G protein-coupled receptors (GPCRs) are the largest family of cell surface receptors, mediating essential physiological responses through diverse intracellular signaling pathways. When coupled to Gs or Gi proteins, GPCR modulates the synthesis of 3′-5′-cyclic adenosine monophosphate (cAMP), which governs a wide array of processes, ranging from cellular growth and survival to metabolic regulation. Studies have highlighted that cAMP is not uniformly distributed within cells but instead is compartmentalized into highly localized nanodomains. These nanodomains, mostly regulated by phosphodiesterases (PDEs), play a critical role in enabling signal precision and functional effects that are specific to individual stimuli. GPCRs can initiate distinct cAMP responses based on their localization within the cell, with evidence showing that both receptors resident at the plasma membrane and intracellular receptors—including endosomal, Golgi, and nuclear GPCRs—elicit unique cAMP signaling profiles. This review examines the mechanisms underlying GPCR signaling through cAMP nanodomains. We focus on the role of PDE-mediated cAMP degradation in shaping local cAMP signals, the emerging views on mechanisms that may contribute to signal compartmentalization, and the role of intracellular membrane compartments. By exploring these aspects, we aim to highlight the complexity of GPCR signaling networks and illustrate some of the implications for the regulation of cellular function.

Membrane receptors and their downstream signaling pathways are integral to the regulation of nearly all biological functions in multicellular organisms. Among these, G protein-coupled receptors (GPCRs) represent the largest class. GPCRs are seven transmembrane receptors. The human genome encodes more than 800 GPCRs [1], underscoring their significance in regulating physiological processes. These receptors respond to diverse signals, including sensory inputs, amines, peptides, proteins, lipids, and nucleotides, reflecting their evolutionary adaptability [2]. GPCR transduces the signal input to the intracellular machinery through second messengers, including cAMP and calcium [2], [3]. To initiate signaling, GPCRs interact with heterotrimeric G proteins, intracellular transducers, which consist of three subunits: Gα, Gβ, and Gγ [4]. G proteins act as molecular switches and, upon agonist binding to a GPCR, the receptor facilitates the exchange of GDP to GTP on the Gα subunit, triggering its dissociation from the Gβγ dimer [5], [4], [6]. This separation generates two active signaling entities: the Gα subunit and the Gβγ complex. The 16 Gα subunits are categorized into four main families: Gs, Gi/o, Gq/11, and G12/13. The Gs and Gi families are known for their opposing regulation of adenylyl cyclase (AC), respectively, stimulating and inhibiting the ability of this enzyme to synthesize cAMP (3′-5′-cyclic adenosine monophosphate) from ATP (Adenosine triphosphate) [3], [7], [8]. The Gq family’s primary effector is phospholipase C, which cleaves phosphatidylinositol 4,5-bisphosphate (PIP2) into inositol 1,4,5-trisphosphate (IP3) and diacylglycerol. This reaction leads to calcium release from the endoplasmic reticulum (ER) and the activation of protein kinase C (PKC). Meanwhile, the G12/13 family predominantly signals through the small G protein Rho, which regulates cytoskeletal dynamics and cell motility [7]. cAMP, the first second messenger identified [9], is a small, hydrophilic molecule present across a wide range of living organisms. A large superfamily of phosphodiesterases (PDEs) hydrolyzes the 3′ phosphate bond of cAMP, converting it into 5′-AMP [10]. This hydrolysis by PDEs is the primary mechanism that terminates cAMP signaling [10]. However, efflux of cAMP into the extracellular space through multidrug-resistance protein (gene name ABCC4) transporters has also been described as a process that reduces intracellular cAMP levels [11]. cAMP exerts its effects by initiating a cascade of signaling events that affect the cellular function through interactions with a limited set of effector proteins, including cAMP-dependent protein kinase A (PKA) [12], exchange proteins activated by cAMP [13], cyclic nucleotide-gated ion channels [14], hyperpolarization-activated cyclic nucleotide-gated channels, and Popeye domain-containing proteins [15]. The most extensively studied downstream effector of cAMP is PKA, a highly promiscuous tetrameric enzyme consisting of two regulatory (R) and two catalytic (C) subunits. Upon cAMP binding to the R subunits, the C subunits are activated to phosphorylate a large variety of target proteins, thereby regulating a vast array of diverse cellular functions [16], [17], [18], [19]. Given that a single cell can express up to 100 different GPCRs, many of which signal via cAMP [20], cells face the challenge of accurately decoding multiple inputs from various GPCRs to ensure specificity in downstream signaling, despite relying on a limited number of effectors [21]. A key mechanism that allows cells to solve this problem is spatial segregation through the organization of cAMP nanodomains located within different subcellular compartments. Within these compartments, A-kinase anchoring proteins (AKAPs) organize multiprotein complexes, or signalosomes, which localize PKA in proximity to specific phosphorylation targets. AKAPs can also anchor other signaling components, including PDEs and phosphatases, enabling precise regulation of the amplitude, duration, and functional outcomes of the cAMP signal [22]. Below, we discuss our current understanding of GPCR signaling via cAMP nanodomains while focusing on the established role played by local degradation of cAMP by PDEs and on emerging modalities that may contribute to cAMP compartmentalization.

GPCR and cAMP

Following GPCR activation, cAMP signaling is primarily initiated at the plasma membrane through the activation of Gα proteins [23]. Notably, numerous receptors can couple to the same Gα protein, and individual receptors can couple to multiple Gα proteins, highlighting the complexity of GPCR signaling [24]. GPCRs are frequently organized into specialized micro- or nanodomains [25], where cytoskeletal scaffold proteins create nanoscopic signaling hotspots that ensure efficient engagement of transducer proteins. These specialized micro- or nanodomains are partially composed of actin fibers, microtubules, and elements of the cytoskeleton, including clathrin-coated pits [26]. Together, these structures form concentrated hotspots on the plasma membrane where receptors and G proteins colocalize. Such organization has been demonstrated with imaging studies using the nanobody Nb37 in CHO[MZ1] (derived from epithelial cells of the Chinese hamster ovary) cells [26]. Nb37 binds to the surface of the alpha-helical (AH) domain of Gαs, where the AH domain is accessible only in the nucleotide-free form of Gαs, representing the catalytic intermediate of G protein activation. In detail, structural studies of Gαs and other Gα subunits have revealed that they consist of two distinct domains: the Ras-like domain (RD)—also known as the GTPase domain—and an AH domain, which is composed of six α-helices embedded within the RD. The RD shares homology with monomeric G-proteins such as p21ras and is a conserved feature across the GTPase superfamily. In contrast, the AH is a unique structural element found exclusively in heterotrimeric G-proteins. Structurally, AH and RD together form a cleft that serves as the binding site for the nucleotide [27]. Nb37 can also be used as a domain marker, allowing tracking of the variable positioning of the small AH region of Gαs [28]. In this study, the α2A-adrenergic receptor (α2A-AR), which strongly couples with the inhibitory G protein (Gi), and a pertussis toxin-insensitive Gαi1 construct were labeled with two distinct organic fluorophores via SNAP5 and CLIP6 tags, respectively. These constructs were transiently expressed in CHO using fast two-color single-molecule microscopy combined with single-particle tracking. The analysis revealed a complex dynamic scenario at the plasma membrane, characterized by high-potential regions (areas where the molecules can easily leave or transition) from which α2A-ARs and Gαi subunits rapidly exited and low-potential regions where they tended to remain trapped. Notably, receptor–G protein interactions occurred preferentially within these shared low-energy areas (regions of the membrane where the energy is minimized, promoting stable and efficient receptor–G protein interactions), or ‘hot spots’ [26]. The organization of GPCR and G protein in hot spots allows direct activation of transducers and subsequent amplification of intracellular signals by precisely co-ordinating a variety of effector proteins and second messengers. For instance, when GPCRs are stimulated at low agonist concentrations, cAMP-rich nanodomains form around the receptors. The spatial confinement of cAMP within these domains is regulated by the cAMP degrading PDEs, as when the PDEs are inhibited, the boundaries between these domains are disrupted [29], [30]. Additionally, scaffolding proteins also contribute to maintaining the spatial precision of GPCR signaling by organizing effector proteins into specialized complexes [25]. This organization facilitates the integration, co-ordination, and processing of cellular signaling events [31]. For example, AKAP5, found to be mostly in association with the plasma membrane, displays the ability to associate with the β2-adrenergic receptor [32]. AKAP5 has been shown to interact with PKA holoenzymes containing RIIα or RIIβ, PKC, calmodulin, calcineurin phosphatase (PP2B), AC type V/VI, L-type calcium channels, and β-adrenergic receptors [33]. AKAP5 primarily functions by binding kinases like PKA and phosphatases such as calcineurin to regulate cAMP signaling effectively [34]. Through this, AKAP5 modulates a wide range of cellular processes, including neuronal and glial cell development, differentiation, and migration; the functioning of neural circuits involved in information processing; and the permeability of the blood–brain barrier [35], where disruption of AKAP interactions or dysregulated expression levels has been shown to contribute to the accelerated pathological progression of ischemic stroke [35]. AKAP5 interaction with phosphatase2B (PP2B) and PKA integrates the cAMP oscillations with Ca²+ signaling in pancreatic β-cells, thereby regulating glucose-stimulated insulin secretion (GSIS) [36,37]. GSIS involves cell membrane depolarization and the influx of Ca²+ through voltage-gated calcium channels [38]. The deletion of AKAP5, or more specifically the disruption of its interaction with PP2B, impairs L-type Ca²+ currents and reduces the cytoplasmic accumulation of Ca²+ and cAMP in β-cells, inhibiting GSIS [36].

