Structural and Functional Characterization of A Calcium-Activated Cation Channel From Tsukamurella Paurometabola

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 11

ARTICLE

Received 13 Apr 2016 | Accepted 25 Jul 2016 | Published 28 Sep 2016 DOI: 10.1038/ncomms12753 OPEN

Structural and functional characterization


of a calcium-activated cation channel from
Tsukamurella paurometabola
Balasundaresan Dhakshnamoorthy1, Ahmed Rohaim1,2, Huan Rui1, Lydia Blachowicz1 & Benoı̂t Roux1

The selectivity filter is an essential functional element of K þ channels that is highly


conserved both in terms of its primary sequence and its three-dimensional structure. Here,
we investigate the properties of an ion channel from the Gram-positive bacterium Tsuka-
murella paurometabola with a selectivity filter formed by an uncommon proline-rich sequence.
Electrophysiological recordings show that it is a non-selective cation channel and that its
activity depends on Ca2 þ concentration. In the crystal structure, the selectivity filter adopts a
novel conformation with Ca2 þ ions bound within the filter near the pore helix where they are
coordinated by backbone oxygen atoms, a recurrent motif found in multiple proteins. The
binding of Ca2 þ ion in the selectivity filter controls the widening of the pore as shown in
crystal structures and in molecular dynamics simulations. The structural, functional and
computational data provide a characterization of this calcium-gated cationic channel.

1 Division
of Biological Sciences, Department of Biochemistry and Molecular Biology, The University of Chicago 929 East 57th Street Chicago, Illinois 60637,
USA. 2 Department of Biophysics, Faculty of Science, Cairo University, Giza, 12613, Egypt. Correspondence and requests for materials should be addressed to
B.R. (email: roux@uchicago.edu).

NATURE COMMUNICATIONS | 7:12753 | DOI: 10.1038/ncomms12753 | www.nature.com/naturecommunications 1


ARTICLE NATURE COMMUNICATIONS | DOI: 10.1038/ncomms12753

otassium (K þ ) channels are tetrameric membrane-

P
with an unusual proline-rich selectivity filter sequence,
spanning proteins that provide a selective pore for the LPMGNGPLSP, attracted our attention. Here, we present a
conductance of K þ across the cell membranes1. The X-ray characterization of this cationic channel, hereafter referred to as
structure of the bacterial channel KcsA from Streptomyces NaKTs. The crystallographic structure determined at 3.15 Å
lividans revealed that the permeation pathway comprises a displays Ca2 þ ions bound to the channel tetramer within the
narrow pore lined exclusively by main chain carbonyl oxygens2. selectivity filter. Functional assays and electrophysiological
Formed by the residues corresponding to the signature sequence recordings show that it is a non-selective cationic channel and
TTVGYG, common to all K þ channels3,4, this essential that Ca2 þ binding is necessary for channel activation and gating.
functional element serves as an ion-conducting selectivity filter, The key Ca2 þ binding motif is formed by an unusual helix-turn
allowing only the passage of nearly dehydrated K þ ions. While structural motif that has not been previously observed in other
the X-ray structures of K þ channels determined subsequently channels. MD simulations show that Ca2 þ ions are stable inside
greatly expanded our knowledge, these efforts have so far revealed the filter and their presence affects the widening of the pore. The
only a small number of possible conformations for the selectivity gating mechanism involving Ca2 þ binding in the selectivity filter
filter. The conformation that is predominantly observed is adds to the molecular repertoire of possible machinery that
the ‘ion conductive state’. First revealed by the X-ray structure nature uses to turn channels on and off.
of KcsA2, this conformation of the selectivity filter has
been subsequently observed in a number of K þ channel Results
pore domains, including KcsA2,5–8, Kv1.2 (refs 9,10),
Homology to the K þ channel family. The NaKTs channel
MthK11,12, KvAP13, KirBac14, and Kir3.1 (ref. 15). Two belongs to the family of channels with fourfold symmetric
additional conformations observed exclusively in crystal
tetrameric pore domains. The subunits display a simple 2-TM
structures of the KcsA channel are the ‘constricted’ and the topology, M1-pore-M2, in which a long pore loop is located
‘flipped’ conformations. The constricted or pinched conformation
between two TM helices. The selectivity filter of the NaKTs
is thought to correspond to a C-type inactivated state of channel is formed by the uncommon proline-rich sequence,
the channel. It has been observed at low [K þ ]5, and in the
LPMGNGPLSP. A search of the non-redundant database (NCBI
presence of tetraethylammonium, a classic quaternary amine and UNIPROT) did not discover other channels with a similar
channel blocker16, as well as in a series of structures in which
sequence, suggesting that this feature is fairly specific to this
the intracellular gate of the channel was open17. The flipped particular channel. Despite the difference, the NaKTs ion channel
conformation has been observed in crystal structures
is nonetheless clearly a member of the broad K þ channel family,
of the non-inactivating mutant E71A7,18. Thus, only three with a high homology to many important channels, as shown by
conformational variants of the selectivity filter have been
the multiple sequence alignment of the NaKTs channel and other
observed, so far, by X-ray crystallography. In contrast, K þ channel family members (Fig. 1a). The pore region of NaKTs
molecular dynamics (MD) simulations suggest that the
has 30% sequence similarity with that of NaK, a channel that can
selectivity filter of K þ channels undergoes considerable thermal be mutated into a KcsA-like conductive pore. A phylogenetic tree
fluctuations and can explore a broad range of conformations.
generated by considering only the sequence of the pore loop
Nevertheless, it is difficult to grasp the significance of these region shows the close relationship of NaKTs to the members of
computational results because only a limited fraction of the
the K þ channel family (Fig. 1b). On the basis of this analysis, it
landscape of conformations accessible to the selectivity filter of a
appears that the pore region of the NaKTs channel is most similar
K þ channel has been mapped out by X-ray crystallography. in terms of primary sequence to that of the CNG, hERG, K2P,
Additional information about the structure of related channels
MthK and Kir channels. This overall similarity suggests that the
with similar sequences can contribute to a better understanding conformational variants adopted by NaKTs are likely to be of
of the conformational landscape of the pore. Implicit in this view
relevance for the entire K þ channel family.
is the assumption that the structural diversity displayed by the
members of a family of homologous proteins is representative of
the structural dynamics displayed by these proteins19. In this Structure determined by X-ray crystallography. The crystal-
regard, the conformational variants displayed by the non-specific lographic structure of the NaKTs channel determined at 3.15 Å
NaK channel, a distant homologue of K þ channels, has helped resolution is shown in Fig. 2. The structural model is complete
expand our knowledge of these systems20,21. Its selectivity filter over all regions, except for residues 1–3 at the N-terminal and
exhibits extensive structural variability20–24, converting in residues 107–123 at the C-terminal regions (the latter might have
response to specific mutations from a fairly wide hydrated been cleaved by chymotrypsin digestion). In its broad features,
pore20 to a KcsA-like conductive conformation22. An important the structure of the NaKTs channel is similar to that of other
concept emerging from these studies is that the selectivity filter tetrameric pore domains, with a TM helical segment-labelled M1
of tetrameric cationic channels is a complex functional element (residues 21–45), a pore helix (residues 49–62), the selectivity
able to display considerable structural polymorphism and filter (residues Pro63-Met64-Gly65-Asn66-Gly67-Pro68) and a
conformational plasticity. While there is an increasing amount M2 helix (residues 73–101). In addition, an amphipathic helical
of structural data about the broad family of tetrameric segment-labelled M0 (residues 5–19) is located at the N-terminal
2-transmembrane (TM) pore domains, the extent of the end of the subunit. The channel is in a conformation with an
complete conformational landscape of the selectivity filter open intracellular gate, with a strong bend of B50° near the
remains unclear. centre of the M2 helix (Fig. 2b). While the open conformation of
Pursuing this strategy, we sought to discover and characterize NaKTs is similar to that observed in previous X-ray structures of
additional related channels by searching the genomic databases. other channels11, the strongest bend in the M2 helix occurs at
We specifically focused on 2-TM bacterial channels forming a the level of residue Val87 rather than at the level of a glycine
tetrameric pore domain with a sequence homology below 30% residue proposed to act as a conserved gating hinge in several
with channels of known structures and with selectivity filters that channels (Supplementary Fig. 1). The narrowest point along the
departed markedly from the highly conserved canonical signature intracellular vestibule occurs at the level of residue Val92, its side
sequence TTVGYGDLMP3,4. One particular channel from the chain pointing towards the central channel axis. The diameter is
bacterium Tsukamurella paurometabola (Uniprot ID: D5UM26) on the order of 10 Å, which is slightly smaller than that observed

