Multiple Modes of A-Type Potassium Current Regulation

advertisement
3178
Current Pharmaceutical Design, 2007, 13, 3178-3184
Multiple Modes of A-Type Potassium Current Regulation
Shi-Qing Cai, Wenchao Li and Federico Sesti*
University of Medicine and Dentistry of New Jersey, Robert Wood Johnson Medical School, Department of Physiology and Biophysics, 683 Hoes Lane, Piscataway, NJ 08854, USA
Abstract: Voltage-dependent potassium (K+) channels (Kv) regulate cell excitability by controlling the movement of K+ ions across the
membrane in response to changes in the cell voltage. The Kv family, which includes A-type channels, constitute the largest group of K+
channel genes within the superfamily of Na+, Ca 2+ and K+ voltage-gated channels. The name “A-type” stems from the typical profile of
these currents that results form the opposing effects of fast activation and inactivation. In neuronal cells, A-type currents (IA), determine
the interval between two consecutive action potentials during repetitive firing. In cardiac muscle, A-type currents (Ito), control the initial
repolarization of the myocardium. Structurally, A-type channels are tetramers of -subunits each containing six putative transmembrane
domains including a voltage-sensor. A-type channels can be modulated by means of protein-protein interactions with so-called -subunits
that control inactivation voltage sensitivity and other properties, and by post-transcriptional modifications such as phosphorylation or
oxidation. Recently a new mode of A-type regulation has been discovered in the form of a class of hybrid -subunits that posses their
own enzymatic activity. Here, we review the biophysical and physiological properties of these multiple modes of A-type channel regulation.
INTRODUCTION
A-type currents were first described in molluscan neurons by
Hagiwara and subsequently found in arthropod muscle and neuron,
and in vertebrate neurons and heart [1-3]. The name A-type derives
from the typical profile of these voltage-dependent K+ currents:
rapid activation at subthreshold voltages followed by fast inactivation. A-type channels are abundantly expressed in neurons and
cardiac and smooth muscle cells. In the nervous system, A-type
currents play a key role in the control of excitability, synaptic input,
and neurotransmitter release. In many neurons, A-type channels are
usually silent at resting membrane potentials. They however, transiently activate during the decay of the after-hyperpolarization
phase of the action potential, delaying depolarization [4]. Thus, Atype currents prolong the period between action potentials and control the neuronal firing frequency by means of inactivation. A-type
currents are present in atrial and ventricular myocytes where they
are called "transient" outward currents (Ito) and govern the initial
repolarization of the ventricular action potential [5, 6]. A-type currents exist in vascular smooth muscle cells, even though their
physiological role is not fully understood [4].
Table 1.
u
rib
The Kv Family of Voltage-Gated K + Channels
t
s
i
D
r
o
F
t
o
N
STRUCTURE
In mammals, genes with A-type properties are found in several
sub-families of voltage-gated K+ channels, including the Kv1
(Shaker-like), Kv3 (Shaw-like) and Kv4 (Shal-like) sub-families
(Table 1). In addition, co-expression of accessory subunits with
certain Kv -subunits confers rapid inactivation to otherwise noninactivating or moderately inactivating delayed rectifier channels
[7-10]. Kv channels exhibit a predicted topology of six transmembrane spans (from S1 to S6) and a selectivity filter located between
S5 and S6. The fourth transmembrane domain, S4, provides the
location of the voltage sensor, composed of several lysine and arginine residues. Functional Kv channels are formed by four subunits packed along a central axis of symmetry, orthogonal to the
plasma membrane. The crystal structures of a bacterial and a mammalian Kv channel have been solved [11, 12]. While, the ionconduction pathway is conserved in these structures and does not
differ form that of other bacterial K+ channels such as KcsA [13],
the putative structure of the voltage-sensing domain (S1-S4) shows
*Address correspondence to this author at the University of Medicine and
Dentistry of New Jersey, Robert Wood Johnson Medical School, Department of Physiology and Biophysics, 683 Hoes Lane, Piscataway, NJ 08854,
USA; Tel: (732) 235 4032; Fax: (732) 235 5038;
E-mail: SESTIFE@UMDNJ.EDU
1381-6128/07 $50.00+.00
n
tio
a cytoplasmic location of S4 when the channel is in the closed state.
This model is presently controversial because it contradicts a large
body of earlier evidence that predicts that S4 spans the plasma
membrane.
Sub-Family
Alternative Name
Kv1.x
Shaker-like
Kv2.x
Kv3.x
Kv4.x
IA, IKUR
Shab-like
Shaw-like
IA
Shal-like
IA
Kv5.x
None
Kv6.x
Kv7.x
Current
None
KCNQ1
Kv8.x
IKs, M-current
None
Kv9.x
None
Kv10.x
eag
Kv11.x
erg
Kv12.x
elk
IKr
The family of voltage-gated K+ channel is composed of 12 sub-families. Sub-families
Kv1, Kv3 and Kv4 encode A-type channels. Kv5.x, Kv6.x, Kv8.x and Kv9.x channels
do not express measurable currents when expressed in heterologous expression systems
suggesting that these genes might encode -subunits that need to associate with other
Kv subunits to conduct currents.
N-TYPE INACTIVATION
In the late 70s, Armstrong and Bezanilla in an effort to explain
the mechanism of A-type inactivation, proposed the “ball-andchain” model, which postulated that a positively charged inactivation particle (the -ball) on a tether, prevents the movement of ions
by physically occluding the pore [14]. Using the Drosophila
channel Shaker, Zagotta, Hoshi and Aldrich, identified the location
and molecular composition of the ball. This is composed, roughly,
of the first 20 amino acids in the amino-terminal—thus the name Ntype inactivation—followed by 40 or more residues constituting the
chain [15, 16]. The inactivation ball possesses two essential
chemical characteristics: the first half is composed of hydrophobic
residues. The rest has a net positive charge which pushes the ball
toward its channel receptor site upon depolarization [17, 18]. The
© 2007 Bentham Science Publishers Ltd.
Multiple Modes of A-Type Potassium Current Regulation
details of the N-inactivation mechanism are now beginning to be
elucidated. Mackinnon and colleagues have proposed that Ninactivation occurs as a sequential reaction: the ball initially binds
to the T1 domain surface by electrostatic interactions, then it enters
through the lateral portals that connect the cytoplasm to the inner
pore, reaches the inner cavity to eventually plug the pore (Fig. 1A)
[12, 19].
