Biochem. J. (2014) 462, 199–213 (Printed in Great Britain)

199

doi:10.1042/BJ20140560

REVIEW ARTICLE Dermot M. F. COOPER*1 and Valentina G. TABBASUM* *Department of Pharmacology, University of Cambridge, Cambridge CB2 1PD, U.K.

Recent advances in the AC (adenylate cyclase)/cAMP field reveal overarching roles for the ACs. Whereas few processes are unaffected by cAMP in eukaryotes, ranging from the rapid modulation of ion channel kinetics to the slowest developmental effects, the large number of cellular processes modulated by only three intermediaries, i.e. PKA (protein kinase A), Epacs (exchange proteins directly activated by cAMP) and CNG (cyclic nucleotide-gated) channels, poses the question of how selectivity and fine control is achieved by cAMP. One answer rests on the number of differently regulated and distinctly expressed AC species. Specific ACs are implicated in processes such as insulin secretion, immunological responses, sino-atrial node pulsatility and memory formation, and specific ACs are linked with particular diseased conditions or predispositions, such as cystic fibrosis, Type 2 diabetes and dysrhythmias. However, much of the selectivity and control exerted by cAMP lies in the sophisticated properties of individual ACs, in terms of their coincident responsiveness, dynamic protein scaffolding and organization of cellular microassemblies. The ACs appear to

be the centre of highly organized microdomains, where both cAMP and Ca2 + , the other major influence on ACs, change in patterns quite discrete from the broad cellular milieu. How these microdomains are organized is beginning to become clear, so that ACs may now be viewed as fundamental signalling centres, whose properties exceed their production of cAMP. In the present review, we summarize how ACs are multiply regulated and the steps that are put in place to ensure discrimination in their signalling. This includes scaffolding of targets and modulators by the ACs and assembling of signalling nexuses in discrete cellular domains. We also stress how these assemblies are cell-specific, context-specific and dynamic, and may be best addressed by targeted biosensors. These perspectives on the organization of ACs uncover new strategies for intervention in systems mediated by cAMP, which promise far more informed specificity than traditional approaches.

ACs CAN BE SUBJECT TO MULTIPLE REGULATORY INFLUENCES

pathways, such as those governed by ERK (extracellular-signalregulated kinase), MAPK (mitogen-activated protein kinase) and β-arrestins [6–9]. Thus there is a clear need to organize these elements precisely in cells. The best understood of these regulatory events, because of its dependence on in vivo responses, and the best characterized at the molecular, cellular and physiological levels, is the interaction between Ca2 + and ACs [12]. This illustrates the organizational abilities of ACs.

Key words: adenylate cyclase, A-kinase-anchoring protein (AKAP), focal adhesion, microdomain, protein kinase, scaffold.

REGULATION OF ACs BY CA2 + IS ELABORATE AND CRITICAL

The organizing powers of ACs are exemplified in their regulation by Ca2 + , in which the concept of AC-recruitment of effectors clearly emerges along with the notion of AC-centred microdomains of cAMP and Ca2 + . Exploration of the regulation of susceptible ACs by Ca2 + in vitro is relatively straightforward, although it does reveal features that are exploited in vivo. Inhibition in vitro

AC5 and AC6 are inhibited directly by submicromolar concentrations of Ca2 + in a manner that exploits the octa-chelating properties of Ca2 + compared with the hexa-chelating powers of the far more abundant cellular cation Mg2 + [13]. Mg2 + cannot

Abbreviations: AC, adenylate cyclase; AKAP, A-kinase-anchoring protein; CaM, calmodulin; CNG, cyclic nucleotide-gated; ECM, extracellular matrix; eNOS, endothelial nitric oxide synthase; Epac, exchange protein directly activated by cAMP; ER, endoplasmic reticulum; ERK, extracellular-signalregulated kinase; FA, focal adhesion; IM, ionomycin; Nt, N-terminal; PDE, phosphodiesterase; PKA, protein kinase A; PKC, protein kinase C; PM, plasma membrane; sAC, soluble AC; SOCE, store-operated Ca2 + entry; TmAC, transmembrane AC; TSH, thyroid-stimulating hormone. 1 To whom correspondence should be addressed ([email protected]).  c The Authors Journal compilation  c 2014 Biochemical Society

Biochemical Journal

The nine mammalian TmACs [transmembrane ACs (adenylate cyclases] are highly conserved in general structural terms, yet they vary considerably outside their catalytic and transmembrane domains (Figure 1), which allows them to display individual properties [1–4]. The complexity of their transmembrane architecture prompts comparisons with ion channels and membrane transporters [5], and, as with ion channels, this may also provide domains for interacting with other membrane proteins. Early studies of overexpressed ACs, coupled with inferences based on structural homologies, led to a consensus view on specific AC regulation as presented in Table 1. Evidence accumulated over 20 years has indicated that that ACs are multiply responsive to, e.g., G-protein subunits (α s , α i and βγ ), PKC (protein kinase C), PKA (protein kinase A) and tyrosine kinases [1,4]. However, the evidence as applied to individual ACs is often less than systematically established [1]. A major goal of the present review is to make the case for determining how the specific cellular context and the cadre of the required regulatory elements will determine how specific ACs will be regulated. Assessing the ACs may be quite challenging in the native contexts in which they occur, as we develop below. By being multiply responsive to, e.g., βγ subunits of G-proteins, PKC and Ca2 + , the ACs can interdigitate the activities of many cellular signalling

www.biochemj.org

Adenylate cyclase-centred microdomains

200

Figure 1

D. M. F. Cooper and V. G. Tabbasum

General structural domains of the membrane-bound ACs

The nine ACs each comprise two transmembrane clusters (TM1 and TM2) of six membrane-spanning domains. TM1 and TM2 are joined by an intracellular loop containing the C1a and C1b regions. Following TM2 is a long intracellular tail, containing the C2a and C2b regions before the C-terminus. The N-terminus, C1b and C2b regions are the most diverse among the ACs in terms of length and composition; the C1a and C2a regions are highly conserved and associate to form the catalytic domain. Reproduced from [1] Halls, M.L. and Cooper, D.M.F. (2011) Regulation by Ca2 + -signaling c pathways of adenylyl cyclases. Cold Spring Harb. Perspect. Biol. 3, a004143, with permission. 2011 Cold Spring Harbor Laboratory Press.

Table 1

Regulation of the adenylate cyclases

The Table is a compilation from the reviews [1,3,4,8,10,11]. CaMK, Ca2 + /calmodulin-dependent protein kinase; RGS, regulator of G-protein signalling; RTK, receptor tyrosine kinase. Regulator G-protein

Gs α Gi α Gβγ

Forskolin Ca2 + Kinase

CaMK II CaMK IV PKA PKC RTK

RGS

AC isoform

Effect

All AC1, AC3, AC5, AC6, AC8, AC9 AC1, AC8 AC2, AC4 All AC1, AC8 (via CaM) AC9 (via calcineurin), AC5, AC6 AC3 AC1 AC5, AC6, AC8 AC2, AC4, AC5, AC6, AC7 AC6, AC9 AC1, AC5, AC6 AC3, AC5, AC6

Stimulation Inhibition Inhibition Stimulation Stimulation Stimulation Inhibition Inhibition Inhibition Inhibition Stimulation Inhibition Stimulation Inhibition

simultaneously chelate both cAMP and pyrophosphate (PPi ) (the inhibitory by-product of cAMP formation), so that PPi is released. However, in the presence of Ca2 + , PPi can remain at the AC catalytic site and inhibit activity. The crystal structure reveals the precise interactions [14].

Stimulation in vitro

AC1 and AC8 are activated by Ca2 + acting via CaM (calmodulin). In vitro studies showed differences in the regulation by Ca2 + /CaM of AC1 and AC8, which suggested distinct applications of these properties in intact cells. AC1 appears to be regulated by Ca2 + /CaM binding directly to the C1b domain in a direct activation mechanism. The critical residues that bind CaM near the catalytic domain have been identified [15,16]. A more complex mechanism is used by AC8, which is reminiscent of the use of CaM by L-type Ca2 + channels [17,18]. CaM is bound tightly by a classic 1-5-8-14 loop at the N-terminus of AC8. Nevertheless, this is not where CaM acts to stimulate  c The Authors Journal compilation  c 2014 Biochemical Society

activity; this site appears to garner limited CaM from its other competitors in the cell [19]. In fact, the N-terminal binding site can be deleted or mutated and the enzyme remains fully stimulable in vitro. The vital role of the N-terminal binding site becomes apparent in vivo (see below). Upon binding Ca2 + by the Cterminal EF-hand pair on CaM, a conformational state is induced that allows the transition of the CaM to an IQ-like domain at the C-terminus of AC8 and a disinhibitory activation ensues [16,20– 24]. The use of shuttling between a classic 1-5-8-14 motif and a ‘Ca2 + -independent’ IQ motif is mirrored in L-type Ca2 + channels [17]. Regulation in vivo Non-excitable cells

The ACs that are regulated by Ca2 + in vitro (AC1, AC5, AC6 and AC8) are regulated by physiological transitions in Ca2 + . In non-excitable cells, this occurs as a consequence of SOCE (store-operated Ca2 + entry) [25] either triggered by physiological means, such as agonist-mediated depletion of inositol 1,4,5trisphosphate-sensitive stores, pharmacologically, as a result of passive store depletion by inhibition of Ca2 + -reuptake (e.g. by thapsigargin), or indirectly, as a result of the depletion of stores by ionophore-driven emptying [26]. Remarkably, neither release of Ca2 + from stores nor the very large accumulation of Ca2 + entry triggered by high concentration of ionophore can regulate these enzymes. Thus the efficacy of Ca2 + is not due to the global cytosolic concentration that is achieved, but to the mechanism by which Ca2 + is elevated. The underlying mechanism in the case of AC8 relies on a direct interaction between the AC and the channel element of SOCE channels, Orai1 (see below) (Figure 2). Strong circumstantial evidence suggests that the same dependence applies to AC1 in blowfly, pancreatic MIN6 cells [27] and the sino-atrial node [28,29]. AC5 and AC6 which show the same requirement for Ca2 + entry over other modes of Ca2 + elevation might be expected to show similar interactions. Another channel, the cGMP-stimulated wild-type rat olfactory CNG (cyclic nucleotide-gated) channel, when expressed by adenovirus, can mediate inhibition of the endogenous AC6 of C6-2B cells [30].

Adenylate cyclase-centred microdomains

Figure 2

201

Binding of AC8 and Orai1 in a cholesterol-rich domain

AC8 enriches in cholesterol-rich domains where it binds to Orai1 via their respective N-termini [41]. The interaction is unaffected by Ca2 + depletion, but is perturbed by disturbing the integrity of lipid rafts (e.g. cholesterol depletion or cytoskeletal disruption) [69,73,74]. STIM, stromal interaction molecule.

Excitable cells

Insensitivity to ionophore 2+

ACs are regulated by L-type Ca channels in excitable cells, both when exogenously and when endogenously expressed. For instance, AC8 when expressed in GH3 and pancreatic MIN6 cells is stimulated [29,31], the endogenous AC5/AC6 in cardiomyocytes is inhibited, and the endogenous AC1/AC8 in hippocampal slices is stimulated [32,33]. The issue of the selectivity of ACs in excitable cells for VGCCs (voltagegated Ca2 + channels) over other forms of Ca2 + entry has not received the same exploration as has SOCE in non-excitable cells. Nevertheless, in MIN6 cells, Kitaguchi et al. [27] showed a regulation of the endogenous AC1 by L-type Ca2 + channels, but not by SOCE or receptor-induced release from stores. It has been suggested that the ACs may form a complex with L-type channel elements and, although the evidence is not as direct as in the case of the AC8–Orai1 complex, ACs and channels do bind shared elements, e.g. AKAPs (A-kinase-anchoring proteins) (see below), so this is quite a viable possibility [34]. Regulation by the emptying of stores

AC activity has been proposed to be stimulated by depletion of Ca2 + stores in a manner that is independent of cytosolic Ca2 + levels, but responsive to the state of filling of the stores [35]. The AC species involved has been proposed to be AC3 [36]. There is no suggestion that the effect involves any regulation by direct effects of Ca2 + ; indeed, in these experiments, Ca2 + entry appears to inhibit the AC3, but the effect is some indirect consequence of store depletion. Given the major cellular consequences that occur upon Ca2 + store depletion, including cytoskeletal movements and ER (endoplasmic reticulum) reorganization [37], it may not be too surprising that the trafficking of membrane-directed proteins (such as AC) might be altered, so that new activity becomes measurable.