PDEs and cAMP

The cAMP signal is terminated predominantly via its hydrolysis by PDEs, whose enzymatic activity was initially used to demonstrate the physiological significance of this cyclic nucleotide [39]. The mammalian PDEs are categorized into 11 families, encoded by 21 distinct genes, each exhibiting multiple splice variants with varying subcellular localizations [10,40]. PDE4, PDE7, and PDE8 selectively hydrolyze cAMP, and PDE1, PDE2, PDE3 PDE10, and PDE11 hydrolyze both cAMP and cGMP [10,40], while PDE5, PDE6, and PDE9 selectively hydrolyze cGMP. This specificity has been attributed to a ‘glutamine switch’, a conserved glutamine residue that regulates the binding of the cyclic nucleotide purine ring within the binding domain [41]. The PDE subfamilies are characterized by distinct functional domains, including a relatively conserved catalytic domain and a variety of family and subfamily regulatory domains [40,42]. Primarily, variability between PDE isoforms is found in the N- and C-terminal domains within each subfamily, which are critical for subcellular localization and activity regulation [40,42]. PDEs play a dynamic and sophisticated role in modulating cAMP signaling. This is achieved, in part, through modulation of their activity by post-translational modifications, including phosphorylation and activation by PKA [43,44]. As cAMP levels increase in the cell, PKA can phosphorylate and activate PDEs (for example, PDE3 and PDE4 isoforms), initiating a feedback loop that terminates the cAMP signal [44]. For instance, the localization and trafficking of the PDE10A splicing variant PDE10A2 are regulated by a balance between phosphorylation at threonine-16 (T16) and palmitoylation at cysteine-11 (C11), which respond to local cAMP fluctuations. In PC12 cells (derived from transplantable rat pheochromocytoma) [45], elevation of cAMP activates PKA, leading to PDE10A2 phosphorylation at T16. This phosphorylation disrupts palmitoylation at C11, resulting in accumulation of PDE10A2 in the cytosol, where it degrades cAMP to restore cellular homeostasis. Conversely, with low cAMP levels, PDE10A2 undergoes palmitoylation at C11, promoting its association with intracellular transport vesicles and facilitating its transport to the plasma membrane, where it regulates dopaminergic and glutamatergic signaling pathways [46]. This dynamic regulatory mechanism controls both membrane association and the efficient trafficking of PDE10A2 to distal dendrites [46], showing the critical role of PDEs in spatially regulating cAMP signaling. A detailed discussion on PDEs is beyond the scope of this review. However, several comprehensive reviews are available that thoroughly examine the role of PDEs. For instance, a recent review explores the kinetic properties of PDEs, including Km and Vmax, as well as their involvement in cardiac pathophysiology [47], highlighting the potential of PDEs as therapeutic drug targets.

cAMP nanodomains

The question of how cells activate stimulus-specific physiological responses despite expressing multiple GPCRs that stimulate similar levels of cAMP has remained unanswered for several decades. The observation that even activation of the same receptor family can elicit different functional effects gave a clear indication of the complexity of the system. A now widely accepted explanation of how hormonal specificity is achieved is the formation of cAMP-dependent ‘signalosomes’, multiprotein complexes that nucleate components of the cAMP signaling pathway to specific subcellular sites and are exposed to spatially restricted pools of cAMP. A prototypical cAMP signalosome may include (i) a cAMP effector, e.g., PKA; (ii) a PDE that degrades cAMP; and (iii) a scaffold protein such as an AKAP, which anchors the complex to a specific subcellular location [48,49]. Under basal conditions, localized PDE activity maintains cAMP levels below the activation threshold for PKA. Upon hormonal stimulation, the elevated cAMP concentration may be sufficient to surpass the rate of PDE-mediated degradation, leading to PKA activation. Subsequently, depending on the specific PDE present, PKA phosphorylation of the PDE may enhance its cAMP hydrolytic activity, restoring cAMP levels to baseline [50]. A good example of how cAMP signalosomes may operate to generate specific cellular effects has been reported for the cAMP response to sympathetic stimulation in the heart. Cardiac myocytes express two isoforms of the β-adrenergic receptor (β-AR), β1-AR and β2-AR. Both are activated by sympathetic hormones (epinephrine/norepinephrine) and initiate Gs-mediated cAMP production but yield distinct functional effects [51,52]. In mouse cardiac myocytes, β1-AR and β2-AR couple with separate cAMP nanodomains, where β1-ARs preferentially engages the PDE4D variant PDE4D9, whereas β2ARs activation recruits the PDE4D5 variant [53]. At resting state, β2AR preferentially associates with PDE4D9. However, upon stimulation, PDE4D9 dissociates from the receptor, while PDE4D5 is recruited to the receptor in a β-arrestin-dependent manner, likely through binding to β2AR/β-arrestin complexes [53]. Under basal conditions, PDE4D9 controls local cAMP concentration and PKA activity near β2-AR, whereas PDE4D5 likely affects β2AR signaling after ligand binding and β-arrestin recruitment. Knockdown of PDE4D9 leads to a substantial increase in the β2-AR-induced cAMP, though the increase remains below that observed with total PDE4 inhibition using rolipram, and results in increased PKA phosphorylation of the receptor under both basal and stimulatory conditions [53]. Overexpression of dominant-negative variant, such as PDE4D-DN, substantially increases the basal contraction rate. These effects of PDE4D9 are thought to be mediated by altered cAMP levels and PKA phosphorylation within β2-AR complexes, while the PDE4D5 effects were suggested to be partially due to the nonspecific increase in cytosolic cAMP and PKA phosphorylation of proteins away from the receptor complexes, as dominant-negative PDE4D5 did not affect PKA phosphorylation of β2-AR [53].

The notion that cAMP has restricted access to distinct intracellular subsets of effectors was proposed over 30 years ago, based on evidence that the two PKA isoenzymes, PKA I and II, localize to different cellular compartments. Specifically, PKA II predominantly associate with the particulate fraction of heart extracts, whereas a larger proportion of PKA I is found in the soluble fraction [54]. These findings led to the hypothesis that epinephrine stimulation selectively activates the holoenzyme in the particulate fraction [51,52]. Although various hormones elevate cAMP in cardiac myocytes, β-adrenergic agonists influence cardiac contractility while prostaglandin does not, supporting the selective activation of a subset of PKA effectors downstream of these two agonists [55,56]. These observations led to the formulation of principles outlining the non-uniform nature of cAMP signaling. Despite being a small, diffusible molecule and therefore expected to equilibrate homogeneously within the cell, it was hypothesized that cAMP does not equally access all PKA isoforms, suggesting the existence of spatially segregated cAMP pools that define distinct signaling compartments [57]. Early experimental evidence consistent with this concept includes the finding that in vitro stimulation of local β-ARs located at one extreme of a cardiac myocyte only enhances Ca2+ currents in proximity of the activated receptors but not at the opposite extreme of the cell, supporting spatial constraints on cAMP signal propagation [58]. Notably, inhibition of PDEs with IBMX (3-isobutyl-1-methylxanthine) led to a global increase of Ca2+ currents across the cell, indicating the key role of cAMP hydrolysis in cAMP compartmentation.

Significant progress in understanding cAMP compartmentalization was made possible by the introduction of molecular tools designed for real-time visualization and monitoring of intracellular cAMP in smooth muscle (BC3Hl) and fibroblast (REF-52) cell lines [59,60]. Unlike previous methods that measured bulk cAMP in cell homogenates at a single time point, these new tools allowed accurate monitoring of the intracellular dynamics of free cAMP with sub-micrometer resolution. Using this approach, direct evidence of cAMP compartmentalization in cardiac myocytes was obtained [61]. This method relies on fluorescence resonance energy transfer (FRET) between a donor and an acceptor fluorophore, each genetically fused to one or two cAMP-binding domains. The sensor design is such that the fluorophores come in close proximity to each other or move further apart upon cAMP binding, altering energy transfer and resulting in a measurable change in emitted fluorescence (ΔFRET). This approach enables detection of signaling events with high temporal and spatial resolution in intact, living cells, preserving the complexity of the intracellular environment [62-66]. The subcellular precision of this technique, combined with its ability to quantitatively track changes in real-time, enabled the first direct demonstration that β-AR activation with norepinephrine does not lead to uniform cAMP distribution within cardiac myocytes. Instead, it forms intracellular gradients, with higher concentrations at specific subcellular sites [61]. In the original study performed in CHO (Chinese hamster ovary) cells, cAMP concentration changes were tracked by monitoring FRET between the PKA RII subunit (linked to a cyan fluorophore) and the PKA C subunit (linked to a yellow fluorophore) [63]. This sensor revealed that the cAMP response induced by norepinephrine was confined to subcellular compartments where PKA is anchored to AKAPs. A variety of cAMP FRET sensors with different designs have been developed since then [62,64,65,67] and multiple studies have confirmed cAMP compartmentalization, both in cardiac myocytes [68-72] and various other cell types. These include for example CFBE (human cystic fibrosis bronchial epithelial cell, derived from a cystic fibrosis patient) [73], mIMCD‐3 (cells exhibiting epithelial morphology, isolated from the kidney of an adult mouse) and HEK293T cells [74], human airway smooth muscle cells (specialized muscle cells found in the airways of the lungs) [75], and pulmonary microvascular endothelial cells (line the small blood vessels (capillaries) in the lungs, forming a crucial barrier for gas exchange) [76]. More recent studies revealed that the size of cAMP compartments can be in the range of nanometers. As their size can be below the resolution limit of standard microscopy (200 nm), targeting the reporters within the nanodomains is necessary to capture the spatial and functional properties of the local cAMP signal [77]. Real-time imaging of cardiac myocytes using targeted reporters demonstrated that distinct subcellular structures, such as the plasmalemma, sarcoplasmic reticulum, and myofilaments—despite being only ∼300 nm apart—experience unique cAMP signaling dynamics in response to catecholamine stimulation [67].

The size of the cAMP pools generated near GPCRs in HEK293 (derived from human embryonic kidney cells) is consistent with the estimated 15–25 nm diameter of a single AKAP:PKA complex [78] and falls within the range where PDEs can effectively regulate cAMP levels [29]. Additionally, the size of these cAMP nanodomains aligns with the dimensions of AKAP:PKA clusters observed at the plasma membrane of CHO cells using super-resolution microscopy, where the average diameter of these clusters is around 100 nm [79]. This suggests that the extension of a cAMP nanodomain is calibrated to be of the size necessary to facilitate the activation of an individual PKA/AKAP signalosomes.

Key role of PDEs in cAMP compartmentalization

Physical barriers, such as membranes or organelles, have been proposed as obvious obstacles to cAMP diffusion. While these barriers may be relevant in some circumstances, a large body of evidence supports a critical role of PDEs in shaping intracellular cAMP gradients via differential degradation of cAMP at different locations [48,80-82]. Two primary models have been proposed to explain how PDEs regulate the formation of cAMP nanodomains: the barrier model and the sink model.

The notion that PDEs act as barriers to cyclic nucleotide diffusion stems from initial observations that PDE inhibition results in the broader spreading of the cAMP signal (Figure 1). The early experiments using whole-cell patch recording in frog ventricular myocytes demonstrated that local β-adrenergic stimulation could only be detected at distal L-type Ca2+ channels following PDE inhibition with IBMX [58], [83] suggesting the presence of a barrier to cAMP diffusion. Similarly, results showing that stimulation of adenosine receptors can activate the cystic fibrosis transmembrane conductance regulator channel at a distance only in the presence of PDE4 inhibitors [84] were interpreted as consistent with the presence of an enzymatic barrier to cAMP diffusion. Several other studies reported extended diffusion of cAMP signals upon PDE inhibition [85], [61], [86]. In neurons, for example, cAMP signals were shown to be restricted to specific regions near lipid rafts, with functional pools estimated to be between 20 and 200 nm [87]. Immunolocalization studies confirmed PDE4 enrichment in these areas [71], [86], supporting the idea that PDEs function as barriers protecting macromolecular complexes or subdomains from cAMP intrusion. Another example where the notion of a barrier has been invoked involves receptor-associated independent cAMP nanodomains (RAINs). At low agonist concentrations, where receptor occupancy is minimal, GPCRs (such as GLP-1R) create localized cAMP pools near the receptor, extending up to 60 nm away from the membrane. This localized cAMP pool directly contributes to receptor-associated PKA activity, forming a self-sustained and independent signaling unit. The study also suggests that PKA regulatory subunits are required within RAINs for the localized binding and buffering of cAMP. This localization ensures that cAMP remains confined, while also allowing localized PDEs to regulate the size of the GPCR-associated independent cAMP nanodomains [29].

Models for mechanism of action of PDEs.
Figure 1:
Models for mechanism of action of PDEs.

The figures were generated using biorender (Created in BioRender. Zaccolo (2025) https://BioRender.com/c75z717).

Figure 1:
Models for mechanism of action of PDEs.

The figures were generated using biorender (Created in BioRender. Zaccolo (2025) https://BioRender.com/c75z717).

