Home Life Sciences Towards improved receptor targeting: anterograde transport, internalization and postendocytic trafficking of neuropeptide Y receptors
Article Publicly Available

Towards improved receptor targeting: anterograde transport, internalization and postendocytic trafficking of neuropeptide Y receptors

  • Stefanie Babilon

    Stefanie Babilon was born in 1985 in Giessen, Germany. She studied Biochemistry at the University of Leipzig in a bachelor’s-master’s program. In 2010, she performed her master thesis under the supervision of Prof. Dr. Annette G. Beck-Sickinger, during which she spent a 3-month research internship at the Vanderbilt University. After her graduation she stayed in the research group of Prof. Dr. Annette G. Beck-Sickinger for her PhD, which focuses on the internalization and postendocytic trafficking mechanisms of human Y receptors. Since 2011, she is supported by a scholarship of the Fonds der Chemischen Industrie.

    , Karin Mörl

    Karin Mörl was born 1967 in Munich and studied Biology at the Ludwig-Maximilians-University, Munich. She graduated from the Max-Planck-Institute for Neurobiology in Martinsried, Germany under the supervision of Prof. Dr. Hans Thoenen and has been working as a postdoctoral researcher with Dr. Michael Meyer in Martinsried and Prof. Dr. Volker Bigl in Leipzig. In 2001 she joined the group of Prof. Dr. Annette G. Beck-Sickinger at the Institute of Biochemistry, University of Leipzig as a staff scientist.

    and Annette G. Beck-Sickinger

    Annette Beck-Sickinger studied Chemistry (Diploma 1986) and Biology (Diploma 1990) at the Eberhard Karls University in Tübingen and graduated with G. Jung in 1989. After fellowships in Zürich (Carafoli), Scripps La Jolla (Houghten), and Copenhagen (Schwartz) she became Assistant Professor of Pharmacology and Biochemistry at ETH Zürich. In 1999 she accepted the full professorship at Leipzig University in the Departmant of Bioorganic Chemistry and Biochemistry. In 2009 she was visiting professor at Vanderbilt University, in Nashville.

    EMAIL logo
Published/Copyright: March 1, 2013

Abstract

The neuropeptide Y system is known to be involved in the regulation of many central physiological and pathophysiological processes, such as energy homeostasis, obesity, cancer, mood disorders and epilepsy. Four Y receptor subtypes have been cloned from human tissue (hY1, hY2, hY4 and hY5) that form a multiligand/multireceptor system together with their three peptidic agonists (NPY, PYY and PP). Addressing this system for medical application requires on the one hand detailed information about the receptor-ligand interaction to design subtype-selective compounds. On the other hand comprehensive knowledge about alternative receptor signaling, as well as desensitization, localization and downregulation is crucial to circumvent the development of undesired side-effects and drug resistance. By bringing such knowledge together, highly potent and long-lasting drugs with minimized side-effects can be engineered. Here, current knowledge about Y receptor export, internalization, recycling, and degradation is summarized, with a focus on the human Y receptor subtypes, and is discussed in terms of its impact on therapeutic application.

Introduction

G protein-coupled receptors (GPCR) represent the largest family of cell surface receptors and are responsible for the majority of signal transduction across cell membranes (Millar and Newton, 2010). Thus, it is not surprising that nearly half of the pharmaceuticals prescribed worldwide act on those receptors (Bridges and Lindsley, 2008). Extensive investigations on GPCR have revealed that the sensitive balance between receptor activation and desensitization is crucial to maintain cellular homeostasis. One component of this balancing act is the agonist-driven desensitization, internalization and recycling process of GPCR. Understanding the mechanism underlying these processes tremendously enlarges the scope of clinical intervention, when GPCR are used as a drug target. One group of GPCR that is being focused on in current investigations are the neuropeptide Y receptors (Y receptors). Besides their very prominent role in food intake and cancer (Korner and Reubi, 2007; Nguyen et al., 2011; Zhang et al., 2011), further involvement of these receptors in ischemic diseases, reproduction, memory retention (Flood et al., 1989), epilepsy (Vezzani and Sperk, 2004; El Bahh et al., 2005), and mood disorders (Heilig, 2004) has been reported during the past few decades. Consequently, these GPCR are extraordinarily interesting in terms of medical applications.

Y receptors belong to the group of rhodopsin-like GPCR and act preferentially via pertussis toxin-sensitive, heterotrimeric Gi/o proteins (Michel et al., 1998). Like all GPCR, the Y receptors consist of seven transmembrane-spanning helices, three intra- and extracellular loops (ECLs), and an external amino(N-)- and an internal carboxy(C-)-terminus. Their activation leads to the inhibition of the adenylate cyclase, the modulation of potassium and calcium channels and, in some cell types, to the phosphorylation of p44/42 mitogen-activated protein kinases (MAPKs) and their downstream effectors (Howell et al., 2005; Lu et al., 2010; Thiriet et al., 2011; Shimada et al., 2012). Five Y receptors have been cloned from mammals, namely Y1, Y2, Y4, Y5 and y6. The last, y6, is not expressed in primates due to a frame-shift mutation resulting in a truncated nonfunctional protein (Rose et al., 1997). Y receptors can be found in various tissues including the central nervous system, blood vessels, intestine and kidneys, where they convey various physiological effects.

The endogenous ligands of Y receptors are the closely-related neuropeptide Y (NPY), peptide YY (PYY) and the pancreatic polypeptide (PP). NPY is the most abundantly expressed neuropeptide in the brain, whereas PYY and PP are gut-derived hormones. All three members of the NPY family consist of a 36-amino-acid(aa)-long peptide chain with an amidated C-terminus. The spatial arrangement of the NPY family peptides was long thought to be a hairpin-like so-called PP-fold, as determined from the crystal structure of avian PP in 1981 (Blundell et al., 1981). This structure comprises an N-terminal polyproline helix (residues 1–8), a consecutive turn and a C-terminal α-helix (residues 14–31) arranged in a U-shaped tertiary structure. Nuclear magnetic resonance (NMR) studies in the presence of membrane-mimicking dodecylphosphocholine confirmed the formation of the C-terminal α-helical structure. However, those studies revealed that the N-termini of all three peptides are more flexible and do not form the hairpin-like fold under physiological conditions (Bader et al., 2001; Lerch et al., 2002, 2004) (Figure 1). Notably, the peptides were found to pre-associate with the membrane via their amphipathic α-helix, and for the PP also with its N-terminus, presumably to get into the correct orientation for specific Y receptor subtype binding (Lerch et al., 2002; Thomas et al., 2005). Y receptors have unique but overlapping ligand binding profiles, thereby constituting a multiligand/multireceptor system. NPY and PYY are both bound by the Y1, Y2 and Y5 receptors with relatively high affinities (Cabrele and Beck-Sickinger, 2000; Lindner et al., 2008a). In contrast, PP, which evolved most recently, binds predominantly to the Y4, and with lower but still nanomolar affinity to the Y5 receptor. Moreover, truncation of NPY and PYY by dipeptidyl peptidase IV leads to Y2/ Y5 selective agonists, namely NPY336 and PYY336 (Mentlein et al., 1993; Borowsky et al., 1998).

Figure 1 (A) Primary and (B) tertiary structure of NPY, PYY and PP, as determined by nuclear magnetic resonance in the presence of membrane mimicking dodecylphosphocholine-micelles (PDB ID: pNPY: 1F8P, pPYY: 1RUU, bPP: 1LJV).
Figure 1

(A) Primary and (B) tertiary structure of NPY, PYY and PP, as determined by nuclear magnetic resonance in the presence of membrane mimicking dodecylphosphocholine-micelles (PDB ID: pNPY: 1F8P, pPYY: 1RUU, bPP: 1LJV).

Targeting such a multiligand/multireceptor system for clinical application is challenging because one distinct Y receptor subtype has to be addressed specifically to favor the desired outcome while minimizing side effects that might occur from the co-activation of other Y receptor subtypes. In order to unravel the similarities and differences in Y receptor structures, and more precisely in their binding pockets, chimeric receptor approaches, peptide alanine-scans (Pedragosa Badia et al., 2013), recombinant receptor expression (Schmidt et al., 2009) and NMR studies were performed. Several peptide residues were identified as being involved in receptor binding, while determination of the receptor counterpart was more difficult. The first direct interaction sites between Y receptors and their ligands were determined by Merten et al. (2007). In those studies, it was shown that the aspartic acid6.59 residue within the ECL3 of Y1 and Y4 receptor forms a salt-bridge with the arginine35 of NPY and PP. A similar interaction was confirmed for aspartic acid6.59 in the Y2 and Y5, however, the counterpart within NPY was found to be arginine33 instead of arginine35. In addition, Y2 and Y5 receptor binding required the presence of arginine25 within the peptide ligands, while Y1 and Y4 were insensitive to the alanine mutation of arginine25. Shortly after, the counterpart of arginine25 was identified to be aspartic acid2.68 in the ECL2 of the Y5 receptor (Lindner et al., 2008b). These studies point towards distinct binding modes for the Y2/Y5 and Y1/Y4 receptors. Based on this knowledge, improved subtype selective peptides as well as small non-peptidic agonists and antagonists have been developed. These ligands can specifically modify receptor signal transduction in order to favor a physiological response (Figure 2).

Figure 2 Internalization and postendocytic trafficking of G protein-coupled receptors.Internalization of radio-ligands enables the cell-specific targeting of tumor cells (left panel). Likewise, the receptor function can be influenced by different ligands, thereby modulating the cellular response (right panel). While classic antagonists do not promote internalization, agonist and functional antagonist lead to receptor endocytosis mediated by arrestin and rab5. Once internalized, the receptor–ligand complex is either degraded in the lysosome or recycled back passing the rab4- (direct recycling) or rab11-dependent (indirect recycling) pathway. Arr, arrestin; EE, early endosome; RE, recycling endosome; SE, sorting endosome.
Figure 2

Internalization and postendocytic trafficking of G protein-coupled receptors.

Internalization of radio-ligands enables the cell-specific targeting of tumor cells (left panel). Likewise, the receptor function can be influenced by different ligands, thereby modulating the cellular response (right panel). While classic antagonists do not promote internalization, agonist and functional antagonist lead to receptor endocytosis mediated by arrestin and rab5. Once internalized, the receptor–ligand complex is either degraded in the lysosome or recycled back passing the rab4- (direct recycling) or rab11-dependent (indirect recycling) pathway. Arr, arrestin; EE, early endosome; RE, recycling endosome; SE, sorting endosome.

Peptide ligands, including the endogenous ligands of Y receptors, and some organic agonists or antagonists are not cell permeable; thus modulation of Y receptor activity with any ligand requires initial receptor availability at the cell surface. This is not only determined by the expression levels of GPCR but also by trafficking processes that comprise:

  • the anterograde transport of the receptor to the cell membrane;

  • the internalization; and

  • the postendocytic trafficking that may lead to receptor recycling or degradation.