The independence of Ca2 + -sensitive ACs from ‘trivial’ forms of Ca2 + rise is a continuing puzzle. Certainly, a major factor must be the extreme closeness of the ACs to Ca2 + channels. But the ineffectiveness of IM (ionomycin)-mediated Ca2 + elevation (when IM-triggered store-depletion is taken into account) is an abiding conundrum. One possibility is that the elevation in cytosolic Ca2 + achieved by IM does not occur in the domain of the ACs. For instance, IM is reported not to dissolve well in lipid membranes that contain cholesterol [38], so that the presence of the Ca2 + -sensitive ACs in cholesterol-rich rafts may distance them from IM-induced transitions in Ca2 + . Another possibility is that the IM-induced Ca2 + rise is precluded from reaching the AC by the presence of Ca2 + buffers or Ca2 + pumps in the vicinity of the AC, which mutes the Ca2 + rise. When Ca2 + was measured in the vicinity of an AC using an attached fluorescent protein sensor for Ca2 + , GCaMP [39,40], a very small IM component, as well as a SOCE component, was detectable in response to IM [40,41], although this did not compare with the Ca2 + rise reported by a cytosolic sensor [40]. So it may well be that insufficient Ca2 + reaches the vicinity of the AC, and the impression conveyed by a cytosolic sensor exaggerates the Ca2 + in the AC domain. Another consideration is that the GCaMP only measures Ca2 + ; this leaves open the possibility that the AC8–GCaMP sensor ‘sees’ the Ca2 + , but the AC does not respond. This option raises the possibility that the nature of the Ca2 + rise achieved by ionophore is different from that achieved by a channel. Clearly ionophores facilitate the transport of ions across membranes in a continuous, concentration-driven, manner, whereas channels open and close and yield rapid rises and falls in very high (up to 100 μM) ion concentration in the immediate vicinity of the channel mouths [42]. Thus it is conceivable that the ACs respond to these increments rather than to the absolute amounts of Ca2 + achieved. An analogous situation has been nicely dissected in  c The Authors Journal compilation  c 2014 Biochemical Society

202 Table 2

D. M. F. Cooper and V. G. Tabbasum AC-interacting proteins and the supporting experimental evidence

A wide range of experimental evidence supports the associations noted in the Table. In some cases, the support is diverse and extensive; in others it is more tentative (as indicated). The ideal breadth of different types of support may not always be feasible logistically, but ideally the evidence should go beyond complexes inferred from co-immunoprecipitation, GST pull-downs or co-localization, given the fact that ACs are membrane-inserted proteins and detergent solubilization is not the most rigorous of techniques and antibody specificities do not always translate from the Western blot to the immunofluorescence platform. Experimental approaches applied: a, CaM overlay; b, mutagenesis of proposed binding sites; c, immunoprecipitation AC assay; d, co-immunoprecipitation; e, pull-down to map interaction sites; f, overexpression; g, shRNA or siRNA knockdown; h, putative domains deleted; i, FRET, FLIM (fluorescence lifetime imaging microscopy); j, yeast two-hybrid; k, disruptor peptides; l, Ca2 + measurements; m, co-localization/immunostaining; n, TIRFM (total internal reflection fluorescence microscopy); o, peptide array; p, truncations; q, overlay assay; r, peptide competition; s, acceptor bleaching FRET; t, FCS (fluorescence correlation spectroscopy); u, FRAP. mAKAP, muscle-specific AKAP. AC isoform AC1

C1b

AC2

Nt

AC3

AC4 AC5

Nt, C1, C2 Nt Nt Nt C2 C1

AC6

AC7 AC8

AC9

Nt Nt

Nt Nt, C2b Nt Nt Nt N/A Nt Nt Nt

Binding partner

Experiment

Reference(s)

CaM Yotiao AKAP150 Yotiao AKAP79 Gβγ Caveolin1 AKAP79/150 Yotiao Gβγ mAKAP79β (Nt, residues 275–340) AKAP79/150-(77–153) Ric8a Gs α Gs βγ PAM-(1028–1231) (protein associated with Myc) RGS2 (Nt, first 19 residues) Caveolin1, phospho-caveolin1 (Tyr14 ) Caveolin3 AKAP79/150-(77–153) Snapin-(33–51) Caveolin1, phospho-caveolin1 (Tyr14 )

a, b c d c, e d, f, g d c c, d c d c, e c, d, e, g, h, i, k, l d, e, j, m d, e d, e, i j, g b, e d, g, m d c, d, e b, g, j, d, e, c, m, k d, g, m

[54–56] [57] [47] [57] [50,58] [59] [60] [47,58] [57] [61] [49] [47,58] [62] [3] [3] [63] [64] [65] [66] [47,58] [67,68] [65]

Orai1 (Nt) CaM actin AKAP79/150 Cholesterol (ergosterol) WFS1?* PP2Ac Yotiao-(1–808) AKAP79-(77–153)

i, g, e, d, n, o, p, b q, b, r, p e, d, h, m, t, u e, d, h, g, i, k s d, f, g j, e c, e c, e

[41] [20,22] [69] [48] [69] [70] [71] [51,57] [58]

*Discordant molecular masses are ascribed to putative AC8 bands.

the case of the control of sperm turning by Ca2 + entry via CNG channels. It turns out that, by modelling the rate of change, not the absolute amount of Ca2 + , it was possible to relate how changes in Ca2 + concentration controlled turning. Alvarez et al. [43] showed that the sperm responded to a time derivative of [Ca2 + ]i which they proposed could be based, for instance, on the different time constants and co-operative Ca2 + binding to the EF-hands of CaM. This situation could readily apply to the Ca2 + /CaM-regulation of AC1 and AC8. Detailed analysis of the effects of the relevant CaM-binding Nt (N-terminal), Ct (C-terminal) or C1b peptides from AC8 and AC1 respectively show quite different effects on Ca2 + -binding kinetics by the N- and C-lobes of CaM [24], which could readily introduce a ‘chemical differentiation’ step [42] that could allow ACs to respond to rapid Ca2 + gradients. The N-lobe of the CaM has extremely high affinity for Ca2 + when bound to the N-terminus of AC8, but this is reduced by two orders of magnitude upon its transition to the C-terminus [24]. Thus both a highly discerning affinity for very high Ca2 + concentrations is built in, as are found near the mouths of ion channels, but not at ionophores, as well as a temporal differentiation step due to the transition of the CaM to the C-terminus, resulting in a reduced affinity for Ca2 + , which delays the process and  c The Authors Journal compilation  c 2014 Biochemical Society

allows a temporal differentiation step to be present (see also [42]).

ACs ARE SCAFFOLDS

ACs are much more than producers of cAMP. Table 2 lists proteins that have been demonstrated to interact with ACs; the evidence varies from extremely robust to suggestive, but it is clear that a wide variety of proteins can be involved. The best-represented group are the AKAPs, which are bound by every AC species [44]. The discovery of AKAPs as devices that minimally recruit PKA, a major target for cAMP, along with, often, an effector of cAMP, as well as a cellular targeting domain was a major conceptual advance in imparting great efficiency to cAMP signalling by minimizing diffusion distances for interacting components [45,46]. The fact that all of the ACs bind some selection of AKAPs allows even more immediacy in co-localizing the source and the target of cAMP. Well-established examples of AC–AKAP complexes include AC5/AC6–AKAP79– L-type channel–PDE4D3 (PDE is phosphodiesterase) [47], AC8– AKAP79–PDE4D3 [48] and AC5–mAKAP (muscle-specific

Adenylate cyclase-centred microdomains

AKAP) in cardiomyocytes [49]. However AC–AKAP complexes are more than just co-localized interacting assemblies of AC– PKA and PKA targets; they are also devices for integrating with other signalling systems. For instance, AKAP79 binds PKC as well as PKA, and this allows AC2 to be regulated by the PKC pathway, as stimulated by muscarinic cholinergic receptors [50]. AC9 through binding the AKAP Yotiao permits the regulation of the I Ks (slow outward K + current) by both calcineurin and PKA in cardiac myocytes [51]. In another example, the relaxin RXFP1 receptor recruits an AC2–AKAP79–PDE4D3 complex which associates with β-arrestin signalling, allowing a novel pathway to be activated [52]. AC–AKAP complexes can also act as bridges to effectively permit the AC recruitment of other targets of PKA such as L-type channels [34] and K + channels [51] to yield other versions of AC-centred complexes. The interaction between the ACs and the AKAPs can be expected to be dynamic and regulated. Indeed, the PKA bound to the AKAP79–AC8 complex phosphorylates the N-terminus of AC8 and reduces its responsiveness to Ca2 + rises, which may provide a device for a feedback inhibition or the generation of cAMP oscillations, depending on the context [53]. ACs bind scaffolds other than AKAPs, such as PP2A (protein phosphatase 2A), a complex multiply interacting scaffold, which organizes numerous pathways [71,72]. In any proteomic analysis or yeast two-hybrid analysis of putative AC-interacting proteins, cytoskeletal elements are generally encountered ([71,72], and K.L. Everett, A. Schotten and D.M.F. Cooper, unpublished work). Such interactions would be expected from either stable interactions or trafficking involving the cytoskeleton (see below). The ACs are encountered in regulatory complexes that include not only the downstream targets of cAMP, such as voltage-gated Ca2 + channels and K + channels, but also regulators of AC such as Orai1 (the likely basis for the dependence of AC8 on SOCE) and L-type channels which can feedback inhibit (or stimulate) the ACs [34]. AC8 binds Orai1 directly via their respective N-termini, as established by peptide array and FRET analysis of truncated and mutated forms, co-immunoprecipitations and functional assessments of Ca2 + and cAMP in the domain of the complex [41] (Figure 2). AC5/AC6 may bind L-type channels indirectly by sharing AKAP79 [34]. Two extensive reviews [3,75] summarized the consequences of AC-knockout and -overexpression studies which implicate ACs in processes as diverse as learning and memory (AC1/AC8), olfaction, nociception (AC3), control of movement (AC5) and cardiac output (AC5/AC6). However, within the context of the present review, one might expect difficulties to arise in distinguishing the importance of the loss of the ability to produce cAMP from the loss of AC scaffolds. The separable roles of AC compared with cAMP in cardiac physiology in terms of its effects on phospholamban phosphorylation, sarcoplasmic reticulum Ca2 + pump activity and L-type Ca2 + channel activity is illuminated by a series of reports on the effect of the knockout of AC6 and the knockin of a catalytically dead AC6 [76,77]. In AC6 knockouts, Ca2 + handling is enhanced in the left ventricle despite the absence of the cAMP that is normally considered to regulate the relevant proteins. However, cardiac function is preserved in knockin AC-dead mice regardless of the fact that the production of cAMP is reduced. The complete deletion of AC6 severely impairs cardiac function [78]. This draws attention to the likely role of the AC as a scaffold. The proteins involved and the mechanisms are still not understood. The latter example highlights the inevitable consequences of knocking out a scaffold rather than a simple source of cAMP. Although the first knockouts of AC1 and AC8 were enormously insightful in establishing the roles of these ACs

203

in learning and memory [79], some questions might now be posed on whether there had been any consequences of the mislocation of the proteins scaffolded by the ACs. Is it possible that the scaffolded proteins had complicating or compensatory actions in the development of those mice, or recruited/scaffolded proteins might be protected from degradation so that the absence of the scaffold would shorten their half-life, dampening any consequences of mislocation. The ideal approach seems to be to knock back in ACs with targeted deletions or mutations, notwithstanding that overexpression studies of dead enzymes may also have unforeseen consequences. For instance, in the study cited above, the mutant AC6 is apparently expressed at the nuclear membrane (by immunohistochemistry [77]) and it is possible that other ACs can be up-regulated so that the cAMP levels are unimpaired. Nevertheless, the role of ACs as scaffolds is an unavoidable experimental consideration. Another significant issue, which is apparent from Table 2, is that the ACs and their binding partners have binding sites for multiple partners, but presumably not at the same time, given that some of the binding regions overlap. Consequently, dynamic binding and release events should be expected. For instance, in AC8, the Nt-binding domain for CaM is the same domain as for AKAP79 and Orai1, as established by extensive mutagenesis, peptide array, functional and FRET studies [22,41,48]. The Orai1binding domain for AC8 can also bind phosphatidylinositol 4,5bisphosphate and be phosphorylated by PKC [80], so it can be anticipated that access to the various targets is regulated. Thus exchange of partners in different settings should be anticipated. In addition, there is a short form of Orai1 [80] (Orai1b) with an alternative start site which lacks the first 60 residues of Orai1a, rendering it incapable of interacting with, or regulating, AC8, so that the relative abundance of these forms will determine which is in complex with AC8. These complex issues underline the challenges and opportunities now presented by ACs.

