Electroneutral cation–chloride cotransporters are widely expressed and perform a variety of physiological roles. A novel gene family of five members, encompassing a Na+–Cl transporter, two Na+–K+–2Cl transporters and two K+–Cl cotransporters, encodes these membrane proteins; homologous genes have also been identified in a prokaryote and a number of lower eukaryotes. The cotransporter proteins share a common predicted membrane topology, with twelve putative transmembrane segments flanked by long hydrophilic N- and C-terminal cytoplasmic domains. The molecular identification of these transporters has had a significant impact on the study of their function, regulation and pathophysiology.

Electroneutral, diuretic-sensitive cotransport of Na+, K+ and Cl was first described in Ehrlich cells by Geck et al. (1980). The basic characteristics of this transport activity include a sensitivity to loop diuretics (most frequently bumetanide and furosemide), a stoichiometry of 1Na+:1K+:2Cl and a dependence on the simultaneous presence of all three transported ions. These functional criteria have been extremely useful, facilitating the identification of bumetanide-sensitive Na+–K+–2Cl transport in a large number of cells and tissues (Kaplan et al. 1996a). Inwardly directed bumetanide-sensitive Na+–K+–2Cl transport is activated by cell shrinkage, implicating this transporter in regulatory volume increase (Hoffmann and Dunham, 1995). In contrast, regulatory volume decrease invokes an outwardly directed Na+-independent K+–Cl cotransport activity which is weakly inhibited by bumetanide and furosemide (Lauf et al. 1992; Hoffmann and Dunham, 1995). A third distinct cation–chloride cotransporter, thiazide-sensitive Na+–Cl transport, is abundant in the renal distal convoluted tubule (DCT) (Costanzo, 1985) and the flounder urinary bladder (Stokes, 1989). The cloning of all three classes of cation–chloride cotransporters has been accomplished over the last 4 years, resulting in the identification of a new five-member gene family (see Tables 1, 2). This review summarizes the subsequent advances in the molecular physiology of the vertebrate cation–chloride transporters.

Table 1.

The five vertebrate cation–chloride transporters

The five vertebrate cation–chloride transporters
The five vertebrate cation–chloride transporters
Table 2.

Full-length invertebrate cation–chloride transporters

Full-length invertebrate cation–chloride transporters
Full-length invertebrate cation–chloride transporters

The first cation–chloride transporter to be characterized at the molecular level was the thiazide-sensitive Na+–Cl cotransporter from winter flounder (flTSC) (Gamba et al. 1993). A 3.7 kilobase flTSC cDNA was isolated by expression cloning in Xenopus laevis oocytes, using flounder urinary bladder as the source of poly(A)+ RNA. The expression of flTSC in Xenopus laevis oocytes produces a thiazide-sensitive Na+–Cl transport with identical kinetics to that of the endogenous transporter. A cDNA encoding the first bumetanide-sensitive Na+–K+–2Cl transporter was also cloned from a fish tissue, in this case shark rectal gland (Xu et al. 1994). This cDNA, termed NKCC1 (Na+K+ClCotransporter-1), was obtained by screening a shark cDNA library with two monoclonal antibodies specific for the purified bumetanide-binding protein. Functional expression in HEK-293 cells revealed the expected characteristics of a Na+–K+–2Cl transporter, i.e. bumetanide-sensitive uptake of 86Rb+ (a substitute for K+) that was dependent on the presence of both Na+ and Cl.

The fish cDNA clones were subsequently utilized to identify the mammalian homologs of the electroneutral cation–chloride transporters (Table 1). A fragment of the flTSC cDNA was used to ‘fish’ out cDNAs encoding the rat thiazide-sensitive cotransporter (rTSC) and apical bumetanide-sensitive Na+–K+–2Cl transporter (rBSC1) from renal cortex and outer medulla, respectively (Gamba et al. 1994). Mouse BSC1 cDNAs (NKCC2 in the alternative nomenclature) have also been identified (Mount et al. 1995; Igarashi et al. 1995). All the full-length rodent TSC and BSC1 clones direct expression of the appropriate transport activity in Xenopus oocytes (Gamba et al. 1994). The mammalian NKCC1/BSC2 cotransporters have been cloned using similar homology-based approaches (Delpire et al. 1994; Payne et al. 1995).

The two K+–Cl cotransporter cDNAs were obtained through the identification of related human ESTs (expressed sequence tags) in the GenBank database. Several ESTs in the database were found to be moderately homologous to NKCC1, suggestive of a new branch of the cotransporter family. Using this information, and a combination of library screening and reverse transcriptase/polymerase chain reaction (RT-PCR), KCC1 cDNAs (K+ClCotransporter-1) were subsequently isolated from rat, human and rabbit (Gillen et al. 1996). A similar protocol resulted in the isolation of another putative K+–Cl cotransporter, KCC2, from rat brain (Payne et al. 1996). Heterologous expression of both KCC1 and KCC2 fulfills the criteria for a K+–Cl cotransporter (see below).

The percentage homology between the five mammalian cotransporters is shown in Fig. 2B. In addition to the multiple vertebrate cDNAs, putative cation–chloride cotransporter cDNAs and/or ESTs have been characterized from Manduca sexta, Drosophila melanogaster, Caenorhabditis elegans, yeast and a cyanobacterium (see Table 2 for selected full-length invertebrate homologs). A homology tree (Fig. 1) indicates the phylogenetic relationships between representative sequences, both vertebrate and invertebrate. Two main branches of the gene family can be appreciated, one consisting of predominantly renal Na+-coupled cotransporters and the other of the K+–Cl cotransporters and homologs from simpler organisms. All the cotransporter proteins share a basic structural motif, predicted by hydropathy analysis of their amino acid sequence (Fig. 2A). This primary structure predicts 12 putative hydrophobic transmembrane (TM) segments flanked by long hydrophilic N- and C-terminal cytoplasmic domains. Homology between family members is strongest within the TM domains, but is also evident in the C-terminal domains and in the predicted intracellular loops between TM segments. Human–shark BSC2 chimeras have demonstrated that the transport characteristics (affinities for Na+, K+, Cl and bumetanide) are encoded by the central 12 transmembrane domains (Isenring and Forbush, 1997). In the Na+-dependent transporters (TSC, BSC1 and BSC2), there is a large extracellular loop between TM-7 and TM-8, with potential sites for N-linked glycosylation. An obvious structural departure is evident in the KCC1 and KCC2 sequences, which predict a glycosylated extracellular loop between TM-5 and TM-6 (Gillen et al. 1996; Payne et al. 1996). Full confirmation of the proposed membrane topology has yet to be published. However, both the N-terminal and C-terminal domains of shark BSC2 are clearly phosphorylated, indicating a cytoplasmic orientation (Lytle and Forbush, 1992; Xu et al. 1994). In addition, mutation of the first of two N-glycosylation sites in rTSC dramatically reduces glycosylation of the protein, confirming that the loop between TM-7 and TM-8 is extracellular (Poch et al. 1996).

Fig. 1.

The electroneutral cation–chloride cotransporter family; a homology tree of representative cotransporters based on amino acid sequences. The two main branches of the gene family are outlined. The length of the horizontal lines approximates evolutionary distance. Generated using the DNAStar program (Madison, WI, USA).

Fig. 1.

The electroneutral cation–chloride cotransporter family; a homology tree of representative cotransporters based on amino acid sequences. The two main branches of the gene family are outlined. The length of the horizontal lines approximates evolutionary distance. Generated using the DNAStar program (Madison, WI, USA).

Fig. 2.

(A) Predicted membrane topology of the Na+-coupled cation–chloride cotransporters (BSC1/NKCC2, BSC2/NKCC1 and TSC/NCC). Analysis of amino acid sequences predicts a total of 12 transmembrane (TM) segments, flanked by large intracellular amino- and carboxy-terminal domains; there is a large predicted extracellular loop between TM-7 and TM-8. The KCC sequences predict a large extracellular loop between TM-5 and TM-6. (B) Diagram showing the percentage homology between the four human cation–chloride transporter proteins and rat KCC2.

Fig. 2.

(A) Predicted membrane topology of the Na+-coupled cation–chloride cotransporters (BSC1/NKCC2, BSC2/NKCC1 and TSC/NCC). Analysis of amino acid sequences predicts a total of 12 transmembrane (TM) segments, flanked by large intracellular amino- and carboxy-terminal domains; there is a large predicted extracellular loop between TM-7 and TM-8. The KCC sequences predict a large extracellular loop between TM-5 and TM-6. (B) Diagram showing the percentage homology between the four human cation–chloride transporter proteins and rat KCC2.