2 NATURE COMMUNICATIONS | 7:12753 | DOI: 10.1038/ncomms12753 | www.nature.com/naturecommunications


NATURE COMMUNICATIONS | DOI: 10.1038/ncomms12753 ARTICLE

a a c
Tsukamurella_ionchannel
KcsA
MthK
Kv1.2_Shaker_Rat
KvAP_Aeropyrum
K2P1_Twik1_Human
KirBac1.1_B.Pseudomallei
Kir2.1_Mouse
Kesv_ESV1-Virus Gly65(D) SCa Ser59(D)
Kcv__PBCV1-Virus
CNG_Rat
ERG_Human
NaK_B.Cereus Pro63(B)
Leu62(D)
b
KirBac3.1 Magnetospirillum
l
nne

b d
seudomallei
cha

Kir3.1 Mouse
ion

use
LysinD.Hafniense Na

Kir .3 Rat

use
lla

Mo
n

an
KirBac1.1 B.P
Tw mure
ma

Mo

m
3.2
Kir3

ick
Hu
Hu

Ki ir2.1

Ch
K
ka

.3
ik1

i Na

.2
r2

r2
Tsu

Ki
S2
MthK

s
iru

s
n

iru
ma

-V
P

-V
K2

V1

V1
Hu

ES

C
PB
G

SCa
t
Ra

sv

aK
ER

V SCa
Ke

Kc lus N
G

il
CN

ac ter
eob re us bac
G e aero
B.C
NaK Sph
6TM
yrum
Aerop
KvAP
Kv1.2 Shaker Rat
D.Baculatum
Nostoc
Kv P
Kcs orphyro
A mon
M2
as

M1 M0
NavRh/NaChBac
10
Figure 2 | X-ray crystal structure of the NaKTs ion channel determined at
Figure 1 | The NaKTs ion channel in relation with other tetrameric cation 3.15 Å resolution. (a) Ribbon representation of top view of the NaKTs
channels. (a) Multiple sequence alignment of the pore helix and selectivity tetramer (chain ABDE—dataset 2), along with bound Ca2 þ ions. The M0
filter regions for the NaKTs ion channel protein along with K þ , Na þ and helix (residues 5–19) is shown in blue, the M1 outer helix (residues 25–45)
non-selective channels. The multiple sequence alignment was done using is shown in magenta, the pore helix (residues 46–61) is shown in green, the
Clustalw and JalView (the colour coding is based on the Taylor scheme selectivity filter (residues 62–69) is shown in yellow, the M2 inner helix is
available in JalView). (b) Phylogenetic tree (scored from pore helix and shown in cyan and the loop regions are shown in green. Calcium ions are
filter region). The phylogenetic tree was generated based on the multiple shown as purple spheres, with the 2Fo  Fc map (grey) at 1s contour and
sequence alignment of the pore helix and the selectivity filter region only Fo  Fc map (green) at 4.5s contour. (b) The TM region of two opposing
using the programs JalView and MEGA 5. monomers of the NaKTs ion channel protein (chain A, chain D) shown with
the 2Fo  Fc density map at 1s contour for the selectivity filter region
(yellow sticks). (c) Fo  Fc map (green) at 4.5s contour and 2Fo  Fc map
in other channels in an open conformation, for example, NaK (grey) at 1s contour showing the location of Ca2 þ ion (purple sphere) at
(10.3 Å at Phe92) and MthK (12.9 Å at Ala88)21. The M2 helix site SCa. The pore helix (green cartoon), the selectivity filter residues
also displays a twist of B12° with respect to the open (yellow sticks) and Ca2 þ coordinating AA’s are labelled accordingly.
conformation of the NaK channel (3E86), but of only (d) Close-up view of two opposing monomers of the NaKTs selectivity filter
B3° with respect to the open conformation of the MthK (chain A, chain D) shown in yellow sticks with bound Ca2 þ ions at sites S2
channel (4HYO) (Supplementary Fig. 2). These structural and SCa.
differences regarding the intracellular activation gate suggest
that some aspect of the gating mechanism of NaKTs may differ
polypeptide chains, each containing a proline-rich sequence
from other channels. Furthermore, an amphipathic helical M0
LPMGNGPLSP that is not commonly observed in other
segment (residues 5–17) docks against the hydrophobic surface of
homologue channels (Figs 1 and 2d). The overall topology of
the pore domain. In this configuration, the N terminus of the
the pore loop is similar to that of other tetrameric channels
protein reaches across the membrane region to interact with
(Supplementary Fig. 4) although there are noticeable structural
Phe41 from a neighbouring subunit near the extracellular side.
differences with other members of the tetrameric channel family
It seems unlikely that M0 affected the overall geometry of the
(Supplementary Table 1). A superimposition of the pore region
pore, as it is very similar to that of other tetrameric channels.
(that is, the pore helix and the filter) of the NaKTs channel with
While this orientation is stabilized by a number of interactions
that of the KcsA channel in the conductive state (PDB id 1K4C)
(Supplementary Fig. 3), it is probably not representative of a
yields a root-mean-square deviation (RMSD) of 2.43 Å for the
membrane-bound conformation of the channel. In MD
backbone atoms. The RMSD with other K þ -selective channels
simulations including the M0 segment, water molecules rushed
(MthK, Kir2.2, Kir3.2, Kv1.2) is also on the order of 3 Å.
into the membrane interior along the pore domain to solvate
For comparison, the RMSD among these various channels is
charged residues (Lys9 and Arg10). Most likely, the amphipathic
typically much smaller, on the order of 0.5–1.0 Å. This shows
M0 segment should be parallel to the membrane interface25.
that the conformation of the pore region for the NaKTs
channel departs from that of other members of the tetrameric
Cation conducting pore. The ion-conducting pore of NaKTs channel family. Despite these differences, NaKTs is a non-
is aligned with the fourfold symmetry axis of the tetramer, selective cationic channel according to electrophysiological
consistent with the structure of other channels in the family2. recordings of the channel reconstituted in lipid membranes.
However, the selectivity filter is formed by four extended The chymotrypsin-treated form of the channel that was

NATURE COMMUNICATIONS | 7:12753 | DOI: 10.1038/ncomms12753 | www.nature.com/naturecommunications 3


ARTICLE NATURE COMMUNICATIONS | DOI: 10.1038/ncomms12753

crystallized conducts both K þ and Na þ , with a measured molecules and does not interact directly with backbone oxygen
conductance of approximately 88 and 110 pS at 200 mM of KCl atoms. The distance between the carbonyl oxygen of Gly67 and
and NaCl, respectively (Fig. 3a–d). the Ca2 þ ion at site 1 is about 4.9 Å, suggesting that the ion binds
in this site in a hydrated form through bound water molecules.
In state III, two Ca2 þ ions are bound diagonally opposite to
Presence of Ca2 þ in the selectivity filter. A salient difference each other 4.1 Å away from the pore axis at a site in the selectivity
between the NaKTs channel and other members of the tetrameric filter referred to as SCa (Fig. 2c), where they are coordinated by
channel family is the presence of Ca2 þ ions, which are observed the backbone carbonyl oxygen atoms of Ser59 (distance of 2.6 Å),
in different regions of the selectivity filter. Three different Leu62 (distance of 2.5 Å) and the backbone nitrogen atom of
diffraction datasets were analyzed to identify different ion Gly65 (distance of 2.9 Å) from one subunit, and the backbone
occupancy states (by virtue of the I4 space group symmetry, the oxygen atom of Pro63 from a neighboring subunit (distance
six monomers yield three crystallographically related tetramers). of 2.9 Å). A third Ca2 þ is located at the centre of the pore in a
A total of nine crystallographically distinct tetramers were site corresponding to S2 in KcsA. The ion is coordinated by the
considered in the analysis, revealing four different occupancy side chain of Asn66, which points towards the pore of the
states, I to IV, with 0, 2, 3 and 6 Ca2 þ ions, respectively (Fig. 4). selectivity filter. The OD1 atom of Asn66 interacts with the Ca2 þ
There are no ions in state I, which serves as the reference for ion and the ND2 atom of Asn66 interacts with the backbone
the unoccupied pore. In state II, there are two Ca2 þ ions near the carbonyl oxygen atom of Leu69. In state IV, there are four Ca2 þ
extracellular side in locations corresponding to sites S0 and S1 in ions in SCa and two additional Ca2 þ ions, one in the site
KcsA. The Ca2 þ ion in site S0 is surrounded by four water corresponding to S3 in KcsA near Gly65 and a second ion near

a b
6 I (pA)
4
+40 mV
2
+20 mV 0
–2 V (mV)
–20 mV
–4
–40 mV –5
50 –30 –10 10 30 50
c d
5 I (pA)
+40 mV 3
+20 mV 1
–1 V (mV)
–20 mV
–3
–40 mV –5
–50 –30 –10 10 30 50
e f
0 mM
0.8 PO
1 mM 0.7
2 mM 0.6
0.5
3 mM 0.4
4 mM 0.3
0.2
5 mM 0.1
0
7.5 mM 0 5 10 15
CaCl2 (mM)
g
10 pA