Close modal

Despite numerous experimental studies [58], [61], [86], [88], [89], [90], [91], [92], [93], [94], along with computational studies [95], [96], indicating that PDEs play a critical role in the compartmentalization of cAMP signaling [72], [97], [98], the model of an enzymatic barrier is not completely satisfactory, as it does not explain how PKA subsets anchored away from the membrane and the receptors that trigger cAMP synthesis could ever become activated. In addition, this model suffers from a major challenge. The core issue lies in the apparent paradox between the rapid diffusivity of cAMP and the relatively slow enzymatic activity of PDEs [99]. The diffusion coefficient of cAMP has been experimentally estimated to be around 40 mm²/s—only about an order of magnitude slower than in water[100] . Given the rapid intracellular diffusion of cAMP, one would expect fast equilibration of cAMP levels throughout the cell, which seems to contradict the concept of cAMP compartmentalization [68], [93], [94]. Considering the reported KM and Vmax values for PDEs, it is difficult to reconcile the relatively low efficiency of PDE binding and hydrolysis of cAMP with the generation of gradients, even at basal cAMP levels, which are estimated to be around 1 µM in several cell types [99], [101], [85]. This challenge becomes even greater when trying to explain how PDEs could compartmentalize high concentrations of rapidly diffusing cAMP generated in response to hormonal stimulation.

An alternative interpretation to the barrier function is that PDEs may act as ‘sinks’, depleting cAMP in specific locations where they are localized. According to this model, PDEs generate subcellular domains, where cAMP concentrations are kept sufficiently low to prevent PKA activation (Figure 1) [68]. Evidence supporting this notion includes the characterization of a complex comprising the relaxin family peptide receptor 1 (RXFP1), the scaffold protein AKAP79, AC, AC2, β-arrestin 2, and PDE4D3 in HEK293 cells overexpressing RXFP1 [102]. In this model, relaxin induces a biphasic increase in cAMP, with the most sensitive response occurring in the femtomolar range. This response is abolished in the presence of a PDE4 inhibitor [102]. These findings suggest that the PDE4D3–PKA–β-arrestin complex maintains a nanodomain with low cAMP levels, which is crucial for a highly sensitive relaxin response [102]. A similar conclusion has been drawn for β1-adrenergic receptor signaling, where antagonists of β1AR signaling disrupt PDE4 complexes, leading to a localized increase in cAMP near the receptor despite a reduction in bulk cAMP levels [103].

Emerging mechanisms for cAMP compartmentalization

cAMP buffering has recently been proposed as a mechanism that contributes to cAMP compartmentalization. cAMP mobility has been measured in the basal state and shown to be minimal [30]. According to these studies, under basal conditions, cells are capable of buffering the majority of their cAMP. Upon minimal activation of ACs, buffering combined with the hydrolyzing activity of PDEs is still sufficient to create a cAMP nanodomain in proximity to the receptor. When PDEs are inhibited and stimuli are more intense, cAMP levels rise sufficiently to surpass the buffering capacity, leading to a progressive increase in cAMP concentrations that eventually affects the entire cell [30].

A recent study in HEK293T cells proposed a crucial driver of cAMP compartmentalization is buffering of cAMP through the formation of biomolecular condensates composed of the RI (regulatory) subunit of PKA [104]. The data show that RIα undergoes liquid–liquid phase separation (LLPS) at endogenous levels [105]. These condensates are enhanced in response to cAMP-raising stimuli, forming RIα bodies that contain high levels of cAMP and PKA activity. This dynamic sequestration of cAMP has been proposed to be essential for cAMP compartmentalization and disruption of these condensates and the release of the cAMP they sequester has been suggested to result in loss of PDE-mediated cAMP gradient formation (Figure 2). Interestingly, LLPS of RIα is disrupted by a PKA catalytic fusion oncoprotein present in fibrolamellar carcinoma, resulting in aberrant cAMP signaling and cell transformation, indicating that RIα biomolecular condensates may play an important functional role [106].

Biomolecular condensation of PKA-PDE-RIα.
Figure 2:
Biomolecular condensation of PKA-PDE-RIα.

The figures were generated using biorender (Created in BioRender. Zaccolo (2025) https://BioRender.com/h51c625).

Figure 2:
Biomolecular condensation of PKA-PDE-RIα.

The figures were generated using biorender (Created in BioRender. Zaccolo (2025) https://BioRender.com/h51c625).

Close modal
PDE8-PKA/substrate channeling

While PDEs, including PDE8, are generally assumed to hydrolyze unbound, free cAMP, studies suggest that PDEs can hydrolyze cAMP bound to its primary receptor, the R-subunit of PKA, through a molecular channeling mechanism [107], [108], [109]. This model offers an alternative mechanism for the dynamic turnover of cAMP levels. The concept of substrate channeling, commonly associated with metabolic enzymes, refers to supramolecular complexes that restrict diffusion of reaction intermediates by forming channels between associated proteins [110]. Channeling allows enzyme reactions to occur in a co-ordinated manner without the release of ligands or intermediates into the surrounding solvent (Figure 3). In the context of cAMP signaling, channeling may support the rapid and precise hydrolysis of cAMP bound to PKA by forming a PDE/PKA-R complex [108], [111]. This mechanism has been described in some detail in vitro for PDE8 and seems to involve high-affinity binding of cAMP to the R-subunit. The molecular channel would enable PDE8 to have direct access to cAMP, when tightly bound to PKA-R subunits and facilitate the translocation of cAMP to the enzyme active site for hydrolysis. This mechanism would contribute to the robustness of the cAMP signaling response, ensuring that the cAMP-PKA pathway is activated only during significant cAMP fluxes, such as those observed during hormonal stimulation, allowing for adaptation to steady-state cAMP levels [112].

cAMP channeling in a PDE8/PKA-RI complex.
Figure 3:
cAMP channeling in a PDE8/PKA-RI complex.

The figures were generated using biorender (Created in BioRender. Zaccolo (2025) https://BioRender.com/j24o028).

Figure 3:
cAMP channeling in a PDE8/PKA-RI complex.

The figures were generated using biorender (Created in BioRender. Zaccolo (2025) https://BioRender.com/j24o028).

Close modal

GPCR in intracellular membrane compartments and their role in cAMP signaling

Traditionally, GPCRs have been thought to function exclusively at the cell surface, transmitting signals in response to extracellular stimuli. However, recent studies indicate that GPCRs are also capable of signaling from intracellular membrane compartments, including endosomes, the Golgi apparatus, and nuclear membranes [24], [113], [114]. Notably, intracellular GPCR signaling often leads to distinct functional outcomes compared with plasma membrane signaling, despite often utilizing the same G protein effectors and second messengers, supporting a role of signaling from internal membranes in determining specificity of response. Within this paradigm, the generation of cAMP from intracellular GPCRs would provide a local source of second messenger that could feed and activate selectively signalosomes localized nearby.

Signaling from endosomal GPCRs

Following activation and internalization from the plasma membrane, many GPCRs enter the endosomal pathway and are sorted within endosomal compartments through a highly regulated process, determining whether they can either recycle back to the plasma membrane or be directed towards lysosomal degradation [115,116]. The sorting process can involve receptor-specific amino acid motifs, post-translational modifications, such as phosphorylation and ubiquitination, and interactions with endosome-associated proteins [117]. The internalization is primarily mediated by clathrin-coated vesicles, facilitated by the interaction of β-arrestin-1 or 2 with the clathrin adaptor protein AP-2 [118,119]. Internalization of GPCRs from the plasma membrane to endosomes was traditionally viewed as a mechanism for signal termination and receptor desensitization [120,121]. However, the discovery that GPCR-mediated G protein signaling can persist from endosomes [122] has revealed a more nuanced role for receptor internalization. One notable example involves the luteinizing hormone receptor (LHR), where sustained cAMP signaling by receptor internalization is essential for the meiotic progression of oocytes within ovarian follicles. LHR expression is predominantly restricted to the two outermost cell layers of the mature ovarian follicle, the theca and mural granulosa cells. Luteinizing hormone (LH) signaling occurs through a GPCR selectively expressed in these external follicular layers, which enables the receptor to trigger distinct temporal and spatial phases of cAMP signaling[123] . The LHR stimulation induces two distinct phases of cAMP signaling in the cells surrounding the oocyte: first a reversible cAMP signal is generated, followed by a persistent increase in cAMP levels after internalization of LHR has occurred. The disruption of LHR internalization abolishes the second, persistent phase of cAMP signaling, which partially inhibits the resumption of oocyte meiosis [123], suggesting that sustained cAMP signaling from internalized LHRs is critical for transmitting LH signals to follicle cells and the oocyte.

Signaling from GPCR localized to the Golgi apparatus

GPCRs can also undergo retrograde transport to the Golgi apparatus or may be retained within the Golgi during biosynthetic trafficking before being delivered to the plasma membrane [124]. Retrograde trafficking from endosomes to the trans-Golgi network (TGN) for the recycling and retrieval of protein cargos/transmembrane proteins is frequently mediated by the retromer complex [124]. The retromer is a highly conserved, cytosolic/peripheral membrane hetero-pentameric protein complex present in all mammalian cells. It consists of two endosomal membrane-bound sorting nexins (Snx1 and Snx2), proteins that contribute to membrane recruitment and formation of recycling tubules [125], and a soluble heterotrimer (Vps26, Vps29, and Vps35) that recognizes cargos [126-129]. GPCRs, such as the parathyroid hormone receptor and thyroid-stimulating hormone receptor (TSHR), have been shown to traffic retrogradely to the TGN [130], [131]. While not all GPCRs are recycled through the TGN, all GPCRs are integral membrane proteins and must transit through the Golgi apparatus, where they undergo crucial post-translational modifications, such as glycosylation, prior to their insertion into the plasma membrane [132]. Although many GPCRs pass transiently through the Golgi, certain receptors, like the delta-opioid receptor (DOR) and the β1-adrenergic receptor, display steady-state localization within the Golgi [133], [134]. The export of GPCRs from the Golgi is tightly regulated by various factors, including receptor-specific amino acid motifs, interactions with trafficking proteins, and lipid compositions of the Golgi membrane [135], [136], [137], [138]. The functional significance of Golgi-localized GPCR signaling is evident from the distinct cellular effects it elicits compared with signals generated by GPCRs located at the plasma membrane. One striking example comes from the studies on TSHR, which demonstrate that cAMP and PKA signaling can persist in the Golgi up to 10 minutes post-receptor internalization, and this sustained signaling can be abrogated by treatment with inhibitors of endocytosis [131]. In thyroid cells, internalized TSHR retrogradely traffics to the TGN and activates endogenous Gs proteins within retromer-coated compartments. This process is crucial for inducing a late-phase cAMP/PKA response localized at the Golgi/TGN [131]. Importantly, this localized signaling at the Golgi is required for efficient activation of downstream targets such as cAMP response element-binding protein (CREB), a key transcription factor. Disruption of this TSH-induced phosphorylation pathway—whether by blocking receptor internalization, inhibiting PKA II or its localization to the Golgi/TGN, silencing retromer, or disrupting Golgi/TGN organization—significantly impairs CREB phosphorylation [131]. These processes highlight that retrograde trafficking to the TGN is not simply a means of receptor recycling but instead a critical process that positions GPCRs near the nucleus to activate local Gs protein-mediated cAMP/PKA signaling.