Usually, these receptor trafficking processes are tightly regulated by a mutual interplay of receptor conformation and consensus sequences on the one site and accessory proteins on the other site. Remarkably, interaction partners involved in receptor internalization were also found to alter GPCR signaling pathways by serving as a scaffold for further enzymes and regulators. This clearly shows an interdependence of receptor localization and signaling following agonist-induced activation and endocytosis. Key players in this phenomenon are G protein-coupled receptor kinases (visual GRK1 and GRK7, non-visual GRK2-6) and arrestins (Arr; visual Arr1 and Arr4, non-visual Arr2 and Arr3) (Gurevich and Gurevich, 2006; Gurevich et al., 2012; Magalhaes et al., 2012). These proteins usually bind to the receptor subsequent to its activation and G protein-dependent signaling and mediate receptor desensitization, endocytosis and alternative G protein-independent signaling pathways. The observation of an Arr-induced alternative signaling pathway led to the identification of a new class of receptor ligands, named biased ligands. Such compounds can trigger a discrete receptor conformation that either corresponds to G protein activation or Arr binding and consecutive internalization. In the latter case, the receptor is removed from the cell surface independent from G protein signaling, which is referred to as functional antagonism (Figure 2).

Besides mediating receptor activation and localization, GPCR trafficking can also be used as a shuttling system when receptors are specifically addressed and overexpressed in cells that are related to cancer. The underlying mechanism is the co-internalization of a ligand with its receptor subsequent to its binding and activation. Thus, radioactive or cytotoxic ligands can be applied that accumulate within the tumor cells subsequent to receptor stimulation and internalization. This enables improved diagnosis and therapy of distinct types of cancer and has been successfully shown for several GPCR, including the bombesin (gastrin-releasing peptide) and the Y1 receptor (Santos-Cuevas et al., 2008; Khan et al., 2010) (Figure 2).

Contrary to the benefits of using GPCR trafficking as a shuttling system, postponed export or simultaneous internalization and degradation of the receptor may lead to drug resistance when Y receptors are addressed for clinical intervention. Consequently, detailed knowledge about GPCR trafficking is essential for the design of highly effective diagnostic and therapeutic tools that do not only modulate classic Y receptor subtype activation but also control receptor localization and signaling pathways. Therefore, molecular mechanisms that are responsible for membrane targeting, internalization and the recycling of Y receptors were investigated in detail to contribute to a better understanding of how to use these receptors as drug targets. Here, recent findings on Y receptor trafficking are reviewed with a focus on the human proteins and discussed with respect to their relevance for addressing those receptors in therapeutic applications.

Human Y receptors

Y1 receptor

The human Y1 (hY1) receptor consists of 384 aa and shows a sequence identity of more than 92% with its orthologs expressed in pig, guinea pig, mouse, rat and Bos taurus. Generally, the sequence identity among the Y receptor subtypes is rather low, but is highest for the Y1 and the Y4 receptor. Hence, hY1 and hY4 (and y6 in other mammals) form one Y receptor superfamily, while hY2 and hY5 each form an autonomous group (Larhammar and Salaneck, 2004). Tissues in which the hY1 receptor can be found are predominantly in the brain (dentate gyrus of the hippocampus) (Dumont et al., 1998) but also in the vascular smooth muscle cells, adipose tissue, kidney and the gastrointestinal tract (Michel et al., 1998). The most prominent role of this Y1 receptor subtype is the induction of food intake and the regulation of energy homeostasis in synergy with Y5 (Nguyen et al., 2012). Moreover, the Y1 receptor modulates vasoconstriction (Hodges et al., 2009) and conveys neuroregenerative effects in vitro and in vivo (Howell et al., 2005; Thiriet et al., 2011), as well as antidepressant and anxiolytic effects in rodents (Wahlestedt et al., 1993; Verma et al., 2012). The hY1 is overexpressed in more than 84% of primary human breast (Reubi et al., 2001) and Ewing sarcoma tumors (Korner et al., 2008), representing a promising diagnostic and therapeutic target for these types of cancer.

Y2 receptor

The hY2 receptor is a 381-aa-protein that is expressed predominately in neuronal tissue, but also in the spleen, liver, blood vessels, gastrointestinal tract, white and brown fat tissue. This receptor shares a sequence identity of more than 92% with its orthologs in mouse, Bos taurus, pig and guinea pig. In the brain, activation of presynaptically-expressed Y2 receptors inhibits neurotransmitter release (Klapstein and Colmers, 1993), including NPY and glutamate, which makes this receptor an interesting target for antiepileptic drugs (Vezzani and Sperk, 2004). Furthermore, the Y2 conveys anorexigenic effects in mice, rats and humans (Batterham et al., 2002) and is involved in memory retention, mood disorders (Verma et al., 2012), angiogenesis (Lee et al., 2003), and the reward system following alcohol consumption (Hayes et al., 2012). In addition, the Y2 receptor is overexpressed in glioblastoma and neuroblastoma, where it induces tumor growth and vascularization (Kitlinska et al., 2005; Korner and Reubi, 2008; Lu et al., 2010). Interestingly, it was recently found that Y2 receptors expressed in the central nervous system and in non-neuronal tissues might have different assignments. Only Y2 receptors of the central nervous system regulate bone mass while peripherally-expressed receptors were not found to contribute to bone metabolism at all (Shi et al., 2011).

Y4 receptor

With a sequence identity of <86%, the hY4 receptor has the lowest proportion of identical sequence compared to its orthologs in relation to other Y receptor subtypes. The hY4 receptor consists of 375 aa and is mainly expressed in the gastrointestinal tract and, to a much lesser extent, in the hippocampus, the hypothalamus and the area postrema, a region with an incomplete blood-brain barrier (Dumont et al., 1998; Lindner et al., 2008a; Bellmann-Sickert et al., 2011). There, the Y4 receptor can receive signals from peripherally-circulating PP and NPY, modulating energy homeostasis and emotional behavior in synergy with the Y2 receptor (Lin et al., 2009; Tasan et al., 2009). Moreover, the Y4 receptor inhibits gastric emptying in humans, followed by a delayed postprandial increase in blood glucose levels and subsequent insulin secretion (Schmidt et al., 2005), thereby inducing anorexigenic effects. As the Y4 receptor also has an inhibitory effect on gastrointestinal secretion, it might be a potent target in diarrheal disorders (Tough et al., 2006). Besides this, the Y4 receptor was found to be expressed in several human colon adenocarcinoma cell lines (Cox et al., 2001), pointing towards this receptor having a role in the development of colon cancer in humans.

Y5 receptor

The hY5 receptor exists in two isoforms with similar pharmacological profiles (Rodriguez et al., 2003). The long isoform of the hY5 receptor consists of 455 aa, while the short isoform constitutes a splice variant that lacks the first 10 aa. With respect to the other Y receptor subtypes, the Y5 receptor possesses a relative large third intracellular loop (ICL3) with approximately 140 aa, while the C-terminus is rather short. The Y5 receptor shares a sequence identity with its orthologs of over 85% and is abundantly expressed in the hypothalamus, thalamus, amygdala and temporal cortex (Durkin et al., 2000). It contributes to energy homeostasis and food intake together with Y1 (Nguyen et al., 2012), and mediates anticonvulsant effects (Benmaamar et al., 2005). Moreover, recent findings suggest that the Y5 receptor also contributes to anxiolysis in rats (Morales-Medina et al., 2012). Similar to the Y1 receptor, the Y5 receptor is expressed in distinct human breast cancer cell lines, where it promotes cell growth and migration (Sheriff et al., 2010).

The anterograde transport of Y receptors

GPCR are translated in the rough endoplasmic reticulum (ER) and transported to the cell surface passing the ER–Golgi intermediate complex, the Golgi apparatus and the trans-Golgi network (Wang and Wu, 2012) (Figure 3C). The anterograde transport is a very well regulated process and ER export requires correct folding, optional N-glycosylation and protein assembly of GPCR. Membrane proteins are then packed into COPII vesicles, which require Sar1 GTPase and Sec23/24 proteins to bud off from the ER. This first step in GPCR trafficking is mediated by a variety of distinct receptor sequence-dependent motifs within the proximal C-terminus (dileucine, triple phenylalanine, diacid, and diphenyl motifs) as reviewed by Duvernay et al. (2005). For some receptors, ER export is also controlled by the N-terminus and its glycosylation (Dong and Wu, 2007). Once the receptor is sorted into COPII vesicles, its trafficking from the ER to the membrane is mediated by ras-related in brain (rab) proteins 1, 2, 6, and 8 (Wang and Wu, 2012). These small GTPases associate with their cargo directly or indirectly and take part in vesicle targeting, tethering and fusion.

Figure 3 Regulation of the anterograde transport of human Y receptors.(A) Comparison of N- and C-terminal sequences of human Y receptors. Underlined sequences represent the deleted region in YR(Δ8th helix) and YR(ΔN+8), respectively. In YR(ΔN), the whole N-terminus as presented here was truncated except for the starting methionine. Asparagines that are proposed to be glycosylated as well as residues that might contribute to the formation of the putative eighth helix are written in bold. (B) Representative fluorescence images show the distribution of wild-type and mutant receptors in HEK293 cells (Lindner et al., 2009; Walther et al., 2012). (C) Schematic illustration of the impact of N- and C-terminal deletion on human Y receptor export. Scale bar: 10 μm; ECL2, extra cellular loop 2; ERGIC, ER-Golgi intermediate complex; wt, wild type; YR, human Y receptor.
Figure 3

Regulation of the anterograde transport of human Y receptors.

(A) Comparison of N- and C-terminal sequences of human Y receptors. Underlined sequences represent the deleted region in YR(Δ8th helix) and YR(ΔN+8), respectively. In YR(ΔN), the whole N-terminus as presented here was truncated except for the starting methionine. Asparagines that are proposed to be glycosylated as well as residues that might contribute to the formation of the putative eighth helix are written in bold. (B) Representative fluorescence images show the distribution of wild-type and mutant receptors in HEK293 cells (Lindner et al., 2009; Walther et al., 2012). (C) Schematic illustration of the impact of N- and C-terminal deletion on human Y receptor export. Scale bar: 10 μm; ECL2, extra cellular loop 2; ERGIC, ER-Golgi intermediate complex; wt, wild type; YR, human Y receptor.

Y receptors are predominately expressed at the cell surface, however, the anterograde transport mechanisms responsible were not known until recently. Pioneer studies in our laboratory were performed to shed light on the membrane-targeting of human Y receptors. A combination of truncated and point mutants of human Y receptors in HEK293 cells revealed that the molecular mechanisms of all four human Y receptors seem to differ dramatically (Lindner et al., 2009; Walther et al., 2012) (Figure 3).

Role of C-terminal sequences

With a focus on hY2 receptor, the sequence motif 332Y(x)3F(x)3F340 located at the very proximal C-terminus was identified to be crucial for ER export (Walther et al., 2012). Complete or partial deletion of the motif as well as point mutations from tyrosine/phenylalanine to alanine led to dramatic ER retention of the receptor. Interestingly, transferring the motif to a more distal region of the hY2 receptor C-terminus did not rescue the anterograde transport and the receptor still accumulated in the ER. As the very proximal C-terminus comprises the putative eighth helix, we concluded that the 332Y(x)3F(x)3F340 motif must be essential for its formation, and thus for the correct receptor conformation. This hypothesis was further confirmed by performing a low-temperature (28°C) experiment to rescue receptor folding. Thereby, only receptors with single mutations that affected the ER export motif (tyrosine/phenylalanine→alanine) or the formation of the eighth helix (338S→P) were partly rescued when cells were cultured at 28°C. In contrast, complete deletion as well as relocation of the motif to a more distal C-terminal region still prevented ER export despite lower temperature cultivation. These findings are in good agreement with the function and position of the triple phenylalanine motif of the dopamine D1 receptor, which was found to bind to the ER chaperone DRiP78 (Bermak et al., 2001).

