Báo cáo khoa học: Protein tyrosine phosphatases: structure–function relationships ppt

16 319 0
Báo cáo khoa học: Protein tyrosine phosphatases: structure–function relationships ppt

Đang tải... (xem toàn văn)

Tài liệu hạn chế xem trước, để xem đầy đủ mời bạn chọn Tải xuống

Thông tin tài liệu

MINIREVIEW Protein tyrosine phosphatases: structure–function relationships Lydia Tabernero 1 , A. Radu Aricescu 2 , E. Yvonne Jones 2 and Stefan E. Szedlacsek 3 1 Faculty of Life Sciences, University of Manchester, UK 2 Wellcome Trust Centre for Human Genetics, University of Oxford, UK 3 Institute of Biochemistry of the Romanian Academy, Bucharest, Romania The most significant trait of the protein tyrosine phos- phatase (PTP) superfamily is conservation of the signa- ture motif CX 5 R, which forms the phosphate-binding loop in the active site (known as the P-loop or PTP- loop). Despite relatively large sequence variations in the X 5 segment, the conformation of the P-loop is strictly conserved and can be easily superimposed from different PTP structures, with minor deviations in the Ca tracing (< 1 A ˚ ). This structurally conserved arrangement ensures that the catalytic Cys, the nucleo- phile in catalysis, and the Arg, involved in phosphate binding, remain in close proximity and form a cradle to hold the phosphate group of the substrate in place for nucleophilic attack. The cysteine Sc-atom is the nucleophile that attacks the substrate phosphorus atom leading to the cysteinyl-phosphate reaction inter- mediate. The arginine is involved both in substrate binding and in stabilization of the reaction intermedi- ate [1]. Further to this, the amide groups in the P-loop point towards the interior of the cradle and form a network of hydrogen bonds to the phosphate oxygens (Fig. 1A). A conserved Ser ⁄ Thr residue in the P-loop has been proposed to play an important role in the stabilization of the thiolate group in the transition state facilitating the breakdown of the phosphoenzyme intermediate [2] (Scheme 1). The catalytic mechanism of PTP reaction requires the participation of a general acid and a general base. Keywords catalytic mechanism; cell adhesion; cell signalling; drug design; ligand binding; NMR; phosphatase inhibitor; protein structure; protein tyrosine phosphatase; structure–function receptor; X-ray crystallography Correspondence L. Tabernero, Faculty of Life Sciences, University of Manchester, Michael Smith Building, Manchester M13 9PT, UK Fax: +44 161275 5082 Tel: +44 1612757794 E-mail: Lydia.Tabernero@manchester.ac.uk (Received 27 October 2007, accepted 18 December 2007) doi:10.1111/j.1742-4658.2008.06251.x Structural analysis of protein tyrosine phosphatases (PTPs) has expanded considerably in the last several years, producing more than 200 structures in this class of enzymes (from 35 different proteins and their complexes with ligands). The small–medium size of the catalytic domain of 280 resi- dues plus a very compact fold makes it amenable to cloning and over- expression in bacterial systems thus facilitating crystallographic analysis. The low molecular weight PTPs being even smaller, 150 residues, are also perfect targets for NMR analysis. The availability of different structures and complexes of PTPs with substrates and inhibitors has provided a wealth of information with profound effects in the way we understand their biological functions. Developments in mammalian expression technology recently led to the first crystal structure of a receptor-like PTP extracellular region. Altogether, the PTP structural work significantly advanced our knowledge regarding the architecture, regulation and substrate specificity of these enzymes. In this review, we compile the most prominent structural traits that characterize PTPs and their complexes with ligands. We discuss how the data can be used to design further functional experiments and as a basis for drug design given that many PTPs are now considered strategic therapeutic targets for human diseases such as diabetes and cancer. Abbreviations KIM, kinase interaction motif; LMW-PTP, low molecular weight protein tyrosine phosphatase; N-SH2, N-terminal SH2 domain; PTP, protein tyrosine phosphatase; RPTP, receptor protein tyrosine phosphatase; YopH, Yersinia PTP. FEBS Journal 275 (2008) 867–882 ª 2008 The Authors Journal compilation ª 2008 FEBS 867 This is provided by a unique aspartic residue situated on the WPD-loop. During formation of the transition state intermediate, the catalytic Asp acts as a general acid protonating the oxygen of the leaving group in the tyrosine residue. In the second catalysis step, the same Asp functions as a general base during hydrolysis of the phospho-enzyme by accepting a proton from the attacking water and assisting in the conversion of the phospho-Cys enzyme to its resting Cys-SH state, thus regenerating the free enzyme [2,3]. Upon substrate binding, the WPD-loop closes over the active site bringing the catalytic Asp near the leaving group. An analogous Asp residue is found in the DPYY-loop of the low molecular weight protein tyrosine phosphatases (LMW-PTPs), although in this enzyme it appears to be less mobile than the WPD-loop and it adopts a fixed position near the active site. We focus our review on the tyrosine-specific PTPs with a Cys-based mechanism of catalysis (class I and class II) as described in the classification by Alonso et al. [4]. Cytoplasmic class I PTPs Cytoplasmic PTPs, also called soluble or non-receptor PTPs, have a modular organization. In addition to the highly conserved catalytic domain they contain non- catalytic regions or domains that play a role in subcel- lular targeting, in regulation of the enzymatic activity or in recruiting specific ligands [4]. Structural characteristics of the PTP catalytic domain The catalytic domain contains 280 amino acids that determine a specific PTP fold with several characteristic Fig. 1. (A) Structure of the phosphate-binding loop (P-loop). Stick representation of the consensus signature motif (CX 5 R) that forms the P- loop present in the active site of PTPs. The P-loop from bovine LMW-PTP (1PNT) [79] is represented and the catalytic Cys12 and Arg18 are labelled. The amide nitrogens form hydrogen-bond interactions (dotted green lines) with the phosphatase bound showing network of interac- tions that involve the catalytic Arg. The cradle-like conformation of the P-loop is conserved in the structures of all PTPs. (B) Structure of PTP1B (C215S mutant) in complex with