Abstract
Soluble adenylyl cyclase (sAC) is a recently identified source of the ubiquitous second messenger cAMP. sAC is distinct from the more widely studied source of cAMP, the transmembrane adenylyl cyclases (tmACs); its activity is uniquely regulated by bicarbonate anions, and it is distributed throughout the cytoplasm and in cellular organelles. Due to its unique localization and regulation, sAC has various functions in a variety of physiological systems which are distinct from tmACs. In this review, we detail the known functions of sAC, and we reassess commonly held views of cAMP signaling inside cells.
Keywords: cyclic AMP, intracellular signal, Intracellular pH
Adenylyl cyclase (AC) is the effector molecule of one of the most widely utilized signal transduction pathways. Its product, cAMP, modulates cell growth and differentiation in organisms from bacteria to higher eukaryotes. In animals, a seemingly ubiquitous membrane-associated AC activity is encoded by a family of transmembrane adenylyl cyclases (tmACs) which mediate cellular responses to extracellular stimuli. In mammals, nine distinct tmAC genes differing in their patterns of expression and regulatory properties have thus far been identified. These tmACs are widely studied in a variety of laboratories.
A second type of AC activity in mammals was described in 1975 1. A soluble source of adenylyl cyclase activity was detected in testis and predicted to be molecularly distinct from tmACs 2,3 The activity was thought to be dependent upon manganese 1 and found to be insensitive to G protein 4 and forskolin 5 regulation. A biochemically related activity was detected in spermatozoa; however, it was loosely associated with membranes 1. This membrane associated, sAC-like activity was thought to be stimulated by sodium bicarbonate 6–9. However, the molecular nature, biochemical regulation, and physiological function of sAC remained unclear until the sAC protein was purified and cloned in 1999 10. The catalytic domains of sAC are related to bicarbonate-sensing adenylyl cyclases from cyanobacteria 10,11, suggesting conservation of function of these cyclases as bicarbonate sensors throughout evolution.
Genomic organization, structure and biochemistry
There is a single functional sAC gene in the human genome (ADCY10), comprising 33 exons which encompasses greater than 100 kb; however, it seems to utilize multiple promoters 12,13, and sAC mRNA undergoes extensive alternative splicing 12–15. Full-length mammalian sAC (sACfl) is comprised by two heterologous catalytic domains (C1 and C2) which constitute the 50 KDa amino terminus of the protein. The additional ~ 140 kDa C terminus of sACfl includes several putative regulatory domains such as an auto inhibitory region 16 and canonical P-loop and leucine zipper sequences 10. The minimal functional sAC variant, termed sACt, is a truncated form almost exclusively comprised of C1 and C2 10,15. While sACt cAMP-forming activity is several times stronger than it is for sACfl, both C1C2-containing sAC proteins are stimulated by HCO3− and are sensitive to all known selective sAC inhibitors (reviewed in 17).
sAC activity requires two divalent metal cations in the catalytic active site of the enzyme to coordinate binding and cyclizing of ATP. sAC is most active in the presence of Mn2+, a distinctive feature that led to its discovery in mammals 1, and which also applies to sea urchin 18, shark 19 and cyanobacterial sAC-like cyclases 11,20,21. However, it is not clear whether the physiological intracellular [Mn2+] concentration would support sAC activity in mammals or in other species. In vivo, Mg2+ sAC activity can be modulated by Ca2+ (which increases the affinity for ATP of mammalian sAC) and by HCO3− (which releases ATP-Mg2+ inhibition and increases Vmax of mammalian sAC)22 (reviewed in 17). The catalytic mechanism of cAMP production by sAC-like enzymes and its activation by HCO3− has been elucidated for CyaC, an adenylyl cyclase from the cyanobacterium Spirulina platensis 23. The key amino acid residues implicated in catalysis are conserved in cyanobacterial and mammalian sAC. In the structure- and kinetics-based model, ATP, with Ca2+-bound to its γ-phosphate, coordinates with specific residues in the sAC catalytic center. This results in an “open sAC state”. Then, the second divalent metal, a Mg2+ ion, binds to the α-phosphate of ATP, leading to a distinct set of catalytic residue interactions referred to as the “closed state”. This change, from the “open” to “closed” states, induces esterification of the alpha phosphate with the ribose in adenosine and the concomitant release of the β- and γphosphates (“cyclizing”). HCO3− stimulates the enzyme’s Vmax by fostering the allosteric change that leads to active site closure, recruitment of the catalytic Mg2+, and rearrangement of the phosphates in the bound ATP 23. A human sACt crystal structure reported in a patent application (WO 2007/010285) is consistent with the proposed catalytic mechanism and bicarbonate regulation.
The source of HCO3− regulating sAC could be external to the cell (body fluids or environment) or metabolically generated (Figure 1). Intra- and extracellular carbonic anhydrases (CAs) are in many cases essential for rapid hydration of CO2 into HCO3− that activates sAC (reviewed in 24). For example, in a number of epithelia and endothelia, elevated [HCO3−] stimulates ion and fluid transport, and in a subset of these examples, sAC has been shown to be involved (see below). But sAC was also shown to be modulated by metabolically generated CO2/HCO3− inside the matrix of the mitochondria 25–27 In diverse systems such as the endocrine pancreas, neuronal cells and neutrophils, sAC activity can be activated by elevations in free intracellular Ca2+.
Figure 1.
Activation of sAC by HCO3− and Ca2+. Cytosolic sAC can be activated by HCO3− derived from carbonic anhydrase (CA)-dependent hydration of (a) external and (b) metabolic CO2; and/or (c) HCO3− that enters via membrane transporting proteins (purple icon) such as anion exchangers, Na+/HCO3− cotransporters (NBC) or CFTRs. sAC can also be activated by (d) Ca2+ entering the cell via membrane transporters (green icon) such as voltage-dependent Ca2+ channels or potentially via Ca2+ release from the ER or mitochondria (not depicted). (e) HCO3− and Ca2+ can potentially activate sAC in the nucleus. (f) sAC inside mitochondria has been shown to be activated by metabolically generated CO2 via CA. See text for details.
Molecular studies predict the existence of sAC isoforms which contain only the C2 catalytic domain generated by alternative splicing and/or alternate promoter utilization 12–14. When heterologously expressed in insect Sf9 cells, some of these sAC variants localized to different regions of the cell, in a manner that suggested that the non-catalytic C-terminal domain favors association to the plasma membrane 28. However, it is not clear how these isoforms could generate cAMP with just one catalytic domain, and recombinant C2-only sAC isoforms had no detectable activity under the conditions tested 13,29.
Specific base substitutions in the human sAC gene have been linked to absorptive hypercalciuria (deficient renal and intestinal Ca2+ absorption) and to low spinal bone density 30. To date, the only reported phenotype of the existing sAC-knockout mouse model is male infertility due to an inability to activate flagellar movement upon ejaculation and failure to undergo a maturation process known as “capacitation” 31–33. However, results from other pH sensing proteins suggest additional phenotypes may be revealed under stressful conditions where sAC is required for sensing and compensation (e.g. 34). In addition, the existing sAC-knockout mouse only removes the exons encoding the C1 domain; it retains the C2 domain and the non-catalytic C-terminus, and it appears to include a putative alternative promoter 12. Therefore, it is possible that C2-containing proteins are responsible for essential functions or that they compensate for the lack of C1C2-containing sAC isoforms. Testing this hypothesis awaits generation of KO mice which specifically disrupt the C2 domain.
sAC orthologs have been functionally characterized in gills of the dogfish shark 19, and potential roles have been suggested for sAC in sea urchin sperm 18,35,36 and in the intestine of teleost (bony) fishes 37. The recent surge in genomic and transcriptomic information confirmed sAC orthologs to be present in most metazoan phylogenetic groups.
Microdomain organization of cAMP signaling
Dr. Earl Sutherland was awarded the Nobel Prize for identifying cAMP as the mediator of cellular control of metabolic activity 38. In the 50 years since he discovered this prototypical second messenger, cAMP signaling has been described in organisms as diverse as bacteria and mammals. However, in a seeming conundrum, cAMP has also been implicated in a wide variety of often-contradictory physiological processes, including different aspects of metabolism, proliferation, apoptosis, differentiation, migration, development, ion transport, pH regulation, and gene expression. Only recently has it become clear how this single second messenger could simultaneously mediate so many processes. In current models of cyclic nucleotide signal transduction, cAMP is locally generated within independently regulated microdomains (most recently reviewed in 39). This new microdomain model depends upon A-kinase anchoring proteins (AKAPs), which tether PKA to specific locations inside cells 40,41, and phosphodiesterases (PDEs), which degrade cAMP and act as barriers for cAMP diffusion 42–44 to avoid unregulated cross-communication between microdomains. The modern cAMP microdomain model also requires multiple sources of cAMP distributed at the cell membrane and throughout the cell (Figure 2).
Figure 2.
Intracellular cAMP signaling microdomains. cAMP signaling occurs in discrete intracellular compartments such as the membrane vicinity, focal cytoplasm points, mitochondria and the nucleus. Each microdomain contains (1) a source of cAMP (soluble adenylyl cyclase -sAC- or transmembrane adenylyl cyclase -tmAC-); (2) phosphodiesterases (PDE) that act as barriers for cAMP diffusion; and (3) cAMP targets such as protein kinase A (PKA) or Exchange proteins activated by cAMP (EPAC) (not illustrated). tmAC cAMP signaling occurs in response to various extracellular ligands and it requires modulation by G-Protein coupled receptors and heterotrimeric G-protein. The most widely described tmAC-dependent microdomain occurs at the cell membrane, but additional intracellular tmAC-dependent microdomains occur in endosomes after internalization. sAC present throughout the cytoplasm and in organelles such as mitochondria, nucleus, mid-bodies and centrioles define other microdomains. Additional regulation might involve the movement of sAC between compartments. See text for details.
A need for cAMP microdomains was first postulated in cardiomyocytes, when it was observed that distinct hormones elicited unique responses via cAMP in a single cell type 45. Subsequently, FRET-based and biophysical methods that enable measuring cAMP concentrations in situ revealed that cAMP levels are not uniform within cells (recently reviewed in 46,47). The existence of membrane-proximal cAMP microdomains was definitively demonstrated in neurons 48 and cardiomyocytes 49, and artificial, localized production of second messenger 50–52 supports the model that cAMP acts in locally restricted microdomains.
In most membrane microdomains, cAMP is likely generated by the classic transmembrane adenylyl cyclases (tmACs), which are regulated by heterotrimeric G proteins downstream from hormonally regulated G protein Coupled Receptors (GPCR). As their name implies, tmACs are obligatory transmembrane-proteins, and their mode of activation demands that they be localized on the plasma membrane in close proximity to the GPCR. Recently, it was found that tmACs in thyroid follicles can still signal as they co-sort with ligand-bound GPCRs on endosomes during receptor internalization 53,54 reviewed in 43,44.
