Keywords: aquaporin-2, kidney, phosphoproteomics, protein kinases, V2 vasopressin receptor
Abstract
Vasopressin controls renal water excretion through actions to regulate aquaporin-2 (AQP2) trafficking, transcription, and degradation. These actions are in part dependent on vasopressin-induced phosphorylation changes in collecting duct cells. Although most efforts have focused on the phosphorylation of AQP2 itself, phosphoproteomic studies have identified many vasopressin-regulated phosphorylation sites in proteins other than AQP2. The goal of this bioinformatics-based review is to create a compendium of vasopressin-regulated phosphorylation sites with a focus on those that are seen in both native rat inner medullary collecting ducts and cultured collecting duct cells from the mouse (mpkCCD), arguing that these sites are the best candidates for roles in AQP2 regulation. This analysis identified 51 vasopressin-regulated phosphorylation sites in 45 proteins. We provide resource web pages at https://esbl.nhlbi.nih.gov/Databases/AVP-Phos/ and https://esbl.nhlbi.nih.gov/AVP-Network/, listing the phosphorylation sites and describing annotated functions of each of the vasopressin-targeted phosphoproteins. Among these sites are 23 consensus protein kinase A (PKA) sites that are increased in response to vasopressin, consistent with a central role for PKA in vasopressin signaling. The remaining sites are predicted to be phosphorylated by other kinases, most notably ERK1/2, which accounts for decreased phosphorylation at sites with a X-p(S/T)-P-X motif. Additional protein kinases that undergo vasopressin-induced changes in phosphorylation are Camkk2, Cdk18, Erbb3, Mink1, and Src, which also may be activated directly or indirectly by PKA. The regulated phosphoproteins are mapped to processes that hypothetically can account for vasopressin-mediated control of AQP2 trafficking, cytoskeletal alterations, and Aqp2 gene expression, providing grist for future studies.
NEW & NOTEWORTHY Vasopressin regulates renal water excretion through control of the aquaporin-2 water channel in collecting duct cells. Studies of vasopressin-induced protein phosphorylation have focused mainly on the phosphorylation of aquaporin-2. This study describes 44 phosphoproteins other than aquaporin-2 that undergo vasopressin-mediated phosphorylation changes and summarizes potential physiological roles of each.
INTRODUCTION
Vasopressin controls the osmotic water permeability of the renal collecting duct through actions in principal cells to regulate the water channel aquaporin-2 (AQP2) (1–10). AQP2 is regulated in at least three ways: 1) vasopressin controls trafficking of AQP2-containing vesicles to and from the plasma membrane (11–14), 2) vasopressin increases the half-life of the AQP2 protein (15, 16), and 3) vasopressin increases the rate of transcription of the Aqp2 gene (4, 17). In addition, extensive evidence indicates that the regulation of trafficking of AQP2 depends on vasopressin-induced rearrangements in the actin cytoskeleton (18–22). These actions depend on the binding of arginine vasopressin to V2 receptors in the basolateral plasma membrane, resulting in Gαs-mediated activation of adenylyl cyclase 6, increasing cAMP levels and protein kinase A (PKA) activity (8, 23). Ultimately, the effects of vasopressin in collecting duct principal cells depend almost entirely on PKA (24–26), either through direct PKA-mediated phosphorylation of proteins or indirectly through PKA-mediated regulation of other protein kinases (27). An example of indirect effects of PKA activation are mitogen-activated protein kinases (MAPKs), ERK1 and ERK2, which are decreased in activity in response to vasopressin in native collecting ducts in part via PKA-dependent Raf1 inhibition (28). In addition to cAMP, an important second messenger is Ca2+, which is increased in collecting duct principal cells in response to V2 receptor occupation (29–31). This response has been proposed to be secondary to PKA-dependent phosphorylation of inositol 1,4,5-trisphosphate receptors Itpr1 and Itpr3, sensitizing them to activation by fluctuations in basal Ca2+ levels (24). Regulation of water permeability in collecting ducts occurs in part through Ca2+/calmodulin-mediated activation of myosin light chain kinase (Mylk), a regulator of conventional nonmuscle myosins in collecting duct cells (20). Beyond its action to phosphorylate myosin regulatory light chain, several noncanonical targets of Mylk have been identified by CRISPR/Cas9-mediated gene deletion followed by phosphoproteomics (32). Beyond Mylk, there are many other kinases that are regulated by Ca2+/calmodulin, including Ca2+/calmodulin-dependent protein kinase type IIδ (Camk2d), which has been proposed to phosphorylate AQP2 at Ser256 (33, 34).
The intracellular actions of vasopressin in the renal collecting duct are shown in Table 1 (4, 11–22, 24, 29–31, 35–42). These actions all appear to be responses to vasopressin-dependent phosphorylation events catalyzed by PKA and PKA-regulated downstream protein kinases. However, the specific phosphorylation events that mediate the actions shown in Table 1 are largely unidentified. We know from phosphoproteomics and immunoblot analysis with phospho-specific antibodies that AQP2 itself undergoes changes in phosphorylation in a cluster of four COOH-terminal serines (Ser256, Ser261, Ser264, and Ser269), which are known as pivotal phosphorylation sites for AQP2 trafficking to the plasma membrane (7, 43–55). Specifically, Ser256, Ser264, and Ser269 undergo increases in phosphorylation in response to vasopressin (45, 47, 48, 56–59), whereas Ser261 undergoes a decrease (48, 57, 60). Recent studies have used powerful quantitative phosphoproteomics methods to identify additional vasopressin-dependent phosphorylation changes in native rat inner medullary collecting duct (IMCD) cells (61) and cultured mpkCCD cells derived from mouse collecting ducts (27). The former identified 156 vasopressin-regulated phosphopeptides, and the latter identified 429 vasopressin-regulated phosphopeptides. A long-term goal is to identify the roles of all of these sites in the vasopressin-regulated processes shown in Table 1. In this paper, to prioritize vasopressin-regulated phosphorylation sites for future study, we identified the vasopressin-regulated sites that are common to both rat native IMCDs and cultured mouse mpkCCD cells, arguing that phosphorylation sites that change in the same way in both models are the most likely to play physiological roles in the regulation of AQP2. Also, we except that the most consistent findings between rat and mouse collecting ducts will be representative of the species-independent behavior of the mammalian collecting duct. Using this approach, we identified 51 phosphorylation sites in 45 phosphoproteins that show the same changes in mpkCCD cells and IMCD cell suspensions in response to the V2 receptor-specific vasopressin analog desmopressin [1-deamino-8-d-arginine vasopressin (dDAVP)]. We mapped these phosphoproteins to known cellular actions of vasopressin (Table 1) and created a data resource to allow users to access the information. Although the major focus of this paper is on the regulation of AQP2, the information shown in Table 1 emphasizes that vasopressin regulates other processes in collecting duct cells (such as apoptosis, cell proliferation, and tight junction permeability) and some of the phosphorylation changes discussed here may be relevant to these other vasopressin-regulated processes.
Table 1.
Intracellular processes regulated by vasopressin in collecting duct cells
Vasopressin Actions | Reference(s) |
---|---|
Increased exocytosis of AQP2-containing vesicles | (11, 12, 14) |
Decreased AQP2 endocytosis | (11, 13) |
Increased AQP2 protein half life | (15, 16) |
Increased Aqp2 gene transcription | (4, 17) |
Intracellular Ca2+ mobilization | (29–31) |
Reorganization of actin filaments | (18–20, 22) |
Depolymerization of filamentous actin | (18, 19, 21) |
Microtubule-dependent AQP2 redistribution | (35) |
Increased tight junction permeability | (36) |
Decreased rate of apoptosis | (37) |
Decreased rate of proliferation | (38) |
Increased principal cell size | (39–42) |
Decreased ERK1/2 activity | (24, 28) |
AQP2, aquaporin-2.
METHODS
Curation of Consensus Vasopressin-Regulated Phosphorylation Sites
Phosphorylation sites that significantly changed in abundance in response to 30-min treatment of the vasopressin analog dDAVP in both cultured mouse mpkCCD cells (27) and in native rat IMCD suspensions (61) were curated into a single data set. Because amino acid numbering in rat versus mouse proteins can differ, curation was done by mapping the four amino acid sequences surrounding the phosphorylation sites (X-X-p[S/T]-X), that is, by comparing two amino acids upstream from a given site and one amino acid downstream from the site. The mapping process is provided in Supplemental Spreadsheet S1 (see https://esbl.nhlbi.nih.gov/Databases/Supplemental-Phosphorylation/). The final list of consensus vasopressin-regulated phosphorylation sites is provided at https://esbl.nhlbi.nih.gov/Databases/AVP-Phos/. Automated Bioinformatic Extractor (ABE; https://esbl.nhlbi.nih.gov/ABE/) was used to create annotations for individual proteins based on official gene symbols. Protein functions were derived from the “[FUNCTION]” fields in UniProt records for individual proteins. Reported amino acid numbering is based on the UniProt Accession numbers provided at https://esbl.nhlbi.nih.gov/Databases/AVP-Phos/. Our laboratory carried out previous quantitative phosphoproteomic analysis in mouse mpkCCD cells (62) and rat IMCD cells (63), but these were not included in the present analysis owing to the fact that earlier LC-MS/MS methodologies were far less sensitive, leading to many fewer sites being quantified. In general, the sites found in these prior studies overlapped the sites reported here and showed the same direction of change.