Signaling from GPCRs localized to the nuclear membrane and endoplasmic reticulum

GPCRs can localize to nuclear membranes [139] and associated ER membranes either following synthesis or after internalization from the plasma membrane [140]. The localization to the nuclear membranes can be mediated by classical nuclear localization sequences (NLS) and/or by interactions with importin proteins [140]. Nuclear targeting of GPCRs via these mechanisms has been observed following internalization, as demonstrated for the protease-activated receptor 2 and the oxytocin receptor [140]. Alternatively, GPCRs, such as the α1-adrenergic receptors and platelet-activating factor receptor may be targeted directly to the nucleus after synthesis [140], [141], [142], [143]. The protease-activated receptor 4 uses classical RXR (arginine-X-arginine, where X can be any amino acid) motifs for retention in the ER. These motifs are recognized by the COP-I (coatomer protein I) protein complex, which plays a role in the retention of GPCRs within the biosynthetic pathway [144], [145]. Interestingly, similar COPI-interacting motifs are also found in GPCRs localized to the Golgi apparatus, such as the DOR [137], [146]. However, the metabotropic glutamate receptor 5 (mGluR5) achieves nuclear localization through non-NLS sequences, a non-canonical mechanism to reach the nucleus, potentially interacting with chromatin to retain its nuclear presence [147]. Signaling from nuclear membrane-localized GPCRs, such as mGluR5 in neurons, provides a compelling example of how intracellular signaling can diverge [148]. For example, stimulation of either cell surface or intracellular mGluR5 triggers the phosphorylation typical of plasma membrane-based signaling. In striatal neurons, both plasma membrane-localized and intracellular mGluR5 signaling are capable of activating the transcription factor CREB [148]. The activation of mGluR5 at these two locations generates different Ca²+ signaling patterns, leading to unique cellular responses of c-Jun N-terminal kinase, Ca²+/calmodulin-dependent protein kinase, and CREB. However, only intracellular mGluR5 activation results in the phosphorylation of extracellular signal-regulated kinases 1 and 2 and Elk-1, a transcriptional regulator [148]. In striatal neurons, intracellular mGluR5 is both present and functional in hippocampal neurons, where it can cause dendritic Ca2+ rises that differ in amplitude from cell surface receptor activation. It has also been shown that intracellular mGluR5 can mediate protein-synthesis-dependent long-term depression (LTD), whereas cell surface mGluR5 is involved in both LTD and Long-term potentiation [148], highlighting the diverse physiological effects that can arise from GPCR signaling at intracellular membranes.

The discovery of novel mechanisms governing cAMP compartmentalization opens several exciting avenues for future research. One critical area of focus is the further exploration of LLPS in the regulation of cAMP nanodomains. The involvement of biomolecular condensates in cellular signaling pathways is a relatively new concept, and future studies will need to investigate how LLPS integrates with other mechanisms, such as PDE activity or PKA-RI subunits interactions with AKAPs, to shape cAMP dynamics. Additionally, understanding the conditions that promote or disrupt LLPS in disease contexts could reveal new therapeutic avenues.

Another promising direction is the study of substrate channeling between PDEs and PKA. While this concept is well-established in metabolic pathways, its application to cAMP signaling is still in its infancy.

Finally, the role of cAMP buffering in regulating cAMP diffusion warrants further exploration. While recent studies have demonstrated its importance, the full extent of its impact on cAMP compartmentalization remains unclear. Future research could focus on identifying additional cAMP-binding proteins involved in buffering and how these interactions change under various physiological and pathological conditions.

As our understanding of cAMP nanodomains continues to evolve, integrating these novel mechanisms with a deeper knowledge of GPCR signaling from different subcellular locations will be crucial for developing a comprehensive model of GPCR signaling via cAMP compartmentalization.

The authors declare that there are no competing interest associated with the manuscript.

This work was supported by funding from British Heart Foundation (RG/17/6/32944, RE/18/3/34214) and the Oxford NIHR Biomedical Research Centre to MZ.

RY: Writing—original draft, Writing—review & editing. MZ: Conceptualization, Supervision, Funding acquisition, Writing—original draft, Writing—review & editing.