Like the hY2 receptor, the hY4 receptor comprises a comparable motif within the proximal C-terminus, namely 329F(x)3I(x)3V337. Interestingly, deletion of this motif did not result in impaired ER export but prevented transport of the receptor from the Golgi apparatus to the cell surface (Walther et al., 2012). Similarly, the motif of the hY1 receptor 327F(x)3L(x)3F335 resembles the motif of the hY2 and hY4 receptors and is also located within the putative eighth helix. Nonetheless, deletion of this motif had no influence on the cell surface targeting of the hY1 receptor (Walther et al., 2012).

As mentioned above, the hY5 receptor exhibits a short C-terminus and a large ICL3. Thus, it seems reasonable that export motifs or structural determinants might be displayed by the ICL3 instead of the C-terminus. Indeed, deletion of the putative eighth helix of the hY5 receptor C-terminus had no influence on cell surface expression of the receptor as it has been observed for the hY1 receptor (Walther et al., 2012). Replacing the ICL3 of the hY5 with the ICL3 of the hY2 receptor, as well as deletion of large parts of Y5 ICL3 sequences (unpublished results), had no effect on membrane targeting either (Bohme et al., 2008). Thus, it can be suggested that the hY5 receptor displays its export motif in some region other than the C-terminus/ICL3.

With respect to palmitoylation, which is also a common posttranslational modification of GPCR stabilizing the eighth helix, mutation of the Cys within the proximal C-terminus of the Y1 and Y2 receptor led to opposite results. While the palmitoylation motif is crucial for the G protein-coupling and desensitization of the Y1 receptor, mutation of the Cys within the Y2 receptor sequence had no effect on receptor signaling properties or cell surface targeting (Holliday and Cox, 2003; Walther et al., 2012).

Role of the N-terminus and its glycosylation motifs

All human Y receptors have an N-terminus of about 40–50 aa, with at least one putative N-glycosylation site. An additional glycosylation site is postulated within the ECL2 of the hY4 receptor. Generally, N-glycosylation was found to be a prerequisite for the association of chaperons like calnexin and calreticulin, which facilitate GPCR folding in the ER (Dong and Wu, 2007). However, not all GPCR require their N-termini and receptor glycosylation for ER export. The Y1, Y2 and Y4 receptors were reported to be glycosylated, however, the impact of this posttranslational modification on membrane targeting was not addressed (Hansen and Sheikh, 1992; Voisin et al., 2000; Holliday and Cox, 2003). To answer the question of whether the N-termini and their potential glycosylation sites of human Y receptors are important for membrane integration, N-terminally truncated mutants were investigated in our laboratory (Lindner et al., 2009). Interestingly, complete deletion of the hY1 and hY4 receptor N-termini prevented membrane integration. Receptors accumulated in intracellular compartments that resembled the Golgi apparatus (Figure 3B). However, N-terminal extension with arbitrary sequences of at least eight residues rescued membrane targeting, despite the absence of any glycosylation motifs. Here, the N-terminal extension seems to stabilize the overall receptor conformation of the hY1 and hY4 receptors. In contrast, complete deletion of the N-terminus was very well tolerated for the hY2 and hY5 receptors and no N-terminal extension was required for membrane integration. Consequently, it can be speculated that glycosylation does not play any role in the membrane targeting of Y receptors. Nonetheless, it cannot be excluded that the hY4 displays a glycosylation within its ECL2 that might promote cell surface expression.

Receptor oligomerization

It is very likely that the N-terminus or the putative eighth helix contribute to the organization of the correct receptor conformation and thus are essential elements required for cell surface expression. Those structural elements may therefore favor receptor dimerization in the ER and promote transport to the membrane and/or enhance the membrane retention time, as has been shown for other GPCR (Salahpour et al., 2004). However, the role of class A GPCR oligomerization is discussed controversially, since its physiological relevance is still in question (Gurevich and Gurevich, 2008). Oligomerization of GPCR is generally thought to be achieved by three types of interaction:

  • the most stable covalent cysteine-bridge like in metabotropic glutamate receptors;

  • coiled-coil interactions favored by the putative eighth helix, such as in GABAB receptors; and

  • non-covalent contacts between transmembrane helices that might lead to the domain swapping of the helices (Breitwieser, 2004).

Non-covalent interactions were recently evaluated in more detail: the crystal structures of different GPCR revealed that receptor protomers contact each other via two types of interfaces. While some receptors, such as rhodopsin and the κ-opioid receptor, form less compact homodimers via their transmembrane helices I, II and their intracellular eighth helix, other GPCR such as CXCR4, tightly interact through its transmembrane helix V and VI also involving intracellular regions of helix III and IV (Filizola, 2013; Katrich, 2013). Moreover, some receptors seem to be capable of interacting via both interfaces, as shown for the μ-opioid receptor.

Although the crystal structures of Y receptors and the corresponding oligomerization interfaces have not been available up to now, Y receptor homodimerization at the cell membrane has been shown in vitro (Berglund et al., 2003a; Dinger et al., 2003) and in vivo (Estes et al., 2008) using ultracentrifugation, bioluminescence and fluorescence resonance energy transfer (BRET/FRET) as well as Western blot techniques. Moreover, Y receptor dimers were found to be pre-associated with heterotrimeric G proteins at the cell surface in a ratio of 2:1 or 2:2 (Estes et al., 2011; Kilpatrick et al., 2012). Pre-association with G protein was also observed for other GPCR, such as the GABA-B2 receptor, and most likely occurs shortly after biosynthesis in the ER (David et al., 2006). Thus, Y receptor homodimerization and pre-formation of the G-protein-receptor complex might be in fact a prerequisite for anterograde transport and full functionality of the receptor at the cell surface. In addition to homodimers, heterodimerization was observed for the Y1 and Y5 receptors (Gehlert et al., 2007), presumably due to the fact that these receptors are often co-expressed in the same tissues, and are located on the same but opposite and overlapping chromosomal region. When stimulating the Y1/Y5 heterodimer, Gehlert et al. observed an increase in Y5-receptor internalization with respect to its mono- and homodimer (Gehlert et al., 2007). These results were contradicted by the findings of later studies by Bohme et al, showing that co-expression of Y1/Y5 receptors did not influence Y5 receptor endocytosis (Bohme et al., 2008). Vice versa, the Y1 receptor was found to be internalized alone after stimulation despite being co-expressed with the Y5 receptor (Bohme et al., 2008). Moreover, while some groups report on Y receptor dimer dissociation upon agonist stimulation (Berglund et al., 2003a; Estes et al., 2008), others have observed an increase in oligomerization following agonist-induced receptor activation (Gehlert et al., 2007). Clearly, the role of Y receptor dimerization, as for many other class A GPCR, is still a matter of debate. Thus, further studies will be necessary to identify additional sequences and interaction partners that promote Y receptor export and putative oligomerization to gain further insights into the exact trafficking mechanisms.

Internalization and postendocytic trafficking of Y receptors

Extensive investigation revealed that many GPCR, upon agonist binding and activation, undergo receptor endocytosis via clathrin-coated pits, resulting in down-regulation or altered receptor signaling. In brief, the common mechanism of GPCR internalization requires agonist-induced receptor activation followed by G protein-mediated signaling. Subsequently, GRKs recognize and bind to the activated GPCR, thereby displacing the G protein from the receptor binding site (receptor desensitization). Once having bound, GRKs phosphorylate the active receptor at the serine/threonine residues within the C-terminus or ICL3, which enables the binding of Arr. Finally, Arr recruits the endocytosis machinery, such as adapter protein 2 (AP2) or clathrin, and serves as a scaffold for components involved in further signaling cascades. Some internalized GPCR are reported to preferentially bind Arr3, from which they dissociate immediately after endocytosis (‘class A’ receptor), while other GPCR bind Arr2 and Arr3 in the same manner and with higher affinity (‘class B’ receptors). Dissociation of Arr in combination with dephosphorylation of GPCR promotes recycling and resensitization either via direct or indirect pathways. In contrast, prolonged Arr association and ubiquitination favors receptor trafficking to lysosomes and subsequent degradation (Figure 2). In addition to the general internalization process, which is briefly explained above, several more mechanisms and accessory proteins were found to be involved in GPCR signaling, trafficking, de- and resensitization. All of these findings point towards a coupling of receptor activity to the endocytic machinery and vice versa and emphasize the importance of comprehensive studies on GPCR trafficking when modulating receptor activity. With respect to Y receptors, detailed knowledge about endocytosis and trafficking is rather limited. The Y1 receptor subtype has been studied in the greatest detail and similar and complementary results have been found. In contrast, trafficking properties of the remaining three human Y receptor subtypes have been more controversial, especially for the Y2 receptor. Taking the latest findings into account, it can be concluded that the hY1, hY2 and hY4 receptors are internalized quickly, while endocytosis of the hY5 receptor is slow. Clearly, different structural components located in the ICL2, the ICL3 and the C-terminus were identified as contributing to the observed subtype-specific desensitization, Arr recruitment and trafficking behavior of the Y receptors.

Y1 receptor

The hY1 receptor was found to be rapidly internalized upon agonist stimulation by different groups (Fabry et al., 2000; Gicquiaux et al., 2002; Bohme et al., 2008; Ouedraogo et al., 2008; Lindner et al., 2009; Kilpatrick et al., 2010; Lecat et al., 2011; Lundell et al., 2011), consistent with the results that were obtained for its orthologs in the rat (Holliday and Cox, 2003; Pheng et al., 2003; Holliday et al., 2005; Kilpatrick et al., 2012) or guinea pig (Parker et al., 2001, 2002b). Several studies on the internalization mechanism revealed that the Y1 receptor undergoes clathrin-dependent endocytosis in an Arr3-dependent fashion and is transported back to the cell membrane, involving direct and indirect recycling mechanisms (Gicquiaux et al., 2002; Berglund et al., Pheng et al., 2003b; Holliday et al., 2005; Ouedraogo et al., 2008; Kilpatrick et al., 2010; Lecat et al., 2011). Beside the interaction with Arr3, the Y1 receptor was also found to bind efficiently to Arr2, suggesting that it belongs to the group of ‘class B’ receptors (Ouedraogo et al., 2008; Kilpatrick et al., 2012). As mentioned before, interaction with Arr is usually a prerequisite for activated and phosphorylated receptors. For the identification of phosphorylation motifs, C-terminally truncated mutants of the rat and hY1 receptor (whose C-terminal sequences only vary in three distal residues) were investigated with respect to G protein coupling, Arr binding and internalization (Holliday et al., 2005; Ouedraogo et al., 2008; Lecat et al., 2011). These studies showed that the truncation of the whole C-terminus is very well tolerated with respect to G protein signaling, supporting the idea that the formation of the correct Y1 receptor structure does not require the presence of the C-terminus (see also anterograde transport). In contrast, truncation of the C-terminus had a dramatic impact on receptor desensitization, internalization and Arr recruitment. Detailed studies on point mutations clarified that individual serine/threonine residues within the C-terminal (S/T)-(S/T)-Φ-H-(S/T)-(E/D)-V-(S/T)-x-T motif (where x represents any aa and Φ a bulky hydrophobic residue) are phosphorylated by GRK2, and thus are required for receptor desensitization and Arr recruitment (Holliday et al., 2005; Ouedraogo et al., 2008; Kilpatrick et al., 2010, 2012) (Figure 4). Within this cluster, the number of serine/threonine residues rather than their actual position was found to be crucial for efficient Arr binding, suggesting that multiple phosphorylation events occur within this sequence (Kilpatrick et al., 2010).