phosphotyrosine (PDB entry 1PTV). Position of the substrate in the active site is illustrated by the phosphotyrosine ligand (blue). Tyr46 within the ‘KNRY’ conserved motif contributes the substrate recognition. Active-site nucleophile Cys215 (grey) (here mutated to Ser) attacks the substrate phosphorus leading to the formation of the cysteinyl-phosphate intermediate. Asp181 within the WPD-loop (cyan), here in the closed conformation, acts as a general acid donating a proton to the phenolate leaving group. (C) Binding of an allosteric inhibitor of PTP1B keeps the catalytic WPD-loop in the open conformation. Structure of PTP1B (cyan) in complex with the allosteric inhibitor 3-(3,5-dibromo-4-hydroxy-benzoyl)-2-ethyl-benzofuran-6-sulfonic acid 4-sulfamoyl-phenyl)-amide (termed ‘com- pound-2’ in [18]) (PDB entry 1T49) overlain on the PTP1B (C215S mutant) structure (red) in complex with a p-Tyr substrate (PDB entry 1PTV). Only the main structural elements involved in allosteric inhibition are represented. In the presence of the allosteric inhibitor (yellow), the C-terminus of PTP1B is disordered while in presence of the phosphotyrosine (green) it adopts the a-helical structure a7. Binding of allo- steric inhibitor impedes the interaction between helices a3, a6 and a7, thus preventing the closure of the WPD-loop. Scheme 1. General mechanism of catalysis of PTPs. PTP structure–function relationship L. Tabernero et al. 868 FEBS Journal 275 (2008) 867–882 ª 2008 The Authors Journal compilation ª 2008 FEBS features. As illustrated by the structure of PTP1B (Fig. 1B), the first reported structure for a PTP [5], this fold is represented by a central, highly twisted b sheet composed of eight b strands forming a mixed b sheet with four parallel strands flanked by antiparallel ones. Six a helices surround the central sheet, four on one side and two on the other [6]. The active site is situated in a 9-A ˚ deep crevice, 3 A ˚ deeper than for dual-speci- ficity phosphatases, thus providing selectivity for phos- photyrosine-containing protein substrates [7]. The signature motif VHCSXGXGR(T ⁄ S)G that forms the PTP-loop [8] between the C-terminus of the central b10 strand and the a4 helix is located at the bottom of the catalytic site. This loop contains the essential cata- lytic residues Cys215 and Arg221 in PTP1B. An essen- tial structural component of the active site is the phosphotyrosine-recognition loop with the conserved motif KNRY (residues 43–46 in PTP1B) [7,8]. This loop determines the depth of the active site cleft and interacts, through its tyrosine residue, with the aro- matic ring of the phosphotyrosine in the substrate. Another key element in catalysis is the WPD-loop (res- idues 179–181 in PTP1B; Fig. 1B). Remarkably, sub- strate binding into the active site triggers a significant movement of 6A ˚ of the essential Asp residue, simultaneous with a conformational switch of the whole WPD-loop from an ‘open’ to a ‘closed’ position [9]. In addition to phosphopeptidic substrates [7], small ligands also induce closure of the WPD loop. Structures of PTPs with small ligands like tungstate [9], sulfate [10] and phosphate [11–13] show evidence of a closed WPD-loop. Apo forms of PTPs generally have the WPD-loop in the open conformation [6]. However, there are a few notable exceptions, for exam- ple, apo-PTP1B has the WPD-loop in the closed form [14], the PTP1B complex with tungstate contains the open form of the WPD-loop [6] and similarly, in the SHP-1 complex with phosphopeptide substrates EDTLTpYADLD or PSFSEpYASVQ the WPD-loop is also in the open form [15]. These apparently contra- dictory findings may in part be explained by assuming that the WPD-loop fluctuates between an open and a closed position and that the binding of ligands deter- mines a much longer residence time of the closed posi- tion [16]. Another catalytically important structural element of the PTP fold is the Q-loop defined by the QTXXQYXF motif [8]. The glutamine residues in this motif are highly conserved and their main role is to position a water molecule in the active site, which is involved in hydrolysis of the thio-phosphate intermediate. Overall, the structure of cytoplasmic PTPs is highly conserved with only minor differences in the main structural core. For example, the structure of SHP-2 contains a short b sheet formed by the N-terminal bA and C-terminal bN strands, which is not encountered in PTP1B but that is present in the structures of kinase interaction motif (KIM)-containing PTPs (PTPRR, HePTP, STEP) and receptor PTPs; also, in the struc- ture of KIM-containing PTPs there are several 3 10 heli- ces which are not observed in the structures of other PTPs [11,13,17]. More prominent differences are visible at the N- and C-termini of the catalytic domains of sev- eral PTPs. Thus, PTP1B contains an additional C-ter- minal helix a7, not present in other PTP structures. This is a particularly important regulatory element for the catalytic activity of PTP1B as it stabilizes closure of WPD-loop through its interaction with helices a3 and a6 [18]. This idea is supported by the recent finding that most of inhibitor–resistant mutants of PTP1B are clustered on helix a7 and its surroundings [19]. A number of reported PTP structures have an addi- tional N-terminal helix (generically termed ‘a0’) with a relatively less-conserved sequence among PTPs, but are apparently important functionally. In SHP-1 this helix a0 is highly mobile: whereas in the peptide-bound form it is positioned far away from the catalytic core, in the ligand free-form it is rotated 60° and is located in the proximity of the PTP domain [15,20]. Experi- mental results support the idea that a0 of SHP-1 plays a cooperative role in substrate recognition [21]. PTPRR, HePTP, STEP and other KIM-containing PTPs also display an N-terminal helix a0. This is mainly stabilized by hydrophobic interactions with helix a5 and the loop following helix a2¢, forming a hydrophobic cavity of 16 A ˚ depth [11,13,17]. The role played by this helix seems to be related to the presence of: (a) its N-terminal residue (Thr253 in PTPRR and Thr45 in HePTP), which is specifically phosphorylated by ERK2 and p38 MAP kinases; and (b) a KIM located 15 residues upstream of a0, which is essential for the interaction with ERK2 and p38. It has been suggested that helix a0 contributes to the positioning of the KIM region for interaction with the docking groove of ERK2 and to the proper directing of Thr253 residue into