Skip to main content

MINI REVIEW article

Front. Physiol., 25 November 2013
Sec. Membrane Physiology and Membrane Biophysics
This article is part of the Research Topic Acid-base sensing and regulation: molecular mechanisms and functional implications in health and disease View all 27 articles

pH sensing via bicarbonate-regulated “soluble” adenylyl cyclase (sAC)

  • Department of Pharmacology, Weill Cornell Medical College, New York, NY, USA

Soluble adenylyl cyclase (sAC) is a source of the second messenger cyclic adenosine 3′, 5′ monophosphate (cAMP). sAC is directly regulated by bicarbonate (HCO3) ions. In living cells, HCO3 ions are in nearly instantaneous equilibrium with carbon dioxide (CO2) and pH due to the ubiquitous presence of carbonic anhydrases. Numerous biological processes are regulated by CO2, HCO3, and/or pH, and in a number of these, sAC has been shown to function as a physiological CO2/HCO3/pH sensor. In this review, we detail the known pH sensing functions of sAC, and we discuss two highly-studied, pH-dependent pathways in which sAC might play a role.

Almost 40 years ago, pioneering work of Theodor Braun described a novel, “soluble” adenylyl cyclase (AC) from rat testis (Braun and Dods, 1975). From the initial reports, it became clear that this cyclase was unique; unlike the more widely studied, G protein-regulated transmembrane adenylyl cyclases (tmACs), soluble AC activity was not associated with the plasma membrane (Braun and Dods, 1975); was predicted to be approximately 48 kDa in size (Gordeladze et al., 1981), was hormone and heterotrimeric G protein insensitive (Braun et al., 1977), and its activity appeared to be dependent upon Mn2+-ATP as substrate (Braun and Dods, 1975). It was also shown that it was distinct from the previously identified soluble guanylyl cyclase (Braun et al., 1977; Neer and Murad, 1979). Despite extensive searching, soluble AC activity had been detected only in testis (Neer, 1978). A membrane-associated activity found in sperm was similarly hormone and G protein insensitive (Braun and Dods, 1975; Stengel and Hanoune, 1984; Garty and Salomon, 1987; Rojas and Bruzzone, 1992), and it was hypothesized that during maturation in the testis and/or during the transport of sperm through the epididymis, a single, novel form of AC would change from being cytosolic to membrane-attached (Braun and Dods, 1975). Subsequent studies confirmed sAC was responsible for soluble activity in testis and the majority of the membrane-associated activity in sperm (Esposito et al., 2004; Hess et al., 2005).

Purification and Cloning of sAC

Numerous attempts to purify this unique cAMP-producing enzyme failed. In 1999, starting from 950-rat testis, our laboratory purified a 48 kDa candidate protein band which eluted with soluble AC activity (Buck et al., 1999, 2002). The amino acid sequences of three peptides derived from this 48 kDa candidate protein revealed it represented a novel protein and were sufficient to isolate its cDNA by degenerate PCR (Buck et al., 1999). The initial screen of a rat testis cDNA library identified three different clones, the longest of which had an open reading frame predicting the full-length protein (sACfl) would be 187 kDa. A second cDNA predicted a smaller, truncated sAC isoform, sACt, which was subsequently shown to be a bona fide splice variant (Jaiswal and Conti, 2001), and which corresponds to the originally purified protein (Buck et al., 1999).

sACt comprises the amino terminal ~50 kDa of sACfl, and it contains two domains (C1 and C2) homologous to the catalytic domains from other Class III nucleotidyl cyclases (Buck et al., 1999). Closely related catalytic domains are in ACs found in Cyanobacteria, and the biochemical properties of mammalian sAC (described below) are conserved from these bacterial orthologs (Steegborn et al., 2005a,b), which are thought to have evolved over 3 billion years ago.

Initial studies using the cloned sAC cDNA suggested it was abundantly expressed only in testis (Buck et al., 1999); however, subsequent studies using more sensitive RT-PCR (Sinclair et al., 2000; Farrell et al., 2008), mRNA microarrays (Geng et al., 2005), Western Blotting (Chen et al., 2000), or immunohistochemistry (Chen et al., 2013), revealed that sAC could be detected in nearly all tissues examined. Currently, sAC expression is considered to be ubiquitous. In individual cells, sAC can be found throughout the cytoplasm, inside or at various organelles and intracellular compartments—mitochondria, nuclei, microtubules, centrioles, mitotic spindles, and midbodies (Zippin et al., 2003). As such, sAC's intracellular localization provides an elegant solution to a long-standing conundrum inherent in the historical models for cAMP signaling. Prior to sAC's discovery and the demonstration that it is widely, if not ubiquitously expressed, and distributed throughout the cell, models for cAMP signaling were dependent upon cAMP production exclusively at the plasma membrane. These models required diffusion of the second messenger through the cytoplasm to its target effector proteins. Current models posit sAC localized at different cellular sites would provide specificity by transducing the cAMP signal exclusively to the appropriate player in close proximity to the cyclase (Wuttke et al., 2001; Zippin et al., 2003; Bundey and Insel, 2004).