Figure 4 Internalization motifs within human Y receptors.(A) Position and sequence of identified and putative motifs are displaced. Motifs within boxes were identified as being responsible for either receptor internalization or postendocytic trafficking. Underlined letters represent key residues in postulated or identified consensus sequences. (B) Localization of human Y receptor fused to the enhanced yellow fluorescent protein at the C-termini in living HEK293 cells prior to (unstimulated) or after agonist stimulation. hY1, hY2 and hY5 were stimulated with 1 μm pNPY, while hY4 was stimulated with 100 nm hPP for 60 min. Scale bar: 10 μm; indep., independent; rec, recycling.
Figure 4

Internalization motifs within human Y receptors.

(A) Position and sequence of identified and putative motifs are displaced. Motifs within boxes were identified as being responsible for either receptor internalization or postendocytic trafficking. Underlined letters represent key residues in postulated or identified consensus sequences. (B) Localization of human Y receptor fused to the enhanced yellow fluorescent protein at the C-termini in living HEK293 cells prior to (unstimulated) or after agonist stimulation. hY1, hY2 and hY5 were stimulated with 1 μm pNPY, while hY4 was stimulated with 100 nm hPP for 60 min. Scale bar: 10 μm; indep., independent; rec, recycling.

Another sequence within the C-terminus, namely the Y-x-x-Φ (YETI) motif, lays upstream of the serine/threonine cluster. This tyrosine-based sequence is known to promote direct interaction with the μ2 subunit of AP2, thereby mediating internalization (Pandey, 2009). In fact, the YETI motif was found to induce constitutive clathrin-dependent internalization of the Y1 receptor once the downstream phosphorylation and Arr binding motif was absent (Holliday et al., 2005; Lecat et al., 2011). Interestingly, mutation of the tyrosine residue (AETI) in combination with the full length Y1 receptor had no impact on internalization but impaired receptor reappearance at the plasma membrane after agonist-induced endocytosis (Lecat et al., 2011). Thus, the YETI motif within the C-terminus of the wild-type Y1 receptor is believed to regulate receptor recycling rather than internalization. In addition, a second Y-x-x-Φ (YIRL) motif is present in the ICL3 of the Y1 receptor and further adapter protein interaction sequences based on dileucine (D/E)-X2-3-L-(L/I) or triple basic motifs (R/K)-(R/K)-(R/K) are predicted at the beginning of ICL2 (ERHQLI) and ICL3 (KRR), respectively (Pandey, 2009; Smith et al., 2012). However, the impact of these motifs on receptor internalization, recycling and Arr binding has not been estimated yet and is therefore speculative.

Beside the relevance of C-terminal sequences, a proline residue located in the ICL2 was identified as acting in concert with the DRY motif as an Arr anchor in the 5-HT2c receptor (Marion et al., 2006). This proline residue can also be found in the ICL2 of the Y1 receptor. However, none of the Y receptors displays a classical DRY motif, and substitution of the proline residue to an alanine only had a minor impact on Y1 receptor internalization and Arr association (Ouedraogo et al., 2008). In conclusion, the C-terminus seems to be most crucial for the regulation of Y1 receptor desensitization and trafficking, as summarized in Figure 4. However, it cannot be ruled out that further motifs within the three ICLs contribute synergistically to receptor internalization and trafficking properties, binding either to Arr or to proteins of the endocytic machinery.

Clearly, those protein-receptor interactions depend on the conformation of both binding partners and the corresponding accessibility of consensus motifs exposed by the cytoplasmic receptor site. According to the hypothesis of ‘functional selectivity or biased agonism’, different ligands can promote discrete receptor conformations and thus favor one downstream effect over another, as it has been observed for several GPCR, including adrenergic receptors and the vasopressin V2 receptor (Rahmeh et al., 2012). Interestingly, Pheng et al. reported that the peptidic ligand GR231118 induced sequestration of Y1 receptor independent of G protein signaling (Pheng et al., 2003), which points towards a biased agonism of this compound. It is worth noting that GR231118 was initially identified to have agonistic effects only for Y4, while displaying a classical antagonistic function towards Y1 (Schober et al., 1998; Dumont and Quirion, 2000). In fact, Y1 receptor internalization induced by GR231118 could not be confirmed in later experiments, even though the receptor species and cell types were consistent with the initial study (Tough et al., 2006). One possible explanation for this might be that internalization experiments by Pheng et al. were based on the internalization of the radio-labeled GR231118, while later studies focused on receptor redistribution subsequent to ligand stimulation. Thus, further investigations will be necessary to unravel how GR231118 is internalized and whether the Y1 receptor can undergo biased signaling.

Y2 receptor

For a long time, the internalization and postendocytic trafficking properties of the Y2 receptor were controversial. Early radio-ligand binding studies on the guinea pig Y2 receptor revealed that this receptor subtype does not undergo sequestration (Parker et al., 2001). Similar studies on the human and rhesus receptor, respectively, confirmed that Y2 is resistant to endocytosis and shows only a weak interaction with Arr3 (Gicquiaux et al., 2002; Berglund et al., 2003b; Ouedraogo et al., 2008). In contrast, Bohme et al. clearly showed that the hY2 receptor desensitizes and undergoes fast internalization in response to agonist stimulation comparable to rates observed for the hY1 and hY4 receptors (Bohme et al., 2008). Recent studies confirmed those results and revealed that internalization of the hY2 receptor strongly depends on the agonist concentration used in the experiment (Lundell et al., 2011). In addition, Parker et al. reported that the internalization rate observed for hY2 is much higher than for its ortholog in the guinea pig (Parker et al., 2008). It is worth noting that the C-terminal sequence of the guinea pig Y2 receptor displays significant differences with respect to the serine/threonine cluster in comparison to other species, including humans. In a chimeric approach Bohme et al. showed that both the ICL3 and the C-terminus of the hY2 receptor can individually promote endocytosis of the usually non-internalizing hY5 receptor, thereby clearly demonstrating that these structural components determine agonist-induced internalization (Bohme et al., 2008). In addition, recent studies on Y1/Y2 chimeras revealed an enhanced internalization rate of the Y1 receptor when its C-terminal sequence is exchanged with the Y2 receptor C-terminus (Lundell et al., 2011).

Detailed investigation of the molecular mechanism underling Y2 receptor internalization unraveled three key motifs within the C-terminus responsible for the regulation of hY2 receptor trafficking (Walther et al., 2010). The first motif within the very distal C-terminus, namely D-S-x-T-E-x-T (DSFTEAT), was found to promote Arr3-dependent internalization subsequent to GRK2 phosphorylation. A second motif, D-x-Φ-H-(S/T)-(E/D)-V-(S/T)-x-T, that resembles the phosphorylation motif of the Y1 receptor, was found in the more proximal C-terminus, comprising residues DAIHSEVSVT. This sequence also induces agonist-mediated internalization, but surprisingly in an Arr3-independent manner (Walther et al., 2010). The Arr3-independent internalization mechanism was only observed when all further distal C-terminal parts were truncated. Conversely, the presence of the intermediate basic sequence FKAKKNLEVRKN impeded Arr3-independent receptor internalization, presuming that this inhibitory motif masks the proximal internalization motif D-x-Φ-H-(S/T)-(E/D)-V-(S/T)-x-T in the full-length receptor (Walther et al., 2010). Moreover, the proximal motif was also found to promote Y2 receptor recycling rather than internalization in the wild type receptor, similar to the function assigned to the YETI motif in the Y1 receptor.

By comparing the Y1 and Y2 subtypes, it becomes clear that the regulation of receptor trafficking strongly depends on sequences within the C-terminus. Importantly, consensus sequences and exact mechanisms seem to vary between those two Y receptor subtypes. However, each receptor displays one primary phosphorylation motif within the C-tail that functions as a molecular switch, thereby regulating Arr recruitment and internalization. In addition, a second latent internalization motif is present that controls receptor recycling and, apart from that, becomes involved in receptor endocytosis under the condition that the primary motif is missing.

Interestingly, while all other Y receptors possess a proline residue within the ICL2 for the proposed Arr interaction, the Y2 receptor displays a histidine at this position. Replacing this histidine with proline resulted in an enhanced Arr binding and internalization of the Y2 receptor (Marion et al., 2006; Kilpatrick et al., 2010, 2012), suggesting that Arr also recognized motifs within the ICL. Clearly, the role of additional trafficking sequences within the cytoplasmic region of the Y2 receptor requires further investigation.

Beside C-terminal residues, the N-terminus of the receptor has been found to contribute to receptor internalization properties. Even though hY2 receptor export is not affected when the entire N-terminus is truncated, ligand binding, receptor activation, and consequently the internalization process were drastically impaired (Lindner et al., 2009). Importantly, elongation of the first transmembrane residues by any eight aa rescued receptor activation and internalization, suggesting a sequence-independent, stabilizing effect of the N-terminus towards the ligand-binding pocket and/or the active receptor conformation (Lindner et al., 2009). Similarly, Parker et al. confirmed that the Y2 receptor N-terminus is not directly involved in agonist binding (Parker et al., 2008). However, this group discovered an acidic and proline-rich motif within the N-terminus contributing to receptor internalization. By mutating one aspartic acid residue within this motif to alanine, the receptor internalization was found to be accelerated. Since no other receptor characteristic was affected by this alanine substitution, it was suggested that the acidic motif within the N-terminus of the Y2 receptor binds to components of the extracellular matrix. This, in turn, might affect Y2 receptor internalization kinetics (Parker et al., 2012).

In conclusion, the hY2 receptor is internalized in an Arr3-dependent manner, but higher agonist concentration seems to be required to promote receptor endocytosis (Walther et al., 2010; Lundell et al., 2011). Even though the serum concentration of NPY and PYY might be too low to drive receptor internalization in the periphery, local concentrations in the synaptic gap might be high enough to induce hY2 endocytosis in the brain. That the Y2 receptor desensitization might indeed have an important physiological relevance was impressively shown by Ortiz et al. (2007). Here, a stabilized PYY(13–36) analog was observed to reduce food intake in mice. In a long-term study, however, food intake returned to the control level after 3days of daily drug administration. This might be due to receptor desensitization or down-regulation.

Y4 receptor

Similar to the Y2 receptor, studies on Y4 receptor desensitization and trafficking revealed opposing results. Initially, the hY4 receptor was reported to be resistant to agonist-promoted desensitization and internalization when being expressed in Chinese hamster ovary cells (Voisin et al., 2000). In contrast, several other groups independently showed that the Y4 receptor is rapidly internalized and subsequently transported back to the cell surface, most probably via the indirect recycling pathway (Parker et al., 2001, 2002a; Tough et al., 2006; Bohme et al., 2008). The determined internalization and recycling rates were either comparable or slightly slower than those observed for the Y1 receptor. With respect to the exact internalization mechanism, Y4 receptor endocytosis was found to be promoted by Arr3, suggesting a clathrin-dependent process (Berglund et al., 2003b). In these studies, however, Arr3 recruitment was less pronounced for the Y4 than for the Y1 receptor, supporting the idea that Y4 receptor internalization and recycling rates might indeed be slightly prolonged.