DYNAMIC CELLULAR ORGANIZATION OF ACs

The foregoing discussion has made the case for considering the ACs as scaffolds, which group responsive proteins in compact arrays. The AC scaffolds are also organized in a broader cellular context, so that discrete lipid and cytoskeletal interactions must be considered. Lipid rafts are seen as cellular devices that can concentrate and promote interactions between signalling targets and effectors [81]. These rafts are domains of the plasma membrane that are rich in cholesterol, sphingolipids and unsaturated side-chain phospholipids, a subset of which are caveolae, which are identifiable structures particularly prominent in endothelial cells [82]. The concept of lipid rafts has evolved significantly with the development of super-resolution microscopic and other live-cell methods and a growing awareness of the heterogeneity and dynamism of lipid elements within the plasma membrane [83]. Consequently, whereas numerous proteins have been found in biochemically prepared rafts by proteomic analyses, individual rafts are seen as often transient ‘nanodomains’ with sizes estimated between 5 and 300 nm, which can accommodate 10–30 protein molecules [84–87]. Some ACs (AC1/AC3/AC5/AC6/AC8) are consistently associated with such rafts and others (AC2/AC4/AC7/AC9) are not [88–90]. The evidence gathered to support this selection ranges from crude biochemical cellular fractionation to more elaborate livecell methods [69,90,91]. The cytosolic regions appear to be responsible for the residence in rafts of AC5 and AC6 [92,93] which suggests that protein–protein interactions might underlie these associations. It is also interesting that AKAP79 can be  c The Authors Journal compilation  c 2014 Biochemical Society

204

D. M. F. Cooper and V. G. Tabbasum

palmitoylated and as a consequence occur in rafts along with AC8 [94,95]. Orai1 and TRPC1 (transient receptor potential canonical 1), channel components of SOCE, are also raft-resident proteins [96,97], as are CNG channels [98] and L-type channels [99]. Identification of the precise targeting mechanisms of proteins to rafts is rendered difficult by the fact that the forces involved in sustaining raft residence must be a summation of low-affinity interactions with cholesterol and other cholesterol-binding or palmitoylated protein partners. Rafts are more dynamic than simple platforms for signalling molecules. Besides diffusing laterally in the bilayer, rafts also undergo intracellular assembly and trafficking to and from the plasma membrane as well as endocytosis and inward trafficking to an array of subcellular compartments [100– 102]. Movement of these nanoplatforms is dictated in part by their sorting signals and maintained through the integrity of various cytoskeletal structures [103–106]. The involvement of cytoskeletal elements in trafficking of rafts to the plasma membrane [107] make it unsurprising that proteins such as actin, actinin, filamin, paxilin, vinculin, myosin and raft markers (e.g. caveolin) are not uncommonly associated with ACs by a range of direct co-immunoprecipitation, proteomic or livecell approaches [65,69,90,108,109]. The associations of ACs with these cellular features can cloud interpretations of specific regulatory interactions from more generic trafficking interactions. Doubtless trafficking is specific and likely to be individual for every protein; nevertheless the overall trafficking process, as might be perturbed by, e.g., brefeldin, is not centred on specific proteins, such as the ACs. Yet specific interactions do seem to occur between actin and ACs and it can be difficult to separate a vital role in permitting the precise cellular placement of ACs from a general role in trafficking and recycling ACs between various destinations [69,90]. Ric8A, which was originally proposed as a guanine-nucleotideexchange factor, bound to the N-terminus of AC5 and inhibited its activity [62]. This may now be a good example of a trafficking rather than a regulatory cofactor. Ric8a has been mooted as a molecular chaperone for G-proteins [110] and so may simply interfere with the PM (plasma membrane) delivery of AC5 or its components. Another protein, snapin, was one of the first proteins shown to interact with AC6 [75] and was originally posited as an element of the SNARE (soluble N-ethylmaleimide-sensitive fusion protein-attachment protein receptor) complex. This protein is now viewed to have a more widespread distribution and to be more widely involved in vesicular trafficking [111]. A related problem arises with caveolin. Caveolin has been considered as a specific regulator of AC5 [90] and, indeed, a peptide derived from the putative AC5-binding sequence of caveolin potently inhibited AC5 activity, which led to the proposal that caveolin exerted a specific inhibitory regulatory influence on AC5 [66]. Developments in eNOS (endothelial nitric oxide synthase) research have cast a different light on this issue. eNOS has been shown to bind caveolin, to be inhibited by caveolin, and this inhibition is relieved and antagonized by Ca2 + /CaM binding to a domain on eNOS. Consequently, a rather elegant mechanism was suggested involving activation/inactivation cycles of caveolin binding/unbinding that were controlled by transitions in Ca2 + [112]. Recently, it has been demonstrated that in the mature eNOS structure, the caveolin-binding domain of eNOS is quite inaccessible to caveolin because of the three-dimensional conformation adopted by the active form of eNOS [100,113]. It turns out that many caveolin-binding domains are similarly inaccessible in mature proteins [114]. Consequently, the previous mechanism for eNOS is suggested to be inadequate. A compromise might be that where, as part of the undoubted  c The Authors Journal compilation  c 2014 Biochemical Society

trafficking and selective residence of eNOS to caveolae, eNOS, although initially unfolded, does form a tight association with caveolin through this binding site; this binding to caveolin ensures its trafficking to caveolae and, upon elevation of Ca2 + levels, caveolin is released and eNOS adopts its active conformation permanently in a manner no longer subject to caveolin. A similar situation may apply to ACs where trafficking to caveolae in endothelial cells or cardiomyocytes requires an explicit (and inactivating) association with caveolin which is relieved upon the delivery of the ACs to its appropriate environment [88,115]. Biogenesis, delivery and endocytosis of ACs is a topic that has not been explicitly explored to date. We should expect that the targeting of Ca2 + -regulated ACs to cholesterol-rich domains will rely on specific sorting signals, interactions with cholesterol, cytoskeletal elements, other raft-targeted proteins and possibly association with chaperones (including caveolin [69,91]). A summary of these interactions is presented in Figure 3, from which it can be observed that disruption of these essential processes at any step can ultimately have consequences for overall AC localization, association with lipid rafts and AC activity. The foregoing discussion proposed that AC scaffolds are organized in their cellular destination as part of a regulated, but continuous, process. A more dynamic situation may occur at FAs (focal adhesions), where a scaffolding AC could play a key regulatory and dynamic structural role. FAs are large macromolecular assemblies through which both mechanical force and regulatory signals are transmitted [118]. They link the cytoskeleton to the ECM (extracellular matrix) and are considered as subcellular macromolecular signalling hubs that mediate the regulatory effects of ECM adhesion on cell movement. As well as ERK signalling processes, both Ca2 + and cAMP play major roles in FA dynamics [119,120]. Transiently forming rafts are also part of the FA cycle of release and attachment to the cellular substrata [121]. The Ca2 + -inhibitable AC6 could be poised to play an important role in such situations given that high cAMP promotes actin depolymerization and low cAMP allows actin to stay polymerized and maintain contact with integrins at the FA. Thus AC6, which binds actin at FAs [65], is susceptible to the influx of Ca2 + to keep cAMP levels low. AKAP220 has already been reported at FAs [122], as have the elements of SOCE [123] in addition to their previously known raft associations [93,124]. Thus it seems a small step to propose that a raft-localized Ca2 + inhibitable AC scaffolding a palmitoylated AKAP and subject to inhibition by a raft-localized SOCE complex could play a key role in maintaining FAs. Removal of the Ca2 + -elevating stimulus would allow cAMP to rise, the actin to depolymerize, the FA to collapse and cell migration to resume (Figure 4). The foregoing discussion aimed to convey that the interactions of ACs will be determined by context, that the organization of ACs in the cell will be dynamic and can be expected to change throughout the behavioural activities of the cell [90]. The ACs may act as foci for other regulatory molecules, which have their own coterie of associated proteins, and the ACs may interact with the cyto-architectural elements, so that they can be viewed as sculptors of their own domains both by recruiting proteins and by producing cAMP, which affects actin polymerization.

ACs MAY BE INTERNALIZED AND SIGNAL INSIDE CELLS

The organization of ACs at the PM, coupled with a proposed firewall of cytosolic PDEs has sometimes prompted speculation as to how cAMP can reach different internal areas of the cell. It has been clearly argued [129] that the placement of PDEs will have a major influence on where cAMP may, or may not, reach. Tethered

Adenylate cyclase-centred microdomains

Figure 3

205

Likely trafficking pathways of raft-localized ACs

AC (green circles) is synthesized in the ER and dimerizes early during secretion [116]. During biogenesis, AC is complexed with N-glycans which act as sorting signals for its retention in the PM [91,117]. AC is then assimilated into lipid-rich particles (blue circles) at some point during transition through the Golgi. Upon exiting the Golgi, the AC–lipid complexes are shuttled to the PM lipid rafts, tracking along the cytoskeleton (red lines), where they may interact with Orai1 or palmitoylated AKAP79 or other partners. Lack of N-linked glycosylation results in retention at the PM for a shorter period of time and fast rejection into sub-plasmalemmal structures (endosomes) [91]. Cytoskeletal remodelling critically influences both localization and association of AC with lipid-rich domains [69]. The integrity of the cytoskeleton is therefore critical for biogenesis, recycling, association with lipid rafts and, consequently, functional activity of AC.

Figure 4

Dynamic association of Ca2 + -regulated ACs at FA sites

FAs are cholesterol-rich dynamic microplatforms that form at the leading edge of cells upon binding of integrins to the ECM (navy broken lines) [125]. Following ligand binding, the integrins cluster, recruit and concentrate an array of signalling molecules, including talin, FAK (focal adhesion kinase), vinculin and actinin. The resultant FA complexes initiate actin stress fibre formation along with a number of signalling pathways [126]. FA sites and rafts are also areas of intense Ca2 + and cAMP signalling, and the Ca2 + -regulated AC5/AC6 (black rectangles) and SOCE apparatus (double vertical grey blocks) enrich in these specialized lipid domains [119,127,128]. High Ca2 + by inhibiting AC5/AC6 keeps cAMP low, and promotes actin polymerization and the integrity of FAs. When Ca2 + declines, the rise in cAMP allows actin to depolymerize and the FAs to collapse [65,90].

 c The Authors Journal compilation  c 2014 Biochemical Society

206

D. M. F. Cooper and V. G. Tabbasum

internal sensors for cAMP do rapidly report cAMP in response to external stimuli so that, apart from simple diffusion, the formal possibility can be considered that there may be internal sources of cAMP [apart from sAC (soluble AC), which is mentioned below]. Against this background, a body of intriguing and provocative papers have suggested that ACs are internalized and can be a source of internal cAMP. The typical experiment traces fluorescently tagged receptors, G-proteins and perhaps BODIPY– forskolin binding (as a surrogate for the catalytic unit of ACs). For instance, Calebiro et al. [130] presented an elegant study showing that, whereas short stimulation by TSH (thyroid-stimulating hormone) of thyroid follicles showed transient increments in cAMP, a longer treatment, e.g. 30 min, resulted in a sustained elevation in cAMP. Associated with this sustained elevation was a sustained internalization of the TSH receptor, Gs α and even AC3, as well as BODIPY–forskolin. Although it is advisable to be mindful of the likely receptor and Gs internalization that always accompanies prolonged receptor stimulation, as well the functionality of the tools, such as the ability of a putative anti-AC3 antibody to report the likely endogenous levels of AC3 in cells, nevertheless the most attractive interpretation of such findings amounts to a major additional signalling option for ACs. A different interpretation of these results comes from Neumann et al. [131] who suggested that inhibition of TSH receptor internalization by dynamin inhibition did not affect persistent cAMP accumulation, and concluded that TSH remaining at the cell surface was responsible even for the prolonged effects. A possible explanation builds on the longstanding observation that extended receptor activation is accompanied by coupling of βarrestins to receptors, which leads to a switch in coupling from Gs to Gi , with consequent activation of the ERK pathway [6]; the internalized β-arrestin activates ERK which in turn can lead to phosphorylation and inhibition of the activity of long PDE isoforms [6]. Thus ERK activation would result in increased cAMP levels independently of acute AC stimulation. Such is the interpretation of the effects of low- compared with highdose isoprenaline (isoproterenol) on acute or sustained cAMP accumulation in cardiac myocytes [132]. Thus, although the possibility that ACs might continue to signal even when (and if) internalized remains an attractive hypothesis deserving continuing reassessment, alternative interpretations of the data such as normal recycling of ACs to and from the PM (Figure 3) or β-arrestin effects on PDEs, along with some reservations on some of the reagents, dilutes enthusiasm for embracing internal ACs as a significant cAMP signalling option at present. An alternative source of internal cAMP can be the sAC which possesses no transmembrane segments. First discovered in high concentrations in sea urchin sperm and responsive to pH [133], it was later found to be widespread in mammalian sperm, where, as part of the acrosomal reaction, it reacted with bicarbonate [134]. It is not stimulated by forskolin or G-proteins, and, for activity to be measured in vitro, high concentrations (2–5 mM) of either Mn2 + or Mg2 + and high millimolar concentrations of ATP are required (K m ∼2 mM compared with 0.2 mM for TmACs). This requirement for millimolar concentrations of ATP has been taken to suggest that sAC may act as a sensor of ATP [135]. For instance, when cells overexpressing sAC are poisoned with mitochondrial toxins such as azide, the cAMP levels fall along with the cellular ATP. Even with mitochondrial poisons, cellular ATP levels would not approach limiting concentrations for the TmACs, but, for sAC, because of its high K m when poisoned, cAMP levels due to overexpressed sAC fall, whereas those stimulated by forskolin (i.e. TmACs) are unaffected. From targeted disruption experiments, there seems little doubt that sAC plays a vital role in sperm motility [136], but the role  c The Authors Journal compilation  c 2014 Biochemical Society

of sAC in tissues where it is expressed at low levels is a difficult experimental challenge. An observation that can be made is that the classic inhibitor of TmACs, 2 ,5 -dideoxyadenosine, can only ever inhibit cAMP accumulation by ∼95 %. (In the context of the present review, it is not surprising that compartmentalized ACs may not be fully accessible to inhibitors, and inhibitors can rarely be expected to reach intracellular concentrations where they can achieve 100 % inhibition.) Under such circumstances, the putative sAC inhibitor KH-7 can be shown to inhibit the residual cAMP accumulation and some downstream events, thus the presence and involvement of sAC can be proposed. A potential problem in such experiments is that KH-7 also inhibits glucose oxidation and so leads to decreased ATP levels [137], so it is difficult to separate the role for a sAC that depends on very high ATP levels from a process that depends on high ATP levels.