sAC Activity

The early characterization of soluble AC activity revealed it was insensitive to the usual regulators of the more widely studied tmACs, including heterotrimeric G proteins (Braun and Dods, 1975) and the plant diterpene forskolin (Stengel et al., 1982; Forte et al., 1983); these properties were confirmed in the heterologously expressed sAC cDNAs (Buck et al., 1999). In partially purified preparations, soluble testis AC activity was reported to have a Km for its substrate, ATP, of ~1–2 mM in the presence of the divalent cation Mn2+ (Gordeladze et al., 1981; Stengel and Hanoune, 1984; Rojas et al., 1993). Other divalent cations were found to support the enzymatic activity, but none supported activity to the same extent as Mn2+ (Braun, 1975). For example, in the presence of Mg2+, the enzyme's Km for substrate ATP was reported to be above 15 mM (Stengel and Hanoune, 1984). In these early studies, there was no real consensus about regulation by Ca2+; in one report, Mn2+-ATP dependent activity was potentiated by Ca2+(Braun et al., 1977), but others found the same activity to be Ca2+ insensitive (Stengel and Hanoune, 1984). The membrane-associated, soluble AC-like activity found in sperm was thought to be stimulated by sodium bicarbonate (HCO3) (Okamura et al., 1985; Garty and Salomon, 1987; Visconti et al., 1990), while in guinea pig sperm, Ca2+ activation was reported to be dependent upon HCO3 (Garbers et al., 1982).

Heterologous expression of full-length (sACfl) and truncated sAC (sACt) isoforms revealed a significant difference in their specific activities (Buck et al., 1999; Jaiswal and Conti, 2003), which was subsequently shown to be due to an autoinhibitory domain present in the longer isoform (Chaloupka et al., 2006). It remains unclear how this autoinhibitory domain regulates sAC activity. sACfl also contains a unique Heme binding domain, but its contribution to activity also remains unclear (Middelhaufe et al., 2012). Despite these differences, sACt and sACfl share the same C1 and C2 catalytic domains (Buck et al., 1999), and much of their regulation is conserved (Buck et al., 1999; Chen et al., 2000; Chaloupka et al., 2006). Characterization of heterologously expressed and purified sACt confirmed many of the biochemical properties identified for purified soluble AC from testis, and they demonstrated that these properties were intrinsic to sAC. Purified sACt exhibits a Km for substrate ATP of ~1 mM in the presence of Mn2+, and while its affinity for Mg2+ATP is in excess of 10 mM, Ca2+ stimulates the enzyme's activity by increasing its affinity for ATP to ~1 mM (Litvin et al., 2003). Its affinity for ATP is close to the concentration found in cells (Traut, 1994), prompting us to postulate that sAC may be sensitive to physiologically relevant changes in ATP (Litvin et al., 2003). We recently demonstrated this to be the case; sAC, but not tmAC, activity inside cells is sensitive to inhibitors which diminished ATP levels (Zippin et al., 2013). Cellular ATP levels increase in high glucose, and we showed that overexpression of sAC converts a cell line which normally does not alter cAMP levels in response to glucose into glucose-responsive cells. In addition, we found that in cells which are normally glucose responsive, pancreatic β cells, sAC is required for the ATP-dependent glucose response (Zippin et al., 2013). Other functions of sAC have been identified based upon phenotypes in sAC knockout mice (Esposito et al., 2004; Hess et al., 2005; Lee et al., 2011; Choi et al., 2012; Zippin et al., 2013) or by taking advantage of pharmacological inhibitors which distinguish between sAC and tmACs (Bitterman et al., 2013), including functions dependent upon cellular Ca2+. Tumor Necrosis Factor (TNF) (Han et al., 2005) and neurotrophin (NGF) signaling (Stessin et al., 2006) and glucose sensing (Ramos et al., 2008) are mediated via Ca2+-dependent regulation of sAC.

sAC is directly stimulated by the bicarbonate anion (HCO3)(Chen et al., 2000). HCO3 stimulates sAC via two mechanisms; it relieves substrate inhibition and elevates Vmax (Litvin et al., 2003) by facilitating closure of the active site (Steegborn et al., 2005b). As previously predicted (Garbers et al., 1982), HCO3stimulation is synergistic with Ca2+ (Litvin et al., 2003). HCO3-dependent regulation of sAC plays a role in sperm activation (Esposito et al., 2004; Hess et al., 2005), activity-dependent metabolic communication between astrocytes and neurons (Choi et al., 2012), and multiple processes in the eye, including aqueous humor formation (Lee et al., 2011), retinal ganglion cell survival (Corredor et al., 2012) and corneal endothelial cell protection (Li et al., 2011).

