
Keywords: adenylyl cyclase, cAMP, G protein-coupled receptors
Abstract
Adenylyl cyclases (ACs) catalyze the conversion of ATP to the ubiquitous second messenger cAMP. Mammals possess nine isoforms of transmembrane ACs, dubbed AC1–9, that serve as major effector enzymes of G protein-coupled receptors (GPCRs). The transmembrane ACs display varying expression patterns across tissues, giving the potential for them to have a wide array of physiological roles. Cells express multiple AC isoforms, implying that ACs have redundant functions. Furthermore, all transmembrane ACs are activated by Gαs, so it was long assumed that all ACs are activated by Gαs-coupled GPCRs. AC isoforms partition to different microdomains of the plasma membrane and form prearranged signaling complexes with specific GPCRs that contribute to cAMP signaling compartments. This compartmentation allows for a diversity of cellular and physiological responses by enabling unique signaling events to be triggered by different pools of cAMP. Isoform-specific pharmacological activators or inhibitors are lacking for most ACs, making knockdown and overexpression the primary tools for examining the physiological roles of a given isoform. Much progress has been made in understanding the physiological effects mediated through individual transmembrane ACs. GPCR-AC-cAMP signaling pathways play significant roles in regulating functions of every cell and tissue, so understanding each AC isoform’s role holds potential for uncovering new approaches for treating a vast array of pathophysiological conditions.
CLINICAL HIGHLIGHTS
The potential physiological roles of adenylyl cyclases (ACs) are enormous, making these enzymes a potential therapeutic target for diseases of every organ and system. Understanding the physiological functions regulated by each AC isoform is therefore essential for drug development. Therapeutic strategies directed toward AC will lack specificity unless a specific AC isoform is targeted.
1. INTRODUCTION
Adenylyl cyclases (ACs) are a family of enzymes that catalyze the conversion of adenosine triphosphate (ATP) into cyclic adenosine monophosphate (cAMP), a ubiquitous second messenger across the kingdoms of life (1–4). The enzymes are expressed throughout most organisms and play key roles in the diversity and complexity of cAMP signaling pathways (5). This review focuses on the mammalian transmembrane ACs, excluding the soluble mammalian adenylyl cyclase (often termed sAC or AC10). sAC is regulated by bicarbonate and calcium, not G proteins, and plays key roles in pH sensing and many specific (patho)physiological functions (6–9). Mammals express nine isoforms of transmembrane ACs, termed AC1–9. Cloning of these isoforms reveals highly conserved catalytic domains but significant sequence diversity outside these domains (3). The major domains of ACs include an intracellular NH2 terminus, a coiled-coil helical domain that separates two clusters of six transmembrane-spanning helixes (for a total of 12 transmembrane-spanning domains) from the large intracellular C1 and C2 cytoplasmic loops (10). The C1 and C2 loops are divided into C1a and C2a subdomains that dimerize to form the catalytic site and C1b and C2b regulatory subdomains. The C1a and C2a subdomains are highly conserved across ACs (1). Divergence outside of these conserved subdomains plays roles in the signaling and regulatory diversity across ACs. AC structure has been recently reviewed (11) and is summarized in FIGURE 1.
FIGURE 1.
General structure of mammalian transmembrane adenylyl cyclases (ACs). TM1 and TM2 are transmembrane domains each consisting of 6 transmembrane α-helices. The second extracellular loop of TM2 contains N-linked glycosylation sites that target some isoforms to lipid rafts (12). C1 and C2 are intracellular cytoplasmic loops, with C1a and C2a forming the catalytic domain and forskolin binding site (in the case of AC1–8). The C1a domain is the site of Gαi binding, whereas the C2a domain interacts with Gαs. C1b and C2b are regulatory subdomains, and the NH2-terminal domain is involved in many protein-protein interactions. The C1 and C2 domains are separated from each transmembrane domain by 2 coiled-coil helices (shown in yellow). These 2 helices and the C1b domain of AC5 contain the indicated mutations associated with dyskinesias (13–16). This structural model is based largely on the recent cryo-EM structure of AC9 (10).
1.1. cAMP Signal Transduction
Whereas ACs produce cAMP from ATP, G protein-coupled receptors (GPCRs), protein kinase A (PKA), A kinase anchoring proteins (AKAPs), and phosphodiesterases (PDEs) all contribute to the initiation, propagation, termination, or organization of cAMP signaling (FIGURE 2). The first step of signal generation is the activation of the G protein by an agonist-occupied GPCR. The stimulatory G protein alpha subunit, Gαs, then activates AC enzymatic activity leading to an increase in cAMP synthesis (17). Other GPCRs preferentially activate the inhibitory G protein alpha subunit, Gαi, which can inhibit AC activity and reduce cAMP synthesis (18).
FIGURE 2.

Schematic diagram of signaling via transmembrane adenylyl cyclases (ACs). Once activated by an agonist, a G protein-coupled receptor (GPCR) will activate coupled G proteins. Gαs activates AC activity, which catalyzes the conversion of ATP to cAMP, whereas Gαi inhibits AC activity. AC isoforms are also regulated by various other signals, as described in TABLE 1. Phosphodiesterases (PDEs) terminate the signal by decyclizing cAMP to 5′-AMP. cAMP activates several effector proteins, including PKA, Epac, Popeye Domain Containing (POPDC), and cyclic nucleotide-gated channels. PKA is anchored by A kinase anchoring proteins (AKAPs) to specific phosphorylation targets. PKA, Epac, cyclic nucleotide-gated channels, and POPDCs elicit various downstream physiological responses depending on the cell type.
The most ubiquitous and well-studied effector molecule of cAMP is PKA, which exists as a holoenzyme of two regulatory (cAMP binding) subunits and two catalytic (kinase) subunits (19–21). PKA catalytic subunits phosphorylate consensus sequences in downstream effector proteins to elicit cellular responses (22). Although in theory thousands of proteins can be phosphorylated by PKA, in practice most PKA action is targeted to specific effector proteins via tethering to AKAPs. AKAPs bind the regulatory subunits of PKA as well as key targets of PKA phosphorylation to enable rapid, high-fidelity coupling of the kinase to downstream effectors (23–25). The AKAP family is large and diverse and so can play the roles of cAMP signaling complex scaffolds and organizers that contribute to the unique characteristics of differentiated cells. Several AKAPs are known to bind to specific AC isoforms (26–28).
PDEs terminate the cAMP signal by decyclizing the cAMP to AMP (29). Eleven gene families and 21 individual genes for PDEs exist, and they differ in terms of substrate specificities (30). PDE4, PDE7, and PDE8 hydrolyze only cAMP, whereas PDE1, PDE2, PDE3, PDE10, and PDE11 hydrolyze both cAMP and cyclic guanosine monophosphate (cGMP). PDE5, PDE6, and PDE9 only hydrolyze cGMP. Certain PDE isoforms specifically associate with particular GPCR-AC or AC-AKAP complexes (31, 32). AKAPs bind specifically with AC isoforms as well as a whole range of other signaling proteins and possess a PKA binding motif to coordinate PKA activation of downstream effectors (28). Some AKAPs also possess a PDE binding domain, which recruits PDEs to AC/GPCR complex sites to regulate cAMP levels and limit diffusion of the cAMP signal (33).
Another important downstream effector of cAMP signaling is exchange protein activated by cAMP, or EPAC (34). EPACs are guanine-nucleotide exchange factor proteins that are directly activated by cAMP (35). EPACs were more recently discovered and are thus less studied than PKA, but EPACs are important effectors in mediating various physiological responses of cAMP signaling (36). Cyclic nucleotide-gated channels, which nonselectively gate cations in response to direct binding by cAMP or cGMP, have specific roles in regulating membrane potential in sensory neurons of the olfactory and retinal systems and are involved in the regulation of heart rate (37–41). Additional cAMP effectors, Popeye Domain Containing (POPDC) proteins, have been described more recently (42, 43). POPDC proteins have diverse and complex functions, but early studies indicate they are widely expressed and play roles in various physiological effects of cAMP signaling. POPDC proteins are known to regulate heart rate and rhythm, epithelial cell tight junction formation, trabecular meshwork contraction and intraocular pressure, cell adhesion, and vesicular fusion (44–47).
1.2. Distinguishing the Functions of AC Isoforms
Complex signaling pathways activated by cAMP and the conservation of the catalytic domain of ACs have made it difficult to understand the specific roles of individual AC isoforms. Limited by a general lack of isoform-specific inhibitors (11), studies to tease out differences among isoforms have largely relied on genetic approaches. The general expression patterns of the various AC isoforms in cells and tissues have been elucidated (48–51). Most cell types express multiple isoforms, but abundant expression of certain isoforms in particular tissues is notable (FIGURE 3). AC1 and AC8 consistently show high expression in areas of the brain such as the hippocampus, cerebellum, cortex, thalamus, and hypothalamus. AC1 also is notably expressed in the heart, kidney, skeletal muscle, pancreas, and reproductive organs (56). AC2 expression is widespread, but it shows particularly high expression in lung and brain tissue. AC3 is most concentrated in olfactory epithelia and the pancreas (1). AC4–7 show high expression in regions throughout the heart and vascular system (57). AC5 expression is high in the basal ganglia and the heart (56, 58). AC6 expression is widespread but notably high in the cerebellum, choroid plexus, and the heart (56, 58). AC9 is widespread but most concentrated in skeletal muscles, the pituitary gland, and cerebral cortex (1, 59).
FIGURE 3.
Tissue distribution of transmembrane adenylyl cyclases (ACs): the predominant AC isoforms expressed by major organs and systems in the human body. AC isoforms are color coded by group (group I in blue, group II in green, group III in red, group IV in brown). The digestive tract expresses all AC isoforms except AC2 (52). The vasculature expresses primarily AC3, AC5, and AC6 (53–55).
The widespread expression of ACs precludes classification of the isoforms by tissue type or other anatomical distribution. Instead, it is more useful to classify the isoforms by their regulatory properties (FIGURE 4). The major direct regulation of ACs is by G protein subunits and Ca2+ via calmodulin (CaM) (1). The Ca2+/CaM-stimulated isoforms are AC1 and AC8, with weak stimulation of AC3 (60, 61). Although inhibition of AC activity by high concentrations (mid micromolar) of free Ca2+ is seen in all isoforms, high-affinity inhibition by Ca2+ is observed in AC5 and AC6 (62, 63). Gβγ, Gαi, and Gαs subunits have stimulatory and/or inhibitory effects on specific AC isoforms. Gαs stimulates all AC isoforms (64). Gαi inhibits AC1, AC5, and AC6 (65, 66) and may have minor influence on the activity of other isoforms that is counteracted by the stimulatory effects of coincident activation of Gβγ. Gβγ inhibits AC1, AC3, and AC8 but conditionally stimulates AC2, AC4, AC5, AC6, and AC7 in the presence of another activator (60, 67–69). The diterpene forskolin directly activates all isoforms of mammalian transmembrane AC (70, 71) with the exception of AC9 (72, 73). An additional important isoform-specific regulatory mechanism is via AKAP-mediated targeting of protein kinase C (PKC) (74). PKC stimulates AC1, AC2, AC3, AC5, and AC7 but inhibits AC4 and AC6, with the impact on AC8 unknown (75–78). Overall, this pattern of regulation categorizes the isoforms into four main groups. Group I consists of AC1, AC3, and AC8, which are all Ca2+ stimulated. Group II consists of Gβγ-stimulated AC2, AC4, and AC7. Group III is Gαi-inhibited and Ca2+-inhibited AC5 and AC6. Group IV contains AC9 (TABLE 1) (50). This type-specific regulation of AC isoforms, as well as the tissue distributions of the isoforms, is well studied and has been reviewed in depth previously (1, 3, 11, 51, 96). AC1, AC8, and AC3 are largely concentrated in the central nervous system, so stimulation by Ca2+/CaM is mainly observed in these tissues (93).
FIGURE 4.
Regulation of mammalian transmembrane adenylyl cyclases (ACs). Numerous signals directly regulate enzymatic activity of specific AC isoforms. AC isoforms are classified into 4 groups (and color coded) based on these regulatory properties. Green arrows denote stimulatory effects, red arrows denote inhibitory effects, and dashed arrows show conditional effects that are dependent upon other, coincident stimuli. NO, nitric oxide.
Table 1.
Physiological regulators of transmembrane adenylyl cyclase isoforms
| AC Isoform | Stimulators | Inhibitors | References |
|---|---|---|---|
| AC1 | GαsCa2+/CaMPKC | GαiGβγ | (18, 64–66, 79–81) |
| AC2 | Gαs(Gβγ)PKCRaf kinase | (60, 64, 82) | |
| AC3 | GαsCa2+/CaMPKC | GβγRGS proteins | (64, 80–83) |
| AC4 | Gαs(Gβγ) | PKC | (60, 64, 76, 84) |
| AC5 | Gαs(Gβγ)PKCRaf kinase | GαiCa2+PKANitric oxideRGS proteinsAnnexin A4 | (64–66, 69, 79, 80, 85–89) |
| AC6 | Gαs(Gβγ)Raf kinase | GαiCa2+PKCPKANitric oxide | (64–66, 68, 69, 79, 80, 86–88, 90, 91) |
| AC7 | GαsPKC(Gβγ) | (60, 64, 92) | |
| AC8 | GαsCa2+/CaM | GβγPKA | (64, 93–95) |
| AC9 | Gαs | (64) |
Parentheses indicate conditional regulators.
1.3. cAMP Signaling Complexes and Compartmentation
To achieve rapid and specific signaling via the diffusible second messenger cAMP, the various components of the signaling pathway arrange in preformed signaling complexes within cells (97, 98). How these complexes form and regulate various cellular responses is not well understood, but direct interactions between certain GPCRs, ACs, AKAPs, and PDEs have been defined (28, 99, 100). The idea that a single AC isoform might have a distinct physiological role has traditionally been difficult to conceptualize. The ACs were initially viewed as largely redundant isoforms because of cells expressing multiple isoforms that have a singular enzymatic output. Appreciation of the distinct regulatory properties caused a shift in viewing these isozymes as central to tailoring the regulation of cAMP by coincident signals in a specific cell type. However, a framework for how these different isoforms might have distinct effects on cell function remained lacking.
Buxton and Brunton’s observations of cAMP signaling compartmentation (101) followed by the appreciation of ACs being stratified across distinct plasma membrane microdomains (54, 102, 103), facilitated the conceptualization of how individual ACs could have distinct physiological roles. AC isoforms are stratified by their localization in either lipid raft or nonraft microdomains (FIGURE 5). AC1, AC3, AC5, AC6, and AC8 localize to lipid raft microdomains, whereas AC2, AC4, AC7, and AC9 localize to nonraft domains (99, 104). The structural determinants of lipid raft localization of certain isoforms appear to depend on N-linked glycosylation of an extracellular loop in TM2 (FIGURE 1) and perhaps also the intracellular C1 and C2 domains, but more work is needed in this area (12, 105, 106). ACs appear stable in their localization to these microdomains, so they can anchor the formation of larger signaling complexes (99, 107). Of the many Gαs-coupled GPCRs in a given cell, certain receptors will preferentially couple to a specific AC isoform based on colocalization or preassembly (108–113). A given AC exists in complex with many other organizing and regulatory proteins, such as AKAPs, PKA, and PDEs, but may also form specific interactions with other proteins upon activation (11, 102, 114, 115). Microdomains other than lipid raft and nonraft plasma membranes likely exist, including cAMP signaling in endosomes (116).
FIGURE 5.
Representation of cAMP signaling compartmentation. Membrane microdomains and prearranged complexes form in most cells that couple specific G protein-coupled receptors (GPCRs) to unique subsets of physiological responses. Transmembrane adenylyl cyclase (AC) isoforms characteristically localize in either lipid raft (AC1, 3, 5, 6, and 8) or nonraft (AC2, 4, 7, and 9) domains, where they form complexes with unique signaling partners [GPCRs, A kinase anchoring proteins (AKAPs), phosphodiesterases (PDEs), and others] to create distinct cAMP signaling compartments. There are 43 AKAP isoforms and 24 PDE genes (plus many splice variants), making the combinatorial possibilities for cAMP signaling complex formation enormous. Diffusion of cAMP is limited by PKA buffering, colocalized PDE isoforms, and cellular barriers to diffusion (represented as a brown matrix). Effector proteins downstream of PKA can also be localized in specific domains such that each compartment can give rise to a unique pattern of cellular and physiological responses. AC isoforms are color coded by group (group I in blue, group II in green, group II in red, group IV in brown).
Various ACs can be used to examine the role of cAMP compartmentation. Specifically, in contrast to transmembrane AC isoforms, sAC gives rise to a subcellular cAMP distribution different from membrane-bound adenylyl cyclase (mAC), with distinct functional differences (117). Intriguingly, sAC has a bacterial equivalent, i.e., the type III secretion protein ExoY from Pseudomonas aeruginosa (118). However, both sAC and ExoY have a very broad substrate specificity, giving rise to cGMP, cCMP, and cUMP in addition to cAMP (119–121), so that sAC and ExoY have only limited value to specifically examine cAMP compartmentation. Nonetheless, sAC does display localized cAMP signaling in intracellular cytosolic compartments (122).
An important consequence of the localization of cAMP signaling complexes is that it allows cAMP to be produced locally within a cell and have specific signaling effects, as opposed to globally diffusing throughout the cytosol and triggering all possible downstream signals (99, 123, 124). Anchoring of cAMP response elements to the cAMP generating system can increase signal specificity dramatically. For example, regulation of the transient receptor potential vanilloid 1 (TRPV1) channel by forskolin-stimulated cAMP/PKA signaling became 100-fold more potent when AC5 and AKAP79 were coexpressed in HEK-293 cells (125). In mouse dorsal root ganglion neurons, disrupting the interaction between AKAP79 and AC5 reduced PGE2 regulation of TRPV1 by more than half (125). Thus, anchoring of the GPCR-AC signaling pathway and compartmentation of cAMP signaling play key roles in allowing for specific cell responses despite the ubiquitous nature of this second messenger and the expression of multiple AC isoforms in a particular cell.
2. PHYSIOLOGICAL ROLES OF AC ISOFORMS
As discussed above, the widespread distribution and ubiquitous nature of their product, cAMP, has made it difficult to tease out the unique physiological roles of the various AC isoforms. However, this research is an emerging field for the development of clinically useful pharmacological targets. Both pharmacological activators and inhibitors of ACs are available, but these pharmacological tools are not sufficiently selective for any given AC isoform to allow reasonable dissection of the function of an individual AC in a (patho)physiological context. This problem is highlighted by the AC activator forskolin and three AC inhibitors in TABLE 2. Shown are the data for representatives of the three transmembrane AC families, AC1, AC2, and AC5. Forskolin is more potent at activating AC1 than AC2 and AC5, but the difference is far too small to allow for discrimination of AC isoforms in a cellular or tissue context (126). The same problem applies to structurally distinct AC inhibitors. In general, AC2 is less sensitive to inhibition than AC1 and AC5, but the differences are too small to be useful in (patho)physiological contexts (127–129).
Table 2.
Primary pharmacological tools for studying adenylyl cyclase isoforms
There are additional problems with the available pharmacological tools. First, forskolin has effects on several pharmacological targets including glucose transporters (130) and nuclear receptors (131), rendering data interpretation difficult. Second, 2′, 3′-(O)-(N-methyl)anthraniloyl (MANT) nucleotides and trinitrophenyl (TNP) nucleotides are far too hydrophilic to cross the plasma membrane (127). Third, SQ22,536 and related compounds interact with adenosine receptors (132). Since adenosine receptors are ubiquitously expressed, interpretation with SQ22,539 in (patho)physiological settings is problematic. Finally, pharmacological data at each individual AC isoform are available for only a few compounds, e.g., SQ22,536 (128). In most cases, only data for representative AC isoforms are available (126, 127, 129). Although not ideal, it is true that different members from a given AC family are pharmacologically similar (128, 133).
Because of a general lack of isoform-specific inhibitors, there has been an emphasis on knockout and overexpression studies to tease out the specific roles of the isoforms in various systems and pathologies (50). These approaches provide the most direct evidence of how a particular AC isoform mediates a specific physiological function. Several studies also take a genetic approach focusing on the association of AC genes, ADCY1–9, with disease conditions. This review focuses on the major studies on the physiological roles of AC isoforms since the last comprehensive review of ACs (11). Research has overwhelmingly focused on AC1, AC8, AC5, and AC6, presumably because of their recognized roles in the heart and brain. Recent progress has been made in defining the functions of AC9 in the heart. Studies examining the physiological roles of AC2 and AC4 are notably lacking.
3. GROUP I (AC1, AC8, AC3)
3.1. Adenylyl Cyclase 1
3.1.1. Learning and behavior.
AC1 studies have largely focused on its role in the nervous system and regulation of neuronal long-term potentiation (LTP), a molecular surrogate for learning. Early studies with the Drosophila Ca2+-sensitive AC1 homolog rutabaga (134, 135), followed by AC1 knockout (AC1−/−) mice, demonstrated that AC1 is imperative for some forms of LTP in several brain regions and that deletion leads to molecular and behavioral impairment of learning (136–138). However, AC1 knockout does not significantly reduce expression of LTP in the CA1 region of the hippocampus, only the rising phase, but the behavioral learning impairment still occurs (136, 139), although a later study demonstrated a deficit in the expression of LTP after one high-frequency stimulation but not two (140). AC1−/− mice are still able to find a platform in the Morris water maze, but unlike wild-type mice they fail to favor the region in which the platform is placed. Thus, they appear to lack a spatial memory component but still display a short-term ability to complete the task.
A reciprocal increase in LTP and certain learning behaviors can be induced by overexpressing AC1 in the forebrain (141). Although there were striking increases in LTP, AC1-overexpressing mice did not have pronounced behavioral changes in all learning paradigms. AC1 overexpression resulted in increases in novel object recognition but not contextual or cue-conditioned learning behaviors or spatial memory. It is critical to note that AC1 appears important for the longer-term storage of spatial and recognition memory; however, it seems dispensable for acquisition and for other conditioned learning paradigms (139, 140). Several studies have expanded on elucidating the differential regulation of AC1 on various learning paradigms. AC1 overexpression in mice failed to alter fear learning but did increase social recognition and object recognition (142). Although the object recognition data were similar to previous studies (141), the work was in contrast to a decrease in social interaction shown in a later study (143). Of significance, spatial memory was significantly decreased in aged AC1-overexpressing mice, and they displayed a behavioral phenotype similar to AC1−/−/AC8−/− double-knockout mice (143). The current data are conflicting, as a separate paper demonstrated increased spatial memory in AC1-overexpressing mice (144).
Two follow-up studies utilizing these AC1 overexpression mice have provided novel insight into other behaviors with a potential AC1 regulatory component. Overexpression of AC1 in the forebrain resulted in mice that were highly hyperactive, lacked behavioral inhibition, and were less social in multiple social behavior paradigms (143). Nighttime home cage activity increased but not daytime activity, similar to a study observing voluntary home cage wheel running in the daytime (145). Although AC1 overexpression led to risky behavior, a lack of behavioral inhibition, and hyperactivity, it seems to be protective during stress situations and can provide molecular and behavioral resilience to stressors (145). In both studies described above, AC1-overexpressing mice had resilience to acute stress paradigms (143, 145). However, AC1−/− mice did not display an increased susceptibility to acute stress (140).
These studies demonstrated an interesting role for AC1 in stress-related behaviors, although more rigorous exploration is needed to decipher the contrasts in the data. Overexpression studies do not necessarily provide insight into normal physiological regulation but provide insight into potential pathophysiology associated with increased AC1 expression or behavioral consequences of therapeutically stimulating AC1. Stress is linked to depression and anxiety phenotypes, and a recent genetic analysis reports that intronic single-nucleotide polymorphisms (SNPs) in ADCY1 are associated with anxiety and depression scores in a small human study (TABLE 3) (222).
Table 3.
Disease-linked polymorphisms in adenylyl cyclase isoforms
| AC Gene | Disease-Linked Polymorphisms | Nonsynonymous Variants | References |
|---|---|---|---|
| Adcy1 | Deafness | R1038X | (146, 147) |
| Adcy2 | Bipolar disorders, autism spectrum disorder, pulmonary diseases, congenital heart disease, Tourette syndrome, developmental disorders | A438T, L498S | (148–155) |
| Adcy3 | Obesity, depression, autism spectrum disorder, schizophrenia, bowel disease | N64I, S107P, G423X, V577M, C948Y, E1008A, I1106X, G1110R, F1118X, L1119S | (156–167) |
| Adcy4 | Unknown | ||
| Adcy5 | Altered glucose metabolism, diabetes and obesity, autism spectrum disorder, movement disorders (chorea, dyskinesia, familial dyskinesia with facial myokymia, dystonia, hypotonia, spastic paraparesis, alternating hemiplegia of childhood), congenital heart defect, possible link to Parkinson’s disease | Y233H, P399L, R418Q/W/G, R438P, A441V, I460F, N467S, I475M, A534T, D588N, R603C, E634D, K691E, R695W, G697V, L720P, A726T, R727K, E908K, R1013C, D1015E, E1025V, M1029K/R, D1060X, R1192X, M1209V | (13, 14, 16, 155, 168–191) |
| Adcy6 | Adhesion of sickle red cells, arthrogryposis multiplex congenital with axoglial defects, autism, cardiac hypertrophy | A674S, R739W, Y992C, E1003K, R1116C | (192–198) |
| Adcy7 | Alcoholism, juvenile idiopathic arthritis, autism spectrum disorder, schizophrenia, ulcerative colitis | V272I, K349M, D439E, S1020G | (199–206) |
| Adcy8 | Autism spectrum disorder, bipolar disorder, dissociative disorder, schizophrenia, depression and alcoholism, myocardial infarction | L327V, V602I, A728P, R968Q | (152, 164, 168, 207–210) |
| Adcy9 | Asthma treatment, pancreatic cancer, body mass index, malaria susceptibility, schizophrenia, autism spectrum disorder, efficacy of dalcetrapib to treat cardiovascular events | K386X, I772M, S782G, N1154S | (154, 211–220) |
Missense and nonsense nonsynonymous variant sites are listed. The Human Gene Mutation Database was used to compile some of the information in this table (221).
3.1.2. Neural development.
AC1−/− mice display deficits in correct neural wiring during development, notably in the primary somatosensory cortex and visual systems (223, 224). Region-specific knockouts later demonstrated that thalamic AC1 deletion had more severe sensorimotor and social deficits than cortical AC1 deletions, suggesting a key role for thalamic AC1 in proper neuronal development (225). The AC1 knockouts, or region-specific knockouts in this study, were devoid of alterations in locomotor activity despite the observation that overexpression of AC1 in the forebrain resulted in 20% increased locomotor activity (143).
As the AC1 overexpression phenotype displayed some of the same behavioral characteristics as in the autism-like mouse Fmr1 knockout model, it was recently hypothesized that knockout of fragile X mental retardation protein (FMRP), protein product of Fmr1, may impact AC1 expression or function (226). Indeed, it was elegantly determined that loss of FMRP leads to increased AC1 mRNA and protein levels (227). Genetic and pharmacological inhibition of AC1 in this mouse model not only reduced the molecular pathology but reversed several of the autism-like behaviors in these mice (227). It should be noted that in a separate study exploring the interaction between FMRP and AC1 in pain, FMRP knockouts did not display increased AC1 levels and were actually protective against increased AC1 expression due to pain (228). However, the pain study was examining the anterior cingulate cortex (ACC), and the autism study explored AC1 expression in the hippocampus.
3.1.3. Pain.
There are significant studies exploring AC1 in pain, and foundational studies utilizing AC1−/− mice demonstrated an essential role for AC1 in chronic pain behaviors (229). AC1−/− mice exhibit reduced allodynia in the complete Freund’s adjuvant and formalin models of inflammatory pain but maintained acute nociceptive responses. Furthermore, recent studies have demonstrated the involvement of AC1 in bone cancer (230), models of diabetic neuropathy pain (231, 232), migraine (233), and visceral pain (234, 235). These studies demonstrated that although AC1 does not regulate acute nociceptive functions, it is critical to longer-term pain syndromes and provides a possible novel therapeutic opportunity for the treatment of chronic pain (FIGURE 6).
FIGURE 6.
The role of adenylyl cyclase (AC)1 in chronic pain-induced allodynia in vivo. In wild-type mice subjected to a chronic pain model, AC1 inhibitors can block allodynia. In wild-type mice that have recovered from a chronic pain model, AC1 inhibitors can block naltrexone (NTX)-induced allodynia (latent sensitization model). It is likely that stress-induced allodynia in chronic pain models will also be blocked by AC1 inhibitors. In AC1 knockout mice, chronic pain models do not produce allodynia.
AC1 plays a critical role in synaptic plasticity, and this function drives its physiological role in chronic pain. As a Ca2+/CaM-sensitive cyclase, increases in intracellular calcium promote increased AC1 activity. During chronic pain states, persistent neuronal signaling through voltage-gated calcium channels and N-methyl-d-aspartate receptors (NMDA-Rs) leads to increased AC1 activity, and these plasticity alterations in the ACC appear to maintain allodynic behaviors (236, 237). Furthermore, these alterations have been proposed to act in a positive feedback manner, as increased AC1 activity can reciprocally drive increases in the GluN2B subunit of NMDA-Rs (234) as well as increase the phosphorylation and expression of α-amino-3-hydroxy-5-methyl-4-isoxazolepropionic acid (AMPA) receptors (235, 238). The cycle of NMDA-R-driven AC1 activity and upregulation of NMDA-Rs and AMPA receptors leads to persistent activity that can drive long-term potentiation and result in “sensitized” neurons (FIGURE 7). Although many studies have investigated AC1 in the ACC for pain, AC1-induced long-term potentiation in the insular cortex is involved in the development of certain chronic pain models (233, 234, 238, 239) and spinal AC1 contributes to the transition and maintenance of long-term pain (240, 241).
FIGURE 7.

Cellular adaptations to chronic nociceptive input promote allodynia through adenylyl cyclase (AC)1. Top: sufficient acute nociceptive input results in activation of voltage-gated calcium (CaV) channels, α-amino-3-hydroxy-5-methyl-4-isoxazolepropionic acid (AMPA) receptors (AMPARs), and N-methyl-d-aspartate (NMDA) receptors (NMDARs). This activation at varying levels of the pain pathways (i.e., spinal cord, anterior cingulate cortex, insular cortex) can produce an acute behavioral response in rodents, typically a response to a noxious mechanical/thermal stimulus. Although calcium influx through these channels will activate AC1, AC1 does not modulate acute nociceptive responses. Bottom: chronic nociceptive input produces sustained calcium influx and AC1 activation. This promotes a positive feedback loop whereby AC1 activation results in phosphorylation of AMPARs and increased expression of GluN2B-containing NMDARs, both of which further promote increased AC1 signaling. This ultimately results in a sensitized system where nonnoxious stimuli produce a behavioral response (allodynia). Both panels are generalized schemes and likely occur at multiple levels of the pain pathways. Mu opioid receptor (MOR) agonists can inhibit AC1, among other functions. AC1 regulation by MORs in the acute nociceptive setting (top) is likely different from that in the chronic setting (bottom). In the chronic setting there is an interplay between NMDARs, MORs, and AC1 that are involved in latent sensitization (see text).
There is also significant evidence for the contributions of AC1 in a phenomenon known as latent sensitization (240, 242). In one study, animals were subjected to inflammatory pain and then allowed to recover from the pain model over several weeks (240). After recovery, injection of naloxone caused the hyperalgesia and allodynia to return, an unmasking of latent pain. This phenomenon is mediated through constitutive activity of the mu opioid receptor inhibiting AC1 activity. It was also demonstrated that this naloxone-precipitated pain could be effectively blocked by inhibiting AC1 activity or the NMDA receptor. This mechanism further emphasizes the intricate interactions between NMDA receptors and AC1 and how the communication between the two signaling molecules regulates the transition to chronic pain in various areas along the pain neuroaxis through synaptic plasticity (FIGURE 7).
The significant contribution of AC1 to the development of chronic pain, but not acute protective nociceptive functions, provides opportunity for novel nonopioid pain therapeutics. Indeed, drug discovery efforts for inhibitors of AC1 have generated two novel compounds (NB001 and ST034307) that can effectively alleviate chronic pain by inhibiting AC1 in the spinal cord and/or the ACC (243, 244). However, it should be noted that the mechanism of action of NB001 appears to be indirect to AC1 and potentially through CaM interactions (128, 244). Redox reactions can alter CaM activity to fine-tune the regulation of AC1, providing an opportunity for the development of small-molecule inhibitors that leverage CaM-AC1 to treat chronic pain (245). Although there is substantial evidence that AC1 plays a critical role in the development of various chronic pain syndromes, all of it is derived from knockout animal studies or pharmacological manipulations. It is critical that new and novel tools be developed to rigorously validate AC1 regulation and function in chronic pain.
