24.12.2012 Views

SLC4 base (HCO3 , CO3 2 ) transporters ... - Renal Physiology

SLC4 base (HCO3 , CO3 2 ) transporters ... - Renal Physiology

SLC4 base (HCO3 , CO3 2 ) transporters ... - Renal Physiology

SHOW MORE
SHOW LESS

You also want an ePaper? Increase the reach of your titles

YUMPU automatically turns print PDFs into web optimized ePapers that Google loves.

<strong>SLC4</strong> <strong>base</strong> (HCO − 2−<br />

3 , <strong>CO3</strong><br />

) <strong>transporters</strong>: classification,<br />

function, structure, genetic diseases, and knockout models<br />

Alexander Pushkin and Ira Kurtz<br />

Am J Physiol <strong>Renal</strong> Physiol 290:F580-F599, 2006. ;<br />

doi: 10.1152/ajprenal.00252.2005<br />

You might find this additional info useful...<br />

This article cites 233 articles, 123 of which you can access for free at:<br />

http://ajprenal.physiology.org/content/290/3/F580.full#ref-list-1<br />

This article has been cited by 29 other HighWire-hosted articles:<br />

http://ajprenal.physiology.org/content/290/3/F580#cited-by<br />

Updated information and services including high resolution figures, can be found at:<br />

http://ajprenal.physiology.org/content/290/3/F580.full<br />

Additional material and information about American Journal of <strong>Physiology</strong> - <strong>Renal</strong> <strong>Physiology</strong> can be<br />

found at:<br />

http://www.the-aps.org/publications/ajprenal<br />

This information is current as of December 23, 2012.<br />

American Journal of <strong>Physiology</strong> - <strong>Renal</strong> <strong>Physiology</strong> publishes original manuscripts on a broad range of subjects<br />

relating to the kidney, urinary tract, and their respective cells and vasculature, as well as to the control of body fluid<br />

volume and composition. It is published 12 times a year (monthly) by the American Physiological Society, 9650<br />

Rockville Pike, Bethesda MD 20814-3991. Copyright © 2006 the American Physiological Society. ISSN:<br />

0363-6127, ESSN: 1522-1466. Visit our website at http://www.the-aps.org/.<br />

Downloaded from<br />

http://ajprenal.physiology.org/<br />

by guest on December 23, 2012


Invited Review<br />

� 2�<br />

<strong>SLC4</strong> <strong>base</strong> (<strong>H<strong>CO3</strong></strong> ,<strong>CO3</strong>) <strong>transporters</strong>: classification, function, structure,<br />

genetic diseases, and knockout models<br />

Alexander Pushkin and Ira Kurtz<br />

Division of Nephrology, David Geffen School of Medicine at UCLA, Los Angeles, California<br />

� 2<br />

Pushkin, Alexander, and Ira Kurtz. <strong>SLC4</strong> <strong>base</strong> (<strong>H<strong>CO3</strong></strong> ,<strong>CO3</strong><br />

�<br />

) <strong>transporters</strong>:<br />

classification, function, structure, genetic diseases, and knockout models. Am J<br />

Physiol <strong>Renal</strong> Physiol 290: F580–F599, 2006; doi:10.1152/ajprenal.00252.2005.—<br />

In prokaryotic and eukaryotic organisms, biochemical and physiological processes<br />

are sensitive to changes in H� activity. For these processes to function optimally,<br />

a variety of proteins have evolved that transport H� /<strong>base</strong> equivalents across cell<br />

and organelle membranes, thereby maintaining the pH of various intracellular and<br />

extracellular compartments within specific limits. The <strong>SLC4</strong> family of <strong>base</strong><br />

� 2�<br />

(<strong>H<strong>CO3</strong></strong> ,<strong>CO3</strong>)<br />

transport proteins plays an essential role in mediating Na�- and/or<br />

Cl�-dependent <strong>base</strong> transport in various tissues and cell types in mammals. In<br />

addition to pH regulation, specific members of this family also contribute to<br />

vectorial transepithelial <strong>base</strong> transport in several organ systems including the<br />

kidney, pancreas, and eye. The importance of these <strong>transporters</strong> in mammalian cell<br />

biology is highlighted by the phenotypic abnormalities resulting from spontaneous<br />

<strong>SLC4</strong> mutations in humans and targeted deletions in murine knockout models. This<br />

review focuses on recent advances in our understanding of the molecular organization<br />

and functional properties of <strong>SLC4</strong> <strong>transporters</strong> and their role in disease.<br />

bicarbonate; transport<br />

� 2<br />

MEMBRANE PROTEINS THAT MEDIATE <strong>base</strong> (<strong>H<strong>CO3</strong></strong> ,<strong>CO3</strong><br />

�<br />

) transport<br />

play an important role in eukaryotes in maintaining both<br />

intracellular pH (pHi) and extracellular pH (pHo) within nar-<br />

�<br />

row limits. This is not surprising given that the <strong>H<strong>CO3</strong></strong> /<br />

2� CO 3<br />

/CO2 buffer system is arguably the most important buffer<br />

in eukaryotic organisms. In mammals, two unrelated multigene<br />

families, <strong>SLC4</strong> and SLC26 <strong>transporters</strong>, have evolved, whose<br />

members transport <strong>base</strong> in both a cell type- and cell membranespecific<br />

fashion. This review focuses specifically on the <strong>SLC4</strong><br />

transporter family. The reader is referred to recent reviews on<br />

SLC26 <strong>transporters</strong> for an update on the role these <strong>transporters</strong><br />

play in health and disease states (46, 131). Following an initial<br />

overview of the classification of the <strong>SLC4</strong> family of transport<br />

proteins, we highlight recent advances in our understanding of<br />

the structure and function of these <strong>transporters</strong> and the role of<br />

protein interactions in the regulation of their functional properties.<br />

Abnormalities in either the function and/or membrane<br />

targeting of certain members of the <strong>SLC4</strong> family are the cause<br />

of various genetic diseases in humans. Mouse knockout models<br />

have also provided a useful adjunct with which to study the<br />

role of <strong>SLC4</strong> family members in specific organ systems.<br />

Advances in our understanding of the role of <strong>SLC4</strong> <strong>transporters</strong><br />

in health and disease are also the focus of this review.<br />

CLASSIFICATION OF <strong>SLC4</strong> TRANSPORTERS<br />

Based on the nature of the specific transport process involved,<br />

e.g., exchange, cotransport, and the specific ions that<br />

are transported, <strong>SLC4</strong> family members can be conveniently<br />

Address for reprint requests and other correspondence: I. Kurtz, Div. of<br />

Nephrology, David Geffen School of Medicine at UCLA, 10833 Le Conte<br />

Ave., Rm. 7–155 Factor Bldg., Los Angeles, CA 90095 (e-mail:<br />

ikurtz@mednet.ucla.edu).<br />

F580<br />

separated into three functional groups with potential transport<br />

� 2<br />

modes for <strong>H<strong>CO3</strong></strong> and/or <strong>CO3</strong><br />

�<br />

(Fig. 1): 1) Na�-independent Cl� �<br />

/<strong>H<strong>CO3</strong></strong> exchangers mediating electroneutral exchange of<br />

Cl� � 2<br />

for <strong>base</strong> (<strong>H<strong>CO3</strong></strong> ,<strong>CO3</strong><br />

�<br />

); 2) Na� �<br />

-<strong>H<strong>CO3</strong></strong> co<strong>transporters</strong><br />

mediating cotransport of Na� � 2<br />

and <strong>base</strong> (<strong>H<strong>CO3</strong></strong> ,<strong>CO3</strong><br />

�<br />

); and 3)<br />

Na�-driven Cl� �<br />

/<strong>H<strong>CO3</strong></strong> exchangers mediating the electroneutral<br />

exchange of Cl� for Na� � 2<br />

and <strong>base</strong> (<strong>H<strong>CO3</strong></strong> ,<strong>CO3</strong><br />

�<br />

). <strong>SLC4</strong><br />

proteins have in common the property of transporting <strong>base</strong><br />

� 2<br />

(<strong>H<strong>CO3</strong></strong> , <strong>CO3</strong><br />

�<br />

) 1 but differ in their ability to mediate the<br />

concomitant transport of Na � and or Cl � . In addition, the<br />

cotransport of Na � by most <strong>SLC4</strong> <strong>transporters</strong> distinguishes<br />

this family from SLC26 proteins, which predominantly mediate<br />

Na � -independent anion transport. The sequences and classification<br />

of human <strong>SLC4</strong> <strong>transporters</strong> are shown in Fig. 2 and<br />

Table 1.<br />

Na�-Independent Cl� �<br />

/<strong>H<strong>CO3</strong></strong> Exchangers<br />

<strong>SLC4</strong>A1 (AE1). AE1 mediates Na�-independent anion exchange<br />

in red blood cells and the renal collecting duct. Two<br />

variants of AE1 have been well characterized (6, 25, 33, 107,<br />

123, 196). Red cell AE1 (eAE1) or band 3 protein (human<br />

�<br />

eAE1, 911 aa) plays an important role in transporting <strong>H<strong>CO3</strong></strong> to<br />

the lung that was derived from metabolically produced CO2.In<br />

the kidney, AE1 (kAE1) is transcribed from an alternate<br />

is �500:1, whereas at a typical pHi of 7.1,<br />

is twice this value, or �1,000:1. It is currently unclear as to<br />

(or both) is transported in various <strong>SLC4</strong> <strong>transporters</strong>.<br />

1 At pHo of 7.4, the <strong>H<strong>CO3</strong></strong> � /<strong>CO3</strong> 2 �<br />

Am J Physiol <strong>Renal</strong> Physiol 290: F580–F599, 2006;<br />

doi:10.1152/ajprenal.00252.2005.<br />

� 2<br />

the <strong>H<strong>CO3</strong></strong> /<strong>CO3</strong> �<br />

� 2<br />

whether <strong>H<strong>CO3</strong></strong> or <strong>CO3</strong> �<br />

The various potential transport modes are depicted in Fig. 1. At the present<br />

�<br />

time, the term “bicarbonate (<strong>H<strong>CO3</strong></strong> )” that is used in naming or describing the<br />

function (see Table 1) of <strong>SLC4</strong> <strong>transporters</strong> should not be viewed as a<br />

definitive description of the type of <strong>base</strong> transported, but rather as a term that<br />

� 2<br />

generically refers to <strong>H<strong>CO3</strong></strong> and/or <strong>CO3</strong> �<br />

until definitive studies are reported<br />

characterizing the specific species transported.<br />

0363-6127/06 $8.00 Copyright © 2006 the American Physiological Society http://www.ajprenal.org<br />

Downloaded from<br />

http://ajprenal.physiology.org/<br />

by guest on December 23, 2012


Fig. 1. <strong>SLC4</strong> <strong>transporters</strong> can be divided functionally into 3 groups: Na�- independent Cl� � � �<br />

/<strong>H<strong>CO3</strong></strong> exchangers (A), Na -<strong>H<strong>CO3</strong></strong> co<strong>transporters</strong> (B), and<br />

Na�-driven Cl� � �<br />

/<strong>H<strong>CO3</strong></strong> exchangers (C). Potential transport modes for <strong>H<strong>CO3</strong></strong><br />

2�<br />

and <strong>CO3</strong> are depicted. In B,Na� �<br />

-<strong>H<strong>CO3</strong></strong> co<strong>transporters</strong> are further subdivided:<br />

(i) electroneutral Na� � � 2�<br />

-<strong>H<strong>CO3</strong></strong> co<strong>transporters</strong> with a <strong>H<strong>CO3</strong></strong> (<strong>CO3</strong> )-to-Na� transport stoichiometry of 1:1, and electrogenic Na� �<br />

-<strong>H<strong>CO3</strong></strong> co<strong>transporters</strong><br />

� 2�<br />

with a <strong>H<strong>CO3</strong></strong> (<strong>CO3</strong><br />

)-to-Na� transport stoichiometry of either 2:1 (ii) or3:1<br />

(iii) are shown. It has not been definitively established whether <strong>SLC4</strong> proteins<br />

� 2�<br />

mediate <strong>H<strong>CO3</strong></strong> and/or <strong>CO3</strong> transport. Depicted are the hypothetical modes of<br />

� 2�<br />

transport, wherein the coupling ratio of <strong>base</strong> (<strong>H<strong>CO3</strong></strong> and/or <strong>CO3</strong> ) transport to<br />

Na � and/or Cl � is equivalent. Note that these transport modes are not<br />

equivalent with regard to the total flux of carbon, hydrogen, and oxygen per<br />

transport cycle.<br />

promoter, and the resultant protein has an NH2-terminal truncation<br />

(first 65 amino acids in human AE1) (25). Kidneyspecific<br />

AE1 is expressed on the basolateral membrane of<br />

collecting duct �-intercalated cells (7, 205), where it plays an<br />

important role in bicarbonate reabsorption and urinary<br />

acidification.<br />

<strong>SLC4</strong>A2 (AE2). AE2 mediates Na � -independent anion exchange<br />

and is more widely expressed than AE1. Five NH2terminal<br />

splice variants of AE2 (a, b1, b2, c1, and c2) have<br />

been described in mammals and are expressed in a tissue- and<br />

cell-specific manner (5, 6, 8–10, 13, 40, 110, 129, 168, 217).<br />

Human AE2 is �300 aa larger than eAE1 and, like AE1, is<br />

glycosylated (6, 233). Deglycosylation of AE2 does not appre-<br />

� 2�<br />

<strong>SLC4</strong> BASE (<strong>H<strong>CO3</strong></strong> ,<strong>CO3</strong>) TRANSPORTERS<br />

AJP-<strong>Renal</strong> Physiol • VOL 290 • MARCH 2006 • www.ajprenal.org<br />

Invited Review<br />

F581<br />

ciably affect its transport function (53). AE2 is widely expressed<br />

in nonexcitable tissues, where it has been proposed to<br />

be a housekeeping Cl � /HCO 3 � . It is also highly expressed on<br />

the basolateral membrane of choroid-plexus epithelial cells (9)<br />

and gastric parietal cells (191). In hepatocytes, AE2a, AE2b1,<br />

and AE2b2 have been localized to the apical membrane (13).<br />

In the kidney, AE2 is expressed in the basolateral membrane of<br />

the thick ascending limb (TAL) (6, 10). Ion transport mediated<br />

by AE2 expressed in Xenopus laevis oocytes (77, 188, 229) and<br />

mammalian cells (91, 115) is pH sensitive. Residues in the<br />

cytoplasmic NH2 terminus of mouse AE2, aa 336–347, 357–<br />

362, and 391–510, have been shown to regulate the pH dependence<br />

of the transporter (188–190, 229).<br />

<strong>SLC4</strong>A3 (AE3). AE3 mediates Na � -independent anion exchange<br />

like AE1 and AE2. Two NH2-terminal splice variants<br />

of AE3, cardiac (cAE3) and brain (bAE3), have been described<br />

(108, 118). The size of bAE3 (1,232 aa in humans) is comparable<br />

to AE2, whereas cAE3 is �200 aa shorter. Transcripts of<br />

both variants were found in different organs and tissues including<br />

the brain and heart (106, 108, 110, 118, 225). bAE3 protein<br />

has also been detected in basolateral membrane vesicles isolated<br />

from human ileal and colonic epithelial cells (11). AE3 is<br />

also expressed in the basal end foot processes of retinal Müller<br />

cells (106). AE3 expressed in X. laevis oocytes and HEK293<br />

cells mediates significantly less Cl � /HCO 3 � exchange than<br />

does AE1 or AE2 (53, 77, 115, 165, 225). Unlike AE2, when<br />

expressed in HEK cells, AE3 does not appear to be glycosylated<br />

(53). The decreased rate of AE3-mediated anion exchange<br />

in HEK cells has been attributed to a low level of<br />

plasma membrane targeting. cAE3 transport is stimulated at an<br />

alkaline pH similar to AE2 (188).<br />

Sodium Bicarbonate Co<strong>transporters</strong><br />

Sodium bicarbonate co<strong>transporters</strong> mediate cotransport of<br />

Na� � 2<br />

and <strong>base</strong> (<strong>H<strong>CO3</strong></strong> ,<strong>CO3</strong><br />

�<br />

). 2 Depending on the flux of the<br />

net charge per transport cycle, these <strong>transporters</strong> are characterized<br />

�<br />

as electroneutral [i.e., transported negative charge (<strong>H<strong>CO3</strong></strong> and/or<br />

2<br />

<strong>CO3</strong> �<br />

)] equals transported positive charge (Na� ), or electro-<br />

� 2<br />

genic [i.e., transported negative charge (<strong>H<strong>CO3</strong></strong> and/or <strong>CO3</strong><br />

�<br />

)]<br />

exceeds transported positive charge (Na � ).<br />

<strong>SLC4</strong>A4 (NBC1, NBCe1). NBC1 is an electrogenic sodium<br />

bicarbonate cotransporter that was first cloned from Ambystoma<br />

tigrinum (166) and human (30) kidney. In humans, two<br />

major NBC1 variants, kNBC1 (NBCe1-A) (1,035 aa) and<br />

pNBC1 (NBCe1-B) (1,079 aa), that differ in their extreme NH2<br />

terminus are differentially expressed in a cell-specific manner<br />

(2) and are derived from alternate promoter usage in the<br />

<strong>SLC4</strong>A4 gene (4). A third NBC1 variant, rb2NBC (NBCe1c),<br />

has been cloned from rat brain and is identical to rat pNBC1<br />

except for 61 COOH-terminal residues and a COOH-terminal<br />

PDZ motif (18). Recently, a similar splice variant of the<br />

<strong>SLC4</strong>A4 gene has been detected in human and porcine vas<br />

2 As depicted in Fig. 1, in electrogenic sodium bicarbonate co<strong>transporters</strong>, a<br />

�<br />

2:1 stoichiometry could result for example from the transport of 2<strong>H<strong>CO3</strong></strong> �<br />

1Na� 2<br />

, 1<strong>CO3</strong> �<br />

� 1Na� � 2<br />

, or 2<strong>H<strong>CO3</strong></strong> � 1<strong>CO3</strong> �<br />

� 2Na� . A 3:1 cotransporter<br />

� � stoichiometry could result from the transport of either 3<strong>H<strong>CO3</strong></strong> � 1Na ,<br />

2<br />

3<strong>CO3</strong> �<br />

� 2Na� � 2<br />

, or 1<strong>H<strong>CO3</strong></strong> � 1<strong>CO3</strong> �<br />

� 1Na� . Although the total charge<br />

transported is identical, it is clear that the total carbon, hydrogen, and oxygen<br />

flux per transport cycle will differ and be dependent on the specific <strong>base</strong><br />

species transported, e.g., HCO 3 � vs. <strong>CO3</strong> 2 �<br />

.<br />

Downloaded from<br />

http://ajprenal.physiology.org/<br />

by guest on December 23, 2012


Invited Review<br />

F582 � 2�<br />

<strong>SLC4</strong> BASE (<strong>H<strong>CO3</strong></strong> ,<strong>CO3</strong>) TRANSPORTERS<br />

Fig. 2. Alignment of human <strong>SLC4</strong> <strong>transporters</strong>: <strong>SLC4</strong>A1 (NP_000333), <strong>SLC4</strong>A2 (NP_003031), <strong>SLC4</strong>A3 (NP_005061), <strong>SLC4</strong>A4 (NP_003750), <strong>SLC4</strong>A5<br />

(NP_597812), <strong>SLC4</strong>A7 (NP_003606), <strong>SLC4</strong>A8 (NP_004849), <strong>SLC4</strong>A9 (NP_113655), and <strong>SLC4</strong>A10 (NP_071341). The alignment was performed using<br />

GeneWorks 2.5.1 software (Oxford Molecular Group, Campbell, CA). Putative transmembrane segments are shown in red (<strong>SLC4</strong>A1) and blue (<strong>SLC4</strong>A4) <strong>base</strong>d<br />

on topological models (198, 232). Amino acids shown to be important for <strong>SLC4</strong>A1- and <strong>SLC4</strong>A4-mediated transport (1, 35, 87, 97, 132, 133, 195, 231) are shown<br />

in bold font. Amino acids important for plasma membrane expression of <strong>SLC4</strong>A1 (54, 231, 232) and <strong>SLC4</strong>A4 (1, 75, 116) are marked with brown stars. Lysine<br />

residues shown to covalently bind DIDS and H2DIDS in <strong>SLC4</strong>A1 (93, 140) or homologous lysine residues in <strong>SLC4</strong>A4 predicted to covalently bind<br />

DIDS/H2DIDS are marked with ❖. Note that AE2 (<strong>SLC4</strong>A2), which has been shown to be inhibitable by DIDS (6), has methionine (M 1181 ) in a second putative<br />

DIDS binding site instead of lysine, whereas NBC3 (<strong>SLC4</strong>A7), which has been shown to be insensitive to DIDS (155), has lysines in both putative DIDS binding<br />

sites. Glycosylation sites for <strong>SLC4</strong>A1 (33) and <strong>SLC4</strong>A4 (38) are depicted with a branched tree symbol. <strong>SLC4</strong>A1, <strong>SLC4</strong>A2, and <strong>SLC4</strong>A4 amino acids that have<br />

been shown to play critical roles in binding of carbonic anhydrase II (CAII) (68, 154, 163, 206–208) are shown in magenta. CAII binding sites are depicted with<br />

a magenta line. The COOH-terminal PKA phosphorylation site in NBC1 is shown in green (arrowhead). Note that the <strong>SLC4</strong>A11 transporter NaBC1 is not<br />

included in the alignment despite its distant homology to the other members of the <strong>SLC4</strong> family, because recent studies have shown that it functions as a<br />

