ArticlePDF AvailableLiterature Review

Structural and functional analysis of the Na+/H+ exchanger

Authors:

Abstract and Figures

The mammalian NHE (Na+/H+ exchanger) is a ubiquitously expressed integral membrane protein that regulates intracellular pH by removing a proton in exchange for an extracellular sodium ion. Of the nine known isoforms of the mammalian NHEs, the first isoform discovered (NHE1) is the most thoroughly characterized. NHE1 is involved in numerous physiological processes in mammals, including regulation of intracellular pH, cell-volume control, cytoskeletal organization, heart disease and cancer. NHE comprises two domains: an N-terminal membrane domain that functions to transport ions, and a C-terminal cytoplasmic regulatory domain that regulates the activity and mediates cytoskeletal interactions. Although the exact mechanism of transport by NHE1 remains elusive, recent studies have identified amino acid residues that are important for NHE function. In addition, progress has been made regarding the elucidation of the structure of NHEs. Specifically, the structure of a single TM (transmembrane) segment from NHE1 has been solved, and the high-resolution structure of the bacterial Na+/H+ antiporter NhaA has recently been elucidated. In this review we discuss what is known about both functional and structural aspects of NHE1. We relate the known structural data for NHE1 to the NhaA structure, where TM IV of NHE1 shows surprising structural similarity with TM IV of NhaA, despite little primary sequence similarity. Further experiments that will be required to fully understand the mechanism of transport and regulation of the NHE1 protein are discussed.
Content may be subject to copyright.
Biochem. J. (2007) 401, 623–633 (Printed in Great Britain) doi:10.1042/BJ20061062 623
REVIEW ARTICLE
Structural and functional analysis of the Na+/H+exchanger
Emily R. SLEPKOV, Jan K. RAINEY, Brian D. SYKES and Larry FLIEGEL1
Department of Biochemistry, University of Alberta, Edmonton, Alberta, Canada T6G 2H7
The mammalian NHE (Na+/H+exchanger) is a ubiquitously ex-
pressed integral membrane protein that regulates intracellular pH
by removing a proton in exchange for an extracellular sodium
ion. Of the nine known isoforms of the mammalian NHEs, the
first isoform discovered (NHE1) is the most thoroughly character-
ized. NHE1 is involved in numerous physiological processes in
mammals, including regulation of intracellular pH, cell-volume
control, cytoskeletal organization, heart disease and cancer. NHE
comprises two domains: an N-terminal membrane domain that
functions to transport ions, and a C-terminal cytoplasmic regul-
atory domain that regulates the activity and mediates cytoskeletal
interactions. Although the exact mechanism of transport by NHE1
remains elusive, recent studies have identified amino acid residues
that are important for NHE function. In addition, progress has
been made regarding the elucidation of the structure of NHEs.
Specifically, the structure of a single TM (transmembrane) seg-
ment from NHE1 has been solved, and the high-resolution struc-
ture of the bacterial Na+/H+antiporter NhaA has recently been
elucidated. In this review we discuss what is known about both
functional and structural aspects of NHE1. We relate the known
structural data for NHE1 to the NhaA structure, where TM IV
of NHE1 shows surprising structural similarity with TM IV of
NhaA, despite little primary sequence similarity. Further experi-
ments that will be required to fully understand the mechanism of
transport and regulation of the NHE1 protein are discussed.
Key words: cation transport, intracellular pH, membrane protein,
Na+/H+exchanger (NHE), NhaA, structure–function analysis.
INTRODUCTION
The mammalian NHE (Na+/H+exchanger) is an integral mem-
brane protein that functions to exchange one intracellular proton
for one extracellular sodium ion. By its involvement in ion fluxes,
the NHE serves to regulate pHi(intracellular pH) and cell volume,
and to initiate changes in the growth or functional state of cells
[1]. Aside from its physiological role, the NHE serves important
roles in human pathology. Transport by the NHE plays a pivo-
tal role in the damage caused to the human myocardium during
and following a myocardial infarction [2], and it is considered to
represent a key step in the oncogenic transformation of cancerous
cells [3].
To date, the structure of the mammalian NHE remains elusive.
However, a great deal is known about many residues that are
required for the activity and regulation of the exchanger. In addi-
tion, the structure of a peptide of a single TM (transmembrane)
segment from the mammalian NHE has been published recently
[4]. This structure has some similarity to the structure of a func-
tionally important TM segment from the Escherichia coli Na+/H+
antiporter NhaA, suggesting that these two proteins may have
a similar structural architecture, although they share little se-
quence similarity.
MAMMALIAN NHE ISOFORMS
To date, nine isoforms (NHE1–NHE9) have been identified within
the mammalian NHE family [5,6]. The isoforms share approx.
25–70%amino acid identity, with calculated relative molecular
masses ranging from approx. 74000 to 93000 [5,7]. Hydropathy
analysis of the exchangers predicts that they have similar mem-
brane topologies, with an N-terminal membrane domain consist-
ing of 12 predicted TM segments and a more divergent C-terminal
cytoplasmic domain [5]. The NHE1 isoform is the ‘housekeeping’
isoform of the exchanger and is ubiquitously expressed in the
plasma membrane of virtually all tissues. It is the primary NHE
isoform found in the plasma membrane of the myocardium [2].
The NHE2–NHE5 isoforms are also localized to the plasma
membrane, but have more restricted tissue distributions. NHE2
and NHE3 are predominantly located in the apical membrane
of epithelia and are highly expressed in kidney and intestine
[8,9]. NHE4 is most abundant in stomach, but is also expressed
in intestine, kidney, brain, uterus and skeletal muscle [8]. NHE5
is expressed predominantly in brain, but may also be present
at low levels in other non-epithelial tissues, including spleen,
testis and skeletal muscle [10,11]. The isoforms NHE6–NHE9
are ubiquitously expressed and are present in intracellular com-
partments [6]. These organellar membrane NHEs are presumed
to regulate luminal pH and the cation concentration of the intra-
cellular compartments [6]. NHE6 expression is highest in heart,
brain and skeletal muscle and is localized to early recycling
endosomes [6,12]. The NHE7 isoform is localized predominantly
to the trans-Golgi network, and differs from the other NHE
isoforms in that it mediates the influx of either Na+or K+in
exchange for H+[13]. The highest levels of NHE8 expression
are found in skeletal muscle and kidney, and this isoform is
mainly localized to the mid- to trans-Golgi compartments [6].
The recently identified NHE9 isoform is localized to late recycling
endosomes [6].
NHE1 TOPOLOGY AND TRANSPORT PROPERTIES
The NHE1 isoform is the most well-characterized isoform of the
NHE family. As shown in Figure 1, NHE1 is 815 amino acids in
Abbreviations used: CAII, carbonic anhydrase II; CaM, calmodulin; CHP, calcineurin homologous protein; EL, extracellular loop; ERK1/2, extracellular-
signal-regulated kinase 1/2; ERM, ezrin, radixin and moesin; HSP70, heat-shock protein 70; IL, intracellular loop; MAPK, mitogen-activated protein kinase;
NHE, Na+/H+exchanger; pHi, intracellular pH; PIP2, phosphatidylinositol 4,5-bisphosphate; SERCA, sarcoplasmic/endoplasmic reticulum Ca2+-ATPase;
TM, transmembrane.
1To whom correspondence should be addressed (email lfliegel@ualberta.ca).
c
2007 Biochemical Society
624 E. R. Slepkov and others
Figure 1 Model of NHE1 showing the membrane and cytoplasmic domains
Upper panel: topology of the membrane domain which functions to transport cations. This
illustration is based on the findings of Wakabayashi et al. [90]. Pink circles, residues implicated
in both ion transport and inhibitor binding; orange circles, residues implicated in ion binding
and transport; yellow circles, residues implicated in inhibitor binding, green circles, residues
implicated in NHE1 folding and targeting to the plasma membrane. Lower panel: representation of
the cytoplasmic domain which functions to regulate the membrane domain through interactions
with signalling molecules.
length, with residues 1–500 comprising the membrane domain and
residues 501–815 comprising the cytoplasmic tail. The membrane
domain of NHE1 is both necessary and sufficient for ion transport,
whereas the cytosolic domain is involved in regulation of the activ-
ity of the exchanger [14]. Ion flux via the exchanger is driven by
the TM Na+gradient and requires no direct metabolic energy
input. NHE1 exhibits a simple Michaelis–Menten dependence on
extracellular Na+, with a reported apparent Kmof 5–50 mM
[15]. Extracellular Li+and H+compete with Na+for binding at
the Na+-binding site, and high extracellular concentrations of K+
inhibit NHE1 [16]. In contrast with the simple Michaelis–Menten
dependence on extracellular Na+, intracellular acidification allo-
sterically increases the activity of NHE, resulting in a rapid in-
crease in pHi.
NHEs are targets for inhibition by the diuretic compound
amiloride and its analogues, and by novel benzoylguanidine
derivatives [17]. Comparisons of the different NHE isoforms show
that they have varying affinities for these inhibitors, with the fol-
lowing order of sensitivity under similar experimental conditions:
NHE1 NHE2 >NHE5 >NHE3 >NHE4 [5,16]. Because NHE1
is the isoform that is most sensitive to inhibition, and is the only
isoform that is present in the plasma membrane of the myo-
cardium, the selective properties of these inhibitors can be ex-
ploited therapeutically.
PHYSIOLOGICAL FUNCTIONS OF NHE1
Na+/H+exchange is critical for a variety of physiological func-
tions. The mammalian NHE protects cells from intracellular
acidification, as shown by the fact that mutant cell lines devoid
of Na+/H+exchange activity are extremely sensitive to acidosis
[18,19]. Because NHE is activated by decreased pHi,when
acidosis occurs it increases NHE1 activity, resulting in the return
of pHito resting values. NHE also serves as a major Na+entry
pathway in many cell types, and as such it regulates both sodium
fluxes and cell volume after osmotic shrinkage [20–22]. Although
the mechanism by which NHE1 regulates cell volume remains
elusive, it is known that the volume- or osmolarity-sensitive site
within NHE1 is located in the N-terminal 566 amino acids of this
protein [23].
In addition to its role in regulating cellular pH and volume,
NHE also initiates shifts in pHithat stimulate changes in the
growth or functional state of cells [18,24]. NHE1 is required
for normal cell growth and proliferation, and cell proliferation is
significantly reduced in cells expressing an inactive NHE1 mutant
[25,26]. NHE1 activity is also important for cell differentiation.
Transcription of NHE1 increases during differentiation in both
human leukaemic cells and P19 embryonal carcinoma cells, and
the P19 cells that have NHE1 inhibited or deleted are markedly
deficient in their ability to differentiate [26–28]. NHE1 also plays
a role in either promoting or inhibiting apoptosis, with its role
varying depending on cell type [29–32].
NHE1 AS A STRUCTURAL ANCHOR
NHE1 acts as a structural anchor that is involved in regulating
cytoskeletal organization. NHE1 is restricted to specialized mem-
brane domains in some cells, such as in the lamellipodia of fibro-
blasts. There, NHE1 associates with the cytoskeleton via a di-
rect interaction with the actin-binding ERM (ezrin, radixin and
moesin) proteins at residues 553–564 in the C-terminal tail of
NHE1 [33,34]. Fibroblasts expressing NHE1 with mutations that
disrupt the interaction show impaired formation of focal adhesions
and actin stress fibres. The cells also have an irregular shape, and,
in wounding assays, migration of the cells is greatly impaired
[33,35,36]. NHE1 therefore has a key role in cell migration
through its interaction with the cytoskeleton and its restricted
location. These observations, plus the fact that NHE1 binds
several other proteins in the cytoplasmic regulatory domain, have
led to the hypothesis that NHE1 can act as a plasma membrane
scaffold that brings together many proteins so they can interact
functionally [37]. For example, in addition to its association
with ERM proteins and the cytoskeleton, NHE1 associates with
signalling molecules such as PIP2(phosphatidylinositol 4,5-
bisphosphate). PIP2binds in the same region as ERM proteins and
ERM proteins require PIP2for F-actin binding. It is hypothesized
that the proximity of PIP2on NHE1 facilitates ERM association
with F-actin [37]. However, further investigation is necessary
to confirm if the binding of NHE1 with a variety of proteins
facilitates their interaction in complicated processes.
NHE1 IN DISEASE
NHE1 has been implicated in the physiology of several diseases,
with the majority of research focusing on the role of NHE1
in heart disease and cancer. In the myocardium, under normal
conditions, NHE1 removes excess intracellular acid in exchange
for extracellular sodium. The increased intracellular sodium is
removed by regulatory membrane proteins, including the Na+/K+
ATPas e a n d t he Na+/Ca2+exchanger. Problems arise in the
myocardium with the increased production of protons that occurs
in the human myocardium during and following a myocardial
infarction [38]. The mechanism by which this occurs has been the
subject of much investigation. In the first part of this mechanism,
ischaemia results in increased anaerobic glycolysis, which leads
to a large increase in the production of protons. This serves to
c
2007 Biochemical Society
Structural and functional analysis of the Na+/H+exchanger 625
activate NHE1, which then exchanges intracellular protons for
extracellular sodium, leading to a rapid accumulation of sodium
in the cell [39–41]. Na+/K+ATPase function may be inhibited by
the decrease in high-energy phosphate stores that occurs during
ischaemia. The increased sodium concentration within the cell
drives an increase in calcium within the cell via reversal of the
Na+/Ca2+exchanger. This results in a buildup of calcium in the cell
that triggers various pathways, leading to cell death. It is
known that inhibiting NHE1 can have beneficial effects on the
myocardium during ischaemia and reperfusion, and the use of
NHE1 inhibitors to protect the heart against ischaemic damage has
been well established in animal studies [42–44]. The importance
of NHE1 in ischaemia–reperfusion injury is further supported by
the fact that genetic ablation of NHE1 in mice protects the heart
from ischaemia–reperfusion injury [45].
