ArticlePDF Available

The function of the two-pore channel TPC1 depends on dimerization of its carboxy-terminal helix

Authors:

Abstract and Figures

Two-pore channels (TPCs) constitute a family of intracellular cation channels with diverse permeation properties and functions in animals and plants. In the model plant Arabidopsis, the vacuolar cation channel TPC1 is involved in propagation of calcium waves and in cation homeostasis. Here, we discovered that the dimerization of a predicted helix within the carboxyl-terminus (CTH) is essential for the activity of TPC1. Bimolecular fluorescence complementation and co-immunoprecipitation demonstrated the interaction of the two C-termini and pointed towards the involvement of the CTH in this process. Synthetic CTH peptides dimerized with a dissociation constant of 3.9 µM. Disruption of this domain in TPC1 either by deletion or point mutations impeded the dimerization and cation transport. The homo-dimerization of the CTH was analyzed in silico using coarse-grained molecular dynamics (MD) simulations for the study of aggregation, followed by atomistic MD simulations. The simulations revealed that the helical region of the wild type, but not a mutated CTH forms a highly stable, antiparallel dimer with characteristics of a coiled-coil. We propose that the voltage- and Ca²⁺-sensitive conformation of TPC1 depends on C-terminal dimerization, adding an additional layer to the complex regulation of two-pore cation channels. Electronic supplementary material The online version of this article (doi:10.1007/s00018-016-2131-3) contains supplementary material, which is available to authorized users.
Point mutations within a conserved C-terminal region reduce TPC1 activity. a–c Representative whole-vacuolar current responses (top) and current–voltage relations (bottom) of corresponding steady-state currents of TPC1-GFP variants as indicated at the top. Applied voltages ranged from −73 to +147 mV in 20 mV intervals, starting from a holding potential of −53 mV; current–voltage curves are shown starting from −43 mV. Additionally shown in b and c are confocal fluorescence overlay images of the GFP (green) and chlorophyll (red) signals of corresponding mesophyll protoplasts (scales represent 5 μm). a TPC1-GFP as in Fig. 1 (closed circles, n = 4), b TPC1-3LP-GFP (open squares, n = 7), c TPC1-3LA-GFP (open circles, n = 8). d Current–voltage relations of excised cytosolic side-out vacuolar membrane patches of TPC1-GFP (closed circles), TPC1-S706A-GFP (upward triangle), and TPC1-S706D-GFP (downward triangle) expressing cells as well as tpc1-2 mutants (open circles), with cytosolic [Ca²⁺] of 1 mM (left), 0.2 mM (middle), or 0.05 mM (right), respectively (n = 3–7). eC-terminal alignment of TPC1 from different plant species. Gray bar indicates the end of transmembrane segment S12, as predicted for AtTPC1 [84]. Green bar denotes the predicted C-terminal helix for AtTPC1. Aly-Arabidopsis lyrata; Ath-Arabidopsis thaliana; Bna-Brassica napus; Bol-Brassica oleracea; Cru-Capsella rubella; Csa-Cucumis sativus; Egr-Eucalyptus grandis; Gma-Glycine max; Gso-Glycine soja; Hvu-Hordeum vulgare; Jcu-Jatropha curcas; Mtr-Medicago truncatula; Nta-Nicotiana tabacum; Osa-Oryza sativa; Ptr-Populus trichocarpa; Sit-Setaria italica; Sly-Solanum lycopersicum; Tae-Triticum aestivum; Tca-Theobroma cacao; Zma-Zea mays
… 
This content is subject to copyright. Terms and conditions apply.
ORIGINAL ARTICLE
The function of the two-pore channel TPC1 depends
on dimerization of its carboxy-terminal helix
Nina Larisch
1
Sonja A. Kirsch
2
Alexandra Schambony
3
Tanja Studtrucker
1
Rainer A. Bo
¨ckmann
2
Petra Dietrich
1
Received: 12 August 2015 / Revised: 7 December 2015 / Accepted: 4 January 2016 / Published online: 18 January 2016
The Author(s) 2016. This article is published with open access at Springerlink.com
Abstract Two-pore channels (TPCs) constitute a family
of intracellular cation channels with diverse permeation
properties and functions in animals and plants. In the
model plant Arabidopsis, the vacuolar cation channel TPC1
is involved in propagation of calcium waves and in cation
homeostasis. Here, we discovered that the dimerization of a
predicted helix within the carboxyl-terminus (CTH) is
essential for the activity of TPC1. Bimolecular fluores-
cence complementation and co-immunoprecipitation
demonstrated the interaction of the two C-termini and
pointed towards the involvement of the CTH in this pro-
cess. Synthetic CTH peptides dimerized with a dissociation
constant of 3.9 lM. Disruption of this domain in TPC1
either by deletion or point mutations impeded the dimer-
ization and cation transport. The homo-dimerization of the
CTH was analyzed in silico using coarse-grained molecular
dynamics (MD) simulations for the study of aggregation,
followed by atomistic MD simulations. The simulations
revealed that the helical region of the wild type, but not a
mutated CTH forms a highly stable, antiparallel dimer with
characteristics of a coiled-coil. We propose that the volt-
age- and Ca
2?
-sensitive conformation of TPC1 depends on
C-terminal dimerization, adding an additional layer to the
complex regulation of two-pore cation channels.
Keywords Vacuole Calcium signaling
MD simulation Patch-clamp
Microscale thermophoresis Mutation
Introduction
Two-pore cation channels (TPCs) are intracellular ion
channels residing in the vacuolar membrane of plant cells,
and in lysosomes and endosomes of mammalian cells [1,
2]. Depending on their biophysical properties and host
species, they play diverse roles, ranging from cation and
pH homeostasis, sensing of the metabolic state, cell-to-cell
signaling, control of the membrane potential and mem-
brane trafficking to pigmentation, and even to the control
of Ebola virus invasion in animal cells [18]. The under-
lying molecular mechanisms are in most cases fairly
unknown.
TPCs belong to the superfamily of voltage-gated ion
channels that are built from Shaker-like domains of 6
transmembrane segments (S1-S6) and a pore-forming helix
between S5 and S6. One TPC monomer contains two
Shaker-like domains connected by a cytosolic linker, and
the channel functions as a dimer [1,9,10]. The cytosolic
N-terminus contains a dileucine motif, which is responsible
for targeting of the channel to the tonoplast in plant cells or
lysosomes in animal cells [11,12]. Two EF-hand motifs in
the linker domain of plant, but not animal TPCs mediate
binding and channel activation by cytosolic calcium ions
[13,14]. In contrast, animal TPC1 and TPC2 are directly
Electronic supplementary material The online version of this
article (doi:10.1007/s00018-016-2131-3) contains supplementary
material, which is available to authorized users.
&Petra Dietrich
petra.dietrich@fau.de
1
Molecular Plant Physiology, Department of Biology,
University of Erlangen-Nu
¨rnberg, Staudtstrasse 5, 91058
Erlangen, Germany
2
Computational Biology, Department of Biology, University
of Erlangen-Nu
¨rnberg, Staudtstrasse 5, 91058 Erlangen,
Germany
3
Developmental Biology, Department of Biology, University
of Erlangen-Nu
¨rnberg, Staudtstrasse 5, 91058 Erlangen,
Germany
Cell. Mol. Life Sci. (2016) 73:2565–2581
DOI 10.1007/s00018-016-2131-3 Cellular and Molecular Life Sciences
123
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
activated by phosphatidyl-inositol-3,5-bisphosphate
PI(3,5)P
2
[6,7,15] and have been identified as mediators
of nicotinic acid adenine dinucleotide phosphate
(NAADP)-induced Ca
2?
-release from endolysosomes [9,
12,1618].
While animal cells express two or more different
channels, in most plant species there is only one TPC
isoform, TPC1 [1,2]. TPC1 encodes the slow vacuolar
(SV) channel [19] and mediates the passage of K
?
and
Na
?
, and other monovalent as well as divalent cations,
although a direct involvement in calcium release from the
vacuole in vivo is strongly debated [2025].
Recently, a role for TPC1 in a rapid, long-distance
signaling system based on Ca
2?
waves has been uncovered
[3]. In response to a localized salt stimulus at the Ara-
bidopsis root, the calcium wave mainly travels through the
cortex and endodermis at speeds of up to 420 lm/s, and is
required for activation of stress responsive genes in the
shoot. Lack of TPC1 largely reduces the travel speed of
this trigger wave, while TPC1 overexpression accelerates
it. TPC1 thus plays an important role in long-distance
signaling in response to salt stress [3]. How the density or
distribution of TPC1 in the vacuolar membrane affects the
speed of the long-distance calcium wave remains largely
unknown.
Deregulation of the voltage-dependent activity of TPC1
leads to imbalanced cation homeostasis in the vacuole of
fou2, which has a threefold higher Ca/K ratio as compared
to the wild type [26]. The expression profile of fou2
resembles that of wild type plants under K starvation, and
fou2 plants produce more oxylipins in response to
wounding [27,28]. The fou2/TPC1D454 N mutation
introduces an amino acid exchange in the binding site for
luminal Ca
2?
, which abolishes the inhibition of TPC1 by
luminal calcium ions and shifts the activity range towards
more negative potentials [26,29]. These results demon-
strate the important role of TPC1 for cation homeostasis
and vacuolar storage function.
A tight regulation of SV channels prevents loss of
potassium and other cations from the vacuole, and many
factors down-regulating or blocking TPC1 have been
identified, including luminal calcium ions [30], protons
[31], sodium ions [32], and polyamines [33]. A further
negative regulation is mediated by polyunsaturated fatty
acids, which like luminal Ca
2?
ions shift the voltage
dependence towards more negative potentials [34]. In
comparison to these ionic and metabolic factors, less is
known about interactions of TPC1 with regulatory proteins
and their sites of interaction. 14-3-3 proteins rapidly reduce
SV currents [35,36], and regulation by kinases and phos-
phatases is most likely [37,38].
Except for a few cases like calcium binding [13,29]or
block by polyamines [33], little is known about the struc-
ture function relation in TPC1. Here, we report that the
predicted carboxy-terminal helical regions of two TPC1
monomers are prone to dimerization. These dimers are
shown to be essential for the function of the (dimerized)
TPC1 channel. We employed wild type coarse-grained and
atomistic simulations, as well as coarse-grained mutant
simulations, which revealed that the wild type C-terminal
domain forms a stable antiparallel coiled-coil. In contrast,
the mutant showed a highly promiscuous dimerization
pattern, pointing to a significantly decreased affinity. We
suggest that the dimerization of the wild type TPC1 car-
boxyl-termini stabilizes the channel in a conformation,
which is sensitive to Ca
2?
-binding and depolarization,
adding an additional layer to the complex regulation of
two-pore cation channels.
Materials and methods
Plant material and growth conditions
Arabidopsis thaliana Col-0 wild type and tpc1-2mutant
plants [20] were used. After seed stratification at 4 C for
3 days, plants were grown on soil in a growth chamber
under 8 h light/16 h dark conditions at 22 C. Nicotiana
benthamiana wild type plants were cultivated on soil in the
greenhouse at 22 C under a 16 h light/8 h dark cycle and
used after 6 weeks for infiltration.
Cloning procedures
Cloning procedures for C-terminal GFP-fusions are
described elsewhere [11], and eGFP was used in all cases.
PCR-primers were designed accordingly with the respec-
tive cloning sites and with or without stop-codon and
purchased from Sigma. For construct details and primer
sequences see Supplemental Table 1 and 2.
For BiFC analyses the gateway entry vector pENTR-D-
TOPO and the destination vectors pDEST–
GW
VYCE and
pDEST–
GW
VYNE were used for C-terminal fusion of the
Venus C- and N-terminus, respectively, and pDEST–
VYCE(R)
GW
and pDEST–VYNE(R)
GW
for N-terminal
tagging [39].
For the Co-IP tests the C-terminally tagged GFP-fusions
were used as templates to amplify the soluble C-terminus
(amino acids 673–733) of TPC1 or TPC1-3LP fused to
GFP with PCR. The constructs were subcloned into the
pENTR-D-TOPO vector and afterwards brought into the
expression vectors with EcoRI and XhoI (New England
2566 N. Larisch et al.
123
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
Biolabs), so that they were N-terminally tagged either with
6x Myc (pCS2?MT) or with the FLAG-tag (pCS2 ?Flag)
[40].
Protoplast isolation and transformation
Expression of GFP-fusion proteins for electrophysiological
measurements or confocal microscopy was performed in A.
thaliana tpc1-2mesophyll protoplasts as described [11,
13]. Briefly, leaves were enzymatically digested to release
mesophyll protoplasts, which were then transformed by the
polyethylene glycol method. Protoplasts were used for
electrophysiological measurements and subcellular local-
ization of GFP fluorescence 2–4 days after transformation.
Confocal microscopy
Fluorescence signals were detected and documented with a
TCS SP2 confocal laser scanning microscope and the Leica
Confocal Software (Leica Microsystems). A 488 nm Argon
laser was used for excitation of GFP and of the auto-flu-
orescence of chlorophyll. GFP-signals were detected from
500 to 556 nm and chlorophyll signals from 675 to
767 nm. Venus was excited with a 543 nm helium–neon
laser and detected in the range from 520 to 556 nm, the
corresponding chlorophyll signals were detected from 637
to 736 nm. Images were processed with Photoshop (Adobe
Systems).
Electrophysiological recordings
Patch-clamp experiments were performed on transformed
vacuoles harboring two-pore channels as identified by GFP
fluorescence. The pipette (luminal) solution consisted of
100 mM K-gluconate, 2 mM MgCl
2
, 10 mM EGTA,
10 mM MES, pH 5.5/Tris. The bath (cytosolic) solution
contained 50 mM K-gluconate, 1 mM CaCl
2
,1mM
MgCl
2
, 2 mM DTT, 10 mM HEPES, pH 7.5/Tris. Solu-
tions were adjusted to 430 mosmol/kg by adding
D-sorbitol. For analysis of the TPC1-S706 mutants two
further calcium concentrations of the bath solution were
used in which 1 mM CaCl
2
was substituted by 0.2 mM or
0.05 mM CaCl
2
, respectively. Competition assays using
the CTH peptide were performed on endogenous TPC1
channels in vacuoles isolated from Col-0 plants. Peptides
were dissolved at 10 mM in bath solution supplemented
with 1 % DMSO and 0.1 % pluronic, and diluted [600-
fold for application in patch-clamp recordings. Currents
were recorded either in the whole-vacuolar or cytosolic
side-out configuration of the patch-clamp technique, using
an EPC10 amplifier and the program PULSE (HEKA
electronics, Lambrecht, Germany). Recordings and data
analysis were performed as described earlier [11]. Relative
open probabilities (P
o
) were determined from tail currents
at -53 mV, following pulses to the different test voltages.
Tail currents were normalized to the maximum values at
1mM Ca
2?
, plotted as a function of the applied test
voltage, and fitted according to the Boltzmann equation:
PO¼A
1þexpzF
RTðVoVÞ
where V
o
and zare parameters for the voltage at half-
maximum open probability and the apparent number of
gating charges, respectively, and Adescribes the maximum
conductance that is reached under the experimental con-
ditions. For comparison of the voltage dependence in the
presence of different calcium concentrations, the gating
charge as determined for TPC1 in the presence of 1 mM
Ca
2?
was held constant for fits at lower Ca
2?