the active site of ERK2 [17]. The structure of PTPL1 provides another example of the N-terminal helix playing a specific role. Its helix a0 is located at a topologically equivalent posi- tion to helix a7 in PTP1B [12], and deletion of this helix results in an enzyme with very low activity. This finding together with the fact that helix a0 interacts with helix a3 in a similar manner as helix a7 does in PTP1B, suggests that helix a0 may be a regulatory structural element, involved in stabilization of WPD- loop in PTPL1. L. Tabernero et al. PTP structure–function relationship FEBS Journal 275 (2008) 867–882 ª 2008 The Authors Journal compilation ª 2008 FEBS 869 Structural determinants for substrate binding and regulation of the catalytic activity The catalytic domains of PTPs often exhibit relatively broad substrate specificities in vitro, although under in vivo conditions full-length PTPs have considerably more stringent specificities. This is in part due to the presence of additional regulatory domains that direct their subcellular localization and interactions with spe- cific substrates. For example, the C-terminal segment of full-length PTP1B is responsible for anchoring PTP1B to the endoplasmic reticulum [22]. Substrate specificity may also be provided by selective substrate binding modules. The small KIM motif composed of 16 amino acids is an example of this strategy, it directs the KIM-containing PTPs to their physiological sub- strates, MAP kinases ERK2 or p38, which are subse- quently dephosphorylated and inactivated [23]. Yersinia PTP (YopH) provides another interesting example of a substrate-targeting mechanism with mul- tiple phosphotyrosine-binding sites. YopH is a viru- lence factor that dephosphorylates several focal adhesion proteins, for example, p130Cas [24] in human epithelial cells. It contains an N-terminal non-catalytic domain that binds tyrosine-phosphorylated proteins [25], but with a different fold than SH2 or PTB domains, thus representing a novel phosphotyrosine- binding domain [26]. Recently, the crystal structure of the YopH PTP domain in complex with a phospho- peptide revealed a second substrate-binding site (in addition to the active site) within the catalytic domain, that is the third phosphotyrosine binding site within the full-length YopH protein [27]. Moreover, it was proved that the two non-catalytic substrate-targeting sites co-operate in binding the p130Cas substrate, which contains 15 phosphorylation sites, thus provid- ing efficiency and specificity for the PTP domain inter- action with its specific substrate. The structures of three other cytoplasmic PTPs, PTP1B, TC-PTP and PTPL1, also reveal the presence of secondary phosphotyrosine-binding sites within their catalytic domains [12,28,29]. These secondary sub- strate-binding sites are represented by a positively charged pocket located close to the active site. In PTP1B, the secondary site is formed by Arg24, Arg254, Met258 and Gly259 and plays the important role of providing specificity for PTP1B action. This is illustrated by the fact that a physiological substrate of PTP1B, the insulin receptor kinase, contains a tandem of phosphotyrosine residues (1162 and 1163) and inter- acts with PTP1B in a characteristic bidentate mode: pTyr1162 is recognized and selectively dephosphoryl- ated by the active site, whereas pTyr1163 is bound to the secondary binding site of PTP1B [30]. The inter- action between the di-phosphorylated IRK peptide and the two substrate binding sites of PTP1B is highly selective, being 70-fold tighter than for the mono- phosphorylated peptides. The secondary site has also been exploited for development of inhibitors against PTP1B [31]. The crystal structures of SHP-1 and SHP-2 reveal a more sophisticated regulatory mechanism controlled by substrate recruitment. SHP-1 and SHP-2 contain an N-terminal tandem of two SH2 domains as well as a C-terminal extension. The apo forms of both SHP-1 and SHP-2 are essentially inactive. However, their cata- lytic activity increases considerably upon binding of phosphopeptides by their SH2 domains [32,33]. Com- parison of the structure of the auto-inhibited form of SHP-2 [34] with the structure of the N-terminal SH2 domain (N-SH2), both free and in complex with phos- phopeptides, shows that N-SH2 plays the role of an allosteric switch: in the absence of a phosphopeptide ligand it binds to the PTP domain blocking closure of the WPD-loop (through the N-SH2 b sheet defined by strands bD¢, bE and bF) but its phosphopeptide- binding site (defined by the loops EF and BG) is not functional; by contrast, binding of a phosphopeptide to the N-SH2 domain triggers concerted conformational transitions (involving EF loop, helix aB and b sheet bD’, bE and bF) which ultimately lead to dissociation of the N-SH2 from the PTP active site and subsequent activation. The other SH2 domain (C-SH2) seems not to be directly involved in regulation of catalytic activity and its role cannot be fully understood based on these structural data. However, it seems reasonable that C-SH2 acts in a cooperative manner with the N-SH2 domain in binding bisphosphorylated peptides. Consis- tent with this idea, it was proved that bisphosphorylat- ed peptides binding both SH2 domains stimulate SHP catalytic activity 100-fold, whereas mono-phosphory- lated peptides stimulate only 10-fold [35]. The structure of SHP-1 [20] supports an activation mechanism similar to that of SHP-2. The difference between the two struc- tures consists mainly in a different orientation of the C- SH2 domain of SHP-1 and also a higher flexibility of this domain as compared to the C-SH2 of SHP-2. A different type of allosteric regulation of the cata- lytic activity was also reported for PTP1B. Benzbroma- rone derivatives are non-competitive inhibitors of PTP1B [18] and they bind in a novel allosteric site in PTP1B, as shown in the crystal structure of the com- plexes (Fig. 1C). This binding site is located 2A ˚ from the active site and is formed by helices a3 and a6. The inhibitory effect seems to result from blocking the interaction between helices a7 and a3–a6, present PTP structure–function relationship L. Tabernero et al. 870 FEBS Journal 275 (2008) 867–882 ª 2008 The Authors Journal compilation ª 2008 FEBS in the closed form of PTP1B, thus preventing closure of the WPD-loop. A truncated form of PTP1B lack- ing a7 is fourfold less active than the native form. This finding, together with those mentioned above, provides additional