Like the majority of ion-channels, A-type channels generally
contain additional regulatory proteins or -subunits that fine-tune
their properties, including trafficking, location and abundance, sensitivity to stimulation, pharmacology and gating. In the case of Atype channels, -subunits play a pivotal role in the control of inactivation, which is probably the most important physiological attribute
of these currents. Four major classes of ancillary subunits of A-type
channels have been described which differ in their sequences, structures, effects and -subunit specificity. Post-translational modifications are also pivotal in A-type K+ channel modulation. Protein
phosphorylation, for instance, represents a widespread mechanism
to regulate cellular excitability by temporarily modifying the function of K+ channels. These two fundamental modes of modulation
can co-exist in the same channel protein. These are channels containing hybrid -subunits that possess their own enzymatic activity.
Here we review the attributes of the Kv2 subunit, a soluble protein
with aldo-keto-reductase (AKR) activity [20] and of a C. elegans
KCNE-related gene, MPS-1 which exhibits kinase activity [21].
These multiple modes of channel regulation are graphically illustrated in (Figs. 1 and 2).
Current Pharmaceutical Design, 2007, Vol. 13, No. 31
brain (Kv1 and Kv2) [7], from ferret (Kv3) [24] and human
heart (hKv3) [25, 26]. Kv-subunits are soluble proteins lacking
any transmembrane domains, leader sequence or N-glycosylation
consensus sites. There are several protein kinase A (PKA) and Casein kinase 2 (CKII) consensus target sites in all Kv sequences.
The N-terminus of the Kv subunits is variable. In some isoforms
the first ± 20 residues constitute the inactivating ball (-ball) which
confers A-type characteristics to non-inactivating Kv -subunits.
Thus, Kv1 and Kv3 subunits interact with the Kv1 family of subunits and induce rapid inactivation in all Kv1 channels [7, 10,
27, 28], except Kv1.6 that has a NIP—an N-type inactivation prevention domain [29]. Kv2 has no inactivation ball therefore it does
not convert their partners into A-type channels although it can accelerate inactivation kinetics [30]. Generally, the -ball and the ball do not have direct interactions [31] although they may share
sequence homology as, for instance, the -balls of Kv1 and Kv3
with the -balls of Kv1.4 and Kv3.4. In contrast, the Kv Cterminus, which is responsible for the binding to the -subunit, is
more conserved. This domain shares sequence homology to AKR,
an enzyme that catalyzes a redox reaction using an NADPH cofactor (the reduced form of nicotinamide adenine dinucleotide phosphate). The catalytic domain is important for channel trafficking.
Thus, immunofluorescence and protein chemistry studies showed
that the expression domain alone was sufficient to increase the surface expression of Kv1.2 in heterologous expression systems [31].
It was shown that mutations in the cofactor NADP(+) binding domain, abolished the ability of several Kv isoforms to alter the trafficking of Kv1.2 [32, 33]. Intriguingly, mutations in a key catalytic
tyrosine residue found in the active site of all AKRs did not affect
trafficking characteristics. In contrast, when residues within the
NADP(+) binding pocket were mutated, they suppressed Kvmediated effects on Kv1.2 trafficking suggesting that the integrity
of the NADP(+) binding pocket, rather than catalytic activity, is
required for intracellular trafficking of Kv-Kv1 channel complexes
[32]. The structures of rat Kv2 [34] alone and in complexes
formed with the T1 domain of rat Kv1.1 [35] and with the entire rat
o
F
t
o
N
n
tio
u
rib
t
s
i
D
r
ANCILLARY SUBUNITS
Kv-Subunits
Kv-subunits were originally identified in bovine rat extracts by
Parcej and colleagues [22]. Two years later, the same group isolated
cDNA encoding a -subunit co-purifying with the bovine brain
DTX acceptor complex [23]. Subsequently, cDNAs encoding three
highly related distinct -subunit isoforms were cloned from rat
3179
Fig. (1). N-type inactivation and its regulation by enzymes. A. Illustration of N-type K+ channel inactivation. The inactivation peptide binds to the T1 domain by electrostatic interactions and then it enters through the later openings to plug the pore. B. Enzymatic modulation of N-type inactivation. Catalytic
incorporation of negatively charged phosphates in the inactivation peptide affects its binding and/or movement by electrostatic mechanisms. C. Metabolic
modification of cysteines is associated with the formation of disulphide bridges between the inactivation peptides of -subunits or -subunits.
3180 Current Pharmaceutical Design, 2007, Vol. 13, No. 31
Kv1.2 [12] have been resolved crystallographically (Fig. 2A, B and
C). These analyses show that the Kv2-subunit binds to the T1
domain in a fourfold symmetric complex, with four T1 domains
facing the transmembrane pore and four Kv2 subunits facing the
cytoplasm. The T1 domain exhibits four prominent loops termed
contact loops that provide a large docking surface for Kv2. The T1
domain is attached to the rest of the channel by four tethers that
define four lateral openings through which K+ ions enter the channel’s pore.
K+ CHANNEL-INTERACTING PROTEINS (KCHIPS SUBUNITS)
The K+ channel interacting proteins, KChIPs, constitute an
important family of soluble auxiliary subunits of Kv4.x channels.
Four KChIPs have been identified to date, named from KChIP1 to
KChIP4. The first three members of this family, were identified by
yeast-two-hybrid screens using the N-terminus of the Kv4 channel
as bait [36]. KchIP4 was cloned later, after identification as a binding partner of presenilin 2 [37]. All four KChIP genes are highly
expressed in the brain, whereas only KChIP2 mRNA is abundant in
the heart [38]. Kv4-KchIP complexes recapitulate some gating
properties of native A-type currents in cardiac myocytes and neurons [39]. Rosati and colleagues showed a gradient of KChIP2 expression across the ventricular wall consistent with the gradient of
Ito [40]. Consistent with these data, KChIP2 knock-out mice exhibited prolonged QT interval and were susceptible to ventricular arrhythmia during electrical stimulation [41]. No description of the
potential phenotype of KChIP2 in the nervous system has been
provided however. KChIPs belong to the recoverin/NCS subfamily
of Ca2+-binding proteins and contain calcium-binding consensus
sites, composed of four EF hands each formed by two -helices
[36], in the conserved C-terminal domain. The physiological significance of these EF hands is still under investigation although
evidence suggests that hands 3 and 4 may act as calcium binders.
