首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 31 毫秒
1.
Ectopic ureters are rare congenital malformations of the renal system that most commonly present in females. It is extremely rare to encounter an ectopic ureter in an older man undergoing radical prostatectomy. We report herein a case of a 66-year-old man with prostate cancer and a complete duplication of the left renal collecting system, with an upper pole ectopic ureter and associated normal functioning renal parenchyma entering into the prostatic urethra. This anomaly was incidentally discovered on preoperative magnetic resonance imaging of the prostate. Open radical retropubic prostatectomy and a left ureteroureterostomy were performed.Key words: Prostate cancer, Ectopic ureter, Prostatic urethra, Prostatectomy, UreteroureterostomyEctopic ureters are rare and occur in about 1 of 1900 live births.1 Over 85% of ectopic ureters are associated with duplicated systems and most commonly present in females. Ectopic ureters present 2 to 12 times less frequently in males than in females, and in males are most commonly associated with a single collecting system.2 Ectopic ureters inserting into the prostatic urethra often present with obstruction and/or urinary tract infections. Few cases of ectopic ureters entering into the prostatic urethra as an incidental finding in men undergoing radical prostatectomy for prostate cancer have been reported in the literature.3,4 This case report describes a patient with prostate cancer and an asymptomatic upper pole ectopic ureter inserting into the prostatic urethra associated with normal functioning renal parenchyma demonstrated on preoperative mercaptoacetyltriglycine (MAG-3) renogram, magnetic resonance imaging (MRI), and computed tomography (CT). We discuss the treatment plan for this patient and give an overview of ectopic ureters.  相似文献   

2.
We recently cared for a patient with adenocarcinoma of the pancreas who presented with ureteral metastasis followed by hydroureteronephrosis long before the appearance of any symptoms related to the primary lesion. The entity is extremely rare; only seven similar cases are on record in the scientific literature. No recent review exists on this topic. This encouraged us to present our case along with the previous cases of adenocarcinoma of the pancreas with ureteral metastasis that have been reported.Key words: Ureter, Metastasis, Ureteral metastasis, Pancreatic adenocarcinoma, HydronephrosisAlthough, ureteral obstruction is relatively common in patients with locally invasive pelvic malignancies, it is extremely rare to find ureteral obstruction due to metastases to the ureter from distant primary tumors.14 Only approximately 400 such cases have been reported.1,2 Most reported cases were diagnosed postmortem. Ureteral metastases from pancreatic adenocarcinoma have been reported in only seven published cases until now. In the case of the tumors in sites other than the pancreas with ureteral metastases, the primary lesions usually become evident before the ureteral lesions are found.5 Three of the seven (43%) reported cases of pancreatic adenocarcinoma presented with initial urologic complaints, and not with complaints related to the primary lesion. We report on a case of pancreatic adenocarcinoma that presented with ureteral metastasis and hydronephrosis. This case also had metastatic invasion of all layers of the ureter which is another infrequent and unique finding.3 We also review the seven previously reported cases here, along with the relevant literature.  相似文献   

3.
Collapsin response mediator protein 2 (CRMP2) is an intracellular protein that mediates signaling of Semaphorin3A (Sema3A), a repulsive axon guidance molecule. Fyn, a Src-type tyrosine kinase, is involved in the Sema3A signaling. However, the relationship between CRMP2 and Fyn in this signaling pathway is still unknown. In our research, we demonstrated that Fyn phosphorylated CRMP2 at Tyr32 residues in HEK293T cells. Immunohistochemical analysis using a phospho-specific antibody at Tyr32 of CRMP showed that Tyr32-phosphorylated CRMP was abundant in the nervous system, including dorsal root ganglion neurons, the molecular and Purkinje cell layer of adult cerebellum, and hippocampal fimbria. Overexpression of a nonphosphorylated mutant (Tyr32 to Phe32) of CRMP2 in dorsal root ganglion neurons interfered with Sema3A-induced growth cone collapse response. These results suggest that Fyn-dependent phosphorylation of CRMP2 at Tyr32 is involved in Sema3A signaling.Collapsin response mediator proteins (CRMPs)4 have been identified as intracellular proteins that mediate Semaphorin3A (Sema3A) signaling in the nervous system (1). CRMP2 is one of the five members of the CRMP family. CRMPs also mediate signal transduction of NT3, Ephrin, and Reelin (24). CRMPs interact with several intracellular molecules, including tubulin, Numb, kinesin1, and Sra1 (58). CRMPs are involved in axon guidance, axonal elongation, cell migration, synapse maturation, and the generation of neuronal polarity (1, 2, 4, 5).CRMP family proteins are known to be the major phosphoproteins in the developing brain (1, 9). CRMP2 is phosphorylated by several Ser/Thr kinases, such as Rho kinase, cyclin-dependent kinase 5 (Cdk5), and glycogen synthase kinase 3β (GSK3β) (2, 1013). The phosphorylation sites of CRMP2 by these kinases are clustered in the C terminus and have already been identified. Rho kinase phosphorylates CRMP2 at Thr555 (10). Cdk5 phosphorylates CRMP2 at Ser522, and this phosphorylation is essential for sequential phosphorylations by GSK3β at Ser518, Thr514, and Thr509 (2, 1113). These phosphorylations disrupt the interaction of CRMP2 with tubulin or Numb (2, 3, 13). The sequential phosphorylation of CRMP2 by Cdk5 and GSK3β is an essential step in Sema3A signaling (11, 13). Furthermore, the neurofibrillary tangles in the brains of people with Alzheimer disease contain hyperphosphorylated CRMP2 at Thr509, Ser518, and Ser522 (14, 15).CRMPs are also substrates of several tyrosine kinases. The phosphorylation of CRMP2 by Fes/Fps and Fer has been shown to be involved in Sema3A signaling (16, 17). Phosphorylation of CRMP2 at Tyr479 by a Src family tyrosine kinase Yes regulates CXCL12-induced T lymphocyte migration (18). We reported previously that Fyn is involved in Sema3A signaling (19). Fyn associates with PlexinA2, one of the components of the Sema3A receptor complex. Fyn also activates Cdk5 through the phosphorylation at Tyr15 of Cdk5 (19). In dorsal root ganglion (DRG) neurons from fyn-deficient mice, Sema3A-induced growth cone collapse response is attenuated compared with control mice (19). Furthermore, we recently found that Fyn phosphorylates CRMP1 and that this phosphorylation is involved in Reelin signaling (4). Although it has been shown that CRMP2 is involved in Sema3A signaling (1, 11, 13), the relationship between Fyn and CRMP2 in Sema3A signaling and the tyrosine phosphorylation site(s) of CRMPs remain unknown.Here, we show that Fyn phosphorylates CRMP2 at Tyr32. Using a phospho-specific antibody against Tyr32, we determined that the residue is phosphorylated in vivo. A nonphosphorylated mutant CRMP2Y32F inhibits Sema3A-induced growth cone collapse. These results indicate that tyrosine phosphorylation by Fyn at Tyr32 is involved in Sema3A signaling.  相似文献   