BSC1: Na+–K+–2Cl cotransport

The renal thick ascending limb of the loop of Henle (TALH) is the main pharmacological target of the loop diuretics (bumetanide, ethacrynic acid, furosemide, piretanide and torsemide). These drugs exert their natriuretic effect by inhibition of electroneutral Na+–K+–2Cl cotransport in the TALH (Greger, 1985). In contrast to secretory epithelia, which generally exhibit basolateral bumetanide-sensitive Na+–K+–2Cl transport, this transport activity is located on the apical membrane of the TALH facing the tubule lumen (see Figs 3, 4). Bumetanide-sensitive Na+–K+–2Cl cotransport in the TALH is encoded by BSC1/NKCC2 (hereafter referred to as BSC1), whereas bumetanide-sensitive Na+–K+–2Cl transport elsewhere is the function of BSC2/NKCC1 (hereafter referred to as BSC2).

Fig. 3.

Immunofluorescence of rat kidney, with antibodies to the kidney-specific cotransporters BSC1/NKCC2 and TSC/NCC. (A) Immunofluorescence (Rhodamine) with the anti-BSC1 antibody (Kaplan et al. 1996c). Red labelling of the apical membrane of the thick ascending limb (TAL) is seen. Apical labelling of taller columnar cells at the macula densa (MD) is also apparent. The green arrows indicate the direction of urine flow. (B) Immunofluorescence (Rhodamine) with the anti-TSC antibody (Plotkin et al. 1996). Labeling of distal tubule; arrows indicate the sharp border between the cortical TAL and the distal convoluted tubule (DCT). (C) Immunofluorescence of medullary TAL (MTAL) with the BSC1 antibody, clearly showing apical labeling (arrow). (D) Immunofluorescence of the DCT with anti-TSC antibody, showing apical labeling of cells (arrow). (E) Drawing of an idealized nephron, showing the cortical–medullary junction (green line) and the locations of medullary and cortical TAL (MTAL and CTAL), macula densa (MD) and distal convoluted tubule (DCT).

Fig. 3.

Immunofluorescence of rat kidney, with antibodies to the kidney-specific cotransporters BSC1/NKCC2 and TSC/NCC. (A) Immunofluorescence (Rhodamine) with the anti-BSC1 antibody (Kaplan et al. 1996c). Red labelling of the apical membrane of the thick ascending limb (TAL) is seen. Apical labelling of taller columnar cells at the macula densa (MD) is also apparent. The green arrows indicate the direction of urine flow. (B) Immunofluorescence (Rhodamine) with the anti-TSC antibody (Plotkin et al. 1996). Labeling of distal tubule; arrows indicate the sharp border between the cortical TAL and the distal convoluted tubule (DCT). (C) Immunofluorescence of medullary TAL (MTAL) with the BSC1 antibody, clearly showing apical labeling (arrow). (D) Immunofluorescence of the DCT with anti-TSC antibody, showing apical labeling of cells (arrow). (E) Drawing of an idealized nephron, showing the cortical–medullary junction (green line) and the locations of medullary and cortical TAL (MTAL and CTAL), macula densa (MD) and distal convoluted tubule (DCT).

Fig. 4.

Localization of BSC2/NKCC1 immunoreactivity in epithelial tissues (kidney, salivary gland and choroid plexus). (A) Co-staining of the mouse juxtaglomerular apparatus (kidney) with anti-BSC2 antibody (Kaplan et al. 1996b) and anti-α-actin antibody (a marker for vascular cells). Green (FITC staining) indicates cells staining only with anti-actin, red (Rhodamine) indicates cells labeled with anti-BSC2 alone, and areas of overlap are in yellow. A thin ribbon of BSC2-positive extraglomerular mesangial (EGM) cells is seen behind the macula densa and thick ascending limb (TAL). Light staining of the glomerulus (Glom) with anti-BSC2 is also evident. (B) A model of the juxtaglomerular apparatus (JGA), showing the relationships between cell types and transporters. Coloring follows the same pattern as in A. EGM, extraglomerular mesangial cells; MD, macula densa; AA, afferent arteriole; EA, efferent arteriole; JGC, juxtaglomerular cells; TAL, thick ascending limb of Henle. A subset of renin-negative (non-JG) cells in the afferent arteriole are BSC2-positive (Kaplan et al. 1996b). (C) A model of the nephron, showing the relative positions of the cortical collecting duct (CCD), outer medullary collecting duct (OMCD) and inner medullary collecting duct (IMCD). (D) Staining of type-A intercalated cells in rat outer medullary collecting duct (OMCD). The apical membrane stains with anti-H+-ATPase antibody (red), a gift from Dr Stephen Gluck. The basolateral membrane stains with anti-BSC2 antibody (green). (E) Staining of mouse inner medullary collecting duct (IMCD) with anti-BSC2 antibody, clearly delineating the basolateral membranes. (F) Immunocytochemistry of rat salivary gland with anti-BSC2 antibody, indicating basolateral staining (brown) of epithelial cells. (G) Immunofluorescence of rat choroid plexus; apical membranes stain red with anti-BSC2 antibody.

Fig. 4.

Localization of BSC2/NKCC1 immunoreactivity in epithelial tissues (kidney, salivary gland and choroid plexus). (A) Co-staining of the mouse juxtaglomerular apparatus (kidney) with anti-BSC2 antibody (Kaplan et al. 1996b) and anti-α-actin antibody (a marker for vascular cells). Green (FITC staining) indicates cells staining only with anti-actin, red (Rhodamine) indicates cells labeled with anti-BSC2 alone, and areas of overlap are in yellow. A thin ribbon of BSC2-positive extraglomerular mesangial (EGM) cells is seen behind the macula densa and thick ascending limb (TAL). Light staining of the glomerulus (Glom) with anti-BSC2 is also evident. (B) A model of the juxtaglomerular apparatus (JGA), showing the relationships between cell types and transporters. Coloring follows the same pattern as in A. EGM, extraglomerular mesangial cells; MD, macula densa; AA, afferent arteriole; EA, efferent arteriole; JGC, juxtaglomerular cells; TAL, thick ascending limb of Henle. A subset of renin-negative (non-JG) cells in the afferent arteriole are BSC2-positive (Kaplan et al. 1996b). (C) A model of the nephron, showing the relative positions of the cortical collecting duct (CCD), outer medullary collecting duct (OMCD) and inner medullary collecting duct (IMCD). (D) Staining of type-A intercalated cells in rat outer medullary collecting duct (OMCD). The apical membrane stains with anti-H+-ATPase antibody (red), a gift from Dr Stephen Gluck. The basolateral membrane stains with anti-BSC2 antibody (green). (E) Staining of mouse inner medullary collecting duct (IMCD) with anti-BSC2 antibody, clearly delineating the basolateral membranes. (F) Immunocytochemistry of rat salivary gland with anti-BSC2 antibody, indicating basolateral staining (brown) of epithelial cells. (G) Immunofluorescence of rat choroid plexus; apical membranes stain red with anti-BSC2 antibody.

BSC1 plays a primary role in transcellular absorption of Na+–Cl by the medullary and cortical TALH, and plays a secondary role in the paracellular transport of Na+, Ca2+ and Mg2+ by this nephron segment (Greger, 1985; Hebert, 1992) (see below). Salt absorption by the TALH is crucial to countercurrent multiplication in the renal medulla and the excretion of a concentrated urine. The TALH also has a function in renal acid excretion, since NH4+ can substitute for K+ in transport by BSC1 (Good, 1994). Ammonium generated by the renal proximal tubule is thus reabsorbed from the tubule lumen in the TALH and excreted from the renal interstitium by more distal nephron segments.

Regulation of salt transport in the TALH, which is endowed with receptors for a large number of hormones, occurs through a combination of both primary and secondary effects on apical Na+–K+–2Cl transport (Kaplan et al. 1996a). A number of hormones activate protein kinase A (PKA) in this nephron segment, resulting in an increase in salt transport. The primary structure of BSC1 predicts phosphorylation sites for both PKA and protein kinase C (PKC) (Gamba et al. 1994; Igarashi et al. 1995). Alternative splicing of the 3′ end of mouse BSC1 also modifies the predicted phosphorylation of the transporter protein by truncating the C terminus and substituting a unique C-terminal 55 amino acid segment (Mount et al. 1995). However, the functional significance of this alternative splicing is still unclear, and there is as yet no direct biochemical evidence for phosphorylation of any isoforms of BSC1.