3.1.4. Other physiological functions.
Besides the more established literature on synaptic plasticity, affective behaviors, and pain, AC1 has also been linked to drugs of abuse. AC1−/− mice had reduced opioid withdrawal symptoms including wet dog shakes, paw tremor, diarrhea, ptosis, and weight loss compared with wild-type mice (246). More recently, it was revealed that AC1−/− mice have a reduction to ethanol-induced locomotor sensitization and reduced GluN2B-NMDA phosphorylation following repeated alcohol administration (247). Other functions of AC1 include being an effector component of a contrast sensitivity circadian rhythm mechanism in retinal ganglion cells (248, 249). The two studies from this group determined that AC1 levels in a subset of retinal ganglion cells are modulated on a circadian rhythm through dopamine D4 receptor (DRD4) activating clock gene Npas1. Knockouts of DRD4 or Npas1 did not have rhythmic AC1 levels, and AC1−/− mice had reductions in sensitivity contrast. A single study has also demonstrated a conserved role for AC1 in hearing, where a mutation in the carboxyl tail caused a mild hearing impairment in humans, and this mutation-induced phenotype can be recapitulated in zebrafish (146).
Beyond the brain, AC1 has been linked to the cholesterol efflux pathway in macrophage foam cells (250). Macrophage foam cells are the main cells found in atherosclerotic lesions. AC1 protein and mRNA levels were consistently upregulated during cholesterol loading and apoA-1 activation. Knockdown with short hairpin RNA (shRNA) of AC1 reduced apoA-1-induced cAMP production and decreased cholesterol efflux while intracellular cholesterol remained high. It was suggested that AC1 is the main effector for apoA-1-mediated cholesterol efflux (transport and exocytosis) in macrophage foam cells. This is significant because cholesterol efflux is important for reducing the risk of atherosclerosis. AC1 has also been implicated in chloride secretion in airway epithelial cells, an important function in the pulmonary system (251). This study provided evidence of the colocalization of GPCRs, cystic fibrosis transmembrane conductance regulator (CFTR), TMEM16A, AC1, and EPAC1 in microdomains. AC1 and EPAC1 were responsible for mediating the cross talk between Ca2+ and cAMP signaling, and this cross talk activated CFTR and TMEM16A to amplify electrolyte secretion in airway epithelial cells. AC1 regulation by CaM appears to play roles in the physiological effects of both Bordetella pertussis (whooping cough) and Bacillus anthracis (anthrax) infections, raising the prospect that inhibitors of CaM stimulation of AC1 could be a viable therapeutic approach for treating these infections (252, 253).
3.2. Adenylyl Cyclase 8
3.2.1. Central nervous system functions.
AC8 is also largely studied for its role in the brain. The synaptic regulation of AC1 and AC8 has been extensively reviewed (254). Briefly, AC8 knockout (AC8−/−) mice, like AC1−/− mice, had slight deficits in mossy fiber LTP in the hippocampus (255). They displayed normal acute synaptic function and normal perforant path LTP. However, AC8−/− mice also displayed deficits in long-term depression (256) and short-term plasticity measures, whereas AC1−/− mice did not (255). They also demonstrated normal contextual and passive avoidance memory and locomotor activity (257). AC8−/− mice had impaired object recognition memory, but their spatial recognition memory appeared intact (257). However, when AC8−/− mice were tested in a more rigorous and rapid acquisition spatial memory task, animals took longer to find the relocated platform in the Morris water maze. This suggests that AC8 has an important role in rapid novel spatial recognition memory and working/episodic-like memory (257).
Early studies with these AC8−/− mice suggested that they display normal baseline anxiety behavior and were resistant to repeated stress-induced anxiety (256, 258, 259). One of these studies confirmed previous findings that AC8−/− mice have reduced stress-induced and anxiety-like behaviors, although the data for social defeat were not significant (259). They also noted an interesting increase in spontaneous activity in the home cage of mice during dark phase (wake hours), a significant decrease in leptin content, reduced stress-induced weight loss and feeding, and a significant enlargement of the adrenal glands through an unknown mechanism. A long-term anxiety-like paradigm also demonstrated a role for AC8 in anxiety but not for AC1 (258). AC8 has a demonstrated role in stress-induced behaviors and anxiety in mouse models, consistent with human studies that have identified SNPs within the AC8 gene that are associated with bipolar disorder (207, 208), posttraumatic stress disorder (209), and alcoholism comorbid with major depressive disorder (168).
3.2.2. Cardiac functions.
Recent studies of AC8 have focused on its role in the sinoatrial node (SAN) of the heart. Via cardiac-specific overexpression of AC8 in mice, it was demonstrated that these mice displayed increased heart rate and decreased heart rate variability (260). They revealed that heart rate and rhythm can be intrinsically driven by AC8 when overexpressed in SAN cells, independent of autonomic drive. It was suggested that the mechanisms behind this cardiac physiology may be due to the modified catecholamine synthesizing and degradation enzyme levels, plasma catecholamine levels, desensitization of β-adrenergic signaling, or blunted vagal input. These results have physiological significance in maintaining heart health and survival, and the authors suggested that further studies may provide evidence for pharmacological intervention to treat arrhythmias and heart failure. It was noted that these modifications due to AC8 overexpression resemble the aging heart, and longer studies are needed to determine whether AC8 overexpression may lead to early heart failure. The age of the mice was not noted in this study. Another study using transgenic AC8 overexpression mice investigated age-related cardiac function (261). AC8-overexpressing mice have increased contractility, reduced strain rate, but accelerated age-related heart dysfunction, early cardiomyocyte hypertrophy, and interstitial fibrosis (261). The early effects of aging were associated with subcellular remodeling of cAMP/PKA pathway. Both studies implicate chronic AC8 overexpression as detrimental to heart function. However, studies using AC8 knockdown are needed to confirm a role for this isoform in cardiac function.
3.2.3. Calcium regulation.
AC8 regulates calcium levels in cells, specifically store-operated calcium (SOC) entry, through interactions with a channel called Orai1 (262, 263). Orai1 functions as an SOC channel, allowing calcium entry into the cell after internal endoplasmic reticulum stores are depleted. AC8 regulated Orai1 through a direct interaction between the NH2 terminus of AC8 and the NH2 terminus of Orai1 that modified the phosphorylation state of Orai1 (263). It was later determined that the interaction of AC8 and Orai1 positions AC8 to be readily activated, activating downstream PKA, which is then positioned to phosphorylate Orai1 to inactivate it (264). A recent study further investigated AC8’s involvement in this pathway in cancer cell lines. This study (265) suggested an interesting countermechanism of AC8 and Orai1 regulation where overexpression of AC8 and Orai1 variants results in AC8-mediated enhancement of SOC, not reduction as previously suggested (264).
3.3. Adenylyl Cyclase 1 and 8 Redundancies
AC1 and AC8 are often studied in tandem via double-knockout studies, which have revealed many overlapping functions between the two isoforms. Both AC1 and AC8 are Ca2+/CaM sensitive and exhibit a similar expression pattern with overlap in cortical layers, the hippocampus, amygdala, thalamus, striatum, and brain stem regions (266). Thus, one isoform can often take over the functions of the other, where the single knockout displayed no phenotype but the double-knockout mice displayed significant alterations. This was clearly demonstrated when observing the outcome of AC1−/−, AC8−/−, or double knockout (DKO) on late-phase long-term potentiation (L-LTP) in the CA1 region of the hippocampus (139). Whereas single-knockout mice alone had no change in L-LTP, the double-knockout mice lost all L-LTP. These deficits were paired with deficits in long-term memory. Double-knockout mice had reduced passive avoidance learning and reduced contextual learning but not cued learning and memory. A further characterization of AC1 and AC8 double-knockout mice revealed no change in anxiety-like behavior in the elevated plus maze, in the light/dark preference test, or in open field (267). Surprisingly, the male double-knockout mice had significantly reduced immobility in the forced swim test (antidepressant-like behavior) whereas females did not, and both males and females interacted more in a social behavior test. Both female and male double-knockout mice had a reduced sucrose preference and reduced body weight. However, it was noted that double-knockout mice were often compared to non-littermate control animals because of the large numbers of mice required, making it difficult to account for differences in maternal behavior.
Double-knockout studies have also revealed roles of AC1 and AC8 in the effects of a multitude of drugs of abuse including ethanol consumption and locomotor responses (247, 268–271), cocaine sensitization (272), opioid tolerance and withdrawal (246, 273), and methamphetamine activity (274). A recent double-knockout study further investigated the role of AC1 and AC8 in the reward areas in the brain (269). Manganese-enhanced magnetic resonance imaging (MEMRI), a novel brain imaging technique in live animals that maps the entry of manganese into active neurons through calcium channels, revealed that AC1/AC8 double-knockout animals had reduced activity in reward areas. These areas included the ACC, nucleus accumbens, medial prefrontal cortex, anterior caudate putamen, medial thalamus, and ventral lateral thalamus. This hypoactivity was linked to a significant decrease in ethanol consumption and preference over time and a range of ethanol concentrations. These results provided an interesting whole brain imaging approach to link previously determined changes in ethanol drinking behavior in double-knockout mice to various brain regions in live animals.
AC1 and AC8 are thought to be present in SAN cells and appear to be involved in SAN pacemaker activity (275–277). The double-knockout mice have altered Ca2+ homeostasis and sensitivity in the SAN (276). These changes altered automaticity and pacemaker current, suggesting the Ca2+-stimulated AC1 and AC8 isoforms are important for coupling calcium signaling to pacemaker current. More specific means of pharmacological inhibition, more specific antibodies, and utility of single AC knockouts are needed to provide more informative measures of AC1 and AC8 regulation in the SAN. The unclear RNA expression profile of AC1 and AC8 in the SAN emphasizes this point. In guinea pigs AC1 is expressed in the SAN, but there is very low expression of AC8 (277). In one mouse strain, C57BL/6, there is no detected AC8, but AC1 is not examined (278). A separate RNA sequencing (RNA seq) study of SAN cells in transgenic mice in the CD1(ICR) background detected AC1 but not AC8 RNA; however, AC1 was not enriched in the SAN compared with the right atrial myocardium (279). As mentioned above, a recent overexpression study determined that overexpression of AC8 in SAN cells can increase heart rate and decrease heart rate variability (260). These results suggest that AC1, and potentially AC8, may be central to maintaining basal pacemaker current, but the overexpression study suggests an important role for either Ca2+/CaM-stimulated cyclase activity or cAMP in general for maintaining heart rate and rhythm and potentially long-term cardiac dysfunction (261).
3.4. Adenylyl Cyclase 3
3.4.1. Olfaction.
AC3 was originally thought to be an olfactory-specific adenylyl cyclase (280) but was later identified in a variety of other cell types (281). It is now accepted that AC3 is a predominantly cilium-specific adenylyl cyclase and can be found on primary cilia throughout the brain (282). AC3 is integral for proper olfaction (283), and this role has led to a number of studies on AC3 in terms of olfaction mechanisms and development (284–287), pheromone detection (288), maternal behavior (289), and olfactory-dependent learning (290, 291).
3.4.2. Depression.
It has been reported that AC3 is the primary adenylyl cyclase in human blood platelets (292), and studies have suggested that reduced AC3 activity in these platelets is a potential biomarker for major depression (293, 294). Although not significant, SNPs in AC3 itself were suggested to be associated with major depressive disorder (MDD) (156), as were lower blood transcript levels of AC3 (295). Recent studies have taken these human studies and determined whether this also translates in animal models of depression-like behavior.
Recently, a study tested AC3−/− mice against a battery of depression-like paradigms (296). AC3 knockouts had increased immobility in both the tail suspension and forced swim tests, decreased sociability, reduced novelty-suppressed feeding and drinking, a reduced coat grooming score, and decreased nesting scores, all indicative of depression-like behaviors. They also exhibited increased rapid eye movement (REM) sleep, altered non-REM sleep, hippocampal atrophy, and reduced signaling, impaired spatial navigation, and depressed L-LTP, all markers of depression-like features in humans. This study also examined an inducible AC3−/− mouse model and a conditional knockout model only in the forebrain. Many of the same depression-like symptoms were recapitulated in these models, as in the germline knockouts. This study further validates that reductions in AC3 may result in depression as the human studies suggested.
A very recent study determined that deletion of AC3 from somatostatin (SST)-positive cortical interneurons was sufficient to induce depression-like and anxiety-like behaviors in mice (297). Deletion of AC3 in parvalbumin (PV)-positive interneurons failed to result in a similar phenotype. Confirming previous findings, AC3−/− mice display a plethora of depression-like symptoms (298). This was paired with a reduction in hippocampal tyrosine hydroxylase (TH), dopamine D1 receptor (DRD1), and dopamine D3 receptor (DRD3) protein levels. Dopamine D2 receptor (DRD2) levels were reduced, but not significantly. Interestingly, when this group conditionally deleted AC3 from the main olfactory epithelium (MOE), mice displayed a depression-like phenotype somewhat similar to the global knockout, although food intake was unaffected. The global AC3 knockout effects on hippocampal protein levels of TH and DRD3, and the trending reduction on DRD2, were recaptured in the MOE AC3 knockout. MOE AC3 knockouts also had reduced GluN2B protein levels. Future studies are needed to rigorously determine the mechanisms of AC3 in depression because multiple AC3 knockout paradigms display a similar phenotype likely affecting different neuronal functions.
3.4.3. Metabolic regulation.
An early study found that AC3−/− mice become significantly obese, with males increasing in mass by 40% and females increasing by 70% (299). These AC3−/− mice had significantly more fat versus lean mass, larger adipocytes, and increases in triglycerides, leptin, and insulin. At young ages, AC3−/− mice were less active and consumed more food, and at old ages they continued to be less active. These physiological changes were hypothesized to be mediated through loss of AC3 expression in primary cilia in the hypothalamus (299), a critical region for food intake regulation (300). This group later verified this hypothesis by deleting AC3 from the hypothalamus and demonstrated an increase in food consumption and body weight (301). A specific ventromedial hypothalamus deletion resulted in the same effect (301), although a different subregion of the hypothalamus, the paraventricular nucleus, has also been implicated in the body weight regulation of AC3 (302).
Recent studies have continued to build on these previous studies demonstrating AC3 regulation of metabolism. Haploinsufficiency of the ADCY3 gene in AC3−/+ mice produced a phenotype similar to AC3−/− but was dependent on the type of diet mice were fed (303). When fed normal mouse chow, heterozygous AC3 mice do not gain weight. When fed a high-fat diet, these mice gained significant weight, while they consumed the same amount of food as wild-type mice. Furthermore, when fed a high-fat diet, AC3−/+ mice have increased triglycerides, leptin, and other hepatic biochemical parameters. Additionally, they display impaired glucose tolerance and insulin resistance. It is intriguing that heterozygous mice display a metabolic profile similar to AC3 knockout mice, but only when fed a high-fat diet. Conversely, mice with a gain-of-function mutant in AC3 display resistance to diet-induced increases in body weight, leptin content, and insulin content (304).
These mouse studies have substantial translational bearing in humans. Variants of the ADCY3 gene have been linked to increased body mass index (BMI, corrected for height) in children (157); increased type 2 diabetes, BMI, fat percentage, fasting plasma glucose, and acute plasma glucose in a Greenlandic cohort (158); and increased BMI in obese adult subjects (159). The human genetic data, coupled with the animal models and AC3 location, provide compelling evidence for AC3 as a unique therapeutic target for obesity disorders. Interestingly, potential activators of AC3 may confer benefit for both obesity and depression.
4. GROUP II (AC2, AC4, AC7)
4.1. Adenylyl Cyclase 2
Reports of AC2 knockout or overexpression are scarce. One previous study in human airway smooth muscle cells suggests a role for AC2 in mediating airway smooth muscle tone in response to prostaglandins (305), but research to directly verify AC2’s role in physiological functions of airways with knockdown or genetic deletion has not been reported. AC2 activation in the brain via Gβγ and Gαs may be associated with chronic opioid use and tolerance (306), but the specific role of AC2 is inferred by the Gβγ and Gαs dependence, so AC4 may also be involved in these responses. One major hurdle to understanding AC2’s physiological role is the expression pattern variation between mice and humans. AC2 expression has been observed in human airway smooth muscle cells, but the same expression pattern is not found in mouse airway smooth muscle (305, 307). Studies of ADCY2 polymorphisms suggest roles for AC2 in tobacco-exposed lung function, chronic obstructive pulmonary disease, bipolar disorder, and schizophrenia (TABLE 3) (148–150, 308), but, as with other cases, specific knockdown studies have not been performed to directly establish mechanistic links between this AC isoform and these diseases. AC2-selective inhibitors have been reported and should aid in studies of the physiological roles of AC2 (126, 309, 310). A study using an AC2-selective inhibitor in a cell model of Lesch–Nyhan disease suggests that AC2 plays a role in the pathogenesis of this neuropsychiatric disorder (311). Inducing a deficiency in hypoxanthine phosphoribosyl transferase in cells, which mimics Lesch–Nyhan disease, leads to an overall reduction in AC activity and a specific loss of AC2 expression (311, 312).
One recent study examined the gene expression profiles of young and adult mice in a schizophrenia model (313). Differential expression of several genes involved in calcium homeostasis and signaling along with ADCY2 were identified. ADCY2 expression was reduced in both juvenile and adult schizophrenic mice compared with control animals. These results imply a role for AC2 in schizophrenia pathology. These investigators predict specifically that ADCY2 plays a role in early psychotic symptoms relating to learning and memory. Specific manipulations of AC2 expression are needed to confirm these mechanistic linkages, and an examination of ADCY2 expression in humans with schizophrenia is required before AC2 can be declared a therapeutic target for the disease.
4.2. Adenylyl Cyclase 4
Reports of knockout and overexpression studies of AC4 are also limited in the literature, as are ADCY4 polymorphism studies. One previous tissue-specific AC4 knockout has been reported in kidney collecting duct principal cells, but the results indicate that AC4 does not play a role in the collecting duct’s handling of water or sodium (314). Interestingly, AC4 deletion left vasopressin-stimulated cAMP production unaltered in this region of the kidney, an indication that vasopressin receptors only couple to one of the other AC isoforms (AC3 and/or AC6) expressed in the collecting duct. AC4 appears to have a role in dopaminergic regulation of prefrontal cortical neurons (315). Dopamine-stimulated cAMP production was potentiated by simultaneous activation of M1 muscarinic acetylcholine receptors in neuronal primary cultures. This potentiation depended upon concomitant stimulation of Gβγ and Gαs, a classical feature of group II ACs. Knockdown of AC4 abrogated this effect, whereas knockdown of AC2 was inconsequential. Although these studies imply that AC4 is involved in attention, cognition, and emotion, there are as yet no direct studies linking AC4 to these functions. AC4 expression is high throughout the heart and vascular system (57), yet no studies of its physiological roles in these systems have been reported.
Another recent study implicates a potential role of AC4 in regulating lipopolysaccharide- or Escherichia coli-mediated sepsis (316). ADCY4 is expressed in macrophages and participates in the α2b-adrenoceptor-PDE8A-PKA signaling pathway that underlies the inhibition of caspase-11 inflammasome activation by epinephrine. AC4 knockdown reduced epinephrine-stimulated cAMP synthesis and inhibition of lipopolysaccharide (LPS)-stimulated inflammasome activation, whereas knockdown of the highly homologous isoform, AC2, lacked these effects. Although AC4 was not directly linked to sepsis in mice, pharmacological inhibition of PDE8A did protect mice from caspase-11-mediated sepsis and septic death, implying a central role of the entire α2b-adrenoceptor-AC4-PDE8A-PKA signaling pathway. Thus, AC4 may be a relevant therapeutic target in sepsis.
A recent epigenetic study may implicate AC4 in breast cancer pathology (317). ADCY4 was significantly downregulated in breast cancer, and high expression of the gene in breast cancer patients correlated with better survival. This suggests that AC4 may act as a tumor suppressor or restrain metastasis. The expression levels of AC4 correlated negatively with methylation of the ADCY4 promoter, with hypermethylation observed in breast cancer patients. This study indicates that AC4 expression could be used as a biomarker for breast cancer, and it warrants further research into AC4 as a possible therapeutic target.
4.3. Adenylyl Cyclase 7
Previous studies have demonstrated roles for AC7 in several different physiological processes and suggest that alterations in the activity of this isoform may lead to diseases. Particular attention has been devoted to studying alterations in the immune system and neuropsychiatric disorders such as alcoholism and depression (318). More recent studies have also indicated that AC7 plays a role in certain cellular processes associated with cancer (319, 320).
The regulatory properties of AC7 are similar to those of other group II adenylyl cyclases. AC7 possesses the Gln-X-X-Glu-Arg motif that should allow for Gβγ interaction and conditional activation of the cyclase by the dimer, but this has not been directly tested in vitro (92, 321–323). Moreover, AC7 is insensitive to Gαi/o, activated by PKC, conditionally activated by Gα12/13, and, interestingly, potentiated by ethanol (92, 318, 322, 324–326). The relationship between AC7 function and ethanol has been the focus of a number of studies. At the molecular level, ethanol’s conditional activation of AC7 is insensitive to pertussis toxin but suppressed by PKC inhibition (327, 328). Ethanol appears to directly interact with AC7 but may require additional proteins for its stimulatory effect in certain cellular environments (329, 330). PKCδ phosphorylates AC7, and this step is required for the potentiation of AC7 activity by ethanol in HEK and HEL cells (327). Transmembrane regions of AC7 are not required, as ethanol can still enhance the activity of mutants of AC7 that were created lacking the transmembrane domains (330, 331).
Even though AC7 knockout is largely lethal to mice (324), several AC7 expression studies have implicated a role for AC7 in alcoholism. For instance, ethanol-induced increases of blood adrenocorticotropin (ACTH) and corticosterone are decreased in heterozygous ADCY7+/− female mice and increased in female mice overexpressing AC7 compared with wild-type animals (332). In addition, ethanol preference is increased in ADCY7+/− female mice compared with wild-type mice (333). In humans, the presence of a SNP in ADCY7 has been correlated with a reduction in the risk of developing alcohol dependence for females (333). However, in contrast to the mouse data, this SNP is linked with lower levels of ADCY7 transcripts in blood and adipose tissue compared with its major allele (333). It would be interesting to measure the levels of AC7 protein expression in the brain of subjects with this SNP to determine whether the same pattern would be observed. AC activity in platelets from men with alcoholism is significantly decreased compared with matched control subjects (334, 335). Although multiple AC isoforms are likely to be expressed in platelets, AC7 is the main isoform in the platelet-derived HEL cells (199). As this study was done with platelets, it presents a remarkable contrast with the AC7 blood transcript levels from the SNP study. However, a constant feature of AC7 studies has been the presence of sex-specific effects. As noted above, the study with human platelets was done in male subjects, whereas in the mouse studies and the human SNP study significant results were only obtained with females (332–335).
The role of AC7 in depressive disorders has recently been reviewed (318). In summary, the available studies indicate that higher AC7 activity is linked to depression. One study used ADCY7+/− mice and mice overexpressing AC7 in the central nervous system (CNS) to demonstrate that in female mice lower AC7 expression correlates with a reduction in behaviors that are linked to depression in the forced swim and tail suspension tests (200). However, no significant effects were observed in male mice. The amygdala is a central component of the limbic system that has an important role for emotion and motivation control (336). Notably, the levels of AC7 mRNA are increased in the amygdala of male humans with familial major depression and in a mouse model of depression that lacks the serotonin transporter compared with their respective control subjects (201). A genetic polymorphism in ADCY7 has also been found to correlate with an increased risk of familial depression in women (TABLE 3) (200). Human carriers of that polymorphism display increased threat-related amygdala reactivity in comparison to control subjects (201).
cAMP plays an important role in the regulation of immune responses. For instance, the activity of T and B cells, neutrophil chemotaxis, and cytokine production have all been related to changes in cAMP levels and, therefore, with AC activity (337–342). Genetic deletion of AC7 in the hematopoietic system results in hypersensitivity to lipopolysaccharide (LPS)-induced endotoxic shock, increased production of the inflammatory cytokine TNF-α by bone marrow-derived macrophages (BMDMs) in response to LPS, and deficient antibody responses to antigens (337). The study also highlighted the importance of AC7 for cAMP responses in B and T cells, particularly in T cell-dependent antibody responses, which are crucial for adaptive immunity (337). As further discussed below, it should be noted that there are other AC isoforms expressed in those cells and AC9 activity has also been implicated in T cell function (338).
Several other studies have established a role for AC7 in the production of TNF-α. When an inflammatory response was triggered, TNF-α production in AC7-deficient BMDMs was markedly increased compared with wild-type BMDMs (343). In other studies, knockdown of AC7 in a monocytic leukemia cell line caused an increase in TNF-α production in response to LPS (344). In humans, this correlated with an LPS-induced increase in plasma TNF-α concentrations and a decrease in AC7 transcripts in primary monocytes from male subjects (344). It is noteworthy that sex differences were also observed in this study, with LPS-induced production of TNF-α being more pronounced in young males than in young females. Furthermore, the drop in AC7 transcripts was not observed in blood samples from females (344). Additional evidence for the role AC7 plays in the human immune system comes from a genomewide association study identifying a single-nucleotide polymorphism (SNP) in one of the intron regions of the ADCY7 gene that is associated with two pediatric-age-of-onset autoimmune diseases, psoriasis and Crohn’s disease (TABLE 3) (202).
The conditional activation of AC7 by Gα12/13 has been implicated in the enzyme’s function in immune responses. Specifically, activation of Gα12/13-coupled receptors potentiates Gαs-stimulated AC7 activity in BMDMs (324, 325). Upon treatment with zymosan, a cell wall extract from Saccharomyces cerevisiae that induces inflammatory responses, BMDMs display a robust increase in cAMP production. Notably, that response largely depends on the expression of AC7 (343). Knockout of Gα12 and Gα13 in those cells causes a profound reduction in PKA activity in response to zymosan. Pertussis toxin treatment also decreases PKA activity in that context, but the decrease is less pronounced (343).
The activity of AC7 has also been linked to certain processes related to cancer. For instance, in solid tumors, inadequate oxygen supply results in activation of the hypoxic pathway. Hypoxia is generally linked to tumor resistance to therapy, robust malignancy, and, thus, poor prognosis (345, 346). Carbonic anhydrase IX (CA IX) is an important mediator of cell survival and malignancy in hypoxic conditions (347). Notably, the activity of PKA is essential for CA IX activation, indicating a possible role for specific AC isoforms in the process (348). In four different carcinoma cell lines, AC6 and AC7 mRNA expression was increased in hypoxic versus normoxic conditions (319). Moreover, the upregulation of AC6 and AC7 is dependent on hypoxia-inducible factor-1 (HIF-1), which is also necessary for the expression of CA IX in hypoxic conditions (347). Knockdown of AC6 or AC7 under hypoxic conditions causes a significant reduction in cAMP accumulation and PKA activation and results in a decrease in migration capacity of the cells (319). These data suggest a role for AC6 and AC7 in the malignancy capacity of the cell lines tested.
AC7 has a role in acute promyelocytic leukemia (APL) cell differentiation (320). Knockdown of AC7 inhibits retinoic acid-induced differentiation of the human acute promyelocytic leukemic cell line NB4. Moreover, a microRNA (miRNA) that has been reported to participate in the progression of acute myeloid leukemia (miR-192) diminishes AC7 expression and lowers cAMP accumulation in NB4 cells (320). The study also showed that knockdown of miR-192 causes upregulation of retinoic acid-induced differentiation of NB4 cells and that upregulation can be disrupted by knocking AC7 down. Notably, in patients with relapsed APL the transcript levels of AC7 in bone marrow cells are significantly lower than the levels in patients with newly diagnosed APL (320). These results correlate with an increase in miR-192 levels in bone marrow cells of patients with relapsed APL versus patients with newly diagnosed APL (320).
It is interesting to contrast the meaning of the data in the manuscripts from Simko et al. (319) and He et al. (320). Simko et al. (319) found that the activity of AC7 is important for tumor cell migration in hypoxia, whereas He et al. (320) indicated that a drop in AC7 expression is linked to APL relapse. These two studies present opposite functions for AC7 in two different processes related to cancer. Whereas one associates increased AC7 activity with metastasis, the other links decreased AC7 activity with cancer relapse (319, 320). These data highlight the complexity of tumorigenic processes and also the selective nature of AC isoform functions within different processes related to the same category of diseases.
5. GROUP III (AC5, AC6)
5.1. Adenylyl Cyclase 5
AC5 is very broadly expressed in mammalian organs (11). Accordingly, it is not surprising that AC5 plays important roles in the regulation of multiple organ functions. These properties render AC5 an attractive drug target, but developing a clinically useful AC5 inhibitor has a number of challenges. First of all, and most importantly, there is a paucity of potent and highly selective AC5 inhibitors despite the efforts of many research groups (11, 227, 349–353). Second, the broad expression of AC5 in multiple organs (11) means that systemic application of any selective AC5 inhibitor would have physiological effects in many organs and numerous adverse effects. A solution to this problem could be the local administration of AC5 inhibitors, but such an approach is applicable only to a few organs such as skin, eye, and nose. Third, in some organs such as the heart AC5 inhibition is assumed to be therapeutically valuable (353), but in other organs such as the brain AC5 inhibition may be deleterious and cause multiple neuropsychiatric problems (354, 355). Although the therapeutic index of AC5 inhibitors may be narrow, targeting a drug to act only peripherally or only in the CNS may have therapeutic utility. A recent study suggests that annexin A4 inhibits AC5 but not AC6 (356), but at the present time it is difficult to envisage how this mechanism can be exploited pharmacologically. Based on these considerations, it is still too premature to seriously consider clinical applications of AC5 inhibitors. It has been proposed that certain AC inhibitors could be used to preferentially inhibit the elevated activity of AC5 mutants in a distinct form of human dyskinesia (357), but the insufficient selectivity of the currently available AC inhibitors is a serious concern (11).
AC5 deletion has been suggested to play a protective role in the heart, by protecting both against β-adrenoceptor (βAR) stimulation and chronic pressure overload as well as against age-related cardiac myopathy (358–362). However, this concept has been challenged (227) and still awaits confirmation by independent research groups. Intriguingly, AC5 deletion does not protect against Gαq-induced cardiac myopathy (363). AC5 mediates the interplay of calcium and cAMP signaling in the heart under both sympathetic and parasympathetic regulation (364). AC5 deletion has also been suggested to protect against metabolic diseases by increasing insulin sensitivity and glucose tolerance, leading to decreased body weight even on a high-fat diet (365). Deletion or inhibition of AC5 may also protect against oxidative stress and reduce aging (366, 367).
In vitro, in silico, ex vivo, and in vivo experiments investigated the effect of increased extracellular glucose levels (hyperglycemia) on a cAMP signaling pathway in arterial myocytes (368). It is already established that PKA and L-type Ca2+ channels are associated with vasoconstriction in response to hyperglycemia, but the mechanism is unclear. This study concludes that the pathway is dependent on AC5 activity. Glucose stimulates cAMP production via AC5 in arterial myocytes, and this activity induces the L-type Ca2+ channel via PKA mediation and is required for vasoconstriction. AC5 and L-type Ca2+ channels are spatially closely associated in arterial myocytes. This pathway was verified in two established diabetic animal models, which implicates AC5 in the vasomotor response in diabetes. The aforementioned study by Ho and coworkers (365), who observed that AC5 knockout mice have increased longevity due to increased energy metabolism, glucose tolerance, and insulin sensitivity, appears to corroborate these findings and support the conclusion that AC5 links glucose to vasoconstriction.
In contrast to the cardiovascular system and the metabolic system, in the central nervous system AC5 deletions tend to be detrimental, specifically in the reward circuitry of the brain and motor systems. AC5 deletions induce impaired stress response mechanisms, impaired striatal learning and plasticity, impaired motor coordination in a Parkinson’s-like phenotype, and increased alcohol consumption (354, 369–372). Of similar or even greater concern, AC5 deletion causes autism-like symptoms (355). Considering the fact that there is no effective pharmacological treatment of autism (373), such an adverse effect as a result of treatment with an AC5 inhibitor would be unacceptable. As a beneficial effect, AC5 knockouts display an analgesic response to both acute and chronic neuropathic pain (374), suggesting that AC5 inhibition could be exploited for pain treatment, at least conceptually.
Human ADCY5 polymorphisms have been associated with metabolic diseases including diabetes and obesity (TABLE 3) (169–174, 375,376). Most prominently, ADCY5 polymorphisms are associated with neuropsychiatric and central nervous system disorders, notably alcoholism, depression, familial dyskinesia, facial myokymia, dystonia, myoclonus, and choreoathetosis (13, 15, 16, 168, 175, 176). Of particular interest are mutations in the regions of AC5 that link the C1a and C2a catalytic/G protein binding domains to the transmembrane (TM) domains, which are associated with familial dyskinesias (FIGURE 1). Mutations at R418 in the coiled-coil helical domain of AC5 that links TM1 to the C1a domain, at A726 in the C1b domain, and at M1029 in the coiled-coil helical domain linking TM2 to the C2a domain all contribute to dyskinesias (13–16). These ADCY5 polymorphisms result in AC5 variants with elevated catalytic activity causing severe neurological symptoms, providing the conceptual basis for selectively abrogating the elevated catalytic activity with appropriate inhibitors (357).