Despite the promising results generated in animal models,
results from clinical studies of NHE1 inhibition have not been
very positive. Two large studies have failed to find any significant
benefit of an NHE1 inhibitor. Any benefits were restricted to a
subset of patients who underwent coronary artery bypass graft
surgery. Additionally, a small study of 100 patients did reveal that
the NHE1 inhibitor cariporide reduced infarct size and improved
left ventricular function in post-infarction patients undergoing
angioplasty [46–48]. A more recent study [49] showed that cari-
poride administration caused a modest, but significant, reduction
in myocardial infarction after coronary artery bypass graft
surgery; however, in this study a concurrent increase in
cerebrovascular events occurred in high-risk patients, and this
risk outweighed the positive results.
It has long been known that NHE is important for tumour
growth, because tumour cells deficient in Na+/H+exchange activ-
ity either fail to grow tumours or show severely retarded growth
when implanted into immune-deprived mice [50]. It is now evi-
dent that NHE1 causes a reversal of the pH gradient in many types
of transformed and/or malignant cells so that the intracellular
environment is alkaline and the extracellular environment is
acidic [3]. This ‘malignant acidosis’ is considered to represent
a key step in oncogenic transformation and is necessary for
the development and maintenance of a transformed phenotype
[31,51,52]. Whereas NHE1 functions normally to regulate pH
in untransformed cell types, in at least some transformed cells
the tumour microenvironment aberrantly activates the protein.
For example, in breast cancer cells, serum deprivation is a
common tumour microenvironmental condition and this results
in abnormal activation of NHE1 in this cell type. The activation
of NHE1 was shown to be mediated by effects of RhoA and
Rac1, which were specific to tumour cells. In cancer cells, serum
deprivation provoked inhibition of RhoA activity in the leading
edge of pseudopodia, whereas in non-tumour cells this did not
occur [53,54]. This resulted in increased motility and invasion,
characteristics that are required for metastasis to occur [3,53]. A
recent study [55] has also demonstrated that the microenvironment
of breast tumours can activate NHE1 through CD44. This results
in acidification of the extracellular environment and promotes
breast cancer progression. Thus NHE1 inhibitors have a potential
use in treating various types of cancers. To date, NHE1 inhibitors
have been shown to induce apoptosis in leukaemic and breast
cancer cell lines [31,56]. The efficacy of NHE1 inhibitors for
treating cancer currently awaits preclinical and clinical trials.
REGULATION OF NHE1
The NHE1 isoform is highly regulated. Intracellular acidosis is the
major stimulus that regulates NHE1 activity, which is negligible
under normal physiological conditions, but is rapidly activated as
the pHidecreases [57]. This activation exhibits a Hill coefficient of
approx. 3, indicating that more than one proton binds to NHE1
during the transport cycle [58]. Thus it has been suggested that
NHE1 contains a non-transporting H+-binding site, sometimes
referred to as a ‘proton-modifier site’, which allows an allosteric
regulatory mechanism to lead to a greater increase in NHE
activity than would be expected based on pHi[58]. In addition to
responding to intracellular protons, NHE1 is regulated by phos-
phorylation by various kinases and by interactions with other
cellular proteins. NHE1 is also regulated at the transcriptional
level, allowing both mRNA levels and the amount of NHE1
protein produced to be controlled [59–62].
REGULATION BY PHOSPHORYLATION
The distal region of the C-terminal tail of NHE1, which comprises
amino acids 700–815, contains a number of serine and threonine
residues that are phosphorylated by protein kinases in response to
sustained acidosis or to hormone and growth-factor stimulation
[63–65]. Phosphorylation of residues in this region moves the set
point of the exchanger, shifting the exchange activity to be more
active at more alkaline pH values. Kinases that phosphorylate
NHE1 and stimulate its activity include: ERK1/2 (extracellular-
signal-regulated kinase 1/2), via the MAPK (mitogen-activated
protein kinase) cascade [66–68]; p90rsk (p90 ribosomal S6 kinase),
a downstream substrate of ERK1/2 [66,69,70]; the Rho-associated
kinase, p160ROCK [71]; NIK (Nck-interacting kinase) [72]; and
CaMKII (Ca2+/calmodulin-dependent kinase II) [73]. NHE1 is
also directly phosphorylated by p38 MAPK [32]. This kinase
inhibits NHE1 activity in response to angiotensin II via inhibition
of ERK1/2 [74], but it may also stimulate NHE1 and induce
alkalinization in an apoptotic pathway [32]. Protein kinases C
and D are also able to regulate the exchanger, but do not appear
to phosphorylate it directly [68,73,75,76]. In addition, NHE1 is
also regulated by dephosphorylation via protein phosphatase 1
[64,77].
INTERACTION WITH SIGNALLING MOLECULES
NHE activity is regulated by interaction with a variety of sig-
nalling molecules, and these interactions are shown in Figure 1
(lower panel). To date, three calcium-binding proteins have been
shown to interact with the exchanger: CaM (calmodulin) and CHP
(calcineurin homologous protein) act to stimulate NHE1, whereas
tescalcin inhibits NHE1. CaM binds to the cytoplasmic C-terminal
tail of NHE1 at two sites: a high-affinity site that is located at
amino acids 636–656 (CaM-A; Kd20 nM) and a low-affinity
site that is located at amino acids 657–700 (CaM-B; Kd350 nM)
[78]. The high-affinity site regulates NHE1 activity by functioning
as an auto-inhibitory domain, and either deletion of this site or
binding of Ca2+/CaM to this site abolishes the inhibitory effect
[79]. A second Ca2+-binding protein that interacts with NHE1
is CHP [80,81]. Endogenous CHP always contains two tightly
bound Ca2+ions when it is associated with NHE1, and it binds
to amino acids 515–530 in the C-terminal cytoplasmic domain
of NHE1 [80,82]. Finally, tescalcin binds to the final 180 amino
acids in the C-terminal tail of NHE1 and functions to inhibit the
activity of the exchanger [83,84].
Other signalling molecules that bind to NHE1 are: CAII
(carbonic anhydrase II), the adaptor protein 14-3-3, PIP2,andthe
mammalian HSP70 (heat-shock protein 70). CAII binds to amino
acids 790–802 at the distal end of the C-terminal tail of NHE1 and
increases the activity of the exchanger [85,86]. Phosphorylation
of NHE1 at a site within amino acids 634–789 causes an increased
interaction between NHE1 and CAII [85,86]. The binding of
c
2007 Biochemical Society
626 E. R. Slepkov and others
14-3-3 protein to NHE1 is also dependent on phosphorylation of
the exchanger, with binding occurring only when Ser703 is phos-
phorylated [87]. In Chinese-hamster fibroblasts, 14-3-3 protein
binding is thought to participate in serum-stimulated exchanger
activation by preventing dephosphorylation of Ser703 and by stabil-
izing an active conformation [87]. The signalling molecule PIP2
binds to NHE1 at two putative PIP2-binding motifs within the
C-terminal domain at residues 513–520 and 556–564, and deletion
of these binding motifs results in reduced transport activity in vivo
[88]. Binding of PIP2to NHE1 is required for optimal activity of
the exchanger, and this interaction, at least partially, accounts
for the ATP dependence of NHE1. Finally, HSP70 binds directly
to the C-terminal regulatory domain of NHE1, an interaction that
is probably involved in the folding and processing of the antiporter
[89].
STRUCTURE OF THE MEMBRANE DOMAIN
Relatively little is known about the structure of NHE1, because of
the inherent difficulties associated with crystallizing membrane
proteins. Thus molecular biology techniques have been used to
gain some understanding of the general structure of NHE1. For
example, the membrane topology of NHE1 shown in Figure 1
(upper panel) was determined experimentally by means of substi-
tuted-cysteine-accessibility analysis [90]. This analysis confirmed
predictions that NHE1 has 12 TM segments, with both the N- and
C-termini located in the cytosol. In addition, it identified three
membrane-associated loop regions, IL2 (intracellular loop 2), IL4
and EL5 (extracellular loop 5) which may be involved in NHE1
function.
The mature form of NHE1 is localized to the plasma membrane
and is glycosylated at both N- and O-linked sites. The N-linked
glycosylation is not necessary for Na+/H+exchange function
and biosynthesis [91,92]. In addition, it is known that NHE1
forms homodimers in intact cells, and that dimer formation is not
required for Na+/H+exchange activity [93–95]. Although it was
originally thought that the membrane domain of NHE1 alone is
sufficient to allow for dimerization, a recent study showed that
the proximal C-termini (amino acids 503–580) have a strong pro-
pensity to interact directly with each other, suggesting that the two
C-termini of the NHE1 dimer may also interact with each other
[94,95]. Finally, structural information has been deduced about
the C-terminal cytoplasmic domain of NHE1 using CD spectro-
scopy. This method revealed that the cytoplasmic tail of NHE1
is 35%α-helix, 17%β-turn and 48%random coil, and that
the structure of the cytoplasmic tail is more compact at regions
proximal to the membrane domain, whereas regions distal to the
membrane domain are more flexible and display calcium-depend-
ent conformational changes [83,96].
We recently published the first high-resolution structure of a TM
segment of the human Na+/H+exchanger [4]. A TM IV peptide
was expressed and purified, and its structure was determined in
a membrane-mimetic environment (Figure 2). From this NMR
structure it is clear that TM IV does not, as a whole, assume a
single conformation. Rather, sections within TM IV have distinct
structural characteristics. Specifically, TM IV is composed of
three sections of four to nine residues that converge structurally
and only one region is α-helical: Asp159–Leu163 form a series of
β-turns; Leu165–Pro168 has an extended structure; and Ile169 –Pro176
form an α-helix. As the structure of the peptide was determined
in a membrane-mimetic solvent in the absence of the balance
of the protein, these three structured regions rotated quite freely
with respect to each other at swivel points that are located at
Phe164 and Pro168/Ile169 (Figure 2). It must be presumed that in
the entire NHE1 protein the interaction of TM IV with other TM
Figure 2 Structure of a TM IV peptide in a membrane -mimetic environment
Convergent stretches (grey) of residues Asp159–Leu163 ,Leu
165–Pro168 , and Ile169–Phe176 in
relation to pivot residues Phe176 and Pro168/Ile169 are shown [137]. The flexible N- and C-termini
are represented by dashed lines. Note that only the proline side chains are indicated. Reproduced
from [4] with permission.
segments would restrict the rotation about these swivel points.
Aside from this TM segment, no other information has been pub-
lished regarding the structure of NHE1. However, in contrast with
results from TM IV, we have recently determined TM VII is a
kinked α-helix [96b].
RESIDUES INVOLVED IN NHE1 FUNCTION
Although the exact mechanisms of transport and inhibitor binding
by NHE1 are not known, specific residues within the membrane
domain of NHE1 have been implicated as being important for ion
binding and transport. The location of these residues is highlighted
in Figure 1 (upper panel), and Table 1 summarizes the effects of
these mutations.
TM IV
Numerous residues in TM IV have been implicated in NHE1
function. For example, a Phe165 Tyr mutation in TM IV of
the hamster NHE1 sequence (corresponding to human Phe161)
causes both an increase in resistance to inhibitors and a decrease
in Vmax for Na+[97]. In addition, a Leu167 Phe mutation
(corresponding to human Leu163) causes increased inhibitor resi-
stance with no effect on Na+transport. In 1997, Counillon et al.
[98] used random mutagenesis and found that a Gly174 Ser
mutation in TM IV causes a modest increase in resistance to
amiloride with no effect on Na+transport. They also made an
NHE1 mutant with a Leu163 Phe/Gly174 Ser double mutation,
and this mutant possessed a strongly reduced affinity for HOE
694 [(3-methylsulfonyl-4-piperidinobenzoyl)guanidine methane-
sulfonate] and a 2-fold decrease in sodium affinity. A Phe162 Ser
substitution in TM IV has been found to cause a dramatic decrease
in affinity for cariporide and a 10-fold decrease in Na+affinity
[99].
c
2007 Biochemical Society
Structural and functional analysis of the Na+/H+exchanger 627
Table 1 Amino acids known to be involved in NHE1 structure and function
Summary of residues in the membrane domain of NHE1 required for ion transport, inhibitor binding and/or expression and targeting of the exchanger. Amino acid numbering corresponds to the human
NHE1. For mutants that were studied in other mammalian models, the species used and the corresponding amino acid number are indicated in parentheses. The location of the residue within the topology
of the exchanger, the specific mutations studied and the effects of the mutations are indicated. EIPA, 5-(
N
-ethyl)-
N
-isopropylamiloride; HOE 694, (3-methylsulfonyl-4-piperidinobenzoyl)guanidine
methanesulfonate; MPA, 5-(
N
-methyl)-
N
-propylamiloride; MTSET, 2-(trimethylammonium)ethyl methanethiosulfonate bromide.