In the mouse medullary TALH, vasopressin activates a switch in the K+-dependence of apical Na+–(K+)–2Cl transport, from a K+-independent Na+–Cl transporter to a K+-dependent Na+–K+–2Cl transporter (Sun et al. 1991). This modulation of apical Na+–K+–2Cl transport is probably of crucial importance for the physiology of the TALH. Recycling of K+ through the Na+–K+–2Cl transporter and apical K+ channels generates a lumen-positive potential difference, which drives the absorption of cations through a cation-selective paracellular pathway. This mechanism is responsible for the reabsorption of divalent cations (Ca2+ and Mg2+) by the TALH (Greger, 1985; Hebert, 1992). In addition, the availability of a paracellular pathway for Na+ doubles Na+–Cl transport by the TALH without an increase in energy expenditure (Sun et al. 1991). K+-independent bumetanide-sensitive Na+–Cl transport has also been found in rabbit and rat TALH (Alvo et al. 1985; Ludens et al. 1995).

As predicted by its functional role, expression of BSC1 is restricted to the kidney, specifically the medullary and cortical TALH. Indeed, Tamm–Horsfall glycoprotein (Yu et al. 1994) and BSC1 (Igarashi et al. 1996) are the only known TALH-specific genes. Both BSC1 transcript (Gamba et al. 1994; Igarashi et al. 1995; Obermüller et al. 1996) and protein (Kaplan et al. 1996c; Ecelbarger et al. 1996) have been localized to the TALH. Affinity-purified antibodies to a C-terminal fusion protein of rBSC1 detect a protein of approximately 150 kDa in rat and mouse kidney (Kaplan et al. 1996c; Ecelbarger et al. 1996), the same size as the major renal bumetanide-binding protein (Haas et al. 1991). Immunofluorescence studies with this antibody label the apical membrane of all cells along the entire length of the TALH (Fig. 3A,C).

The availability of molecular probes for BSC1 and other transport proteins promises to refine significantly our understanding of the TALH phenotype. The low-conductance apical K+ channel of the mammalian TALH is encoded by ROMK (rat outer medulla K+ channel), the prototypical inwardly rectifying K+ channel (Ho et al. 1993). Double-immunofluorescence studies with anti-ROMK and anti-BSC1 antibodies indicate a distinct heterogeneity of rat TALH, such that not all BSC1-positive TALH cells are positive for the ROMK protein (Xu et al. 1997). Electron microscopy of rat TALH has previously identified two subtypes of TALH cells, with both rough and smooth apical membranes differing in the abundance of apical microvilli (Allen and Tisher, 1976). In the hamster TALH, the relative frequency of these two cell types correlates closely with the proportion of cells having a high and low apical K+ conductance (Tsuruoka et al. 1994), and the important issue of which morphological subtype of TALH cell expresses ROMK is under investigation.

Alternative splicing of BSC1 may also generate heterogeneity of ion transport in the TALH. Three mutually exclusive cassette exons have been described near the 5′ end of rabbit, mouse and human BSC1 (Igarashi et al. 1995; Payne and Forbush, 1994; Mount et al. 1997). The cassette exons are predicted to encode 31 amino acids, spanning most of the second transmembrane domain and extending 10–13 amino acids into the cytoplasm. Expression of the individual cassettes is spatially restricted along the TALH, in the inner and outer stripe of outer medulla and in the cortical TALH (Igarashi et al. 1995; Payne and Forbush, 1994). The functional significance of this alternative splicing is not known. However, given the probable role of the transmembrane segments in ion binding and transport, it is highly likely that the cassette exons affect the relative affinity of apical cation–chloride transport along the TALH.

BSC1 protein (Kaplan et al. 1996b,c) and transcript (Obermüller et al. 1996) are also detected in the macula densa, a specialized group of tubular cells situated at the point of apposition of renal tubules with their parent glomeruli (Fig. 3A). The macula densa controls two important physiological processes, tubuloglomerular feedback and tubular regulation of renin release. An increase in luminal Na+–Cl activity at the macula densa decreases glomerular filtration by inducing constriction of the afferent renal arteriole; this tubuloglomerular feedback mechanism is blocked by luminal furosemide (Ito and Carretero, 1990). Increases in luminal Na+–Cl activity at the macula densa also inhibit renin release from juxtaglomerular cells in the afferent arteriole, an effect that is blocked by luminal bumetanide and furosemide (Lorenz et al. 1991). Macula densa cells have been shown to possess apical bumetanide-sensitive Na+–K+–2Cl transport with a Km for Cl of 32.5 mmol l−1 (Lapointe et al. 1995). This Km is close to the half-maximal concentration of tubular Cl for inhibition of renin release (He et al. 1995). Indeed, it appears that the macula densa responds to changes in tubular Cl concentration, since variations in tubular Na+ concentration have no effect on tubuloglomerular feedback or renin secretion (Lorenz et al. 1991).

BSC2: Na+–K+–2Cl cotransport

Bumetanide-sensitive Na+–K+–2Cl transport is a feature of perhaps all cultured mammalian cells. However, recent developments suggest that data from tissue-specific cell lines cannot be extrapolated to native tissue, since a number of cell types which express bumetanide-sensitive Na+–K+–2Cl transport in culture do not appear to express BSC2 in vivo (Kaplan et al. 1996b; Plotkin et al. 1997a). BSC2 transcript and transport activity are not detected in freshly isolated proximal tubule cells, aortic endothelial cells and vascular smooth muscle cells, but are induced after more prolonged culture (Raat et al. 1996). Although low-level expression of BSC2 protein is detectable in freshly isolated endothelial cells (Yerby et al. 1997), evidence for in vivo expression in this cell type has not been published. Induction of BSC2 in culture may be involved in the increase in cell volume necessary for cell growth and progression through the cell cycle. In NIH 3T3 cells, for example, K+ content and cell volume increase after serum stimulation, and bumetanide has an anti-proliferative effect (Bussolati et al. 1996). The influx of Na+ through amiloride- and bumetanide-sensitive pathways is also an early event after serum stimulation of 3T3 cells (Berman et al. 1995).

A more direct role for cation–chloride transporters in cell growth and signal transduction is suggested by genetic manipulation of tobacco protoplasts (Harling et al. 1997). One of several auxin-independence genes isolated by activation T-DNA tagging, axi-4, encodes a plant protein with significant homology to cation–chloride cotransporters. Overexpression of full-length axi-4 or shark NKCC1 in protoplasts confers auxin-independent cell growth. However, only a portion of the axi-4 gene was isolated in the initial T-DNA tagging, and the C-terminal 237 amino acids of the axi-4 protein are sufficient to confer auxin-independent growth (Harling et al. 1997). This suggests that the C-terminal domains of the cation–chloride transporters, which share a moderate degree of homology, play a direct role in cellular growth.

The expression pattern of BSC2 in the kidney and the nervous system has been examined in detail (Kaplan et al. 1996b; (Plotkin et al. 1997a). These studies utilized a polyclonal antibody specific for a C-terminal mBSC2 fusion protein. This antibody recognizes two proteins of approximately 145 kDa and 155 kDa, which are probably differentially glycosylated forms of BSC2. Within the brain, BSC2 transcript is most abundant in the choroid plexus, followed by the cerebellum and brain stem. Immunofluorescence and functional studies localize BSC2 to the apical cell membrane of the choroid plexus (Plotkin et al. 1997a) (Fig. 4G). The precise role of BSC2 in the choroid plexus is unclear. However, a component of cerebrospinal fluid production is bumetanide-sensitive (Javaheri and Wagner, 1993).

BSC2 protein is also detected in some neuronal cell bodies and dendrites in the brain and in dorsal root ganglion sensory neurons in the peripheral nervous system (Plotkin et al. 1997a). This observation is particularly important, since the magnitude of the Cl electrochemical gradient plays an important role in the response of neurons to stimuli that affect Cl conductance (Misgeld et al. 1986; Staley et al. 1996). In neurons with strong inward transport of Cl, intracellular Cl concentration is high and GABAA receptor stimulation, which activates a Cl channel, is depolarizing and excitatory. In contrast, in cells demonstrating outward Cl transport, GABAA activation is hyperpolarizing and inhibitory. Both inward and outward transport of Cl in neurons are diuretic-sensitive and are probably encoded by BSC2 and KCC1/KCC2 (see below), respectively. BSC2 in the central nervous system is developmentally regulated, with high levels of expression at birth and during the first few postnatal days, and decreased levels of expression thereafter (Plotkin et al. 1997b). This correlates very well with the switch of the GABA effect, from depolarizing to hyperpolarizing, during the first weeks of postnatal life.