, as this
parameter does not dependent on the cytosolic Ca
2?
con-
centration [23].
BiFC, tobacco infiltration, and fluorescence
quantification
BiFC studies were performed as described [41] using a
Split Venus system [39]. The Agrobacterium strain C58C1
[42] harboring the plasmid of interest and the helper strain
p19 [43] were grown in LB medium supplemented with
100 lg/ml Kanamycin at 29 C overnight. The cultures
were brought to an optical density (OD
600
) of 1.0 in infil-
tration buffer (10 mM MES, 10 mM MgCl
2
, 100 lM
acetosyringone, pH 5.7/KOH) and the combinations to test
for co-expression mixed at equal amounts. The p19 helper
strain (OD
600
=1.0) was added to these mixtures in a 1:1
ratio. Suspensions were incubated for 2 h before infiltration
into the abaxial side of 6-week-old N. benthamiana leaves.
Each leaf was infiltrated with several plasmid combinations
and a positive control (TPC1/pDEST-
GW
VYCE and TPC1/
pDEST-
GW
VYNE) as well as a negative control (p19
alone) at separated areas to monitor comparability of the
transient expression and to define background fluorescence.
For fluorescence quantification leaf disks 6 mm)
were cut out of the infiltrated areas with a cork borer 2.5
day after infiltration. Leaf disks were placed upside-down
in a black 96-well plate prefilled with 70 llH
2
O per well
to minimize dehydration effects. Fluorescence was mea-
sured with a plate-reader (Infinite F200, Tecan) with an
excitation wavelength of 485 ±20 nm and an emission
wavelength of 525 ±25 nm [41].
Cell culture, transfection, Co-IP and western blot
analysis
HEK293T cells were propagated in DMEM GIBCO Glu-
tamax (Life Technologies) supplemented with 10 % fetal
The function of the two-pore channel TPC1 depends on dimerization of its carboxy-terminal helix 2567
123
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
calf serum. Cells were seeded in 6–well plates (20–30 %
confluency) and transfected the next day using 10 ll Roti-
Fect (Carl Roth) per 2 lg total DNA (1 lg DNA per
plasmid). Two days after transfection cells were washed
with 2 ml PBS buffer (Life Technologies) and harvested
for immunoprecipitation in 150 ll cold lysis buffer
(10 mM HEPES, 150 mM NaCl, 1 % Nonidet P40, 5 %
glycerol, pH 7,4/KOH) supplemented with protease inhi-
bitors and phosphatase inhibitors (Complete mini EDTA
free and PhosStop, Roche). Cells were disrupted mechan-
ically with a syringe (cannula diameter 0.55 mm). The
lysate was cleared by centrifugation at 13.500 rpm and
4C for 10 min. Total protein concentrations were mea-
sured with a bicinchoninic acid protein assay (Applichem),
typically the total protein amount was 500–600 lg. Load-
ing controls of 50 lg of the lysate were taken accordingly
and filled up to 20 ll with lysis buffer before adding 5 ll
of 5x sample buffer (250 mM Tris, 5 % SDS, 25 % glyc-
erol, 0.25 % bromophenol blue, 250 mM b
mercaptoethanol, pH 6.8/HCl) and boiling. For immuno-
precipitation, 3 lg of mouse anti-c-Myc Epitope antibody
9E10 (sc-40, Santa Cruz) was used per sample. Samples
were incubated with the antibody for 45 min at 4 C with a
rotator, then 10 ll of protein G Dynabeads (Life Tech-
nologies) equilibrated in the lysis buffer was added to each
sample and incubation continued for another 45 min.
Samples were washed three times with cold lysis buffer
before eluting three times with 20 ll of 100 mM glycine
(pH 2.5/HCl). After neutralizing the pH 15 ll of 5x sample
buffer was added and samples were boiled.
Proteins of the loading controls and samples were sep-
arated with SDS-PAGE (12 % gel) and transferred to a
PVDF membrane in duplicate. Membranes were blocked
for 1 h with blocking reagent (Roche) and incubated over
night at 4 C with rabbit anti-Myc or anti-Flag antibodies.
After washing three times with TBST buffer, the mem-
branes were incubated for 1 h at RT with anti-rabbit
antibody coupled to alkaline phosphatase. Membranes
were washed three times with TBST, followed by two
times with alkaline phosphatase buffer (100 mM NaCl,
100 mM Tris, pH 9.5/HCl). Colorimetric detection of the
proteins was performed with NBT/BCIP substrate (Roche)
in alkaline phosphatase buffer until a sufficient staining
was achieved. All antibodies used for the western blots
were purchased from Cell Signaling Technology.
Molecular dynamics simulations
All simulations were carried out using GROMACS 4.6.x
[44]. The secondary structure of the TPC1 C-terminal part
(679–733) was addressed applying different prediction
tools (Table 1). The consensus helical sequence
RSQRVDTLLHHMLGDEL, i.e., those residues that were
predicted by at least three out of four prediction tools, was
converted into an a-helical structure using PyMOL [45],
and the atomistic structure was changed to coarse-grained
representation with the aid of martinize [46]. Aggregation
of two TPC helices was studied from 100 simulations with
randomized starting structures, using the previously
developed DAFT algorithm [47], in conjunction with the
polarizable Martini [48] force field version 2.2; the two a-
helices were put in a dodecahedron simulation box at a
specified distance to each other. Finally, the solvent was
added using insane [49,50]. In this manner, each generated
system contained two 17 amino acids long helices and
approximately 5300 polarizable water beads. The usage of
a polarizable water model instead of a standard water
model was shown to be more appropriate for adsorption
studies of small peptides to a bilayer since the standard
water resulted in overestimation of binding, probably due
to a difference in electrostatic screening [51]. In addition to
the wild type TPC1 systems, coarse-grained systems of the
TPC1-3LA mutant were set up in the same manner as
described above and simulations were performed under the
same simulation conditions.
Five atomistic [force field Amber 14SB, 52] starting
configurations were generated from selected coarse-grained
frames of the wild type simulations using the backward
method [53]. The K
?
-ion concentration of 100 mM was
chosen similar to the cytoplasmic salt concentrations in
plants [54]. Additionally, the unit cells were altered to
smaller ones, containing approximately 5700 TIP3P [55]
water molecules as well as counterions (Cl
-
).
The modeled sequence is a small part of a larger protein
sequence. To describe the helices accurately, the termini
were kept neutral in all coarse-grained simulations or
acetylated and amidated, respectively, in the atomistic
simulations.
Simulation details
After automatically generating each 100 coarse-grained
(CG) starting structures using DAFT, the wild type and
mutant systems were subjected to a relaxation and equili-
bration process using martinate [56]. The energy was
minimized using steepest-descent (500 steps) and a 10 ps
position-restrained simulation with a time step of 2 fs.
Subsequently, the systems were relaxed in a 100 ps NpT
simulation with an integration time step of 20 fs under
constant pressure pand temperature T. In all simulations,
the isotropic pressure of 1 bar was kept constant using a
weak coupling scheme [57] with a 3 ps time constant. The
temperature was maintained at 310 K with the v-rescale
thermostat and a time constant of 1 ps [58]. The relative
dielectric constant was globally set to 2.5 and long-range
electrostatic interactions were treated using particle-mesh
2568 N. Larisch et al.
123
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
Ewald [59] summation with a real-space cutoff of 1.2 nm.
Dispersion interactions were described by a Lennard-Jones
12–6 potential that was shifted to zero between 0.9 nm and
1.2 nm. The production simulations were run for 250 ns
each, using a time step of 20 fs.
Selected backmapped coarse-grained (CG) structures of
the wild type were studied at atomistic resolution at 300 K
and 1 bar using the Amber14SB force field. The simulation
length was 200 ns (4 simulations) and 270 ns (1 simulation),
respectively, with an integration time step of 2 fs. The
temperature was kept constant using the Nose
´-Hoover [60,
61] thermostat with a time constant of 0.5 ps. The pressure
was modulated isotropically with the Parinello-Rahman [62]
barostat and a time constant of 10 ps. The relative permit-
tivity was set to 1. Long-range Coulomb interactions were
calculated with the particle-mesh Ewald summation and a
real-space cutoff of 1 nm. Van der Waals interactions were
treated with a cutoff of 1 nm as typical for Amber force fields
[63]. Furthermore, long-range dispersion corrections were
applied for the energy and pressure.
Data analysis
Relative orientations of the C-terminal helices of AtTPC1
and mutated AtTPC1-3LA were described using Euler
angles (Fig. 7b), which characterize the relative orientation
of two peptides [47,64]. The matrix with Euler angles was
obtained by least-square fitting a structure on a reference
structure. To investigate the stability of formed dimers, the
tilt angle between the dimers was determined over the last
50 ns of each simulation and its distribution plotted in a
histogram. The tilt, ranging from 0to 180, was then
divided into three subintervals (0–50,50–130, and
130–180) in the case of wild type dimers and four
subintervals (0–35,35–100, 100–150, and 150
180) in the case of mutant dimers. These intervals were
chosen according to the minima in the histogram
(Fig. S6a). Subsequently, the number of transitions
between these intervals was counted for each simulation
and further evaluated.
MicroScale thermophoresis (MST) binding assay
For MST experiments, peptides corresponding to the con-
sensus helical sequence (RSQRVDTLLHHMLGDEL) and
the 3LA mutant (RSQRVDTAAHHMAGDEL) were syn-
thesized (Peptide Speciality Laboratories GmbH,
Heidelberg, Germany). To allow for label-free MST
binding experiments, additional wild type and mutant
peptides were synthesized, which were C-terminally
extended by addition of two tryptophanes and a short linker
(AAWW).
Labelfree MicroScale Thermophoresis binding experi-
ments were performed in cooperation with the 2bind
GmbH (Regensburg, Germany), using 750 nM tryptophane
containing target peptide in PBS pH 7.5, 1 % DMSO,
0.1 % Pluronic with varied concentrations of the ligand
peptide at 80 % MST power, 100 % LED power in
hydrophilic zero background capillaries on a Monolith
NT.labelfree device at 25 C (NanoTemper Technologies,
Munich, Germany). Normalized fluorescence data sets
(WT peptide and MT peptide) were analyzed in the ther-
mophoresis and temperature jump. For determination of the
binding affinity (K
D
) of the wild type peptide, the recorded
fluorescence was normalized to the fraction bound
(0 =unbound, 1 =bound), and fitted using the K
D
fit
formula derived from the law of mass action. Technical
duplicates were performed for each experimental setup.
Sequence information
Sequence data from this article can be found in the EMBL/
GenBank data libraries under the following accession
numbers: Aly-Arabidopsis lyrata (D7M2M4); Ath-
Table 1 Secondary structure prediction of the TPC1 carboxyl-terminus
CCoil, Hhelical, Esheet, Tbeta-turn
a
Yang et al. [85], Roy et al. [86], Zhang [87]
b
Xu, Zhang [88]
c
Remmert et al. [89], So
¨ding [90], So
¨ding et al. [91]
d
Kelley, Sternberg [92]
The function of the two-pore channel TPC1 depends on dimerization of its carboxy-terminal helix 2569
123
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
Arabidopsis thaliana (B9DFD5); Bna-Brassica napus
(A0A078G686); Bol-Brassica oleracea (A0A0D3E0B2);
Cru-Capsella rubella (R0GT49); Csa-Cucumis sativus
(A0A0A0K5Q7); Egr-Eucalyptus grandis (A0A059A094);
Gma-Glycine max (1M3S8); Gso-Glycine
soja(A0A0B2R1M3); Hvu-Hordeum vulgare (Q6S5H8);
Jcu-Jatropha curcas (A0A067JJP4); Mtr-Medicago trun-
catula (A0A072VBZ7); Nta-Nicotiana tabacum
(Q75VR1); Osa-Oryza sativa (Q5QM84); Ptr-Populus
trichocarpa (U5FYB3); Sit-Setaria italica (K3XEV7); Sly-
Solanum lycopersicum (K4CFU2); Tae-Triticum aestivum
(Q6YLX9); Tca-Theobroma cacao (A0A061E309); Zma-
Zea mays (B6SP34).
Results
AC-terminal region is essential for TPC1 function
We previously reported that deletion of the last 55 amino
acids of AtTPC1, corresponding to the cytosolic carboxyl-
terminus, resulted in a mutant (TPCDC) which is correctly
targeted to the tonoplast, but lacks channel function [11].
This indicates an essential role of this region for the activity
of TPC1, but not for the targeting or trafficking process.
These previous results were obtained from electrophysi-
ological recordings of excised patches and thus, due to the
small membrane area, a little residual activity of TPC1DC
may have escaped observation. We now compared slow
vacuolar (SV) currents in the whole-vacuolar configuration
of the patch-clamp technique after expression of wild type
TPC1 and the TPC1DC mutant, respectively. Both channel
variants were expressed as GFP-fusions in the tpc1-2
knockout background, allowing localization and electro-
physiological analysis of the introduced TPC1 versions as
homo-dimers [11,13]. In the presence of 1 mM CaCl
2
in the
bath solution, which fully activates the TPC1 wild type
(Fig. 1a), no currents were obtained from vacuoles
expressing the truncated channel (Fig. 1d). This result sup-
ports the conclusion that the C-terminus is necessary for the
voltage- and Ca
2?
-dependent activity of TPC1 [11].
To investigate whether a specific sub-region of the
carboxyl-terminus is involved in this regulation, two
additional truncated channel versions were created:
TPC1DC8 contained the residues 1–725, and TPC1DC29
the residues 1–704. Both mutants were localized in the
tonoplast, indicating their expression, efficient ER export,
and correct targeting (Fig. 1b, c). The lack of the last 8
amino acids did not interfere with the activity of TPC1DC8
(Fig. 1b), and the amplitudes and current–voltage behavior
of the mutant were wild type like (Fig. 1b, e). In contrast,
although TPC1DC29 was expressed normally (Fig. 1c,
Fig. S1), this mutant stayed silent like the TPC1DC mutant
lacking the whole carboxyl-terminus (Fig. 1c–e). These
results identified amino acids 705–725 to include a region
indispensable for channel function.
Mutation in a predicted helical domain abolishes
TPC1 activity
So far, no structural information is available about TPCs of
plant or animal origin. To address the secondary structure
of the C-terminus, different prediction tools were applied,
TPC1
TPC1 C8
TPC1 C29
TPC1 C
2 nA
200 ms
2 nA
200 ms
2 nA
200 ms
2 nA
200 ms
a
b
c
d
e
300
200
100
Iss (pA/pF)
150100500
-50
V (mV)
Fig. 1 The C-terminus is essential for the function of TPC1. a
dConfocal fluorescence overlay images of the GFP (green) and
chlorophyll (red) signals (scales represent 5 lm), and representative
whole-vacuolar current responses of TPC1-GFP variants as indicated
on the left. Applied voltages ranged from -73 to ?147 mV in 20 mV
intervals, starting from a holding potential of -53 mV. eCurrent–
voltage relations of whole-vacuolar steady-state currents determined
from traces as shown in (ad) for TPC1-GFP (closed circles,n=4),
TPC1DC8-GFP (open circles,n=5), TPC1DC29-GFP (open
rhombi,n=6), and TPC1DC-GFP (closed rhombi,n=4) normal-
ized to the vacuolar membrane capacitance. Data represent mean
values ±SE
2570 N. Larisch et al.
123
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
which revealed a consensus helical sequence to span amino
acids 707–723 (Table 1).
