Departments of 1 Ophthalmology and Visual Sciences, 2 Physiology, and 3 Human Genetics, University of Michigan, Ann Arbor, Michigan 48105; and 4 Department of Genetics, Yale University, New Haven, Connecticut 06520
![]() |
ABSTRACT |
---|
![]() ![]() ![]() ![]() ![]() ![]() ![]() |
---|
To identify novel potassium channel genes expressed in the retina, we screened a human retina cDNA library with an EST sequence showing partial homology to inwardly rectifying potassium (Kir) channel genes. The isolated cDNA yielded a 2,961-base pair sequence with the predicted open reading frame showing strong homology to the rat Kir2.4 (rKir2.4). Northern analysis of mRNA from human and bovine tissues showed preferential expression of Kir2.4 in the neural retina. In situ hybridization to sections of monkey retina detected Kir2.4 transcript in most retinal neurons. Somatic hybridization analysis and dual-color in situ hybridization to metaphase chromosomes mapped Kir2.4 to human chromosome 19 q13.1-q13.3. Expression of human Kir2.4 cRNA in Xenopus oocytes generated strong, inwardly rectifying K+ currents that were enhanced by extracellular alkalinization. We conclude that human Kir2.4 encodes an inwardly rectifying K+ channel that is preferentially expressed in the neural retina and that is sensitive to physiological changes in extracellular pH.
potassium channels; nucleotide sequence; chromosomal localization
![]() |
INTRODUCTION |
---|
![]() ![]() ![]() ![]() ![]() ![]() ![]() |
---|
INWARDLY RECTIFYING K+ channels (Kir) comprise a diverse subset of K+-selective channels that preferentially conduct K+ movement in the inward direction. This widely expressed group of plasma membrane proteins serves both homeostatic and specialized functions. In neurons, Kir channels are important in establishing membrane excitability and shaping action potentials, whereas in other cell types, such as epithelial and glial cells, they help set the membrane potential and play an integral role in ion transport processes.
To date, more than a dozen Kir subunits have been cloned and characterized (for review, see Ref. 22). Although these Kir subunits exhibit structural and functional differences, they share a highly conserved motif consisting of two membrane-spanning domains (M1 and M2) separated by a putative pore region (H5). Kir channels are formed by the assembly of four subunits in the plasma membrane (16) and can comprise either identical or two different types of Kir subunit (12). The ability of different subunits to form heterotetramers may be part of the explanation for the great diversity that exists in the kinetic and conductive properties of native Kir channels found in various cells and tissues.
Electrophysiological studies in cells isolated from the vertebrate retina have identified inwardly rectifying K+ currents in horizontal cells (44, 50), Müller cells (21, 30), and the retinal pigment epithelium (RPE) (20). Relatively little is known, however, about the molecular identity of the Kir subunits that make up the underlying conductances. To date, the only Kir subunit for which expression in the neural retina has been demonstrated is Kir4.1, a subunit originally cloned from rat brain (45) and later localized by in situ hybridization and immunohistochemistry to Müller cells (21) and the RPE (27). Functional differences between the cloned Kir4.1 channel and the inwardly rectifying K+ conductance present in horizontal cells suggest that other Kir subunits may be expressed in the retina.
Here, we report the cloning of the cDNA for an inwardly rectifying K+ channel subunit, human Kir2.4 (hKir2.4), from a human retinal cDNA library. We have demonstrated that transcripts for hKir2.4 are preferentially expressed in the neural retina and are present in most retinal neurons, and that its gene is localized on human chromosome 19 at q13.1-q13.3. We also have shown that expression of hKir2.4 cRNA in Xenopus oocytes generates a K+ conductance exhibiting strong inward rectification and sensitivity to extracellular pH. Preliminary accounts of this work have been presented in abstract form (25, 26).
![]() |
METHODS |
---|
![]() ![]() ![]() ![]() ![]() ![]() ![]() |
---|
Isolation of the cDNAs and Sequence Analysis
Methods used for routine recombinant DNA analysis were essentially as described in laboratory manuals (3, 35). The oligonucleotide primers used for sequencing and PCR experiments were synthesized by Genosys (Woodlands, TX). The Expressed Sequence Tag (EST) database was searched using BLAST 2.0 software (2) for sequences derived from human retinal cDNA libraries that showed homology to Kir channel subunits. The cDNA clone (I.M.A.G.E. clone no. 586857) for one of the identified ESTs (AA504908) was obtained from the repository of American Type Culture Collection (Bethesda, MD). A 1-kb fragment (EcoR I-EcoR I) representing the 3'-untranslated region of the EST was used as a probe to screen a human adult retina directional cDNA library (42) in Charon BS(Northern Analysis
For Northern analysis, we obtained a multiple tissue RNA blot from Clontech (Palo Alto, CA) containing 2 µg of poly(A+) human RNA per lane and prepared a second blot by loading 4 µg of poly(A+) RNA from adult human retina in one lane and 50 µg of total RNA from human fetal brain, placenta, fetus plus placenta, lung, thymus, and HeLa cells in the others. A third blot was prepared by loading 4 µg of poly(A+) RNA from bovine retina and RPE sheets. Poly(A+) RNA was obtained by extracting total RNA from neural retina or washed, dispase-dissociated RPE sheets with TRIzol (Life Technologies, Rockville, MD) and then applying it to an oligo(dT) column (Pharmacia, Piscataway, New Jersey). The blots were sequentially hybridized with a 1-kb probe (EcoR I-EcoR I fragment of EST AA504908) containing the 3'-untranslated region (nucleotides 1,962-2,961) of hKir2.4 cDNA and aIn Situ Hybridization
Adult rhesus monkey and bovine eyes were obtained within 1 h of death and transported to the laboratory on ice. After the anterior segment was dissected away, the vitreous was carefully removed by blunt dissection. Eyecups were sectioned into quarters and placed in 4% paraformaldehyde at 4°C overnight. Tissue pieces were then rinsed in 0.1 M phosphate buffer and cryoprotected as described elsewhere (4). Sections (13 µm) were cut atIn situ hybridization to monkey and bovine retinal sections was carried out using digoxigenin-labeled cRNA probes and a commercially available DIG Nucleic Acid Detection Kit (Boehringer Mannheim, Indianapolis, IN). For riboprobes, 700-bp Xho I-EcoR I and 1-kb EcoR I-EcoR I fragments of the EST plasmid (AA504908), representing the COOH-terminal coding region (nucleotides 1,321-1,962) and the 3'-untranslated region of hKir2.4 (nucleotides 1,962-2,961), respectively, were cloned into Bluescript vectors (pKS+ or pSK+). Sense and antisense riboprobes were generated from these plasmids in the presence of 11-digoxigenin UTP (Boehringer Mannheim) using either T3 or T7 RNA polymerase.