AC

adenylyl cyclase

AH

alpha-helical

AR

adrenergic receptor

C11

cysteine-11

CREB

cAMP response element-binding protein

DOR

delta-opioid receptor

ER

endoplasmic reticulum

FRET

Fluorescence Resonance Energy Transfer

GPCR

G protein-coupled receptors

GSIS

glucose-stimulated insulin secretion

LH

Luteinizing hormone

LHR

luteinizing hormone receptor

LLPS

liquid–liquid phase separation

LTD

long-term depression

NLS

nuclear localization sequences

PDE

phosphodiesterases

PKA

protein kinase A

PKC

protein kinase C

PP2B

phosphatase2B

RAINs

receptor-associated independent cAMP nanodomains

RD

Ras-like domain

RXFP1

relaxin family peptide receptor 1

T16

threonine-16

TGN

trans-Golgi network

TSHR

thyroid-stimulating hormone receptor

mGluR5

metabotropic glutamate receptor 5

α2A-AR

α2A-adrenergic receptor

β-AR

β-adrenergic receptor

1.
Fredriksson
,
R.
,
Lagerström
,
M.C.
,
Lundin
,
L.G.
and
Schiöth
,
H.B
. (
2003
)
The G-protein-coupled receptors in the human genome form five main families. Phylogenetic analysis, paralogon groups, and fingerprints
.
Mol. Pharmacol.
63
,
1256
1272
https://doi.org/10.1124/mol.63.6.1256
2.
Rosenbaum
,
D.M.
,
Rasmussen
,
S.G.F.
and
Kobilka
,
B.K
. (
2009
)
The structure and function of G-protein-coupled receptors
.
Nature
459
,
356
363
https://doi.org/10.1038/nature08144
3.
Weis
,
W.I.
and
Kobilka
,
B.K
. (
2018
)
The molecular basis of G protein-coupled receptor activation
.
Annu. Rev. Biochem.
87
,
897
919
https://doi.org/10.1146/annurev-biochem-060614-033910
4.
Bourne
,
H.R.
,
Sanders
,
D.A.
and
McCormick
,
F
. (
1990
)
The GTPase superfamily: a conserved switch for diverse cell functions
.
Nature
348
,
125
132
https://doi.org/10.1038/348125a0
5.
Gilman
,
A.G
. (
1987
)
G proteins: transducers of receptor-generated signals
.
Annu. Rev. Biochem.
56
,
615
649
https://doi.org/10.1146/annurev.bi.56.070187.003151
6.
Simon
,
M.I.
,
Strathmann
,
M.P.
and
Gautam
,
N
. (
1991
)
Diversity of G proteins in signal transduction
.
Science
252
,
802
808
https://doi.org/10.1126/science.1902986
7.
Calebiro
,
D.
,
Koszegi
,
Z.
,
Lanoiselée
,
Y.
,
Miljus
,
T.
and
O’Brien
,
S
. (
2021
)
G protein-coupled receptor-G protein interactions: a single-molecule perspective
.
Physiol. Rev.
101
,
857
906
https://doi.org/10.1152/physrev.00021.2020
8.
Seyedabadi
,
M.
,
Gharghabi
,
M.
,
Gurevich
,
E.V.
and
Gurevich
,
V.V
. (
2022
)
Structural basis of GPCR coupling to distinct signal transducers: implications for biased signaling
.
Trends Biochem. Sci.
47
,
570
581
https://doi.org/10.1016/j.tibs.2022.03.009
9.
Sutherland
,
E.W.
and
Rall
,
T.W
. (
1958
)
Fractionation and characterization of a cyclic adenine ribonucleotide formed by tissue particles
.
J. Biol. Chem.
232
,
1077
1091
https://doi.org/10.1016/S0021-9258(19)77423-7
10.
Azevedo
,
M.F
. (
2014
)
Clinical and molecular genetics of the phosphodiesterases (PDEs)
.
Endocr. Rev.
35
,
195
233
https://doi.org/10.1210/er.2013-1053
11.
Sassi
,
Y.
,
Abi‐Gerges
,
A.
,
Fauconnier
,
J.
,
Mougenot
,
N.
,
Reiken
,
S.
,
Haghighi
,
K.
et al
. (
2012
)
Regulation of cAMP homeostasis by the efflux protein MRP4 in cardiac myocytes
.
FASEB j.
26
,
1009
1017
https://doi.org/10.1096/fj.11-194027
12.
Walsh
,
I.A.
,
I-’eriuns
,
J.P.
and
Krexs
,
E.G
. (
1968
)
Communications an adenosine 3’) 5’-monophosphate-dependant protein kinase from rabbit skeletal muscle (Received for publication, April 2, 19G8)
.
J. Biol. Chem.
243
, 1068 https://doi.org/10.1016/S0021-9258(19)34204-8
13.
de Rooij
,
J.
,
Zwartkruis
,
F.J.T.
,
Verheijen
,
M.H.G.
,
Cool
,
R.H.
,
Nijman
,
S.M.B.
,
Wittinghofer
,
A.
et al
. (
1998
)
Epac is a Rap1 guanine-nucleotide-exchange factor directly activated by cyclic AMP
.
Nature
396
,
474
477
https://doi.org/10.1038/24884
14.
Fesenko
,
E.E.
,
Kolesnikov
,
S.S.
and
Lyubarsky
,
A.L
. (
1985
)
Induction by cyclic GMP of cationic conductance in plasma membrane of retinal rod outer segment
.
Nature
313
,
310
313
https://doi.org/10.1038/313310a0
15.
Brand
,
T
. (
2005
)
The Popeye domain-containing gene family
.
Cell Biochem. Biophys.
43
,
95
103
https://doi.org/10.1385/CBB:43:1:095/METRICS
16.
Jhala
,
U.S.
,
Canettieri
,
G.
,
Screaton
,
R.A.
,
Kulkarni
,
R.N.
,
Krajewski
,
S.
,
Reed
,
J.
et al
. (
2003
)
cAMP promotes pancreatic beta-cell survival via CREB-mediated induction of IRS2
.
Genes Dev.
17
,
1575
1580
https://doi.org/10.1101/gad.1097103
17.
Li
,
M.
,
Wang
,
X.
,
Meintzer
,
M.K.
,
Laessig
,
T.
,
Birnbaum
,
M.J.
and
Heidenreich
,
K.A
. (
2000
)
Cyclic AMP promotes neuronal survival by phosphorylation of glycogen synthase kinase 3beta
.
Mol. Cell. Biol.
20
,
9356
9363
https://doi.org/10.1128/MCB.20.24.9356-9363.2000
18.
Boularan
,
C.
and
Gales
,
C
. (
2015
)
Cardiac cAMP: production, hydrolysis, modulation and detection
.
Front. Pharmacol
6
, 203 https://doi.org/10.3389/fphar.2015.00203
19.
Kandel
,
E.R
. (
2012
)
The molecular biology of memory: cAMP, PKA, CRE, CREB-1, CREB-2, and CPEB
.
Mol. Brain
5
, 14 https://doi.org/10.1186/1756-6606-5-14
20.
Tang
,
C.M.
and
Insel
,
P.A
. (
2004
)
GPCR expression in the heart; “new” receptors in myocytes and fibroblasts
.
Trends Cardiovasc. Med.
14
,
94
99
https://doi.org/10.1016/j.tcm.2003.12.007
21.
Demby
,
A.
and
Zaccolo
,
M
. (
2024
)
GPCR signalling: Yet another variant route in a highly complex road map
.
Cell Calcium
124
, S0143-4160(24)00123-4 https://doi.org/10.1016/j.ceca.2024.102965
22.
Smith
,
F.D.
,
Langeberg
,
L.K.
and
Scott
,
J.D
. (
2006
)
The where’s and when’s of kinase anchoring
.
Trends Biochem. Sci.
31
,
316
323
https://doi.org/10.1016/j.tibs.2006.04.009
23.
Lohse
,
M.J.
,
Nuber
,
S.
and
Hoffmann
,
C
. (
2012
)
Fluorescence/bioluminescence resonance energy transfer techniques to study G-protein-coupled receptor activation and signaling
.
Pharmacol. Rev.
64
,
299
336
https://doi.org/10.1124/pr.110.004309
24.
Eichel
,
K.
and
Von Zastrow
,
M
. (
2018
)
Subcellular organization of GPCR signaling
.
Trends Pharmacol. Sci.
39
,
200
208
https://doi.org/10.1016/j.tips.2017.11.009
25.
Lohse
,
M.J.
,
Bock
,
A.
and
Zaccolo
,
M
. (
2024
)
G protein-coupled receptor signaling: new insights define cellular nanodomains
.
Annu. Rev. Pharmacol. Toxicol.
64
,
387
415
https://doi.org/10.1146/annurev-pharmtox-040623-115054
26.
Sungkaworn
,
T.
,
Jobin
,
M.L.
,
Burnecki
,
K.
,
Weron
,
A.
,
Lohse
,
M.J.
and
Calebiro
,
D
. (
2017
)
Single-molecule imaging reveals receptor-G protein interactions at cell surface hot spots
.
Nature
550
,
543
547
https://doi.org/10.1038/nature24264
27.
Shnerb
,
T.
,
Lin
,
N.
and
Shurki
,
A
. (
2007
)
What is the role of the helical domain of Gsα in the GTPase reaction
.
Biochemistry
46
,
10875
10885
https://doi.org/10.1021/BI700585W/SUPPL_FILE/BI700585W-FILE002.PDF
28.
Westfield
,
G.H
. (
2011
)
Structural flexibility of the Gαs α-helical domain in the β 2-adrenoceptor Gs complex
.
Proc. Natl. Acad. Sci. U.S.A.
108
,
16086
16091
https://doi.org/10.1073/PNAS.1113645108/SUPPL_FILE/SAPP.PDF
29.
Anton
,
S.E.
,
Kayser
,
C.
,
Maiellaro
,
I.
,
Nemec
,
K.
,
Möller
,
J.
,
Koschinski
,
A.
et al
. (
2022
)
Receptor-associated independent cAMP nanodomains mediate spatiotemporal specificity of GPCR signaling
.
Cell
185
,
1130
1142
. https://doi.org/10.1016/j.cell.2022.02.011
30.
Bock
,
A.
,
Annibale
,
P.
,
Konrad
,
C.
,
Hannawacker
,
A.
,
Anton
,
S.E.
,
Maiellaro
,
I.
et al
. (
2020
)
Optical Mapping of cAMP Signaling at the Nanometer Scale
.
Cell
182
,
1519
1530
. https://doi.org/10.1016/j.cell.2020.07.035
31.
Scott
,
J.D.
,
Dessauer
,
C.W.
and
Taskén
,
K
. (
2013
)
Creating order from chaos: cellular regulation by kinase anchoring
.
Annu. Rev. Pharmacol. Toxicol.
53
,
187
210
https://doi.org/10.1146/ANNUREV-PHARMTOX-011112-140204/CITE/REFWORKS
32.
Chen
,
M.H.
and
Malbon
,
C.C
. (
2009
)
G-protein-coupled receptor-associated A-kinase anchoring proteins AKAP5 and AKAP12: differential trafficking and distribution
.
Cell. Signal.
21
,
136
142
https://doi.org/10.1016/j.cellsig.2008.09.019
33.
Weisenhaus
,
M.
,
Allen
,
M.L.
,
Yang
,
L.
,
Lu
,
Y.
,
Nichols
,
C.B.
,
Su
,
T.
et al
. (
2010
)
Mutations in AKAP5 disrupt dendritic signaling complexes and lead to electrophysiological and behavioral phenotypes in mice
.
PLoS ONE
5
, e10325 https://doi.org/10.1371/journal.pone.0010325
34.
Gildart
,
M.
,
Kapiloff
,
M.S.
and
Dodge-Kafka
,
K.L
. (
2020
)
Calcineurin-AKAP interactions: therapeutic targeting of a pleiotropic enzyme with a little help from its friends
.
J. Physiol. (Lond.)
598
,
3029
3042
https://doi.org/10.1113/JP276756
35.
He
,
Z.
,
Xie
,
L.
,
Liu
,
J.
,
Wei
,
X.
,
Zhang
,
W.
and
Mei
,
Z
. (
2024
)
Novel insight into the role of A-kinase anchoring proteins (AKAPs) in ischemic stroke and therapeutic potentials
.
Biomedicine & Pharmacotherapy
175
,
116715
https://doi.org/10.1016/j.biopha.2024.116715
36.
Hinke
,
S.A
, et al
. (
2012
)
Anchored phosphatases modulate glucose homeostasis
.
EMBO J.
31
,
3991
4004
https://doi.org/10.1038/EMBOJ.2012.244/SUPPL_FILE/EMBJ2012244.REVIEWER_COMMENTS.PDF
37.
Marzook
,
A.
,
Tomas
,
A.
and
Jones
,
B
. (
2021
)