AC5’s unique regulatory pattern of Gαi inhibition and Gαs stimulation is an important mediator in AC5’s physiological effects. Both G proteins have the ability to independently bind with AC5, opening up the possibility of a ternary complex. Recent studies used molecular dynamics modeling and simulations to examine the ternary complex of interactions of AC5 with Gαi and Gαs (377). Gαi profoundly affects structure and flexibility of AC (378, 379). The ternary complex of AC with Gαi and Gαs resembles an inactive conformation, suggesting that effects of Gαi are stronger than those of Gαs (377) The ternary Gαi:AC5:Gαs complex may contribute to coincident detection, particularly in corticostriatal synapses important for reinforcement learning and LTP (380). Modeling the dynamics of this complex reveals that it creates a low catalytic activity of the enzyme in the basal state, facilitating larger and synergistic increases in cAMP that enhance detection of coincident signals. Such coincident detection by AC5 is important for corticostriatal synaptic plasticity, especially for reinforcement and reward-based learning. However, as a caveat, it needs to be emphasized that molecular dynamics studies regarding the effect of Gαi on AC5 do not replace authentic structural (crystallographic) analyses. The latter studies still need to be performed.
Another recent study presents novel evidence of another physiologically relevant regulatory heterotetramer of AC5. The proposed precoupled complex would notably account for the coincident signal detection by Gαi and Gαs (381). Although there is previous evidence of G protein and AC precoupling, this study reports precoupling of transmembrane domains of ACs with transmembrane domains of a GPCR. Using a peptide-based approach, Navarro et al. propose a heterotetramer of an adenosine receptor (A2AR), a dopamine receptor (DRD2), two AC5 molecules, Gαi, and Gαs. When activated, there is rearrangement of membrane domains of the precomplex. The heterotetramer also allows for the canonical antagonistic interactions of Gαi and Gαs with AC5. These results are physiologically significant because AC5 plays key roles in both the A2AR and DRD2 pathways, which are important for neuropsychiatric drugs (multiple GPCR antagonists; previously referred to as antipsychotics) (382). Thus, AC5 may be inhibited indirectly via targeting the GPCRs proximally located in the signaling cascade.
This A2AR, DRD2, and AC5 homodimer heterotetramer complex has a proposed role in striatal neurons (383). The AC5 homodimer allows for the canonical antagonistic Gαi/Gαs interactions on AC5. The heterotetramer integrates both adenosine and dopamine signaling in the brain. The model can explain the physiological and behavioral responses of adenosine and dopamine in the striatopallidal neurons, including caffeine psychostimulation, Parkinson’s disease, schizophrenia, and substance use disorders.
AC5 also may play a role in polycystic kidney disease (PKD) (384). cAMP is known to contribute to PKD by stimulating cell proliferation and fluid secretion. Knockdown of AC5 decreased cAMP levels in renal epithelial cells. In collecting duct-specific Pkd-2 (PKD gene)-deficient mice, there is an increase in cAMP and AC5 mRNA levels in the kidney. AC5 and Pkd-2 double-mutant mice had decreased kidney injury, decreased kidney enlargement, decreased cyst index, and improved kidney function. cAMP levels were reduced in the kidneys of the double mutants. In renal epithelial cells, cilium length was decreased upon ablation of AC5 in the Pkd-2 mutant mice. Overall, these results imply that AC5 inhibition is a viable therapeutic approach for PKD, but, as stated above, selectivity of AC5 inhibitors and the widespread expression of this isoform make adverse effects a serious concern.
5.2. Adenylyl Cyclase 6
Structurally, AC5 and AC6 are very closely related to each other (11). Accordingly, it is not surprising at all that pharmacologically it has been very difficult if not impossible to develop AC6 inhibitors with selectivity relative to AC5 (385). Like AC5, AC6 is also broadly expressed in heart, brain, kidney, pancreas, and bone. Interestingly, despite the large structural similarities between AC5 and AC6 (11), AC5 and AC6 possess distinct functions. Specifically, AC6 deletion in the heart limits β1AR-stimulated ventricular contractions and calcium signaling (386–388). Although AC5 deletion was protective, AC6 overexpression is protective of heart failure, specifically by increasing cardiac response and contraction and decreasing detrimental left ventricular remodeling (389–392). This difference between AC5 and AC6 in the heart may be due to the fact that AC6 is uniquely regulated by constitutive Gαi activity. To help tease out these isoform-specific effects on cardiac contractility, AC5 and AC6 knockout mice were compared (393). In particular, it was studied whether Gαi inhibits cAMP-regulated contractile force in the ventricle in an AC isoform-specific manner (393). With the use of Förster resonance energy transfer (FRET) to determine cAMP levels, in hearts from AC5 but not AC6 knockouts cAMP increased after inhibition of PDE3 and PDE4 in cardiomyocytes. This indicates that AC6 is regulated by Gαi activity and helps maintain basal cAMP in the compartment that regulates contractility of the rat ventricle. When PDE3, PDE4, and Gαi are inhibited in the AC6 knockout mice, an increase in contractility occurs. These data demonstrate that the inotropic and diastolic effects are dependent on AC6 and that AC6 controls contractile function in a localized cellular compartment. The study also shows that Gαi inhibition of AC6 limits the cAMP increase via β1AR.
The beneficial effects of AC6 overexpression are not just due to increased cAMP signaling, as AC6 also improves cardiac myocyte calcium handling (394), likely through interaction with a specific phosphatase, PH-domain leucine-rich protein phosphatase (PHLPP), that regulates Akt (395). Human gene therapy trials delivering AC6-expressing adeno-associated virus to patients with heart failure improved left ventricular function, particularly in patients with nonischemic failure (396). This is at odds with studies in mice, where transgenic overexpression is not protective against chronic pressure overload, particularly in female mice, and where AC6 deletion is beneficial (397, 398). These differences have not been reconciled but may reflect critical differences between humans and mice. Another explanation may come from recent studies in which just the catalytic domains of AC6 were expressed in transgenic mice (399, 400). These mice were protected from pressure overload-induced systolic and diastolic dysfunction as well as the deleterious effects of isoproterenol infusion. This same group of investigators expressed a mutant form of AC6 that lacks catalytic activity (401). These mice also displayed less dysfunction after isoproterenol infusion. Thus, the physiological effects of AC6 in the heart appear to be more complex than just its ability to catalyze the synthesis of cAMP.
β2AR agonists are used to reduce airway smooth muscle tone for treatment in asthma and chronic obstructive pulmonary disease, and AC6 has been linked as the primary isoform mediating these effects by studies of AC6 knockout mice (402). In the brain, AC6 is also implicated in hippocampal learning processes and age-related neurodegenerative diseases (403). In bone, AC6 in the primary cilia is involved in loading-induced bone formation (404). In the pancreas, AC6 regulates amylase and fluid secretion (405). In marked contrast to ADCY5, ADCY6 polymorphism studies are limited and have only linked AC6 to sickle red cell adhesion (TABLE 3) (406).
In the kidney, AC6 functions in water homeostasis, renin secretion, urine osmolarity and output, and cyst growth in PKD (407–414). More recently, it was found that AC6 also plays an important role in regulating acid-base homeostasis (415). Globally deleted AC6 mice and renal tubule and collecting duct-specific AC6 knockout mice were used to study AC6’s role in acid-base homeostasis. AC6 knockout mice have lower urinary pH and higher blood pH at baseline, implying that AC6 is required for normal acid-base homeostasis. However, when renal tubule and collecting duct-specific AC6 knockout mice were examined there was no disturbance in baseline acid-base homeostasis, suggesting that extrarenal AC6 has a role in energy expenditure. When the investigators challenged mice with they observed alterations in blood pH, supporting the idea that AC6 plays specific roles in the kidney tubule and collecting duct that enable acid-base homeostasis.
AC6 may also play a role in inducing inflammatory responses and chronic pain in various diseases (416). Knockout of AC6 increased TNF release and inflammatory responses by BMDMs, indicating that AC6 is important in inhibiting TNF release. More recently, alpha7 nicotinic acetylcholine receptors (CHRNA7) were found to directly interact with and signal through AC6 to induce anti-inflammatory effects (417). CHRNA7-selective agonists reduced production of inflammatory mediators by macrophages stimulated with LPS. Knockdown of AC6 prevented the anti-inflammatory response to CHRNA7 agonists, and overexpression of AC6 promoted these responses. The authors were able to observe that AC6 activation by CHRNA7 agonists led to degradation of Toll-like receptor 4, accounting for the anti-inflammatory action. Prolonged activation of the inflammatory response is implicated in many chronic diseases such as rheumatoid arthritis and Crohn’s disease, so characterization of AC6’s role in regulating inflammatory pathways may be useful in developing new treatment options.
AC6 is also important in regulating ciliary length, with consequences for both the airway and intestinal systems (418). AC6 knockout mice have longer cilium length and decreased flow in the epithelial cells compared with wild type. The AC6 knockouts exhibit decreased levels of Kif19A, a depolymerizing kinesin that controls cilium length. In vitro studies teased out a pathway in which AC6 inhibits AMP-activated kinase (AMPK), suggesting that AMPK can be more readily activated in the AC6 knockout (418). The AMPK targets Kif19A for autophagy, and the decreased levels of Kif19A then lead to increased cilium length. This study implicates AC6 in an important pathway for regulation of cilial length and motility that could play a role in cilia-related disorders specifically of the respiratory tract, reproduction tract, brain ventricle, and middle ear function. However, the role of AC6 has yet to be directly examined in ciliopathies of these various organs. In addition, a recent study implies that AC3, not AC6, is the primary resident isoform in primary cilia (419).
In intestinal epithelial cells, AC6 is a necessary component of cholera toxin-induced diarrhea (420, 421). In cholera, diarrhea results from increased cAMP levels, which activate cystic fibrosis transmembrane conductance regulator (CFTR)-mediated chloride transport. Thomas et al. (420) first proposed a role of AC6 in cholera toxin-induced diarrhea by using RNA seq to identify AC6 as the predominant isoform that associates with CFTR in epithelial intestinal cells. These investigators then created epithelial AC6 knockout mice and observed an abolishment of the CFTR-dependent fluid secretion. Intestinal epithelial cell-specific AC6 knockout mice displayed similar effects (421). AC6 was essential for cAMP production leading to cholera toxin-induced diarrhea via the movement of water into the intestinal lumen. Together, these results suggest that AC6-CFTR complexes may mediate diarrhea in cholera, and inhibition of AC6 may be useful as a treatment.
Intestinal epithelial cell differentiation appears to be regulated by AC expression. Treatment of colon epithelial cells in culture with sodium butyrate induces epithelial differentiation, which coincides with a large decrease in expression of AC3, AC4, AC6, and AC7 (but not AC5 or AC9) (422). Similar expression patterns were also observed in colon epithelia from mice, albeit involving some different AC isoforms that likely reflect the species differences. Expression of Gαs was also reduced in these studies, implying that global cAMP signaling is altered during differentiation. The differentiation of epithelia appears to require a reduction in cAMP generating capacity, likely to liberate cells from the antiapoptotic effects of cAMP (422).
5.3. Adenylyl Cyclase 5 and 6 Redundancies
There are many similarities between the expression patterns and functions of AC5 and AC6 (11), leading them to be studied together in double knockouts. AC5 and AC6 are the predominant isoforms in the heart, so most functional redundancies have been studied in this tissue. Both AC5 and AC6 have been previously reported to act in the β1AR pathway. A recent study used both AC5 and AC6 knockout mice to investigate their roles in β1AR-mediated inotropic responses by the heart (423). The results indicate a functional redundancy between AC5 and AC6 in mediating β1AR inotropic responses. Pertussis toxin inactivation of Gαi reveals that both isoforms are subject to tonic inhibition by Gαi. In the AC6 knockout, β1AR switched from predominant AC6 coupling to AC5 coupling. This switch was associated with alteration in the PDE isoform regulating the cAMP pool, indicating that AC6 knockout results in reorganized signaling compartments.
AC5 and AC6 knockout mice have also been used to investigate their respective localizations in ventricular myocytes and their effect on L-type Ca2+ current (424). AC6 is differentially localized to the plasma membrane outside of the T-tubular region, and it enhances β1AR-mediated L-type Ca2+ current. In contrast, AC5 is localized to the membrane in T-tubular regions, and its influence on L-type Ca2+ current is limited by PDEs. AC5 is responsible for the β1AR enhancement of L-type Ca2+ current, further supporting the hypothesis of compartmentation of AC isoforms in the heart. AKAP5 participates in this βAR-AC5-PKA complex to regulate L-type Ca2+ current in the T tubule (424). Cardiac myocytes also express AKAP6 (also known as mAKAP), which interacts with AC5 to coordinate cAMP, calcium, and MAP kinase pathways to regulate cellular hypertrophy (26). These specific AKAP-AC interactions are critical to the spatial regulation of cAMP signaling and encoding specific cellular responses such as cardiac repolarization, calcium release, and cell stress responses (FIGURE 8). There is also evidence for caveolin-3 as a scaffolding protein for AC5 (426). This scaffolding interaction may be important for compartmentalizing the β1AR signaling pathway.
FIGURE 8.

Adenylyl cyclase (AC)-A kinase anchoring protein (AKAP) signaling complexes regulate specific cardiac functions. AC isoforms associate with specific AKAPs to form signaling complexes. These complexes are localized in specific structures and spaces within the cardiac myocyte, likely based on localization of AC. AKAP interactions with downstream effector and regulator proteins allow a complex to regulate unique physiological functions. Cardiac repolarization is facilitated by an AKAP9 (Yotiao) complex that includes KCNQ1, PKA, protein phosphatase 1 (PP1), and phosphodiesterase 4DE3 (PDE4D3). Disruption of the complex prolongs the action potential. Calcium-stimulated calcium release is regulated by a macromolecular complex consisting of AC5/6 and AKAP5 (AKAP79/150). AKAP5 anchors a cAMP regulatory unit in the transverse tubule, where it is in register with phospholamban (PLN), ryanodine receptors (RyRs), and the sarco(endo)plasmic reticulum calcium ATPase (SERCA). Molecules involved in cardiac hypertrophy, including protein phosphatase 2B (PP2B), PDE4D3, and phospholipase Cε (PLCε), are anchored by AKAP6 (mAKAP) to both AC5 in the transverse tubule and the nuclear membrane via nesprin. The close proximity of the transverse tubules and the nucleus allows functional coupling of the AC5 complex to nuclear calcium signaling. The cardiac stress response is further regulated by binding of AC9 to heat shock protein (HSP)20, facilitating PKA regulation of HSP20 phosphorylation and promoting cardioprotection. LTCC, L-type Ca2+ channel. Figure adapted from Ref. 425, with permission from Journal of Cardiovascular Development and Disease.
Collectively, both AC5 and AC6 are broadly expressed AC isoforms. Despite their structural similarities (11), they possess different functions, most prominently in the heart. The different functions of AC5 and AC6 may be attributable, at least in part, to differential cellular localization. Unfortunately, it is still impossible to therapeutically exploit the different functions of AC5 and AC6 because of the lack of AC6 inhibitors with high selectivity relative to AC5 (and, of course, AC isoforms from other AC families) (11). Several studies have exploited the Ca2+-inhibitable nature of AC5 and AC6 (427) to infer roles for these isoforms. Such an approach is unable to distinguish between AC5 and AC6 but does establish the involvement of these group III ACs. Endothelial cell gap junction formation and vascular permeability (428–431), regulation of vascular tone (432), and catecholamine-regulated renin secretion (408) are a few prominent examples.
A general problem in the AC field, and a problem particularly relevant for the AC5/6 field, is the paucity of well-characterized and AC isoform-specific antibodies. Some commercial AC5 and AC6 antibodies are available and have been used in certain studies (433, 434). However, in general, authors rely on company descriptions of selectivity. Rarely, antibodies are validated by studying corresponding AC knockout animals, AC knockdown experiments, or recombinant AC isoforms. This is the major reason for the fact that most studies on the (patho)physiological functions of AC5 and AC6 rely on AC5 and AC6 knockout animal studies. An issue not sufficiently known in the scientific community and contributing to the difficulties of valid studies with AC5 and AC antibodies is that the expression levels of these signaling proteins are very low compared with GPCRs and, particularly, G proteins (11).
6. GROUP IV (AC9)
6.1. Adenylyl Cyclase 9
Until recently, what little was known about physiological functions of AC9 was gleaned from ADCY9 polymorphisms and studies of microRNAs that potentially target AC9. Human miR-142-3p and miRNA-181-5p reduce AC9 expression and have implicated AC9 in sepsis (TABLE 3) (435), cervical cancer (436), promyelocytic leukemia (437), and immune function (438). It is of note that AC9 is expressed in many immune cells, including regulatory T cells, neutrophils, and monocytes, and reducing AC9 expression decreases Gβγ-mediated neutrophil chemotaxis (338, 439–442). ADCY9 polymorphism studies have also associated AC9 with asthma (211, 212), pancreatic cancer (213), liver cancer (443), mood disorders (444), body weight (214), malaria susceptibility (215, 216), and efficacy of the cholesteryl ester transfer protein (CETP) inhibitor dalcetrapib (217).
The link between the serum lipid-altering drug dalcetrapib and AC9 is complicated. Clinically, dalcetrapib (and to a lesser extent evacetrapib) had better efficacy and 20% fewer cardiac events when a particular polymorphism of AC9 was present in the second intron of the ADCY9 gene (217, 445). However, the presence of this polymorphism had no effect on the related drug anacetrapib within the REVEAL trial (446). A larger trial, Dal-GenE, will hopefully determine the actual benefit-risk ratio for this genotype (447). In mice, inactivation of ADCY9 by gene-trapping (AC9KO) gave rise to reduced atherosclerosis, macrophage accumulation, and plaque formation after treatment with an atherogenic protocol that consisted of AAV8 viral expression of a gain-of-function mutant of Pcsk9 followed by a high-cholesterol diet to induce hypercholesterolemia (448). AC9KO also increased body mass and fat mass after the atherogenic protocol. However, effects on at These results have implications both for protection from atherosclerosis and for treatment of atherosclerosis.
Mouse studies of AC9KO have also recently uncovered important roles for AC9 in heart, including regulation of stress responses, repolarization, and heart rate control (448, 449). AC9 activity is detected in SAN, and two different laboratories report bradycardia for AC9KO mice, suggesting a role in heart rate control (449). AC9 is also recognized for its ability to associate with the AKAP Yotiao, which is the smallest of the splice variants of AKAP9 (450). Yotiao also anchors PKA and the alpha subunit (KCNQ1) of the slowly activating delayed-rectifier K+ current (IKs), critical components for the late-phase repolarization of the cardiac action potential in humans during sympathetic stimulation (451). Mutations of KCNQ1 or Yotiao that disrupt their association can cause potentially fatal arrhythmias, known as long QT syndrome. In a transgenic mouse line that expresses the KCNQ1/KCNE1 components of IKs, AC9 was the only AC isoform to associate with Yotiao and KCNQ1 (452). Deletion of AC9 resulted in reduced isoproterenol-stimulated KCNQ1 phosphorylation in vivo and in decreased IKs currents in adult cardiomyocytes, suggesting that AC9 is necessary for the regulation of the IKs potassium current. By bringing together AC9, PKA, and IKs, Yotiao facilitates the generation of local pools of cAMP to activate anchored PKA, increase PKA phosphorylation of the channel, and enhance potassium currents necessary for cardiac repolarization (115) (FIGURE 8).
AC9 may also play an important role in basal cardiac stress responses, regulating the PKA phosphorylation of heat shock protein 20 (Hsp20) at baseline but not upon sympathetic stimulation (319). AC9 activity associates with Hsp20 in heart, suggesting that these proteins are in complex with one another. The close interaction of AC9 with downstream targets of PKA (e.g., KCNQ1 and Hsp20) may be required because of the very low levels of AC9 activity in heart compared with the major cardiac isoforms, AC5 and AC6 (319).
AC9 was originally categorized as forskolin insensitive (group IV), but several studies demonstrate that AC9 is weakly conditionally stimulated by forskolin in the presence of Gαs (72, 453). Based on cellular studies, AC9 is currently thought to be stimulated by Gαs, PKC βII, and calcium calmodulin kinase II (CaMKII) and inhibited by Gαi/o, novel PKC isoforms, and calcium calcineurin (78, 454–456). However, direct regulation of AC9 was limited to Gαs (although regulation by calcineurin could not be tested) (72). The other regulators likely either act indirectly or require additional scaffolding proteins in order to facilitate regulation. AC9 can also form homodimers and can heterodimerize with Gαi-regulated AC5 and AC6 when overexpressed, potentially adding to the complication of interpreting cellular overexpression studies of AC9 (72).
Interestingly, a regulatory method of autoinhibition for AC9 has been proposed with physiological implications (457). A short motif in the C2b regulatory domain of AC9 was identified that reduced Gαs activation. Suppression was lost upon proteolytic cleavage of AC9 at the COOH terminus. The resulting truncated AC9 protein was prevalent in the left ventricle of the heart, as detected by antibodies. This unusual mechanism was further elucidated with the solution of the cryogenic electron microscopy structure of full-length AC9 bound to Gαs (10). Density from a region from the COOH-terminal C2b domain of AC9 (aa 1246–1275) was present within the active and forskolin binding sites. By comparing a truncated AC9 with the full-length protein, it was clear that this region had little effect on basal activity, but its presence dramatically increased the Km for ATP by 18-fold when the enzyme was activated by Gαs (48 vs. 880 μM ATP). Thus, proteolytic cleavage of AC9 may regulate GPCR activation of AC9 activation in heart (457, 458).
Another mechanism of regulating GPCR activity relates to the trafficking of GPCRs. All transmembrane ACs are present on the plasma membrane (PM), although localization to internal membranes has certainly been reported (reviewed in Refs. 11, 459). This is consistent with the need for GPCRs to couple G protein activation and AC regulation at the PM. However, upon ligand binding, many GPCRs internalize yet continue to activate G proteins and enhance cAMP accumulation at endomembrane sites (131) (reviewed in Ref. 116). AC9, but not AC1, can internalize with hormone-stimulated β2AR (460). Although both the GPCR and AC9 utilize a dynamin-dependent membrane pathway, AC9 traffics independently of the GPCR. Gαs activation was sufficient to promote AC9 accumulation in endosomes; the trafficking of AC9 did not require β-arrestins or AC activity (FIGURE 9). Thus AC9, and potentially other AC isoforms, may promote cAMP generation from internal sites to differentially regulate downstream events.
FIGURE 9.
Internalization and endosomal signaling by adenylyl cyclase (AC)9 but not AC1. β2 adrenergic receptors (β2ARs) activated by agonist stimulate cAMP at the plasma membrane but then internalize via arrestin-mediated endocytosis into endosomes. When complexed with AC1, the receptor ceases cAMP signaling once internalized, creating a short-duration cAMP signal delimited to the plasma membrane. When complexed with AC9, the receptor internalizes with the AC and continues to generate cAMP in the endosomal compartment (131, 460). AC9 internalization occurs independently of the receptor and does not require arrestins or AC activity (460).
7. GENETIC REGULATION OF AC ISOFORMS
The genetic structure and chromosomal location of the different AC isoform genes have been studied (6, 56, 461–466). However, details about the genetic regulation and promoters of those genes are still largely missing (48, 56). Studies employing sequence homology and expressed sequence tags (ESTs) have shown that human AC genes are formed of 11–26 exons and have a length of 16–430 kb (56). The genetic organization of AC isoforms appears to be well conserved among group II and III ACs, with ADCY5 and ADCY6 presenting 21 exons and ADCY2, ADCY4, and ADCY7 being composed of 25, 25, and 26 exons, respectively. The numbers of exons and introns are less conserved among group I ACs, with 20, 22, and 18 exons for ADCY1, ADCY3, and ADCY8, respectively. ADCY9 has 11 exons (56).
Regarding the promoter regions of human AC genes, ADCY2, ADCY3, ADCY5, ADCY6, ADCY8, and ADCY9 do not contain a TATA box, suggesting that the expression of these isoforms is not regulated by TATA box-binding factors (56). ADCY1, ADCY2, ADCY3, ADCY4, ADCY5, and ADCY9 present putative GC boxes, which may be regulated by transcription factor specificity protein 1 (Sp1) (56, 467). In addition, the translation initiation ATG codons are only within the context of good Kozak sequences for ADCY6 and ADCY9, which may result in lower expression levels of the other AC isoforms. The promoter regions of rat ADCY3 and mouse ADCY8 possess a CRE motif and therefore can be regulated by changes in intracellular cAMP (468–470). Notably, point mutations in the promoter region of ADCY3 (in sites different from the CRE motif) have been linked to decreased glucose-induced insulin secretion in a rat model of diabetes (469).
In rats, the expression of AC8 is also enhanced by the farnesoid X receptor (FXR), a nuclear receptor that binds to the promoter region of ADCY8 (471). The expression levels of both AC8 and FXR are reduced in a rat model of diabetes compared with control animals. A more detailed study analyzed the genetic regulation of mouse AC1. In addition to reporting the lack of a TATA box, promoter sequences for potential binding of glucocorticoid receptors, POU transcription factors, the immediate-early gene zif268, Gcn4, E-box-binding factors, as well as the previously mentioned Sp1 were identified (472).
The methylation state of the promoter regions of ADCY3 and ADCY4 have been studied. For ADCY3, hypomethylation has been detected in gastric cancer cell lines, where AC3 transcript levels are increased (473). For ADCY4, hypermethylation of the promoter region has been shown in patients with breast cancer compared with control subjects (317). This profile was correlated with a decrease in AC4 mRNA transcripts in those patients.
Processing of mRNA through alternative splicing represents another natural method for regulating enzymatic activity. Notably, splice variants for AC isoforms have also been reported (56, 474–478). Some variants, such as those from AC4, AC5, AC6, and AC8, lack certain exon regions and may present different activity compared with the full enzymes (56). In the case of AC8, short variants dimerize with full-length AC8 to downregulate cAMP production (475). Notably, splice variants for AC1, AC5, and AC9 were reported to encode for nearly half of the AC (56, 474). These near-half variants may serve as another way to regulate AC activity and cAMP signaling.
8. CONCLUDING REMARKS
ACs have the potential to regulate a vast array of physiological responses since their enzymatic product, cAMP, regulates an immense number of cell functions. The fact that most cells express several of the nine AC isoforms creates a redundancy of function that makes it difficult to discern the role of one specific isoform. Drugs to specifically stimulate or inhibit a single AC isoform have largely been lacking, making knockdown and overexpression the mainstay of understanding the physiological roles of an AC. A number of early biochemical studies of ACs defined how these isoforms are regulated by different coincident signals, allowing a framework for understanding how different isoforms could be central to the physiological functions of highly differentiated cells. More recent studies make clear that these isoforms exist in distinct signaling complexes where they couple to specific upstream receptors and generate compartmentalized cAMP pools downstream. The concept that ACs are not redundant but are capable of playing roles in specific physiological responses at the cellular, tissue, and organismal levels is now widely accepted. As such, studies to examine these specific responses are beginning to accelerate.
Resolving AC and cAMP compartmentation in a spatial and temporal manner will help further tease out the physiological roles of the AC isoforms. Historically, measures of cAMP levels in cells have been limited to end-point assays that, in most cases, required the inclusion of broad PDE inhibitors so that cAMP could accumulate to levels sufficient for detection. These whole cell measures of cAMP do not reflect the biochemical reality that the second messenger activates local effectors, is degraded locally, and does not freely diffuse throughout the cell (479–481). cAMP biosensors, most of which are based on FRET, have been developed recently that allow the detection of cAMP in living cells and can be monitored in real time (482–484). FRET-based approaches are limited by relatively low efficiency, so other approaches are needed for more sensitive and rapid readouts. One approach uses cyclic nucleotide-gated channels to report cAMP levels in near-membrane regions (485, 486), whereas others have developed fluorescent biosensors that quench or unquench upon binding cAMP (487). Targeting these various biosensors to discrete subcellular compartments can report localized cAMP dynamics, representing a major advance in understanding compartmentalized GPCR-AC-cAMP signaling (108, 113, 488–493). Combining these powerful new cAMP detection tools with molecular interventions that disrupt expression of specific AC isoforms should accelerate the discovery of how individual ACs regulate specific physiological processes.
GRANTS
Support was provided by a New Investigator Award from the American Association of Colleges of Pharmacy and an IntegraConnect Grant from the Lloyd L. Gregory School of Pharmacy at Palm Beach Atlantic University (T.F.B.); NIH Grant GM60419 (C.W.D.); NIH Grants NS119917 and NS111070 and the Purdue University College of Pharmacy (V.J.W.); and NIH Grant GM107094 (R.S.O.).
DISCLOSURES
No conflicts of interest, financial or otherwise, are declared by the authors.
AUTHOR CONTRIBUTIONS
K.F.O., T.F.B., V.J.W., and R.S.O. prepared figures; K.F.O., J.E.L., T.F.B., R.S., C.W.D., V.J.W., and R.S.O. drafted manuscript; K.F.O., J.E.L., T.F.B., R.S., C.W.D., V.J.W., and R.S.O. edited and revised manuscript; K.F.O., J.E.L., T.F.B., R.S., C.W.D., V.J.W., and R.S.O. approved final version of manuscript.
ACKNOWLEDGMENTS
The authors thank Katherine Lyon for assistance with literature searches. Figures and graphical abstract were created with BioRender.com.