The KChIP subunits enhance the Kv4.2 trafficking from the endoplasmic reticulum to the plasma membrane roughly eight-fold [36,
42]. KChIP2 and 3 are targeted to the plasma membrane by palmitoylation and their plasma membrane localization is enhanced by
co-expression with Kv4.3 [43]. In contrast, KChIP1 is not palmitoylated but it is predicted to be myristoylated at its N terminus,
which contributes to the trafficking of Kv4 channels [44]. When
heterologously co-expressed, KChIPs dramatically alter the inactivation kinetics and the rate of recovery from inactivation of Kv4
channels [36]. Isoforms KchIP1 through 3 modulate Kv4 channels
in a similar fashion, and their co-expression decelerates the normally rapid rate of voltage-dependent inactivation and recovery
from inactivation [45, 46]. KchIP4, however, delays KchIP.Kv4
channel opening and it abolishes fast voltage-dependent inactivation due to a unique inactivation suppressor domain (KIS domain )
in the N-terminus which effectively converts Kv4 to a slowly inactivating delayed rectifier channel [47]. It is interesting to note that
regulatory domains such as NIP and KIS are not a prerogative of
accessory subunits. Recently we showed that a C. elegans poreforming subunit, KVS-1, possesses an N-inactivating ball preceded
by a N-Inactivation_Regulatory_Domain (NIRD) that acts to slow
down N-inactivation through steric mechanisms [48].
The structure of the KChIP1–Kv4.3 T1 complex was resolved
by the means of crystallography (Fig. 2D) [49]. It shows a crossshaped octamer having the T1 tetramer at the center and individual
KChIPs extending radially. Ten KChIP1 -helices form a clamshell-like structure organized in four hands each composed by two
helices. The tenth helix, H10, anchors the KChIP1 subunit to the
channel by interacting with its N-terminus. There are two principal
interaction sites. A Kv4 channel N-terminal hydrophobic segment
interacts with the KChIP1 H10 helix. The second point of contact
occurs at the level of a T1 assembly domain loop and the H2 helix
of KChIP1. Functional and biochemical studies indicate that the
first site controls channel trafficking, whereas the second affects
Cai et al.
channel gating. These data are consistent with previous crystallographic studies of complexes formed by KChIP1 and partial Kv4
N-terminus domains [50, 51]. Moreover, the T1 docking loop provides specificity of interaction with Kv4 channels. When this domain was inserted into the Kv1.2 N terminus, KChIP1 could assemble with this mutant channel [51]. The stoichiometry of Kv4.2KChIP2 channel complex was determined by Kim and colleagues
as 1 to 1 ratio [52]. Using electron microscopy, the same group
visualized the Kv4.2-KchIP2 channel complex at 21 Å resolution
[53]. Four KChIP2 subunits fit into four peripheral columns that
extend from the membrane embedded portion of Kv4.2.
KCNE ANCILLARY SUBUNITS
KCNE genes encode integral membrane proteins that function
as ancillary subunits of Kv channels. KCNE genes are found in
invertebrates (C. elegans), amphibians (Xenopus laevis) and mammals, including homo sapiens [8, 9, 54, 55]. KCNE proteins range
from approx 10-30 kDa and have a single transmembrane span
which is the most conserved domain in these proteins. KCNE proteins have been shown to modulate mammalian A-type currents.
Thus, KCNE2, which is expressed in human left ventricle assembles with Kv4.2 in heterologous expression systems suggesting that
KCNE2 might be a component of Ito native channels [56]. In addition, KCNE subunits can interact with non-inactivating or moderately inactivating Kv channels which however can be converted
into A-type channels by KCNE and/or Kv subunits [57-61]. The
importance of KCNE genes in regulating A-type currents is emphasized in C. elegans. Four KCNE-related genes, termed mps, operate
in the nervous system of the animal [8, 9]. MPS proteins have been
shown to interact with a moderately inactivating potassium channel,
KVS-1, in vitro and in vivo. When expressed in heterologous expression systems each MPS protein acts to alter KVS-1 attributes.
MPS-1 and MPS-4 convert KVS-1 into an A-type channel. In contrast, MPS-3 suppresses KVS-1 inactivation. It is interesting to note
that MPS proteins are a means to generate heterogeneity of Kv
currents in the same neuron. In the chemosensory ADF neurons,
MPS-1, MPS-2 and MPS-3 combine with KVS-1 to form two distinct channel complexes: binary complexes containing only MPS-1
subunits (MPS-1-KVS-1) and ternary complexes containing MPS2, MPS-3 and KVS-1. These complexes appear to control specific
neuronal functions. The MPS-2-MPS-3-KVS-1 ternary complex in
particular, is required to fine tune the taste of the animal for sodium
ions. [8]. The role of the MPS-1-KVS-1 complex in ADF neurons is
more elusive. This protein is expressed in several other chemosensory neurons therefore its destabilization does not impact a single
neuron type but rather causes multiple sensory defects [9].
u
rib
t
s
i
D
r
o
F
t
o
N
n
tio
DIPEPTIDYL PEPTIDASE-LIKE PROTEINS
The fourth class of ancillary subunits of A-type channels discussed in this review is represented by the peptidase-like proteins.
The name of these subunits stems form their sequence homology to
the peptidyl aminoprotease CD26/DPP4 [62]. Despite this similarity, they lack enzymatic activity, due to a S to D substitution in the
DDH catalytic triad [63]. DPP6 was originally identified by copurification with Kv4.2 from rat brain membranes [62]. DPP6 facilitates trafficking and membrane targeting, accelerates inactivation
kinetics, alters the voltage dependence of activation and inactivation and accelerates recovery from inactivation thereby conferring
ISA-like properties to the current. At present two DPP-like subunits,
DPP6 and DPP10, have been shown to be essential components of
Kv4 and Kv1 potassium channels [62, 64, 65]. Structurally, DPP
subunits have a single transmembrane domain a short N-terminus
and a large C-terminus containing the aminopeptidase-like domain.