AC-CENTRED MICRODOMAINS AND KINETICS OF cAMP AND CA2 + ARE REVEALED WITH TARGETED SENSORS

The foregoing discussion of AC-based signalling hubs implies that cAMP dynamics might be expected to vary in different AC microdomains. Support for such concepts required the development of live-cell sensors for cAMP. Traditional methods that measured the cAMP produced by a multitude of cells at a single time point constrained thoughts on this messenger to the conceptual doldrums of no spatial and limited temporal resolution. Following the clever, but technically demanding, efforts of patchcramming of CNG channels [138] and the microinjection of chemically fluorescently labelled PKA subunits [139], the leap to genetically encoded sensors brought temporal and spatial awareness of intracellular cAMP as a signal to the same level of sophisticated insight and framing of hypotheses as had been the case for Ca2 + . The development of FRET-based cAMP sensors built around Epacs (exchange proteins directly activated by cAMP) allowed the single polypeptide to be targeted to discrete domains of the cell by the use of PM-targeting sequences derived from, e.g., Lyn kinase, or directly attached to ACs, to sample cAMP in the domain from which it emanated or to which it might diffuse from other sources [140–143]. (The placing of a substantial Epac tag on either the N- or C-terminus of AC8 has no deleterious effect on AC responsiveness either in vitro or in vivo.) Although compartmentalization had been proposed as the only likely explanation for phenomena such as the fact that two agents which caused equivalent elevations in cellular cAMP did not cause equivalent physiological outcomes [144], the concept was not universally embraced, given the prevailing ignorance at the time of how cells, and ACs, were organized. Twenty years ago, the apparent differences in diffusion between Ca2 + and cAMP [145] were invoked to explain presumed targeted signalling by Ca2 + , but not by cAMP. With the growing realization of mechanisms such as PDEs to limit cAMP diffusion [129], the organization by scaffolding of cAMP targets, combined with single-cell imaging of dynamic cAMP kinetics, these inevitably na¨ıve discussions have become moot. At the same time, it remains critical to acknowledge that any measurements (and concepts derived therefrom) do reflect the measuring methodology rather than the absolute entity being measured. In the case of Ca2 + , the ready acknowledgement was instructive that, since fluorescent derivatives of EGTA were the basis for quin-2, fura-2 and indo-1 measurements, the apparent cellular dynamics of Ca2 + must reflect the presence of the fluorescent Ca2 + reporter, its concentration and its ability to bind and release Ca2 + [146]. Probe concentration and kinetics have profound implications for

Adenylate cyclase-centred microdomains

Figure 5

207

Distinct cAMP events in the domain of a pituitary-derived GH3 cell endogenously expressing a Ca2 + -stimulable and a Ca2 + -inhibitable AC

IP3 , inositol 1,4,5-trisphosphate; PLC, phospholipase C; TRH-R, Gq -coupled TRH receptor; VIP-R, Gs -coupled VIP receptor; + , stimulatory effect of Ca2 + ions on AC8 and Gs on global ACs; − , inhibitory effect of Ca2 + ions on AC5/AC6; red, cAMP emanating from global ACs; green, cAMP emanating from AC8. Adapted with permission from The Company of Biologists Ltd, from [143] Wachten, S., Masada, N., Ayling, L.J., Ciruela, A., Nikolaev, V.O., Lohse, M.J. and Cooper, D.M.F. (2010) Distinct pools of cAMP centre on different isoforms of adenylyl cyclase in pituitary-derived GH3 B6 cells. J. Cell Sci. 123, 95–106; permission conveyed through Copyright Clearance Center, Inc.

the manifestation of Ca2 + blips, puffs and sparks. Analogous reservations should attend the measurement of cAMP. Microdomains of cAMP have now been demonstrated using targeted sensors such as PKA, Epac and CNG channels [142,147,148]. Even more striking are the microdomains of both cAMP and Ca2 + that are manifest when sensors are targeted directly to the ACs, which reveal unique signals surrounding differently regulated ACs [41,143] (Figure 5). Targeted sensors are obviously the new key to exploring these domains. More than one AC is expressed in cells, with different responsive properties; hence microdomains can be expected to be the key level of organization for ACs, since their targets may also be extremely local. The likely local or targeted signalling by cAMP implies that a global elevation of cAMP or a generic inhibition of AC may give quite unintended and uninsightful consequences. Coincident with this reservation is the confusion attending elevation of cAMP by forskolin (which can activate every AC in a cell) the use of cAMP analogues with unknown permeation properties or PDE inhibitors that globally elevate cAMP to mimic the effects of hormones. Targeted cAMP and Ca2 + sensors revealed that, in fact, the ACs detect selective changes not only in cAMP in their domains in response to different signals, but also in Ca2 + , that are quite distinct from the dynamics in the global cytosol [40,73] (Figure 6). A particular power of AC-tagged sensors could be to probe the composition of AC microdomains, or the proteins which affect cAMP or Ca2 + in those domains. For instance, peptide or smallmolecule disruptors of scaffolding interactions or AC channel interactions, or siRNAs against components, could all be usefully assessed using targeted compared with global sensors of the second messengers. AC-CENTRED MICRODOMAINS ALLOW cAMP LEVELS TO OSCILLATE AND THE OSCILLATIONS CAN BE DECODED BY PKA

The information content provided to Ca2 + signalling by the manifestation and development of Ca2 + oscillations is widely acknowledged, particularly in the context of CaMKIII (Ca2 + /CaM

kinase III) [149–151] and in the different transcriptional regulation afforded by distinct temporal patterns of Ca2 + rises [152]. However, whether the opportunities offered by digital frequency-encoded signalling could apply to cAMP has, until recently, been more theoretical. It had been predicted that cAMP levels could oscillate as a function of feedback loops between ACs and Ca2 + channels under their control [12]. Indeed, over 40 years ago, Brooker [153] had shown that cAMP levels oscillated in the beating frog heart. That system had the advantage of a synchronized organ and the ability to freeze and extract cAMP at different parts of the contractile cycle. Clearly, in independent cells, the diffusion of cAMP away from its site of synthesis along with the influences that give rise to oscillations are unlikely to sustain the fidelity of cAMP oscillations distant from the enzyme source, a consideration which draws attention to the potential added value of localized cAMP microdomains. Using live-cell sensors, cAMP oscillations have been shown to be inducible artificially, or to occur spontaneously. [154–157]. In addition, the absolute need for these oscillations as distinct from a simple rise in cAMP has been demonstrated in the case of neuronal path-finding. Gaspar and colleagues’ exuberant axon pathfinding paradigm is an explicit example of the temporal dynamics of localized cAMP signalling [158]. In that system, pulses of cAMP determined by the entry of Ca2 + via L-type channels or released by pulsatile uncaging exerted regulatory effects, but global elevation of cAMP or inhibition of PDE induced spatially uninformative cAMP rises of no regulatory consequence [158]. Those effects were apparently mediated by AC1 and were not demonstrable in AC1-knockout mice. cAMP in those domains might be expected to oscillate if measurements could be made using a cAMP-sensor tagged AC1. Curiously, when AC1 and AC8 were expressed in the non-excitable HEK (human embryonic kidney)-293 cell line, the Ca2 + oscillations induced by muscarinic cholinergic receptor stimulation of SOCE could induce measurable cAMP oscillations using a global cAMP sensor in cells expressing AC8, but not AC1 [16]. These data were interpreted to reflect the different mechanism of AC activation by Ca2 + /CaM mentioned above. However, it is not inconceivable that AC1 could generate cAMP oscillations in the AC1 microdomain in response to  c The Authors Journal compilation  c 2014 Biochemical Society

208

Figure 6

D. M. F. Cooper and V. G. Tabbasum

Distinct Ca2 + microenvironments surround a Ca2 + -insensitive AC2 and a Ca2 + -stimulable AC8

Based on experiments with AC constructs tagged with the Ca2 + -sensor, GCaMPs. AC2 and AC8 associate with different sites of Ca2 + elevation. IP3 , inositol 1,4,5-trisphosphate; mACh, muscarinic acetylcholine receptor; PLC, phospholipase C; SERCA, sarcoplasmic/endoplasmic reticulum Ca2 + -ATPase. Reproduced with permission from The Company of Biologists Ltd, from [40] Willoughby, D., Wachten, S., Masada, N. and Cooper, D.M.F. (2010) Direct demonstration of discrete Ca2 + microdomains associated with different isoforms of adenylyl cyclase. J. Cell Sci. 123, 107–117; permission conveyed through Copyright Clearance Center, Inc.

the Ca2 + transients induced in exuberant neurons by L-type channels. Even though the fact that cAMP oscillations occur and that they can have explicit consequences has been shown, until very recently how these oscillations could be read by PKA had not been understood. Traditional, textbook views of the regulation of PKA had represented the dissociation of the PKA tetrameric holoenzyme in response to high concentrations of cAMP, such as those that are achieved by forskolin, the global activator of ACs and IBMX (3-isobutyl 1-methylxanthine) a global inhibitor of (most) PDEs. Obviously (as we have noted above), such experiments cannot address potential spatiotemporal properties of cAMP. But Taylor et al. [159] have now shown how low concentrations of cAMP can lead to activation of PKA without full dissociation of the subunits. By such means, they propose that PKA can decode temporally complex patterns in cAMP [159].

AC INHIBITORS IN THE LIGHT OF MORE THAN ONE AC PER CELL TYPE

Specific inhibitors of ACs still hold promise as therapeutic agents or as tools for dissecting the contribution of particular ACs to defined cellular events [160]. The problem with inhibitors, apart from their expected ‘off-target’ effects, when they are nucleotide or sugar derivatives, is magnified by the fact that no cell expresses only one AC, which exacerbates the problem of specificity. A recent screen of supposedly selective AC inhibitors against a full panel of ACs revealed little specificity among sugar-modified purine derivatives. A structure-based virtual screen identified a potential non-adenine-binding pocket in the catalytic site that showed promise for selective compounds [161]. The issue of access to the cells, and to the compartment where the ACs reside, magnifies this challenge enormously. In vitro comparisons just cannot anticipate in vivo results and the apparent efficacy of some of these inhibitors in intact cells may not even have to  c The Authors Journal compilation  c 2014 Biochemical Society

do with their action on ACs. In one example, the purportedly AC5-selective compound vidarabine was fully effective against residual AC activity in AC5-knockout mice [162]. The continued efficacy of vidarabine in AC5-knockout mice may reflect off-target (non-AC) effects as much as poor selectivity for AC subtypes. Although investigators acknowledge the possibility of off-target effects, little effort is made to exclude such effects at early stages of inhibitor development. Another insightful study of the widely used AC inhibitor SQ 22,536 shows that not only is it a potent inhibitor of AC activity, but it also inhibits the downstream effects of cAMP (or dibutyryl-cAMP) on PKA-mediated CREB (cAMP-response-element-binding protein) phosphorylation [163]. A recent study used an intact cell screening method for inhibitors against PMA-stimulated AC2 activity and later validated the efficacy of the optimal compounds against AC2 activity in in vitro assays [164]. Other approaches that target isoform-specific properties of the ACs such as the binding sites for CaM might also be of interest, though obviously ridden with an equal array of possibly off-target effects on other CaM-binding proteins. Given the problems associated with AC inhibitors, within the concept of the present review scaffold disruptors seem particularly worth exploring. Clearly, the involvement of AKAP complexes in numerous disease predispositions makes them viable and precise therapeutic targets as discussed recently [165]. Indeed, the first successful attempts at a small-molecule inhibitor of cAMPdirected PKA/AKAP disruptors in cardiac myocytes have been reported [166,167]. Drugs which aim to disrupt the AC–AKAP interaction may offer even greater precision.