RESULTS AND DISCUSSION
We curated a list of 51 phosphorylation sites in 45 proteins that significantly changed in abundance in response to the vasopressin analog dDAVP in both cultured mouse mpkCCD cells (27) and in native rat IMCD suspensions (61) (organized by “Molecular Function” in Fig. 1). The curation process is detailed in Supplemental Spreadsheet S1. Information about these sites is archived for browsing and download at https://esbl.nhlbi.nih.gov/Databases/AVP-Phos/. The vasopressin-regulated phosphoproteins are mapped to known effects of vasopressin at https://esbl.nhlbi.nih.gov/AVP-Network/. This webpage displays documentation from prior literature about each node by hovering over the node. Although the annotations provided focus mainly on regulation of AQP2-mediated water transport, we acknowledge that vasopressin regulates sodium transport in the cortical collecting duct and regulates urea transport in the IMCD and findings may be relevant to these regulatory events as well.
Figure 1.
Consensus vasopressin-regulated protein phosphorylation sites in collecting duct cells. Two separate datasets, one from the cultured collecting duct cells from the mouse (mpkCCD) cell line treated with 0.1 nM desmopressin [1-deamino-8-d-arginine vasopressin (dDAVP)] for 30 min and one from an inner medullary collecting duct (IMCD) suspension treated with 1 nM dDAVP for 30 min, were combined to identify the phosphorylation sites that changed significantly in response to vasopressin in both studies (27, 61). 51 consensus dDAVP-responsive phosphosites are identified.
Figure 2 shows a graphical display of the vasopressin-regulated phosphorylation sites mapped to molecular functions. Most of the vasopressin-regulated phosphorylation sites have not been studied in the context of the collecting duct, and their listing in Figs. 1 and 2 provide a convenient list of hypotheses for future investigations.
Figure 2.
Relational graph showing the vasopressin-regulated phosphoproteins mapped to molecular functions. An annotated version with popups showing descriptions of each node can be found online at https://esbl.nhlbi.nih.gov/AVP-Network/. The 51 phosphosites are categorized into 11 vasopressin-associated cellular processes (see Table 1). Each cellular process is known to play roles in the regulation of collecting duct water permeability.
The molecular functions highlighted in Figs. 1 and 2 include several categories related to known aspects of vasopressin-mediated signaling in collecting duct cells, viz. “cAMP-dependent signaling,” “calcium signaling,” “regulation of GTPase activity,” and “protein phosphorylation.” Notably, in the latter category, there were several protein kinases, namely misshapen-like kinase 1 (Mink1), Ca2+/calmodulin-dependent protein kinase kinase 2 (Camkk2), cyclin-dependent kinase 18 (Cdk18), Erbb3, and Src, which play likely roles downstream from PKA. The remaining molecular function categories shown in Figs. 1 and 2 also correspond closely to cellular level effects of vasopressin shown in Table 1, viz. “vesicle-mediated transport,” “regulation of transcription,” “actin cytoskeleton,” “microtubule cytoskeleton,” and “cell polarity.” An additional category, “RNA processing,” was not anticipated from prior studies.
Phosphorylation Changes in Response to Vasopressin Are Dependent on PKA Activity
We hypothesized that the demonstrated vasopressin-regulated phosphorylation sites shown in Fig. 1 are either PKA targets or are downstream from PKA actions, e.g., through regulation of non-PKA kinases secondary to PKA activity changes. If so, the phosphorylation changes seen with deletion of PKA should be opposite to the changes seen in response to vasopressin. The PKA catalytic subunit is coded by two separate genes expressing PKA-Cα and PKA-Cβ proteins. Isobe et al. (24) deleted both (PKA double knockout) in mouse mpkCCD cells and carried out quantitative phosphoproteomics. Figure 3A shows a plot of phosphorylation changes in response to vasopressin in mpkCCD cells versus changes in response to PKA deletion for all sites quantified in both studies. In general, there was a strong negative correlation, confirming the hypothesis. Specifically, this provides further support for the prior observation that almost all vasopressin-mediated signaling in collecting duct cells is dependent on PKA activation (25). One exception is the regulatory subunit of PKA, Prkar2a, which inhibits PKA activity by binding to the catalytic site when cAMP is absent. Its phosphorylation increases at Ser96 when bound to the catalytic subunit, i.e., in the absence of cAMP (64). The complete set of phosphorylation changes in response to PKA deletion are shown as a volcano plot in Fig. 3B. The two catalytic subunit proteins, PKA-Cα and PKA-Cβ, have extremely similar amino acid sequences in their catalytic regions and would be expected to phosphorylate the same proteins if they were similarly located in the cell. As shown in Fig. 3C, these two PKA proteins have different effects on phosphorylation when deleted individually, implying that they may be compartmentalized to interact with different substrates (26). Specifically, it was concluded that PKA-Cα deletion chiefly affected phosphorylation of proteins associated with cell membranes and membrane vesicles, whereas target proteins in PKA-Cβ-null cells were largely associated with the actin cytoskeleton and cell junctions (26). Among the phosphorylation sites that stand out in Fig. 3C are Bin1, Sec22b, Slc9a3r1, Agfg1, and Ralgapa2, which change in opposite directions in response to deletion of the two PKA catalytic proteins. Overall, we conclude that vasopressin-dependent signaling is almost entirely PKA dependent, but the two PKA catalytic proteins perform different functions, presumably due to differences in subcellular localization.
Figure 3.
Role of protein kinase A (PKA) in vasopressin-mediated phosphorylation. A: correlation between the response to vasopressin for the sites shown in Fig. 1 and response to deletion of both PKA catalytic genes in cultured collecting duct cells from mouse (mpkCCD) cells. The plot shows a strong negative correlation (P < 0.01). B: volcano plot showing changes in phosphorylation for all sites in mpkCCD cells in response to deletion of both PKA catalytic genes, highlighting the phosphorylation sites shown in Fig. 1. C: desmopressin [1-deamino-8-d-arginine vasopressin (dDAVP)]-responsive phosphosites show differing responses to PKA-Cα single knockout (KO) versus PKA-Cβ single KO mpkCCD cells. Many of the dDAVP-responsive phosphosites showed opposite phosphorylation changes in PKA-Cα KO versus PKA-Cβ KO cells, suggesting that the two PKA catalytic subunits have different functions in the cell. Green indicates sites with increased phosphorylation in response to dDAVP; pink indicates sites with decreased phosphorylation in response to dDAVP. Ctrl, control.
Some Vasopressin-Regulated Phosphorylation Events Have Known Effects on Protein Function
Among the 51 phosphorylation sites regulated by vasopressin, eight sites have been investigated in prior studies in noncollecting duct tissues with regard to the effect of phosphorylation on the activities of the phosphoproteins. These are shown in Table 2 (65–78). The following paragraphs link these sites to possible roles in collecting duct principal cells.
Table 2.
Effect of phosphorylation on protein activity for vasopressin-regulated phosphorylation sites
Gene | Phosphosite | Effect of Phosphorylation | Reference(s) |
---|---|---|---|
Arhgef2 | Ser885 | PKA-mediated phosphorylation at Ser885 inhibits the GEF activity | (65, 66) |
Exoc7 | Ser250 | ERK1/2-mediated phosphorylation stimulates exocyst complex assembly and exocytosis | (67) |
Hdac4 | Ser245 | CaMK- and Mark kinase-mediated phosphorylation at Ser245 promote the interaction with 14-3-3, which induces nuclear export and cytoplasmic accumulation | (68, 69) |
Itpr1 | Ser1588, Ser1755 | PKA-mediated phosphorylation at Ser1588 and Ser1755 enhances Ca2+ release | (70, 71) |
Nsfl1c | Ser176 | PKA-mediated phosphorylation at Ser176 induces Nsfl1c activity | (72) |
Src | Ser17 | PKA-mediated phosphorylation at Ser17 induces Src activity | (73) |
Stim1 | Ser575 | ERK1/2-mediated phosphorylation at Ser575 activates Stim1-mediated store-operated Ca2+ entry | (74, 75) |
Zfp36l1 | Ser54 | MAPK-activated protein kinase 2-mediated phosphorylation at Ser54 inhibits mRNA decay. Site is also phosphorylated by PKA | (76–78) |
Rho guanine nucleotide exchange factor 2 (Arhgef2; also known as Lfc and GEF-H1) plays a role in the regulation of the state of actin polymerization (65, 66). Vasopressin increases phosphorylation at Ser885 (Fig. 1), likely mediated directly by PKA. It has been found that phosphorylation of Arhgef2 at Ser885 inhibits its GEF activity and its ability to activate RhoA (65, 66). Given the prior observations that vasopressin action is associated with the depolymerization of F-actin (18, 19, 21), thought to be critical to apical trafficking of AQP2, and that the depolymerization is in part dependent on RhoA (19, 79, 80), it appears plausible that phosphorylation of Arhgef2 at Ser885 plays an important causal role in F-actin depolymerization and AQP2 trafficking to the apical plasma membrane.