In addition to these findings, the rate of Y4 receptor sequestration was found to be highly sensitive to agonist affinity and efficacy, i.e., full agonists induced a higher internalization rate than partial agonists (Parker et al., 2002a; Tough et al., 2006). This is in agreement with the common mechanism of GPCR activation followed by sequestration, as described above, and underlines the fact that the endogenous Y4 ligands generate two conformations, which are equally important for G protein and GRK/Arr binding. It is worth noting that the Y4 receptor and its ligand PP emerged last and presumably quite quickly in evolution, leading to greater divergences in peptide and receptor sequences across mammalian species (Yulyaningsih et al., 2011). Thus, a combination of receptors and ligands originating from different species was observed to result in a decreased internalization rate due to reduced peptide efficacy (Tough et al., 2006).

Very little information is available about consensus sequences that drive internalization and recycling processes of the Y4 receptor. As the Y4 receptor is most closely related to the Y1 receptor, it is not surprising that a variant of the serine/threonine cluster, which is known to promote Arr binding and internalization of the Y1 receptor subtype, also occurs within the distal C-tail of the Y4 receptor (STVHTEVSKG). Further sequences that might be involved in Y receptor trafficking include the proline-based Arr-binding motif and the tyrosine-based AP-interaction sequence within ICL2 and ICL3, respectively. Even though, all these motifs are present in the Y4 receptor, their role in internalization and trafficking is still a matter of investigation.

Y5 receptor

The regulation of Y5 receptor trafficking is not well understood. Microscopic studies using EYFP-fused receptors and radio-ligand internalization experiments revealed an extremely slow internalization rate for both ligand and receptor (Parker et al., 2003; Bohme et al., 2008). Nonetheless, the receptor was still found to desensitize in inositol phosphate (IP)-accumulation assays (Bohme et al., 2008). The relatively small fraction of internalized receptor was even smaller in the presence of phenylarsine oxide, an inhibitor of clathrin-coated pit formation (Parker et al., 2003), suggesting a clathrin-dependent internalization mechanism. Studies on Arr recruitment are rather limited but we were unable to detect a pronounced Arr recruitment subsequent to Y5 receptor stimulation in our lab (data unpublished). Contrasting with this, BRET studies showed that the Y5 receptor can recruit Arr3 after stimulation, even though the signals were much lower than observed for the Y1 receptor (Berglund et al., 2003b).

Comprehensive studies are necessary to unravel the structural determinants that modulate and inhibit Y5 receptor internalization. It is reasonable that these structures being distinct to all other Y receptors, namely the long ICL3 and the short C-terminus, have an impact on Y5 receptor trafficking. The Y5 receptor chimera bearing either the ICL3 or C-terminus of the Y2 receptor showed an accelerated internalization rate (Bohme et al., 2008). Still, it cannot be determined whether this effect is due to the deletion of inhibitory segments lying in the Y5 receptor ICL3 and C-terminus or to the addition of internalization motif present in the Y2 receptor sequence. Thus, further studies are required to estimate the mechanisms underlying Y5 receptor trafficking.

Taking everything into account, it can be concluded that Y receptor subtypes internalize with rates in the following order: Y1>Y2≥Y4>>Y5. All receptors seem to undergo clathrin-mediated endocytosis, involving Arr in case of the Y1, Y2 and Y4 receptors. The three latter receptors were also found to be recycled subsequent to endocytosis.

Conclusion and future directions

The NPY system has frequently been shown to be a promising target for the treatment of several diseases, with an emphasis on obesity and cancer. Beside the Y1 receptor-selective agonist for breast cancer diagnosis (Khan et al., 2010), various Y receptor agonists and antagonists have been developed for therapeutic applications. Among those are several Y5 receptor antagonists designed for antiobesity medication. Two such antagonists, MK-0557 and Velneperit, have already been tested in Phase II trials, but failed to show a clinically-meaningful effect (Sato et al., 2009). In contrast, another antiobesity approach using the dual Y2/Y4 receptor peptidic agonist Obinepitide (produced by 7TM Pharma) caused a significant reduction in food intake during Phase I and II trials. Despite this desired clinical outcome of Obinepitide, side-effects were reported such as nausea (Sato et al., 2009).

Thus, the development of specific agonists, antagonists and even biased ligands is still a major issue when using GPCR as drug targets. However, the examples provided above emphasize that engineering specific ligands acting on GPCR can only be the first step in the development of well-tolerated drugs with high potentials.

Here, we report on specific Y receptor-trafficking processes, such as anterograde transport, internalization and recycling, which clearly intervene with receptor response and signaling. These processes might be responsible for the undesired side effects. They can influence the response via the modulation of cell surface availability of receptors. In order to understand and avoid desensitization mechanisms and drug resistance, it will be crucial to unravel the mechanisms of subcellular receptor trafficking. Also for the receptor-mediated uptake of cytostatic compounds for cancer therapy, it will be essential to elucidate internalization or degradation pathways and their regulation. Intervention in alternative signaling pathways related to internalization might open further possibilities of specifically modifying Y receptor signaling. Consequently, intracellular receptor trafficking should be included in a second step in drug development.

With respect to the Y receptor system, it is to be noted that despite the formation of one multiligand/multipeptide system, each Y receptor subtype apparently has a distinct molecular mechanism responsible for the regulation of its export or internalization. This provides the possibility to specifically modulate the localization/signaling of one subtype receptor while leaving others unaffected. Surely, further studies will be helpful in generating an overall picture that, in combination with subtype-selective ligands, will provide a powerful tool in the engineering of highly effective and long-acting pharmaceuticals that address Y receptors.


Corresponding author: Annette G. Beck-Sickinger, Institute of Biochemistry, Faculty of Biosciences, Pharmacy and Psychology, Leipzig University, Brüderstraße 34, D-04103 Leipzig, Germany

About the authors

Stefanie Babilon

Stefanie Babilon was born in 1985 in Giessen, Germany. She studied Biochemistry at the University of Leipzig in a bachelor’s-master’s program. In 2010, she performed her master thesis under the supervision of Prof. Dr. Annette G. Beck-Sickinger, during which she spent a 3-month research internship at the Vanderbilt University. After her graduation she stayed in the research group of Prof. Dr. Annette G. Beck-Sickinger for her PhD, which focuses on the internalization and postendocytic trafficking mechanisms of human Y receptors. Since 2011, she is supported by a scholarship of the Fonds der Chemischen Industrie.

Karin Mörl

Karin Mörl was born 1967 in Munich and studied Biology at the Ludwig-Maximilians-University, Munich. She graduated from the Max-Planck-Institute for Neurobiology in Martinsried, Germany under the supervision of Prof. Dr. Hans Thoenen and has been working as a postdoctoral researcher with Dr. Michael Meyer in Martinsried and Prof. Dr. Volker Bigl in Leipzig. In 2001 she joined the group of Prof. Dr. Annette G. Beck-Sickinger at the Institute of Biochemistry, University of Leipzig as a staff scientist.

Annette G. Beck-Sickinger

Annette Beck-Sickinger studied Chemistry (Diploma 1986) and Biology (Diploma 1990) at the Eberhard Karls University in Tübingen and graduated with G. Jung in 1989. After fellowships in Zürich (Carafoli), Scripps La Jolla (Houghten), and Copenhagen (Schwartz) she became Assistant Professor of Pharmacology and Biochemistry at ETH Zürich. In 1999 she accepted the full professorship at Leipzig University in the Departmant of Bioorganic Chemistry and Biochemistry. In 2009 she was visiting professor at Vanderbilt University, in Nashville.

The review summarizes the work that was achieved within the SFB 610, TP A1. The German Science Foundation (DFG) is kindly acknowledged for continued financial funding. S.B. receives a stipend from the Fond der Chemischen Industrie and kindly acknowledges financial contribution.

References

Bader, R., Bettio, A., Beck-Sickinger, A.G., and Zerbe, O. (2001). Structure and dynamics of micelle-bound neuropeptide Y: comparison with unligated NPY and implications for receptor selection. J. Mol. Biol. 305, 307–329.10.1006/jmbi.2000.4264Search in Google Scholar

Batterham, R.L., Cowley, M.A., Small, C.J., Herzog, H., Cohen, M.A., Dakin, C.L., Wren, A.M., Brynes, A.E., Low, M.J., Ghatei, M.A., et al. (2002). Gut hormone PYY(3-36) physiologically inhibits food intake. Nature 418, 650–654.10.1038/nature00887Search in Google Scholar

Bellmann-Sickert, K., Elling, C.E., Madsen, A.N., Little, P.B., Lundgren, K., Gerlach, L.O., Bergmann, R., Holst, B., Schwartz, T.W., and Beck-Sickinger, A.G. (2011). Long-acting lipidated analogue of human pancreatic polypeptide is slowly released into circulation. J. Med. Chem. 54, 2658–2667.10.1021/jm101357eSearch in Google Scholar

Benmaamar, R., Richichi, C., Gobbi, M., Daniels, A.J., Beck-Sickinger, A.G., and Vezzani, A. (2005). Neuropeptide Y Y5 receptors inhibit kindling acquisition in rats. Regul. Pept. 125, 79–83.10.1016/j.regpep.2004.07.029Search in Google Scholar

Berglund, M.M., Schober, D.A., Esterman, M.A., and Gehlert, D.R. (2003a). Neuropeptide Y Y4 receptor homodimers dissociate upon agonist stimulation. J. Pharmacol. Exp. Ther. 307, 1120–1126.10.1124/jpet.103.055673Search in Google Scholar

Berglund, M.M., Schober, D.A., Statnick, M.A., McDonald, P.H., and Gehlert, D.R. (2003b). The use of bioluminescence resonance energy transfer 2 to study neuropeptide Y receptor agonist-induced β-arrestin 2 interaction. J. Pharmacol. Exp. Ther. 306, 147–156.10.1124/jpet.103.051227Search in Google Scholar

Bermak, J.C., Li, M., Bullock, C., and Zhou, Q.Y. (2001). Regulation of transport of the dopamine D1 receptor by a new membrane-associated ER protein. Nat. Cell Biol. 3, 492–498.10.1038/35074561Search in Google Scholar

Blundell, T.L., Pitts, J.E., Tickle, I.J., Wood, S.P., and Wu, C.W. (1981). X-ray analysis (1. 4-Å resolution) of avian pancreatic polypeptide: small globular protein hormone. Proc. Natl. Acad. Sci. USA 78, 4175–4179.10.1073/pnas.78.7.4175Search in Google Scholar

Bohme, I., Stichel, J., Walther, C., Morl, K., and Beck-Sickinger, A.G. (2008). Agonist induced receptor internalization of neuropeptide Y receptor subtypes depends on third intracellular loop and C-terminus. Cell. Signal. 20, 1740–1749.10.1016/j.cellsig.2008.05.017Search in Google Scholar

Borowsky, B., Walker, M.W., Bard, J., Weinshank, R.L., Laz, T.M., Vaysse, P., Branchek, T.A., and Gerald, C. (1998). Molecular biology and pharmacology of multiple NPY Y5 receptor species homologs. Regul. Pept. 7576, 45–53.10.1016/S0167-0115(98)00052-4Search in Google Scholar

Breitwieser, G.E. (2004). G protein-coupled receptor oligomerization: implications for G protein activation and cell. signal. Circ. Res. 94, 17–27.10.1161/01.RES.0000110420.68526.19Search in Google Scholar

Bridges, T.M. and Lindsley, C.W. (2008). G-protein-coupled receptors: from classical modes of modulation to allosteric mechanisms. ACS Chem. Biol. 3, 530–541.10.1021/cb800116fSearch in Google Scholar