The interplay of glucagon-like peptide-1 receptor trafficking and signalling in pancreatic beta cells
.
Front. Endocrinol. (Lausanne)
12
, 678055 https://doi.org/10.3389/fendo.2021.678055
38.
Rutter
,
G.A.
,
Hodson
,
D.J.
,
Chabosseau
,
P.
,
Haythorne
,
E.
,
Pullen
,
T.J.
and
Leclerc
,
I
. (
2017
)
Local and regional control of calcium dynamics in the pancreatic islet
.
Diabetes. Obes. Metab.
19 Suppl 1
,
30
41
https://doi.org/10.1111/dom.12990
40.
Bender
,
A.T.
and
Beavo
,
J.A
. (
2006
)
Cyclic nucleotide phosphodiesterases: molecular regulation to clinical use
.
Pharmacol. Rev.
58
,
488
520
https://doi.org/10.1124/pr.58.3.5
41.
Zhang
,
K.Y.J.
,
Card
,
G.L.
,
Suzuki
,
Y.
,
Artis
,
D.R.
,
Fong
,
D.
,
Gillette
,
S.
et al
. (
2004
)
A glutamine switch mechanism for nucleotide selectivity by phosphodiesterases
.
Mol. Cell
15
,
279
286
https://doi.org/10.1016/j.molcel.2004.07.005
42.
Thul
,
P.J.
,
Åkesson
,
L.
,
Wiking
,
M.
,
Mahdessian
,
D.
,
Geladaki
,
A.
,
Ait Blal
,
H.
et al
. (
2017
)
A subcellular map of the human proteome
.
Science
356
, eaal3321 https://doi.org/10.1126/science.aal3321
43.
Conti
,
M.
and
Jin
,
S.L
. (
1999
)
The molecular biology of cyclic nucleotide phosphodiesterases
.
Prog. Nucleic Acid Res. Mol. Biol.
63
,
1
38
https://doi.org/10.1016/s0079-6603(08)60718-7
44.
Sette
,
C.
and
Conti
,
M
. (
1996
)
Phosphorylation and activation of a cAMP-specific phosphodiesterase by the cAMP-dependent protein kinase. Involvement of serine 54 in the enzyme activation
.
J. Biol. Chem.
271
,
16526
16534
https://doi.org/10.1074/jbc.271.28.16526
45.
Wiatrak
,
B.
,
Kubis-Kubiak
,
A.
,
Piwowar
,
A.
and
Barg
,
E
. (
2020
)
PC12 cell line: cell types, coating of culture vessels, differentiation and other culture conditions
.
Cells
9
, 958 https://doi.org/10.3390/cells9040958
46.
Charych
,
E.I.
,
Jiang
,
L.X.
,
Lo
,
F.
,
Sullivan
,
K.
and
Brandon
,
N.J
. (
2010
)
Interplay of palmitoylation and phosphorylation in the trafficking and localization of phosphodiesterase 10A: implications for the treatment of schizophrenia
.
J. Neurosci.
30
,
9027
9037
https://doi.org/10.1523/JNEUROSCI.1635-10.2010
47.
Kelly
,
M.P
, et al
. (
2025
)
Cyclic nucleotide phosphodiesterases as drug targets
.
Pharmacol. Rev.
77
, 100042 https://doi.org/10.1016/J.PHARMR.2025.100042/ASSET/EE6520D4-77A6-4245-A879-45294FE577AB/MAIN.ASSETS/GR11.JPG
48.
Zaccolo
,
M.
and
Kovanich
,
D
. (
2025
)
Nanodomain cAMP signaling in cardiac pathophysiology: potential for developing targeted therapeutic interventions
.
Physiol. Rev.
105
,
541
591
https://doi.org/10.1152/physrev.00013.2024
49.
Monterisi
,
S.
and
Zaccolo
,
M
. (
2017
)
Components of the mitochondrial cAMP signalosome
.
Biochem. Soc. Trans.
45
,
269
274
https://doi.org/10.1042/BST20160394
50.
Dodge
,
K.L.
,
Khouangsathiene
,
S.
,
Kapiloff
,
M.S.
,
Mouton
,
R.
,
Hill
,
E.V.
,
Houslay
,
M.D.
et al
. (
2001
)
mAKAP assembles A protein kinase A/PDE4 phosphodiesterase cAMP signaling module
.
EMBO J.
20
,
1921
1930
https://doi.org/10.1093/emboj/20.8.1921
51.
Chruscinski
,
A.J.
,
Rohrer
,
D.K.
,
Schauble
,
E.
,
Desai
,
K.H.
,
Bernstein
,
D.
and
Kobilka
,
B.K
. (
1999
)
Targeted disruption of the beta2 adrenergic receptor gene
.
J. Biol. Chem.
274
,
16694
16700
https://doi.org/10.1074/jbc.274.24.16694
52.
Rohrer
,
D.K.
,
Desai
,
K.H.
,
Jasper
,
J.R.
,
Stevens
,
M.E.
,
Regula
,
D.P.
Jr
,
Barsh
,
G.S.
et al
. (
1996
)
Targeted disruption of the mouse beta1-adrenergic receptor gene: developmental and cardiovascular effects
.
Proc. Natl. Acad. Sci. U.S.A.
93
,
7375
7380
https://doi.org/10.1073/pnas.93.14.7375
53.
De Arcangelis
,
V.
,
Liu
,
R.
,
Soto
,
D.
and
Xiang
,
Y
. (
2009
)
Differential association of phosphodiesterase 4D isoforms with beta2-adrenoceptor in cardiac myocytes
.
J. Biol. Chem.
284
,
33824
33832
https://doi.org/10.1074/jbc.M109.020388
54.
Di Benedetto
,
G.
,
Zoccarato
,
A.
,
Lissandron
,
V.
,
Terrin
,
A.
,
Li
,
X.
,
Houslay
,
M.D.
et al
. (
2008
)
Protein kinase A type I and type II define distinct intracellular signaling compartments
.
Circ. Res.
103
,
836
844
https://doi.org/10.1161/CIRCRESAHA.108.174813
55.
Hayes
,
J.S.
,
Brunton
,
L.L.
,
Brown
,
J.H.
,
Reese
,
J.B.
and
Mayer
,
S.E
. (
1979
)
Hormonally specific expression of cardiac protein kinase activity
.
Proc. Natl. Acad. Sci. U.S.A.
76
,
1570
1574
https://doi.org/10.1073/pnas.76.4.1570
56.
Buxton
,
I.L.O.
and
Brunton
,
L.L
. (
1983
)
Compartments of cyclic AMP and protein kinase in mammalian cardiomyocytes
.
Journal of Biological Chemistry
258
,
10233
10239
https://doi.org/10.1016/S0021-9258(17)44447-4
57.
Steinberg
,
S.F.
and
Brunton
,
L.L
. (
2001
)
Compartmentation of G protein-coupled signaling pathways in cardiac myocytes
.
Annu. Rev. Pharmacol. Toxicol.
41
,
751
773
https://doi.org/10.1146/annurev.pharmtox.41.1.751
58.
Jurevicius
,
J.
and
Fischmeister
,
R
. (
1996
)
cAMP compartmentation is responsible for a local activation of cardiac Ca2+ channels by beta-adrenergic agonists
.
Proc. Natl. Acad. Sci. U.S.A.
93
,
295
299
https://doi.org/10.1073/pnas.93.1.295
59.
Bacskai
,
B.J.
,
Hochner
,
B.
,
Mahaut-Smith
,
M.
,
Adams
,
S.R.
,
Kaang
,
B.K.
,
Kandel
,
E.R.
et al
. (
1993
)
Spatially resolved dynamics of cAMP and protein kinase A subunits in Aplysia sensory neurons
.
Science
260
,
222
226
https://doi.org/10.1126/science.7682336
60.
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
https://doi.org/10.1038/349694a0
61.
Zaccolo
,
M.
and
Pozzan
,
T
. (
2002
)
Discrete microdomains with high concentration of cAMP in stimulated rat neonatal cardiac myocytes
.
Science
295
,
1711
1715
https://doi.org/10.1126/science.1069982
62.
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
https://doi.org/10.1073/PNAS.0405973101/ASSET/628FCF9F-376B-4DC8-9359-8EDB073E98E7/ASSETS/GRAPHIC/ZPQ0480465860004.JPEG
63.
Zaccolo
,
M.
,
De Giorgi
,
F.
,
Cho
,
C.Y.
,
Feng
,
L.
,
Knapp
,
T.
,
Negulescu
,
P.A.
et al
. (
2000
)
A genetically encoded, fluorescent indicator for cyclic AMP in living cells
.
Nat. Cell Biol.
2
,
25
29
https://doi.org/10.1038/71345
64.
Ponsioen
,
B
, et al
. (
2004
)
Detecting cAMP-induced epac activation by fluorescence resonance energy transfer: Epac as a novel cAMP indicator
.
EMBO Rep.
5
,
1176
1180
https://doi.org/10.1038/SJ.EMBOR.7400290/SUPPL_FILE/EMBR7400290-SUP-0001.PDF
65.
Nikolaev
,
V.O.
,
Bünemann
,
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
https://doi.org/10.1074/jbc.C400302200
66.
Demby
,
A.
and
Zaccolo
,
M
. (
2023
)
Investigating G-protein coupled receptor signalling with light-emitting biosensors
.
Front. Physiol.
14
,
1310197
https://doi.org/10.3389/FPHYS.2023.1310197/XML/NLM
67.
Surdo
,
N.C.
,
Berrera
,
M.
,
Koschinski
,
A.
,
Brescia
,
M.
,
Machado
,
M.R.
,
Carr
,
C.
et al
. (
2017
)
FRET biosensor uncovers cAMP nano-domains at β-adrenergic targets that dictate precise tuning of cardiac contractility
.
Nat. Commun.
8
, 15031 https://doi.org/10.1038/ncomms15031
68.
Terrin
,
A.
,
Di Benedetto
,
G.
,
Pertegato
,
V.
,
Cheung
,
Y.F.
,
Baillie
,
G.
,
Lynch
,
M.J.
et al
. (
2006
)
PGE(1) stimulation of HEK293 cells generates multiple contiguous domains with different [cAMP]: role of compartmentalized phosphodiesterases
.
J. Cell Biol.
175
,
441
451
https://doi.org/10.1083/jcb.200605050
69.
Iancu
,
R.V.
,
Jones
,
S.W.
and
Harvey
,
R.D
. (
2007
)
Compartmentation of cAMP signaling in cardiac myocytes: a computational study
.
Biophys. J.
92
,
3317
3331
https://doi.org/10.1529/biophysj.106.095356
70.
Shcherbakova
,
O.G.
,
Hurt
,
C.M.
,
Xiang
,
Y.
,
Dell’Acqua
,
M.L.
,
Zhang
,
Q.
,
Tsien
,
R.W.
et al
. (
2007
)
Organization of beta-adrenoceptor signaling compartments by sympathetic innervation of cardiac myocytes
.
J. Cell Biol.
176
,
521
533
https://doi.org/10.1083/jcb.200604167
71.
Leroy
,
J.
,
Richter
,
W.
,
Mika
,
D.
,
Castro
,
L.R.V.
,
Abi-Gerges
,
A.
and
Xie
,
M
. (
2011
)
Phosphodiesterase 4B in the cardiac L-type Ca
.
J. Clin. Invest.
121
,
2651
2661
https://doi.org/10.1172/JCI44747
72.
Feinstein
,
W.P.
,
Zhu
,
B.
,
Leavesley
,
S.J.
,
Sayner
,
S.L.
and
Rich
,
T.C
. (
2012
)
Assessment of cellular mechanisms contributing to cAMP compartmentalization in pulmonary microvascular endothelial cells
.
Am. J. Physiol., Cell Physiol.
302
,
C839
52
https://doi.org/10.1152/ajpcell.00361.2011
73.
Maiellaro
,
I.
,
Lohse
,
M.J.
,
Kittel
,
R.J.
and
Calebiro
,
D
. (
2016
)
cAMP signals in drosophila motor neurons are confined to single synaptic boutons
.
Cell Rep.
17
,
1238
1246
https://doi.org/10.1016/j.celrep.2016.09.090
74.
Monterisi
,
S.
,
Favia
,
M.
,
Guerra
,
L.
,
Cardone
,
R.A.
,
Marzulli
,
D.
,
Reshkin
,
S.J.
et al
. (
2012
)
CFTR regulation in human airway epithelial cells requires integrity of the actin cytoskeleton and compartmentalized cAMP and PKA activity
.
J. Cell. Sci.
125
,
1106
1117
https://doi.org/10.1242/jcs.089086
75.
Hansen
,
J.N
, et al
. (
2022
)
A cAMP signalosome in primary cilia drives gene expression and kidney cyst formation
.
EMBO Rep.
23
,
8
https://doi.org/10.15252/EMBR.202154315/SUPPL_FILE/EMBR202154315-SUP-0003-DATASETEV1.XLSX
76.
Agarwal
,
S.R.
,
Miyashiro
,
K.
,
Latt
,
H.
,
Ostrom
,
R.S.
and
Harvey
,
R.D
. (
2017
)
