. [DOI] [PubMed] [Google Scholar]
- 37.Sutter J.U. Selective mobility and sensitivity to SNAREs is exhibited by the Arabidopsis KAT1K + channel at the plasma membrane. Plant Cell. 2006;18:935–954. doi: 10.1105/tpc.105.038950. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 38.Kwon C. Co-option of a default secretory pathway for plant immune responses. Nature. 2008;451:835–840. doi: 10.1038/nature06545. [DOI] [PubMed] [Google Scholar]
- 39.Blatt M.R. Concepts and techniques in plant membrane physiology. In: Blatt M.R., editor. Membrane Transport in Plants. Blackwell; Oxford: 2004. pp. 1–39. [Google Scholar]
- 40.Amtmann A., Blatt M.R. Regulation of macronutrient transport. New Phytol. 2009;181:35–52. doi: 10.1111/j.1469-8137.2008.02666.x. [DOI] [PubMed] [Google Scholar]
- 41.Maathuis F.J.M., Sanders D. Mechanisms of potassium absorption by higher-plant roots. Physiol. Plant. 1996;96:158–168. [Google Scholar]
- 42.Very A-A. Molecular biology of K+ transport across the plant cell membrane: What do we learn from comparison between plant species? J. Plant Physiol. 2014;171:748–769. doi: 10.1016/j.jplph.2014.01.011. [DOI] [PubMed] [Google Scholar]
- 43.Marschner H. vol. 2. Academic Press; New York: 1995. (Mineral Nutrition of Higher Plants). [Google Scholar]
- 44.Shabala S. Oscillations in plant membrane transport: model predictions, experimental validation, and physiological implications. J. Exp. Bot. 2006;57:171–184. doi: 10.1093/jxb/erj022. [DOI] [PubMed] [Google Scholar]
- 45.Blatt M.R., Thiel G. K + channels of stomatal guard cells: bimodal control of the K + inward-rectifier evoked by auxin. Plant J. 1994;5:55–68. doi: 10.1046/j.1365-313x.1994.5010055.x. [DOI] [PubMed] [Google Scholar]
- 46.Gradmann D. Electrocoupling of ion transporters in plants. J. Membr. Biol. 1993;136:327–332. doi: 10.1007/BF00233671. [DOI] [PubMed] [Google Scholar]
- 47.Zonia L. Oscillatory chloride efflux at the pollen tube apex has a role in growth and cell volume regulation and is targeted by inositol 3,4,5,6-tetrakisphosphate. Plant Cell. 2002;14:2233–2249. [PubMed] [Google Scholar]
- 48.Minguet-Parramona C. An optimal frequency in Ca2+ oscillations for stomatal closure is an emergent property of ion transport in guard cells. Plant Physiol. 2016;170:32–45. doi: 10.1104/pp.15.01607. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 49.Hills A. OnGuard, a computational platform for quantitative kinetic modeling of guard cell physiology. Plant Physiol. 2012;159:1026–1042. doi: 10.1104/pp.112.197244. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 50.Chen Z.H. Systems dynamic modeling of the stomatal guard cell predicts emergent behaviors in transport, signaling, and volume control. Plant Physiol. 2012;159:1235–1251. doi: 10.1104/pp.112.197350. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 51.Garcia-Mata C. Nitric oxide regulates K + and Cl - channels in guard cells through a subset of abscisic acid-evoked signaling pathways. Proc. Nat. Acad. Sci. U.S.A. 2003;100:11116–11121. doi: 10.1073/pnas.1434381100. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 52.Grabov A., Blatt M.R. Parallel control of the inward-rectifier K + channel by cytosolic-free Ca 2+ and pH in Vicia guard cells. Planta. 1997;201:84–95. [Google Scholar]
- 53.Sokolovski S. Protein phosphorylation is a prerequisite for intracellular Ca 2+ release and ion channel control by nitric oxide and abscisic acid in guard cells. Plant J. 2005;43:520–529. doi: 10.1111/j.1365-313X.2005.02471.x. [DOI] [PubMed] [Google Scholar]
- 54.Duby G. AtKC1, a conditionally targeted Shaker-type subunit, regulates the activity of plant K + channels. Plant J. 2008;53:115–123. doi: 10.1111/j.1365-313X.2007.03324.x. [DOI] [PubMed] [Google Scholar]