FUTURE DIRECTIONS AND APPROACHES

The central role played by cAMP in physiological homoeostasis via its downstream targets, PKA, CNG channels and Epac has been appreciated for many years, nevertheless the sources

Adenylate cyclase-centred microdomains

of the cAMP, the ACs, offer even more insights into the elaborate devices used in these pathways. Specific ACs are now identified with distinct physiological processes and disease states, and unravelling their important contributions poses sizeable challenges, not least because most, if not all, cells express more than one AC, and the ACs can act as scaffolds. How or whether individual ACs with demonstrated susceptibility to indirect regulation, such as PKC, βγ subunits of G-proteins and tyrosine kinases, are regulated in cells which possess more than one AC within restricted cellular domains is a particular challenge. This issue may be most easily be pursued with targeted sensors and, of course, selective inhibitors. The mechanism of stimulation of AC8 by Ca2 + entry in non-excitable cells seems to be solved by the AC8–Orai1 complex. However, whether the other similarly discriminating Ca2 + -sensitive ACs (AC1 and AC5/AC6) form similar complexes remains to be determined. In excitable cells, the mode of interaction between all of the Ca2 + -sensitive ACs with L-type Ca2 + channels remains to be clarified. The full range of AC partners scaffolded by ACs, particularly in explicit physiological contexts, is unknown. Although a daunting challenge, this is the level of complexity at which the ACs operate. Inferences based on simple AC knockdowns are ambiguous; far more insight is likely from knockbacks with defined mutations in regions scaffolding particular partners. Measurements of the consequences for cAMP dynamics in microdomains upon the knockout of putative partners will also be insightful. The dynamic behaviour of ACs is another demanding question, calling for elaborate methodologies. Single-molecule analyses of ACs labelled with photoactivatable probes along with putative partners [168] could allow kinetic assessments of encounters, combined with high-resolution microscopic techniques [169,170]. Whether ACs signal internally still seems be an open question. The notion that internalized β-arrestin via ERK inactivates PDEs and thereby allows a sustained accumulation of cAMP seems the more likely interpretation of the experiments described to date. However, whether acceleration of AC internalization provides a means of allowing cAMP to originate from internal zones remains worth considering. ACs tagged with cAMP or Ca2 + sensors are a major opportunity for studying the enzymes in their own microenvironment and to address the intimate signalling dynamics of specific ACs among a medley of different ACs in a cell. Such experiments are challenging and, whereas some ACs tolerate N-terminal modifications, not all may; significant optimization, calibration and extensive controls will be needed. Nevertheless, if ACcentred microdomains are the base cAMP-signalling unit, the challenge may need to be embraced or alternative avenues into these domains devised [171,172]. The fact that mechanisms have been proposed that allow PKA to respond to cAMP oscillations should stimulate a major exploration of the possible role of cAMP oscillations. Techniques such as generating pulses of cAMP by flash photolysis of caged cAMP will be valuable. Even less disruptive would be to impose cAMP oscillations by controlled membrane depolarization in excitable cells to yield predictable Ca2 + oscillations linked to cAMP via Ca2 + -dependent ACs. Such experiments will be highly insightful in systems such as cardiomyocytes, islet β-cells and neurons. The development of live-cell sensors for cAMP (and Ca2 + ) has allowed significant maturing of the cAMP field. The theoretical postulate of cAMP compartments has materialized and the dynamic behaviour of cAMP in cellular spaces and over time has been demonstrated. The underpinnings of this behaviour in

209

the shape of AC-centred and sculpted microdomains is beginning to emerge, and this new awareness promises new routes to precise tools to manipulate this centrally important physiological regulator. ACKNOWLEDGEMENT

Space constraints dictated that this overview could not fully represent the work of the many investigators in the relevant areas. FUNDING

We thank the Newton Trust and the Gates Cambridge Trust for support. REFERENCES 1 Halls, M. L. and Cooper, D. M. F. (2011) Regulation by Ca2 + -signaling pathways of adenylyl cyclases. Cold Spring Harb. Perspect. Biol. 3, a004143 CrossRef PubMed 2 Sunahara, R. K., Dessauer, C. W. and Gilman, A. G. (1996) Complexity and diversity of mammalian adenylyl cyclases. Annu. Rev. Pharmacol. Toxicol. 36, 461–480 CrossRef PubMed 3 Sadana, R. and Dessauer, C. W. (2009) Physiological roles for G protein-regulated adenylyl cyclase isoforms: insights from knockout and overexpression studies. Neurosignals 17, 5–22 CrossRef PubMed 4 Ostrom, R. S., Bogard, A. S., Gros, R. and Feldman, R. D. (2012) Choreographing the adenylyl cyclase signalosome: sorting out the partners and the steps. Naunyn Schmiedebergs Arch. Pharmacol. 385, 5–12 CrossRef PubMed 5 Krupinski, J., Coussen, F., Bakalyar, H. A., Tang, W. J., Feinstein, P. G., Orth, K., Slaughter, C., Reed, R. R. and Gilman, A. G. (1989) Adenylyl cyclase amino acid sequence: possible channel- or transporter-like structure. Science 244, 1558–1564 CrossRef PubMed 6 Baillie, G. S. and Houslay, M. D. (2005) Arrestin times for compartmentalised cAMP signalling and phosphodiesterase-4 enzymes. Curr. Opin. Cell Biol. 17, 129–134 CrossRef PubMed 7 MacKenzie, S. J., Baillie, G. S., McPhee, I., Bolger, G. B. and Houslay, M. D. (2000) ERK2 mitogen-activated protein kinase binding, phosphorylation, and regulation of the PDE4D cAMP-specific phosphodiesterases: the involvement of COOH-terminal docking sites and NH2 -terminal UCR regions. J. Biol. Chem. 275, 16609–16617 CrossRef PubMed 8 Edwards, H. V., Christian, F. and Baillie, G. S. (2012) cAMP: novel concepts in compartmentalised signalling. Semin. Cell Dev. Biol. 23, 181–190 CrossRef PubMed 9 Willis, M. J. and Baillie, G. S. (2014) Arrestin-dependent localization of phosphodiesterases. Handb. Exp. Pharmacol. 219, 293–307 CrossRef PubMed 10 Cooper, D. M. F. (2003) Regulation and organization of adenylyl cyclases and cAMP. Biochem. J. 375, 517–529 CrossRef PubMed 11 Beazely, M. A. and Watts, V. J. (2006) Regulatory properties of adenylate cyclases type 5 and 6: a progress report. Eur. J. Pharmacol. 535, 1–12 CrossRef PubMed 12 Cooper, D. M. F., Mons, N. and Karpen, J. W. (1995) Adenylyl cyclases and the interaction between calcium and cAMP signalling. Nature 374, 421–424 CrossRef PubMed 13 Falke, J. J., Drake, S. K., Hazard, A. L. and Peersen, O. B. (1994) Molecular tuning of ion binding to calcium signaling proteins. Q. Rev. Biophys. 27, 219–290 CrossRef PubMed 14 Mou, T. C., Masada, N., Cooper, D. M. F. and Sprang, S. R. (2009) Structural basis for inhibition of mammalian adenylyl cyclase by calcium. Biochemistry 48, 3387–3397 CrossRef PubMed 15 Wu, Z., Wong, S. T. and Storms, D. R. (1993) Modification of the calcium and calmodulin sensitivity of the type I adenylyl cyclase by mutagenesis of its calmodulin binding domain. J. Biol. Chem. 268, 23766–23768 PubMed 16 Masada, N., Ciruela, A., Macdougall, D. A. and Cooper, D. M. F. (2009) Distinct mechanisms of regulation by Ca2 + /calmodulin of type 1 and 8 adenylyl cyclases support their different physiological roles. J. Biol. Chem. 284, 4451–4463 CrossRef PubMed 17 Catterall, W. A. (2011) Voltage-gated calcium channels. Cold Spring Harb. Perspect. Biol. 3, a003947 CrossRef PubMed 18 Mori, M. X., Erickson, M. G. and Yue, D. T. (2004) Functional stoichiometry and local enrichment of calmodulin interacting with Ca2 + channels. Science 304, 432–435 CrossRef PubMed 19 Tran, Q. K., Black, D. J. and Persechini, A. (2003) Intracellular coupling via limiting calmodulin. J. Biol. Chem. 278, 24247–24250 CrossRef PubMed  c The Authors Journal compilation  c 2014 Biochemical Society

210

D. M. F. Cooper and V. G. Tabbasum

20 Gu, C. and Cooper, D. M. F. (1999) Calmodulin-binding sites on adenylyl cyclase type VIII. J. Biol. Chem. 274, 8012–8021 CrossRef PubMed 21 Simpson, R. E., Ciruela, A. and Cooper, D. M. F. (2006) The role of calmodulin recruitment in Ca2 + -stimulation of adenylyl cyclase type 8. J. Biol. Chem. 281, 17379–17389 CrossRef PubMed 22 Macdougall, D. A., Wachten, S., Ciruela, A., Sinz, A. and Cooper, D. M. F. (2009) Separate elements within a single IQ-like motif in adenylyl cyclase type 8 impart Ca2 + /calmodulin binding and autoinhibition. J. Biol. Chem. 284, 15573–15588 CrossRef PubMed 23 Herbst, S., Masada, N., Pfennig, S., Ihling, C. H., Cooper, D. M. F. and Sinz, A. (2013) Structural insights into calmodulin/adenylyl cyclase 8 interaction. Anal. Bioanal. Chem. 405, 9333–9342 CrossRef PubMed 24 Masada, N., Schaks, S., Jackson, S. E., Sinz, A. and Cooper, D. M. F. (2012) Distinct mechanisms of calmodulin binding and regulation of adenylyl cyclases 1 and 8. Biochemistry 51, 7917–7929 CrossRef PubMed 25 Putney, Jr, J. W. (1991) The capacitative model for receptor-activated calcium entry. Adv. Pharmacol. 22, 251–269 CrossRef PubMed 26 Willoughby, D. and Cooper, D. M. F. (2007) Organization and Ca2 + regulation of adenylyl cyclases in cAMP microdomains. Physiol. Rev. 87, 965–1010 CrossRef PubMed 27 Kitaguchi, T., Oya, M., Wada, Y., Tsuboi, T. and Miyawaki, A. (2013) Extracellular calcium influx activates adenylate cyclase 1 and potentiates insulin secretion in MIN6 cells. Biochem. J. 450, 365–373 CrossRef PubMed 28 Kryukova, Y. N., Protas, L. and Robinson, R. B. (2012) Ca2 + -activated adenylyl cyclase 1 introduces Ca2 + -dependence to β-adrenergic stimulation of HCN2 current. J. Mol. Cell. Cardiol. 52, 1233–1239 CrossRef PubMed 29 Fagan, K. A., Graf, R. A., Tolman, S., Schaack, J. and Cooper, D. M. F. (2000) Regulation of a Ca2 + -sensitive adenylyl cyclase in an excitable cell: role of voltage-gated versus capacitative Ca2 + entry. J. Biol. Chem. 275, 40187–40194 CrossRef PubMed 30 Fagan, K. A., Rich, T. C., Tolman, S., Schaack, J., Karpen, J. W. and Cooper, D. M. F. (1999) Adenovirus-mediated expression of an olfactory cyclic nucleotide-gated channel regulates the endogenous Ca2 + -inhibitable adenylyl cyclase in C6-2B glioma cells. J. Biol. Chem. 274, 12445–12453 CrossRef PubMed 31 Everett, K. L. and Cooper, D. M. F. (2013) An improved targeted cAMP sensor to study the regulation of adenylyl cyclase 8 by Ca2 + entry through voltage-gated channels. PLoS ONE 8, e75942 CrossRef PubMed 32 Yu, H. J., Ma, H. and Green, R. D. (1993) Calcium entry via L-type calcium channels acts as a negative regulator of adenylyl cyclase activity and cyclic AMP levels in cardiac myocytes. Mol. Pharmacol. 44, 689–693 PubMed 33 Chetkovich, D. M., Gray, R., Johnston, D. and Sweatt, J. D. (1991) N -methyl-D-aspartate receptor activation increases cAMP levels and voltage-gated Ca2 + channel activity in area Ca1 of hippocampus. Proc. Natl. Acad. Sci. U.S.A. 88, 6467–6471 CrossRef PubMed 34 Dai, S., Hall, D. D. and Hell, J. W. (2009) Supramolecular assemblies and localized regulation of voltage-gated ion channels. Physiol. Rev. 89, 411–452 CrossRef PubMed 35 Lefkimmiatis, K., Srikanthan, M., Maiellaro, I., Moyer, M. P., Curci, S. and Hofer, A. M. (2009) Store-operated cyclic AMP signalling mediated by STIM1. Nat. Cell Biol. 11, 433–442 CrossRef PubMed 36 Maiellaro, I., Lefkimmiatis, K., Moyer, M. P., Curci, S. and Hofer, A. M. (2012) Termination and activation of store-operated cyclic AMP production. J. Cell. Mol. Med. 16, 2715–2725 CrossRef PubMed 37 Grigoriev, I., Gouveia, S. M., van der Vaart, B., Demmers, J., Smyth, J. T., Honnappa, S., Splinter, D., Steinmetz, M. O., Putney, Jr, J. W., Hoogenraad, C. C. and Akhmanova, A. (2008) STIM1 is a MT-plus-end-tracking protein involved in remodeling of the ER. Curr. Biol. 18, 177–182 CrossRef PubMed 38 Blau, L. and Weissmann, G. (1988) Transmembrane calcium movements mediated by ionomycin and phosphatidate in liposomes with Fura 2 entrapped. Biochemistry 27, 5661–5666 CrossRef PubMed 39 Tallini, Y. N., Ohkura, M., Choi, B. R., Ji, G., Imoto, K., Doran, R., Lee, J., Plan, P., Wilson, J., Xin, H. B. et al. (2006) Imaging cellular signals in the heart in vivo : cardiac expression of the high-signal Ca2 + indicator GCaMP2. Proc. Natl. Acad. Sci. U.S.A. 103, 4753–4758 CrossRef PubMed 40 Willoughby, D., Wachten, S., Masada, N. and Cooper, D. M. F. (2010) Direct demonstration of discrete Ca2 + microdomains associated with different isoforms of adenylyl cyclase. J. Cell Sci. 123, 107–117 CrossRef PubMed 41 Willoughby, D., Everett, K. L., Halls, M. L., Pacheco, J., Skroblin, P., Vaca, L., Klussmann, E. and Cooper, D. M. F. (2012) Direct binding between Orai1 and AC8 mediates dynamic interplay between Ca2 + and cAMPs. Sci. Signal. 5, ra29 PubMed 42 Tadross, M. R., Dick, I. E. and Yue, D. T. (2008) Mechanism of local and global Ca2 + sensing by calmodulin in complex with a Ca2 + channel. Cell 133, 1228–1240 CrossRef PubMed  c The Authors Journal compilation  c 2014 Biochemical Society