Exocyst complex component 7 (Exoc7) is an element of the exocyst complex that consists of Sec3, Sec5, Sec6, Sec8, Sec10, Sec15, Exoc7, and Exoc8. Vasopressin decreases phosphorylation of Exoc7 at Ser250 (Fig. 1), a likely target of ERK1 or ERK2. The exocyst complex is associated with vesicle trafficking via post-Golgi vesicle targeting to the basolateral plasma membrane (81, 82). As a direct substrate of ERK1/2, phosphorylation of Exoc7 at Ser250 stimulates exocytosis to the basolateral plasma membrane by promoting assembly of the exocyst complex (67). Potentially, therefore, vasopressin’s action to reduce Ser250 phosphorylation could reduce exocytosis of AQP2 to the basolateral plasma membrane in lieu of the apical plasma membrane.
Histone deacetylase 4 (Hdac4) plays a role in transcriptional processes by regulating chromatin accessibility. Phosphorylation at Ser245 (along with two neighboring sites) is required for the binding of Hdac4 to 14-3-3 and sequestration in the cytoplasm (83). The sequestration results in the prevention of histone H3 deacetylation at Lys27 (H3K27Ac), an activation chromatin mark (84). Vasopressin decreases phosphorylation of Hdac4 at Ser245 (Fig. 1), so it is predicted to translocate to the nucleus and deacetylate histone H3 at Lys27 at the transcriptional start sites of target genes. Since histone H3K27Ac increases at the AQP2 promoter (85), this change is unlikely to be the consequence of decreased phosphorylation of Hdac4 at Ser245, which should cause the opposite effect. However, vasopressin action results in the repression of many genes (17, 86), which could potentially be the result of the observed phosphorylation change. The protein kinase responsible for phosphorylation of Ser245 of Hdac4 has been proposed to be Ca2+/calmodulin-dependent protein kinase (CaMK) (68, 87). In addition, polarity kinases [microtubule affinity regulating kinases (Mark2 and Mark3)] have also been shown to be capable of phosphorylating Ser245 in Hdac4 (69). Ser245 is found in a classical SNF1 family motif (L-X-(R/K)-X-X-p(S/T)-X-X-X-L), typical of microtubule affinity regulating kinases (88).
Itpr1 is an intracellular Ca2+ channel in the endoplasmic reticulum (ER) membrane, which mediates Ca2+ release from the ER (70). It has been reported that the Ca2+-releasing activity of Itpr1 is upregulated by PKA-mediated phosphorylation and substitution of two PKA consensus sequences (R-R-X-p[S/T]) at Ser1588 and Ser1755 to alanine inhibits Ca2+ mobilization (70, 71). As shown in Fig. 1, both of these sites display increases in phosphorylation in response to vasopressin, providing an explanation for the observation that vasopressin action through the V2 receptor increases the intracellular Ca2+ concentration (29–31). Working via calmodulin, vasopressin-dependent Ca2+ mobilization can activate certain protein kinases, such as Mylk (16, 20, 32) and calmodulin-dependent kinase IIδ, that are involved in the regulation of AQP2.
NSFL1 (p97) cofactor p47 (Nsfl1c) is a cofactor of valosin-containing protein (VCP; also known as p97), which has multiple functions in cellular processes, such as protein degradation and membrane fusion/trafficking (72, 89). Phosphorylation of Nsfl1c at Ser176 is mediated by PKA and the phosphorylation enhances the activity of VCP/p47 in mouse neuronal cells (72). As shown in Fig. 1, vasopressin increases phosphorylation of Nsfl1c at Ser176, possibly playing a role in vasopressin-mediated vesicle fusion and AQP2 trafficking to the plasma membrane.
The tyrosine-directed protein kinase Src (Src) is involved in diverse cellular signaling pathways. Vasopressin causes an increase in Src phosphorylation at Ser17 (Fig. 1), thought to be a PKA site (90). Src phosphorylation at Ser17 in the amino terminus facilitates its activity and activates the small G protein Rap1 (73, 90). In collecting duct cells, Rap1 activation inhibits AQP2 endocytosis, increasing retention of AQP2 in the apical plasma membrane (55). Thus, PKA-mediated phosphorylation of Src at Ser17 may cause AQP2 plasma membrane retention by activation of Rap1. Against this hypothesis is the observation that application of a tyrosine kinase inhibitor, dasatinib, is associated with increased accumulation of AQP2 in the apical region of IMCD cells (91). In addition, Rap1 plays an important role of inhibiting the ability of Ras to activate the ERK cascade (92), so vasopressin-mediated phosphorylation of Src at Ser17 may account for the ability of vasopressin to reduce ERK activity (24, 28). Overall, the conflicting observations described earlier indicate that more work will be required to fully understand the role of Src and other tyrosine kinases, including growth factor receptors, and should be a priority for future studies.
Stromal interaction molecule 1 (Stim1) is a key regulator of Ca2+ signaling, which activates store-operated Ca2+ entry by sensing Ca2+ (93). In response to Ca2+ depletion in the ER, Stim1 activates store-operated Ca2+ channels in the plasma membrane, such as Orai1, and promotes Ca2+ influx (94). Vasopressin decreases phosphorylation of Stim1 at Ser575 (Fig. 1), a likely target site for ERK1 or ERK2. Prior studies have shown that ERK1/2 phosphorylation of Stim1 at Ser575 facilitates Stim1 multimerization, which is a necessary prerequisite for its role in promoting Ca2+ entry (74, 75). Thus, the overall effect of vasopressin in the collecting duct is predicted to be inhibition of plasma membrane uptake of Ca2+.
mRNA decay activator protein ZFP36L1 (Zfp36l1) is an mRNA-destabilizing protein interacting with AU-rich elements (AREs) located in the 3′-untranslated regions of target mRNAs (95). Thus, Zfp36l1-dependent mRNA degradation is an important determinant of the abundances of target mRNAs that may exert control of gene expression in a transcription-independent manner. Vasopressin increases phosphorylation of Zfp36l1 at Ser54, present in a motif compatible with phosphorylation by PKA (R-R-X-pS-V) (Fig. 1). Prior studies in other tissues have indeed shown that PKA (78) can phosphorylate Ser54. Also, MAPK-activated protein kinase 2 (MAPKAPK2) is capable of phosphorylating Zpf36l1 at this site (76, 77). Phosphorylation at Ser54 inhibits binding to the target mRNA and would be expected to increase target mRNA abundances. This action is unlikely to play a role in the vasopressin-mediated increase in Aqp2 mRNA (Table 1), given that the AQP2 transcript lacks an ARE motif in its 3′-untranslated region (17). However, a number of other vasopressin-regulated transcripts do contain ARE motifs (17) and are candidates for regulation by Zfp36l1-mediated mRNA degradation. Interestingly, Zfp36l1 is also regulated by vasopressin in another way. Cai et al. (96) found that the vasopressin analog dDAVP given for 7 days decreased mRNA abundance of Zfp36l1 protein in the mouse renal inner medulla to 66% of the control value. The relationship between the short-term phosphorylation response and the long-term protein abundance response is so far unexplored.
Time Course of Vasopressin-Induced Phosphorylation Changes
The data used to develop Fig. 1 were obtained after 30 min of vasopressin addition in both mpkCCD cells and the IMCD. To provide additional information about earlier time points, we cite a recent study that identified the time courses of phosphorylation changes in rat IMCD suspensions in response to the vasopressin analog dDAVP in the timeframe of 1–15 min (97). Figure 4 shows the time courses for phosphorylation sites identified in Fig. 1, categorized according to the molecular functions regulated by vasopressin. In general, the phosphorylation responses were complete within 2–5 min including the responses in the “vesicle-mediated transport” group of phosphoproteins. Typically, the water permeability response in isolated perfused collecting ducts displays an initial increase requiring ∼10 min followed by a slower secondary response lasting at least 30 additional min (11, 98). Thus, phosphorylation changes appear to be complete before water permeability changes. These data suggest that phosphorylation changes are not rate limiting for the water permeability response and that the dynamics may be due to secondary changes following the phosphorylation changes such as other posttranslational modifications.
Figure 4.