REFERENCES
- 1.Cooper DM. Regulation and organization of adenylyl cyclases and cAMP. Biochem J 375: 517–529, 2003. doi: 10.1042/bj20031061. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 2.Tang WJ, Yan S, Drum CL. Class III adenylyl cyclases: regulation and underlying mechanisms. Adv Second Messenger Phosphoprotein Res 32: 137–151, 1998. doi: 10.1016/s1040-7952(98)80009-8. [DOI] [PubMed] [Google Scholar]
- 3.Sunahara RK, Dessauer CW, Gilman AG. Complexity and diversity of mammalian adenylyl cyclases. Annu Rev Pharmacol Toxicol 36: 461–480, 1996. doi: 10.1146/annurev.pa.36.040196.002333. [DOI] [PubMed] [Google Scholar]
- 4.Khannpnavar B, Mehta V, Qi C, Korkhov V. Structure and function of adenylyl cyclases, key enzymes in cellular signaling. Curr Opin Struct Biol 63: 34–41, 2020. doi: 10.1016/j.sbi.2020.03.003. [DOI] [PubMed] [Google Scholar]
- 5.Linder JU, Schultz JE. The class III adenylyl cyclases: multi-purpose signalling modules. Cell Signal 15: 1081–1089, 2003. doi: 10.1016/S0898-6568(03)00130-X. [DOI] [PubMed] [Google Scholar]
- 6.Buck J, Sinclair ML, Schapal L, Cann MJ, Levin LR. Cytosolic adenylyl cyclase defines a unique signaling molecule in mammals. Proc Natl Acad Sci USA 96: 79–84, 1999. doi: 10.1073/pnas.96.1.79. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 7.Chen Y, Cann MJ, Litvin TN, Iourgenko V, Sinclair ML, Levin LR, Buck J. Soluble adenylyl cyclase as an evolutionarily conserved bicarbonate sensor. Science 289: 625–628, 2000. doi: 10.1126/science.289.5479.625. [DOI] [PubMed] [Google Scholar]
- 8.Levin LR, Buck J. Physiological roles of acid-base sensors. Annu Rev Physiol 77: 347–362, 2015. doi: 10.1146/annurev-physiol-021014-071821. [DOI] [PubMed] [Google Scholar]
- 9.Schmid A, Meili D, Salathe M. Soluble adenylyl cyclase in health and disease. Biochim Biophys Acta 1842: 2584–2592, 2014. doi: 10.1016/j.bbadis.2014.07.010. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 10.Qi C, Sorrentino S, Medalia O, Korkhov VM. The structure of a membrane adenylyl cyclase bound to an activated stimulatory G protein. Science 364: 389–394, 2019. doi: 10.1126/science.aav0778. [DOI] [PubMed] [Google Scholar]
- 11.Dessauer CW, Watts VJ, Ostrom RS, Conti M, Dove S, Seifert R. International Union of Basic and Clinical Pharmacology. CI. Structures and small molecule modulators of mammalian adenylyl cyclases. Pharmacol Rev 69: 93–139, 2017. doi: 10.1124/pr.116.013078. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 12.Tabbasum VG, Cooper DM. Structural and functional determinants of AC8 trafficking, targeting and responsiveness in lipid raft microdomains. J Membr Biol 252: 159–172, 2019. doi: 10.1007/s00232-019-00060-x. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 13.Chen YZ, Friedman JR, Chen DH, Chan GC, Bloss CS, Hisama FM, Topol SE, Carson AR, Pham PH, Bonkowski ES, Scott ER, Lee JK, Zhang G, Oliveira G, Xu J, Scott-Van Zeeland AA, Chen Q, Levy S, Topol EJ, Storm D, Swanson PD, Bird TD, Schork NJ, Raskind WH, Torkamani A. Gain-of-function ADCY5 mutations in familial dyskinesia with facial myokymia. Ann Neurol 75: 542–549, 2014. doi: 10.1002/ana.24119. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 14.Douglas AG, Andreoletti G, Talbot K, Hammans SR, Singh J, Whitney A, Ennis S, Foulds NC. ADCY5-related dyskinesia presenting as familial myoclonus-dystonia. Neurogenetics 18: 111–117, 2017. doi: 10.1007/s10048-017-0510-z. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 15.Chen DH, Méneret A, Friedman JR, Korvatska O, Gad A, Bonkowski ES, et al. ADCY5-related dyskinesia: broader spectrum and genotype-phenotype correlations. Neurology 85: 2026–2035, 2015. doi: 10.1212/WNL.0000000000002058. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 16.Chen YZ, Matsushita MM, Robertson P, Rieder M, Girirajan S, Antonacci F, Lipe H, Eichler EE, Nickerson DA, Bird TD, Raskind WH. Autosomal dominant familial dyskinesia and facial myokymia: single exome sequencing identifies a mutation in adenylyl cyclase 5. Arch Neurol 69: 630–635, 2012. doi: 10.1001/archneurol.2012.54. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 17.Robishaw JD, Russell DW, Harris BA, Smigel MD, Gilman AG. Deduced primary structure of the alpha subunit of the GTP-binding stimulatory protein of adenylate cyclase. Proc Natl Acad Sci USA 83: 1251–1255, 1986. doi: 10.1073/pnas.83.5.1251. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 18.Taussig R, Iñiguez-Lluhi JA, Gilman AG. Inhibition of adenylyl cyclase by Gi alpha. Science 261: 218–221, 1993. doi: 10.1126/science.8327893. [DOI] [PubMed] [Google Scholar]
- 19.Gilman AG. A protein binding assay for adenosine 3':5'-cyclic monophosphate. Proc Natl Acad Sci USA 67: 305–312, 1970. doi: 10.1073/pnas.67.1.305. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 20.Gill GN, Garren LD. Role of the receptor in the mechanism of action of adenosine 3':5'-cyclic monophosphate. Proc Natl Acad Sci USA 68: 786–790, 1971. [PM C]doi: 10.1073/pnas.68.4.786. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 21.Taylor SS, Stafford PH. Characterization of adenosine 3':5'-monophosphate-dependent protein kinase and its dissociated subunits from porcine skeletal muscle. J Biol Chem 253: 2284–2287, 1978. doi: 10.1016/S0021-9258(17)38070-5. [DOI] [PubMed] [Google Scholar]
- 22.Kennelly PJ, Krebs EG. Consensus sequences as substrate specificity determinants for protein kinases and protein phosphatases. J Biol Chem 266: 15555–15558, 1991. doi: 10.1016/S0021-9258(18)98436-X. [DOI] [PubMed] [Google Scholar]
- 23.Carr DW, Stofko-Hahn RE, Fraser ID, Bishop SM, Acott TS, Brennan RG, Scott JD. Interaction of the regulatory subunit (RII) of cAMP-dependent protein kinase with RII-anchoring proteins occurs through an amphipathic helix binding motif. J Biol Chem 266: 14188–14192, 1991. doi: 10.1016/S0021-9258(18)98665-5. [DOI] [PubMed] [Google Scholar]
- 24.Hirsch AH, Glantz SB, Li Y, You Y, Rubin CS. Cloning and expression of an intron-less gene for AKAP 75, an anchor protein for the regulatory subunit of cAMP-dependent protein kinase II beta. J Biol Chem 267: 2131–2134, 1992. doi: 10.1016/S0021-9258(18)45852-8. [DOI] [PubMed] [Google Scholar]
- 25.Omar MH, Scott JD. AKAP signaling islands: venues for precision pharmacology. Trends Pharmacol Sci 41: 933–946, 2020. doi: 10.1016/j.tips.2020.09.007. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 26.Kapiloff MS, Piggott LA, Sadana R, Li J, Heredia LA, Henson E, Efendiev R, Dessauer CW. An adenylyl cyclase-mAKAPbeta signaling complex regulates cAMP levels in cardiac myocytes. J Biol Chem 284: 23540–23546, 2009. doi: 10.1074/jbc.M109.030072. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 27.Efendiev R, Samelson BK, Nguyen BT, Phatarpekar PV, Baameur F, Scott JD, Dessauer CW. AKAP79 interacts with multiple adenylyl cyclase (AC) isoforms and scaffolds AC5 and -6 to alpha-amino-3-hydroxyl-5-methyl-4-isoxazole-propionate (AMPA) receptors. J Biol Chem 285: 14450–14458, 2010. doi: 10.1074/jbc.M110.109769. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 28.Dessauer CW. Adenylyl cyclase-A-kinase anchoring protein complexes: the next dimension in cAMP signaling. Mol Pharmacol 76: 935–941, 2009. doi: 10.1124/mol.109.059345. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 29.Beavo JA. Cyclic nucleotide phosphodiesterases: functional implications of multiple isoforms. Physiol Rev 75: 725–748, 1995. doi: 10.1152/physrev.1995.75.4.725. [DOI] [PubMed] [Google Scholar]
- 30.Soderling SH, Beavo JA. Regulation of cAMP and cGMP signaling: new phosphodiesterases and new functions. Curr Opin Cell Biol 12: 174–179, 2000. doi: 10.1016/S0955-0674(99)00073-3. [DOI] [PubMed] [Google Scholar]
- 31.Schmidt M, Cattani-Cavalieri I, Nuñez FJ, Ostrom RS. Phosphodiesterase isoforms and cAMP compartments in the development of new therapies for obstructive pulmonary diseases. Curr Opin Pharmacol 51: 34–42, 2020. doi: 10.1016/j.coph.2020.05.002. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 32.Lynch MJ, Hill EV, Houslay MD. Intracellular targeting of phosphodiesterase-4 underpins compartmentalized cAMP signaling. Curr Top Dev Biol 75: 225–259, 2006. doi: 10.1016/S0070-2153(06)75007-4. [DOI] [PubMed] [Google Scholar]
- 33.Halls ML, Cooper DM. Adenylyl cyclase signalling complexes-pharmacological challenges and opportunities. Pharmacol Ther 172: 171–180, 2017. doi: 10.1016/j.pharmthera.2017.01.001. [DOI] [PubMed] [Google Scholar]
- 34.Robichaux WG 3rd, Cheng X. Intracellular cAMP sensor EPAC: physiology, pathophysiology, and therapeutics development. Physiol Rev 98: 919–1053, 2018. doi: 10.1152/physrev.00025.2017. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 35.de Rooij J, Zwartkruis FJ, Verheijen MH, Cool RH, Nijman SM, Wittinghofer A, Bos JL. Epac is a Rap1 guanine-nucleotide-exchange factor directly activated by cyclic AMP. Nature 396: 474–477, 1998. doi: 10.1038/24884. [DOI] [PubMed] [Google Scholar]
- 36.Kawasaki H, Springett GM, Mochizuki N, Toki S, Nakaya M, Matsuda M, Housman DE, Graybiel AM. A family of cAMP-binding proteins that directly activate Rap1. Science 282: 2275–2279, 1998. [PM C]doi: 10.1126/science.282.5397.2275. [DOI] [PubMed] [Google Scholar]
- 37.Kaupp UB, Seifert R. Cyclic nucleotide-gated ion channels. Physiol Rev 82: 769–824, 2002. doi: 10.1152/physrev.00008.2002. [DOI] [PubMed] [Google Scholar]
- 38.Biel M, Michalakis S. Cyclic nucleotide-gated channels. Handb Exp Pharmacol 191: 111–136, 2009. doi: 10.1007/978-3-540-68964-5_7. [DOI] [PubMed] [Google Scholar]
- 39.Stieber J, Herrmann S, Feil S, Löster J, Feil R, Biel M, Hofmann F, Ludwig A. The hyperpolarization-activated channel HCN4 is required for the generation of pacemaker action potentials in the embryonic heart. Proc Natl Acad Sci USA 100: 15235–15240, 2003. doi: 10.1073/pnas.2434235100. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 40.Alig J, Marger L, Mesirca P, Ehmke H, Mangoni ME, Isbrandt D. Control of heart rate by cAMP sensitivity of HCN channels. Proc Natl Acad Sci USA 106: 12189–12194, 2009. doi: 10.1073/pnas.0810332106. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 41.Milanesi R, Baruscotti M, Gnecchi-Ruscone T, DiFrancesco D. Familial sinus bradycardia associated with a mutation in the cardiac pacemaker channel. N Engl J Med 354: 151–157, 2006. doi: 10.1056/NEJMoa052475. [DOI] [PubMed] [Google Scholar]
- 42.Schindler RF, Brand T. The Popeye domain containing protein family—a novel class of cAMP effectors with important functions in multiple tissues. Prog Biophys Mol Biol 120: 28–36, 2016. doi: 10.1016/j.pbiomolbio.2016.01.001. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 43.Brand T, Schindler R. New kids on the block: the Popeye domain containing (POPDC) protein family acting as a novel class of cAMP effector proteins in striated muscle. Cell Signal 40: 156–165, 2017. doi: 10.1016/j.cellsig.2017.09.015. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 44.Russ PK, Kupperman AI, Presley SH, Haselton FR, Chang MS. Inhibition of RhoA signaling with increased Bves in trabecular meshwork cells. Invest Ophthalmol Vis Sci 51: 223–230, 2010. doi: 10.1167/iovs.09-3539. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 45.Schindler RF, Scotton C, Zhang J, Passarelli C, Ortiz-Bonnin B, Simrick S, et al. POPDC1(S201F) causes muscular dystrophy and arrhythmia by affecting protein trafficking. J Clin Invest 126: 239–253, 2016. doi: 10.1172/JCI79562. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 46.Rinné S, Ortiz-Bonnin B, Stallmeyer B, Kiper AK, Fortmüller L, Schindler RF, Herbort-Brand U, Kabir NS, Dittmann S, Friedrich C, Zumhagen S, Gualandi F, Selvatici R, Rapezzi C, Arbustini E, Ferlini A, Fabritz L, Schulze-Bahr E, Brand T, Decher N. POPDC2 a novel susceptibility gene for conduction disorders. J Mol Cell Cardiol 145: 74–83, 2020. doi: 10.1016/j.yjmcc.2020.06.005. [DOI] [PubMed] [Google Scholar]
- 47.Froese A, Breher SS, Waldeyer C, Schindler RF, Nikolaev VO, Rinné S, Wischmeyer E, Schlueter J, Becher J, Simrick S, Vauti F, Kuhtz J, Meister P, Kreissl S, Torlopp A, Liebig SK, Laakmann S, Müller TD, Neumann J, Stieber J, Ludwig A, Maier SK, Decher N, Arnold HH, Kirchhof P, Fabritz L, Brand T. Popeye domain containing proteins are essential for stress-mediated modulation of cardiac pacemaking in mice. J Clin Invest 122: 1119–1130, 2012. doi: 10.1172/JCI59410. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 48.Defer N, Best-Belpomme M, Hanoune J. Tissue specificity and physiological relevance of various isoforms of adenylyl cyclase. Am J Physiol Renal Physiol 279: F400–F416, 2000. doi: 10.1152/ajprenal.2000.279.3.F400. [DOI] [PubMed] [Google Scholar]
- 49.Willoughby D, Cooper DM. Organization and Ca2+ regulation of adenylyl cyclases in cAMP microdomains. Physiol Rev 87: 965–1010, 2007. doi: 10.1152/physrev.00049.2006. [DOI] [PubMed] [Google Scholar]
- 50.Sadana R, Dessauer CW. Physiological roles for G protein-regulated adenylyl cyclase isoforms: insights from knockout and overexpression studies. Neurosignals 17: 5–22, 2009. doi: 10.1159/000166277. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 51.Tang WJ, Gilman AG. Adenylyl cyclases. Cell 70: 869–872, 1992. doi: 10.1016/0092-8674(92)90236-6. [DOI] [PubMed] [Google Scholar]
- 52.Sabbatini ME, Gorelick F, Glaser S. Adenylyl cyclases in the digestive system. Cell Signal 26: 1173–1181, 2014. doi: 10.1016/j.cellsig.2014.01.033. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 53.Gros R, Ding Q, Chorazyczewski J, Pickering JG, Limbird LE, Feldman RD. Adenylyl cyclase isoform-selective regulation of vascular smooth muscle proliferation and cytoskeletal reorganization. Circ Res 99: 845–852, 2006. doi: 10.1161/01.RES.0000245189.21703.c0. [DOI] [PubMed] [Google Scholar]
- 54.Ostrom RS, Liu X, Head BP, Gregorian C, Seasholtz TM, Insel PA. Localization of adenylyl cyclase isoforms and G protein-coupled receptors in vascular smooth muscle cells: expression in caveolin-rich and noncaveolin domains. Mol Pharmacol 62: 983–992, 2002. doi: 10.1124/mol.62.5.983. [DOI] [PubMed] [Google Scholar]
- 55.Webb JG, Yates PW, Yang Q, Mukhin YV, Lanier SM. Adenylyl cyclase isoforms and signal integration in models of vascular smooth muscle cells. Am J Physiol Heart Circ Physiol 281: H1545–H1552, 2001. doi: 10.1152/ajpheart.2001.281.4.H1545. [DOI] [PubMed] [Google Scholar]
- 56.Ludwig MG, Seuwen K. Characterization of the human adenylyl cyclase gene family: cDNA, gene structure, and tissue distribution of the nine isoforms. J Recept Signal Transduct Res 22: 79–110, 2002. doi: 10.1081/RRS-120014589. [DOI] [PubMed] [Google Scholar]
- 57.Wang T, Brown MJ. Differential expression of adenylyl cyclase subtypes in human cardiovascular system. Mol Cell Endocrinol 223: 55–62, 2004. doi: 10.1016/j.mce.2004.05.012. [DOI] [PubMed] [Google Scholar]
- 58.Sanabra C, Mengod G. Neuroanatomical distribution and neurochemical characterization of cells expressing adenylyl cyclase isoforms in mouse and rat brain. J Chem Neuroanat 41: 43–54, 2011. doi: 10.1016/j.jchemneu.2010.11.001. [DOI] [PubMed] [Google Scholar]
- 59.Premont RT, Matsuoka I, Mattei MG, Pouille Y, Defer N, Hanoune J. Identification and characterization of a widely expressed form of adenylyl cyclase. J Biol Chem 271: 13900–13907, 1996. doi: 10.1074/jbc.271.23.13900. [DOI] [PubMed] [Google Scholar]
- 60.Tang WJ, Gilman AG. Type-specific regulation of adenylyl cyclase by G protein beta gamma subunits. Science 254: 1500–1503, 1991. doi: 10.1126/science.1962211. [DOI] [PubMed] [Google Scholar]
- 61.Wayman GA, Impey S, Wu Z, Kindsvogel W, Prichard L, Storm DR. Synergistic activation of the type I adenylyl cyclase by Ca2+ and Gs-coupled receptors in vivo. J Biol Chem 269: 25400–25405, 1994. doi: 10.1016/S0021-9258(18)47263-8. [DOI] [PubMed] [Google Scholar]
- 62.Yoshimura M, Cooper DM. Cloning and expression of a Ca2+-inhibitable adenylyl cyclase from NCB-20 cells. Proc Natl Acad Sci USA 89: 6716–6720, 1992. doi: 10.1073/pnas.89.15.6716. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 63.Katsushika S, Chen L, Kawabe J, Nilakantan R, Halnon NJ, Homcy CJ, Ishikawa Y. Cloning and characterization of a sixth adenylyl cyclase isoform: types V and VI constitute a subgroup within the mammalian adenylyl cyclase family. Proc Natl Acad Sci USA 89: 8774–8778, 1992. doi: 10.1073/pnas.89.18.8774. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 64.Iyengar R. Molecular and functional diversity of mammalian Gs-stimulated adenylyl cyclases. FASEB J 7: 768–775, 1993. doi: 10.1096/fasebj.7.9.8330684. [DOI] [PubMed] [Google Scholar]
- 65.Chen J, Iyengar R. Inhibition of cloned adenylyl cyclases by mutant-activated Gi-alpha and specific suppression of type 2 adenylyl cyclase inhibition by phorbol ester treatment. J Biol Chem 268: 12253–12256, 1993. doi: 10.1016/S0021-9258(18)31381-4. [DOI] [PubMed] [Google Scholar]
- 66.Federman AD, Conklin BR, Schrader KA, Reed RR, Bourne HR. Hormonal stimulation of adenylyl cyclase through Gi-protein beta gamma subunits. Nature 356: 159–161, 1992. doi: 10.1038/356159a0. [DOI] [PubMed] [Google Scholar]
- 67.Gao BN, Gilman AG. Cloning and expression of a widely distributed (type IV) adenylyl cyclase. Proc Natl Acad Sci USA 88: 10178–10182, 1991. doi: 10.1073/pnas.88.22.10178. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 68.Thomas JM, Hoffman BB. Isoform-specific sensitization of adenylyl cyclase activity by prior activation of inhibitory receptors: role of beta gamma subunits in transducing enhanced activity of the type VI isoform. Mol Pharmacol 49: 907–914, 1996. [PubMed] [Google Scholar]
- 69.Gao X, Sadana R, Dessauer CW, Patel TB. Conditional stimulation of type V and VI adenylyl cyclases by G protein betagamma subunits. J Biol Chem 282: 294–302, 2007. doi: 10.1074/jbc.M607522200. [DOI] [PubMed] [Google Scholar]
- 70.Seamon KB, Padgett W, Daly JW. Forskolin: unique diterpene activator of adenylate cyclase in membranes and in intact cells. Proc Natl Acad Sci USA 78: 3363–3367, 1981. doi: 10.1073/pnas.78.6.3363. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 71.Insel PA, Ostrom RS. Forskolin as a tool for examining adenylyl cyclase expression, regulation, and G protein signaling. Cell Mol Neurobiol 23: 305–314, 2003. doi: 10.1023/A:1023684503883. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 72.Baldwin TA, Li Y, Brand CS, Watts VJ, Dessauer CW. Insights into the regulatory properties of human adenylyl cyclase type 9. Mol Pharmacol 95: 349–360, 2019. doi: 10.1124/mol.118.114595. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 73.Hacker BM, Tomlinson JE, Wayman GA, Sultana R, Chan G, Villacres E, Disteche C, Storm DR. Cloning, chromosomal mapping, and regulatory properties of the human type 9 adenylyl cyclase (ADCY9). Genomics 50: 97–104, 1998. doi: 10.1006/geno.1998.5293. [DOI] [PubMed] [Google Scholar]
- 74.Levin LR, Reed RR. Identification of functional domains of adenylyl cyclase using in vivo chimeras. J Biol Chem 270: 7573–7579, 1995. doi: 10.1074/jbc.270.13.7573. [DOI] [PubMed] [Google Scholar]
- 75.Lai HL, Yang TH, Messing RO, Ching YH, Lin SC, Chern Y. Protein kinase C inhibits adenylyl cyclase type VI activity during desensitization of the A2a-adenosine receptor-mediated cAMP response. J Biol Chem 272: 4970–4977, 1997. doi: 10.1074/jbc.272.8.4970. [DOI] [PubMed] [Google Scholar]
- 76.Zimmermann G, Taussig R. Protein kinase C alters the responsiveness of adenylyl cyclases to G protein alpha and betagamma subunits. J Biol Chem 271: 27161–27166, 1996. doi: 10.1074/jbc.271.43.27161. [DOI] [PubMed] [Google Scholar]
- 77.Shen JX, Cooper DM. AKAP79, PKC, PKA and PDE4 participate in a Gq-linked muscarinic receptor and adenylate cyclase 2 cAMP signalling complex. Biochem J 455: 47–56, 2013. doi: 10.1042/BJ20130359. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 78.Cumbay MG, Watts VJ. Novel regulatory properties of human type 9 adenylate cyclase. J Pharmacol Exp Ther 310: 108–115, 2004. doi: 10.1124/jpet.104.065748. [DOI] [PubMed] [Google Scholar]
- 79.Taussig R, Tang WJ, Hepler JR, Gilman AG. Distinct patterns of bidirectional regulation of mammalian adenylyl cyclases. J Biol Chem 269: 6093–6100, 1994. doi: 10.1016/S0021-9258(17)37574-9. [DOI] [PubMed] [Google Scholar]
- 80.Taussig R, Quarmby LM, Gilman AG. Regulation of purified type I and type II adenylylcyclases by G protein beta gamma subunits. J Biol Chem 268: 9–12, 1993. doi: 10.1016/S0021-9258(18)54106-5. [DOI] [PubMed] [Google Scholar]
- 81.Tang WJ, Krupinski J, Gilman AG. Expression and characterization of calmodulin-activated (type I) adenylylcyclase. J Biol Chem 266: 8595–8603, 1991. doi: 10.1016/S0021-9258(18)93016-4. [DOI] [PubMed] [Google Scholar]
- 82.Diel S, Klass K, Wittig B, Kleuss C. Gbetagamma activation site in adenylyl cyclase type II. Adenylyl cyclase type III is inhibited by Gbetagamma. J Biol Chem 281: 288–294, 2006. doi: 10.1074/jbc.M511045200. [DOI] [PubMed] [Google Scholar]
- 83.Choi EJ, Xia Z, Storm DR. Stimulation of the type III olfactory adenylyl cyclase by calcium and calmodulin. Biochemistry 31: 6492–6498, 1992. doi: 10.1021/bi00143a019. [DOI] [PubMed] [Google Scholar]
- 84.Gao BN, Gilman AG. Cloning and expression of a widely distributed (type IV) adenylyl cyclase. Proc Natl Acad Sci USA 88: 10178–10182, 1991. doi: 10.1073/pnas.88.22.10178. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 85.Brand CS, Sadana R, Malik S, Smrcka AV, Dessauer CW. Adenylyl cyclase 5 regulation by Gbetagamma involves isoform-specific use of multiple interaction sites. Mol Pharmacol 88: 758–767, 2015. doi: 10.1124/mol.115.099556. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 86.Yoshimura M, Cooper DM. Cloning and expression of a Ca2+-inhibitable adenylyl cyclase from NCB-20 cells. Proc Natl Acad Sci USA 89: 6716–6720, 1992. doi: 10.1073/pnas.89.15.6716. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 87.Katsushika S, Chen L, Kawabe J, Nilakantan R, Halnon NJ, Homcy CJ, Ishikawa Y. Cloning and characterization of a sixth adenylyl cyclase isoform: types V and VI constitute a subgroup within the mammalian adenylyl cyclase family. Proc Natl Acad Sci USA 89: 8774–8778, 1992. doi: 10.1073/pnas.89.18.8774. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 88.Iwami G, Kawabe J, Ebina T, Cannon PJ, Homcy CJ, Ishikawa Y. Regulation of adenylyl cyclase by protein kinase A. J Biol Chem 270: 12481–12484, 1995. doi: 10.1074/jbc.270.21.12481. [DOI] [PubMed] [Google Scholar]
- 89.Heinick A, Husser X, Himmler K, Kirchhefer U, Nunes F, Schulte JS, Seidl MD, Rolfes C, Dedman JR, Kaetzel MA, Gerke V, Schmitz W, Müller F. Annexin A4 is a novel direct regulator of adenylyl cyclase type 5. FASEB J 29: 3773–3787, 2015. doi: 10.1096/fj.14-269837. [DOI] [PubMed] [Google Scholar]
- 90.Lai HL, Yang TH, Messing RO, Ching YH, Lin SC, Chern Y. Protein kinase C inhibits adenylyl cyclase type VI activity during desensitization of the A2a-adenosine receptor-mediated cAMP response. J Biol Chem 272: 4970–4977, 1997. doi: 10.1074/jbc.272.8.4970. [DOI] [PubMed] [Google Scholar]
- 91.Chen Y, Harry A, Li J, Smit MJ, Bai X, Magnusson R, Pieroni JP, Weng G, Iyengar R. Adenylyl cyclase 6 is selectively regulated by protein kinase A phosphorylation in a region involved in Galphas stimulation. Proc Natl Acad Sci USA 94: 14100–14104, 1997. doi: 10.1073/pnas.94.25.14100. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 92.Yoshimura M, Ikeda H, Tabakoff B. mu-Opioid receptors inhibit dopamine-stimulated activity of type V adenylyl cyclase but enhance dopamine-stimulated activity of type VII adenylyl cyclase. Mol Pharmacol 50: 43–51, 1996. [PubMed] [Google Scholar]
- 93.Cali JJ, Zwaagstra JC, Mons N, Cooper DM, Krupinski J. Type VIII adenylyl cyclase. A Ca2+/calmodulin-stimulated enzyme expressed in discrete regions of rat brain. J Biol Chem 269: 12190–12195, 1994. doi: 10.1016/S0021-9258(17)32700-X. [DOI] [PubMed] [Google Scholar]
- 94.Willoughby D, Halls ML, Everett KL, Ciruela A, Skroblin P, Klussmann E, Cooper DM. A key phosphorylation site in AC8 mediates regulation of Ca2+-dependent cAMP dynamics by an AC8-AKAP79-PKA signalling complex. J Cell Sci 125: 5850–5859, 2012. doi: 10.1242/jcs.111427. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 95.Steiner D, Saya D, Schallmach E, Simonds WF, Vogel Z. Adenylyl cyclase type-VIII activity is regulated by G(betagamma) subunits. Cell Signal 18: 62–68, 2006. doi: 10.1016/j.cellsig.2005.03.014. [DOI] [PubMed] [Google Scholar]
- 96.Hanoune J, Pouille Y, Tzavara E, Shen T, Lipskaya L, Miyamoto N, Suzuki Y, Defer N. Adenylyl cyclases: structure, regulation and function in an enzyme superfamily. Mol Cell Endocrinol 128: 179–194, 1997. doi: 10.1016/S0303-7207(97)04013-6. [DOI] [PubMed] [Google Scholar]
- 97.Arora K, Sinha C, Zhang W, Ren A, Moon CS, Yarlagadda S, Naren AP. Compartmentalization of cyclic nucleotide signaling: a question of when, where, and why? Pflugers Arch 465: 1397–1407, 2013. doi: 10.1007/s00424-013-1280-6. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 98.Ostrom RS, Insel PA. The evolving role of lipid rafts and caveolae in G protein-coupled receptor signaling: implications for molecular pharmacology. Br J Pharmacol 143: 235–245, 2004. doi: 10.1038/sj.bjp.0705930. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 99.Johnstone TB, Agarwal SR, Harvey RD, Ostrom RS. cAMP signaling compartmentation: adenylyl cyclases as anchors of dynamic signaling complexes. Mol Pharmacol 93: 270–276, 2018. doi: 10.1124/mol.117.110825. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 100.Houslay MD. Underpinning compartmentalised cAMP signalling through targeted cAMP breakdown. Trends Biochem Sci 35: 91–100, 2010. doi: 10.1016/j.tibs.2009.09.007. [DOI] [PubMed] [Google Scholar]
- 101.Buxton IL, Brunton LL. Compartments of cyclic AMP and protein kinase in mammalian cardiomyocytes. J Biol Chem 258: 10233–10239, 1983. doi: 10.1016/S0021-9258(17)44447-4. [DOI] [PubMed] [Google Scholar]
- 102.Ostrom RS, Bogard AS, Gros R, Feldman RD. Choreographing the adenylyl cyclase signalosome: sorting out the partners and the steps. Naunyn-Schmiedebergs Arch Pharmacol 385: 5–12, 2012. doi: 10.1007/s00210-011-0696-9. [DOI] [PubMed] [Google Scholar]
- 103.Schwencke C, Yamamoto M, Okumura S, Toya Y, Kim SJ, Ishikawa Y. Compartmentation of cyclic adenosine 3',5'-monophosphate signaling in caveolae. Mol Endocrinol 13: 1061–1070, 1999. doi: 10.1210/mend.13.7.0304. [DOI] [PubMed] [Google Scholar]
- 104.Cooper DM, Tabbasum VG. Adenylate cyclase-centred microdomains. Biochem J 462: 199–213, 2014. doi: 10.1042/BJ20140560. [DOI] [PubMed] [Google Scholar]
- 105.Thangavel M, Liu X, Sun SQ, Kaminsky J, Ostrom RS. The C1 and C2 domains target human type 6 adenylyl cyclase to lipid rafts and caveolae. Cell Signal 21: 301–308, 2009. doi: 10.1016/j.cellsig.2008.10.017. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 106.Antoni FA, Wiegand UK, Black J, Simpson J. Cellular localisation of adenylyl cyclase: a post-genome perspective. Neurochem Res 31: 287–295, 2006. doi: 10.1007/s11064-005-9019-1. [DOI] [PubMed] [Google Scholar]
- 107.Ostrom RS, Bundey RA, Insel PA. Nitric oxide inhibition of adenylyl cyclase type 6 activity is dependent upon lipid rafts and caveolin signaling complexes. J Biol Chem 279: 19846–19853, 2004. doi: 10.1074/jbc.M313440200. [DOI] [PubMed] [Google Scholar]
- 108.Agarwal SR, Fiore C, Miyashiro K, Ostrom RS, Harvey RD. Effect of adenylyl cyclase type 6 on localized production of cAMP by beta-2 adrenoceptors in human airway smooth-muscle cells. J Pharmacol Exp Ther 370: 104–110, 2019. doi: 10.1124/jpet.119.256594. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 109.Bogard AS, Adris P, Ostrom RS. Adenylyl cyclase 2 selectively couples to E prostanoid type 2 receptors, whereas adenylyl cyclase 3 is not receptor-regulated in airway smooth muscle. J Pharmacol Exp Ther 342: 586–595, 2012. doi: 10.1124/jpet.112.193425. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 110.Liu X, Thangavel M, Sun SQ, Kaminsky J, Mahautmr P, Stitham J, Hwa J, Ostrom RS. Adenylyl cyclase type 6 overexpression selectively enhances beta-adrenergic and prostacyclin receptor-mediated inhibition of cardiac fibroblast function because of colocalization in lipid rafts. Naunyn-Schmiedebergs Arch Pharmacol 377: 359–369, 2008. doi: 10.1007/s00210-007-0196-0. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 111.Ostrom RS, Gregorian C, Drenan RM, Xiang Y, Regan JW, Insel PA. Receptor number and caveolar co-localization determine receptor coupling efficiency to adenylyl cyclase. J Biol Chem 276: 42063–42069, 2001. doi: 10.1074/jbc.M105348200. [DOI] [PubMed] [Google Scholar]
- 112.Ostrom RS, Violin JD, Coleman S, Insel PA. Selective enhancement of beta-adrenergic receptor signaling by overexpression of adenylyl cyclase type 6: colocalization of receptor and adenylyl cyclase in caveolae of cardiac myocytes. Mol Pharmacol 57: 1075–1079, 2000. [PubMed] [Google Scholar]