Due to the ubiquitous presence of carbonic anhydrases (CAs), HCO3regulation means that sAC activity inside cells will be modulated by changes in CO2 and/or pHi (Tresguerres et al., 2010a; Buck and Levin, 2011). CO2-dependent regulation of sAC plays a role inside the mitochondrial matrix, where metabolically generated CO2 modulates the synthesis of ATP by the electron transport chain via sAC-generated cAMP (Acin-Perez et al., 2009, 2011; Di Benedetto et al., 2013; Lefkimmiatis et al., 2013), and in airway cilia, where CO2 regulates the ciliary beat frequency via sAC (Schmid et al., 2007, 2010).

sAC as pH Sensor

Although its in vitro activity is insensitive to physiologically relevant pH changes (Chen et al., 2000), the pH-dependent equilibrium between CO2 and HCO3 means that cellular sAC activity will be regulated by local changes in pH. In fact, sAC has been shown to play a role in pH dependent movements of the electrogenic, proton pumping vacuolar–ATPase (V-ATPase) in a number of physiological contexts (Pastor-Soler et al., 2003; Paunescu et al., 2008, 2010; Tresguerres et al., 2010b). In epididymis and proximal tubules of the kidney, V-ATPases translocate to the apical surface of the cell in response to a pH change in the corresponding lumen. sAC is in a complex with V-ATPase (Paunescu et al., 2008), and the V-ATPase translocation is mediated by sAC, in a CA dependent manner (Pastor-Soler et al., 2003, 2008). This pH dependent signaling is evolutionarily conserved; a similar mechanism, involving CA, sAC, and V-ATPase, is responsible for organismal pH homeostasis in shark (Tresguerres et al., 2010b). Interestingly, in shark gills the V-ATPase translocates to the basolateral side of the pH sensing cells. Thus, sAC-dependent V-ATPase mobilization is a conserved mechanism of pH sensing. The appreciation that pH sensing can be achieved via HCO3 regulation of sAC instead of being exclusively dependent upon proton sensing suggests there may be additional pH-dependent physiological processes dependent upon sAC. We discuss two possible examples below.

pH Sensing in the Endosomal-Lysosomal Pathway

The endo-lysosomal system is central to the processes of autophagy and endocytosis (Klionsky, 2007; Mizushima, 2007), and there is growing appreciation of its involvement in a broad range of diseases (Futerman and Van Meer, 2004; Nixon et al., 2008). As internalized materials pass from early to late endosomes and finally to lysosomes, the lumen of the endocytic organelles become more acidic. Lysosomes are the terminal compartment of both endocytic and autophagic pathways, and within lysosomes, acid hydrolase enzymes degrade proteins, lipids, and polysaccharides. The pH of the lysosome lumen is maintained between 4 and 5 (Pillay et al., 2002), which is the optimal pH for lysosomal enzyme activity.

Regulation of lysosomal pH is a complex process involving multiple channels and transporters. Acidification of lysosomes is accomplished by the electrogenic V-ATPase, which pumps protons into the lysosomal lumen (Forgac, 2007). Chloride movement through an opposite conductance pathway (Jentsch, 2007) (mediated at least in part via CLC7) and efflux of cations (Steinberg et al., 2010) facilitate vesicle acidification by neutralizing the positive charge and reducing the membrane potential caused by the pumped protons. Little is known about how the V-ATPase sets the pH or how these parallel ion transports are regulated. In particular, no pH-sensitive signaling cascades have been implicated.

Cyclic AMP has been shown to modulate lysosomal pH in macrophages (Di et al., 2006), microglia (Majumdar et al., 2007) and retinal pigment epithelium (RPE) cells (Liu et al., 2008). The cAMP effector, Protein Kinase A (PKA) increases chloride conductance (Bae and Verkman, 1990), possibly via the chloride channel CLC7. Lysosomal acidification in microglia is enhanced by upregulation of CLC7 (Majumdar et al., 2011), in what is thought to be a PKA dependent process (Majumdar et al., 2007). However, how the cAMP “second messenger” is made and whether cAMP levels are dependent upon pH remains unknown. It is tempting to postulate that sAC is the pH regulated source of cAMP regulating these processes.

Like lysosomes, both early and late endosomes are maintained within certain pH ranges; early endosomes range between pH ~5.9-6.8 whereas late endosomes range between pH ~4.9 and 6.0 (Maxfield and Yamashiro, 1987). Endosomal acidification is linked to intracellular trafficking, but it remains unknown how early endosomes “set” luminal pH to ~6.5 and late endosome/lysosomal set their luminal pH to ~5. Endosomal pH is maintained via similar proteins as control lysosomal pH, but endosomes use distinct isoforms of V-ATPases and chloride channels [for a complete review, see Forgac (2007), Stauber and Jentsch (2013)]. In such cases, different isoforms need to be trafficked to the endosomes or their activity modulated in order to establish and maintain the proper pH. sAC has already been shown to modulate the pH-dependent translocation of the V-ATPase to plasma membranes (Pastor-Soler et al., 2003; Tresguerres et al., 2010b); might sAC-generated cAMP play a role in trafficking V-ATPases or other chloride channels to endosomal/lysosomal membranes and hence establishing intra-vesicular pH?

pH Sensing in Osteoclasts

Bone remodeling is a tightly regulated process involving bone formation by osteoblasts and bone resorption by osteoclasts. Imbalances in either result in bone diseases. For example, excessive resorption of bone due to increased activity or number of osteoclasts leads to osteoporosis, while deficient osteoclastic bone resorption leads to osteopetrosis (high bone mineral density) (Lazner et al., 1999; Qin et al., 2012). Bicarbonate and pH are known regulators of osteoclasts. Bone resorption is inhibited by high HCO3 or alkaline pH and stimulated by low HCO3 or acidic pH (Kraut et al., 1986; Goldhaber and Rabadjija, 1987; Arnett and Dempster, 1990; Krieger et al., 1992; Geng et al., 2009). In vitro, high concentrations of HCO3 decreased osteoclast formation and growth. The effects of HCO3 were blocked by inhibitors of sAC and recapitulated by addition of exogenous cAMP (Geng et al., 2009), suggesting sAC activity inhibits osteoclast formation. Consistent with this potential role, human sAC was identified as a candidate gene in the locus tightly associated with hyperabsorptive hypercalciuria (Reed et al., 1999), a disease characterized by low bone mineral density.