Cabrele, C. and Beck-Sickinger, A.G. (2000). Molecular characterization of the ligand-receptor interaction of the neuropeptide Y family. J. Peptide Sci. 6, 97–122.10.1002/(SICI)1099-1387(200003)6:3<97::AID-PSC236>3.0.CO;2-ESearch in Google Scholar

Cox, H.M., Tough, I.R., Zandvliet, D.W., and Holliday, N.D. (2001). Constitutive neuropeptide Y Y4 receptor expression in human colonic adenocarcinoma cell lines. Br. J. Pharmacol. 132, 345–353.10.1038/sj.bjp.0703815Search in Google Scholar

David, M., Richer, M., Mamarbachi, A.M., Villeneuve, L.R., Dupre, D.J., and Hebert, T.E. (2006). Interactions between GABA-B1 receptors and Kir 3 inwardly rectifying potassium channels. Cell. Signal. 18, 2172–2181.10.1016/j.cellsig.2006.05.014Search in Google Scholar

Dinger, M.C., Bader, J.E., Kobor, A.D., Kretzschmar, A.K., and Beck-Sickinger, A.G. (2003). Homodimerization of neuropeptide y receptors investigated by fluorescence resonance energy transfer in living cells. J. Biol. Chem. 278, 10562–10571.10.1074/jbc.M205747200Search in Google Scholar

Dong, C. and Wu, G. (2007). Regulation of anterograde transport of adrenergic and angiotensin II receptors by Rab2 and Rab6 GTPases. Cell. Signal. 19, 2388–2399.10.1016/j.cellsig.2007.07.017Search in Google Scholar

Dumont, Y. and Quirion, R. (2000). [125I]-GR231118: a high affinity radioligand to investigate neuropeptide Y Y1 and Y4 receptors. Br. J. Pharmacol. 129, 37–46.10.1038/sj.bjp.0702983Search in Google Scholar

Dumont, Y., Jacques, D., Bouchard, P., and Quirion, R. (1998). Species differences in the expression and distribution of the neuropeptide Y Y1, Y2, Y4, and Y5 receptors in rodents, guinea pig, and primates brains. J. Comp. Neurol. 402, 372–384.10.1002/(SICI)1096-9861(19981221)402:3<372::AID-CNE6>3.0.CO;2-2Search in Google Scholar

Durkin, M.M., Walker, M.W., Smith, K.E., Gustafson, E.L., Gerald, C., and Branchek, T.A. (2000). Expression of a novel neuropeptide Y receptor subtype involved in food intake: an in situ hybridization study of Y5 mRNA distribution in rat brain. Exp. Neurol. 165, 90–100.10.1006/exnr.2000.7446Search in Google Scholar

Duvernay, M.T., Filipeanu, C.M., and Wu, G. (2005). The regulatory mechanisms of export trafficking of G protein-coupled receptors. Cell. Signal. 17, 1457–1465.10.1016/j.cellsig.2005.05.020Search in Google Scholar

El Bahh, B., Balosso, S., Hamilton, T., Herzog, H., Beck-Sickinger, A.G., Sperk, G., Gehlert, D.R., Vezzani, A., and Colmers, W.F. (2005). The anti-epileptic actions of neuropeptide Y in the hippocampus are mediated by Y and not Y receptors. Eur. J. Neurosci. 22, 1417–1430.10.1111/j.1460-9568.2005.04338.xSearch in Google Scholar

Estes, A.M., Wong, Y.Y., Parker, M.S., Sallee, F.R., Balasubramaniam, A., and Parker, S.L. (2008). Neuropeptide Y (NPY) Y2 receptors of rabbit kidney cortex are largely dimeric. Regul. Pept. 150, 88–94.10.1016/j.regpep.2008.06.002Search in Google Scholar

Estes, A.M., McAllen, K., Parker, M.S., Sah, R., Sweatman, T., Park, E.A., Balasubramaniam, A., Sallee, F.R., Walker, M.W., and Parker, S.L. (2011). Maintenance of Y receptor dimers in epithelial cells depends on interaction with G-protein heterotrimers. Amino Acids 40, 371–380.10.1007/s00726-010-0642-zSearch in Google Scholar

Fabry, M., Langer, M., Rothen-Rutishauser, B., Wunderli-Allenspach, H., Hocker, H., and Beck-Sickinger, A.G. (2000). Monitoring of the internalization of neuropeptide Y on neuroblastoma cell line SK-N-MC. Eur. J. Biochem. 267, 5631–5637.10.1046/j.1432-1327.2000.01631.xSearch in Google Scholar

Filizola, M. and Devi, L.A. (2013). Grand opening of structure-guided design for novel opioids. Trends Pharmacol. Sci. 34, 6–12.10.1016/j.tips.2012.10.002Search in Google Scholar

Flood, J.F., Baker, M.L., Hernandez, E.N., and Morley, J.E. (1989). Modulation of memory processing by neuropeptide Y varies with brain injection site. Brain Res. 503, 73–82.10.1016/0006-8993(89)91706-XSearch in Google Scholar

Gehlert, D.R., Schober, D.A., Morin, M., and Berglund, M.M. (2007). Co-expression of neuropeptide Y Y1 and Y5 receptors results in heterodimerization and altered functional properties. Biochem. Pharmacol. 74, 1652–1664.10.1016/j.bcp.2007.08.017Search in Google Scholar PubMed

Gicquiaux, H., Lecat, S., Gaire, M., Dieterlen, A., Mely, Y., Takeda, K., Bucher, B., and Galzi, J.L. (2002). Rapid internalization and recycling of the human neuropeptide Y Y1 receptor. J. Biol. Chem. 277, 6645–6655.10.1074/jbc.M107224200Search in Google Scholar PubMed

Gurevich, E.V. and Gurevich, V.V. (2006). Arrestins: ubiquitous regulators of cellular signaling pathways. Genome Biol. 7, 236.10.1186/gb-2006-7-9-236Search in Google Scholar PubMed PubMed Central

Gurevich, V.V. and Gurevich, E.V. (2008). GPCR monomers and oligomers: it takes all kinds. Trends Neurosci. 31, 74–81.10.1016/j.tins.2007.11.007Search in Google Scholar

Gurevich, E.V., Tesmer, J.J., Mushegian, A., and Gurevich, V.V. (2012). G protein-coupled receptor kinases: more than just kinases and not only for GPCRs. Pharmacol. Ther. 133, 40–69.10.1016/j.pharmthera.2011.08.001Search in Google Scholar

Hansen, A.P. and Sheikh, S.P. (1992). Y2 receptor proteins for peptide YY and neuropeptide Y. Characterization as N-linked complex glycoproteins. FEBS Lett. 306, 147–150.10.1016/0014-5793(92)80987-RSearch in Google Scholar

Hayes, D.M., Fee, J.R., McCown, T.J., Knapp, D.J., Breese, G.R., Cubero, I., Carvajal, F., Lerma-Cabrera, J.M., Navarro, M., and Thiele, T.E. (2012). Neuropeptide Y signaling modulates the expression of ethanol-induced behavioral sensitization in mice. Addict. Biol. 17, 338–350.10.1111/j.1369-1600.2011.00336.xSearch in Google Scholar PubMed PubMed Central

Heilig, M. (2004). The NPY system in stress, anxiety and depression. Neuropeptides 38, 213–224.10.1016/j.npep.2004.05.002Search in Google Scholar PubMed

Hodges, G.J., Jackson, D.N., Mattar, L., Johnson, J.M., and Shoemaker, J.K. (2009). Neuropeptide Y and neurovascular control in skeletal muscle and skin. Am. J. Physiol. Regul. Integr. Comp. Physiol. 297, 546–555.10.1152/ajpregu.00157.2009Search in Google Scholar PubMed PubMed Central

Holliday, N.D. and Cox, H.M. (2003). Control of signalling efficacy by palmitoylation of the rat Y1 receptor. Br. J. Pharmacol. 139, 501–512.10.1038/sj.bjp.0705276Search in Google Scholar PubMed PubMed Central

Holliday, N.D., Lam, C.W., Tough, I.R., and Cox, H.M. (2005). Role of the C-terminus in neuropeptide Y Y1 receptor desensitization and internalization. Mol. Pharmacol. 67, 655–664.10.1124/mol.104.006114Search in Google Scholar PubMed

Howell, O.W., Doyle, K., Goodman, J.H., Scharfman, H.E., Herzog, H., Pringle, A., Beck-Sickinger, A.G., and Gray, W.P. (2005). Neuropeptide Y stimulates neuronal precursor proliferation in the post-natal and adult dentate gyrus. J. Neurochem. 93, 560–570.10.1111/j.1471-4159.2005.03057.xSearch in Google Scholar PubMed

Katritch, V., Cherezov, V., and Stevens, R.C. (2013). Structure-function of the G protein-coupled receptor superfamily. Annu. Rev. Pharmacol. Toxicol. 53, 531–556.10.1146/annurev-pharmtox-032112-135923Search in Google Scholar PubMed PubMed Central

Khan, I.U., Zwanziger, D., Bohme, I., Javed, M., Naseer, H., Hyder, S.W., and Beck-Sickinger, A.G. (2010). Breast-cancer diagnosis by neuropeptide Y analogues: from synthesis to clinical application. Angew. Chem. Int. Ed. Engl. 49, 1155–1158.10.1002/anie.200905008Search in Google Scholar PubMed

Kilpatrick, L.E., Briddon, S.J., Hill, S.J., and Holliday, N.D. (2010). Quantitative analysis of neuropeptide Y receptor association with β-arrestin2 measured by bimolecular fluorescence complementation. Br. J. Pharmacol. 160, 892–906.10.1111/j.1476-5381.2010.00676.xSearch in Google Scholar PubMed PubMed Central

Kilpatrick, L.E., Briddon, S.J., and Holliday, N.D. (2012). Fluorescence correlation spectroscopy, combined with bimolecular fluorescence complementation, reveals the effects of β-arrestin complexes and endocytic targeting on the membrane mobility of neuropeptide Y receptors. BBA-Mol. Cell Res. 1823, 1068–1081.10.1016/j.bbamcr.2012.03.002Search in Google Scholar PubMed PubMed Central

Kitlinska, J., Abe, K., Kuo, L., Pons, J., Yu, M., Li, L., Tilan, J., Everhart, L., Lee, E.W., Zukowska, Z., et al. (2005). Differential effects of neuropeptide Y on the growth and vascularization of neural crest-derived tumors. Cancer Res. 65, 1719–1728.10.1158/0008-5472.CAN-04-2192Search in Google Scholar PubMed

Klapstein, G.J. and Colmers, W.F. (1993). On the sites of presynaptic inhibition by neuropeptide Y in rat hippocampus in vitro. Hippocampus 3, 103–111.10.1002/hipo.450030111Search in Google Scholar PubMed

Korner, M. and Reubi, J.C. (2007). NPY receptors in human cancer: a review of current knowledge. Peptides 28, 419–425.10.1016/j.peptides.2006.08.037Search in Google Scholar PubMed

Korner, M. and Reubi, J.C. (2008). Neuropeptide Y receptors in primary human brain tumors: overexpression in high-grade tumors. J. Neuropathol. Exp. Neurol. 67, 741–749.10.1097/NEN.0b013e318180e618Search in Google Scholar PubMed