43 Alvarez, L., Dai, L., Friedrich, B. M., Kashikar, N. D., Gregor, I., Pascal, R. and Kaupp, U. B. (2012) The rate of change in Ca2 + concentration controls sperm chemotaxis. J. Cell Biol. 196, 653–663 CrossRef PubMed 44 Dessauer, C. W. (2009) Adenylyl cyclase–A-kinase anchoring protein complexes: the next dimension in cAMP signaling. Mol. Pharmacol. 76, 935–941 CrossRef PubMed 45 Rubin, C. S. (1994) A kinase anchor proteins and the intracellular targeting of signals carried by cyclic AMP. Biochim. Biophys. Acta 1224, 467–479 PubMed 46 Scott, J. D., Dessauer, C. W. and Task´en, K. (2012) Creating order from chaos: cellular regulation by kinase anchoring. Annu. Rev. Pharmacol. Toxicol. 53, 187–210 CrossRef PubMed 47 Bauman, A. L., Soughayer, J., Nguyen, B. T., Willoughby, D., Carnegie, G. K., Wong, W., Hoshi, N., Langeberg, L. K., Cooper, D. M. F., Dessauer, C. W. and Scott, J. D. (2006) Dynamic regulation of cAMP synthesis through anchored PKA–adenylyl cyclase V/VI complexes. Mol. Cell 23, 925–931 CrossRef PubMed 48 Willoughby, D., Masada, N., Wachten, S., Pagano, M., Halls, M. L., Everett, K. L., Ciruela, A. and Cooper, D. M. F. (2010) AKAP79/150 interacts with AC8 and regulates Ca2 + -dependent cAMP synthesis in pancreatic and neuronal systems. J. Biol. Chem. 285, 20328–20342 CrossRef PubMed 49 Kapiloff, M. S., Piggott, L. A., Sadana, R., Li, J., Heredia, L. A., Henson, E., Efendiev, R. and Dessauer, C. W. (2009) An adenylyl cyclase–mAKAPβ signaling complex regulates cAMP levels in cardiac myocytes. J. Biol. Chem. 284, 23540–23546 CrossRef PubMed 50 Shen, J. X. and Cooper, D. M. F. (2013) AKAP79, PKC, PKA and PDE4 participate in a Gq -linked muscarinic receptor and adenylate cyclase 2 cAMP signalling complex. Biochem. J. 455, 47–56 CrossRef PubMed 51 Li, Y., Chen, L., Kass, R. S. and Dessauer, C. W. (2012) 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 CrossRef PubMed 52 Halls, M. L. and Cooper, D. M. F. (2010) Sub-picomolar relaxin signalling by a pre-assembled RXFP1, AKAP79, AC2, β-arrestin 2, PDE4D3 complex. EMBO J. 29, 2772–2787 CrossRef PubMed 53 Willoughby, D., Halls, M. L., Everett, K. L., Ciruela, A., Skroblin, P., Klussmann, E. and Cooper, D. M. F. (2012) 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 CrossRef PubMed 54 Levin, L. R. and Reed, R. R. (1995) Identification of functional domains of adenylyl cyclase using in vivo chimeras. J. Biol. Chem. 270, 7573–7579 CrossRef PubMed 55 Wu, Z., Wong, S. T. and Storm, D. R. (1993) Modification of the calcium and calmodulin sensitivity of the type I adenylyl cyclase by mutagenesis of its calmodulin binding domain. J. Biol. Chem. 268, 23766–23768 PubMed 56 Vorherr, T., Knopfel, L., Hofmann, F., Mollner, S., Pfeuffer, T. and Carafoli, E. (1993) The calmodulin binding domain of nitric oxide synthase and adenylyl cyclase. Biochemistry 32, 6081–6088 CrossRef PubMed 57 Piggott, L. A., Bauman, A. L., Scott, J. D. and Dessauer, C. W. (2008) The A-kinase anchoring protein Yotiao binds and regulates adenylyl cyclase in brain. Proc. Natl. Acad. Sci. U.S.A. 105, 13835–13840 CrossRef PubMed 58 Efendiev, R., Samelson, B. K., Nguyen, B. T., Phatarpekar, P. V., Baameur, F., Scott, J. D. and Dessauer, C. W. (2010) AKAP79 interacts with multiple adenylyl cyclase (AC) isoforms and scaffolds AC5 and -6 to α-amino-3-hydroxyl-5-methyl-4isoxazole-propionate (AMPA) receptors. J. Biol. Chem. 285, 14450–14458 CrossRef PubMed 59 Wang, H. Y. and Burns, L. H. (2006) Gβγ that interacts with adenylyl cyclase in opioid tolerance originates from a Gs protein. J. Neurobiol. 66, 1302–1310 CrossRef PubMed 60 Schwencke, C., Yamamoto, M., Okumura, S., Toya, Y., Kim, S. J. and Ishikawa, Y. (1999) Compartmentation of cyclic adenosine 3 ,5 -monophosphate signaling in caveolae. Mol. Endocrinol. 13, 1061–1070 PubMed 61 Wang, Z., Balet Sindreu, C., Li, V., Nudelman, A., Chan, G. C. and Storm, D. R. (2006) Pheromone detection in male mice depends on signaling through the type 3 adenylyl cyclase in the main olfactory epithelium. J. Neurosci. 26, 7375–7379 CrossRef PubMed 62 Wang, S. C., Lai, H. L., Chiu, Y. T., Ou, R., Huang, C. L. and Chern, Y. (2007) Regulation of type V adenylyl cyclase by Ric8a, a guanine nucleotide exchange factor. Biochem. J. 406, 383–388 CrossRef PubMed 63 Scholich, K., Pierre, S. and Patel, T. B. (2001) Protein associated with Myc (PAM) is a potent inhibitor of adenylyl cyclases. J. Biol. Chem. 276, 47583–47589 CrossRef PubMed 64 Salim, S., Sinnarajah, S., Kehrl, J. H. and Dessauer, C. W. (2003) Identification of RGS2 and type V adenylyl cyclase interaction sites. J. Biol. Chem. 278, 15842–15849 CrossRef PubMed 65 Swaney, J. S., Patel, H. H., Yokoyama, U., Head, B. P., Roth, D. M. and Insel, P. A. (2006) Focal adhesions in (myo)fibroblasts scaffold adenylyl cyclase with phosphorylated caveolin. J. Biol. Chem. 281, 17173–17179 CrossRef PubMed

Adenylate cyclase-centred microdomains 66 Schwencke, C., Oka, N., Toya, Y. and Ishikawa, Y. (1997) Isoform-specific interaction of adenylyl cyclase with caveolin. FASEB J. 2, A323 67 Wu, C. S., Lin, J. T., Chien, C. L., Chang, W. C., Lai, H. L., Chang, C. P. and Chern, Y. (2011) Type VI adenylyl cyclase regulates neurite extension by binding to Snapin and Snap25. Mol. Cell. Biol. 31, 4874–4886 CrossRef PubMed 68 Chou, J. L., Huang, C. L., Lai, H. L., Hung, A. C., Chien, C. L., Kao, Y. Y. and Chern, Y. (2004) Regulation of type VI adenylyl cyclase by Snapin, a SNAP25-binding protein. J. Biol. Chem. 279, 46271–46279 CrossRef PubMed 69 Ayling, L. J., Briddon, S. J., Halls, M. L., Hammond, G. R., Vaca, L., Pacheco, J., Hill, S. J. and Cooper, D. M. F. (2012) Adenylyl cyclase AC8 directly controls its micro-environment by recruiting the actin cytoskeleton in a cholesterol-rich milieu. J. Cell Sci. 125, 869–886 CrossRef PubMed 70 Fonseca, S. G., Urano, F., Weir, G. C., Gromada, J. and Burcin, M. (2012) Wolfram syndrome 1 and adenylyl cyclase 8 interact at the plasma membrane to regulate insulin production and secretion. Nat. Cell Biol. 14, 1105–1112 CrossRef PubMed 71 Crossthwaite, A. J., Ciruela, A., Rayner, T. F. and Cooper, D. M. F. (2006) A direct interaction between the N terminus of adenylyl cyclase AC8 and the catalytic subunit of protein phosphatase 2A. Mol. Pharmacol. 69, 608–617 CrossRef PubMed 72 Herzog, F., Kahraman, A., Boehringer, D., Mak, R., Bracher, A., Walzthoeni, T., Leitner, A., Beck, M., Hartl, F. U., Ban, N. et al. (2012) Structural probing of a protein phosphatase 2A network by chemical cross-linking and mass spectrometry. Science 337, 1348–1352 CrossRef PubMed 73 Willoughby, D., Everett, K. L., Halls, M. L., Pacheco, J., Skroblin, P., Vaca, L., Klussmann, E. and Cooper, D. M. F. (2012) Direct binding between Orai1 and AC8 mediates dynamic interplay between Ca2 + and cAMP signaling. Sci. Signal. 5, ra29 PubMed 74 Smith, K. E., Gu, C., Fagan, K. A., Hu, B. and Cooper, D. M. F. (2002) Residence of adenylyl cyclase type 8 in caveolae is necessary but not sufficient for regulation by capacitative Ca2 + entry. J. Biol. Chem. 277, 6025–6031 CrossRef PubMed 75 Wang, S. C., Lin, J. T. and Chern, Y. (2009) Novel regulation of adenylyl cyclases by direct protein–protein interactions: insights from snapin and ric8a. Neurosignals 17, 169–180 CrossRef PubMed 76 Gao, M. H., Tang, T., Lai, N. C., Miyanohara, A., Guo, T., Tang, R., Firth, A. L., Yuan, J. X. and Hammond, H. K. (2011) Beneficial effects of adenylyl cyclase type 6 (AC6) expression persist using a catalytically inactive AC6 mutant. Mol. Pharmacol. 79, 381–388 CrossRef PubMed 77 Gao, M. H., Lai, N. C., Tang, T., Guo, T., Tang, R., Chun, B. J., Wang, H., Dalton, N. N., Suarez, J., Dillmann, W. H. and Hammond, H. K. (2013) Preserved cardiac function despite marked impairment of cAMP generation. PLoS ONE 8, e72151 CrossRef PubMed 78 Tang, T., Lai, N. C., Hammond, H. K., Roth, D. M., Yang, Y., Guo, T. and Gao, M. H. (2010) Adenylyl cyclase 6 deletion reduces left ventricular hypertrophy, dilation, dysfunction, and fibrosis in pressure-overloaded female mice. J. Am. Coll. Cardiol. 55, 1476–1486 CrossRef PubMed 79 Wang, H. and Storm, D. R. (2003) Calmodulin-regulated adenylyl cyclases: cross-talk and plasticity in the central nervous system. Mol. Pharmacol. 63, 463–468 CrossRef PubMed 80 Putney, J. W. (2013) Alternative forms of the store-operated calcium entry mediators, STIM1 and Orai1. Curr. Top. Membr. 71, 109–123 CrossRef PubMed 81 Rajendran, L. and Simons, K. (2005) Lipid rafts and membrane dynamics. J. Cell Sci. 118, 1099–1102 CrossRef PubMed 82 Ostrom, R. S. and Insel, P. A. (2004) The evolving role of lipid rafts and caveolae in G protein-coupled receptor signaling: implications for molecular pharmacology. Br. J. Pharmacol. 143, 235–245 CrossRef PubMed 83 Owen, D. M., Magenau, A., Williamson, D. and Gaus, K. (2012) The lipid raft hypothesis revisited: new insights on raft composition and function from super-resolution fluorescence microscopy. BioEssays 34, 739–747 CrossRef PubMed 84 Janosi, L., Li, Z., Hancock, J. F. and Gorfe, A. A. (2012) Organization, dynamics, and segregation of Ras nanoclusters in membrane domains. Proc. Natl. Acad. Sci. U.S.A. 109, 8097–8102 CrossRef PubMed 85 Foster, L. J., De Hoog, C. L. and Mann, M. (2003) Unbiased quantitative proteomics of lipid rafts reveals high specificity for signaling factors. Proc. Natl. Acad. Sci. U.S.A. 100, 5813–5818 CrossRef PubMed 86 Simons, K. and Toomre, D. (2000) Lipid rafts and signal transduction. Nat. Rev. Mol. Cell Biol. 1, 31–39 CrossRef PubMed 87 Jacobson, K., Mouritsen, O. G. and Anderson, R. G. (2007) Lipid rafts: at a crossroad between cell biology and physics. Nat. Cell Biol. 9, 7–14 CrossRef PubMed 88 Patel, H. H., Murray, F. and Insel, P. A. (2008) Caveolae as organizers of pharmacologically relevant signal transduction molecules. Annu. Rev. Pharmacol. Toxicol. 48, 359–391 CrossRef PubMed 89 Cooper, D. M. F. and Crossthwaite, A. J. (2006) Higher-order organization and regulation of adenylyl cyclases. Trends Pharmacol. Sci. 27, 426–431 CrossRef PubMed