Within mouse kidney, heavy expression of BSC2 protein is detected at the basolateral membrane of inner medullary collecting duct cells (IMCD) (Kaplan et al. 1996b) (Fig. 4E). This result was expected, since there is in vivo evidence for hormone-sensitive basolateral Na+–K+–2Cl transport in rodent IMCD (Rocha and Kudo, 1990). Within the glomerulus, light staining of mesangial cells is also evident (Fig. 4A). Cultured mesangial cells have a well-characterized Na+–K+–2Cl transport system which is activated by factors such as angiotensin II (Homma and Harris, 1991). Unexpected staining of BSC2 is also found in juxtaglomerular cells in the afferent arteriole, suggesting a role in the regulation of renin secretion by these cells (Fig. 4A,B). The so-called ‘extraglomerular’ mesangial cells, located between the glomerulus and the macula densa, are also labeled with BSC2 antibody (Fig. 4A). The function of BSC2 in the glomerulus and juxtaglomerular apparatus is unknown. However, the dominant role of extracellular Cl in the function of the juxtaglomerular apparatus and associated cells is intriguing (Tsukahara et al. 1994; Matsunaga et al. 1994). As in neurons, BSC2 may modulate the transmembrane Cl gradient, altering the response of renal cells to agents that affect Cl conductance (Ling et al. 1995).

The distribution of BSC2 in rat kidney has also been studied, using the antibody described above (Fig. 4D) and an antibody generated against a C-terminal rat BSC2 peptide (Ginns et al. 1996). These studies replicate the findings in mouse kidney; however, cells within the rat IMCD are not labeled. In the rat outer medullary collecting duct, however, the basolateral membranes of type A intercalated cells are stained by the rBSC2 antibody (Ginns et al. 1996) (Fig. 4D). The type A intercalated cell, defined by expression of an apical H+-ATPase and a basolateral Cl–HCO3 exchanger, functions in acid secretion (Brown and Breton, 1996). It has been demonstrated that basolateral BSC2 in a mouse IMCD cell line is capable of transporting NH4+ (Wall et al. 1995). Therefore, both BSC1 (see above) and BSC2 are part of the pathway for renal excretion of NH4+.

Within secretory epithelia, BSC2 protein has been localized to the basolateral membrane of acinar cells in rat submandibular gland (Fig. 4F) (He et al. 1997) and parotid gland (Lytle et al. 1995). Heterogeneous staining of the basolateral membrane of intralobular duct cells in submandibular gland has also been observed (He et al. 1997). BSC2 is heavily expressed in stomach, and the level of transcript in the Necturus maculosus gastric fundus essentially doubles after feeding (Soybel et al. 1995). Functional experiments in this model strongly support a role for BSC2 in the secretion of stomach acid, probably by providing a basolateral entry pathway for Cl (Soybel et al. 1995).

The complex functional regulation of BSC2 has previously been reviewed in detail (Haas, 1994; Kaplan et al. 1996a; Mount et al. 1997). Although all the cation–chloride transporters are predicted to be substrates for protein kinases, BSC2 is the only member of the family for which there is published evidence of phosphorylation (Xu et al. 1994; Tanimura et al. 1995; Plotkin et al. 1997a). The activation of PKA in rat parotid gland stimulates phosphorylation of one of at least three phosphorylation sites in the BSC2 protein, indicating a direct role for protein phosphorylation in the regulation of BSC2 (Tanimura et al. 1995). However, the modulation of BSC2 by PKA is evidently very complex, involving alternative splicing, secondary activation of other protein kinases and protein–protein interactions. The solitary predicted PKA site in BSC2 is encoded by a short exon that is removed by alternative splicing in several mouse tissues, such that a fraction of BSC2 is not predicted to be a substrate for PKA (Randall et al. 1997). In shark rectal gland, activation of PKA results in phosphorylation of BSC2 on a threonine that is not part of a consensus PKA site (Lytle and Forbush, 1992). This phosphorylation can be blocked by raising internal Cl concentration, suggesting that the effect of PKA is mediated by a Cl-sensitive phosphorylation event that does not directly involve this kinase (Lytle and Forbush, 1996). Cl may thus regulate its own transport by affecting phosphorylation of BSC2, and this phenomenon has also been reported in mammalian epithelial cells (Haas et al. 1995). Finally, in the T84 cell line, BSC2 co-immunoprecipitates with two integral membrane proteins of 130 and 160 kDa. Activation of PKA increases the surface expression of these co-precipitated proteins, with a lesser effect on the surface expression of BSC2 (D’Andrea et al. 1996).

TSC: Na+–Cl cotransport

In the distal convoluted tubule (DCT) of the mammalian kidney, the primary apical entry pathway for Na+ is through the thiazide-sensitive Na+–Cl cotransporter (Mount et al. 1997). The natriuretic effect of the thiazide diuretics underlies their therapeutic effect in edema states and hypertension. In addition, luminal thiazide in the DCT has a hypocalciuric effect, implicating TSC in the fine-tuning of Ca2+ excretion (Costanzo, 1985). Although microperfusion identified the DCT as the primary site of action of the thiazides, newer molecular probes have been invaluable in defining the phenotype of cells in this segment of the kidney. A TSC-specific antibody raised against an amino-terminal rTSC fusion protein detects the TSC protein at the apical cell membrane of the DCT (Fig. 3D). The transition between the BSC1-positive cells of the cortical TALH and the TSC-positive DCT is dramatic (Fig. 3B). However, in the rat and human kidney (Plotkin et al. 1996; Obermüller et al. 1995), the distal border of the DCT is less distinct, with a mixture of TSC-negative and TSC-positive cells in the connecting segment between the DCT and the cortical collecting duct. Consistent with the role of TSC in Ca2+ homeostasis, all cells that express this transporter in the DCT and beyond are also positive for calbindin D28 (Plotkin et al. 1996), an intracellular Ca2+-binding protein implicated in transcellular Ca2+ transport.

KCC1 and KCC2: K+–Clcotransport

Electroneutral cotransport of K+–Cl is detected in a wide range of cells functioning in regulatory volume decrease, transepithelial salt transport and the maintenance of transmembrane Cl gradients. Consistent with this broad functional distribution, KCC1 transcript is heavily expressed in many tissues, including brain (Gillen et al. 1996). Precise localization awaits the development of a KCC1-specific antibody. In contrast to KCC1, KCC2 expression is restricted to the brain, and in situ hybridization reveals KCC2 transcript in neurons throughout the central nervous system (Payne et al. 1996). A BLAST search of the EST database with rKCC2 also detects a significant number of related ESTs in a cDNA library from human retina (D. B. Mount, unpublished results).

Stable expression of full-length rabbit KCC1 in HEK-293 cells reveals the expected characteristics of a K+–Cl cotransporter (Lauf et al. 1992): 86Rb+ efflux in these cells is stimulated by cell swelling to a much greater extent than in untransfected cells. Cl-dependent influx of 86Rb+ in stable transfectants is activated by treatment with N-ethylmaleimide (NEM), a reagent that probably exerts its effect by modification of sulfhydryl groups on the transporter and/or associated proteins. This influx is weakly furosemide-and bumetanide-sensitive (Gillen et al. 1996), and kinetic analysis also indicates a low affinity of the transporter for both K+ (Km>25 mmol l−1) and Cl (Km>50 mmol l−1).

In contrast to KCC1, stable expression of KCC2 in HEK-293 cells results in a K+–Cl cotransporter which is not activated by cell swelling (Payne, 1997). KCC2 also displays a much higher affinity for extracellular K+ (Km≈5.2 mmol l−1). On the basis of thermodynamic considerations, KCC2 may function in the buffering of external K+ within the central nervous system, in addition to maintaining the transmembrane Cl gradient (Payne, 1997).