Classically, research on cAMP signaling relied on the use of membrane permeant cAMP analogs, the potent, pan-tmAC activator, forskolin, and/or the broad specificity phosphodiesterase inhibitor, IBMX. In light of the cAMP microdomain model, results with each of these reagents needs re-evaluation because their effects do not reflect biologically meaningful second messenger responses. Cyclic nucleotide analogs have multiple targets and they competitively inhibit phosphodiesterases 55, which will disrupt the integrity of signaling microdomains. Forskolin stimulation will produce unphysiologically potent and prolonged cAMP changes. For example, in rat insulinoma INS-1 cells, forskolin induces a continuous increase in cAMP accumulation, reaching maximum values of ~10-fold after 30 min 56. This contrasts sharply with the effect of glucagon-like peptide-1 (GLP-1), a physiological activator, that induces a ~2.5-fold activation after 5–10 minutes, but which returns to basal levels after 30 min 57. Similarly, the use of IBMX will eliminate the inter-microdomain PDE barriers permitting cAMP diffusion. Use of forskolin and IBMX is more akin to the toxic effects of cAMP elevating toxins, such as the Edema Toxin of anthrax or cholera or pertussis toxins. Inhibition of PDEs with IBMX includes an additional complication because it prevents degradation of both cAMP and cGMP. Therefore, any effect obtained with IBMX could be due to increased activity of cAMP or cGMP effectors from any microdomain, or even to inhibition of cross-regulation between cAMP and cGMP pathways 58.
Mammalian sAC is distributed to discrete locations throughout the cell. It is found diffusely distributed in the cytoplasm and at the centrioles and mid-body 59, and inside the nucleus 59–62 and mitochondria 26,27,59,63. Each of these locations contain known targets for cAMP; thus sAC represents an additional source of cAMP inside cells that can produce the second messenger locally and activate nearby cAMP effectors 64–66. The nuclear and mitochondrial sAC-defined cAMP microdomains have been functionally characterized. Nuclear sAC is capable of phosphorylating CREB 61,67, and this microdomain might be related to the gene regulation in solid tumors cells observed with alkaline cytoplasmic pH (reviewed in 68). Mitochondrial sAC modulates oxidative phosphorylation in response to CO2/HCO3− generated by the TCA cycle 26,27, and cytoplasmic sAC translocates to the mitochondria during acidosis/ischemia to promote the mitochondrial apoptotic pathway 63
sAC roles throughout the body
Testis and sperm
The most widely accepted role of sAC in mammals is in male fertility 32,33. During spermatogenesis, sAC mRNA is first detectable in mid-pachytene spermatocytes and shows a strong upregulation in the later stages of spermiogenesis 69. sACfl protein is only detectable by immunohistochemistry in late pachytene spermatocytes (end of meiotic prophase I) 70. However, because the antibodies utilized were against the carboxy-terminus of sACfl, it is possible that other sAC variants are present and play a role at earlier stages. Sperm continue maturation during transit through the epididymis, and they are stored in the cauda epididymis, where [HCO3−] is significantly lower than in plasma and seminal fluids 71. The diminished luminal bicarbonate concentration in the epididymis is maintained by sAC regulation of V-type H+-ATPases (VHAs) in a process described in greater detail below. Although morphologically mature, these stored epididymal sperm still do not have the “capacity” to fertilize an egg 72. Upon ejaculation, the stored sperm are mixed with seminal and prostatic fluids where [HCO3−] suddenly rises to ~25 mM.
In mature sperm, sAC is the sole producer of cAMP in response to elevations in [HCO3−] 32,33, which varies dramatically in the environments sperm encounter during the reproductive process 71. The ejaculated sperm acquire fertilization-competence during transit through the female reproductive tract. This critical process involves many changes that are collectively grouped under a single term, “capacitation”. Among the first definable events in capacitation is the entrance of Ca2+ and HCO3− into sperm, which activate cAMP production by sAC. In the short term, this activates asymmetrical flagellar beat frequency (FBF) resulting in vigorous forward sperm motility. Extended activation of protein kinase A (PKA) in the presence of cholesterol acceptors leads to the prototypical pattern of tyrosine phosphorylation that represents a molecularly defined hallmark of capacitation 73–78. Physiologically, the end results of capacitation are sperm hyperactivation of motility and the ability to perforate the egg’s zona pellucida via the acrosome reaction.
The existing sAC knockout mouse model, which deletes the exons encoding the C1 domain (sAC-C1 KO), removes the two characterized C1C2 containing isoforms, sACt and sACfl 33. The phenotype of sAC C1 KO sperm includes defects in flagellum movement resulting in lack of motility 31–33, an aberrant tyrosine phosphorylation pattern during capacitating conditions 33, and the inability to fertilize an egg 32,33. Loss of these C1C2 isoforms also resulted in a morphological aberration, termed flagellar angulation (tail bending) 32,33, which may be the result of diminished metabolic capacity 31.
In addition to responding to HCO3−, sAC is essential for the acceleration of FBF in response to adenosine analogs and catecholamine agonists 79. The catecholamine stimulatory mechanism is not clear, but it does not appear to be mediated via bicarbonate as it occurs when HCO3− is omitted from the medium. Both agonists and antagonists of β-adrenergic receptors stimulated FBF and there was no discrimination between l-(−) and d-(+) catecholamine isomers, arguing against a conventional (i.e., G protein coupled) adrenergic receptor. Modulation by adenosine also seems independent from GPCRs and tmACs 79,80, and instead may rely on the entrance of adenosine via specific transporters to promote sAC-generated cAMP accumulation via unknown direct or indirect mechanisms 80.
Somatic functions of sAC
Initially, sAC’s role in sperm biology was thought to be its only function. This was largely due to the relatively low expression of sAC mRNA and protein in other mammalian tissues 12,69,81, to the absence of any other overt phenotype besides male infertility in the sAC C1 KO mice 32 and to the absence of sAC in the sequenced genomes of fruit fly (Drosophila melanogaster) and roundworm (Caenorhabditis elegans) 82. However, more sensitive mRNA and protein analytical techniques and the explosion in genome sequencing during the last decade, reveal that sAC is widely expressed in animals and is found in virtually every animal Phyla. Thus, the putative existence of sAC C2 isoforms coupled with the likely need to stress the system to unmask additional sAC functions, presumably explain the lack of additional obvious sAC C1 KO phenotypes.
Kidney
The presence of sAC in the kidney was hinted in an early study that described HCO3−-stimulated cAMP forming activity in rat kidney (medulla > cortex) homogenates 83. Subsequent to its molecular isolation 10, sAC mRNA was detected in kidney by RNA Array 13 and RT-PCR 12,13,84, and sAC protein was identified in kidneys by Western blotting 11–13 and immunohistochemistry 85,86. Taken together, these studies suggested the presence of several sAC splice variants in kidney. The immunohistochemistry revealed sAC (or at least a subset of the sAC variants) are preferentially expressed in cells of the medullary and cortical thick ascending loop of Henle (TAL), in cells of the distal tubule (DT) and in cells of the collecting duct (CD) 85,86.
Renal corpuscle and proximal convoluted tubule
Immunohistochemical studies using the monoclonal antibody R21 (directed against an epitope in coding sequence Exon 5) did not detect sAC either in the renal corpuscle or in the proximal convoluted tubule (PCT) 85,86. However, unpublished results from our lab using polyclonal antibodies against the C terminus of sAC revealed strong immunostaining in the PCT, and positive immunostaining in some glomeruli (L.R.L. and J.B., unpublished observations). Studies of sAC function in the PCT await confirmation of its presence in these regions of the nephron, for example, by laser capture micro-dissection followed by transcriptomic or proteomic studies. Potential roles for sAC in the PCT include interaction with other pH/CO2/HCO3− responsive enzymes proposed to regulate salt and fluid absorption in the PCT, such as Pyk2 or an as yet undefined tyrosine kinase 87–89. pH/CO2/HCO3− sensing throughout the body, including kidney, was recently reviewed elsewhere 24.
Thick ascending loop of Henle
The thick ascending loop of Henle (TAL) actively absorbs NaCl (and to a much lesser extent, Ca2+ and Mg2+), and it is responsible for urine concentration during antidiuresis and for urine dilution during diuresis (reviewed in 90). The bulk of NaCl absorption across cells of the TAL takes place via apical Na+/Cl−/2Cl− cotransporters (NKCC) 91–94, energized by basolateral Na+/K+-ATPases (recently reviewed in 95). sAC is present in the TAL 85,86, in both medullary and cortical segments 85, and even though there are, as yet, no functional studies of sAC in native TAL, there is evidence supporting its ability to regulate both Na+/K+-ATPase (see below for description of sAC regulation in mpkCCDc14 cells, immortalized cells derived from the mouse cortical collecting duct) and NKCCs.
The intestine of marine teleost fish absorbs NaCl via cellular mechanisms similar to those in the TAL, and it has traditionally been used as a model for NaCl absorption 96–100. In the toadfish intestine, it was recently shown that HCO3− stimulates NaCl absorption (estimated from short-circuit current measurements), seemingly via sAC-dependent regulation of NKCC2 and/or Na+/K+-ATPase 37. Several lines of evidence suggest sAC may play a similar role in the TAL; (1) mild metabolic alkalosis reduces the diuretic, natriuretic and chloruretic effects of bumetanide by ~40%, ~21% and ~25%, respectively 101, suggesting a stimulatory effect of HCO3− on NaCl and water absorption; and (2) NKCC2 has been demonstrated to be phosphorylated and shuttled into the apical membrane of the TAL in response to vasopressin 102 and cAMP 103,104, in a PKA-dependent manner 104.
Distal tubule
In this review, we use “Distal tubule” (DT) to collectively refer to the distal convoluted tubule, connecting tubule and cortical collecting duct 105. Overall, the DT reabsorbs ~10 % of the filtered NaCl, and it is also important for Mg2+ and Ca2+ reabsorption and for K+ secretion and reabsorption 105. In addition, A- and B-type intercalated cells (ICs) are responsible for metabolic compensation of systemic (“blood”) acid/base (A/B) status 106.
The pioneering research about sAC as a sensor and regulator of A/B-related ion transport was performed in the epididymis, which shares its embryonic origin with the renal nephron 107. In addition, both epithelia have acid-secreting cells that are functionally similar (termed “clear cells” in the epididymis and A-type ICs in the nephron DT). The epididymis is often used for functional studies on acid secretion as a surrogate model for the DT because it is easier to isolate and perfuse compared to the nephron, and it is simpler due to the absence of any cell type corresponding to the countering B-type ICs.