Chromosomal Localization
For somatic cell hybrid panel analysis, mapping panel no. 2, consisting of 24 human-rodent somatic cell hybrids, was obtained from the National Institute of General Medical Sciences (NIGMS) cell repository. Characterization and human chromosome content of these hybrids are described in detail in the NIGMS catalog. DNA samples were digested with Pst I, separated in 1% agarose gels by electrophoresis, transferred to nylon filters, and hybridized to a 32P-labeled hKir2.4 probe (the entire p17B clone sequence) as previously described (56) to identify human-specific genomic fragments that were then scored in the human-rodent hybrid panel. Regional assignment of the hKir2.4 gene was carried out by the fluorescence in situ hybridization (FISH) of metaphase chromosomes procedure, as previously described (37). Briefly, plasmid containing p17B cDNA was labeled by nick translation with digoxigenin-dUTP and hybridized to human metaphase chromosomes at a concentration of 50 ng/µl. Hybridization signal was detected by rhodamine-conjugated anti-digoxigenin antibody, and chromosomes were counterstained with 4,6-diamino-2-phenylindole. Localization was confirmed by dual-color FISH using p17B and P1 clone RMC19P009 (obtained from Joe W. Gray, University of California, San Francisco), which maps to the distal short arm of chromosome 19. The P1 clone was labeled with biotin-dUTP and detected by fluorescein avidin, whereas the plasmid was labeled with digoxigenin-dUTP and detected by rhodamine-anti-digoxigenin antibody.Expression in Oocytes
For the construction of a transcription plasmid, the coding region of hKir2.4 cDNA (nucleotides 395-1,760) was PCR amplified with the Expand High-Fidelity PCR system (Boehringer Mannheim) and subcloned into the polyadenylating transcription vector pBSTA (17) at the Bgl II site using a blunt-end ligation procedure. The resulting plasmid contained the cloned cDNA insert flanked by the 3'- and 5'-untranslated regions of theElectrophysiology
Xenopus laevis oocytes were surgically removed from adult female frogs and defolliculated by incubating clusters of oocytes in 0.2% type IV collagenase (Sigma Chemical) in calcium-free ND96 solution (96 mM NaCl, 2 mM KCl, 1 mM MgCl2, 10 mM HEPES, pH 7.4, 200 µg/ml gentamicin, and 550 µg/ml sodium pyruvate). Healthy-looking stage V-VI oocytes were collected and stored overnight at 18°C in ND96 solution plus 1 mM CaCl2.Defolliculated oocytes were injected with 0.1-5 ng of capped
hKir2.4 cRNA in 50 nl and incubated at 18°C for 1-3 days before electrophysiological experiments. Oocytes injected with the same volume
of DEPC-treated water served as controls. Whole cell currents were recorded using the two-electrode voltage-clamp technique (39). Thick-wall borosilicate glass microelectrodes
having impedance of 0.5-1.5 M when filled with 3 M KCl were
used as voltage-sensing and current-passing electrodes. Signals from
the current-passing electrode were amplified with a GeneClamp 500 amplifier (Axon Instruments, Burlingame, CA) and stored on a computer
hard disk for later analysis with pCLAMP 6.0 software (Axon
Instruments). The standard bath solutions for recording comprised
modified ND96 solution containing 96 mM NaCl, 2 mM KCl, 1 mM
CaCl2, 1 mM MgCl2, and 10 mM
2-(N-morpholino)ethanesulfonic acid (pH 5.5, 6.0, and 6.5),
10 mM HEPES (pH 7.0, 7.4, and 8.0), or 10 mM
3-([1,1-dimethyl-2-hydroxyethyl)amino]-2-hydroxypropanesulfonic acid
(pH 8.5 and 9.0). In some experiments, 300 µM niflumic acid was added to block endogenous Ca2+-activated
Cl
currents. In experiments investigating the
relationship between Kir2.4 current and extracellular K+
concentration, NaCl was replaced with equimolar amounts of KCl. For
determination of blocker sensitivity, various amounts of
Ba2+ or Cs+ were directly added to 98 mM KCl Ringer.
![]() |
RESULTS |
---|
![]() ![]() ![]() ![]() ![]() ![]() ![]() |
---|
Identification and Sequence Analysis of hKir2.4
A search of the EST database led to the identification of two overlapping human retinal EST clones (W25800 and AA504908) that showed partial homology with known Kir channel subunits. EST AA504908 was then used as a probe to screen >500,000 plaques of an adult human retina cDNA library, resulting in 13 positive phage clones after tertiary screening. Plasmid clones obtained from five of the positive phage clones were characterized by restriction analysis and nucleotide sequencing. All were found to contain the EST AA504908 sequence in their cDNA inserts, which ranged in size from 1.5 to 2.3 kb. One of the cDNA clones, p17B, contained the largest cDNA insert (2.3 kb) and was completely sequenced. Analysis of the sequencing data indicated that p17B contained the entire coding sequence for a Kir2.4 channel subunit as well as a partial 3' end sequence (Fig. 1A). Partial sequencing of the remaining four clones at the 3' and 5' ends revealed that they were smaller fragments of the same transcript from which p17B was derived. Sequencing of EST AA504908 cDNA revealed that it contained a partial coding sequence that overlapped completely with the 3' end of p17B as well as the complete 3' end sequence that included a large poly(A+) tail (Fig. 1A). The composite sequence of the hKir2.4 cDNA was obtained by assembling the p17B and EST AA504908 sequences using Lasergene software (DNASTAR). The full-length hKir2.4 cDNA was 2,961 bp long, with a 1,308-bp coding region flanked at the 5' and 3' ends by 406- and 1,247-bp untranslated regions, respectively (Fig. 1B).
|
Analysis of hKir2.4 cDNA indicated that it codes for a 436-amino
acid-long protein (Fig. 1B) containing major hydrophobic regions that correspond to two membrane-spanning domains, M1 and M2,
and an intervening H5 pore-forming region (Fig. 1C), which are the hallmarks of Kir subunit proteins (22). A
computer-assisted search for protein motifs in hKir2.4 revealed that it
contains one putative Asn glycosylation site at N193, one cAMP- and
cGMP-dependent phosphorylation site at S11, four protein kinase C (PKC)
phosphorylation sites at T263, S362, S376, and S422, and one tyrosine
kinase phosphorylation site at R240. Comparison of the predicted
protein sequence of the clone with that of other Kir channel subunits
indicated that it is most closely related to members of the Kir2
subfamily of K+ channels (Fig.