Amino acid Location Mutation Effect of mutation Reference
Gly148 (rat: Gly152) EL2 Gly Ala Increase in
K
ifor EIPA [7]
Pro153/Pro154 (rat: Pro157 /Pro158) EL2 Pro Ser/Pro Phe Increase in
K
ifor EIPA [7]
Decrease in transport
Phe161 (hamster: Phe165)TMIVPheTyr Increase in
K
ifor amiloride and MPA [97]
Decrease in transport
Phe161 TM IV Phe Cys Lines ion-transport pore [4]
Phe162 TM IV Phe Ser 1500-fold increase in
K
ifor cariporide [99]
Increase in
K
mfor Na+
Leu163 (hamster: Leu167)TMIVLeuPhe, Ala, Arg, Trp Increase in
K
ifor amiloride, MPA and HOE 694 [97]
Leu Tyr Eliminates Na+/H+transport
Pro167 TM IV Pro Gly, Ala, Cys Decreased Na+/H+transport, expression [100]
and plasma membrane targeting
Pro168 TM IV Pro Gly, Ala, Cys Decreased Na+/H+transport [100]
Gly174 TM IV Gly Ser,Asp Increase in
K
ifor amiloride and HOE 694 [98]
Leu163/Gly174 TM IV Leu Phe/Gly Ser Increase in
K
ifor HOE 694 [98]
Increase in
K
mfor Na+
Arg180 IL2 Arg Cys MTSET treatment decreases activity [90]
Gln181 IL2 Gln Cys MTSET treatment decreases activity [90]
Glu262 TM VII Glu Gln Eliminates Na+/H+transport [102]
Glu Asp Restores partial Na+/H+transport
Increases
K
mfor Li+
Asp267 TM VII Asp Asn Eliminates Na+/H+transport [102]
Asp Glu Restores Na+/H+transport
His349 TM IX His Gly, Leu Increase in
K
ifor amiloride [103]
His Tyr, Phe Decrease in
K
ifor amiloride
Glu346 (rat: Glu350)TMIXGluAsp Increase in
K
ifor EIPA and HOE 694 [7,105]
Decrease in transport
Increase in
K
mfor Na+
Glu Asn, Gln Increase in
K
ifor EIPA
Decrease in transport
Gly352 (rat: Gly356)TMIXGlyAla, Ser, Asp Increase in
K
ifor EIPA [7]
Decrease in transport
Glu391 EL5 Glu Gln Decreased Na+/H+transport [102]
Glu Asp Restores Na+/H+transport
Arg440 IL5 Arg Cys, Lys, His, Asp, Glu, Leu Shifts pHidependence to acidic side [107]
Tyr454 TM XI Tyr Cys Retained in endoplasmic reticulum [106]
Gly455 TM XI Gly Cys, Gln, Thr, Val Shift pHidependence to alkaline side [107]
Gly Ala, Asp, Asn, Ser No effect on pHidependence
Gly456 TM XI Gly Cys Shifts pHidependence to alkaline side [107]
Arg458 TM XI Tyr Cys Retained in endoplasmic reticulum [106]
We have been investigating the importance of residues in TM
IV of NHE1 for several years, and we have found numerous resi-
dues in this TM segment that are required for normal Na+/H+
exchange activity. For example, we found that both Pro167 and
Pro168 are required for normal NHE1 activity, whereas Pro178 is
not [100]. In addition, mutation of Pro167 affects the expression
and membrane targeting of NHE1 [100]. Thus both Pro167 and
Pro168 are critical for normal NHE1 function, and may be required
to directly interact with transported cations, or to allow TM IV to
assume a unique structure or undergo a conformational change
that is required for NHE1 activity. Mutation of these prolines
may also affect the structure or folding of NHE1 that may cause
aberrant targeting of the protein.
We also used the substituted-cysteine-accessibility method
[101] to examine both functional and structural aspects of TM
IV in NHE1. We found that TM IV is exceptionally sensitive to
mutation, with each of the 23 single-cysteine mutations resulting
in a significant decrease in exchanger activity [4]. In addition, we
examined the sensitivity of the active single-cysteine mutants to
thiol-reactive reagents and found that the mutant Phe161 Cys
was significantly inhibited by this treatment [4]. Therefore, in
addition to being extremely sensitive to mutation, TM IV, and
specifically Phe161, line the ion-transport pore of NHE1.
The loop regions at either end of TM IV also contain residues
that are important for NHE1 function. Mutation of three residues
in the second exomembrane loop at the N-terminal end of TM
IV (EL2) affects both the drug sensitivity and the activity of the
exchanger [7]. IL2, at the C-terminal end of TM IV, also contains
residues that may line the ion-transport pore, because treating the
mutants Arg180 Cys or Gln181 Cys at IL2 with membrane-
impermeant thiol reagents severely inhibits transport [90].
TM VII
TM VII is clearly involved in the ion-binding and transport
capabilities of NHE1, because both Glu262 and Asp267 are essential
for NHE1 activity [102]. Both of the mutations Glu262 Gln and
Asp267 Asn abolished Na+/H+exchange activity, whereas the
c
2007 Biochemical Society
628 E. R. Slepkov and others
conservative mutations Glu262 Asp and Asp267 Glu, which
retain the acidic side chain, restored Na+/H+exchange activity.
In addition, the mutant Glu262 Asp had a lower affinity for Li+
than the wild-type exchanger. Because Li+has a smaller ionic
radius than Na+, this decreased affinity may be due to the shorter
side-chain length of the aspartate residue reducing the ability of
the exchanger to co-ordinate the smaller Li+ion. Thus TM VII is
involved in the cation-exchange mechanism of NHE1, and Glu262
is a likely target for directly interacting with transported cations.
TM IX
The importance of TM IX for NHE1 activity was first demon-
strated when it was determined that mutating His349 affects the
amiloride sensitivity of NHE1 [103]. Moreover, chimaeric NHE
proteins made by interchanging a 66-amino-acid segment con-
taining TM IX and its adjacent loops from NHE1 and NHE3
displayed reciprocal alterations in their sensitivities to inhibitors,
but retained normal Na+transport properties [104]. Recently, two
residues within TM IX of rat NHE1, Glu350 and Gly356 (corres-
ponding to Glu346 and Gly352 in human NHE1), were identified as
major determinants of drug sensitivity [7]. In addition, mutation
of Glu346 also affects the Na+affinity of NHE1 [105].
TM XI
A number of residues in TM XI may be involved in either ion
transport or proper targeting to the plasma membrane, as shown
by the fact that mutation of these residues alters NHE1 function
[106]. Specifically, two mutants in TM XI, Tyr454 Cys and
Arg458 Cys, are retained in the endoplasmic reticulum [106]. In
addition, the mutations Gly455 Cys and Gly456 Cys in TM XI
shift the pHidependence of the exchanger to a more alkaline value
(pK7), whereas the mutation Arg440 Cys in IL5 at the N-
terminal end of TM XI shifts the pHidependence to a more acidic
value (pK<6.2) [107]. The shift in pKobserved with mutation
of Gly455 was dependent on the size of the substituted side-chain
residue, implying that these mutations may perturb the structure
of TM XI, thereby indirectly affecting the H+-modifier site. In the
case of Arg440, mutation to a lysine residue had a more modest
effect on the pKthan other mutations, indicating that a positive
charge at this site may be important for normal pHisensitivity.
Further experiments measuring 22Na+efflux from cells in which
the Na+/H+exchanger was functioning in the reverse mode also
support the conclusion that both IL5 and TM XI play a crucial
role in the proper functioning of the H+-modifier site [107].
Other proximal membrane regions involved in NHE1 function
The first extracellular loop (EL1), the membrane-associated seg-
ment (EL5) and the proximal region of the cytoplasmic domain
all contain residues that are involved in NHE1 function. EL1 is
involved in the differences in volume sensitivity between isoforms
of the Na+/H+exchanger. Specifically, amino acids 41–53 are res-
ponsible for inhibiting hyperosmolarity-induced activation of
NHE2, resulting in the lack of RVI (regulatory volume increase)
that is evident in this isoform [108]. However, the N-terminal
region (amino acids 1–95) of NHE1 do not appear to be involved
in the volume-sensing mechanism of this isoform [108].
The possible functional involvement of the membrane-asso-
ciated segment within NHE1 is of interest, because this segment
contains several polar amino acids and is reminiscent of the
selectivity filter of potassium channels [102]. It has been demon-
strated that Glu391, located in the membrane-associated segment,
Figure 3 Structure of the
E. coli
Na+/H+antiporter NhaA
Ribbon representation of NhaA viewed in parallel with the membrane. The 12 TM segments
are labelled with Roman numerals. TM segments IV and XI have a helix-extended chain–helix
conformation. The cytoplasmic (upper) and periplasmic (lower) faces of the membrane are
indicated by broken lines. TM segments are colour coded as follows: I, pink; II, cyan; III, blue;
IV, red; V, grey; VI, green; VII, yellow; VIII, orange; IX, green; X, pale yellow; XI, brown; XII cyan.
This Figure is adapted from [112] with permission from
Nature
c
2005 Macmillan Magazines
Ltd. (http://www.nature.com/).
is important for activity [102]. A Glu391 Gln mutation resulted
in a partial reduction in activity, and a Glu391 Asp mutation,
an alternative acidic residue, restored Na+/H+exchanger activity.
Thus the membrane-associated segment plays a role in the ion-
binding and transport properties of NHE1.
A highly conserved histidine-rich sequence of amino acids in
the proximal region of the cytoplasmic domain, H540YGHHH545,
is also involved in NHE1 function. Mutation of this sequence to
H540HHHHH545 has no effect on the activation of NHE1 by pro-
tons, but did cause a decrease in the maximal velocity of the ex-
changer [109]. Thus this conserved sequence is involved in NHE1
function, but is not involved in proton sensing.
STRUCTURE OF PROKARYOTIC Na+/H+ANTIPORTERS
NHEs are ubiquitous throughout the Animal Kingdom. In bac-
teria, Na+/H+exchange serves a role in osmotic regulation and
removes an internal sodium ion in exchange for external H+.
E. coli has two antiporters, NhaA and NhaB. They exchange
Na+or Li+for H+. NhaA is indispensable for transport and is
electrogenic with a stoichiometry of 2H+/Na+[110]. It is the
most well studied of the prokaryotic Na+/H+exchangers and
has been well characterized by site-specific mutagenesis, and has
also been overexpressed, purified and crystallized. Despite their
similarity in function with the mammalian NHEs, NhaA, NhaB
and other Na+/H+exchangers, such as those found in yeast, do
not share a large amount of similarity in their primary sequence.
It may be that some critical amino acids involved in transport
are conserved in their position and in their function in cation
co-ordination [111]; however, this has yet to be determined.
Crystal structure of the Na+/H+antiporter from
E. coli
Recently, the crystal structure of the Na+/H+antiporter, NhaA,
from E. coli has been solved. Although NhaA shares little
sequence homology with NHE1, these proteins share a similar
basic topology with 12 membrane-spanning segments and both
the C- and N-termini in the cytoplasm. Figure 3 shows the 3.45 Å
c
2007 Biochemical Society
Structural and functional analysis of the Na+/H+exchanger 629
Figure 4 Proposed mechanism of transport by the
E. coli
Na+/H+antiporter NhaA
The TM IV–TM XI assembly and its interaction with TM IX is shown. (A) Acidic pH-locked conformation. TM IX is bent, and the conformation of the TM IV–TM XI assembly only partly exposes the
Na+-binding site. (B) Alkaline pH causes a conformational change in helix IX that results in a reorientation of the TM IV–TM XI assembly. This exposes the Na+-binding site (yellow circle) to
the cytoplasmic funnel (red broken lines and red circle) and blocks it from the periplasm (orange line). (C)Na
+binding causes the cation-loaded binding site to be exposed to the periplasm. Upon
release of the cation, key aspartic acid residues are protonated, shifting NhaA back into the cytoplasm-exposed conformation in (B). This Figure is adapted from [112] with permission from
Nature
c
2005 Macmillan Magazines Ltd. (http://www.nature.com/).
resolution structure of NhaA in an acid-locked conformation. This
structure reveals a negatively charged funnel that opens to the
cytoplasm and ends in the middle of the membrane at the putative
ion-binding site [112]. Of the 12 TM segments, ten span the
bilayer as α-helices, whereas two (TMs IV and XI) are composed
of a short helix, an extended polypeptide chain and a short helix
(Figure 3). It should also be noted that, despite weak sequence
similarity, TM segments III, IV and V form a bundle with strong
structural similarity to the bundle formed by TM segments X, XI
and XII. These bundles are in the opposite orientation relative to
the membrane and bring TM segments IV and XI into close prox-
imity at the ion-binding site while providing a balanced electro-
static environment. Extended (also referred to as ‘unwound’)
segments in pairs of TM segments at active sites have also been
observed in the SERCA1a (sarcoplasmic/endoplasmic reticulum
Ca2+-ATPase 1a); [113] and in a bacterial homologue of Na+/Cl-
dependent neurotransmitter transporters (LeuTAA) [114]. In each
case, the extended portions of otherwise helical TM segments
are in close proximity to each other and form ion- and substrate-
binding sites.
The mechanism of NhaA regulation is thought to be dependent
on TM IX, which is known to contribute to the ‘pH sensor’ and
undergo a pH-induced conformational change. In the structure,
TM IX is a distorted helix that is in contact with the TM IV–TM
XI assembly at the centre of the membrane. Thus it is thought that
at alkaline pH the conformation of TM IX changes, resulting in
a reorientation of the TM IV–TM XI assembly that would fully
expose the Na+-binding site to the cytoplasm. Subsequent binding
of Na+would then trigger another small movement of the TM
IV–TM XI assembly, thereby exposing the cation-loaded binding
site to the periplasm. Figure 4 shows this proposed mechanism
of pH regulation and ion translocation. This mechanism, which
requires only a small conformational change, is consistent with
the extremely high catalytic activity of NhaA.
STRUCTURAL SIMILARITIES BETWEEN NHE1 AND NHAA
It has been proposed that the three-dimensional architecture of
NhaA may be similar to that of the mammalian NHEs, despite the
fact that these proteins share little sequence similarity in align-
ments of primary structure and have different transport stoi-
chiometry [111,115,116]. In a number of membrane proteins,
intramolecular structural homology has been observed within
three-dimensional structures for regions having limited primary
sequence similarity [112,114,117–122], including the case of
NhaA, as described above [112,119,120]. Despite differences in
transport mechanism, such as differing stoichiometery, the dis-
tantly related proteins NHE1 and NhaA do share a number of simi-
lar characteristics. In each of these proteins the pH regulatory site
is different from the active site, a loop participates in the pH
response and each protein exists as an oligomer within the mem-
brane [95,115,123,124].