K+–Cl cotransport in red blood cells is stimulated by kinase inhibition and inactivated by phosphatase inhibition (Flatman et al. 1996; Lauf et al. 1992). Genetic evidence has recently implicated the cytoplasmic tyrosine kinases Fgr and Hck in the regulation of erythrocyte K+–Cl cotransport (De Franceschi et al. 1997). Thus, red cell K+ content is lower in fgr−/−hck−/− mice owing to an activation of K+–Cl cotransport. Okadaic acid is capable of inhibiting K+–Cl cotransport in double-negative red cells, suggesting that the Fgr and Hck tyrosine kinases negatively regulate a phosphatase, which in turn inhibits KCC1.

Pathophysiology of cation–chloride cotransport

Given the physiological importance of both thiazide- and bumetanide-sensitive salt transport by the kidney, it is not surprising that altered activity of these transporters has been implicated in human disease. Even before the availability of cDNA clones, the human BSC1 and TSC genes were leading candidate loci for two forms of inherited metabolic alkalosis, Bartter’s syndrome and Gitelman’s syndrome (Stein, 1985). In addition to a hypokalemic alkalosis, patients with Bartter’s syndrome have a decreased urinary concentrating ability and polyuria, and generally exhibit increased urinary excretion of Ca2+ with a normal serum Mg2+ level (Bettinelli et al. 1992). This constellation of findings is most compatible with a primary defect in the TALH. As anticipated, mutations in the human BSC1 gene have recently been reported in several kindreds with Bartter’s syndrome (Simon et al. 1996a). However, other families with Bartter’s syndrome do not exhibit linkage to BSC1, suggesting genetic heterogeneity. Mutations in the human ROMK K+ channel gene have also been identified in families with Bartter’s syndrome (Simon et al. 1996b; International Collaborative Study Group, 1997), underscoring the functional coupling between apical K+ channels and apical Na+–K+–2Cl transport in the mammalian TALH. More recently, a third gene, CLC-KNB, has been implicated in the genesis of Bartter’s syndrome in a third subset of families with the disease (Simon et al. 1997). CLC-KNB is a member of the CLC family of Cl channels and probably encodes a component of the basolateral Cl conductance in TALH cells.

In contrast to Bartter’s syndrome, patients with Gitelman’s syndrome do not have a defect in urinary concentrating ability. Another major distinguishing feature is the presence of marked hypomagnesemia and hypocalcuria in Gitelman’s syndrome (Bettinelli et al. 1992). The primary defect in this syndrome appears to be in the distal tubule, and the lack of a diuretic response to thiazides in some patients pointed to mutations in TSC. Linkage of the human TSC gene to Gitelman’s syndrome has recently been reported, along with the direct characterization of non-conservative mutations in affected patients (Simon et al. 1996c; Mastroianni et al. 1996; Lemmink et al. 1996; Takeuchi et al. 1996; Pollak et al. 1996). The other cation–chloride cotransporter genes are not obvious candidates for monogenic disorders, but probably play a role in human disease. For example, excessive K+–Cl cotransport and cellular dehydration may underlie the destruction of red cells in hemolytic anemias, particularly in sickle cell anemia (Franco et al. 1995). In addition, the importance of the transmembrane Cl gradient in neuronal excitability implicates BSC2 and the K+–Cl cotransporters in seizure disorders (Hochman et al. 1995; Payne, 1997).

The cloning of the cation–chloride transporters was a significant advance and has already begun to clarify a previously confusing field. It is highly likely that other members of this rapidly expanding gene family remain to be discovered. Localization studies have been informative, and the cotransporter antibodies are proving to be valuable phenotypic and functional markers. The ability to specifically inhibit individual genes by both germline inactivation and anti-sense technology should help clarify some of the remaining physiological issues, such as the roles of BSC1 and BSC2 in tubuloglomerular feedback and tubular regulation of renin release. Within cells, the C termini of cation–chloride transporters may play a fundamental role in the regulation of cell growth, independent of the effect of cation–chloride cotransport on intracellular ion concentrations (Harling et al. 1997). Conversely, the level of intracellular Na+ and K+ modulates the activation of apoptosis; the role of the cation–chloride transporters in this process is not known and is an intriguing area for future investigation (Bortner et al. 1997; Hughes et al. 1997). At the molecular level, the identification of ion translocation sites, phosphorylation sites and transporter-associated proteins is pending. This information will help resolve issues such as the differences in ionic affinity between transporters (Payne et al. 1995; Payne, 1997; Isenring and Forbush, 1997) and the inverse effect of kinases, phosphatases and cell volume on the activity of BSC2 and KCC1 (Hoffmann and Dunham, 1995).

Original research was supported by grants from the National Institues of Health to S.C.H. (DK45792 and DK36803), to D.B.M. (DK02328) and to E.D. (HL49251) and from the Mexican Council of Science and Technology to G.G. (CONACYT, M3840). E.D. is an Established Investigator of the American Heart Association and G.G. is an International Research Scholar of the Howard Hughes Medical Institute.