Dynamics of phosphorylation responses to vasopressin for different phosphorylation sites in native inner medullary collecting duct (IMCD) cells. Data are replotted from Leo et al. (97). A: sites associated with vesicle-mediated transport: Bin1 (Ser304), Exoc7 (Ser250), Nsfl1c (Ser176), and Sec22b (Ser137). B: sites associated with the regulation of GTPase activity/actin cytoskeleton/microtubule cytoskeleton: Agfg1 (Ser181), Arfgef1 (Ser1566), Arhgef2 (Ser885), Specc1l (Ser385), and Rmdn3 (Ser46). C: sites associated with the regulation of transcription: Ctnnb1 (Ser552 and Thr551), Lrrfip1 (Ser88), and Tsc22d4 (Thr223). D: sites associated with protein phosphorylation: Camkk2 (Ser495 and Ser511), CDK18 (Ser66), Erbb3 (Ser980), and Src (Ser17). E: sites associated with cAMP-dependent signaling/Ca2+ signaling: Prkar2a (Ser96), Itpr1 (Ser1588), Itpr3 (Ser1832), and Stim1 (Ser575). F: sites associated with cell polarity: Igsf5 (Ser335), Luzp (Ser261), and Slc9a3r1 (Ser275). Some sites shown in Fig. 1 did not yield time course data in Leo et al. (97). In general, phosphorylation responses were faster than observed water permeability responses, indicating that phosphorylation is not rate limiting for the overall response (see text).
Hypotheses about Roles of Vasopressin-Mediated Phosphorylation Changes in AQP2 Trafficking and Increases in Aqp2 Gene Expression
In Figs. 5, 6, and 7, we relate the phosphoproteins whose phosphorylation states are regulated by vasopressin to known physiological actions of vasopressin at a cellular level. These provide hypotheses for future studies of the mechanisms of AQP2 regulation in collecting duct principal cells.
Figure 5.
Vasopressin-regulated phosphoproteins with hypothetical roles in “vesicle-mediated transport.” Listed phosphoproteins are likely to be involved in three biological processes, “SNARE-mediated vesicle fusion,” “endosomal biogenesis,” and “interaction with dynamins,” which could regulate vasopressin-mediated aquaporin-2 (AQP2) trafficking. Cdk18, cyclin-dependent kinase 18; PKA, protein kinase A.
Figure 6.
Vasopressin-regulated phosphoproteins with hypothetical roles in the regulation of the “actin cytoskeleton,” “microtubule cytoskeleton,” and “GTPase activity.” The phosphorylation responses could potentially be involved in the regulation of aquaporin-2 (AQP2) trafficking. PKA, protein kinase A.
Figure 7.
Vasopressin-regulated phosphoproteins with hypothetical roles in the “regulation of transcription” and “RNA processing.” The phosphorylation responses could potentially be involved in the regulation of the abundance of aquaporin-2 (Aqp2) mRNA and secondarily AQP2 protein. PKA, protein kinase A.
Figure 5 shows the phosphoproteins that map to “vesicle-mediated transport” in Fig. 2. These can be categorized in three groups, viz. those that mediate “SNARE-mediated vesicle fusion,” “endosomal biogenesis,” and “interaction with dynamin.” Five of the phosphoproteins have recognized roles in SNARE-mediated vesicle fusion and therefore are candidates for roles in vasopressin-stimulated AQP2 exocytosis (11, 12, 14). Two of the phosphoproteins are known to interact with dynamin and therefore are candidates for roles in vasopressin’s action to inhibit endocytosis (11, 13). The other two phosphoproteins have roles in endosome biogenesis and therefore could be involved in AQP2 recycling or AQP2 degradation via late endosomes and lysosomes. Six of these proteins are likely direct targets of PKA, whereas two are likely targets of ERK1 or ERK2, both of which are downregulated through the action of PKA (24, 28). The remaining protein, amphiphysin-1, is a likely target of Cdk18 (or PCTAIRE-3). Cdk18 is activated by PKA-mediated phosphorylation at three sites, including Ser66 (99). In addition, Cdk18 mRNA abundance is also upregulated in collecting duct mpkCCD cells in response to vasopressin (27, 86). In addition, Cdk18 protein may be regulated through control of its degradation, as it was found in a complex with the ubiquitin ligase Stub1 and regulatory subunits of PKA bound to AQP2 (100).
Figure 6 shows the phosphoproteins that map to three processes related to regulation of the cytoskeleton (Fig. 2), namely, “actin cytoskeleton,” “microtubule cytoskeleton,” and “regulation of GTPase activity.” All of these have relevance to AQP2 trafficking to and from the plasma membrane (Table 1). Five of these phosphoproteins are involved in the regulation of activities of Rho, Rac, and CDC42, small GTP-binding proteins that play central roles in the regulation of the actin cytoskeleton (101). RhoA in particular has been shown to play key roles in the regulation of AQP2 trafficking in collecting duct cells (19, 79, 80). Three phosphoproteins (with 4 phosphorylation sites) are directly involved in “actin cytoskeleton rearrangement,” which is a well-described cellular process in response to vasopressin (20, 22). Finally, two phosphoproteins (Clip1 and Rmdn3) are involved in “microtubule cytoskeleton organization.” The microtubule cytoskeleton has been shown to have roles in the determination of AQP2 localization within collecting duct cells (35, 102–106), and phosphorylation of Clip1 and Rmdn3 can be hypothesized to be involved in these processes.
Figure 7 shows the phosphoproteins that map to “regulation of transcription” and “RNA processing,” which together determine cellular levels of individual mRNAs (Fig. 2). One phosphoprotein, Zfp36l1, regulates mRNA degradation, as detailed earlier. Four phosphoproteins (5 phosphorylation sites) are transcription factors and transcriptional coregulators that are directly associated with control of RNA polymerase II-mediated transcription. Hypothetically, vasopressin-mediated phosphorylation changes in these proteins are involved in vasopressin-mediated regulation of AQP2 transcription (4, 17). Finally, an additional phosphoprotein, Hdac4, is involved in the regulation of DNA accessibility by control of histone H3 acetylation, as detailed earlier.
Perspectives and Significance
Vasopressin regulates renal water excretion through control of the AQP2 water channel in collecting duct cells. Studies of vasopressin-induced protein phosphorylation have focused mainly on the phosphorylation of AQP2. This study describes 44 phosphoproteins other than AQP2 that undergo vasopressin-mediated phosphorylation changes and summarizes potential physiological roles for each. Few of these sites have been studied in detail. We have laid out a bioinformatic analysis, highlighting potential roles for these sites, both in the regulation of AQP2 and in the mediation of other actions of vasopressin highlighted in Table 1. Although much has been gained through the systems/-omics approach, the full identification of physiological roles of each of the phosphorylation sites will likely require a reductionist/hypothesis-driven approach. New techniques of genome-editing (e.g., CRISPR/Cas9) have been developed, which will allow each of the 51 phosphorylation sites to be mutated, in either animals or cultured cells, to learn their roles in vasopressin-mediated regulation of collecting duct function.
DATA AVAILABILITY
Data access: https://esbl.nhlbi.nih.gov/Databases/AVP-Phos/ and https://esbl.nhlbi.nih.gov/AVP-Network/.
SUPPLEMENTAL DATA
Supplemental Spreadsheet S1: https://esbl.nhlbi.nih.gov/Databases/Supplemental-Phosphorylation/.
GRANTS
This work was funded by the Division of Intramural Research, National Heart, Lung, and Blood Institute Grants ZIAHL001285 and ZIAHL006129 (to M.A.K.).
DISCLOSURES
No conflicts of interest, financial or otherwise, are declared by the authors.
AUTHOR CONTRIBUTIONS
M.A.K. conceived and designed research; E.P., C-R.Y., V.R., V.D., A.D., B.G.P., K.T.L., H.K., L.C., C-L.C., and M.A.K. analyzed data; E.P., C-R.Y., V.R., V.D., A.D., B.G.P., K.T.L., H.K., L.C., C-L.C., and M.A.K., interpreted results of experiments; E.P. and M.A.K. prepared figures; E.P., C-R.Y., V.R., L.C., C-L.C., and M.A.K. drafted manuscript; E.P., C-R.Y., V.R., V.D., A.D., B.G.P., K.T.L., H.K., L.C., C-L.C., and M.A.K. edited and revised manuscript; E.P., C-R.Y., V.R., V.D., A.D., B.G.P., K.T.L., H.K., L.C., C-L.C., and M.A.K. approved final version of manuscript.