- 113.Agarwal SR, Miyashiro K, Latt H, Ostrom RS, Harvey RD. Compartmentalized cAMP responses to prostaglandin EP2 receptor activation in human airway smooth muscle cells. Br J Pharmacol 174: 2784–2796, 2017. doi: 10.1111/bph.13904. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 114.Johnstone TB, Smith KH, Koziol-White CJ, Li F, Kazarian AG, Corpuz ML, Shumyatcher M, Ehlert FJ, Himes BE, Panettieri RA Jr, Ostrom RS. PDE8 is expressed in human airway smooth muscle and selectively regulates cAMP signaling by beta2-adrenergic receptors and adenylyl cyclase 6. Am J Respir Cell Mol Biol 58: 530–541, 2018. doi: 10.1165/rcmb.2017-0294OC. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 115.Marsden AN, Dessauer CW. Nanometric targeting of type 9 adenylyl cyclase in heart. Biochem Soc Trans 47: 1749–1756, 2019. doi: 10.1042/BST20190227. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 116.Eichel K, von Zastrow M. Subcellular organization of GPCR signaling. Trends Pharmacol Sci 39: 200–208, 2018. doi: 10.1016/j.tips.2017.11.009. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 117.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 98: 675–681, 2006. doi: 10.1161/01.RES.0000209516.84815.3e. [DOI] [PubMed] [Google Scholar]
- 118.Stevens TC, Ochoa CD, Morrow KA, Robson MJ, Prasain N, Zhou C, Alvarez DF, Frank DW, Balczon R, Stevens T. The Pseudomonas aeruginosa exoenzyme Y impairs endothelial cell proliferation and vascular repair following lung injury. Am J Physiol Lung Cell Mol Physiol 306: L915–L924, 2014. doi: 10.1152/ajplung.00135.2013. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 119.Hasan A, Danker KY, Wolter S, Bähre H, Kaever V, Seifert R. Soluble adenylyl cyclase accounts for high basal cCMP and cUMP concentrations in HEK293 and B103 cells. Biochem Biophys Res Commun 448: 236–240, 2014. doi: 10.1016/j.bbrc.2014.04.099. [DOI] [PubMed] [Google Scholar]
- 120.Beckert U, Wolter S, Hartwig C, Bähre H, Kaever V, Ladant D, Frank DW, Seifert R. ExoY from Pseudomonas aeruginosa is a nucleotidyl cyclase with preference for cGMP and cUMP formation. Biochem Biophys Res Commun 450: 870–874, 2014. doi: 10.1016/j.bbrc.2014.06.088. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 121.Morrow KA, Seifert R, Kaever V, Britain AL, Sayner SL, Ochoa CD, Cioffi EA, Frank DW, Rich TC, Stevens T. Heterogeneity of pulmonary endothelial cyclic nucleotide response to Pseudomonas aeruginosa ExoY infection. Am J Physiol Lung Cell Mol Physiol 309: L1199–L1207, 2015. doi: 10.1152/ajplung.00165.2015. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 122.Zippin JH, Chen Y, Nahirney P, Kamenetsky M, Wuttke MS, Fischman DA, Levin LR, Buck J. Compartmentalization of bicarbonate-sensitive adenylyl cyclase in distinct signaling microdomains. FASEB J 17: 82–84, 2003. doi: 10.1096/fj.02-0598fje. [DOI] [PubMed] [Google Scholar]
- 123.Agarwal SR, Yang PC, Rice M, Singer CA, Nikolaev VO, Lohse MJ, Clancy CE, Harvey RD. Role of membrane microdomains in compartmentation of cAMP signaling. PLoS One 9: e95835, 2014. doi: 10.1371/journal.pone.0095835. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 124.Antoni FA. New paradigms in cAMP signalling. Mol Cell Endocrinol 353: 3–9, 2012. doi: 10.1016/j.mce.2011.10.034. [DOI] [PubMed] [Google Scholar]
- 125.Efendiev R, Bavencoffe A, Hu H, Zhu MX, Dessauer CW. Scaffolding by A-kinase anchoring protein enhances functional coupling between adenylyl cyclase and TRPV1 channel. J Biol Chem 288: 3929–3937, 2013. doi: 10.1074/jbc.M112.428144. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 126.Erdorf M, Mou TC, Seifert R. Impact of divalent metal ions on regulation of adenylyl cyclase isoforms by forskolin analogs. Biochem Pharmacol 82: 1673–1681, 2011. doi: 10.1016/j.bcp.2011.07.099. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 127.Pinto C, Lushington GH, Richter M, Gille A, Geduhn J, König B, Mou TC, Sprang SR, Seifert R. Structure-activity relationships for the interactions of 2'- and 3'-(O)-(N-methyl)anthraniloyl-substituted purine and pyrimidine nucleotides with mammalian adenylyl cyclases. Biochem Pharmacol 82: 358–370, 2011. doi: 10.1016/j.bcp.2011.05.010. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 128.Brand CS, Hocker HJ, Gorfe AA, Cavasotto CN, Dessauer CW. Isoform selectivity of adenylyl cyclase inhibitors: characterization of known and novel compounds. J Pharmacol Exp Ther 347: 265–275, 2013. doi: 10.1124/jpet.113.208157. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 129.Suryanarayana S, Göttle M, Hübner M, Gille A, Mou TC, Sprang SR, Richter M, Seifert R. Differential inhibition of various adenylyl cyclase isoforms and soluble guanylyl cyclase by 2',3'-O-(2,4,6-trinitrophenyl)-substituted nucleoside 5'-triphosphates. J Pharmacol Exp Ther 330: 687–695, 2009. doi: 10.1124/jpet.109.155432. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 130.Wang J, Wang Z, Yu J, Zhang Y, Zeng Y, Gu Z. A forskolin-conjugated insulin analog targeting endogenous glucose-transporter for glucose-responsive insulin delivery. Biomater Sci 7: 4508–4513, 2019. doi: 10.1039/C9BM01283D. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 131.Irannejad R, Tomshine JC, Tomshine JR, Chevalier M, Mahoney JP, Steyaert J, Rasmussen SG, Sunahara RK, El-Samad H, Huang B, von Zastrow M. Conformational biosensors reveal GPCR signalling from endosomes. Nature 495: 534–538, 2013. doi: 10.1038/nature12000. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 132.Klotz KN, Kachler S. Inhibitors of membranous adenylyl cyclases with affinity for adenosine receptors. Naunyn-Schmiedebergs Arch Pharmacol 389: 349–352, 2016. doi: 10.1007/s00210-015-1197-z. [DOI] [PubMed] [Google Scholar]
- 133.Gille A, Lushington GH, Mou TC, Doughty MB, Johnson RA, Seifert R. Differential inhibition of adenylyl cyclase isoforms and soluble guanylyl cyclase by purine and pyrimidine nucleotides. J Biol Chem 279: 19955–19969, 2004. doi: 10.1074/jbc.M312560200. [DOI] [PubMed] [Google Scholar]
- 134.Levin LR, Han PL, Hwang PM, Feinstein PG, Davis RL, Reed RR. The Drosophila learning and memory gene rutabaga encodes a Ca2+/calmodulin-responsive adenylyl cyclase. Cell 68: 479–489, 1992. doi: 10.1016/0092-8674(92)90185-F. [DOI] [PubMed] [Google Scholar]
- 135.Livingstone MS, Sziber PP, Quinn WG. Loss of calcium/calmodulin responsiveness in adenylate cyclase of rutabaga, a Drosophila learning mutant. Cell 37: 205–215, 1984. doi: 10.1016/0092-8674(84)90316-7. [DOI] [PubMed] [Google Scholar]
- 136.Wu ZL, Thomas SA, Villacres EC, Xia Z, Simmons ML, Chavkin C, Palmiter RD, Storm DR. Altered behavior and long-term potentiation in type I adenylyl cyclase mutant mice. Proc Natl Acad Sci USA 92: 220–224, 1995. doi: 10.1073/pnas.92.1.220. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 137.Storm DR, Hansel C, Hacker B, Parent A, Linden DJ. Impaired cerebellar long-term potentiation in type I adenylyl cyclase mutant mice. Neuron 20: 1199–1210, 1998. doi: 10.1016/S0896-6273(00)80500-0. [DOI] [PubMed] [Google Scholar]
- 138.Villacres EC, Wong ST, Chavkin C, Storm DR. Type I adenylyl cyclase mutant mice have impaired mossy fiber long-term potentiation. J Neurosci 18: 3186–3194, 1998. doi: 10.1523/JNEUROSCI.18-09-03186.1998. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 139.Wong ST, Athos J, Figueroa XA, Pineda VV, Schaefer ML, Chavkin CC, Muglia LJ, Storm DR. Calcium-stimulated adenylyl cyclase activity is critical for hippocampus-dependent long-term memory and late phase LTP. Neuron 23: 787–798, 1999. doi: 10.1016/S0896-6273(01)80036-2. [DOI] [PubMed] [Google Scholar]
- 140.Zheng F, Zhang M, Ding Q, Sethna F, Yan L, Moon C, Yang M, Wang H. Voluntary running depreciates the requirement of Ca2+-stimulated cAMP signaling in synaptic potentiation and memory formation. Learn Mem 23: 442–449, 2016. doi: 10.1101/lm.040642.115. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 141.Wang H, Ferguson GD, Pineda VV, Cundiff PE, Storm DR. Overexpression of type-1 adenylyl cyclase in mouse forebrain enhances recognition memory and LTP. Nat Neurosci 7: 635–642, 2004. doi: 10.1038/nn1248. [DOI] [PubMed] [Google Scholar]
- 142.Garelick MG, Chan GC, DiRocco DP, Storm DR. Overexpression of type I adenylyl cyclase in the forebrain impairs spatial memory in aged but not young mice. J Neurosci 29: 10835–10842, 2009. doi: 10.1523/JNEUROSCI.0553-09.2009. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 143.Chen X, Cao H, Saraf A, Zweifel LS, Storm DR. Overexpression of the type 1 adenylyl cyclase in the forebrain leads to deficits of behavioral inhibition. J Neurosci 35: 339–351, 2015. doi: 10.1523/JNEUROSCI.2478-14.2015. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 144.Zhang M, Wang H. Mice overexpressing type 1 adenylyl cyclase show enhanced spatial memory flexibility in the absence of intact synaptic long-term depression. Learn Mem 20: 352–357, 2013. [P MC]doi: 10.1101/lm.030114.112. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 145.Yang M, Ding Q, Zhang M, Moon C, Wang H. Forebrain overexpression of type 1 adenylyl cyclase promotes molecular stability and behavioral resilience to physical stress. Neurobiol Stress 13: 100237, 2020. doi: 10.1016/j.ynstr.2020.100237. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 146.Santos-Cortez RL, Lee K, Giese AP, Ansar M, Amin-Ud-Din M, Rehn K, Wang X, Aziz A, Chiu I, Ali RH, Smith JD; University of Washington Center For Mendelian Genetics, Shendure J, Bamshad M, Nickerson DA, Ahmed ZM, Ahmad W, Riazuddin S, Leal SM. Adenylate cyclase 1 (ADCY1) mutations cause recessive hearing impairment in humans and defects in hair cell function and hearing in zebrafish. Hum Mol Genet 23: 3289–3298, 2014. doi: 10.1093/hmg/ddu042. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 147.Ansar M, Chahrour MH, Amin Ud Din M, Arshad M, Haque S, Pham TL, Yan K, Ahmad W, Leal SM. DFNB44, a novel autosomal recessive non-syndromic hearing impairment locus, maps to chromosome 7p14.1-q11.22. Hum Hered 57: 195–199, 2004. doi: 10.1159/000081446. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 148.Muhleisen TW, Leber M, Schulze TG, Strohmaier J, Degenhardt F, Treutlein J, Mattheisen M, Forstner AJ, et al. Genome-wide association study reveals two new risk loci for bipolar disorder. Nat Commun 5: 3339, 2014. doi: 10.1038/ncomms4339. [DOI] [PubMed] [Google Scholar]
- 149.Hardin M, Zielinski J, Wan ES, Hersh CP, Castaldi PJ, Schwinder E, Hawrylkiewicz I, Sliwinski P, Cho MH, Silverman EK. CHRNA3/5, IREB2, and ADCY2 are associated with severe chronic obstructive pulmonary disease in Poland. Am J Respir Cell Mol Biol 47: 203–208, 2012. doi: 10.1165/rcmb.2012-0011OC. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 150.Panasevich S, Melén E, Hallberg J, Bergström A, Svartengren M, Pershagen G, Nyberg F. Investigation of novel genes for lung function in children and their interaction with tobacco smoke exposure: a preliminary report. Acta Paediatr 102: 498–503, 2013. doi: 10.1111/apa.12204. [DOI] [PubMed] [Google Scholar]
- 151.Willsey AJ, Fernandez TV, Yu D, King RA, Dietrich A, Xing J, Sanders SJ, Mandell JD, Huang AY, Richer P, Smith L, Dong S, Samocha KE; Tourette International Collaborative Genetics (TIC), Tourette Syndrome Association International Consortium for Genetics (TSAICG), Neale BM, Coppola G, Mathews CA, Tischfield JA, Scharf JM, State MW., Heiman GA. De novo coding variants are strongly associated with Tourette disorder. Neuron 94: 486–499.e9, 2017. doi: 10.1016/j.neuron.2017.04.024. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 152.Satterstrom FK, Kosmicki JA, Wang J, Breen MS, De Rubeis S, An JY, et al. Large-scale exome sequencing study implicates both developmental and functional changes in the neurobiology of autism. Cell 180: 568–584.e23, 2020. doi: 10.1016/j.cell.2019.12.036. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 153.Deciphering Developmental Disorders Study. Prevalence and architecture of de novo mutations in developmental disorders. Nature 542: 433–438, 2017. doi: 10.1038/nature21062. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 154.Kosmicki JA, Samocha KE, Howrigan DP, Sanders SJ, Slowikowski K, Lek M, Karczewski KJ, Cutler DJ, Devlin B, Roeder K, Buxbaum JD, Neale BM, MacArthur DG, Wall DP, Robinson EB, Daly MJ. Refining the role of de novo protein-truncating variants in neurodevelopmental disorders by using population reference samples. Nat Genet 49: 504–510, 2017. doi: 10.1038/ng.3789. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 155.Izarzugaza JM, Ellesøe SG, Doganli C, Ehlers NS, Dalgaard MD, Audain E, Dombrowsky G, Banasik K, Sifrim A, Wilsdon A, Thienpont B, Breckpot J, Gewillig M; Competence Network For Congenital Heart Defects, Germany, Brook JD, Hitz MP, Larsen LA, Brunak S. Systems genetics analysis identifies calcium-signaling defects as novel cause of congenital heart disease. Genome Med 12: 76, 2020. doi: 10.1186/s13073-020-00772-z. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 156.Wray NR, Pergadia ML, Blackwood DH, Penninx BW, Gordon SD, Nyholt DR, et al. Genome-wide association study of major depressive disorder: new results, meta-analysis, and lessons learned. Mol Psychiatry 17: 36–48, 2012. doi: 10.1038/mp.2010.109. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 157.Stergiakouli E, Gaillard R, Tavare JM, Balthasar N, Loos RJ, Taal HR, Evans DM, Rivadeneira F, St Pourcain B, Uitterlinden AG, Kemp JP, Hofman A, Ring SM, Cole TJ, Jaddoe VW, Smith GD, Timpson NJ. Genome-wide association study of height-adjusted BMI in childhood identifies functional variant in ADCY3. Obesity (Silver Spring) 22: 2252–2259, 2014. doi: 10.1002/oby.20840. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 158.Grarup N, Moltke I, Andersen MK, Dalby M, Vitting-Seerup K, Kern T, Mahendran Y, Jørsboe E, Larsen CV, Dahl-Petersen IK, Gilly A, Suveges D, Dedoussis G, Zeggini E, Pedersen O, Andersson R, Bjerregaard P, Jørgensen ME, Albrechtsen A, Hansen T. Loss-of-function variants in ADCY3 increase risk of obesity and type 2 diabetes. Nat Genet 50: 172–174, 2018. doi: 10.1038/s41588-017-0022-7. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 159.Wang H, Wu M, Zhu W, Shen J, Shi X, Yang J, Zhao Q, Ni C, Xu Y, Shen H, Shen C, Gu HF. Evaluation of the association between the AC3 genetic polymorphisms and obesity in a Chinese Han population. PLoS One 5: e13851, 2010. doi: 10.1371/journal.pone.0013851. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 160.Hulur I, Gamazon ER, Skol AD, Xicola RM, Llor X, Onel K, Ellis NA, Kupfer SS. Enrichment of inflammatory bowel disease and colorectal cancer risk variants in colon expression quantitative trait loci. BMC Genomics 16: 138, 2015. doi: 10.1186/s12864-015-1292-z. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 161.Kaymaz N, Drukker M, Lieb R, Wittchen HU, Werbeloff N, Weiser M, Lataster T, van Os J. Do subthreshold psychotic experiences predict clinical outcomes in unselected non-help-seeking population-based samples? A systematic review and meta-analysis, enriched with new results. Psychol Med 42: 2239–2253, 2012. doi: 10.1017/S0033291711002911. [DOI] [PubMed] [Google Scholar]
- 162.Nordman S, Abulaiti A, Hilding A, Langberg EC, Humphreys K, Ostenson CG, Efendic S, Gu HF. Genetic variation of the adenylyl cyclase 3 (AC3) locus and its influence on type 2 diabetes and obesity susceptibility in Swedish men. Int J Obes 32: 407–412, 2008. doi: 10.1038/sj.ijo.0803742. [DOI] [PubMed] [Google Scholar]
- 163.Yuen RK, Merico D, Bookman M, Howe JL, Thiruvahindrapuram B, Patel RV, et al. Whole genome sequencing resource identifies 18 new candidate genes for autism spectrum disorder. Nat Neurosci 20: 602–611, 2017. doi: 10.1038/nn.4524. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 164.Magri C, Giacopuzzi E, La Via L, Bonini D, Ravasio V, Elhussiny ME, Orizio F, Gangemi F, Valsecchi P, Bresciani R, Barbon A, Vita A, Gennarelli M. A novel homozygous mutation in GAD1 gene described in a schizophrenic patient impairs activity and dimerization of GAD67 enzyme. Sci Rep 8: 15470, 2018. doi: 10.1038/s41598-018-33924-8. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 165.Lim ET, Uddin M, De Rubeis S, Chan Y, Kamumbu AS, Zhang X, D’Gama AM, Kim SN, Hill RS, Goldberg AP, Poultney C, Minshew NJ, Kushima I, Aleksic B, Ozaki N, Parellada M, Arango C, Penzol MJ, Carracedo A, Kolevzon A, Hultman CM, Weiss LA, Fromer M, Chiocchetti AG, Freitag CM; Autism Sequencing Consortium, Church GM, Scherer SW, Buxbaum JD, Walsh CA. Rates, distribution and implications of postzygotic mosaic mutations in autism spectrum disorder. Nat Neurosci 20: 1217–1224, 2017. doi: 10.1038/nn.4598. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 166.Loid P, Mustila T, Makitie RE, Viljakainen H, Kampe A, Tossavainen P, Lipsanen-Nyman M, Pekkinen M, Makitie O. Rare variants in genes linked to appetite control and hypothalamic development in early-onset severe obesity. Front Endocrinol (Lausanne) 11: 81, 2020. doi: 10.3389/fendo.2020.00081. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 167.Saeed S, Arslan M, Manzoor J, Din SM, Janjua QM, Ayesha H, Ain QT, Inam L, Lobbens S, Vaillant E, Durand E, Derhourhi M, Amanzougarene S, Badreddine A, Berberian L, Gaget S, Khan WI, Butt TA, Bonnefond A, Froguel P. Genetic causes of severe childhood obesity: a remarkably high prevalence in an inbred population of Pakistan. Diabetes 69: 1424–1438, 2020. doi: 10.2337/db19-1238. [DOI] [PubMed] [Google Scholar]
- 168.Procopio DO, Saba LM, Walter H, Lesch O, Skala K, Schlaff G, Vanderlinden L, Clapp P, Hoffman PL, Tabakoff B. Genetic markers of comorbid depression and alcoholism in women. Alcohol Clin Exp Res 37: 896–904, 2013. doi: 10.1111/acer.12060. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 169.Barker A, Sharp SJ, Timpson NJ, Bouatia-Naji N, Warrington NM, Kanoni S, et al. Association of genetic Loci with glucose levels in childhood and adolescence: a meta-analysis of over 6,000 children. Diabetes 60: 1805–1812, 2011. doi: 10.2337/db10-1575. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 170.Rees SD, Hydrie MZ, O’Hare JP, Kumar S, Shera AS, Basit A, Barnett AH, Kelly MA. Effects of 16 genetic variants on fasting glucose and type 2 diabetes in South Asians: ADCY5 and GLIS3 variants may predispose to type 2 diabetes. PLoS One 6: e24710, 2011. doi: 10.1371/journal.pone.0024710. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 171.Wagner R, Dudziak K, Herzberg-Schäfer SA, Machicao F, Stefan N, Staiger H, Häring HU, Fritsche A. Glucose-raising genetic variants in MADD and ADCY5 impair conversion of proinsulin to insulin. PLoS One 6: e23639, 2011. doi: 10.1371/journal.pone.0023639. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 172.Windholz J, Kovacs P, Tönjes A, Dittrich K, Blüher S, Kiess W, Stumvoll M, Körner A. Effects of genetic variants in ADCY5, GIPR, GCKR and VPS13C on early impairment of glucose and insulin metabolism in children. PLoS One 6: e22101, 2011. doi: 10.1371/journal.pone.0022101. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 173.Ng MC, Saxena R, Li J, Palmer ND, Dimitrov L, Xu J, Rasmussen-Torvik LJ, Zmuda JM, Siscovick DS, Patel SR, Crook ED, Sims M, Chen YD, Bertoni AG, Li M, Grant SF, Dupuis J, Meigs JB, Psaty BM, Pankow JS, Langefeld CD, Freedman BI, Rotter JI, Wilson JG, Bowden DW. Transferability and fine mapping of type 2 diabetes loci in African Americans: the Candidate Gene Association Resource Plus Study. Diabetes 62: 965–976, 2013. doi: 10.2337/db12-0266. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 174.Knigge A, Klöting N, Schön MR, Dietrich A, Fasshauer M, Gärtner D, Lohmann T, Dreßler M, Stumvoll M, Kovacs P, Blüher M. ADCY5 gene expression in adipose tissue is related to obesity in men and mice. PLoS One 10: e0120742, 2015. doi: 10.1371/journal.pone.0120742. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 175.Chang FC, Westenberger A, Dale RC, Smith M, Pall HS, Perez-Dueñas B, Grattan-Smith P, Ouvrier RA, Mahant N, Hanna BC, Hunter M, Lawson JA, Max C, Sachdev R, Meyer E, Crimmins D, Pryor D, Morris JG, Münchau A, Grozeva D, Carss KJ, Raymond L, Kurian MA, Klein C, Fung VS. Phenotypic insights into ADCY5-associated disease. Mov Disord 31: 1033–1040, 2016. doi: 10.1002/mds.26598. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 176.Carapito R, Paul N, Untrau M, Le Gentil M, Ott L, Alsaleh G, Jochem P, Radosavljevic M, Le Caignec C, David A, Damier P, Isidor B, Bahram S. A de novo ADCY5 mutation causes early-onset autosomal dominant chorea and dystonia. Mov Disord 30: 423–427, 2015. doi: 10.1002/mds.26115. [DOI] [PubMed] [Google Scholar]
- 177.Bohlega SA, Al-Foghom NB. Drug-induced Parkinson’s disease. A clinical review. Neurosciences (Riyadh) 18: 215–221, 2013. [PubMed] [Google Scholar]
- 178.Dean M, Messiaen L, Cooper GM, Amaral MD, Rashid S, Korf BR, Standaert DG. Child neurology: spastic paraparesis and dystonia with a novel ADCY5 mutation. Neurology 93: 510–514, 2019. doi: 10.1212/WNL.0000000000008089. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 179.Friedman JR, Méneret A, Chen DH, Trouillard O, Vidailhet M, Raskind WH, Roze E. ADCY5 mutation carriers display pleiotropic paroxysmal day and nighttime dyskinesias. Mov Disord 31: 147–148, 2016. doi: 10.1002/mds.26494. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 180.Liu Y, Chang X, Glessner J, Qu H, Tian L, Li D, Nguyen K, Sleiman PM, Hakonarson H. Association of rare recurrent copy number variants with congenital heart defects based on next-generation sequencing data from family trios. Front Genet 10: 819, 2019. doi: 10.3389/fgene.2019.00819. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 181.O’Roak BJ, Vives L, Fu W, Egertson JD, Stanaway IB, Phelps IG, Carvill G, Kumar A, Lee C, Ankenman K, Munson J, Hiatt JB, Turner EH, Levy R, O’Day DR, Krumm N, Coe BP, Martin BK, Borenstein E, Nickerson DA, Mefford HC, Doherty D, Akey JM, Bernier R, Eichler EE, Shendure J. Multiplex targeted sequencing identifies recurrently mutated genes in autism spectrum disorders. Science 338: 1619–1622, 2012. doi: 10.1126/science.1227764. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 182.O’Roak BJ, Vives L, Girirajan S, Karakoc E, Krumm N, Coe BP, Levy R, Ko A, Lee C, Smith JD, Turner EH, Stanaway IB, Vernot B, Malig M, Baker C, Reilly B, Akey JM, Borenstein E, Rieder MJ, Nickerson DA, Bernier R, Shendure J, Eichler EE. Sporadic autism exomes reveal a highly interconnected protein network of de novo mutations. Nature 485: 246–250, 2012. doi: 10.1038/nature10989. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 183.Powis Z, Towne MC, Hagman KD, Blanco K, Palmaer E, Castro A, Sajan SA, Radtke K, Feyma TJ, Juliette K, Tang S, Sidiropoulos C. Clinical diagnostic exome sequencing in dystonia: Genetic testing challenges for complex conditions. Clin Genet 97: 305–311, 2020. doi: 10.1111/cge.13657. [DOI] [PubMed] [Google Scholar]
- 184.Roman TS, Cannon ME, Vadlamudi S, Buchkovich ML, Wolford BN, Welch RP, Morken MA, Kwon GJ, Varshney A, Kursawe R, Wu Y, Jackson AU; National Institutes of Health Intramural Sequencing Center (NISC) Comparative Sequencing Program, Erdos MR, Kuusisto J, Laakso M, Scott LJ, Boehnke M, Collins FS, Parker SC, Stitzel ML, Mohlke KL. A type 2 diabetes-associated functional regulatory variant in a pancreatic islet enhancer at the ADCY5 locus. Diabetes 66: 2521–2530, 2017. doi: 10.2337/db17-0464. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 185.Westenberger A, Max C, Brüggemann N, Domingo A, Grütz K, Pawlack H, Weissbach A, Kühn AA, Spiegler J, Lang AE, Sperner J, Fung VS, Schallner J, Gillessen-Kaesbach G, Münchau A, Klein C. Alternating hemiplegia of childhood as a new presentation of adenylate cyclase 5-mutation-associated disease: a report of two cases. J Pediatr 181: 306–308.e1, 2017. doi: 10.1016/j.jpeds.2016.10.079. [DOI] [PubMed] [Google Scholar]
- 186.Zech M, Boesch S, Jochim A, Weber S, Meindl T, Schormair B, Wieland T, Lunetta C, Sansone V, Messner M, Mueller J, Ceballos-Baumann A, Strom TM, Colombo R, Poewe W, Haslinger B, Winkelmann J. Clinical exome sequencing in early-onset generalized dystonia and large-scale resequencing follow-up. Mov Disord 32: 549–559, 2017. doi: 10.1002/mds.26808. [DOI] [PubMed] [Google Scholar]
- 187.Zech M, Jech R, Boesch S, Skorvanek M, Weber S, Wagner M, et al. Monogenic variants in dystonia: an exome-wide sequencing study. Lancet Neurol 19: 908–918, 2020. doi: 10.1016/S1474-4422(20)30312-4. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 188.Graziola F, Garone G, Stregapede F, Bosco L, Vigevano F, Curatolo P, Bertini E, Travaglini L, Capuano A. Diagnostic yield of a targeted next-generation sequencing gene panel for pediatric-onset movement disorders: a 3-year cohort study. Front Genet 10: 1026, 2019. doi: 10.3389/fgene.2019.01026. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 189.Retterer K, Juusola J, Cho MT, Vitazka P, Millan F, Gibellini F, Vertino-Bell A, Smaoui N, Neidich J, Monaghan KG, McKnight D, Bai R, Suchy S, Friedman B, Tahiliani J, Pineda-Alvarez D, Richard G, Brandt T, Haverfield E, Chung WK, Bale S. Clinical application of whole-exome sequencing across clinical indications. Genet Med 18: 696–704, 2016. doi: 10.1038/gim.2015.148. [DOI] [PubMed] [Google Scholar]
- 190.Waalkens AJ, Vansenne F, van der Hout AH, Zutt R, Mourmans J, Tolosa E, de Koning TJ, Tijssen MA. Expanding the ADCY5 phenotype toward spastic paraparesis: a mutation in the M2 domain. Neurol Genet 4: e214, 2018. doi: 10.1212/NXG.0000000000000214. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 191.Vissers L, van Nimwegen KJ, Schieving JH, Kamsteeg EJ, Kleefstra T, Yntema HG, Pfundt R, van der Wilt GJ, Krabbenborg L, Brunner HG, van der Burg S, Grutters J, Veltman JA, Willemsen MA. A clinical utility study of exome sequencing versus conventional genetic testing in pediatric neurology. Genet Med 19: 1055–1063, 2017. doi: 10.1038/gim.2017.1. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 192.Gros R, Ding Q, Cao H, Main T, Hegele RA, Feldman RD. Identification of a dysfunctional missense single nucleotide variant of human adenylyl cyclase VI. Clin Pharmacol Ther 77: 271–278, 2005. doi: 10.1016/j.clpt.2004.11.005. [DOI] [PubMed] [Google Scholar]
- 193.Gros R, Van Uum S, Hutchinson-Jaffe A, Ding Q, Pickering JG, Hegele RA, Feldman RD. Increased enzyme activity and beta-adrenergic mediated vasodilation in subjects expressing a single-nucleotide variant of human adenylyl cyclase 6. Arterioscler Thromb Vasc Biol 27: 2657–2663, 2007. doi: 10.1161/ATVBAHA.107.145557. [DOI] [PubMed] [Google Scholar]
- 194.Gonzaga-Jauregui C, Harel T, Gambin T, Kousi M, Griffin LB, Francescatto L; Baylor-Hopkins Center for Mendelian Genomics, et al. Exome sequence analysis suggests that genetic burden contributes to phenotypic variability and complex neuropathy. Cell Rep 12: 1169–1183, 2015. doi: 10.1016/j.celrep.2015.07.023. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 195.Laquerriere A, Maluenda J, Camus A, Fontenas L, Dieterich K, Nolent F, et al. Mutations in CNTNAP1 and ADCY6 are responsible for severe arthrogryposis multiplex congenita with axoglial defects. Hum Mol Genet 23: 2279–2289, 2014. doi: 10.1093/hmg/ddt618. [DOI] [PubMed] [Google Scholar]
- 196.Yuen RK, Merico D, Cao H, Pellecchia G, Alipanahi B, Thiruvahindrapuram B, et al. Genome-wide characteristics of de novo mutations in autism. NPJ Genom Med 1: 160271–1602710, 2016. doi: 10.1038/npjgenmed.2016.27. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 197.Agolini E, Cherchi C, Bellacchio E, Martinelli D, Cocciadiferro D, Cutrera R, Chiarini Testa MB, Barone C, Bianca S, Novelli A. Expanding the clinical and molecular spectrum of lethal congenital contracture syndrome 8 associated with biallelic variants of ADCY6. Clin Genet 97: 649–654, 2020. doi: 10.1111/cge.13691. [DOI] [PubMed] [Google Scholar]
- 198.Ikoma E, Tsunematsu T, Nakazawa I, Shiwa T, Hibi K, Ebina T, Mochida Y, Toya Y, Hori H, Uchino K, Minamisawa S, Kimura K, Umemura S, Ishikawa Y. Polymorphism of the type 6 adenylyl cyclase gene and cardiac hypertrophy. J Cardiovasc Pharmacol 42: S27–S32, 2003. doi: 10.1097/00005344-200312001-00008. [DOI] [PubMed] [Google Scholar]
- 199.Hellevuo K, Welborn R, Menninger JA, Tabakoff B. Human adenylyl cyclase type 7 contains polymorphic repeats in the 3' untranslated region: investigations of association with alcoholism. Am J Med Genet 74: 95–98, 1997. doi:. [DOI] [PubMed] [Google Scholar]