+20 mV

–20 mV 10 ms

Figure 3 | Results from single-channel recordings. (a) Representative current traces of NaKTs ion channel Na þ conductance at different TM potentials
(internal solution: 20 mM HEPES, pH 7.5; 0.2 M NaCl and external solution: 20 mM MOPS, pH 4; 0.2 M NaCl; 5 mM CaCl2). (b) I/V relationship of NaKTs
ion channel in the presence of NaCl. The NaKTs ion channel Na þ conductance is 110 pS. (c) Representative current traces of NaKTs ion channel K þ
conductance at different TM potentials (internal solution: 20 mM HEPES, pH 7.5; 0.2 M KCl and external solution: 20 mM MOPS, pH 4; 0.2 M KCl; 5 mM
CaCl2). (d) I/V relationship of NaKTs ion channel in the presence of KCl. The NaKTs ion channel K þ conductance is 88 pS. (e) Representative current
traces of the NaKTs ion channel Na þ conductance as a function of Ca2 þ in the external solution at a fixed membrane potential of  40 mV.
(f) Open-channel probability of NaKTs ion channel, Po, as a function of Ca2 þ in the external solution at a holding membrane potential of þ 40 mV (internal
solution: 20 mM HEPES, pH 7.5; 0.2 M NaCl and external solution: 20 mM MOPS, pH 4; 0.2 M NaCl with CaCl2 ranging from 250 mM to 20 mM). The error
bars indicate the mean-square deviation from the average (fraction of open time over a total time interval) based on 3–5 separate time intervals.
(g) Representative current traces of the chimeric NaKTs-SFKcsA channel at 200 mM KCl with 0 mM CaCl2. The single-channel conductance of the
chimera is on the order of 300 pS.

4 NATURE COMMUNICATIONS | 7:12753 | DOI: 10.1038/ncomms12753 | www.nature.com/naturecommunications


NATURE COMMUNICATIONS | DOI: 10.1038/ncomms12753 ARTICLE

S0
20
a

S1 16

Douter (Å)
12

State I State II 8

b
S2 10
S2
SCa SCa SCa

Dinner (Å)
SCa
S3 8

State III State IV

Figure 4 | Occupancy states of the selectivity filter of NaKTs. Snapshots 6


of the selectivity filter of the NaKTs ion channel structure (yellow sticks),
describing the location of Ca2 þ ions along the ion-conduction pathway,
categorized into four different states. Selectivity filter of state I is seen in 0 1 2 4
three tetramers (chain CGHI of datasets 1 and 3 and chain FJKL of dataset NCa
1), selectivity filter of state II is seen in one tetramer (chain FJKL of dataset
3), selectivity filter of state III is seen in 2 tetramers (chain ABDE of Figure 5 | Outer and inner diameters of the selectivity filter pore in the
datasets 1 and 2), selectivity filter of state IV is seen in two tetramers (chain NaKTs channel. (a) The outer diameter Douter taken as the Ca–Ca distance
ABDE of dataset 3 and chain FJKL of dataset 2). Calcium ions are shown as between the diagonally placed Pro68 is shown. (b) The inner diameter
purple spheres, along with their respective anomalous density maps Dinner taken as the Ca–Ca distance between the diagonally placed Pro63 is
(orange) at 5–10s contour and Fo  Fc map (green) at 4s contour. shown. Supplementary Fig. 4 shows the relative positions of the proline
residues. The diameters from the NaKTs channel crystal structures are
Pro63 (a different sub-state is seen with the ion position shifted presented by black symbols. They are grouped by the number of Ca2 þ
by 2.1 and 1.1 Å towards the intracellular surface). Lastly, there is at the selectivity filter SCa site. States I ( and n) and II (&) have no
a tetramer with a single Ca2 þ ion located at the centre of the Ca2 þ at the SCa site. Two structures are used for State I. The difference
intracellular vestibule with no ions observed in the pore, as between the two lies in that the second structure (n) has a Ca2 þ in the
observed in the KcsA and NaK structures, assigned to State I. centre of the intracellular vestibule. State III ( and ~) has two Ca2 þ at
More details about the locations of Ca2 þ ions in the 3.4 Å dataset the diagonally place SCa sites; State IV ( has all four SCa sites occupied.
3 (3.4 Å resolution) and 3.6 Å dataset 1 (3.6 Å resolution) crystal The averaged Douter and Dinner in the MD simulations (  500 mV) are
structures are provided in Supplementary Fig. 5. The Ca2 þ ion plotted with red symbols. The error bars indicate the standard deviations of
occupancy along the selectivity filter pore and SCa site were the distances calculated from the trajectories. The distances are calculated
refined using SAD anomalous data and their respective ion from two pairs of diagonally placed subunits. In all cases, filled symbols
occupancy is given in Supplementary Table 2. indicate that the distance is taken from subunits with the SCa sites occupied
by Ca2 þ . The Douter and Dinner in the NaK channel (2AHY, solid line) and
KcsA (1K4C, dotted line) are shown as references.
Structural and functional response to the binding of Ca2 þ .
The crystallographic and MD data indicate that the diameter of
the pore at its inner and outer ends, defined respectively from the the open-channel probability (Fig. 3e,f). The observed mean
backbone Ca–Ca distance of residues Pro63 and Pro68 from open-channel probability appears to be slightly sensitive to the
diagonally-opposing subunits, is sensitive to occupancy by Ca2 þ membrane potential (around 0.25 at negative voltages and
ions in the pore (Fig. 5). The diameter of the inner pore increases 0.6 at positive voltages with 2.5 mM [Ca2 þ ]). The average
by 1–1.2 Å in response to the binding of a Ca2 þ ion at site SCa, in single-channel conductance is not significantly affected by the
the vicinity of Pro63, Gly65, Ser59 and Leu62. Once a Ca2 þ ion presence of Ca2 þ .
binds to this position, Pro63 sways by about 0.5 Å away from the To probe the functional significance of the SCa binding site on
ion-conduction pathway. If one assumes that widening the channel activation, a chimera of the NaKTs channel with its
selectivity filter transforms the channel from a non-conductive selectivity filter (PMGNGP) replaced by the corresponding
(or poorly conductive) state to a conductive state, Pro63 thus residues from the KcsA channel (TVGYGD) was engineered
appears as a key residue responsible for channel activation upon and purified for functional assays. Whereas no channel activity
the binding of Ca2 þ at the SCa sites. Of relevance to these can be observed with the wild-type NaKTs unless Ca2 þ is present
observations, functional measurements of the channel recon- (Fig. 3e), it is observed that this chimeric NaKTs-SFKcsA channel
stituted in lipid bilayers show that Ca2 þ is absolutely required for conducts K þ in the absence of any Ca2 þ (Fig. 3g). This
ion-conduction activity, with a K1/2 of B4 mM determined from observation strengthens the conclusion that the binding of Ca2 þ

NATURE COMMUNICATIONS | 7:12753 | DOI: 10.1038/ncomms12753 | www.nature.com/naturecommunications 5