Allen
,
F.
and
Tisher
,
C. C.
(
1976
).
Morphology of the ascending thick limb of Henle
.
Kidney Int
.
9
,
8
22
.
Alvo
,
M.
,
Calamia
,
J.
and
Eveloff
,
J.
(
1985
).
Lack of potassium effect on Na+–Cl cotransport in the medullary thick ascending limb
.
Am. J. Physiol
.
249
,
F34
F39
.
Berman
,
E.
,
Sharon
,
I.
and
Atlan
,
H.
(
1995
).
An early transient increase of intracellular Na+ may be one of the first components of the mitogenic signal. Direct detection by 23Na+-NMR spectroscopy in quiescent 3T3 mouse fibroblasts stimulated by growth factors
.
Biochim. biophys. Acta
1239
,
177
185
.
Bettinelli
,
A.
,
Bianchetti
,
M. G.
,
Girardin
,
E.
,
Caringella
,
A.
,
Cecconi
,
M.
,
Appiani
,
A. C.
,
Pavanello
,
L.
,
Gastaldi
,
R.
,
Isimbaldi
,
C.
,
Lama
,
G.
,
Marchesoni
,
C.
,
Matteuci
,
C.
,
Patriacca
,
P.
,
Di Natale
,
B.
,
Setzu
,
C.
and
Vitucci
,
P.
(
1992
).
Use of calcium excretion values to distinguish two forms of primary renal tubular hypokalemic alkalosis: Bartter and Gitelman syndromes
.
J. Pediatr
.
120
,
38
43
.
Bortner
,
C. D.
,
Hughes
,
F. M. J.
and
Cidlowski
,
J. A.
(
1997
).
A primary role for K+ and Na+ efflux in the activation of apoptosis
.
J. biol. Chem
.
272
,
32436
32442
.
Brown
,
D.
and
Breton
,
S.
(
1996
).
Mitochondria-rich, proton-secreting epithelial cells
.
J. exp. Biol
.
199
,
2345
2358
.
Bussolati
,
O.
,
Uggeri
,
J.
,
Belletti
,
S.
,
Dall’asta
,
V.
and
Gazzola
,
G. C.
(
1996
).
The stimulation of Na+,K+,Cl cotransport and of system A for neutral amino acid transport is a mechanism for cell volume increase during the cell cycle
.
FASEB J
.
10
,
920
926
.
Costanzo
,
L. S.
(
1985
).
Localization of diuretic action in microperfused rat distal tubules: Ca2+ and Na+ transport
.
Am. J. Physiol
.
248
,
F527
F535
.
D’andrea
,
L.
,
Lytle
,
C.
,
Matthews
,
J. B.
,
Hofman
,
P.
,
Forbush Iii
,
B.
R. and
Madara
,
J. L.
(
1996
).
Na+–K+–2Cl cotransporter (NKCC) of intestinal epithelial cells. Surface expression in response to cyclic AMP
.
J. biol. Chem
.
271
,
28969
28976
.
De Franceschi
,
L.
,
Fumagalli
,
L.
,
Olivieri
,
O.
,
Corrocher
,
R.
,
Lowell
,
C. A.
and
Berton
,
G.
(
1997
).
Deficiency of Src family kinases Fgr and Hck results in activation of erythrocyte K+–Cl cotransport
.
J. clin. Invest
.
99
,
220
227
.
Delpire
,
E.
,
Rauchman
,
M. I.
,
Beier
,
D. R.
,
Hebert
,
S. C.
and
Gullans
,
S. R.
(
1994
).
Molecular cloning and chromosome localization of a putative basolateral Na+–K+–2Cl cotransporter from mouse inner medullary collecting duct (mIMCD-3) cells
.
J. biol. Chem
.
269
,
25677
25683
.
Ecelbarger
,
C. A.
,
Terris
,
J.
,
Hoyer
,
J. R.
,
Nielsen
,
S.
,
Wade
,
J. B.
and
Knepper
,
M. A.
(
1996
).
Localization and regulation of the rat renal Na+–K+–2Cl cotransporter, BSC-1
.
Am. J. Physiol
.
271
,
F619
F628
.
Flatman
,
P. W.
,
Adragna
,
N. C.
and
Lauf
,
P. K.
(
1996
).
Role of protein kinases in regulating sheep erythrocyte K+–Cl cotransport
.
Am. J. Physiol
.
271
,
C255
C263
.
Franco
,
R. S.
,
Palascak
,
M.
,
Thompson
,
H.
and
Joiner
,
C. H.
(
1995
).
K+–Cl cotransport activity in light versus dense transferrin receptor-positive sickle reticulocytes
.
J. clin. Invest
.
95
,
2573
2580
.
Gamba
,
G.
,
Miyanoshita
,
A.
,
Lombardi
,
M.
,
Lytton
,
J.
,
Lee
,
W. S.
,
Hediger
,
M. A.
and
Hebert
,
S. C.
(
1994
).
Molecular cloning, primary structure and characterization of two members of the mammalian electroneutral sodium–(potassium)–chloride cotransporter family expressed in kidney
.
J. biol. Chem
.
269
,
17713
17722
.
Gamba
,
G.
,
Saltzberg
,
S. N.
,
Lombardi
,
M.
,
Miyanoshita
,
A.
,
Lytton
,
J.
,
Hediger
,
M. A.
,
Brenner
,
B. M.
and
Hebert
,
S. C.
(
1993
).
Primary structure and functional expression of a cDNA encoding the thiazide-sensitive, electroneutral sodium–chloride cotransporter
.
Proc. natn. Acad. Sci. U.S.A
.
90
,
2749
2753
.
Geck
,
P.
,
Pietrzyk
,
C.
,
Burckhardt
,
B. C.
,
Pfeiffer
,
B.
and
Heinz
,
E.
(
1980
).
Electrically silent cotransport on Na+, K+ and Cl in Ehrlich cells
.
Biochim. biophys. Acta
600
,
432
447
.
Gillen
,
C. M.
,
Brill
,
S.
,
Payne
,
J. A.
and
Forbush
III,
B.
(
1996
).
Molecular cloning and functional expression of the K+–Cl cotransporter from rabbit, rat and human. A new member of the cation–chloride cotransporter family
.
J. biol. Chem
.
271
,
16237
16244
.
Ginns
,
S. M.
,
Knepper
,
M. A.
,
Ecelbarger
,
C. A.
,
Terris
,
J.
,
He
,
X.
,
Coleman
,
R. A.
and
Wade
,
J. B.
(
1996
).
Immunolocalization of the secretory isoform of Na+–K+–Cl cotransporter in rat renal intercalated cells
.
J. Am. Soc. Nephrol
.
7
,
2533
2542
.
Good
,
D. W.
(
1994
).
Ammonium transport by the thick ascending limb of Henle’s loop
.
A. Rev. Physiol
.
56
,
623
647
.
Greger
,
R.
(
1985
).
Ion transport mechanisms in thick ascending limb of Henle’s loop of mammalian nephron
.
Physiol. Rev
.
65
,
760
797
.
Haas
,
M.
(
1994
).
The Na+–K+–Cl cotransporters
.
Am. J. Physiol
.
267
,
C869
C885
.
Haas
,
M.
,
Dunham
,
P. B.
and
Forbush
,
B. R.
(
1991
).
[3H]-bumetanide binding to mouse kidney membranes: identification of corresponding membrane proteins
.
Am. J. Physiol
.
260
,
C791
C804
.
Haas
,
M.
,
Mcbrayer
,
D.
and
Lytle
,
C.
(
1995
).
[Cl]i-dependent phosphorylation of the Na+–K+–Cl cotransport protein of dog tracheal epithelial cells
.
J. biol. Chem
.
270
,
28955
28961
.
Harling
,
H.
,
Czaja
,
I.
,
Schell
,
J.
and
Walden
,
R.
(
1997
).
A plant cation–chloride co-transporter promoting auxin-independent tobacco protoplast division
.
EMBO J
.
16
,
5855
5866
.
He
,
X.-R.
,
Greenberg
,
S. G.
,
Briggs
,
J. P.
and
Schnermann
,
J.
(
1995
).
Effects of furosemide and verapamil on the Na+,Cl dependency of macula densa-mediated renin secretion
.
Hypertension
26
,
137
142
.
He
,
X.
,
Tse
,
C. M.
,
Donowitz
,
M.
,
Alper
,
S. L.
,
Gabriel
,
S. E.
and
Baum
,
B. J.
(
1997
).
Polarized distribution of key membrane transport proteins in the rat submandibular gland
.
Pflügers Arch
.
433
,
260
268
.
Hebert
,
S. C.
(
1992
).
Nephron heterogeneity
.
In Handbook of Physiology – Renal Physiology
, vol.
1
, chapter 20, pp.
875
925
.
Oxford
:
Oxford University Press
.
Ho
,
K.
,
Nichols
,
C. G.
,
Lederer
,
W. J.
,
Lytton
,
J.
,
Vassilev
,
P. M.
,
Kanazirska
,
M. V.
and
Hebert
,
S. C.
(
1993
).
Cloning and expression of an inwardly rectifying ATP-regulated potassium channel
.
Nature
362
,
31
38
.
Hochman
,
D. W.
,
Baraban
,
S. C.
,
Owens
,
J. W.
and
Schwartzkroin
,
P. A.
(
1995
).
Dissociation of synchronization and excitability in furosemide blockade of epileptiform activity