Na � -dependent borate transporter (142, 143, 145, 194).<br />

AJP-<strong>Renal</strong> Physiol • VOL 290 • MARCH 2006 • www.ajprenal.org<br />

Downloaded from<br />

http://ajprenal.physiology.org/<br />

by guest on December 23, 2012


Table 1. Characteristics of <strong>SLC4</strong> <strong>transporters</strong><br />

deferens epithelial cells (148). NBC1 proteins have been<br />

shown to be glycosylated (38). kNBC1 is highly expressed on<br />

the basolateral membrane of the renal proximal tubule where it<br />

mediates sodium bicarbonate efflux (21, 127, 178, 223). In<br />

humans, the pNBC1 transcript is highly expressed in the<br />

pancreas and to a lesser extent in various other tissues (2).<br />

Unlike the kidney, wherein all studies agree on the localization<br />

of kNBC1, the expression pattern of pNBC1 in the<br />

pancreas is complex. NBC1 was localized with antibodies<br />

common to kNBC1 and pNBC1 to the basolateral membrane of<br />

large ducts in the human pancreas (125). In the rat pancreas,<br />

NBC1 was localized to the basolateral membrane of acinar<br />

cells, the apical and basolateral membranes of centroacinar<br />

cells, intra- and extralobular ducts, and main pancreatic duct<br />

cells (199). Localization of the pNBC1 variant to the basolateral<br />

membrane of pancreatic duct cells was further confirmed<br />

with a pNBC1-specific antibody (21). Satoh and coauthors<br />

(177), using kNBC1- and pNBC1-specific antibodies, reported<br />

that in human pancreatic ducts pNBC1 was expressed basolaterally<br />

and acinar cells were unstained, whereas in the rat<br />

pancreas acinar cells were stained basolaterally, and ductal<br />

cells expressed pNBC1 both apically and basolaterally. In<br />

addition, kNBC1 was localized apically in some ducts. Recently,<br />

in a similar study, Roussa and coauthors (169) found<br />

that pNBC1 was localized to the basolateral membrane of rat<br />

ductal and acinar cells and that pNBC1 and kNBC1 were<br />

localized together to the apical membrane of certain ductal<br />

cells. Moreover, these authors reported that in rat kidney<br />

pNBC1, in addition to kNBC1, was also localized to the<br />

basolateral membrane of S1 and S2 segments of the proximal<br />

tubule. It is clear that additional studies are needed to determine<br />

the reason for the varying results among different studies.<br />

The expression pattern of pNBC1 has also been characterized<br />

in corneal endothelium, duodenum, epididymis, vas deferens,<br />

and parotid and submandibular glands, where it is<br />

thought to mediate basolateral sodium bicarbonate influx (21,<br />

48, 89, 150, 170, 192, 193, 203). Species differences may<br />

determine differences shown for relative distribution of<br />

kNBC1 and pNBC1. For example, pNBC1 is expressed in<br />

various rat eye tissues including cornea, conjunctiva, lens,<br />

ciliary body, and retina, whereas the expression of kNBC1 is<br />

� 2�<br />

<strong>SLC4</strong> BASE (<strong>H<strong>CO3</strong></strong> ,<strong>CO3</strong>) TRANSPORTERS<br />

Invited Review<br />

Human Gene Protein Name Function Stoichiometry Electrogenicity<br />

<strong>SLC4</strong>A1 AE1, band 3 Cl � /HCO 3 � exchange 1:1 Electroneutral<br />

<strong>SLC4</strong>A2 AE2 Cl � /HCO 3 � exchange 1:1 Electroneutral<br />

<strong>SLC4</strong>A3 AE3 Cl � /HCO 3 � exchange 1:1 Electroneutral<br />

<strong>SLC4</strong>A4 NBC1, NBCe1 Na � -HCO 3 � cotransport 2:1;3:1 Electrogenic<br />

<strong>SLC4</strong>A5 NBC4*, NBCe2 Na � -HCO 3 � cotransport 2:1;3:1 Electrogenic<br />

<strong>SLC4</strong>A7 NBC3, NBCn1 Na � -HCO 3 � cotransport 1:1 Electroneutral<br />

<strong>SLC4</strong>A8 NDCBE Na � -driven Cl � /HCO 3 � exchange 2:1:1 Electroneutral<br />

<strong>SLC4</strong>A9 AE4† Cl � /HCO 3 � exchange 1:1 Electroneutral<br />

Na � -HCO 3 � cotransport 1:1 Electroneutral<br />

<strong>SLC4</strong>A10 NCBE† Na � -driven Cl � /HCO 3 � exchange 2:1:1 Electroneutral<br />

Na � -HCO 3 � cotransport 1:1 Electroneutral<br />

Names used in the literature for <strong>SLC4</strong> <strong>transporters</strong> are as follows: <strong>SLC4</strong>A1: eAE1, (band 3), kAE1; <strong>SLC4</strong>A2: AE2a,b1,b2,c1,c2; <strong>SLC4</strong>A3: bAE3, cAE3;<br />

<strong>SLC4</strong>A4: kNBC1 (NBC1a, NBCe1-A); pNBC1 (NBCe1-B, NBC1b, dNBC1, hhNBC, rb1NBC); rb2NBC (NBCel-C); <strong>SLC4</strong>A5: NBC4a, NBC4b, NBC4c<br />

(NBCe2-C), NBC4d, NBC4e, NBC4f; <strong>SLC4</strong>A7: NBC3 (NBCn1-A), NBCn1-B, NBCn1-C, NBCn1-D, NBCn1-E; <strong>SLC4</strong>A8: NDCBE, kNBC-3; <strong>SLC4</strong>A9: AE4a,<br />

AE4b; <strong>SLC4</strong>A10: rb1NCBE (NCBE-B), rb2NCBE, NCBE-A. *The NBC4c splice variant functions in mammalian cells as an electrogenic sodium bicarbonate<br />

cotransporter (176). The NBC4e splice variant has been reported to function in Xenopus laevis as an electroneutral sodium bicarbonate cotransporter (222). †The<br />

functional properties of these <strong>transporters</strong> is controversial (see text for details and references). The stoichiometry for <strong>SLC4</strong>A4 and <strong>SLC4</strong>A5 <strong>transporters</strong> reflects<br />

the HCO 3 � :Na � coupling ratio. For <strong>SLC4</strong>A8 and <strong>SLC4</strong>A10 <strong>transporters</strong>, the 2:1:1 stoichiometry reflects the <strong>H<strong>CO3</strong></strong> � :Na:Cl � coupling ratio.<br />

AJP-<strong>Renal</strong> Physiol • VOL 290 • MARCH 2006 • www.ajprenal.org<br />

F583<br />

restricted to the conjunctiva (21). In the human eye, kNBC1<br />

and pNBC1 appear to be more evenly distributed (202).<br />

<strong>SLC4</strong>A5 (NBC4, NBCe2). NBC4 was originally cloned from<br />

a human cardiac cDNA library (156). Six human NBC4<br />

COOH-terminal splice variants (NBC4a–f: 1,137, 1,074,<br />

1,121, 1,040, 1,051, and 760 aa, respectively) (156, 157, 222)<br />

have been cloned. The NBC4c splice variant has been shown to<br />

mediate electrogenic sodium bicarbonate cotransport when<br />

expressed in mammalian mPCT cells (176) and X. laevis<br />

oocytes (210). NBC4e mediated electroneutral sodium bicarbonate<br />

cotransport when expressed in X. laevis oocytes (222).<br />

NBC4e was identical to NBC4c except for deletion of the last<br />

70 aa and 2 aa substitutions (H612Y and A699T) that are not<br />

present in the human genome data<strong>base</strong>. If confirmed, NBC4e<br />

would be the first known example of a sodium bicarbonate<br />

cotransporter whose transport stoichiometry and electrogenic<br />

properties change as a result of a deletion/modification of<br />

specific residues. NBC4 transcripts are expressed in several<br />

organs and tissues (156). In rat hepatocytes, NBC4c protein has<br />

been localized to the basolateral (sinusoidal) plasma membrane,<br />

whereas intrahepatic cholangiocytes stain apically (3).<br />

In the kidney, NBC4c is also expressed on the apical membrane<br />

of uroepithelial cells lining the renal pelvis (3), where the<br />

cotransporter may play an important role in protecting the renal<br />

parenchyma from alterations in urine pH. Kristensen and<br />

colleagues (109) have recently reported that NBC4 potentially<br />

regulates the pH of sarcolemmal vesicles in skeletal muscle.<br />

<strong>SLC4</strong>A7 (NBC3, NBCn1). The electroneutral sodium bicarbonate<br />

cotransporter NBC3 was originally cloned from human<br />

skeletal muscle (155). Human NBC3 consists of 1,214 aa and<br />

is the first known stilbene-insensitive member of the <strong>SLC4</strong><br />

family. Subsequently, the rat ortholog of NBC3, the electroneutral<br />

sodium bicarbonate cotransporter NBCn1, was cloned<br />

from rat aorta (37). NBC3 (114) and NBCn1 (211) are apparently<br />

glycosylated to a greater extent than NBC1 proteins.<br />

Three rat NBCn1 splice variants named NBCn1-(B–D) were<br />

also cloned from rat aorta (37).<br />

<strong>SLC4</strong>A7 proteins are expressed in several tissues. NBC3<br />

protein is expressed in rat renal connecting tubule (114) and in<br />

rat and rabbit renal collecting ducts in the apical membrane of<br />

�-intercalated cells and the basolateral membrane of �-inter-<br />

Downloaded from<br />

http://ajprenal.physiology.org/<br />

by guest on December 23, 2012


Invited Review<br />

F584 � 2�<br />

<strong>SLC4</strong> BASE (<strong>H<strong>CO3</strong></strong> ,<strong>CO3</strong>) TRANSPORTERS<br />

calated cells (114, 161). NBCn1 was detected in intercalated<br />

cells in the medulla and, in addition, at the basolateral membrane<br />

of the rat medullary TAL and inner medullary collecting<br />

duct cells (151, 211). NBC3 was also highly expressed on the<br />

apical membrane of narrow cells in caput epididymis and light<br />

(clear) cells in corpus and cauda epididymis (158). The cotransporter<br />

apparently plays a role in pHi regulation in the outer<br />

medullary and inner medullary collecting ducts (113, 151,<br />

227). In addition, the cotransporter plays a role in basolateral<br />

bicarbonate transport in the medullary TAL (23, 82, 139),<br />

proximal duodenal villus cells (150), the basolateral domain of<br />

choroid plexus epithelial cells (152), and submandibular glands<br />

(58, 122).<br />

Cooper and colleagues (42) recently reported a deletion<br />

variant (NBCn1-E) in rat hippocampal neurons that lacks 123<br />

amino acids in the cytoplasmic NH2-terminal domain. The<br />

cotransporter may play an important role in regulating hippocampal<br />

neuronal activity.<br />

Na�-Driven Cl� �<br />

/<strong>H<strong>CO3</strong></strong> Exchangers<br />

<strong>SLC4</strong>A8 (NDCBE). The first Na�-driven Cl� �<br />

/<strong>H<strong>CO3</strong></strong> exchanger<br />

(NDCBE) was cloned by Romero and coauthors (167)<br />

from Drosophila and named Na�-driven anion exchanger 1<br />

(NDAE1). Its transport was blocked by stilbenedisulfonates<br />

and might not be strictly electroneutral. In addition, NDAE1<br />

was able to use OH� instead of bicarbonate. NDCBE (1,044<br />

aa) cloned from the human brain by Grichtchenko and colleagues<br />

(60) when expressed in X. laevis oocytes functions as<br />

aNa�-driven Cl� �<br />

/<strong>H<strong>CO3</strong></strong> exchanger, has a strict requirement<br />

� �<br />

for <strong>H<strong>CO3</strong></strong> , and is DIDS sensitive. Although CO2/<strong>H<strong>CO3</strong></strong> stimulates<br />

NDAE1-induced pHi changes, it is not clear whether<br />

� �<br />

NDAE1 mediates <strong>H<strong>CO3</strong></strong> transport per se because CO2/<strong>H<strong>CO3</strong></strong> could simply dissipate OH� in the adjacent unstirred layer.<br />

NDAE1 expressed in X. laevis oocytes also differs from human<br />

NDCBE in that it is associated with a Cl� current, as well as<br />

�<br />

an inward current caused by CO2/<strong>H<strong>CO3</strong></strong> . NDCBE transcripts<br />

are present in the brain, testis, kidney, and ovary (60).<br />

A splice variant of NDCBE (originally named kNBC-3) was<br />

cloned from mouse inner medullary collecting duct<br />

(mIMCD-3) cells and when expressed in X. laevis oocytes<br />

mediated electroneutral sodium bicarbonate cotransport (212).<br />

The chloride dependence was not studied, and therefore its<br />

functional properties remain unclear. More recently, Virkki<br />

and coauthors (209) reported the cloning of a Na�-driven Cl� �<br />

/<strong>H<strong>CO3</strong></strong> exchanger of �1,200 aa from squid giant fiber lobe<br />

(sqNDCBE). Unlike the transport process described in native<br />

squid axon (22), reversal of the Na� gradient alters the direction<br />

sqNDCBE transport, and Li� can support pHi recovery<br />

from an acid load nearly as well as Na� .<br />

<strong>SLC4</strong> Transporters Whose Function is<br />

Incompletely Characterized<br />

<strong>SLC4</strong>A9 (AE4). AE4 was originally cloned (AE4a, 955 aa,<br />

and AE4b, 939 aa) from rabbit �-intercalated cells (202) and<br />

was localized to the apical membrane of renal collecting duct<br />

�-intercalated cell. Rabbit AE4a expressed in X. laevis oocytes<br />

and COS cells mediated Na � -independent, DIDS-insensitive<br />

Cl � /HCO 3 � exchange. In a separate study, AE4 was localized<br />

to apical and lateral membranes of collecting duct �-intercalated<br />

cells in the rabbit (105). However, in rat and mouse<br />

AJP-<strong>Renal</strong> Physiol • VOL 290 • MARCH 2006 • www.ajprenal.org<br />

collecting ducts, AE4 was detected on the basolateral membrane<br />

of �-intercalated cells (105). AE4 is also expressed on<br />

the basolateral membrane of mouse submandibular gland duct<br />

(105). In contrast to the data of Tsuganezawa and colleagues<br />

(201), rat AE4 expressed in LLC-PK1 and HEK293 cells<br />

mediated DIDS-sensitive Cl � /HCO 3 � exchange (105). Parker<br />

and colleagues (144) characterized human AE4 (945 aa) expressed<br />

in X. laevis oocytes as an electroneutral sodium bicarbonate<br />

cotransporter without Cl � /HCO 3 � exchange activity.<br />

Whether these discrepancies arise from species-mediated differences<br />

in amino acid composition or expression systemspecific<br />

protein(s) potentially interacting with AE4 requires<br />

further study. Based on its amino acid sequence, AE4 is most<br />

homologous to electrogenic sodium bicarbonate co<strong>transporters</strong><br />

rather than AE1–3.<br />

<strong>SLC4</strong>A10 (NCBE). NCBE (1,088 aa) was originally cloned<br />

from a mouse insulinoma cell line and when expressed in X.<br />

laevis oocytes and HEK293 cells was reported to mediate<br />

Na � -driven Cl � /HCO 3 � exchange (213). Two splice variants<br />

have been subsequently cloned from rat brain and functionally<br />

expressed (57). rb1NCBE was similar to mouse NCBE except<br />

for a 30-aa insert in the NH2 terminus and was highly expressed<br />

in spinal cord and brain beginning in the embryo.<br />

rb2NCBE has an additional 18 aa in its COOH terminus, which<br />

ends with a PDZ motif. rb2NCBE is more highly expressed in<br />

astrocytes than rb1NCBE. Both transcripts mediated Na � -<br />

driven Cl � /HCO 3 � exchange when expressed in mouse fibroblast<br />

3T3 cells, but rb2NCBE was more active in pH regulation<br />

(57). A human ortholog of rb1NCBE, NCBE-B (1,118 aa), and<br />

its splice variant missing 30 aa in the NH2 terminus, NCBE-A,<br />

have been reported (39).<br />

NCBE transcripts have been detected in the peripheral nervous<br />

system and in nonneuronal tissues such as the choroid<br />

plexus, the dura, and some epithelia including the acid-secreting<br />

epithelium of the stomach and the duodenal epithelium (76)<br />

and kidney (153). rb1NCBE protein was highly expressed on<br />

the basolateral membrane of mouse and rat choroid plexus cells<br />

(152, 153). rb2NCBE protein has also been localized in preliminary<br />

studies to the basolateral membrane of mTAL intercalated<br />

cells in the inner medullary collecting duct and duodenal<br />

villus cells (153). In a preliminary study, Choi and coauthors<br />

(39) demonstrated that the human NCBE-B splice variant<br />

functions in X. laevis oocytes as an electroneutral Na � -HCO 3 �<br />

cotransporter. Whether experimental differences in the extent<br />

of Cl � depletion and/or expression systems account for these<br />

findings is currently unknown.<br />

<strong>SLC4</strong> Transporters That Do Not Transport Bicarbonate<br />

<strong>SLC4</strong>A11 (NaBC1). <strong>SLC4</strong>A11 (initially called BTR1) (145)<br />

was originally included in the <strong>SLC4</strong> family because of distant<br />

sequence homology to other family members. BTR1 has been<br />

renamed NaBC1 because it has been shown to function as a<br />

Na�-dependent borate transporter (143, 145). In the absence of<br />

borate, NaBC1 conducts Na� and OH� (H� ), whereas in the<br />

presence of borate, NaBC1 functions as a voltage-regulated,<br />

electrogenic Na� �<br />

-B(OH) 4 cotransporter with a stoichiometry<br />

of Na� �<br />

:B(OH) 4 �2. NaBC1 is the mammalian ortholog of the<br />

Arabidopsis thaliana borate transporter AtBor1 (194).<br />

Downloaded from<br />

http://ajprenal.physiology.org/<br />

by guest on December 23, 2012


CHANNEL-LIKE PROPERTIES OF <strong>SLC4</strong> TRANSPORTERS<br />

Several electroneutral <strong>SLC4</strong> <strong>transporters</strong> also have channellike<br />

properties. AE1 in native X. laevis oocytes has an associated<br />

small conductive anion current (51). The Na � -driven<br />

Cl � /HCO 3 � exchanger cloned from Drosophila NDAE1 also<br />

mediates a small Cl � current and a HCO 3 � current (167). Choi<br />

and colleagues (37, 42) have shown that two rat ortholog splice<br />

variants of the human NBC3 (NBCn1-B and NBCn1-E) in both<br />

X. laevis oocytes and HEK293 cells have an associated ion<br />

conductance that, as they suggested, is likely mediated by the<br />

cotransporter per se rather than an associated protein, although<br />

the latter possibility has not been definitely ruled out. The<br />

associated current was stimulated by DIDS, is not bicarbonate<br />

dependent, and Na � appeared to be the predominant ion<br />

responsible for the current. The physiological importance of<br />

these data, as well as potential implications to the transport<br />

mechanism, is unclear.<br />

TRANSPORT STOICHIOMETRY<br />

The ion transport stoichiometry of <strong>SLC4</strong> <strong>transporters</strong> determines<br />

whether they are electroneutral or electrogenic (Table 1)<br />

(67, 112). The flux through electrogenic <strong>SLC4</strong> <strong>transporters</strong> is<br />

affected by both the membrane potential and the activity of the<br />

transported ion species. AE1–3 proteins exchange equivalent<br />

amounts of Cl� for <strong>base</strong> and are electroneutral (6, 33). Sodium<br />

bicarbonate co<strong>transporters</strong> mediate the transport of Na� and<br />

<strong>base</strong> with different stoichiometries per transport cycle. In<br />