In osteoclasts, V-ATPase is an essential player in bone reabsorption. Acidification of the resorptive lacunae (to a pH of ~4.5) is achieved by the combined actions of the V-ATPase and CLC-7 clustered at high density in the ruffled border (Kornak et al., 2001; Qin et al., 2012). A continuous flow of H+ and Cl dissolves the mineral component of bone and enhance the activity of enzymes that digest the organic matrix. The important role played by the V-ATPase/CLC-7 complex suggests potential involvement of sAC, and might indicate that sAC contributes to multiple pH-dependent functions in bone.

Conclusions

sAC functions as a pH sensor in numerous physiological systems. Historically, pH sensors were assumed to be proteins directly regulated by protons. With the discovery of sAC, it is clear that this dogma needs revision; pH sensing can also be achieved via a protein directly regulated by HCO3.

Conflict of Interest Statement

Levin and Buck own equity interest in CEP Biotech which has licensed commercialization of a panel of monoclonal antibodies directed against sAC.

References

Acin-Perez, R., Gatti, D. L., Bai, Y., and Manfredi, G. (2011). Protein phosphorylation and prevention of cytochrome oxidase inhibition by ATP: coupled mechanisms of energy metabolism regulation. Cell Metab. 13, 712–719. doi: 10.1016/j.cmet.2011.03.024

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Acin-Perez, R., Salazar, E., Kamenetsky, M., Buck, J., Levin, L. R., and Manfredi, G. (2009). Cyclic AMP produced inside mitochondria regulates oxidative phosphorylation. Cell Metab. 9, 265–276. doi: 10.1016/j.cmet.2009.01.012

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Arnett, T. R., and Dempster, D. W. (1990). Protons and osteoclasts. J. Bone Miner. Res. 5, 1099–1103. doi: 10.1002/jbmr.5650051102

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Bae, H. R., and Verkman, A. S. (1990). Protein kinase A regulates chloride conductance in endocytic vesicles from proximal tubule. Nature 348, 637–639. doi: 10.1038/348637a0

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Bitterman, J. L., Ramos-Espiritu, L., Diaz, A., Levin, L. R., and Buck, J. (2013). Pharmacological distinction between soluble and transmembrane Adenylyl Cyclases. J. Pharmacol. Exp. Ther. doi: 10.1124/jpet.113.208496. [Epub ahead of print].

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Braun, T. (1975). The effect of divalent cations on bovine spermatozoal adenylate cyclase activity. J. Cyclic Nucleotide Res. 1, 271–281.

Pubmed Abstract | Pubmed Full Text

Braun, T., and Dods, R. F. (1975). Development of a Mn-2+-sensitive, “soluble” adenylate cyclase in rat testis. Proc. Natl. Acad. Sci. U.S.A. 72, 1097–1101. doi: 10.1073/pnas.72.3.1097

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Braun, T., Frank, H., Dods, R., and Sepsenwol, S. (1977). Mn2+-sensitive, soluble adenylate cyclase in rat testis. Differentiation from other testicular nucleotide cyclases. Biochim. Biophys. Acta 481, 227–235. doi: 10.1016/0005-2744(77)90155-3

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Buck, J., and Levin, L. R. (2011). Physiological sensing of carbon dioxide/bicarbonate/pH via cyclic nucleotide signaling. Sensors 11, 2112–2128. doi: 10.3390/s110202112

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Buck, J., Sinclair, M. L., and Levin, L. R. (2002). Purification of soluble adenylyl cyclase. Meth. Enzymol. 345, 95–105. doi: 10.1016/S0076-6879(02)45009-4

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Buck, J., Sinclair, M. L., Schapal, L., Cann, M. J., and Levin, L. R. (1999). Cytosolic adenylyl cyclase defines a unique signaling molecule in mammals. Proc. Natl. Acad. Sci. U.S. A. 96, 79–84. doi: 10.1073/pnas.96.1.79

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Bundey, R. A., and Insel, P. A. (2004). Discrete intracellular signaling domains of soluble adenylyl cyclase: camps of cAMP? Sci. STKE 2004, pe19. doi: 10.1126/stke.2312004pe19

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Chaloupka, J. A., Bullock, S. A., Iourgenko, V., Levin, L. R., and Buck, J. (2006). Autoinhibitory regulation of soluble adenylyl cyclase. Mol. Reprod. Dev. 73, 361–368. doi: 10.1002/mrd.20409

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Chen, J., Martinez, J., Milner, T. A., Buck, J., and Levin, L. R. (2013). Neuronal expression of soluble adenylyl cyclase in the mammalian brain. Brain Res. 1518, 1–8. doi: 10.1016/j.brainres.2013.04.027