.
Science
270
,
99
102
.
Hoffmann
,
E. K.
and
Dunham
,
P. B.
(
1995
).
Membrane mechanisms and intracellular signalling in cell volume regulation
.
Int. Rev. Cytol
.
161
,
173
262
.
Homma
,
T.
and
Harris
,
R. C.
(
1991
).
Time-dependent biphasic regulation of Na+–K+–Cl cotransport in rat glomerular mesangial cells
.
J. biol. Chem
.
266
,
13553
13559
.
Hughes
,
F. M. J.
,
Bortner
,
C. D.
,
Purdy
,
G. D.
and
Cidlowski
,
J. A.
(
1997
).
Intracellular K+ suppresses the activation of apoptosis in lymphocytes
.
J. biol. Chem
.
272
,
30567
30576
.
Igarashi
,
P.
,
Vanden Heuvel
,
G.
B.,
Payne
,
J. A.
and
Forbush
III,
B.
(
1995
).
Cloning, embryonic expression and alternative splicing of a murine kidney-specific Na+–K+–Cl cotransporter
.
Am. J. Physiol
.
269
,
F405
F418
.
Igarashi
,
P.
,
Whyte
,
D. A.
,
Li
,
K.
and
Nagami
,
G. T.
(
1996
).
Cloning and kidney cell-specific activity of the promoter of the murine renal Na+–K+–Cl cotransporter gene
.
J. biol. Chem
.
271
,
9666
9674
.
International Collaborative Study Group
(
1997
).
Mutations in the gene encoding the inwardly-rectifying renal potassium channel, ROMK, cause the antenatal variant of Bartter syndrome: evidence for genetic heterogeneity
.
Human molec. Genet
.
6
,
17
26
.
Isenring
,
P.
and
Forbush
III,
B.
(
1997
).
Ion and bumetanide binding by the Na+–K+–Cl cotransporter. Importance of transmembrane domains
.
J. biol. Chem
.
272
,
24556
24562
.
Ito
,
S.
and
Carretero
,
O. A.
(
1990
).
An in vitro approach to the study of macula densa-mediated glomerular hemodynamics
.
Kidney Int
.
38
,
1206
1210
.
Javaheri
,
S.
and
Wagner
,
K. R.
(
1993
).
Bumetanide decreases canine cerebrospinal fluid production. In vivo evidence for Na+–Cl cotransport in the central nervous system
.
J. clin. Invest
.
92
,
2257
2261
.
Kaplan
,
M. R.
,
Mount
,
D. B.
,
Delpire
,
E.
,
Gamba
,
G.
and
Hebert
,
S. C.
(
1996a
).
Molecular mechanisms of Na+,Cl cotransport
.
A. Rev. Physiol
.
58
,
649
668
.
Kaplan
,
M. R.
,
Plotkin
,
M. D.
,
Brown
,
D.
,
Hebert
,
S. C.
and
Delpire
,
E.
(
1996b
).
Expression of the mouse Na+–K+–Cl cotransporter, mBSC2, in the terminal inner medullary collecting duct, the glomerular and extraglomerular mesangium and the glomerular afferent arteriole
.
J. clin. Invest
.
98
,
723
730
.
Kaplan
,
M. R.
,
Plotkin
,
M. D.
,
Lee
,
W. S.
,
Xu
,
Z. C.
,
Lytton
,
J.
and
Hebert
,
S. C.
(
1996c
).
Apical localization of the Na+–K+–Clcotransporter, rBSC1, on rat thick ascending limbs
.
Kidney Int
.
49
,
40
47
.
Lapointe
,
J. Y.
,
Laamarti
,
A.
,
Hurst
,
A. M.
,
Fowler
,
B. C.
and
Bell
,
P. D.
(
1995
).
Activation of Na+–K+–2Cl cotransport by luminal chloride in macula densa cells
.
Kidney Int
.
47
,
752
757
.
Lauf
,
P. K.
,
Bauer
,
J.
,
Adragna
,
N. C.
,
Fujise
,
H.
,
Zade-Oppen
,
A. M.
,
Ryu
,
K. H.
and
Delpire
,
E.
(
1992
).
Erythrocyte K+–Cl cotransport: properties and regulation
.
Am. J. Physiol
.
263
,
C917
C132
.
Lemmink
,
H. H.
,
Van Den Heuvel
,
L. P.
,
Van Dijk
,
H. A.
,
Merkx
,
G. F.
,
Smilde
,
T. J.
,
Taschner
,
P. E.
,
Monnens
,
L. A.
,
Hebert
,
S. C.
and
Knoers
,
N. V.
(
1996
).
Linkage of Gitelman syndrome to the thiazide-sensitive sodium–chloride cotransporter gene with identification of mutations in Dutch families
.
Pediatric Nephrol
.
10
,
403
407
.
Ling
,
B. N.
,
Matsunaga
,
H.
,
Ma
,
H.
and
Eaton
,
D. C.
(
1995
).
Role of growth factors in mesangial cell ion channel regulation
.
Kidney Int
.
48
,
1158
1166
.
Lorenz
,
J. N.
,
Weihprecht
,
H.
,
Schnermann
,
J.
,
Skott
,
O.
and
Briggs
,
J. P.
(
1991
).
Renin release from isolated juxtaglomerular apparatus depends on macula densa chloride transport
.
Am. J. Physiol
.
260
,
F486
F493
.
Ludens
,
J. H.
,
Clark
,
M. A.
and
Lawson
,
J. A.
(
1995
).
Does ADH alter cotransporter properties in conscious rats?: evidence for a shift from K+-independent to K+-dependent cotransport
.
J. Am. Soc. Nephrol
.
6
,
344
.
Lytle
,
C.
and
Forbush
III,
B.
(
1992
).
The Na+–K+–Cl cotransport protein of shark rectal gland. II. Regulation by direct phosphorylation
.
J. biol. Chem
.
267
,
25438
25443
.
Lytle
,
C.
and
Forbush
III,
B.
(
1996
).
Regulatory phosphorylation of the secretory Na+–K+–Cl cotransporter: modulation by cytoplasmic Cl
.
Am. J. Physiol
.
270
,
C437
C448
.
Lytle
,
C.
,
Xu
,
J. C.
,
Biemesderfer
,
D.
and
Forbush
III
, B
. (
1995
).
Distribution and diversity of Na+–K+–Cl cotransport proteins: a study with monoclonal antibodies
.
Am. J. Physiol
.
269
,
C1496
C1505
.
Mastroianni
,
N.
,
Bettinelli
,
A.
,
Bianchetti
,
M.
,
Colussi
,
G.
,
De Fusco
,
M.
,
Sereni
,
F.
,
Ballabio
,
A.
and
Casari
,
G.
(
1996
).
Novel molecular variants of the Na+–Cl cotransporter gene are responsible for Gitelman syndrome
.
Am. J. Human Genet
.
59
,
1019
1026
.
Matsunaga
,
H.
,
Yamashita
,
N.
,
Okuda
,
T.
and
Kurokawa
,
K.
(
1994
).
Mesangial cell ion transport and tubuloglomerular feedback
.
Curr. Opin. Nephrol. Hypertens
.
3
,
518
522
.
Misgeld
,
U.
,
Deisz
,
R. A.
,
Dodt
,
H. U.
and
Lux
,
H. D.
(
1986
).
The role of chloride transport in postsynaptic inhibition of hippocampal neurons
.
Science
232
,
1413
1415
.
Mount
,
D. B.
,
Hall
,
A. E.
,
Plata
,
C.
,
Vilanueva
,
Y.
,
Kaplan
,
M. R.
,
Gamba
,
G.
and
Hebert
,
S. C.
(
1995
).
Characterization of alternatively spliced transcripts of the murine apical bumetanide-sensitive Na+–K+–Cl cotransporter gene
.
J. Am. Soc. Nephrol
.
6
,
347
.
Mount
,
D. B.
,
Hoover
,
R. S.
and
Hebert
,
S. C.
(
1997
).
The molecular physiology of electroneutral cation–chloride cotransport
.
J. Membr. Biol
.
158
,
177
186
.
Obermüller
,
N.
,
Bernstein
,
P.
,
Velazquez
,
H.
,
Reilly
,
R.
,
Moser
,
D.
,
Ellison
,
D. H.
and
Bachmann
,
S.
(
1995
).
Expression of the thiazide-sensitive Na+–Cl cotransporter in rat and human kidney
.
Am. J. Physiol
.
269
,
F900
F910
.
Obermüller
,
N.
,
Kunchaparty
,
S.
,
Ellison
,
D. H.
and
Bachmann
,
S.
(
1996
).
Expression of the Na+–K+–2Cl cotransporter by macula densa and thick ascending limb cells of rat and rabbit nephron
.
J. clin. Invest
.
98
,
635
640
.
Payne
,
J. A.
(
1997
).
Functional characterization of the neuronal-specific K+–Cl cotransporter: implications for [K+]o regulation
.
Am. J. Physiol
.
273
,
C1516
C1525
.
Payne
,
J. A.
and
Forbush
III
, B
. (
1994
).
Alternatively spliced isoforms of the putative renal Na+–K+–Cl cotransporter are differentially distributed within the rabbit kidney
.
Proc. natn. Acad. Sci. U.S.A
.
91
,
4544
4548
.
Payne
,
J. A.
,
Stevenson
,
T. J.
and
Donaldson
,
L. F.
(
1996
).