4.
Human concentrative nucleoside transporter 3 (hCNT3) utilizes electrochemical gradients of both Na+ and H+ to accumulate pyrimidine and purine nucleosides within cells. We have employed radioisotope flux and electrophysiological techniques in combination with site-directed mutagenesis and heterologous expression in Xenopus oocytes to identify two conserved pore-lining glutamate residues (Glu-343 and Glu-519) with essential roles in hCNT3 Na+/nucleoside and H+/nucleoside cotransport. Mutation of Glu-343 and Glu-519 to aspartate, glutamine, and cysteine severely compromised hCNT3 transport function, and changes included altered nucleoside and cation activation kinetics (all mutants), loss or impairment of H+ dependence (all mutants), shift in Na+:nucleoside stoichiometry from 2:1 to 1:1 (E519C), complete loss of catalytic activity (E519Q) and, similar to the corresponding mutant in Na+-specific hCNT1, uncoupled Na+ currents (E343Q). Consistent with close-proximity integration of cation/solute-binding sites within a common cation/permeant translocation pore, mutation of Glu-343 and Glu-519 also altered hCNT3 nucleoside transport selectivity. Both residues were accessible to the external medium and inhibited by p-chloromercuribenzene sulfonate when converted to cysteine.Physiologic nucleosides and the majority of synthetic nucleoside analogs with antineoplastic and/or antiviral activity are hydrophilic molecules that require specialized plasma membrane nucleoside transporter (NT)3 proteins for transport into or out of cells (14). NT-mediated transport is required for nucleoside metabolism by salvage pathways and is a critical determinant of the pharmacologic actions of nucleoside drugs (36). By regulating adenosine availability to purinoreceptors, NTs also modulate a diverse array of physiological processes, including neurotransmission, immune responses, platelet aggregation, renal function, and coronary vasodilation (4, 6, 7). Two structurally unrelated NT families of integral membrane proteins exist in human and other mammalian cells and tissues as follows: the SLC28 concentrative nucleoside transporter (CNT) family and the SLC29 equilibrative nucleoside transporter (ENT) family (3, 4, 6, 8, 9). ENTs are normally present in most, possibly all, cell types (4, 6, 8). CNTs, in contrast, are found predominantly in intestinal and renal epithelia and other specialized cell types, where they have important roles in absorption, secretion, distribution, and elimination of nucleosides and nucleoside drugs (13, 5, 6, 9).The CNT protein family in humans is represented by three members, hCNT1, hCNT2, and hCNT3. Belonging to a CNT subfamily phylogenetically distinct from hCNT1/2, hCNT3 utilizes electrochemical gradients of both Na+ and H+ to accumulate a broad range of pyrimidine and purine nucleosides and nucleoside drugs within cells (10, 11). hCNT1 and hCNT2, in contrast, are Na+-specific and transport pyrimidine and purine nucleosides, respectively (1113). Together, hCNT1–3 account for the three major concentrative nucleoside transport processes of human and other mammalian cells. Nonmammalian members of the CNT protein family that have been characterized functionally include hfCNT, a second member of the CNT3 subfamily from the ancient marine prevertebrate the Pacific hagfish Eptatretus stouti (14), CeCNT3 from Caenorhabditis elegans (15), CaCNT from Candida albicans (16), and the bacterial nucleoside transporter NupC from Escherichia coli (17). hfCNT is Na+- but not H+-coupled, whereas CeCNT3, CaCNT, and NupC are exclusively H+-coupled. Na+:nucleoside coupling stoichiometries are 1:1 for hCNT1 and hCNT2 and 2:1 for hCNT3 and hfCNT3 (11, 14). H+:nucleoside coupling ratios for hCNT3 and CaCNT are 1:1 (11, 16).Although much progress has been made in molecular studies of ENT proteins (4, 6, 8), studies of structurally and functionally important regions and residues within the CNT protein family are still at an early stage. Topological investigations suggest that hCNT1–3 and other eukaryote CNT family members have a 13 (or possibly 15)-transmembrane helix (TM) architecture, and multiple alignments reveal strong sequence similarities within the C-terminal half of the proteins (18). Prokaryotic CNTs lack the first three TMs of their eukaryotic counterparts, and functional expression of N-terminally truncated human and rat CNT1 in Xenopus oocytes has established that these three TMs are not required for Na+-dependent uridine transport activity (18). Consistent with this finding, chimeric studies involving hCNT1 and hfCNT (14) and hCNT1 and hCNT3 (19) have demonstrated that residues involved in Na+- and H+-coupling reside in the C-terminal half of the protein. Present in this region of the transporter, but of unknown function, is a highly conserved (G/A)XKX3NEFVA(Y/M/F) motif common to all eukaryote and prokaryote CNTs.By virtue of their negative charge and consequent ability to interact directly with coupling cations and/or participate in cation-induced and other protein conformational transitions, glutamate and aspartate residues play key functional and structural roles in a broad spectrum of mammalian and bacterial cation-coupled transporters (2030). Little, however, is known about their role in CNTs. This study builds upon a recent mutagenesis study of conserved glutamate and aspartate residues in hCNT1 (31) to undertake a parallel in depth investigation of corresponding residues in hCNT3. By employing the multifunctional capability of hCNT3 as a template for these studies, this study provides novel mechanistic insights into the molecular mechanism(s) of CNT-mediated cation/nucleoside cotransport, including the role of the (G/A)XKX3NEFVA(Y/M/F) motif.  相似文献   

5.
6.
An improved algorithm for the piecewise-smooth Mumford and Shah functional is presented. Compared to the previous work of Chan and Vese, and Choi et al., extensions of the key functions are replaced by updating the level set function based on an artificial image that is composed of the diffused image and the original image. The low convergence problem of the classical algorithm is efficiently solved in the proposed approach. The resulting algorithm has also been demonstrated by several cases.[1,2,3,4,5,6,7,8,9,10,11]  相似文献   

7.
Men with 47,XYY syndrome present with varying physical attributes and degrees of infertility. A retrospective chart review was performed on a male infertility and genetic anomaly database. Three patients with 47,XYY were found. Each presented with > 2 years of infertility. All were tall with elevated body mass indices. Scrotal findings ranged from normal to atrophic testicles. Semen analyses demonstrated oligospermia and varying endocrine profiles. Because of the diverse phenotype and potential lack of symptoms, identification and diagnosis of men with 47,XYY syndrome may be difficult. We recommend careful screening of 47,XYY patients and referral to primary physicians for long-term follow-up for increased incidence of health-related comorbidities.Key words: Infertility syndromesThe 47,XYY sex chromosome variation is the most common sex chromosome anomaly after Klinefelter syndrome (47,XXY),13 occurring in approximately 1 out of 1000 live male births.4,5 Parental nondisjunction at meiosis II resulting in an extra Y chromosome produces a 47,XYY karyotype in the affected offspring.68 46,XY/47,XYY mosaics from parental nondisjunction during cell division after postzygotic mitosis can result in addition of the extra Y chromosome in early embryonic development.6,8Most patients with 47,XYY have a delayed diagnosis, with a median age of 17.1 years at diagnosis, as was shown in a Danish cohort study.9 Although most have no phenotypic abnormalities, XYY boys are at greater risk for behavioral problems, mild learning disability, delayed speech and language development, and tall stature.10 Studies have increasingly reported an association between 47,XYY and fertility problems, noting an increased incidence of chromosomally abnormal spermatozoa in the semen of men with 47,XXY syndrome.7,1115 This greater prevalence of hyperhaploid sperm results in an increased risk of passing the extra Y chromosome to offspring.14 Men with 47,XXY syndrome can have variable sperm counts, ranging from normal to azoospermia.3,8,14,1618Here we review pertinent findings on physical examination and laboratory evaluation in three men with 47,XXY syndrome diagnosed during infertility evaluation as well as review the available literature on the subject, with special emphasis on male fertility effects.  相似文献   