- 200.Hines LM, Hoffman PL, Bhave S, Saba L, Kaiser A, Snell L, Goncharov I, LeGault L, Dongier M, Grant B, Pronko S, Martinez L, Yoshimura M, Tabakoff B; World Health Organization/International Society for Biomedical Research on Alcoholism Study on State and Trait Markers of Alcohol Use and Dependence Investigators. A sex-specific role of type VII adenylyl cyclase in depression. J Neurosci 26: 12609–12619, 2006. doi: 10.1523/JNEUROSCI.1040-06.2006. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 201.Joeyen-Waldorf J, Nikolova YS, Edgar N, Walsh C, Kota R, Lewis DA, Ferrell R, Manuck SB, Hariri AR, Sibille E. Adenylate cyclase 7 is implicated in the biology of depression and modulation of affective neural circuitry. Biol Psychiatry 71: 627–632, 2012. doi: 10.1016/j.biopsych.2011.11.029. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 202.Li YR, Li J, Zhao SD, Bradfield JP, Mentch FD, Maggadottir SM, Hou C, Abrams DJ, et al. Meta-analysis of shared genetic architecture across ten pediatric autoimmune diseases. Nat Med 21: 1018–1027, 2015. doi: 10.1038/nm.3933. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 203.Luo Y, de Lange KM, Jostins L, Moutsianas L, Randall J, Kennedy NA, Lamb CA, McCarthy S, Ahmad T, Edwards C, Serra EG, Hart A, Hawkey C, Mansfield JC, Mowat C, Newman WG, Nichols S, Pollard M, Satsangi J, Simmons A, Tremelling M, Uhlig H, Wilson DC, Lee JC, Prescott NJ, Lees CW, Mathew CG, Parkes M, Barrett JC, Anderson CA. Exploring the genetic architecture of inflammatory bowel disease by whole-genome sequencing identifies association at ADCY7. Nat Genet 49: 186–192, 2017. doi: 10.1038/ng.3761. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 204.Xu B, Roos JL, Dexheimer P, Boone B, Plummer B, Levy S, Gogos JA, Karayiorgou M. Exome sequencing supports a de novo mutational paradigm for schizophrenia. Nat Genet 43: 864–868, 2011. doi: 10.1038/ng.902. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 205.Meng X, Hou X, Wang P, Glessner JT, Qu HQ, March ME, Zhang S, Qi X, Zhu C, Nguyen K, Gao X, Li X, Liu Y, Zhou W, Zhang S, Li J, Sun Y, Yang J, Sleiman PM, Xia Q, Hakonarson H, Li J. Association of novel rare coding variants with juvenile idiopathic arthritis. Ann Rheum Dis 80: 626–631, 2021. doi: 10.1136/annrheumdis-2020-218359. [DOI] [PubMed] [Google Scholar]
- 206.Kim N, Kim KH, Lim WJ, Kim J, Kim SA, Yoo HJ. Whole exome sequencing identifies novel de novo variants interacting with six gene networks in autism spectrum disorder . Genes (Basel) 12: 1, 2020. doi: 10.3390/genes12010001. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 207.Zhang P, Xiang N, Chen Y, Sliwerska E, McInnis MG, Burmeister M, Zöllner S. Family-based association analysis to finemap bipolar linkage peak on chromosome 8q24 using 2,500 genotyped SNPs and 15,000 imputed SNPs. Bipolar Disord 12: 786–792, 2010. doi: 10.1111/j.1399-5618.2010.00883.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 208.de Mooij-van Malsen AJ, van Lith HA, Oppelaar H, Hendriks J, de Wit M, Kostrzewa E, Breen G, Collier DA, Olivier B, Kas MJ. Interspecies trait genetics reveals association of Adcy8 with mouse avoidance behavior and a human mood disorder. Biol Psychiatry 66: 1123–1130, 2009. doi: 10.1016/j.biopsych.2009.06.016. [DOI] [PubMed] [Google Scholar]
- 209.Wolf EJ, Rasmusson AM, Mitchell KS, Logue MW, Baldwin CT, Miller MW. A genome-wide association study of clinical symptoms of dissociation in a trauma-exposed sample. Depress Anxiety 31: 352–360, 2014. doi: 10.1002/da.22260. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 210.Medack A, Brænne I, Stark K, Hengstenberg C, Strom T, Meitinger T, Schunkert H, Erdmann J. Missense mutation in Adcy8 linked to myocardial infarction (Abstract). Circulation 126: A14242, 2012. [Google Scholar]
- 211.Small KM, Brown KM, Theiss CT, Seman CA, Weiss ST, Liggett SB. An Ile to Met polymorphism in the catalytic domain of adenylyl cyclase type 9 confers reduced beta2-adrenergic receptor stimulation. Pharmacogenetics 13: 535–541, 2003. doi: 10.1097/00008571-200309000-00002. [DOI] [PubMed] [Google Scholar]
- 212.Kim SH, Ye YM, Lee HY, Sin HJ, Park HS. Combined pharmacogenetic effect of ADCY9 and ADRB2 gene polymorphisms on the bronchodilator response to inhaled combination therapy. J Clin Pharm Ther 36: 399–405, 2011. doi: 10.1111/j.1365-2710.2010.01196.x. [DOI] [PubMed] [Google Scholar]
- 213.Lee YH, Gyu Song G. Genome-wide pathway analysis in pancreatic cancer. J BUON 20: 1565–1575, 2015. ] [PubMed] [Google Scholar]
- 214.Berndt SI, Gustafsson S, Magi R, Ganna A, Wheeler E, Feitosa MF, Justice AE, et al. Genome-wide meta-analysis identifies 11 new loci for anthropometric traits and provides insights into genetic architecture. Nat Genet 45: 501–512, 2013. doi: 10.1038/ng.2606. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 215.Manjurano A, Clark TG, Nadjm B, Mtove G, Wangai H, Sepulveda N, Campino SG, Maxwell C, Olomi R, Rockett KR, Jeffreys A; MalariaGen Consortium, Riley EM, Reyburn H, Drakeley C. Candidate human genetic polymorphisms and severe malaria in a Tanzanian population. PLoS One 7: e47463, 2012. doi: 10.1371/journal.pone.0047463. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 216.Maiga B, Dolo A, Touré O, Dara V, Tapily A, Campino S, Sepulveda N, Risley P, Silva N, Corran P, Rockett KA, Kwiatkowski D; MalariaGEN Consortium, Clark TG, Troye-Blomberg M, Doumbo OK. Human candidate polymorphisms in sympatric ethnic groups differing in malaria susceptibility in Mali. PLoS One 8: e75675, 2013. doi: 10.1371/journal.pone.0075675. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 217.Tardif JC, Rhéaume E, Lemieux Perreault LP, Grégoire JC, Feroz Zada Y, Asselin G, Provost S, Barhdadi A, Rhainds D, L’Allier PL, Ibrahim R, Upmanyu R, Niesor EJ, Benghozi R, Suchankova G, Laghrissi-Thode F, Guertin MC, Olsson AG, Mongrain I, Schwartz GG, Dubé MP. Pharmacogenomic determinants of the cardiovascular effects of dalcetrapib. Circ Cardiovasc Genet 8: 372–382, 2015. doi: 10.1161/CIRCGENETICS.114.000663. [DOI] [PubMed] [Google Scholar]
- 218.Gulsuner S, Walsh T, Watts AC, Lee MK, Thornton AM, Casadei S, Rippey C, Shahin H; Consortium on the Genetics of Schizophrenia (COGS), PAARTNERS Study Group, Nimgaonkar VL, Go RC, Savage RM, Swerdlow NR, Gur RE, Braff DL, King MC, McClellan JM. Spatial and temporal mapping of de novo mutations in schizophrenia to a fetal prefrontal cortical network. Cell 154: 518–529, 2013. doi: 10.1016/j.cell.2013.06.049. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 219.Ghanbari M, Franco OH, de Looper HW, Hofman A, Erkeland SJ, Dehghan A. Genetic variations in microRNA-binding sites affect microRNA-mediated regulation of several genes associated with cardio-metabolic phenotypes. Circ Cardiovasc Genet 8: 473–486, 2015. doi: 10.1161/CIRCGENETICS.114.000968. [DOI] [PubMed] [Google Scholar]
- 220.Wang T, Guo H, Xiong B, Stessman HA, Wu H, Coe BP, Turner TN, Liu Y, Zhao W, Hoekzema K, Vives L, Xia L, Tang M, Ou J, Chen B, Shen Y, Xun G, Long M, Lin J, Kronenberg ZN, Peng Y, Bai T, Li H, Ke X, Hu Z, Zhao J, Zou X, Xia K, Eichler EE. De novo genic mutations among a Chinese autism spectrum disorder cohort. Nat Commun 7: 13316, 2016. doi: 10.1038/ncomms13316. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 221.Stenson PD, Mort M, Ball EV, Chapman M, Evans K, Azevedo L, Hayden M, Heywood S, Millar DS, Phillips AD, Cooper DN. The Human Gene Mutation Database (HGMDR): optimizing its use in a clinical diagnostic or research setting. Hum Genet 139: 1197–1207, 2020. doi: 10.1007/s00439-020-02199-3. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 222.Lazary J, Eszlari N, Kriko E, Tozser D, Dome P, Deakin JF, Juhasz G, Bagdy G. Genetic analyses of the endocannabinoid pathway in association with affective phenotypic variants. Neurosci Lett 744: 135600, 2021. doi: 10.1016/j.neulet.2020.135600. [DOI] [PubMed] [Google Scholar]
- 223.Ravary A, Muzerelle A, Hervé D, Pascoli V, Ba-Charvet KN, Girault JA, Welker E, Gaspar P. Adenylate cyclase 1 as a key actor in the refinement of retinal projection maps. J Neurosci 23: 2228–2238, 2003. doi: 10.1523/JNEUROSCI.23-06-02228.2003. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 224.Abdel-Majid RM, Leong WL, Schalkwyk LC, Smallman DS, Wong ST, Storm DR, Fine A, Dobson MJ, Guernsey DL, Neumann PE. Loss of adenylyl cyclase I activity disrupts patterning of mouse somatosensory cortex. Nat Genet 19: 289–291, 1998. doi: 10.1038/980. [DOI] [PubMed] [Google Scholar]
- 225.Arakawa H, Akkentli F, Erzurumlu RS. Region-specific disruption of adenylate cyclase type 1 gene differentially affects somatosensorimotor behaviors in mice(1,2,3). eNeuro 1: ENEURO.0007-14.2014, 2014. doi: 10.1523/ENEURO.0007-14.2014. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 226.Sethna F, Feng W, Ding Q, Robison AJ, Feng Y, Wang H. Enhanced expression of ADCY1 underlies aberrant neuronal signalling and behaviour in a syndromic autism model. Nat Commun 8: 14359, 2017. doi: 10.1038/ncomms14359. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 227.Braeunig JH, Schweda F, Han PL, Seifert R. Similarly potent inhibition of adenylyl cyclase by P-site inhibitors in hearts from wild type and AC5 knockout mice. PLoS One 8: e68009, 2013. doi: 10.1371/journal.pone.0068009. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 228.Yang LK, Lu L, Feng B, Wang XS, Yue J, Li XB, Zhuo M, Liu SB. FMRP acts as a key messenger for visceral pain modulation. Mol Pain 16: 1744806920972241, 2020. doi: 10.1177/1744806920972241. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 229.Wei F, Qiu CS, Kim SJ, Muglia L, Maas JW, Pineda VV, Xu HM, Chen ZF, Storm DR, Muglia LJ, Zhuo M. Genetic elimination of behavioral sensitization in mice lacking calmodulin-stimulated adenylyl cyclases. Neuron 36: 713–726, 2002. doi: 10.1016/S0896-6273(02)01019-X. [DOI] [PubMed] [Google Scholar]
- 230.Kang WB, Yang Q, Guo YY, Wang L, Wang DS, Cheng Q, Li XM, Tang J, Zhao JN, Liu G, Zhuo M, Zhao MG. Analgesic effects of adenylyl cyclase inhibitor NB001 on bone cancer pain in a mouse model. Mol Pain 12: 1744806916652409, 2016. doi: 10.1177/1744806916652409. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 231.Griggs RB, Laird DE, Donahue RR, Fu W, Taylor BK. Methylglyoxal requires AC1 and TRPA1 to produce pain and spinal neuron activation. Front Neurosci 11: 679, 2017. doi: 10.3389/fnins.2017.00679. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 232.Griggs RB, Santos DF, Laird DE, Doolen S, Donahue RR, Wessel CR, Fu W, Sinha GP, Wang P, Zhou J, Brings S, Fleming T, Nawroth PP, Susuki K, Taylor BK. Methylglyoxal and a spinal TRPA1-AC1-Epac cascade facilitate pain in the db/db mouse model of type 2 diabetes. Neurobiol Dis 127: 76–86, 2019. doi: 10.1016/j.nbd.2019.02.019. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 233.Liu Y, Chen QY, Lee JH, Li XH, Yu S, Zhuo M. Cortical potentiation induced by calcitonin gene-related peptide (CGRP) in the insular cortex of adult mice. Mol Brain 13: 36, 2020. doi: 10.1186/s13041-020-00580-x. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 234.Liu SB, Wang XS, Yue J, Yang L, Li XH, Hu LN, Lu JS, Song Q, Zhang K, Yang Q, Zhang MM, Bernabucci M, Zhao MG, Zhuo M. Cyclic AMP-dependent positive feedback signaling pathways in the cortex contributes to visceral pain. J Neurochem 153: 252–263, 2020. doi: 10.1111/jnc.14903. [DOI] [PubMed] [Google Scholar]
- 235.Liu SB, Zhang MM, Cheng LF, Shi J, Lu JS, Zhuo M. Long-term upregulation of cortical glutamatergic AMPA receptors in a mouse model of chronic visceral pain. Mol Brain 8: 76, 2015. doi: 10.1186/s13041-015-0169-z. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 236.Zhao MG, Ko SW, Wu LJ, Toyoda H, Xu H, Quan J, Li J, Jia Y, Ren M, Xu ZC, Zhuo M. Enhanced presynaptic neurotransmitter release in the anterior cingulate cortex of mice with chronic pain. J Neurosci 26: 8923–8930, 2006. doi: 10.1523/JNEUROSCI.2103-06.2006. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 237.Xu H, Wu LJ, Wang H, Zhang X, Vadakkan KI, Kim SS, Steenland HW, Zhuo M. Presynaptic and postsynaptic amplifications of neuropathic pain in the anterior cingulate cortex. J Neurosci 28: 7445–7453, 2008. doi: 10.1523/JNEUROSCI.1812-08.2008. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 238.Qiu S, Zhang M, Liu Y, Guo Y, Zhao H, Song Q, Zhao M, Huganir RL, Luo J, Xu H, Zhuo M. GluA1 phosphorylation contributes to postsynaptic amplification of neuropathic pain in the insular cortex. J Neurosci 34: 13505–13515, 2014. doi: 10.1523/JNEUROSCI.1431-14.2014. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 239.Miao HH, Li XH, Chen QY, Zhuo M. Calcium-stimulated adenylyl cyclase subtype 1 is required for presynaptic long-term potentiation in the insular cortex of adult mice. Mol Pain 15: 1744806919842961, 2019. doi: 10.1177/1744806919842961. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 240.Corder G, Doolen S, Donahue RR, Winter MK, Jutras BL, He Y, Hu X, Wieskopf JS, Mogil JS, Storm DR, Wang ZJ, McCarson KE, Taylor BK. Constitutive mu-opioid receptor activity leads to long-term endogenous analgesia and dependence. Science 341: 1394–1399, 2013. doi: 10.1126/science.1239403. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 241.Wei F, Vadakkan KI, Toyoda H, Wu LJ, Zhao MG, Xu H, Shum FW, Jia YH, Zhuo M. Calcium calmodulin-stimulated adenylyl cyclases contribute to activation of extracellular signal-regulated kinase in spinal dorsal horn neurons in adult rats and mice. J Neurosci 26: 851–861, 2006. doi: 10.1523/JNEUROSCI.3292-05.2006. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 242.Fu W, Wessel CR, Taylor BK. Neuropeptide Y tonically inhibits an NMDARAC1TRPA1/TRPV1 mechanism of the affective dimension of chronic neuropathic pain. Neuropeptides 80: 102024, 2020. doi: 10.1016/j.npep.2020.102024. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 243.Wang H, Xu H, Wu LJ, Kim SS, Chen T, Koga K, Descalzi G, Gong B, Vadakkan KI, Zhang X, Kaang BK, Zhuo M. Identification of an adenylyl cyclase inhibitor for treating neuropathic and inflammatory pain. Sci Transl Med 3: 65ra3, 2011. doi: 10.1126/scitranslmed.3001269. [DOI] [PubMed] [Google Scholar]
- 244.Brust TF, Alongkronrusmee D, Soto-Velasquez M, Baldwin TA, Ye Z, Dai M, Dessauer CW, van Rijn RM, Watts VJ. Identification of a selective small-molecule inhibitor of type 1 adenylyl cyclase activity with analgesic properties. Sci Signal 10: eaah5381, 2017. doi: 10.1126/scisignal.aah5381. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 245.Lübker C, Urbauer RJ, Moskovitz J, Dove S, Weisemann J, Fedorova M, Urbauer JL, Seifert R. Membranous adenylyl cyclase 1 activation is regulated by oxidation of N- and C-terminal methionine residues in calmodulin. Biochem Pharmacol 93: 196–209, 2015. doi: 10.1016/j.bcp.2014.11.007. [DOI] [PubMed] [Google Scholar]
- 246.Zachariou V, Liu R, LaPlant Q, Xiao G, Renthal W, Chan GC, Storm DR, Aghajanian G, Nestler EJ. Distinct roles of adenylyl cyclases 1 and 8 in opiate dependence: behavioral, electrophysiological, and molecular studies. Biol Psychiatry 63: 1013–1021, 2008. doi: 10.1016/j.biopsych.2007.11.021. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 247.Bosse KE, Oginsky MF, Susick LL, Ramalingam S, Ferrario CR, Conti AC. Adenylyl cyclase 1 is required for ethanol-induced locomotor sensitization and associated increases in NMDA receptor phosphorylation and function in the dorsal medial striatum. J Pharmacol Exp Ther 363: 148–155, 2017. doi: 10.1124/jpet.117.242321. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 248.Hwang CK, Chaurasia SS, Jackson CR, Chan GC, Storm DR, Iuvone PM. Circadian rhythm of contrast sensitivity is regulated by a dopamine-neuronal PAS-domain protein 2-adenylyl cyclase 1 signaling pathway in retinal ganglion cells. J Neurosci 33: 14989–14997, 2013. doi: 10.1523/JNEUROSCI.2039-13.2013. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 249.Jackson CR, Chaurasia SS, Hwang CK, Iuvone PM. Dopamine D4 receptor activation controls circadian timing of the adenylyl cyclase 1/cyclic AMP signaling system in mouse retina. Eur J Neurosci 34: 57–64, 2011. doi: 10.1111/j.1460-9568.2011.07734.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 250.Tang W, Ma W, Ding H, Lin M, Xiang L, Lin G, Zhang Z. Adenylyl cyclase 1 as a major isoform to generate cAMP signaling for apoA-1-mediated cholesterol efflux pathway. J Lipid Res 59: 635–645, 2018. doi: 10.1194/jlr.M082297. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 251.Lérias J, Pinto M, Benedetto R, Schreiber R, Amaral M, Aureli M, Kunzelmann K. Compartmentalized crosstalk of CFTR and TMEM16A (ANO1) through EPAC1 and ADCY1. Cell Signal 44: 10–19, 2018. doi: 10.1016/j.cellsig.2018.01.008. [DOI] [PubMed] [Google Scholar]
- 252.Schuler D, Lübker C, Lushington GH, Tang WJ, Shen Y, Richter M, Seifert R. Interactions of Bordetella pertussis adenylyl cyclase toxin CyaA with calmodulin mutants and calmodulin antagonists: comparison with membranous adenylyl cyclase I. Biochem Pharmacol 83: 839–848, 2012. doi: 10.1016/j.bcp.2012.01.005. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 253.Lübker C, Seifert R. Effects of 39 Compounds on calmodulin-regulated adenylyl cyclases AC1 and Bacillus anthracis edema factor. PLoS One 10: e0124017, 2015. doi: 10.1371/journal.pone.0124017. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 254.Wang H, Zhang M. The role of Ca2+-stimulated adenylyl cyclases in bidirectional synaptic plasticity and brain function. Rev Neurosci 23: 67–78, 2012. [DOI] [PubMed] [Google Scholar]
- 255.Wang H, Pineda VV, Chan GC, Wong ST, Muglia LJ, Storm DR. Type 8 adenylyl cyclase is targeted to excitatory synapses and required for mossy fiber long-term potentiation. J Neurosci 23: 9710–9718, 2003. doi: 10.1523/JNEUROSCI.23-30-09710.2003. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 256.Schaefer ML, Wong ST, Wozniak DF, Muglia LM, Liauw JA, Zhuo M, Nardi A, Hartman RE, Vogt SK, Luedke CE, Storm DR, Muglia LJ. Altered stress-induced anxiety in adenylyl cyclase type VIII-deficient mice. J Neurosci 20: 4809–4820, 2000. doi: 10.1523/JNEUROSCI.20-13-04809.2000. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 257.Zhang M, Moon C, Chan GC, Yang L, Zheng F, Conti AC, Muglia L, Muglia LJ, Storm DR, Wang H. Ca-stimulated type 8 adenylyl cyclase is required for rapid acquisition of novel spatial information and for working/episodic-like memory. J Neurosci 28: 4736–4744, 2008. doi: 10.1523/JNEUROSCI.1177-08.2008. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 258.Bernabucci M, Zhuo M. Calcium activated adenylyl cyclase AC8 but not AC1 is required for prolonged behavioral anxiety. Mol Brain 9: 60, 2016. doi: 10.1186/s13041-016-0239-x. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 259.Razzoli M, Andreoli M, Maraia G, Di Francesco C, Arban R. Functional role of calcium-stimulated adenylyl cyclase 8 in adaptations to psychological stressors in the mouse: implications for mood disorders. Neuroscience 170: 429–440, 2010. doi: 10.1016/j.neuroscience.2010.07.022. [DOI] [PubMed] [Google Scholar]
- 260.Moen JM, Matt MG, Ramirez C, Tarasov KV, Chakir K, Tarasova YS, Lukyanenko Y, Tsutsui K, Monfredi O, Morrell CH, Tagirova S, Yaniv Y, Huynh T, Pacak K, Ahmet I, Lakatta EG. Overexpression of a neuronal type adenylyl cyclase (type 8) in sinoatrial node markedly impacts heart rate and rhythm. Front Neurosci 13: 615, 2019. doi: 10.3389/fnins.2019.00615. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 261.Mougenot N, Mika D, Czibik G, Marcos E, Abid S, Houssaini A, Vallin B, Guellich A, Mehel H, Sawaki D, Vandecasteele G, Fischmeister R, Hajjar RJ, Dubois-Randé JL, Limon I, Adnot S, Derumeaux G, Lipskaia L. Cardiac adenylyl cyclase overexpression precipitates and aggravates age-related myocardial dysfunction. Cardiovasc Res 115: 1778–1790, 2019. doi: 10.1093/cvr/cvy306. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 262.Martin AC, Willoughby D, Ciruela A, Ayling LJ, Pagano M, Wachten S, Tengholm A, Cooper DM. Capacitative Ca2+ entry via Orai1 and stromal interacting molecule 1 (STIM1) regulates adenylyl cyclase type 8. Mol Pharmacol 75: 830–842, 2009. doi: 10.1124/mol.108.051748. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 263.Willoughby D, Everett KL, Halls ML, Pacheco J, Skroblin P, Vaca L, Klussmann E, Cooper DM. Direct binding between Orai1 and AC8 mediates dynamic interplay between Ca2+ and cAMP signaling. Sci Signal 5: ra29, 2012. doi: 10.1126/scisignal.2002299. [DOI] [PubMed] [Google Scholar]
- 264.Zhang X, Pathak T, Yoast R, Emrich S, Xin P, Nwokonko RM, Johnson M, Wu S, Delierneux C, Gueguinou M, Hempel N, Putney JW Jr, Gill DL, Trebak M. A calcium/cAMP signaling loop at the ORAI1 mouth drives channel inactivation to shape NFAT induction. Nat Commun 10: 1971, 2019. doi: 10.1038/s41467-019-09593-0. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 265.Sanchez-Collado J, Lopez JJ, Jardin I, Camello PJ, Falcon D, Regodon S, Salido GM, Smani T, Rosado JA. Adenylyl cyclase type 8 overexpression impairs phosphorylation-dependent Orai1 inactivation and promotes migration in MDA-MB-231 breast cancer cells. Cancers (Basel) 11: 1624, 2019. doi: 10.3390/cancers11111624. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 266.Visel A, Alvarez-Bolado G, Thaller C, Eichele G. Comprehensive analysis of the expression patterns of the adenylate cyclase gene family in the developing and adult mouse brain. J Comp Neurol 496: 684–697, 2006. doi: 10.1002/cne.20953. [DOI] [PubMed] [Google Scholar]
- 267.Krishnan V, Graham A, Mazei-Robison MS, Lagace DC, Kim KS, Birnbaum S, Eisch AJ, Han PL, Storm DR, Zachariou V, Nestler EJ. Calcium-sensitive adenylyl cyclases in depression and anxiety: behavioral and biochemical consequences of isoform targeting. Biol Psychiatry 64: 336–343, 2008. doi: 10.1016/j.biopsych.2008.03.026. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 268.Conti AC, Lowing JL, Susick LL, Bowen SE. Investigation of calcium-stimulated adenylyl cyclases 1 and 8 on toluene and ethanol neurobehavioral actions. Neurotoxicol Teratol 34: 481–488, 2012. doi: 10.1016/j.ntt.2012.06.005. [DOI] [PubMed] [Google Scholar]
- 269.Bosse KE, Ghoddoussi F, Eapen AT, Charlton JL, Susick LL, Desai K, Berkowitz BA, Perrine SA, Conti AC. Calcium/calmodulin-stimulated adenylyl cyclases 1 and 8 regulate reward-related brain activity and ethanol consumption. Brain Imaging Behav 13: 396–407, 2019. doi: 10.1007/s11682-018-9856-6. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 270.Maas JW Jr, Vogt SK, Chan GC, Pineda VV, Storm DR, Muglia LJ. Calcium-stimulated adenylyl cyclases are critical modulators of neuronal ethanol sensitivity. J Neurosci 25: 4118–4126, 2005. doi: 10.1523/JNEUROSCI.4273-04.2005. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 271.Susick LL, Lowing JL, Bosse KE, Hildebrandt CC, Chrumka AC, Conti AC. Adenylyl cyclases 1 and 8 mediate select striatal-dependent behaviors and sensitivity to ethanol stimulation in the adolescent period following acute neonatal ethanol exposure. Behav Brain Res 269: 66–74, 2014. doi: 10.1016/j.bbr.2014.04.031. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 272.DiRocco DP, Scheiner ZS, Sindreu CB, Chan GC, Storm DR. A role for calmodulin-stimulated adenylyl cyclases in cocaine sensitization. J Neurosci 29: 2393–2403, 2009. doi: 10.1523/JNEUROSCI.4356-08.2009. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 273.Li S, Lee ML, Bruchas MR, Chan GC, Storm DR, Chavkin C. Calmodulin-stimulated adenylyl cyclase gene deletion affects morphine responses. Mol Pharmacol 70: 1742–1749, 2006. doi: 10.1124/mol.106.025783. [DOI] [PubMed] [Google Scholar]
- 274.Bosse KE, Charlton JL, Susick LL, Newman B, Eagle AL, Mathews TA, Perrine SA, Conti AC. Deficits in behavioral sensitization and dopaminergic responses to methamphetamine in adenylyl cyclase 1/8-deficient mice. J Neurochem 135: 1218–1231, 2015. doi: 10.1111/jnc.13235. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 275.Capel RA, Bose SJ, Collins TP, Rajasundaram S, Ayagama T, Zaccolo M, Burton RB, Terrar DA. IP3-mediated Ca2+ release regulates atrial Ca2+ transients and pacemaker function by stimulation of adenylyl cyclases. Am J Physiol Heart Circ Physiol 320: H95–H107, 2021. doi: 10.1152/ajpheart.00380.2020. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 276.Robinson RB, Dun W, Boyden PA. Autonomic modulation of sinoatrial node: role of pacemaker current and calcium sensitive adenylyl cyclase isoforms. Prog Biophys Mol Biol 166: 22–28, 2021. doi: 10.1016/j.pbiomolbio.2020.08.004. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 277.Mattick P, Parrington J, Odia E, Simpson A, Collins T, Terrar D. Ca2+-stimulated adenylyl cyclase isoform AC1 is preferentially expressed in guinea-pig sino-atrial node cells and modulates the If pacemaker current. J Physiol 582: 1195–1203, 2007. doi: 10.1113/jphysiol.2007.133439. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 278.Lipskaia L, Defer N, Esposito G, Hajar I, Garel MC, Rockman HA, Hanoune J. Enhanced cardiac function in transgenic mice expressing a Ca2+-stimulated adenylyl cyclase. Circ Res 86: 795–801, 2000. doi: 10.1161/01.RES.86.7.795. [DOI] [PubMed] [Google Scholar]
- 279.Vedantham V, Galang G, Evangelista M, Deo RC, Srivastava D. RNA sequencing of mouse sinoatrial node reveals an upstream regulatory role for Islet-1 in cardiac pacemaker cells. Circ Res 116: 797–803, 2015. doi: 10.1161/CIRCRESAHA.116.305913. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 280.Bakalyar HA, Reed RR. Identification of a specialized adenylyl cyclase that may mediate odorant detection. Science 250: 1403–1406, 1990. doi: 10.1126/science.2255909. [DOI] [PubMed] [Google Scholar]
- 281.Xia Z, Choi EJ, Wang F, Storm DR. The type III calcium/calmodulin-sensitive adenylyl cyclase is not specific to olfactory sensory neurons. Neurosci Lett 144: 169–173, 1992. doi: 10.1016/0304-3940(92)90742-P. [DOI] [PubMed] [Google Scholar]
- 282.Bishop GA, Berbari NF, Lewis J, Mykytyn K. Type III adenylyl cyclase localizes to primary cilia throughout the adult mouse brain. J Comp Neurol 505: 562–571, 2007. doi: 10.1002/cne.21510. [DOI] [PubMed] [Google Scholar]
- 283.Wong ST, Trinh K, Hacker B, Chan GC, Lowe G, Gaggar A, Xia Z, Gold GH, Storm DR. Disruption of the type III adenylyl cyclase gene leads to peripheral and behavioral anosmia in transgenic mice. Neuron 27: 487–497, 2000. doi: 10.1016/S0896-6273(00)00060-X. [DOI] [PubMed] [Google Scholar]
- 284.Col JA, Matsuo T, Storm DR, Rodriguez I. Adenylyl cyclase-dependent axonal targeting in the olfactory system. Development 134: 2481–2489, 2007. doi: 10.1242/dev.006346. [DOI] [PubMed] [Google Scholar]
- 285.Zhang Z, Yang D, Zhang M, Zhu N, Zhou Y, Storm DR, Wang Z. Deletion of type 3 adenylyl cyclase perturbs the postnatal maturation of olfactory sensory neurons and olfactory cilium ultrastructure in mice. Front Cell Neurosci 11: 1, 2017. doi: 10.3389/fncel.2017.00001. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 286.Sinnarajah S, Dessauer CW, Srikumar D, Chen J, Yuen J, Yilma S, Dennis JC, Morrison EE, Vodyanoy V, Kehrl JH. RGS2 regulates signal transduction in olfactory neurons by attenuating activation of adenylyl cyclase III. Nature 409: 1051–1055, 2001. doi: 10.1038/35059104. [DOI] [PubMed] [Google Scholar]
- 287.Chen X, Xia Z, Storm DR. Stimulation of electro-olfactogram responses in the main olfactory epithelia by airflow depends on the type 3 adenylyl cyclase. J Neurosci 32: 15769–15778, 2012. doi: 10.1523/JNEUROSCI.2180-12.2012. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 288.Wang Z, Balet Sindreu C, Li V, Nudelman A, Chan GC, Storm DR. Pheromone detection in male mice depends on signaling through the type 3 adenylyl cyclase in the main olfactory epithelium. J Neurosci 26: 7375–7379, 2006. doi: 10.1523/JNEUROSCI.1967-06.2006. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 289.Wang Z, Storm DR. Maternal behavior is impaired in female mice lacking type 3 adenylyl cyclase. Neuropsychopharmacology 36: 772–781, 2011. doi: 10.1038/npp.2010.211. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 290.Wang Z, Phan T, Storm DR. The type 3 adenylyl cyclase is required for novel object learning and extinction of contextual memory: role of cAMP signaling in primary cilia. J Neurosci 31: 5557–5561, 2011. doi: 10.1523/JNEUROSCI.6561-10.2011. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 291.Liu X, Zhou Y, Yang D, Li S, Liu X, Wang Z. Type 3 adenylyl cyclase in the MOE is involved in learning and memory in mice. Behav Brain Res 383: 112533, 2020. doi: 10.1016/j.bbr.2020.112533. [DOI] [PubMed] [Google Scholar]
- 292.Katsel PL, Tagliente TM, Schwarz TE, Craddock-Royal BD, Patel ND, Maayani S. Molecular and biochemical evidence for the presence of type III adenylyl cyclase in human platelets. Platelets 14: 21–33, 2003. doi: 10.1080/0953710021000062905. [DOI] [PubMed] [Google Scholar]
- 293.Hines LM, Tabakoff B; WHO/ISBRA Study on State and Trait Markers of Alcohol Use and Dependence Investigators. Platelet adenylyl cyclase activity: a biological marker for major depression and recent drug use. Biol Psychiatry 58: 955–962, 2005. doi: 10.1016/j.biopsych.2005.05.040. [DOI] [PubMed] [Google Scholar]
- 294.Menninger JA, Tabakoff B. Forskolin-stimulated platelet adenylyl cyclase activity is lower in persons with major depression. Biol Psychiatry 42: 30–38, 1997. doi: 10.1016/S0006-3223(96)00245-4. [DOI] [PubMed] [Google Scholar]