In order to maintain sperm quiescence, the lumen of the epididymis has low pH and low [HCO3−] 71. This is achieved by H+ pumping by apical V-type H+-ATPases (VHA) 108. Elevations in luminal pH or [HCO3−] are transmitted to the inside of the clear cells either by Na+/HCO3− cotransporters (NBCs) or by hydration/dehydration into CO2, catalyzed by extracellular and intracellular carbonic anhydrases (CA) 85,109. Intracellular HCO3− activates sAC, which promotes the insertion of VHAs into the apical membrane and the development of extensive apical microvili, leading to increased apical H+ secretion into the epididymis lumen 85. Thus, sAC in clear cells senses elevations in luminal pH and [HCO3−] and restores, via VHA translocation and H+ pumping, the original low pH and [HCO3−] luminal values. Based on the effects of pharmacological inhibitors, the immediate downstream target of sAC-generated cAMP appears to be protein kinase A (PKA), while exchange protein directly activated by cAMP (Epac) does not seem to play a role on the VHA apical translocation 110. Downstream of PKA, the A subunit of the VHA is a potential phosphorylation target 111. The stimulatory effect of sAC/cAMP/PKA on VHA apical accumulation is counterbalanced by an inhibitory effect of AMP-activated kinase (AMPK) 111 (Figure 3). This is not the only time where the effects of PKA and AMPK are antagonistic 111–114, and if the two kinases are localized within the same signaling microdomain, cAMP and its degradation product, AMP, could function as a timing mechanism. AMPK is regulated by an increase of AMP at the expense of an ATP; in a two-step reaction catalyzed by any ATPase and adenylate kinase, one ATP is converted into one AMP, and this change stimulates AMPK activity (reviewed in 115). The same change in the AMP/ATP ratio is effected, also in two steps, in cAMP signaling cascades. In the first step, an ATP is converted into cAMP via an adenylyl cyclase, and in the second step, cAMP is hydrolyzed into AMP by a phosphodiesterase (PDE). Thus, PKA and AMPK regulation of VHA translocation may function as a timer. Activation by cAMP and PKA could be automatically terminated by AMPK subsequent to PDE hydrolysis of the cAMP into AMP. This “clock” may function in other systems where PKA and AMPK are antagonistic. AMPK is a known modulator of metabolic pathways, and we have shown that sAC generated cAMP modulates metabolic activity in pancreatic beta cells 57 and astrocytes 116 and that it functions as a metabolic sensor inside mammalian mitochondria 26,27.
Figure 3.
Regulation of V-type H+-ATPase translocation by sAC and AMPK. (1) Extracellular HCO3− enters the cell via transporter proteins (purple icon) or is dehydrated into CO2, a reaction catalyzed by extracellular carbonic anhydrase (CAIV). CO2 would then diffuse into the cell, where it is hydrated into H+ and HCO3− by intracellular carbonic anhydrase (CAII). (2) The elevated intracellular [HCO3−] activates sAC (3) to produce cAMP, which promotes (via PKA) (4) the insertion of VHA-containing vesicles into the cell membrane. (5) Membrane inserted VHAs secrete H+, which counteract the original alkalosis. (6) cAMP is hydrolyzed by phosphodiesterase (PDE) into AMP, which can (7) via stimulation of AMPK, inhibit the PKA mediated effects. This hypothetical mechanism, involving sequential activation of PKA and AMPK, could serve as a self-regulating circuit.
The role of sAC was subsequently examined in the renal DT. sAC is present in A- and B-type ICs 84, as well as in principal cells 84,86. Immunofluorescent and immunogold staining demonstrates sAC is most abundantly present in the apical pole of A-type ICs, and in the basolateral and apical poles of B-type ICs 84. sAC and VHA not only co-localize in both types of ICs, but they also co-immunoprecipitate from rat kidney homogenates 84. Functional studies in renal A-type ICs basically mimicked those from the clear cells from the epididymis; i.e., apical VHA accumulation and microvilli elongation dependent on cAMP and PKA 113,117, possibly involving direct phosphorylation of the VHA A subunit 118. Importantly, equivalent results were obtained in kidney slices 113, after intravenous cAMP infusion via the femoral vein 117, and in isolated ICs 117. Although direct effects of increased external pH and/or [HCO3−] on these processes have not yet been demonstrated, chronic CA inhibition with acetazolamide, which increases HCO3− delivery to the DT 119,120, did stimulate the apical microvilli elongation (as well as the number of A-type ICs) 119. Interestingly, as in epididymal clear cells, AMPK activity opposes the role played by sAC-generated cAMP; it inhibits the VHA apical translocation and the development of microvilli in A-type ICs 113.
The physiological role of sAC in renal B-type ICs is less clear, especially since intravenous cAMP infusion had no clear effect on VHA intracellular localization 117. However, sAC has been shown to play a regulatory role on VHA translocation to the basolateral membrane of cells involved in a base-secreting physiological process in a non-mammalian system 19. In aquatic animals, the gills (and not the kidneys) are the principal A/B regulatory organs 121, and the cellular and molecular mechanisms governing H+ and HCO3− transport are remarkably similar. In the dogfish shark gill epithelium, base-secreting cells normally have VHA in cytoplasmic vesicles, where they are presumably inactive 122. Upon blood alkalosis, VHA containing vesicles exhibit sAC- 19, CA- 123, and microtubule-dependent 124 VHA translocation to the basolateral membrane, where VHA absorbs H+ into the blood and energizes HCO3− secretion. Pharmacological inhibition of sAC prevents VHA translocation both in vitro and in vivo 19, suggesting dogfish sAC (dfsAC) is both necessary and sufficient as a sensor and regulator of systemic blood A/B homeostasis. Elucidating this mechanism was possible because of the ultrastructure of the basolateral membrane of shark base-secreting cells (which is heavily infolded but lacks an elaborated tubovesicular system 122,125) and because of the feeding physiology of the dogfish shark, which involves a pronounced post-feeding blood alkalosis 123,126.
Another, yet unexplored, potential target of sAC activity in B-type ICs is the anion exchanger Pendrin, which colocalizes with sAC at the apical region 84. Interestingly, dogfish pendrin also seems to be present in shark gill base-secreting cells 121.
The potential role of sAC in principal cells of the DT has been studied in confluent polarized mouse cortical collecting duct (mpkCCDc14) cells. These cells, which express ENaC and Na+/K+-ATPase in their apical and basolateral membranes, respectively, and exhibit hormonal regulation of Na+ transport similar to in vivo models, are considered most similar to DT principal cells 127,128. Pharmacological and sAC siRNA manipulations revealed a potential role of sAC in regulating transepithelial Na+ transport in these cultured cells, both in basal and in forskolin- or aldosterone-stimulated conditions 86.
Finally, specific base substitutions in the sAC gene correlate with familial absorptive hypercalcyuria 30, a genetically inherited disease characterized by an excess of Ca2+ in urine due to inadequate reabsorption in the distal tubule and/or in the intestine 129. However, a defined role for sAC in Ca2+ absorption has not yet been investigated in cells or organs.
Conclusions kidney
The A/B status of plasma and filtrated fluid affect several ion-transporting processes in the nephron, several of which are regulated by cAMP. Because sAC is present throughout the nephron, sAC is a good candidate to integrate external (tubular fluid) and internal (plasma, renal interstitium) cues with their appropriate responses via cAMP signaling. Plus, confirmed or proposed roles for sAC in other tissues raises the possibility that sAC regulation might be involved in additional processes including Ca2+ signaling and gene expression.
Eye
In both corneal endothelium 130 and cilliary body 131, HCO3− stimulates fluid secretion. A role for sAC was first suggested by the observation that HCO3− stimulates cAMP production in homogenates from both tissues 83. Subsequent to its molecular isolation, sAC was confirmed to be present in primary cultures of bovine corneal endothelial cells, and sAC activation increased CFTR dependent secretion of Cl−, HCO3− and/or ATP 132. Although these studies were performed prior to the advent of sAC-selective inhibitors, all data suggest that cAMP produced by sAC stimulates PKA phosphorylation of apical CFTR, thus increasing apical Cl− permeability 132,133. It was also demonstrated in cultured corneal endothelial cells that higher [HCO3−] in the cultured medium increased sAC expression 133.
A role for sAC has also been proposed in retinal ganglion cells (RGCs). In a subset of RGCs, periodic depolarizations, acting via a Ca2+ dependent cAMP/PKA cascade, are critical for proper circuit development 134. Mice lacking the Ca2+-sensitive tmACs (AC1 and AC8) still displayed depolarization-induced Ca2+-dependent PKA transients, which were only inhibited after pharmacological inhibition of all mammalian adenylyl cyclases, including sAC 135.
Airways
Calu-3 cells, a cancer cell line derived from bronchial submucosal glands 136 express sACfl and sACt mRNA, although only a ~50 KDa band (consistent with sACt) is detectable at the protein level 137. cAMP production in Calu-3 cells is stimulated by HCO3−, an activity which is inhibited by a selective sAC inhibitor (2-CE) but not by a tmAC-selective, P site inhibitor 137,138 (see Table 1 for a summary of adenylyl cyclase inhibitors). In cell-attached patch clamp experiments, HCO3− stimulated CFTR single-channel activity in a 2-CE sensitive manner, suggesting sAC regulates CFTR in Calu-3 cells 137. In addition, switching from HCO3−-free to HCO3−-containing buffer significantly increases CFTR mRNA and protein levels in a sAC-dependent manner 138. These conditions also increased phosphorylation of nuclear CREB. In vivo, regulation of CFTR by sAC in airway glands is probably related to the secretion of airway surface liquid and mucus.
TABLE 1.
Pharmacological inhibitors of adenylyl cyclases
Airway epithelial cells express multiple sAC mRNA and protein variants, including putative sAC-C2 only isoforms 14. One particular sAC variant (of ~50 KDa) was demonstrated by a combination of Western Blotting and immunocytochemistry to be present in cilia. Apical application of HCO3− to culture-differentiated human airway epithelial cells increased cAMP production via sAC; this sAC-generated cAMP stimulated PKA which increased ciliary beat frequency 14. Thus, sAC seems poised to sense changes in CO2/HCO3− concentration in airways during normal and disease conditions and coordinate the clearance of mucus. For example, this form of sAC-dependent regulation appears to be adversely affected in airway epithelia from cystic fibrosis patients, possibly contributing to their mucociliary dysfunction 139. Finally, stimulation of ciliary beat frequency by ethanol also depends on PKA and sAC 140, although the link between ethanol and sAC activation is not yet clear.
