Acta Biochim Biophys Sin, 2015, 47(1), 29–38 doi: 10.1093/abbs/gmu107 Advance Access Publication Date: 4 December 2014 Review

Review

Structural dissection of Hippo signaling Zhubing Shi1,2, Shi Jiao1, and Zhaocai Zhou1,* 1

*Correspondence address. Tel/Fax: +86-21-54921291; E-mail: [email protected] Received 16 September 2014; Accepted 26 October 2014

Abstract The Hippo pathway controls cell number and organ size by restricting cell proliferation and promoting apoptosis, and thus is a key regulator in development and homeostasis. Dysfunction of the Hippo pathway correlates with many pathological conditions, especially cancer. Hippo signaling also plays important roles in tissue regeneration and stem cell biology. Therefore, the Hippo pathway is recognized as a crucial target for cancer therapy and regeneration medicine. To date, structures of several key components in Hippo signaling have been determined. In this review, we summarize current available structural studies of the Hippo pathway, which may help to improve our understanding of its regulatory mechanisms, as well as to facilitate further functional studies and potential therapeutic interventions. Key words: Hippo pathway, three-dimensional structure, protein–protein/DNA interaction, regulatory mechanism

Introduction Multiple signaling pathways participate in the development and tissue homeostasis, and their aberrant regulation can cause various diseases such as cancer. Among them, the Hippo pathway is an emerging regulator of tissue growth and cell fate [1–6]. The Hippo pathway regulates cell number and organ size through inhibiting cell growth and proliferation, as well as promoting cell death. Core components of the Hippo pathway were firstly identified in Drosophila by genetic screen. The activation of Hippo (Hpo), Salvador (Sav), Warts (Wts), or Mps one binder kinase activator-like 1 (Mob1; also known as Mob as tumor suppressor protein 1 or Mats) causes overgrowth of Drosophila tissues [7–16]. Hpo, Sav, Wts, and Mob1 form a kinase cascade, in which the Hpo–Sav complex phosphorylates Wts and Mob1, and then activated Wts–Mob1 complex further phosphorylates and inactivates transcriptional coactivator Yorkie (Yki) [11,16–22]. The Hippo pathway is highly conserved in mammals, and its corresponding components are mammalian sterile 20 (STE20)-like protein kinase 1/2 (MST1/2), Salvador homolog 1 (SAV1, also known as 45 kDa WW domain protein or WW45), large tumor suppressor homolog 1/2 (LATS1/2), MOB kinase activator 1A/B (MOB1A/B), and Yesassociated protein (YAP) or transcriptional coactivator with PDZ-binding motif (TAZ, also known as WW domain-containing

transcription regulator protein 1 or WWTR1) [19,23,24]. Phosphorylated Yki, YAP, or TAZ is recruited by 14-3-3 proteins and therefore sequestered in cytoplasm [23–25]. Then phosphorylated YAP/TAZ by LATS1/2 are further phosphorylated by casein kinase I isoforms δ/ɛ (CK1δ/ɛ) and degraded by proteasome dependent on F-box/WD repeat-containing protein (β-TrCP) [26,27]. In the absence of Hippo signaling, unphosphorylated Yki, YAP, or TAZ translocates to nucleus and acts as transcriptional coactivator to activate Scalloped (Sd in Drosophila)-dependent or TEA domain family member 1 to 4 (TEAD1–4 in mammal, also known as TEF-1, TEF-4, TEF-5, and TEF-3, respectively)-dependent transcription of pro-proliferative and antiapoptotic genes. Many regulators of the Hippo pathway have been identified [28– 31]. The protein components involved in the apical-basal polarity, such as neurofibromin-2, kidney and brain protein (KIBRA), crumbs homolog (CRB), angiomotin (AMOT), scribble homolog (SCRIB), and disks large homolog (DLG), regulate the Hippo pathway via modulating subcellular localization of Hippo components. Planar cell polarity can activate or suppress Hippo signaling. In Drosophila, Fat (Ft) and Dachsous (Ds) affect Wts and Yki activity. Mechanical cues sensed by extracellular matrix and cytoskeleton can influence YAP/TAZ activity [32]. Other regulators of the Hippo pathway