NBC3 (NBCn1), the transport stoichiometry is 1:1, and, like<br />

AE1–3, this transport process is electroneutral (37, 155).<br />

NBC1 mediates the cotransport of one positive charge and<br />

either two or three negative charges per transport cycle (2, 30,<br />

62–64, 67, 68, 112, 166). NBC4 is also electrogenic and has<br />

been reported to have a 3:1 transport stoichiometry when<br />

expressed in mammalian cells (176) and a 2:1 stoichiometry in<br />

X. laevis oocytes (211). Human NDCBE exchanges Cl� for the<br />

equivalent of one Na� �<br />

and two <strong>H<strong>CO3</strong></strong> ions, and the process is<br />

electroneutral (60) (see Fig. 1 for other potential modes of<br />

transport). Although all studies agree that AE4 and NCBE are<br />

electroneutral <strong>transporters</strong>, their Cl� and Na� dependence are<br />

controversial (39, 57, 105, 144, 201, 213), and additional<br />

studies are required to determine their mode of transport.<br />

NBC1 and NBC4 <strong>transporters</strong> differ from other <strong>SLC4</strong> family<br />

members in that they are electrogenic. The net charge carried<br />

by NBC1 and NBC4 varies with their transport stoichiometry<br />

and is an important determinant of the direction of Na� and<br />

� 2<br />

<strong>H<strong>CO3</strong></strong> (or <strong>CO3</strong><br />

�<br />

) flux (Fig. 3) (67, 112). In addition, because<br />

the cell membrane potential induces changes in the flux of<br />

electrogenic sodium bicarbonate co<strong>transporters</strong>, the effective<br />

cell buffer capacity will also vary with the membrane potential<br />

(16). This provides an interesting mechanism for coupling<br />

unrelated transport processes such as the electroneutral monocarboxylate<br />

transporter via MCT1 (which transports H� and<br />

moncarboxylates) to changes in membrane potential in cells<br />

expressing kNBC1 (16).<br />

It had been originally thought that the transport stoichiometry<br />

of <strong>SLC4</strong> proteins is an inherent property of each transporter<br />

that remains fixed. However, Chernova and colleagues<br />

(35) showed that an E699Q mutation in mouse AE1 blocked<br />

electroneutral Cl� /Cl� anion exchange and created an electrogenic<br />

Cl� 2�<br />

/SO4 exchange process. These data indicated that a<br />

� 2�<br />

<strong>SLC4</strong> BASE (<strong>H<strong>CO3</strong></strong> ,<strong>CO3</strong>) TRANSPORTERS<br />

AJP-<strong>Renal</strong> Physiol • VOL 290 • MARCH 2006 • www.ajprenal.org<br />

Fig. 3. HCO 3 � :Na � transport stoichiometry and ion gradients of the cotransported<br />

species determine the reversal potential (Erev) and therefore the direction<br />

of transport mediated by electrogenic <strong>SLC4</strong> <strong>transporters</strong> (67, 112).<br />

Depicted is the effect of changes in either intracellular-to-extracellular Na �<br />

concentration ([Na � ]i/[Na � ]o; A) or intracellular-to-extracellular HCO 3 � concentration<br />

([HCO 3 � ]i/[ <strong>H<strong>CO3</strong></strong> � ]o ; B) ratio on Erev of an electrogenic Na � -<br />

HCO 3 � cotransporter functioning with a <strong>H<strong>CO3</strong></strong> � :Na � cotransport stoichiometry<br />

of either 2:1 or 3:1. The plot was generated using the following equation:<br />

E rev � RT<br />

F�n � 1� ln �Na� i<br />

�HCO 3 � � i n<br />

� n<br />

�Na�o �<strong>H<strong>CO3</strong></strong> � o<br />

Invited Review<br />

F585<br />

where R, T, and F have their usual meaning, and n is the HCO 3 � :Na � transport<br />

stoichiometry. In A, [HCO 3 � ]i � 15 mM and [<strong>H<strong>CO3</strong></strong> � ]o � 25 mM. In B,<br />

[Na � ]i � 20 mM and [Na � ]o � 150 mM. When the membrane potential is<br />

more negative than the Erev, the cotransporter functions in the efflux mode,<br />

thereby contributing to transepithelial HCO 3 � absorption as in the renal<br />

proximal tubule. Although it is usually assumed that a cotransporter functioning<br />

in the 2:1 mode mediates HCO 3 � influx as in the pancreas, in the rabbit<br />

proximal tubule when the cotransporter functions in the 2:1 mode, at a<br />

basolateral membrane potential of about �50 mV, basolateral HCO 3 � efflux<br />

can be driven by a high [Na � ]i of �60 mM (175, 181).<br />

single amino acid substitution could shift an electroneutral<br />

transport process to an electrogenic one. In addition, following<br />

the cloning of the NH2-terminal variants of NBC1 (kNBC1 and<br />

pNBC1), it had been thought that kNBC1 functions in a 3:1<br />

(efflux) mode and pNBC1 functions in 2:1 (influx) mode.<br />

Gross and colleagues (63), however, showed that both <strong>transporters</strong><br />

can function in either a 2:1 or 3:1 mode depending on<br />

the cell type in which they were exogenously expressed. In<br />

addition, it was shown that the HCO 3 � :Na � transport stoichiometry<br />

of NBC1 (kNBC1 and pNBC1) can shift from 3:1 to<br />

2:1 following cAMP-induced PKA-dependent phosphorylation<br />

of kNBC1-Ser 982 /pNBC1-Ser 1026 (62, 64). Moreover, Asp 986<br />

and Asp 988 were also shown to be required for the cAMPinduced<br />

shift of kNBC1 stoichiometry (68). These residues are<br />

located in close proximity to the PKA phosphorylation site,<br />

K 979 KGS, and are part of a putative D 986 NDD motif that, in<br />

addition to a second motif, L 958 DDV, was shown to be involved<br />

in carbonic anhydrase II (CAII) binding (154).<br />

Based on these considerations, Gross and colleagues (67, 68)<br />

originally hypothesized that a potential mechanism for the<br />

cAMP-induced shift in stoichiometry of kNBC1 requires the<br />

binding/dissociation of CAII. Against this hypothesis was the<br />

subsequent finding by Pushkin and colleagues (154) that the<br />

PKA-dependent phosphorylation of Ser 982 does not decrease<br />

CAII binding in vitro. In addition to PKA-dependent phosphorylation,<br />

an increase in intracellular Ca 2� concentration shifts<br />

the stoichiometry of kNBC1 expressed in X. laevis oocytes<br />

from 2:1 to 3:1 (132). The mechanism for the Ca 2� -induced<br />

shift in stoichiometry is currently unknown. Additional studies<br />

have shown that PKA-dependent phosphorylation of pNBC1<br />

Downloaded from<br />

http://ajprenal.physiology.org/<br />

by guest on December 23, 2012


Invited Review<br />

F586 � 2�<br />

<strong>SLC4</strong> BASE (<strong>H<strong>CO3</strong></strong> ,<strong>CO3</strong>) TRANSPORTERS<br />

(KRKT 49 ) does not alter the transport stoichiometry but enhances<br />

flux through the cotransporter, thereby potentially playing<br />

an important role in secretin-induced stimulation of pancreatic<br />

bicarbonate secretion (62).<br />

TRANSPORTED BASE: HCO 3 � VS. <strong>CO3</strong> 2�<br />

It has not been definitively established whether <strong>SLC4</strong> pro-<br />

� 2<br />

teins mediate <strong>H<strong>CO3</strong></strong> and/or <strong>CO3</strong><br />

�<br />

transport. Depicted in Fig. 1<br />

are the hypothetical modes of transport wherein the net charge<br />

� 2<br />

per transport cycle by <strong>H<strong>CO3</strong></strong> and/or <strong>CO3</strong><br />

�<br />

with Na� and/or<br />

Cl� is equivalent. Müller-Berger and colleagues (135) suggested<br />

that the finding that carbonic anhydrase (CA) inhibition<br />

only altered the flux through the basolateral proximal tubule<br />

electrogenic sodium bicarbonate cotransporter (kNBC1) functioning<br />

in the 3:1 mode and not in the 2:1 mode could be used<br />

� 2<br />

to determine whether <strong>H<strong>CO3</strong></strong> and/or <strong>CO3</strong><br />

�<br />

was the transported<br />

species. These authors argued on theoretical grounds that when<br />

CA inhibitors decrease the flux through the cotransporter,<br />

2<br />

<strong>CO3</strong> �<br />

must be one of the transported species and that, in the 3:1<br />

2<br />

mode, 1<strong>CO3</strong> � � � �1<strong>H<strong>CO3</strong></strong> � 1Na are transported. Moreover,<br />

the finding that CA inhibition did not alter the flux in the 2:1<br />

2<br />

mode argued against <strong>CO3</strong> �<br />

� �<br />

flux in favor of 2<strong>H<strong>CO3</strong></strong> � 1Na<br />

being transported. More recently, it has been suggested in<br />

preliminary experiments that even in the 2:1 mode, kNBC1<br />

transports 1Na� 2<br />

and1<strong>CO3</strong> �<br />

(59).<br />

2<br />

Another potential approach to distinguish <strong>CO3</strong> �<br />

�<br />

from <strong>H<strong>CO3</strong></strong> transport is the use of equilibrium thermodynamics (112). In<br />

the presence of a stable PCO2 difference across a membrane<br />

with a functional kNBC1 for example (e.g., lipid bilayer), it<br />

was previously demonstrated that theoretically under these<br />

conditions, the respective reversal potential values for a<br />

� � 2<br />

3<strong>H<strong>CO3</strong></strong> � 1Na transport mode vs. a1<strong>CO3</strong><br />

�<br />

�<br />

� 1<strong>H<strong>CO3</strong></strong> �<br />

1Na� transport mode would differ. Another potential approach<br />

by which to distinguish these modes of transport would be to<br />

establish a stable pK difference across a membrane (112). This<br />

� 2<br />

is possible to accomplish because the <strong>H<strong>CO3</strong></strong> 3 <strong>CO3</strong><br />

�<br />

� H� reaction varies with ionic strength. There are, however, technical<br />

limitations that have thus far not been solved that have<br />

prevented these approaches from being used experimentally.<br />

Fig. 4. Phylogenetic tree of <strong>SLC4</strong> <strong>transporters</strong> created using<br />

the UPGMA analysis. Numbers on top of each line represent<br />

the relative difference between the specific sequences using<br />

GeneWorks software.<br />

STRUCTURE<br />

Primary Structure<br />

AJP-<strong>Renal</strong> Physiol • VOL 290 • MARCH 2006 • www.ajprenal.org<br />

Comparison of the primary structures of <strong>SLC4</strong> <strong>transporters</strong><br />

shows significant sequence homology especially in putative<br />

membrane domains (Fig. 2). On the basis of their amino acid<br />

sequence homology, <strong>SLC4</strong> <strong>transporters</strong> appear to be separated<br />

structurally into three groups (Fig. 4) that likely have an<br />

evolutionary basis: 1) <strong>SLC4</strong>A1, <strong>SLC4</strong>A2, and <strong>SLC4</strong>A3; 2)<br />

<strong>SLC4</strong>A4, <strong>SLC4</strong>A5, and <strong>SLC4</strong>A9; and 3) <strong>SLC4</strong>A7, <strong>SLC4</strong>A8,<br />

and <strong>SLC4</strong>A10. These groups are distinct from the known<br />

functional categorization of <strong>SLC4</strong> <strong>transporters</strong>, indicating that<br />

primary structure may not be a sufficient predictor of the<br />

functional properties of each transporter.<br />

Posttranslational Modifications<br />

Although all <strong>SLC4</strong> <strong>transporters</strong> have various sites for potential<br />

posttranslational modification, thus far AE1 and NBC1<br />

<strong>transporters</strong> have been studied in some detail. Human erythrocyte<br />

AE1 contains a single site of N-glycosylation (Asn642 )in<br />

the fourth exofacial loop (Figs. 2 and 5) that contains either a<br />

short complex oligosaccharide or an extended polylactosaminyl<br />

oligosaccharide (33). Approximately equal amounts of the<br />

different glycosylated forms of AE1 are found in human red<br />

blood cells. Recently, a novel glycosylation site at Asn555 in<br />

the third exofacial loop has been described in addition (36).<br />

kNBC1 has been shown to contain two normally glycosylated<br />

sites at Asn597 and Asn617 (38) located in the second exofacial<br />

loop (Fig. 6) (198). Human AE1 was also palmitoylated at<br />

Cys843 (36). Interestingly, this modification did not affect<br />

trafficking of the anion exchanger.<br />

kNBC1 and pNBC1 are phosphorylated in their common<br />

COOH-terminal PKA phosphorylation site, KKGS (62, 64,<br />

68). The PKA-mediated phosphorylation of Ser982 in kNBC1<br />

(Ser1026 in pNBC1) shifts the transport stoichiometry of both<br />

NBC1 variants from 3:1 to 2:1 (62, 64, 68). pNBC1 possesses<br />

an additional PKA site (KRKT49 ), involved in the regulation of<br />

pNBC1-mediated flux but not the cotransporter stoichiometry<br />

(62).<br />

Downloaded from<br />

http://ajprenal.physiology.org/<br />

by guest on December 23, 2012


Topological Models<br />

Different algorithms predict 9–15 transmembrane segments<br />

in members of the <strong>SLC4</strong> family. Topological models have been<br />

experimentally evaluated in two <strong>SLC4</strong> <strong>transporters</strong>. The most<br />

extensively studied <strong>SLC4</strong> transporter, the anion exchanger<br />

AE1, was proposed to possess 12 (149), 13 (33, 54, 232), or 14<br />

transmembrane segments (70, 94), of which the first 8 are<br />

generally accepted to exist (Fig. 5). The models fit calculations<br />

of AE1 transmembrane area <strong>base</strong>d on a low-resolution twodimensional<br />

crystallography data of dimeric COOH-terminal<br />

membrane domain of AE1 (214). All AE1 models predict that<br />

the NH2 and COOH termini are localized intracellularly. Controversial<br />

experimental data regarding putative COOH-terminal<br />

transmembrane segments require use of high-resolution<br />

� 2�<br />

<strong>SLC4</strong> BASE (<strong>H<strong>CO3</strong></strong> ,<strong>CO3</strong>) TRANSPORTERS<br />

Invited Review<br />

Fig. 5. Human AE1 topology model <strong>base</strong>d on the work of Zhu and colleagues (232). Amino acids that are important for AE1-mediated transport are shown in<br />

red (acidic), blue (basic), and black (others). Green stars depict the residues involved in human disease.<br />

X-ray crystallography to precisely identify the number and<br />

location of AE1 transmembrane segments.<br />

Several studies indicated that the integrity of several cytoplasmic<br />

loops of AE1 is not essential for assembly into an<br />

anion transport-active structure. Coexpression of two pairs of<br />

complementary fragments of AE1 resulted in the generation of<br />

the stilbene disulfonate-sensitive uptake of Cl � into X. laevis<br />

oocytes, similar to expressed intact AE1 (69). For example, the<br />

pairs of fragments, which contained the first 8 and the last 6<br />

membrane spans, and the first 12 and the last 2 membrane<br />

spans, separated within cytoplasmic surface loops of the membrane<br />

domain, were able to generate stilbene disulfonatesensitive<br />

Cl � uptake. In contrast, the pairs separated in the<br />

second cytoplasmic loop were not active (216). Omission of<br />

Fig. 6. Human NBC1 (NBCe1) topological model (198). Amino acids that are important for NBC1-mediated transport are shown in red (acidic), blue (basic),<br />

and black (others). Green stars depict the residues involved in human disease. TM, transmembrane.<br />

AJP-<strong>Renal</strong> Physiol • VOL 290 • MARCH 2006 • www.ajprenal.org<br />

F587<br />

Downloaded from<br />

http://ajprenal.physiology.org/<br />

by guest on December 23, 2012


Invited Review<br />

F588 � 2�<br />

<strong>SLC4</strong> BASE (<strong>H<strong>CO3</strong></strong> ,<strong>CO3</strong>) TRANSPORTERS<br />

any of single transmembrane segments except of spans 6 and 7<br />

caused complete loss of functional expression (70). The unusually<br />

stable association between the AE1 fragments divided<br />

in the second exofacial loop was suggested to indicate an<br />

existence of interactions between polar surfaces on amphiphilic<br />

portions of the third and fifth transmembrane spans. Most AE1<br />

fragments contained the necessary information to fold in the<br />

membrane (71).<br />

Tatishchev and colleagues (198) demonstrated that NBC1<br />

has 10 transmembrane segments, with the NH2 and COOH<br />

termini localized intracellularly (Fig. 6). Whether the difference<br />

in the number of transmembrane segments in AE1 and<br />

NBC1 is related to their specific functional properties remains<br />

to be determined.<br />

Tertiary Structure<br />

The tertiary structure of the members of <strong>SLC4</strong> family has<br />

only been partially determined. A low-resolution (20 Å) structure<br />

of the membrane domain of AE1 has been reported <strong>base</strong>d<br />

on electron microscopy and three-dimensional image reconstruction<br />

of negatively stained two-dimensional crystals (215).<br />

The dimeric membrane domain revealed a U-shaped structure<br />

with dimensions of 60 � 110 and a thickness of 80 Å. The<br />

structure is open on the top and at the sides, with the monomers<br />

in close contact at the <strong>base</strong>, and the basal domain was 40 Å<br />

thick. The upper part of the dimer consisted of two elongated<br />

protrusions, which formed the sides of a canyon, enclosing a<br />

wide space that narrowed down and converged into a depression<br />

at the center of the dimer on the top of the basal domain.<br />

This depression might represent the opening to a transport<br />

channel located at the dimer interface.<br />

The crystal structure of the NH2-terminal cytoplasmic domain<br />

of the human AE1 (aa 1–379) has been reported at a<br />

much higher (2.6 Å) resolution (230). A tight symmetrical<br />

dimer stabilized by interlocked dimerization arms contributed<br />

by both monomers was reported. The dimer fit within a<br />

rectangular prism of approximate dimensions 75 � 55 � 45 Å.<br />

Each subunit also included a larger peripheral protein binding<br />

domain with an ���-fold. The binding sites of ankyrin,<br />

deoxyhemoglobin, aldolase, and glyceraldehyde-3-phosphate<br />

dehydrogenase were localized in the structure.<br />

The COOH-terminal cytoplasmic domain of human NBC3<br />

(aa 1127–1214) expressed in Escherichia coli cells and analyzed<br />

by gel permeation chromatography and sedimentation<br />

velocity centrifugation had a Stokes radius of 26 and 30 Å,<br />

respectively (119). Shape modeling suggested that the COOH<br />

terminus had a rodlike shape with the dimensions of 16 � 190<br />

Å, which was supposed to promote binding of regulatory<br />

proteins (119).<br />

Oligomeric Structure<br />

It is not known whether <strong>SLC4</strong> <strong>transporters</strong> are functionally<br />

active in monomeric form or whether oligomerization is required<br />

for their activity. Numerous biochemical studies of<br />

mammalian AE1 have identified monomers, dimers, trimers,<br />

tetramers, and higher order oligomers and their mixtures (32,<br />

41, 50, 136). The relative abundance of these forms was<br />

dependent on the detergent used for isolation of AE1 and the<br />

temperature of isolation. In two-dimensional crystals of AE1<br />

(50), the protein subunits, each consisting of two AE1 mono-<br />

mers, were arranged with threefold symmetry, suggesting that<br />

a functionally active form of AE1 might be a dimer or/and even<br />

a hexamer. Based on these structural data and because a<br />

mixture of dimers and tetramers has been shown to represent<br />

the majority of AE1 in red blood cells (19, 32, 41, 50, 136, 141,<br />

215), it is widely accepted that a dimer is an active oligomeric<br />

form of AE1. On the basis of coimmunoprecipitation and<br />

functional expression in X. laevis oocytes of noncontiguous<br />

and overlapping pairs of membrane domains of human AE1,<br />

Groves and Tanner (70) hypothesized that transmembrane<br />

spans 1–5 and 9–12 are involved in, and spans 6–8, 13, and 14<br />

are not important for, dimerization of AE1.<br />

Dimers and to a lesser extent tetramers were also shown to<br />

be major components of bAE3 purified from rabbit renal<br />

microsomal membranes in the presence of Triton X-100 (160).<br />

Tetramers of bAE3 were further characterized by transmission<br />

electron microscopy of negatively stained ion-exchanger molecules<br />

(160). Image analysis of the tetramer micrographs<br />

revealed an anion exchanger molecule in which monomers<br />

were arranged with fourfold symmetry around a cavity in the<br />

center of the molecule, which could play a role of an iontransporting<br />

channel. Oligomers of human NBC1 and NBC3<br />

were also detected when these proteins were expressed in<br />

HEK293 cells (159).<br />

TRANSPORT MECHANISM<br />

AJP-<strong>Renal</strong> Physiol • VOL 290 • MARCH 2006 • www.ajprenal.org<br />