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Chen, Y., Cann, M. J., Litvin, T. N., Iourgenko, V., Sinclair, M. L., Levin, L. R., et al. (2000). Soluble adenylyl cyclase as an evolutionarily conserved bicarbonate sensor. Science 289, 625–628. doi: 10.1126/science.289.5479.625

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Choi, H. B., Gordon, G. R., Zhou, N., Tai, C., Rungta, R. L., Martinez, J., et al. (2012). Metabolic Communication between astrocytes and neurons via bicarbonate-responsive soluble adenylyl cyclase. Neuron 75, 1094–1104. doi: 10.1016/j.neuron.2012.08.032

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Corredor, R. G., Trakhtenberg, E. F., Pita-Thomas, W., Jin, X., Hu, Y., and Goldberg, J. L. (2012). Soluble adenylyl cyclase activity is necessary for retinal ganglion cell survival and axon growth. J. Neurosci. 32, 7734–7744. doi: 10.1523/JNEUROSCI.5288-11.2012

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Di, A., Brown, M. E., Deriy, L. V., Li, C., Szeto, F. L., Chen, Y., et al. (2006). CFTR regulates phagosome acidification in macrophages and alters bactericidal activity. Nat. Cell Biol. 8, 933–944. doi: 10.1038/ncb1456

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Di Benedetto, G., Scalzotto, E., Mongillo, M., and Pozzan, T. (2013). Mitochondrial Ca(2+) uptake induces cyclic AMP generation in the matrix and modulates organelle ATP levels. Cell Metab. 17, 965–975. doi: 10.1016/j.cmet.2013.05.003

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Esposito, G., Jaiswal, B. S., Xie, F., Krajnc-Franken, M. A., Robben, T. J., Strik, A. M., et al. (2004). Mice deficient for soluble adenylyl cyclase are infertile because of a severe sperm-motility defect. Proc. Natl. Acad. Sci. U.S.A. 101, 2993–2998. doi: 10.1073/pnas.0400050101

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Farrell, J., Ramos, L., Tresguerres, M., Kamenetsky, M., Levin, L. R., and Buck, J. (2008). Somatic ‘soluble’ adenylyl cyclase isoforms are unaffected in Sacy tm1Lex/Sacy tm1Lex ‘knockout’ mice. PLoS ONE 3:e3251. doi: 10.1371/journal.pone.0003251

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Forgac, M. (2007). Vacuolar ATPases: rotary proton pumps in physiology and pathophysiology. Nat. Rev. Mol. Cell Biol. 8, 917–929. doi: 10.1038/nrm2272

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Forte, L. R., Bylund, D. B., and Zahler, W. L. (1983). Forskolin does not activate sperm adenylate cyclase. Mol. Pharmacol. 24, 42–47.

Pubmed Abstract | Pubmed Full Text

Futerman, A. H., and Van Meer, G. (2004). The cell biology of lysosomal storage disorders. Nat. Rev. Mol. Cell Biol. 5, 554–565. doi: 10.1038/nrm1423

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Garbers, D. L., Tubb, D. J., and Hyne, R. V. (1982). A requirement of bicarbonate for Ca2+-induced elevations of cyclic AMP in guinea pig spermatozoa. J. Biol. Chem. 257, 8980–8984.

Pubmed Abstract | Pubmed Full Text

Garty, N. B., and Salomon, Y. (1987). Stimulation of partially purified adenylate cyclase from bull sperm by bicarbonate. FEBS Lett. 218, 148–152. doi: 10.1016/0014-5793(87)81036-0

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Geng, W., Hill, K., Zerwekh, J. E., Kohler, T., Muller, R., and Moe, O. W. (2009). Inhibition of osteoclast formation and function by bicarbonate: Role of soluble adenylyl cyclase. J. Cell Physiol. 220, 332–340. doi: 10.1002/jcp.21767

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Geng, W., Wang, Z., Zhang, J., Reed, B. Y., Pak, C. Y., and Moe, O. W. (2005). Cloning and characterization of the human soluble adenylyl cyclase. Am. J. Physiol. Cell Physiol. 288, C1305–C1316. doi: 10.1152/ajpcell.00584.2004

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Goldhaber, P., and Rabadjija, L. (1987). H+ stimulation of cell-mediated bone resorption in tissue culture. Am. J. Physiol. 253, E90–E98.

Pubmed Abstract | Pubmed Full Text

Gordeladze, J. O., Andersen, D., and Hansson, V. (1981). Physicochemical and kinetic properties of the Mn2+-dependent adenylyl cyclase of the human testis. J. Clin. Endocrinol. Metab. 53, 465–471. doi: 10.1210/jcem-53-3-465

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Han, H., Stessin, A., Roberts, J., Hess, K., Gautam, N., Kamenetsky, M., et al. (2005). Calcium-sensing soluble adenylyl cyclase mediates TNF signal transduction in human neutrophils. J. Exp. Med. 202, 353–361. doi: 10.1084/jem.20050778

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Hess, K. C., Jones, B. H., Marquez, B., Chen, Y., Ord, T. S., Kamenetsky, M., et al. (2005). The “soluble” adenylyl cyclase in sperm mediates multiple signaling events required for fertilization. Dev. Cell 9, 249–259. doi: 10.1016/j.devcel.2005.06.007