The protein is oriented in the membrane with the N-terminus in the
cytoplasm. The crystal structure of the amino-peptidase-like site of
DPP6 (extracellular domain) has been solved at 3 Å resolution (Fig.
2E) [63]. It is a dimer formed by two identical monomers. Each
monomer is composed by a eight-bladed -propeller and a / hy-
Multiple Modes of A-Type Potassium Current Regulation
Current Pharmaceutical Design, 2007, Vol. 13, No. 31
n
tio
u
rib
t
s
i
D
r
o
F
t
o
N
3181
Fig. (2). Experimental structures of -subunts.
A. Crystal structure of Kv2 (amino acids 36 to 367) at 2.8 Å resolution. The AKR domain is depicted in turquoise. Adapted from [34].
B. Surface representation of the Kv1.2-Kv2 complex at 2.9 Å resolution. Basic and acidic residues are depicted, respectively in blue and red colors. The
membrane spanning portion of the channel, (TM], the T1 domain and Kv2, (), are indicated. Adapted from [12].
C. Hypothetical model for the binding of the -ball to the channel. An inactivation peptide corresponding to the 22 N-terminal residues of Kv1 was modeled
in an extended conformation to enter a side portal and plug the inner pore of the channel. Adapted from [12].
D. Model of the Kv4.3–KChIP complex based on the Kv1.2 structure. KChIP1 (cyan), Kv4.3 T1 domain (orange) and the transmembrane domains (TM, dark
red) are indicated. Adapted from [49].
E. Crystal structure the extracellular domain of the DPP6 dimer at 3.0 Å resolution. The color changes from blue (N terminus) to red (C terminus). Carbohydrates with glycosylation sites are shown in ball-and-stick representation and the “catalytic triad” is shown in CPK representation. Adapted from [63].
drolase domain. The dimer interface is formed by the last -strand
of the propeller domain and two helices of the / hydrolase and by
the -hairpin insertion motif of the -propeller domain. The active
site is located at the interface of the / domain and the propeller.
DPP subunits exhibit a slightly higher abundance of histidine residues than average suggesting that they might have a role in regulating channels’ susceptibility to extracellular pH.
POST-TRANSCRIPTIONAL MODIFICATIONS
A large body of evidence supports the notion that A-type K+
currents in both neuronal and cardiac cells can be modulated by
phosphorylation [21, 39, 66-68]. Often, phosphorylation of these
channels results in altered N-type inactivation. This is not surprising if one considers the key role of electrostatic charge in the
mechanisms underlying N-type inactivation as well the physiological role of this feature (Fig. 1B). Thus, phosphorylation of serine
residues in the -ball in Kv3.4 abolished N-inactivation [67]. Structural analysis by NMR indicated that this mechanism stems from a
structural rearrangement of the ball which distinctly alters inactivation kinetics [66]. Phosphorylation of a tyrosine in the inactivation
ball of the Shaker channel [69], and phosphorylation of the Nterminus of Kv1.3 [70] also removed the fast inactivation of these
currents. Kv4.x channels have been reported to be phosphorylated
by PKA [71], PKC [72], MAPK [73-75] and SGK [76] in vivo and
in vitro. Activation by PKC suppresses Kv4.2 or Kv4.3 current in
Xenopus oocytes and PKA suppresses Kv4.2 current in hippocampal neurons; on the contrary, co-expression with SGK1 increases
Kv4.3 channel current in heterologous expression systems. Phosphorylation does not always induce detectable functional modifications. For instance, two PKA sites and three ERK/MAPK phosphorylation sites [73] in Kv4.2 are phosphorylated in vivo [71].
However, PKA phosphorylation of only one site—serine-552—has
been found to affect channel’s function: it shifts the voltage dependence of activation of the channel [75]. Some protein kinases
such as ERK/MAPK act to regulate their channel targets directly
[74]. Other kinases need anchor proteins to assist them to modulate
potassium channel currents. Schradder and colleagues found that
PKA regulation of Kv4.2 currents required the presence of a KChIP
subunit [75]. Despite the ability of PKA to phosphorylate the Kv4.2
-subunit in the absence of KChIP3, PKA was unable to alter the
3182 Current Pharmaceutical Design, 2007, Vol. 13, No. 31
biophysical properties of the channel by this mechanism alone.
These findings reveal an unexpected complexity to the structurefunction relationships for kinase regulation of membrane potassium
channels.
A variety of agents normally present in biological cells can
oxidize A-type channels leading to removal or introduction of Ntype inactivation. Thus, a cloned mammalian A type potassium
channel, RCK, lost its N-type component of inactivation after oxidation of a cysteine and consequent formation of disulphide bridges
in the inactivation ball domain (Fig. 1C) [77]. On the other side of
the spectrum, oxidation of histidines and cysteines in the Nterminus of mouse Kv1.7 induced fast inactivation, whereas reducing conditions reversed this effect [78]. The same phenomenon was
observed in the potassium channels Kv1.4, Kv3.4 and that associate
with Kv subunits that also have a conserved cysteine in the -ball
[79, 80]. Oxidation of residues other than cysteines has also been
shown to affect N-type inactivation. Conversion of methionine into
methionine sulfoxide in the inactivation ball of the ShC/ShB channel, suppressed N-type inactivation [81, 82]. Unlike cysteine, oxidation of methionine cannot be reversed by reducing agents and in
these cases enzymes such as the peptide methionine sulfoxide reductase mediate the reversibility of these reactions. It is not clear
whether oxidation of methionine is a common mechanism underlying potassium channel regulation. For example, oxidation of methionine in the N-terminus of mouse KV1.7 did not affect Ninactivation [78].
In addition to phosphorylation and oxidation other common
post-transcriptional modifications such as N-glycosylation and
palmitoylation have been shown to influence A-type function, but
little is known about the impact of these mechanisms on native Atype channels [43, 83]. Finally, Oliver and others showed that
phospholipids can control A-type inactivation in vitro [84]. Thus,
phosphoinositides can bind the -ball of Kv1.1 and Kv3.4 removing
N-type inactivation whereas arachidonic acid induces N-inactivation in Kv3.1. Given the second messenger role of PIP2, these findings may underscore a new mode of enzymatic modulation—by
phospholipases—of A-type currents.