2). Amino acids D175 and E227, which have
been shown to be critical for conferring strong inward rectification in
all homotetrameric Kir2 channels via the interactions with Mg2+ and polyamines (13), are also conserved
in hKir2.4. In addition, hKir2.4 contains amino acids C127, H130, and
C159, which correspond to residues in Kir2.3 that confer sensitivity to
extracellular pH (8). The presence of these sequence
motifs suggests that the assembly of Kir2.4 subunits should give rise
to a strong inwardly rectifying potassium channel that is sensitive to
changes in extracellular pH (see pH sensitivity).
|
In addition to an extracellular pH sensor, Kir2.3 has three amino acids in the NH2 terminus (T53, Y57, M60) that have been shown to be responsible for intracellular pH sensitivity (34). This motif, however, is absent from hKir2.4. Unlike all other Kir2 subunits, the COOH terminus of hKir2.4 lacks an RRESXI domain, which contains both a protein kinase A (PKA) phosphorylation site that is involved in channel gating and a PDZ-binding domain that allows channel clustering (7, 51). Although hKir2.4 has a unique PKA phosphorylation consensus site at S11 that could potentially play a role in channel regulation, the lack of an RRESXI sequence seems to suggest that hKir2.4 might have a subcellular distribution different from that of other members of the Kir2 subfamily.
While these studies were in progress, the rat homolog of hKir2.4, rKir2.4, was identified from a rat brain cDNA library (46). Alignment of hKir2.4 and rKir2.4 protein sequences demonstrated a 92% identity (Fig. 2). The pore-forming region H5 and membrane-spanning domains M1 and M2 of these homologs are identical except for a conserved substitution of A105 in the M1 domain of hKir2.4 with T103 in rKir2.4. There is also a substitution of Q138 in rKir2.4 for H130 in hKir2.4, but it can be inferred from the results of site-directed mutagenesis studies on Kir2.3 that this amino acid difference would not affect pH sensitivity (8). Except for an extra PKC phosphorylation site at S422 and five nonconserved amino acid substitutions in nonfunctional regions, all of the other substitutions are either conserved or similar. Alignment of cDNA sequences reveals that whereas the coding regions of hKir2.4 cDNA and rKir2.4 cDNA are 86.9% identical, their 5'- and 3'-untranslated regions share only 29.1% and 39% sequence identities (data not shown).
Expression Pattern of hKir2.4
The expression of hKir2.4 mRNA was evaluated by Northern blot analysis. RNA blots of several human tissues/cell lines, including adult retina, were hybridized with a 32P-labeled probe, which was generated from a 1-kb fragment representing the untranslated region of hKir2.4 (nucleotides 1,962-2,961) (see Fig. 1). The hKir2.4 probe hybridized to a major transcript at ~3 kb in the retina RNA lane only (Fig. 3, A and B, top), which corresponds to the size of the full-length cDNA clone (2.9 kb). A second signal at ~5 kb was also detected, suggesting the possibility of alternative polyadenylation or splicing. The integrity and the relative amounts of RNA in the total RNA samples were established by the hybridization of the same blot with a 32P-labeled
|
Localization of Kir2.4 Transcripts in the Retina
To determine the cellular distribution of hKir2.4 in the retina, we performed in situ hybridization on embedded adult monkey and bovine retinal sections using sense and antisense probes recognizing the 3'-untranslated region or the COOH-terminal coding sequence. Figure 4 shows a Nomarski photomicrograph of monkey peripheral retina hybridized with antisense probes corresponding to the COOH-terminal coding region of hKir2.4. A strong hybridization signal was observed in the outer nuclear layer, which contains the cell bodies of rod and cone photoreceptors, in the inner nuclear layer, which includes the cell bodies of bipolar, amacrine, horizontal, and Müller cells, and also in some of the cell bodies located in the ganglion cell layer (Fig. 4A). In situ hybridization with the complementary sense probe demonstrated no detectable signal (Fig. 4B). A similar pattern of hybridization was observed with the antisense and sense probes for the 3'-untranslated region of the hKir2.4 in monkey and bovine retina sections, confirming the specificity of hybridization to Kir2.4 transcripts (data not shown). Hence, Kir2.4 transcript appears to be present in most retinal neurons and possibly Müller cells as well.
|
Chromosomal Localization of the Kir2.4 Gene
Southern blot hybridization of a 32P-labeled hKir2.4 probe with Pst I-digested genomic DNA from a human-rodent somatic cell hybrid panel was carried out to identify human-specific genomic fragments that were scored in the human-rodent hybrid panel (data not shown). Three human-specific fragments of 5.3, 3.2, and 3 kb, a mouse-specific band of 1.35 kb, and a Chinese hamster band of 1.8 kb were detected. All three human-specific fragments showed perfect segregation with human chromosome 19. These results of discordance analysis indicated the segregation of the hKir2.4 gene with the human chromosome 19. Regional assignment of the hKir2.4 gene was accomplished by dual-color FISH experiments using the probe for hKir2.4 and a probe specific for 19p. The results indicated that hKir2.4 maps to human chromosome 19q13.1-q13.3 (Fig. 5).
|
Electrophysiology of Heterologously Expressed hKir2.4 in Xenopus Oocytes
Xenopus oocytes injected with hKir2.4 cRNA generally developed large negative membrane potentials, indicating the expression of functional Kir channels. Oocytes injected with 0.1-10 ng of hKir2.4 cRNA had an average zero-current potential (V0) of
|
Dependence on extracellular K+
concentration.
As has been well documented for other Kir channels (24,
45), the magnitude of hKir2.4 currents was strongly dependent on
the extracellular K+ concentration
([K+]o). Figure
7A shows two families of
hKir2.4 currents recorded in the same oocyte superfused first with 2 mM
K+ and then with 98 mM K+. Elevation of
[K+]o produced more than a 10-fold increase
in inward current. Figure 7B summarizes the results of
experiments on six hKir2.4 cRNA-injected oocytes, each bathed with a
series of [K+]o concentrations
(Na+ replacement). Increasing
[K+]o produced a marked increase in inward
but not outward currents. The I-V relationship also became
more linear at higher [K+]o as a result of a
decrease in voltage-dependent block by extracellular Na+.