Molecular characterization of a putative K+–Cl cotransporter in rat brain. A neuronal-specific isoform
.
J. biol. Chem
.
271
,
16245
16252
.
Payne
,
J. A.
,
Xu
,
J. C.
,
Haas
,
M.
,
Lytle
,
C. Y.
,
Ward
,
D.
and
Forbush
III
, B
. (
1995
).
Primary structure, functional expression and chromosomal localization of the bumetanide-sensitive Na+–K+–Cl cotransporter in human colon
.
J. biol. Chem
.
270
,
17977
17985
.
Plotkin
,
M. D.
,
Kaplan
,
M. R.
,
Verlander
,
J. W.
,
Lee
,
W.-S.
,
Brown
,
D.
,
Poch
,
E.
,
Gullans
,
S. R.
and
Hebert
,
S. C.
(
1996
).
Localization of the thiazide-sensitive Na+–Cl cotransporter, rTSC1, in the rat kidney
.
Kidney Int
.
50
,
174
183
.
Plotkin
,
M. D.
,
Kaplan
,
M. R.
,
Peterson
,
L. N.
,
Gullans
,
S. R.
,
Hebert
,
S. C.
and
Delpire
,
E.
(
1997a
).
Expression of the Na+–K+–2Cl cotransporter, BSC2, in the nervous system
.
Am. J. Physiol
.
272
,
C173
C183
.
Plotkin
,
M. D.
,
Snyder
,
E. Y.
,
Hebert
,
S. C.
and
Delpire
,
E.
(
1997b
).
Expression of the Na–K–2Cl cotransporter is developmentally regulated in postnatal rat brains: a possible mechanism underlying GABA’s excitatory role in immature brain
.
J. Neurobiol
.
30
,
781
795
.
Poch
,
E.
,
Suastegui
,
R.
,
Gamba
,
G.
and
Hebert
,
S. C.
(
1996
).
Role of N-linked glycosylation in rat thiazide-sensitive Na+–Cl cotransporter
.
J. Am. Soc. Nephrol
.
7
,
1288
.
Pollak
,
M. R.
,
Delaney
,
V. B.
,
Graham
,
R. M.
and
Hebert
,
S. C.
(
1996
).
Gitelman’s syndrome (Bartter’s variant) maps to the thiazide-sensitive cotransporter gene locus on chromosome 16q13 in a large kindred
.
J. Am. Soc. Nephrol
.
7
,
2244
2248
.
Raat
,
N. J.
,
Delpire
,
E.
,
Van Os
,
C. H.
and
Bindels
,
R. J.
(
1996
).
Culturing induced expression of basolateral Na+–K+–2Cl cotransporter BSC2 in proximal tubule, aortic endothelium and vascular smooth muscle
.
Pflügers Arch
.
431
,
458
460
.
Randall
,
J.
,
Thorne
,
T.
and
Delpire
,
E.
(
1997
).
Partial cloning and characterization of Slc12a2: the gene encoding the secretory Na+–K+–2Cl cotransporter
.
Am. J. Physiol
.
273
,
C1267
C1277
.
Rocha
,
A. S.
and
Kudo
,
L. H.
(
1990
).
Atrial peptide and cGMP effects on Na+–Cl transport in inner medullary collecting duct
.
Am. J. Physiol
.
259
,
F258
F268
.
Simon
,
D. B.
,
Bindra
,
R. S.
,
Mansfield
,
T. A.
,
Nelson-Williams
,
C.
,
Mendonca
,
E.
,
Stone
,
R.
,
Schurman
,
S.
,
Nayir
,
A.
,
Alpay
,
H.
,
Bakkaloglu
,
A.
,
Rodriguez-Soriano
,
J.
,
Morales
,
J. M.
,
Sanjad
,
S. A.
,
Taylor
,
C. M.
,
Pilz
,
D.
,
Brem
,
A.
,
Trachtman
,
H.
,
Griswold
,
W.
,
Richard
,
G. A.
,
John
,
E.
and
Lifton
,
R. P.
(
1997
).
Mutations in the chloride channel gene, CLCNKB, cause Bartter’s syndrome type III
.
Nature Genet
.
17
,
171
178
.
Simon
,
D. B.
,
Karet
,
F. E.
,
Hamdan
,
J. M.
,
Dipietro
,
A.
,
Sanjad
,
S. A.
and
Lifton
,
R. P.
(
1996a
).
Bartter’s syndrome, hypokalaemic alkalosis with hypercalciuria, is caused by mutations in the Na–K–2Cl cotransporter NKCC2
.
Nature Genet
.
13
,
183
188
.
Simon
,
D. B.
,
Karet
,
F. E.
,
Rodriquez-Soriano
,
J.
,
Hamdan
,
J. H.
,
Dipietro
,
A.
,
Trachtman
,
H.
,
Sanjad
,
S. A.
and
Lifton
,
R. P.
(
1996b
).
Genetic heterogeneity of Bartter’s syndrome revealed by mutations in the K+ channel, ROMK
.
Nature Genet
.
14
,
152
156
.
Simon
,
D. B.
,
Nelson-Williams
,
C.
,
Bia
,
M. J.
,
Ellison
,
D.
,
Karet
,
F. E.
,
Molina
,
A. M.
,
Vaara
,
I.
,
Iwata
,
F.
,
Cushner
,
H. M.
,
Koolen
,
M.
,
Gainza
,
F. J.
,
Gitleman
,
H. J.
and
Lifton
,
R. P.
(
1996c
).
Gitelman’s variant of Bartter’s syndrome, inherited hypokalaemic alkalosis, is caused by mutations in the thiazide-sensitive Na–Cl cotransporter
.
Nature Genet
.
12
,
24
30
.
Soybel
,
D. I.
,
Gullans
,
S. R.
,
Maxwell
,
F.
and
Delpire
,
E.
(
1995
).
Role of basolateral Na+–K+–Cl cotransport in HCl secretion by amphibian gastric mucosa
.
Am. J. Physiol
.
269
,
C242
C249
.
Staley
,
K.
,
Smith
,
R.
,
Schaack
,
R.
,
Wilcox
,
C.
and
Jentsch
,
T. J.
(
1996
).
Alteration of GABAA receptor function following gene transfer of the CLC-2 chloride channel
.
Neuron
17
,
543
551
.
Stein
,
J. H.
(
1985
).
The pathogenetic spectrum of Bartter’s syndrome (clinical conference)
.
Kidney Int
.
28
,
85
93
.
Stokes
,
J. B.
(
1989
).
Electroneutral Na+–Cl transport in the distal tubule
.
Kidney Int
.
36
,
427
433
.
Sun
,
A.
,
Grossman
,
E. B.
,
Lombardi
,
M.
and
Hebert
,
S. C.
(
1991
).
Vasopressin alters the mechanism of apical Cl entry from Na+–Cl to Na+–K+–2Cl cotransport in mouse medullary thick ascending limb
.
J. Membr. Biol
.
120
,
83
94
.
Takeuchi
,
K.
,
Kure
,
S.
,
Kato
,
T.
,
Taniyama
,
Y.
,
Takahashi
,
N.
,
Ikeda
,
Y.
,
Abe
,
T.
,
Narisawa
,
K.
,
Muramatsu
,
Y.
and
Abe
,
K.
(
1996
).
Association of a mutation in thiazide-sensitive Na+–Cl?cotransporter with familial Gitelman’s syndrome
.
J. clin. Endocr. Metab
.
81
,
4496
4499
.
Tanimura
,
A.
,
Kurihara
,
K.
,
Reshkin
,
S. J.
and
Turner
,
R. J.
(
1995
).
Involvement of direct phosphorylation in the regulation of the rat parotid Na+–K+–2Cl cotransporter
.
J. biol. Chem
.
270
,
25252
25258
.
Tsukahara
,
H.
,
Krivenko
,
Y.
,
Moore
,
L. C.
and
Goligorsky
,
M. S.
(
1994
).
Decrease in ambient [Cl] stimulates nitric oxide release from cultured rat mesangial cells
.
Am. J. Physiol
.
267
,
F190
F195
.
Tsuruoka
,
S.
,
Koseki
,
C.
,
Muto
,
S.
,
Tabei
,
K.
and
Imai
,
M.
(
1994
).
Axial heterogeneity of potassium transport across hamster thick ascending limb of Henle’s loop
.
Am. J. Physiol
.
267
,
F121
F129
.
Wall
,
S. M.
,
Trinh
,
H. N.
and
Woodward
,
K. E.
(
1995
).
Heterogeneity of NH4 transport in mouse inner medullary collecting duct cells
.
Am. J. Physiol
.
38
,
F536
F544
.
Xu
,
J. Z.
,
Hall
,
A. E.
,
Peterson
,
L. N.
,
Bienkowski
,
M. J.
,
Eessalu
,
T. E.
and
Hebert
,
S. C.
(
1997
).
Localization of the ROMK protein on apical membranes of rat kidney nephron segments
.
Am. J. Physiol
.
273
,
F739
F748
.
Xu
,
J. C.
,
Lytle
,
C.
,
Zhu
,
T. T.
,
Payne
,
J. A.
,
Benz
,
E.
, Jr
and
Forbush
III
, B
. (
1994
).
Molecular cloning and functional expression of the bumetanide-sensitive Na+–K+–Cl cotransporter
.
Proc. natn. Acad. Sci. U.S.A
.
91
,
2201
2205
.
Yerby
,
T. R.
,
Vibat
,
C. R. T.
,
Sun
,
D.
,
Payne
,
J. A.
and
O’donnell
,
M. E.
(
1997
).
Molecular characterization of the Na+–K+–Cl cotransporter of bovine aortic endothelial cells
.
Am. J. Physiol
.
273
,
C188
C197
.
Yu
,
H.
,
Papa
,
F.
and
Sukhatme
,
V. P.
(
1994
).
Bovine and rodent Tamm–Horsfall protein (THP) genes: cloning, structural analysis and promoter identification
.
Gene Expr
.
4
,
63
75
.