REFERENCES
- 1. Nielsen S, Frokiaer J, Marples D, Kwon TH, Agre P, Knepper MA. Aquaporins in the kidney: from molecules to medicine. Physiol Rev 82: 205–244, 2002. doi: 10.1152/physrev.00024.2001. [DOI] [PubMed] [Google Scholar]
- 2. Noda Y, Sasaki S. Trafficking mechanism of water channel aquaporin-2. Biol Cell 97: 885–892, 2005. doi: 10.1042/BC20040120. [DOI] [PubMed] [Google Scholar]
- 3. Valenti G, Procino G, Tamma G, Carmosino M, Svelto M. Minireview: aquaporin 2 trafficking. Endocrinology 146: 5063–5070, 2005. doi: 10.1210/en.2005-0868. [DOI] [PubMed] [Google Scholar]
- 4. Hasler U, Leroy V, Martin PY, Feraille E. Aquaporin-2 abundance in the renal collecting duct: new insights from cultured cell models. Am J Physiol Renal Physiol 297: F10–F18, 2009. doi: 10.1152/ajprenal.00053.2009. [DOI] [PubMed] [Google Scholar]
- 5. Hoffert JD, Chou CL, Knepper MA. Aquaporin-2 in the “-omics” era. J Biol Chem 284: 14683–14687, 2009. doi: 10.1074/jbc.R900006200. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 6. Brown D, Bouley R, Păunescu TG, Breton S, Lu HAJ. New insights into the dynamic regulation of water and acid-base balance by renal epithelial cells. Am J Physiol Cell Physiol 302: C1421–C1433, 2012. doi: 10.1152/ajpcell.00085.2012. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 7. Park EJ, Kwon TH. A minireview on vasopressin-regulated aquaporin-2 in kidney collecting duct cells. Electrolyte Blood Press 13: 1–6, 2015. doi: 10.5049/EBP.2015.13.1.1. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 8. Pearce D, Soundararajan R, Trimpert C, Kashlan OB, Deen PM, Kohan DE. Collecting duct principal cell transport processes and their regulation. Clin J Am Soc Nephrol 10: 135–146, 2015. doi: 10.2215/CJN.05760513. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 9. Jung HJ, Kwon TH. Molecular mechanisms regulating aquaporin-2 in kidney collecting duct. Am J Physiol Renal Physiol 311: F1318–F1328, 2016. doi: 10.1152/ajprenal.00485.2016. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 10. Olesen ET, Fenton RA. Aquaporin-2 membrane targeting: still a conundrum. Am J Physiol Renal Physiol 312: F744–F747, 2017. doi: 10.1152/ajprenal.00010.2017. [DOI] [PubMed] [Google Scholar]
- 11. Nielsen S, Knepper MA. Vasopressin activates collecting duct urea transporters and water channels by distinct physical processes. Am J Physiol Renal Physiol 265: F204–F213, 1993. doi: 10.1152/ajprenal.1993.265.2.F204. [DOI] [PubMed] [Google Scholar]
- 12. Nielsen S, Chou CL, Marples D, Christensen EI, Kishore BK, Knepper MA. Vasopressin increases water permeability of kidney collecting duct by inducing translocation of aquaporin-CD water channels to plasma membrane. Proc Natl Acad Sci USA 92: 1013–1017, 1995. doi: 10.1073/pnas.92.4.1013. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 13. Brown D. The ins and outs of aquaporin-2 trafficking. Am J Physiol Renal Physiol 284: F893–F901, 2003. doi: 10.1152/ajprenal.00387.2002. [DOI] [PubMed] [Google Scholar]
- 14. Nunes P, Hasler U, McKee M, Lu HA, Bouley R, Brown D. A fluorimetry-based ssYFP secretion assay to monitor vasopressin-induced exocytosis in LLC-PK1 cells expressing aquaporin-2. Am J Physiol Cell Physiol 295: C1476–C1487, 2008. doi: 10.1152/ajpcell.00344.2008. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 15. Nedvetsky PI, Tabor V, Tamma G, Beulshausen S, Skroblin P, Kirschner A, Mutig K, Boltzen M, Petrucci O, Vossenkamper A, Wiesner B, Bachmann S, Rosenthal W, Klussmann E. Reciprocal regulation of aquaporin-2 abundance and degradation by protein kinase A and p38-MAP kinase. J Am Soc Nephrol 21: 1645–1656, 2010. doi: 10.1681/ASN.2009111190. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 16. Sandoval PC, Slentz DH, Pisitkun T, Saeed F, Hoffert JD, Knepper MA. Proteome-wide measurement of protein half-lives and translation rates in vasopressin-sensitive collecting duct cells. J Am Soc Nephrol 24: 1793–1805, 2013. doi: 10.1681/ASN.2013030279. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 17. Sandoval PC, Claxton JS, Lee JW, Saeed F, Hoffert JD, Knepper MA. Systems-level analysis reveals selective regulation of Aqp2 gene expression by vasopressin. Sci Rep 6: 34863, 2016. doi: 10.1038/srep34863. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 18. Simon H, Gao Y, Franki N, Hays RM. Vasopressin depolymerizes apical F-actin in rat inner medullary collecting duct. Am J Physiol Cell Physiol 265: C757–C762, 1993. doi: 10.1152/ajpcell.1993.265.3.C757. [DOI] [PubMed] [Google Scholar]
- 19. Tamma G, Klussmann E, Maric K, Aktories K, Svelto M, Rosenthal W, Valenti G. Rho inhibits cAMP-induced translocation of aquaporin-2 into the apical membrane of renal cells. Am J Physiol Renal Physiol 281: F1092–F1101, 2001. doi: 10.1152/ajprenal.0091.2001. [DOI] [PubMed] [Google Scholar]
- 20. Chou CL, Christensen BM, Frische S, Vorum H, Desai RA, Hoffert JD, de Lanerolle P, Nielsen S, Knepper MA. Non-muscle myosin II and myosin light chain kinase are downstream targets for vasopressin signaling in the renal collecting duct. J Biol Chem 279: 49026–49035, 2004. doi: 10.1074/jbc.M408565200. [DOI] [PubMed] [Google Scholar]
- 21. Noda Y, Horikawa S, Kanda E, Yamashita M, Meng H, Eto K, Li Y, Kuwahara M, Hirai K, Pack C, Kinjo M, Okabe S, Sasaki S. Reciprocal interaction with G-actin and tropomyosin is essential for aquaporin-2 trafficking. J Cell Biol 182: 587–601, 2008. doi: 10.1083/jcb.200709177. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 22. Loo CS, Chen CW, Wang PJ, Chen PY, Lin SY, Khoo KH, Fenton RA, Knepper MA, Yu MJ. Quantitative apical membrane proteomics reveals vasopressin-induced actin dynamics in collecting duct cells. Proc Natl Acad Sci USA 110: 17119–17124, 2013. doi: 10.1073/pnas.1309219110. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 23. Mehta YR, Lewis SA, Leo KT, Chen L, Park E, Raghuram V, Chou CL, Yang CR, Kikuchi H, Khundmiri S, Poll BG, Knepper MA. ADPKD-“omics”: determinants of cyclic AMP levels in renal epithelial cells. Kidney Int 101: 47–62, 2022. doi: 10.1016/j.kint.2021.10.014. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 24. Isobe K, Jung HJ, Yang CR, Claxton J, Sandoval P, Burg MB, Raghuram V, Knepper MA. Systems-level identification of PKA-dependent signaling in epithelial cells. Proc Natl Acad Sci USA 114: E8875–E8884, 2017. doi: 10.1073/pnas.1709123114. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 25. Datta A, Yang CR, Limbutara K, Chou CL, Rinschen MM, Raghuram V, Knepper MA. PKA-independent vasopressin signaling in renal collecting duct. FASEB J 34: 6129–6146, 2020. doi: 10.1096/fj.201902982R. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 26. Raghuram V, Salhadar K, Limbutara K, Park E, Yang CR, Knepper MA. Protein kinase A catalytic-α and catalytic-β proteins have nonredundant regulatory functions. Am J Physiol Renal Physiol 319: F848–F862, 2020. doi: 10.1152/ajprenal.00383.2020. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 27. Datta A, Yang CR, Salhadar K, Park E, Chou CL, Raghuram V, Knepper MA. Phosphoproteomic identification of vasopressin-regulated protein kinases in collecting duct cells. Br J Pharmacol 178: 1426–1444, 2021. [Erratum in Br J Pharmacol 178: 2027, 2021]. doi: 10.1111/bph.15352. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 28. Pisitkun T, Jacob V, Schleicher SM, Chou CL, Yu MJ, Knepper MA. Akt and ERK1/2 pathways are components of the vasopressin signaling network in rat native IMCD. Am J Physiol Renal Physiol 295: F1030–F1043, 2008. doi: 10.1152/ajprenal.90339.2008. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 29. Star RA, Nonoguchi H, Balaban R, Knepper MA. Calcium and cyclic adenosine monophosphate as second messengers for vasopressin in the rat inner medullary collecting duct. J Clin Invest 81: 1879–1888, 1988. doi: 10.1172/JCI113534. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 30. Champigneulle A, Siga E, Vassent G, Imbert-Teboul M. V2-like vasopressin receptor mobilizes intracellular Ca2+ in rat medullary collecting tubules. Am J Physiol Renal Physiol 265: F35–F45, 1993. doi: 10.1152/ajprenal.1993.265.1.F35. [DOI] [PubMed] [Google Scholar]