8.
Two methods were used to compare the biodegradation of six polychlorinated biphenyl (PCB) congeners by 12 white rot fungi. Four fungi were found to be more active than Phanerochaete chrysosporium ATCC 24725. Biodegradation of the following congeners was monitored by gas chromatography: 2,3-dichlorobiphenyl, 4,4′-dichlorobiphenyl, 2,4′,5-trichlorobiphenyl (2,4′,5-TCB), 2,2′,4,4′-tetrachlorobiphenyl, 2,2′,5,5′-tetrachlorobiphenyl, and 2,2′,4,4′,5,5′-hexachlorobiphenyl. The congener tested for mineralization was 2,4′,5-[U-14C]TCB. Culture supernatants were also assayed for lignin peroxidase and manganese peroxidase activities. Of the fungi tested, two strains of Bjerkandera adusta (UAMH 8258 and UAMH 7308), one strain of Pleurotus ostreatus (UAMH 7964), and Trametes versicolor UAMH 8272 gave the highest biodegradation and mineralization. P. chrysosporium ATCC 24725, a strain frequently used in studies of PCB degradation, gave the lowest mineralization and biodegradation activities of the 12 fungi reported here. Low but detectable levels of lignin peroxidase and manganese peroxidase activity were present in culture supernatants, but no correlation was observed among any combination of PCB congener biodegradation, mineralization, and lignin peroxidase or manganese peroxidase activity. With the exception of P. chrysosporium, congener loss ranged from 40 to 96%; however, these values varied due to nonspecific congener binding to fungal biomass and glassware. Mineralization was much lower, ≤11%, because it measures a complete oxidation of at least part of the congener molecule but the results were more consistent and therefore more reliable in assessment of PCB biodegradation.

Polychlorinated biphenyls (PCBs) are produced by chlorination of biphenyl, resulting in up to 209 different congeners. Commercial mixtures range from light oily fluids to waxes, and their physical properties make them useful as heat transfer fluids, hydraulic fluids, solvent extenders, plasticizers, flame retardants, organic diluents, and dielectric fluids (1, 21). Approximately 24 million lb are in the North American environment (19). The stability and hydrophobic nature of these compounds make them a persistent environmental hazard.To date, bacterial transformations have been the main focus of PCB degradation research. Aerobic bacteria use a biphenyl-induced dioxygenase enzyme system to attack less-chlorinated congeners (mono- to hexachlorobiphenyls) (1, 5, 7, 8, 22). Although more-chlorinated congeners are recalcitrant to aerobic bacterial degradation, microorganisms in anaerobic river sediments reductively dechlorinate these compounds, mainly removing the meta and para chlorines (1, 6, 10, 33, 34).The degradation of PCBs by white rot fungi has been known since 1985 (11, 18). Many fungi have been tested for their ability to degrade PCBs, including the white rot fungi Coriolus versicolor (18), Coriolopsis polysona (41), Funalia gallica (18), Hirneola nigricans (35), Lentinus edodes (35), Phanerochaete chrysosporium (3, 11, 14, 17, 18, 35, 39, 4143), Phlebia brevispora (18), Pleurotus ostreatus (35, 43), Poria cinerescens (18), Px strain (possibly Lentinus tigrinus) (35), and Trametes versicolor (41, 43). There have also been studies of PCB metabolism by ectomycorrhizal fungi (17) and other fungi such as Aspergillus flavus (32), Aspergillus niger (15), Aureobasidium pullulans (18), Candida boidinii (35), Candida lipolytica (35), Cunninghamella elegans (16), and Saccharomyces cerevisiae (18, 38). The mechanism of PCB biodegradation has not been definitively determined for any fungi. White rot fungi produce several nonspecific extracellular enzymes which have been the subject of extensive research. These nonspecific peroxidases are normally involved in lignin degradation but can oxidize a wide range of aromatic compounds including polycyclic aromatic hydrocarbons (37). Two peroxidases, lignin peroxidase (LiP) and Mn peroxidase (MnP), are secreted into the environment of the fungus under conditions of nitrogen limitation in P. chrysosporium (23, 25, 27, 29) but are not stress related in fungi such as Bjerkandera adusta or T. versicolor (12, 30).Two approaches have been used to determine the biodegradability of PCBs by fungi: (i) loss of the parent congener analyzed by gas chromatography (GC) (17, 32, 35, 42, 43) and (ii) mineralization experiments in which the 14C of the universally labeled 14C parent congener is recovered as 14CO2 (11, 14, 18, 39, 41). In the first method, the loss of a peak on a chromatogram makes it difficult to decide whether the PCB is being partly degraded, mineralized, adsorbed to the fungal biomass, or bound to glassware, soil particles, or wood chips. Even when experiments with killed-cell and abiotic controls are performed, the extraction efficiency and standard error can make data difficult to interpret. For example, recoveries can range anywhere from 40 to 100% depending on the congener used and the fungus being investigated (17). On the other hand, recovery of significant amounts of 14CO2 from the cultures incubated with a 14C substrate provides definitive proof of fungal metabolism. There appears to be only one report relating data from these two techniques (18), and in that study, [U-14C]Aroclor 1254, rather than an individual congener, was used.In this study, we examined the ability of 12 white rot fungal strains to metabolize selected PCB congeners to determine which strains were the most active degraders. Included in this group was P. chrysosporium ATCC 24725, a strain used extensively in PCB studies (3, 14, 18, 35, 39, 4143). Six PCB congeners were selected to give a range of chlorine substitutions and therefore a range of potential biodegradability which was monitored by GC. One of the chosen congeners was 14C labeled and used in studies to compare the results from a mineralization method with those from the GC method.  相似文献   