Because HCO3− stimulation of ciliary beat frequency is diminished in cultured cells from cystic fibrosis patients 139, HCO3− probably enters ciliated cells via CFTR. Therefore, the relationship between sAC and CFTR in airway epithelial cells is complex and it may involve sAC regulation of CFTR expression and activity (as shown in Calu-3 cells) as well as CFTR regulation of sAC (as shown in ciliated cells).
Pancreas
There are proposed roles for sAC in both exocrine and endocrine pancreas. In intrahepatic trees, sAC is preferentially expressed in cholangiocytes of large bile ducts 141, which are specialized bile-secreting, secretin-responsive, epithelial cells. Pharmacological inhibition (by KH7) or siRNA downregulation of sAC significantly abolishes the HCO3−-induced stimulation of fluid secretion 141. Similar inhibitions were found upon application of acetazolamide and H89, suggesting that CA and PKA are upstream and downstream, respectively, of sAC. These authors proposed that sAC sustains basal levels of cAMP and fluid secretion during the interdigestive phase, while gastrointestinal hormones and cholinergic and β-adrenergic agonists acting via GPCRs-tmACs mediate regulated phases of cAMP and fluid secretion. Because Cl− secretion by cholangiocytes of large ducts depends on apical CFTR 142, this might represent another case of CFTR regulation by sAC.
β-cells in the endocrine pancreas release insulin in response to various stimuli including hormones, neurotransmitters and blood glucose levels. It was known for decades that an increase in external glucose concentration stimulates cAMP production while modulating the release of insulin 143. However, the source of this cAMP remained unknown 144 until recently. Based on studies on INS-1E cells using selective inhibitors and siRNA, sAC is responsible for the glucose-induced cAMP production, while tmAC(s) mediates responses to incretins such as GLP-1 57. Glucose induced activation of sAC is dependent upon entry of Ca2+ into the cell 57, which may be synergistic with glucose metabolism-dependent elevations in intracellular [HCO3−] and/or ATP. sAC-generated cAMP is essential for the increased activation of ERK1 and 2 observed during high glucose conditions; the contribution of sAC-generated cAMP on insulin release has not yet been reported.
Digestive tract
Transient and sustained Cl− and K+ secretions in the distal colonic epithelium, which determine the rate of fluid secretion in relation to food digestion, water conservation and intestinal flushing, are subject to sympathetic modulation. Addition of epinephrine to isolated colonic mucosa induces rapid and transient Cl− secretion, which is followed by sustained K+ secretion. Based on differential responses with inhibitors selective for sAC (KH7) or for tmACs (ddAdo), the rapid response is dependent on β2-adrenergic receptors and tmACs, while the sustained response likely relies on HCO3−, β1- and β2-adrenergic receptors and sAC 145.
In marine bony fish, intestinal HCO3− secretion and NaCl and water absorption are essential for hypo-osmoregulation 146,147. sAC has been proposed to coordinate HCO3− secretion with NaCl absorption by monitoring intracellular levels of carbonic anhydrase-generated HCO3− and activating membrane ion-transporting proteins 37. As explained above, a similar mechanism might be occurring in the mammalian TAL.
Brain and nervous system
sAC-dependent processes have been hinted at or established in choroid plexus, neurons and astrocytes. In choroid plexus, CO2 metabolism has long been linked to cerebrospinal fluid (CSF) secretion 148. sAC mRNA 149, protein 11 and activity 83 have been demonstrated in chorioid plexus, and it is straightforward to hypothesize that bicarbonate regulation of sAC plays a role in CSF homeostasis.
Astrocytes express several sAC splice variants 116, (some or all of) which are involved in a novel mechanism of metabolic coupling between neurons and astrocytes. Elevation of [K+] at the extracellular space caused by neuronal activity depolarizes the cell membrane of nearby astrocytes and induces HCO3− entry via electrogenic Na+/HCO3− cotransporters (NBC). The elevation in [HCO3−]i activates sAC, which leads to glycogen breakdown, enhanced glycolysis and generation and release of lactate for use by the neighboring ‘active’ neurons for energy.
sAC has been shown to be present in developing neurons, where, depending on the origin of the neuron, it was located in cell bodies, dendrites, axons and/or growth cones 150. One proposed role for sAC in developing neurons is to regulate growth cones and promote axonal growth. In cultured dorsal root ganglion and spinal commissural neurons, sAC inhibition, either by small molecule inhibitors selective for sAC (KH7 or catechol estrogens) or sAC specific RNAi, blocked netrin-1 induced growth cone elaboration and axonal growth 150. Both responses were mimicked by sAC overexpression. In an apparent conundrum, the existing sAC knockout mouse model (sAC-C1 KO) does not display any of the phenotypic defects in the ventral spinal commissure which are a hallmark of netrin-1 deficiency 81,150.
sAC was also shown to be essential for responses to the prototypical neurotrophin, nerve growth factor (NGF). PC12 cells, which are derived from rat adrenal medulla, are used as a model for neuronal differentiation because they develop neuron-like characteristics when treated with NGF or with pituitary adenylyl cyclase-activating peptide (PACAP). Both NGF and PACAP stimulate axon generation via stimulation of the small G protein Rap1. It had long been established that PACAP stimulates Rap1 and axonegenesis via cAMP generated by the GPCR-G protein-tmAC pathway 151, but the NGF-stimulating mechanism, and whether it involved cAMP, remained unclear 152. The confirmation that PC12 cells express sAC and the observation that sAC inhibition by small molecules or siRNA blocked NGF induced activation of Rap1 suggested that sAC is also involved in axon growth in response to NGF 153. Similar to the studies demonstrating calcium involvement in glucose induced sAC activation in pancreatic beta cells, NGF stimulation of sAC in PC12 cells is dependent upon calcium.
Immune cells
Calcium regulation of sAC and signaling via the effector Rap1 were also found to play a role in the inflammatory response in neutrophils. sAC mRNA was found to be abundant in human leukocytes 13, and neutrophils represent the most abundant type of white blood cell. Immunostaining and Western blotting of highly purified neutrophils confirmed sAC protein presence 154. sAC in neutrophils was essential for tumor necrosis factor (TNF)-induced release of H2O2 (respiratory outburst). And, as seen in PC12 cells in response to NGF, sAC activation was shown to be dependent upon elevated intracellular Ca2+, and the proximal target of sAC generated cAMP was Rap1.
Bone
Several sAC splice variants are present in osteoclasts and osteoblasts 13, and mutations in the human sAC gene correlate with low spinal bone density 30. Calcification by osteoblasts is intrinsically connected to HCO3− and Ca2+, and the acid/base status greatly influences mineralization 155–157. Although these elements suggest a key role for sAC in bone biology, to date the only reported role of sAC in bone is in osteoclastogenesis 28. Differentiation of RAW264.7 cells into osteoclasts (estimated from TRAP staining and activity) is maximum in the absence of HCO3− in the medium, and it is sharply inhibited in the presence of 12 or 24 mM HCO3−. Experiments using the sAC inhibitor 2CE or siRNA suggested sAC is important for inhibiting osteoclast differentiation in high external [HCO3−], although interpretation of results is somewhat confused by similar inhibitory effects during (non-physiological) HCO3−-free conditions. Bone density in cultured mouse calvaria was similarly promoted by high [HCO3−] in a 2CE-sensitive manner 28, indicating HCO3−-sensing sAC is a physiologically relevant regulator of bone formation and/or reabsorption.
sAC expression in other tissues, of as yet unknown functions
sAC mRNA and/or protein has been reported in almost every other tissue 13,69. However, apart from the systems described above, in most cases the role of sAC has not been elucidated yet. Some interesting cases include placenta 13,30,158, carotid body 159(potential roles reviewed in 24) and embryos 69,160. Other organs and tissues where sAC mRNA has been detected include liver, muscle, thymus, spleen 13,69 and ovary 69.
Conclusions
sAC is the most recently identified source of cAMP inside animal cells and it is directly modulated by HCO3− and Ca2+. This allows sAC to act as a sensor of the external and intracellular A/B status, as well as a sensor of metabolically generated HCO3− from CO2. In addition, sAC can be secondarily modulated by hormones that lead to increase intracellular [Ca2] or [HCO3−]. Because cAMP is a ubiquitous intracellular signaling messenger, the potential physiological effects subjected to sAC modulation are multiple and they include protein directly sensitive to cAMP like cyclic nucleotide-gated ion channels, exchange proteins activated by cAMP (EPAC), as well as proteins sensitive to EPAC signaling and PKA-phosphorylation.
List of Abbreviations
- sAC
soluble adenylyl cyclase
- tmAC
transmembrane adenylyl cyclase
- AC
adenylyl cyclase
- sACfl
full-length isoform of sAC
- sACt
truncated isoform of sAC
- C1
first catalytic domain of sAC
- C2
second catalytic domain of sAC
- KO
knockout
- PDE
phosphodiesterase
- FRET
Förster (or fluorescence) resonance energy transfer
- GPCR
G protein coupled receptor
- IBMX
3-isobutyl-1-methylxanthine (broad specificity PDE inhibitor)
- GLP-1
glucagon like peptide
- CREB
cAMP response element binding protein
- TCA cycle
tricarboxylic acid (or Krebs) cycle
- VHA
V-type H+ ATPase
- FBF
flagellar beat frequency
- PKA
Protein Kinase A (cAMP-dependent protein kinase)
- DAL
thick ascending loop of Henle
- DT
distal tubule
- CD
collecting duct
- PCT
proximal convoluted tubule
- NKCC
Na+/Cl−/2Cl− cotransporter
- IC
intercalated cells
- NBC
Na+/HCO3− cotransporters
- CA
carbonic anhydrase
- AMPK
AMP-activated kinase
- A/B
acid/base
- dfsAC
dogfish sAC
- CFTR
Cystic fibrosis transmembrane conductance regulator
- RGC
retinal ganglion cells
- CSF
cerebrospinal fluid
- NGF
Nerve Growth Factor
- PACAP
pituitary adenylyl cyclase activating peptide
- TNF
tumor necrosis factor
- TRAP
tartrate-resistant acid phosphatase
Footnotes
DISCLOSURE STATEMENT:
The authors declare they have no conflicts of interest.