- 55.Geiger D. Heteromeric AtKC1.AKT1 channels in Arabidopsis roots facilitate growth under K+-limiting conditions. J. Biol. Chem. 2009;284:21288–21295. doi: 10.1074/jbc.M109.017574. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 56.Honsbein A. A tripartite SNARE-K + channel complex mediates in channel-dependent K + nutrition in Arabidopsis. Plant Cell. 2009;21:2859–2877. doi: 10.1105/tpc.109.066118. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 57.Honsbein A. A molecular framework for coupling cellular volume and osmotic solute transport control. J. Exp. Bot. 2011;62:2363–2370. doi: 10.1093/jxb/erq386. [DOI] [PubMed] [Google Scholar]
- 58.Grefen C. A novel motif essential for SNARE interaction with the K + channel KC1 and channel gating in Arabidopsis. Plant Cell. 2010;22:3076–3092. doi: 10.1105/tpc.110.077768. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 59.Zhang B. The Arabidopsis R-SNARE VAMP721 interacts with KAT1 and KC1 K+ channels to moderate K+ current at the plasma membrane. Plant Cell. 2015;27:1697–1717. doi: 10.1105/tpc.15.00305. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 60.Grefen C. A vesicle-trafficking protein commandeers Kv channel voltage sensors for voltage-dependent secretion. Nat. Plants. 2015;1:15108–15119. doi: 10.1038/nplants.2015.108. [DOI] [PubMed] [Google Scholar]
- 61.Dreyer I., Blatt M.R. What makes a gate? The ins and outs of Kv-like K + channels in plants. Trends Plant Sci. 2009;14:383–390. doi: 10.1016/j.tplants.2009.04.001. [DOI] [PubMed] [Google Scholar]
- 62.Palovcak E. Evolutionary imprint of activation: The design principles of VSDs. J. Gen. Physiol. 2014;143:145–156. doi: 10.1085/jgp.201311103. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 63.Lefoulon C. Voltage-sensor transitions of the inward-rectifying K + channel KAT1 indicate a closed-to-open latching mechanism that is biased by hydration around the S4 alpha-helix. Plant Physiol. 2014;166:960–975. doi: 10.1104/pp.114.244319. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 64.Labro A.J. Molecular mechanism for depolarization-induced modulation of Kv channel closure. J. Gen. Physiol. 2012;140:481–493. doi: 10.1085/jgp.201210817. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 65.Lai H.C. The S4 voltage sensor packs against the pore domain in the KAT1 voltage-gated potassium channel. Neuron. 2005;47:395–406. doi: 10.1016/j.neuron.2005.06.019. [DOI] [PubMed] [Google Scholar]
- 66.Tyrrell M. Selective targeting of plasma membrane and tonoplast traffic by inhibitory (dominant-negative) SNARE fragments. Plant J. 2007;51:1099–1115. doi: 10.1111/j.1365-313X.2007.03206.x. [DOI] [PubMed] [Google Scholar]
- 67.Geelen D. The abscisic acid-related SNARE homolog NtSyr1 contributes to secretion and growth: Evidence from competition with its cytosolic domain. Plant Cell. 2002;14:387–406. doi: 10.1105/tpc.010328. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 68.Sokolovski S. Functional interaction of the SNARE protein NtSyp121 in Ca 2+ channel gating, Ca 2+ transients and ABA signalling of stomatal guard cells. Mol. Plant. 2008;1:347–358. doi: 10.1093/mp/ssm029. [DOI] [PubMed] [Google Scholar]
- 69.Lawson T., Blatt M.R. Stomatal size, speed, and responsiveness impact on photosynthesis and water use efficiency. Plant Physiol. 2014;164:1556–1570. doi: 10.1104/pp.114.237107. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 70.Chaumont F., Tyerman S.D. Aquaporins: Highly regulated channels controlling plant water relations. Plant Physiol. 2014;164:1600–1618. doi: 10.1104/pp.113.233791. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 71.Sack L., Holbrook N.M. Leaf hydraulics. Annu. Rev. Plant Biol. 2006:361–381. doi: 10.1146/annurev.arplant.56.032604.144141. [DOI] [PubMed] [Google Scholar]