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Jaiswal, B. S., and Conti, M. (2001). Identification and functional analysis of splice variants of the germ cell soluble adenylyl cyclase. J. Biol. Chem. 276, 31698–31708. doi: 10.1074/jbc.M011698200

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Jaiswal, B. S., and Conti, M. (2003). Calcium regulation of the soluble adenylyl cyclase expressed in mammalian spermatozoa. Proc. Natl. Acad. Sci. U.S.A. 100, 10676–10681. doi: 10.1073/pnas.1831008100

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Jentsch, T. J. (2007). Chloride and the endosomal-lysosomal pathway: emerging roles of CLC chloride transporters. J. Physiol. 578, 633–640. doi: 10.1113/jphysiol.2006.124719

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Klionsky, D. J. (2007). Autophagy: from phenomenology to molecular understanding in less than a decade. Nat. Rev. Mol. Cell Biol. 8, 931–937. doi: 10.1038/nrm2245

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Kornak, U., Kasper, D., Bosl, M. R., Kaiser, E., Schweizer, M., Schulz, A., et al. (2001). Loss of the ClC-7 chloride channel leads to osteopetrosis in mice and man. Cell 104, 205–215. doi: 10.1016/S0092-8674(01)00206-9

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Kraut, J. A., Mishler, D. R., Singer, F. R., and Goodman, W. G. (1986). The effects of metabolic acidosis on bone formation and bone resorption in the rat. Kidney Int. 30, 694–700. doi: 10.1038/ki.1986.242

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Krieger, N. S., Sessler, N. E., and Bushinsky, D. A. (1992). Acidosis inhibits osteoblastic and stimulates osteoclastic activity in vitro. Am. J. Physiol. 262, F442–F448.

Pubmed Abstract | Pubmed Full Text

Lazner, F., Gowen, M., Pavasovic, D., and Kola, I. (1999). Osteopetrosis and osteoporosis: two sides of the same coin. Hum. Mol. Genet. 8, 1839–1846. doi: 10.1093/hmg/8.10.1839

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Lee, Y. S., Tresguerres, M., Hess, K., Marmorstein, L. Y., Levin, L. R., Buck, J., et al. (2011). Regulation of anterior chamber drainage by bicarbonate-sensitive soluble adenylyl cyclase in the ciliary body. J. Biol. Chem. 286, 41353–41358. doi: 10.1074/jbc.M111.284679

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Lefkimmiatis, K., Leronni, D., and Hofer, A. M. (2013). The inner and outer compartments of mitochondria are sites of distinct cAMP/PKA signaling dynamics. J. Cell Biol. 202, 453–462. doi: 10.1083/jcb.201303159

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Li, S., Allen, K. T., and Bonanno, J. A. (2011). Soluble adenylyl cyclase mediates bicarbonate-dependent corneal endothelial cell protection. Am. J. Physiol. Cell Physiol. 300, C368–C374. doi: 10.1152/ajpcell.00314.2010

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Litvin, T. N., Kamenetsky, M., Zarifyan, A., Buck, J., and Levin, L. R. (2003). Kinetic properties of “soluble” adenylyl cyclase. Synergism between calcium and bicarbonate. J. Biol. Chem. 278, 15922–15926. doi: 10.1074/jbc.M212475200

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Liu, J., Lu, W., Reigada, D., Nguyen, J., Laties, A. M., and Mitchell, C. H. (2008). Restoration of lysosomal pH in RPE cells from cultured human and ABCA4(-/-) mice: pharmacologic approaches and functional recovery. Invest. Ophthalmol. Vis. Sci. 49, 772–780. doi: 10.1167/iovs.07-0675

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Majumdar, A., Capetillo-Zarate, E., Cruz, D., Gouras, G. K., and Maxfield, F. R. (2011). Degradation of Alzheimer's amyloid fibrils by microglia requires delivery of ClC-7 to lysosomes. Mol. Biol. Cell 22, 1664–1676. doi: 10.1091/mbc.E10-09-0745

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Majumdar, A., Cruz, D., Asamoah, N., Buxbaum, A., Sohar, I., Lobel, P., et al. (2007). Activation of microglia acidifies lysosomes and leads to degradation of Alzheimer amyloid fibrils. Mol. Biol. Cell 18, 1490–1496. doi: 10.1091/mbc.E06-10-0975

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Maxfield, F. R., and Yamashiro, D. J. (1987). Endosome acidification and the pathways of receptor-mediated endocytosis. Adv. Exp. Med. Biol. 225, 189–198. doi: 10.1007/978-1-4684-5442-0_16

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Middelhaufe, S., Leipelt, M., Levin, L. R., Buck, J., and Steegborn, C. (2012). Identification of a haem domain in human soluble adenylate cyclase. Biosci. Rep. 32, 491–499. doi: 10.1042/BSR20120051

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Mizushima, N. (2007). Autophagy: process and function. Genes Dev. 21, 2861–2873. doi: 10.1101/gad.1599207

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Neer, E. J. (1978). Physical and functional properties of adenylate cyclase from mature rat testis. J. Biol. Chem. 253, 5808–5812.