Cai et al.
elements that can act directly to alter their own function. This leads
us to speculate that Kv and MPS-1 may provide examples of novel
mechanisms of dynamic regulation of electrical signaling in the
nervous system. Further studies will address these fundamental
questions.
FUTURE DIRECTIONS AND THE C. ELEGANS MODEL
Even though it is well known that the regulation of A-type potassium channels by protein-protein interactions or posttranslational modifications are critical for life processes, the big
challenge is to determine where and when these regulations occur
in vivo. Invertebrate animal models such as C. elegans can significantly contribute to this effort. All major families of K+ channel
genes are represented in C. elegans [85]. To date, two Kv channels
in C. elegans have been shown to produce A-type currents [9, 86]
and a sub-family of KCNE genes was recently identified [8, 9]. The
possibility to perform genetic manipulations and to assess the effect
of these manipulations in vivo along with new techniques that have
remarkably advanced the ability to electrophysiologically record K +
currents in native cells [87] make C. elegans an unprecedented tool
for the study of the A-type K+ channels. C. elegans seems a reasonable system to overcome the limitation of higher organisms without
losing the ability to work in true physiological context.
u
rib
o
F
t
o
N
t
s
i
D
r
ENZYMATIC MODULATION OF A-TYPE CURRENTS BY
-SUBUNITS
Recent observations indicate that the two fundamental modes of
A-type K+ channel regulation—assembly with -subunits and enzymatic modulation—can co-exist in the same protein [20, 21], as
bifunctional -subunits which posses enzymatic activity. We mentioned before that the C-terminus of the Kv subunits shares sequence homology to AKR. The role of the putative catalytic activity
of Kv, as well its biological function, has been a mystery for a
long time, until recently it was demonstrated that Kv2 possesses
AKR activity that controls inactivation kinetics [20]. Using KV1.4
as substrate, Weng and colleagues showed that Kv2 modulates
Kv1.4 channel inactivation through oxidizing the bound NADPH
cofactor. Since the subunit can assemble with a series of poreforming subunits in different tissues, the finding may provide a link
between modulation of potassium channels and the redox state in
vivo. The second example is represented by the C. elegans KCNErelated gene, MPS-1 [21]. Nine characteristic protein kinase catalytic subdomains can be recognized in the C-terminus of MPS-1
(Shi-Qing Cai unpublished observations). Recombinant MPS-1 was
shown to catalyze incorporation of 32P into myelin basic protein (a
standard S/T kinase substrate) and more importantly, MPS-1 was
found to phosphorylate one of its physiological substrates the KVS1 K+ channel. As with Kv subunits, MPS-1 also acts to modulate
KVS-1 function through independent mechanisms: the kinase activity decreases the steady-state open probability of the complex, resulting in reduced macroscopic current, whereas its -subunit nature confers A-type characteristics. These bifunctional -subunits
convert their K+ channel partners from passive elements into active
n
tio
ACKNOWLEDGEMENTS
We are deeply indebted to Dr. John Lenard for critical reading
of the manuscript. This work was supported by an NIH grant
(R01GM68581) to FS.
REFERENCES
References 88-90 are related articles recently published in Current
Pharmaceutical Design.
[1]
[2]
[3]
[4]
[5]
[6]
[7]
[8]
[9]
[10]
[11]
[12]
[13]
Hagiwara S, Kusano K, Saito N. Membrane changes of Onchidium
nerve cell in potassium-rich media. J Physiol 1961; 155: 470-89.
Llinas RR. The intrinsic electrophysiological properties of mammalian neurons: insights into central nervous system function. Science 1988; 242: 1654-64.
Rudy B, Hoger JH, Lester HA, Davidson N. At least two mRNA
species contribute to the properties of rat brain A-type potassium
channels expressed in Xenopus oocytes. Neuron 1988; 1: 649-58.
Amberg GC, Koh SD, Imaizumi Y, Ohya S, Sanders KM. A-type
potassium currents in smooth muscle. Am J Physiol Cell Physiol
2003; 284: C583-95.
Wang SY, Yoshino M, Sui JL, Wakui M, Kao PN, Kao CY. Potassium currents in freshly dissociated uterine myocytes from nonpregnant and late-pregnant rats. J Gen Physiol 1998; 112: 737-56.
Wang Z, Fermini B, Nattel S. Rapid and slow components of delayed rectifier current in human atrial myocytes. Cardiovasc Res
1994; 28: 1540-6.
Rettig J, Heinemann SH, Wunder F, Lorra C, Parcej DN, Dolly JO,
et al. Inactivation properties of voltage-gated K+ channels altered
by presence of beta-subunit. Nature 1994; 369: 289-94.
Park KH, Hernandez L, Cai SQ, Wang Y, Sesti F. A Family of K+
Channel Ancillary Subunits Regulate Taste Sensitivity in
Caenorhabditis elegans. J Biol Chem 2005; 280: 21893-9.
Bianchi L, Kwok SM, Driscoll M, Sesti F. A potassium channelMiRP complex controls neurosensory function in Caenorhabditis
elegans. J Biol Chem 2003; 278: 12415-24.
Li Y, Um SY, McDonald TV. Voltage-gated potassium channels:
regulation by accessory subunits. Neuroscientist 2006; 12: 199210.
Jiang Y, Lee A, Chen J, Ruta V, Cadene M, Chait B, et al. X-ray
structure of a voltage-dependent K+ channel. Nature 2003; 423: 3341.
Long SB, Campbell EB, Mackinnon R. Crystal structure of a
mammalian voltage-dependent Shaker family K+ channel. Science
2005; 309: 897-903.
Doyle D, Morais Cabral J, Pfuetzner R, Kuo A, Gulbis J, Cohen S,
et al. The structure of the potassium channel: molecular bases for
K+ conduction and selectivity. Science 1998; 280: 69-77.
Multiple Modes of A-Type Potassium Current Regulation
[14]
[15]
[16]
[17]
[18]
[19]
[20]
[21]
[22]
[23]
[24]
[25]
Armstrong CM, Bezanilla F. Inactivation of the sodium channel. II.
Gating current experiments. J Gen Physiol 1977; 70: 567-90.