ARTICLE NATURE COMMUNICATIONS | DOI: 10.1038/ncomms12753

at the site SCa is directly implicated in the activation of the NaKTs B4 mM determined for the NaKTs channel on the basis of the
channel. open-channel probability (Fig. 3f).
Comparison with other channels shows some similarities but
also important differences. A Ca2 þ -dependent gating located
MD simulations. To gain further insight about the structural
within the selectivity filter was also shown to control the K þ
conduction in the MthK channel26. However, while the sequence sensitivity of the selectivity filter to the Ca2 þ ions in the SCa
binding sites, MD simulations of the NaKTs channel tetramer
similarity of NaKTs and MthK is high (Fig. 1a), the latter does
not possess the proline (Pro63) that forms part of the SCa binding (ABDE) embedded in fully hydrated POPC bilayer membranes
were carried out. Four different systems were simulated (0, 1, 2
site observed in NaKTs and the filter gating observed in the MthK
channel is believed to be allosterically triggered by the binding of and 4), comprising respectively zero, one, two and four Ca2 þ
ions bound at site SCa in the selectivity filter using the polarizable
Ca2 þ to the cytoplasmic RCK domains at the intracellular side26.
A recent study based on chimeric channels that replaced the Drude force field31–35 with recently refined parameters for
Ca2 þ ions36. Equivalent simulations carried out using the
selectivity filter of the bacterial non-selective cation channel
NaK, TVGDGNFS, with a cyclic-nucleotide gated channel filter nonpolarizable additive CHARMM36 (C36) force field37,38 were
unstable (See METHODS for details). The system with two Ca2 þ
sequence, TIGETPPP also displayed Ca2 þ block (B14 mM)27.
However, there are important differences because the key prolines at the SCa site is shown in Fig. 7a,b. The simulations with double
Ca2 þ occupancy display relatively small deviation from the
in those chimeras were located toward the extracellular end of the
filter, whereas the SCa binding site in NaKTs is located at its crystal structure when compared with the other systems
(Supplementary Fig. 6b) and also retain a water-filled selectivity
intracellular end.
filter (Fig. 7a,d). In the simulations, the number of Ca2 þ ions
bound at the SCa sites in the selectivity filter appears to control the
Structural similarity of the SCa site with other systems. The opening of the pore. This is observed from the distance
unusual helix-turn Ca2 þ-binding motif forming the SCa site is a (Dinner and Douter) between the Ca atoms (Fig. 5) and the
structural feature that has not been previously observed in radius profile of the pore restriction zone (Rmin), which is near
other crystallographic structures of tetrameric ion channels. residue Pro63 at the intracellular end of the filter (Supplementary
Interestingly, a large number of similar helix-turn Ca2 þ -binding Fig. 6c). In the simulations, the diameter measured across the
motifs could be identified from a search via the Metal3 pore between the diagonally placed Pro63 (Dinner in Fig. 5)
webserver28 (Fig. 6). Interestingly, Ca2 þ appears to be increases as the SCa sites become occupied by Ca2 þ . This trend
functionally important for several systems where the helix-turn continues until two of the SCa sites are filled with Ca2 þ . Further
motif is found, for example, it triggers the assembly process of increasing the number of Ca2 þ at the SCa sites does not lead to a
SbsB, a layer-forming component of the cell-wall in archaea and more open pore. Instead, it causes structural distortions of the
many bacteria29, it is essential for the enzyme activation of filter and closure of the intracellular gate formed by M2
human transglutaminase 3, and it is required for the toxin activity (Supplementary Fig. 6a). The Rmin profiles in Supplementary
of Clostridium botulinum neurotoxins30. It should be noted, Fig. 6c corroborate the notion that Ca2 þ binding at the SCa sites
however, that a large fraction of the structures with a helix-turn promotes the opening of the selectivity filter. Increasing the
binding motif include coordination by negatively charged side number of Ca2 þ ions in the selectivity filter from zero to two
chains, which are absent from the SCa site of the NaKTs channel. leads to more and longer incidents of transient pore openings
As a result, the ion binding affinity to SCa is likely to be weak. (Supplementary Fig. 6c). The correlation offers one explanation
Weak ion binding affinity is generally thought to be consistent from a structural point of view for the relationship between
with the function of a channel, as strong binding would oppose [Ca2 þ ] and the channel open probability Po determined from
rapid conduction. While the role of Ca2 þ in NaKTs is not as a single-channel recordings (Fig. 3f). However, no ion-conduction
permeant ion but rather to activate the channel, low binding events were observed during the MD simulations with an
affinity for the site SCa is nonetheless reasonable, as it is accessible applied voltage. Further analysis indicates that the unfavourable
via the selectivity filter. In the case of the layer-forming protein electrostatic repulsion resulting from Ca2 þ ion bound in the
SbsB (PDB id 4AQ1), the measured K1/2 falls within a range of sites Sca likely blocks the passage of other cations through the
0.9–110 mM, a value that is considerably smaller than the K1/2 of selectivity filter.

a b Discussion
Knowledge of the landscape of conformations accessible to the
Pore helix
selectivity filter of tetrameric K þ channels remains limited. To
make further progress, we present here a first characterization of
a novel tetrameric ion channel, referred to as NaKTs, with an
Sca
unusual proline-rich selectivity filter sequence using X-ray
crystallography, electrophysiology and MD simulations. Sequence
comparison suggests that, despite the differences, this channel is
closely related to the members of the broad family of tetrameric
2-TM channels. The NaKTs channel was crystallized with its
Figure 6 | The helix-turn SCa binding motif (a) Superimposition of the intracellular gate in an open state. The X-ray structures revealed
Ca2 þ binding motif of 47 PDB structures (grey cartoon) out of 369 hits the presence of Ca2 þ binding sites, SCa, formed by main chain
obtained by analysis using the Metal3 database43, and the pore helix atoms of residues Pro63, Gly65, Ser59 and Leu62, near the
(green), selectivity filter (yellow) and the calcium ion at site SCa of intracellular end of the selectivity filter. A database search reveals
the NaKTs structure. (b) Superimposition of the NaKTs structure (pore that similar helix-turn Ca2 þ -binding motifs are found in
helix—green; main chain atoms of selectivity filter—yellow sticks; Ca2 þ at multiple proteins, with a broad variation in the type of
site SCa—purple sphere) with representative motifs of PDB id 3K28 and coordinating ligands involved. Single-channel recordings demon-
2IZ7 (grey cartoon) with o1.0 Å RMSD and their respective Ca2 þ ions are strate that the NaKTs channel is a Ca2 þ -activated non-selective
shown in green spheres. cation channel. Analysis of crystallographic data indicates that the

6 NATURE COMMUNICATIONS | 7:12753 | DOI: 10.1038/ncomms12753 | www.nature.com/naturecommunications


NATURE COMMUNICATIONS | DOI: 10.1038/ncomms12753 ARTICLE

a b Paradoxically, even though the Ca2 þ bound in the site SCa


kept the selectivity filter in the crystallographic open state,
no ion-conduction events were observed during the MD
simulations with an applied voltage due to the unfavourable
electrostatic repulsion from the bound Ca2 þ . It is also possible
that the binding of Ca2 þ could induce additional conformational
changes not yet observed, or that the X-ray structure
obtained from a chymotrypsin-treated truncated form of the
protein is not fully representative of the channel probed by the
electrophysiological experiments. For example, the intracellular
gate of KcsA truncated is more wide-open than KcsA full-
length17,39. The situation may be different in the case of the
c d NaKTs channel, as the conductance of the truncated channel
(Fig. 3a) is smaller than that of the full-length channel
(Supplementary Fig. 7a). To be as consistent as possible with
the crystal structure, the functional experiments were carried out
with the truncated chymotrypsin-digested form of the NaKTs
channel. However, additional experiments with the full-length
channel NaKTs channel indicate that it exhibits a complex
functional behaviour, with two major conductance levels.
Furthermore, the full-length channel undergoes some
degradation over time at room temperature, which may partly
explain its complex functional behaviour. The smaller
e f conductance level is observed more frequently with the
12 12
truncated form of the channel while the larger level occurred
10 10 more frequently with the full-length form of the channel
(Supplementary Fig. 7a). Nonetheless, while the truncated form
Dinner

8 8 of the channel displays a more reliable functional behaviour than


the full-length form, it remains somewhat complex
6 6
(Supplementary Fig. 7b). It is also noticeable that the kinetic
4 4 behaviour of the full-length channel is slower than that of the
0 5 10 15 20 25 30 0 5 10 15 20 25 30 truncated channel. These observations suggest that there might be
Time (ns) Time (ns) two open-channel conformations, and their relative probability is
Figure 7 | MD simulations. (a) The top view and (b) the side view of the
different in the chymotrypsin-digested and full-length form of the
two Ca2 þ bound membrane-channel system before simulations using the
channel.
Drude polarizable force field. After 30 ns simulation, the pore closes in the
While Ca2 þ binding at the site SCa is key for channel
system (c) without any Ca2 þ at the SCa site, whereas the selectivity filter
activation, a complex interplay between gating and conduction
pore remains open and permeable to water in the system (d) with two
might take place because the permeating cations and Ca2 þ must
Ca2 þ bound at the SCa site. The NaKTs channel is shown in yellow cartoon
translocate through the same passageway. This suggests various
presentation with the front and back subunits hidden in (b–d) for the sake
mechanistic scenarios whereby the occurrence of ‘open’
of clarity. POPC lipids are shown as grey lines with their phosphates in
conformations of the pore could be simultaneously promoted
orange spheres. Ca2 þ ions are indicated by purple spheres. The other ions
by the reversible binding of Ca2 þ as well as blocked for ion
including K þ (light pink), Na þ (yellow) and Cl  (green) are shown as
conduction. In other words, the channel could display sustained
well. Surface rendering (marine) is used to show the water accessible
periods of conduction if the lifetime of the open state persists for
regions in the system. The molecular presentation is created using PyMOL.
sufficiently long time after the dissociation of the activating Ca2 þ
(e,f) show the time series of Dinner in the zero and two SCa Ca2 þ systems
ion(s). The intriguing mechanism by which Ca2 þ -activation and
with  500 mV TM potential, respectively. The lighter-coloured points are
ion conduction are interrelated in the NaKTs channel will require
instantaneous values of Dinner and the lines show the running averages. The
further investigation.
two Dinner in one tetramer channel are distinguished by different colours.
In closing, while the primary objective of the present study
was to characterize the structure and function of a novel
tetrameric 2-TM cationic channel in relation to other K þ
presence of bound Ca2 þ at diagonally-opposing sites inside the channels, the striking structural differences presented by the
selectivity filter controls the diameter of the conducting pore. pore of the NaKTs channel (its uncommon proline-rich sequence
Pro63, which contributes to the SCa binding site, appears to be a LPMGNGPLSP is not found in human channels) could
key residue responsible for the channel activation triggered by the potentially offer a useful target for the discovery of specific
binding of Ca2 þ . The direct functional implication of the SCa inhibitors that may be of therapeutic interest. Tsukamurella
binding site on channel activation is confirmed by engineering a species are Gram-positive bacilli that have been isolated from
chimera protein in which the classic selectivity filter of the KcsA environmental samples and sludge, that can lead to human
channel has been inserted in the NaKTs channel. While no diseases40. Catheter related bloodstream infections (CR-BSI),
channel activity is observed with the wild-type NaKTs unless the most common presentation of Tsukamurella species
Ca2 þ is present (Fig. 3e), the chimeric NaKTs-SFKcsA channel infections, primarily occur in patients with indwelling central
conducts K þ in the absence of any Ca2 þ (Fig. 3g). The lines and/or immune suppression41. Additional clinical
importance of Pro63 is supported by the results from MD malfunctions include bacteremia42,43, meningitis44, peritonitis45,
simulations with a polarizable force field. The results from MD and cutaneous, lung and knee prosthesis infection43,46.
strengthen the view that the Ca2 þ binding motif located at the Tsukamurella species is also recognized as a potential pathogen
base of the selectivity filter may act as an activating switch. in immunocompromised patients47.