It is unusual that the same mutigene family contains transport<br />

proteins that function as exchangers (e.g., AE1, AE2,<br />

AE3, NDCBE) and co<strong>transporters</strong> (e.g., NBC1, NBC4) (Fig.<br />

1). AE1 exchanges equal amounts of Cl � for HCO 3 � by a<br />

ping-pong mechanism, wherein the transporter can be in either<br />

the inward-facing or the outward-facing state (52, 55, 72, 88).<br />

According to Jennings et al. (88), the AE1 catalytic cycle<br />

consists of one pair of exchanging anions per anion exchanger<br />

monomer. The transport rate is governed by a single AE1<br />

conformational change, which leads to the transfer of a single<br />

anion across the membrane (146). The dissociation/association<br />

of ions with AE1 occurs more quickly than the AE1 conformational<br />

changes leading to the translocation of the bound<br />

anion. Human AE1 has an intrinsic functional asymmetry;<br />

therefore, at near neutral pH and equal extra- and intracellular<br />

concentrations of Cl � , most of the anion exchanger molecules<br />

are in the inward-facing state (104).<br />

The stilbene inhibitor DIDS has been shown to bind preferentially<br />

to the outward-facing state of AE1 (88) and is inhibited<br />

by DIDS from outside (31, 96). In contrast to AE1, kNBC1<br />

was inhibited equally by intra- or extracellularly applied DIDS<br />

(74). In addition, H2DIDS has been shown to covalently bind<br />

Lys 539 and Lys 851 in human AE1 at physiological pH (140),<br />

whereas DIDS binds to Lys 851 only at pH �12 (93). According<br />

to current AE1 topological models, both lysine residues are<br />

located in transmembrane segments (54, 70, 111, 232). kNBC1<br />

(30, 166) and the pancreatic form of NBC1, pNBC1 (2), as<br />

well as all members the <strong>SLC4</strong> family with the exception of the<br />

electroneutral sodium bicarbonate cotransporter NBC3 (155)<br />

and rabbit AE4a (202) are stilbene inhibitable, supporting an<br />

idea of a possible shared mechanism(s) of ion transport. However,<br />

against this assumption is the finding that stilbene inhibition<br />

is not limited to the members of <strong>SLC4</strong> family, suggesting<br />

that the mechanism of this inhibition is nonspecific and<br />

Downloaded from<br />

http://ajprenal.physiology.org/<br />

by guest on December 23, 2012


involves binding of stilbene to exposed lysine residues, leading<br />

to steric blocking of normal ion transport mechanism. For<br />

example, various members of the SLC26 anion exchanger<br />

family, which have shared no sequence homology with <strong>SLC4</strong><br />

<strong>transporters</strong>, are also stilbene sensitive (131).<br />

Kinetic studies of the electrogenic sodium bicarbonate cotransporter<br />

NBC1 suggest that contrary to AE1, binding of<br />

both sodium and bicarbonate to the cotransporter is required<br />

before a transport event (65, 90). Gross and Hopfer (66)<br />

developed a kinetic model in which bicarbonate and sodium<br />

bind to NBC1 in an ordered rather than a random manner.<br />

Using additional simplifying assumptions, they fitted the<br />

model to experimental current-voltage relationships obtained at<br />

different concentrations of Na � and HCO 3 � and predicted that<br />

binding of the ions to the cotransporter is voltage dependent,<br />

with an electrical coefficient of 0.2 at pH 7.5. The later<br />

suggested that HCO 3 � “senses” �20% of the membrane’s<br />

electric field on binding to the cotransporter or, in other words,<br />

that the binding site for HCO 3 � is located about one-fifth of the<br />

electrical distance into the membrane. Two basic amino acid<br />

residues, Lys 854 and His 907 , in transmembrane segments of<br />

kNBC1 (Fig. 6) (198) potentially involved in binding of<br />

bicarbonate (1) are located �20 and �10% of membrane width<br />

from the cytoplasmic side. Therefore, binding of bicarbonate to<br />

kNBC1 in the inward-facing conformation may fit the prediction<br />

of Gross and Hopfer (66).<br />

ION SELECTIVITY AND TRANSLOCATION<br />

In proteins that function as exchangers and co<strong>transporters</strong>,<br />

there are residues usually located outside transmembrane segments<br />

mediating ion selectivity, and residues usually located<br />

within transmembrane segments that are involved in ion binding<br />

and translocation (ion binding/translocation site) (44, 172,<br />

220). In the absence of the NH2-terminal domain, the COOHterminal<br />

domain of AE1 (61) and AE3 (108) was able to<br />

mediate Cl � /HCO 3 � exchange, suggesting that the NH2-terminal<br />

domain is not necessary for ion exchange. Several amino<br />

acid residues have been identified that might be involved in ion<br />

selectivity and binding/translocation of AE1. Specifically,<br />

treatment of human red cell AE1 with Woodward’s reagent K<br />

labeled only glutamate but not aspartate residues (86) and<br />

inhibited Cl � /Br � exchange, suggesting that glutamate residue(s)<br />

is involved in AE1-mediated ion transport. Pretreatment<br />

with 4,4�-dinitrostilbene-2,2�-disulfonate (DNDS) protected<br />

AE1 from binding and inactivation by Woodward’s reagent K,<br />

suggesting that the stilbene-binding site and the glutamate<br />

residue(s) are located close to each other. In addition to<br />

glutamate, lysine, arginine, and histidine residues have been<br />

shown to be essential for AE1 transport activity (73, 84, 126,<br />

140, 146, 219, 228).<br />

The importance of glutamate residue(s) in AE1-mediated<br />

transport was further confirmed when Glu 681 in transmembrane<br />

segment 8 of human red cell AE1 (Glu 699 in mouse AE1) was<br />

shown to be involved in the ion exchange function of human<br />

(86, 87) and mouse AE1 (35). Mutation of Glu 681 in human<br />

AE1 and a homologous residue in human AE2 (182) to neutral<br />

or basic amino acids blocked transport of monovalent anions<br />

(Cl � and bicarbonate) but did not change maximum sulfate/<br />

sulfate exchange. Interestingly, replacement of mouse AE1<br />

Glu 699 with aspartate blocked transport of both monovalent<br />

� 2�<br />

<strong>SLC4</strong> BASE (<strong>H<strong>CO3</strong></strong> ,<strong>CO3</strong>) TRANSPORTERS<br />

AJP-<strong>Renal</strong> Physiol • VOL 290 • MARCH 2006 • www.ajprenal.org<br />

Invited Review<br />

F589<br />

and divalent anions (133). Replacement of mouse AE1 Glu 699<br />

with glutamine inhibited Cl � /HCO 3 � exchange, but the mutant<br />

anion exchanger mediated an electrogenic Cl � /SO 4 2� exchange<br />

when expressed in X. laevis oocytes (35). It was suggested (87)<br />

that Glu 681 is the binding site for H � , which is transported with<br />

SO 4 2� during AE1-mediated H � -SO4 2� cotransport, and can be<br />

alternately exposed to the intracellular and extracellular media.<br />

In addition, interaction of mouse AE1 Glu 699 with His 752 was<br />

shown to be important for pH sensitivity of Cl � transport at<br />

low pH (133). Recently, it was shown that an additional<br />

glutamate residue is involved in Cl � binding and is part of a<br />

second Cl � binding/transport site in the AE1 monomer (85).<br />

In human kNBC1 (Fig. 6), Asp 754 (glutamate in AE1–3)<br />

flanks transmembrane segment 6 and plays a key role in<br />

kNBC1-mediated sodium bicarbonate cotransport, suggesting<br />

that the decreased length of aspartic acid compared with<br />

glutamic acid may be related to sodium dependence of ion<br />

transport (1). In addition, Asp 555 , Glu 559 , Asp 647 , Asp 685 , and<br />

Asp 778 are important for kNBC1 function (1). Asp 685 , conserved<br />

in Na � -dependent <strong>SLC4</strong> <strong>transporters</strong> and AE4, is potentially<br />

involved in sodium binding/translocation (1). Asp 555 ,<br />

conserved in all <strong>SLC4</strong> <strong>transporters</strong>, is located in close proximity<br />

to Lys 559 , which has been suggested (1) to participate in<br />

bicarbonate binding and translocation.<br />

The involvement of mouse AE1 Lys 558 (Lys 539 in human<br />

AE1, Fig. 5) in AE1-mediated Cl � transport was hypothesized<br />

<strong>base</strong>d on site-directed mutagenesis data (134). This amino acid<br />

residue located in a transmembrane span is also a target for<br />

covalent binding of H2DIDS and DIDS (93, 140). Both<br />

H2DIDS (DIDS at pH �12) and pyridoxal phosphate bind<br />

another lysine residue, Lys 851 (100, 140) and inhibit AE1.<br />

Several lysine residues (Lys 559 , Lys 667 , Lys 668 , Lys 681 , Lys 770 ,<br />

Lys 771 , Lys 854 , and Lys 924 ) are functionally important for<br />

kNBC1 (1). Lys 854 is localized in transmembrane segment 8,<br />

which is conserved in Na � -dependent <strong>SLC4</strong> <strong>transporters</strong> and<br />

potentially involved in ion binding/translocation mediated by<br />

NBC1. Human kNBC1 Lys 559 and Lys 924 are potential targets<br />

for DIDS inhibition. These residues are located extracellularly<br />

(Fig. 6); however, it is known that stilbenes inhibit kNBC1<br />

from both extra- and intracellular locations (74). Based on<br />

these data, Abuladze and coauthors (1) hypothesized that this<br />

part of the extracellular loop 2 in NBC1 is able to migrate<br />

between extra- and intracellular compartments. In addition,<br />

these authors suggested that this region is involved in bicarbonate<br />

binding in NBC1. The second half of this loop contains<br />

naturally glycosylated Asn 597 and Asn 617 (38) and therefore is<br />

located extracellularly. Lys 667 , Lys 668 , Lys 681 , Lys 770 , Lys 771 ,<br />

and Lys 854 are probably involved in ion selectivity in NBC1.<br />

Four histidine residues located in transmembrane segments<br />

of mouse AE1, His 721 , His 752 , His 837 , and His 852 (His 703 ,<br />

His 734 , His 819 , and His 834 in human AE1, Fig. 5) were shown<br />

by site-directed mutagenesis to participate in AE1-mediated<br />

ion exchange (133, 134). Mouse AE1 His 852 was shown to play<br />

a key role in the control of pH dependence of Cl � transport<br />

(133). It was suggested that the formation of a hydrogen bond<br />

between His 852 and Glu 699 (Glu 681 in human AE1) is essential<br />

for the decrease in Cl � transport at low pH. In addition, His 834<br />

in human AE1 was shown to play an essential role in the<br />

protein conformational changes during the ion exchange and is<br />

located in close proximity to the anion translocation site (92).<br />

It was also hypothesized that all these histidine residues inter-<br />

Downloaded from<br />

http://ajprenal.physiology.org/<br />

by guest on December 23, 2012


Invited Review<br />

F590 � 2�<br />

<strong>SLC4</strong> BASE (<strong>H<strong>CO3</strong></strong> ,<strong>CO3</strong>) TRANSPORTERS<br />

act with mouse AE1 Lys 558 (Lys 539 human AE1), the target of<br />

covalent binding of DIDS (93, 140). All these histidine residues<br />

except His 819 are conserved in <strong>SLC4</strong> <strong>transporters</strong> (Fig. 2).<br />

The only NBC1 homolog of these AE1 histidine residues<br />

located in transmembrane segments, kNBC1-His 907 (human<br />

AE1-His 834 ), was shown to be important for NBC1-mediated<br />

ion transport (1), suggesting that it can participate in bicarbonate<br />

binding/translocation. Human kNBC1-His 776 (His 734 in<br />

human AE1) located in the third intracellular loop is also<br />

important for kNBC1 function and apparently involved in<br />

bicarbonate selection (1), whereas the role of His 807 (His 734 in<br />

human AE1) has not yet been evaluated.<br />

Mutating Arg 509 and Arg 748 in mouse AE1 to lysine, threonine,<br />

and cysteine or lysine and glutamine, respectively,<br />

abolished Cl � /Cl � exchange activity in X. laevis oocytes (97).<br />

The Arg 734 mutant was supposed to be functionally inactive,<br />

whereas the Arg 509 mutant was possibly active but abnormally<br />

folded and therefore subsequently degraded. Because Arg 734 is<br />

conserved in AE1–3 and missing in Na � -dependent members<br />

of the <strong>SLC4</strong> family, including AE4, this residue may be<br />

responsible for Cl � binding/translocation in anion exchangers.<br />

In addition, the kNBC1 Arg 538 , Arg 680 , Arg 722 , Arg 881 , and<br />

Arg 943 conserved in the <strong>SLC4</strong> family are also important for<br />

cotransporter function (1, 75).<br />

Several uncharged amino acids were detected in kNBC1<br />

transmembrane segments (Ser 427 , Tyr 433 , Phe 656 , Ser 684 , and<br />

Ala 799 ) and in loops (Thr 485 , Ala 556 , Thr 671 , Thr 677 , and<br />

Thr 815 ) that are important for cotransporter function (1, 49, 75).<br />

Numerous uncharged amino acids in the membrane domain of<br />

AE1 have also been shown to be important for anion exchange.<br />

Tang and coauthors (195) mutated to cysteine the amino acids<br />

within and in close proximity to transmembrane segment 8.<br />

The cysteine mutants of this segment, Ala 666 , Ser 667 , Leu 669 ,<br />

Leu 673 , Leu 677 , and Leu 680 , and in addition the intracellular<br />

mutants of Ile 684 and Ile 688 were inhibitable by sulfhydryl<br />

reagents. It was suggested that these amino acids could be a<br />

part of a AE1 translocation pore (195). Analysis of single<br />

cysteine mutants in the COOH-terminal area of human AE1<br />

(Phe 806 -Cys 885 ) in the last two putative transmembrane segments<br />

identified certain key residues in the Val 849 -Leu 863<br />

region (231). The putative ion selectivity filter was located at<br />

Ser 852 -Leu 857 (231).<br />

INTERACTION WITH CA: A TRANSPORT METABOLON<br />

CAs are a multigene family of enzymes that catalyze the<br />

reversible synthesis of bicarbonate from CO2 and H2O (24,<br />

180, 183, 200). Inhibition of CA activity affects <strong>SLC4</strong> anion<br />

exchange and sodium bicarbonate cotransport function (28, 29,<br />

128, 174, 181, 186). Kifor and coauthors (102) on the basis of<br />

indirect evidence suggested first that red cell AE1 might<br />

interact with a cytosolic CA. Reithmeier and colleagues (163,<br />

206–208) demonstrated that the COOH terminus of AE1 binds<br />

CAII. Functional studies have demonstrated that CAII stimulates<br />

the transport function of AE1, AE2, and AE3 (45, 186).<br />

As a model for the CAII-mediated enhancement of HCO 3 �<br />

transport through AE1, it was hypothesized that a complex of<br />

AE1 and CAII functions in red blood cells as a transport<br />

metabolon. In this model, the efficiency of bicarbonate transport<br />

via AE1 is enhanced by the intermolecular transfer of<br />

AJP-<strong>Renal</strong> Physiol • VOL 290 • MARCH 2006 • www.ajprenal.org<br />

bicarbonate between CAII and AE1, thereby eliminating the<br />

slower cytoplasmic diffusion of HCO 3 � between the two proteins<br />

(163, 186, 187, 206–208). In addition to CAII, CAIV was<br />

shown to bind AE1–3 (185).<br />

It is known that inhibition of CA activity in the renal<br />

proximal tubule of the kidney significantly decreases the rate of<br />

transepithelial bicarbonate absorption and basolateral sodium<br />

bicarbonate efflux (28, 29, 128, 174, 181). Complete inhibition<br />

of CA by 0.1 mM acetazolamide (ACTZ) decreased the flux<br />

through exogenously expressed kNBC1 by �65% when the<br />

transporter functioned with a HCO 3 � :Na � stoichiometry of 3:1<br />

(68) as in the in vivo proximal tubule (28) without altering its<br />

stoichiometry. ACTZ does not inhibit kNBC1-mediated flux<br />

when the cotransporter functions in the 2:1 stoichiometry mode<br />

(68, 181).<br />

The COOH terminus of kNBC1 (kNBC1-ct) was shown to<br />

bind CAII in vitro with high (Kd � 0.2 �M) affinity (68). Two<br />

kNBC1 acidic motifs, L 958 DDV and D 986 NDD, are involved in<br />

this binding (154). The complex of kNBC1 with CAII functioned<br />

as a transport metabolon (154). In pancreatic ducts,<br />

where pNBC1 mediates the basolateral influx of bicarbonate,<br />

CAII is highly expressed (137, 138). pNBC1 has a COOH<br />

terminus identical to kNBC1, and transfection of HEK cells<br />

with nonfunctional CAII has a dominant-negative effect on<br />

pNBC1 function (12), suggesting that pNBC1 also forms a<br />

functional complex with CAII. pNBC1 also binds CAIV (12).<br />

The COOH terminus of electroneutral sodium bicarbonate<br />

cotransporter NBC3, like NBC1 and AE1-AE3, binds CAII<br />

with high affinity (Kd �100 nM) (119). The human NBC3-<br />

D 1135 -D 1136 residues were essential for the interaction between<br />

the two proteins. In �-intercalated cells in the outer medullary<br />

collecting duct, the interaction between apical NBC3 and CAII<br />

may play an important role in pHi regulation.<br />

In addition to CA, certain <strong>SLC4</strong> <strong>transporters</strong> interact with<br />

other cellular proteins. The cytoplasmic domain of AE1 contains<br />

sites for binding ankyrin, 4.1 and 4.2 proteins, glyceraldehyde-3-phosphate<br />

dehydrogenase, phosphofructokinase, deoxyhemoglobin,<br />

p72syk protein tyrosine kinase, and<br />

hemichromes (230) and functions as an anchoring site for these<br />

membrane-associated proteins. These interactions are important<br />

for regulation of cell flexibility and shape (121), glucose<br />

metabolism (120), ion transport (124), and cell life span (95).<br />

The AE1 NH2-terminal cytoplasmic domain is stabilized by<br />

interlocked dimerization arms contributed by both monomers<br />

(230).<br />

GENETIC DISEASES AND KNOCKOUT MODELS<br />

The crucial importance of <strong>SLC4</strong> <strong>transporters</strong> in regulating<br />

intracellular pH and transporting bicarbonate in various cell<br />

types is best exemplified by the phenotype resulting from an<br />

abnormality in the functional properties of several members of<br />

the family (Table 2).<br />

<strong>SLC4</strong>A1<br />

Patients with AE1 mutations tend to have either hereditary<br />

spherocytosis (HS) or distal renal tubular acidosis (dRTA) but<br />

not both abnormalities. The reason for the lack of overlap<br />

between these syndromes is not understood. The majority of<br />

patients with HS-associated AE1 mutations have an autosomal<br />

dominant inheritance with missense mutations that are distrib-<br />

Downloaded from<br />

http://ajprenal.physiology.org/<br />

by guest on December 23, 2012


Table 2. Genetic diseases and mouse knockout models<br />

Gene Species Phenotype Inheritance Reference(s)<br />

<strong>SLC4</strong>A1 Human Anemia, spherocytosis, distal RTA, hypokalemia,<br />

hypercalciuria, nephrocalcinosis, nephrolithiasis<br />

uted throughout the transporter. The underlying abnormality<br />

thus far appears to be abnormal targeting to the plasma membrane<br />

due to misfolding of mutant <strong>transporters</strong>. In addition, a<br />

dominant-negative effect results in a decrease in wild-type (wt)<br />

AE1 membrane expression. Patients with autosomal dominant<br />

dRTA due to mutant AE1 have missense mutations in AE1-<br />

R589 (R589H, R589C, and R589S), S613F, A889X, G609R,<br />

and a truncation of 11 COOH-terminal residues, R901X (26,<br />

34, 83, 99, 171). The syndrome is characterized by abnormal<br />

urinary acidification, metabolic acidosis, nephrocalcinosis, hypercalciuria,<br />

and hypokalemia. Within a given family, the<br />

phenotypic expression can vary, likely because of background<br />

genes that are thus far unidentified (218).<br />

Studies of either red blood cells or heterologous expression<br />

systems demonstrated only modest changes in function that<br />

could not account for the urinary acidification abnormality in<br />

patients (26, 83). wt-AE1 in vivo and in polarized Madin-<br />

Darby canine kidney (MDCK) cells is targeted to the basolateral<br />

membrane. The predominant abnormality in the mutants<br />

characterized thus far in polarized MDCK cells (G609R and<br />

R901X) is the abnormal targeting of kAE1 to both the apical<br />

and basolateral membranes (47, 171). It is speculated that<br />

luminal bicarbonate secretion via mislocalized apical AE1<br />

decreases urinary acidification in these patients and that oligomerization<br />

of mutant AE1 with the wt-AE1 results in a<br />

dominant-negative effect by mistargeting wt-AE1 to the apical<br />

membrane. The magnitude and direction of apical AE1-mediated<br />

transport would be dependent on the respective cell/lumen<br />

HCO 3 � and Cl � gradients. Whether patients with abnormal<br />

luminal bicarbonate secretion can be distinguished by a higher<br />

than urinary PCO2 compared with patients with dRTA due to<br />

abnormal H � -ATPase function remains to be determined.<br />

Autosomal recessive dRTA is prevalent in Southeast Asia.<br />

Unlike patients with autosomal dominant dRTA, autosomal<br />

recessive dRTA presents earlier and tends to be of greater<br />

severity (14, 17, 98). These patients have been reported to have<br />

missense mutations in AE1 G701D, or various compound<br />

heterozygous combinations, including patients with G701D<br />

and the second allele having an in-frame AE1-400–408 deletion<br />

[called South Asian ovalocytosis (SAO)], (G701D/SAO),<br />

A858D/SAO, �V850/SAO, A858D/�V850, and G701D/<br />

S773P (27, 197, 204, 226). The G701D mutant functions and<br />

is targeted normally in red blood cells but not in �-intercalated<br />

cells, which lack the targeting of glycophorin A (GPA) that is<br />

expressed in the red cell membrane (197). AE1 SAO is unable<br />

� 2�<br />

<strong>SLC4</strong> BASE (<strong>H<strong>CO3</strong></strong> ,<strong>CO3</strong>) TRANSPORTERS<br />