9.
Calyceal diverticula are rare outpouchings of the upper collecting system that likely have a congenital origin. Stones can be found in up to 50% of calyceal diverticula, although, over the combined reported series, 96% of patients presented with stones. Diagnosis is best made by intravenous urography or computed tomography urogram. Shock wave lithotripsy (SWL) is an option for first-line therapy in patients with stone-bearing diverticula that have radiologically patent necks in mid- to upper-pole diverticula and small stone burdens. Stone-free rates are the lowest with SWL, although patients report being asymptomatic following therapy in up to 75% of cases with extended follow-up. Ureteroscopy (URS) is best suited for management of anteriorly located mid- to upperpole diverticular stones. Drawbacks to URS include difficulty in identifying the ostium and low rate of obliteration. Percutaneous management is best used in posteriorly located mid- to lower-pole stones, and offers the ability to directly ablate the diverticulum. Percutaneous nephrolithotomy remains effective in the management of upperpole diverticula, but carries the risk of pulmonary complications unless subcostal access strategies such as triangulation or renal displacement are used. Laparoscopic surgery provides definitive management, but should be reserved for cases with large stones in anteriorly located diverticula with thin overlying parenchyma, and cases that are refractory to other treatment. This article reviews the current theories on the pathogenesis of calyceal diverticula. The current classification is examined in addition to the current diagnostic methods. Here we summarize an extensive review of the literature on the outcomes of the different treatment approaches.Key words: Calyceal diverticula, Percutaneous nephrostolithotomy, Laparoscopic surgery, Shock wave lithotripsy, UreterorenoscopyCalyceal diverticula are eventrations of the upper collecting system lying within the renal parenchyma.1 These nonsecretory outpouchings are lined by transitional cell epithelium and communicate with the main collecting system via a narrow channel, allowing for passive filling with urine. They were first described in 1841 by Rayer in “Traitements des maladies des reins.”2 Thought to be either cysts or localized hydronephrosis, he used the term kyste urinaire to describe his finding of intrarenal urine-containing cavities that communicate with calyces. Other investigators reported similar findings and—depending on location and postulated etiology—described them as pelvic cysts,3 peripelvic cysts,4 pyelorenal cysts,5 pyelosynaptic cysts,6 pyelogenous cysts,7 hydrocalicosis,8 cystic dilatations of the calyx,9 congenital cortical cysts,10 congenital cystic dysplasia,4,11 calyceal pseudocysts,12 juxta-calyceal cysts,13 pelvic diverticula, 14 congenital diverticula of the calyx,15 and finally, calyceal diverticula.1618 Prather is credited with coining the term and the definition of calyceal diverticulum that we use today.  相似文献   

10.
SLC26A7 (human)/Slc26a7 (mouse) is a recently identified chloride-base exchanger and/or chloride transporter that is expressed on the basolateral membrane of acid-secreting cells in the renal outer medullary collecting duct (OMCD) and in gastric parietal cells. Here, we show that mice with genetic deletion of Slc26a7 expression develop distal renal tubular acidosis, as manifested by metabolic acidosis and alkaline urine pH. In the kidney, basolateral Cl/HCO3 exchange activity in acid-secreting intercalated cells in the OMCD was significantly decreased in hypertonic medium (a normal milieu for the medulla) but was reduced only mildly in isotonic medium. Changing from a hypertonic to isotonic medium (relative hypotonicity) decreased the membrane abundance of Slc26a7 in kidney cells in vivo and in vitro. In the stomach, stimulated acid secretion was significantly impaired in isolated gastric mucosa and in the intact organ. We propose that SLC26A7 dysfunction should be investigated as a potential cause of unexplained distal renal tubular acidosis or decreased gastric acid secretion in humans.The collecting duct segment of the distal kidney nephron plays a major role in systemic acid base homeostasis by acid secretion and bicarbonate absorption. The acid secretion occurs via H+-ATPase and H-K-ATPase into the lumen and bicarbonate is absorbed via basolateral Cl/HCO3 exchangers (14). The tubules, which are located within the outer medullary region of the kidney collecting duct (OMCD),2 have the highest rate of acid secretion among the distal tubule segments and are therefore essential to the maintenance of acid base balance (2).The gastric parietal cell is the site of generation of acid and bicarbonate through the action of cytosolic carbonic anhydrase II (5, 6). The intracellular acid is secreted into the lumen via gastric H-K-ATPase, which works in conjunction with a chloride channel and a K+ recycling pathway (710). The intracellular bicarbonate is transported to the blood via basolateral Cl/HCO3 exchangers (1114).SLC26 (human)/Slc26 (mouse) isoforms are members of a conserved family of anion transporters that display tissue-specific patterns of expression in epithelial cells (1524). Several SLC26 members can function as chloride/bicarbonate exchangers. These include SLC26A3 (DRA), SLC26A4 (pendrin), SLC26A6 (PAT1 or CFEX), SLC26A7, and SLC26A9 (2531). SLC26A7 and SLC26A9 can also function as chloride channels (3234).SLC26A7/Slc26a7 is predominantly expressed in the kidney and stomach (28, 29). In the kidney, Slc26a7 co-localizes with AE1, a well-known Cl/HCO3 exchanger, on the basolateral membrane of (acid-secreting) A-intercalated cells in OMCD cells (29, 35, 36) (supplemental Fig. 1). In the stomach, Slc26a7 co-localizes with AE2, a major Cl/HCO3 exchanger, on the basolateral membrane of acid secreting parietal cells (28). To address the physiological function of Slc26a7 in the intact mouse, we have generated Slc26a7 ko mice. We report here that Slc26a7 ko mice exhibit distal renal tubular acidosis and impaired gastric acidification in the absence of morphological abnormalities in kidney or stomach.  相似文献   