NATURE COMMUNICATIONS | 7:12753 | DOI: 10.1038/ncomms12753 | www.nature.com/naturecommunications 7


ARTICLE NATURE COMMUNICATIONS | DOI: 10.1038/ncomms12753

Methods outliers in disallowed regions of the Ramachandran plot. Two additional crystal
Expression and purification. Genomic DNA from T. paurometabola (DSM 20,162) structures, at 3.4 Å (dataset 3) and 3.6 Å (dataset 1) resolution, were determined by
was obtained from the DSMZ (German Collection of Microorganisms and Cell molecular replacement, using the NaKTs 3.15 Å crystal structure as an initial
Cultures, Braunschweig, Germany). The genomic databases search leading to the search model followed by repeated cycles of model building in COOT and
identification of a 2-TM tetrameric channel from T. paurometabola used PsiBlast48, refinement with REFMAC 5.5 in the CCP4 suite of programs. For dataset
followed by manual analysis to focus on the sequence in the selectivity filter region. 3 (PDB ID: 5CBH), the N-terminal residues 1–4 and C-terminal residues 107–123
The Tpau_1687 gene encoding residues 1–123 of an ion transport two domain were disordered in the crystal structure. The final model contains six monomers
protein was amplified by PCR and cloned into the pQE60 expression vector (A, B, C, D, E and F) of residues 5–106 and 8 Ca2 þ ions refined to Rcryst and Rfree
(Qiagen). The primers used to clone the full-length NaKTs channel into pQE60 of 23.2 and 26.2%, respectively, with no outliers in disallowed regions of the
used the NcoI and BamHI sites of pQE60. The forward primer is 50 -CGTATAAC Ramachandran plot. Tetramer 1 (Chain ABDE) contains four Ca2 þ ions located at
ATGTTGGGTCTCACATTGATGTTCAAGAGATTC-30 and the reverse primer is: site SCa and two Ca2 þ ions at sites 2 and 3. Tetramer 2 (Chain CGHI) has no ions
50 -CTTGTAGGATCCGTCCTCCGCCTCGGCCGAAC-30 (restriction sites in the pore. Tetramer 3 (Chain FJKL) contains two Ca2 þ ions located at sites S0
highlighted in bold and the NaKTs coding sequence is underlined). The NaKTs ion and 1. For dataset 1 (PDB ID: 5CBH), the N-terminal residues 1–4 and C-terminal
channel protein was expressed in E. coli strain XL-1 Blue cells on induction with residues 107–123 were disordered in the crystal structure. The final model contains
0.5 mM isopropyl-b-D-thiogalactopyranoside (IPTG) at 37 °C for 3 h at an OD600 of six monomers (A, B, C, D, E and F) of residues 5–106 and three Ca2 þ ions and
0.6–0.8. Cell pellets were resuspended in buffer A (20 mM Tris, pH 8; 0.2 M NaCl) refined to Rcryst and Rfree of 23.87 and 27.9%, respectively, with no outliers in
plus 1 mM PMSF and DNase I. Cells were lysed by homogenization. The cell lysate disallowed regions of the Ramachandran plot. Tetramer 1 (Chain ABDE) contains
was clarified by centrifugation at 107,000g for 35 min at 4 °C. and the membrane two Ca2 þ ions bound diagonally opposite to each other at site SCa and 1 Ca2 þ ion
pellet was solubilized by incubation at 4 °C in Buffer B (Buffer A plus 10 mM bound at site 2. Tetramer 2 (Chain CGHI) and Tetramer 3 (Chain FJKL) have no
DDM (n-Dodecyl-b-D-maltoside) or 40 mM DM (n-Decyl-b-D-maltoside). The ions in the pore.
supernatant was then purified using a cobalt metal affinity column pre-washed with
buffer A and 0.5 mM b-DDM. The affinity-bound protein was eluted with Buffer A Determination of calcium ions. Calcium ions (Ca2 þ ) were identified based on
containing 500 mM imidazole and 0.5 mM b-DDM. The 6  His C-terminal tag the Fo  Fc map at 4.5s cutoff. A cluster of Ca2 þ ions was observed at the centre of
along with the unstable cytoplasmic domain was removed from the eluted protein by the hexamer of tetramer assembly (overall molecular packing). Calcium positions
incubation with chymotrypsin for 2 h at room temperature. The cleaved protein was in all of the three crystal structures were initially identified based on the Fo  Fc
further purified by gel filtration (buffer: 20 mM HEPES pH 8, 200 mM NaCl and difference map peak at 4.5s cutoff and confirmed by calculating an anomalous
0.3 mM b-DDM). The peak fraction was concentrated to 8–25 mg ml  1 and used for density map using SAD phasing in CNS program53,54. Anomalous maps were
crystallization. To confirm the role of the SCa binding site, a chimera NaKTs-SFKcsA calculated by calcium anomalous scattering values (f 0 ¼ 0.3002 and f 00 ¼ 0.5654) at
channel was engineered and purified, where the selectivity filter of NaKTs l ¼ 0.9792 Å using calcium positions obtained from the refined NaKTs ion channel
(PMGNGP) was replaced by the corresponding residues from the selectivity filter of structures at 3.15, 3.4 and 3.6 Å resolution and further confirmed by random
KcsA (TVGYGD) using QuikChange Site-Directed Mutagenesis (Agilent anomalous peak search procedure. Anomalous density peaks 410.0s cutoff were
Technologies). The primers used to make the NaKTs-SFKcsA chimera are as follow. also observed for the cluster of calcium ions located at the centre of the hexamer of
To first make a NaKTs without a selectivity filter (NaKTS-noSF), the forward primer tetramer assembly. The calcium ion sub-structure occupancy refinement was
is 50 -CTCGGTGGGATTGCTCAGCCCCACGC-30 and the reverse primer is performed using SAD data in REFMAC 5.5 in the CCP4 suite of programs. The
50 -GCGTGGGGCTGAGCAATCCCACCGAG-30 . To insert the KcsA selectivity calcium ion occupancy along the ion-conduction pathway and their corresponding
filter into NaKTs-noSF, the forward primer is 50 -CTCGGTGGGATTGACCGTC B-factor values are reported in Supplementary Table 2. All structure figures were
GGCTACGGCGACCTCAGCCCCACGC-30 and the reverse primer is 50 -GCGTG generated in PyMOL (http://www.pymol.org/).
GGGCTGAGGTCGCCGTAGCCGACGGTCAATCCCACCGAG-0 3 (insertion in
bold). The NaKTs-SFKcsA chimera was expressed and purified according to the same Functional measurements. All single-channel recording and analysis experiments
protocol as the wild-type protein. were carried out using Port-a-Patch patch clamp system (Nanion Technologies
GmbH, Munich, Germany). Planar lipid bilayers were obtained from GUVs
Crystallization and data collection. Trials were carried out in the presence prepared by using the electroformation method55 in an indium tin oxide (ITO)
and absence of 5 mM CaCl2 in sitting drop 96 well, 3 drop crystallization plates coated glass chamber connected to the Nanion Vesicle Prep Pro setup (Nanion
(Hampton Research), using various commercially available crystallization kits Technologies GmbH, Munich, Germany). A mixture of 10 mM DPhPC
(JCSG þ suite (Qiagen), Wizard I&II (Emerald Bio), HR2 110, HR2 112, HR2 136, (2-diphytanoyl-sn-glycero-3-Phosphocholine, Avanti Polar Lipids Inc.) and 1 mM
PEG I & II (Hampton Research), HTS Screen II (Jena Biosciences)). Crystals were cholesterol, dissolved in chloroform, was deposited on the ITO-coated glass surface
obtained in JCSG þ suite conditions 18 (50% PEG 200/5% PEG 1,000; 0.1 M and air-dried. After complete evaporation of the solvent, the lipid film was
phosphate/citrate, pH 4.2) and 22 (50% PEG 200, 0.2 M MgCl2, 0.1 M sodium rehydrated with 210 ml of 1 M sorbitol. GUVs were formed by electroswelling using
cacodylate, pH 6.5). More than 150 crystals obtained at varying conditions were an alternating electric field of amplitude 3 V, 5 Hz for 2 h. GUVs were collected and
screened at  180 °C under a nitrogen stream. All datasets were collected on the used directly for the reconstitution of the ion channel. Purified chymotrypsin-
Northeastern Collaborative Access Team beamline NECAT 24-ID-C at Advanced treated NaKTs channel was reconstituted into the GUVs solution to a final
Photon Source (APS) of the Argonne National Laboratory (ANL) on a PILATUS concentration of 40 mg ml  1. Giga-ohmic seals were formed within a few seconds
6MF detector at a wavelength of 0.97920 Å. Chymotrypsin-digested NaKTs ion by pipetting 5 ml of the protein-GUVs on the micro-structured glass patch clamp
channel crystals obtained in JCSG þ suite condition 18 diffracted only to 7 Å NPC-1 Chips by Nanion. Channel activity was recorded under symmetric
resolution. The best achievable diffraction data (3.15 Å resolution) were obtained in conditions (unless otherwise stated) using 20 mM HEPES, pH 7.5, 0.2 M NaCl/KCl
JCSG þ condition 22. Three different crystallographic datasets were obtained at as the internal solution and 20 mM MOPS, pH 4, 0.2 M NaCl/KCl, as the external
3.15 Å (dataset 2), 3.36 Å (dataset 3) and 3.6 Å (dataset 1) resolution and the data solution. The channel is active only at pH 4 (MOPS & HEPES); no current was
were processed using the HKL2000 suite49; see Table 1. NaKTs ion channels detected at pH 7.5 (HEPES). The data were filtered at 3 kHz (Bessel filter, HEKA
crystallized in the I4 space group and contained 6 molecules per asymmetric unit. amplifier) digitized at a sampling rate of 50 kHz and analyzed with ClampFit (Axon
Instruments, part of Molecular Devices, USA). Ionic currents were observed only
after adding Ca2 þ in the external solution. All I/V curves were determined using
Structural refinement. Three datasets were collected, labelled chronologically 1, a voltage sweep procedure in which the membrane potential was changed from
2 and 3, at 3.6, 3.15 and 3.4 Å resolution, respectively. For each dataset, the  40 to þ 40 mV in steps of 20 mV lasting 20 ms. Open-channel probability was
asymmetric unit contains six distinct subunits (monomers labelled A, B, C, D, measured as a function of Ca2 þ concentration, with internal solution containing
E, F), which can be used to generate three complete tetramers: Tetramer 1 is 20 mM HEPES, pH 7.5, 0.2 M NaCl and the external solution containing 20 mM
formed by subunits A, B, D and E; tetramer 2 is formed by chain C and chains G, MOPS, pH 4, 0.2 M NaCl. External CaCl2 was applied using a multichannel
H and I generated by crystallographic fourfold symmetry from chain C; tetramer 3 perfusion system linked to the Port-a-Patch. The membrane potential was held at
is formed by chain F and chains J, K and L generated by crystallographic fourfold  40 mV and the current recordings were measured using CaCl2 concentrations of
symmetry from chain F. For dataset 2 (PDB ID: 5CBG), the structure solution was 1, 2, 3, 4, 5, 7.5, 10, 12.5 and 15 mM. Single-channel currents were recorded from
determined by molecular replacement with PHASER in the PHENIX program50 multiple patches for 1–3 min per Ca2 þ concentration using continuous acquisition
using the pore helix (residues 49–61) of NaK (PDB ID: 2AHY) structure as mode. The Po were calculated as the fraction of open time over a total time interval
an initial model. Automated model building was performed using the PHENIX- during which there is channel activity. For each Ca2 þ concentration, about 3–5
Autobuild program and repeated cycles of model building were done using separate time intervals of B200 ms were analyzed and the error on the Po was
COOT51. Structure refinement was done using REFMAC 5.5 in the CCP4 suite of estimated as the mean-square deviation from the average. Single-channel
programs52. Calcium ions were located based on Fo  Fc difference map peak at recordings on the chimeric NaKTs-SFKcsA channel were carried out according to
4.5s cutoff and water molecules at 3s cutoff. The water molecules were also the same procedure.
located based on the Coot:find water facility in REFMAC (CCP4 6.5.001) and
finally checked manually using the electron density map in COOT51. The Building the NaKTs channel–membrane systems. The POPC bilayers systems
N-terminal residues 1–4 and C-terminal residues 107–123 were disordered. The for MD simulations were built from the crystal structure of NaKTs using the
final model contains six monomers (A, B, C, D, E and F) of residues 5–106 and 8 Membrane Builder module56 in CHARMM-GUI57. The N-terminal M0 helix was
Ca2 þ ions refined to Rcryst and Rfree of 23.6 and 29.3%, respectively, with no not included; its inclusion resulted in water penetrating the lipid bilayers according