© The Author 2014. Published by ABBS Editorial Office in association with Oxford University Press on behalf of the Institute of Biochemistry and Cell Biology, Shanghai Institutes for Biological Sciences, Chinese Academy of Sciences. 29

Downloaded from http://abbs.oxfordjournals.org/ at UNIVERSITY OF PITTSBURGH on February 11, 2015

National Center for Protein Science Shanghai, State Key Laboratory of Cell Biology, Institute of Biochemistry and Cell Biology, Shanghai Institutes for Biological Sciences, Chinese Academy of Sciences, Shanghai 200031, China, and 2 School of Life Sciences and Technology, Tongji University, Shanghai 200092, China

30

Core Kinase Cascade MST1/2–SAV1/RASSFs complex MST1/2 kinases belong to STE20 germinal center kinase (GCK) II subfamily. They consist of an amino (N)-terminal kinase domain, an inhibitory region and a carboxy (C)-terminal SARAH (Sav/Rassf/ Hpo) domain [36,37]. SAV1 contains two WW domains and a SARAH domain in the C-terminal region [38]. RASSF family contains 10 members, namely RASSF1 to RASSF10 [39,40]. RASSF5 is also known as new Ras effector 1 (NORE1) and regulator for cell adhesion and polarization enriched in lymphoid tissues (RAPL). They share a conserved Ras association (RA) domain in their N- or C-terminal regions. A SARAH domain is present in the C-terminal regions of RASSF1–6, but not in those of RASSF7–10. The SARAH domain mediates the homo- and heterodimerization of MST1/2, SAV1, and RASSFs [19,36,37,41–43]. The kinase domains of MST1/2 adopt a canonical kinase fold, with N- and C-lobes (Fig. 1A) [44]. The solved structure of MST1 kinase domain exhibits an active conformation (Protein Data Bank [PDB] code: 3COM). Its activation loop is phosphorylated at residues Thr177 and Thr183 and adopts an extended conformation, with the phosphate group of phosphorylated Thr183 forming electrostatic interaction with Arg148 in the His-Arg-Asp (HRD) motif and Arg181. The residue Thr183 in the activation loop is a primary autophosphorylation site, and may be autophosphorylated in a similar manner observed for other GCK members, such as MST4 and oxidative stress-responsive 1 [45–47]. In contrast, the structure of MST2 kinase domain, in which residue Asp146 in the DFG (Asp-Phe-Gly) motif is mutated to asparagine, adopts an inactive conformation (PDB code: 4LG4) [44]. The activation loop of MST2 is unphosphorylated and folds into a helix. Both MST1 and MST2 kinase domains form homodimers in the crystals, which are mainly mediated by the αEF and αG helices in their C-lobes. In the inactive MST2 kinase domain, the activation loop is also involved in the dimerization. However, the relative orientations of two monomers are distinct in MST1 and MST2 homodimers, which may be caused by crystal packing. Although mutations in the dimeric interface of MST2 kinase domain do not affect its