To assess the role of this a-helix we investigated a
mutant with an altered secondary structure. Substitution of
three leucines (L714, 715, and 719) by prolines as a potent
helix-breaker [65] resulted in the mutant TPC1-3LP,
which, according to the secondary structure prediction
tools listed in Table 1, lacks the a-helix. In a comple-
mentary experiment, a second point mutant, TPC1-3LA,
was chosen to replace bulky hydrophobic leucines by the
comparably small alanine favoring an a-helical confor-
mation [66]. The prediction tools as listed in Table 1
supported the assumption of a helical secondary structure
in the C-terminus of TPC1-3LA. The fluorescence emerg-
ing from the GFP-tagged mutant was detected in the
tonoplast of transformed cells, but whole-vacuolar currents
of TPC1-3LP produced only about 7 % of the wild type
amplitude (Fig. 2a, b). Interestingly, TPC1-3LA resulted in
a loss of SV currents to a similar extent as TPC1-3LP
(Fig. 2c). Comparable expression levels between wild type
and mutants were indicated by similar GFP intensities, both
on the single cell level and for the protoplast suspensions
(Fig. 2b–c, Fig. S1). Residual currents produced by both
mutants displayed the typical voltage- and time-dependent
characteristics of the SV channel. In excised patches, no
macroscopic currents were resolved, but the presence of
single-channels with a conductance of 58 pS, similar to
that of the wild type [14,20,67], showed that the single
channel amplitudes were not affected by the mutation
(Fig. S2).
A current reduction was also obtained, when a point
mutation was introduced at Ser706, which is located at the
N-terminal end of the helix (Table 1). Since Ser706 rep-
resents a putative phosphorylation site, it was replaced
either by the phosphorylation-mimicking Asp, introducing
a negative charge, or by Ala. In both cases, the whole-
vacuolar current amplitudes were largely reduced, by 58 %
and 53 % for TPC1-S706D and TPC1-S706A, respectively
(Fig. 2d, Fig. S3). The current reductions were also
observed at non-saturating Ca
2?
-concentrations of 0.2 mM
and 0.05 mM (Fig. 2d). Similar open probabilities of wild
type and mutants at saturating Ca
2?
concentrations (1 mM)
and 0.2 mM, corresponding to the half-maximal activation
concentration [13], showed that the current reductions did
not result from a largely impaired Ca
2?
-dependent shift of
the voltage dependence (Fig. S3). Besides shifting the
voltage dependence to less negative potentials, an elevation
of the cytosolic Ca
2?
concentration also increases the
maximum conductance [13]. While TPC1-S706 wild type
currents were reduced to 53 ±10 % at 107 mV following
the reduction of Ca
2?
from 1 mM to 200 lM, this value
was 28 ±3 % for TPC1-S706A and 24 ±4 % for TPC1-
S706D (Fig. 2d). The S706 mutation therefore appears to
modestly affect the Ca
2?
-sensitivity of channels via the
link between the Ca
2?
-binding and change in the number
of voltage-sensitive channels.
Together, these results show that Ser706 appears to have
a structural role instead of being involved in channel reg-
ulation by phosphorylation. Furthermore we conclude that
mutations in or near the C-terminal helix reduce the
number of open channels rather than affecting the voltage
dependence itself.
An alignment of the TPC1 carboxyl-termini of different
species shows that the C-terminal helix (CTH), including
S706 as well as L714, L715, and L719, is highly conserved
among plants (Fig. 2e), implicating an important function
of this domain. Our results strongly suggest that vacuoles
harboring mutations in the C-terminal helix formed less
functional two-pore channels. As helical structures are
often involved in mediating protein–protein interactions,
the CTH of TPC1 may be involved in protein–protein
interactions required as a prerequisite for channel gating or
stabilization of the open state.
TPC1 dimerizes via its C-termini
One possibility for a CTH-mediated interaction would be a
dimerization of the cytosolic C-termini of two TPC1 sub-
units. To test this possibility, interaction studies were
performed using bimolecular fluorescence complementa-
tion (BiFC) and co-immunoprecipitation (Co-IP).
Split Venus BiFC was tested for the full-length chan-
nels, either the wild type TPC1 or the TPC1-3LP mutant.
BiFC signals from wild type channels could be observed
when two C-terminally tagged TPC1 subunits were co-
infiltrated in Nicotiana benthamiana (Fig. 3a). The fluo-
rescence signals emerged from the vacuolar membrane and
in few cells additionally from the preceding endomem-
branes of the secretory pathway, such as endoplasmic
reticulum or Golgi apparatus. This shows that the C-ter-
minal parts of TPC1 come into close contact to one
another, allowing for fluorophore formation. Patch-clamp
analysis of tpc1-2 Arabidopsis cells co-transformed with
the C-terminally tagged TPC1 BiFC-constructs resolved
typical SV currents, supporting the hypothesis that a
dimerization via the C-termini results in voltage-sensitive
channels (Fig. 3b).
TPC1-3LP C-terminally tagged with the Venus halves
produced BiFC signals that were reduced compared to the
wild type (Fig. 3d). Confocal fluorescence microscopy
revealed that compared to the wild type, TPC1-3LP tagged
with the Venus halves was less efficiently transported to
the vacuole (Fig. 3c), since the fluorescence emerged not
only from the vacuolar membrane, but also to significant
amounts from other endomembranes, mostly the ER. This
localization pattern of TPC1-3LP with BiFC was not seen
The function of the two-pore channel TPC1 depends on dimerization of its carboxy-terminal helix 2571
123
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
with GFP-tagged TPC1-3LP in Arabidopsis tpc1-2cells
(Fig. 2b, Fig. S1) and might hint to a slowed trafficking
associated with the tobacco expression system compared to
the Arabidopsis mesophyll protoplasts. Nevertheless, the
reduced BiFC signals of the helix-breaking point mutant
TPC1-3LP (Fig. 3c) in comparison to the wild type may
indicate that the dimer formation via the C-terminus was
hampered. However, the flexibility of the C-termini and
Venus halves would still allow the formation of the
fluorophore.
200 ms
2 nA
200 ms
2 nA
200 ms
2 nA
TPC1 TPC1-3LATPC1-3LP
abc
d
200
150
100
50
Iss (pA/pF)
150100500
-50
V (mV)
200
150
100
50
Iss (pA/pF)
150100500
-50
V (mV)
200
150
100
50
Iss (pA/pF)
150100
50
0-50
V (mV)
Sly NLIVAFVLEAFQAEMDLEAAANCADGDD--KESRSERRRNVGTKTRSQRVDFLLHHMLSSELTECSHDDP---- 739
Nta NLIVAFVLEAFQAEVDLEASARCVDGDD--KEAKSERRRNVGTKTRSQRVDFLLHHMLRSELTECSNENP---- 735
Cru NLVVAFVLEAFFTELDLEEEEKCQGQD---SQERRNRRRSAGSKSRSQRVDTLLHHMLGDELSKPECSTSDT-- 732
Ath NLVVAFVLEAFFTELDLEEEEKCQGQD---SQEKRNRRRSAGSKSRSQRVDTLLHHMLGDELSKPECSTSDT-- 733
Aly NLVVAFVLEAFFAELDLEEEENCQGED---SQERRNRRRSAGTKSRSQRVDTLLHHMLGDELSKPECSTSDT-- 732
Bol NLIVAFVLEAFFTELDLEEEEKCEGQD---SQERRNRRRSAGSKSRSQRVDTLLHHMLGDELSKPECSTTATDT 736
Bna NLIVAFVLEAFFTELDLEEEEKCEGQD---SQERRNRRRSAGSKSRSQRVDTLLHHMLGDELSKPECSTTATDT 738
Gso NLIIAFVLEAFFAEMELESSETCEGNDKEVEGDKY-RKRSIGTKTRSQRVDALLHHMLSAELCQNESSSTQTS- 739
Gma NLIIAFVLEAFFAEMELESSETCEGNGKEVEGDKY-RKRSIGTKTRSQRVDALLHHMLSAELCQNEPSSTQTS- 738
Mtr NLIIAFVLEAFFAEIELEEAETGDGNDKEVAGERYPRRRALGTKSRSQRVDALLHHMLSAELGQNQTSST---- 737
Ptr NLVMAFVLEGFFAEMELETAEKCEAEDKE-GSNSKSRRRSVGTKTRSQRVDNLLHHMLSAELEKPECSNA---- 726
Jcu NLVVAFVLEAFFAEMDLEKPEECEDED---ARTSKPRR--IGTKSRSQRVDILLHHMLSAELHDNQSSNA---- 735
Egr NLVVAFVLEAFFAEMDLESSENCEGQDEE-IRGR--RSRSVGTKTRSQRVDVLLHHMLSAELDKAKCTCPQP-- 737
Tca NLVVAFVLEAFFTEMDLETSGNCEEDDKD-AGSGKYRRRLVGTKTRSQRIDILLHHMLSAELDKGQSSASSTP- 737
Csa NLVVAFVLEAFFAELDIESSENGEEQDQD-KDSRKDRPRFVGTKTRSRKVDILLHHMLSAELDDKDSD------ 738
Zma NLIVAFVLEAFFAEMELEKAGEADTQDS--TPQGRNKRRSMRARTKGTMVDILLHHMLSNELDGSQNSD----- 749
Sit NLIVAFVLEAFFAEMELEKAGESDMQDS--TPQGRNKRRSMRVKTKGTMVDILLHHMLSNELDGSQNTDQ---- 745
Osa NLIVAFVLEAFFAEMELEKDGEADIQDP--TLEGRNRRRSVRVRTKGTMVDILLHHMLSNELDGSQNRDQ---- 757
Tae NLIVAFVLEAFFAEMELEKGEEVDIQNP--TSGGIKKRRSMRVRSKGTMVDILLHHMLSNELDGSQNS------ 742
Hvu NLIVAFVLEAFFAEMELEKGEEVDIQSP--TSGGIKKRRSMRVRSKGTMVDILLHHMLSNELDGSQNS------ 742
**::*****.* :*:::* . : :::. :******* ** .
e
2572 N. Larisch et al.
123
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
In contrast, combinations with N-terminally tagged
TPC1 subunits did not result in Venus-fluorescence making
intramolecular interactions of the N-terminus seem unli-
kely (Fig. 3d). Functional expression and tonoplast
localization of an N-terminally tagged TPC1 were verified
using a GFP tag (Fig. S4).
In an independent approach, the soluble C-termini
(residues 673–732) of wild type TPC1 and of TPC1-3LP
were assayed for co-immunoprecipitation in a heterologous
system. The proteins were N-terminally tagged with a
6xMyc-tag for precipitation or a FLAG-tag for detection of
co-precipitated proteins, and in addition C-terminally
enlarged by a GFP-fusion to provide better handling, and
co-expressed in HEK293T cells. These experiments con-
firmed that the TPC1 C-terminus bound to itself and could
be detected with both antibodies in the immuno-precipitate
(Fig. 4). As a control, the empty vector pCS2 or with
6xMyc-GFP was co-transfected with the Flag- or Myc-
tagged C-terminus, respectively, since GFP is known to
have a tendency to dimerize itself. Compared to the wild
type, the C-terminus of the TPC1-3LP-mutant showed a
reduced ability to dimerize, reflected by a weaker signal
bFig. 2 Point mutations within a conserved C-terminal region reduce
TPC1 activity. acRepresentative whole-vacuolar current responses
(top) and current–voltage relations (bottom) of corresponding steady-
state currents of TPC1-GFP variants as indicated at the top. Applied
voltages ranged from -73 to ?147 mV in 20 mV intervals, starting
from a holding potential of -53 mV; current–voltage curves are
shown starting from -43 mV. Additionally shown in band care
confocal fluorescence overlay images of the GFP (green) and
chlorophyll (red) signals of corresponding mesophyll protoplasts
(scales represent 5 lm). aTPC1-GFP as in Fig. 1(closed circles,
n=4), bTPC1-3LP-GFP (open squares,n=7), cTPC1-3LA-GFP
(open circles,n=8). dCurrent–voltage relations of excised cytosolic
side-out vacuolar membrane patches of TPC1-GFP (closed circles),
TPC1-S706A-GFP (upward triangle), and TPC1-S706D-GFP (down-
ward triangle) expressing cells as well as tpc1-2mutants (open
circles), with cytosolic [Ca
2?
] of 1 mM (left), 0.2 mM (middle), or
0.05 mM (right), respectively (n=3–7). eC-terminal alignment of
TPC1 from different plant species. Gray bar indicates the end of
transmembrane segment S12, as predicted for AtTPC1 [84]. Green
bar denotes the predicted C-terminal helix for AtTPC1. Aly-
Arabidopsis lyrata;Ath-Arabidopsis thaliana;Bna-Brassica napus;
Bol-Brassica oleracea;Cru-Capsella rubella;Csa-Cucumis sativus;
Egr-Eucalyptus grandis;Gma-Glycine max;Gso-Glycine soja;Hvu-
Hordeum vulgare;Jcu-Jatropha curcas;Mtr-Medicago truncatula;
Nta-Nicotiana tabacum;Osa-Oryza sativa;Ptr-Populus trichocarpa;
Sit-Setaria italica;Sly-Solanum lycopersicum;Tae-Triticum aes-
tivum;Tca-Theobroma cacao;Zma-Zea mays
ab
cd
Fig. 3 The C-termini of two TPC1 monomers interact with each
other. aConfocal fluorescence overlay images of the Venus (yellow)
and chlorophyll (red) signals of a tobacco cell co-expressing TPC1
fused to Venus-Ct (TPC1-VC) and TPC1 fused to Venus-Nt (TPC1-
VN), shown at two different magnifications (scales represent 5 lm).
The BiFC signals emerge from the tonoplast. bWhole-vacuolar
current response of a tpc1-2cell co-expressing the same constructs as
in (a). cConfocal fluorescence overlay image of the YFP (yellow) and
chlorophyll (red) signals of a tobacco cell co-expressing TPC1-3LP-
VC and TPC1-3LP-VN (scales represent 5 lm). The BiFC signals
emerge from endomembranes and tonoplast. dFluorescence signal
intensities of BiFC experiments in relative units, determined from leaf
disks with a fluorescence reader. VC-TPC1 and VN-TPC1 (n=23),
TPC1-VC and VN-TPC1 (n=23), VC-TPC1 and TPC1-VN
(n=23), TPC1-VC and TPC1-VN (n=38), TPC1-3LP-VC and
TPC1-3LP-VN (n=11). As negative controls leaf disks only
expressing the helper strain p19 (n=28) and not infiltrated leaves
(n=39) were used
The function of the two-pore channel TPC1 depends on dimerization of its carboxy-terminal helix 2573
123
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
from the Co-IP, while the expression level was comparable
to the wild type (Fig. 4).