AD, AR 14,17,26,27,34,47,83,98,<br />

99,103,164,171,179,<br />

197,204,218,221,226<br />

Bovine Anemia, spherocytosis, distal RTA AD 80<br />

Mouse knockout Anemia, spherocytosis, runting AR 147,184<br />

<strong>SLC4</strong>A2 Mouse knockout Growth retardation, achlorhydria, edentulous, abnormal<br />

spermatogenesis, embryonically lethal<br />

AR 56,130<br />

<strong>SLC4</strong>A3 Human (polymorphism) Generalized seizures 173<br />

<strong>SLC4</strong>A4 Human Short stature, basal ganglia calcification, mental retardation,<br />

cataracts, band keratopathy, proximal RTA, hypokalemia<br />

AR 49,75,78,79,81,116,177,202<br />

<strong>SLC4</strong>A5 Human (polymorphism) Hypertension 15<br />

<strong>SLC4</strong>A7 Mouse knockout Retinal degeneration, mild auditory impairment AR 20<br />

RTA, renal tubular acidosis; AD, autosomal dominant; AR, autosomal recessive.<br />

AJP-<strong>Renal</strong> Physiol • VOL 290 • MARCH 2006 • www.ajprenal.org<br />

Invited Review<br />

F591<br />

to transport anions (179, 221). The A858D mutant functions<br />

abnormally, and its function is only partially rescued by GPA.<br />

In addition, the functioning of A858D is significantly decreased<br />

when coexpressed with �V850 or SAO (27). Similarly,<br />

the abnormal function of the �V850 mutant is not rescued<br />

completely by GPA (27). Recent detailed studies of the S773P<br />

mutant show that it is expressed at a low level, has a shorter<br />

half-life, is misfolded, and is targeted to the proteosome for<br />

degradation (103).<br />

Kittanakom and colleagues (103) have a proposed a model<br />

wherein AE1 mutants that cause autosomal dominant dRTA<br />

form heteroligomers with wt-AE1 that are retained in the<br />

endoplasmic reticulum, resulting in a dominant-negative effect.<br />

In contrast, AE1 mutants that cause autosomal recessive dRTA<br />

form homodimers which are targeted to the proteosome for<br />

degradation. Unlike mutants causing autosomal dominant<br />

dRTA, the S773P/wt-AE1 and G701D/wt-AE1 oligomers are<br />

targeted normally to the plasma membrane in HEK293 cells.<br />

Whether, as demonstrated with autosomal G609R and R901X<br />

mutations, homoligomers S773P and G701D homoligomers<br />

are targeted abnormally to the apical membrane in polarized<br />

cells remains to be determined. Interestingly, although patients<br />

with autosomal recessive dRTA typically have no red cell<br />

phenotype, patients with homozygous AE1-V488M have neonatal<br />

severe anemia and dRTA (164). Similarly, cattle homozygous<br />

for the AE1 R646X mutation have spherocytosis and<br />

dRTA (80). AE1 knockout mice have runting and severe<br />

anemia (147, 184).<br />

<strong>SLC4</strong>A2<br />

Although there are no known diseases in humans attributed<br />

to loss of AE2 function, targeted disruption of <strong>SLC4</strong>A2 in mice<br />

has recently been reported. Using a Cre/loxP strategy to disrupt<br />

expression of three of the four variants of AE2, Medina and<br />

coauthors (130) reported failure of spermiogenesis and infertility<br />

in male AE2�/� mice. Gawenis and colleagues (56)<br />

reported a more severe phenotype in AE2�/� mice related<br />

most likely to a different targeting strategy that resulted in the<br />

complete loss of AE2 expression. In this study, in some mice,<br />

loss of AE2 function was lethal to the embryo. In those<br />

AE2�/� mice that survived, most of them died around the time<br />

of weaning. The variability in survival suggests a role for<br />

background genes in modifying the severity of the phenotype.<br />

AE2�/� mice exhibited achlorhydria, moderate dilation of the<br />

Downloaded from<br />

http://ajprenal.physiology.org/<br />

by guest on December 23, 2012


Invited Review<br />

F592 � 2�<br />

<strong>SLC4</strong> BASE (<strong>H<strong>CO3</strong></strong> ,<strong>CO3</strong>) TRANSPORTERS<br />

gastric gland lumens, and a reduction in the number of parietal<br />

cells. Ultrastructural analysis revealed abnormal parietal cell<br />

structure, with severely impaired development of secretory<br />

canaliculi and few tubulovesicles but normal apical microvilli.<br />

In addition, the mice had severe growth retardation and were<br />

edentulous, ataxic, and almost deaf. These studies demonstrate<br />

the requirement for normal AE2 function in various organ<br />

systems and, specifically, the requirement of AE2 for normal<br />

gastric acid secretion and for normal development of secretory<br />

canalicular and tubulovesicular membranes in mouse parietal<br />

cells. In humans, it is probable that that total loss of AE2<br />

function is embryonically lethal.<br />

<strong>SLC4</strong>A3<br />

No genetic disease has been described for AE3. A significantly<br />

increased 2600 C/A polymorphism variant (A867D<br />

substitution) was detected in patients with idiopathic generalized<br />

epilepsies (173).<br />

<strong>SLC4</strong>A4<br />

Igarashi and colleagues (78) first reported two unrelated<br />

patients with homozygous mutations in the <strong>SLC4</strong>A4 gene. The<br />

patients were mentally retarded, with short stature and severe<br />

proximal renal tubular acidosis (pRTA), systemic academia,<br />

and hypokalemia. Ocular abnormalities included glaucoma,<br />

cataracts, and band keratopathy. The first patient had a missense<br />

R298S mutation (kNBC1 numbering) whereas the second<br />

patient had a R510H substitution. These mutations are<br />

predicted to affect both the kNBC1 and pNBC1 variants of the<br />

NBC1 (<strong>SLC4</strong>A4) gene. The serum amylase was elevated, and<br />

thyroid abnormalities were documented. A head computed<br />

tomographic scan revealed bilateral calcification of the basal<br />

ganglia. In a subsequent patient reported by the same group<br />

with a Q29X nonsense mutation affecting only kNBC1, the<br />

patient had severe pRTA and glaucoma without band keratopathy<br />

or cataracts (79). Both kNBC1 and pNBC1 are expressed in<br />

the eye in a cell type-specific fashion. In the rat eye, Bok and<br />

colleagues (21) detected pNBC1 in the ciliary body, conjunctival<br />

surface cells, cornea, lens epithelium, and retina, whereas<br />

kNBC1 expression was restricted to conjunctival basal cells.<br />

Usui and colleagues (202) reported a more widespread expression<br />

pattern of kNBC1 in various cell types in the human eye.<br />

The finding that the patient with the Q29X mutation only had<br />

glaucoma might suggest that band keratopathy and cataracts<br />

are due to pNBC1 mutations, whereas glaucoma is due to loss<br />

of function of kNBC1. However, other causes of congenital<br />

pRTA and systemic acidemia such as Lowe’s syndrome (43)<br />

also cause glaucoma and cataracts in addition to mental retardation<br />

and growth abnormalities, indicating that one should be<br />

cautious at this point in attributing a particular ocular phenotype<br />

to loss of function of either NBC1 variant in patients with<br />

<strong>SLC4</strong>A4 mutations. Dinour and colleagues (49) reported a<br />

patient with a homozygous S427L <strong>SLC4</strong>A4 missense mutation.<br />

The patient had severe pRTA, glaucoma, corneal opacities,<br />

cataracts, and abnormal dentition. A computed tomographic<br />

scan showed no calcifications, and the patient was of normal<br />

intelligence. Inatomi et al. (81) also recently reported a patient<br />

with a homozygous deletion of A2311 (kNBC1 numbering),<br />

resulting in anomalous transcription of 29 amino acids followed<br />

by a stop codon. The patient had severe pRTA, renal<br />

AJP-<strong>Renal</strong> Physiol • VOL 290 • MARCH 2006 • www.ajprenal.org<br />

insufficiency, elevated lipase and amylase levels, normal intelligence,<br />

glaucoma, corneal opacities, cataracts, tooth enamel<br />

defects, growth retardation, calcification of the basal ganglia<br />

and the left frontoparietal region, mild osteopenia, and normal<br />

thyroid function.<br />

Of the known NBC1 mutations causing pRTA in humans,<br />

several have been studied in heterologous expression system.<br />

The R298S and R510H missense mutations when assayed in<br />

ECV304 cells decreased the functional activity of kNBC1 by<br />

�55% (78). The same mutations in pNBC1 (R342S, R554H)<br />

also decreased the functional activity of pNBC1 by �50%<br />

(177). The S427L mutant that is targeted normally to the<br />

plasma membrane in X. laevis oocytes has significantly abnormal<br />

voltage and Na � -dependent transport (49). Interestingly,<br />

this mutant has no reversal potential accounting for the failure<br />

of kNBC1 to mediate sodium bicarbonate efflux in the kidney<br />

and eye. Horita and coauthors (75) recently identified three<br />

additional homozygous mutations (T485S, A799V, and<br />

R881C) in patients with severe pRTA and ocular abnormalities.<br />

Functional studies of these mutants in addition to the<br />

R298S and R510H mutants suggested that a 50% or greater<br />

reduction in kNBC1-mediated transport is required to cause<br />

severe pRTA (plasma HCO 3 � concentration � 13 mM).<br />

Li and coauthors (116) recently demonstrated that the kNBC1-<br />

R510H and S427L mutations not only have reduced intrinsic<br />

activity but are also mistargeted when expressed in polarized<br />

MDCK cells. The R510H mutant was predominantly retained in<br />

the cytoplasm, whereas the S427L mutant was mistargeted to the<br />

apical membrane. They also identified a kNBC1 COOH-terminal<br />

motif, QQPFLS (aa 1010–1015), is required for targeting of<br />

NBC1 to the basolateral membrane in MDCK cells (117). Within<br />

this motif, a kNBC1-F1013A mutant was abnormally targeted to<br />

the apical membrane. In a large-scale mutagenesis screen, Abuladze<br />

and coauthors (1) have shown that residues in several loops<br />

and transmembrane segments in kNBC1 are required for normal<br />

plasma membrane trageting. Studies of the kNBC1-A2311 deletion<br />

mutant failed to show any plasma membrane expression and<br />

functional activity in X. laevis oocytes (81).<br />

Recently, it has been suggested that the phenotype in patients<br />

with rod and cone degeneration (RP17 form of retinitis pigmentosa)<br />

due to missense mutations in CAIV results from an impaired<br />

wt-NBC1/CAIV transport metabolon in the endothelium of the<br />

choriocapillaris (224). The biologically active CAIV-R14W mutant<br />

did not stimulate NBC1 activity by failing to bind to NBC1,<br />

whereas the CAIV-R219S mutant did not stimulate NBC1 activity<br />

due to its enzymatic inactivity. Additional studies are needed to<br />

determine whether patients with NBC1 mutations have degenerated<br />

rods/cones and whether the ocular phenotype resulting from<br />

mutant CAIV can occur in the absence of functioning NBC1. In<br />

a separate study, Rebello and coauthors (162) have reported a<br />

different underlying mechanism for the RP17 form of retinitis<br />

pigmentosa. In patients with the CAIV-R14W mutation, their<br />

finding suggests that the mutant enzyme in endothelial cells in the<br />

choriocapillaris leads to ER stress due to an accumulation of<br />

unfolded protein, apoptosis, and ischemia of the retina.<br />

<strong>SLC4</strong>A5<br />

Polymorphisms in this gene are associated with hypertension<br />

(15). Further studies are needed to clarify the basis for this<br />

association.<br />

Downloaded from<br />

http://ajprenal.physiology.org/<br />

by guest on December 23, 2012


<strong>SLC4</strong>A7<br />

Recently, the importance of NBC3 in sensory transduction<br />

has been demonstrated in an <strong>SLC4</strong>A7 knockout mouse (20).<br />

<strong>SLC4</strong>A7 knockout mice have a slow degeneration and ultimate<br />

loss of photoreceptors in the eye and a mild hearing defect due<br />

to the loss of inner and outer hair cells in the inner ear. These<br />

findings indicate that <strong>SLC4</strong>A7 knockout mice are a model for<br />

Usher syndrome, a group of diseases characterized by visual<br />

loss and auditory impairment in humans (101).<br />

CONCLUSION<br />

The <strong>SLC4</strong> family contains a functionally diverse group of<br />

transport proteins that play an essential role in the transport of<br />

� 2<br />

<strong>base</strong> (<strong>H<strong>CO3</strong></strong> ,<strong>CO3</strong><br />

�<br />

) in various tissues in mammals. Although<br />

there are have been numerous studies of the structure and<br />

function of the initial member of the <strong>SLC4</strong> family, AE1, more<br />

recently, additional members of the <strong>SLC4</strong> family have been<br />

investigated. Because of the efforts of several laboratories,<br />

various heterogeneous diseases can now be attributed to members<br />

of the <strong>SLC4</strong> family. Nevertheless, although much progress<br />

has been made, a thorough understanding of the structure and<br />

function of both wild-type and mutated <strong>transporters</strong> at a molecular<br />

level is lacking. The current topological models of AE1<br />

are still controversial because of the various methodologies<br />

used by different investigators to address this question. Because<br />

of these complexities, one cannot satisfactorily determine<br />

whether the larger number of transmembrane regions in<br />

AE1 compared with NBC1 represents differences in their ion<br />

binding sites/mechanism of transport. Nevertheless, recent<br />

studies on specific amino acid residues critical for NBC1mediated<br />

transport provide a structural basis for the molecular<br />

comparison of NBC1 and AE1. Results obtained thus far<br />

suggest that the molecular architecture and the transport mechanisms<br />

of NBC1 and AE1 have certain similar features, and at<br />

the same time, there are unique differences that might account<br />

for the separate transport modes (exchanger vs. cotransporter).<br />

In the future, successful crystallization of <strong>SLC4</strong> proteins will<br />

contribute to an understanding of their transport mechanism<br />

and help address these and other currently unresolved questions.<br />

GRANTS<br />

This work was supported by National Institutes of Health (NIH) Grants<br />

DK-63125, DK-58563, DK-07789, and ES-012935-A1, the Max Factor Family<br />

Foundation, the Richard and Hinda Rosenthal Foundation, the Fredrika<br />

Taubitz Fund, National Kidney Foundation of Southern California Grant<br />

J891002, and American Heart Association (Western State Affiliate) Grant<br />

0365022Y.<br />

REFERENCES<br />

1. Abuladze N, Azimov R, Newman D, Liu W, Tatishchev S, Pushkin A,<br />

and Kurtz I. Critical amino acid residues involved in the electrogenic<br />

sodium bicarbonate cotransporter kNBC1-mediated transport. J Physiol<br />

565: 717–730, 2005.<br />

2. Abuladze N, Lee I, Newman D, Hwang J, Boorer K, Pushkin A, and<br />

Kurtz I. Molecular cloning, chromosomal localization, tissue distribution,<br />

and functional expression of the human pancreatic sodium bicarbonate<br />

cotransporter. J Biol Chem 273: 17689–17695, 1998.<br />

3. Abuladze N, Pushkin A, Tatishchev S, Newman D, Sassani P, and<br />

Kurtz I. Expression and localization of rat NBC4c in liver and renal<br />

uroepithelium. Am J Physiol Cell Physiol 287: C781–C789, 2004.<br />

4. Abuladze N, Song M, Pushkin A, Newman D, Lee I, Nicholas S, and<br />

Kurtz I. Structural organization of the human NBC1 gene: kNBC1 is<br />

� 2�<br />

<strong>SLC4</strong> BASE (<strong>H<strong>CO3</strong></strong> ,<strong>CO3</strong>) TRANSPORTERS<br />

AJP-<strong>Renal</strong> Physiol • VOL 290 • MARCH 2006 • www.ajprenal.org<br />

Invited Review<br />

F593<br />

transcribed from an alternative promoter in intron 3. Gene 251: 109–122,<br />

2000.<br />

5. Alper SL. Genetic diseases of acid-<strong>base</strong> <strong>transporters</strong>. Annu Rev Physiol<br />

64: 899–923, 2002.<br />

6. Alper SL, Darman RB, Chernova MN, and Dahlberg CG. The AE<br />

gene family of Cl � /HCO 3 � exchangers. J Nephrol 15: S41–S53, 2002.<br />

7. Alper SL, Natale J, Gluck S, Lodish HF, and Brown D. Subtypes of<br />

intercalated cells in rat kidney collecting duct defined by antibodies<br />

against erythroid band 3 and renal vacuolar H � -ATPase. Proc Natl Acad<br />

Sci USA 86: 5429–5433, 1989.<br />

8. Alper SL, Rossmann H, Wilhelm S, Stuart-Tilley AK, Shmukler BE,<br />

and Seidler U. Expression of AE2 anion exchanger in mouse intestine.<br />

Am J Physiol Gastrointest Liver Physiol 277: G321–G322, 1999.<br />

9. Alper SL, Stuart-Tilley A, Simmons CF, Brown D, and Drenckhahn<br />

S. The fodrin-ankyrin cytoskeleton of choroid plexus preferentially<br />

colocalizes with apical Na � ,K � -ATPase rather than with basolateral<br />

anion exchanger AE2. J Clin Invest 93: 1430–1438, 1994.<br />

10. Alper SL, Stuart-Tilley A, Biemesderfer D, Shmukler B, and Brown<br />

D. Immunolocalization of AE2 anion exchanger in rat kidney. Am J<br />

Physiol <strong>Renal</strong> Physiol 273: F601–F614, 1997.<br />

11. Alrefai WA, Tyagi S, Nazir TM, Barakat J, Anwar SS, Hadjiagapiou<br />

C, Bavishi D, Sahi J, Malik P, Goldstein J, Layden TJ, Ramaswamy<br />

K, and Dudeja PK. Human intestinal anion exchanger isoforms: expression,<br />

distribution, and membrane localization. Biochim Biophys Acta<br />

1511: 17–27 2001.<br />

12. Alvarez BV, Loiselle FB, Supuran CT, Schwartz GJ, and Casey JR.<br />

Direct extracellular interaction between carbonic anhydrase IV and the<br />

human NBC1 sodium/bicarbonate co-transporter. Biochemistry 42:<br />

12321–12329, 2003.<br />

13. Aranda V, Martinez I, Melero S, Lecanda J, Banales JM, Prieto J,<br />

and Medina JF. Shared apical sorting of anion exchanger isoforms<br />

AE2a, AE2b1, and AE2b2 in primary hepatocytes. Biochem Biophys Res<br />

Commun 319: 1040–1046, 2004.<br />

14. Bajaj G and Quan A. <strong>Renal</strong> tubular acidosis and deafness: report of a<br />

large family. Am J Kidney Dis 27: 880–882, 1996.<br />

15. Barkley RA, Chakravarti A, Cooper RS, Ellison RC, Hunt SC,<br />

Province MA, Turner ST, Weder AB, and Boerwinkle E on behalf of<br />

the Family Blood Pressure Program. Positional identification of hypertension<br />

susceptibility genes on chromosome 2. Hypertension 43:<br />

477–482, 2004.<br />

16. Becker HM and Deitmer JW. Voltage dependence of H � buffering<br />

mediated by sodium bicarbonate cotransport expressed in Xenopus oocytes.<br />

J Biol Chem 279: 28057–28062, 2004.<br />

17. Bentur L, Alon U, Mandel H, Pery M, and Berant M. Familial distal<br />

renal tubular acidosis with neurosensory deafness: early nephrocalcinosis.<br />

Am J Nephrol 9: 470–474, 1989.<br />

18. Bevensee MO, Schmitt BM, Choi I, Romero MF, and Boron WF. An<br />

electrogenic Na � -HCO 3 � cotransporter (NBC) with a novel COOH-terminus,<br />

cloned from rat brain. Am J Physiol Cell Physiol 278: C1200–<br />

C1211, 2000.<br />

19. Blackman SM, Piston DW, and Beth AH. Oligomeric state of human<br />

erythrocyte band 3 measured by fluorescence resonance energy homotransfer.<br />

Biophys J 75: 1117–1130, 1998.<br />

20. Bok D, Galbraith G, Lopez I, Woodruff M, Nusinowitz S, BeltrandelRio<br />

H, Huang W, Zhao S, Geske R, Montgomery C, Van Sligtenhorst<br />

I, Friddle C, Platt K, Sparks MJ, Pushkin A, Abuladze N,<br />

Ishiyama A, Dukkipati R, Liu W, and Kurtz I. Blindness and auditory<br />

impairment caused by loss of the sodium bicarbonate cotransporter<br />

NBC3. Nat Genet 34: 313–319, 2003.<br />

21. Bok D, Schibler MJ, Pushkin A, Sassani P, Abuladze N, Naser Z, and<br />

Kurtz I. Immunolocalization of electrogenic sodium-bicarbonate co<strong>transporters</strong><br />

pNBC1 and kNBC1 in the rat eye. Am J Physiol <strong>Renal</strong><br />

Physiol 281: F920–F935, 2001.<br />

22. Boron WF and Russel JM. Stoichiometry and ion dependencies of the<br />

intracellular pH-regulating mechanism in squid giant axons. J Gen<br />

Physiol 81: 373–399, 1983.<br />

23. Bourgeois S, Masse S, Paillard M, and Houillier P. Basolateral<br />

membrane Cl � -, Na � -, and K � -coupled <strong>base</strong> transport mechanisms in rat<br />