- 295.Redei EE, Andrus BM, Kwasny MJ, Seok J, Cai X, Ho J, Mohr DC. Blood transcriptomic biomarkers in adult primary care patients with major depressive disorder undergoing cognitive behavioral therapy. Transl Psychiatry 4: e442, 2014. doi: 10.1038/tp.2014.66. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 296.Chen X, Luo J, Leng Y, Yang Y, Zweifel LS, Palmiter RD, Storm DR. Ablation of type III adenylyl cyclase in mice causes reduced neuronal activity, altered sleep pattern, and depression-like phenotypes. Biol Psychiatry 80: 836–848, 2016. doi: 10.1016/j.biopsych.2015.12.012. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 297.Yang XY, Ma ZL, Storm DR, Cao H, Zhang YQ. Selective ablation of type 3 adenylyl cyclase in somatostatin-positive interneurons produces anxiety- and depression-like behaviors in mice. World J Psychiatry 11: 35–49, 2021. doi: 10.5498/wjp.v11.i2.35. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 298.Liu X, Zhou Y, Li S, Yang D, Jiao M, Liu X, Wang Z. Type 3 adenylyl cyclase in the main olfactory epithelium participates in depression-like and anxiety-like behaviours. J Affect Disord 268: 28–38, 2020. doi: 10.1016/j.jad.2020.02.041. [DOI] [PubMed] [Google Scholar]
- 299.Wang Z, Li V, Chan GC, Phan T, Nudelman AS, Xia Z, Storm DR. Adult type 3 adenylyl cyclase-deficient mice are obese. PLoS One 4: e6979, 2009. doi: 10.1371/journal.pone.0006979. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 300.Palkovits M. Hypothalamic regulation of food intake. Ideggyogy Sz 56: 288–302, 2003. [PubMed] [Google Scholar]
- 301.Cao H, Chen X, Yang Y, Storm DR. Disruption of type 3 adenylyl cyclase expression in the hypothalamus leads to obesity. Integr Obes Diabetes 2: 225–228, 2016. doi: 10.15761/IOD.1000149. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 302.Siljee JE, Wang Y, Bernard AA, Ersoy BA, Zhang S, Marley A, Von Zastrow M, Reiter JF, Vaisse C. Subcellular localization of MC4R with ADCY3 at neuronal primary cilia underlies a common pathway for genetic predisposition to obesity. Nat Genet 50: 180–185, 2018. doi: 10.1038/s41588-017-0020-9. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 303.Tong T, Shen Y, Lee HW, Yu R, Park T. Adenylyl cyclase 3 haploinsufficiency confers susceptibility to diet-induced obesity and insulin resistance in mice. Sci Rep 6: 34179, 2016. doi: 10.1038/srep34179. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 304.Pitman JL, Wheeler MC, Lloyd DJ, Walker JR, Glynne RJ, Gekakis N. A gain-of-function mutation in adenylate cyclase 3 protects mice from diet-induced obesity. PLoS One 9: e110226, 2014. doi: 10.1371/journal.pone.0110226. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 305.Bogard AS, Xu C, Ostrom RS. Human bronchial smooth muscle cells express adenylyl cyclase isoforms 2, 4, and 6 in distinct membrane microdomains. J Pharmacol Exp Ther 337: 209–217, 2011. doi: 10.1124/jpet.110.177923. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 306.Wang HY, Burns LH. Gbetagamma that interacts with adenylyl cyclase in opioid tolerance originates from a Gs protein. J Neurobiol 66: 1302–1310, 2006. doi: 10.1002/neu.20286. [DOI] [PubMed] [Google Scholar]
- 307.Bogard AS, Adris P, Ostrom R. Adenylyl cyclase 2 (AC2) selectively couples to EP2 receptors while adenylyl cyclase 3 (AC3) is not receptor regulated in airway smooth muscle. J Pharmacol Exp Ther 342: 586–595, 2012. doi: 10.1124/jpet.112.193425. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 308.Suárez-Rama JJ, Arrojo M, Sobrino B, Amigo J, Brenlla J, Agra S, Paz E, Brión M, Carracedo Á, Páramo M, Costas J. Resequencing and association analysis of coding regions at twenty candidate genes suggest a role for rare risk variation at AKAP9 and protective variation at NRXN1 in schizophrenia susceptibility. J Psychiatr Res 66-67: 38–44, 2015. doi: 10.1016/j.jpsychires.2015.04.013. [DOI] [PubMed] [Google Scholar]
- 309.Conley JM, Brand CS, Bogard AS, Pratt EP, Xu R, Hockerman GH, Ostrom RS, Dessauer CW, Watts VJ. Development of a high-throughput screening paradigm for the discovery of small-molecule modulators of adenylyl cyclase: identification of an adenylyl cyclase 2 inhibitor. J Pharmacol Exp Ther 347: 276–287, 2013. doi: 10.1124/jpet.113.207449. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 310.Pinto C, Papa D, Hübner M, Mou TC, Lushington GH, Seifert R. Activation and inhibition of adenylyl cyclase isoforms by forskolin analogs. J Pharmacol Exp Ther 325: 27–36, 2008. doi: 10.1124/jpet.107.131904. [DOI] [PubMed] [Google Scholar]
- 311.Kinast L, von der Ohe J, Burhenne H, Seifert R. Impairment of adenylyl cyclase 2 function and expression in hypoxanthine phosphoribosyltransferase-deficient rat B103 neuroblastoma cells as model for Lesch-Nyhan disease: BODIPY-forskolin as pharmacological tool. Naunyn Schmiedebergs Arch Pharmacol 385: 671–683, 2012. doi: 10.1007/s00210-012-0759-6. [DOI] [PubMed] [Google Scholar]
- 312.Pinto CS, Seifert R. Decreased GTP-stimulated adenylyl cyclase activity in HPRT-deficient human and mouse fibroblast and rat B103 neuroblastoma cell membranes. J Neurochem 96: 454–459, 2006. doi: 10.1111/j.1471-4159.2005.03570.x. [DOI] [PubMed] [Google Scholar]
- 313.Genis-Mendoza A, Gallegos-Silva I, Tovilla-Zarate CA, López-Narvaez L, González-Castro TB, Hernández-Diáz Y, López-Casamichana M, Nicolini H, Morales-Mulia S. Comparative analysis of gene expression profiles involved in calcium signaling pathways using the NLVH animal model of schizophrenia. J Mol Neurosci 64: 111–116, 2018. doi: 10.1007/s12031-017-1013-y. [DOI] [PubMed] [Google Scholar]
- 314.Kittikulsuth W, Stuart D, Kohan DE. Adenylyl cyclase 4 does not regulate collecting duct water and sodium handling. Physiol Rep 2: e00277, 2014. doi: 10.1002/phy2.277. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 315.Olianas MC, Dedoni S, Onali P. Coincidence signaling of dopamine D1-like and M1 muscarinic receptors in the regulation of cyclic AMP formation and CREB phosphorylation in mouse prefrontal cortex. Neurosignals 21: 61–74, 2013. doi: 10.1159/000335208. [DOI] [PubMed] [Google Scholar]
- 316.Chen R, Zeng L, Zhu S, Liu J, Zeh HJ, Kroemer G, Wang H, Billiar TR, Jiang J, Tang D, Kang R. cAMP metabolism controls caspase-11 inflammasome activation and pyroptosis in sepsis. Sci Adv 5: eaav5562, 2019. doi: 10.1126/sciadv.aav5562. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 317.Fan Y, Mu J, Huang M, Imani S, Wang Y, Lin S, Fan J, Wen Q. Epigenetic identification of ADCY4 as a biomarker for breast cancer: an integrated analysis of adenylate cyclases. Epigenomics 11: 1561–1579, 2019. doi: 10.2217/epi-2019-0207. [DOI] [PubMed] [Google Scholar]
- 318.Price T, Brust TF. Adenylyl cyclase 7 and neuropsychiatric disorders: a new target for depression? Pharmacol Res 143: 106–112, 2019. doi: 10.1016/j.phrs.2019.03.015. [DOI] [PubMed] [Google Scholar]
- 319.Simko V, Iuliano F, Sevcikova A, Labudova M, Barathova M, Radvak P, Pastorekova S, Pastorek J, Csaderova L. Hypoxia induces cancer-associated cAMP/PKA signalling through HIF-mediated transcriptional control of adenylyl cyclases VI and VII. Sci Rep 7: 10121, 2017. doi: 10.1038/s41598-017-09549-8. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 320.He B, Chang Y, Yang C, Zhang Z, Xu G, Feng X, Zhuang L. Adenylate cyclase 7 regulated by miR-192 promotes ATRA-induced differentiation of acute promyelocytic leukemia cells. Biochem Biophys Res Commun 506: 543–547, 2018. doi: 10.1016/j.bbrc.2018.10.125. [DOI] [PubMed] [Google Scholar]
- 321.Chen J, DeVivo M, Dingus J, Harry A, Li J, Sui J, Carty DJ, Blank JL, Exton JH, Stoffel RH. A region of adenylyl cyclase 2 critical for regulation by G protein beta gamma subunits. Science 268: 1166–1169, 1995. doi: 10.1126/science.7761832. [DOI] [PubMed] [Google Scholar]
- 322.Avidor-Reiss T, Nevo I, Saya D, Bayewitch M, Vogel Z. Opiate-induced adenylyl cyclase superactivation is isozyme-specific. J Biol Chem 272: 5040–5047, 1997. doi: 10.1074/jbc.272.8.5040. [DOI] [PubMed] [Google Scholar]
- 323.Ehrlich AT, Furuyashiki T, Kitaoka S, Kakizuka A, Narumiya S. Prostaglandin E receptor EP1 forms a complex with dopamine D1 receptor and directs D1-induced cAMP production to adenylyl cyclase 7 through mobilizing G(betagamma) subunits in human embryonic kidney 293T cells. Mol Pharmacol 84: 476–486, 2013. doi: 10.1124/mol.113.087288. [DOI] [PubMed] [Google Scholar]
- 324.Jiang LI, Collins J, Davis R, Fraser ID, Sternweis PC. Regulation of cAMP responses by the G12/13 pathway converges on adenylyl cyclase VII. J Biol Chem 283: 23429–23439, 2008. doi: 10.1074/jbc.M803281200. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 325.Jiang LI, Wang JE, Sternweis PC. Regions on adenylyl cyclase VII required for selective regulation by the G13 pathway. Mol Pharmacol 83: 587–593, 2013. doi: 10.1124/mol.112.082446. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 326.Yoshimura M, Tabakoff B. Selective effects of ethanol on the generation of cAMP by particular members of the adenylyl cyclase family. Alcohol Clin Exp Res 19: 1435–1440, 1995. doi: 10.1111/j.1530-0277.1995.tb01004.x. [DOI] [PubMed] [Google Scholar]
- 327.Nelson EJ, Hellevuo K, Yoshimura M, Tabakoff B. Ethanol-induced phosphorylation and potentiation of the activity of type 7 adenylyl cyclase. Involvement of protein kinase C delta. J Biol Chem 278: 4552–4560, 2003. doi: 10.1074/jbc.M210386200. [DOI] [PubMed] [Google Scholar]
- 328.Rabbani M, Nelson EJ, Hoffman PL, Tabakoff B. Role of protein kinase C in ethanol-induced activation of adenylyl cyclase. Alcohol Clin Exp Res 23: 77–86, 1999. doi: 10.1111/j.1530-0277.1999.tb04026.x. [DOI] [PubMed] [Google Scholar]
- 329.Yoshimura M, Pearson S, Kadota Y, Gonzalez CE. Identification of ethanol responsive domains of adenylyl cyclase. Alcohol Clin Exp Res 30: 1824–1832, 2006. doi: 10.1111/j.1530-0277.2006.00219.x. [DOI] [PubMed] [Google Scholar]
- 330.Qualls-Creekmore E, Gupta R, Yoshimura M. The effect of alcohol on recombinant proteins derived from mammalian adenylyl cyclase. Biochem Biophys Rep 10: 157–164, 2017. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 331.Dokphrom U, Qualls-Creekmore E, Yoshimura M. Effects of alcohols on recombinant adenylyl cyclase type 7 expressed in bacteria. Alcohol Clin Exp Res 35: 1915–1922, 2011. doi: 10.1111/j.1530-0277.2011.01542.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 332.Pronko SP, Saba LM, Hoffman PL, Tabakoff B. Type 7 adenylyl cyclase-mediated hypothalamic-pituitary-adrenal axis responsiveness: influence of ethanol and sex. J Pharmacol Exp Ther 334: 44–52, 2010. doi: 10.1124/jpet.110.166793. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 333.Desrivières S, Pronko SP, Lourdusamy A, Ducci F, Hoffman PL, Wodarz N, Ridinger M, Rietschel M, Zelenika D, Lathrop M, Schumann G, Tabakoff B. Sex-specific role for adenylyl cyclase type 7 in alcohol dependence. Biol Psychiatry 69: 1100–1108, 2011. doi: 10.1016/j.biopsych.2011.01.037. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 334.Parsian A, Todd RD, Cloninger CR, Hoffman PL, Ovchinnikova L, Ikeda H, Tabakoff B. Platelet adenylyl cyclase activity in alcoholics and subtypes of alcoholics. WHO/ISBRA Study Clinical Centers. Alcohol Clin Exp Res 20: 745–751, 1996. doi: 10.1111/j.1530-0277.1996.tb01681.x. [DOI] [PubMed] [Google Scholar]
- 335.Tabakoff B, Hoffman PL, Lee JM, Saito T, Willard B, De Leon-Jones F. Differences in platelet enzyme activity between alcoholics and nonalcoholics. N Engl J Med 318: 134–139, 1988. doi: 10.1056/NEJM198801213180302. [DOI] [PubMed] [Google Scholar]
- 336.Janak PH, Tye KM. From circuits to behaviour in the amygdala. Nature 517: 284–292, 2015. doi: 10.1038/nature14188. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 337.Duan B, Davis R, Sadat EL, Collins J, Sternweis PC, Yuan D, Jiang LI. Distinct roles of adenylyl cyclase VII in regulating the immune responses in mice. J Immunol 185: 335–344, 2010. doi: 10.4049/jimmunol.0903474. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 338.Huang B, Zhao J, Lei Z, Shen S, Li D, Shen GX, Zhang GM, Feng ZH. miR-142-3p restricts cAMP production in CD4+CD25- T cells and CD4+CD25+ TREG cells by targeting AC9 mRNA. EMBO Rep 10: 180–185, 2009. doi: 10.1038/embor.2008.224. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 339.Taskén K, Stokka AJ. The molecular machinery for cAMP-dependent immunomodulation in T-cells. Biochem Soc Trans 34: 476–479, 2006. doi: 10.1042/BST0340476. [DOI] [PubMed] [Google Scholar]
- 340.Essayan DM. Cyclic nucleotide phosphodiesterases. J Allergy Clin Immunol 108: 671–680, 2001. doi: 10.1067/mai.2001.119555. [DOI] [PubMed] [Google Scholar]
- 341.Roper RL, Ludlow JW, Phipps RP. Prostaglandin E2 inhibits B lymphocyte activation by a cAMP-dependent mechanism: PGE-inducible regulatory proteins. Cell Immunol 154: 296–308, 1994. doi: 10.1006/cimm.1994.1079. [DOI] [PubMed] [Google Scholar]
- 342.Wall EA, Zavzavadjian JR, Chang MS, Randhawa B, Zhu X, Hsueh RC, Liu J, Driver A, Bao XR, Sternweis PC, Simon MI, Fraser ID. Suppression of LPS-induced TNF-alpha production in macrophages by cAMP is mediated by PKA-AKAP95-p105. Sci Signal 2: ra28, 2009. doi: 10.1126/scisignal.2000202. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 343.Jiang LI, Sternweis PC, Wang JE. Zymosan activates protein kinase A via adenylyl cyclase VII to modulate innate immune responses during inflammation. Mol Immunol 54: 14–22, 2013. doi: 10.1016/j.molimm.2012.10.027. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 344.Risøe PK, Rutkovskiy A, Ågren J, Kolseth IB, Kjeldsen SF, Valen G, Vaage J, Dahle MK. Higher TNFalpha responses in young males compared to females are associated with attenuation of monocyte adenylyl cyclase expression. Hum Immunol 76: 427–430, 2015. doi: 10.1016/j.humimm.2015.03.018. [DOI] [PubMed] [Google Scholar]
- 345.Jing X, Yang F, Shao C, Wei K, Xie M, Shen H, Shu Y. Role of hypoxia in cancer therapy by regulating the tumor microenvironment. Mol Cancer 18: 157, 2019. doi: 10.1186/s12943-019-1089-9. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 346.Nordsmark M, Bentzen SM, Rudat V, Brizel D, Lartigau E, Stadler P, Becker A, Adam M, Molls M, Dunst J, Terris DJ, Overgaard J. Prognostic value of tumor oxygenation in 397 head and neck tumors after primary radiation therapy. An international multi-center study. Radiother Oncol 77: 18–24, 2005. doi: 10.1016/j.radonc.2005.06.038. [DOI] [PubMed] [Google Scholar]
- 347.Pastorekova S, Zatovicova M, Pastorek J. Cancer-associated carbonic anhydrases and their inhibition. Curr Pharm Des 14: 685–698, 2008. doi: 10.2174/138161208783877893. [DOI] [PubMed] [Google Scholar]
- 348.Ditte P, Dequiedt F, Svastova E, Hulikova A, Ohradanova-Repic A, Zatovicova M, Csaderova L, Kopacek J, Supuran CT, Pastorekova S, Pastorek J. Phosphorylation of carbonic anhydrase IX controls its ability to mediate extracellular acidification in hypoxic tumors. Cancer Res 71: 7558–7567, 2011. doi: 10.1158/0008-5472.CAN-11-2520. [DOI] [PubMed] [Google Scholar]
- 349.Mann L, Heldman E, Bersudsky Y, Vatner SF, Ishikawa Y, Almog O, Belmaker RH, Agam G. Inhibition of specific adenylyl cyclase isoforms by lithium and carbamazepine, but not valproate, may be related to their antidepressant effect. Bipolar Disord 11: 885–896, 2009. doi: 10.1111/j.1399-5618.2009.00762.x. [DOI] [PubMed] [Google Scholar]
- 350.Seifert R. Vidarabine is neither a potent nor a selective AC5 inhibitor. Biochem Pharmacol 87: 543–546, 2014. doi: 10.1016/j.bcp.2013.12.025. [DOI] [PubMed] [Google Scholar]
- 351.Seifert R. Does vidarabine mediate cardioprotection via inhibition of AC5? J Pharmacol Exp Ther 358: 242–243, 2016. doi: 10.1124/jpet.116.234245. [DOI] [PubMed] [Google Scholar]
- 352.Bravo CA, Vatner DE, Pachon R, Zhang J, Vatner SF. A Food and Drug Administration-approved antiviral agent that inhibits adenylyl cyclase type 5 protects the ischemic heart even when administered after reperfusion. J Pharmacol Exp Ther 357: 331–336, 2016. doi: 10.1124/jpet.116.232538. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 353.Zhang J, Levy D, Oydanich M, Bravo CA, Yoon S, Vatner DE, Vatner SF. A novel adenylyl cyclase type 5 inhibitor that reduces myocardial infarct size even when administered after coronary artery reperfusion. J Mol Cell Cardiol 121: 13–15, 2018. doi: 10.1016/j.yjmcc.2018.05.014. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 354.Iwamoto T, Okumura S, Iwatsubo K, Kawabe J, Ohtsu K, Sakai I, Hashimoto Y, Izumitani A, Sango K, Ajiki K, Toya Y, Umemura S, Goshima Y, Arai N, Vatner SF, Ishikawa Y. Motor dysfunction in type 5 adenylyl cyclase-null mice. J Biol Chem 278: 16936–16940, 2003. doi: 10.1074/jbc.C300075200. [DOI] [PubMed] [Google Scholar]
- 355.Kim H, Lee Y, Park JY, Kim JE, Kim TK, Choi J, Lee JE, Lee EH, Kim D, Kim KS, Han PL. Loss of adenylyl cyclase type-5 in the dorsal striatum produces autistic-like behaviors. Mol Neurobiol 54: 7994–8008, 2017. doi: 10.1007/s12035-016-0256-x. [DOI] [PubMed] [Google Scholar]
- 356.Heinick A, Pluteanu F, Hermes C, Klemme A, Domnik M, Husser X, Gerke V, Schmitz W, Müller FU. Annexin A4 N-terminal peptide inhibits adenylyl cyclase 5 and limits beta-adrenoceptor-mediated prolongation of cardiac action potential. FASEB J 34: 10489–10504, 2020. doi: 10.1096/fj.201902094RR. [DOI] [PubMed] [Google Scholar]
- 357.Doyle TB, Hayes MP, Chen DH, Raskind WH, Watts VJ. Functional characterization of AC5 gain-of-function variants: impact on the molecular basis of ADCY5-related dyskinesia. Biochem Pharmacol 163: 169–177, 2019. doi: 10.1016/j.bcp.2019.02.005. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 358.Tang T, Lai NC, Roth DM, Drumm J, Guo T, Lee KW, Han PL, Dalton N, Gao MH. Adenylyl cyclase type V deletion increases basal left ventricular function and reduces left ventricular contractile responsiveness to beta-adrenergic stimulation. Basic Res Cardiol 101: 117–126, 2006. doi: 10.1007/s00395-005-0559-y. [DOI] [PubMed] [Google Scholar]
- 359.Okumura S, Takagi G, Kawabe J, Yang G, Lee MC, Hong C, Liu J, Vatner DE, Sadoshima J, Vatner SF, Ishikawa Y. Disruption of type 5 adenylyl cyclase gene preserves cardiac function against pressure overload. Proc Natl Acad Sci USA 100: 9986–9990, 2003. doi: 10.1073/pnas.1733772100. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 360.Okumura S, Vatner DE, Kurotani R, Bai Y, Gao S, Yuan Z, Iwatsubo K, Ulucan C, Kawabe J, Ghosh K, Vatner SF, Ishikawa Y. Disruption of type 5 adenylyl cyclase enhances desensitization of cyclic adenosine monophosphate signal and increases Akt signal with chronic catecholamine stress. Circulation 116: 1776–1783, 2007. doi: 10.1161/CIRCULATIONAHA.107.698662. [DOI] [PubMed] [Google Scholar]
- 361.Yan L, Vatner DE, O’Connor JP, Ivessa A, Ge H, Chen W, Hirotani S, Ishikawa Y, Sadoshima J, Vatner SF. Type 5 adenylyl cyclase disruption increases longevity and protects against stress. Cell 130: 247–258, 2007. doi: 10.1016/j.cell.2007.05.038. [DOI] [PubMed] [Google Scholar]
- 362.Yan L, Vatner SF, Vatner DE. Disruption of type 5 adenylyl cyclase prevents beta-adrenergic receptor cardiomyopathy: a novel approach to beta-adrenergic receptor blockade. Am J Physiol Heart Circ Physiol 307: H1521–H1528, 2014. doi: 10.1152/ajpheart.00491.2014. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 363.Timofeyev V, Porter CA, Tuteja D, Qiu H, Li N, Tang T, Singapuri A, Han PL, Lopez JE, Hammond HK, Chiamvimonvat N. Disruption of adenylyl cyclase type V does not rescue the phenotype of cardiac-specific overexpression of Galphaq protein-induced cardiomyopathy. Am J Physiol Heart Circ Physiol 299: H1459–H1467, 2010. doi: 10.1152/ajpheart.01208.2009. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 364.Okumura S, Kawabe J, Yatani A, Takagi G, Lee MC, Hong C, Liu J, Takagi I, Sadoshima J, Vatner DE, Vatner SF, Ishikawa Y. Type 5 adenylyl cyclase disruption alters not only sympathetic but also parasympathetic and calcium-mediated cardiac regulation. Circ Res 93: 364–371, 2003. doi: 10.1161/01.RES.0000086986.35568.63. [DOI] [PubMed] [Google Scholar]
- 365.Ho D, Zhao X, Yan L, Yuan C, Zong H, Vatner DE, Pessin JE, Vatner SF. Adenylyl cyclase type 5 deficiency protects against diet-induced obesity and insulin resistance. Diabetes 64: 2636–2645, 2015. doi: 10.2337/db14-0494. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 366.Vatner SF, Pachon RE, Vatner DE. Inhibition of adenylyl cyclase type 5 increases longevity and healthful aging through oxidative stress protection. Oxid Med Cell Longev 2015: 250310, 201 5. doi: 10.1155/2015/250310. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 367.Vatner SF, Park M, Yan L, Lee GJ, Lai L, Iwatsubo K, Ishikawa Y, Pessin J, Vatner DE. Adenylyl cyclase type 5 in cardiac disease, metabolism, and aging. Am J Physiol Heart Circ Physiol 305: H1–H8, 2013. doi: 10.1152/ajpheart.00080.2013. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 368.Syed AU, Reddy GR, Ghosh D, Prada MP, Nystoriak MA, Morotti S, Grandi E, Sirish P, Chiamvimonvat N, Hell JW, Santana LF, Xiang YK, Nieves-Cintrón M, Navedo MF. Adenylyl cyclase 5-generated cAMP controls cerebral vascular reactivity during diabetic hyperglycemia. J Clin Invest 129: 3140–3152, 2019. doi: 10.1172/JCI124705. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 369.Kim KS, Han PL. Mice lacking adenylyl cyclase-5 cope badly with repeated restraint stress. J Neurosci Res 87: 2983–2993, 2009. doi: 10.1002/jnr.22119. [DOI] [PubMed] [Google Scholar]
- 370.Kheirbek MA, Britt JP, Beeler JA, Ishikawa Y, McGehee DS, Zhuang X. Adenylyl cyclase type 5 contributes to corticostriatal plasticity and striatum-dependent learning. J Neurosci 29: 12115–12124, 2009. doi: 10.1523/JNEUROSCI.3343-09.2009. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 371.Kim KS, Kim H, Baek IS, Lee KW, Han PL. Mice lacking adenylyl cyclase type 5 (AC5) show increased ethanol consumption and reduced ethanol sensitivity. Psychopharmacology (Berl) 215: 391–398, 2011. doi: 10.1007/s00213-010-2143-x. [DOI] [PubMed] [Google Scholar]
- 372.Kim KS, Kim H, Park SK, Han PL. The dorsal striatum expressing adenylyl cyclase-5 controls behavioral sensitivity of the righting reflex to high-dose ethanol. Brain Res 1489: 27–36, 2012. doi: 10.1016/j.brainres.2012.10.016. [DOI] [PubMed] [Google Scholar]
- 373.Howes OD, Rogdaki M, Findon JL, Wichers RH, Charman T, King BH, Loth E, McAlonan GM, McCracken JT, Parr JR, Povey C, Santosh P, Wallace S, Simonoff E, Murphy DG. Autism spectrum disorder: consensus guidelines on assessment, treatment and research from the British Association for Psychopharmacology. J Psychopharmacol 32: 3–29, 2018. doi: 10.1177/0269881117741766. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 374.Kim KS, Kim J, Back SK, Im JY, Na HS, Han PL. Markedly attenuated acute and chronic pain responses in mice lacking adenylyl cyclase-5. Genes Brain Behav 6: 120–127, 2007. doi: 10.1111/j.1601-183X.2006.00238.x. [DOI] [PubMed] [Google Scholar]
- 375.Holstein JD, Patzer O, Körner A, Stumvoll M, Kovacs P, Holstein A. Genetic variants in GCKR, GIPR, ADCY5 and VPS13C and the risk of severe sulfonylurea-induced hypoglycaemia in patients with type 2 diabetes. Exp Clin Endocrinol Diabetes 121: 54–57, 2013. doi: 10.1055/s-0032-1321834. [DOI] [PubMed] [Google Scholar]
- 376.Hodson DJ, Mitchell RK, Marselli L, Pullen TJ, Gimeno Brias S, Semplici F, Everett KL, Cooper DM, Bugliani M, Marchetti P, Lavallard V, Bosco D, Piemonti L, Johnson PR, Hughes SJ, Li D, Li WH, Shapiro AM, Rutter GA. ADCY5 couples glucose to insulin secretion in human islets. Diabetes 63: 3009–3021, 2014. doi: 10.2337/db13-1607. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 377.van Keulen SC, Narzi D, Rothlisberger U. Association of both inhibitory and stimulatory Galpha subunits implies adenylyl cyclase 5 deactivation. Biochemistry 58: 4317–4324, 2019. doi: 10.1021/acs.biochem.9b00662. [DOI] [PubMed] [Google Scholar]
- 378.Frezza E, Amans TM, Martin J. Allosteric inhibition of adenylyl cyclase type 5 by G-protein: a molecular dynamics study. Biomolecules 10: 1330, 2020. doi: 10.3390/biom10091330. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 379.Narzi D, van Keulen SC, Röthlisberger U. Galphai1 inhibition mechanism of ATP-bound adenylyl cyclase type 5. PLoS One 16: e0245197, 2021. doi: 10.1371/journal.pone.0245197. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 380.Bruce NJ, Narzi D, Trpevski D, van Keulen SC, Nair AG, Röthlisberger U, Wade RC, Carloni P, Hellgren Kotaleski J. Regulation of adenylyl cyclase 5 in striatal neurons confers the ability to detect coincident neuromodulatory signals. PLoS Comput Biol 15: e1007382, 2019. doi: 10.1371/journal.pcbi.1007382. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 381.Navarro G, Cordomi A, Casado-Anguera V, Moreno E, Cai NS, Cortes A, Canela EI, Dessauer CW, Casado V, Pardo L, Lluis C, Ferre S. Evidence for functional pre-coupled complexes of receptor heteromers and adenylyl cyclase. Nat Commun 9: 1242, 2018. doi: 10.1038/s41467-018-03522-3. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 382.Seifert R, Schirmer B. A simple mechanistic terminology of psychoactive drugs: a proposal. Naunyn-Schmiedebergs Arch Pharmacol 393: 1331–1339, 2020. doi: 10.1007/s00210-020-01918-x. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 383.Ferré S, Bonaventura J, Zhu W, Hatcher-Solis C, Taura J, Quiroz C, Cai NS, Moreno E, Casadó-Anguera V, Kravitz AV, Thompson KR, Tomasi DG, Navarro G, Cordomí A, Pardo L, Lluís C, Dessauer CW, Volkow ND, Casadó V, Ciruela F, Logothetis DE, Zwilling D. Essential control of the function of the striatopallidal neuron by pre-coupled complexes of adenosine A2A-dopamine D2 receptor heterotetramers and adenylyl cyclase. Front Pharmacol 9: 243, 2018. doi: 10.3389/fphar.2018.00243. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 384.Wang Q, Cobo-Stark P, Patel V, Somlo S, Han PL, Igarashi P. Adenylyl cyclase 5 deficiency reduces renal cyclic AMP and cyst growth in an orthologous mouse model of polycystic kidney disease. Kidney Int 93: 403–415, 2018. doi: 10.1016/j.kint.2017.08.005. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 385.Seifert R, Lushington GH, Mou TC, Gille A, Sprang SR. Inhibitors of membranous adenylyl cyclases. Trends Pharmacol Sci 33: 64–78, 2012. doi: 10.1016/j.tips.2011.10.006. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 386.Tang T, Gao MH, Lai NC, Firth AL, Takahashi T, Guo T, Yuan JX, Roth DM, Hammond HK. Adenylyl cyclase type 6 deletion decreases left ventricular function via impaired calcium handling. Circulation 117: 61–69, 2008. doi: 10.1161/CIRCULATIONAHA.107.730069. [DOI] [PubMed] [Google Scholar]
- 387.Gao M, Ping P, Post S, Insel PA, Tang R, Hammond HK. Increased expression of adenylylcyclase type VI proportionately increases beta-adrenergic receptor-stimulated production of cAMP in neonatal rat cardiac myocytes. Proc Natl Acad Sci USA 95: 1038–1043, 1998. doi: 10.1073/pnas.95.3.1038. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 388.Tepe NM, Lorenz JN, Yatani A, Dash R, Kranias EG, Dorn GW 2nd, Liggett SB. Altering the receptor-effector ratio by transgenic overexpression of type V adenylyl cyclase: enhanced basal catalytic activity and function without increased cardiomyocyte beta-adrenergic signalling. Biochemistry 38: 16706–16713, 1999. doi: 10.1021/bi991619k. [DOI] [PubMed] [Google Scholar]
- 389.Roth DM, Gao MH, Lai NC, Drumm J, Dalton N, Zhou JY, Zhu J, Entrikin D, Hammond HK. Cardiac-directed adenylyl cyclase expression improves heart function in murine cardiomyopathy. Circulation 99: 3099–3102, 1999. doi: 10.1161/01.CIR.99.24.3099. [DOI] [PubMed] [Google Scholar]
- 390.Roth DM, Bayat H, Drumm JD, Gao MH, Swaney JS, Ander A, Hammond HK. Adenylyl cyclase increases survival in cardiomyopathy. Circulation 105: 1989–1994, 2002. doi: 10.1161/01.CIR.0000014968.54967.D3. [DOI] [PubMed] [Google Scholar]
- 391.Lai NC, Roth DM, Gao MH, Fine S, Head BP, Zhu J, McKirnan MD, Kwong C, Dalton N, Urasawa K, Roth DA, Hammond HK. Intracoronary delivery of adenovirus encoding adenylyl cyclase VI increases left ventricular function and cAMP-generating capacity. Circulation 102: 2396–2401, 2000. doi: 10.1161/01.CIR.102.19.2396. [DOI] [PubMed] [Google Scholar]
- 392.Lai NC, Roth DM, Gao MH, Tang T, Dalton N, Lai YY, Spellman M, Clopton P, Hammond HK. Intracoronary adenovirus encoding adenylyl cyclase VI increases left ventricular function in heart failure. Circulation 110: 330–336, 2004. doi: 10.1161/01.CIR.0000136033.21777.4D. [DOI] [PubMed] [Google Scholar]