Compartmentalized cAMP responses to prostaglandin EP 2 receptor activation in human airway smooth muscle cells
.
Br. J. Pharmacol.
174
,
2784
2796
https://doi.org/10.1111/bph.13904
77.
Zaccolo
,
M.
,
Zerio
,
A.
and
Lobo
,
M.J
. (
2021
)
Subcellular organization of the cAMP signaling pathway
.
Pharmacol. Rev.
73
,
278
309
https://doi.org/10.1124/pharmrev.120.000086
78.
Smith
,
F.D.
,
Esseltine
,
J.L.
,
Nygren
,
P.J.
,
Veesler
,
D.
,
Byrne
,
D.P.
,
Vonderach
,
M.
et al
. (
2017
)
Local protein kinase A action proceeds through intact holoenzymes
.
Science
356
,
1288
1293
https://doi.org/10.1126/science.aaj1669
79.
Mo
,
G.C.H.
,
Ross
,
B.
,
Hertel
,
F.
,
Manna
,
P.
,
Yang
,
X.
,
Greenwald
,
E.
et al
. (
2017
)
Genetically encoded biosensors for visualizing live-cell biochemical activity at super-resolution
.
Nat. Methods
14
,
427
434
https://doi.org/10.1038/nmeth.4221
80.
Conti
,
M.
,
Mika
,
D.
and
Richter
,
W
. (
2014
)
Cyclic AMP compartments and signaling specificity: role of cyclic nucleotide phosphodiesterases
.
J. Gen. Physiol.
143
,
29
38
https://doi.org/10.1085/jgp.201311083
81.
Brescia
,
M.
and
Zaccolo
,
M
. (
2016
)
Modulation of compartmentalised cyclic nucleotide signalling via local inhibition of phosphodiesterase activity
.
Int. J. Mol. Sci.
17
, 1672 https://doi.org/10.3390/ijms17101672
82.
Mika
,
D.
,
Leroy
,
J.Ô.
,
Vandecasteele
,
G.
and
Fischmeister
,
R
. (
2012
)
PDEs create local domains of cAMP signaling
.
J. Mol. Cell. Cardiol.
52
,
323
329
https://doi.org/10.1016/j.yjmcc.2011.08.016
83.
Fischmeister
,
R.
,
Castro
,
L.R.V.
,
Abi-Gerges
,
A.
,
Rochais
,
F.
,
Jurevicius
,
J.
,
Leroy
,
J.
et al
. (
2006
)
Compartmentation of cyclic nucleotide signaling in the heart: the role of cyclic nucleotide phosphodiesterases
.
Circ. Res.
99
,
816
828
https://doi.org/10.1161/01.RES.0000246118.98832.04
84.
Barnes
,
A.P.
,
Livera
,
G.
,
Huang
,
P.
,
Sun
,
C.
,
O’Neal
,
W.K.
,
Conti
,
M.
et al
. (
2005
)
Phosphodiesterase 4D forms a cAMP diffusion barrier at the apical membrane of the airway epithelium
.
J. Biol. Chem.
280
,
7997
8003
https://doi.org/10.1074/jbc.M407521200
85.
Börner
,
S.
,
Schwede
,
F.
,
Schlipp
,
A.
,
Berisha
,
F.
,
Calebiro
,
D.
,
Lohse
,
M.J.
et al
. (
2011
)
FRET measurements of intracellular cAMP concentrations and cAMP analog permeability in intact cells
.
Nat. Protoc.
6
,
427
438
https://doi.org/10.1038/nprot.2010.198
86.
Mongillo
,
M.
,
McSorley
,
T.
,
Evellin
,
S.
,
Sood
,
A.
,
Lissandron
,
V.
,
Terrin
,
A.
et al
. (
2004
)
Fluorescence resonance energy transfer-based analysis of cAMP dynamics in live neonatal rat cardiac myocytes reveals distinct functions of compartmentalized phosphodiesterases
.
Circ. Res.
95
,
67
75
https://doi.org/10.1161/01.RES.0000134629.84732.11
87.
Averaimo
,
S.
,
Assali
,
A.
,
Ros
,
O.
,
Couvet
,
S.
,
Zagar
,
Y.
,
Genescu
,
I.
et al
. (
2016
)
A plasma membrane microdomain compartmentalizes ephrin-generated cAMP signals to prune developing retinal axon arbors
.
Nat. Commun.
7
,
12896
https://doi.org/10.1038/ncomms12896
88.
Rich
,
T.C.
,
Fagan
,
K.A.
,
Tse
,
T.E.
,
Schaack
,
J.
,
Cooper
,
D.M.F.
and
Karpen
,
J.W
. (
2001
)
A uniform extracellular stimulus triggers distinct cAMP signals in different compartments of a simple cell
.
Proc. Natl. Acad. Sci. U.S.A
98
,
13049
13054
https://doi.org/10.1073/pnas.221381398
89.
Houslay
,
M.D.
,
Schafer
,
P.
and
Zhang
,
K.Y.J
. (
2005
)
Keynote review: phosphodiesterase-4 as a therapeutic target
.
Drug Discov. Today
10
,
1503
1519
https://doi.org/10.1016/S1359-6446(05)03622-6
90.
Baillie
,
G.S
. (
2009
)
Compartmentalized signalling: spatial regulation of cAMP by the action of compartmentalized phosphodiesterases
.
FEBS J.
276
,
1790
1799
https://doi.org/10.1111/j.1742-4658.2009.06926.x
91.
Houslay
,
M.D
. (
2010
)
Underpinning compartmentalised cAMP signalling through targeted cAMP breakdown
.
Trends Biochem. Sci.
35
,
91
100
https://doi.org/10.1016/j.tibs.2009.09.007
92.
Perera
,
R.K.
and
Nikolaev
,
V.O
. (
2013
)
Compartmentation of cAMP signalling in cardiomyocytes in health and disease
.
Acta Physiol. (Oxf).
207
,
650
662
https://doi.org/10.1111/apha.12077
93.
Lefkimmiatis
,
K.
and
Zaccolo
,
M
. (
2014
)
cAMP signaling in subcellular compartments
.
Pharmacol. Ther.
143
,
295
304
https://doi.org/10.1016/j.pharmthera.2014.03.008
94.
Kokkonen
,
K.
and
Kass
,
D.A
. (
2017
)
Nanodomain regulation of cardiac cyclic nucleotide signaling by phosphodiesterases
.
Annu. Rev. Pharmacol. Toxicol.
57
,
455
479
https://doi.org/10.1146/annurev-pharmtox-010716-104756
95.
Chay
,
A.
,
Zamparo
,
I.
,
Koschinski
,
A.
,
Zaccolo
,
M.
and
Blackwell
,
K.T
. (
2016
)
Control of βAR- and N-methyl-D-aspartate (NMDA) receptor-dependent cAMP dynamics in hippocampal neurons
.
Plos Comput. Biol.
12
, e1004735 https://doi.org/10.1371/journal.pcbi.1004735
96.
Oliveira
,
R.F.
,
Terrin
,
A.
,
Di Benedetto
,
G.
,
Cannon
,
R.C.
,
Koh
,
W.
,
Kim
,
M.
et al
. (
2010
)
The role of type 4 phosphodiesterases in generating microdomains of cAMP: large scale stochastic simulations
.
Plos One
5
, e11725 https://doi.org/10.1371/journal.pone.0011725
97.
Saucerman
,
J.J.
,
Greenwald
,
E.C.
and
Polanowska-Grabowska
,
R
. (
2014
)
Mechanisms of cyclic AMP compartmentation revealed by computational models
.
J. Gen. Physiol.
143
,
39
48
https://doi.org/10.1085/jgp.201311044
98.
Lohse
,
C.
,
Bock
,
A.
,
Maiellaro
,
I.
,
Hannawacker
,
A.
,
Schad
,
L.R.
,
Lohse
,
M.J.
et al
. (
2017
)
Experimental and mathematical analysis of cAMP nanodomains
.
Plos One
12
, e0174856 https://doi.org/10.1371/journal.pone.0174856
99.
Koschinski
,
A.
and
Zaccolo
,
M
. (
2017
)
Activation of PKA in cell requires higher concentration of cAMP than in vitro: implications for compartmentalization of cAMP signalling
.
Sci. Rep.
7
, 14090 https://doi.org/10.1038/s41598-017-13021-y
100.
Richards
,
M.
,
Lomas
,
O.
,
Jalink
,
K.
,
Ford
,
K.L.
,
Vaughan-Jones
,
R.D.
,
Lefkimmiatis
,
K.
et al
. (
2016
)
Intracellular tortuosity underlies slow cAMP diffusion in adult ventricular myocytes
.
Cardiovasc. Res.
110
,
395
407
https://doi.org/10.1093/cvr/cvw080
101.
Beavo
,
J.A.
and
Brunton
,
L.L
. (
2002
)
Cyclic nucleotide research -- still expanding after half a century
.
Nat. Rev. Mol. Cell Biol.
3
,
710
718
https://doi.org/10.1038/nrm911
102.
Halls
,
M.L.
and
Cooper
,
D.M.F
. (
2010
)
Sub-picomolar relaxin signalling by a pre-assembled RXFP1, AKAP79, AC2, beta-arrestin 2, PDE4D3 complex
.
EMBO J.
29
,
2772
2787
https://doi.org/10.1038/emboj.2010.168
103.
Richter
,
W.
,
Mika
,
D.
,
Blanchard
,
E.
,
Day
,
P.
and
Conti
,
M
. (
2013
)
β1-adrenergic receptor antagonists signal via PDE4 translocation
.
EMBO Rep.
14
,
276
283
https://doi.org/10.1038/embor.2013.4
104.
Zhang
,
J.Z.
,
Lu
,
T.W.
,
Stolerman
,
L.M.
,
Tenner
,
B.
,
Yang
,
J.R.
,
Zhang
,
J.F
, et al
. (
2020
)
Phase separation of a PKA regulatory subunit controls cAMP compartmentation and oncogenic signaling
.
Cell
182
,
1531
1544
https://doi.org/10.1016/j.cell.2020.07.043
105.
Folkmanaite
,
M.
and
Zaccolo
,
M
. (
2025
)
Compartmentalisation in cAMP signalling: a phase separation perspective
.
Preprints
. https://doi.org/10.22541/au.174232810.05470792/v1
106.
Honeyman
,
J.N.
,
Simon
,
E.P.
,
Robine
,
N.
,
Chiaroni-Clarke
,
R.
,
Darcy
,
D.G.
,
Lim
,
I.I.P.
et al
. (
2014
)
Detection of a recurrent DNAJB1-PRKACA chimeric transcript in fibrolamellar hepatocellular carcinoma
.
Science
343
,
1010
1014
https://doi.org/10.1126/science.1249484
107.
Krishnamurthy
,
S.
,
Moorthy
,
B.S.
,
Liqin
,
L.
and
Anand
,
G.S
. (
2013
)
Dynamics of phosphodiesterase-induced cAMP dissociation from protein kinase A: capturing transient ternary complexes by HDXMS
.
Biochim. Biophys. Acta
1834
,
1215
1221
https://doi.org/10.1016/j.bbapap.2013.02.028
108.
Krishnamurthy
,
S.
,
Moorthy
,
B.S.
,
Xin Xiang
,
L.
,
Xin Shan
,
L.
,
Bharatham
,
K.
,
Tulsian
,
N.K
, et al
. (
2014
)
Active site coupling in PDE:PKA complexes promotes resetting of mammalian cAMP signaling
.
Biophys. J.
107
,
1426
1440
https://doi.org/10.1016/j.bpj.2014.07.050
109.
Moorthy
,
B.S.
,
Gao
,
Y.
and
Anand
,
G.S
. (
2011
)
Phosphodiesterases catalyze hydrolysis of cAMP-bound to regulatory subunit of protein kinase A and mediate signal termination
.
Mol. Cell Proteomics
10
, M110.002295 https://doi.org/10.1074/mcp.M110.002295
110.
Spivey
,
H.O.
and
Ovádi
,
J
. (
1999
)
Substrate channeling
.
Methods
19
,
306
321
https://doi.org/10.1006/meth.1999.0858
111.
Krishnamurthy
,
S.
,
Tulsian
,
N.K.
,
Chandramohan
,
A.
and
Anand
,
G.S
. (
2015
)
Parallel allostery by cAMP and PDE coordinates activation and termination phases in cAMP signaling
.
Biophys. J.
109
,
1251
1263
https://doi.org/10.1016/j.bpj.2015.06.067
112.
Houslay
,
M.D
. (
1998
)
Adaptation in cyclic AMP signalling processes: a central role for cyclic AMP phosphodiesterases
.
Semin. Cell Dev. Biol.
9
,
161
167
https://doi.org/10.1006/scdb.1997.0221
113.
Zhang
,
X.
,
Bedigian
,
A.V.
,
Wang
,
W.
and
Eggert
,
U.S
. (
2012
)
G protein‐coupled receptors participate in cytokinesis
.
Cytoskeleton (Hoboken).
69
,
810
818
https://doi.org/10.1002/cm.21055
114.
O’Malley
,
K.L.
,
Jong
,
Y.J.I.
,
Gonchar
,
Y.