8 NATURE COMMUNICATIONS | 7:12753 | DOI: 10.1038/ncomms12753 | www.nature.com/naturecommunications


NATURE COMMUNICATIONS | DOI: 10.1038/ncomms12753 ARTICLE

Table 1 | Data collection and refinement statistics.

Dataset 1 (5CBF) Dataset 2 (5CBG) Dataset 3 (5CBH)


Data collection
Space group I4 I4 I4
Cell dimensions
a, b, c (Å) 116.05, 116.05, 132.58 115.57, 115.57, 127.49 116.47, 116.47, 128.13
a, b, g (°) 90, 90, 90 90, 90, 90 90, 90, 90
Resolution (Å) 50–3.60 (3.66–3.60)* 50–3.15 (3.20–3.15)* 50–3.37 (3.42–3.37)*
Rsym or Rmerge 14.4 (75) 16 (66.5) 20.9 (86.7)
I/sI 3 (2.85) 3.11 (3) 3 (2.42)
Completeness (%) 99.3 (99.6) 98.7 (92.2) 96.2 (98.3)
Redundancy 4.8 (4.5) 4.9 (4.4) 3 (3.1)

Refinement
Resolution (Å) 50–3.6 50–3.15 50–3.37
No. reflections 9,567 11,407 13,791
Rwork/Rfree 23.87 (27.9) 23.68 (29.8) 23.22 (26.2)
No. atoms 4,627 4,771 4,633
Protein 4,620 4,620 4,620
Detergent/Ca2 þ 0/3 99/8 0/8
Water — 36 1
B-factors
Protein 114.5 65.5 88
Detergent/ Ca2 þ 0/125 46.7/83.8 0/75.9
Water — 38.4 8.83
R.m.s. deviations
Bond lengths (Å) 0.015 0.013 0.014
Bond angles (°) 2.23 1.88 1.95
R.m.s., root-mean-square.
*Values in parentheses are for highest-resolution shell.