In the NMR structure of TM IV from NHE1 (Figure 2), an ex-
tended structure in the middle of the segment can be observed
which is reminiscent of those seen in the ion-/substrate-binding
sites of NhaA, SERCA1a [113] and LeuTAA [114]. Although it
is certainly a possibility that an isolated TM peptide in mem-
brane-mimetic conditions will assume a non-physiologically rel-
evant structure, a number of isolated TM segments under such
conditions have been shown to be both functional and properly
structured where a structure of the full-length protein exists [125–
132]. Beyond the general motif of paired TM segments in close
proximity with extended regions interrupting structure, direct
comparison of the crystal structures of NhaA, SERCA1a and
LeuTAA shows that neither TM IV or TM XI of NhaA has strong
structural similarity with M4 or M6 of SERCA1a or to TM I
or TM VI of LeuTAA (results not shown). At best, a five-residue
stretch of M4 of the extended portion of SERCA1a has 0.98 Å
CαRMSD (root mean square deviation) overlap with TM IV
of NhaA. Comparison of TM IV of NHE1 with TM IV and
TM XI of NhaA demonstrates strong structural similarity, parti-
cularly between the pair of TM IV segments.
Specifically, when aligned as shown in Figure 5, 14 residues
of the TM IV segments of each exchanger are structurally
homologous (Figure 5; see also detailed superimpositions in
the Supplementary Figures, http://www.BiochemJ.org/bj/401/
bj4010623add.htm). Most notably, the extended segment in NHE1
at residues 165–168 superimposes extremely well on to the
c
2007 Biochemical Society
630 E. R. Slepkov and others
Figure 5 Structural similarity between TM IV segments of NHE1 and NhaA
Upper panel: sequence alignment of TM IVs of NHE1 and NhaA with arbitrary coloration. Lower
panel: representative NHE1 TM IV structure [4] (arbitrary orientation at swivel point Phe164)
shown alongside the TM IV segment of NhaA [112], with colouring as indicated in the sequence
alignment. Despite little sequence similarity, alignment of the TM IV segments of NHE1 and
NhaA over the residues illustrated allows structural superimposition at the 14 pairs of residues
indicated by arrows. Differences in structure between the Leu163–Phe164 swivel point of NHE1
and the crystal structure of inactive NhaA at Ile128–Pro129 mean that the entire segment does
not superimpose well. However, Asp159–Phe162 of NHE1 shows extremely similar structure to
Ile121–Trp126 of NhaA, and a subset of the NMR structures with the appropriate Pro168/Ile169 swivel
point orientation gives excellent superimposition of Leu165–Gly174 of NHE1 on Ale130 –Gly139 of
NhaA.
extended region of NhaA at residues 130–133. There would
be little chance of predicting this structural similarity based on
primary sequence, since in NHE1 this region has the sequence
Leu-Leu-Pro-Pro versus Ala-Ala-Thr-Asp in NhaA. As would be
expected, the helical segment 169–174 of NHE1 shows a good
superimposition on to the helical segment TM IVc of NhaA, but
the flexibility we observed in the NMR structures at the swivel
point between Pro168 and Ile169 (Figure 2) means that only a subset
of the TM IV structures of NHE1 superimposes well over both the
extended segment and the C-terminal helical region IVc of NhaA.
An example of another structure showing good superimposition
over this entire region is presented in Figure 5. Finally, the turn
structure observed over Asp159–Phe162 of NHE1 superimposes
exceptionally well on to Glu124–Ala127 of NhaA. This covers the
C-terminus of the TM IVp helix, which extends from Ile121–Trp126
in the X-ray structure of the pH-inactivated form of NhaA, and the
initiation of the β-bend observed at Ala127–Ala130 .TMIVofNHE1
also shows a lesser structural similarity to TM XI of NhaA (results
not shown), which would be expected given the quasi-symmetrical
symmetry between TM IV and XI in the NhaA crystal structure
[112,119,120]. Finally, TM IV of NHE1 does not superimpose
well on to the M4 segment of SERCA1a, even though the M4
segment shows some structural similarity to TM IV of NhaA.
Therefore, despite extremely low amino acid sequence similarity,
theNMRstructureofTMIVofNHE1overresiduesAsp
159–Gly174
shows strong structural similarity to the TM IV segment of NhaA
over residues Glu124 –Gly139, with the exception of the swivel point
region Leu163–Phe164 .
Because biochemical studies have shown that TM IV in NHE1
lines the ion-transport pore and contains numerous residues that
are important for NHE1 function [4], it is reasonable to assume
that this TM segment is likely to play a central role in the mech-
anism of NHE1. As shown in Figure 4, conformational change of
TM IV in NhaA about the extended non-helical region is proposed
to play a major role in exposing the Na+-binding site in the
active versus inactive protein. Given the structural similarity of
the extended segment of TM IV in NHE1, alongside its inherent
flexibility, as exhibited by the pair of swivel points immediately
N- and C-terminal to the extended segment, it is very possible that
TM IV of NHE1 is involved in a similar mechanistically important
conformational change. Furthermore, despite an apparent lack of
sequence similarity, it is possible that further structural similarity
exists and that these proteins share a similar tertiary fold. However,
a definitive analysis of the comparative structure of NhaA and
NHE1 will await the resolution of the structure of the entire
membrane domain of NHE1.
CONCLUSIONS AND FUTURE DIRECTIONS
In recent years, there have been many advances made in our under-
standing of the function and regulation of the mammalian NHE.
Several regions of the exchanger have been shown to be involved
in ion-exchange activity, inhibitor binding, interaction with sig-
nalling molecules and regulation by phosphorylation. However,
many unanswered questions still remain. For example, although
it is evident that TM IV lines the ion-transport pore of NHE1
and contains many residues that are important for NHE1 function
and inhibitor binding, less is known about other TM segments.
Thus further mutagenesis studies are required to identify pore-
lining and functionally important residues in other TM segments
of NHE1. In addition, once further pore-lining residues are
identified in TM segments other than TM IV, site-directed chemi-
cal cross-linking experiments may be of use to elucidate structural
information about NHE1. This technique has been used to probe
the three-dimensional structures of many polytopic membrane
proteins and can be used to develop a model for the arrangement
of TM segments [133–136]. Finally, although some structural
information can be deduced about NHE1 based on the NMR
structure of TM IV and similarities between NHE1 and NhaA,
a major goal in the field of NHE research is the elucidation of
high-resolution structural information about NHE1. One way to
accomplish this is to determine the structure of TM segment
peptides. It may also be possible to express and purify larger sec-
tions of the NHE1 protein that encompass several TM segments.
The structures of larger sections of the NHE1 protein would
be especially useful for understanding results from biochemical
studies, because interactions between the TM segments may limit
rotation within the TM segments, allowing the segments to adopt
the conformation that they have in full-length NHE1. Finally,
we await the structure of full-length NHE1, which will allow us
to understand at a molecular level how this protein binds and
transports cations and interacts with inhibitors.
E.R.S. was supported by an AHFMR (Alberta Heritage Foundation for Medical Research)
Studentship. J.K.R. is supported by a Fellowship from the CIHR (Canadian Institutes of
Health Research) Strategic Training Program in Membrane Proteins and Cardiovascular
Disease. B. D.S. receives support as a Canada Research Chair in Structural Biology. L. F. is
supported by an AHFMR Senior Scientist Award. Research by L. F. in this area is supported
by the CIHR.
c
2007 Biochemical Society
Structural and functional analysis of the Na+/H+exchanger 631
REFERENCES
1 Orlowski, J. and Grinstein, S. (1997) Na+/H+exchangers of mammalian cells.
J. Biol. Chem. 272, 22373–22376
2 Karmazyn, M., Gan, T., Humphreys, R. A., Yoshida, H. and Kusumoto, K. (1999) The
myocardial Na+–H+exchange. Structure, regulation, and its role in heart disease.
Circ. Res. 85, 777–786
3 Cardone, R. A., Casavola, V. and Reshkin, S. J. (2005) The role of disturbed pH
dynamics and the Na+/H+exchanger in metastasis. Nat. Rev. Cancer 5, 786–795
4 Slepkov, E. R., Rainey, J. K., Li, X., Liu, Y., Cheng, F. J., Lindhout, D. A., Sykes, B. D. and
Fliegel, L. (2005) Structural and functional characterization of transmembrane segment
IV of the NHE1 isoform of the Na+/H+exchanger. J. Biol. Chem. 280, 17863–17872
5 Orlowski, J. and Grinstein, S. (2004) Diversity of the mammalian sodium/proton
exchanger SLC9 gene family. P߬
ugers Arch. 447, 549–565
6 Nakamura, N., Tanaka, S., Teko, Y., Mitsui, K. and Kanazawa, H. (2005) Four Na+/H+
exchanger isoforms are distributed to Golgi and post-Golgi compartments and are
involved in organelle pH regulation. J. Biol. Chem. 280, 1561–1572
7 Khadilkar, A., Iannuzzi, P. and Orlowski, J. (2001) Identification of sites in the second
exomembrane loop and ninth transmembrane helix of the mammalian Na+/H+
exchanger important for drug recognition and cation translocation. J. Biol. Chem. 276,
43792–43800
8 Orlowski, J., Kandasamy, R. A. and Shull, G. E. (1992) Molecular cloning of putative
members of the Na+/H+exchanger gene family. J. Biol. Chem. 267, 9331–9339
9 Noel, J., Roux, D. and Pouyssegur, J. (1996) Differential localization of Na+/H+
exchanger isoforms (NHE1 and NHE3) in polarized epithelial cell lines. J. Cell Sci. 109,
929–939
10 Baird, N. R., Orlowski, J., Szabo, E. Z., Zaun, H. C., Schultheis, P. J., Menon, A. G. and
Shull, G. E. (1999) Molecular cloning, genomic organization, and functional expression
of Na+/H+exchanger isoform 5 (NHE5) from human brain. J. Biol. Chem. 274,
4377–4382
11 Attaphitaya, S., Park, K. and Melvin, J. E. (1999) Molecular cloning and functional
expression of a rat Na+/H+exchanger (NHE5) highly expressed in brain. J. Biol. Chem.
274, 4383–4388
12 Brett, C. L., Wei, Y., Donowitz, M. and Rao, R. (2002) Human Na+/H+exchanger
isoform 6 is found in recycling endosomes of cells, not in mitochondria. Am. J. Physiol.
Cell Physiol. 282, C1031–C1041
13 Numata, M. and Orlowski, J. (2001) Molecular cloning and characterization of a novel
(Na+,K+)/H+exchanger localized to the
trans
-Golgi network. J. Biol. Chem. 276,
17387–17394
14 Wakabayashi, S., Fafournoux, P., Sardet, C. and Pouyssegur, J. (1992) The Na+/H+
antiporter cytoplasmic domain mediates growth factor signals and controls
‘H+-sensing’. Proc. Natl. Acad. Sci. U.S.A. 89, 2424–2428
15 Pedersen, S. F., O’Donnell, M. E., Anderson, S. E. and Cala, P. M. (2006) Physiology and
pathophysiology of Na+/H+exchange and Na+–K+–2Clcotransport in the heart,
brain, and blood. Am. J. Physiol. Regul. Integr. Comp. Physiol. 291, R1–R25
16 Masereel, B., Pochet, L. and Laeckmann, D. (2003) An overview of inhibitors of Na+/H+
exchanger. Eur. J. Med. Chem. 38, 547–554
17 Counillon, L., Scholz, W., Lang, H. L. and Pouyss´
egur, J. (1993) Pharmacological
characterization of stably transfected Na+/H+antiporter isoforms using amiloride
analogs and a new inhibitor exhibiting anti-ischemic properties. Mol. Pharmacol. 44,
1041–1045
18 Grinstein, S., Rotin, D. and Mason, M. J. (1989) Na+/H+exchange and growth
factor-induced cytosolic pH changes. Role in cellular proliferation.
Biochim. Biophys. Acta 988, 73–97
19 Pouyssegur, J., Sardet, C., Franchi, A., L’Allemain, G. and Paris, S. (1984) A specific
mutation abolishing Na+/H+antiport activity in hamster fibroblasts precludes growth at
neutral and acidic pH. Proc. Natl. Acad. Sci. U.S.A. 81, 4833–4837
20 Wakabayashi, S., Sardet, C., Fafournoux, P., Counillon, L., Meloche, S., Pages, G. and
Pouyssegur, J. (1992) Structure function of the growth factor-activatable Na+/H+
exchanger (NHE1). Rev. Physiol. Biochem. Pharmacol. 119, 157–186
21 Shrode, L., Cabado, A., Goss, G. and Grinstein, S. (1996) Role of the Na+/H+antiporter
isoforms in cell volume regulation. In The Na+/H+Exchanger (Fliegel, L., ed.),
pp. 101–122, R.G. Landes Company, Austin, TX
22 Rotin, D. and Grinstein, S. (1989) Impaired cell volume regulation in Na+-H+
exchange-deficient mutants. Am. J. Phyiol. 257, C1158–C1165
23 Bianchini, L., Kapus, A., Lukacs, G., Wasan, S., Wakabayashi, S., Pouyssegur, J., Yu,
F. H., Orlowski, J. and Grinstein, S. (1995) Responsiveness of mutants of NHE1 isoform
of Na+/H+antiport to osmotic stress. Am. J. Physiol. 269, C998–C1007
24 Hoffmann, E. and Simonsen, L. O. (1989) Membrane mechanisms in volume and pH
regulation in vertebrate cells. Physiol. Rev. 69, 315–382
25 Putney, L. K. and Barber, D. L. (2003) Na–H exchange-dependent increase in
intracellular pH times G2/M entry and transition. J. Biol. Chem. 278, 44645–44649
26 Wang, J., Singh, D. and Fliegel, L. (1997) The Na+/H+exchanger potentiates growth
and retinoic acid induced differentiation of P19 embryonal carcinoma cells.
J. Biol. Chem. 272, 26545–26549
27 Rao, G. N., Sardet, C., Pouyssegur, J. and Berk, B. C. (1992) Na+/H+antiporter gene
expression increases during retinoic acid-induced granulocytic differentiation of HL60
cells. J. Cell. Physiol. 151, 361–366
28 Yang, W., Dyck, J. R. B. and Fliegel, L. (1996) Regulation of NHE1 expression in L6
muscle cells. Biochim. Biophys. Acta 1306, 107–113
29 Wu, K. L., Khan, S., Lakhe-Reddy, S., Jarad, G., Mukherjee, A., Obejero-Paz, C. A.,
Konieczkowski, M., Sedor, J. R. and Schelling, J. R. (2004) The NHE1 Na+/H+
exchanger recruits ezrin/radixin/moesin proteins to regulate Akt-dependent cell survival.