As has been found with other strong inwardly rectifying K+
channels (22), the inward slope conductance of hKir2.4 was roughly proportional to the square root of
[K+]o (Fig. 7C). A plot of
V0 as a function of log [K+]
yielded a slope of 49 mV per decade change in [K+] (Fig.
7D), which is somewhat less than the 58 mV predicted by the
Nernst equation. Because V0 was not affected by
the substitution of extracellular Na+ with
NMDG+ (not shown), this lower value does not reflect a
significant Na+ permeability of hKir2.4 channels but
probably resulted from the presence of endogenous Cl
channels. Hence, we conclude that hKir2.4 is a K+-selective
channel.
|
Time course of
[K+]o-induced changes in
hKir2.4 current.
When the solution bathing the oocytes expressing hKir2.4 was
switched from 2 mM to 98 mM K+, the inward current
typically increased in a biphasic manner (Fig.
8A, top). The first
phase, which occurred within the first minute of the solution change,
coincided with the change in V0 (Fig.
8A, bottom) and reflected the time course of the
change in [K+] outside the oocyte. This was followed by a
slower, second phase of current increase that continued for the next 5 min, resulting in nearly a doubling of current (Fig.
8B). This phenomenon appears to be particular to
the hKir2.4 channel because in identical experiments on oocytes
injected with cRNA encoding Kir7.1, exposure to high [K+]o produced only a single, rapid increase
in inward current (data not shown). After the bathing solution was
briefly returned to 2 mM K+, the oocyte was exposed to 98 mM K+ a second time. The inward current rapidly increased
to a level that was slightly greater than that before the superfusion
of 2 mM K+ and then continued to increase slowly. These and
similar results in other oocytes are consistent with the idea that
exposure to 98 mM K+ triggers an increase in macroscopic
hKir2.4 conductance by a mechanism that is not rapidly reversible.
Therefore, in subsequent experiments testing the effects of
Cs+, Ba2+, or H+, each test
measurement was bracketed by measurements in control solution to
correct for time-dependent increases in hKir2.4 current.
|
Inhibition by Cs+.
Inwardly rectifying K+ channels are generally blocked to
various degrees by extracellular Cs+ and Ba2+.
To determine the pharmacological sensitivity of the hKir2.4 channel, we
measured currents in hKir2.4 cRNA-injected oocytes bathed with various
concentrations of blocker ions. Figure
9A shows a family of currents
recorded in an oocyte bathed with 100 µM Cs+. Currents
were evoked by voltage steps from a holding potential of 0 mV to
membrane potentials in the range between +30 and 130 mV. Stepping the
membrane potential from 0 mV to more negative potentials produced a
rapid block with a time course that was faster than the capacitive
transient (<2 ms). Figure 9B summarizes the results of
experiments on five oocytes and plots normalized steady-state currents
measured in the presence and absence of Cs+ as a function
of membrane voltage. The block by Cs+ was strongly voltage
dependent, with the fraction of current blocked increasing as the
membrane was made more negative. At potentials negative to
100 mV,
the fractional block decreased, perhaps as a result of Cs+
permeation through the channel. To quantify the concentration and
voltage dependence of Cs+ block, we plotted the ratio of
the steady-state currents measured in the presence and absence of
Cs+ as a function of Cs+ concentration. Figure
9C shows a family of dose-response curves for the
Cs+-induced block of hKir2.4 current measured at various
potentials. The symbols represent mean values, and the smooth
curves are the least-squares fits of the data to the first-order
equation
![]() |
(1) |
![]() |
(2) |
|
Inhibition by Ba2+.
Compared with the block by Cs+, the
Ba2+-induced block of hKir2.4 currents was more weakly
voltage dependent. Figure
10A shows that 100 µM
Ba2+ in the bath produced a time-dependent block of inward
currents. Figure 10B summarizes the results of experiments
on five oocytes and plots the normalized steady-state I-V
relationships in the presence of various concentrations of
Ba2+. The block by Ba2+ was mildly voltage
dependent, resulting in a slight curvature in the I-V
relationship at voltages more negative than 80 mV. Figure
10C summarizes the concentration and voltage dependence of
the Ba2+-induced block of hKir2.4 currents. The symbols
represent mean values of normalized current, and the smooth curves are
the least-squares fits of the data to Eq. 1. The
KD at
100 mV was 116.0 ± 8.9 µM, and
it increased slightly at more positive potentials, as can be seen by
the rightward shift in the dose-response curve. Figure 10D
plots the average value of log KD as a function
of voltage, and the straight line is the least-squares fit of the data
to Eq. 2. The results indicate a 10-fold change in
KD per 166-mV change in membrane voltage and
imply that the fractional distance of the binding site in the electric
field is 0.18. Hence, Ba2+ appears to act at some
superficial site near the outside of the channel.
|
pH sensitivity.
As mentioned in Identification and Sequence Analysis of
hKir2.4, analysis of the predicted amino acid sequence of
hKir2.4 revealed that it contains residues in the extracellular linker between the M1 and H5 domains that correspond to residues in human Kir2.3 that are thought to confer extracellular pH sensitivity (8). To test the possibility that hKir2.4 channels are
also sensitive to changes in external pH, we measured hKir2.4 currents while superfusing oocytes with solutions buffered to a range of pH
values. We carried out these experiments with 10 mM K+ in
the bath so that possible effects of extracellular pH on endogenous currents could be detected as a change in V0. In
addition, Na+ was omitted (NMDG substitution) to avoid
possible complications arising from interactions between protons and
the Na+ binding site. Figure
11A shows currents recorded
in response to a series of voltage steps from a holding potential of 0 mV to membrane potentials in the range from +50 to 150 mV. Compared with currents recorded in our standard solution buffered to pH 7.4, hKir2.4 currents were significantly smaller at pH 6.5 and larger at pH
8.5. In contrast, manipulations expected to alter intracellular pH,
such as exposure to 10 mM NH4+, 50 mM acetate, or
CO2/HCO3
at constant extracellular pH,
had no significant effect on hKir2.4 currents. The time course of
currents following voltage steps was not affected by pH, indicating
that modulation of hKir2.4 currents by extracellular protons was
essentially voltage independent. This conclusion is supported by the
finding that the fraction of current increased or decreased by the
extracellular pH change was essentially independent of voltage (Fig.