The two independent protein–protein interaction assays
demonstrated the dimerization of the TPC1 C-terminus and
suggest that this interaction is mediated by the C-terminal a-
helix (CTH). The ability of the CTH to homo-dimerize was
further directly measured in microscale thermophoresis
(MST) experiments. Titration of the wild type peptide cor-
responding to the CTH (Table 1) against the wild type
peptide labeled by tryptophane induced a dose-dependent
change in mobility (Fig. 5a), indicative of the dimer for-
mation. From the change in fluorescence a binding affinity of
3.85 ±1.1 lM was determined (Fig. 5c). In contrast, titra-
tion curves of the 3-LA peptide revealed no interactions of
this mutant peptide in a concentration range from 3 nM to
100 lM (Fig. 5b,c).
The CTH mediates formation of an antiparallel
coiled-coil dimer
The CTH-mediated dimerization was further analyzed in
silico, both for the wild type and for the CTH-3LA mutant.
To this end, the dimerization was first analyzed from 100
aggregation simulations (250 ns each) for each system.
Each dimerization simulation started from two C-terminal
TPC a-helices at an initial peptide center of mass (COM)
distance between 5.5 and 6.5 nm solvated in a box of water
at coarse-grained resolution. Dimers were formed for both
systems in all simulations within tens of nanoseconds,
reflecting the observed enhanced stickiness of the coarse-
grained MARTINI force field [47].
Fig. 4 The C-terminus of TPC1 interacts with itself. Co-immuno-
precipitation of TPC1 C-terminal proteins co-expressed in HEK 293T
cells (combinations as indicated) and analyzed by western blotting.
Precipitation was achieved via the Myc-tag (IP anti-Myc), and
proteins detected via antibodies against the Myc- (WB anti-Myc) or
the Flag-tag (WB anti-Flag). The Flag-tagged C-terminus of TPC1
(Flag-Ct) was co-precipitated with the Myc-tagged C-terminus of
TPC1 (Myc-Ct, lane 4), whereas only background interaction was
observed with Myc-tagged GFP (lanes 2 and 5). Introduction of the
3LP mutation resulted in visibly weaker interaction of the two C-
termini (lane 8). Similar results for the interaction of the TPC1 C-
termini were obtained in 3 independent experiments. The presence of
the proteins was verified in the cell lysate. Calculated protein weights
were Myc-Ct (46 kDa), Flag-Ct (37 kDa), 3LP-Myc-Ct (47 kDa),
3LP-Flag-Ct (38 kDa), Myc-eGFP (37 kDa), Myc (10 kDa), Flag
(1 kDa)
a
c
b
Fig. 5 Synthetic CTH peptides dimerize. a,bMST time traces
(normalized fluorescence) of 16 capillaries containing 750 nM
tryptophane-labeled wild type (a) or 3LA mutant (b) CTH peptide
and unlabeled wild type (a) or 3LA mutant (b) CTH peptide at
concentrations between 3 nM and 100 lM. Thermodiffusion is
reduced with increasing peptide concentrations. cThe normalized
fluorescence of the MST traces was converted to the fraction bound
and plotted against the concentration of the ligand. A K
D
of
3.85 ±1.1 lM was determined for the interaction of the wild type
CTH peptides (open circles), while no interaction was measured for
the 3LA mutant (closed circles)
2574 N. Larisch et al.
123
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
However, the dimerized CTH-3LA adopted significantly
different relative orientations as compared to wild type
dimers: Overall, the mutant exhibited more flexibility in
the bound state (Fig. 6, Fig. S5). The angle between the
helical axes, called tilt, describes the parallel or antiparallel
orientation of the bound peptides (Fig. 7b). The distribu-
tion of this angle (Fig. 6), calculated over the last 50 ns of
the trajectories, is in the case of the CTH-3LA dimers very
diffuse, pointing to a drastically enhanced flexibility and
thus less stable dimers as compared to the wild type
dimers. The number of transitions between the orientations
(see ‘Data Analysis’) strengthens this conclusion: While
the majority of wild type dimers were stable in their ori-
entation (median of orientation transitions is 0), the number
of transitions in mutant dimers was significantly increased
(median is 4.5, Fig. S6). This flexibility in the relative
orientation of mutated helices most likely results from a
reduction in steric hindrance. Leucines consist of two
coarse-grained beads (four residual methyl groups), while
alanine exhibits only one backbone bead (one residual
methyl group). Since the majority of wild type dimers
ended up in a conformation, where leucines are buried in
the interface (see below), this exchange to the smaller
alanines caused an increased flexibility in the tilt angle.
Furthermore, alanines are not as hydrophobic as leucines
and thus, they are more prone to face the solvent.
The relative instability and promiscuity of the CTH-3LA
dimers observed in MD simulations corresponds to the lack
of dimer formation in MST experiments. In the following,
we focused on the analysis of dimer formation for the CTH
wild type peptide.
Within approximately 50 ns, the wild type peptides
approached each other in all simulations to a distance
below 1 nm, reflecting dimerization (81 % of the mono-
mers dimerized within 25 ns). Figure 7a shows the center
of mass distances for each simulation. With increasing
simulation time, cluster formation is seen, with one cluster
at a COM distance of &1 nm (c1) and a second cluster at
&1.45 nm (c2). Peptides in dimers of cluster c2 adopted a
shifted configuration (see below).
The orientation of the C-terminal dimers of TPC1 was
analyzed using Euler angles [47,64]. A tilt angle below 50
indicates that the helices are in a parallel conformation,
while an angle above 130reflects an antiparallel orienta-
tion (Fig. 7b). The latter configuration was adopted in
86 % of the coarse-grained simulations, characterized by
an average tilt angle of 160. In 7 % of the simulations, the
dimers ended up in a parallel orientation (average tilt of
18), the remaining simulations ended up in diverse inter-
mediate configurations. To reach the antiparallel
conformation most of the dimers followed a specific
dimerization pathway (Fig. 7c): At larger intermolecular
distances (3–7 nm), i.e., at the beginning of the simulation,
the distribution of tilt angles was random (Fig. 7c). Upon
approach, the tilt was constrained to values of 10–50, the
peptides thus oriented in a parallel head-to-tail configura-
tion. Subsequently, the peptides reoriented and ended up in
the antiparallel configuration (130–180, Fig. 7c).
To evaluate the frequency of the different dimer con-
formations at the end of the simulations, combinations of
the binding position band rotation angle /(compare
Fig. 7b) obtained from the trajectories were plotted as a
two-dimensional kernel density map (Fig. 7d). While the
position of one monomer to a reference monomer is
defined by the position b, the exact binding site that faces
the reference structure is given by the phase /, the rotation
around its helical axis (Fig. 7b).
In the bound state, i.e., at the end of the simulations,
positively and negatively charged amino acids come into
close proximity, in particular in the antiparallel configu-
ration (Fig. 7d). These antiparallel dimers (cluster A in
Fig. 7d) are electrostatically favored by proximity of the N-
terminal arginines of one monomer and the C-terminal
glutamic and aspartic acids of the second monomer.
Additionally, the hydrophobic residues leucine and valine
are positioned at the dimer interface.
Besides the main cluster A, four additional clusters (B–
E) were identified. The second-largest cluster B consists of
antiparallel dimers that interact as well through electro-
static interactions. Due to a decreased number of
hydrophobic residues at the dimer interface, this confor-
mation is expected to be metastable. Parallel dimers were
found in clusters C and D. In these arrangements, the
helices are shifted with respect to each other, allowing for a
0.000 0.025
0 20 40 60 80 100 120 140 160 180
0.000 0.010
0 20 40 60 80 100 120 140 160 180
Wild type
Mutant
DensityDensity
Tilt
Tilt
Fig. 6 CTH dimer tilt distribution reflects an increased flexibility of
mutant dimers. Tilt angles between coarse-grained dimers were
determined over the last 50 ns from each simulation and plotted in a
histogram (x-axis: tilt, y-axis: density). The density distribution of tilt
angles between mutant dimers (bottom) is compared to wild type
dimers (top) very diffuse, pointing to an increased flexibility of
mutant dimers
The function of the two-pore channel TPC1 depends on dimerization of its carboxy-terminal helix 2575
123
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
better packing of hydrophobic residues (parallel dimer,
Fig. 7d). Therefore, the center of mass distances between
the corresponding monomers is increased as compared to
the antiparallel dimers (compare c2 in Fig. 7a). For cluster
E (tilt from 50–130) the inner leucines are in close
proximity to each other. Furthermore, in this configuration
the central aspartic acids are able to interact with the
arginines.
To evaluate the binding strength, the non-bonded
interaction energy was calculated for the different confor-
mations. In agreement with the high frequency of
antiparallel dimers (cluster A and B), their mean interac-
tion energy was lowest among all dimers with a value of
&-600 kJ/mol, while the parallel dimers interact with
&-420 kJ/mol, and the rest found in cluster E with
&-500 kJ/mol.
For further analysis of the stability of the dimers formed
at coarse-grained (CG) resolution, five CG conformations
corresponding to the main clusters A–E as indicated in
Fig. 7d were backmapped to atomistic resolution and
simulated for 200 ns (cluster A, C–E) or 270 ns (cluster B).
The two initial parallel dimer configurations (C and D)
changed in the atomistic simulations to an antiparallel
orientation within less than 10 ns, indicating a relative
instability of the parallel configurations. The dimers of
clusters D and E adopted an antiparallel configuration
(position d,ein Fig. 7d) during the atomistic simulations,
with the three inner leucines and valine at the helical
interface stabilized by two salt bridges at the termini
(Fig. 8a, iv and v). This orientation is indicative of coiled-
coil packing, burying hydrophobic residues within the core
flanked by stabilizing ionic interactions at the helical sides.
Less stable conformations were observed for the other
dimers (clusters A, B, C). Here, leucines and valines were
tilt
tilt
0 50 100 150 200 250
0
1
2
3
4
5
6
7
Distance (nm)
Time (ns)
a
b
c
c2
c1
d
Parallel
Antiparallel
COM Distance (nm)
012345 67
180
20
40
60
80
100
120
140
160
-150 -100 150100500-50
β
Φ
50 3002502001501000 350
bFig. 7 In silico dimerization study of the carboxy-terminus of TPC1.
aCenter of mass (COM) distance between the two helices (residues
707–723), with initial values between 5.5 and 6.5 nm, as a function of
simulation time. Different colored lines represent 100 independent
coarse-grained (CG) simulations. c2 and c1 indicate two populations
of dimers distinguished by their COM distance. bSchematic repre-
sentation of angles used to analyze the orientation of bound helices. b
describes the position of the binding partner to the reference structure
(colored in gray), /is the rotation around the helical axis, and the tilt
angle depicts the tilt between the peptides. Lower tilt values (B50)
point to parallel binding and larger (C130) to antiparallel binding.
Helices were colored in rainbow scheme from N-terminus (blue)toC-
terminus (red). cCenter of mass distances of two monomers forming
homo-dimers in antiparallel arrangement, shown for all simulations
resulting in antiparallel configurations. dLeft: Kernel density for b
and /angle combinations in the last 50 ns of the CG simulations,
describing the relative orientation of peptides. The yellow area
indicates a high density, white color a low density. Antiparallel
dimers were found in clusters A and B, parallel dimers in C and D,
others in E. Two selected conformations from clusters D and E were
finally observed at positions in phase space d and e, respectively, after
backmapping to atomistic resolution followed by atomic-scale
molecular dynamics simulations. Right Orientation for the antiparallel
and parallel homo-dimers. Glutamic and aspartic acids are colored
red, arginine green, and hydrophobic residues light brown. For
clarity, only interacting residues are colored
2576 N. Larisch et al.
123
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
partially exposed to the solvent (Fig. 8a, ii and iii) or
hydrogen bonds within the helix were broken and conse-
quently the helix disrupted (Fig. 8a, i and ii). Additionally,
hydrophilic histidines were partially found at the interface.
The combined coarse-grained and atomistic study thus
strongly suggests that the wild type, but not the mutant
CTHs of two TPC1 monomers form a stable helical dimer
in an antiparallel coiled-coil conformation (Fig. 8b).
Together with the results of the mutagenesis study, we
propose the formation of an antiparallel CTH dimer as a
prerequisite for TPC1 activity. This conclusion is further
supported by the ability of the synthetic CTH peptide to
rapidly inhibit TPC1 currents, when applied to the active
channels in electrophysiological recordings (Fig. S7).
a
b
i ii iii
iv v
C
C
N
N
Fig. 8 CTH dimer configurations observed in combined coarse-
grained/atomistic simulations. aMetastable (i,ii,iii) and stable (iv,
v) dimer configurations. For the stable dimer arrangement, the
hydrophobic residues are found mainly at the interfacial region, and
stabilizing salt bridges at the termini. For clarity, interfacial
methionines are hidden. The less stable configurations show partially
exposed hydrophobic (light brown) amino acids and a disrupted
helical conformation. Glutamic and aspartic acids are colored red and
arginine green.bModel for the functional TPC1 with the antiparallel
CTH dimer enlarged (Shaker domains I and II are differently colored
in each monomer). A head-to-tail conformation is assumed for the
TPC1 dimer, as suggested by Rietdorf et al. [10]
The function of the two-pore channel TPC1 depends on dimerization of its carboxy-terminal helix 2577
123
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
Discussion
Here, we identified the presence of a carboxy-terminal
helix (CTH) in TPC1 and documented its role for channel
dimerization and function. Loss of the CTH rendered the
channel inactive while point mutations within the CTH
resulted in severely reduced channel activity. Corre-
spondingly, BiFC analysis and Co-IP experiments reported
on the interaction of the C-terminus, which was reduced by
mutations within the CTH. A residual current of 10 % in
the TPC1-3LA and -3LP mutant may indicate that in
contrast to the CTH peptide the complete C-terminus may
dimerize to a small extent even with three leucines
replaced. Although we cannot exclude the possibility that
in addition to the CTH other parts of the C-terminus are
involved in the dimerization process, the lack of the CTH
was sufficient to render the channel silent.
Dimerization of the CTH was directly shown in MST
assays and MD simulations. The sequential multiscale
simulations revealed that the wild type CTH dimer
preferably adopts an antiparallel coiled-coil conformation.
This antiparallel orientation is in accordance with an
assumed head-to-tail configuration of the channels [10].
The fact that leucine to alanine mutations in the inner helix
drastically impaired channel function is in accordance with
the observation that although alanine is a residue of high
helix propensity, it can decrease the stability of the helix
(dimer) [66,68]. Indeed, coarse-grained MD simulations
discovered mutated dimers that exhibit a significant
increase of flexibility, pointing to less stable dimers. These
results were corroborated by MST analyses, which showed
that the wild type CTH dimerized with a K
d
of 3.9 lM
while the mutants did not interact. Apparently, the CTH is
not required for correct assembly of the channel in the
membrane. However, our results suggest that the CTH is
crucial for stabilization of the TPC1 dimer and coupled
with dimerization also for channel gating.