11.
12.
13.
Stathmin is an important regulator of microtubule polymerization and dynamics. When unphosphorylated it destabilizes microtubules in two ways, by reducing the microtubule polymer mass through sequestration of soluble tubulin into an assembly-incompetent T2S complex (two α:β tubulin dimers per molecule of stathmin), and by increasing the switching frequency (catastrophe frequency) from growth to shortening at plus and minus ends by binding directly to the microtubules. Phosphorylation of stathmin on one or more of its four serine residues (Ser16, Ser25, Ser38, and Ser63) reduces its microtubule-destabilizing activity. However, the effects of phosphorylation of the individual serine residues of stathmin on microtubule dynamic instability have not been investigated systematically. Here we analyzed the effects of stathmin singly phosphorylated at Ser16 or Ser63, and doubly phosphorylated at Ser25 and Ser38, on its ability to modulate microtubule dynamic instability at steady-state in vitro. Phosphorylation at either Ser16 or Ser63 strongly reduced or abolished the ability of stathmin to bind to and sequester soluble tubulin and its ability to act as a catastrophe factor by directly binding to the microtubules. In contrast, double phosphorylation of Ser25 and Ser38 did not affect the binding of stathmin to tubulin or microtubules or its catastrophe-promoting activity. Our results indicate that the effects of stathmin on dynamic instability are strongly but differently attenuated by phosphorylation at Ser16 and Ser63 and support the hypothesis that selective targeting by Ser16-specific or Ser63-specific kinases provides complimentary mechanisms for regulating microtubule function.Stathmin is an 18-kDa ubiquitously expressed microtubule-destabilizing phosphoprotein whose activity is modulated by phosphorylation of its four serine residues, Ser16, Ser25, Ser38, and Ser63 (17). Several classes of kinases have been identified that phosphorylate stathmin, including kinases associated with cell growth and differentiation such as members of the mitogen-activated protein kinase (MAPK)2 family, cAMP-dependent protein kinase (15, 811), and kinases associated with cell cycle regulation such as cyclin-dependent kinase 1 (3, 1214). Phosphorylation of stathmin is required for cell cycle progression through mitosis and for proper assembly/function of the mitotic spindle (3, 1316). Inhibition of stathmin phosphorylation produces strong mitotic phenotypes characterized by disassembly and disorganization of mitotic spindles and abnormal chromosome distributions (3, 1314).Stathmin is known to destabilize microtubules in two ways. One is by binding to soluble tubulin and forming a stable complex that cannot polymerize into microtubules, consisting of one molecule of stathmin and two molecules of tubulin (T2S complex) (1724). Addition of stathmin to microtubules in equilibrium with soluble tubulin results in sequestration of the tubulin and a reduction in the level of microtubule polymer (1718, 22, 2528). In addition to reducing the amount of assembled polymer, tubulin sequestration by stathmin has been shown to increase the switching frequency at microtubule plus ends from growth to shortening (called the catastrophe frequency) as the microtubules relax to a new steady state (17, 29). The second way is by binding directly to microtubules (2730). The direct binding of stathmin to microtubules increases the catastrophe frequency at both ends of the microtubules and considerably more strongly at minus ends than at plus ends (27). Consistent with its strong catastrophe-promoting activity at minus ends, stathmin increases the treadmilling rate of steady-state microtubules in vitro (27). These results have led to the suggestion that stathmin might be an important cellular regulator of minus-end microtubule dynamics (27).Phosphorylation of stathmin diminishes its ability to regulate microtubule polymerization (3, 14, 2526). Phosphorylation of Ser16 or Ser63 appears to be more critical than phosphorylation of Ser25 and Ser38 for the ability of stathmin to bind to soluble tubulin and to inhibit microtubule assembly in vitro (3, 25). Inhibition of stathmin phosphorylation induces defects in spindle assembly and organization (3, 14) suggesting that not only soluble tubulin-microtubule levels are regulated by phosphorylation of stathmin, but the dynamics of microtubules could also be regulated in a phosphorylation-dependent manner.It is not known how phosphorylation at any of the four serine residues of stathmin affects its ability to regulate microtubule dynamics and, specifically, its ability to increase the catastrophe frequency at plus and minus ends due to its direct interaction with microtubules. Thus, we determined the effects of stathmin individually phosphorylated at either Ser16 or Ser63 and doubly phosphorylated at both Ser25 and Ser38 on dynamic instability at plus and minus ends in vitro at microtubule polymer steady state and physiological pH (pH 7.2). We find that phosphorylation of Ser16 strongly reduces the direct catastrophe-promoting activity of stathmin at plus ends and abolishes it at minus ends, whereas phosphorylation of Ser63 abolishes the activity at both ends. The effects of phosphorylation of individual serines correlated well with stathmin''s reduced abilities to form stable T2S complexes, to inhibit microtubule polymerization, and to bind to microtubules. In contrast, double phosphorylation of Ser25 and Ser38 did not alter the ability of stathmin to modulate dynamic instability at the microtubule ends, its ability to form a stable T2S complex, or its ability to bind to microtubules. The data further support the hypotheses that phosphorylation of stathmin on either Ser16 or Ser63 plays a critical role in regulating microtubule polymerization and dynamics in cells.  相似文献   

14.
Time series incubations were conducted to provide estimates for the size selectivities and rates of protistan grazing that may be occurring in a sandy, contaminated aquifer. The experiments involved four size classes of fluorescently labeled groundwater bacteria (FLB) and 2- to 3-μm-long nanoflagellates, primarily Spumella guttula (Ehrenberg) Kent, that were isolated from contaminated aquifer sediments (Cape Cod, Mass.). The greatest uptake and clearance rates (0.77 bacteria · flagellate−1 · h−1 and 1.4 nl · flagellate−1 · h−1, respectively) were observed for 0.8- to 1.5-μm-long FLB (0.21-μm3 average cell volume), which represent the fastest growing bacteria within the pore fluids of the contaminated aquifer sediments. The 19:1 to 67:1 volume ratios of nanoflagellate predators to preferred bacterial prey were in the lower end of the range commonly reported for other aquatic habitats. The grazing data suggest that the aquifer nanoflagellates can consume as much as 12 to 74% of the unattached bacterial community in 1 day and are likely to have a substantive effect upon bacterial degradation of organic groundwater contaminants.While heterotrophic protists have been found in pristine and contaminated aquifers (3, 29, 34, 37, 5457), very little research has been performed to elucidate their role in the subsurface. In other environments (e.g., surface and marine waters, topsoil, and wastewater treatment plants), it is well documented that they typically consume bacteria (2, 11, 15, 41, 42, 47), although some have been observed to consume high-molecular-weight organics (48, 59) and even viruses (20, 39). Protists typically graze selectively, depending upon the size (1, 9, 17, 25, 52), growth condition (18, 53), species (16, 17, 35), and motility (18) of their prey. In carbon-limited environments, protists decrease bacterial competition, resulting in a greater bacterial uptake rate for organic substrate per unit of bacterial biomass (27). Based upon indirect field observations, it is also hypothesized that this may be the role they play in organically contaminated aquifers (31). In nutrient-limited environments, protists may release nitrogen or phosphorus needed by bacteria (10, 28, 44, 61).Studies at the U.S. Geological Survey’s (USGS) Toxic Substances Hydrology Program research site at the Massachusetts Military Reservation (MMR) on Cape Cod, Mass., have shown that sandy aquifer sediments can harbor large protistan populations even at relatively low levels (≤2 mg/liter) of dissolved organic matter (30). Protistan abundances in the MMR aquifer plume range from 1 × 104 to 7 × 104 g (dry weight)−1 (30) and consist primarily of nanoflagellates (2 to 3 μm in length) (29) that belong to the genera Bodo, Cercomonas, Cryptaulax, Cyanthomonas, Goniomonas, and Spumella, along with some undescribed species (37). A few amoebae (63) and no ciliates (29) have been observed.Results of a principal-component factor analysis of protistan and bacterial abundances and chemical constituents in the MMR plume suggested that the flagellates were preying upon unattached bacteria (30). Additional evidence of predation was obtained from flowthrough columns of aquifer sediment from which fluorescently labeled unattached bacteria eluted at much lower rates than they did from sterile (protist-free) controls (31). However, these results provide only indirect evidence of predation because no enumerations of the bacteria within the flagellates were performed.The purpose of the research reported in this paper was to directly determine whether the MMR nanoflagellates can consume unattached bacteria in the plume and the extent to which they engage in size-selective grazing. Rates of bacterivory (grazing and clearance rates) were estimated in the laboratory by using fluorescently labeled, monodispersed bacteria (FLB) and nanoflagellates that had been isolated from the MMR aquifer plume. Although other methods exist (19, 26, 38, 43, 45, 62, 6466), we chose to use fluorescent labeling to study flagellate bacterivory because this procedure requires shorter incubation times and relies upon direct visual observation of the prey within the predator. In addition, experiments could be designed with different sizes of FLB to determine if the nanoflagellates can discriminate between prey. This involved using FLB with cell lengths of 0.1 to 0.5, 0.5 to 0.8, 0.8 to 1.5, and >1.5 μm (average cell volumes of 0.06, 0.14, 0.21, and 0.87 μm3, respectively) in the grazing experiments.  相似文献   