Korner, M., Waser, B., and Reubi, J.C. (2008). High expression of neuropeptide Y1 receptors in ewing sarcoma tumors. Clin. Cancer. Res. 14, 5043–5049.10.1158/1078-0432.CCR-07-4551Search in Google Scholar PubMed

Larhammar, D. and Salaneck, E. (2004). Molecular evolution of NPY receptor subtypes. Neuropeptides 38, 141–151.10.1016/j.npep.2004.06.002Search in Google Scholar PubMed

Lecat, S., Ouedraogo, M., Cherrier, T., Noulet, F., Ronde, P., Glasser, N., Galzi, J.L., Mely, Y., Takeda, K., and Bucher, B. (2011). Contribution of a tyrosine-based motif to cellular trafficking of wild-type and truncated NPY Y1 receptors. Cell. Signal. 23, 228–238.10.1016/j.cellsig.2010.09.007Search in Google Scholar PubMed

Lee, E.W., Michalkiewicz, M., Kitlinska, J., Kalezic, I., Switalska, H., Yoo, P., Sangkharat, A., Ji, H., Li, L., Michalkiewicz, T., et al. (2003). Neuropeptide Y induces ischemic angiogenesis and restores function of ischemic skeletal muscles. J. Clin. Invest. 111, 1853–1862.10.1172/JCI16929Search in Google Scholar

Lerch, M., Gafner, V., Bader, R., Christen, B., Folkers, G., and Zerbe, O. (2002). Bovine pancreatic polypeptide (bPP) undergoes significant changes in conformation and dynamics upon binding to DPC micelles. J. Mol. Biol. 322, 1117–1133.10.1016/S0022-2836(02)00889-6Search in Google Scholar

Lerch, M., Mayrhofer, M., and Zerbe, O. (2004). Structural similarities of micelle-bound peptide YY (PYY) and neuropeptide Y (NPY) are related to their affinity profiles at the Y receptors. J. Mol. Biol. 339, 1153–1168.10.1016/j.jmb.2004.04.032Search in Google Scholar PubMed

Lin, S., Shi, Y.C., Yulyaningsih, E., Aljanova, A., Zhang, L., Macia, L., Nguyen, A.D., Lin, E.J., During, M.J., Herzog, H., and Sainsbury, A. (2009). Critical role of arcuate Y4 receptors and the melanocortin system in pancreatic polypeptide-induced reduction in food intake in mice. PLoS One 4, e8488.10.1371/journal.pone.0008488Search in Google Scholar PubMed PubMed Central

Lindner, D., Stichel, J., and Beck-Sickinger, A.G. (2008a). Molecular recognition of the NPY hormone family by their receptors. Nutrition 24, 907–917.10.1016/j.nut.2008.06.025Search in Google Scholar PubMed

Lindner, D., van Dieck, J., Merten, N., Morl, K., Gunther, R., Hofmann, H.J., and Beck-Sickinger, A.G. (2008b). GPC receptors and not ligands decide the binding mode in neuropeptide Y multireceptor/multiligand system. Biochemistry 47, 5905–5914.10.1021/bi800181kSearch in Google Scholar PubMed

Lindner, D., Walther, C., Tennemann, A., and Beck-Sickinger, A.G. (2009). Functional role of the extracellular N-terminal domain of neuropeptide Y subfamily receptors in membrane integration and agonist-stimulated internalization. Cell. Signal. 21, 61–68.10.1016/j.cellsig.2008.09.007Search in Google Scholar PubMed

Lu, C., Everhart, L., Tilan, J., Kuo, L., Sun, C.C., Munivenkatappa, R.B., Jonsson-Rylander, A.C., Sun, J., Kuan-Celarier, A., Li, L., et al. (2010). Neuropeptide Y and its Y2 receptor: potential targets in neuroblastoma therapy. Oncogene 29, 5630–5642.10.1038/onc.2010.301Search in Google Scholar PubMed PubMed Central

Lundell, I., Rabe Bernhardt, N., Johnsson, A.K., and Larhammar, D. (2011). Internalization studies of chimeric neuropeptide Y receptors Y1 and Y2 suggest complex interactions between cytoplasmic domains. Regul. Pept. 168, 50–58.10.1016/j.regpep.2011.03.004Search in Google Scholar PubMed

Magalhaes, A.C., Dunn, H., and Ferguson, S.S. (2012). Regulation of GPCR activity, trafficking and localization by GPCR-interacting proteins. Br. J. Pharmacol. 165, 1717–1736.10.1111/j.1476-5381.2011.01552.xSearch in Google Scholar PubMed PubMed Central

Marion, S., Oakley, R.H., Kim, K.M., Caron, M.G., and Barak, L.S. (2006). A β-arrestin binding determinant common to the second intracellular loops of rhodopsin family G protein-coupled receptors. J. Biol. Chem. 281, 2932–2938.10.1074/jbc.M508074200Search in Google Scholar

Mentlein, R., Dahms, P., Grandt, D., and Kruger, R. (1993). Proteolytic processing of neuropeptide Y and peptide YY by dipeptidyl peptidase IV. Regul. Pept. 49, 133–144.10.1016/0167-0115(93)90435-BSearch in Google Scholar

Merten, N., Lindner, D., Rabe, N., Rompler, H., Morl, K., Schoneberg, T., and Beck-Sickinger, A.G. (2007). Receptor subtype-specific docking of Asp6.59 with C-terminal arginine residues in Y receptor ligands. J. Biol. Chem. 282, 7543–7551.10.1074/jbc.M608902200Search in Google Scholar PubMed

Michel, M.C., Beck-Sickinger, A., Cox, H., Doods, H.N., Herzog, H., Larhammar, D., Quirion, R., Schwartz, T., and Westfall, T. (1998). XVI. International Union of Pharmacology recommendations for the nomenclature of neuropeptide Y, peptide YY, and pancreatic polypeptide receptors. Pharmacol. Rev. 50, 143–150.Search in Google Scholar

Millar, R.P. and Newton, C.L. (2010). The year in G protein-coupled receptor research. Mol. Endocrinol. 24, 261–274.10.1210/me.2009-0473Search in Google Scholar PubMed PubMed Central

Morales-Medina, J.C., Dominguez-Lopez, S., Gobbi, G., Beck-Sickinger, A.G., and Quirion, R. (2012). The selective neuropeptide Y Y5 agonist [cPP(1-7),NPY(19-23),Ala31,Aib32,Gln34]hPP differently modulates emotional processes and body weight in the rat. Behav. Brain Res. 233, 298–304.10.1016/j.bbr.2012.05.015Search in Google Scholar PubMed

Nguyen, A.D., Herzog, H., and Sainsbury, A. (2011). Neuropeptide Y and peptide YY: important regulators of energy metabolism. Curr. Opin. Endocrinol. Diabetes Obes. 18, 56–60.10.1097/MED.0b013e3283422f0aSearch in Google Scholar PubMed

Nguyen, A.D., Mitchell, N.F., Lin, S., Macia, L., Yulyaningsih, E., Baldock, P.A., Enriquez, R.F., Zhang, L., Shi, Y.C., Zolotukhin, S., et al. (2012). Y1 and Y5 receptors are both required for the regulation of food intake and energy homeostasis in mice. PLoS One 7, e40191.10.1371/journal.pone.0040191Search in Google Scholar PubMed PubMed Central

Ortiz, A.A., Milardo, L.F., DeCarr, L.B., Buckholz, T.M., Mays, M.R., Claus, T.H., Livingston, J.N., Mahle, C.D., and Lumb, K.J. (2007). A novel long-acting selective neuropeptide Y2 receptor polyethylene glycol-conjugated peptide agonist reduces food intake and body weight and improves glucose metabolism in rodents. J. Pharmacol. Exp. Ther. 323, 692–700.10.1124/jpet.107.125211Search in Google Scholar PubMed

Ouedraogo, M., Lecat, S., Rochdi, M.D., Hachet-Haas, M., Matthes, H., Gicquiaux, H., Verrier, S., Gaire, M., Glasser, N., Mely, Y., et al. (2008). Distinct motifs of neuropeptide Y receptors differentially regulate trafficking and desensitization. Traffic 9, 305–324.10.1111/j.1600-0854.2007.00691.xSearch in Google Scholar PubMed

Pandey, K.N. (2009). Functional roles of short sequence motifs in the endocytosis of membrane receptors. Front. Biosci. 14, 5339–5360.10.2741/3599Search in Google Scholar PubMed PubMed Central

Parker, S.L., Kane, J.K., Parker, M.S., Berglund, M.M., Lundell, I.A., and Li, M.D. (2001). Cloned neuropeptide Y (NPY) Y1 and pancreatic polypeptide Y4 receptors expressed in Chinese hamster ovary cells show considerable agonist-driven internalization, in contrast to the NPY Y2 receptor. Eur. J. Biochem. 268, 877–886.10.1046/j.1432-1327.2001.01966.xSearch in Google Scholar

Parker, M.S., Lundell, I., and Parker, S.L. (2002a). Internalization of pancreatic polypeptide Y4 receptors: correlation of receptor intake and affinity. Eur. J. Pharmacol. 452, 279–287.10.1016/S0014-2999(02)02339-7Search in Google Scholar

Parker, S.L., Parker, M.S., Lundell, I., Balasubramaniam, A., Buschauer, A., Kane, J.K., Yalcin, A., and Berglund, M.M. (2002b). Agonist internalization by cloned Y1 neuropeptide Y (NPY) receptor in Chinese hamster ovary cells shows strong preference for NPY, endosome-linked entry and fast receptor recycling. Regul. Pept. 107, 49–62.10.1016/S0167-0115(02)00094-0Search in Google Scholar

Parker, S.L., Parker, M.S., Buschauer, A., and Balasubramaniam, A. (2003). Ligand internalization by cloned neuropeptide Y Y5 receptors excludes Y2 and Y4 receptor-selective peptides. Eur. J. Pharmacol. 474, 31–42.10.1016/S0014-2999(03)02039-9Search in Google Scholar

Parker, S.L., Parker, M.S., Wong, Y.Y., Sah, R., Balasubramaniam, A., and Sallee, F. (2008). Importance of a N-terminal aspartate in the internalization of the neuropeptide Y Y2 receptor. Eur. J. Pharmacol. 594, 26–31.10.1016/j.ejphar.2008.07.038Search in Google Scholar PubMed PubMed Central

Parker, M.S., Sah, R., and Parker, S.L. (2012). Surface masking shapes the traffic of the neuropeptide Y Y2 receptor. Peptides 37, 40–48.10.1016/j.peptides.2012.06.008Search in Google Scholar PubMed PubMed Central

Pedragosa Badia, X., Stichel, J., and Beck-Sickinger, A.G. (2013). Neuropeptide Y receptors: how to get subtype selectivity. Frontiers in Endocrinology 4, 5.Search in Google Scholar

Pheng, L.H., Dumont, Y., Fournier, A., Chabot, J.G., Beaudet, A., and Quirion, R. (2003). Agonist- and antagonist-induced sequestration/internalization of neuropeptide Y Y1 receptors in HEK293 cells. Br. J. Pharmacol. 139, 695–704.10.1038/sj.bjp.0705306Search in Google Scholar PubMed PubMed Central

Rahmeh, R., Damian, M., Cottet, M., Orcel, H., Mendre, C., Durroux, T., Sharma, K.S., Durand, G., Pucci, B., Trinquet, E., et al. (2012). Structural insights into biased G protein-coupled receptor signaling revealed by fluorescence spectroscopy. Proc. Natl. Acad. Sci. USA 109, 6733–6738.10.1073/pnas.1201093109Search in Google Scholar PubMed PubMed Central