- 393.Bull Melsom C, Cosson MV, Ørstavik Ø, Lai NC, Hammond HK, Osnes JB, Skomedal T, Nikolaev V, Levy FO, Krobert KA. Constitutive inhibitory G protein activity upon adenylyl cyclase-dependent cardiac contractility is limited to adenylyl cyclase type 6. PLoS One 14: e0218110, 2019. doi: 10.1371/journal.pone.0218110. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 394.Gao MH, Lai NC, Tang T, Guo T, Tang R, Chun BJ, Wang H, Dalton NN, Suarez J, Dillmann WH, Hammond HK. Preserved cardiac function despite marked impairment of cAMP generation. PLoS One 8: e72151, 2013. doi: 10.1371/journal.pone.0072151. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 395.Gao MH, Miyanohara A, Feramisco JR, Tang T. Activation of PH-domain leucine-rich protein phosphatase 2 (PHLPP2) by agonist stimulation in cardiac myocytes expressing adenylyl cyclase type 6. Biochem Biophys Res Commun 384: 193–198, 2009. doi: 10.1016/j.bbrc.2009.04.110. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 396.Hammond HK, Penny WF, Traverse JH, Henry TD, Watkins MW, Yancy CW, Sweis RN, Adler ED, Patel AN, Murray DR, Ross RS, Bhargava V, Maisel A, Barnard DD, Lai NC, Dalton ND, Lee ML, Narayan SM, Blanchard DG, Gao MH. Intracoronary gene transfer of adenylyl cyclase 6 in patients with heart failure: a randomized clinical trial. JAMA Cardiol 1: 163–171, 2016. doi: 10.1001/jamacardio.2016.0008. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 397.Guellich A, Gao S, Hong C, Yan L, Wagner TE, Dhar SK, Ghaleh B, Hittinger L, Iwatsubo K, Ishikawa Y, Vatner SF, Vatner DE. Effects of cardiac overexpression of type 6 adenylyl cyclase affects on the response to chronic pressure overload. Am J Physiol Heart Circ Physiol 299: H707–H712, 2010. doi: 10.1152/ajpheart.00148.2010. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 398.Tang T, Lai NC, Hammond HK, Roth DM, Yang Y, Guo T, Gao MH. Adenylyl cyclase 6 deletion reduces left ventricular hypertrophy, dilation, dysfunction, and fibrosis in pressure-overloaded female mice. J Am Coll Cardiol 55: 1476–1486, 2010. doi: 10.1016/j.jacc.2009.11.066. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 399.Tan Z, Giamouridis D, Lai NC, Kim YC, Guo T, Xia B, Gao MH, Hammond HK. Cardiac-directed expression of adenylyl cyclase catalytic domain (C1C2) attenuates deleterious effects of pressure overload. Hum Gene Ther 30: 682–692, 2019. doi: 10.1089/hum.2018.176. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 400.Gao MH, Lai NC, Giamouridis D, Kim YC, Tan Z, Guo T, Dillmann WH, Suarez J, Hammond HK. Cardiac-directed expression of adenylyl cyclase catalytic domain reverses cardiac dysfunction caused by sustained beta-adrenergic receptor stimulation. JACC Basic Transl Sci 1: 617–629, 2016. doi: 10.1016/j.jacbts.2016.08.004. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 401.Gao MH, Lai NC, Giamouridis D, Kim YC, Guo T, Hammond HK. Cardiac-directed expression of a catalytically inactive adenylyl cyclase 6 protects the heart from sustained beta-adrenergic stimulation. PLoS One 12: e0181282, 2017. doi: 10.1371/journal.pone.0181282. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 402.Birrell MA, Bonvini SJ, Wortley MA, Buckley J, Yew-Booth L, Maher SA, Dale N, Dubuis ED, Belvisi MG. The role of adenylyl cyclase isoform 6 in beta-adrenoceptor signalling in murine airways. Br J Pharmacol 172: 131–141, 2015. doi: 10.1111/bph.12905. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 403.Chang CP, Lee CT, Hou WH, Lin MS, Lai HL, Chien CL, Chang C, Cheng PL, Lien CC, Chern Y. Type VI adenylyl cyclase negatively regulates GluN2B-mediated LTD and spatial reversal learning. Sci Rep 6: 22529, 2016. doi: 10.1038/srep22529. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 404.Lee KL, Hoey DA, Spasic M, Tang T, Hammond HK, Jacobs CR. Adenylyl cyclase 6 mediates loading-induced bone adaptation in vivo. FASEB J 28: 1157–1165, 2014. doi: 10.1096/fj.13-240432. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 405.Sabbatini ME, D’Alecy L, Lentz SI, Tang T, Williams JA. Adenylyl cyclase 6 mediates the action of cyclic AMP-dependent secretagogues in mouse pancreatic exocrine cells via protein kinase A pathway activation. J Physiol 591: 3693–3707, 2013. doi: 10.1113/jphysiol.2012.249698. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 406.Eyler CE, Jackson T, Elliott LE, De Castro LM, Jonassaint J, Ashley-Koch A, Telen MJ. beta(2)-Adrenergic receptor and adenylate cyclase gene polymorphisms affect sickle red cell adhesion. Br J Haematol 141: 105–108, 2008. doi: 10.1111/j.1365-2141.2008.07008.x. [DOI] [PubMed] [Google Scholar]
- 407.Chien CL, Wu YS, Lai HL, Chen YH, Jiang ST, Shih CM, Lin SS, Chang C, Chern Y. Impaired water reabsorption in mice deficient in the type VI adenylyl cyclase (AC6). FEBS Lett 584: 2883–2890, 2010. doi: 10.1016/j.febslet.2010.05.004. [DOI] [PubMed] [Google Scholar]
- 408.Aldehni F, Tang T, Madsen K, Plattner M, Schreiber A, Friis UG, Hammond HK, Han PL, Schweda F. Stimulation of renin secretion by catecholamines is dependent on adenylyl cyclases 5 and 6. Hypertension 57: 460–468, 2011. doi: 10.1161/HYPERTENSIONAHA.110.167130. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 409.Rieg T, Tang T, Murray F, Schroth J, Insel PA, Fenton RA, Hammond HK, Vallon V. Adenylate cyclase 6 determines cAMP formation and aquaporin-2 phosphorylation and trafficking in inner medulla. J Am Soc Nephrol 21: 2059–2068, 2010. doi: 10.1681/ASN.2010040409. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 410.Rieg T, Tang T, Uchida S, Hammond HK, Fenton RA, Vallon V. Adenylyl cyclase 6 enhances NKCC2 expression and mediates vasopressin-induced phosphorylation of NKCC2 and NCC. Am J Pathol 182: 96–106, 2013. doi: 10.1016/j.ajpath.2012.09.014. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 411.Roos KP, Strait KA, Raphael KL, Blount MA, Kohan DE. Collecting duct-specific knockout of adenylyl cyclase type VI causes a urinary concentration defect in mice. Am J Physiol Renal Physiol 302: F78–F84, 2012. doi: 10.1152/ajprenal.00397.2011. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 412.Roos KP, Bugaj V, Mironova E, Stockand JD, Ramkumar N, Rees S, Kohan DE. Adenylyl cyclase VI mediates vasopressin-stimulated ENaC activity. J Am Soc Nephrol 24: 218–227, 2013. doi: 10.1681/ASN.2012050449. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 413.Rees S, Kittikulsuth W, Roos K, Strait KA, Van Hoek A, Kohan DE. Adenylyl cyclase 6 deficiency ameliorates polycystic kidney disease. J Am Soc Nephrol 25: 232–237, 2014. doi: 10.1681/ASN.2013010077. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 414.Choi YH, Suzuki A, Hajarnis S, Ma Z, Chapin HC, Caplan MJ, Pontoglio M, Somlo S, Igarashi P. Polycystin-2 and phosphodiesterase 4C are components of a ciliary A-kinase anchoring protein complex that is disrupted in cystic kidney diseases. Proc Natl Acad Sci USA 108: 10679–10684, 2011. doi: 10.1073/pnas.1016214108. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 415.Poulsen SB, Marin De Evsikova C, Murali SK, Praetorius J, Chern Y, Fenton RA, Rieg T. Adenylyl cyclase 6 is required for maintaining acid-base homeostasis. Clin Sci (Lond) 132: 1779–1796, 2018. doi: 10.1042/CS20180060. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 416.Tarnawski L, Reardon C, Caravaca AS, Rosas-Ballina M, Tusche MW, Drake AR, Hudson LK, Hanes WM, Li JH, Parrish WR, Ojamaa K, Al-Abed Y, Faltys M, Pavlov VA, Andersson U, Chavan SS, Levine YA, Mak TW, Tracey KJ, Olofsson PS. Adenylyl cyclase 6 mediates inhibition of TNF in the inflammatory reflex. Front Immunol 9: 2648, 2018. doi: 10.3389/fimmu.2018.02648. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 417.Zhu S, Huang S, Xia G, Wu J, Shen Y, Wang Y, Ostrom RS, Du A, Shen C, Xu C. Anti-inflammatory effects of alpha 7 nicotinic acetylcholine receptors through interacting with adenylyl cyclase 6. Br J Pharmacol 178: 2324–2338, 2021. doi: 10.1111/bph.15412. [DOI] [PubMed] [Google Scholar]
- 418.Arora K, Lund JR, Naren NA, Zingarelli B, Naren AP. AC6 regulates the microtubule-depolymerizing kinesin KIF19A to control ciliary length in mammals. J Biol Chem 295: 14250–14259, 2020. doi: 10.1074/jbc.RA120.013703. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 419.Sherpa RT, Mohieldin AM, Pala R, Wachten D, Ostrom RS, Nauli SM. Author correction: sensory primary cilium is a responsive cAMP microdomain in renal epithelia. Sci Rep 10: 1731, 2020. doi: 10.1038/s41598-020-58754-5. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 420.Thomas A, Ramananda Y, Mun K, Naren AP, Arora K. AC6 is the major adenylate cyclase forming a diarrheagenic protein complex with cystic fibrosis transmembrane conductance regulator in cholera. J Biol Chem 293: 12949–12959, 2018. doi: 10.1074/jbc.RA118.003378. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 421.Fenton RA, Murali SK, Kaji I, Akiba Y, Kaunitz JD, Kristensen TB, Poulsen SB, Dominguez Rieg JA, Rieg T. Adenylyl cyclase 6 expression is essential for cholera toxin-induced diarrhea. J Infect Dis 220: 1719–1728, 2019. doi: 10.1093/infdis/jiz013. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 422.Choi LJ, Jenikova G, Hanson E, Spehlmann ME, Boehling NS, Kirstein SL, Bundey RA, Smith JR, Insel PA, Eckmann L. Coordinate down-regulation of adenylyl cyclase isoforms and the stimulatory G protein (Gs) in intestinal epithelial cell differentiation. J Biol Chem 285: 12504–12511, 2010. doi: 10.1074/jbc.M109.059741. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 423.Cosson MV, Hiis HG, Moltzau LR, Levy FO, Krobert KA. Knockout of adenylyl cyclase isoform 5 or 6 differentially modifies the beta1-adrenoceptor-mediated inotropic response. J Mol Cell Cardiol 131: 132–145, 2019. doi: 10.1016/j.yjmcc.2019.04.017. [DOI] [PubMed] [Google Scholar]
- 424.Timofeyev V, Myers RE, Kim HJ, Woltz RL, Sirish P, Heiserman JP, Li N, Singapuri A, Tang T, Yarov-Yarovoy V, Yamoah EN, Hammond HK, Chiamvimonvat N. Adenylyl cyclase subtype-specific compartmentalization: differential regulation of L-type Ca2+ current in ventricular myocytes. Circ Res 112: 1567–1576, 2013. doi: 10.1161/CIRCRESAHA.112.300370. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 425.Baldwin TA, Dessauer CW. Function of adenylyl cyclase in heart: the AKAP connection. J Cardiovasc Dev Dis 5: 2, 2018. doi: 10.3390/jcdd5010002. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 426.Head BP, Patel HH, Roth DM, Lai NC, Niesman IR, Farquhar MG, Insel PA. G-protein-coupled receptor signaling components localize in both sarcolemmal and intracellular caveolin-3-associated microdomains in adult cardiac myocytes. J Biol Chem 280: 31036–31044, 2005. doi: 10.1074/jbc.M502540200. [DOI] [PubMed] [Google Scholar]
- 427.Mou TC, Masada N, Cooper DM, Sprang SR. Structural basis for inhibition of mammalian adenylyl cyclase by calcium. Biochemistry 48: 3387–3397, 2009. doi: 10.1021/bi802122k. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 428.Moore TM, Chetham PM, Kelly JJ, Stevens T. Signal transduction and regulation of lung endothelial cell permeability. Interaction between calcium and cAMP. Am J Physiol Lung Cell Mol Physiol 275: L203–L222, 1998. doi: 10.1152/ajplung.1998.275.2.L203. [DOI] [PubMed] [Google Scholar]
- 429.Cioffi DL, Moore TM, Schaack J, Creighton JR, Cooper DM, Stevens T. Dominant regulation of interendothelial cell gap formation by calcium-inhibited type 6 adenylyl cyclase. J Cell Biol 157: 1267–1278, 2002. doi: 10.1083/jcb.200204022. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 430.Stevens T, Nakahashi Y, Cornfield DN, McMurtry IF, Cooper DM, Rodman DM. Ca2+-inhibitable adenylyl cyclase modulates pulmonary artery endothelial cell cAMP content and barrier function. Proc Natl Acad Sci USA 92: 2696–2700, 1995. doi: 10.1073/pnas.92.7.2696. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 431.Werthmann RC, von Hayn K, Nikolaev VO, Lohse MJ, Bünemann M. Real-time monitoring of cAMP levels in living endothelial cells: thrombin transiently inhibits adenylyl cyclase 6. J Physiol 587: 4091–4104, 2009. doi: 10.1113/jphysiol.2009.172957. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 432.von Hayn K, Werthmann RC, Nikolaev VO, Hommers LG, Lohse MJ, Bünemann M. Gq-mediated Ca2+ signals inhibit adenylyl cyclases 5/6 in vascular smooth muscle cells. Am J Physiol Cell Physiol 298: C324–C332, 2010. doi: 10.1152/ajpcell.00197.2009. [DOI] [PubMed] [Google Scholar]
- 433.Johnson GP, Stavenschi E, Eichholz KF, Corrigan MA, Fair S, Hoey DA. Mesenchymal stem cell mechanotransduction is cAMP dependent and regulated by adenylyl cyclase 6 and the primary cilium. J Cell Sci 131: jcs222737, 2018. doi: 10.1242/jcs.222737. [DOI] [PubMed] [Google Scholar]
- 434.Romero-Calvo I, Ocon B, Gamez-Belmonte R, Hernandez-Chirlaque C, de Jonge HR, Bijvelds MJ, Martinez-Augustin O, Sanchez de Medina F. Adenylyl cyclase 6 is involved in the hyposecretory status of experimental colitis. Pflugers Arch 470: 1705–1717, 2018. doi: 10.1007/s00424-018-2187-z. [DOI] [PubMed] [Google Scholar]
- 435.Risøe PK, Ryg U, Wang YY, Rutkovskiy A, Smedsrød B, Valen G, Dahle MK. Cecal ligation and puncture sepsis is associated with attenuated expression of adenylyl cyclase 9 and increased miR142-3p. Shock 36: 390–395, 2011. doi: 10.1097/SHK.0b013e318228ec6f. [DOI] [PubMed] [Google Scholar]
- 436.Yang L, Wang YL, Liu S, Zhang PP, Chen Z, Liu M, Tang H. miR-181b promotes cell proliferation and reduces apoptosis by repressing the expression of adenylyl cyclase 9 (AC9) in cervical cancer cells. FEBS Lett 588: 124–130, 2014. doi: 10.1016/j.febslet.2013.11.019. [DOI] [PubMed] [Google Scholar]
- 437.Zhuang LK, Xu GP, Pan XR, Lou YJ, Zou QP, Xia D, Yan WW, Zhang YT, Jia PM, Tong JH. MicroRNA-181a-mediated downregulation of AC9 protein decreases intracellular cAMP level and inhibits ATRA-induced APL cell differentiation. Cell Death Dis 5: e1161, 2014. doi: 10.1038/cddis.2014.130. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 438.Cortes-Troncoso J, Jang SI, Perez P, Hidalgo J, Ikeuchi T, Greenwell-Wild T, Warner BM, Moutsopoulos NM, Alevizos I. T cell exosome-derived miR-142-3p impairs glandular cell function in Sjogren’s syndrome. JCI Insight 5: e133497, 2020. doi: 10.1172/jci.insight.133497. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 439.Liu L, Das S, Losert W, Parent CA. mTORC2 regulates neutrophil chemotaxis in a cAMP- and RhoA-dependent fashion. Dev Cell 19: 845–857, 2010. doi: 10.1016/j.devcel.2010.11.004. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 440.Mahadeo DC, Janka-Junttila M, Smoot RL, Roselova P, Parent CA. A chemoattractant-mediated Gi-coupled pathway activates adenylyl cyclase in human neutrophils. Mol Biol Cell 18: 512–522, 2007. doi: 10.1091/mbc.e06-05-0418. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 441.Brunskole Hummel I, Reinartz MT, Kalble S, Burhenne H, Schwede F, Buschauer A, Seifert R. Dissociations in the effects of beta2-adrenergic receptor agonists on cAMP formation and superoxide production in human neutrophils: support for the concept of functional selectivity. PLoS One 8: e64556, 2013. doi: 10.1371/journal.pone.0064556. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 442.Werner K, Neumann D, Seifert R. Analysis of the histamine H2-receptor in human monocytes. Biochem Pharmacol 92: 369–379, 2014. doi: 10.1016/j.bcp.2014.08.028. [DOI] [PubMed] [Google Scholar]
- 443.Chao X, Jia Y, Feng X, Wang G, Wang X, Shi H, Zhao F, Jiang C. A case-control study of ADCY9 gene polymorphisms and the risk of hepatocellular carcinoma in the Chinese Han population. Front Oncol 10: 1450, 2020. doi: 10.3389/fonc.2020.01450. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 444.Toyota T, Hattori E, Meerabux J, Yamada K, Saito K, Shibuya H, Nankai M, Yoshikawa T. Molecular analysis, mutation screening, and association study of adenylate cyclase type 9 gene (ADCY9) in mood disorders. Am J Med Genet 114: 84–92, 2002. doi: 10.1002/ajmg.10117. [DOI] [PubMed] [Google Scholar]
- 445.Hicks KA, Mahaffey KW, Mehran R, Nissen SE, Wiviott SD, Dunn B, et al. 2017 Cardiovascular and stroke endpoint definitions for clinical trials. Circulation 137: 961–972, 2018. doi: 10.1161/CIRCULATIONAHA.117.033502. [DOI] [PubMed] [Google Scholar]
- 446.Hopewell JC, Ibrahim M, Hill M, Shaw PM, Braunwald E, Blaustein RO, Bowman L, Landray MJ, Sabatine MS, Collins R; HPS3/TIMI55–REVEAL Collaborative Group. Impact of ADCY9 genotype on response to anacetrapib. Circulation 140: 891–898, 2019. doi: 10.1161/CIRCULATIONAHA.119.041546. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 447.Tardif JC, Dube MP, Pfeffer MA, Waters DD, Koenig W, Maggioni AP, McMurray JJ, Mooser V, White HD, Heinonen T, Black DM, Guertin MC; Dal-GenE Investigators. Study design of Dal-GenE, a pharmacogenetic trial targeting reduction of cardiovascular events with dalcetrapib. Am Heart J 222: 157–165, 2020. doi: 10.1016/j.ahj.2020.01.007. [DOI] [PubMed] [Google Scholar]
- 448.Rautureau Y, Deschambault V, Higgins ME, Rivas D, Mecteau M, Geoffroy P, Miquel G, et al. ADCY9 (adenylate cyclase type 9) inactivation protects from atherosclerosis only in the absence of CETP (cholesteryl ester transfer protein). Circulation 138: 1677–1692, 2018. doi: 10.1161/CIRCULATIONAHA.117.031134. [DOI] [PubMed] [Google Scholar]
- 449.Li Y, Baldwin TA, Wang Y, Subramaniam J, Carbajal AG, Brand CS, Cunha SR, Dessauer CW. Loss of type 9 adenylyl cyclase triggers reduced phosphorylation of Hsp20 and diastolic dysfunction. Sci Rep 7: 5522, 2017. doi: 10.1038/s41598-017-05816-w. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 450.Li Y, Chen L, Kass RS, Dessauer CW. The A-kinase anchoring protein Yotiao facilitates complex formation between adenylyl cyclase type 9 and the IKs potassium channel in heart. J Biol Chem 287: 29815–29824, 2012. doi: 10.1074/jbc.M112.380568. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 451.Marx SO, Kurokawa J, Reiken S, Motoike H, D’Armiento J, Marks AR, Kass RS. Requirement of a macromolecular signaling complex for beta adrenergic receptor modulation of the KCNQ1-KCNE1 potassium channel. Science 295: 496–499, 2002. doi: 10.1126/science.1066843. [DOI] [PubMed] [Google Scholar]
- 452.Li Y, Hof T, Baldwin TA, Chen L, Kass RS, Dessauer CW. Regulation of IKs potassium current by isoproterenol in adult cardiomyocytes requires type 9 adenylyl cyclase. Cells 8: 981, 2019. doi: 10.3390/cells8090981. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 453.Zhang B, Zhou X, Qiu Y, Feng F, Feng J, Jia Y, Zhu H, Hu K, Liu J, Liu Z, Wang S, Gong Y, Zhou C, Zhu T, Cheng Y, Liu Z, Deng H, Tao F, Ren Y, Cheng B, Gao L, Wu X, Yu L, Huang Z, Mao Z, Song Q, Zhu B, Wang J. Clinical characteristics of 82 death cases with COVID-19 (Preprint). medRxiv 2020.2002.2026.20028191, 2020. doi: 10.1101/2020.02.26.20028191. [DOI]
- 454.Cumbay MG, Watts VJ. Galphaq potentiation of adenylate cyclase type 9 activity through a Ca2+/calmodulin-dependent pathway. Biochem Pharmacol 69: 1247–1256, 2005. doi: 10.1016/j.bcp.2005.02.001. [DOI] [PubMed] [Google Scholar]
- 455.Liu L, Gritz D, Parent CA. PKCbetaII acts downstream of chemoattractant receptors and mTORC2 to regulate cAMP production and myosin II activity in neutrophils. Mol Biol Cell 25: 1446–1457, 2014. doi: 10.1091/mbc.e14-01-0037. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 456.Antoni FA, Palkovits M, Simpson J, Smith SM, Leitch AL, Rosie R, Fink G, Paterson JM. Ca2+/calcineurin-inhibited adenylyl cyclase, highly abundant in forebrain regions, is important for learning and memory. J Neurosci 18: 9650–9661, 1998. doi: 10.1523/JNEUROSCI.18-23-09650.1998. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 457.Pálvölgyi A, Simpson J, Bodnár I, Biró J, Palkovits M, Radovits T, Skehel P, Antoni FA. Auto-inhibition of adenylyl cyclase 9 (AC9) by an isoform-specific motif in the carboxyl-terminal region. Cell Signal 51: 266–275, 2018. doi: 10.1016/j.cellsig.2018.08.010. [DOI] [PubMed] [Google Scholar]
- 458.Surinkaew S, Aflaki M, Takawale A, Chen Y, Qi XY, Gillis MA, Shi YF, Tardif JC, Chattipakorn N, Nattel S. Exchange protein activated by cyclic-adenosine monophosphate (Epac) regulates atrial fibroblast function and controls cardiac remodelling. Cardiovasc Res 115: 94–106, 2019. doi: 10.1093/cvr/cvy173. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 459.Antoni FA. The chilling of adenylyl cyclase 9 and its translational potential. Cell Signal 70: 109589, 2020. doi: 10.1016/j.cellsig.2020.109589. [DOI] [PubMed] [Google Scholar]
- 460.Lazar AM, Irannejad R, Baldwin TA, Sundaram AB, Gutkind JS, Inoue A, Dessauer CW, Von Zastrow M. G protein-regulated endocytic trafficking of adenylyl cyclase type 9. Elife 9: e58039, 2020. doi: 10.7554/eLife.58039. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 461.Villacres EC, Xia Z, Bookbinder LH, Edelhoff S, Disteche CM, Storm DR. Cloning, chromosomal mapping, and expression of human fetal brain type I adenylyl cyclase. Genomics 16: 473–478, 1993. doi: 10.1006/geno.1993.1213. [DOI] [PubMed] [Google Scholar]
- 462.Stengel D, Parma J, Gannagé MH, Roeckel N, Mattei MG, Barouki R, Hanoune J. Different chromosomal localization of two adenylyl cyclase genes expressed in human brain. Hum Genet 90: 126–130, 1992. [DOI] [PubMed] [Google Scholar]
- 463.Haber N, Stengel D, Defer N, Roeckel N, Mattei MG, Hanoune J. Chromosomal mapping of human adenylyl cyclase genes type III, type V and type VI. Hum Genet 94: 69–73, 1994. doi: 10.1007/BF02272844. [DOI] [PubMed] [Google Scholar]
- 464.Edelhoff S, Villacres EC, Storm DR, Disteche CM. Mapping of adenylyl cyclase genes type I, II, III, IV, V, and VI in mouse. Mamm Genome 6: 111–113, 1995. doi: 10.1007/BF00303253. [DOI] [PubMed] [Google Scholar]
- 465.Hellevuo K, Berry R, Sikela JM, Tabakoff B. Localization of the gene for a novel human adenylyl cyclase (ADCY7) to chromosome 16. Hum Genet 95: 197–200, 1995. doi: 10.1007/BF00209401. [DOI] [PubMed] [Google Scholar]
- 466.Premont RT, Matsuoka I, Mattei MG, Pouille Y, Defer N, Hanoune J. Identification and characterization of a widely expressed form of adenylyl cyclase. J Biol Chem 271: 13900–13907, 1996. doi: 10.1074/jbc.271.23.13900. [DOI] [PubMed] [Google Scholar]
- 467.Kuwahara J, Yonezawa A, Futamura M, Sugiura Y. Binding of transcription factor Sp1 to GC box DNA revealed by footprinting analysis: different contact of three zinc fingers and sequence recognition mode. Biochemistry 32: 5994–6001, 1993. doi: 10.1021/bi00074a010. [DOI] [PubMed] [Google Scholar]
- 468.Chao JR, Ni YG, Bolaños CA, Rahman Z, DiLeone RJ, Nestler EJ. Characterization of the mouse adenylyl cyclase type VIII gene promoter: regulation by cAMP and CREB. Eur J Neurosci 16: 1284–1294, 2002. doi: 10.1046/j.1460-9568.2002.02186.x. [DOI] [PubMed] [Google Scholar]
- 469.Abdel-Halim SM, Guenifi A, He B, Yang B, Mustafa M, Hojeberg B, Hillert J, Bakhiet M, Efendić S. Mutations in the promoter of adenylyl cyclase (AC)-III gene, overexpression of AC-III mRNA, and enhanced cAMP generation in islets from the spontaneously diabetic GK rat model of type 2 diabetes. Diabetes 47: 498–504, 1998. doi: 10.2337/diabetes.47.3.498. [DOI] [PubMed] [Google Scholar]
- 470.Lane-Ladd SB, Pineda J, Boundy VA, Pfeuffer T, Krupinski J, Aghajanian GK, Nestler EJ. CREB (cAMP response element-binding protein) in the locus coeruleus: biochemical, physiological, and behavioral evidence for a role in opiate dependence. J Neurosci 17: 7890–7901, 1997. doi: 10.1523/JNEUROSCI.17-20-07890.1997. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 471.Kong X, Li B, Deng Y, Ma X. FXR mediates adenylyl cyclase 8 expression in pancreatic beta-cells. J Diabetes Res 2019: 8915818, 2019. doi: 10.1155/2019/8915818. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 472.Chan GC, Lernmark U, Xia Z, Storm DR. DNA elements of the type 1 adenylyl cyclase gene locus enhance reporter gene expression in neurons and pinealocytes. Eur J Neurosci 13: 2054–2066, 2001. doi: 10.1046/j.0953-816x.2001.01578.x. [DOI] [PubMed] [Google Scholar]
- 473.Hong SH, Goh SH, Lee SJ, Hwang JA, Lee J, Choi IJ, Seo H, Park JH, Suzuki H, Yamamoto E, Kim IH, Jeong JS, Ju MH, Lee DH, Lee YS. Upregulation of adenylate cyclase 3 (ADCY3) increases the tumorigenic potential of cells by activating the CREB pathway. Oncotarget 4: 1791–1803, 2013. doi: 10.18632/oncotarget.1324. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 474.Katsushika S, Kawabe J, Homcy CJ, Ishikawa Y. In vivo generation of an adenylylcyclase isoform with a half-molecule motif. J Biol Chem 268: 2273–2276, 1993. doi: 10.1016/S0021-9258(18)53766-2. [DOI] [PubMed] [Google Scholar]
- 475.Vallin B, Legueux-Cajgfinger Y, Clément N, Glorian M, Duca L, Vincent P, Limon I, Blaise R. Novel short isoforms of adenylyl cyclase as negative regulators of cAMP production. Biochim Biophys Acta Mol Cell Res 1865: 1326–1340, 2018. doi: 10.1016/j.bbamcr.2018.06.012. [DOI] [PubMed] [Google Scholar]
- 476.Emala CW, Kumasaka D, Hirshman CA, Lindeman KS. Adenylyl cyclase messenger ribonucleic acid in myometrium: splice variant of type IV. Biol Reprod 59: 169–175, 1998. doi: 10.1095/biolreprod59.1.169. [DOI] [PubMed] [Google Scholar]
- 477.Cali JJ, Parekh RS, Krupinski J. Splice variants of type VIII adenylyl cyclase. Differences in glycosylation and regulation by Ca2+/calmodulin. J Biol Chem 271: 1089–1095, 1996. doi: 10.1074/jbc.271.2.1089. [DOI] [PubMed] [Google Scholar]
- 478.Steiner D, Avidor-Reiss T, Schallmach E, Butovsky E, Lev N, Vogel Z. Regulation of adenylate cyclase type VIII splice variants by acute and chronic Gi/o-coupled receptor activation. Biochem J 386: 341–348, 2005. doi: 10.1042/BJ20041670. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 479.Bers DM, Xiang YK, Zaccolo M. Whole-cell cAMP and PKA activity are epiphenomena, nanodomain signaling matters. Physiology (Bethesda) 34: 240–249, 2019. doi: 10.1152/physiol.00002.2019. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 480.Agarwal SR, Clancy CE, Harvey RD. Mechanisms restricting diffusion of intracellular cAMP. Sci Rep 6: 19577, 2016. doi: 10.1038/srep19577. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 481.Yang PC, Boras BW, Jeng MT, Docken SS, Lewis TJ, McCulloch AD, Harvey RD, Clancy CE. A computational modeling and simulation approach to investigate mechanisms of subcellular cAMP compartmentation. PLoS Comput Biol 12: e1005005, 2016. doi: 10.1371/journal.pcbi.1005005. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 482.Adams SR, Harootunian AT, Buechler YJ, Taylor SS, Tsien RY. Fluorescence ratio imaging of cyclic AMP in single cells. Nature 349: 694–697, 1991. doi: 10.1038/349694a0. [DOI] [PubMed] [Google Scholar]
- 483.Zaccolo M, De Giorgi F, Cho CY, Feng L, Knapp T, Negulescu PA, Taylor SS, Tsien RY, Pozzan T. A genetically encoded, fluorescent indicator for cyclic AMP in living cells. Nat Cell Biol 2: 25–29, 2000. doi: 10.1038/71345. [DOI] [PubMed] [Google Scholar]
- 484.Ponsioen B, Zhao J, Riedl J, Zwartkruis F, van der Krogt G, Zaccolo M, Moolenaar WH, Bos JL, Jalink K. Detecting cAMP-induced Epac activation by fluorescence resonance energy transfer: Epac as a novel cAMP indicator. EMBO Rep 5: 1176–1180, 2004. doi: 10.1038/sj.embor.7400290. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 485.Rich TC, Fagan KA, Nakata H, Schaack J, Cooper DM, Karpen JW. Cyclic nucleotide-gated channels colocalize with adenylyl cyclase in regions of restricted cAMP diffusion. J Gen Physiol 116: 147–161, 2000. doi: 10.1085/jgp.116.2.147. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 486.Brady JD, Rich TC, Le X, Stafford K, Fowler CJ, Lynch L, Karpen JW, Brown RL, Martens JR. Functional role of lipid raft microdomains in cyclic nucleotide-gated channel activation. Mol Pharmacol 65: 503–511, 2004. doi: 10.1124/mol.65.3.503. [DOI] [PubMed] [Google Scholar]
- 487.Tewson PH, Martinka S, Shaner NC, Hughes TE, Quinn AM. New DAG and cAMP sensors optimized for live-cell assays in automated laboratories. J Biomol Screen 21: 298–305, 2016. doi: 10.1177/1087057115618608. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 488.Stangherlin A, Zaccolo M. Phosphodiesterases and subcellular compartmentalized cAMP signaling in the cardiovascular system. Am J Physiol Heart Circ Physiol 302: H379–H390, 2012. doi: 10.1152/ajpheart.00766.2011. [DOI] [PubMed] [Google Scholar]
- 489.Terrin A, Monterisi S, Stangherlin A, Zoccarato A, Koschinski A, Surdo NC, Mongillo M, Sawa A, Jordanides NE, Mountford JC, Zaccolo M. PKA and PDE4D3 anchoring to AKAP9 provides distinct regulation of cAMP signals at the centrosome. J Cell Biol 198: 607–621, 2012. doi: 10.1083/jcb.201201059. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 490.Moore BS, Stepanchick AN, Tewson PH, Hartle CM, Zhang J, Quinn AM, Hughes TE, Mirshahi T. Cilia have high cAMP levels that are inhibited by Sonic Hedgehog-regulated calcium dynamics. Proc Natl Acad Sci USA 113: 13069–13074, 2016. doi: 10.1073/pnas.1602393113. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 491.Agarwal SR, Gratwohl J, Cozad M, Yang PC, Clancy CE, Harvey RD. Compartmentalized cAMP signaling associated with lipid raft and non-raft membrane domains in adult ventricular myocytes. Front Pharmacol 9: 332, 2018. doi: 10.3389/fphar.2018.00332. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 492.Surdo NC, Berrera M, Koschinski A, Brescia M, Machado MR, Carr C, Wright P, Gorelik J, Morotti S, Grandi E, Bers DM, Pantano S, Zaccolo M. FRET biosensor uncovers cAMP nano-domains at beta-adrenergic targets that dictate precise tuning of cardiac contractility. Nat Commun 8: 15031, 2017. doi: 10.1038/ncomms15031. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 493.Musheshe N, Schmidt M, Zaccolo M. cAMP: from long-range second messenger to nanodomain signalling. Trends Pharmacol Sci 39: 209–222, 2018. doi: 10.1016/j.tips.2017.11.006. [DOI] [PubMed] [Google Scholar]