Pubmed Abstract | Pubmed Full Text

Neer, E. J., and Murad, F. (1979). Separation of soluble adenylate and guanylate cyclases from the mature rat testis. Biochim. Biophys. Acta 583, 531–534. doi: 10.1016/0304-4165(79)90070-9

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Nixon, R. A., Yang, D. S., and Lee, J. H. (2008). Neurodegenerative lysosomal disorders: a continuum from development to late age. Autophagy 4, 590–599. doi: 10.4161/auto.6259

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Okamura, N., Tajima, Y., Soejima, A., Masuda, H., and Sugita, Y. (1985). Sodium bicarbonate in seminal plasma stimulates the motility of mammalian spermatozoa through direct activation of adenylate cyclase. J. Biol. Chem. 260, 9699–9705.

Pubmed Abstract | Pubmed Full Text

Pastor-Soler, N., Beaulieu, V., Litvin, T. N., Da Silva, N., Chen, Y., Brown, D., et al. (2003). Bicarbonate-regulated adenylyl cyclase (sAC) is a sensor that regulates pH-dependent V-ATPase recycling. J. Biol. Chem. 278, 49523–49529. doi: 10.1074/jbc.M309543200

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Pastor-Soler, N. M., Hallows, K. R., Smolak, C., Gong, F., Brown, D., and Breton, S. (2008). Alkaline pH- and cAMP-induced V-ATPase membrane accumulation is mediated by protein kinase A in epididymal clear cells. Am. J. Physiol. Cell Physiol. 294, C488–C494. doi: 10.1152/ajpcell.00537.2007

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Paunescu, T. G., Da Silva, N., Russo, L. M., Mckee, M., Lu, H. A., Breton, S., et al. (2008). Association of soluble adenylyl cyclase with the V-ATPase in renal epithelial cells. Am. J. Physiol. Renal Physiol. 294, F130–F138. doi: 10.1152/ajprenal.00406.2007

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Paunescu, T. G., Ljubojevic, M., Russo, L. M., Winter, C., McLaughlin, M. M., Wagner, C. A., et al. (2010). cAMP stimulates apical V-ATPase accumulation, microvillar elongation, and proton extrusion in kidney collecting duct A-intercalated cells. Am. J. Physiol. Renal Physiol. 298, F643–F654. doi: 10.1152/ajprenal.00584.2009

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Pillay, C. S., Elliott, E., and Dennison, C. (2002). Endolysosomal proteolysis and its regulation. Biochem. J. 363, 417–429. doi: 10.1042/0264-6021:3630417

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Qin, A., Cheng, T. S., Pavlos, N. J., Lin, Z., Dai, K. R., and Zheng, M. H. (2012). V-ATPases in osteoclasts: structure, function and potential inhibitors of bone resorption. Int. J. Biochem. Cell Biol. 44, 1422–1435. doi: 10.1016/j.biocel.2012.05.014

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Ramos, L. S., Zippin, J. H., Kamenetsky, M., Buck, J., and Levin, L. R. (2008). Glucose and GLP-1 stimulate cAMP production via distinct adenylyl cyclases in INS-1E insulinoma cells. J. Gen. Physiol. 132, 329–338. doi: 10.1085/jgp.200810044

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Reed, B. Y., Heller, H. J., Gitomer, W. L., and Pak, C. Y. (1999). Mapping a gene defect in absorptive hypercalciuria to chromosome 1q23.3- q24. J. Clin. Endocrinol. Metab. 84, 3907–3913. doi: 10.1210/jc.84.11.3907

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Rojas, F. J., and Bruzzone, M. E. (1992). Regulation of cyclic adenosine monophosphate synthesis in human ejaculated spermatozoa. I. Experimental conditions to quantitate membrane-bound adenylyl cyclase activity. Hum. Reprod. 7, 1126–1130.

Pubmed Abstract | Pubmed Full Text

Rojas, F. J., Patrizio, P., Do, J., Silber, S., Asch, R. H., and Moretti-Rojas, I. (1993). Evidence for a novel adenylyl cyclase in human epididymal sperm. Endocrinology 133, 3030–3033. doi: 10.1210/en.133.6.3030

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Schmid, A., Sutto, Z., Nlend, M. C., Horvath, G., Schmid, N., Buck, J., et al. (2007). Soluble adenylyl cyclase is localized to cilia and contributes to ciliary beat frequency regulation via production of cAMP. J. Gen. Physiol. 130, 99–109. doi: 10.1085/jgp.200709784

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Schmid, A., Sutto, Z., Schmid, N., Novak, L., Ivonnet, P., Horvath, G., et al. (2010). Decreased soluble adenylyl cyclase activity in cystic fibrosis is related to defective apical bicarbonate exchange and affects ciliary beat frequency regulation. J. Biol. Chem. 285, 29998–30007. doi: 10.1074/jbc.M110.113621

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Sinclair, M. L., Wang, X. Y., Mattia, M., Conti, M., Buck, J., Wolgemuth, D. J., et al. (2000). Specific expression of soluble adenylyl cyclase in male germ cells. Mol. Reprod. Dev. 56, 6–11. doi: 10.1002/(SICI)1098-2795(200005)56:1<6::AID-MRD2>3.0.CO;2-M

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Stauber, T., and Jentsch, T. J. (2013). Chloride in vesicular trafficking and function. Annu. Rev. Physiol. 75, 453–477. doi: 10.1146/annurev-physiol-030212-183702