References
- 1.Braun T, Dods RF. Development of a Mn-2+-sensitive, “soluble” adenylate cyclase in rat testis. Proc Natl Acad Sci U S A. 1975;72:1097–1101. doi: 10.1073/pnas.72.3.1097. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 2.Braun T. Purification of soluble form of adenylyl cyclase from testes. Methods Enzymol. 1991;195:130–136. doi: 10.1016/0076-6879(91)95160-l. [DOI] [PubMed] [Google Scholar]
- 3.Neer EJ. Physical and functional properties of adenylate cyclase from mature rat testis. J Biol Chem. 1978;253:5808–5812. [PubMed] [Google Scholar]
- 4.Braun T, Frank H, Dods R, Sepsenwol S. Mn2+-sensitive, soluble adenylate cyclase in rat testis. Differentiation from other testicular nucleotide cyclases. Biochim Biophys Acta. 1977;481:227–235. doi: 10.1016/0005-2744(77)90155-3. [DOI] [PubMed] [Google Scholar]
- 5.Forte LR, Bylund DB, Zahler WL. Forskolin does not activate sperm adenylate cyclase. Mol Pharmacol. 1983;24:42–47. [PubMed] [Google Scholar]
- 6.Garbers DL, Tubb DJ, Hyne RV. A requirement of bicarbonate for Ca2+-induced elevations of cyclic AMP in guinea pig spermatozoa. J Biol Chem. 1982;257:8980–8984. [PubMed] [Google Scholar]
- 7.Garty NB, Salomon Y. Stimulation of partially purified adenylate cyclase from bull sperm by bicarbonate. FEBS Lett. 1987;218:148–152. doi: 10.1016/0014-5793(87)81036-0. [DOI] [PubMed] [Google Scholar]
- 8.Okamura N, Tajima Y, Soejima A, Masuda H, Sugita Y. Sodium bicarbonate in seminal plasma stimulates the motility of mammalian spermatozoa through direct activation of adenylate cyclase. J Biol Chem. 1985;260:9699–9705. [PubMed] [Google Scholar]
- 9.Visconti PE, Muschietti JP, Flawia MM, Tezon JG. Bicarbonate dependence of cAMP accumulation induced by phorbol esters in hamster spermatozoa. Biochim Biophys Acta. 1990;1054:231–236. doi: 10.1016/0167-4889(90)90246-a. [DOI] [PubMed] [Google Scholar]
- 10.Buck J, Sinclair ML, Schapal L, Cann MJ, Levin LR. Cytosolic adenylyl cyclase defines a unique signaling molecule in mammals. Proc Natl Acad Sci U S A. 1999;96:79–84. doi: 10.1073/pnas.96.1.79. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 11.Chen Y, et al. Soluble adenylyl cyclase as an evolutionarily conserved bicarbonate sensor. Science. 2000;289:625–628. doi: 10.1126/science.289.5479.625. [DOI] [PubMed] [Google Scholar]
- 12.Farrell J, et al. Somatic “soluble” adenylyl cyclase isoforms are unaffected in Sacytm1Lex/Sacytm1Lex “knockout” mice. PLoS ONE. 2008;3:e3251. doi: 10.1371/journal.pone.0003251. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 13.Geng W, et al. Cloning and characterization of the human soluble adenylyl cyclase. Am J Physiol Cell Physiol. 2005;288:C1305–1316. doi: 10.1152/ajpcell.00584.2004. [DOI] [PubMed] [Google Scholar]
- 14.Schmid A, et al. Soluble adenylyl cyclase is localized to cilia and contributes to ciliary beat frequency regulation via production of cAMP. J Gen Physiol. 2007;130:99–109. doi: 10.1085/jgp.200709784. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 15.Jaiswal BS, Conti M. Identification and functional analysis of splice variants of the germ cell soluble adenylyl cyclase. J Biol Chem. 2001;276:31698–31708. doi: 10.1074/jbc.M011698200. [DOI] [PubMed] [Google Scholar]
- 16.Chaloupka JA, Bullock SA, Iourgenko V, Levin LR, Buck J. Autoinhibitory regulation of soluble adenylyl cyclase. Mol Reprod Dev. 2006;73:361–368. doi: 10.1002/mrd.20409. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 17.Kamenetsky M, et al. Molecular details of cAMP generation in Mammalian cells: a tale of two systems. J Mol Biol. 2006;362:623–639. doi: 10.1016/j.jmb.2006.07.045. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 18.Nomura M, Beltran C, Darszon A, Vacquier VD. A soluble adenylyl cyclase from sea urchin spermatozoa. Gene. 2005;353:231–238. doi: 10.1016/j.gene.2005.04.034. [DOI] [PubMed] [Google Scholar]
- 19.Tresguerres M, et al. Bicarbonate-sensing soluble adenylyl cyclase is an essential sensor for acid/base homeostasis. Proc Natl Acad Sci USA. 2010;107:442–447. doi: 10.1073/pnas.0911790107. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 20.Cann MJ, Hammer A, Zhou J, Kanacher T. A defined subset of adenylyl cyclases is regulated by bicarbonate ion. J Biol Chem. 2003;278:35033–35038. doi: 10.1074/jbc.M303025200. [DOI] [PubMed] [Google Scholar]
- 21.Kasahara M, Yashiro K, Sakamoto T, Ohmori M. The Spirulina platensis adenylate cyclase gene, cyaC, encodes a novel signal transduction protein. Plant Cell Physiol. 1997;38:828–836. doi: 10.1093/oxfordjournals.pcp.a029241. [DOI] [PubMed] [Google Scholar]
- 22.Litvin TN, Kamenetsky M, Zarifyan A, Buck J, Levin LR. Kinetic properties of “soluble” adenylyl cyclase. Synergism between calcium and bicarbonate. J Biol Chem. 2003;278:15922–15926. doi: 10.1074/jbc.M212475200. [DOI] [PubMed] [Google Scholar]
- 23.Steegborn C, Litvin TN, Levin LR, Buck J, Wu H. Bicarbonate activation of adenylyl cyclase via promotion of catalytic active site closure and metal recruitment. Nat Struct Mol Biol. 2005;12:32–37. doi: 10.1038/nsmb880. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 24.Tresguerres M, Buck J, Levin L. Physiological carbon dioxide, bicarbonate, and pH sensing. Pflügers Archiv European Journal of Physiology. 2010;460:953–964. doi: 10.1007/s00424-010-0865-6. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 25.Acin-Perez R, et al. Control of oxidative phosphorylation by vitamin A illuminates a fundamental role in mitochondrial energy homoeostasis. FASEB J. 2010;24:627–636. doi: 10.1096/fj.09-142281. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 26.Acin-Perez R, et al. Modulation of mitochondrial protein phosphorylation by soluble adenylyl cyclase ameliorates cytochrome oxidase defects. Embo Molecular Medicine. 2009;1:392–406. doi: 10.1002/emmm.200900046. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 27.Acin-Perez R, et al. Cyclic AMP Produced inside Mitochondria Regulates Oxidative Phosphorylation. Cell Metabolism. 2009;9:265–276. doi: 10.1016/j.cmet.2009.01.012. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 28.Geng W, et al. Inhibition of osteoclast formation and function by bicarbonate: role of soluble adenylyl cyclase. J Cell Physiol. 2009;220:332–340. doi: 10.1002/jcp.21767. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 29.Farrell J. PhD thesis. Weill Cornell Medical College; 2007. The molecular identity of soluble adenylyl cyclase. [Google Scholar]
- 30.Reed BY, et al. Identification and characterization of a gene with base substitutions associated with the absorptive hypercalciuria phenotype and low spinal bone density. J Clin Endocrinol Metab. 2002;87:1476–1485. doi: 10.1210/jcem.87.4.8300. [DOI] [PubMed] [Google Scholar]
- 31.Xie F, et al. Soluble adenylyl cyclase (sAC) is indispensable for sperm function and fertilization. Dev Biol. 2006;296:353–362. doi: 10.1016/j.ydbio.2006.05.038. [DOI] [PubMed] [Google Scholar]
- 32.Esposito G, et al. Mice deficient for soluble adenylyl cyclase are infertile because of a severe sperm-motility defect. Proc Natl Acad Sci U S A. 2004;101:2993–2998. doi: 10.1073/pnas.0400050101. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 33.Hess KC, et al. The “soluble” adenylyl cyclase in sperm mediates multiple signaling events required for fertilization. Dev Cell. 2005;9:249–259. doi: 10.1016/j.devcel.2005.06.007. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 34.Finberg KE, et al. The B1-subunit of the H+ ATPase is required for maximal urinary acidification. Proc Natl Acad Sci U S A. 2005;102:13616–13621. doi: 10.1073/pnas.0506769102. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 35.Beltran C, et al. Particulate and soluble adenylyl cyclases participate in the sperm acrosome reaction. Biochem Biophys Res Commun. 2007;358:1128–1135. doi: 10.1016/j.bbrc.2007.05.061. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 36.Nomura M, Vacquier VD. Proteins associated with soluble adenylyl cyclase in sea urchin sperm flagella. Cell Motil Cytoskeleton. 2006;63:582–590. doi: 10.1002/cm.20147. [DOI] [PubMed] [Google Scholar]
- 37.Tresguerres M, Levin LR, Buck J, Grosell M. Modulation of NaCl absorption by [HCO3−] in the marine teleost intestine is mediated by soluble adenylyl cyclase. Am J Physiol Regul Integr Comp Physiol. 2010;299:62–71. doi: 10.1152/ajpregu.00761.2009. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 38.Robison GA, Butcher RW, Sutherland EW. Cyclic AMP. Annu Rev Biochem. 1968;37:149–174. doi: 10.1146/annurev.bi.37.070168.001053. [DOI] [PubMed] [Google Scholar]
- 39.Zaccolo M. cAMP signal transduction in the heart: understanding spatial control for the development of novel therapeutic strategies. Br J Pharmacol. 2009;158:50–60. doi: 10.1111/j.1476-5381.2009.00185.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 40.Beene DL, Scott JD. A-kinase anchoring proteins take shape. Curr Opin Cell Biol. 2007;19:192–198. doi: 10.1016/j.ceb.2007.02.011. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 41.Carnegie GK, Means CK, Scott JD. A-kinase anchoring proteins: from protein complexes to physiology and disease. IUBMB Life. 2009;61:394–406. doi: 10.1002/iub.168. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 42.Houslay MD. Underpinning compartmentalised cAMP signalling through targeted cAMP breakdown. Trends Biochem Sci. 2010;35:91–100. doi: 10.1016/j.tibs.2009.09.007. [DOI] [PubMed] [Google Scholar]