MTALH. Am J Physiol <strong>Renal</strong> Physiol 282: F655–F668, 2002.<br />

24. Breton S. The cellular physiology of carbonic anhydrases. JOP 2:<br />

159–164, 2001.<br />

25. Brosius FC III, Alper SL, Garcia AM, and Lodish HF. The major<br />

kidney band 3 gene transcript predicts an amino-terminal truncated band<br />

3 polypeptide. J Biol Chem 264: 7784–7787, 1989.<br />

Downloaded from<br />

http://ajprenal.physiology.org/<br />

by guest on December 23, 2012


Invited Review<br />

F594 � 2�<br />

<strong>SLC4</strong> BASE (<strong>H<strong>CO3</strong></strong> ,<strong>CO3</strong>) TRANSPORTERS<br />

26. Bruce LJ, Cope DL, Jones GK, Schofield AE, Burley M, Povey S,<br />

Unwin RJ, Wrong O, and Tanner MJA. Familial distal renal tubular<br />

acidosis is associated with mutations in the red cell anion exchanger<br />

(band 3, AE1) gene. J Clin Invest 100: 1693–1701, 1997.<br />

27. Bruce LJ, Wrong O, Toye AM, Young MT, Ogle G, Ismail Z, Sinha<br />

AK, McMaster P, Hwaihwanje I, Nash GB, Hart S, Lavu E, Palmer<br />

R, Othman A, Unwin RJ, and Tanner MJ. Band 3 mutations, renal<br />

tubular acidosis and South-East Asian ovalocytosis in Malaysia and<br />

Papua New Guinea: loss of up to 95% band 3 transport in red cells.<br />

Biochem J 350: 41–51, 2000.<br />

28. Burckhardt BC, Sato K, and Fromter E. Electrophysiological analysis<br />

of bicarbonate permeation across the peritubular cell membrane of rat<br />

kidney proximal tubule. I. Basic observations. Pflügers Arch 401: 34–42,<br />

1984.<br />

29. Burg M and Green N. Bicarbonate transport by isolated perfused rabbit<br />

proximal convoluted tubules. Am J Physiol <strong>Renal</strong> Fluid Electrolyte<br />

Physiol 233: F307–F314, 1977.<br />

30. Burnham CE, Amlal H, Wang Z, Shull GE, and Soleimani M.<br />

Cloning and functional expression of a human kidney Na � :HCO 3 �<br />

cotransporter. J Biol Chem 272: 19111–19114, 1997.<br />

31. Cabantchik ZI and Greger R. Chemical probes for anion <strong>transporters</strong><br />

of mammalian cell membranes. Am J Physiol Cell Physiol 262: C803–<br />

C827, 1992.<br />

32. Casey JR and Reithmeier RAF. Analysis of the oligomeric state of<br />

band 3, the anion transport protein of the human erythrocyte membrane,<br />

by size exclusion high performance liquid chromatography. J Biol Chem<br />

266: 15726–15737, 1991.<br />

33. Casey JR and Reithmeier RAF. Anion exchangers in the red cell and<br />

beyond. Biochem Cell Biol 76: 709–713, 1998.<br />

34. Cheidde L, Vieira TC, Lima PR, Saad ST, and Heilberg IP. A novel<br />

mutation in the anion exchanger 1 gene is associated with familial distal<br />

renal tubular acidosis and nephrocalcinosis. Pediatrics 112: 1361–1367,<br />

2003.<br />

35. Chernova MN, Jiang L, Crest M, Hand M, Vandorpe DH, Strange<br />

K, and Alper SL. Electrogenic sulfate/chloride exchange in Xenopus<br />

oocytes mediated by murine AE1 E699Q. J Gen Physiol 109: 345–360,<br />

1997.<br />

36. Cheung JC and Reithmeier RA. Palmitoylation is not required for<br />

trafficking of human anion exchanger 1 to the cell surface. Biochem J<br />

378: 1015–1021, 2004.<br />

37. Choi I, Aalkjaer C, Boulpaep EL, and Boron WF. An electroneutral<br />

sodium/bicarbonate cotransporter NBCn1 and associated sodium channel.<br />

Nature 405: 571–575, 2000.<br />

38. Choi I, Hu L, Rojas JD, Schmidtt BM, and Boron WF. Role of<br />

glycosylation in the renal electrogenic Na � -HCO 3 � cotransporter<br />

(NBCe1). Am J Physiol <strong>Renal</strong> Physiol 284: F1199–F1206, 2003.<br />

39. Choi I, Rojas JD, Kobayashi C, and Boron WF. Functional characterization<br />

of “NCBE”, an electroneutral Na/<strong>H<strong>CO3</strong></strong> cotransporter (Abstract).<br />

FASEB J 16: A796, 2002.<br />

40. Chow A, Dobbins JW, Aronson PS, and Igarashi P. cDNA cloning<br />

and localization of a band 3-related protein from ileum. Am J Physiol<br />

Gastrointest Liver Physiol 263: G345–G352, 1992.<br />

41. Clarke S. The size and detergent binding of membrane proteins. J Biol<br />

Chem 250: 5459–5469, 1975.<br />

42. Cooper DS, Saxena NC, Yang HS, Lee HJ, Moring AG, Lee A, and<br />

Choi I. Molecular and functional characterization of the electroneutral<br />

Na/<strong>H<strong>CO3</strong></strong> cotransporter NBCn1 in rat hippocampal neurons. J Biol Chem<br />

280: 17823–17830, 2005.<br />

43. Cyvin KB, Weidemann J, and Bathen J. Lowes syndrome. Acta<br />

Paediatr Scand 62: 309–312, 1973.<br />

44. DeFelice LJ. Transporter structure and mechanism. TRENDS Neurosci<br />

27: 352–359, 2004.<br />

45. Dahl NK, Jiang L, Chernova MN, Stuart-Tilley AK, Shmukler BE,<br />

and Alper SL. Deficient HCO 3 � transport in an AE1 mutant with normal<br />

Cl � transport can be rescued by carbonic anhydrase II presented on an<br />

adjacent AE1 protomer. J Biol Chem 278: 44949–44958, 2003.<br />

46. Dawson PA and Markovich D. Pathogenetics of the human SLC26<br />

<strong>transporters</strong>. Curr Med Chem 12: 385–396, 2005.<br />

47. Devonald MA, Smith AN, Poon JP, Ihrke G, and Karet FE. Nonpolarized<br />

targeting of AE1 causes autosomal dominant distal renal<br />

tubular acidosis. Nat Genet 33: 125–127, 2003.<br />

48. Diecke FP, Wen Q, Sanchez JM, Kuang K, and Fischbarg J. Immunocytochemical<br />

localization of Na � -HCO 3 � co<strong>transporters</strong> and carbonic<br />

anhydrase dependence of fluid transport in corneal endothelial cells. Am J<br />

Physiol Cell Physiol 286: C1434–C1442, 2004.<br />

49. Dinour D, Chang MH, Satoh JI, Smith BL, Angle N, Knecht A,<br />

Serban I, Holtzman EJ, and Romero MF. A novel missense mutation<br />

in the sodium bicarbonate cotransporter (NBCe1/<strong>SLC4</strong>A4) causes proximal<br />

tubular acidosis and glaucoma through ion transport defects. J Biol<br />

Chem 279: 52238–52246, 2004.<br />

50. Dodler M, Walz T, Hefti A, and Engel A. Human erythrocyte band 3.<br />

Solubilization and reconstitution into two-dimensional crystals. J Mol<br />

Biol 231: 119–132, 1993.<br />

51. Freedman JC and Novak TS. Electrodiffusion, barrier, and gating<br />

analysis of DIDS-insensitive chloride conductance in human red blood<br />

cells treated with valinomycin or gramicidin. J Gen Physiol 109: 201–<br />

216, 1997.<br />

52. Frohlich O and Gunn RB. Erythrocyte anion transport: the kinetics of<br />

a single-site obligatory exchange system. Biochim Biophys Acta 864:<br />

169–194, 1986.<br />

53. Fuginaga J, Loiselle FB, and Casey JR. Transport activity of chimeric<br />

AE2-AE3 chloride/bicarbonate anion exchange proteins. Biochem J 371:<br />

687–696, 2003.<br />

54. Fujinaga J, Tang XB, and Casey JR. Topology of the membrane<br />

domain of human erythrocyte anion exchange protein, AE1. J Biol Chem<br />

274: 6626–6633, 1999.<br />

55. Furuya W, Tarshis T, Law FY, and Knauf PA. Transmembrane<br />

effects of intracellular chloride on the inhibitory potency of extracellular<br />

H2DIDS. Evidence for two conformations of the transport site of the<br />

human erythrocyte anion exchange protein. J Gen Physiol 83: 657–681,<br />

1984.<br />

56. Gawenis LR, Ledoussal C, Judd LM, Prasad V, Alper SL, Stuart-<br />

Tilley A, Woo AL, Grisham C, Sanford LP, Doetschman T, Miller<br />

ML, and Shull GE. Mice with a targeted disruption of the AE2<br />

Cl� �<br />

/<strong>H<strong>CO3</strong></strong> exchanger are achlorhydric. J Biol Chem 279: 30531–30539,<br />

2004.<br />

57. Giffard RG, Lee YS, Ouyang YB, Murphy SL, and Monyer H. Two<br />

variants of the rat brain sodium-driven chloride bicarbonate exchanger<br />

(NCBE): developmental expression and addition of a PDZ motif. Eur<br />

J Neurosci 18: 2935–2945, 2003.<br />

58. Gresz V, Kwon TH, Vorum H, Zelles T, Kurtz I, Steward MC,<br />

Aalkjaer C, and Nielsen S. Immunolocalization of electroneutral Na�- �<br />

<strong>H<strong>CO3</strong></strong> co<strong>transporters</strong> in human and rat salivary glands. Am J Physiol<br />

Gastrointest Liver Physiol 283: G473–G480, 2002.<br />

59. Grichtchenko II and Boron WF. Surface-pH measurements in voltageclamped<br />

Xenopus oocytes co-expressing NBCe1 and CAIV: evidence for<br />

CO 3 2�<br />

AJP-<strong>Renal</strong> Physiol • VOL 290 • MARCH 2006 • www.ajprenal.org<br />

transport (Abstract). FASEB J 16: A795, 2002.<br />

60. Grichtchenko II, Choi I, Zhong X, Bray-Ward P, Russell JM, and<br />

Boron WF. Cloning, characterization and chromosomal mapping of a<br />

human electroneutral Na � -driven Cl-<strong>H<strong>CO3</strong></strong> exchanger. J Biol Chem 276:<br />

8358–8363, 2001.<br />

61. Grinstein S, McCulloch L, and Rothstein A. Transmembrane effects of<br />

irreversible inhibitors of anion transport in red blood cells. Evidence for<br />

mobile transport sites. J Gen Physiol 73: 493–514, 1979.<br />

62. Gross E, Fedotoff O, Pushkin A, Abuladze N, Newman D, and Kurtz<br />

I. Phosphorylation-induced modulation of pNBC1 function: distinct roles<br />

for the amino- and carboxy-termini. J Physiol 549: 673–682, 2003.<br />

63. Gross E, Hawkins K, Abuladze N, Pushkin A, Cotton CU, Hopfer U,<br />

and Kurtz I. The stoichiometry of the electrogenic sodium bicarbonate<br />

cotransporter NBC1 is cell-type dependent. J Physiol 531: 597–603,<br />

2001.<br />

64. Gross E, Hawkins K, Pushkin A, Abuladze N, Hopfer U, Sassani P,<br />

Dukkipati R, and Kurtz I. Phosphorylation of Ser982 in kNBC1 shifts<br />

the HCO 3 � :Na � stoichiometry from 3:1 to 2:1 in proximal tubule cells.<br />

J Physiol 537: 659–665, 2001.<br />

65. Gross E and Hopfer U. Activity and stoichiometry of Na � :HCO 3 �<br />

cotransport in immortalized renal proximal tubule cells. J Membr Biol<br />

152: 245–252, 1996.<br />

66. Gross E and Hopfer U. Voltage and cosubstrate dependence of the<br />

Na-<strong>H<strong>CO3</strong></strong> cotransporter kinetics in renal proximal tubule cells. Biophys J<br />

75: 810–824, 1998.<br />

67. Gross E and Kurtz I. Structural determinants and significance of<br />

regulation of electrogenic Na � -HCO 3 � cotransporter stoichiometry. Am J<br />

Physiol <strong>Renal</strong> Physiol 283: F876–F887, 2002.<br />

68. Gross E, Pushkin A, Abuladze N, Fedotoff O, and Kurtz I. Regulation<br />

of the sodium bicarbonate cotransporter kNBC1 function: role of Asp 986 ,<br />

Downloaded from<br />

http://ajprenal.physiology.org/<br />

by guest on December 23, 2012


Asp 988 and kNBC1-carbonic anhydrase II binding. J Physiol 544: 679–<br />

685, 2002.<br />

69. Groves JD and Tanner MJ. Co-expressed complementary fragments of<br />

the human red cell anion exchanger (band 3, AE1) generate stilbene<br />

disulfonate-sensitive anion transport. J Biol Chem 270: 9097–9105,<br />

1995.<br />

70. Groves JD and Tanner MJ. Structural model for the organization of the<br />

transmembrane spans of the human red-cell anion exchanger (band 3;<br />

AE1). Biochem J 344: 699–711, 1999.<br />

71. Groves JD, Wang L, and Tanner MJ. Functional reassembly of the<br />

anion transport domain of human red cell band 3 (AE1) from multiple<br />

and non-complementary fragments. FEBS Lett 433: 223–227, 1998.<br />

72. Gunn RB and Frohlich O. Asymmetry in the mechanism for anion<br />

exchange in human red blood cell membranes. Evidence for reciprocating<br />

sites that react with one transported anion at a time. J Gen Physiol 74:<br />

351–374, 1979.<br />

73. Hamasaki N, Izuhara K, Okubo K, Kanazawa Y, Omachi A, and<br />

Kleps RA. Inhibition of chloride binding to the anion transport site by<br />

diethylpyrocarbonate modification of band 3. J Membr Biol 116: 87–91,<br />

1990.<br />

74. Heyer M, Muller-Berger S, Romero MF, Boron WF, and Fromter E.<br />

Stoichiometry of the rat kidney Na � -HCO 3 � cotransporter expressed in<br />

Xenopus laevis oocytes. Pflügers Arch 438: 322–329, 1999.<br />

75. Horita S, Yamada H, Inatomi J, Moriyama N, Sekine T, Igarashi T,<br />

Endo Y, Dasouki M, Ekim M, Al-Gazali L, Shimadzu M, Seki G, and<br />

Fujita T. Functional analysis of NBC1 mutants associated with proximal<br />

renal tubular acidosis and ocular abnormalities. J Am Soc Nephrol 16:<br />

2270–2278, 2005.<br />

76. Hubner CA, Hentschke M, Jacobs S, and Hermans-Borgmeyer I.<br />

Expression of the sodium-driven chloride bicarbonate exchanger NCBE<br />

during prenatal mouse development. Gene Expr Patterns 5: 219–223,<br />

2004.<br />

77. Humphreys BD, Jiang L, Chernova MN, and Alper SL. Hypertonic<br />

activation of AE2 anion exchanger in Xenopus oocytes via NHE-mediated<br />

intracellular alkalinization. Am J Physiol Cell Physiol 268: C201–<br />

C209, 1995.<br />

78. Igarashi T, Inatomi J, Sekine T, Cha SH, Kanai Y, Kunimi M,<br />

Tsukamoto K, Satoh H, Shimadzu M, Tozawa F, Mori T, Shiobara<br />

M, Seki G, and Endou H. Mutations in <strong>SLC4</strong>A4 cause permanent<br />

isolated proximal renal tubular acidosis with ocular abnormalities. Nat<br />

Genet 23: 264–266, 1999.<br />

79. Igarashi T, Inatomi J, Sekine T, Seki G, Shimadzu M, Tozawa F,<br />

Takeshima Y, Takumi T, Takahashi T, Yoshikawa N, Nakamura H,<br />

and Endou H. Novel nonsense mutation in the Na � /HCO 3 � cotransporter<br />

gene (<strong>SLC4</strong>A4) in a patient with permanent isolated proximal renal<br />

tubular acidosis and bilateral glaucoma. J Am Soc Nephrol 12: 713–718,<br />

2001.<br />

80. Inaba M, Yawata A, Koshino I, Sato K, Takeuchi M, Takakuwa Y,<br />

Manno S, Yawata Y, Kanzaki A, Sakai J, Ban A, Ono K, and Maede<br />

Y. Defective anion transport and marked spherocytosis with membrane<br />

instability caused by hereditary total deficiency of red cell band 3 in cattle<br />

due to a nonsense mutation. J Clin Invest 97: 1804–1817, 1996.<br />

81. Inatomi J, Horita S, Braverman N, Sekine T, Yamada H, Suzuki Y,<br />

Kawahara K, Moriyama N, Kudo A, Kawakami H, Shimadzu M,<br />

Endou H, Fujita T, Seki G, and Igarashi T. Mutational and functional<br />

analysis of <strong>SLC4</strong>A4 in a patient with proximal renal tubular acidosis.<br />

Pflügers Arch 448: 438–444, 2004.<br />

82. Jakobsen JK, Odgaard E, Wang W, Elkjaer ML, Nielsen S, Aalkjaer<br />

C, and Leipziger J. Functional up-regulation of basolateral Na � -dependent<br />

HCO 3 � transporter NBCn1 in medullary thick ascending limb of<br />

K � -depleted rats. Pflügers Arch 448: 571–578, 2004.<br />

83. Jarolim P, Shayakul C, Prabakaran D, Jiang L, Stuart-Tilley A,<br />

Rubin HL, Simova S, Zavadil J, Herrin JT, Brouillette J, Somers<br />

MJG, Seemanova E, Brugnara C, Guay-Woodford LM, and Alper<br />

SL. Autosomal dominant distal renal tubular acidosis is associated in<br />

three families with heterozygosity for the R589H mutation in the AE1<br />

(band 3) Cl � /HCO 3 � exchanger. J Biol Chem 273: 6380–6388, 1998.<br />

84. Jennings ML. Stoichiometry of a half-turnover of band 3, the chloride<br />

transport protein of human erythrocytes. J Gen Physiol 79: 169–185,<br />

1982.<br />

85. Jennings ML. Evidence for a second binding/transport site for chloride<br />

in erythrocyte anion transporter AE1 modified at glutamate 681. Biophys<br />

J 88: 2681–2691, 2005.<br />

� 2�<br />

<strong>SLC4</strong> BASE (<strong>H<strong>CO3</strong></strong> ,<strong>CO3</strong>) TRANSPORTERS<br />

AJP-<strong>Renal</strong> Physiol • VOL 290 • MARCH 2006 • www.ajprenal.org<br />