Gating of ion channels by their soluble termini is a
common feature of ion channels [6973], and often mod-
ification of the C-terminus induces alterations of the
voltage dependence [7477]. In comparison, mutations
within and near the CTH did not feedback onto the voltage
dependence of TPC1. In this respect, TPC1 shares struc-
tural and functional properties with voltage-dependent
sodium (Na
V
) channels. Deletion of the complete or distal
part of the C-terminus from a prokaryotic Na
V
resulted in a
complete or almost complete loss of sodium currents,
respectively, while the trafficking of the channel and the
voltage dependence of activation were unaffected [70].
Combined MD simulations and electron paramagnetic
resonance spectroscopy showed that a C-terminal helix of
this channel forms a coiled-coil bundle involving four
subunits, and that this tetramerization is essential for cou-
pling to channel opening via a proximal C-terminal linker
following S6. Most notably, the linker contains a nega-
tively charged cluster, and several glutamate residues are
also conserved in the plant TPC C-termini following S6
(Fig. 2). A role of the coiled-coil in stabilizing the sodium
channel tetramer or dimer in case of TPC1, and in enabling
the opening and closing of the pore during gating without
disrupting the quaternary structure may thus represent a
mechanism also valid for two-pore channels, which has to
be further evaluated in future studies.
Besides voltage changes, TPC1 is activated by binding
of Ca
2?
ions to the EF-hands in the central linker domain
between transmembrane S6 and S7, which apparently sta-
bilizes the open state [13,78,79]. The rabbit skeletal
muscle type 1 ryanodine receptor RyR1 is an intracellular
Ca
2?
release channel, which also belongs to the six-
transmembrane superfamily. Similar to TPC1, activation of
RyR1 by Ca
2?
involves cytosolic EF-hands, here present in
the N-terminus, while the C-terminus homodimerizes. An
essential function of the RyR1 C-terminus requires the last
15 amino acids, deletion of which abolishes channel
function [80]. Recently, the crystal structure of the RyR1
revealed a putative mechanism for Ca
2?
-mediated gating
involving the C-terminus [73,81]. In RyR1 Ca
2?
-depen-
dent changes in the conformation of the N-terminal EF-
hand containing domain are transmitted to the pore via
contacts with the C-terminal domain, inducing a change of
the cytosolic aperture of the channel and stabilizing the
open state [73,81]. A RyR1-like allosteric mechanism may
therefore also account for the role of the C-terminus in the
Ca
2?
- and voltage-dependent activity of TPC1.
TPC1 is a large-conductance channel and is therefore
tightly regulated to prevent ion leakage from the vacuole
[reviewed in 2]. The many ionic cytosolic regulators, such as
H
?
,Ca
2?
,Mg
2?
,K
?
, and Na
?
, exert their effects via
alteration of the voltage dependence of the channel. In
contrast, 14-3-3 proteins inhibit TPC1 activity by about
90 % within 10 s without any changes in the voltage
dependence [36]. A similar reduction in channel activity
without affecting the voltage dependence was observed for
the CTH mutants described here. Likewise, application of
the synthetic CTH peptide to the vacuolar membrane rapidly
inhibited TPC1 currents (Fig. S7). It is therefore tempting to
speculate that the 14-3-3 protein GRF6, which regulates
TPC1 in Arabidopsis [35], interferes with the dimerization of
the carboxyl-termini. A region partly overlapping with the
CTH contains two serines (S706, S708) and constitutes a
putative 14-3-3 binding site [705-KSRSQR, 82]. Accumu-
lation of GRF6 dimers may thus push the CTHs apart, which
will destabilize the TPC1 dimer and induce a rapid closure of
the channel pore. Alternatively, 14-3-3 proteins may act
2578 N. Larisch et al.
123
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
more indirectly, because a second putative 14-3-3 binding
site is predicted to be located in the linker domain between
the two EF-hands at T359 [83]. Given a RyR1-like allosteric
mechanism, 14-3-3 binding in the linker may interfere with
the coordinated mechanism, which links Ca
2?
binding to the
activation of the channel via the C-terminal a-helical region.
In any case, the coiled-coil mediated dimerization of wild
type CTH’s as a prerequisite for channel opening adds a very
rapid regulatory mechanism to shut down this large vacuolar
conductance, either by destabilizing the TPC1 dimer, or by
preventing the allosteric coupling to the channel pore.
Acknowledgments This work was supported by the German Sci-
ence Foundation (DFG): Research Training Group 1962-Dynamic
Interactions at Biological Membranes. Computer time was provided
by the Computing Center of the University Erlangen-Nu
¨rnberg
(RRZE). We would like to thank Joanna Bogdanska-Urbaniak and
Gudrun Steingra
¨ber for technical assistance, and Norbert Sauer (FAU
Erlangen-Nu
¨rnberg) for sharing the confocal microscope.
Open Access This article is distributed under the terms of the
Creative Commons Attribution 4.0 International License (http://
creativecommons.org/licenses/by/4.0/), which permits unrestricted
use, distribution, and reproduction in any medium, provided you give
appropriate credit to the original author(s) and the source, provide a
link to the Creative Commons license, and indicate if changes were
made.
References
1. Zhu MX, Ma J, Parrington J, Galione A, Evans AM (2010) TPCs:
endolysosomal channels for Ca
2?
mobilization from acidic
organelles triggered by NAADP. FEBS Lett 584(10):1966–1974.
doi:10.1016/j.febslet.2010.02.028
2. Hedrich R, Marten I (2011) TPC1-SV channels gain shape. Mol
Plant 4(3):428–441. doi:10.1093/mp/ssr017
3. Choi WG, Toyota M, Kim SH, Hilleary R, Gilroy S (2014) Salt
stress-induced Ca
2?
waves are associated with rapid, long-dis-
tance root-to-shoot signaling in plants. Proc Natl Acad Sci USA
111(17):6497–6502. doi:10.1073/pnas.1319955111
4. Lin-Moshier Y, Keebler MV, Hooper R, Boulware MJ, Liu X,
Churamani D, Abood ME, Walseth TF, Brailoiu E, Patel S,
Marchant JS (2014) The two-pore channel (TPC) interactome
unmasks isoform-specific roles for TPCs in endolysosomal
morphology and cell pigmentation. Proc Natl Acad Sci USA
111(36):13087–13092. doi:10.1073/pnas.1407004111
5. Sakurai Y, Kolokoltsov AA, Chen C-C, Tidwell MW, Bauta WE,
Klugbauer N, Grimm C, Wahl-Schott C, Biel M, Davey RA
(2015) Two-pore channels control Ebola virus host cell entry and
are drug targets for disease treatment. Science
347(6225):995–998. doi:10.1126/science.1258758
6. Wang X, Zhang X, Dong XP, Samie M, Li X, Cheng X, Goschka
A, Shen D, Zhou Y, Harlow J, Zhu MX, Clapham DE, Ren D, Xu
H (2012) TPC proteins are phosphoinositide- activated sodium-
selective ion channels in endosomes and lysosomes. Cell
151(2):372–383. doi:10.1016/j.cell.2012.08.036
7. Cang C, Bekele B, Ren D (2014) The voltage-gated sodium
channel TPC1 confers endolysosomal excitability. Nat Chem
Biol 10(6):463–469. doi:10.1038/nchembio.1522
8. Galione A (2011) NAADP Receptors. In: Perspectives in Biology,
vol 3, 1. Cold Spring Harbor. doi:10.1101/cshperspect.a004036
9. Zong X, Schieder M, Cuny H, Fenske S, Gruner C, Rotzer K,
Griesbeck O, Harz H, Biel M, Wahl-Schott C (2009) The two-
pore channel TPCN2 mediates NAADP-dependent Ca
2?
-release
from lysosomal stores. Pflugers Arch 458(5):891–899. doi:10.
1007/s00424-009-0690-y
10. Rietdorf K, Funnell TM, Ruas M, Heinemann J, Parrington J,
Galione A (2011) Two-pore channels form homo- and hetero-
dimers. J Biol Chem 286(43):37058–37062. doi:10.1074/jbc.
C111.289835
11. Larisch N, Schulze C, Galione A, Dietrich P (2012) An N-ter-
minal dileucine motif directs two-pore channels to the tonoplast
of plant cells. Traffic 13(7):1012–1022. doi:10.1111/j.1600-0854.
2012.01366.x
12. Brailoiu E, Rahman T, Churamani D, Prole DL, Brailoiu GC,
Hooper R, Taylor CW, Patel S (2010) An NAADP-gated two-
pore channel targeted to the plasma membrane uncouples trig-
gering from amplifying Ca
2?
signals. J Biol Chem
285(49):38511–38516. doi:10.1074/jbc.M110.162073
13. Schulze C, Sticht H, Meyerhoff P, Dietrich P (2011) Differential
Contribution of EF-hands to the Ca
2?
-dependent activation in the
plant two-pore channel TPC1. Plant J. doi:10.1111/j.1365-313X.
2011.04697.x
14. Hedrich R, Neher E (1987) Cytoplasmic calcium regulates volt-
age-dependent ion channels in plant vacuoles. Nature
329(6142):833–836
15. Boccaccio A, Scholz-Starke J, Hamamoto S, Larisch N, Festa M,
Gutla P, Costa A, Dietrich P, Uozumi N, Carpaneto A (2014) The
phosphoinositide PI(3,5)P
2
mediates activation of mammalian but
not plant TPC proteins: functional expression of endolysosomal
channels in yeast and plant cells. Cell Mol Life Sci. doi:10.1007/
s00018-014-1623-2
16. Pitt SJ, Funnell TM, Sitsapesan M, Venturi E, Rietdorf K, Ruas
M, Ganesan A, Gosain R, Churchill GC, Zhu MX, Parrington J,
Galione A, Sitsapesan R (2010) TPC2 is a novel NAADP-sen-
sitive Ca
2?
release channel, operating as a dual sensor of luminal
pH and Ca
2?
. J Biol Chem 285(45):35039–35046. doi:10.1074/
jbc.M110.156927
17. Brailoiu GC, Brailoiu E, Parkesh R, Galione A, Churchill GC,
Patel S, Dun NJ (2009) NAADP-mediated channel ‘chatter’ in
neurons of the rat medulla oblongata. Biochem J 419(1):91–97.
doi:10.1042/BJ20081138
18. Calcraft PJ, Ruas M, Pan Z, Cheng X, Arredouani A, Hao X,
Tang J, Rietdorf K, Teboul L, Chuang KT, Lin P, Xiao R, Wang
C, Zhu Y, Lin Y, Wyatt CN, Parrington J, Ma J, Evans AM,
Galione A, Zhu MX (2009) NAADP mobilizes calcium from
acidic organelles through two-pore channels. Nature
459(7246):596–600. doi:10.1038/nature08030
19. Peiter E, Maathuis FJ, Mills LN, Knight H, Pelloux J, Hether-
ington AM, Sanders D (2005) The vacuolar Ca
2?
-activated
channel TPC1 regulates germination and stomatal movement.
Nature 434(7031):404–408. doi:10.1038/nature03381
20. Ranf S, Wu
¨nnenberg P, Lee J, Becker D, Dunkel M, Hedrich R,
Scheel D, Dietrich P (2008) Loss of the vacuolar cation channel,
AtTPC1, does not impair Ca
2?
signals induced by abiotic and
biotic stresses. Plant J 53(2):287–299. doi:10.1111/j.1365-313X.
2007.03342.x
21. Ward JM, Schroeder JI (1994) Calcium-activated K
?
channels and
calcium-induced calcium release by slow vacuolar ion channels in
guard cell vacuoles implicated in the control of stomatal closure.
Plant Cell 6(5):669–683. doi:10.1105/tpc.6.5.669
22. Pottosin II, Dobrovinskaya OR, Mun
˜iz J (2001) Conduction of
monovalent and divalent cations in the slow vacuolar channel.
J Membr Biol 181(1):55–65. doi:10.1007/s0023200100073
The function of the two-pore channel TPC1 depends on dimerization of its carboxy-terminal helix 2579
123
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
23. Pottosin II, Tikhonova LI, Hedrich R, Scho
¨nknecht G (1997)
Slowly activating vacuolar channels can not mediate Ca
2?
-in-
duced Ca
2?
release. Plant J 12(6):1387–1398. doi:10.1046/j.
1365-313x.1997.12061387.x
24. Pottosin I, Wherrett T, Shabala S (2009) SV channels dominate
the vacuolar Ca2?release during intracellular signaling. FEBS
Lett 583(5):921–926. doi:10.1016/j.febslet.2009.02.009
25. Islam MM, Munemasa S, Hossain MA, Nakamura Y, Mori IC,
Murata Y (2010) Roles of AtTPC1, vacuolar two pore channel 1,
in Arabidopsis stomatal closure. Plant Cell Physiol
51(2):302–311. doi:10.1093/pcp/pcq001
26. Beyhl D, Ho
¨rtensteiner S, Martinoia E, Farmer EE, Fromm J,
Marten I, Hedrich R (2009) The fou2 mutation in the major
vacuolar cation channel TPC1 confers tolerance to inhibitory
luminal calcium. Plant J 58(5):715–723. doi:10.1111/j.1365-
313X.2009.03820.x
27. Bonaventure G, Gfeller A, Proebsting WM, Ho
¨rtensteiner S,
Che
´telat A, Martinoia E, Farmer EE (2007) A gain-of-function
allele of TPC1 activates oxylipin biogenesis after leaf wounding
in Arabidopsis. Plant J 49(5):889–898. doi:10.1111/j.1365-313X.
2006.03002.x
28. Bonaventure G, Gfeller A, Rodrı
´guez VM, Armand F, Farmer EE
(2007) The fou2 gain-of-function allele and the wild-type allele
of Two Pore Channel 1 contribute to different extents or by
different mechanisms to defense gene expression in Arabidopsis.
Plant Cell Physiol 48(12):1775–1789. doi:10.1093/pcp/pcm151
29. Dadacz-Narloch B, Beyhl D, Larisch C, Lopez-Sanjurjo EJ,
Reski R, Kuchitsu K, Mu
¨ller TD, Becker D, Scho
¨nknecht G,
Hedrich R (2011) A novel calcium binding site in the slow
vacuolar cation channel TPC1 senses luminal calcium levels.
Plant Cell 23(7):2696–2707. doi:10.1105/tpc.111.086751
30. Pottosin I, Martı
´nez-Este
´vez M, Dobrovinskaya O, Mun
˜iz J,
Scho
¨nknecht G (2004) Mechanism of luminal Ca
2?
and Mg
2?
action on the vacuolar slowly activating channels. Planta
219(6):1057–1070. doi:10.1007/s00425-004-1293-7
31. Schulz-Lessdorf B, Hedrich R (1995) Protons and calcium
modulate SV-type channels in the vacuolar-lysosomal compart-
ment—channel interaction with calmodulin inhibitors. Planta
197(4):655–671. doi:10.1007/BF00191574
32. Ivashikina N, Hedrich R (2005) K
?
currents through SV-type
vacuolar channels are sensitive to elevated luminal sodium levels.