- 72.Caldeira C.F. Circadian rhythms of hydraulic conductance and growth are enhanced by drought and improve plant performance. Nat. Commun. 2014;5:5365. doi: 10.1038/ncomms6365. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 73.Wang Y. Systems analysis of guard cell membrane transport for enhanced stomatal dynamics and water use efficiency. Plant Physiol. 2014;164:1593–1599. doi: 10.1104/pp.113.233403. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 74.Wang Y. Overexpression of plasma membrane H + -ATPase in guard cells promotes light-induced stomatal opening and enhances plant growth. Proc. Nat. Acad. Sci. U.S.A. 2014;111:533–538. doi: 10.1073/pnas.1305438111. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 75.Grefen C., Blatt M.R. SNAREs - molecular governors in signalling and development. Curr. Opin. Plant Biol. 2008;11:600–609. doi: 10.1016/j.pbi.2008.08.006. [DOI] [PubMed] [Google Scholar]
- 76.Maxwell J.C. On Governors. Proc. R. Soc. Lond. 1868;16:270–283. [Google Scholar]
- 77.Sudhof T.C., Rothman J.E. Membrane fusion: Grappling with SNARE and SM proteins. Science. 2009;323:474–477. doi: 10.1126/science.1161748. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 78.Hu S.H. Possible roles for Munc18-1 domain 3a and Syntaxin1 N-peptide and C-terminal anchor in SNARE complex formation. Proc. Nat. Acad. Sci. U.S.A. 2011;108:1040–1045. doi: 10.1073/pnas.0914906108. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 79.Dulubova I. Munc18-1 binds directly to the neuronal SNARE complex. Proc. Nat. Acad. Sci. U.S.A. 2007;104:2697–2702. doi: 10.1073/pnas.0611318104. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 80.Rizo J. Staging membrane fusion. Science. 2012;337:1300–1301. doi: 10.1126/science.1228654. [DOI] [PubMed] [Google Scholar]
- 81.Ma C. Reconstitution of the vital functions of Munc18 and Munc13 in neurotransmitter release. Science. 2013;339:421–425. doi: 10.1126/science.1230473. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 82.Baker R.W., Hughson F.M. Chaperoning SNARE assembly and disassembly. Nat. Rev. Mol. Cell Biol. 2016;17 doi: 10.1038/nrm.2016.65. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 83.Brose N. Mammalian homologs of Caenorhabditis elegans unc-13 gene define novel family of C2-domain proteins. J. Biol. Chem. 1995;270:25273–25280. doi: 10.1074/jbc.270.42.25273. [DOI] [PubMed] [Google Scholar]
- 84.Zarsky V. Exocyst complexes multiple functions in plant cells secretory pathways. Curr. Opin. Plant Biol. 2013;16:726–733. doi: 10.1016/j.pbi.2013.10.013. [DOI] [PubMed] [Google Scholar]
- 85.Zarsky V. Exocytosis and cell polarity in plants - exocyst and recycling domains. New Phytol. 2009;183:255–272. doi: 10.1111/j.1469-8137.2009.02880.x. [DOI] [PubMed] [Google Scholar]
- 86.Hala M. An exocyst complex functions in plant cell growth in Arabidopsis and tobacco. Plant Cell. 2008;20:1330–1345. doi: 10.1105/tpc.108.059105. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 87.Assaad F.F. The cytokinesis gene KEULE encodes a Sec1 protein that binds the syntaxin KNOLLE. J. Cell Biol. 2001;152:531–543. doi: 10.1083/jcb.152.3.531. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 88.Karnik R. Binding of SEC11 indicates its role in SNARE recycling after vesicle fusion and identifies two pathways for vesicular traffic to the plasma membrane. Plant Cell. 2015;27:675–694. doi: 10.1105/tpc.114.134429. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 89.Hashimoto-Sugimoto M. A Munc13-like protein in Arabidopsis mediates H+–ATPase translocation that is essential for stomatal responses. Nat. Commun. 2013;4:2215. doi: 10.1038/ncomms3215. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 90.Becker J.D. Transcriptional profiling of Arabidopsis tissues reveals the unique characteristics of the pollen transcriptome. Plant Physiol. 2003;133:713–725. doi: 10.1104/pp.103.028241. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 91.Wang Y. Transcriptome analyses show changes in gene expression to accompany pollen germination and tube growth in Arabidopsis. Plant Physiol. 2008;148:1201–1211. doi: 10.1104/pp.108.126375. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 92.Dal Bosco C. Inactivation of the chloroplast ATP synthase gamma subunit results in high non-photochemical fluorescence quenching and altered nuclear gene expression in Arabidopsis thaliana. J. Biol. Chem. 2004;279:1060–1069. doi: 10.1074/jbc.M308435200. [DOI] [PubMed] [Google Scholar]