- 43.Calebiro D, Nikolaev VO, Lohse MJ. Imaging of persistent cAMP signaling by internalized G protein-coupled receptors. J Mol Endocrinol. 2010;45:1–8. doi: 10.1677/JME-10-0014. [DOI] [PubMed] [Google Scholar]
- 44.Calebiro D, Nikolaev VO, Persani L, Lohse MJ. Signaling by internalized G-protein-coupled receptors. Trends Pharmacol Sci. 2010;31:221–228. doi: 10.1016/j.tips.2010.02.002. [DOI] [PubMed] [Google Scholar]
- 45.Buxton IL, Brunton LL. Compartments of cyclic AMP and protein kinase in mammalian cardiomyocytes. J Biol Chem. 1983;258:10233–10239. [PubMed] [Google Scholar]
- 46.Berrera M, et al. In: Protein-protein interactions as new drug targets. Klussmann E, Scott J, editors. XV. Springer-Verlag; 2008. pp. 286–298. [Google Scholar]
- 47.Willoughby D, Cooper DMF. Live-cell imaging of cAMP dynamics. Nat Meth. 2008;5:29–36. doi: 10.1038/nmeth1135. [DOI] [PubMed] [Google Scholar]
- 48.Davare MA, et al. A beta2 adrenergic receptor signaling complex assembled with the Ca2+ channel Cav1.2. Science. 2001;293:98–101. doi: 10.1126/science.293.5527.98. [DOI] [PubMed] [Google Scholar]
- 49.Marx SO, et al. Requirement of a macromolecular signaling complex for beta adrenergic receptor modulation of the KCNQ1-KCNE1 potassium channel. Science. 2002;295:496–499. doi: 10.1126/science.1066843. [DOI] [PubMed] [Google Scholar]
- 50.Sayner S, Stevens T. Soluble adenylate cyclase reveals the significance of compartmentalized cAMP on endothelial cell barrier function. Biochem Soc Trans. 2006;34:492–494. doi: 10.1042/BST0340492. [DOI] [PubMed] [Google Scholar]
- 51.Sayner SL, Alexeyev M, Dessauer CW, Stevens T. Soluble adenylyl cyclase reveals the significance of cAMP compartmentation on pulmonary microvascular endothelial cell barrier. Circ Res. 2006;98:675–681. doi: 10.1161/01.RES.0000209516.84815.3e. [DOI] [PubMed] [Google Scholar]
- 52.Creighton J, Zhu B, Alexeyev M, Stevens T. Spectrin-anchored phosphodiesterase 4D4 restricts cAMP from disrupting microtubules and inducing endothelial cell gap formation. J Cell Sci. 2008;121:110–119. doi: 10.1242/jcs.011692. [DOI] [PubMed] [Google Scholar]
- 53.Calebiro D, et al. Persistent cAMP-signals triggered by internalized G-protein-coupled rReceptors. PLoS Biol. 2009;7:e1000172. doi: 10.1371/journal.pbio.1000172. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 54.Ferrandon S, et al. Sustained cyclic AMP production by parathyroid hormone receptor endocytosis. Nat Chem Biol. 2009;5:734–742. doi: 10.1038/nchembio.206. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 55.Poppe H, et al. Cyclic nucleotide analogs as probes of signaling pathways. Nat Meth. 2008;5:277–278. doi: 10.1038/nmeth0408-277. [DOI] [PubMed] [Google Scholar]
- 56.Ding WQ, et al. Forskolin suppresses insulin gene transcription in islet [beta]-cells through a protein kinase A-independent pathway. Cell Signal. 2003;15:27–35. doi: 10.1016/s0898-6568(02)00051-7. [DOI] [PubMed] [Google Scholar]
- 57.Ramos LS, Zippin JH, Kamenetsky M, Buck J, Levin LR. Glucose and GLP-1 stimulate cAMP production via distinct adenylyl cyclases in INS-1E insulinoma cells. J Gen Physiol. 2008;132:329–338. doi: 10.1085/jgp.200810044. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 58.Shelly M, et al. Local and long-range reciprocal regulation of cAMP and cGMP in axon/dendrite formation. Science. 2010;327:547–552. doi: 10.1126/science.1179735. [DOI] [PubMed] [Google Scholar]
- 59.Zippin JH, et al. Compartmentalization of bicarbonate-sensitive adenylyl cyclase in distinct signaling microdomains. FASEB J. 2003;17:82–84. doi: 10.1096/fj.02-0598fje. [DOI] [PubMed] [Google Scholar]
- 60.Zippin JH, Chadwick PA, Levin LR, Buck J, Magro CM. Soluble Adenylyl Cyclase Defines a Nuclear cAMP Microdomain in Keratinocyte Hyperproliferative Skin Diseases. J Invest Dermatol. 2010;130:1279–1287. doi: 10.1038/jid.2009.440. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 61.Zippin JH, et al. Bicarbonate-responsive “soluble” adenylyl cyclase defines a nuclear cAMP microdomain. J Cell Biol. 2004;164:527–534. doi: 10.1083/jcb.200311119. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 62.Feng Q, et al. Two domains are critical for the nuclear localization of soluble adenylyl cyclase. Biochimie. 2006;88:319–328. doi: 10.1016/j.biochi.2005.09.003. [DOI] [PubMed] [Google Scholar]
- 63.Kumar S, Kostin S, Flacke JP, Reusch HP, Ladilov Y. Soluble adenylyl cyclase controls mitochondria-dependent apoptosis in coronary endothelial cells. J Biol Chem. 2009 doi: 10.1074/jbc.M900925200. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 64.Bundey RA, Insel PA. Discrete intracellular signaling domains of soluble adenylyl cyclase: camps of cAMP? Sci STKE. 2004;2004:pe19. doi: 10.1126/stke.2312004pe19. [DOI] [PubMed] [Google Scholar]
- 65.Wuttke MS, Buck J, Levin LR. Bicarbonate-regulated soluble adenylyl cyclase. Jop. 2001;2:154–158. [PubMed] [Google Scholar]
- 66.Zippin JH, Levin LR, Buck J. CO2/HCO3−-responsive soluble adenylyl cyclase as a putative metabolic sensor. Trends Endocrinol Metab. 2001;12:366–370. doi: 10.1016/s1043-2760(01)00454-4. [DOI] [PubMed] [Google Scholar]
- 67.Mardones P, Medina JF, Elferink RPJO. Activation of cyclic AMP signaling in Ae2-deficient mouse fibroblasts. J Biol Chem. 2008;283:12146–12153. doi: 10.1074/jbc.M710590200. [DOI] [PubMed] [Google Scholar]
- 68.Parks SK, Chiche J, Pouysségur J. pH control mechanisms of tumor survival and growth. J Cell Physiol. 2011;226:299–308. doi: 10.1002/jcp.22400. [DOI] [PubMed] [Google Scholar]
- 69.Sinclair ML, et al. Specific expression of soluble adenylyl cyclase in male germ cells. Mol Reprod Dev. 2000;56:6–11. doi: 10.1002/(SICI)1098-2795(200005)56:1<6::AID-MRD2>3.0.CO;2-M. [DOI] [PubMed] [Google Scholar]
- 70.Xie F, Conti M. Expression of the soluble adenylyl cyclase during rat spermatogenesis: evidence for cytoplasmic sites of cAMP production in germ cells. Dev Biol. 2004;265:196–206. doi: 10.1016/j.ydbio.2003.09.020. [DOI] [PubMed] [Google Scholar]
- 71.Levine N, Marsh DJ. Micropuncture studies of electrochemical aspects of fluid and electrolyte transport in individual seminiferous tubules, epididymis and vas deferens in rats. J Physiol. 1971;213:557. doi: 10.1113/jphysiol.1971.sp009400. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 72.Yanagimachi R. In: The Physiology of Reproduction. Knobil E, Neill JD, editors. Raven Press, Ltd; 1994. pp. 189–317. [Google Scholar]
- 73.Osheroff JE, et al. Regulation of human sperm capacitation by a cholesterol efflux-stimulated signal transduction pathway leading to protein kinase A-mediated up-regulation of protein tyrosine phosphorylation. Mol Hum Reprod. 1999;5:1017–1026. doi: 10.1093/molehr/5.11.1017. [DOI] [PubMed] [Google Scholar]
- 74.Visconti PE. Understanding the molecular basis of sperm capacitation through kinase design. Proc Natl Acad Sci U S A. 2009;106:667–668. doi: 10.1073/pnas.0811895106. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 75.Visconti PE, et al. Capacitation of mouse spermatozoa. I. Correlation between the capacitation state and protein tyrosine phosphorylation. Development. 1995;121:1129–1137. doi: 10.1242/dev.121.4.1129. [DOI] [PubMed] [Google Scholar]
- 76.Visconti PE, et al. The molecular basis of sperm capacitation. J Androl. 1998;19:242–248. [PubMed] [Google Scholar]
- 77.Visconti PE, et al. Capacitation of mouse spermatozoa. II. Protein tyrosine phosphorylation and capacitation are regulated by a cAMP-dependent pathway. Development. 1995;121:1139–1150. doi: 10.1242/dev.121.4.1139. [DOI] [PubMed] [Google Scholar]
- 78.Visconti PE, et al. Roles of bicarbonate, cAMP, and protein tyrosine phosphorylation on capacitation and the spontaneous acrosome reaction of hamster sperm. Biol Reprod. 1999;61:76–84. doi: 10.1095/biolreprod61.1.76. [DOI] [PubMed] [Google Scholar]
- 79.Schuh SM, et al. Signaling pathways for modulation of mouse sperm motility by adenosine and catecholamine agonists. Biol Reprod. 2006;74:492–500. doi: 10.1095/biolreprod.105.047837. [DOI] [PubMed] [Google Scholar]
- 80.Schuh SM, Hille B, Babcock DF. Adenosine and catecholamine agonists speed the flagellar beat of mammalian sperm by a non-receptor-mediated mechanism. Biol Reprod. 2007;77:960–969. doi: 10.1095/biolreprod.107.062562. [DOI] [PubMed] [Google Scholar]
- 81.Moore SW, et al. Soluble adenylyl cyclase is not required for axon guidance to netrin-1. J Neurosci. 2008;28:3920–3924. doi: 10.1523/JNEUROSCI.0547-08.2008. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 82.Roelofs J, Van Haastert PJ. Deducing the origin of soluble adenylyl cyclase, a gene lost in multiple lineages. Mol Biol Evol. 2002;19:2239–2246. doi: 10.1093/oxfordjournals.molbev.a004047. [DOI] [PubMed] [Google Scholar]
- 83.Mittag TW, Guo WB, Kobayashi K. Bicarbonate-activated adenylyl cyclase in fluid-transporting tissues. Am J Physiol. 1993;264:F1060–1064. doi: 10.1152/ajprenal.1993.264.6.F1060. [DOI] [PubMed] [Google Scholar]
- 84.Paunescu TG, et al. Association of soluble adenylyl cyclase with the V-ATPase in renal epithelial cells. Am J Physiol Renal Physiol. 2008;294:F130–138. doi: 10.1152/ajprenal.00406.2007. [DOI] [PubMed] [Google Scholar]