J. Biol. Chem. 279, 26280–26286
30 Behr, T. M., Berova, M., Doe, C. P., Ju, H., Angermann, C. E., Boehm, J. and Willette,
R. N. (2003) p38 Mitogen-activated protein kinase inhibitors for the treatment of chronic
cardiovascular disease. Curr. Opin. Investig. Drugs 4, 1059–1064
31 Rich, I. N., Worthington-White, D., Garden,O. A. and Musk, P. (2000) Apoptosis of
leukemic cells accompanies reduction in intracellular pH after targeted inhibition of the
Na+/H+exchanger. Blood 95, 1427–1434
32 Khaled, A. R., Moor, A. N., Li, A., Kim, K., Ferris, D. K., Muegge, K., Fisher, R. J.,
Fliegel, L. and Durum, S. K. (2001) Trophic factor withdrawal: p38 mitogen-activated
protein kinase activates NHE1, which induces intracellular alkalinization. Mol. Cell. Biol.
21, 7545–7557
33 Denker, S. P., Huang, D. C., Orlowski, J., Furthmayr, H. and Barber, D. L. (2000) Direct
binding of the Na-H exchager NHE1 to ERM proteins regulates the cortical cytoskeleton
and cell shape independently of H+translocation. Mol. Cell 8, 1425–1436
34 Karmazyn, M., Sawyer, M. and Fliegel, L. (2005) The Na+/H+exchanger: a target for
cardiac therapeutic intervention. Curr. Drug Targets Cardiovasc. Haematol. Disord. 5,
323–335
35 Denker, S. P. and Barber, D. L. (2002) Cell migration requires both ion translocation and
cytoskeletal anchoring by the Na-H exchanger NHE1. J. Cell Biol. 159, 1087–1096
36 Putney, L. K., Denker, S. P. and Barber, D. L. (2002) The changing face of the Na+/H+
exchanger, NHE1: structure, regulation, and cellular actions. Annu. Rev.
Pharmacol. Toxicol. 42, 527–552
37 Baumgartner, M. and Patel H., Barber D. L. (2004) Na+/H+exchanger NHE1 as plasma
membrane scaffold in the assembly of signaling complexes. Am. J. Physiol.
Cell Physiol. 287, C844–C850
38 Fliegel, L. and Wang, H. (1997) Regulation of the Na+/H+exchanger in the mammalian
myocardium. J. Mol. Cell. Cardiol. 29, 1991–1999
39 Allen, D. G. and Xiao, X. H. (2003) Role of the cardiac Na+/H+exchanger during
ischemia and reperfusion. Cardiovasc. Res. 57, 934–941
40 Lazdunski, M., Frelin, C. and Vigne, P. (1985) The sodium/hydrogen exchange system in
cardiac cells. Its biochemical and pharmacological properties and its role in regulating
internal concentrations of sodium and internal pH. J. Mol. Cell. Cardiol. 17,
1029–1042
41 Avkiran, M. (2001) Protection of the ischaemic myocardium by Na+/H+exchange
inhibitors: potential mechanisms of action. Basic Res. Cardiol. 96, 306–311
42 Gumina, R. J., Buerger, E., Eickmeier, C., Moore, J., Daemmgen, J. and Gross, G. J.
(1999) Inhibition of the Na+/H+exchanger confers greater cardioprotection against 90
minutes of myocardial ischemia than ischemic preconditioning in dogs. Circulation
100, 2519–2526
43 Spitznagel, H., Chung, O., Xia, Q., Rossius, B., Illner, S., Jahnichen, G., Sandmann, S.,
Reinecke, A., Daemen, M. J. and Unger, T. (2000) Cardioprotective effects of the
Na+/H+-exchange inhibitor cariporide in infarct-induced heart failure. Cardiovasc. Res.
46, 102–110
44 Chen, L., Chen, C. X., Gan, X. T., Beier, N., Scholz, W. and Karmazyn, M. (2004)
Inhibition and reversal of myocardial infarction-induced hypertrophy and heart failure by
NHE-1 inhibition. Am. J. Physiol. Heart Circ. Physiol. 286, H381–H387
45 Wang, Y., Meyer, J. W., Ashraf, M. and Shull, G. E. (2003) Mice with a null mutation in
the NHE1 Na+–H+exchanger are resistant to cardiac ischemia–reperfusion injury.
Circ. Res. 93, 776–782
46 Theroux, P., Chaitman, B. R., Danchin, N., Erhardt, L., Meinertz, T., Schroeder, J. S.,
Tognoni, G., White, H. D., Willerson, J. T. and Jessel, A. (2000) Inhibition of the
sodium-hydrogen exchanger with cariporide to prevent myocardial infarction in
high-risk ischemic situations. Main results of the GUARDIAN trial. Guard during
ischemia against necrosis (GUARDIAN) Investigators. Circulation 102,
3032–3038
47 Zeymer, U., Suryapranata, H., Monassier, J. P., Opolski, G., Davies, J., Rasmanis, G.,
Linssen, G., Tebbe, U., Schroder, R., Tiemann, R. et al. (2001) The Na+/H+exchange
inhibitor eniporide as an adjunct to early reperfusion therapy for acute myocardial
infarction. Results of the evaluation of the safety and cardioprotective effects of eniporide
in acute myocardial infarction (ESCAMI) trial. J. Am. Coll. Cardiol. 38,
1644–1650
c
2007 Biochemical Society
632 E. R. Slepkov and others
48 Rupprecht, H. J., vom Dahl, J., Terres, W., Seyfarth, K. M., Richardt, G., Schultheibeta,
H. P., Buerke, M., Sheehan, F. H. and Drexler, H. (2000) Cardioprotective effects of the
Na+/H+exchange inhibitor cariporide in patients with acute anterior myocardial
infarction undergoing direct PTCA. Circulation 101, 2902–2908
49 Mentzer, Jr, R. M. (2003) Effects of Na+/H+exchange inhibition by cariporide on death
and nonfatal myocardial infarction in patients undergoing coronary artery bypass graft
surgery: the Expedition study. Circulation 108, 2723
50 Rotin, D., Steele-Norwood, D., Grinstein, S. and Tannock, I. (1989) Requirement of the
Na+/H+exchanger for tumor growth. Cancer Res. 49, 205–211
51 Reshkin, S. J., Bellizzi, A., Caldeira, S., Albarani, V., Malanchi, I., Poignee, M.,
Alunni-Fabbroni, M., Casavola, V. and Tommasino, M. (2000) Na+/H+exchanger-
dependent intracellular alkalinization is an early event in malignant transformation and
plays an essential role in the development of subsequent transformation-associated
phenotypes. FASEB J. 14, 2185–2197
52 Harguindey, S., Orive, G., Luis Pedraz, J., Paradiso, A. and Reshkin, S. J. (2005) The
role of pH dynamics and the Na+/H+antiporter in the etiopathogenesis and treatment of
cancer. Two faces of the same coin one single nature. Biochim. Biophys. Acta 1756,
1–24
53 Cardone, R. A., Bagorda, A., Bellizzi, A., Busco, G., Guerra, L., Paradiso, A.,
Casavola, V., Zaccolo, M. and Reshkin, S. J. (2005) Protein kinase A gating of a
pseudopodial-located RhoA/ROCK/p38/NHE1 signal module regulates invasion in
breast cancer cell lines. Mol. Biol. Cell 16, 3117–3127
54 Paradiso, A., Cardone, R. A., Bellizzi, A., Bagorda, A., Guerra, L., Tommasino, M.,
Casavola, V. and Reshkin, S. J. (2004) The Na+–H+exchanger-1 induces cytoskeletal
changes involving reciprocal RhoA and Rac1 signaling, resulting in motility and
invasion in MDA-MB-435 cells. Breast Cancer Res. 6, R616–R628
55 Bourguignon, L. Y., Singleton, P. A., Diedrich, F., Stern, R. and Gilad, E. (2004) CD44
interaction with Na+–H+exchanger (NHE1) creates acidic microenvironments leading
to hyaluronidase-2 and cathepsin B activation and breast tumor cell invasion.
J. Biol. Chem. 279, 26991–27007
56 Reshkin, S. J., Bellizzi, A., Cardone, R. A., Tommasino, M., Casavola, V. and Paradiso, A.
(2003) Paclitaxel induces apoptosis via protein kinase A- and p38 mitogen-activated
protein-dependent inhibition of the Na+/H+exchanger (NHE) NHE isoform 1 in human
breast cancer cells. Clin. Cancer Res. 9, 2366–2373
57 Karmazyn, M., Sostaric, J. V. and Gan, X. T. (2001) The myocardial Na+/H+exchanger:
a potential therapeutic target for the prevention of myocardial ischaemic and reperfusion
injury and attenuation of postinfarction heart failure. Drugs 61, 375–389
58 Avkiran, M. and Haworth, R. S. (2003) Regulatory effects of G protein-coupled receptors
on cardiac sarcolemmal Na+/H+exchanger activity: signalling and significance.
Cardiovasc. Res. 57, 942–952
59 Dyck, J. R. B., Maddaford, T., Pierce, G. N. and Fliegel, L. (1995) Induction of expression
of the sodium–hydrogen exchanger in rat myocardium. Cardiovasc. Res. 29, 203–208
60 Fernandez-Rachubinski, F. and Fliegel, L. (2001) COUP-TF I and COUP-TFII regulate
expression of the NHE through a nuclear hormone responsive element with enhancer
activity. Eur. J. Biochem. 268, 620–634
61 Slepkov, E. and Fliegel, L. (2004) Regulation of expression of the Na+/H+exchanger by
thyroid hormone. Vitam. Horm. 69, 249–269
62 Yang, W., Wang, H. and Fliegel, L. (1996) Regulation of Na+/H+exchanger gene
expression. Role of a novel poly(dA:dT) element in regulation of the NHE1 promoter.
J. Biol. Chem. 271, 20444–20449
63 Sardet, C., Counillon, L., Franchi, A. and Pouyssegur, J. (1990) Growth factors induce
phosphorylation of the Na+/H+antiporter, glycoprotein of 110 kD. Science 247,
723–726
64 Sardet, C., Fafournoux, P. and Pouyssegur, J. (1991) α-Thrombin, epidermal growth
factor, and okadaic acid activate the Na+/H+exchanger, NHE-1, by phosphorylating a
set of common sites. J. Biol. Chem. 266, 19166–19171
65 Haworth, R. S., McCann, C., Snabaitis, A. K., Roberts, N. A. and Avkiran, M. (2003)
Stimulation of the plasma membrane Na+/H+exchanger NHE1 by sustained
intracellular acidosis. Evidence for a novel mechanism mediated by the ERK pathway.
J. Biol. Chem. 278, 31676–31684
66 Moor, A. N. and Fliegel, L. (1999) Protein kinase mediated regulation of the Na+/H+
exchanger in the rat myocardium by MAP-kinase-dependent pathways. J. Biol. Chem.
274, 22985–22992
67 Wang, H., Silva, N. L.C. L., Lucchesi, P. A., Haworth, R., Wang, K., Michalak, M.,
Pelech, S. and Fliegel, L. (1997) Phosphorylation and regulation of the Na+/H+
exchanger through mitogen-activated protein kinase. Biochemistry 36, 9151–9158
68 Snabaitis, A. K., Yokoyama, H. and Avkiran, M. (2000) Roles of mitogen-activated
protein kinases and protein kinase C in α1A-adrenoreceptor-mediated stimulation of the
sarcolemmal Na+–H+exchanger. Circ. Res. 86, 214–220
69 Takahashi, E., Abe, J.-i. and Berk, B. C. (1997) Angiotensin II stimulates p90rsk in
vascular smooth muscle cells: a potential Na+/H+exchanger kinase. Hypertension 29,
1265–1272
70 Takahashi, E., Abe, J., Gallis, B., Aebersold, R., Spring, D. J., Krebs, E. G. and Berk,
B. C. (1999) p90RSK is a serum-stimulated Na+/H+exchanger isoform-1 kinase.
Regulatory phosphorylation of serine 703 of Na+/H+exchanger isoform-1.
J. Biol. Chem. 274, 20206–20214
71 Tominaga, T., Ishizaki, T., Narumiya, S. and Barber, D. L. (1998) p160ROCK mediates
RhoA activation of Na–H exchange. EMBO J. 17, 4712–4722
72 Yan, W., Nehrke, K., Choi, J. and Barber, D. L. (2001) The Nck-interacting kinase (NIK)
phosphorylates the Na+–H+exchanger NHE1 and regulates NHE1 activation by
platelet-derived growth factor. J. Biol. Chem. 276, 31349–31356
73 Fliegel, L., Walsh, M. P., Singh, D., Wong, C. and Barr, A. (1992) Phosphorylation of the
carboxyl-terminal domain of the Na+/H+exchanger by Ca2+/calmodulin-dependent
protein kinase II. Biochem. J. 282, 139–145
74 Kusuhara, M., Takahashi, E., Peterson, T. E., Abe, J., Ishida, M., Han, J., Ulevitch, R. and
Berk, B. C. (1998) p38 Kinase is a negative regulator of angiotensin II signal
transduction in vascular smooth muscle cells. Effects on Na+/H+exchange and ERK1/2.