Hoshi T, Zagotta WN, Aldrich RW. Biophysical and molecular
mechanisms of Shaker potassium channel inactivation. Science
1990; 250: 533-8.
Zagotta WN, Hoshi T, Aldrich RW. Restoration of inactivation in
mutants of Shaker potassium channels by a peptide derived from
ShB. Science 1990; 250: 568-71.
Murrell-Lagnado RD, Aldrich RW. Interactions of amino terminal
domains of Shaker K channels with a pore blocking site studied
with synthetic peptides. J Gen Physiol 1993; 102: 949-75.
Murrell-Lagnado RD, Aldrich RW. Energetics of Shaker K channels block by inactivation peptides. J Gen Physiol 1993; 102: 9771003.
Zhou M, Morais-Cabral JH, Mann S, MacKinnon R. Potassium
channel receptor site for the inactivation gate and quaternary amine
inhibitors. Nature 2001; 411: 657-61.
Weng J, Cao Y, Moss N, Zhou M. Modulation of voltagedependent Shaker family potassium channels by an aldo-keto reductase. J Biol Chem 2006; 281: 15194-200.
Cai SQ, Hernandez L, Wang Y, Park KH, Sesti F. MPS-1 is a K(+)
channel beta-subunit and a serine/threonine kinase. Nat. Neurosci.
2005; 8(11): 1503-9.
Parcej DN, Scott VE, Dolly JO. Oligomeric properties of alphadendrotoxin-sensitive potassium ion channels purified from bovine
brain. Biochemistry 1992; 31: 11084-8.
Scott VE, Rettig J, Parcej DN, Keen JN, Findlay JB, Pongs O, et al.
Primary structure of a beta subunit of alpha-dendrotoxin-sensitive
K+ channels from bovine brain. Proc Natl Acad Sci USA 1994; 91:
1637-41.
Morales MJ, Castellino RC, Crews AL, Rasmusson RL, Strauss
HC. A novel beta subunit increases rate of inactivation of specific
voltage-gated potassium channel alpha subunits. J Biol Chem 1995;
270: 6272-7.
England SK, Uebele VN, Shear H, Kodali J, Bennett PB, Tamkun
MM. Characterization of a voltage-gated K+ channel beta subunit
expressed in human heart. Proc Natl Acad Sci USA 1995; 92:
6309-13.
Majumder K, De Biasi M, Wang Z, Wible BA. Molecular cloning
and functional expression of a novel potassium channel betasubunit from human atrium. FEBS Lett 1995; 361: 13-6.
Heinemann SH, Rettig J, Wunder F, Pongs O. Molecular and functional characterization of a rat brain Kv beta 3 potassium channel
subunit. FEBS Lett 1995; 377: 383-9.
Leicher T, Bahring R, Isbrandt D, Pongs O. Coexpression of the
KCNA3B gene product with Kv1.5 leads to a novel A-type potassium channel. J Biol Chem 1998; 273: 35095-101.
Roeper J, Sewing S, Zhang Y, Sommer T, Wanner SG, Pongs O.
NIP domain prevents N-type inactivation in voltage-gated potassium channels. Nature 1998; 391: 390-3.
McIntosh P, Southan AP, Akhtar S, Sidera C, Ushkaryov Y, Dolly
JO, et al. Modification of rat brain Kv1.4 channel gating by association with accessory Kvbeta1.1 and beta2.1 subunits. Pflugers
Arch 1997; 435: 43-54.
Accili EA, Kiehn J, Yang Q, Wang Z, Brown AM, Wible BA.
Separable Kvbeta subunit domains alter expression and gating of
potassium channels. J Biol Chem 1997; 272: 25824-31.
Campomanes CR, Carroll KI, Manganas LN, Hershberger ME,
Gong B, Antonucci DE, et al. Kv beta subunit oxidoreductase activity and Kv1 potassium channel trafficking. J Biol Chem 2002;
277: 8298-305.
Tipparaju SM, Saxena N, Liu SQ, Kumar R, Bhatnagar A. Differential regulation of voltage-gated K+ channels by oxidized and reduced pyridine nucleotide coenzymes. Am J Physiol Cell Physiol
2005; 288: C366-76.
Gulbis JM, Mann S, MacKinnon R. Structure of a voltagedependent K+ channel beta subunit. Cell 1999; 97: 943-52.
Gulbis JM, Zhou M, Mann S, MacKinnon R. Structure of the cytoplasmic beta subunit-T1 assembly of voltage-dependent K+ channels. Science 2000; 289: 123-7.
An WF, Bowlby MR, Betty M, Cao J, Ling HP, Mendoza G, et al.
Modulation of A-type potassium channels by a family of calcium
sensors. Nature 2000; 403: 553-6.
[27]
[28]
[29]
[30]
[31]
[32]
[33]
[34]
[35]
[36]
[37]
[38]
[39]
[40]
[41]
[42]
[43]
[44]
[45]
[46]
[47]
[48]
[49]
[50]
[51]
[52]
[53]
[54]
[55]
[56]
3183
Morohashi Y, Hatano N, Ohya S, Takikawa R, Watabiki T, Takasugi N, et al. Molecular cloning and characterization of
CALP/KChIP4, a novel EF-hand protein interacting with presenilin
2 and voltage-gated potassium channel subunit Kv4. J Biol Chem
2002; 277: 14965-75.
Takimoto K, Ren X. KChIPs (Kv channel-interacting proteins)--a
few surprises and another. J Physiol 2002; 545: 3.
Birnbaum SG, Varga AW, Yuan LL, Anderson AE, Sweatt JD,
Schrader LA. Structure and function of Kv4-family transient potassium channels. Physiol Rev 2004; 84: 803-33.
Rosati B, Pan Z, Lypen S, Wang HS, Cohen I, Dixon JE, et al.
Regulation of KChIP2 potassium channel beta subunit gene expression underlies the gradient of transient outward current in canine
and human ventricle. J Physiol 2001; 533: 119-25.
Kuo HC, Cheng CF, Clark RB, Lin JJ, Lin JL, Hoshijima M, et al.
A defect in the Kv channel-interacting protein 2 (KChIP2) gene
leads to a complete loss of I(to) and confers susceptibility to ventricular tachycardia. Cell 2001; 107: 801-13.