211

90 Head, B. P., Patel, H. H., Roth, D. M., Murray, F., Swaney, J. S., Niesman, I. R., Farquhar, M. G. and Insel, P. A. (2006) Microtubules and actin microfilaments regulate lipid raft/caveolae localization of adenylyl cyclase signaling components. J. Biol. Chem. 281, 26391–26399 CrossRef PubMed 91 Pagano, M., Clynes, M. A., Masada, N., Ciruela, A., Ayling, L. J., Wachten, S. and Cooper, D. M. F. (2009) Insights into the residence in lipid rafts of adenylyl cyclase AC8 and its regulation by capacitative calcium entry. Am. J. Physiol. Cell Physiol. 296, C607–C619 CrossRef PubMed 92 Crossthwaite, A. J., Seebacher, T., Masada, N., Ciruela, A., Dufraux, K., Schultz, J. E. and Cooper, D. M. F. (2005) The cytosolic domains of Ca2 + -sensitive adenylyl cyclases dictate their targeting to plasma membrane lipid rafts. J. Biol. Chem. 280, 6380–6391 CrossRef PubMed 93 Thangavel, M., Liu, X., Sun, S. Q., Kaminsky, J. and Ostrom, R. S. (2009) The C1 and C2 domains target human type 6 adenylyl cyclase to lipid rafts and caveolae. Cell. Signal. 21, 301–308 CrossRef PubMed 94 Delint-Ramirez, I., Willoughby, D., Hammond, G. V., Ayling, L. J. and Cooper, D. M. F. (2011) Palmitoylation targets AKAP79 protein to lipid rafts and promotes its regulation of calcium-sensitive adenylyl cyclase type 8. J. Biol. Chem. 286, 32962–32975 CrossRef PubMed 95 Keith, D. J., Sanderson, J. L., Gibson, E. S., Woolfrey, K. M., Robertson, H. R., Olszewski, K., Kang, R., El-Husseini, A. and Dell’acqua, M. L. (2012) Palmitoylation of A-kinase anchoring protein 79/150 regulates dendritic endosomal targeting and synaptic plasticity mechanisms. J. Neurosci. 32, 7119–7136 CrossRef PubMed 96 Pani, B., Ong, H. L., Liu, X., Rauser, K., Ambudkar, I. S. and Singh, B. B. (2008) Lipid rafts determine clustering of STIM1 in endoplasmic reticulum-plasma membrane junctions and regulation of store-operated Ca2 + entry (SOCE). J. Biol. Chem. 283, 17333–17340 CrossRef PubMed 97 Martin, A. C., Willoughby, D., Ciruela, A., Ayling, L. J., Pagano, M., Wachten, S., Tengholm, A. and Cooper, D. M. F. (2009) Capacitative Ca2 + entry via Orai1 and stromal interacting molecule 1 (STIM1) regulates adenylyl cyclase type 8. Mol. Pharmacol. 75, 830–842 CrossRef PubMed 98 Brady, J. D., Rich, T. C., Le, X., Stafford, K., Fowler, C. J., Lynch, L., Karpen, J. W., Brown, R. L. and Martens, J. R. (2004) Functional role of lipid raft microdomains in cyclic nucleotide-gated channel activation. Mol. Pharmacol. 65, 503–511 CrossRef PubMed 99 Balijepalli, R. C., Foell, J. D., Hall, D. D., Hell, J. W. and Kamp, T. J. (2006) Localization of cardiac L-type Ca2 + channels to a caveolar macromolecular signaling complex is required for β 2 -adrenergic regulation. Proc. Natl. Acad. Sci. U.S.A. 103, 7500–7505 CrossRef PubMed 100 Parton, R. G. and del Pozo, M. A. (2013) Caveolae as plasma membrane sensors, protectors and organizers. Nat. Rev. Mol. Cell Biol. 14, 98–112 CrossRef PubMed 101 Helms, J. B. and Zurzolo, C. (2004) Lipids as targeting signals: lipid rafts and intracellular trafficking. Traffic 5, 247–254 CrossRef PubMed 102 Simons, K. and Ehehalt, R. (2002) Cholesterol, lipid rafts, and disease. J. Clin. Invest. 110, 597–603 CrossRef PubMed 103 Ikonen, E. (2001) Roles of lipid rafts in membrane transport. Curr. Opin. Cell Biol. 13, 470–477 CrossRef PubMed 104 Ikonen, E. and Simons, K. (1998) Protein and lipid sorting from the trans -Golgi network to the plasma membrane in polarized cells. Semin. Cell Dev. Biol. 9, 503–509 CrossRef PubMed 105 Mundy, D. I., Machleidt, T., Ying, Y. S., Anderson, R. G. and Bloom, G. S. (2002) Dual control of caveolar membrane traffic by microtubules and the actin cytoskeleton. J. Cell Sci. 115, 4327–4339 CrossRef PubMed 106 Kusumi, A. and Sako, Y. (1996) Cell surface organization by the membrane skeleton. Curr. Opin. Cell Biol. 8, 566–574 CrossRef PubMed 107 Muriel, O., Echarri, A., Hellriegel, C., Pavon, D. M., Beccari, L. and Del Pozo, M. A. (2011) Phosphorylated filamin A regulates actin-linked caveolae dynamics. J. Cell Sci. 124, 2763–2776 CrossRef PubMed 108 Sayner, S. L., Balczon, R., Frank, D. W., Cooper, D. M. F. and Stevens, T. (2011) Filamin A is a phosphorylation target of membrane but not cytosolic adenylyl cyclase activity. Am. J. Physiol. Lung Cell. Mol. Physiol. 301, L117–L124 CrossRef PubMed 109 Rybin, V. O., Xu, X., Lisanti, M. P. and Steinberg, S. F. (2000) Differential targeting of β-adrenergic receptor subtypes and adenylyl cyclase to cardiomyocyte caveolae: a mechanism to functionally regulate the cAMP signaling pathway. J. Biol. Chem. 275, 41447–41457 CrossRef PubMed 110 Gabay, M., Pinter, M. E., Wright, F. A., Chan, P., Murphy, A. J., Valenzuela, D. M., Yancopoulos, G. D. and Tall, G. G. (2011) Ric-8 proteins are molecular chaperones that direct nascent G protein α subunit membrane association. Sci. Signal. 4, ra79 PubMed 111 Ye, X. and Cai, Q. (2014) Snapin-mediated BACE1 retrograde transport is essential for its degradation in lysosomes and regulation of APP processing in neurons. Cell Rep. 6, 24–31 CrossRef PubMed  c The Authors Journal compilation  c 2014 Biochemical Society

212

D. M. F. Cooper and V. G. Tabbasum

112 Feron, O., Michel, J. B., Sase, K. and Michel, T. (1998) Dynamic regulation of endothelial nitric oxide synthase: complementary roles of dual acylation and caveolin interactions. Biochemistry 37, 193–200 CrossRef PubMed 113 Collins, B. M., Davis, M. J., Hancock, J. F. and Parton, R. G. (2012) Structure-based reassessment of the caveolin signaling model: do caveolae regulate signaling through caveolin–protein interactions? Dev. Cell 23, 11–20 CrossRef PubMed 114 Byrne, D. P., Dart, C. and Rigden, D. J. (2012) Evaluating caveolin interactions: do proteins interact with the caveolin scaffolding domain through a widespread aromatic residue-rich motif? PLoS ONE 7, e44879 CrossRef PubMed 115 Ostrom, R. S., Bundey, R. A. and Insel, P. A. (2004) Nitric oxide inhibition of adenylyl cyclase type 6 activity is dependent upon lipid rafts and caveolin signaling complexes. J. Biol. Chem. 279, 19846–19853 CrossRef PubMed 116 Gu, C., Cali, J. J. and Cooper, D. M. F. (2002) Dimerization of mammalian adenylate cyclases. Eur. J. Biochem. 269, 413–421 CrossRef PubMed 117 Gu, C., Sorkin, A. and Cooper, D. M. F. (2001) Persistent interactions between the two transmembrane clusters dictate the targeting and functional assembly of adenylyl cyclase. Curr. Biol. 11, 185–190 CrossRef PubMed 118 Wehrle-Haller, B. (2012) Assembly and disassembly of cell matrix adhesions. Curr. Opin. Cell Biol. 24, 569–581 CrossRef PubMed 119 Howe, A. K. (2011) Cross-talk between calcium and protein kinase A in the regulation of cell migration. Curr. Opin. Cell Biol. 23, 554–561 CrossRef PubMed 120 Tojima, T., Hines, J. H., Henley, J. R. and Kamiguchi, H. (2011) Second messengers and membrane trafficking direct and organize growth cone steering. Nat. Rev. Neurosci. 12, 191–203 CrossRef PubMed 121 Norambuena, A. and Schwartz, M. A. (2011) Effects of integrin-mediated cell adhesion on plasma membrane lipid raft components and signaling. Mol. Biol. Cell 22, 3456–3464 CrossRef PubMed 122 Logue, J. S., Whiting, J. L., Tunquist, B., Sacks, D. B., Langeberg, L. K., Wordeman, L. and Scott, J. D. (2011) AKAP220 protein organizes signaling elements that impact cell migration. J. Biol. Chem. 286, 39269–39281 CrossRef PubMed 123 Schafer, C., Rymarczyk, G., Ding, L., Kirber, M. T. and Bolotina, V. M. (2012) Role of molecular determinants of store-operated Ca2 + entry (Orai1, phospholipase A2 group 6, and STIM1) in focal adhesion formation and cell migration. J. Biol. Chem. 287, 40745–40757 CrossRef PubMed 124 Fagan, K. A., Smith, K. E. and Cooper, D. M. F. (2000) Regulation of the Ca2 + -inhibitable adenylyl cyclase type VI by capacitative Ca2 + entry requires localization in cholesterol-rich domains. J. Biol. Chem. 275, 26530–26537 CrossRef PubMed 125 Geiger, B., Spatz, J. P. and Bershadsky, A. D. (2009) Environmental sensing through focal adhesions. Nat. Rev. Mol. Cell Biol. 10, 21–33 CrossRef PubMed 126 Harburger, D. S. and Calderwood, D. A. (2009) Integrin signalling at a glance. J. Cell Sci. 122, 159–163 CrossRef PubMed 127 Giannone, G., Ronde, P., Gaire, M., Beaudouin, J., Haiech, J., Ellenberg, J. and Takeda, K. (2004) Calcium rises locally trigger focal adhesion disassembly and enhance residency of focal adhesion kinase at focal adhesions. J. Biol. Chem. 279, 28715–28723 CrossRef PubMed 128 Howe, A. K. (2004) Regulation of actin-based cell migration by cAMP/PKA. Biochim. Biophys. Acta 1692, 159–174 CrossRef PubMed 129 Houslay, M. D. and Milligan, G. (1997) Tailoring cAMP-signalling responses through isoform multiplicity. Trends Biochem. Sci. 22, 217–224 CrossRef PubMed 130 Calebiro, D., Nikolaev, V. O., Gagliani, M. C., de Filippis, T., Dees, C., Tacchetti, C., Persani, L. and Lohse, M. J. (2009) Persistent cAMP-signals triggered by internalized G-protein-coupled receptors. PLoS Biol. 7, e1000172 CrossRef PubMed 131 Neumann, S., Geras-Raaka, E., Marcus-Samuels, B. and Gershengorn, M. C. (2010) Persistent cAMP signaling by thyrotropin (TSH) receptors is not dependent on internalization. FASEB J. 24, 3992–3999 CrossRef PubMed 132 De Arcangelis, V., Liu, S., Zhang, D., Soto, D. and Xiang, Y. K. (2010) Equilibrium between adenylyl cyclase and phosphodiesterase patterns adrenergic agonist dose-dependent spatiotemporal cAMP/protein kinase A activities in cardiomyocytes. Mol. Pharmacol. 78, 340–349 CrossRef PubMed 133 Garbers, D. L. (1981) The elevation of cyclic AMP concentrations in flagella-less sea urchin sperm heads. J. Biol. Chem. 256, 620–624 PubMed 134 Stengel, D. and Hanoune, J. (1984) The sperm adenylate cyclase. Ann. N.Y. Acad. Sci. 438, 18–28 CrossRef PubMed 135 Zippin, J. H., Chen, Y., Straub, S. G., Hess, K. C., Diaz, A., Lee, D., Tso, P., Holz, G. G., Sharp, G. W., Levin, L. R. and Buck, J. (2013) CO2 /HCO3 − - and calcium-regulated soluble adenylyl cyclase as a physiological ATP sensor. J. Biol. Chem. 288, 33283–33291 CrossRef PubMed 136 Esposito, G., Jaiswal, B. S., Xie, F., Krajnc-Franken, M. A., Robben, T. J., Strik, A. M., Kuil, C., Philipsen, R. L., van Duin, M., Conti, M. and Gossen, J. A. (2004) Mice deficient for soluble adenylyl cyclase are infertile because of a severe sperm-motility defect. Proc. Natl. Acad. Sci. U.S.A. 101, 2993–2998 CrossRef PubMed  c The Authors Journal compilation  c 2014 Biochemical Society