Reubi, J.C., Gugger, M., Waser, B., and Schaer, J.C. (2001). Y1-mediated effect of neuropeptide Y in cancer: breast carcinomas as targets. Cancer Res. 61, 4636–4641.Search in Google Scholar

Rodriguez, M., Audinot, V., Dromaint, S., Macia, C., Lamamy, V., Beauverger, P., Rique, H., Imbert, J., Nicolas, J.P., Boutin, J.A., et al. (2003). Molecular identification of the long isoform of the human neuropeptide Y Y5 receptor and pharmacological comparison with the short Y5 receptor isoform. Biochem. J. 369, 667–673.10.1042/bj20020739Search in Google Scholar PubMed PubMed Central

Rose, P.M., Lynch, J.S., Frazier, S.T., Fisher, S.M., Chung, W., Battaglino, P., Fathi, Z., Leibel, R., and Fernandes, P. (1997). Molecular genetic analysis of a human neuropeptide Y receptor. The human homolog of the murine’Y5’ receptor may be a pseudogene. J. Biol. Chem. 272, 3622–3627.10.1074/jbc.272.6.3622Search in Google Scholar PubMed

Salahpour, A., Angers, S., Mercier, J.F., Lagace, M., Marullo, S., and Bouvier, M. (2004). Homodimerization of the β2-adrenergic receptor as a prerequisite for cell surface targeting. J. Biol. Chem. 279, 33390–33397.10.1074/jbc.M403363200Search in Google Scholar

Santos-Cuevas, C.L., Ferro-Flores, G., Arteaga de Murphy, C., and Pichardo-Romero, P.A. (2008). Targeted imaging of gastrin-releasing peptide receptors with 99mTc-EDDA/HYNIC-[Lys3]-bombesin: biokinetics and dosimetry in women. Nucl. Med. Commun. 29, 741–747.10.1097/MNM.0b013e3282ffb45cSearch in Google Scholar

Sato, N., Ogino, Y., Mashiko, S., and Ando, M. (2009). Modulation of neuropeptide Y receptors for the treatment of obesity. Expert Opin. Ther. Pat. 19, 1401–1415.10.1517/13543770903251722Search in Google Scholar

Schmidt, P.T., Naslund, E., Gryback, P., Jacobsson, H., Holst, J.J., Hilsted, L., and Hellstrom, P.M. (2005). A role for pancreatic polypeptide in the regulation of gastric emptying and short-term metabolic control. J. Clin. Endocrinol. Metab. 90, 5241–5246.10.1210/jc.2004-2089Search in Google Scholar

Schmidt, P., Lindner, D., Montag, C., Berndt, S., Beck-Sickinger, A.G., Rudolph, R., and Huster, D. (2009). Prokaryotic expression, in vitro folding, and molecular pharmacological characterization of the neuropeptide Y receptor type 2. Biotechnol. Prog. 25, 1732–1739.10.1002/btpr.266Search in Google Scholar

Schober, D.A., Van Abbema, A.M., Smiley, D.L., Bruns, R.F., and Gehlert, D.R. (1998). The neuropeptide Y Y1 antagonist, 1229U91, a potent agonist for the human pancreatic polypeptide-preferring (NPY Y4) receptor. Peptides 19, 537–542.10.1016/S0196-9781(97)00455-5Search in Google Scholar

Sheriff, S., Ali, M., Yahya, A., Haider, K.H., Balasubramaniam, A., and Amlal, H. (2010). Neuropeptide Y Y5 receptor promotes cell growth through extracellular signal-regulated kinase signaling and cyclic AMP inhibition in a human breast cancer cell line. Mol. Cancer Res. 8, 604–614.10.1158/1541-7786.MCR-09-0301Search in Google Scholar PubMed

Shi, Y.C., Lin, S., Castillo, L., Aljanova, A., Enriquez, R.F., Nguyen, A.D., Baldock, P.A., Zhang, L., Bijker, M.S., Macia, L., et al. (2011). Peripheral-specific Y2 receptor knockdown protects mice from high-fat diet-induced obesity. Obesity (Silver Spring) 19, 2137–2148.10.1038/oby.2011.99Search in Google Scholar PubMed

Shimada, K., Ohno, Y., Okamatsu-Ogura, Y., Suzuki, M., Kamikawa, A., Terao, A., and Kimura, K. (2012). Neuropeptide Y activates phosphorylation of ERK and STAT3 in stromal vascular cells from brown adipose tissue, but fails to affect thermogenic function of brown adipocytes. Peptides 34, 336–342.10.1016/j.peptides.2012.02.012Search in Google Scholar PubMed

Smith, K.R., Muir, J., Rao, Y., Browarski, M., Gruenig, M.C., Sheehan, D.F., Haucke, V., and Kittler, J.T. (2012). Stabilization of GABAA receptors at endocytic zones is mediated by an AP2 binding motif within the GABAA receptor β3 subunit. J. Neurosci. 32, 2485–2498.10.1523/JNEUROSCI.1622-11.2011Search in Google Scholar PubMed PubMed Central

Tasan, R.O., Lin, S., Hetzenauer, A., Singewald, N., Herzog, H., and Sperk, G. (2009). Increased novelty-induced motor activity and reduced depression-like behavior in neuropeptide Y (NPY)-Y4 receptor knockout mice. Neuroscience 158, 1717–1730.10.1016/j.neuroscience.2008.11.048Search in Google Scholar PubMed PubMed Central

Thiriet, N., Agasse, F., Nicoleau, C., Guegan, C., Vallette, F., Cadet, J.L., Jaber, M., Malva, J.O., and Coronas, V. (2011). NPY promotes chemokinesis and neurogenesis in the rat subventricular zone. J. Neurochem. 116, 1018–1027.10.1111/j.1471-4159.2010.07154.xSearch in Google Scholar PubMed

Thomas, L., Scheidt, H.A., Bettio, A., Huster, D., Beck-Sickinger, A.G., Arnold, K., and Zschornig, O. (2005). Membrane interaction of neuropeptide Y detected by EPR and NMR spectroscopy. Biochim Biophys Acta Biomembr 1714, 103–113.10.1016/j.bbamem.2005.06.012Search in Google Scholar PubMed

Tough, I.R., Holliday, N.D., and Cox, H.M. (2006). Y4 receptors mediate the inhibitory responses of pancreatic polypeptide in human and mouse colon mucosa. J. Pharmacol. Exp. Ther. 319, 20–30.10.1124/jpet.106.106500Search in Google Scholar PubMed

Vezzani, A. and Sperk, G. (2004). Overexpression of NPY and Y2 receptors in epileptic brain tissue: an endogenous neuroprotective mechanism in temporal lobe epilepsy? Neuropeptides 38, 245–252.10.1016/j.npep.2004.05.004Search in Google Scholar PubMed

Verma, D., Tasan, R.O., Herzog, H., and Sperk, G. (2012). NPY controls fear conditioning and fear extinction by combined action on Y1 and Y2 receptors. Br. J. Pharmacol. 166, 1461–1473.10.1111/j.1476-5381.2012.01872.xSearch in Google Scholar PubMed PubMed Central

Voisin, T., Goumain, M., Lorinet, A.M., Maoret, J.J., and Laburthe, M. (2000). Functional and molecular properties of the human recombinant Y4 receptor: resistance to agonist-promoted desensitization. J. Pharmacol. Exp. Ther. 292, 638–646.Search in Google Scholar

Wahlestedt, C., Pich, E.M., Koob, G.F., Yee, F., and Heilig, M. (1993). Modulation of anxiety and neuropeptide Y-Y1 receptors by antisense oligodeoxynucleotides. Science 259, 528–531.10.1126/science.8380941Search in Google Scholar PubMed

Walther, C., Nagel, S., Gimenez, L.E., Morl, K., Gurevich, V.V., and Beck-Sickinger, A.G. (2010). Ligand-induced internalization and recycling of the human neuropeptide Y2 receptor is regulated by its carboxyl-terminal tail. J. Biol. Chem. 285, 41578–41590.10.1074/jbc.M110.162156Search in Google Scholar PubMed PubMed Central

Walther, C., Lotze, J., Beck-Sickinger, A.G., and Morl, K. (2012). The anterograde transport of the human neuropeptide Y2 receptor is regulated by a subtype specific mechanism mediated by the C-terminus. Neuropeptides 46, 335–343.10.1016/j.npep.2012.08.011Search in Google Scholar PubMed

Wang, G. and Wu, G. (2012). Small GTPase regulation of GPCR anterograde trafficking. Trends Pharmacol. Sci. 33, 28–34.10.1016/j.tips.2011.09.002Search in Google Scholar PubMed PubMed Central

Yulyaningsih, E., Zhang, L., Herzog, H., and Sainsbury, A. (2011). NPY receptors as potential targets for anti-obesity drug development. Br. J. Pharmacol. 163, 1170–1202.10.1111/j.1476-5381.2011.01363.xSearch in Google Scholar PubMed PubMed Central

Zhang, L., Bijker, M.S., and Herzog, H. (2011). The neuropeptide Y system: pathophysiological and therapeutic implications in obesity and cancer. Pharmacol. Ther. 131, 91–113.10.1016/j.pharmthera.2011.03.011Search in Google Scholar PubMed

Received: 2013-1-19
Accepted: 2013-2-22
Published Online: 2013-03-01
Published in Print: 2013-08-01

©2013 by Walter de Gruyter Berlin Boston

Articles in the same Issue

  1. Masthead
  2. Masthead
  3. Guest Editorial
  4. Highlight: Protein states with cell biological and medicinal relevance
  5. HIGHLIGHT: PROTEIN STATES WITH CELL BIOLOGICAL AND MEDICAL RELEVANCE
  6. Towards improved receptor targeting: anterograde transport, internalization and postendocytic trafficking of neuropeptide Y receptors
  7. Progress in demystification of adhesion G protein-coupled receptors
  8. The unresolved puzzle why alanine extensions cause disease
  9. Molecular function of the prolyl cis/trans isomerase and metallochaperone SlyD
  10. Structure and allosteric regulation of eukaryotic 6-phosphofructokinases
  11. Polyionic and cysteine-containing fusion peptides as versatile protein tags
  12. p0071/PKP4, a multifunctional protein coordinating cell adhesion with cytoskeletal organization
  13. Lysine-specific histone demethylase LSD1 and the dynamic control of chromatin
  14. Methylation of the nuclear poly(A)-binding protein by type I protein arginine methyltransferases – how and why
  15. Oxidative in vitro folding of a cysteine deficient variant of the G protein-coupled neuropeptide Y receptor type 2 improves stability at high concentration
  16. Identification of prolyl oligopeptidase as a cyclosporine-sensitive protease by screening of mouse liver extracts
  17. In vitro maturation of Drosophila melanogaster Spätzle protein with refolded Easter reveals a novel cleavage site within the prodomain
  18. Subcellular localization and RNP formation of IGF2BPs (IGF2 mRNA-binding proteins) is modulated by distinct RNA-binding domains
  19. High level expression of the Drosophila Toll receptor ectodomain and crystallization of its complex with the morphogen Spätzle
Downloaded on 15.1.2026 from https://www.degruyterbrill.com/document/doi/10.1515/hsz-2013-0123/html
Scroll to top button