11B) and suggests that the regulation of hKir2.4 channel
involves the binding of protons to some external site on the channel
protein.
|
![]() |
DISCUSSION |
---|
![]() ![]() ![]() ![]() ![]() ![]() ![]() |
---|
We have described the cloning, expression pattern, chromosomal localization, and functional characterization of a retinal inwardly rectifying K+ channel. Analysis of the predicted amino acid sequence indicated that it is a member of the Kir2 subfamily of K+ channel subunits. Subsequent to our initial report (25), the rat homolog of this gene (rKir2.4) was identified from a rat brain cDNA library (46). The human and rat homologs of Kir2.4 share a high degree of identity (92%), with some differences in their amino and carboxy-terminal regions. Because the amino acid substitutions between these homologs are mostly conserved or lie outside known functional motifs, it seems unlikely that these differences have any functional significance.
The heterologous expression of hKir2.4 in Xenopus oocytes
resulted in functional channels with properties similar to those of
other members of the Kir2 subfamily, including strong inward rectification, a conductance that is roughly proportional to the square
root of the extracellular K+ concentration, and
susceptibility to block by extracellular Cs+ and
Ba2+. As reported previously for rKir2.4 (46),
the block of hKir2.4 currents by Cs+ was more voltage
dependent than the block by Ba2+. Specifically, a 10-fold
change in KD was produced by a 35-mV change in
membrane potential for the Cs+ block, but an equivalent
change in the KD for the Ba2+ block
required a 166-mV change in membrane potential. We also confirmed that,
compared with other Kir2 channels, hKir2.4 channels required a 20-fold
higher concentration Ba2+ to block 50% of the inward
current; this result suggests a binding site with an unusually low
affinity for this ion. In contrast, the concentration dependence of the
Cs+ block of hKir2.4 channels at 80 mV indicated an
affinity similar to that of Kir2.1 channels (24, 46). In
this regard, our results differ from those of Topert et al.
(46), who reported a Cs+ affinity for rKir2.4
channels that is lower by a factor of 30-50 compared with that for
other Kir2 channels. The reason for this difference between hKir2.4 and
rKir2.4 channels is not known.
A previously unrecognized property of the Kir2.4 channel is its modulation by changes in extracellular pH in the physiological range. We found that extracellular alkalinization enhanced, whereas extracellular acidification diminished, macroscopic hKir2.4 currents, with an apparent pKa of 7.14. This behavior is similar to that of Kir2.3 (HIR), for which the molecular determinant of pH sensitivity has been shown to be a histidine residue (H117) in the M1-to-H5 linker that influences one of two titratable cysteine residues (114 and 146) located in the M1-to-H5 and H5-to-M2 linkers (8). It has also been shown that extracellular Zn2+ blocks Kir2.3 currents in a voltage-independent and pH-sensitive manner, which indicates that Zn2+ binds to the same residue as H+. It seems likely that the modulation of hKir2.4 by extracellular pH involves a similar mechanism, because its predicted amino acid sequence contains the same critical residues in corresponding locations (C127, H130, and C159). In support of this idea, we have found that hKir2.4 currents are also blocked by extracellular Zn2+ in micromolar concentrations (data not shown).
Another remarkable characteristic of hKir2.4 channels expressed in oocytes is that exposure to high [K+]o solution triggers a slow increase in macroscopic current. This property has not been described for any other Kir2 channel (including rKir2.4), but it superficially resembles the high [K+]-induced activation of Kir1.1 (9) and Kir4.2 channels (33) expressed in oocytes. Although the activation of these other cloned channels is thought to involve allosteric regulation by external K+, the mechanism by which hKir2.4 currents are activated may be different, because the conductance increase was not rapidly reversed on return to low [K+]o (see Fig. 8).
Northern blot analysis of human RNA indicated the presence of Kir2.4 transcripts only in the retina, with no detectable signal in any other tissues. To our knowledge, this is the first report of an inwardly rectifying K+ channel with this particular tissue distribution. Our results contrast with those of Topert and colleagues (46), who recently reported that Kir2.4 transcript is expressed predominantly in human and rat brain but not in significant amounts in the retina. This discrepancy may be due to differences in both the selection of probes and the hybridization conditions. Whereas we hybridized under very stringent conditions (65°C in Express-Hyb solution) using a highly selective probe synthesized using the 3'-untranslated region of hKir2.4 cDNA, Topert et al. used a probe representing the coding region of Kir2.4 cDNA and hybridized under lower stringency conditions (42°C in Express-Hyb solution). Hence, it is possible that Topert et al. detected transcripts for other K+ channel subtypes sharing homology with Kir2.4. Regardless, our finding that hKir2.4 probe hybridizes to bovine as well as human retinal poly(A+) mRNA strengthens our conclusion that Kir2.4 transcript is expressed in the retina.
The hybridization of hKir2.4 probe revealed 3- and 5-kb bands in Northern blots of human retinal mRNA, suggesting the possibility that different isoforms of Kir2.4 may be expressed in the retina. Topert et al. (46) reached a similar conclusion from their observation that their probe hybridized to two transcripts, 3 and 4 kb in size. Although there is no evidence for alternative splicing of other members of the Kir2 subfamily, alternatively spliced variants of ROMK (Kir1.1) channels with varying expression pattern and channel properties have been demonstrated (5).
In situ hybridization to sections of monkey retina indicated the presence of Kir2.4 transcripts in the perinuclear region of most cells in the neural retina. On the basis of this expression pattern, one might expect that retinal neurons should exhibit inwardly rectifying K+ channels with properties similar to those of the cloned Kir2.4 channel, i.e., strong inward rectification, high selectivity to K+, block by Cs+ and Ba2+, and sensitivity to changes in extracellular pH. The inwardly rectifying K+ current is a major component of the Müller cell membrane properties (30), but recent evidence strongly suggests that the underlying conductance is comprised of Kir4.1 channels (21). Inwardly rectifying K+ currents displaying many of the same properties as Kir2.4 have been documented in horizontal cells, where they help shape light-evoked responses (50). Takahashi and Copenhagen (44) have demonstrated that the inwardly rectifying K+ current in isolated catfish horizontal cells is enhanced by intracellular alkalinization, but there is no published information on the sensitivity of this current to changes in extracellular pH. Further studies are needed to determine the relationship between Kir2.4 and the inwardly rectifying K+ channel in horizontal cells.