Circ. Res. 83, 824–831
75 Haworth, R. S., Sinnett-Smith, J., Rozengurt, E. and Avkiran, M. (1999) Protein kinase D
inhibits plasma membrane Na+/H+exchanger activity. Am. J. Physiol. 277,
C1202–C1209
76 Sauvage, M., Maziere, P., Fathallah, H. and Giraud, F. (2000) Insulin stimulates NHE1
activity by sequential activation of phosphatidylinositol 3-kinase and protein kinase Cζ
in human erythrocytes. Eur. J. Biochem. 267, 955–962
77 Misik, A. J., Perreault, K., Holmes, C. F. and Fliegel, L. (2005) Protein phosphatase
regulation of Na+/H+exchanger isoform I. Biochemistry 44, 5842–5852
78 Bertrand, B., Wakabayashi, S., Ikeda, T., Pouyssegur, J. and Shigekawa, M. (1994) The
Na+/H+exchanger isoform 1 (NHE1) is a novel member of the calmodulin-binding
proteins. J. Biol. Chem. 269, 13703–13709
79 Wakabayashi, S., Bertrand, B., Ikeda, T., Pouyssegur, J. and Shigekawa, M. (1994)
Mutation of calmodulin-binding site renders the Na+/H+exchanger (NHE1) highly
H+-sensitive and Ca2+regulation-defective. J. Biol. Chem. 269, 13710–13715
80 Pang, T., Su, X., Wakabayashi, S. and Shigekawa, M. (2001) Calcineurin homologous
protein as an essential cofactor for Na+/H+exchangers. J. Biol. Chem. 276,
17367–17372
81 Lin, X. and Barber, D. L. (1996) A calcineurin homologous protein inhibits
GTPase-stimulated Na–H exchange. J. Biol. Chem. 93, 12631–12636
82 Pang, T. H. T., Mori, H., Shigekawa, M. and Wakabayashi, S. (2004) Role of
calcineurin B homologous protein in pH regulation by the Na+/H+exchanger 1:
tightly bound Ca2+ions as important structural elements. Biochemistry 43,
3628–3636
83 Li, X., Liu, Y., Kay, C. M., Muller-Esterl, W. and Fliegel, L. (2003) The Na+/H+
exchanger cytoplasmic tail: structure, function, and interactions with tescalcin.
Biochemistry 42, 7448–7456
84 Mailander, J., Muller-Esterl, W. and Dedio, J. (2001) Human homolog of mouse
tescalcin associates with Na+/H+exchanger type-1. FEBS Lett. 507, 331–335
85 Li, X., Liu, Y., Alvarez, B. V., Casey, J. R. and Fliegel, L. (2006)A novel carbonic
anhydrase II binding site regulates NHE1 activity. Biochemistry 45, 2414–2424
86 Li, X., Alvarez, B., Casey, J. R., Reithmeier, R. A. and Fliegel, L. (2002) Carbonic
anhydrase II binds to and enhances activity of the Na+/H+exchanger. J. Biol. Chem.
277, 36085–36091
87 Lehoux, S., Abe, J., Florian, J. A. and Berk, B. C. (2001) 14-3-3 Binding to
Na+/H+exchanger isoform-1 is associated with serum-dependent activation
of Na+/H+exchange. J. Biol. Chem. 276, 15794–15800
88 Aharonovitz, O., Zaun, H. C., Balla, T., York, J. D., Orlowski, J. and Grinstein, S. (2000)
Intracellular pH regulation by Na+/H+exchange requires phosphatidylinositol
4,5-bisphosphate. J. Cell Biol. 150, 213–224
89 Silva, N. L., Haworth, R. S., Singh, D. and Fliegel, L. (1995) The carboxyl-terminal
region of the Na+/H+exchanger interacts with mammalian heat shock protein.
Biochemistry 34, 10412–10420
90 Wakabayashi, S., Pang, T., Su, X. and Shigekawa, M. (2000) A novel topology
model of the human Na+/H+exchanger isoform 1. J. Biol. Chem. 275,
7942–7949
91 Counillon, L., Pouyss´
egur, J. and Reithmeier, R. A. F. (1994) The Na+/H+exchanger
NHE-1 possesses N- and O-linked glycosylation restricted to the first N-terminal
extracellular domain. Biochemistry 33, 10463–10469
92 Haworth, R. S., Frohlich, O. and Fliegel, L. (1993) Multiple carbohydrate moieties on the
Na+/H+exchanger. Biochem. J. 289, 637–640
93 Fliegel, L., Haworth, R. S. and Dyck, J. R. B. (1993) Characterization of the placental
brush border membrane Na+/H+exchanger: identification of thiol-dependent
transitions in apparent molecular size. Biochem. J. 289, 101–107
94 Fafournoux, P., Noel, J. and Pouyss ´
egur, J. (1994) Evidence that Na+/H+exchanger
isoforms NHE1 and NHE3 exist as stable dimers in membranes with a high degree of
specificity for homodimers. J. Biol. Chem. 269, 2589–2596
c
2007 Biochemical Society
Structural and functional analysis of the Na+/H+exchanger 633
95 Hisamitsu, T., Pang, T., Shigekawa, M. and Wakabayashi, S. (2004) Dimeric interaction
between the cytoplasmic domains of the Na+/H+exchanger NHE1 revealed by
symmetrical intermolecular cross-linking and selective co-immunoprecipitation.
Biochemistry 43, 11135–11143
96 Gebreselassie, D., Rajarathnam, K. and Fliegel, L. (1998) Expression, purification, and
characterization of the carboxyl-terminal region of the Na+/H+exchanger.
Biochem. Cell Biol. 76, 837–842
96b Ding, J., Rainey, J. K, Xu, C., Sykes, B. D. and Fliegel, L. (2006) Structural and
functional characterization of transmembrane segment VII of the Na+/H+exchanger
isoform 1. J. Biol. Chem. 281, 29817–29829
97 Counillon, L., Franchi, A. and Pouyssegur, J. (1993) A point mutation of the Na+/H+
exchanger gene (NHE1) and amplification of the mutated allele confer amiloride
resistance upon chronic acidosis. Proc. Natl. Acad. Sci. U.S.A. 90, 4508–4512
98 Counillon, L., Noel, J., Reithmeier, R. A. F. and Pouyssegur, J. (1997) Random
mutagenesis reveals a novel site involved in inhibitor interaction within the fourth
transmembrane segment of the Na+/H+exchanger-1. Biochemistry 36, 2951–2959
99 Touret, N., Poujeol, P. and Counillon, L. (2001) Second-site revertants of a
low-sodium-affinity mutant of the Na+/H+exchanger reveal the participation of TM4
into a highly constrained sodium-binding site. Biochemistry 40, 5095–5101
100 Slepkov, E. R., Chow, S., Lemieux, M. J. and Fliegel, L. (2004) Proline residues in
transmembrane segment IV are critical for activity, expression and targeting of the
Na+/H+exchanger isoform 1. Biochem. J. 379, 31–38
101 Terzic, A., Puceat, M., Clement, O., Scamps, F. and Vassort, G. (1982) α1-Adrenergic
effects on intracellular pH and calcium and on myofilaments in single rat cardiac cells.
J. Physiol. 447, 275–292
102 Murtazina, B., Booth, B. J., Bullis, B. L., Singh, D. N. and Fliegel, L. (2001) Functional
analysis of polar amino acid residues in membrane associated regions of the NHE1
isoform of the Na+/H+exchanger. Eur. J. Biochem. 268,113
103 Wang, D., Balkovetz, D. F. and Warnock, D. G. (1995) Mutational analysis of
transmembrane histidines in the amiloride-sensitive Na+/H+exchanger. Am. J. Physiol.
269, C392–C402
104 Orlowski, J. and Kandasamy, R. A. (1996) Delineation of transmembrane domains of the
Na+/H+exchanger that confer sensitivity to pharmacological antagonists.
J. Biol. Chem. 271, 19922–19927
105 Noel, J., Germain, D. and Vadnais, J. (2003) Glutamate 346 of human Na+–H+
exchanger NHE1 is crucial for modulating both the affinity for Na+and the interaction
with amiloride derivatives. Biochemistry 42, 15361–15368
106 Wakabayashi, S., Pang, T., Su, X. and Shigekawa, M. (2000) Second mutations rescue
point mutant of the Na+/H+exchanger NHE1 showing defective surface expression.
FEBS Lett 487, 257–261
107 Wakabayashi, S., Hisamitsu, T., Pang, T. and Shigekawa, M. (2003) Mutations of Arg440
and Gly455/Gly456 oppositely change pH sensing of Na+/H+exchanger 1. J. Biol. Chem.
278, 11828–11835
108 Su, X., Pang, T., Wakabayashi, S. and Shigekawa, M. (2003) Evidence for involvement of
the putative first extracellular loop in differential volume sensitivity of the Na+/H+
exchangers NHE1 and NHE2. Biochemistry 42, 1086–1094
109 Dibrov, P., Murtazina, R., Kinsella, J. and Fliegel, L. (2000) Characterization of a
histidine rich cluster of amino acids in the cytoplasmic domain of the Na+/H+
exchanger. Biosci. Rep. 20, 185–197
110 Padan, E., Rimon, A., Tzubery, T., Muller, M., Katia, H. and Galili, L. (2003) NhaA
Na+/H+antiporter: structure, mechanism and function in homeostasis of Na+and pH.
In The Na+/H+Exchanger: From Molecule to Its Role in Disease (Karmazyn, M.,
Avkiran, M. and Fliegel, L., eds.), pp. 91–108, Kluwer Academic Publishers
111 Wiebe, C. A., DiBattista, E. R. and Fliegel, L. (2001) Functional role of amino acid
residues in Na+/H+exchangers. Biochem. J. 357,110
112 Hunte, C., Screpanti, E., Venturi, M., Rimon, A., Padan, E. and Michel, H. (2005)
Structure of a Na+/H+antiporter and insights into mechanism of action and regulation
by pH. Nature 435, 1197–1202
113 Toyoshima, C., Nakasako, M., Nomura, H. and Ogawa, H. (2000) Crystal structure of the
calcium pump of sarcoplasmic reticulum at 2.6 A
˚resolution. Nature 405, 647–655
114 Yamashita, A., Singh, S. K., Kawate, T., Jin, Y. and Gouaux,E. (2005) Crystal structure of
a bacterial homologue of Na+/Cl-dependent neurotransmitter transporters. Nature
437, 215–223
115 Padan, E., Tzubery, T., Herz, K., Kozachkov, L., Rimon, A. and Galili, L. (2004)
NhaA of
Escherichia coli
, as a model of a pH-regulated Na+/H+antiporter.
Biochim. Biophys. Acta 1658,213
116 Dibrov, P. and Fliegel, L. (1998) Comparative molecular analysis of Na+/H+
exchangers: a unified model for Na+/H+antiport? FEBS Lett. 424,15
117 Abramson, J., Smirnova, I., Kasho, V., Verner, G., Kaback, H. R. and Iwata, S. (2003)
Structure and mechanism of the lactose permease of
Escherichia coli
. Science 301,
610–615
118 Andrade, S. L., Dickmanns, A., Ficner, R. and Einsle, O. (2005) Crystal structure of the
archaeal ammonium transporter Amt-1 from
Archaeoglobus fulgidus
. Proc. Natl. Acad.
Sci. U.S.A. 102, 14994–14999
119 Dutzler, R., Campbell, E. B., Cadene, M., Chait, B. T. and MacKinnon, R. (2002) X-ray
structure of a ClC chloride channel at 3.0 A
˚reveals the molecular basis of anion
selectivity. Nature 415, 287–294
120 Khademi, S., O’Connell, 3rd, J., Remis, J., Robles-Colmenares, Y., Miercke, L. J. and
Stroud, R.M. (2004) Mechanism of ammonia transport by Amt/MEP/Rh: structure of
AmtB at 1.35 A
˚. Science 305, 1587–1594
121 Locher, K. P., Lee, A. T. and Rees, D. C. (2002) The
E
.
coli
BtuCD structure:
a framework for ABC transporter architecture and mechanism. Science 296,
1091–1098
122 Van den Berg, B., Clemons, Jr, W. M., Collinson, I., Modis, Y., Hartmann, E., Harrison,
S. C. and Rapoport, T. A. (2004) X-ray structure of a protein-conducting channel. Nature
427, 36–44
123 Kinsella, J. L., Heller, P. and Froehlich, J. P. (1998) Na+/H+exchanger: proton modifier
site regulation of activity. Biochem. Cell Biol. 76, 743–749
124 Gerchman, Y., Rimon, A., Venturi, M. and Padan, E. (2001) Oligomerization of NhaA, the
Na+/H+antiporter of
Escherichia coli
in the membrane and its functional and structural
consequences. Biochemistry 40, 3403–3412
125 Hunt, J. F., Earnest, T. N., Bousche, O., Kalghatgi, K., Reilly, K., Horvath, C., Rothschild,
K. J. and Engelman, D. M. (1997) A biophysical study of integral membrane protein
folding. Biochemistry 36, 15156–15176
126 Katragadda, M., Chopra, A., Bennett, M., Alderfer, J. L., Yeagle, P. L. and Albert, A. D.
(2001) Structures of the transmembrane helices of the G-protein coupled receptor,
rhodopsin. J. Pept. Res. 58, 79–89
127 Oblatt-Montal, M., Reddy, G. L., Iwamoto, T., Tomich, J. M. and Montal, M. (1994)
Identification of an ion channel-forming motif in the primary structure of CFTR, the
cystic fibrosis chloride channel. Proc. Natl. Acad. Sci. U.S.A. 91,
1495–1499
128 Katragadda, M., Alderfer, J. L. and Yeagle, P. L. (2001) Assembly of a polytopic
membrane protein structure from the solution structures of overlapping peptide
fragments of bacteriorhodopsin. Biophys. J. 81, 1029–1036
129 Wigley, W. C., Vijayakumar, S., Jones, J. D., Slaughter, C. and Thomas, P. J. (1998)
Transmembrane domain of cystic fibrosis transmembrane conductance regulator:
design, characterization, and secondary structure of synthetic peptides m1–m6.