15.
Obstructive nephropathy is a frequently encountered situation in newborns. In previous studies, the urinary peptidome has been analyzed for the identification of clinically useful biomarkers of obstructive nephropathy. However, the urinary proteome has not been explored yet and should allow additional insight into the pathophysiology of the disease. We have analyzed the urinary proteome of newborns (n = 5/group) with obstructive nephropathy using label free quantitative nanoLC-MS/MS allowing the identification and quantification of 970 urinary proteins. We next focused on proteins exclusively regulated in severe obstructive nephropathy and identified Arginase 1 as a potential candidate molecule involved in the development of obstructive nephropathy, located at the crossroad of pro- and antifibrotic pathways. The reduced urinary abundance of Arginase 1 in obstructive nephropathy was verified in independent clinical samples using both Western blot and MRM analysis. These data were confirmed in situ in kidneys obtained from a mouse obstructive nephropathy model. In addition, we also observed increased expression of Arginase 2 and increased total arginase activity in obstructed mouse kidneys. mRNA expression analysis of the related arginase pathways indicated that the pro-fibrotic arginase-related pathway is activated during obstructive nephropathy. Taken together we have identified a new actor in the development of obstructive nephropathy in newborns using quantitative urinary proteomics and shown its involvement in an in vivo model of disease. The present study demonstrates the relevance of such a quantitative urinary proteomics approach with clinical samples for a better understanding of the pathophysiology and for the discovery of potential therapeutic targets.Congenital obstructive nephropathy is the main cause of end stage renal disease (ESRD) in children (1). The most frequently found cause of congenital obstructive nephropathy is ureteropelvic junction (UPJ)1 obstruction with an estimated incidence of 1 in 1000–1500 births. Milder forms of UPJ obstruction often progress to the spontaneous resolution of the pathology over time. This has led to a watchful waiting approach with surgical intervention only if renal deterioration is detected (2). Although this medical surveillance prevents unnecessary surgery, it mostly relies on invasive follow-up. Consequently with the aim to reduce this invasive follow-up, several groups have initiated research to identify noninvasive urinary biomarkers of UPJ obstruction using both targeted and nontargeted (e.g. proteome analysis based) strategies. Targeted strategies including urinary cytokine expression analyses failed to clearly determine the need for surgery in UPJ obstruction (3, 4). On the other hand, untargeted strategies have been more successful and by using urinary proteomics, biomarkers for renal and non-renal diseases have been identified (59). Using urinary peptidome analysis, we identified and validated a urinary peptide panel that predicted the clinical outcome of newborns with UPJ obstruction with 97% accuracy several months in advance (3, 10). An independent small-scale study confirmed the efficiency of this biomarker panel (7). These studies indicate the potential of urinary proteomics to predict the clinical fate of patients with UPJ obstruction. Although these endogenous urinary peptide biomarkers are of great potential clinical value, sequencing of these biomarkers mainly identified collagen fragments that are less informative on the pathophysiology of the disease. In contrast, studies of the high molecular weight urinary proteome (i.e. proteins) might be more informative on the pathophysiology of disease. Different approaches have been used in the past to characterize the urinary proteome, either by 2D-gel electrophoresis coupled to mass spectrometry (11, 12) or reverse phase liquid chromatography coupled to tandem mass spectrometry (LC-MS/MS) analysis (1316). In-depth proteome analysis using extensive fractionation of the sample and high resolution, fast sequencing mass spectrometers have reported the identification of >2000 proteins in normal human urine (13, 15, 16). Here, we applied quantitative high-resolution label free LC-MS/MS analysis for the identification of urinary proteins associated to UPJ obstruction in newborns. Among a number of proteins uniquely associated with severe UPJ obstruction, we identified Arginase 1, not previously recognized in UPJ obstruction. Using an independent larger cohort, we further verified reduced urinary abundance of Arginase 1 using both Western blot and multiple reaction monitoring (MRM). Using the mouse model of obstructive nephropathy, we observed that the expression of arginases is modulated in situ in obstructed kidneys. Further gene expression analysis of the arginase pathway allowed us to hypothesize for a role of arginases in the development of fibrotic lesions in obstructive nephropathy.  相似文献   