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Steegborn, C., Litvin, T. N., Hess, K. C., Capper, A. B., Taussig, R., Buck, J., et al. (2005a). A novel mechanism for adenylyl cyclase inhibition from the crystal structure of its complex with catechol estrogen. J. Biol. Chem. 280, 31754–31759. doi: 10.1074/jbc.M507144200

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Steegborn, C., Litvin, T. N., Levin, L. R., Buck, J., and Wu, H. (2005b). Bicarbonate activation of adenylyl cyclase via promotion of catalytic active site closure and metal recruitment. Nat. Struct. Mol. Biol. 12, 32–37. doi: 10.1038/nsmb880

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Steinberg, B. E., Huynh, K. K., Brodovitch, A., Jabs, S., Stauber, T., Jentsch, T. J., et al. (2010). A cation counterflux supports lysosomal acidification. J. Cell Biol. 189, 1171–1186. doi: 10.1083/jcb.200911083

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Stengel, D., Guenet, L., Desmier, M., Insel, P., and Hanoune, J. (1982). Forskolin requires more than the catalytic unit to activate adenylate cyclase. Mol. Cell. Endocrinol. 28, 681–690. doi: 10.1016/0303-7207(82)90155-1

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Stengel, D., and Hanoune, J. (1984). The sperm adenylate cyclase. Ann. N.Y. Acad. Sci. 438, 18–28. doi: 10.1111/j.1749-6632.1984.tb38272.x

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Stessin, A. M., Zippin, J. H., Kamenetsky, M., Hess, K. C., Buck, J., and Levin, L. R. (2006). Soluble adenylyl cyclase mediates nerve growth factor-induced activation of Rap1. J. Biol. Chem. 281, 17253–17258. doi: 10.1074/jbc.M603500200

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Traut, T. W. (1994). Physiological concentrations of purines and pyrimidines. Mol. Cell. Biochem. 140, 1–22. doi: 10.1007/BF00928361

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Tresguerres, M., Buck, J., and Levin, L. R. (2010a). Physiological carbon dioxide, bicarbonate, and pH sensing. Pflugers Arch. 460, 953–964. doi: 10.1007/s00424-010-0865-6

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Tresguerres, M., Parks, S. K., Salazar, E., Levin, L. R., Goss, G. G., and Buck, J. (2010b). Bicarbonate-sensing soluble adenylyl cyclase is an essential sensor for acid/base homeostasis. Proc. Natl. Acad. Sci. U.S.A. 107, 442–447. doi: 10.1073/pnas.0911790107

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Visconti, P. E., Muschietti, J. P., Flawia, M. M., and Tezon, J. G. (1990). Bicarbonate dependence of cAMP accumulation induced by phorbol esters in hamster spermatozoa. Biochim. Biophys. Acta 1054, 231–236. doi: 10.1016/0167-4889(90)90246-A

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Wuttke, M. S., Buck, J., and Levin, L. R. (2001). Bicarbonate-regulated soluble adenylyl cyclase. JOP 2, 154–158.

Pubmed Abstract | Pubmed Full Text

Zippin, J. H., Chen, Y., Nahirney, P., Kamenetsky, M., Wuttke, M. S., Fischman, D. A., et al. (2003). Compartmentalization of bicarbonate-sensitive adenylyl cyclase in distinct signaling microdomains. FASEB J. 17, 82–84. doi: 10.1096/fj.02-0598fje

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Zippin, J. H., Chen, Y., Straub, S. G., Hess, K. C., Diaz, A., Lee, D., et al. (2013). CO2/HCO3- and calcium regulated soluble adenylyl cyclase as a physiological ATP Sensor. J. Biol. Chem. doi: 10.1074/jbc.M113.510073. [Epub ahead of print].

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Keywords: cyclic AMP, soluble adenylyl cyclase, bicarbonate, carbonic anhydrase, pH

Citation: Rahman N, Buck J and Levin LR (2013) pH sensing via bicarbonate-regulated “soluble” adenylyl cyclase (sAC). Front. Physiol. 4:343. doi: 10.3389/fphys.2013.00343

Received: 06 September 2013; Paper pending published: 12 October 2013;
Accepted: 06 November 2013; Published online: 25 November 2013.

Edited by:

Mark O. Bevensee, University of Alabama at Birmingham, USA

Reviewed by:

Sarah L. Sayner, University of South Alabama, USA
Joseph A. Bonanno, Indiana University School of Optometry, USA

Copyright © 2013 Rahman, Buck and Levin. This is an open-access article distributed under the terms of the Creative Commons Attribution License (CC BY). The use, distribution or reproduction in other forums is permitted, provided the original author(s) or licensor are credited and that the original publication in this journal is cited, in accordance with accepted academic practice. No use, distribution or reproduction is permitted which does not comply with these terms.

*Correspondence: Jochen Buck, Department of Pharmacology, Weill Cornell Medical College, 1300 York Avenue, New York, NY 10065, USA. e-mail: jobuck@med.cornell.edu

Disclaimer: All claims expressed in this article are solely those of the authors and do not necessarily represent those of their affiliated organizations, or those of the publisher, the editors and the reviewers. Any product that may be evaluated in this article or claim that may be made by its manufacturer is not guaranteed or endorsed by the publisher.