Plant J 41(4):606–614. doi:10.1111/j.1365-313X.2004.02324.x
33. Dobrovinskaya O, Muniz J, Pottosin I (1999) Inhibition of vac-
uolar ion channels by polyamines. J Membr Biol 167(2):127–140
34. Gutla PV, Boccaccio A, De Angeli A, Gambale F, Carpaneto A
(2012) Modulation of plant TPC channels by polyunsaturated
fatty acids. J Exp Bot 63(17):6187–6197. doi:10.1093/jxb/ers272
35. Latz A, Becker D, Hekman M, Mu
¨ller T, Beyhl D, Marten I, Eing
C, Fischer A, Dunkel M, Bertl A, Rapp UR, Hedrich R (2007)
TPK1, a Ca
2?
-regulated Arabidopsis vacuole two-pore K
?
channel is activated by 14-3-3 proteins. Plant J 52(3):449–459.
doi:10.1111/j.1365-313X.2007.03255.x
36. van den Wijngaard PW, Bunney TD, Roobeek I, Scho
¨nknecht G, de
Boer AH (2001)Slow vacuolar channelsfrom barley mesophyllcells
are regulated by 14-3-3 proteins. FEBS Lett 488(1–2):100–104
37. Allen GJ, Sanders D (1995) Calcineurin, a Type 2B Protein
Phosphatase, Modulates the Ca
2?
-Permeable Slow Vacuolar Ion
Channel of Stomatal Guard Cells. Plant Cell 7(9):1473–1483.
doi:10.1105/tpc.7.9.1473
38. Bethke PC, Jones RL (1997) Reversible protein phosphorylation
regulates the activity of the slow-vacuolar ion channel. Plant J
11(6):1227–1235. doi:10.1046/j.1365-313X.1997.11061227.x
39. Gehl C, Waadt R, Kudla J, Mendel RR, Ha
¨nsch R (2009) New
GATEWAY vectors for high throughput analyses of protein-
protein interactions by bimolecular fluorescence complementa-
tion. Mol Plant 2(5):1051–1058. doi:10.1093/mp/ssp040
40. Rupp RA, Snider L, Weintraub H (1994) Xenopus embryos
regulate the nuclear localization of XMyoD. Genes Dev
8(11):1311–1323
41. Hartmann J, Fischer C, Dietrich P, Sauter M (2014) Kinase
activity and calmodulin binding are essential for growth signaling
by the phytosulfokine receptor PSKR1. Plant J 78(2):192–202.
doi:10.1111/tpj.12460
42. Deblaere R, Bytebier B, De Greve H, Deboeck F, Schell J, Van
Montagu M, Leemans J (1985) Efficient octopine Ti plasmid-
derived vectors for agrobacterium-mediated gene transfer to
plants. Nucl Acids Res 13(13):4777–4788
43. Voinnet O, Rivas S, Mestre P, Baulcombe D (2003) An enhanced
transient expression system in plants based on suppression of
gene silencing by the p19 protein of tomato bushy stunt virus.
Plant J 33(5):949–956
44. Hess B, Kutzner C, van der Spoel D, Lindahl E (2008) GRO-
MACS 4: algorithms for highly efficient, load-balanced, and
scalable molecular simulation. J Chem Theory Comput
4(3):435–447. doi:10.1021/ct700301q
45. Schro
¨dinger, LLC (2010) The PyMOL Molecular Graphics Sys-
tem, Version 1.3r1
46. de Jong DH, Singh G, Bennett WFD, Arnarez C, Wassenaar TA,
Scha
¨fer LV, Periole X, Tieleman DP, Marrink SJ (2013) Improved
parameters for the martini coarse-grained protein force field.
J Chem Theory Comput 9(1):687–697. doi:10.1021/ct300646g
47. Wassenaar TA, Pluhackova K, Moussatova A, Sengupta D,
Marrink SJ, Tieleman DP, Bo
¨ckmann RA (2015) High-through-
put simulations of dimer and trimer assembly of membrane
proteins. The DAFT approach. J Chem Theory Comput. doi:10.
1021/ct5010092
48. Yesylevskyy SO, Scha
¨fer LV, Sengupta D, Marrink SJ (2010)
Polarizable water model for the coarse-grained MARTINI force
field. PLoS Comput Biol 6(6):e1000810. doi:10.1371/journal.
pcbi.1000810
49. Pluhackova K, Wassenaar TA, Bo
¨ckmann RA (2013) Molecular
Dynamics Simulations of Membrane Proteins. In: Rapaport D,
Herrmann JM (eds) Membr Biogen, vol 1033, 1st edn. Humana
Press, New york, pp 85–101. doi:10.1007/978-1-62703-487-6
50. Wassenaar TA, Ingo
´lfsson HI, Bo
¨ckmann RA, Tieleman DP, Mar-
rink SJ (2015) Computational lipidomics with insane: a versatiletool
for generating custom membranes for molecular simulations.
JChemTheoryComput.doi:10.1021/acs.jctc.5b00209
51. Pluhackova K, Wassenaar TA, Kirsch S, Bo
¨ckmann RA (2015)
Spontaneous adsorption of coiled-coil model peptides K and E to
a mixed lipid bilayer. J Phys Chem B 119(12):4396–4408. doi:10.
1021/acs.jpcb.5b00434
52. Case DA, Babin V, Berryman JT, Betz RM, Cai Q, Cerutti DS,
Cheatham TE, Darden TA, Duke RE, Gohlke H, Goetz AW,
Gusarov S, Homeyer N, Janowski P, Kaus J, Kolossva
´ry I,
Kovalenko A, Lee TS, LeGrand S, Luchko T, Luo R, Madej B,
Merz KM, Paesani F, Roe DR, Roitberg A, Sagui C, Salomon-
Ferrer R, Seabra G, Simmerling CL, Smith W, Swails J, Walker
RC, Wang J, Wolf RM, Wu X, Kollman PA (2014) AMBER 14.
University of California, oakland
53. Wassenaar TA, Pluhackova K, Bo
¨ckmann RA, Marrink SJ,
Tieleman DP (2014) Going backward: a flexible geometric
approach to reverse transformation from coarse grained to ato-
mistic models. J Chem Theory Comput 10(2):676–690. doi:10.
1021/ct400617g
54. Niu X, Bressan RA, Hasegawa PM, Pardo JM (1995) Ion
homeostasis in NaCl stress environments. Plant Physiol
109(3):735–742
55. Jorgensen WL, Chandrasekhar J, Madura JD, Impey RW, Klein
ML (1983) Comparison of simple potential functions for simu-
lating liquid water. J Chem Phys 79(2):926–935. doi:10.1063/1.
445869
2580 N. Larisch et al.
123
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
56. Wassenaar TA, Ingo
´lfsson HI, Prieß M, Marrink SJ, Scha
¨fer LV
(2013) Mixing MARTINI: electrostatic coupling in hybrid ato-
mistic–coarse-grained biomolecular simulations. J Phys Chem B
117(13):3516–3530. doi:10.1021/jp311533p
57. Berendsen HJC, Postma JPM, van Gunsteren WF, DiNola A,
Haak JR (1984) Molecular dynamics with coupling to an external
bath. J Chem Physics 81(8):3684–3690. doi:10.1063/1.448118
58. Bussi G, Donadio D, Parrinello M (2007) Canonical sampling
through velocity rescaling. J Chem Physics 126(1):014101.
doi:10.1063/1.2408420
59. Darden T, York D, Pedersen L (1993) Particle mesh Ewald: an
Nlog(N) method for Ewald sums in large systems. J Chem
Physics 98(12):10089–10092. doi:10.1063/1.464397
60. Nose
´S (1984) A molecular dynamics method for simulations in
the canonical ensemble. Mol Phys 52(2):255–268. doi:10.1080/
00268978400101201
61. Hoover WG (1985) Canonical dynamics: equilibrium phase-
space distributions. Phys Rev A 31(3):1695–1697
62. Parrinello M, Rahman A (1981) Polymorphic transitions in single
crystals: a new molecular dynamics method. J Appl Phys
52(12):7182–7190. doi:10.1063/1.328693
63. Lindorff-Larsen K, Piana S, Palmo K, Maragakis P, Klepeis JL,
Dror RO, Shaw DE (2010) Improved side-chain torsion potentials
for the Amber ff99SB protein force field. Proteins
78(8):1950–1958. doi:10.1002/prot.22711
64. Wassenaar TA (2006) Molecular dynamics of sense and sensi-
bility in processing and analysis of data. University Library
Groningen, Groningen
65. Kotelchuk D, Scheraga HA (1968) The influence of short-range
interactions on protein conformation, I. Side chain-backbone inter-
actions within a single peptide unit. Proc Natl Acad Sci USA 61:4
66. Zhang XJ, Baase WA, Matthews BW (1992) Multiple alanine
replacements within alpha-helix 126-134 of T4 lysozyme have
independent, additive effects on both structure and stability.
Protein Sci 1(6):761–776
67. Pottosin I-I, Tikhonova L-I, Hedrich R, Schoenknecht G (1997)
Slowly activating vacuolar channels can not mediate Ca
2?
-in-
duced Ca
2?
release. Plant J 12(6):1387–1398
68. Chou PY, GD F (1973) Structural and functional role of Leu
residues in proteins. J Mol Biol 74:263–281
69. Gustina AS, Trudeau MC (2011) hERG potassium channel gating
is mediated by N- and C-terminal region interactions. J Gen
Physiol 137(3):315–325. doi:10.1085/jgp.201010582
70. Bagneris C, Decaen PG, Hall BA, Naylor CE, Clapham DE, Kay
CW, Wallace BA (2013) Role of the C-terminal domain in the
structure and function of tetrameric sodium channels. Nat Com-
mun 4:2465. doi:10.1038/ncomms3465
71. Nilius B, Prenen J, Tang J, Wang C, Owsianik G, Janssens A,
Voets T, Zhu MX (2005) Regulation of the Ca
2?
sensitivity of the
nonselective cation channel TRPM4. J Biol Chem
280(8):6423–6433. doi:10.1074/jbc.M411089200
72. Koschak A (2010) Impact of gating modulation in CaV1.3 L-type
calcium channels. Channels (Austin) 4(6):523–525
73. Yan Z, Bai XC, Yan C, Wu J, Li Z, Xie T, Peng W, Yin CC, Li
X, Scheres SH, Shi Y, Yan N (2015) Structure of the rabbit
ryanodine receptor RyR1 at near-atomic resolution. Nature
517(7532):50–55. doi:10.1038/nature14063
74. Mumm P, Imes D, Martinoia E, Al-Rasheid KA, Geiger D,
Marten I, Hedrich R (2013) C-terminus-mediated voltage gating
of Arabidopsis guard cell anion channel QUAC1. Mol Plant
6(5):1550–1563. doi:10.1093/mp/sst008
75. Marten I, Hoshi T (1997) Voltage-dependent gating characteris-
tics of the K
?
channel KAT1 depend on the Nand Ctermini. Proc
Natl Acad Sci USA 94(7):3448–3453
76. Hatano N, Ohya S, Muraki K, Clark RB, Giles WR, Imaizumi Y
(2004) Two arginines in the cytoplasmic C-terminal domain are
essential for voltage-dependent regulation of A-type K?current
in the Kv4 channel subfamily. J Biol Chem 279(7):5450–5459.
doi:10.1074/jbc.M302034200
77. Nilius B, Talavera K, Owsianik G, Prenen J, Droogmans G, Voets
T (2005) Gating of TRP channels: a voltage connection? J Phys-
iol 567(Pt 1):35–44. doi:10.1113/jphysiol.2005.088377
78. Pei ZM, Ward JM, Schroeder JI (1999) Magnesium sensitizes
slow vacuolar channels to physiological cytosolic calcium and
inhibits fast vacuolar channels in fava bean guard cell vacuoles.
Plant Phys 121(3):977–986
79. Pottosin I, Scho
¨nknecht G (2007) Vacuolar calcium channels.
J Exp Bot 58(7):1559–1569. doi:10.1093/jxb/erm035
80. Lee EH, Allen PD (2007) Homo-dimerization of RyR1 C-ter-
minus via charged residues in random coils or in an alpha-helix.
Exp Mol Med 39(5):594–602. doi:10.1038/emm.2007.65
81. Zalk R, Clarke OB, des Georges A, Grassucci RA, Reiken S,
Mancia F, Hendrickson WA, Frank J, Marks AR (2015) Structure
of a mammalian ryanodine receptor. Nature 517(7532):44–49.
doi:10.1038/nature13950
82. Dinkel H, Van Roey K, Michael S, Davey NE, Weatheritt RJ,
Born D, Speck T, Kru
¨ger D, Grebnev G, Kuban M, Strumillo M,
Uyar B, Budd A, Altenberg B, Seiler M, Chemes LB, Glavina J,
Sanchez IE, Diella F, Gibson TJ (2014) The eukaryotic linear
motif resource ELM: 10 years and counting. Nucl Acids Res
42:D259–D266. doi:10.1093/nar/gkt1047 (Database issue)
83. Obenauer JC, Cantley LC, Yaffe MB (2003) Scansite 2.0: Pro-
teome-wide prediction of cell signaling interactions using short
sequence motifs. Nucl Acids Res 31(13):3635–3641
84. Schwacke R, Schneider A, van der Graaff E, Fischer K, Catoni E,
Desimone M, Frommer WB, Flu
¨gge U-I, Kunze R (2003)
ARAMEMNON, a novel database for Arabidopsis integral
membrane proteins. Plant Physiol 131(1):16–26. doi:10.1104/pp.
011577
85. Yang J, Yan R, Roy A, Xu D, Poisson J, Zhang Y (2015) The
I-TASSER Suite: protein structure and function prediction. Nat
Meth 12(1):7–8. doi:10.1038/nmeth.3213
86. Roy A, Kucukural A, Zhang Y (2010) I-TASSER: a unified
platform for automated protein structure and function prediction.
Nat Protoc 5(4):725–738. doi:10.1038/nprot.2010.5
87. Zhang Y (2008) I-TASSER server for protein 3D structure pre-
diction. BMC Bioinformatics 9:40. doi:10.1186/1471-2105-9-40
88. Xu D, Zhang Y (2012) Ab initio protein structure assembly using
continuous structure fragments and optimized knowledge-based
force field. Proteins 80(7):1715–1735. doi:10.1002/prot.24065
89. Remmert M, Biegert A, Hauser A, Soding J (2012) HHblits:
lightning-fast iterative protein sequence searching by HMM-
HMM alignment. Nat Meth 9 (2):173–175. doi:http://www.
nature.com/nmeth/journal/v9/n2/abs/nmeth.1818.html—supple-
mentary-information
90. So
¨ding J (2005) Protein homology detection by HMM-HMM
comparison. Bioinformatics 21(7):951–960. doi:10.1093/
bioinformatics/bti125
91. So
¨ding J, Biegert A, Lupas AN (2005) The HHpred interactive
server for protein homology detection and structure prediction.
Nucl Acids Res 33:W244–W248. doi:10.1093/nar/gki408 (Web
Server issue)
92. Kelley LA, Sternberg MJ (2009) Protein structure prediction on
the web: a case study using the phyre server. Nat Protoc
4(3):363–371. doi:10.1038/nprot.2009.2
93. Karimi M, Inze D, Depicker A (2002) GATEWAY vectors for
Agrobacterium-mediated plant transformation. Trends Plant Sci
7(5):193–195
94. Furuichi T, Cunningham KW, Muto S (2001) A putative two pore
channel AtTPC1 mediates Ca
2?
flux in Arabidopsis leaf cells.