- 85.Pastor-Soler N, et al. Bicarbonate-regulated adenylyl cyclase (sAC) is a sensor that regulates pH-dependent V-ATPase recycling. J Biol Chem. 2003;278:49523–49529. doi: 10.1074/jbc.M309543200. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 86.Hallows KR, et al. Regulation of epithelial Na+ transport by soluble adenylyl cyclase in kidney collecting duct cells. J Biol Chem. 2009;284:5774–5783. doi: 10.1074/jbc.M805501200. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 87.Li SY, Sato S, Yang XJ, Preisig PA, Alpern RJ. Pyk2 activation is integral to acid stimulation of sodium/hydrogen exchanger 3. J Clin Invest. 2004;114:1782–1789. doi: 10.1172/JCI18046. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 88.Zhou Y, Bouyer P, Boron WF. Role of a tyrosine kinase in the CO2-induced stimulation of HCO3- reabsorption by rabbit S2 proximal tubules. Am J Physiol Renal Physiol. 2006;291:F358–367. doi: 10.1152/ajprenal.00520.2005. [DOI] [PubMed] [Google Scholar]
- 89.Zhou Y, Zhao J, Bouyer P, Boron WF. Evidence from renal proximal tubules that HCO3− and solute reabsorption are acutely regulated not by pH but by basolateral HCO3− and CO2. Proc Natl Acad Sci U S A. 2005;102:3875–3880. doi: 10.1073/pnas.0500423102. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 90.Greger R. Ion transport mechanisms in thick ascending limb of Henle’s loop of mammalian nephron. Physiol Rev. 1985;65:760–797. doi: 10.1152/physrev.1985.65.3.760. [DOI] [PubMed] [Google Scholar]
- 91.Ecelbarger CA, et al. Localization and regulation of the rat renal Na+-K+-2Cl− cotransporter, BSC-1. Am J Physiol Renal Physiol. 1996;271:F619–628. doi: 10.1152/ajprenal.1996.271.3.F619. [DOI] [PubMed] [Google Scholar]
- 92.Kaplan MR, et al. Apical localization of the Na-K-Cl cotransporter, rBSC1, on rat thick ascending limbs. Kidney Int. 1996;49:40–47. doi: 10.1038/ki.1996.6. [DOI] [PubMed] [Google Scholar]
- 93.Lytle C, Xu JC, Biemesderfer D, Forbush B., 3rd Distribution and diversity of Na-K-Cl cotransport proteins: a study with monoclonal antibodies. Am J Physiol Cell Physiol. 1995;269:C1496–1505. doi: 10.1152/ajpcell.1995.269.6.C1496. [DOI] [PubMed] [Google Scholar]
- 94.Xu JC, et al. Molecular cloning and functional expression of the bumetanide-sensitive Na-K-Cl cotransporter. Proc Natl Acad Sci U S A. 1994;91:2201–2205. doi: 10.1073/pnas.91.6.2201. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 95.Fenton RA, Knepper MA. Mouse models and the urinary concentrating mechanism in the new millennium. Physiol Rev. 2007;87:1083–1112. doi: 10.1152/physrev.00053.2006. [DOI] [PubMed] [Google Scholar]
- 96.Musch MW, et al. Na+ -K+-Cl− co-transport in the intestine of a marine teleost. Nature. 1982;300:351–353. doi: 10.1038/300351a0. [DOI] [PubMed] [Google Scholar]
- 97.Frizzell RA, Halm DR, Musch MW, Stewart CP, Field M. Potassium transport by flounder intestinal mucosa. Am J Physiol Renal Physiol. 1984;246:F946–951. doi: 10.1152/ajprenal.1984.246.6.F946. [DOI] [PubMed] [Google Scholar]
- 98.O’Grady SM, Palfrey HC, Field M. Na–K–2Cl cotransport in winter flounder intestine and bovine kidney outer medulla: [3H] bumetanide binding and effects of furosemide analogues. J Memb Biol. 1987;96:11–18. doi: 10.1007/BF01869330. [DOI] [PubMed] [Google Scholar]
- 99.Musch MW, O’Grady SM, Field M, Sidney F, Becca F. Methods Enzymol. Vol. 192. Academic Press; 1990. pp. 746–753. [DOI] [PubMed] [Google Scholar]
- 100.Field M, Karnaky KJ, Smith PL, Bolton JE, Kinter WB. Ion transport across the isolated intestinal mucosa of the winter flounder, Pseudopleuronectes americanus. J Memb Biol. 1978;41:265–293. doi: 10.1007/BF01870433. [DOI] [PubMed] [Google Scholar]
- 101.Loon NR, Wilcox CS. Mild metabolic alkalosis impairs the natriuretic response to bumetanide in normal human subjects. Clin Sci. 1998;94:287–292. doi: 10.1042/cs0940287. [DOI] [PubMed] [Google Scholar]
- 102.Gimenez I, Forbush B. Short-term stimulation of the renal Na-K-Cl cotransporter (NKCC2) by vasopressin involves phosphorylation and membrane translocation of the protein. J Biol Chem. 2003;278:26946–26951. doi: 10.1074/jbc.M303435200. [DOI] [PubMed] [Google Scholar]
- 103.Ortiz PA. cAMP increases surface expression of NKCC2 in rat thick ascending limbs: role of VAMP. Am J Physiol Renal Physiol. 2006;290:F608–616. doi: 10.1152/ajprenal.00248.2005. [DOI] [PubMed] [Google Scholar]
- 104.Caceres PS, Ares GR, Ortiz PA. cAMP stimulates apical exocytosis of the renal Na+-K+-2Cl− cotransporter NKCC2 in the thick ascending limb. J Biol Chem. 2009;284:24965–24971. doi: 10.1074/jbc.M109.037135. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 105.Reilly RF, Ellison DH. Mammalian distal tubule: physiology, pathophysiology, and molecular anatomy. Physiol Rev. 2000;80:277–313. doi: 10.1152/physrev.2000.80.1.277. [DOI] [PubMed] [Google Scholar]
- 106.Wagner CA, et al. Renal vacuolar H+-ATPase. Physiol Rev. 2004;84:1263–1314. doi: 10.1152/physrev.00045.2003. [DOI] [PubMed] [Google Scholar]
- 107.Hinton B, Turner T. Is the epididymis a kidney analogue? News Physiol Sci. 1988;3:28–31. [Google Scholar]
- 108.Breton S, Smith PJS, Lui B, Brown D. Acidification of the male reproductive tract by a proton pumping H+-ATPase. Nat Med. 1996;2:470–472. doi: 10.1038/nm0496-470. [DOI] [PubMed] [Google Scholar]
- 109.Pastor-Soler N, Pietrement C, Breton S. Role of acid/base transporters in the male reproductive tract and potential consequences of their malfunction. Physiology. 2005;20:417–428. doi: 10.1152/physiol.00036.2005. [DOI] [PubMed] [Google Scholar]
- 110.Pastor-Soler NM, et al. Alkaline pH- and cAMP-induced V-ATPase membrane accumulation is mediated by protein kinase A in epididymal clear cells. Am J Physiol Cell Physiol. 2008;294:C488–C494. doi: 10.1152/ajpcell.00537.2007. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 111.Hallows KR, et al. AMP-activated protein kinase inhibits alkaline pH- and PKA-induced apical vacuolar H+-ATPase accumulation in epididymal clear cells. American Journal of Physiology-Cell Physiology. 2009;296:C672–C681. doi: 10.1152/ajpcell.00004.2009. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 112.Hurley RL, et al. Regulation of AMP-activated protein kinase by multisite phosphorylation in response to agents that elevate cellular cAMP. J Biol Chem. 2006;281:36662–36672. doi: 10.1074/jbc.M606676200. [DOI] [PubMed] [Google Scholar]
- 113.Gong F, et al. Vacuolar H+-ATPase apical accumulation in kidney intercalated cells is regulated by PKA and AMP-activated protein kinase. American Journal of Physiology-Renal Physiology. 2010;298:F1162–F1169. doi: 10.1152/ajprenal.00645.2009. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 114.Djouder N, et al. PKA phosphorylates and inactivates AMPK[alpha] to promote efficient lipolysis. EMBO J. 2010;29:469–481. doi: 10.1038/emboj.2009.339. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 115.Kemp BE, et al. AMP-activated protein kinase, super metabolic regulator. Biochem Soc Trans. 2003;31:162–168. doi: 10.1042/bst0310162. [DOI] [PubMed] [Google Scholar]
- 116.Choi HB, et al. Metabolic communication between astrocytes and neurons via bicarbonate responsive soluble adenylyl cyclase. Nat Neurosci. doi: 10.1016/j.neuron.2012.08.032. (in revision) [DOI] [PMC free article] [PubMed] [Google Scholar]
- 117.Paunescu TG, et al. cAMP stimulates apical V-ATPase accumulation, microvillar elongation, and proton extrusion in kidney collecting duct A-intercalated cells. American Journal of Physiology-Renal Physiology. 2010;298:F643–F654. doi: 10.1152/ajprenal.00584.2009. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 118.Alzamora R, et al. PKA Regulates Vacuolar H+-ATPase Localization and Activity via Direct Phosphorylation of the A Subunit in Kidney Cells. J Biol Chem. 2010;285:24676–24685. doi: 10.1074/jbc.M110.106278. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 119.Bagnis C, Marshansky V, Breton S, Brown D. Remodeling the cellular profile of collecting ducts by chronic carbonic anhydrase inhibition. Am J Physiol Renal Physiol. 2001;280:F437–448. doi: 10.1152/ajprenal.2001.280.3.F437. [DOI] [PubMed] [Google Scholar]
- 120.Chou SY, et al. Effects of acetazolamide on proximal tubule Cl, Na, and HCO3 transport in normal and acidotic dogs during distal blockade. J Clin Invest. 1977;60:162–170. doi: 10.1172/JCI108752. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 121.Evans DH, Piermarini PM, Choe KP. The multifunctional fish gill: dominant site of gas exchange, osmoregulation, acid-base regulation, and excretion of nitrogenous waste. Physiol Rev. 2005;85:97–177. doi: 10.1152/physrev.00050.2003. [DOI] [PubMed] [Google Scholar]
- 122.Tresguerres M, Katoh F, Fenton H, Jasinska E, Goss GG. Regulation of branchial V-H+-ATPase Na+/K+-ATPase and NHE2 in response to acid and base infusions in the Pacific spiny dogfish (Squalus acanthias) J Exp Biol. 2005;208:345–354. doi: 10.1242/jeb.01382. [DOI] [PubMed] [Google Scholar]
- 123.Tresguerres M, Parks SK, Wood CM, Goss GG. V-H+-ATPase translocation during blood alkalosis in dogfish gills: interaction with carbonic anhydrase and involvement in the postfeeding alkaline tide. Am J Physiol Reg Int Comp Physiol. 2007;292:R2012–R2019. doi: 10.1152/ajpregu.00814.2006. [DOI] [PubMed] [Google Scholar]