,
Burkhalter
,
A.
and
Romano
,
C
. (
2003
)
Activation of metabotropic glutamate receptor mGlu5 on nuclear membranes mediates intranuclear Ca2+ changes in heterologous cell types and neurons
.
J. Biol. Chem.
278
,
28210
28219
https://doi.org/10.1074/jbc.M300792200
115.
Puthenveedu
,
M.A.
,
Lauffer
,
B.
,
Temkin
,
P.
,
Vistein
,
R.
,
Carlton
,
P.
,
Thorn
,
K.
et al
. (
2010
)
Sequence-dependent sorting of recycling proteins by actin-stabilized endosomal microdomains
.
Cell
143
,
761
773
https://doi.org/10.1016/j.cell.2010.10.003
116.
Bowman
,
S.L.
and
Puthenveedu
,
M.A
. (
2015
)
Postendocytic sorting of adrenergic and opioid receptors: new mechanisms and functions
.
Prog. Mol. Biol. Transl. Sci.
132
,
189
206
https://doi.org/10.1016/bs.pmbts.2015.03.005
117.
Hanyaloglu
,
A.C.
and
Von Zastrow
,
M.
. (
2008
)
Regulation of GPCRs by endocytic membrane trafficking and its potential implications
.
Annu. Rev. Pharmacol. Toxicol.
48
,
537
568
https://doi.org/10.1146/annurev.pharmtox.48.113006.094830
118.
Laporte
,
S.A.
,
Oakley
,
R.H.
,
Zhang
,
J.
,
Holt
,
J.A.
,
Ferguson
,
S.S.G.
,
Caron
,
M.G.
et al
. (
1999
)
The β 2 -adrenergic receptor/βarrestin complex recruits the clathrin adaptor AP-2 during endocytosis
.
Proc. Natl. Acad. Sci. U.S.A
96
,
3712
3717
https://doi.org/10.1073/pnas.96.7.3712
119.
Weinberg
,
Z.Y.
and
Puthenveedu
,
M.A
. (
2019
)
Regulation of G protein‐coupled receptor signaling by plasma membrane organization and endocytosis
.
Traffic
20
,
121
129
https://doi.org/10.1111/tra.12628
120.
Lefkowitz
,
R.J.
,
Wessels
,
M.R.
and
Stadel
,
J.M
. (
1980
)
Hormones, receptors, and cyclic AMP: their role in target cell refractoriness
.
Curr. Top. Cell. Regul.
17
,
205
230
https://doi.org/10.1016/b978-0-12-152817-1.50011-0
121.
Hertel
,
C.
and
Perkins
,
J.P
. (
1984
)
Receptor-specific mechanisms of desensitization of beta-adrenergic receptor function
.
Mol. Cell. Endocrinol.
37
,
245
256
https://doi.org/10.1016/0303-7207(84)90094-7
122.
Crilly
,
S.E.
and
Puthenveedu
,
M.A
. (
2021
)
Compartmentalized GPCR signaling from intracellular membranes
.
J. Membr. Biol.
254
,
259
271
https://doi.org/10.1007/s00232-020-00158-7
123.
Lyga
,
S.
,
Volpe
,
S.
,
Werthmann
,
R.C.
,
Götz
,
K.
,
Sungkaworn
,
T.
,
Lohse
,
M.J.
et al
. (
2016
)
Persistent cAMP signaling by internalized LH receptors in ovarian follicles
.
Endocrinology
157
,
1613
1621
https://doi.org/10.1210/en.2015-1945
124.
Chen
,
K.E.
,
Healy
,
M.D.
and
Collins
,
B.M
. (
2019
)
Towards a molecular understanding of endosomal trafficking by Retromer and Retriever
.
Traffic
20
,
465
478
https://doi.org/10.1111/tra.12649
125.
Rojas
,
R.
,
Kametaka
,
S.
,
Haft
,
C.R.
and
Bonifacino
,
J.S
. (
2007
)
Interchangeable but essential functions of SNX1 and SNX2 in the association of retromer with endosomes and the trafficking of mannose 6-phosphate receptors
.
Mol. Cell. Biol.
27
,
1112
1124
https://doi.org/10.1128/MCB.00156-06
126.
Vagnozzi
,
A.N.
and
Praticò
,
D
. (
2019
)
Endosomal sorting and trafficking, the retromer complex and neurodegeneration
.
Mol. Psychiatry
24
,
857
868
https://doi.org/10.1038/s41380-018-0221-3
127.
Bonifacino
,
J.S.
and
Rojas
,
R
. (
2006
)
Retrograde transport from endosomes to the trans-Golgi network
.
Nat. Rev. Mol. Cell Biol.
7
,
568
579
https://doi.org/10.1038/nrm1985
128.
Collins
,
B.M.
,
Norwood
,
S.J.
,
Kerr
,
M.C.
,
Mahony
,
D.
,
Seaman
,
M.N.J.
,
Teasdale
,
R.D.
et al
. (
2008
)
Structure of Vps26B and mapping of its interaction with the retromer protein complex
.
Traffic
9
,
366
379
https://doi.org/10.1111/j.1600-0854.2007.00688.x
129.
Collins
,
B.M
. (
2008
)
The structure and function of the retromer protein complex
.
Traffic
9
,
1811
1822
https://doi.org/10.1111/j.1600-0854.2008.00777.x
130.
Feinstein
,
T.N.
,
Wehbi
,
V.L.
,
Ardura
,
J.A.
,
Wheeler
,
D.S.
,
Ferrandon
,
S.
,
Gardella
,
T.J.
et al
. (
2011
)
Retromer terminates the generation of cAMP by internalized PTH receptors
.
Nat. Chem. Biol.
7
,
278
284
https://doi.org/10.1038/nchembio.545
131.
Godbole
,
A.
,
Lyga
,
S.
,
Lohse
,
M.J.
and
Calebiro
,
D
. (
2017
)
Internalized TSH receptors en route to the TGN induce local Gs-protein signaling and gene transcription
.
Nat. Commun.
8
, 443 https://doi.org/10.1038/s41467-017-00357-2
132.
Dong
,
C.
,
Filipeanu
,
C.M.
,
Duvernay
,
M.T.
and
Wu
,
G
. (
2007
)
Regulation of G protein-coupled receptor export trafficking
.
Biochimica et Biophysica Acta (BBA) - Biomembranes
1768
,
853
870
https://doi.org/10.1016/j.bbamem.2006.09.008
133.
Irannejad
,
R.
,
Pessino
,
V.
,
Mika
,
D.
,
Huang
,
B.
,
Wedegaertner
,
P.B.
,
Conti
,
M.
et al
. (
2017
)
Functional selectivity of GPCR-directed drug action through location bias
.
Nat. Chem. Biol.
13
,
799
806
https://doi.org/10.1038/nchembio.2389
134.
Shiwarski
,
D.J.
,
Darr
,
M.
,
Telmer
,
C.A.
,
Bruchez
,
M.P.
and
Puthenveedu
,
M.A
. (
2017
)
PI3K class II α regulates δ-opioid receptor export from the trans -Golgi network
.
Mol. Biol. Cell
28
,
2202
2219
https://doi.org/10.1091/mbc.e17-01-0030
135.
Mittal
,
N.
,
Roberts
,
K.
,
Pal
,
K.
,
Bentolila
,
L.A.
,
Fultz
,
E.
,
Minasyan
,
A
, et al
. (
2013
)
Select G-protein-coupled receptors modulate agonist-induced signaling via a ROCK, LIMK, and β-arrestin 1 pathway
.
Cell Rep.
5
,
1010
1021
https://doi.org/10.1016/j.celrep.2013.10.015
136.
Koliwer
,
J.
,
Park
,
M.
,
Bauch
,
C.
,
Von Zastrow
,
M.
and
Kreienkamp
,
H.J
. (
2015
)
The golgi-associated PDZ domain protein PIST/GOPC stabilizes the β1-adrenergic receptor in intracellular compartments after internalization
.
J. Biol. Chem.
290
,
6120
6129
https://doi.org/10.1074/jbc.M114.605725
137.
St-Louis
,
É.
,
Degrandmaison
,
J.
,
Grastilleur
,
S.
,
Génier
,
S.
,
Blais
,
V.
,
Lavoie
,
C
, et al
. (
2017
)
Involvement of the coatomer protein complex I in the intracellular traffic of the delta opioid receptor
.
Mol. Cell. Neurosci.
79
,
53
63
https://doi.org/10.1016/j.mcn.2016.12.005
138.
Shiwarski
,
D.J.
,
Tipton
,
A.
,
Giraldo
,
M.D.
,
Schmidt
,
B.F.
,
Gold
,
M.S.
,
Pradhan
,
A.A.
et al
. (
2017
)
A PTEN-regulated checkpoint controls surface delivery of δ opioid receptors
.
J. Neurosci.
37
,
3741
3752
https://doi.org/10.1523/JNEUROSCI.2923-16.2017
139.
Boivin
,
B.
,
Lavoie
,
C.
,
Vaniotis
,
G.
,
Baragli
,
A.
,
Villeneuve
,
L.
,
Ethier
,
N.
et al
. (
2006
)
Functional β-adrenergic receptor signalling on nuclear membranes in adult rat and mouse ventricular cardiomyocytes
.
Cardiovasc. Res.
71
,
69
78
https://doi.org/10.1016/j.cardiores.2006.03.015
140.
Di Benedetto
,
A.
,
Sun
,
L.
,
Zambonin
,
C.G.
,
Tamma
,
R.
,
Nico
,
B.
,
Calvano
,
C.D.
et al
. (
2014
)
Osteoblast regulation via ligand-activated nuclear trafficking of the oxytocin receptor
.
Proc. Natl. Acad. Sci. U.S.A
111
,
16502
16507
https://doi.org/10.1073/pnas.1419349111
141.
Wright
,
C.D.
,
Wu
,
S.C.
,
Dahl
,
E.F.
,
Sazama
,
A.J.
and
O’Connell
,
T.D
. (
2012
)
Nuclear localization drives α1-adrenergic receptor oligomerization and signaling in cardiac myocytes
.
Cell. Signal.
24
,
794
802
142.
Joyal
,
J.S.
,
Nim
,
S.
,
Zhu
,
T.
,
Sitaras
,
N.
,
Rivera
,
J.C.
,
Shao
,
Z.
et al
. (
2014
)
Subcellular localization of coagulation factor II receptor-like 1 in neurons governs angiogenesis
.
Nat. Med.
20
,
1165
1173
https://doi.org/10.1038/nm.3669
143.
Bhosle
,
V.K.
,
Rivera
,
J.C.
,
Zhou
,
T.E.
,
Omri
,
S.
,
Sanchez
,
M.
,
Hamel
,
D.
et al
. (
2016
)
Nuclear localization of platelet-activating factor receptor controls retinal neovascularization
.
Cell Discov.
2
,
16017
https://doi.org/10.1038/celldisc.2016.17
144.
Brock
,
C.
,
Boudier
,
L.
,
Maurel
,
D.
,
Blahos
,
J.
and
Pin
,
J.P
. (
2005
)
Assembly-dependent surface targeting of the heterodimeric GABAB Receptor is controlled by COPI but not 14-3-3
.
Mol. Biol. Cell
16
,
5572
5578
https://doi.org/10.1091/mbc.e05-05-0400
145.
Cunningham
,
M.R.
,
McIntosh
,
K.A.
,
Pediani
,
J.D.
,
Robben
,
J.
,
Cooke
,
A.E.
,
Nilsson
,
M.
et al
. (
2012
)
Novel role for proteinase-activated receptor 2 (PAR2) in membrane trafficking of proteinase-activated receptor 4 (PAR4)
.
J. Biol. Chem.
287
,
16656
16669
https://doi.org/10.1074/jbc.M111.315911
146.
Shiwarski
,
D.J.
,
Crilly
,
S.E.
,
Dates
,
A.
and
Puthenveedu
,
M.A
. (
2019
)
Dual RXR motifs regulate nerve growth factor–mediated intracellular retention of the delta opioid receptor
.
Mol. Biol. Cell
30
,
680
690
https://doi.org/10.1091/mbc.E18-05-0292
147.
Sergin
,
I.
,
Jong
,
Y.J.I.
,
Harmon
,
S.K.
,
Kumar
,
V.
and
O’Malley
,
K.L
. (
2017
)
Sequences within the C terminus of the metabotropic glutamate receptor 5 (mGluR5) are responsible for inner nuclear membrane localization
.
J. Biol. Chem.
292
,
3637
3655
https://doi.org/10.1074/jbc.M116.757724
148.
Jong
,
Y.-J.I.
,
Kumar
,
V.
and
O’Malley
,
K.L
. (
2009
)
Intracellular metabotropic glutamate receptor 5 (mGluR5) activates signaling cascades distinct from cell surface counterparts
.
J. Biol. Chem.
284
,
35827
35838
https://doi.org/10.1074/jbc.M109.046276
This is an open access article published by Portland Press Limited on behalf of the Biochemical Society and distributed under the Creative Commons Attribution License 4.0 (CC BY).