Apart from horizontal cells, no other retinal neuron displays the classic inwardly rectifying K+ current. Another current exhibiting inward rectification, the hyperpolarization-activated current, or Ih, has been identified in photoreceptor inner segments (18, 28, 53), amacrine cells (14), bipolar cells (23), and retinal ganglion cells (43). The channels underlying these currents, however, are distinct from Kir channels in that they are blocked by Cs+ but not by Ba2+ and that they are permeable to both K+ and Na+. How does one reconcile the presence of Kir2.4 transcript in retinal neurons that apparently lack inward rectifier K+ currents? One possible explanation is that Kir2.4 channels are localized to synaptic terminals or are expressed in such low abundance that their contribution to whole cell currents is relatively small and therefore difficult to detect. Another possibility is that Kir2.4 channel subunits assemble with an inhibitory subunit. Negative interactions have been demonstrated between Kir2.2 and Kir2.2v subunits (29) and between Kir4.1 and Kir3.4 subunits (47); in both cases, the expression of homomeric channels with one of type of subunit alone (Kir2.2 or Kir4.1) produces inwardly rectifying currents, but coexpression of one subunit with the other produced no current. It will be interesting to learn what other Kir channel subunits are expressed in retinal neurons and whether they can have inhibitory interactions with Kir2.4.
Extracellular pH in the retina changes as a function of light adaptation (54) and hypoxia (55). Our finding that hKir2.4 currents are steeply extracellular pH dependent with a pKa of about pH 7.14 implies that the function of horizontal cells (or any other retinal cell expressing Kir2.4) may be modulated under physiological and pathophysiological conditions.
Mutations in the K+ channel genes have been shown to cause several inherited diseases (10, 36), including long QT syndrome (48, 49), Bartter syndrome (38), and episodic ataxia/myokymia syndrome (6). In addition, mutations in the H5 pore region of the Kir3.2 result in degeneration of cerebellar granule cells and cause ataxia in weaver mice (32). The importance of Kir2.4 channels in determining membrane properties and its expression in retinal cells suggest that mutations in Kir2.4 might lead to hereditary retinal disease. The localization of hKir2.4 gene to human chromosome 19 at q13.1-13.3 makes it an attractive candidate for inherited eye diseases that map to this chromosomal region. The genetic loci for recessive optic atrophy with ataxia (31), dominant cone-rod dystrophy (CORD2) (11), and a form of retinitis pigmentosa (RP11) (1) have been mapped to 19q13. Although the mutations in the cone-rod homeobox (CRX) gene have been identified in families with cone-rod dystrophy (15, 40) and Leber congenital amaurosis (41), the original CORD2 family did not reveal a CRX mutation, indicating that mutations in another gene at 19q13 may also cause this disease. PCR-based analysis, however, indicates that Kir2.4 gene is not present in the YAC Contig spanning RP11 locus (S. S. Bhattacharya, personal communication). Whether Kir2.4 is a candidate gene for CORD2 or any other genetic disorder remains to be evaluated.
![]() |
ACKNOWLEDGEMENTS |
---|
We thank Dr. Peter F. Hitchcock, Deborah C. Otteson, and Mitchell Gillett for help with in situ hybridization experiments and Dr. Alan Goldin (University of California, San Diego) for kindly providing the pBSTA plasmid.
![]() |
FOOTNOTES |
---|
This work was supported by National Eye Institute Grants EY-08850 and EY-07703 (to B. A. Hughes), the Michigan Eye Bank (to G. Kumar), Fight for Sight (to G. Kumar), a Research to Prevent Blindness Lew Wasserman Award (to A. Swaroop), and the Foundation Fighting Blindness (to A. Swaroop and B. A. Hughes).
Address for reprint requests and other correspondence: B. A. Hughes, Dept. of Ophthalmology, Univ. of Michigan, Kellogg Eye Center, 1000 Wall St., Ann Arbor, MI 48105 (E-mail: bhughes{at}umich.edu).
The costs of publication of this article were defrayed in part by the payment of page charges. The article must therefore be hereby marked "advertisement" in accordance with 18 U.S.C. §1734 solely to indicate this fact.
Received 22 January 2000; accepted in final form 3 April 2000.
![]() |
REFERENCES |
---|
![]() ![]() ![]() ![]() ![]() ![]() ![]() |
---|
1.
Al-Maghtheh, M,
Inglehearn CF,
Keen TJ,
Evans K,
Moore AT,
Jay M,
Bird AC,
and
Bhattacharya SS.
Identification of a sixth locus for autosomal dominant retinitis pigmentosa on chromosome 19.
Hum Mol Genet
3:
351-354,
1994[Abstract].
2.
Altschul, SF,
Gish W,
Miller W,
Myers EW,
and
Lipman DJ.
Basic local alignment search tool.
J Mol Biol
215:
403-410,
1990[ISI][Medline].
3.
Ausubel, FM,
Brent R,
Kingston RE,
Moore DD,
Seidman JG,
Smith JA,
and
Struhl K.
Current Protocols in Molecular Biology. New York: Wiley, 1996.
4.
Barthel, LK,
and
Raymond PA.
Subcellular localization of tubulin and opsin mRNA in the goldfish retina using digoxigenin-labeled cRNA probes detected by alkaline phosphatase and HRP histochemistry.
J Neurosci Methods
50:
145-152,
1993[ISI][Medline].
5.
Boim, MA,
Ho K,
Shuck ME,
Bienkowski MJ,
Block JH,
Slightom JL,
Yang Y,
Brenner BM,
and
Hebert SC.
ROMK inwardly rectifying ATP-sensitive K+ channel. II. Cloning and distribution of alternative forms.
Am J Physiol Renal Fluid Electrolyte Physiol
268:
F1132-F1140,
1995
6.
Browne, DL,
Gancher ST,
Nutt JG,
Brunt ER,
Smith EA,
Kramer P,
and
Litt M.
Episodic ataxia/myokymia syndrome is associated with point mutations in the human potassium channel gene, KCNA1.
Nat Genet
8:
136-140,
1994[ISI][Medline].
7.
Cohen, NA,
Brenman JE,
Snyder SH,
and
Bredt DS.
Binding of the inward rectifier K+ channel Kir2.3 to PSD-95 is regulated by protein kinase A phosphorylation.
Neuron
17:
759-767,
1996[ISI][Medline].
8.
Coulter, KL,
Perier F,
Radeke CM,
and
Vandenberg CA.
Identification and molecular localization of a pH-sensing domain for the inward rectifier potassium channel HIR.
Neuron
15:
1157-1168,
1995[ISI][Medline].
9.
Doi, T,
Fakler B,
Schultz JH,
Schulte U,
Brandle U,
Weidemann S,
Zenner HP,
Lang F,
and
Ruppersberg JP.