16.
Prion propagation involves a conformational transition of the cellular form of prion protein (PrPC) to a disease-specific isomer (PrPSc), shifting from a predominantly α-helical conformation to one dominated by β-sheet structure. This conformational transition is of critical importance in understanding the molecular basis for prion disease. Here, we elucidate the conformational properties of a disulfide-reduced fragment of human PrP spanning residues 91–231 under acidic conditions, using a combination of heteronuclear NMR, analytical ultracentrifugation, and circular dichroism. We find that this form of the protein, which similarly to PrPSc, is a potent inhibitor of the 26 S proteasome, assembles into soluble oligomers that have significant β-sheet content. The monomeric precursor to these oligomers exhibits many of the characteristics of a molten globule intermediate with some helical character in regions that form helices I and III in the PrPC conformation, whereas helix II exhibits little evidence for adopting a helical conformation, suggesting that this region is a likely source of interaction within the initial phases of the transformation to a β-rich conformation. This precursor state is almost as compact as the folded PrPC structure and, as it assembles, only residues 126–227 are immobilized within the oligomeric structure, leaving the remainder in a mobile, random-coil state.Prion diseases, such as Creutzfeldt-Jacob and Gerstmann-Sträussler-Scheinker in humans, scrapie in sheep, and bovine spongiform encephalopathy in cattle, are fatal neurological disorders associated with the deposition of an abnormally folded form of a host-encoded glycoprotein, prion (PrP)2 (1). These diseases may be inherited, arise sporadically, or be acquired through the transmission of an infectious agent (2, 3). The disease-associated form of the protein, termed the scrapie form or PrPSc, differs from the normal cellular form (PrPC) through a conformational change, resulting in a significant increase in the β-sheet content and protease resistance of the protein (3, 4). PrPC, in contrast, consists of a predominantly α-helical structured domain and an unstructured N-terminal domain, which is capable of binding a number of divalent metals (512). A single disulfide bond links two of the main α-helices and forms an integral part of the core of the structured domain (13, 14).According to the protein-only hypothesis (15), the infectious agent is composed of a conformational isomer of PrP (16) that is able to convert other isoforms to the infectious isomer in an autocatalytic manner. Despite numerous studies, little is known about the mechanism of conversion of PrPC to PrPSc. The most coherent and general model proposed thus far is that PrPC fluctuates between the dominant native state and minor conformations, one or a set of which can self-associate in an ordered manner to produce a stable supramolecular structure composed of misfolded PrP monomers (3, 17). This stable, oligomeric species can then bind to, and stabilize, rare non-native monomer conformations that are structurally complementary. In this manner, new monomeric chains are recruited and the system can propagate.In view of the above model, considerable effort has been devoted to generating and characterizing alternative, possibly PrPSc-like, conformations in the hope of identifying common properties or features that facilitate the formation of amyloid oligomers. This has been accomplished either through PrPSc-dependent conversion reactions (1820) or through conversion of PrPC in the absence of a PrPSc template (2125). The latter approach, using mainly disulfide-oxidized recombinant PrP, has generated a wide range of novel conformations formed under non-physiological conditions where the native state is relatively destabilized. These conformations have ranged from near-native (14, 26, 27), to those that display significant β-sheet content (21, 23, 2833). The majority of these latter species have shown a high propensity for aggregation, although not all are on-pathway to the formation of amyloid. Many of these non-native states also display some of the characteristics of PrPSc, such as increased β-sheet content, protease resistance, and a propensity for oligomerization (28, 29, 31) and some have been claimed to be associated with the disease process (34).One such PrP folding intermediate, termed β-PrP, differs from the majority of studied PrP intermediate states in that it is formed by refolding the PrP molecule from the native α-helical conformation (here termed α-PrP), at acidic pH in a reduced state, with the disulfide bond broken (22, 35). Although no covalent differences between the PrPC and PrPSc have been consistently identified to date, the role of the disulfide bond in prion propagation remains disputed (25, 3639). β-PrP is rich in β-sheet structure (22, 35), and displays many of the characteristics of a PrPSc-like precursor molecule, such as partial resistance to proteinase K digestion, and the ability to form amyloid fibrils in the presence of physiological concentrations of salts (40).The β-PrP species previously characterized, spanning residues 91–231 of PrP, was soluble at low ionic strength buffers and monomeric, according to elution volume on gel filtration (22). NMR analysis showed that it displayed radically different spectra to those of α-PrP, with considerably fewer observable peaks and markedly reduced chemical shift dispersion. Data from circular dichroism experiments showed that fixed side chain (tertiary) interactions were lost, in contrast to the well defined β-sheet secondary structure, and thus in conjunction with the NMR data, indicated that β-PrP possessed a number of characteristics associated with a “molten globule” folding intermediate (22). Such states have been proposed to be important in amyloid and fibril formation (41). Indeed, antibodies raised against β-PrP (e.g. ICSM33) are capable of recognizing native PrPSc (but not PrPC) (4244). Subsequently, a related study examining the role of the disulfide bond in PrP folding confirmed that a monomeric molten globule-like form of PrP was formed on refolding the disulfide-reduced protein at acidic pH, but reported that, under their conditions, the circular dichroism response interpreted as β-sheet structure was associated with protein oligomerization (45). Indeed, atomic force microscopy on oligomeric full-length β-PrP (residues 23–231) shows small, round particles, showing that it is capable of formation of oligomers without forming fibrils (35). Notably, however, salt-induced oligomeric β-PrP has been shown to be a potent inhibitor of the 26 S proteasome, in a similar manner to PrPSc (46). Impairment of the ubiquitin-proteasome system in vivo has been linked to prion neuropathology in prion-infected mice (46).Although the global properties of several PrP intermediate states have been determined (3032, 35), no information on their conformational properties on a sequence-specific basis has been obtained. Their conformational properties are considered important, as the elucidation of the chain conformation may provide information on the way in which these chains pack in the assembly process, and also potentially provide clues on the mechanism of amyloid assembly and the phenomenon of prion strains. As the conformational fluctuations and heterogeneity of molten globule states give rise to broad NMR spectra that preclude direct observation of their conformational properties by NMR (4750), here we use denaturant titration experiments to determine the conformational properties of β-PrP, through the population of the unfolded state that is visible by NMR. In addition, we use circular dichroism and analytical ultracentrifugation to examine the global structural properties, and the distribution of multimeric species that are formed from β-PrP.  相似文献   

17.
18.
STIM1 and ORAI1, the two limiting components in the Ca2+ release-activated Ca2+ (CRAC) signaling cascade, have been reported to interact upon store depletion, culminating in CRAC current activation. We have recently identified a modulatory domain between amino acids 474 and 485 in the cytosolic part of STIM1 that comprises 7 negatively charged residues. A STIM1 C-terminal fragment lacking this domain exhibits enhanced interaction with ORAI1 and 2–3-fold higher ORAI1/CRAC current densities. Here we focused on the role of this CRAC modulatory domain (CMD) in the fast inactivation of ORAI1/CRAC channels, utilizing the whole-cell patch clamp technique. STIM1 mutants either with C-terminal deletions including CMD or with 7 alanines replacing the negative amino acids within CMD gave rise to ORAI1 currents that displayed significantly reduced or even abolished inactivation when compared with STIM1 mutants with preserved CMD. Consistent results were obtained with cytosolic C-terminal fragments of STIM1, both in ORAI1-expressing HEK 293 cells and in RBL-2H3 mast cells containing endogenous CRAC channels. Inactivation of the latter, however, was much more pronounced than that of ORAI1. The extent of inactivation of ORAI3 channels, which is also considerably more prominent than that of ORAI1, was also substantially reduced by co-expression of STIM1 constructs missing CMD. Regarding the dependence of inactivation on Ca2+, a decrease in intracellular Ca2+ chelator concentrations promoted ORAI1 current fast inactivation, whereas Ba2+ substitution for extracellular Ca2+ completely abrogated it. In summary, CMD within the STIM1 cytosolic part provides a negative feedback signal to Ca2+ entry by triggering fast Ca2+-dependent inactivation of ORAI/CRAC channels.The Ca2+ release-activated Ca2+ (CRAC)5 channel is one of the best characterized store-operated entry pathways (17). Substantial efforts have led to identification of two key components of the CRAC channel machinery: the stromal interaction molecule 1 (STIM1), which is located in the endoplasmic reticulum and acts as a Ca2+ sensor (810), and ORAI1/CRACM1, the pore-forming subunit of the CRAC channel (1113). Besides ORAI1, two further homologues named ORAI2 and ORAI3 belong to the ORAI channel family (12, 14).STIM1 senses endoplasmic reticulum store depletion primarily by its luminal EF-hand in its N terminus (8, 15), redistributes close to the plasma membrane, where it forms puncta-like structures, and co-clusters with ORAI1, leading to inward Ca2+ currents (12, 1619). The STIM1 C terminus, located in the cytosol, contains two coiled-coil regions overlapping with an ezrin-radixin-moesin (ERM)-like domain followed by a serine/proline- and a lysine-rich region (2, 8, 2022). Three recent studies have described the essential ORAI-activating region within the ERM domain, termed SOAR (Stim ORAI-activating region) (23), OASF (ORAI-activating small fragment) (24), and CAD (CRAC-activating domain) (25), including the second coiled coil domain and the following ∼55 amino acids. We and others have provided evidence that store depletion leads to a dynamic coupling of STIM1 to ORAI1 (2628) that is mediated by a direct interaction of the STIM1 C terminus with ORAI1 C terminus probably involving the putative coiled-coil domain in the latter (27).Furthermore, different groups have proven that the C terminus of STIM1 is sufficient to activate CRAC as well as ORAI1 channels independent of store depletion (2225, 27, 29). We have identified that OASF-(233–474) or shorter fragments exhibit further enhanced coupling to ORAI1 resulting in 3-fold increased constitutive Ca2+ currents. A STIM1 fragment containing an additional cluster of anionic amino acids C-terminal to position 474 displays weaker interaction with ORAI1 as well as reduced Ca2+ current comparable with that mediated by wild-type STIM1 C terminus. Hence, we have suggested that these 11 amino acids (474–485) act in a modulatory manner onto ORAI1; however, their detailed mechanistic impact within the STIM1/ORAI1 signaling machinery has remained so far unclear.In this study, we focused on the impact of this negative cluster on fast inactivation of STIM1-mediated ORAI Ca2+ currents. Lis et al. (30) have shown that all three ORAI homologues display distinct inactivation profiles, where ORAI2 and ORAI3 show a much more pronounced fast inactivation than ORAI1. Moreover, it has been reported (31) that different expression levels of STIM1 to ORAI1 affect the properties of CRAC current inactivation. Yamashita et al. (32) have demonstrated a linkage between the selectivity filter of ORAI1 and its Ca2+-dependent fast inactivation. Here we provide evidence that a cluster of acidic residues within the C terminus of STIM1 is involved in the fast inactivation of ORAI1 and further promotes that of ORAI3 and native CRAC currents.  相似文献   