support for the particular significance of helix a7 in controlling the catalytic activity of PTP1B. Reversible oxidation of the catalytic cysteine is a characteristic regulatory mechanism of PTPs [36], and therefore, there was significant interest in establishing its molecular basis. Two structural analyses evidenced that a potential key intermediate in this process is a sulfenyl-amide in which the sulfur atom of active site cysteine (Cys215 PTP1B ) forms a covalent bridge with the amide nitrogen of the neighbouring residue (Ser216 PTP1B ) [37,38]. Formation of the sulfenyl-amide causes conformational modifications in the catalytic site, and has a double role: first, it protects the cata- lytic cysteine from irreversible oxidation to sulfonic acid and, second, it facilitates reactivation of PTP by biological reducing agents, since this is a reversible reaction. In addition, the substantial conformational changes associated with sulfenyl-amide formation, may represent a signal that the given PTP is in a temporary inactive state. Cytoplasmic PTPs as drug targets Many PTP genes in the human genome have been implicated in human diseases, leading to a special interest in selecting PTPs as drug targets [39–41]. The finding that the PTP1B knockout mouse was highly responsive to insulin and was resistant to diet-induced obesity [42] triggered intense efforts to identify specific inhibitors of PTP1B to develop drugs against type II diabetes and obesity [43]. Numerous structures of PTP1B complexes with single-site, double-site or allo- steric inhibitors are now available [43] (see also the database of reported PTP structures at: http:// ptp.cshl.edu or http://science.novonordisk.com/ptp). One promising drug candidate, ertiprotafib (Wyeth Research, Cambridge, MA, USA) successfully entered phase II clinical trials for treatment of type II diabetes; however, due to unsatisfactory efficacy and dose-limit- ing effects the trial was terminated in 2002 [39]. Despite this partial failure, trials to find new inhibitors continued and new research directions to develop effi- cient PTP inhibitors were identified. Structural infor- mation on PTPs has substantially contributed to the success of the fragment-based approach to identify new inhibitory compounds. Using this procedure, dou- ble-site PTP1B inhibitors were designed to bind both the active site and the second phosphotyrosine binding site of PTP1B [31,44]. The novel series of potent inhib- itors exhibited sixfold selectivity over the highly homologous TCPTP and high selectivity over other phosphatases [44]. Class I receptor-like PTPs Twenty-one of the 38 classical PTPs identified in the human genome [4] are type I membrane proteins and, because of their domain organization, were termed ‘receptor-like’ PTPs (RPTPs) well before any ligands had been identified [45]. Typically, an RPTP has an N-terminal extracellular region (lengths vary from 100 to > 1000 residues), a single transmembrane region and one or two intracellular catalytic domains, highly con- served within the family and with other classical PTPs [46]. Given this architecture, RPTPs appear ideally built to transduce signals across the plasma membrane, trig- gered by ligand binding to the extracellular region. Structure and role of the extracellular region The remarkable structural variety of RPTP ectodo- mains offers, first of all, a convenient criterion for classification [4,46]. In most cases, several types of domains are combined to produce modular arrange- ments: commonly used folds include meprin ⁄ A5 ⁄ RPTPl (MAM), Ig-like, fibronectin type III-like, carbonic anhydrase-like and cysteine-rich regions. In addition, alternative splicing and post-translational modifications (mainly N- and O-linked glycosylation) play important regulatory roles [46–48] and potentially contribute to a diversification of epitopes available for ligand binding. However, despite almost two decades of sustained efforts, the number of RPTP ligands identified remains surprisingly low (see review by den Hertog, O ¨ stman and Bohmer in this issue). The few notable examples are heparan sulfate proteoglycans that bind type IIa RPTPs (agrin and collagen XVIII for RPTPr [49], syndecan and Dallylike for LAR in Drosophila [50]), the trans homophilic interactions of type IIb RPTPs RPTPl [51,52] and RPTPj [53], and pleiotrophin, a ligand for the type V RPTPf [54]. A direct effect on catalytic activity has, so far, only been demonstrated for RPTPf where pleiotrophin binding downregulates the catalytic activity; the exact mecha- nism of inhibition still remains unclear. Recently, structural work provided important insights into the ectodomain-dependent mechanisms that regulate type IIb RPTPs [55]. RPTPl is a homo- philic cell-adhesion molecule, expressed at high levels by neurons and vascular endothelia, and causing clus- tering when expressed on the surface of normally non- adherent cells in culture [51,52]; this activity is entirely L. Tabernero et al. PTP structure–function relationship FEBS Journal 275 (2008) 867–882 ª 2008 The Authors Journal compilation ª 2008 FEBS 871 driven by the extracellular region [56]. In confluent cell cultures, RPTPl surface expression is significantly increased, post translationally, and the protein appears to be trapped at cell–cell contact areas, presumably via trans homophilic interactions. A crystal structure of the full-length RPTPl ectodomain revealed an unex- pectedly extended and rigid architecture, with residues from all four N-terminal domains contributing to a large adhesive interface (covering around 1630 A ˚ 2 per monomer) [55]. The RPTPl trans dimer matches the dimensions of adherens junctions and, importantly, the length of cadherin trans dimers [57]. Moreover, cell- surface expression of RPTPl ectodomain deletion con- structs that still preserve the adhesive activity, induce intercellular spacings that correlate directly with the construct length, thus providing further evidence for existence of an extended ectodomain conformation as seen in the crystal structure. It was therefore suggested that the RPTPl extracellular region plays a fundamen- tal regulatory role, acting as a distance gauge and locking the phosphatase to its appropriate functional location (Fig. 2A), in proximity of the cadherin ⁄ cate- nin complex (one of RPTPl physiological substrates) [55,58]. In addition to the trans interaction, the RPTPl ectodomain can also form lateral (cis) dimers [59]. Two MAM domain loops may be