Plant Cell Physiol 42(9):900–905
The function of the two-pore channel TPC1 depends on dimerization of its carboxy-terminal helix 2581
123
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
1.
2.
3.
4.
5.
6.
Terms and Conditions
Springer Nature journal content, brought to you courtesy of Springer Nature Customer Service Center GmbH (“Springer Nature”).
Springer Nature supports a reasonable amount of sharing of research papers by authors, subscribers and authorised users (“Users”), for small-
scale personal, non-commercial use provided that all copyright, trade and service marks and other proprietary notices are maintained. By
accessing, sharing, receiving or otherwise using the Springer Nature journal content you agree to these terms of use (“Terms”). For these
purposes, Springer Nature considers academic use (by researchers and students) to be non-commercial.
These Terms are supplementary and will apply in addition to any applicable website terms and conditions, a relevant site licence or a personal
subscription. These Terms will prevail over any conflict or ambiguity with regards to the relevant terms, a site licence or a personal subscription
(to the extent of the conflict or ambiguity only). For Creative Commons-licensed articles, the terms of the Creative Commons license used will
apply.
We collect and use personal data to provide access to the Springer Nature journal content. We may also use these personal data internally within
ResearchGate and Springer Nature and as agreed share it, in an anonymised way, for purposes of tracking, analysis and reporting. We will not
otherwise disclose your personal data outside the ResearchGate or the Springer Nature group of companies unless we have your permission as
detailed in the Privacy Policy.
While Users may use the Springer Nature journal content for small scale, personal non-commercial use, it is important to note that Users may
not:
use such content for the purpose of providing other users with access on a regular or large scale basis or as a means to circumvent access
control;
use such content where to do so would be considered a criminal or statutory offence in any jurisdiction, or gives rise to civil liability, or is
otherwise unlawful;
falsely or misleadingly imply or suggest endorsement, approval , sponsorship, or association unless explicitly agreed to by Springer Nature in
writing;
use bots or other automated methods to access the content or redirect messages
override any security feature or exclusionary protocol; or
share the content in order to create substitute for Springer Nature products or services or a systematic database of Springer Nature journal
content.
In line with the restriction against commercial use, Springer Nature does not permit the creation of a product or service that creates revenue,
royalties, rent or income from our content or its inclusion as part of a paid for service or for other commercial gain. Springer Nature journal
content cannot be used for inter-library loans and librarians may not upload Springer Nature journal content on a large scale into their, or any
other, institutional repository.
These terms of use are reviewed regularly and may be amended at any time. Springer Nature is not obligated to publish any information or
content on this website and may remove it or features or functionality at our sole discretion, at any time with or without notice. Springer Nature
may revoke this licence to you at any time and remove access to any copies of the Springer Nature journal content which have been saved.
To the fullest extent permitted by law, Springer Nature makes no warranties, representations or guarantees to Users, either express or implied
with respect to the Springer nature journal content and all parties disclaim and waive any implied warranties or warranties imposed by law,
including merchantability or fitness for any particular purpose.
Please note that these rights do not automatically extend to content, data or other material published by Springer Nature that may be licensed
from third parties.
If you would like to use or distribute our Springer Nature journal content to a wider audience or on a regular basis or in any other manner not
expressly permitted by these Terms, please contact Springer Nature at
onlineservice@springernature.com
... A predicted helical domain in the C terminus (S U in Figure 1) was suggested to be essential for plant TPC assembly and function [46]. Both plant and mammalian TPC1 channels are voltagedependent, conferring electrical excitability to the endolysosomal/vacuolar membrane [47,48]. ...
... The N-terminal SN helix lies on top of helix S0 (domain naming is given in Figure 1). In this conformational state the C terminus (i.e., the helix in the cytoplasm following IIS6, S U ), which is indispensable for channel activation [46], interacts with and packs against the EF hand (i.e., the E2-F2 linker), thereby possibly keeping the channel pore closed. When the second voltage-sensor domain switches into its active conformation, as seen from the AtTPC1 DDE mutant, the globular structure formed by these cytoplasmic N-terminal and C-terminal domains (i.e., the S0 and the S U helices) dissociates leading to a rearrangement in the EF hands, which now bind three calcium ions. ...
Article
Across phyla, voltage-gated ion channels (VGICs) allow excitability. The vacuolar two-pore channel AtTPC1 from the tiny mustard plant Arabidopsis thaliana has emerged as a paradigm for deciphering the role of voltage and calcium signals in membrane excitation. Among the numerous experimentally determined structures of VGICs, AtTPC1 was the first to be revealed in a closed and resting state, fueling speculation about structural rearrangements during channel activation. Two independent reports on the structure of a partially opened AtTPC1 channel protein have led to working models that offer promising insights into the molecular switches associated with the gating process. We review new structure-function models and also discuss the evolutionary impact of two-pore channels (TPCs) on K+ homeostasis and vacuolar excitability.
... Two-pore calcium channel protein 1 (TPC1), a voltagegated Ca 2+ channel in plant vacuolar membrane, plays an important role in various signaling responses under abiotic stress (Choi et al. 2014;Larisch et al. 2016). The drought stress-induced expression of TPC1 by fourfold compared to the control plants, whereas steep induction (9.4-fold) in the expression was observed in WRS7-inoculated plants. ...
Article
Full-text available
Some plant growth-promoting bacteria (PGPR) can ameliorate abiotic stressors like drought stress and promote plant growth. The present study investigated various drought-tolerant mechanisms of Enterobacter bugandensis WRS7, a rhizospheric isolate, by which it alleviates the deleterious effects of drought stress in wheat plants (Triticum aestivum L). The isolate WRS7 showed different plant growth-promoting properties, including nitrogen fixation, phosphate solubilization, siderophore production, phytohormone (indole acetic acid and gibberellic acid) production, exopolysaccharide secretion, and ACC deaminase activity. Its inoculation to wheat plants improved plant growth in terms of root/shoot growth and chlorophyll content. Its inoculation also exhibited drought stress ameliorating properties, including increased osmolyte content (proline and total soluble sugar), relative water content, catalase and superoxide dismutase activity, and decreased lipid peroxidation compared to non-inoculated plants. Our biochemical data were coherent with gene expression analysis of WRS7-treated plants, which showed altered expression of genes encoding antioxidant enzymes (CAT, APX, and GPX), osmolyte synthesis (P5CS, P5CR, and TPS1), biosynthesis of stress hormone genes (NCED, WZE, SAMS, ACS1, and ACO encoding proteins for the biosynthesis of abscisic acid and ethylene), and calcium transporter (TPC1) in the wheat plant. The regulation of the ethylene biosynthesis gene and modulation of TPC1 gene expression by PGPR E. bugandensis WRS7 in wheat plants highlights its additional role in alleviating drought stress. The colonization study demonstrated the successful colonization of E. bugandensis WRS7 in wheat plants. Overall, the present study indicates that E. bugandensis WRS7 alleviates drought stress in wheat plants by differentially regulating various metabolic genes in treated plants.
... Salt stress-induced long-distance Ca 2+ wave is dependent on the activity of the vacuolar cation channel protein twopore channel 1 (TPC1) channel protein, which appears to contribute to whole-plant stress tolerance. In Arabidopsis, TPC1 plays an important role in cation homeostasis and vacuolar storage, mediating the passage of Na + and K + [54]. This study showed that two TPC1 (TPC1A and TPCB) genes showed up-regulation under alkali stress. ...
Article
Full-text available
Background Soil salinization and alkalization are widespread environmental problems that limit grapevine (Vitis vinifera L.) growth and yield. However, little is known about the response of grapevine to alkali stress. This study investigated the differences in physiological characteristics, chloroplast structure, transcriptome, and metabolome in grapevine plants under salt stress and alkali stress. Results We found that grapevine plants under salt stress and alkali stress showed leaf chlorosis, a decline in photosynthetic capacity, a decrease in chlorophyll content and Rubisco activity, an imbalance of Na⁺ and K⁺, and damaged chloroplast ultrastructure. Fv/Fm decreased under salt stress and alkali stress. NPQ increased under salt stress whereas decreased under alkali stress. Gene Ontology (GO) and Kyoto Encyclopedia of Genes and Genomes (KEGG) enrichment showed the differentially expressed genes (DEGs) induced by salt stress and alkali stress were involved in different biological processes and have varied molecular functions. The expression of stress genes involved in the ABA and MAPK signaling pathways was markedly altered by salt stress and alkali stress. The genes encoding ion transporter (AKT1, HKT1, NHX1, NHX2, TPC1A, TPC1B) were up-regulated under salt stress and alkali stress. Down-regulation in the expression of numerous genes in the ‘Porphyrin and chlorophyll metabolism’, ‘Photosynthesis-antenna proteins’, and ‘Photosynthesis’ pathways were observed under alkali stress. Many genes in the ‘Carbon fixation in photosynthetic organisms’ pathway in salt stress and alkali stress were down-regulated. Metabolome showed that 431 and 378 differentially accumulated metabolites (DAMs) were identified in salt stress and alkali stress, respectively. L-Glutamic acid and 5-Aminolevulinate involved in chlorophyll synthesis decreased under salt stress and alkali stress. The abundance of 19 DAMs under salt stress related to photosynthesis decreased. The abundance of 16 organic acids in salt stress and 22 in alkali stress increased respectively. Conclusions Our findings suggested that alkali stress had more adverse effects on grapevine leaves, chloroplast structure, ion balance, and photosynthesis than salt stress. Transcriptional and metabolic profiling showed that there were significant differences in the effects of salt stress and alkali stress on the expression of key genes and the abundance of pivotal metabolites in grapevine plants.
... This argued for a short distance communication, either via Ca 2+ flux or via a direct physical interaction. Notably, a dimerization of the TPC1 channels via C-termini was observed, and its disruption resulted in silent channels, not responding to cytosolic Ca 2+ and voltage (Larisch et al., 2016). ...
Article
Full-text available
Two-pore cation channel, TPC1, is ubiquitous in the vacuolar membrane of terrestrial plants and mediates the long distance signaling upon biotic and abiotic stresses. It possesses a wide pore, which transports small mono- and divalent cations. K ⁺ is transported more than 10-fold faster than Ca ²⁺ , which binds with a higher affinity within the pore. Key pore residues, responsible for Ca ²⁺ binding, have been recently identified. There is also a substantial progress in the mechanistic and structural understanding of the plant TPC1 gating by membrane voltage and cytosolic and luminal Ca ²⁺ . Collectively, these gating factors at resting conditions strongly reduce the potentially lethal Ca ²⁺ leak from the vacuole. Such tight control is impressive, bearing in mind high unitary conductance of the TPC1 and its abundance, with thousands of active channel copies per vacuole. But it remains a mystery how this high threshold is overcome upon signaling, and what type of signal is emitted by TPC1, whether it is Ca ²⁺ or electrical one, or a transduction via protein conformational change, independent on ion conductance. Here we discuss non-exclusive scenarios for the TPC1 integration into Ca ²⁺ , ROS and electrical signaling.
... In this study, the results showed that both salt stress and alkali stress induced the up-regulation of AKT1 gene in grapevine leave. The similar results was also found in soybean leaves and roots (Feng et al. 2021 (Larisch et al. 2016). In this study, the results showed that two TPC1 (TPC1A and TPCB) genes were showed up-regulation under alkali stress. ...
Preprint
Full-text available
Salt stress and alkali stress were two abiotic stressors which largely impact plant growth and development in grapevine ( Vitis vinifera ) plants. However, little is known about the mechanisms of grapevine plants response to salt stress and alkaline stress, and whether salt stress and alkali stress have differences in grapevine plants is still unidentified. Thus, we measured physiological parameters, identified chloroplast structure, and analyzed transcriptomics and metabolomics data to characterize the acclimation and difference of grapevine plant under neutral salt stress and alkali salt stress. The results showed that grapevine plants under salt stress and alkali stress were both accompanied by chlorosis of leaf, a decline in photosynthetic capacity, decrease in chlorophyll content and Rubisco activity, imbalance of Na ⁺ and K ⁺ , and poorer chloroplast ultrastructure, compared with control. The chlorophyll fluorescence index (Fv/Fm, NPQ) showed the different response to salt stress and alkali stress. Fv/Fm were both decrease in salt stress and alkali stress, whereas NPQ were only decreased in alkali stress. Alkali stress led more serious damage to phenotype, chloroplast microstructure and physiological characteristics in grapevine plants than salt stress. Transcriptomic data showed that salt stress and alkali stress induced the differential expression of stress genes involved in ABA signal and MAPK signal pathways as well as up-regulation of gene encoding ion transporter such as AKT1 , HKT1 , NHX1 , NHX2 , TPC1A and TPC1B. However, alkali stress induced more down-regulated genes in ‘Porphyrin and chlorophyll metabolism’, ‘Photosynthesis-antenna proteins’ and ‘Photosynthesis’ pathways compared with salt stress, and both two salt stresses induced many down-regulated DEGs in ‘Carbon fixation in photosynthetic organisms’ pathway. Metabolomics data showed that two substrate involved in chlorophyll synthesis including L-Glutamic acid were down-regulated both under salt stress and alkali stress, but 5-Aminolevulinate (ALA) were only down-regulated under salt stress. In the ‘Carbon fixation in photosynthetic organisms’ pathway, two salt stresses commonly induced the increase of key metabolites including D-Sedoheptuiose 7-phosphate, D-Erythrose-4-phosphate and Dihydroxyacetone phosphate involved in Calvin cycle, but 3-Phospho-D-glyceric acid, D-Fructose 6-Phosphate and Sedoheptulose showed converse change. Many sugar metabolites showed decreased levels under salt stress. However, alkali stress induced an increase in 7 sugar metabolites including Rhamnose, Sedoheptulose, D-Sedoheptulose 7-phosphate, D-Fructose 6-Phosphate, D-Glucose 6-phosphate, Solatriose and Raffinose. Taken together, our results showed that salt stress and alkali stress affected the plant phenotype, chloroplast structure, gene expression and metabolite abundance, and alkali stress resulted in more serious damage to grapevine plants than that of salt stress.
... We suspect that this density comes from a predicted C-terminal helix after IIS6, the only missing structured region in the model. In fact, the predicted C-terminal helix appears to be essential for AtTPC1 activity as its deletion mutation resulted in the loss of channel function (22). Following F2 helix, the loop (L1-2, Gln400-Pro405) links EF-hand domain and VSDII by forming interactions with F1-E2 linker, S0, and IIS2-IIS3 linker (Fig. 2D). ...
Article
Full-text available
Significance How membrane potential drives conformational change at the voltage-sensing domain (VSD) and regulates channel gating is a central theme for voltage-gated ion channels. To elucidate voltage-gating mechanisms, one needs to capture VSD structures in both the activated state and resting state, the latter of which is difficult to obtain because VSDs tend to adopt an activated state in the absence of membrane potential in most in vitro experiments. Here, we determined AtTPC1 structures in a closed conformation with a resting VSDII and an unbound EF-hand domain and in a partially open conformation with an activated VSDII and a Ca ²⁺ -activated EF-hand domain, elucidating the structural mechanisms of voltage gating, cytosolic Ca ²⁺ activation, and their coupling in AtTPC1.