Extracellular K+ and intracellular pH allosterically regulate renal Kir1.1 channels.
J Biol Chem
271:
17261-17266,
1996
10.
Doyle, JL,
and
Stubbs L.
Ataxia, arrhythmia and ion-channel defects.
Trends Genet
14:
92-98,
1998[ISI][Medline].
11.
Evans, K,
Fryer A,
Inglehearn C,
Duvall-Young J,
Whittaker JL,
Gregory CY,
Butler R,
Ebenezer N,
Hunt DM,
and
Bhattacharya S.
Genetic linkage of cone-rod retinal dystrophy to chromosome 19q and evidence for segregation distortion.
Nat Genet
6:
210-213,
1994[ISI][Medline].
12.
Fakler, B,
Bond CT,
Adelman JP,
and
Ruppersberg JP.
Heterooligomeric assembly of inward-rectifier K channels from subunits of different subfamilies: Kir2.1 (IRK1) and Kir41 (BIR10).
Pflügers Arch
433:
77-83,
1995[ISI].
13.
Fakler, B,
Brandle U,
Glowatzki E,
Weidemann HP,
and
Ruppersberg JP.
Strong voltage dependent inward rectification of inward rectifier K+ channels is caused by intracellular spermine.
Cell
80:
149-154,
1995[ISI][Medline].
14.
Feigenspan, A,
Gustincich S,
Bean BP,
and
Raviola E.
Spontaneous activity of solitary dopaminergic cells of the retina.
J Neurosci
18:
6776-6789,
1998
15.
Freund, CL,
Gregory-Evans CY,
Furukawa T,
Papaioannou M,
Looser J,
Ploder L,
Bellingham J,
Ng D,
Herbrick JA,
Duncan A,
Scherer SW,
Tsui LC,
Loutradis-Anagnostou A,
Jacobson SG,
Cepko CL,
Bhattacharya SS,
and
McInnes RR.
Cone-rod dystrophy due to mutations in a novel photoreceptor-specific homeobox gene (CRX) essential for maintenance of the photoreceptor.
Cell
91:
543-553,
1997[ISI][Medline].
16.
Glowatzki, E,
Fakler G,
Brandle U,
Rexhaausen U,
Zenner HP,
Ruppersberg JP,
and
Fakler B.
Subunit-dependent assembly of inward-rectifier K+ channels.
Proc R Soc Lond B Biol Sci
261:
251-261,
1995[ISI][Medline].
17.
Goldin, AL.
Maintenance of Xenopus laevis and oocyte injection.
Methods Enzymol
207:
266-279,
1992[ISI][Medline].
18.
Hestrin, S.
The properties and function of inward rectification in rod photoreceptors of the tiger salamander.
J Physiol (Lond)
390:
319-333,
1987[Abstract].
19.
Hille, B.
Ionic Channels in Excitable Membranes. Sunderland, MA: Sinauer, 1992.
20.
Hughes, BA,
and
Takahira M.
Inwardly rectifying K+ currents in isolated human retinal pigment epithelial cells.
Invest Ophthalmol Vis Sci
37:
1125-1139,
1996[Abstract].
21.
Ishii, M,
Horio Y,
Tada Y,
Hibino H,
Inanobe A,
Ito M,
Yamada M,
Gotow T,
Uchiyama Y,
and
Kurachi Y.
Expression and clustered distribution of an inwardly rectifying potassium channel, Kir 4.1, on mammalian retinal Mueller cell membrane: their regulation by insulin and laminin signals.
J Neurosci
17:
7725-7735,
1997
22.
Isomoto, S,
Kondo C,
and
Kurachi Y.
Inwardly rectifying potassium channels: their molecular heterogeneity and function.
Jpn J Physiol
47:
11-39,
1997[ISI][Medline].
23.
Kaneko, A,
and
Tachibana MA.
Voltage-clamp analysis of membrane currents in solitary bipolar cells dissociated from Carassius auratus.
J Physiol (Lond)
358:
131-152,
1985[Abstract].
24.
Kubo, Y,
Baldwin TJ,
Jan YN,
and
Jan LY.
Primary structure and functional expression of a mouse inward rectifier channel.
Nature
362:
127-133,
1993[ISI][Medline].
25.
Kumar, G,
Swaminathan A,
Swaroop A,
and
Hughes BA.
Characterization of a putative potassium channel gene expressed in adult human retina (Abstract).
Invest Ophthalmol Vis Sci
39:
S42,
1998.
26.
Kumar, G,
Yan D,
Swaminathan A,
Yuan Y,
Sharma A,
Swaroop A,
and
Hughes BA.
Localization and functional characterization of a human inward rectifier K+ channel, hKir2.4 (Abstract).
Invest Ophthalmol Vis Sci
40:
S214,
1999[ISI].
27.
Kusaka, S,
Horio Y,
Fujita A,
Matsushita K,
Inanobe A,
Gotow T,
Uchiyama Y,
Tano Y,
and
Kurachi Y.
Expression and polarized distribution of an inwardly rectifying K+ channel, Kir4.1, in rat retinal pigment epithelium.
J Physiol (Lond)
520:
373-381,
1999
28.
Maricq, AV,
and
Korenbrot JI.
Inward rectification in the inner segment of single retinal cone photoreceptors.
J Neurophysiol
64:
1917-1928,
1990
29.
Namba, N,
Inagaki N,
Gonoi T,
Seino Y,
and
Seino S.
Kir2.2v: a possible negative regulator of the inwardly rectifying K+ channel Kir2.2.
FEBS Lett
386:
211-214,
1996[ISI][Medline].
30.
Newman, EA.
Inward-rectifying potassium channels in retinal glial (Müller) cells.
J Neurosci
13:
3333-3345,
1993[Abstract].
31.
Nystuen, A,
Costeff H,
Elpeleg ON,
Apter N,
Bonné-Tamir B,
Mohrenweiser H,
Haider N,
Stone EM,
and
Sheffield VC.
Iraqi-Jewish kindreds with optic atrophy plus (3-methylglutaconic aciduria type 3) demonstrate linkage disequilibrium with the CTG repeat in the 3' untranslated region of the myotonic dystrophy protein kinase gene.
Hum Mol Genet
6:
563-569,
1997
32.
Patil, N,
Cox DR,
Bhat D,
Faham M,
Myers RM,
and
Peterson AS.
A potassium channel mutation in weaver mice implicates membrane excitability in granule cell differentiation.
Nat Genet
11:
126-129,
1995[ISI][Medline].