to preliminary simulations (5 ns). The protein-membrane complex was solvated by protocols were analogous to those during the initial equilibration with constant
an equal molar mixture of KCl and NaCl solutions with a cation concentration of pressure (1 atm along the Z axis) and temperature (300 K), except that a dual
1 M (that is, [K þ ] ¼ [Na þ ] ¼ 0.5 M). Each of the four protein subunits was Langevin thermostat was used to keep the Drude particle at 1 K (ref. 64). The first
patched with an acetylated N terminus and an amidated C terminus. The resulting 2.5 ns of each production was TM potential (Vmp) free. Following it was a 30 ns
system comprises B50,000 atoms (80  80  78 Å3). The ABDE tetramer in production run with a Vmp of either 500 or  500 mV to investigate ion
dataset 2 has two Ca2 þ placed diagonally against each other at the SCa site. It was permeation. The MD simulations based on the polarizable force field were stable
used to generate the double SCa occupancy system while also serving as the within a 30-ns duration. This is in stark contrast with the restraint-free NaKTs
prototype for making the zero, single and full (that is, Ca2 þ at all four SCa sites) SCa channel simulations based on the nonpolarizable force field, in which the Ca2 þ
occupancy systems. ions at the SCa site started to dissociate in a much shorter time scale, leading to a
rapid collapse of the selectivity filter.
Equilibration of the NaKTs channel systems. Initial equilibration simulations
were performed in the NPT ensemble with constant pressure (1 atm along the Z
axis) and temperature (300 K) with the NAMD program58 version 2.9. The time Simulating the hydration of pore Ca2 þ ions. To examine the hydration
duration for each equilibration simulation is 10 ns. The pressure of the system was structure of Ca2 þ ions within the selectivity filter pore, four additional systems
controlled by the hybrid Nosé–Hoover Langevin piston method. Langevin were constructed starting from the crystal structures showing four occupancy states
dynamics thermostat with a 1 ps  1 damping coefficient was used to keep the (I to IV in Fig. 4). The setup procedure was the same as described above. Each
temperature constant. The van der Waals interactions were smoothly switched off system was equilibrated for 375 ps with reducing harmonic restraints asserted on
at 10–12 Å by a switching function and the electrostatic interactions were protein heavy atoms and the Ca2 þ . At the end of the equilibration, the force
calculated using the particle-mesh Ewald method with a mesh size of B1 Å. constant of the harmonic restraints was 2 kcal mol  1 Å  2. It was kept the same in
Harmonic positional restraints were applied to the protein backbone atoms and the the 30-ns production simulations. A Langevin thermostat and a Monte Carlo
bound Ca2 þ , keeping the structural deviation from the crystal structure to a barostat were used to maintain the system temperature and pressure at 300 K and
minimum. This was necessary because it prevents the channel from closing. The 1 atm. Both the equilibration and the production were performed with the
force constant of the harmonic potential was set to 5 kcal mol  1 Å  2 Further OpenMM6.2 program65 on GeForce GTX 780 Ti GPU processors.
equilibration of 100 ns for each system was performed on Anton with the volume
and temperature of the system kept constant. Harmonic positional restraints with a
force constant set to 5 kcal mol  1 Å  2 were applied to the backbone and the SCa Data availability. Coordinates and structure factors have been deposited in the
Ca2 þ atoms to prevent the filter from collapsing. The Nosé–Hoover thermostat Protein Data Bank under accession codes 5CBF (dataset1), 5CBG (dataset2) and
was used to maintain the system temperature at 300 K59. The lengths of all bonds 5CBH (dataset3). Remaining data are available from the corresponding author
involving hydrogen atoms were constrained using M-SHAKE60. The cutoff of the upon reasonable request.
van der Waals and short-range electrostatic interactions was set to 16.21 Å as
suggested by the guesser script. Long-range electrostatic interactions were
evaluated using the k-space Gaussian split Ewald method61 with a 64  64  64 References
mesh. The r-RESPA integration method62 was employed and long-range 1. Hille, B. Ion Channels of Excitable Membranes 3rd edn (Sinauer, 2001).
electrostatics were evaluated every 6 fs. The additive C36 force field37,63 was used in 2. Doyle, D. A. et al. The structure of the potassium channel: molecular basis of
both the NAMD and Anton equilibration simulations. K þ conduction and selectivity. Science 280, 69–77 (1998).
3. Heginbotham, L., Abramson, T. & MacKinnon, R. A functional connection
Simulation of the Drude model systems. After equilibration, each system was between the pores of distantly related ion channels as revealed by mutant
then converted to its Drude model representation using the Drude Prepper module K þ channels. Science 258, 1152 (1992).
in CHARMM-GUI57 with the TIP3P water molecules replaced by the polarizable 4. Heginbotham, L., Lu, Z., Abramson, T. & Mackinnon, R. Mutations in the
water SWM4-DNP (ref. 32). The generated Drude model systems of the NaKTs K þ channel signature sequence. Biophys. J. 66, 1061–1067 (1994).
channel were equilibrated using the input scripts provided in the CHARMM-GUI 5. Zhou, M., Morais-Cabral, J. H., Mann, S. & MacKinnon, R. Potassium channel
Drude Prepper module. The production afterwards was carried out using receptor site for the inactivation gate and quaternary amine inhibitors. Nature
NAMD2.9 with the Drude polarizable model implemented64. The simulation 411, 657–661 (2001).

NATURE COMMUNICATIONS | 7:12753 | DOI: 10.1038/ncomms12753 | www.nature.com/naturecommunications 9