involved in such interactions which, together with the trans adhesive dimers described above, may lead to the formation of 2D receptor arrays [56]. The principle of ectodomain-driven and size-con- trolled subcellular localization appears to apply to yet another RPTP, the type I CD45. The narrow spacings (about 15 nm) at local zones of cell contacts between T cells and antigen-presenting cells force the exclusion of CD45 (a protein with a large ectodomain) from the proximity of the MHC-TCR complex. This allows an increased phosphorylation of the TCR, a key step in signal transduction [60]. Current efforts are directed towards the identifica- tion of additional RPTP ligands [61], as well as struc- tural characterization of ligand–RPTP ectodomain complexes described to date. Other ectodomain-depen- dent processes such as shedding [62,63] and cis-oligo- merization [47,64] may affect catalytic activity either indirectly (via subcellular relocation) or directly (steric contacts). Further structural and functional evidence will be required to fully understand and validate such models. Structure and role of the intracellular region With few exceptions (type VIII RPTPs that are cata- lytically inactive and RPTPa, where both domains have catalytic activity) the phosphatase activity of RPTPs is restricted to the membrane-proximal (and sometimes only) domain, termed D1 [46]. Its size, fold and catalytic mechanism are essentially the same as described above for cytoplasmic PTPs. Unusual fea- tures are the KIM present in type VII RPTPs (also discussed above in the context of MAP kinase phos- phatases, since both PTPRR and PTPN5 can be expressed as receptor-like and cytoplasmic variants) and an 100 residues-long juxtamembrane region (sometimes referred as ‘cadherin-like’, although the similarity is very low) in type IIb RPTPs, a putative docking station for intracellular proteins. The D1 crystal structures of RPTPa [65] and RPTPl [66] ignited a controversy still to be settled regarding the role of an N-terminal ‘wedge’ motif (a helix-turn-helix stabilized by a two-stranded b sheet, Fig. 2B). RPTPa D1 crystallized as a dimer in three different space groups, with the wedge motif of one monomer occluding the catalytic site of a dyad-related molecule. This observation received support from func- tional studies in both RPTPa and CD45 [46], but never replicated in any of the subsequent D1 or D1 + D2 structures. Two structural genomics initia- tives, the Oxford-based Structural Genomics Consor- tium (http://www.sgc.ox.ac.uk/) and the New York Structural Genomics Research Consortium (http:// www.nysgrc.org/) have targeted intracellular RPTP domains and, to date, elucidated most of the D1 struc- tures: all have the wedge-like motif but the catalytic sites were never occluded. Four structures of full-length (D1 + D2) intracellu- lar regions have been reported to date: LAR [67], CD45 [68], RPTPr (NYSGC, PDB ID: 2fh7) and RPTPc (Oxford SGC, PDB ID: 2nlk). The LAR and CD45 structures clearly exclude the wedge-based model of RPTP inhibition because of steric constraints [67,68]. The RPTPr architecture is very similar to LAR (rmsd 1.53 A ˚ over 537 equivalent Ca atoms), while RPTPc, interestingly, reveals a considerable dif- ference in the D2 domain orientation relative to CD45 and LAR. Examination of the RPTPc crystal packing reveals a novel (putative) dimeric arrangement, sup- ported by an extensive interface (1200 A ˚ 2 per mono- mer), with the b10–b11 loop of D2 appearing to block the catalytic site of D1 in a ‘head-to-tail’ arrangement (Fig. 2B). The functional relevance of this arrangement is currently being assessed (A Barr, personal communi- cation). Importantly, the work on LAR intracellular region [67] revealed for the first time the structure of a D2 domain and the relative D1 ⁄ D2 domain arrange- ment (Fig. 2B, tandem phosphatase domains are a PTP structure–function relationship L. Tabernero et al. 872 FEBS Journal 275 (2008) 867–882 ª 2008 The Authors Journal compilation ª 2008 FEBS AB Fig. 2. The architecture and regulation of receptor protein tyrosine phosphatases. (A) Structural basis for ectodomain-controlled localization of RPTPl at cell contacts. The crystal structure of a full-length RPTPl ectodomain (rainbow coloured, PDB accession number 2V5Y) revealed a trans dimer with a rigid and extended conformation, matching in dimensions the intracellular spacings characteristic of adherens (cadherin- driven) junctions [55]. The ectodomain length and the large adhesive interface (1630 A ˚ 2 per monomer) are essential for controlling the sub- cellular localization of RPTPl, bringing it in the proximity of its physiological substrates. Plasma membranes of two opposing cells (cell 1 and 2) are shown by gray rectangles, the full-length C-cadherin ectodomain structure is in blue (Ca 2+ atoms in red, PDB accession number 1L3W). The C-terminal fibronectin type III domain of the RPTPl extracellular region is largely disordered (indicated by dotted black oval), while the juxtamembrane and transmembrane regions are schematically shown by black lines. To give an idea of relative size, crystal struc- tures of the RPTP LAR intracellular region (PDB accession number 1LAR, predicted to have a similar architecture with the corresponding region of type IIb RPTPs) and the complex between the b-catenin armadillo repeat region and an intracellular E-cadherin fragment (PDB accession number 1I7X) are shown in cyan and yellow ⁄ purple, respectively. Other components of the cadherin–catenin complex (and regions missing in the 1I7X structure) are indicated by the irregular green shape. For simplicity, these components are only shown in cell 2. (B) Crystal structures of RPTP intracellular regions and phosphatase regulation by sterical hindrance. The membrane proximal catalytic domains (D1) are shown in cyan, the distal ones (D2) in yellow. Catalytic sites in D1 and equivalent positions in D2 are marked by purple and green spheres, respectively. The ‘inhibitory wedge’ in D1 and the equivalent structure in D2 are shown in orange. Additional features in the CD45 D2 are the ‘acidic’ and ‘basic’ loops (largely disordered in the crystal structures 1YGR and 1YGU) are shown schematically in red and blue, respectively. In the RPTPa D1 crystal structure (PDB accession number 1YFO) the ‘inhibitory wedge’ blocks access to the catalytic site of a dyad-related molecule. In contrast, the D1 + D2 structures of LAR and CD45 are monomeric, casting doubt over the ‘wedge’ model. The recently solved structure of the RPTPc D1 + D2 region (PDB accession number 2NLK) reveals a novel arrangement where the D1 catalytic site is blocked by the D2 of a symmetry-related molecule. The physiological significance of this arrangement remains to be determined. L. Tabernero et al. PTP structure–function relationship FEBS Journal 275 (2008) 867–882 ª 2008 The Authors Journal compilation ª 2008 FEBS 873 unique feature of RPTPs). Both domains have the same tertiary fold, with an rmsd of 1.3 A ˚ between all equivalent Ca positions. The two domains are con- nected by a short, four-residue linker and an extensive network of interactions stabilizes the interdomain interface. The active site architecture of the two domains is very similar to each other and to other known PTPs, yet despite the conservation of all key residues in the PTP-loop, LAR D2 exhibits < 0.001% of the enzyme activity. The structure pointed towards just two residues that might have been responsible for this low activity, one in the WPD-loop and the other in the KNRY phosphotyrosine recognition loop. Site- directed mutagenesis of these residues (Gly1779Asp and Leu1644Tyr) spectacularly restored activity to D1 levels, and the authors propose that, under physiologi- cal conditions, LAR D2 may indeed be active on specific substrates [67]. However, this cannot be extrapolated to all RPTP D2 domains: in cases such as RPTPf and RPTPc, for example, the catalytic Cys residues are mutated to Asp, abrogating enzymatic activity. The CD45 intracellular region structure [68], although maintaining the overall LAR architecture, revealed further details about the function of D2. First, the CD45 D2 is, most likely, not an active phos- phatase despite preserving the catalytic Cys. The large number of substitutions in the PTP signature motif, including the essential Arg residue, results in a signifi- cantly altered shape of the active site pocket that impairs binding to the phosphoryl group [68]. More- over, mutation of the general acid ⁄ base Asp residue in the WPD-loop to Val, and of the conserved Tyr resi- due in the phosphotyrosine recognition loop to Asn, explain why the CD45 D2 cannot be an active phos- phatase under any conditions. Nevertheless, the CD45 D2 structure reveals features not present in any PTP domain: a 20 residue ‘acid’ loop between the b1 and b2 strands and an 11-residue ‘basic’ loop between the a3 helix and the b12 strand (Fig. 2B). Both loops (lar- gely disordered in the crystal structures) are located in the proximity of the D1 active site and, as such, are likely to play an important role in substrate recogni- tion and binding. Another important mechanism in the regulation of RPTP signalling is the reversible oxidation of the cata- lytic Cys residue by reactive oxygen specie) [36,46], also discussed above for cytoplasmic PTPs. Work on RPTPa, the best studied example in this respect, revealed unexpectedly that the D2 catalytic Cys (Cys723) appears to be more sensitive to oxidation than the D1 counterpart [69]. Moreover, oxidation of Cys723 results in the stabilization of observed full- length RPTPa dimers and, presumably via a confor- mational change, induces a relative rotation of the two molecules in the dimer that is detectable on the extra- cellular side of the receptor [70]. Crystallographic anal- ysis revealed important consequences of the RPTPa D2 oxidation [71]. Cys723 forms a five-atom ring structure, termed cyclic sulfenamide, with the main chain nitrogen of the adjacent Ser724, showing similar- ity to the PTP1B case described above. This oxidation is associated with conformational changes in the PTP loop, although not to the extent observed in PTP1B [37,38,71], which adopts an open conformation. It is still unclear, however, how this change can cause the reactive oxygen species-mediated stabilization of RPTPa dimers and influence the relative orientation of receptor dimers. Further work, perhaps in the context of a full intracellular region or involving crystallization of a pre-oxidized RPTPa D2 (as opposed as oxidation within the crystals as performed previously [71], where the lattice may constrain movements) will be required in order to better understand the oxidation-dependent regulation of RPTPa. RPTPs as drug targets The RPTPs are prime targets for drug design, given the importance of phosphotyrosine signalling at the plasma membrane. The steady increase in understand- ing of their biology and specific role in various diseases (see accompanying review by Hendriks et al.) [46,72] have strongly driven both structural genomics (see above) and structure-based drug design efforts. The state of structure-based development of membrane-per- meable RPTP inhibitors (a focus of the pharmaceutical industry) is difficult to gauge at the moment, given the novelty and proprietary nature of such work. Never- theless, a recent report on RPTPb revealed how engi- neering of the catalytic domain, based on its structure, has greatly helped crystallization of inhibitor com- plexes [73]. Considering the structural diversity of RPTP ectodomains and the potential of controlling the RPTP catalytic activity from the extracellular side via ligand interactions with the ectodomain, it is likely that the focus of structural efforts in the future will shift to this region. Class II PTPs (LMW-PTP) Low molecular weight PTPs are a family of small enzymes (18 kDa) involved in the regulation of cell growth, adhesion and cytoskeleton organization in mammalian cells. They share very low sequence homology to the rest of PTPs, except for the consensus PTP structure–function relationship L. Tabernero et al. 874 FEBS Journal 275 (2008) 867–882 ª 2008 The Authors Journal compilation ª 2008 FEBS active site motif CX 5 R, and a similar mechanism of catalysis where the key elements include the catalytic residues Cys12 human and Arg18 human in the P-loop and the general acid Asp129 human in the DPYY-loop (anal- ogous to the WPD-loop). Human LMW-PTP is expressed in most tissues [74] and is the predominant tyrosine phosphatase expressed in lens [75]. Two main active isoforms, A and B, have been characterized that originate by alternative splicing [76]. A third inactive isoform, C has also been reported [77,78] that lacks part of the catalytic domain. The two active isoforms differ only in the region spanning residues 40–73, also known as the ‘variable region’. Structure similarities albeit no sequence homology to classic PTPs A main difference