- 124.Tresguerres M, Parks SK, Katoh F, Goss GG. Microtubule-dependent relocation of branchial V-H+-ATPase to the basolateral membrane in the Pacific spiny dogfish (Squalus acanthias): a role in base secretion. J Exp Biol. 2006;209:599–609. doi: 10.1242/jeb.02059. [DOI] [PubMed] [Google Scholar]
- 125.Laurent P. In: Fish Physiology. Hoar WS, Randall DW, editors. 10A. Academic Press; 1984. pp. 73–183. [Google Scholar]
- 126.Wood CM, Kajimura M, Mommsen TP, Walsh PJ. Alkaline tide and nitrogen conservation after feeding in an elasmobranch (Squalus acanthias) J Exp Biol. 2005;208:2693–2705. doi: 10.1242/jeb.01678. [DOI] [PubMed] [Google Scholar]
- 127.Bens M, et al. Corticosteroid-dependent sodium transport in a novel immortalized mouse collecting duct principal cell line. J Am Soc Nephrol. 1999;10:923–934. doi: 10.1681/ASN.V105923. [DOI] [PubMed] [Google Scholar]
- 128.Vandewalle A. Immortalized renal proximal and collecting duct cell lines derived from transgenic mice harboring L-type pyruvate kinase promoters as tools for pharmacological and toxicological studies. Cell Biol Toxicol. 2002;18:321–328. doi: 10.1023/a:1019584014243. [DOI] [PubMed] [Google Scholar]
- 129.Sutton RAL, Wong NLM, Dirks JH. Effects of metabolic acidosis and alkalosis on sodium and calcium transport in the dog kidney. Kidney Int. 1979;15:520–533. doi: 10.1038/ki.1979.67. [DOI] [PubMed] [Google Scholar]
- 130.Kuang KY, Xu M, Koniarek JP, Fischbarg J. Effects of ambient bicarbonate, phosphate and carbonic anhydrase inhibitors on fluid transport across rabbit corneal endothelium. Exp Eye Res. 1990;50:487–493. doi: 10.1016/0014-4835(90)90037-u. [DOI] [PubMed] [Google Scholar]
- 131.Kishida K, Sasabe T, Iizuka S, Manabe R, Otori T. Sodium and chloride transport across the isolated rabbit ciliary body. Curr Eye Res. 1982;2:149–152. doi: 10.3109/02713688208997688. [DOI] [PubMed] [Google Scholar]
- 132.Sun XC, et al. HCO3−-dependent soluble adenylyl cyclase activates cystic fibrosis transmembrane conductance regulator in corneal endothelium. Am J Physiol Cell Physiol. 2003;284:C1114–1122. doi: 10.1152/ajpcell.00400.2002. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 133.Sun XC, Cui M, Bonanno JA. [HCO3−]-regulated expression and activity of soluble adenylyl cyclase in corneal endothelial and Calu-3 cells. BMC Physiol. 2004;4:8. doi: 10.1186/1472-6793-4-8. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 134.Dunn TA, et al. Imaging of cAMP levels and protein kinase A activity reveals that retinal waves drive oscillations in second-messenger cascades. J Neurosci. 2006;26:12807–12815. doi: 10.1523/JNEUROSCI.3238-06.2006. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 135.Dunn TA, Storm DR, Feller MB. Calcium-dependent increases in protein kinase-A activity in mouse retinal ganglion cells are mediated by multiple adenylate cyclases. PLoS ONE. 2009;4:e7877. doi: 10.1371/journal.pone.0007877. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 136.Shen BQ, Finkbeiner WE, Wine JJ, Mrsny RJ, Widdicombe JH. Calu-3: a human airway epithelial cell line that shows cAMP-dependent Cl- secretion. Am J Physiol Lung Cell Mol Physiol. 1994;266:L493–501. doi: 10.1152/ajplung.1994.266.5.L493. [DOI] [PubMed] [Google Scholar]
- 137.Wang Y, et al. Regulation of CFTR channels by HCO3−-sensitive soluble adenylyl cyclase in human airway epithelial cells. Am J Physiol Cell Physiol. 2005;289:C1145–1151. doi: 10.1152/ajpcell.00627.2004. [DOI] [PubMed] [Google Scholar]
- 138.Baudouin-Legros M, et al. Control of basal CFTR gene expression by bicarbonate-sensitive adenylyl cyclase in human pulmonary cells. Cell Physiol Biochem. 2008;21:075–086. doi: 10.1159/000113749. [DOI] [PubMed] [Google Scholar]
- 139.Schmid A, et al. Decreased soluble adenylyl cyclase activity in cystic fibrosis Is related to defective apical bicarbonate exchange and affects ciliary beat frequency regulation. J Biol Chem. 2010;285:29998–30007. doi: 10.1074/jbc.M110.113621. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 140.Sisson JH, Pavlik JA, Wyatt TA. Alcohol stimulates ciliary motility of isolated airway axonemes through a nitric oxide, cyclase, and cyclic nucleotide-dependent kinase mechanism. Alcoholism: Clinical and Experimental Research. 2009;33:610–616. doi: 10.1111/j.1530-0277.2008.00875.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 141.Strazzabosco M, et al. Diferentially expressed adenylyl cyclase isoforms mediate secretory functions in cholangiocyte subpopulation. Hepatology. 2009;50:244–252. doi: 10.1002/hep.22926. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 142.Alpini G, et al. Morphological, molecular, and functional heterogeneity of cholangiocytes from normal rat liver. Gastroenterology. 1996;110:1636–1643. doi: 10.1053/gast.1996.v110.pm8613073. [DOI] [PubMed] [Google Scholar]
- 143.Charles MA, Fanska R, Schmid FG, Forsham PH, Grodsky GM. Adenosine 3′,5′-monophosphate in pancreatic islets: glucose-induced insulin release. Science. 1973;179:569–571. doi: 10.1126/science.179.4073.569. [DOI] [PubMed] [Google Scholar]
- 144.Rutter GA. Nutrient-secretion coupling in the pancreatic islet beta-cell: recent advances. Mol Aspects Med. 2001;22:247–284. doi: 10.1016/s0098-2997(01)00013-9. [DOI] [PubMed] [Google Scholar]
- 145.Halm ST, Zhang J, Halm DR. β-adrenergic activation of electrogenic K+ and Cl− secretion in guinea pig distal colonic epithelium proceeds via separate cAMP signaling pathways. Am J Physiol Gastrointest Liver Physiol. 2010;299:81–95. doi: 10.1152/ajpgi.00035.2010. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 146.Grosell M. Intestinal anion exchange in marine fish osmoregulation. J Exp Biol. 2006;209:2813–2827. doi: 10.1242/jeb.02345. [DOI] [PubMed] [Google Scholar]
- 147.Wilson RW, et al. Contribution of fish to the marine inorganic carbon cycle. Science. 2009;323:359–362. doi: 10.1126/science.1157972. [DOI] [PubMed] [Google Scholar]
- 148.Ames A, Higashi K, Nesbett FB. Effects of PCO2, acetazolamide and ouabain on volume and composition of choroid-plexus fluid. The Journal of Physiology. 1965;181:516–524. doi: 10.1113/jphysiol.1965.sp007780. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 149.Lein ES, et al. Genome-wide atlas of gene expression in the adult mouse brain. Nature. 2007;445:168–176. doi: 10.1038/nature05453. [DOI] [PubMed] [Google Scholar]
- 150.Wu KY, et al. Soluble adenylyl cyclase is required for netrin-1 signaling in nerve growth cones. Nat Neurosci. 2006;9:1257–1264. doi: 10.1038/nn1767. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 151.Spengler D, et al. Differential signal transduction by five splice variants of the PACAP receptor. Nature. 1993;365:170–175. doi: 10.1038/365170a0. [DOI] [PubMed] [Google Scholar]
- 152.Vaudry D, Stork PJ, Lazarovici P, Eiden LE. Signaling pathways for PC12 cell differentiation: making the right connections. Science. 2002;296:1648–1649. doi: 10.1126/science.1071552. [DOI] [PubMed] [Google Scholar]
- 153.Stessin AM, et al. Soluble adenylyl cyclase mediates nerve growth factor-induced activation of Rap1. J Biol Chem. 2006;281:17253–17258. doi: 10.1074/jbc.M603500200. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 154.Han H, et al. Calcium-sensing soluble adenylyl cyclase mediates TNF signal transduction in human neutrophils. J Exp Med. 2005;202:353–361. doi: 10.1084/jem.20050778. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 155.Huggins C. The composition of bone and the function of the bone cell. Physiol Rev. 1937;17:119–143. [Google Scholar]
- 156.Bushinsky DA. Acid-base imbalance and the skeleton. Eur J Nutr. 2001;40:238–244. doi: 10.1007/s394-001-8351-5. [DOI] [PubMed] [Google Scholar]
- 157.Arnett T. Regulation of bone cell function by acid?base balance. Proc Nutr Soc. 2003;62:511–520. doi: 10.1079/pns2003268. [DOI] [PubMed] [Google Scholar]
- 158.Bernatchez R, et al. Differential expression of membrane and soluble adenylyl cyclase isoforms in cytotrophoblast cells and syncytiotrophoblasts of human placenta. Placenta. 2003;24:648–657. doi: 10.1016/s0143-4004(03)00060-2. [DOI] [PubMed] [Google Scholar]
- 159.Nunes AR, Monteiro EC, Johnson SM, Ganda EB. Bicarbonate-Regulated Soluble Adenylyl Cyclase (sAC) mRNA Expression and Activity in Peripheral Chemoreceptors. Arterial Chemoreceptors. 2009;648:235–241. 455. doi: 10.1007/978-90-481-2259-2_27. [DOI] [PubMed] [Google Scholar]
- 160.Chen MH, et al. Involvement of CFTR in oviductal HCO3− secretion and its effect on soluble adenylate cyclase-dependent early embryo development. Hum Reprod. 2010;25:1744–1754. doi: 10.1093/humrep/deq094. [DOI] [PubMed] [Google Scholar]
- 161.Steegborn C, et al. A novel mechanism for adenylyl cyclase inhibition from the crystal structure of its complex with catechol estrogen. J Biol Chem. 2005;280:31754–31759. doi: 10.1074/jbc.M507144200. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 162.Gille A, et al. Differential inhibition of adenylyl cyclase isoforms and soluble guanylyl cyclase by purine and pyrimidine nucleotides. J Biol Chem. 2004;279:19955–19969. doi: 10.1074/jbc.M312560200. [DOI] [PubMed] [Google Scholar]