dimerization in solution, MST2 kinase activity is indeed impaired in some cases, suggesting that the dimeric conformation is at least partially required for MST2 activation [44]. The inhibitory domain between the kinase domain and the SARAH domain of MST1/2 is an intrinsically unfolded region and structurally disordered [44,48,49]. This region contains caspase recognition sites [50]. Caspase cleavage removes the inhibitory domain of MST1/2, which leads to MST1/2 activation in apoptosis. Unexpectedly, the MST1 inhibitory domain increases the thermodynamic stability of its SARAH domain dimer and the homodimer affinity [49]. However, the inhibitory mechanism of the MST1/2 inhibitory domain is still poorly understood. The SARAH domain of MST1/2 mediates its homodimerization and heterodimeric interactions with other SARAH domain-containing proteins such as SAV1 and RASSFs [19,36,41–43]. The SARAH domain-mediated homodimerization of MST1/2 is required for their full activation [51]. The monomeric state of MST1/2 SARAH domain is thermodynamically unstable and its dimerization has been coupled with structural folding [49,52]. The structures of MST1/2 and RASSF5 SARAH homodimers or heterodimers have been determined (Fig. 1B) [44,48,52–54]. The SARAH domain of MST1/2 or RASSF5 alone forms a symmetric homodimer (PDB codes: 2JO8, 2YMY, 4HKD, 4L0N, 4NR2, and 4OH9). Each monomer of MST1/2 SARAH homodimer contains two helices, H1 and H2, while that of RASSF5 has only one helix, corresponding to the H2 helix of MST1/2 SARAH. In one type of heterodimeric interactions, the RASSF5 SARAH and the H2 helix of MST1/2 SARAH fold into antiparallel coiled coil. In general, most coiled coils are structurally organized as multiple ‘heptad repeat (a–g positions)’. However, the heptad register is interrupted by two stutters in the MST1/2 and RASSF5 SARAH domains, and thus the coiled coils have two transitions in periodicity from heptad (a–g) to tetrad (d–g), resulting in a low degree of supercoiling. Each H1 helix of MST1/2 SARAH contacts H2 helices from the two monomers. Hydrophobic interactions dominate the SARAH domain homo- or heterodimerization, while electrostatic interactions and hydrogen bonds further stabilize dimeric interfaces. Although the overall structures of these SARAH domains are similar, their local conformations vary, even in different structures of the same SARAH domain, implying a dynamic nature of the SARAH domain. Sequence conservation indicates that these structural features exist in all SARAH domains [52,54]. SAV1 binding enhances MST1/2 kinase activity, while the influence of RASSFs on MST1/2 activity is uncertain [41–43,51,55,56]. RASSF5 suppresses MST1 kinase activity in vitro, whereas it appears to promote MST1 activation in mouse pre-B cell line BAF cells [42,43,51]. The SARAH domain of MST1/2 forms heterodimer with that of RASSF5 in a manner similar to their homodimerization (Fig. 1B) (PDB codes: 4LGD and 4OH8) [44,53,57]. The interface of MST1/2–RASSF5 heterodimer also resembles those of their homodimers, but the heterodimers have more intermolecular hydrogen bonds and more extensive hydrophobic interactions than the homodimers, which makes them more stable than the homodimers, as supported by results from urea-induced denaturation experiment [53]. Therefore, SARAH domains function as a universal dimerization module to form homodimer or heterodimer with a preference for heterodimer. The SAV1 SARAH domain is expected to adopt a similar structural fold when forming homodimer or heterodimer with MST1/2. In this regard, the differential effects of SAV1 and RASSFs on MST1/2 activation may be caused by recruiting distinct partners and are dependent on different cellular contexts. Besides the SARAH domain, all RASSFs have a RA domain, which binds to Ras protein in a GTP-dependent manner [58,59]. In the

Downloaded from http://abbs.oxfordjournals.org/ at UNIVERSITY OF PITTSBURGH on February 11, 2015

include Ras association domain-containing proteins (RASSFs), G protein-coupled receptor (GPCR), thousand and one amino acid protein kinase 1 (TAO1), microtubule-associated protein/microtubule affinity-regulating kinase 1 (MARK1, also known as PAR1), saltinducible kinase 1 (SIK1), ankyrin repeat and KH domain-containing protein 1 (MASK), serine/threonine protein phosphatase 2A (PP2A). Some regulators of the Hippo pathway are also important components of other signaling pathways, indicating extensive crosstalk between Hippo and these pathways [33]. For example, several components of Wnt signaling, including Dishevelled (DVL) and β-catenin, are associated with YAP/TAZ to regulate their localization, degradation, and activity [34]. YAP/TAZ cooperate with SMAD family members to maintain stem cell pluripotency or induces specific stem cell differentiation [35]. During the past decade, growing biochemical and structural studies were carried out to better understand the nature of the Hippo signaling transduction. A prominent example is that several important modular protein domains, especially the WW domain and its Pro-Pro-Xaa-Tyr (PPxY) motif, are revealed to widely mediate protein–protein interactions in the Hippo pathway. Here, we summarize the structural studies of the Hippo core kinase complexes and TEAD transcriptional complexes and discuss the regulatory mechanism derived from their structural aspect.