from classic PTPs is that in LMW- PTPs, the P-loop is found at the N-terminus of the protein (residues 12–19), whereas in the rest of PTPs it is found towards the C-terminus. The overall fold of the LMW-PTP, as first revealed from X-ray crystallo- graphic structures of the bovine and human enzymes [79,80] displays a twisted central parallel b sheet with four strands and five a helices packed on both sides, reminiscent of a classic dinucleotide-binding or Ross- mann fold. This fold is similar to the one described for class I PTPs, except that it is smaller, representing the minimal size for a functional protein phosphatase enzyme. Like in other PTPs, the active site P-loop lies at the bottom of a crevice and connects the C-terminus of the b1 strand with the N-terminus of the a1 helix (Fig. 3A) Although the amino acid sequence in this loop lacks the Gly residues conserved in the canonical phosphate-binding motif, GXGXXG, it adopts a simi- lar conformation with all the backbone amide groups oriented toward the phosphate ion (Fig. 1). This is possible because Asn15 (conserved in all LMW-PTPs) adopts a left-handed helical conformation stabilized by a hydrogen bond network with three other conserved residues, Ser19, Ser43, and His72. Thus, Asn15, Ser19, His72, and Ser43 serve structural functions that allow the active site to adopt an optimal geometry for substrate binding and transition state stabilization. Supporting evidence was provided by site-directed mutagenesis and kinetic measurements [81]. Mutations at Ser19 result in an enzyme with altered kinetic prop- erties, impaired catalysis and changes in the pK a of the neighbouring His72. It was proposed that Ser19 acts to facilitate the ionization and orientation of Cys12 for optimal reaction as a nucleophile and as a leaving group. The Asn15 to Ala mutation appears to disrupt the hydrogen-bonding network, with an accompanying alteration of the geometry of the P-loop. The X-ray structure of the S19A mutant enzyme shows that the general conformation of the P-loop is preserved. How- ever, changes in the loop containing His72 result in a displacement of the His72 side chain that may explain Fig. 3. Structure of human LMW-PTP. (A) Labelled are the active site P-loop, the variable region V-loop and the DPYY-loop that con- tains the catalytic Asp residue and the phosphorylation sites Tyr131 and Tyr132 [46]. (B) Detail of the human LMW-PTPA (PDB acces- sion code 5PNT) with a molecule of MES bound into the active site. Sticks represent the catalytic residues Cys12, Arg18, Asp129. A second Cys residue in the P-loop important in redox regulation, Cys17 is also labelled. The side chains of Tyr131 and Tyr49 stack against the ring of the MES molecule stabilizing its binding in the active site. All figures were prepared using PYMOL (2003, DeLano Scientific Ltd, Palo Alto, CA, USA). L. Tabernero et al. PTP structure–function relationship FEBS Journal 275 (2008) 867–882 ª 2008 The Authors Journal compilation ª 2008 FEBS 875 the shift in the pK a and the altered kinetic behaviour of the mutant enzyme [82]. The variable region forms a long loop connecting b2 and a2 and surrounding the P-loop (Fig. 3A). Despite the sequence variation between the A and B isoforms in this region, the overall conformation is greatly conserved and contains two consecutive b turns (G. Redshaw, unpublished data). This characteristic and unique b turn tandem is also present in the struc- tures of bacterial and yeast LMW-PTPs. The first b turn in the tandem contributes to form a deep active site cleft, analogous to the structural role of the tyro- sine recognition loop (the KNRY motif) in class I PTPs. Structures of the A and B isoforms Site-directed mutagenesis, steady-state kinetics, and effector studies of the two human LMW-PTP iso- enzymes (A form, HCPTPA, and B form, HCPTPB) indicated that residues 49 and 50 play important roles in determining the specificities and activity of the two enzymes [83,84]. These two residues, Trp49 and Asn50 in the B form, are located in a loop (the variable loop or V-loop) at the outer rim of the active site (Fig. 3A). Different responses of the A and B forms of the enzyme toward activators and inhibitors have also been reported [85,86]. For example, residues 49 and 50 are involved in the strong activation of the B form by guanosine and cGMP. Mutations of Trp49 to its equivalent in the A form, a tyrosine, and of Asn50 to a glutamate, result in enzymes with kinetic properties of the A form for cGMP activation [86]. The molecu- lar basis for such differences in kinetic properties is still unclear. The differences in substrate specificity of both isoforms could be explained, in part, by the dif- ferent charge distribution around the opening of the active site, where a negative patch created by Glu50 in the A form, is substituted by positive patch with Arg53 in the B form [80]. Evidence for the importance of residues at position 50, 53 and 49 in providing sub- strate selectivity has been also reported for the rat LMW-PTP, ACP1[87]. Structures of LMW-PTP with different ligands have been reported. The B form was crystallized with phos- phate and vanadate [79,88] and showed that both inhibitory compounds bind in the active site forming hydrogen bond interactions with the amide nitrogens in the P-loop and the catalytic Arg residue. The B form bovine enzyme (BPTP) and the human A form (HCPTPA) were also crystallized in the presence of HEPES and MES, aryl sulfonate inhibitory com- pounds that fill the active site with the sulfonate group bound in the P-loop [80]. In the HCPTPA structure, a molecule of MES is bound to the active site with the sulfonate group sitting in the phosphate-binding pocket and with the aryl ring stacking between Tyr131 and Tyr49 at the opening of the active site (Fig. 3B). Similarly, in the BPTP structure, the sulfonate group of a molecule of HEPES sits in the active site forming hydrogen-bond interactions with the amide groups of the P-loop and the catalytic Arg18. The six-member ring of HEPES packs parallel to the ring of Tyr131 and perpendicular to Trp49. Trp49 ⁄ Tyr49 appears to act as a gatekeeper residue whose aromatic ring fluctu- ates between a closed position at the entrance of the active site when only small ligands (sulfate, phosphate) are present, to an open position when larger ligands like MES, HEPES are bound. Stacking interactions of Tyr131 and Trp49 ⁄ Tyr49 with the aryl ring are pre- sumably similar to those expected for a