Shibata R, Misonou H, Campomanes CR, Anderson AE, Schrader
LA, Doliveira LC, et al. A fundamental role for KChIPs in determining the molecular properties and trafficking of Kv4.2 potassium
channels. J Biol Chem 2003; 278: 36445-54.
Takimoto K, Yang EK, Conforti L. Palmitoylation of KChIP splicing variants is required for efficient cell surface expression of
Kv4.3 channels. J Biol Chem 2002; 277: 26904-11.
O'Callaghan DW, Hasdemir B, Leighton M, Burgoyne RD. Residues within the myristoylation motif determine intracellular targeting of the neuronal Ca2+ sensor protein KChIP1 to post-ER transport vesicles and traffic of Kv4 K+ channels. J Cell Sci 2003; 116:
4833-45.
Boland LM, Jiang M, Lee SY, Fahrenkrug SC, Harnett MT,
O'Grady SM. Functional properties of a brain-specific NH2terminally spliced modulator of Kv4 channels. Am J Physiol Cell
Physiol 2003; 285: C161-70.
Van Hoorick D, Raes A, Keysers W, Mayeur E, Snyders DJ. Differential modulation of Kv4 kinetics by KCHIP1 splice variants.
Mol Cell Neurosci 2003; 24: 357-66.
Holmqvist MH, Cao J, Hernandez-Pineda R, Jacobson MD, Carroll
KI, Sung MA, et al. Elimination of fast inactivation in Kv4 A-type
potassium channels by an auxiliary subunit domain. Proc Natl Acad
Sci USA 2002; 99: 1035-40.
Cai SQ, Sesti F. A new mode of regulation of N-type inactivation
in a Caenorhabditis elegans voltage-gated potassium channel. J
Biol Chem 2007; 282: 18597-601.
Pioletti M, Findeisen F, Hura GL, Minor DL, Jr. Three-dimensional
structure of the KChIP1-Kv4.3 T1 complex reveals a cross-shaped
octamer. Nat Struct Mol Biol 2006; 13: 987-95.
Zhou W, Qian Y, Kunjilwar K, Pfaffinger PJ, Choe S. Structural
insights into the functional interaction of KChIP1 with Shal-type
K(+) channels. Neuron 2004; 41: 573-86.
Scannevin RH, Wang K, Jow F, Megules J, Kopsco DC, Edris W,
et al. Two N-terminal domains of Kv4 K(+) channels regulate
binding to and modulation by KChIP1. Neuron 2004; 41: 587-98.
Kim LA, Furst J, Butler MH, Xu S, Grigorieff N, Goldstein SA. Ito
channels are octomeric complexes with four subunits of each Kv4.2
and K+ channel-interacting protein 2. J Biol Chem 2004; 279:
5549-54.
Kim LA, Furst J, Gutierrez D, Butler MH, Xu S, Goldstein SA, et
al. Three-dimensional structure of I(to); Kv4.2-KChIP2 ion channels by electron microscopy at 21 Angstrom resolution. Neuron
2004; 41: 513-9.
Abbott G, Sesti F, Splawski I, Buck M, Lehman M, Timothy K, et
al. MiRP1 forms IKr potassium channels with HERG and is associated with cardiac arrhythmia. Cell 1999; 97: 175-87.
Anantharam A, Lewis A, Panaghie G, Gordon E, McCrossan ZA,
Lerner DJ, et al. RNA interference reveals that endogenous
xenopus MinK-related peptides govern mammalian K+ channel
function in oocyte expression studies. J Biol Chem 2003; 278:
11739-45.
Zhang M, Jiang M, Tseng G. minK-related peptide 1 associates
with Kv4.2 and modulates its gating function: potential role as beta
subunit of cardiac transient outward channel? Circ Res 2001; 88:
1012-9.
n
tio
u
rib
t
s
i
D
r
o
F
t
o
N
[26]
Current Pharmaceutical Design, 2007, Vol. 13, No. 31
3184 Current Pharmaceutical Design, 2007, Vol. 13, No. 31
[57]
[58]
[59]
[60]
[61]
[62]
[63]
[64]
[65]
[66]
[67]
Lewis A, McCrossan ZA, Abbott GW. MinK, MiRP1, and MiRP2
diversify Kv3.1 and Kv3.2 potassium channel gating. J Biol Chem
2004; 279: 7884-92.
McCrossan ZA, Lewis A, Panaghie G, Jordan PN, Christini DJ,
Lerner DJ, et al. MinK-related peptide 2 modulates Kv2.1 and
Kv3.1 potassium channels in mammalian brain. J Neurosci 2003;
23: 8077-91.
Grunnet M, Rasmussen HB, Hay-Schmidt A, Rosenstierne M,
Klaerke DA, Olesen SP, et al. KCNE4 is an inhibitory subunit to
Kv1.1 and Kv1.3 potassium channels. Biophys J 2003; 85: 152537.
Gordon E, Roepke TK, Abbott GW. Endogenous KCNE subunits
govern Kv2.1 K+ channel activation kinetics in Xenopus oocyte
studies. Biophys J 2006; 90: 1223-31.
Abbott G, Butler M, Bendahhou S, Dalaks M, Ptacek L, Goldstein
S. MiRP2 forms potassium channels in skeletal muscle with Kv3.4
and is associated with periodic paralysis. Cell 2001; 104: 217-31.
Nadal MS, Ozaita A, Amarillo Y, Vega-Saenz de Miera E, Ma Y,
Mo W, et al. The CD26-related dipeptidyl aminopeptidase-like protein DPPX is a critical component of neuronal A-type K+ channels.
Neuron 2003; 37: 449-61.
Strop P, Bankovich AJ, Hansen KC, Garcia KC, Brunger AT.
Structure of a human A-type potassium channel interacting protein
DPPX, a member of the dipeptidyl aminopeptidase family. J Mol
Biol 2004; 343: 1055-65.
Zagha E, Ozaita A, Chang SY, Nadal MS, Lin U, Saganich MJ, et
al. DPP10 modulates Kv4-mediated A-type potassium channels. J
Biol Chem 2005; 280: 18853-61.
Li HL, Qu YJ, Lu YC, Bondarenko VE, Wang S, Skerrett IM, et al.