- 31. Chou CL, Yip KP, Michea L, Kador K, Ferraris JD, Wade JB, Knepper MA. Regulation of aquaporin-2 trafficking by vasopressin in the renal collecting duct. Roles of ryanodine-sensitive Ca2+ stores and calmodulin. J Biol Chem 275: 36839–36846, 2000. doi: 10.1074/jbc.M005552200. [DOI] [PubMed] [Google Scholar]
- 32. Isobe K, Raghuram V, Krishnan L, Chou CL, Yang CR, Knepper MA. CRISPR-Cas9/phosphoproteomics identifies multiple noncanonical targets of myosin light chain kinase. Am J Physiol Renal Physiol 318: F600–F616, 2020. doi: 10.1152/ajprenal.00431.2019. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 33. Bradford D, Raghuram V, Wilson JL, Chou CL, Hoffert JD, Knepper MA, Pisitkun T. Use of LC-MS/MS and Bayes' theorem to identify protein kinases that phosphorylate aquaporin-2 at Ser256. Am J Physiol Cell Physiol 307: C123–C139, 2014. doi: 10.1152/ajpcell.00377.2012. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 34. Yang CR, Raghuram V, Emamian M, Sandoval PC, Knepper MA. Deep proteomic profiling of vasopressin-sensitive collecting duct cells. II. Bioinformatic analysis of vasopressin signaling. Am J Physiol Cell Physiol 309: C799–C812, 2015. doi: 10.1152/ajpcell.00214.2015. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 35. Sabolic I, Katsura T, Verbavatz JM, Brown D. The AQP2 water channel: effect of vasopressin treatment, microtubule disruption, and distribution in neonatal rats. J Membr Biol 143: 165–175, 1995. [Erratum in J Membr Biol 145: 107–108, 1995]. doi: 10.1007/BF00233445. [DOI] [PubMed] [Google Scholar]
- 36. Mishler DR, Kraut JA, Nagami GT. AVP reduces transepithelial resistance across IMCD cell monolayers. Am J Physiol Renal Physiol 258: F1561–F1568, 1990. doi: 10.1152/ajprenal.1990.258.6.F1561. [DOI] [PubMed] [Google Scholar]
- 37. Miller RL, Sandoval PC, Pisitkun T, Knepper MA, Hoffert JD. Vasopressin inhibits apoptosis in renal collecting duct cells. Am J Physiol Renal Physiol 304: F177–F188, 2013. doi: 10.1152/ajprenal.00431.2012. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 38. Yamaguchi T, Nagao S, Wallace DP, Belibi FA, Cowley BD, Pelling JC, Grantham JJ. Cyclic AMP activates B-Raf and ERK in cyst epithelial cells from autosomal-dominant polycystic kidneys. Kidney Int 63: 1983–1994, 2003. doi: 10.1046/j.1523-1755.2003.00023.x. [DOI] [PubMed] [Google Scholar]
- 39. Nielsen S, Muller J, Knepper MA. Vasopressin- and cAMP-induced changes in ultrastructure of isolated perfused inner medullary collecting ducts. Am J Physiol Renal Physiol 265: F225–F238, 1993. doi: 10.1152/ajprenal.1993.265.2.F225. [DOI] [PubMed] [Google Scholar]
- 40. Ganote CE, Grantham JJ, Moses HL, Burg MB, Orloff J. Ultrastructural studies of vasopressin effect on isolated perfused renal collecting tubules of the rabbit. J Cell Biol 36: 355–367, 1968. doi: 10.1083/jcb.36.2.355. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 41. Kirk KL, DiBona DR, Schafer JA. Morphologic response of the rabbit cortical collecting tubule to peritubular hypotonicity: quantitative examination with differential interference contrast microscopy. J Membr Biol 79: 53–64, 1984. doi: 10.1007/BF01868526. [DOI] [PubMed] [Google Scholar]
- 42. Chou CL, Yu MJ, Kassai EM, Morris RG, Hoffert JD, Wall SM, Knepper MA. Roles of basolateral solute uptake via NKCC1 and of myosin II in vasopressin-induced cell swelling in inner medullary collecting duct. Am J Physiol Renal Physiol 295: F192–F201, 2008. doi: 10.1152/ajprenal.00011.2008. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 43. Fushimi K, Sasaki S, Marumo F. Phosphorylation of serine 256 is required for cAMP-dependent regulatory exocytosis of the aquaporin-2 water channel. J Biol Chem 272: 14800–14804, 1997. doi: 10.1074/jbc.272.23.14800. [DOI] [PubMed] [Google Scholar]
- 44. Katsura T, Gustafson CE, Ausiello DA, Brown D. Protein kinase A phosphorylation is involved in regulated exocytosis of aquaporin-2 in transfected LLC-PK1 cells. Am J Physiol Renal Physiol 272: F817–F822, 1997. [PubMed] [Google Scholar]
- 45. Christensen BM, Zelenina M, Aperia A, Nielsen S. Localization and regulation of PKA-phosphorylated AQP2 in response to V2-receptor agonist/antagonist treatment. Am J Physiol Renal Physiol 278: F29–F42, 2000. doi: 10.1152/ajprenal.2000.278.1.F29. [DOI] [PubMed] [Google Scholar]
- 46. van Balkom BW, Savelkoul PJ, Markovich D, Hofman E, Nielsen S, van der Sluijs P, Deen PM. The role of putative phosphorylation sites in the targeting and shuttling of the aquaporin-2 water channel. J Biol Chem 277: 41473–41479, 2002. doi: 10.1074/jbc.M207525200. [DOI] [PubMed] [Google Scholar]
- 47. Fenton RA, Moeller HB, Hoffert JD, Yu MJ, Nielsen S, Knepper MA. Acute regulation of aquaporin-2 phosphorylation at Ser-264 by vasopressin. Proc Natl Acad Sci USA 105: 3134–3139, 2008. doi: 10.1073/pnas.0712338105. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 48. Hoffert JD, Fenton RA, Moeller HB, Simons B, Tchapyjnikov D, McDill BW, Yu MJ, Pisitkun T, Chen F, Knepper MA. Vasopressin-stimulated increase in phosphorylation at Ser269 potentiates plasma membrane retention of aquaporin-2. J Biol Chem 283: 24617–24627, 2008. doi: 10.1074/jbc.M803074200. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 49. Lu HJ, Matsuzaki T, Bouley R, Hasler U, Qin QH, Brown D. The phosphorylation state of serine 256 is dominant over that of serine 261 in the regulation of AQP2 trafficking in renal epithelial cells. Am J Physiol Renal Physiol 295: F290–F294, 2008. doi: 10.1152/ajprenal.00072.2008. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 50. Moeller HB, Praetorius J, Rutzler MR, Fenton RA. Phosphorylation of aquaporin-2 regulates its endocytosis and protein-protein interactions. Proc Natl Acad Sci USA 107: 424–429, 2010. doi: 10.1073/pnas.0910683107. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 51. Tamma G, Robben JH, Trimpert C, Boone M, Deen PM. Regulation of AQP2 localization by S256 and S261 phosphorylation and ubiquitination. Am J Physiol Cell Physiol 300: C636–C646, 2011. doi: 10.1152/ajpcell.00433.2009. [DOI] [PubMed] [Google Scholar]
- 52. Rice WL, Zhang Y, Chen Y, Matsuzaki T, Brown D, Lu HA. Differential, phosphorylation dependent trafficking of AQP2 in LLC-PK1 cells. PLoS One 7: e32843, 2012. doi: 10.1371/journal.pone.0032843. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 53. Moeller HB, Aroankins TS, Slengerik-Hansen J, Pisitkun T, Fenton RA. Phosphorylation and ubiquitylation are opposing processes that regulate endocytosis of the water channel aquaporin-2. J Cell Sci 127: 3174–3183, 2014. doi: 10.1242/jcs.150680. [DOI] [PubMed] [Google Scholar]
- 54. Arthur J, Huang J, Nomura N, Jin WW, Li W, Cheng X, Brown D, Lu HJ. Characterization of the putative phosphorylation sites of the AQP2 C terminus and their role in AQP2 trafficking in LLC-PK1 cells. Am J Physiol Renal Physiol 309: F673–F679, 2015. doi: 10.1152/ajprenal.00152.2015. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 55. Wang PJ, Lin ST, Liu SH, Kuo KT, Hsu CH, Knepper MA, Yu MJ. Vasopressin-induced serine 269 phosphorylation reduces Sipa1l1 (signal-induced proliferation-associated 1 like 1)-mediated aquaporin-2 endocytosis. J Biol Chem 292: 7984–7993, 2017. doi: 10.1074/jbc.M117.779611. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 56. Nishimoto G, Zelenina M, Li D, Yasui M, Aperia A, Nielsen S, Nairn AC. Arginine vasopressin stimulates phosphorylation of aquaporin-2 in rat renal tissue. Am J Physiol Renal Physiol 276: F254–F259, 1999. doi: 10.1152/ajprenal.1999.276.2.F254. [DOI] [PubMed] [Google Scholar]