phosphotyro- sine residue from a biological substrate. The structure of the yeast LMW-PTP (LTP1) with HEPES or p-nitrophenyl phosphate (pNPP) molecules bound to the active site shows again packing of the aryl rings with an aromatic side chain Trp134, analogous to Tyr131 [89]. Aromatic residues are conserved at position 131 for all bacterial and eukaryotic LMW-PTPs, with prefer- ence for Tyr (Phe in Bacillus subtilis and Trp in Saccharomyces cerevisiae), suggesting that the stacking interactions with the substrate are important in all LMW-PTPs. On the other hand, the aromatic residues at position 49 (Trp ⁄ Tyr) found in eukarytotic LMW- PTPs, are not strictly conserved in prokaryotes. A subtype of bacterial LMW-PTPs (type II), present a large variation of residues at this position. For instance, B. subtilis YwlE has Ser42 [90] and Escheri- chia coli Wzb has Leu40 [91] instead of Tyr or Trp res- idues. These changes imply a different mechanism of substrate recognition and specificity, which could also explain why these phosphatases show a much lower binding constant towards pNPP than the eukaryotic enzymes [91]. In addition, the prokaryote type II LMW-PTPs do not have Tyr at position 132 as found in the eukaryote enzymes. This residue is critical in regulation by phosphorylation of both the enzymatic activity and the localization of the eukaryotic LMW-PTP. This suggests that the type II bacterial LMW-PTPs may have different ways of regulation. Structural determinants for regulation of catalytic activity in LMW-PTPs Oxidation of the catalytic cysteine thiol inhibits enzymatic activity of the LMW-PTP in a reversible PTP structure–function relationship L. Tabernero et al. 876 FEBS Journal 275 (2008) 867–882 ª 2008 The Authors Journal compilation ª 2008 FEBS [...]... BG, Pei Z, Hutchins CW, Ballaron SJ et al (2003) Selective protein tyrosine phosphatase 1B inhibitors: targeting the second phosphotyrosine binding site with non-carboxylic acid-containing ligands J Med Chem 46, 3437–3440 Hunter T (1989) Protein- tyrosine phosphatases: the other side of the coin Cell 58, 1013–1016 Tonks NK (2006) Protein tyrosine phosphatases: from genes, to function, to disease Nat Rev... human protein tyrosine phosphatase 1B Science 263, 1397–1404 7 Jia Z, Barford D, Flint AJ & Tonks NK (1995) Structural basis for phosphotyrosine peptide recognition by protein tyrosine phosphatase 1B Science 268, 1754– 1758 8 Andersen JN, Mortensen OH, Peters GH, Drake PG, Iversen LF, Olsen OH, Jansen PG, Andersen HS, Tonks NK & Moller NP (2001) Structural and evolutionary relationships among protein tyrosine. .. (1994) Crystal structure of ˚ Yersinia protein tyrosine phosphatase at 2.5 A and the complex with tungstate Nature 370, 571–575 10 Schubert HL, Fauman EB, Stuckey JA, Dixon JE & Saper MA (1995) A ligand-induced conformational change in the Yersinia protein tyrosine phosphatase Protein Sci 4, 1904–1913 11 Mustelin T, Tautz L & Page R (2005) Structure of the hematopoietic tyrosine phosphatase (HePTP) catalytic... (2006) Residues distant from the active site influence protein- tyrosine phosphatase 1B inhibitor binding J Biol Chem 281, 5258–5266 20 Yang J, Liu L, He D, Song X, Liang X, Zhao ZJ & Zhou GW (2003) Crystal structure of human proteintyrosine phosphatase SHP-1 J Biol Chem 278, 6516– 6520 21 Yang J, Cheng Z, Niu T, Liang X, Zhao ZJ & Zhou GW (2001) Protein tyrosine phosphatase SHP-1 specifically recognizes... regulation of protein- tyrosine phosphatases Arch Biochem Biophys 434, 11–15 37 Salmeen A, Andersen JN, Myers MP, Meng TC, Hinks JA, Tonks NK & Barford D (2003) Redox regulation of protein tyrosine phosphatase 1B involves a sulfenylamide intermediate Nature 423, 769–773 38 van Montfort RL, Congreve M, Tisi D, Carr R & Jhoti H (2003) Oxidation state of the active-site cysteine in protein tyrosine phosphatase... structures in protein- tyrosine phosphatase catalysis Proc Natl Acad Sci USA 93, 2493–2498 878 4 Alonso A, Sasin J, Bottini N, Friedberg I, Osterman A, Godzik A, Hunter T, Dixon J & Mustelin T (2004) Protein tyrosine phosphatases in the human genome Cell 117, 699–711 5 Barford D, Keller JC, Flint AJ & Tonks NK (1994) Purification and crystallization of the catalytic domain of human protein tyrosine phosphatase... Targeting the PTPome in human disease Expert Opin Ther Targets 10, 157–177 40 Bialy L & Waldmann H (2005) Inhibitors of protein tyrosine phosphatases: next-generation drugs? Angew Chem Int Ed Engl 44, 3814–3839 41 van Huijsduijnen RH, Bombrun A & Swinnen D (2002) Selecting protein tyrosine phosphatases as drug targets Drug Discov Today 7, 1013–1019 42 Elchebly M, Payette P, Michaliszyn E, Cromlish W,... substrate-targeting sites in the Yersinia protein tyrosine phosphatase co-operate to promote bacterial virulence Mol Microbiol 55, 1346–1356 28 Puius YA, Zhao Y, Sullivan M, Lawrence DS, Almo SC & Zhang ZY (1997) Identification of a second aryl phosphate-binding site in protein- tyrosine phosphatase 1B: a paradigm for inhibitor design Proc Natl Acad Sci USA 94, 13420–13425 PTP structure–function relationship 29... PTP mu, a receptor-type protein tyrosine phosphatase, can mediate cell–cell aggregation J Cell Biol 122, 961–972 Gebbink MF, Zondag GC, Wubbolts RW, Beijersbergen RL, van Etten I & Moolenaar WH (1993) Cell–cell adhesion mediated by a receptor-like protein tyrosine phosphatase J Biol Chem 268, 16101–16104 Sap J, Jiang YP, Friedlander D, Grumet M & Schlessinger J (1994) Receptor tyrosine phosphatase R-PTPkappa... Molecular analysis of receptor protein tyrosine phosphatase mu-mediated cell adhesion EMBO J 25, 701–712 Boggon TJ, Murray J, Chappuis-Flament S, Wong E, Gumbiner BM & Shapiro L (2002) C-cadherin ectodomain structure and implications for cell adhesion mechanisms Science 296, 1308–1313 Sallee JL, Wittchen ES & Burridge K (2006) Regulation of cell adhesion by protein- tyrosine phosphatases: II Cell–cell adhesion . motif; LMW-PTP, low molecular weight protein tyrosine phosphatase; N-SH2, N-terminal SH2 domain; PTP, protein tyrosine phosphatase; RPTP, receptor protein tyrosine phosphatase; YopH, Yersinia. MINIREVIEW Protein tyrosine phosphatases: structure–function relationships Lydia Tabernero 1 , A. Radu Aricescu 2 , E. Yvonne Jones 2 and. human protein tyrosine phosphatase 1B. Science 263, 1397–1404. 7 Jia Z, Barford D, Flint AJ & Tonks NK (1995) Struc- tural basis for phosphotyrosine peptide recognition by protein tyrosine

Ngày đăng: 30/03/2014, 04:20

Tài liệu cùng người dùng

Tài liệu liên quan