DPP10 is an inactivation modulatory protein of Kv4.3 and Kv1.4.
Am J Physiol Cell Physiol 2006; 291: C966-76.
Antz C, Bauer T, Kalbacher H, Frank R, Covarrubias M, Kalbitzer
HR, et al. Control of K+ channel gating by protein phosphorylation: structural switches of the inactivation gate. Nat Struct Biol
1999; 6: 146-50.
Covarrubias M, Wei A, Salkoff L, Vyas TB. Elimination of rapid
potassium channel inactivation by phosphorylation of the inactivation gate. Neuron 1994; 13: 1403-12.
Drain P, Dubin AE, Aldrich RW. Regulation of Shaker K+ channel
inactivation gating by the cAMP-dependent protein kinase. Neuron
1994; 12: 1097-109.
Nitabach MN, Llamas DA, Thompson IJ, Collins KA, Holmes TC.
Phosphorylation-dependent
and
phosphorylation-independent
modes of modulation of shaker family voltage-gated potassium
channels by SRC family protein tyrosine kinases. J Neurosci 2002;
22: 7913-22.
Kwak YG, Hu N, Wei J, George AL, Jr., Grobaski TD, Tamkun
MM, et al. Protein kinase A phosphorylation alters Kvbeta1.3
subunit-mediated inactivation of the Kv1.5 potassium channel. J
Biol Chem 1999; 274: 13928-32.
Anderson AE, Adams JP, Qian Y, Cook RG, Pfaffinger PJ, Sweatt
JD. Kv4.2 phosphorylation by cyclic AMP-dependent protein
kinase. J Biol Chem 2000; 275: 5337-46.
Nakamura TY, Coetzee WA, Vega-Saenz De Miera E, Artman M,
Rudy B. Modulation of Kv4 channels, key components of rat ventricular transient outward K+ current, by PKC. Am J Physiol 1997;
273: H1775-86.
[69]
[70]
[71]
[72]
[73]
[74]
[75]
[76]
[77]
[78]
[79]
[80]
[81]
Adams JP, Anderson AE, Varga AW, Dineley KT, Cook RG,
Pfaffinger PJ, et al. The A-type potassium channel Kv4.2 is a substrate for the mitogen-activated protein kinase ERK. J Neurochem
2000; 75: 2277-87.
Schrader LA, Birnbaum SG, Nadin BM, Ren Y, Bui D, Anderson
AE, et al. ERK/MAPK regulates the Kv4.2 potassium channel by
direct phosphorylation of the pore-forming subunit. Am J Physiol
Cell Physiol 2006; 290: C852-61.
Schrader LA, Anderson AE, Mayne A, Pfaffinger PJ, Sweatt JD.
PKA modulation of Kv4.2-encoded A-type potassium channels requires formation of a supramolecular complex. J Neurosci 2002;
22: 10123-33.
Baltaev R, Strutz-Seebohm N, Korniychuk G, Myssina S, Lang F,
Seebohm G. Regulation of cardiac shal-related potassium channel
Kv 4.3 by serum- and glucocorticoid-inducible kinase isoforms in
Xenopus oocytes. Pflugers Arch 2005; 450: 26-33.
Ruppersberg JP, Stocker M, Pongs O, Heinemann SH, Frank R,
Koenen M. Regulation of fast inactivation of cloned mammalian
IK(A) channels by cysteine oxidation. Nature 1991; 352: 711-4.
Finol-Urdaneta RK, Struver N, Terlau H. Molecular and Functional
Differences between Heart mKv1.7 Channel Isoforms. J Gen
Physiol 2006; 128: 133-45.
Stephens GJ, Owen DG, Robertson B. Cysteine-modifying reagents
alter the gating of the rat cloned potassium channel Kv1.4. Pflugers
Arch 1996; 431: 435-42.
Duprat F, Guillemare E, Romey G, Fink M, Lesage F, Lazdunski
M, et al. Susceptibility of cloned K+ channels to reactive oxygen
species. Proc Natl Acad Sci USA 1995; 92: 11796-800.
Ciorba MA, Heinemann SH, Weissbach H, Brot N, Hoshi T. Regulation of voltage-dependent K+ channels by methionine oxidation:
effect of nitric oxide and vitamin C. FEBS Lett 1999; 442: 48-52.
Ciorba MA, Heinemann SH, Weissbach H, Brot N, Hoshi T.
Modulation of potassium channel function by methionine oxidation
and reduction. Proc Natl Acad Sci USA 1997; 94: 9932-7.
Santacruz-Toloza L, Huang Y, John SA, Papazian DM. Glycosylation of shaker potassium channel protein in insect cell culture and
in Xenopus oocytes. Biochemistry 1994; 33: 5607-13.
Oliver D, Lien CC, Soom M, Baukrowitz T, Jonas P, Fakler B.
Functional conversion between A-type and delayed rectifier K+
channels by membrane lipids. Science 2004; 304: 265-70.
Bargmann C. Neurobiology of the C. elegans genome. Science
1998; 282: 2028-33.
Fawcett GL, Santi CM, Butler A, Harris T, Covarrubias M, Salkoff
L. Mutant analysis of the Shal (Kv4) voltage-gated fast transient
K+ channel in Caenorhabditis elegans. J Biol Chem 2006; 281:
30725-35.
Christensen M, Estevez A, Yin X, Fox R, Morrison R, McDonnell
M, et al. A primary culture system for functional analysis of C.
elegans neurons and muscle cells. Neuron 2002; 33: 503-14.
Abbott GW. Molecular mechanisms of cardiac voltage-gated potassium channelopathies. Curr Pharm Des 2006; 12(28): 3631-44.
Panaghie G, Abbott GW. The impact of ancillary subunits on
small-molecule interactions with voltage-gated potassium channels.
Curr Pharm Des 2006; 12(18): 2285-302.
Panyi G, Possani LD, Rodriguez de la Vega RC, Gaspar R, Varga
Z. K+ channel blockers: novel tools to inhibit T cell activation
leading to specific immunosuppression. Curr Pharm Des 2006;
12(18): 2199-220.
[82]
[83]
[84]
[85]
[86]
[87]
[88]
[89]
[90]
n
tio
u
rib
t
s
i
D
r
o
F
t
o
N
[68]
Cai et al.
Download