137 Tian, G., Sandler, S., Gylfe, E. and Tengholm, A. (2011) Glucose- and hormone-induced cAMP oscillations in α- and β-cells within intact pancreatic islets. Diabetes 60, 1535–1543 CrossRef PubMed 138 Kramer, R. H. (1990) Patch cramming: monitoring intracellular messengers in intact cells with membrane patches containing detector ion channels. Neuron 4, 335–341 CrossRef PubMed 139 Adams, S. R., Harootunian, A. T., Buechler, Y. J., Taylor, S. S. and Tsien, R. Y. (1991) Fluorescence ratio imaging of cyclic AMP in single cells. Nature 349, 694–697 CrossRef PubMed 140 Nikolaev, V. O., Bunemann, M., Hein, L., Hannawacker, A. and Lohse, M. J. (2004) Novel single chain cAMP sensors for receptor-induced signal propagation. J. Biol. Chem. 279, 37215–37218 CrossRef PubMed 141 Ponsioen, B., Zhao, J., Riedl, J., Zwartkruis, F., van der Krogt, G., Zaccolo, M., Moolenaar, W. H., Bos, J. L. and Jalink, K. (2004) Detecting cAMP-induced Epac activation by fluorescence resonance energy transfer: Epac as a novel cAMP indicator. EMBO Rep. 5, 1176–1180 CrossRef PubMed 142 DiPilato, L. M., Cheng, X. and Zhang, J. (2004) Fluorescent indicators of cAMP and Epac activation reveal differential dynamics of cAMP signaling within discrete subcellular compartments. Proc. Natl. Acad. Sci. U.S.A. 101, 16513–16518 CrossRef PubMed 143 Wachten, S., Masada, N., Ayling, L. J., Ciruela, A., Nikolaev, V. O., Lohse, M. J. and Cooper, D. M. F. (2010) Distinct pools of cAMP centre on different isoforms of adenylyl cyclase in pituitary-derived GH3 B6 cells. J. Cell Sci. 123, 95–106 CrossRef PubMed 144 Buxton, I. L. and Brunton, L. L. (1983) Compartments of cyclic AMP and protein kinase in mammalian cardiomyocytes. J. Biol. Chem. 258, 10233–10239 PubMed 145 Allbritton, N. L., Meyer, T. and Stryer, L. (1992) Range of messenger action of calcium ion and inositol 1,4,5-trisphosphate. Science 258, 1812–1815 CrossRef PubMed 146 Tsien, R. Y., Pozzan, T. and Rink, T. J. (1982) Calcium homeostasis in intact lymphocytes: cytoplasmic free calcium monitored with a new, intracellularly trapped fluorescent indicator. J. Cell Biol. 94, 325–334 CrossRef PubMed 147 Zaccolo, M. and Pozzan, T. (2002) Discrete microdomains with high concentration of cAMP in stimulated rat neonatal cardiac myocytes. Science 295, 1711–1715 CrossRef PubMed 148 Rich, T. C., Fagan, K. A., Nakata, H., Schaack, J., Cooper, D. M. F. and Karpen, J. W. (2000) Cyclic nucleotide-gated channels colocalize with adenylyl cyclase in regions of restricted cAMP diffusion. J. Gen. Physiol. 116, 147–161 CrossRef PubMed 149 Berridge, M. J. and Dupont, G. (1994) Spatial and temporal signalling by calcium. Curr. Opin. Cell Biol. 6, 267–274 CrossRef PubMed 150 Dupont, G., Houart, G. and De Koninck, P. (2003) Sensitivity of CaM kinase II to the frequency of Ca2 + oscillations: a simple model. Cell Calcium 34, 485–497 CrossRef PubMed 151 De Koninck, P. and Schulman, H. (1998) Sensitivity of CaM kinase II to the frequency of Ca2 + oscillations. Science 279, 227–230 CrossRef PubMed 152 Dolmetsch, R. E., Lewis, R. S., Goodnow, C. C. and Healy, J. I. (1997) Differential activation of transcription factors induced by Ca2 + response amplitude and duration. Nature 386, 855–858 CrossRef PubMed 153 Brooker, G. (1973) Oscillation of cyclic adenosine monophosphate concentration during the myocardial contraction cycle. Science 182, 933–934 CrossRef PubMed 154 Gorbunova, Y. V. and Spitzer, N. C. (2002) Dynamic interactions of cyclic AMP transients and spontaneous Ca2 + spikes. Nature 418, 93–96 CrossRef PubMed 155 Nicol, X., Hong, K. P. and Spitzer, N. C. (2011) Spatial and temporal second messenger codes for growth cone turning. Proc. Natl. Acad. Sci. U.S.A. 108, 13776–13781 CrossRef PubMed 156 Dyachok, O., Isakov, Y., Sagetorp, J. and Tengholm, A. (2006) Oscillations of cyclic AMP in hormone-stimulated insulin-secreting β-cells. Nature 439, 349–352 CrossRef PubMed 157 Willoughby, D. and Cooper, D. M. F. (2006) Ca2 + stimulation of adenylyl cyclase generates dynamic oscillations in cyclic AMP. J. Cell Sci. 119, 828–836 CrossRef PubMed 158 Nicol, X., Voyatzis, S., Muzerelle, A., Narboux-Neme, N., Sudhof, T. C., Miles, R. and Gaspar, P. (2007) cAMP oscillations and retinal activity are permissive for ephrin signaling during the establishment of the retinotopic map. Nat. Neurosci. 10, 340–347 CrossRef PubMed 159 Taylor, S. S., Ilouz, R., Zhang, P. and Kornev, A. P. (2013) Assembly of allosteric macromolecular switches: lessons from PKA. Nat. Rev. Mol. Cell Biol. 13, 646–658 CrossRef 160 Pierre, S., Eschenhagen, T., Geisslinger, G. and Scholich, K. (2009) Capturing adenylyl cyclases as potential drug targets. Nat. Rev. Drug Discov. 8, 321–335 CrossRef PubMed

Adenylate cyclase-centred microdomains 161 Brand, C. S., Hocker, H. J., Gorfe, A. A., Cavasotto, C. N. and Dessauer, C. W. (2013) Isoform selectivity of adenylyl cyclase inhibitors: characterization of known and novel compounds. J. Pharmacol. Exp. Ther. 347, 265–275 CrossRef PubMed 162 Braeunig, J. H., Schweda, F., Han, P. L. and Seifert, R. (2013) Similarly potent inhibition of adenylyl cyclase by P-site inhibitors in hearts from wild type and AC5 knockout mice. PLoS ONE 8, e68009 CrossRef PubMed 163 Emery, A. C., Eiden, M. V. and Eiden, L. E. (2013) A new site and mechanism of action for the widely used adenylate cyclase inhibitor SQ22,536. Mol. Pharmacol. 83, 95–105 CrossRef PubMed 164 Conley, J. M., Brand, C. S., Bogard, A. S., Pratt, E. P., Xu, R., Hockerman, G. H., Ostrom, R. S., Dessauer, C. W. and Watts, V. J. (2013) 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 CrossRef PubMed 165 Esseltine, J. L. and Scott, J. D. (2013) AKAP signaling complexes: pointing towards the next generation of therapeutic targets? Trends Pharmacol. Sci. 34, 648–655 CrossRef PubMed 166 Sch¨afer, G., Mili´c, J., Eldahshan, A., G¨otz, F., Z¨uhlke, K., Schillinger, C., Kreuchwig, A., Elkins, J. M., Abdul Azeez, K. R., Oder, A. et al. (2013) Highly functionalized terpyridines as competitive inhibitors of AKAP–PKA interactions. Angew. Chem. Int. Ed. Engl. 52, 12187–12191 CrossRef PubMed

213

167 Christian, F., Szasz´ak, M., Friedl, S., Drewianka, S., Lorenz, D., Goncalves, A., Furkert, J., Vargas, C., Schmieder, P., G¨otz, F. et al. (2012) Small molecule AKAP-protein kinase A (PKA) interaction disruptors that activate PKA interfere with compartmentalized cAMP signaling in cardiac myocytes. J. Biol. Chem. 286, 9079–9096 CrossRef 168 Groc, L., Lafourcade, M., Heine, M., Renner, M., Racine, V., Sibarita, J. B., Lounis, B., Choquet, D. and Cognet, L. (2007) Surface trafficking of neurotransmitter receptor: comparison between single-molecule/quantum dot strategies. J. Neurosci. 27, 12433–12437 CrossRef PubMed 169 Kanchanawong, P., Shtengel, G., Pasapera, A. M., Ramko, E. B., Davidson, M. W., Hess, H. F. and Waterman, C. M. Nanoscale architecture of integrin-based cell adhesions. Nature 468, 580–584 CrossRef 170 Brameshuber, M. and Schutz, G. J. Detection and quantification of biomolecular association in living cells using single-molecule microscopy. Methods Enzymol. 505, 159–186 CrossRef 171 Dunn, T. A., Wang, C. T., Colicos, M. A., Zaccolo, M., DiPilato, L. M., Zhang, J., Tsien, R. Y. and Feller, M. B. (2006) Imaging of cAMP levels and protein kinase A activity reveals that retinal waves drive oscillations in second-messenger cascades. J. Neurosci. 26, 12807–12815 CrossRef PubMed 172 Zhang, J., Bal, M., Bierbower, S., Zaika, O. and Shapiro, M. S. (2011) AKAP79/150 signal complexes in G-protein modulation of neuronal ion channels. J. Neurosci. 31, 7199–7211 CrossRef PubMed

Received 30 April 2014/10 June 2014; accepted 24 June 2014 Published on the Internet 7 August 2014, doi:10.1042/BJ20140560

 c The Authors Journal compilation  c 2014 Biochemical Society

Adenylate cyclase-centred microdomains.

Recent advances in the AC (adenylate cyclase)/cAMP field reveal overarching roles for the ACs. Whereas few processes are unaffected by cAMP in eukaryo...
1MB Sizes 1 Downloads 8 Views