Invited Review<br />

F595<br />

86. Jennings ML and Anderson MP. Chemical modification and labeling<br />

of glutamate residues at the stilbenedisulfonate site of human red blood<br />

cell band 3 protein. J Biol Chem 262: 1691–1697, 1987.<br />

87. Jennings ML and Smith JS. Anion-proton cotransport through the<br />

human red blood cell band 3 protein. Role of glutamate 681. J Biol Chem<br />

267: 13964–13971, 1992.<br />

88. Jennings ML, Whitlock J, and Shinde A. Pre-steady state transport by<br />

erythrocyte band 3 protein: uphill countertransport induced by the impermeant<br />

inhibitor H2DIDS. Biochem Cell Biol 76: 807–813, 1998.<br />

89. Jensen LJ, Schmitt BM, Berger UV, Nsumu NN, Boron WF, Hediger<br />

MA, Brown D, and Breton S. Localization of sodium bicarbonate<br />

cotransporter (NBC) protein and messenger ribonucleic acid in rat<br />

epididymis. Biol Reprod 60: 573–579, 1999.<br />

90. Jentsch TJ, Schwartz P, Schill BS, Langner B, Lepple AP, Keller SK,<br />

and Wiederholt M. Kinetic properties of the sodium bicarbonate (carbonate)<br />

symport in monkey kidney epithelial cells (BSC-1). J Biol Chem<br />

261: 10673–10679, 1986.<br />

91. Jiang L, Stuart-Tilley A, Parkash J, and Alper SL. pHi and serum<br />

regulate AE2-mediated Cl � /HCO 3 � exchange in CHOP cells of defined<br />

transient transfection status. Am J Physiol Cell Physiol 267: C845–C856,<br />

1994.<br />

92. Jin XR, Abe Y, Li CY, and Hamasaki N. Histidine-834 of human<br />

erythrocyte band 3 has an essential role in the conformational changes<br />

that occur during the band 3-mediated anion exchange. Biochemistry 42:<br />

12927–12932, 2003.<br />

93. Kang D, Okubo K, Hamasaki N, Kuroda N, and Shiraki H. A<br />

structural study of the membrane domain of band 3 by tryptic digestion.<br />

Conformational change of band 3 in situ induced by alkali treatment.<br />

J Biol Chem 267: 19211–19217, 1992.<br />

94. Kanki T, Sakaguchi M, Kitamura A, Sato T, Mihara K, and Hamasaki<br />

N. The tenth membrane region of band 3 is initially exposed to<br />

the luminal side of the endoplasmic reticulum and then integrated into a<br />

partially folded band 3 intermediate. Biochemistry 41: 13973–13981,<br />

2002.<br />

95. Kannan R, Yuan J, and Low PS. Isolation and partial characterization<br />

of antibody- and globin-enriched complexes from membranes of dense<br />

human erythrocytes. Biochem J 278: 57–62, 1991.<br />

96. Kaplan JH, Scorah K, Fasold H, and Passow H. Sidedness of the<br />

inhibition action of disulfonic acids on chloride equilibrium exchange<br />

and net transport across the human erythrocyte membrane. FEBS Lett 62:<br />

182–185, 1976.<br />

97. Karbach D, Staub M, Wood PG, and Passow H. Effect of site-directed<br />

mutagenesis of the arginine residues 509 and 748 on mouse band 3<br />

protein-mediated anion transport. Biochim Biophys Acta 1371: 114–122,<br />

1998.<br />

98. Karet FE, Finberg KE, Nelson RD, Nayir A, Mocan H, Sanjad SA,<br />

Rodriguez-Soriano J, Santos F, Cremers CW, Di Pietro A, Hoffbrand<br />

BI, Winiarski J, Bakkaloglu A, Ozen S, Dusunsel R, Goodyer<br />

P, Hulton SA, Wu DK, Skvorak AB, Morton CC, Cunningham MJ,<br />

Jha V, and Lifton RP. Mutations in the gene encoding B1 subunit of<br />

H � -ATPase cause renal tubular acidosis with sensorineural deafness. Nat<br />

Genet 21: 84–90, 1999.<br />

99. Karet FE, Gainza FJ, Gyory AZ, Unwin RJ, Wrong O, Tanner MJ,<br />

Nayir A, Alpay H, Santos F, Hulton SA, Bakkaloglu A, Ozen S,<br />

Cunningham MJ, di Pietro A, Walker WG, and Lifton RP. Mutations<br />

in the chloride-bicarbonate exchanger gene AE1 cause autosomal dominant<br />

but not autosomal recessive distal renal tubular acidosis. Proc Natl<br />

Acad Sci USA 95: 6337–6342, 1998.<br />

100. Kawano Y, Okubo K, Tokunaga F, Miyata T, Iwanaga S, and<br />

Hamasaki N. Localization of the pyridoxal phosphate binding site at the<br />

COOH-terminal region of erythrocyte band 3 protein. J Biol Chem 263:<br />

8232–8238, 1988.<br />

101. Keats BJ and Savas S. Genetic heterogeneity in Usher syndrome. Am J<br />

Med Genet 130: 13–16, 2004.<br />

102. Kifor G, Toon MR, Janoshazi A, and Solomon AK. Interaction<br />

between red cell band 3 and cytosolic carbonic anhydrase. J Membr Biol<br />

134: 169–179, 1993.<br />

103. Kittanakom S, Cordat E, Akkarapatumwong V, Yenchitsomanus<br />

PT, and Reithmeier RA. Trafficking defects of a novel autosomal<br />

recessive distal renal tubular acidosis mutant (S773P) of the human<br />

kidney anion exchanger (kAE1). J Biol Chem 279: 40960–40971, 2004.<br />

104. Knauf PA, Spinelli LJ, and Mann NA. Flufenamic acid senses conformation<br />

and asymmetry of human erythrocyte band 3 anion transport<br />

protein. Am J Physiol Cell Physiol 257: C277–C289, 1989.<br />

Downloaded from<br />

http://ajprenal.physiology.org/<br />

by guest on December 23, 2012


Invited Review<br />

F596 � 2�<br />

<strong>SLC4</strong> BASE (<strong>H<strong>CO3</strong></strong> ,<strong>CO3</strong>) TRANSPORTERS<br />

105. Ko SB, Luo X, Hager H, Rojek A, Choi JY, Licht C, Suzuki M,<br />

Muallem S, Nielsen S, and Ishibashi K. AE4 is a DIDS-sensitive<br />

Cl� �<br />

/<strong>H<strong>CO3</strong></strong> exchanger in the basolateral membrane of the renal CCD and<br />

the SMG duct. Am J Physiol Cell Physiol 283: C1206–C1218, 2002.<br />

106. Kobayashi S, Morgans CW, Casey JR, and Kopito RR. AE3 anion<br />

exchanger isoforms in the vertebrate retina: developmental regulation<br />

and differential expression in neurons and glia. J Neurosci 14: 6266–<br />

6279, 1994.<br />

107. Kopito RR and Lodish HF. Primary structure and transmembrane<br />

orientation of the murine anion exchange protein. Nature 316: 234–238,<br />

1985.<br />

108. Kopito RR, Lee BS, Simmons DM, Lindsey AE, Morgans CW, and<br />

Schneider K. Regulation of intracellular pH by a neuronal homologue of<br />

the erythrocyte anion exchanger. Cell 59: 927–937, 1989.<br />

109. Kristensen JM, Kristensen M, and Juel C. Expression of Na� �<br />

/<strong>H<strong>CO3</strong></strong> co-transporter proteins (NBCs) in rat and human skeletal muscle. Acta<br />

Physiol Scand 182: 69–76, 2004.<br />

110. Kudrycki KE, Newman PR, and Schull GE. cDNA cloning and tissue<br />

distribution of mRNAs for two proteins that are related to the band 3<br />

Cl� �<br />

/<strong>H<strong>CO3</strong></strong> exchanger. J Biol Chem 265: 462–471, 1990.<br />

111. Kuma H, Shinde AA, Howren TR, and Jennings ML. Topology of the<br />

anion exchange protein AE1: the controversial sidedness of lysine 743.<br />

Biochemistry 41: 3380–3388, 2002.<br />

112. Kurtz I, Petrasek D, and Tatishchev S. Molecular mechanisms of<br />

electrogenic sodium bicarbonate cotransport: structural and equilibrium<br />

thermodynamic considerations. J Membr Biol 197: 77–90, 2004.<br />

113. Kwon TH, Fulton C, Wang W, Kurtz I, Frøkiær J, Aalkjaer C, and<br />

Nielsen S. Chronic metabolic acidosis upregulates rat kidney Na-<strong>H<strong>CO3</strong></strong><br />

co<strong>transporters</strong> NBCn1 and NBC3 but not NBC1. Am J Physiol <strong>Renal</strong><br />

Physiol 282: F341–F351, 2002.<br />

114. Kwon TH, Pushkin A, Abuladze N, Nielsen S, and Kurtz I. Immunoelectron<br />

microscopic localization of NBC3 sodium-bicarbonate cotransporter<br />

in rat kidney. Am J Physiol <strong>Renal</strong> Physiol 278: F327–F336,<br />

2000.<br />

115. Lee BS, Gunn RB, and Kopito RR. Functional differences among<br />

nonerythroid anion exchangers expressed in a transfected human cell<br />

line. J Biol Chem 266: 11448–11454, 1991.<br />

116. Li HC, Szigligeti P, Worrell RT, Matthews JB, Conforti L, and<br />

Soleimani M. Missense mutations in Na� �<br />

:<strong>H<strong>CO3</strong></strong> cotransporter NBC1<br />

show abnormal trafficking in polarized kidney cells: a basis of proximal<br />

renal tubular acidosis. Am J Physiol <strong>Renal</strong> Physiol 289: F61–F71, 2005.<br />

117. Li HC, Worrell RT, Matthews JB, Husseinzadeh H, Neumeier L,<br />

Petrovic S, Conforti L, and Soleimani M. Identification of a carboxylterminal<br />

motif essential for the targeting of Na� �<br />

-<strong>H<strong>CO3</strong></strong> cotransporter<br />

NBC1 to the basolateral membrane. J Biol Chem 279: 43190–43197,<br />

2004.<br />

118. Linn SC, Kudrycki KE, and Shull GE. The predicted translation<br />

product of a cardiac AE3 mRNA contains an N terminus distinct from<br />

that of the brain AE3 Cl� �<br />

/<strong>H<strong>CO3</strong></strong> exchanger. Cloning of a cardiac AE3<br />

cDNA, organization of the AE3 gene, and identification of an alternative<br />

transcription initiation site. J Biol Chem 267: 7927–7935, 1992.<br />

119. Loiselle FB, Jaschke P, and Casey JR. Structural and functional<br />

characterization of the human NBC3 sodium/bicarbonate co-transporter<br />

carboxyl-terminal cytoplasmic domain. Mol Membr Biol 20: 307–317,<br />

2003.<br />

120. Low PS, Rathinavelu P, and Harrison ML. Regulation of glycolysis<br />

via enzyme binding to the membrane band 3. J Biol Chem 268: 14627–<br />

14631, 1993.<br />

121. Low PS, Willardson BM, Mohandas N, Rossi M, and Shohet S.<br />

Contribution of band 3-ankyrin interaction to erythrocyte membrane<br />

mechanical stability. Blood 77: 1581–1586, 1991.<br />

122. Luo X, Choi JY, Ko SB, Pushkin A, Kurtz I, Ahn W, Lee MG, and<br />

�<br />

Muallem S. <strong>H<strong>CO3</strong></strong> salvage mechanisms in the submandibular gland<br />

acinar, and duct cells. J Biol Chem 276: 9808–9816, 2001.<br />

123. Lux SE, John KM, Kopito RR, and Lodish HF. Cloning and characterization<br />

of band 3, the human erythrocyte anion-exchange protein<br />

(AE1). Proc Natl Acad Sci USA 86: 9089–9093, 1989.<br />

124. Malik S, Sami M, and Watts A. A role for band 4.2 in human<br />

erythrocyte band 3 mediated anion transport. Biochemistry 32: 10078–<br />

10084, 1993.<br />

125. Marino CR, Jeanes V, Boron WF, and Schmitt BM. Expression and<br />

distribution of the Na� �<br />

-<strong>H<strong>CO3</strong></strong> cotransporter in human pancreas. Am J<br />

Physiol Gastrointest Liver Physiol 277: G487–G494, 1999.<br />

AJP-<strong>Renal</strong> Physiol • VOL 290 • MARCH 2006 • www.ajprenal.org<br />

126. Matsuyama H, Kawano Y, and Hamasaki N. Involvement of a histidine<br />

residue in inorganic phosphate and phosphoenolpyruvate transport<br />

across the human erythrocyte membrane. J Biochem 99: 495–501, 1986.<br />

127. Maunsbach AB, Vorum H, Kwon TH, Nielsen S, Simonsen B, Choi<br />

I, Schmitt BM, Boron WF, and Aalkjaer C. Immunoelectron microscopic<br />

localization of the electrogenic Na/<strong>H<strong>CO3</strong></strong> cotransporter in rat and<br />

Ambystoma kidney. J Am Soc Nephrol 11: 2179–2189, 2000.<br />

128. McKinney TD and Burg MB. Biocarbonate and fluid absorption by<br />

renal proximal straight tubules. Kidney Int 12: 1–8, 1977.<br />

129. Medina JF, Lecanda J, Acin A, Ciesielczyk P, and Prieto J. Tissuespecific<br />

N-terminal isoforms from overlapping alternate promoters of the<br />

human AE2 anion exchanger gene. Biochem Biophys Res Commun 267:<br />

228–235, 2000.<br />

130. Medina JF, Recalde S, Prieto J, Lecanda J, Saez E, Funk CD, Vecino<br />

P, van Roon MA, Ottenhoff R, Bosma PJ, Bakker CT, and Elferink<br />

RP. Anion exchanger 2 is essential for spermiogenesis in mice. Proc Natl<br />

Acad Sci USA 100: 15847–14852, 2003.<br />

131. Mount DB and Romero MF. The SLC26 gene family of multifunctional<br />

anion exchangers. Pflügers Arch 447: 710–721, 2004.<br />

132. Müller-Berger S, Ducoudret O, Diakov A, and Fromter E. The renal<br />

Na-HCO 3 � cotransporter expressed in Xenopus laevis oocytes: change in<br />

stoichiometry in response to elevation of cytosolic Ca 2� concentration.<br />

Pflügers Arch 442: 718–728, 2001.<br />

133. Müller-Berger S, Karbach D, Kang D, Aranibar N, Wood PG,<br />

Ruterjans H, and Passow H. Roles of histidine 752 and glutamate 699<br />

in the pH dependence of mouse band 3 protein-mediated anion transport.<br />

Biochemistry 34: 9325–9332, 1995.<br />

134. Müller-Berger S, Karbach D, Konig J, Lepke S, Wood PG, Appelhans<br />

H, and Passow H. Inhibition of mouse erythroid band 3-mediated<br />

chloride transport by site-directed mutagenesis of histidine residues and<br />

its reversal by second site mutation of Lys 558, the locus of covalent<br />

H2DIDS binding. Biochemistry 34: 9315–9324, 1995.<br />

135. Müller-Berger S, Nesterov VV, and Fromter E. Partial recovery of in<br />

vivo function by improved incubation conditions of isolated renal proximal<br />

tubule. II. Change of Na-<strong>H<strong>CO3</strong></strong> cotransport stoichiometry and of<br />

response to acetazolamide. Pflügers Arch 434: 383–391, 1997.<br />

136. Nakashima H and Makino S. Purification and characterization of band<br />

3, the major intrinsic membrane protein of the bovine erythrocyte<br />

membrane. J Biochem 87: 899–910, 1983.<br />

137. Nishimori I, Fujikawa Adachi K, Onishi S, and Hollingsworth MA.<br />

Carbonic anhydrase in human pancreas: hypotheses for the pathophysiological<br />

roles of CA isozymes. Ann NY Acad Sci 880: 5–16, 1999.<br />

138. Nishimori I and Onishi S. Carbonic anhydrase isozymes in the human<br />

pancreas. Dig Liver Dis 33: 68–74, 2001.<br />

139. Odgaard E, Jakobsen JK, Frische S, Praetorius J, Nielsen S, Aalkjaer<br />

C, and Leipziger J. Basolateral Na � -dependent HCO 3 � transporter<br />

NBCn1-mediated HCO 3 � influx in rat medullary thick ascending limb.<br />

J Physiol 555: 205–218, 2004.<br />

140. Okubo K, Kang D, Hamasaki N, and Jennings ML. Red blood cell<br />

band 3. Lysine 539 and lysine 851 react with the same H2DIDS<br />

(4,4�-diisothiocyanodihydrostilbene-2,2�-disulfonic acid) molecule.<br />

J Biol Chem 269: 1918–1926, 1994.<br />

141. Pappert G and Schubert D. The state of association of band 3 protein<br />

of the human erythrocyte membrane in solutions of nonionic detergents.<br />

Mol Membr Biol 15: 153–158, 1983.<br />

142. Park M, Li Q, Shcheynikov N, Muallem S, and Zeng W. Borate<br />

transport and cell growth and proliferation. Not only in plants. Cell Cycle<br />

4: 024–026, 2005.<br />

143. Park M, Li Q, Shcheynikov N, Zeng W, and Muallem S. NaBC1 is a<br />

ubiquitous electrogenic Na � -coupled borate transporter essential for<br />

cellular boron homeostasis and cell growth and proliferation. Mol Cell<br />

16: 331–341, 2004.<br />

144. Parker MD, Boron WF, and Tanner MJ. Characterization of human<br />

“AE4” as an electroneutral sodium bicarbonate cotransporter (Abstract).<br />

FASEB J 16: A796, 2002.<br />

145. Parker MD, Ourmozdi EP, and Tanner MJ. Human BTR1, a new<br />

bicarbonate transporter superfamily member and human AE4 from kidney.<br />

Biochem Biophys Res Commun 282: 1103–1109, 2001.<br />

146. Passow H. Molecular aspects of band 3 protein-mediated anion transport<br />

across the red blood cell membrane. Rev Physiol Biochem Pharmacol<br />

103: 61–203, 1986.<br />

147. Peters LL, Shivdasani RA, Liu SC, Hanspal M, John KM, Gonzalez<br />

JM, Brugnara C, Gwynn B, Mohandas N, Alper SL, Orkin SH, and<br />

Lux SE. Anion exchanger 1 (band 3) is required to prevent erythrocyte<br />

Downloaded from<br />

http://ajprenal.physiology.org/<br />

by guest on December 23, 2012


membrane surface loss but not to form the membrane skeleton. Cell 86:<br />

917–927, 1996.<br />

148. Pierucci-Alves F, O’Leary T, Lorencen M, and Schultz BD. Multiple<br />

NBCe1 splice variants are expressed in vas deferens epithelial cells.<br />

FASEB J 19: A141, 2005.<br />

149. Popov M, Li J, and Reithmeier RAF. Transmembrane folding of the<br />

human erythrocyte anion exchanger (AE1, Band 3) determined by<br />

scanning and insertional N-glycosylation mutagenesis. Biochem J 339:<br />

269–279, 1999.<br />

150. Praetorius J, Hager H, Nielsen S, Aalkjaer C, Friis UG, Ainsworth<br />

MA, and Johansen T. Molecular and functional evidence for electrogenic<br />

and electroneutral Na � -HCO 3 � co<strong>transporters</strong> in murine duodenum.<br />

Am J Physiol Gastrointest Liver Physiol 280: G332–G343, 2001.<br />

151. Praetorius J, Kim YH, Bouzinova EV, Frische S, Rojek A, Aalkjaer<br />

C, and Nielsen S. NBCn1 is a basolateral Na � -HCO 3 � cotransporter in<br />

rat kidney inner medullary collecting ducts. Am J Physiol <strong>Renal</strong> Physiol<br />

286: F903–F912, 2004.<br />

152. Praetorius J, Nejsum LN, and Nielsen S. A SCL4A10 gene product<br />

maps selectively to the basolateral plasma membrane of choroid plexus<br />

epithelial cells. Am J Physiol Cell Physiol 286: C601–C610, 2004.<br />

153. Praetorius J and Nielsen S. <strong>SLC4</strong>a10/NCBE is expressed in the choroid<br />

plexus, kidney tubules and duodenum (Abstract). FASEB J 19: A140,<br />

2005.<br />

154. Pushkin A, Abuladze N, Gross E, Newman D, Tatishchev S, Lee I,<br />

Fedotoff O, Bondar G, Azimov R, Ngyuen M, and Kurtz I. Molecular<br />

mechanism of kNBC1-carbonic anhydrase II interaction in proximal<br />

tubule cells. J Physiol 559: 55–65, 2004.<br />

155. Pushkin A, Abuladze N, Lee I, Newman D, Hwang J, and Kurtz I.<br />

Cloning, tissue distribution, genomic organization, and functional characterization<br />

of NBC3, a new member of the sodium bicarbonate cotransporter<br />

family. J Biol Chem 274: 16569–16575, 1999.<br />

156. Pushkin A, Abuladze N, Newman D, Lee I, Xu G, and Kurtz I.<br />

Cloning, characterization and chromosomal assignment of NBC4, a new<br />

member of the sodium bicarbonate cotransporter family. Biochim Biophys<br />

Acta 1493: 215–218, 2000.<br />

157. Pushkin A, Abuladze N, Newman D, Lee I, Xu G, and Kurtz I. Two<br />

C-terminal variants of NBC4, a new member of the sodium bicarbonate<br />

cotransporter family: cloning, characterization, and localization. IUBMB<br />

Life 50: 13–19, 2000.<br />

158. Pushkin A, Clark I, Kwon TH, Nielsen S, and Kurtz I. Immunolocalization<br />

of NBC3 and NHE3 in the rat epididymis: colocalization of<br />

NBC3 and the vacuolar H � -ATPase. J Androl 21: 708–720, 2000.<br />

159. Pushkin A, Sassani P, Abuladze N, Newman D, Tatishchev S, and<br />

Kurtz I. Oligomeric structure of electrogenic sodium bicarbonate co<strong>transporters</strong>.<br />

J Am Soc Nephrol 12: 8A, 2001.<br />

160. Pushkin AV, Tsuprun VL, Abuladze NK, Newman D, and Kurtz I.<br />

Oligomeric structure of bAE3 protein. IUBMB Life 50: 397–401, 2000.<br />

161. Pushkin A, Yip KP, Clark I, Abuladze N, Kwon TH, Tsuruoka S,<br />

Schwartz GJ, Nielsen S, and Kurtz I. NBC3 expression in rabbit<br />

collecting duct: colocalization with vacuolar H � -ATPase. Am J Physiol<br />

<strong>Renal</strong> Physiol 277: F974–F981, 1999.<br />

162. Rebello G, Rajkumar R, Vorste V, Roberts L, Ehrenreich L, Oppon<br />

E, Gama D, Bardien S, Greenberg J, Bonapace G, Waheed A, Shah<br />

GN, and Sly WS. Apoptosis-inducing signal sequence mutation in<br />

carbonic anhydrase IV identified in patients with the RP17 form of<br />

retinitis pigmentosa. Proc Natl Acad Sci USA 101: 6617–6622, 2004.<br />

163. Reithmeier RA. A membrane metabolon linking carbonic anhydrase<br />

with chloride/bicarbonate anion exchangers. Blood Cells Mol Dis 27:<br />

85–89, 2001.<br />

164. Ribeiro ML, Alloisio N, Almeida H, Gomes C, Texier P, Lemos C,<br />

Mimoso G, Morle L, Bey-Cabet F, Rudigoz RC, Delaunay J, and<br />

Tamagnini G. Severe hereditary spherocytosis and distal renal tubular<br />

acidosis associated with the total absence of band 3. Blood 96: 1602–<br />

1604, 2000.<br />

165. Richards SM, Jaconi ME, Vassort G, and Puceat M. A spliced variant<br />

of AE1 gene encodes a truncated form of Band 3 in heart: the predominant<br />

anion exchanger in ventricular myocytes. J Cell Sci 112: 1519–<br />

1528, 1999.<br />

166. Romero MF, Hediger MA, Boulpaep EL, and Boron WF. Expression<br />

cloning and characterization of a renal electrogenic Na � /HCO 3 � cotransporter.<br />