Biochemistry 37, 844–853
130 Naider, F., Khare, S., Arshava, B., Severino, B., Russo, J. and Becker, J. M. (2005)
Synthetic peptides as probes for conformational preferences of domains of membrane
receptors. Biopolymers 80, 199–213
131 Choi, G., Landin, J., Galan, J. F., Birge, R. R., Albert, A. D. and Yeagle, P. L. (2002)
Structural studies of metarhodopsin II, the activated form of the G-protein coupled
receptor, rhodopsin. Biochemistry 41, 7318–7324
132 Yeagle, P. L., Danis, C., Choi, G., Alderfer, J. L. and Albert, A. D. (2000) Three
dimensional structure of the seventh transmembrane helical domain of the G-protein
receptor, rhodopsin. Mol. Vis. 6, 125–131
133 Qiu, Z., Nicoll, D. A. and Philipson, K. D. (2001) Helix packing of functionally
important regions of the cardiac Na+–Ca2+exchanger. J. Biol. Chem. 276,
194–199
134 Loo, T. W. and Clarke, D. M. (2000) The packing of the transmembrane segments of
human multidrug resistance P-glycoprotein is revealed by disulfide cross-linking
analysis. J. Biol. Chem. 275, 5253–5256
135 Wu, J. and Kaback, H. R. (1996) A general method for determining helix packing in
membrane protein
in situ
: helices I and II are close to helix VII in the lactose permease of
Escherichia coli
. Proc. Natl. Acad. Sci. U.S.A. 93, 14498–14502
136 Sorgen, P. L., Hu, Y., Guan, L., Kaback, H. R. and Girvin, M. E. (2002) An approach to
membrane protein structure without crystals. Proc. Natl. Acad. Sci. U.S.A. 99,
14037–14040
137 Bruschweiler, R., Blackledge, M. and Ernst, R. R. (1991) Multi-conformational peptide
dynamics derived from NMR data: a new search algorithm and its application to
antamanide. J. Biomol. NMR 1,311
Received 13 July 2006/28 September 2006; accepted 30 October 2006
Published on the Internet 12 January 2007, doi:10.1042/BJ20061062
c
2007 Biochemical Society
... Some of the members of the SLC26A family show evidence of interaction with the CA II isoform, based on experiments with CA blockers like acetazolamide that reduces the exchange activity, or truncation experiments showing the CA II binding sites (Sterling et al., 2002;Alvarez et al., 2003). The mammalian Na + /H + exchangers (NHEs) belong to the SLC9 gene family, and so far nine separate isoforms of these exchangers have been found (Slepkov et al., 2007). Their primary role is to catalyze the exchange of Na + or K + with H + down their concentration gradients. ...
... Their primary role is to catalyze the exchange of Na + or K + with H + down their concentration gradients. Of the nine isoforms of NHEs only NHE1 (Slepkov et al., 2007) and NHE3 (Krishnan et al., 2015) have been shown to interact with CA. The NHE1 isoform of NHEs is ubiquitously expressed in the plasma membranes of nearly all tissue, is highly regulated by pH, and very important for regulating cell volume. ...
Article
Full-text available
It has been known for some time that Carbonic Anhydrase (CA, EC 4.2.1.1) plays a complex role in vascular function, and in the regulation of vascular tone. Clinically employed CA inhibitors (CAIs) are used primarily to lower intraocular pressure in glaucoma, and also to affect retinal blood flow and oxygen saturation. CAIs have been shown to dilate vessels and increase blood flow in both the cerebral and ocular vasculature. Similar effects of CAIs on vascular function have been observed in the liver, brain and kidney, while vessels in abdominal muscle and the stomach are unaffected. Most of the studies on the vascular effects of CAIs have been focused on the cerebral and ocular vasculatures, and in particular the retinal vasculature, where vasodilation of its vessels, after intravenous infusion of sulfonamide-based CAIs can be easily observed and measured from the fundus of the eye. The mechanism by which CAIs exert their effects on the vasculature is still unclear, but the classic sulfonamide-based inhibitors have been found to directly dilate isolated vessel segments when applied to the extracellular fluid. Modification of the structure of CAI compounds affects their efficacy and potency as vasodilators. CAIs of the coumarin type, which generally are less effective in inhibiting the catalytically dominant isoform hCA II and unable to accept NO, have comparable vasodilatory effects as the primary sulfonamides on pre-contracted retinal arteriolar vessel segments, providing insights into which CA isoforms are involved. Alterations of the lipophilicity of CAI compounds affect their potency as vasodilators, and CAIs that are membrane impermeant do not act as vasodilators of isolated vessel segments. Experiments with CAIs, that shed light on the role of CA in the regulation of vascular tone of vessels, will be discussed in this review. The role of CA in vascular function will be discussed, with specific emphasis on findings with the effects of CA inhibitors (CAI).
... Other investigators have shown that NHE activity can be modulated in many types of cells by a variety of signal transduction pathways and a number of regulatory domains exist on the exchanger (reviewed in Refs. [38][39][40][41][42]. Among these regulatory regions are sites for binding calmodulin. ...
Article
Full-text available
Previous work has shown that activation of tiger salamander retinal radial glial cells by extracellular ATP induces a pronounced extracellular acidification, which has been proposed to be a potent modulator of neurotransmitter release. This study demonstrates that low micromolar concentrations of extracellular ATP similarly induce significant H+ effluxes from Müller cells isolated from the axolotl retina. Müller cells were enzymatically isolated from axolotl retina and H+ fluxes measured from individual cells using self-referencing H+-selective microelectrodes. The increased H+ efflux from axolotl Müller cells induced by extracellular ATP required activation of metabotropic purinergic receptors and was dependent upon calcium released from internal stores. We further found that ATP-evoked increases in H+ efflux from Müller cells of both tiger salamander and axolotl were sensitive to pharmacological agents known to interrupt calmodulin and protein kinase C (PKC) activity: chlorpromazine (CLP), trifluoperazine (TFP), and W-7 (all calmodulin inhibitors) and chelerythrine, a PKC inhibitor, all attenuated ATP-elicited increases in H+ efflux. ATP-initiated H+ fluxes of axolotl Müller cells were also significantly reduced by amiloride, suggesting a significant contribution by sodium-hydrogen exchangers (NHE). In addition, α-cyano-4-hydroxycinnamate (4-cin), a monocarboxylate transport (MCT) inhibitor, also reduced the ATP-induced increase in H+ efflux in both axolotl and tiger salamander Müller cells, and when combined with amiloride, abolished ATP-evoked increase in H+ efflux. These data suggest that axolotl Müller cells are likely to be an excellent model system in which to understand the cell-signaling pathways regulating H+ release from glia and the role this may play in modulating neuronal signaling.
... NHEs are widespread in neuronal and non-neuronal cells. Of the 9 NHE isoforms identified, NHE1 is the most ubiquitous and expressed in nerve cells (Slepkov et al., 2007;Torres-Lopez et al., 2013). NHE inhibitors concentration-dependently produce intracellular acidosis and can even cause cell death (Schneider et al., 2004). ...
Article
Full-text available
Tissue acidification causes sustained activation of primary nociceptors, which causes pain. In mammals, acid-sensing ion channels (ASICs) are the primary acid sensors; however, Na ⁺ /H ⁺ exchangers (NHEs) and TRPV1 receptors also contribute to tissue acidification sensing. ASICs, NHEs, and TRPV1 receptors are found to be expressed in nociceptive nerve fibers. ASIC inhibitors reduce peripheral acid-induced hyperalgesia and suppress inflammatory pain. Also, it was shown that pharmacological inhibition of NHE1 promotes nociceptive behavior in acute pain models, whereas inhibition of TRPV1 receptors gives relief. The murine skin-nerve preparation was used in this study to assess the activation of native polymodal nociceptors by mild acidification (pH 6.1). We have found that diminazene, a well-known antagonist of ASICs did not suppress pH-induced activation of CMH-fibers at concentrations as high as 25 μM. Moreover, at 100 μM, it induces the potentiation of the fibers’ response to acidic pH. At the same time, this concentration virtually completely inhibited ASIC currents in mouse dorsal root ganglia (DRG) neurons (IC 50 = 17.0 ± 4.5 μM). Non-selective ASICs and NHEs inhibitor EIPA (5-(N-ethyl-N-isopropyl)amiloride) at 10 μM, as well as selective NHE1 inhibitor zoniporide at 0.5 μM induced qualitatively the same effects as 100 μM of diminazene. Our results indicate that excitation of afferent nerve terminals induced by mild acidification occurs mainly due to the NHE1, rather than acid-sensing ion channels. At high concentrations, diminazene acts as a weak blocker of the NHE. It lacks chemical similarity with amiloride, EIPA, and zoniporide, so it may represent a novel structural motif for the development of NHE antagonists. However, the effect of diminazene on the acid-induced excitation of primary nociceptors remains enigmatic and requires additional investigations.
... NHE1 had previously been shown to be localized only to the sperm midpiece in rats, not in the sperm head [22]. NHE1 is a ubiquitously expressed NHE that has been suggested to be a "housekeeping" NHE isoform, providing basal pH regulation in all cells [7,42] and its expression in the equatorial region would provide this function to the head of the sperm. Of note, both male and female NHE1 knockout mice are reported to be fertile [8], and our own unpublished data suggest no clear defects in sperm motility in homozygous NHE1 knockout males; therefore, it appears that NHE1 may not be critical for sperm physiology. ...
Article
Full-text available
Na+/H+ exchangers (NHEs) are a family of ion transporters that regulate the pH of various cell compartments across an array of cell types. In eukaryotes, NHEs are encoded by the SLC9 gene family comprising 13 genes. SLC9C2, which encodes the NHE11 protein, is the only one of the SLC9 genes that is essentially uncharacterized. Here, we show that SLC9C2 exhibits testis/sperm-restricted expression in rats and humans, akin to its paralog SLC9C1 (NHE10). Similar to NHE10, NHE11 is predicted to contain an NHE domain, a voltage sensing domain, and finally an intracellular cyclic nucleotide binding domain. An immunofluorescence analysis of testis sections reveals that NHE11 localizes with developing acrosomal granules in spermiogenic cells in both rat and human testes. Most interestingly, NHE11 localizes to the sperm head, likely the plasma membrane overlaying the acrosome, in mature sperm from rats and humans. Therefore, NHE11 is the only known NHE to localize to the acrosomal region of the head in mature sperm cells. The physiological role of NHE11 has yet to be demonstrated but its predicted functional domains and unique localization suggests that it could modulate intracellular pH of the sperm head in response to changes in membrane potential and cyclic nucleotide concentrations that are a result of sperm capacitation events. If NHE11 is shown to be important for male fertility, it will be an attractive target for male contraceptive drugs due to its exclusive testis/sperm-specific expression.
... SLC9s proteins can form homodimers through their C-terminus, although this formation is not required for Na + /H + exchanger activity [222,223]. Several inhibitors are available for SLC9s transporters, and the most commonly used are amiloride derivates [224,225]. SLC9s transporters are involved in a myriad of cellular and/or systemic processes (e. g. regulation of pH, cell volume, systemic control of electrolytes, acid metabolism, homeostasis of fluid volumes), among which pHi regulation is their primary role [193]. ...
Thesis
The construction and maintenance of coral reefs primarily depends on the calcification of corals, which produce a rigid skeleton made of CaCO3 in the crystalline form of aragonite. Most reef-building corals live in symbiosis with photosynthetic dinoflagellates of the Symbiodiniaceae family, which provide the coral host with energy and nutrients. Given their ecological importance, much progress has been made in identifying key elements of the mechanisms underlying coral calcification. Nevertheless, there are still significant gaps in our understanding. Foremost is the characterization of ion transporters, used by the coral calcifying cells to promote calcification. To contribute to this lack of knowledge, targeted and broad approaches, coupled with molecular and bioinformatics tools, have been used throughout this thesis. Using the targeted approach, ion transporter proteins, previously reported to be involved in calcification of other calcifying species, have been identified for the first time in the genome and transcriptome of the symbiotic coral Stylophora pistillata. Whereas, using a broad approach, novel candidate genes for roles in calcification have been identified in the non-symbiotic coral Tubastraea spp. Overall, both approaches contributed to a better understanding of the ion transporting mechanisms used by the coral calcifying cells to promote calcification in this ecologically important group of marine animals.
... For example, the sodium proton exchanger NHE1 is phosphorylated by p90 ribosomal S6 kinases, downstream of BRAF and the ERK1/2 cascade. This alters the activation profile of NHE1 and is associated with increased contractility [42,43]. Notably, expression of activated NHE1 in mice enhances the degree of hypertrophy induced by phenylephrine [44] suggesting that it is a contributing factor to the overall response. ...
Preprint
Full-text available
Cardiac hypertrophy is necessary for the heart to accommodate an increase in workload. Physiological, compensated hypertrophy (e.g. with exercise) is reversible and largely due to cardiomyocyte hypertrophy. Pathological hypertrophy (e.g. with hypertension) is associated with additional features including increased fibrosis and can lead to heart failure. RAF kinases (ARAF/BRAF/RAF1) integrate signals into the ERK1/2 cascade, a pathway implicated in cardiac hypertrophy, and activation of BRAF in cardiomyocytes promotes compensated hypertrophy. Here, we used mice with tamoxifen-inducible cardiomyocyte-specific BRAF knockout (CM-BRAFKO) to assess the role of BRAF in hypertension-associated cardiac hypertrophy induced by angiotensin II (AngII; 0.8 mg/kg/d, 7 d) and physiological hypertrophy induced by phenylephrine (40 mg/kg/d, 7 d). Cardiac dimensions/function were assessed by echocardiography with histological assessment of cellular changes. AngII promoted cardiomyocyte hypertrophy and increased fibrosis within the myocardium (interstitial) and around the arterioles (perivascular) in male mice; cardiomyocyte hypertrophy and interstitial (but not perivascular) fibrosis were inhibited in mice with CM-BRAFKO. Phenylephrine had a limited effect on fibrosis, but promoted cardiomyocyte hypertrophy and increased contractility in male mice; cardiomyocyte hypertrophy was unaffected in mice with CM-BRAFKO, but the increase in contractility was suppressed and fibrosis increased. Phenylephrine induced a modest hypertrophic response in female mice and, in contrast to the males, tamoxifen-induced loss of cardiomyocyte BRAF reduced cardiomyocyte size, had no effect on fibrosis and increased contractility. The data identify BRAF as a key signalling intermediate in both physiological and pathological hypertrophy in male mice, and highlight the need for independent assessment of gene function in females. Clinical perspectives Background . BRAF is a key signalling intermediate that causes cancer and is upregulated in heart failure, but its role in physiological and pathological cardiac hypertrophy remains to be established. Summary . Cardiomyocyte BRAF is required in male mice for hypertrophy and contributes to interstitial fibrosis in hypertension induced by angiotensin II, but it increases contractility and suppresses fibrosis in physiological hypertrophy induced by α 1 -adrenergic receptor stimulation with phenylephrine. Differences between males and females are highlighted in the phenylephrine response. Potential significance of results to human health and disease . BRAF is a key signalling node in both pathological and physiological hypertrophy: inhibiting BRAF may be beneficial in pathological hypertrophy and the data have implications for repurposing of RAF inhibitors developed for cancer; inhibiting BRAF in physiological hypertrophy may result in increased fibrosis and using RAF inhibitors in this context could be detrimental in the longer term.