- 57. Hoffert JD, Pisitkun T, Wang G, Shen RF, Knepper MA. Quantitative phosphoproteomics of vasopressin-sensitive renal cells: regulation of aquaporin-2 phosphorylation at two sites. Proc Natl Acad Sci USA 103: 7159–7164, 2006. doi: 10.1073/pnas.0600895103. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 58. Moeller HB, Knepper MA, Fenton RA. Serine 269 phosphorylated aquaporin-2 is targeted to the apical membrane of collecting duct principal cells. Kidney Int 75: 295–303, 2009. doi: 10.1038/ki.2008.505. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 59. Xie L, Hoffert JD, Chou CL, Yu MJ, Pisitkun T, Knepper MA, Fenton RA. Quantitative analysis of aquaporin-2 phosphorylation. Am J Physiol Renal Physiol 298: F1018–F1023, 2010. doi: 10.1152/ajprenal.00580.2009. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 60. Hoffert JD, Nielsen J, Yu MJ, Pisitkun T, Schleicher SM, Nielsen S, Knepper MA. Dynamics of aquaporin-2 serine-261 phosphorylation in response to short-term vasopressin treatment in collecting duct. Am J Physiol Renal Physiol 292: F691–F700, 2007. doi: 10.1152/ajprenal.00284.2006. [DOI] [PubMed] [Google Scholar]
- 61. Deshpande V, Kao A, Raghuram V, Datta A, Chou CL, Knepper MA. Phosphoproteomic identification of vasopressin V2 receptor-dependent signaling in the renal collecting duct. Am J Physiol Renal Physiol 317: F789–F804, 2019. doi: 10.1152/ajprenal.00281.2019. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 62. Rinschen MM, Yu MJ, Wang G, Boja ES, Hoffert JD, Pisitkun T, Knepper MA. Quantitative phosphoproteomic analysis reveals vasopressin V2-receptor-dependent signaling pathways in renal collecting duct cells. Proc Natl Acad Sci USA 107: 3882–3887, 2010. doi: 10.1073/pnas.0910646107. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 63. Bansal AD, Hoffert JD, Pisitkun T, Hwang S, Chou CL, Boja ES, Wang G, Knepper MA. Phosphoproteomic profiling reveals vasopressin-regulated phosphorylation sites in collecting duct. J Am Soc Nephrol 21: 303–315, 2010. doi: 10.1681/ASN.2009070728. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 64. Manni S, Mauban JH, Ward CW, Bond M. Phosphorylation of the cAMP-dependent protein kinase (PKA) regulatory subunit modulates PKA-AKAP interaction, substrate phosphorylation, and calcium signaling in cardiac cells. J Biol Chem 283: 24145–24154, 2008. doi: 10.1074/jbc.M802278200. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 65. Meiri D, Greeve MA, Brunet A, Finan D, Wells CD, LaRose J, Rottapel R. Modulation of Rho guanine exchange factor Lfc activity by protein kinase A-mediated phosphorylation. Mol Cell Biol 29: 5963–5973, 2009. doi: 10.1128/MCB.01268-08. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 66. Meiri D, Marshall CB, Greeve MA, Kim B, Balan M, Suarez F, Bakal C, Wu C, Larose J, Fine N, Ikura M, Rottapel R. Mechanistic insight into the microtubule and actin cytoskeleton coupling through dynein-dependent RhoGEF inhibition. Mol Cell 45: 642–655, 2012. [Erratum in Mol Cell 45: 844, 2012]. doi: 10.1016/j.molcel.2012.01.027. [DOI] [PubMed] [Google Scholar]
- 67. Ren J, Guo W. ERK1/2 regulate exocytosis through direct phosphorylation of the exocyst component Exo70. Dev Cell 22: 967–978, 2012. doi: 10.1016/j.devcel.2012.03.005. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 68. McKinsey TA, Zhang CL, Lu J, Olson EN. Signal-dependent nuclear export of a histone deacetylase regulates muscle differentiation. Nature 408: 106–111, 2000. doi: 10.1038/35040593. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 69. Dequiedt F, Martin M, Von Blume J, Vertommen D, Lecomte E, Mari N, Heinen MF, Bachmann M, Twizere JC, Huang MC, Rider MH, Piwnica-Worms H, Seufferlein T, Kettmann R. New role for hPar-1 kinases EMK and C-TAK1 in regulating localization and activity of class IIa histone deacetylases. Mol Cell Biol 26: 7086–7102, 2006. doi: 10.1128/MCB.00231-06. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 70. Wagner LE II, Li WH, Yule DI. Phosphorylation of type-1 inositol 1,4,5-trisphosphate receptors by cyclic nucleotide-dependent protein kinases: a mutational analysis of the functionally important sites in the S2+ and S2− splice variants. J Biol Chem 278: 45811–45817, 2003. doi: 10.1074/jbc.M306270200. [DOI] [PubMed] [Google Scholar]
- 71. Wang Y, Li G, Goode J, Paz JC, Ouyang K, Screaton R, Fischer WH, Chen J, Tabas I, Montminy M. Inositol-1,4,5-trisphosphate receptor regulates hepatic gluconeogenesis in fasting and diabetes. Nature 485: 128–132, 2012. doi: 10.1038/nature10988. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 72. Wang KZQ, Steer E, Otero PA, Bateman NW, Cheng MH, Scott AL, Wu C, Bahar I, Shih YT, Hsueh YP, Chu CT. PINK1 interacts with VCP/p97 and activates PKA to promote NSFL1C/p47 phosphorylation and dendritic arborization in neurons. eNeuro 5: ENEURO.0466-18.2018, 2018. doi: 10.1523/ENEURO.0466-18.2018. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 73. Obara Y, Labudda K, Dillon TJ, Stork PJ. PKA phosphorylation of Src mediates Rap1 activation in NGF and cAMP signaling in PC12 cells. J Cell Sci 117: 6085–6094, 2004. doi: 10.1242/jcs.01527. [DOI] [PubMed] [Google Scholar]
- 74. Pozo-Guisado E, Campbell DG, Deak M, Alvarez-Barrientos A, Morrice NA, Alvarez IS, Alessi DR, Martin-Romero FJ. Phosphorylation of STIM1 at ERK1/2 target sites modulates store-operated calcium entry. J Cell Sci 123: 3084–3093, 2010. doi: 10.1242/jcs.067215. [DOI] [PubMed] [Google Scholar]
- 75. Pozo-Guisado E, Casas-Rua V, Tomas-Martin P, Lopez-Guerrero AM, Alvarez-Barrientos A, Martin-Romero FJ. Phosphorylation of STIM1 at ERK1/2 target sites regulates interaction with the microtubule plus-end binding protein EB1. J Cell Sci 126: 3170–3180, 2013. doi: 10.1242/jcs.125054. [DOI] [PubMed] [Google Scholar]
- 76. Maitra S, Chou CF, Luber CA, Lee KY, Mann M, Chen CY. The AU-rich element mRNA decay-promoting activity of BRF1 is regulated by mitogen-activated protein kinase-activated protein kinase 2. RNA 14: 950–959, 2008. doi: 10.1261/rna.983708. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 77. Herranz N, Gallage S, Mellone M, Wuestefeld T, Klotz S, Hanley CJ, Raguz S, Acosta JC, Innes AJ, Banito A, Georgilis A, Montoya A, Wolter K, Dharmalingam G, Faull P, Carroll T, Martinez-Barbera JP, Cutillas P, Reisinger F, Heikenwalder M, Miller RA, Withers D, Zender L, Thomas GJ, Gil J. mTOR regulates MAPKAPK2 translation to control the senescence-associated secretory phenotype. Nat Cell Biol 17: 1205–1217, 2015. [Erratum in Nat Cell Biol 17: 1370, 2015]. doi: 10.1038/ncb3225. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 78. Rataj F, Planel S, Desroches-Castan A, Le Douce J, Lamribet K, Denis J, Feige JJ, Cherradi N. The cAMP pathway regulates mRNA decay through phosphorylation of the RNA-binding protein TIS11b/BRF1. Mol Biol Cell 27: 3841–3854, 2016. doi: 10.1091/mbc.E16-06-0379. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 79. Klussmann E, Tamma G, Lorenz D, Wiesner B, Maric K, Hofmann F, Aktories K, Valenti G, Rosenthal W. An inhibitory role of Rho in the vasopressin-mediated translocation of aquaporin-2 into cell membranes of renal principal cells. J Biol Chem 276: 20451–20457, 2001. doi: 10.1074/jbc.M010270200. [DOI] [PubMed] [Google Scholar]
- 80. Tamma G, Klussmann E, Procino G, Svelto M, Rosenthal W, Valenti G. cAMP-induced AQP2 translocation is associated with RhoA inhibition through RhoA phosphorylation and interaction with RhoGDI. J Cell Sci 116: 1519–1525, 2003. doi: 10.1242/jcs.00355. [DOI] [PubMed] [Google Scholar]