- 93.Besserer A. Selective regulation of maize plasma membrane aquaporin trafficking and activity by the SNARE SYP121. Plant Cell. 2012;24:3463–3481. doi: 10.1105/tpc.112.101758. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 94.Park M. Sec1/Munc18 protein stabilizes fusion-competent syntaxin for membrane fusion in Arabidopsis cytokinesis. Dev. Cell. 2012;22:989–1000. doi: 10.1016/j.devcel.2012.03.002. [DOI] [PubMed] [Google Scholar]
- 95.Scheible W.R. An Arabidopsis mutant resistant to thaxtomin A, a cellulose synthesis inhibitor from Streptomyces species. Plant Cell. 2003;15:1781–1794. doi: 10.1105/tpc.013342. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 96.Siemens J. Extracellular invertase is involved in the regulation of clubroot disease in Arabidopsis thaliana. Mol. Plant Pathol. 2011;12:247–262. doi: 10.1111/j.1364-3703.2010.00667.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 97.Grefen C. Ion transport, membrane traffic and cellular volume control. Curr. Opin. Plant Biol. 2011;14:332–339. doi: 10.1016/j.pbi.2011.03.017. [DOI] [PubMed] [Google Scholar]
- 98.Hachez C. Arabidopsis SNAREs SYP61 and SYP121 coordinate the trafficking of plasma membrane aquaporin PIP2;7 to modulate the cell membrane water permeability. Plant Cell. 2014;26:3132–3147. doi: 10.1105/tpc.114.127159. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 99.Chao C.C.T. SNARE protein regulation of cardiac potassium channels and atrial natriuretic factor secretion. J. Mol. Cell. Cardiol. 2011;50:401–407. doi: 10.1016/j.yjmcc.2010.11.018. [DOI] [PubMed] [Google Scholar]
- 100.Fasshauer D. Conserved structural features of the synaptic fusion complex: SNARE proteins reclassified as Q- and R-SNAREs. Proc. Nat. Acad. Sci. U.S.A. 1998;95:15781–15786. doi: 10.1073/pnas.95.26.15781. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 101.Bock J.B. A genomic perspective on membrane compartment organization. Nature. 2001;409:839–841. doi: 10.1038/35057024. [DOI] [PubMed] [Google Scholar]
- 102.Bock J.B., Scheller R.H. SNARE proteins mediate lipid bilayer fusion. Proc. Nat. Acad. Sci. U.S.A. 1999;96:12227–12229. doi: 10.1073/pnas.96.22.12227. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 103.Pratelli R. A new catch to the SNARE. Trends Plant Sci. 2004;9:187–195. doi: 10.1016/j.tplants.2004.02.007. [DOI] [PubMed] [Google Scholar]
- 104.Wang Y. Systems dynamic modelling of a guard cell Cl - channel mutant uncovers an emergent homeostatic network regulating stomatal transpiration. Plant Physiol. 2012;160:1956–1972. doi: 10.1104/pp.112.207704. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 105.Blatt M.R. Membrane transport in plants. In: Blatt M.R., editor. Membrane Transport in Plants. Blackwell; Oxford: 2004. pp. 1–256. [Google Scholar]
- 106.Blatt M.R. Cellular signaling and volume control in stomatal movements in plants. Annu. Rev. Cell Dev. Biol. 2000;16:221–241. doi: 10.1146/annurev.cellbio.16.1.221. [DOI] [PubMed] [Google Scholar]
- 107.Gajdanowicz P. Distributed structures determine K + and voltage dependent gating of the K in channel KAT1 and the K out channel SKOR. New Phytol. 2009;182:380–391. [Google Scholar]
- 108.Phillips J.C. Scalable molecular dynamics with NAMD. J. Comput. Chem. 2005;26:1781–1802. doi: 10.1002/jcc.20289. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 109.Humphrey W. VMD: Visual molecular dynamics. J. Mol. Graph. 1996;14:33. doi: 10.1016/0263-7855(96)00018-5. [DOI] [PubMed] [Google Scholar]