... Ischemic injury Tissue ischemia disrupts oxygen and glucose delivery, resulting in metabolic disorders 204 . Hypoxia directly leads to lactate accumulation, which lowers pH to activate ion transporters 205 , primarily Na + /H + exchange (NHE) proteins that regulate intracellular pH by exchanging protons for extracellular sodium ions 206 . In acute tissue ischemia, NHE1 elevates intracellular Na + levels, leading to increased Ca 2+ -Na + exchange and intracellular calcium overload. ...
Article
Full-text available
The current understanding of lactate extends from its origins as a byproduct of glycolysis to its role in tumor metabolism, as identified by studies on the Warburg effect. The lactate shuttle hypothesis suggests that lactate plays an important role as a bridging signaling molecule that coordinates signaling among different cells, organs and tissues. Lactylation is a posttranslational modification initially reported by Professor Yingming Zhao’s research group in 2019. Subsequent studies confirmed that lactylation is a vital component of lactate function and is involved in tumor proliferation, neural excitation, inflammation and other biological processes. An indispensable substance for various physiological cellular functions, lactate plays a regulatory role in different aspects of energy metabolism and signal transduction. Therefore, a comprehensive review and summary of lactate is presented to clarify the role of lactate in disease and to provide a reference and direction for future research. This review offers a systematic overview of lactate homeostasis and its roles in physiological and pathological processes, as well as a comprehensive overview of the effects of lactylation in various diseases, particularly inflammation and cancer.
Chapter
Lethal ventricular arrhythmias are one of the primary causes of high mortality among patients with a variety of cardiac diseases, like acute ischemia-reperfusion injury, chronic ischemic heart disease (IHD), all major forms of heart failure (HF), and congenital cardiomyopathies. Perturbations in intracellular Ca2+ and Na+ homeostasis leading to massive intracellular Ca2+ overload are frequently found in the background of cardiac arrhythmogenesis. Recent data explicitly demonstrate that these perturbations are tightly connected to enhanced function and overexpression of two cardiac sarcolemmal ion transporters, the Na+/Ca2+ exchanger (NCX1), a major regulator of the intracellular Ca2+ concentration ([Ca2+]i), and the Na+/H+ exchanger (NHE1), the primary regulator of intracellular pH. Recent efforts to delineate the real therapeutic value of NCX1 modulation were still hampered by the absence of specific and effective inhibitors and the total absence of specific activators. On the other side, several well-established inhibitors of the NHE1 are already in clinical practice or under evaluation but human data—promising in limiting the size of the ischemic injury—are much less convincing in preventing lethal ventricular arrhythmias. In this chapter, we aim to discuss the involvement of these two ion transporters in the regulation of the [Ca2+]i homeostasis in the healthy and diseased heart, to underline their principal contribution to the generation of ventricular arrhythmia, and to summarize experimental results obtained in studies in the direction of their antiarrhythmic and cardioprotective modulation.
Chapter
A small amount of Ca2+ enters the cell through the sarcolemmal membrane, releases an additional amount of Ca2+ from the sarcoplasmic reticulum, and initiates myocardial contraction. The contraction process is terminated upon lowering the cytosolic Ca2+ by accumulation in the sarcoplasmic reticulum as well as removal into the extracellular space through the sarcolemmal membrane. This article describes sarcolemmal L-type Ca2+ channel, Na+–Ca2+ exchanger, store-operated Ca2+ channels, Ca2+-pump ATPase, Ca2+/Mg2+ecto-ATPase, Na+-H+ exchanger, and Na+-K+ ATPase, which directly or indirectly regulate the movements of Ca2+ in cardiomyocytes. The pharmacological modulation of these Ca2+-handling proteins has been indicated to gain information regarding the regulation of intracellular Ca2+ in cardiomyocytes. Furthermore, alterations in the sarcolemmal Ca2+-handling proteins in different cardiovascular pathologies have been identified to emphasize the Ca2+-handling abnormalities in cardiomyocytes during the development of cardiac dysfunction in heart disease.
Article
Low oxygen bone marrow (BM) niches (~1%-4% low O2 ) provide critical signals for hematopoietic stem/progenitor cells (HSC/HSPCs). Our presented data are the first to investigate live, sorted HSC/HSPCs in their native low O2 conditions. Transcriptional and proteomic analysis uncovered differential Ca2+ regulation that correlated with overlapping phenotypic populations consisting of robust increases of cytosolic and mitochondrial Ca2+ , ABC transporter (ABCG2) expression and sodium/hydrogen exchanger (NHE1) expression in live, HSC/HSPCs remaining in constant low O2. We identified a novel Ca2+ high population in HSPCs predominantly detected in low O2 that displayed enhanced frequency of phenotypic LSK/LSKCD150 in low O2 replating assays compared to Ca2+ low populations. Inhibition of the Ca2+ regulator NHE1 (Cariporide) resulted in attenuation of both the low O2 induced Ca2+ high population and subsequent enhanced maintenance of phenotypic LSK and LSKCD150 during low O2 replating. These data reveal multiple levels of differential Ca2+ regulation in low O2 resulting in phenotypic, signaling, and functional consequences in HSC/HSPCs.
Article
Full-text available
The Na+/H+ exchange activity (NHE1 human isoform) is rapidly activated in response to growth factors and hyperosmotic stress. To get insight into the mechanism of NHE1 activation, we studied the direct interaction of a ubiquitous Ca2+-dependent regulatory factor, calmodulin (CaM) with NHE1. Binding experiments with CaM-Sepharose, as well as fluorescence measurements with dansylated CaM, revealed that the NHE1 cytoplasmic domain strongly binds CaM in a Ca2+-dependent manner. Fusion protein analysis with deletion mutants provided evidence for high (K-d similar to 20 nM) and intermediate (K-d similar to 350 nM) affinity CaM-binding sites located in neighboring regions of NHE1 (amino acids 636-656 and 657-700). To assess a regulatory role of CaM-binding sites, several cDNAs having deletion and point mutations in the high affinity site were generated and expressed in the exchanger-deficient fibroblast cell line PS120. Deletion and point mutations of positively charged residues of the high affinity CaM-binding site resulted in up to 50 and 80% reductions of cytoplasmic alkalinization caused by growth factors (alpha-thrombin, etc.) and 100 mM sucrose, respectively. In these mutants, the reduction in alkalinization was apparently in proportion to that of the CaM-binding ability. These results suggest that binding of Ca2+/CaM to the high affinity site is involved at least partly in the activation of NHE1 in response to different extracellular signals.
Article
Full-text available
To isolate a cDNA encoding Na+/H+ exchanger isoform 5 (NHE5), we screened a human spleen library using exon sequences of theNHE5 gene. Clones spanning 2.9 kilobase pairs were isolated; however, they contained several introns and were missing coding sequences at both the 5′ and 3′ ends. The missing 5′ sequences were obtained by 5′-rapid amplification of cDNA ends and by analysis of an NHE5 genomic clone, and the missing 3′ sequences were obtained by 3′-rapid amplification of cDNA ends. Polymerase chain reaction amplification of brain cDNA yielded products in which each of the introns had been correctly excised, whereas the introns were retained in products from spleen and testis, suggesting that the NHE5 transcripts expressed in these organs do not encode a functional transporter. The intron/exon organization of theNHE5 gene was analyzed and found to be very similar to that of the NHE3 gene. The NHE5 cDNA, which encodes an 896-amino acid protein that is most closely related to NHE3, was expressed in Na+/H+ exchanger-deficient fibroblasts and shown to mediate Na+/H+exchange activity. Northern blot analysis demonstrated that the mRNA encoding NHE5 is expressed in multiple regions of the brain, including hippocampus, consistent with the possibility that it regulates intracellular pH in hippocampal and other neurons.
Article
The ABC transporters are ubiquitous membrane proteins that couple adenosine triphosphate (ATP) hydrolysis to the translocation of diverse substrates across cell membranes. Clinically relevant examples are associated with cystic fibrosis and with multidrug resistance of pathogenic bacteria and cancer cells. Here, we report the crystal structure at 3.2 angstrom resolution of the Escherichia coli BtuCD protein, an ABC transporter mediating vitamin B_(12) uptake. The two ATP-binding cassettes (BtuD) are in close contact with each other, as are the two membrane-spanning subunits (BtuC); this arrangement is distinct from that observed for the E. coli lipid flippase MsbA. The BtuC subunits provide 20 transmembrane helices grouped around a translocation pathway that is closed to the cytoplasm by a gate region whereas the dimer arrangement of the BtuD subunits resembles the ATP-bound form of the Rad50 DNA repair enzyme. A prominent cytoplasmic loop of BtuC forms the contact region with the ATP-binding cassette and appears to represent a conserved motif among the ABC transporters.
Chapter
Every cell, whether prokaryotic or eukaryotic, uses the most common ions, H+ and Na+, to store and transduce energy. Since proteins can endure only certain concentrations of H+ and Na+, these ions, very often, become very potent stressors to all cells. Therefore every cell has efficient homeostatic mechanisms for Na+ and H+ (1). Proteins that play a primary role in these homeostatic mechanisms are Na+/H+ antiporters (2).
Article
Background —The transmembrane sodium/hydrogen exchanger maintains myocardial cell pH integrity during myocardial ischemia but paradoxically may precipitate cell necrosis. The development of cariporide, a potent and specific inhibitor of the exchanger, prompted this investigation of the potential of the drug to prevent myocardial cell necrosis. Methods and Results —A total of 11 590 patients with unstable angina or non–ST-elevation myocardial infarction (MI) or undergoing high-risk percutaneous or surgical revascularization were randomized to receive placebo or 1 of 3 doses of cariporide for the period of risk. The trial failed to document benefit of cariporide over placebo on the primary end point of death or MI assessed after 36 days. Doses of 20 and 80 mg every 8 hours had no effect, whereas a dose of 120 mg was associated with a 10% risk reduction (98% CI 5.5% to 23.4%, P =0.12). With this dose, benefit was limited to patients undergoing bypass surgery (risk reduction 25%, 95% CI 3.1% to 41.5%, P =0.03) and was maintained after 6 months. No effect was seen on mortality. The rate of Q-wave MI was reduced by 32% across all entry diagnostic groups (2.6% versus 1.8%, P =0.03), but the rate of non–Q-wave MI was reduced only in patients undergoing surgery (7.1% versus 3.8%, P =0.005). There were no increases in clinically serious adverse events. Conclusions —No significant benefit of cariporide could be demonstrated across a wide range of clinical situations of risk. The trial documented safety of the drug and suggested that a high degree of inhibition of the exchanger could prevent cell necrosis in settings of ischemia-reperfusion.
Article
The myocardial Na+/H+ exchange (NHE) represents a major mechanism for pH regulation during normal physiological processes but especially during ischaemia and early reperfusion. However, there is now very compelling evidence that its activation contributes to paradoxical induction of cell injury. The mechanism for this most probably reflects the fact that activation of the exchanger is closely coupled to Na+ influx and therefore to elevation in intracellular Ca2+ concentrations through the Na+/Ca2+exchange. The NHE is exquisitely sensitive to intracellular acidosis; however, other factors can also exhibit stimulatory effects viaphosphorylation-dependent processes. These generally represent various autocrine and paracrine as well as hormonal factors such as endothelin-1, angiotensin II and α1-adrenoceptor agonists, which probably act through receptor-signal transduction processes. Thus far, 6 NHE isoforms have been identified and designated as NHE1 through NHE6. All except NHE6, which is located intracellularly, are restricted to the sarcolemmal membrane. In the mammalian myocardium the NHE1 subtype is the predominant isoform, although NHE6 has also been identified in the heart. The predominance of NHE1 in the myocardium is of some importance since, as discussed in this review, pharmacological development of NHE inhibitors for cardiac therapeutics has concentrated specifically on those agents which are selective for NHE1. These agents, as well as the earlier nonspecific amiloride derivatives have now been extensively demonstrated to possess excellent cardioprotective properties, which appear to be superior to other strategies, including the extensively studied phenomenon of ischaemic preconditioning. Moreover, the salutary effects of NHE inhibitors have been demonstrated using a variety of experimental models as well as animal species suggesting that the role of the NHE in mediating injury is not species specific. The success of NHE inhibitors in experimental studies has led to clinical trials for the evaluation of these agents in high risk patients with coronary artery disease as well as in patients with acute myocardial infarction (MI). Recent evidence also suggests that NHE inhibition may be conducive to attenuating the remodelling process after MI, independently of infarct size reduction, and attenuation of subsequent postinfarction heart failure. As such, inhibitors of NHE offer substantial promise for clinical development for attenuation of both acute responses to myocardial as well as chronic postinfarction responses resulting in the evolution to heart failure.