- 81. Mostov KE, Verges M, Altschuler Y. Membrane traffic in polarized epithelial cells. Curr Opin Cell Biol 12: 483–490, 2000. doi: 10.1016/s0955-0674(00)00120-4. [DOI] [PubMed] [Google Scholar]
- 82. Folsch H. The building blocks for basolateral vesicles in polarized epithelial cells. Trends Cell Biol 15: 222–228, 2005. doi: 10.1016/j.tcb.2005.02.006. [DOI] [PubMed] [Google Scholar]
- 83. Grozinger CM, Schreiber SL. Regulation of histone deacetylase 4 and 5 and transcriptional activity by 14-3-3-dependent cellular localization. Proc Natl Acad Sci USA 97: 7835–7840, 2000. doi: 10.1073/pnas.140199597. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 84. Di Giorgio E, Dalla E, Franforte E, Paluvai H, Minisini M, Trevisanut M, Picco R, Brancolini C. Different class IIa HDACs repressive complexes regulate specific epigenetic responses related to cell survival in leiomyosarcoma cells. Nucleic Acids Res 48: 646–664, 2020. doi: 10.1093/nar/gkz1120. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 85. Jung HJ, Raghuram V, Lee JW, Knepper MA. Genome-wide mapping of DNA accessibility and binding sites for CREB and C/EBPβ in vasopressin-sensitive collecting duct cells. J Am Soc Nephrol 29: 1490–1500, 2018. doi: 10.1681/ASN.2017050545. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 86. Khositseth S, Pisitkun T, Slentz DH, Wang G, Hoffert JD, Knepper MA, Yu MJ. Quantitative protein and mRNA profiling shows selective post-transcriptional control of protein expression by vasopressin in kidney cells. Mol Cell Proteomics 10: M110.004036, 2011. doi: 10.1074/mcp.M110.004036. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 87. McKinsey TA, Zhang CL, Olson EN. Identification of a signal-responsive nuclear export sequence in class II histone deacetylases. Mol Cell Biol 21: 6312–6321, 2001. doi: 10.1128/MCB.21.18.6312-6321.2001. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 88. Hardie DG, Schaffer BE, Brunet A. AMPK: an energy-sensing pathway with multiple inputs and outputs. Trends Cell Biol 26: 190–201, 2016. doi: 10.1016/j.tcb.2015.10.013. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 89. Kondo H, Rabouille C, Newman R, Levine TP, Pappin D, Freemont P, Warren G. p47 is a cofactor for p97-mediated membrane fusion. Nature 388: 75–78, 1997. doi: 10.1038/40411. [DOI] [PubMed] [Google Scholar]
- 90. Schmitt JM, Stork PJ. Galpha and Gbeta gamma require distinct Src-dependent pathways to activate Rap1 and Ras. J Biol Chem 277: 43024–43032, 2002. doi: 10.1074/jbc.M204006200. [DOI] [PubMed] [Google Scholar]
- 91. Cheung PW, Terlouw A, Janssen SA, Brown D, Bouley R. Inhibition of non-receptor tyrosine kinase Src induces phosphoserine 256-independent aquaporin-2 membrane accumulation. J Physiol 597: 1627–1642, 2019. doi: 10.1113/JP277024. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 92. Cook SJ, Rubinfeld B, Albert I, McCormick F. RapV12 antagonizes Ras-dependent activation of ERK1 and ERK2 by LPA and EGF in Rat-1 fibroblasts. EMBO J 12: 3475–3485, 1993. doi: 10.1002/j.1460-2075.1993.tb06022.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 93. Roos J, DiGregorio PJ, Yeromin AV, Ohlsen K, Lioudyno M, Zhang S, Safrina O, Kozak JA, Wagner SL, Cahalan MD, Velicelebi G, Stauderman KA. STIM1, an essential and conserved component of store-operated Ca2+ channel function. J Cell Biol 169: 435–445, 2005. doi: 10.1083/jcb.200502019. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 94. Cahalan MD. STIMulating store-operated Ca2+ entry. Nat Cell Biol 11: 669–677, 2009. doi: 10.1038/ncb0609-669. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 95. Lykke-Andersen J, Wagner E. Recruitment and activation of mRNA decay enzymes by two ARE-mediated decay activation domains in the proteins TTP and BRF-1. Genes Dev 19: 351–361, 2005. doi: 10.1101/gad.1282305. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 96. Cai Q, McReynolds MR, Keck M, Greer KA, Hoying JB, Brooks HL. Vasopressin receptor subtype 2 activation increases cell proliferation in the renal medulla of AQP1 null mice. Am J Physiol Renal Physiol 293: F1858–F1864, 2007. doi: 10.1152/ajprenal.00068.2007. [DOI] [PubMed] [Google Scholar]
- 97. Leo KT, Chou CL, Yang CR, Park E, Raghuram V, Knepper MA. Bayesian analysis of dynamic phosphoproteomic data identifies protein kinases mediating GPCR responses. Cell Commun Signal 20: 80, 2022. doi: 10.1186/s12964-022-00892-6. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 98. Wall SM, Han JS, Chou CL, Knepper MA. Kinetics of urea and water permeability activation by vasopressin in rat terminal IMCD. Am J Physiol Renal Physiol 262: F989–F998, 1992. doi: 10.1152/ajprenal.1992.262.6.F989. [DOI] [PubMed] [Google Scholar]
- 99. Matsuda S, Kominato K, Koide-Yoshida S, Miyamoto K, Isshiki K, Tsuji A, Yuasa K. PCTAIRE kinase 3/cyclin-dependent kinase 18 is activated through association with cyclin A and/or phosphorylation by protein kinase A. J Biol Chem 289: 18387–18400, 2014. doi: 10.1074/jbc.M113.542936. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 100. Dema A, Faust D, Lazarow K, Wippich M, Neuenschwander M, Zühlke K, Geelhaar A, Pallien T, Hallscheidt E, Eichhorst J, Wiesner B, Černecká H, Popp O, Mertins P, Dittmar G, von Kries JP, Klussmann E. Cyclin-dependent kinase 18 controls trafficking of aquaporin-2 and its abundance through ubiquitin ligase STUB1, which functions as an AKAP. Cells 9: 673, 2020. doi: 10.3390/cells9030673. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 101. Tapon N, Hall A. Rho, Rac and Cdc42 GTPases regulate the organization of the actin cytoskeleton. Curr Opin Cell Biol 9: 86–92, 1997. doi: 10.1016/s0955-0674(97)80156-1. [DOI] [PubMed] [Google Scholar]
- 102. Marples D, Schroer TA, Ahrens N, Taylor A, Knepper MA, Nielsen S. Dynein and dynactin colocalize with AQP2 water channels in intracellular vesicles from kidney collecting duct. Am J Physiol Renal Physiol 274: F384–F394, 1998. doi: 10.1152/ajprenal.1998.274.2.F384. [DOI] [PubMed] [Google Scholar]
- 103. Tajika Y, Matsuzaki T, Suzuki T, Ablimit A, Aoki T, Hagiwara H, Kuwahara M, Sasaki S, Takata K. Differential regulation of AQP2 trafficking in endosomes by microtubules and actin filaments. Histochem Cell Biol 124: 1–12, 2005. doi: 10.1007/s00418-005-0010-3. [DOI] [PubMed] [Google Scholar]
- 104. Vossenkamper A, Nedvetsky PI, Wiesner B, Furkert J, Rosenthal W, Klussmann E. Microtubules are needed for the perinuclear positioning of aquaporin-2 after its endocytic retrieval in renal principal cells. Am J Physiol Cell Physiol 293: C1129–C1138, 2007. doi: 10.1152/ajpcell.00628.2006. [DOI] [PubMed] [Google Scholar]
- 105. Zhao H, Yao X, Wang TX, Jin WM, Ji QQ, Yang X, Duan QH, Yao LJ. PKCalpha regulates vasopressin-induced aquaporin-2 trafficking in mouse kidney collecting duct cells in vitro via altering microtubule assembly. Acta Pharmacol Sin 33: 230–236, 2012. doi: 10.1038/aps.2011.160. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 106. Yui N, Lu HA, Chen Y, Nomura N, Bouley R, Brown D. Basolateral targeting and microtubule-dependent transcytosis of the aquaporin-2 water channel. Am J Physiol Cell Physiol 304: C38–C48, 2013. doi: 10.1152/ajpcell.00109.2012. [DOI] [PMC free article] [PubMed] [Google Scholar]
Associated Data
This section collects any data citations, data availability statements, or supplementary materials included in this article.
Supplementary Materials
Supplemental Spreadsheet S1: https://esbl.nhlbi.nih.gov/Databases/Supplemental-Phosphorylation/.
Data Availability Statement
Data access: https://esbl.nhlbi.nih.gov/Databases/AVP-Phos/ and https://esbl.nhlbi.nih.gov/AVP-Network/.