19.
20.
The cis-trans peptidylprolyl isomerase Pin1 plays a critical role in regulating a subset of phosphoproteins by catalyzing conformational changes on the phosphorylated Ser/Thr-Pro motifs. The phosphorylation-directed ubiquitination is one of the major mechanisms to regulate the abundance of p27Kip1. In this study, we demonstrate that Pin1 catalyzes the cis-trans conformational changes of p27Kip1 and further mediates its stability through the polyubiquitination mechanism. Our results show that the phosphorylated Thr-187-Pro motif in p27Kip1 is a key Pin1-binding site. In addition, NMR analyses show that this phosphorylated Thr-187-Pro site undergoes conformational change catalyzed by Pin1. Moreover, in Pin1 knock-out mouse embryonic fibroblasts, p27Kip1 has a shorter lifetime and displays a higher degree of polyubiquitination than in Pin1 wild-type mouse embryonic fibroblasts, suggesting that Pin1 plays a critical role in regulating p27Kip1 degradation. Additionally, Pin1 dramatically reduces the interaction between p27Kip1 and Cks1, possibly via isomerizing the cis-trans conformation of p27Kip1. Our study thus reveals a novel regulatory mechanism for p27Kip1 stability and sheds new light on the biological function of Pin1 as a general regulator of protein stability.Cellular differentiation and cell cycle inhibition are tightly controlled via sensitive molecular mechanisms. p27Kip1, a member of the Cip/Kip family, is an essential cell cycle inhibitor that functions largely during the G0/G1 phase where it promotes the assembly of the cyclin D1-CDK4 complex and inhibits the kinase activity of the cyclin E-CDK2 complex in the G1-S phase (14). Several review articles have elegantly summarized and discussed the detailed cellular functions of p27Kip1 (16). p27Kip1 is also a phosphoprotein with multiple Ser/Thr phosphorylation sites, including Ser-10, Ser-178, and Thr-187, followed by a proline residue. Hence, these motifs are potential substrate sites for proline-directed kinases (5, 6). Compared with Ser-178, which has not yet been well studied, the phosphorylation of Ser-10 and Thr-187 has been well characterized to be important for the regulation of p27Kip1 function. For instance, Ser-10 has been found to be the major phosphorylation site of p27Kip1 (7) and to play an important role in regulating cell migration (810), although the regulation of Ser-10 phosphorylation is still not completely defined (11, 12).In contrast to Ser-10 and Thr-178, Thr-187 is the best characterized phosphorylation site on p27Kip1 and is known to regulate the complex formation of p27Kip1-cyclin E-CDK2 (12). In addition, it is also widely accepted that Thr-187 plays a crucial role in determining the abundance of mature p27Kip1 proteins. The phosphorylation of Thr-187 directs p27Kip1 to an SCFSkp2 ubiquitin ligase complex (consisting of Skp2-Skp1-Cks1-Cul1-Roc1), which in turn promotes the polyubiquitination and degradation of p27Kip1 (13, 14). The crystal structure of the Skp1-Skp2-Cks1-p27Kip1 phosphopeptide complex shows that p27Kip1 binds both Cks1 and Skp2 and that the C terminus of Skp2 and Cks1 forms the substrate recognition core of the SCF complex (15). Furthermore, the structure of this complex has revealed that the phosphorylation of Thr-187 in p27Kip1 is recognized by the phosphate-binding site of Cks1, indicating that Cks1 is not only a facilitator but also an indispensable component in p27Kip1 degradation machinery (15).Pin1 is a unique peptidyl-prolyl isomerase (PPIase)2 that recognizes only the phosphorylated Ser/Thr motif preceding a proline residue (16). In addition, Pin1 is very prominent in isomerizing the cis-trans conformation of prolyl-peptidyl bonds in its substrates, resulting in either the modification of their function (e.g. c-Jun (17), β-catenin (18), Bax (19), and Notch1 (20)) or modulation of their stability (e.g. cyclin D1 (21), p53 (22, 23), and NF-κB (24)). Loss of Pin1 in mice results in several phenotypes similar to those of cyclin D1-null mice (21) and neuronal degenerative phenotypes (2528), suggesting the conformational changes mediated by Pin1 may be crucial for the normal functioning of cells. Additionally, Pin1 also plays important roles in cancer and other cellular events, which have been extensively discussed in several recent review articles (2933).In this study, we show that Pin1 binds to p27Kip1, mainly through the phosphorylated Thr-187-Pro motif, and causes subsequent prolyl isomerization of this cell cycle protein. Moreover, we also find that Pin1 can protect p27Kip1 from degradation. Importantly, we demonstrate that by catalyzing conformational changes in p27Kip1, Pin1 hinders its association with Cks1, resulting in a reduction of polyubiquitination of p27Kip1 and protecting its degradation by SCFSkp2 complexes. Our results suggest that the cis-trans isomerization catalyzed by Pin1 represents a novel regulatory mechanism during post-phosphorylation of proteins and polyubiquitination-directed degradation pathways.  相似文献   

设为首页 | 免责声明 | 关于勤云 | 加入收藏

Copyright©北京勤云科技发展有限公司  京ICP备09084417号