Nature 387: 409–413, 1997.<br />

167. Romero MF, Henry D, Nelson S, Harte PJ, Dillon AK, and Sciortino<br />

CM. Cloning and characterization of a Na � -driven anion exchanger<br />

� 2�<br />

<strong>SLC4</strong> BASE (<strong>H<strong>CO3</strong></strong> ,<strong>CO3</strong>) TRANSPORTERS<br />

AJP-<strong>Renal</strong> Physiol • VOL 290 • MARCH 2006 • www.ajprenal.org<br />

Invited Review<br />

F597<br />

(NDAE1). A new bicarbonate transporter. J Biol Chem 275: 24552–<br />

24559, 2000.<br />

168. Rossmann H, Bachmann O, Wang Z, Shull GE, Obermaier B,<br />

Stuart-Tilley A, Alper SL, and Seidler U. Differential expression and<br />

regulation of AE2 anion exchanger subtypes in rabbit parietal and<br />

mucous cells. J Physiol 534: 837–848, 2001.<br />

169. Roussa E, Nastainczyk W, and Thevenod F. Differential expression of<br />

electrogenic NBC1 (<strong>SLC4</strong>A4) variants in rat kidney and pancreas.<br />

Biochem Biophys Res Commun 314: 382–389, 2004.<br />

170. Roussa E, Romero MF, Schmitt BM, Boron WF, Alper SL, and<br />

Thevenod F. Immunolocalization of anion exchanger AE2 and Na � -<br />

HCO 3 � cotransporter in rat parotid and submandibular glands. Am J<br />

Physiol Gastrointest Liver Physiol 277: G1288–G1296, 1999.<br />

171. Rungroj N, Devonald MA, Cuthbert AW, Reimann F, Akkarapatumwong<br />

V, Yenchitsomanus PT, Bennett WM, and Karet FE. A<br />

novel missense mutation in AE1 causing autosomal dominant distal renal<br />

tubular acidosis retains normal transport function but is mistargeted in<br />

polarized epithelial cells. J Biol Chem 279: 13833–13838, 2004.<br />

172. Saier MH Jr. A functional-phylogenetic classification system for transmembrane<br />

solute <strong>transporters</strong>. Microbiol Mol Biol Rev 64: 354–411,<br />

2000.<br />

173. Sander T, Toliat MR, Heils A, Leschik G, Becker C, Ruschendorf F,<br />

Rohde K, Mundlos S, and Nurnberg P. Association of the 867 Asp<br />

variant of the human anion exchanger 3 gene with common subtypes of<br />

idiopathic generalized epilepsy. Epilepsy Res 51: 249–255, 2002.<br />

174. Sasaki S and Marumo F. Effects of carbonic anhydrase inhibitors on<br />

basolateral <strong>base</strong> transport of rabbit proximal straight tubule. Am J Physiol<br />

<strong>Renal</strong> Fluid Electrolyte Physiol 257: F947–F952, 1989.<br />

175. Sasaki S, Shiigai T, Yoshiyama N, and Takeuchi J. Mechanism of<br />

bicarbonate exit across basolateral membrane of rabbit proximal straight<br />

tubule. Am J Physiol <strong>Renal</strong> Fluid Electrolyte Physiol 252: F11–F18,<br />

1987.<br />

176. Sassani P, Pushkin A, Gross E, Gomer A, Abuladze N, Dukkipati R,<br />

Carpenito G, and Kurtz I. Functional characterization of NBC4: a new<br />

electrogenic sodium-bicarbonate cotransporter. Am J Physiol Cell<br />

Physiol 282: C408–C416, 2002.<br />

177. Satoh H, Moriyama N, Hara C, Yamada H, Horita S, Kunimi M,<br />

Tsukamoto K, Iso-o N, Inatomi J, Kawakami H, Kudo A, Endou H,<br />

Igarashi T, Goto A, Fujita T, and Seki G. Localization of Na � -HCO 3 �<br />

cotransporter (NBC-1) variants in rat and human pancreas. Am J Physiol<br />

Cell Physiol 284: C729–C737, 2003.<br />

178. Schmitt BM, Biemesderfer D, Romero MF, Boulpaep EL, and Boron<br />

WF. Immunolocalization of the electrogenic Na � -HCO 3 � cotransporter<br />

in mammalian and amphibian kidney. Am J Physiol <strong>Renal</strong> Physiol 276:<br />

F27–F38, 1999.<br />

179. Schofield AE, Reardon DM, and Tanner MJ. Defective anion transport<br />

activity of the abnormal band 3 in hereditary ovalocytic red blood<br />

cells. Nature 355: 836–838, 1992.<br />

180. Schwartz GJ. <strong>Physiology</strong> and molecular biology of renal carbonic<br />

anhydrase. J Nephrol 15: S61–S74, 2002.<br />

181. Seki G and Fromter E. Acetazolamide inhibition of basolateral <strong>base</strong><br />

exit in rabbit renal proximal tubule S2 segment. Pflügers Arch 422:<br />

60–65, 1992.<br />

182. Sekler I, Lo RS, and Kopito RR. A conserved glutamate is responsible<br />

for ion selectivity and pH dependence of the mammalian anion exchangers<br />

AE1 and AE2. J Biol Chem 270: 28751–28758, 1995.<br />

183. Sly WS. The membrane carbonic anhydrases: from CO2 transport to<br />

tumor markers. EXS 90: 95–104, 2000.<br />

184. Southgate CD, Chishti AH, Mitchell B, Yi SJ, and Palek J. Targeted<br />

disruption of the murine erythroid band 3 gene results in spherocytosis<br />

and severe haemolytic anaemia despite a normal membrane skeleton. Nat<br />

Genet 14: 227–230, 1996.<br />

185. Sterling D, Alvarez BV, and Casey JR. The extracellular component of<br />

a transport metabolon. Extracellular loop 4 of the human AE1 Cl � /<br />

HCO 3 � exchanger binds carbonic anhydrase IV. J Biol Chem 277:<br />

25239–25246, 2002.<br />

186. Sterling D, Reithmeier RA, and Casey JR. A transport metabolon.<br />

Functional interaction of carbonic anhydrase II and chloride/bicarbonate<br />

exchangers. J Biol Chem 276: 47886–47894, 2001.<br />

187. Sterling D, Reithmeier RA, and Casey JR. Carbonic anhydrase: in the<br />

driver’s seat for bicarbonate transport. JOP 2: 165–170, 2001.<br />

188. Stewart AK, Chernova MN, Kunes YZ, and Alper SL. Regulation of<br />

AE2 anion exchanger by intracellular pH: critical regions of the NH2-<br />

Downloaded from<br />

http://ajprenal.physiology.org/<br />

by guest on December 23, 2012


Invited Review<br />

F598 � 2�<br />

<strong>SLC4</strong> BASE (<strong>H<strong>CO3</strong></strong> ,<strong>CO3</strong>) TRANSPORTERS<br />

terminal cytoplasmic domain. Am J Physiol Cell Physiol 281: C1344–<br />

C1354, 2001.<br />

189. Stewart AK, Chernova MN, Shmukler BE, Wilhelm S, and Alper SL.<br />

Regulation of AE2-mediated Cl � transport by intracellular or by extracellular<br />

pH requires highly conserved amino acid residues of AE2<br />

NH2-terminal cytoplasmic domain. J Gen Physiol 120: 707–722, 2002.<br />

190. Stewart AK, Kerr N, Chernova MN, Alper SL, and Vaughan-Jones<br />

RD. Acute pH-dependent regulation of AE2-mediated anion exchange<br />

involves discrete local surfaces of the NH2-terminal cytoplasmic domain.<br />

J Biol Chem 279: 52664–52676, 2004.<br />

191. Stuart-Tilley A, Sardet C, Pouyssegur J, Schwartz MA, Brown D,<br />

and Alper SL. Immunolocalization of anion exchanger AE2 and cation<br />

exchanger NHE1 in distinct, adjacent cells of gastric mucosa. Am J<br />

Physiol Cell Physiol 266: C559–C568, 1994.<br />

192. Sun XC and Bonanno JA. Identification and cloning of the Na/HCO 3 �<br />

cotransporter (NBC) in human corneal endothelium. Exp Eye Res 77:<br />

287–295, 2003.<br />

193. Sun XC, Bonanno JA, Jelamskii S, and Xie Q. Expression and<br />

localization of Na � -HCO 3 � cotransporter in bovine corneal endothelium.<br />

Am J Physiol Cell Physiol 279: C1648–C1655, 2000.<br />

194. Takano J, Noguchi K, Yasumori M, Kobayashi M, Gajdos Z, Miwa<br />

K, Hayashi H, Yoneyama T, and Fujiwara T. Arabidopsis boron<br />

transporter for xylem loading. Nature 420: 337–340, 2002.<br />

195. Tang XB, Kovacs M, Sterling D, and Casey JR. Identification of<br />

residues lining the translocation pore of human AE1, plasma membrane<br />

anion exchange protein. J Biol Chem 274: 3557–3564, 1999.<br />

196. Tanner MJ, Martin PG, and High S. The complete amino acid<br />

sequence of the human erythrocyte membrane anion-transport protein<br />

deduced from the cDNA sequence. Biochem J 256: 703–712, 1988.<br />

197. Tanphaichitr VS, Sumboonnanonda A, Ideguchi H, Shayakul C,<br />

Brugnara C, Takao M, Veerakul G, and Alper SL. Novel AE1<br />

mutations in recessive distal renal tubular acidosis. Loss-of-function is<br />

rescued by glycophorin A. J Clin Invest 102: 2173–2179, 1998.<br />

198. Tatishchev S, Abuladze N, Pushkin A, Newman D, Liu W, Weeks D,<br />

Sachs G, and Kurtz I. Identification of membrane topography of the<br />

electrogenic sodium bicarbonate cotransporter pNBC1 by in vitro transcription/translation.<br />

Biochemistry 42: 755–765, 2003.<br />

199. Thevenod F, Roussa E, Schmitt BM, and Romero MF. Cloning and<br />

immunolocalization of a rat pancreatic Na � bicarbonate cotransporter.<br />

Biochem Biophys Res Commun 264: 291–298, 1999.<br />

200. Tripp BC, Smith K, and Ferry JG. Carbonic anhydrase: new insights<br />

for an ancient enzyme. J Biol Chem 276: 48615–48618, 2001.<br />

201. Tsuganezawa H, Kobayashi K, Iyori M, Araki T, Koizumi A,<br />

Watanabe SI, Kaneko A, Fukao T, Monkawa T, Yoshida T, Kim<br />

DK, Kanai Y, Endou H, Hayashi M, and Saruta T. A new member<br />

of the HCO 3 � transporter superfamily is an apical anion exchanger of<br />

�-intercalated cells in the kidney. J Biol Chem 276: 8180–8189,<br />

2000.<br />

202. Usui T, Hara M, Satoh H, Moriyama N, Kagaya H, Amano S,<br />

Oshika T, Ishii Y, Ibaraki N, Hara C, Kunimi M, Noiri E,<br />

Tsukamoto K, Inatomi J, Kawakami H, Endou H, Igarashi T,<br />

Goto A, Fujita T, Araie M, and Seki G. Molecular basis of ocular<br />

abnormalities associated with proximal renal tubular acidosis. J Clin<br />

Invest 108: 107–115, 2001.<br />

203. Usui T, Seki G, Amano S, Oshika T, Miyata K, Kunimi M, Taniguchi<br />

S, Uwatoko S, Fujita T, and Araie M. Functional and molecular<br />

evidence for Na � -HCO 3 � cotransporter in human corneal endothelial<br />

cells. Pflügers Arch 438: 458–462, 1999.<br />

204. Vasuvattakul S, Yenchitsomanus PT, Vachuanichsanong P, Thuwajit<br />

P, Kaitwatcharachai C, Laosombat V, Malasit P, Wilairat P,<br />

and Nimmannit S. Autosomal recessive distal renal tubular acidosis<br />

associated with Southeast Asian ovalocytosis. Kidney Int 56: 1674–<br />

1682, 1999.<br />

205. Verlander JW, Madsen KM, Low PS, Allen DP, and Tisher CC.<br />

Immunocytochemical localization of band 3 protein in the rat collecting<br />

duct. Am J Physiol <strong>Renal</strong> Fluid Electrolyte Physiol 255: F115–F125, 1988.<br />

206. Vince JW, Carlsson U, and Reithmeier RA. Localization of the<br />

Cl � /HCO 3 � anion exchanger binding site to the amino-terminal region of<br />

carbonic anhydrase II. Biochemistry 39: 13344–13349, 2000.<br />

207. Vince JW and Reithmeier RAF. Carbonic anhydrase II binds to the<br />

carboxyl-terminus of human band 3, the erythrocyte Cl � /HCO 3 � exchanger.<br />

J Biol Chem 273: 28430–28437, 1998.<br />

AJP-<strong>Renal</strong> Physiol • VOL 290 • MARCH 2006 • www.ajprenal.org<br />

208. Vince JW and Reithmeier RAF. Identification of the carbonic anhydrase<br />

II binding site in the Cl � /HCO 3 � anion exchanger AE1. Biochemistry<br />

39: 5527–5533, 2000.<br />

209. Virkki LV, Choi I, Davis BA, and Boron WF. Cloning of a Na � -driven<br />

Cl/<strong>H<strong>CO3</strong></strong> exchanger from squid giant fiber lobe. Am J Physiol Cell<br />

Physiol 285: C771–C780, 2003.<br />

210. Virkki LV, Wilson DA, Vaughan-Jones RD, and Boron WF. Functional<br />

characterization of human NBC4 as an electrogenic Na � -HCO 3 �<br />

cotransporter (NBCe2). Am J Physiol Cell Physiol 282: C1278–C1289,<br />

2002.<br />

211. Vorum H, Kwon TH, Fulton C, Simonsen B, Choi I, Boron W,<br />

Maonsbach AB, Nielsen S, and Aalkajaer C. Immunolocalization of<br />

electroneutral Na-HCO 3 � cotransporter in rat kidney. Am J Physiol <strong>Renal</strong><br />

Physiol 279: F901–F909, 2000.<br />

212. Wang CZ, Conforti L, Petrovic S, Amlal H, Burnham CE, and<br />

Soleimani M. Mouse Na � :HCO 3 � cotransporter isoform NBC-3 (kNBC-<br />

3): cloning, tissue distribution, and functional characterization. Kidney<br />

Int 59: 1405–1414, 2001.<br />

213. Wang CZ, Yano H, Nagashima K, and Seino S. The Na � -driven<br />

Cl � /HCO 3 � exchanger. Cloning, tissue distribution, and functional characterization.<br />

J Biol Chem 275: 35486–35490, 2000.<br />

214. Wang DN, Kuhlbrandt W, Sarabia VE, and Reithmeier RA. Twodimensional<br />

structure of the membrane domain of human band 3, the<br />

anion transport protein of the erythrocyte membrane. EMBO J 12:<br />

2233–2239, 1993.<br />

215. Wang DN, Sarabia VE, Reithmeier RA, and Kuhlbrandt W. Threedimensional<br />

map of the dimeric membrane domain of the human erythrocyte<br />

anion exchanger, band 3. EMBO J 13: 3230–3235, 1994.<br />

216. Wang L, Groves JD, Mawby WJ, and Tanner MJ. Complementation<br />

studies with co-expressed fragments of the human red cell anion transporter<br />

(band 3; AE1). The role of some exofacial loops in anion transport.<br />

J Biol Chem 272: 10631–10638, 1997.<br />

217. Wang Z, Schultheis PJ, and Shull GE. Three N-terminal variants of<br />

AE2 Cl � /HCO 3 � exchanger are encoded by mRNAs transcribed from<br />

alternative promoters. J Biol Chem 271: 7835–7843, 1996.<br />

218. Weber S, Soergel M, Jeck N, and Konrad M. Atypical distal renal<br />

tubular acidosis confirmed by mutation analysis. Pediatr Nephrol 15:<br />

201–204, 2000.<br />

219. Wieth JO, Bjerrum PJ, and Borders CL Jr. Irreversible inactivation of<br />

red cell chloride exchange with phenylglyoxal, and arginine-specific<br />

reagent. J Gen Physiol 79: 283–312, 1982.<br />

220. Wright EM. <strong>Renal</strong> Na � -glucose <strong>transporters</strong>. Am J Physiol <strong>Renal</strong><br />

Physiol 280: F10–F18, 2001.<br />

221. Wrong O, Bruce LJ, Unwin RJ, Toye AM, and Tanner MJ. Band 3<br />

mutations, distal renal tubular acidosis, and Southeast Asian ovalocytosis.<br />

Kidney Int 62: 10–19, 2002.<br />

222. Xu J, Wang Z, Barone S, Petrovich M, Amlal H, Conforti L,<br />

Petrovich S, and Soleimani M. Expression of the Na � -HCO 3 � cotransporter<br />

NBC4 in rat kidney and characterization of a novel NBC4 variant.<br />

Am J Physiol <strong>Renal</strong> Physiol 284: F41–F50, 2003.<br />

223. Yamada H, Yamazaki S, Moriyama N, Hara C, Horita S, Enomoto<br />

Y, Kudo A, Kawakami H, Tanaka Y, Fujita T, and Seki G. Localization<br />

of NBC-1 variants in human kidney and renal cell carcinoma.<br />

Biochem Biophys Res Commun 310: 1213–1218, 2003.<br />

224. Yang Z, Alvarez BV, Chakarova C, Jiang L, Karan G, Frederick<br />

JM, Zhao Y, Sauve Y, Li X, Zrenner E, Wissinger B, Hollander AI,<br />

Katz B, Baehr W, Cremers FP, Casey JR, Bhattacharya SS, and<br />

Zhang K. Mutant carbonic anhydrase 4 impairs pH regulation and causes<br />

retinal photoreceptor degeneration. Hum Mol Genet 14: 255–265, 2005.<br />

225. Yannoukakos D, Stuart-Tilley A, Fernandez HA, Fey P, Duyk G, and<br />

Alper SL. Molecular cloning, expression, and chromosomal localization<br />

of two isoforms of the AE3 anion exchanger from human heart. Circ Res<br />

75: 603–614, 1994.<br />

226. Yenchitsomanus PT, Vasuvattakul S, Kirdpon S, Wasanawatana S,<br />

Susaengrat W, Sreethiphayawan S, Chuawatana D, Mingkum S,<br />

Sawasdee N, Thuwajit P, Wilairat P, Malasit P, and Nimmannit S.<br />

Autosomal recessive distal renal tubular acidosis caused by G701D<br />

mutation of anion exchanger 1 gene. Am J Kidney Dis 40: 21–29, 2002.<br />

227. Yip KP, Tsuruoka S, Schwartz GJ, and Kurtz I. Apical H � /<strong>base</strong><br />

<strong>transporters</strong> mediating bicarbonate absorption and pHi regulation in the<br />

OMCD. Am J Physiol <strong>Renal</strong> Physiol 283: F1098–F1104, 2002.<br />

Downloaded from<br />

http://ajprenal.physiology.org/<br />

by guest on December 23, 2012


228. Zaki L. Anion transport in red blood cells and arginine specific reagents.<br />

1. Effect of chloride and sulfate ions on phenylglyoxal sensitive sites in<br />

the red blood cell membrane. Biochem Biophys Res Commun 110:<br />

616–624, 1983.<br />

229. Zhang Y, Chernova MN, Stuart-Tilley AK, Jiang L, and Alper SL.<br />

The cytoplasmic and transmembrane domains of AE2 both contribute to<br />

regulation of anion exchange by pH. J Biol Chem 271: 5741–5749, 1996.<br />

230. Zhang D, Kiyatkin A, Bolin JT, and Low PS. Crystallographic structure<br />

and functional interpretation of the cytoplasmic domain of erythrocyte<br />

membrane band 3. Blood 96: 2925–2933, 2000.<br />

� 2�<br />

<strong>SLC4</strong> BASE (<strong>H<strong>CO3</strong></strong> ,<strong>CO3</strong>) TRANSPORTERS<br />

AJP-<strong>Renal</strong> Physiol • VOL 290 • MARCH 2006 • www.ajprenal.org<br />

Invited Review<br />

F599<br />

231. Zhu Q and Casey JR. The substrate anion selectivity filter in the human<br />

erythrocyte Cl � /HCO 3 � exchange protein, AE1. J Biol Chem 279:<br />

23565–23573, 2004.<br />

232. Zhu Q, Lee DW, and Casey JR. Novel topology in C-terminal region<br />

of the human plasma membrane anion exchanger, AE1. J Biol Chem 278:<br />

3112–3120, 2003.<br />

233. Zolotarev AS, Townsend RR, Stuart-Tilley A, and Alper SL. HCO 3 � -<br />

dependent conformational change in gastric parietal cell AE2, a glycoprotein<br />

naturally lacking sialic acid. Am J Physiol Gastrointest Liver<br />

Physiol 271: G311–G321, 1996.<br />

Downloaded from<br />

http://ajprenal.physiology.org/<br />

by guest on December 23, 2012

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!