... For the model plant Arabidopsis thaliana, it was shown that the SV channel belongs to the class of Two Pore Channels (TPCs) and is encoded by the TPC1 gene [2]. Cryo-EM and crystal structure analyses indicate that these channels are characterized by a tandem Shaker cassette structure that dimerizes via a C-terminal dimerization domain [3] to form a channel with 4 × 6 transmembrane helices [4,5]. In each Shaker cassette, the first four helices form a voltage-sensing domain (VSD), and helices 5 and 6 constitute the pore domain (P). ...
Article
Full-text available
Two Pore Channels (TPCs) are cation-selective voltage- and ligand-gated ion channels in membranes of intracellular organelles of eukaryotic cells. In plants, the TPC1 subtype forms the slowly activating vacuolar (SV) channel, the most dominant ion channel in the vacuolar membrane. Controversial reports about the permeability properties of plant SV channels fueled speculations about the physiological roles of this channel type. TPC1 is thought to have high Ca2+ permeability, a conclusion derived from relative permeability analyses using the Goldman–Hodgkin–Katz (GHK) equation. Here, we investigated in computational analyses the properties of the permeation pathway of TPC1 from Arabidopsis thaliana. Using the crystal structure of AtTPC1, protein modeling, molecular dynamics (MD) simulations, and free energy calculations, we identified a free energy minimum for Ca2+, but not for K+, at the luminal side next to the selectivity filter. Residues D269 and E637 coordinate in particular Ca2+ as demonstrated in in silico mutagenesis experiments. Such a Ca2+-specific coordination site in the pore explains contradicting data for the relative Ca2+/K+ permeability and strongly suggests that the Ca2+ permeability of SV channels is largely overestimated from relative permeability analyses. This conclusion was further supported by in silico electrophysiological studies showing a remarkable permeation of K+ but not Ca2+ through the open channel.
... Protonation of these histidine residues leads to structural rearrangements (Figures S4B, S4D, and S4E), resulting in dissociation of the SLAH3-SLAH3 inactive dimeric complex to its active monomeric form ( Figures 3C, 3D, and 5D). Thus, in contrast to other membrane proteins that activate upon dimerization, 52 SLAH3 gates open upon monomerization, a mechanism also found for the phosphorylation-controlled dimerization switch of NRT1.1. 53 The importance of the H +mediated stoichiometry change for SLAH3 activation is underlined by our finding that the H + -dependent increase in current amplitude is independent of the single-channel conductance ( Figures 2E and 2F) or rel. ...
Article
Full-text available
Plants, as sessile organisms, gained the ability to sense and respond to biotic and abiotic stressors to survive severe changes in their environments. The change in our climate comes with extreme dry periods but also episodes of flooding. The latter stress condition causes anaerobiosis-triggered cytosolic acidosis and impairs plant function. The molecular mechanism that enables plant cells to sense acidity and convey this signal via membrane depolarization was previously unknown. Here, we show that acidosis-induced anion efflux from Arabidopsis (Arabidopsis thaliana) roots is dependent on the S-type anion channel AtSLAH3. Heterologous expression of SLAH3 in Xenopus oocytes revealed that the anion channel is directly activated by a small, physiological drop in cytosolic pH. Acidosis-triggered activation of SLAH3 is mediated by protonation of histidine 330 and 454. Super-resolution microscopy analysis showed that the increase in cellular proton concentration switches SLAH3 from an electrically silent channel dimer into its active monomeric form. Our results show that, upon acidification, protons directly switch SLAH3 to its open configuration, bypassing kinase-dependent activation. Moreover, under flooding conditions, the stress response of Arabidopsis wild-type (WT) plants was significantly higher compared to SLAH3 loss-of-function mutants. Our genetic evidence of SLAH3 pH sensor function may guide the development of crop varieties with improved stress tolerance.
... TPC proteins present two re-entrant pore loops between TM5-TM6 and TM11-TM12, a large cytosolic linker connecting the two TM domains and intracellular amino-and carboxy-termini Hooper et al., 2011;Rietdorf et al., 2011). For this reason, TPCs are regarded as the molecular intermediate between tetrameric one-domain (six TM) channels, such as the α subunit of voltage-gated K + channels (KV), CatSper or TRP channels, and monomeric four-domain (four x 6TM) channels, such as voltage-gated Na + (NaV) and Ca 2+ (CaV) channels (Larisch et al., 2016;Patel, 2015;Rahman et al., 2014). High-resolution X-ray crystallography and cryo-electron microscopy have recently elucidated the molecular structures of TPC1 (Guo et al., 2016;Kintzer and Stroud, 2016) and TPC2 (She et al., 2019). ...
Chapter
An increase in intracellular Ca2 + concentration ([Ca2 +]i) regulates a plethora of functions in the cardiovascular (CV) system, including contraction in cardiomyocytes and vascular smooth muscle cells (VSMCs), and angiogenesis in vascular endothelial cells and endothelial colony forming cells. The sarco/endoplasmic reticulum (SR/ER) represents the largest endogenous Ca2 + store, which releases Ca2 + through ryanodine receptors (RyRs) and/or inositol-1,4,5-trisphosphate receptors (InsP3Rs) upon extracellular stimulation. The acidic vesicles of the endolysosomal (EL) compartment represent an additional endogenous Ca2 + store, which is targeted by several second messengers, including nicotinic acid adenine dinucleotide phosphate (NAADP) and phosphatidylinositol 3,5-bisphosphate [PI(3,5)P2], and may release intraluminal Ca2 + through multiple Ca2 + permeable channels, including two-pore channels 1 and 2 (TPC1–2) and Transient Receptor Potential Mucolipin 1 (TRPML1). Herein, we discuss the emerging, pathophysiological role of EL Ca2 + signaling in the CV system. We describe the role of cardiac TPCs in β-adrenoceptor stimulation, arrhythmia, hypertrophy, and ischemia-reperfusion injury. We then illustrate the role of EL Ca2 + signaling in VSMCs, where TPCs promote vasoconstriction and contribute to pulmonary artery hypertension and atherosclerosis, whereas TRPML1 sustains vasodilation and is also involved in atherosclerosis. Subsequently, we describe the mechanisms whereby endothelial TPCs promote vasodilation, contribute to neurovascular coupling in the brain and stimulate angiogenesis and vasculogenesis. Finally, we discuss about the possibility to target TPCs, which are likely to mediate CV cell infection by the Severe Acute Respiratory Disease-Coronavirus-2, with Food and Drug Administration-approved drugs to alleviate the detrimental effects of Coronavirus Disease-19 on the CV system.
Chapter
Human TPC channels are an emerging family of intracellular proteins fundamental for cell physiology and involved in various severe pathologies. Their localization in the membranes of endo-lysosomes, intracellular compartments of submicrometric dimensions, makes their study difficult with usual electrophysiological techniques. In this work, we show how the plant vacuole, a versatile organelle that can occupy up to 90% of the volume in mature plant cells, can be used as a heterologous system of expression for functional characterization. For this purpose, the use of vacuoles isolated from mesophyll cells of the Arabidopsis thaliana mutant lacking the endogenous TPC avoids unwanted interferences. The patch-clamp technique can be successfully applied to plant vacuoles in all different configuration modes; of note, the whole-vacuole configuration allows to study channel modulation by cytosolic factors. The combination of patch-clamp with fluorescence techniques, for example, by using fluorescent probes sensitive to specific ions of interest, represents a useful extension to investigate the selectivity properties of the channels. Therefore, the plant vacuole, similar to Xenopus oocytes for ion channels and transporters localized in the plasma membrane, has the capability to become a model system for functional studies on intracellular ion channels and transporters.
Article
Full-text available
Nose has modified Newtonian dynamics so as to reproduce both the canonical and the isothermal-isobaric probability densities in the phase space of an N-body system. He did this by scaling time (with s) and distance (with V¹D/ in D dimensions) through Lagrangian equations of motion. The dynamical equations describe the evolution of these two scaling variables and their two conjugate momenta p/sub s/ and p/sub v/. Here we develop a slightly different set of equations, free of time scaling. We find the dynamical steady-state probability density in an extended phase space with variables x, p/sub x/, V, epsilon-dot, and zeta, where the x are reduced distances and the two variables epsilon-dot and zeta act as thermodynamic friction coefficients. We find that these friction coefficients have Gaussian distributions. From the distributions the extent of small-system non-Newtonian behavior can be estimated. We illustrate the dynamical equations by considering their application to the simplest possible case, a one-dimensional classical harmonic oscillator.
Article
Full-text available
Vacuolar ion channels in guard cells play important roles during stomatal movement and are regulated by many factors including Ca²⁺, calmodulin, protein kinases, and phosphatases. We report that physiological cytosolic and luminal Mg²⁺ levels strongly regulate vacuolar ion channels in fava bean (Vicia faba) guard cells. Luminal Mg²⁺inhibited fast vacuolar (FV) currents with aK i of approximately 0.23 mm in a voltage-dependent manner at positive potentials on the cytoplasmic side. Cytosolic Mg²⁺ at 1 mm also inhibited FV currents. Furthermore, in the absence of cytosolic Mg²⁺, cytosolic Ca²⁺ at less than 10 μm did not activate slow vacuolar (SV) currents. However, when cytosolic Mg²⁺ was present, submicromolar concentrations of cytosolic Ca²⁺ activated SV currents with aK d of approximately 227 nm, suggesting a synergistic Mg²⁺-Ca²⁺ effect. The activation potential of SV currents was shifted toward physiological potentials in the presence of cytosolic Mg²⁺concentrations. The direction of SV currents could also be changed from outward to both outward and inward currents. Our data predict a model for SV channel regulation, including a cytosolic binding site for Ca²⁺ with an affinity in the submicromolar range and a cytosolic low-affinity Mg²⁺-Ca²⁺ binding site. SV channels are predicted to contain a third binding site on the vacuolar luminal side, which binds Ca²⁺ and is inhibitory. In conclusion, cytosolic Mg²⁺ sensitizes SV channels to physiological cytosolic Ca²⁺ elevations. Furthermore, we propose that cytosolic and vacuolar Mg²⁺ concentrations ensure that FV channels do not function as a continuous vacuolar K⁺ leak, which would prohibit stomatal opening.
Article
Molecular dynamics simulations are a powerful tool for complementing experimental studies, providing insights in biological processes at the molecular and atomistic level, at timescales from picoseconds to microseconds. Simulations are useful for testing hypotheses and can provide explanations for experimental observations as well as suggestions for further experiments. This does require that the simulation setup allows assessment of the question addressed. For example, it is evident that for simulation of a protein in its functional state the protein model and the environment have to mimic the biological situation as close as possible. In this chapter, a general strategy is presented for setting up and running simulations of membrane proteins of known structure in biological membranes of diverse composition and size.
Article
For simulations of membranes and membrane proteins, the generation of the lipid bilayer is a critical step in the setup of the system. Membranes comprising multiple components pose a particular challenge, because the relative abundances need to be controlled and the equilibration of the system may take several microseconds. Here we present a comprehensive method for building membrane-containing systems, characterized by simplicity and versatility. The program uses preset, coarse-grain lipid templates to build the membrane, and also allows on-the-fly generation of simple lipid types by specifying the headgroup, linker and lipid tails on the command line. The resulting models can be equilibrated, after which a relaxed atomistic model can be obtained by reverse transformation. For multi-component membranes, this provides an efficient means for generating equilibrated atomistic models. The method is called insane, an acronym for INSert membrANE. The program has been made available, together with the complementary method for reverse transformation, at http://cgmartini.nl/. This work highlights the key features of insane and presents a survey of properties for a large range of lipids as a start of a computational lipidomics project.
Article
Interactions between membrane proteins are of great biological significance, and are con- sequently an important target for pharmacological intervention. Unfortunately, it is still difficult to obtain detailed views on such interactions, both experimentally, where the environment hampers atomic resolution investigation, as well as computationally, where the time and length scales are problematic. Coarse grain simulations have alleviated the later issue, but the slow movement through the bilayer, coupled to the long life times of non-optimal dimers still stands in the way of characterizing binding distributions. In this work, we present DAFT, a Docking Assay For Transmembrane components, developed to identify preferred binding orientations. The method builds on a program developed recently for generating custom membranes, called insane (INSert membrANE). The key feature of DAFT is the setup of starting structures, for which optimal periodic boundary conditions are devised. The purpose of DAFT is to perform a large number of simulations with different components, starting from unbiased non-interacting initial states, such that the simulations evolve collectively, in a manner reflecting the underlying energy landscape of interaction. The implementation and characteristic features of DAFT are explained, and the efficacy and relaxation properties of the method are explored for oligomerization of glycophorin A dimers, polyleucine dimers and trimers, MS1 trimers and rhodopsin dimers. The results suggest that for simple helices, such as GpA and polyleucine, in POPC/DOPC membranes series of 500 simulations of 500 ns each allow characterization of the helix dimer orientations, and allow comparing associating and non-associating components. However, the results also demonstrate that short simulations may suffer significantly from non-convergence of the ensemble and that using too few simulations may obscure or distort features of the interaction distribution. For trimers, simulation times exceeding several microseconds appear needed, due to the increased complexity. Similarly, characterization of larger proteins, such as rhodopsin, takes longer time scales due to the slower diffusion and the increased complexity of binding interfaces. DAFT and its auxiliary programs have been made available from http://cgmartini.nl/, together with a working example.
Article
A molecular description of the lipid-protein interactions underlying the adsorption of proteins to membranes is crucial for understanding e.g. the specificity of adsorption, the binding strength of a protein to a bilayer, or for characterizing protein-induced changes of membrane properties. In this paper we extend an automated in silico assay (DAFT) for binding studies and apply it to characterize the adsorption of the model fusion peptides E and K to a mixed phospholipid/cholesterol membrane using coarse-grained molecular dynamics simulations. In addition, we couple the coarse-grained protocol to reverse transformation to atomistic resolution, thereby allowing to study molecular interactions with high detail. The experimentally observed differential binding of the peptides E and K to membranes, as well as the increased binding affinity of helical over unstructered peptides, could be well reproduced using the polarizable MARTINI coarse-grained (CG) force field. Binding to neutral membranes is shown to be dominated by initial binding of the positively charged N-terminus to the phospholipid headgroup region, followed by membrane surface-aligned insertion of the peptide at the interface between the hydrophobic core of the membrane and its polar headgroup region. Both coarse-grained and atomistic simulations confirm a before hypothesized snorkeling of lysine side chains for the membrane-bound state of the peptide K. Cholesterol was found to be enriched in peptide vicinity, which is probably of importance for the mechanism of membrane fusion. The applied sequential multiscale method - using coarse-grained simulations for the slow adsorption process of peptides to membranes followed by backward transformation to atomistic detail and subsequent atomistic simulations of the preformed peptide-lipid complexes - is shown to be a versatile approach to study the interactions of peptides or proteins with biomembranes.