ARTICLE NATURE COMMUNICATIONS | DOI: 10.1038/ncomms12753

6. Zhou, M. & MacKinnon, R. A mutant KcsA K( þ ) channel with altered conduction 38. MacKerell, J. A., Feig, M. & Brooks, 3rd C. Improved treatment of the
properties and selectivity filter ion distribution. J. Mol. Biol. 338, 839–846 (2004). protein backbone in empirical force fields. J. Am. Chem. Soc. 126, 698–699
7. Cordero-Morales, J. F. et al. Molecular determinants of gating at the potassium- (2004).
channel selectivity filter. Nat. Struct. Mol. Biol. 13, 311–318 (2006). 39. Uysal, S. et al. Mechanism of activation gating in the full-length KcsA K þ
8. Uysal, S. et al. Crystal structure of full-length KcsA in its closed conformation. channel. Proc. Natl Acad. Sci. USA 108, 11896–11899 (2011).
Proc. Natl Acad. Sci. USA 106, 6644–6649 (2009). 40. Collins, M. D., Smida, J., Dorsch, M. & Stackebrandt, E. Tsukamurella gen. nov.
9. Long, S. B., Campbell, E. B. & Mackinnon, R. Crystal structure of a mammalian harboring Corynebacterium paurometabolum and Rhodococcus aurantiacus.
voltage-dependent Shaker family K þ channel. Science 309, 897–903 (2005). Int. J. Syst. Bacteriol. 38, 385–391 (1988).
10. Long, S. B., Tao, X., Campbell, E. B. & MacKinnon, R. Atomic structure of a 41. Liu, C. Y. et al. Clinical characteristics of infections caused by Tsukamurella
voltage-dependent K þ channel in a lipid membrane-like environment. Nature spp. and antimicrobial susceptibilities of the isolates. Int. J. Antimicrob. Agents
450, 376–382 (2007). 38, 534–537 (2011).
11. Jiang, Y. et al. The open pore conformation of potassium channels. Nature 417, 42. Woo, P. C., Ngan, A. H., Lau, S. K. & Yuen, K. Y. Tsukamurella conjunctivitis: a
523–526 (2002). novel clinical syndrome. J. Clin. Microbiol. 41, 3368–3371 (2003).
12. Ye, S., Li, Y. & Jiang, Y. Novel insights into K þ selectivity from high-resolution 43. Maalouf, R., Mierau, S. B., Moore, T. A. & Kaul, A. First case report of
structures of an open K þ channel pore. Nat. Struct. Mol. Biol. 17, 1019–1023 community-acquired pneumonia due to Tsukamurella pulmonis. Ann. Intern.
(2010). Med. 150, 147–148 (2009).
13. Jiang, Y. et al. X-ray structure of a voltage-dependent K þ channel. Nature 423, 44. Bouza, E. et al. Tsukamurella: a cause of catheter-related bloodstream
33–41 (2003). infections. Eur. J. Clin. Microbiol. Infect. Dis. 28, 203–210 (2009).
14. Kuo, A. et al. Crystal structure of the potassium channel KirBac1.1 in the closed 45. Elshibly, S. et al. Central line-related bacteraemia due to Tsukamurella
state. Science 300, 1922–1926 (2003). tyrosinosolvens in a haematology patient. Ulster Med. J. 74, 43–46 (2005).
15. Nishida, M., Cadene, M., Chait, B. T. & MacKinnon, R. Crystal structure of a 46. Sheridan, E. A. et al. Tsukamurella tyrosinosolvens intravascular catheter
infection identified using 16S ribosomal DNA sequencing. Clin. Infect. Dis. 36,
Kir3.1-prokaryotic Kir channel chimera. EMBO J. 26, 4005–4015 (2007).
16. Lenaeus, M. J., Vamvouka, M., Focia, P. J. & Gross, A. Structural basis of TEA e69–e70 (2003).
47. Maertens, J. et al. Catheter-related bacteremia due to Tsukamurella pulmonis.
blockade in a model potassium channel. Nat. Struct. Mol. Biol. 12, 454–459 (2005).
Clin. Microbiol. Infect. 4, 51–53 (1998).
17. Cuello, L. G., Jogini, V., Cortes, D. M. & Perozo, E. Structural mechanism of
48. Boratyn, G. M. et al. BLAST: a more efficient report with usability
C-type inactivation in K( þ ) channels. Nature 466, 203–208 (2010).
improvements. Nucleic Acids Res. 41, W29–W33 (2013).
18. Cheng, W. W., McCoy, J. G., Thompson, A. N., Nichols, C. G. & Nimigean, C.
49. Otwinowski, Z. & Minor., W. in Method in Enzymology, Macromolecular
M. Mechanism for selectivity-inactivation coupling in KcsA potassium
Crystallography Part A, Vol. 276 (eds Jr. Carter, C.W. and Sweet, R.M.)
channels. Proc. Natl Acad. Sci. U S A 108, 5272–5277 (2011). 307–326 (Academic, 1997).
19. Sali, A. Modeling mutations and homologous proteins. Curr. Opin. Biotechnol. 50. Adams, P. D. et al. PHENIX: a comprehensive Python-based system for
6, 437–451 (1995). macromolecular structure solution. Acta Crystallogr. Sect. D Biol. Cryst. 66,
20. Shi, N., Ye, S., Alam, A., Chen, L. & Jiang, Y. Atomic structure of a Na þ - and 213–221 (2010).
K þ -conducting channel. Nature 440, 570–574 (2006). 51. Emsley, P. & Cowtan, K. Coot: model-building tools for molecular graphics.
21. Alam, A. & Jiang, Y. High-resolution structure of the open NaK channel. Acta. Crystallogr. D Biol. Cryst. 60, 2126–2132 (2004).
Nat. Struct. Mol. Biol. 16, 30–34 (2009). 52. Potterton, E., Briggs, P., Turkenburg, M. & Dodson, E. A graphical user
22. Sauer, D. B., Zeng, W., Canty, J., Lam, Y. & Jiang, Y. Sodium and potassium interface to the CCP4 program suite. Acta Crystallogr. Sect. D Biol. Cryst. 59,
competition in potassium-selective and non-selective channels. Nat. Commun. 1131–1137 (2003).
4, 2721 (2013). 53. Brunger, A. et al. Crystallography & amp; NMR system: a new software suite
23. Derebe, M. G. et al. Tuning the ion selectivity of tetrameric cation channels for macromolecular structure determination. Acta Crystallogr. D Biol. Cryst. 54,
by changing the number of ion binding sites. Proc. Natl Acad. Sci. USA 108, 905–921 (1998).
598–602 (2011). 54. Brunger, A. T. Version 1.2 of the Crystallography and NMR System.
24. Liu, S. & Lockless, S. W. Equilibrium selectivity alone does not create K( þ )- Nature Protocols 2, 2728–2733 (2007).
selective ion conduction in K( þ ) channels communications. Nature 4, 2746 (2013). 55. Angelova, M. I. in Liposome Electroformation, in Perspectives in Supramolecular
25. Cortes, D. M., Cuello, L. G. & Perozo, E. Molecular architecture of full-length Chemistry: Giant Vesicles Vol. 6 (eds P. L. Luisi and P. Walde) 26–36 (John
KcsA: role of cytoplasmic domains in ion permeation and activation gating. Wiley & Sons, Ltd., Chichester 1999).
J. Gen. Physiol. 117, 165–180 (2001). 56. Wu, E. L. et al. CHARMM-GUI Membrane Builder toward realistic biological
26. Posson, D. J., Rusinova, R., Andersen, O. S. & Nimigean, C. M. Calcium ions membrane simulations. J. Comp. Chem. 35, 1997–2004 (2014).
open a selectivity filter gate during activation of the MthK potassium channel. 57. Jo, S., Kim, T., Iyer, V. G. & Im, W. CHARMM-GUI: a web-based
Nat. Commun. 6, 8342 (2015). graphical user interface for CHARMM. J. Comp. Chem. 29, 1859–1865
27. Lam, Y. L., Zeng, W., Derebe, M. G. & Jiang, Y. Structural implications of weak (2008).
Ca2 þ block in Drosophila cyclic nucleotide-gated channels. J. Gen. Physiol. 58. Phillips, J. C. et al. Scalable molecular dynamics with NAMD. J. Comput. Chem.
146, 255–263 (2015). 26, 1781–1802 (2005).
28. Valasatava, Y., Rosato, A., Cavallaro, G. & Andreini, C. MetalS(3), 59. Martyna, G. J., Klein, M. L. & Tuckerman, M. Nose-Hoover chains—the
a database-mining tool for the identification of structurally similar metal sites. canonical ensemble via continuous dynamics. J. Chem. Phys. 97, 2635–2643
J. Biol. Inorg. Chem. 19, 937–945 (2014). (1992).
29. Baranova, E. et al. SbsB structure and lattice reconstruction unveil Ca2 þ 60. Krautler, V., Van Gunsteren, W. F. & Hunenberger, P. H. A fast SHAKE:
triggered S-layer assembly. Nature 487, 119–122 (2012). Algorithm to solve distance constraint equations for small molecules in
30. Eswaramoorthy, S., Kumaran, D., Keller, J. & Swaminathan, S. Role of metals in molecular dynamics simulations. J. Comp. Chem. 22, 501–508 (2001).
the biological activity of Clostridium botulinum neurotoxins. Biochemistry 43, 61. Shan, Y., Klepeis, J. L., Eastwood, M. P., Dror, R. O. & Shaw, D. E. Gaussian
2209–2216 (2004). split Ewald: A fast Ewald mesh method for molecular simulation. J. Chem. Phys.
31. Lamoureux, G. & Roux, B. Modeling induced polarization with classical Drude 122, 054101 (2005).
oscillators: theory and molecular dynamics simulation algorithm. J. Chem. 62. Tuckerman, M., Berne, B. J. & Martyna, G. J. Reversible multiple time scale
Phys. 119, 3025–3039 (2003). molecular-dynamics. J. Chem. Phys. 97, 1990–2001 (1992).
32. Lamoureux, G., Harder, E., Vorobyov, I. V., Roux, B. & Mackerell, Jr. A. D. A 63. Best, R. B. et al. Optimization of the additive CHARMM all-atom protein
polarizable model of water for molecular dynamics simulations of force field targeting improved sampling of the backbone phi, psi and
biomolecules. Chem. Phys. Lett. 418, 245–249 (2006). side-chain chi(1) and chi(2) dihedral angles. J. Chem. Theo. Comp. 8,
33. Yu, H. et al. Simulating monovalent and divalent ions in aqueous solution using 3257–3273 (2012).
64. Jiang, W. et al. High-performance scalable molecular dynamics simulations of a
a drude polarizable force field. J. Chem. Theo. Comp. 6, 774–786 (2010).
polarizable force field based on classical Drude oscillators in NAMD. J. Phys.
34. Chowdhary, J. et al. A polarizable force field of dipalmitoylphosphatidylcholine
Chem. Lett. 2, 87–92 (2011).
based on the classical drude model for molecular dynamics simulations of
65. Eastman, P. et al. OpenMM 4: a reusable, extensible, hardware independent
lipids. J. Phys. Chem. B 117, 9142–9160 (2013).
library for high performance molecular simulation. J. Chem. Theo. Comp. 9,
35. Lopes, P. E. et al. Force field for peptides and proteins based on the classical
461–469 (2013).
drude oscillator. J. Chem. Theo. Comp. 9, 5430–5449 (2013).
36. Li, H. et al. Representation of ion–protein interactions using the drude
polarizable force-field. J. Phys. Chem. B 119, 9401–9416 (2015). Acknowledgements
37. Klauda, J. B. et al. Update of the CHARMM all-atom additive force field for This research was supported by NIH-NIGMS through Grant R01-GM062342. This
lipids: validation on six lipid types. J. Phys. Chem. B 114, 7830–7843 (2010). work is based on research conducted at the Advanced Photon Source (APS) on the

10 NATURE COMMUNICATIONS | 7:12753 | DOI: 10.1038/ncomms12753 | www.nature.com/naturecommunications


NATURE COMMUNICATIONS | DOI: 10.1038/ncomms12753 ARTICLE

Northeastern Collaborative Access Team beam lines, which are supported by a grant Additional information
from the National Institute of General Medical Sciences (P41 GM103403) from the Supplementary Information accompanies this paper at http://www.nature.com/
National Institutes of Health. Use of the Advanced Photon Source, an Office of Science naturecommunications
User Facility operated for the U.S. Department of Energy Office of Science by
Competing financial interests: The authors declare no competing financial interests.
Argonne National Laboratory, was supported by the U.S. DOE under Contract No.
DE-AC02-06CH11357. We wish to express the gratitude to the staff at NE-CAT-24-ID-C Reprints and permission information is available online at http://npg.nature.com/
and E beam lines at the Advanced Photon Source in the Argonne National Laboratory. reprintsandpermissions/
Anton computer time was provided by the National Center for Multiscale Modeling of
Biological Systems (MMBioS) through Grant P41GM103712-1 from the National How to cite this article: Dhakshnamoorthy, B. et al. Structural and functional
Institutes of Health and the Pittsburgh Supercomputing Center (PSC). The Anton characterization of a calcium-activated cation channel from Tsukamurella
machine at PSC was generously made available by D.E. Shaw Research. paurometabola. Nat. Commun. 7:12753 doi: 10.1038/ncomms12753 (2016).

This work is licensed under a Creative Commons Attribution 4.0


International License. The images or other third party material in this
Authors contributions article are included in the article’s Creative Commons license, unless indicated otherwise
B.D. identified the channel from genomic databases, solved the crystal structures and in the credit line; if the material is not included under the Creative Commons license,
carried out electrophysiological experiments. A.R. carried out electrophysiological users will need to obtain permission from the license holder to reproduce the material.
experiments and completed the refinement of the crystal structures. L.B. did all the To view a copy of this license, visit http://creativecommons.org/licenses/by/4.0/
molecular biology cloning. H.R. carried out structural analysis and generated the MD
simulations. B.R. designed and supervised the research. All authors contributed to
writing the manuscript. r The Author(s) 2016

NATURE COMMUNICATIONS | 7:12753 | DOI: 10.1038/ncomms12753 | www.nature.com/naturecommunications 11

You might also like