33.
Pearson, WL,
Dourado M,
Schreiber M,
Salkoff L,
and
Nichols CG.
Expression of functional Kir4 family inward rectifier K+ channel from a gene cloned from mouse liver.
J Physiol (Lond)
514:
639-653,
1999
34.
Qu, Z,
Zhu G,
Yang S,
Cui N,
Li Y,
Chanchenvalap S,
Sulaiman S,
Haynie H,
and
Jiang C.
Identification of a critical motif responsible for gating of Kir2.3 channel by intracellular protons.
J Biol Chem
274:
13783-13789,
1999
35.
Sambrook, J,
Fritsch EF,
and
Maniatis T.
Molecular Cloning, A Laboratory Manual. Cold Spring Harbor, NY: Cold Spring Harbor Laboratory, 1989.
36.
Sanguinetti, MC,
and
Spector PS. .
Potassium channelopathies.
Neuropharmacology
36:
755-762,
1997[ISI][Medline].
37.
Shaper, NL,
Lin SP,
Joziasse DH,
Kim DY,
and
Yang-Feng TL.
Assignment of two human alpha-1,3-galactosyltransferase gene sequences (GGTA1 and GGTA1P) to chromosomes 9q33-q34 and 12q14-q15.
Genomics
12:
613-615,
1992[ISI][Medline].
38.
Simon, DB,
Karet FE,
Rodriguez-Soriano J,
Hamdan JH,
DiPietro A,
Trachtman H,
Sanjad SA,
and
Lifton RP.
Genetic heterogeneity of Bartter's syndrome revealed by mutations in the K+ channel, ROMK.
Nat Genet
14:
152-156,
1996[ISI][Medline].
39.
Stühmer, W.
Electrophysiological recording in Xenopus oocytes.
Methods Enzymol
207:
319-339,
1992[ISI][Medline].
40.
Swain, PK,
Chen S,
Wang QL,
Affatigato LM,
Coats CL,
Brady KD,
Fishman GA,
Jacobson SG,
Swaroop A,
Stone E,
Sieving PA,
and
Zack DJ.
Mutations in the cone-rod homeobox gene are associated with the cone-rod dystrophy photoreceptor degeneration.
Neuron
19:
1329-1336,
1997[ISI][Medline].
41.
Swaroop, A,
Wang QL,
Wu W,
Cook J,
Coats C,
Xu S,
Chen S,
Zack DJ,
and
Sieving PA.
Leber congenital amaurosis caused by a homozygous mutation (R90W) in the homeodomain of the retinal transcription factor CRX: direct evidence for the involvement of CRX in the development of photoreceptor function.
Hum Mol Genet
8:
299-305,
1999
42.
Swaroop, A,
and
Xu J.
cDNA libraries from human tissues and cell lines.
Cytogenet Cell Genet
64:
292-294,
1993[ISI][Medline].
43.
Tabata, T,
and
Ishida AT.
Transient and sustained depolarization of retinal ganglion cells by Ih.
J Neurophysiol
75:
1982-1943,
1996
44.
Takahashi, K,
and
Copenhagen DR.
Intracellular alkalinization enhances inward rectifier K+ current in retinal horizontal cells of catfish.
Zoolog Sci
12:
29-34,
1995[ISI][Medline].
45.
Takumi, T,
Ishii T,
Horio Y,
Morishige KI,
Takahashi N,
Yamada M,
Yamashita T,
Kiyama H,
Sohmiya K,
Nakanishi S,
and
Kurachi Y.
A novel ATP-dependent inward rectifier potassium channel expressed predominately in glial cells.
J Biol Chem
270:
16339-16346,
1995
46.
Topert, C,
Doring F,
Wischmeyer E,
Karschin C,
Brockhaus J,
Ballanyi K,
Derest C,
and
Karschin A.
Kir2.4: a novel K+ inward rectifier channel associated with motoneurons of cranial nerve nuclei.
J Neurosci
18:
4096-4105,
1998
47.
Tucker, SJ,
Bond CT,
Herson P,
Pessia M,
and
Adelman JP.
Inhibitory interactions between two inward rectifier K+ channel subunits mediated by the transmembrane domains.
J Biol Chem
271:
5866-5870,
1996
48.
Wang, Q,
Curran ME,
Splawski I,
Burn TC,
Millholland JM,
VanRaay TJ,
Shen J,
Timothy KW,
Vincent GM,
de Jager T,
Schwartz PJ,
Toubin JA,
Moss AJ,
Atkinson DL,
Landes GM,
Connors TD,
and
Keating MT.
Positional cloning of a novel potassium channel gene: KVLQT1 mutations cause cardiac arrhythmias.
Nat Genet
12:
17-23,
1996[ISI][Medline].
49.
Wang, Q,
Shen J,
Splawski I,
Atkinson D,
Li Z,
Robinson JL,
Moss AJ,
Towbin JA,
and
Keating MT.
SCN5A mutations associated with an inherited cardiac arrhythmia, long QT syndrome.
Cell
80:
805-811,
1995[ISI][Medline].
50.
Werblin, FS.
Anomalous rectification in horizontal cells.
J Physiol (Lond)
244:
639-657,
1975[Abstract].
51.
Wischmeyer, G,
and
Karschin A.
Receptor stimulation causes slow inhibition of IRK1 inwardly rectifying K+ channels by direct protein kinase A-mediated phosphorylation.
Proc Natl Acad Sci USA
93:
5819-5823,
1996
52.
Woodhull, AM.
Ionic blockage of sodium channels in nerve.
J Gen Physiol
61:
687-708,
1973
53.
Yagi, T,
and
Macleish PR.
Ionic conductances of monkey solitary cone inner segments.
J Neurophysiol
71:
656-665,
1994
54.
Yamamoto, F,
Borgula GA,
and
Steinberg RH.
Effects of light and darkness on pH outside rod photoreceptors in the cat retina.
Exp Eye Res
54:
685-697,
1992[ISI][Medline].
55.
Yamamoto, F,
and
Steinberg RH.
Effects of systemic hypoxia on pH outside rod photoreceptors in the cat retina.
Exp Eye Res
54:
699-709,
1992[ISI][Medline].
56.
Yang-Feng, TL,
Floyd-Smith G,
Nemer M,
Drouin J,
and
Franke U.
The pronatriodilatin gene is located on the distal short arm of human chromosome 1 and on mouse chromosome 4.
Am J Hum Genet
37:
1117-1128,
1985[ISI][Medline].