Structural dissection of Hippo signaling

Structural dissection of Hippo signaling

31

structure of RASSF5 RA domain in complex with Ras, Ras binds a GTP analog guanyl-5′-yl imidodiphosphate (GppNHp) and a magnesium ion (PDB code: 3DDC) [60]. The RASSF5 RA domain has a ubiquitin fold similar to Ras binding domains (RBDs) of other Ras effectors, but contains additional N-terminal region and an insertion between β1 and β2. The switch I region of Ras forms an intermolecular β-sheet with the strand β2 of RASSF5 RA domain. Unlike other Ras effectors, the N-terminus of RASSF5 RA domain contributes to extra contact with the switch II region of Ras via hydrophobic interactions. RASSF1 and RASSF5 have a C1 domain in the N-terminal part, which is a phorbol-ester/diacylglycerol (DAG)-type zinc finger. The C1 domain of RASSF5 consists of a distorted five-stranded, antiparallel β-sheet with a helix (PDB codes: 1RFH and 2FNF) [61]. Two zinc ions exist in the RASSF5 C1 domain, and each zinc ion is coordinated by three cysteines and one histidine to stabilize the C1 structure. The RASSF5 C1 domain can form an intramolecular interaction with the

RA domain, and Ras-GTP binding to the RA domain disrupts this interaction and displaces the C1 domain. The free C1 domain may bind phosphatidylinositol 3-phosphate to localize toward membrane. SAV1 has two WW domains, WW1 and WW2, which may bind to LATS1/2 [15]. The WW domain was named so as it contains two conserved tryptophan residues [62]. Like typical WW domain, the SAV1 WW1 domain is a monomer consisting of a triple-stranded antiparallel β-sheet (PDB code: 2YSB) (Fig. 2A). The WW1 domain of SAV1 may associate with PPxY motifs to mediate its interaction with other proteins (see the section ‘LATS–YAP/TAZ complex’), but the specific partner has yet to be identified. The second conserved tryptophan residue important for partner binding is substituted by a tyrosine in the SAV1 WW2 domain and this tyrosine residue is evolutionarily conserved [63]. Although the SAV1 WW2 domain has a WW-like fold, unlike the canonical monomeric WW domain, it forms a symmetric β-clam-like homodimer (PDB code: 2DWV) [63]. The molecular surface corresponding to that of the canonical WW domain used for

Downloaded from http://abbs.oxfordjournals.org/ at UNIVERSITY OF PITTSBURGH on February 11, 2015

Figure 1. Structures of MST1/2 kinases and their complexes with RASSF5 (A) Structures of MST1/2 kinase domains. MST1/2 kinase domains form homodimers. (B) Structures of MST1/2 and RASSF5 SARAH homodimers and heterodimers. Following structures in PDB were used: 3COM (MST1_KD), 4LG4 (MST2_KD), 4NR2 (MST1_SARAH), 4HKD (MST2_SARAH), 2YMY (RASSF5_SARAH), 4OH8 (MST1-RASSF5_SARAH) and 4LGD (MST2-RASSF5_SARAH). Structure images were generated by PyMOL (www.pymol.org). KD represents kinase domain.

32

Structural dissection of Hippo signaling

Figure 3. Structures of MOB1 (A) Structures of human MOB1 and yeast Mob1p (PDB codes: 1PI1 and 2HJN). (B) Structure of yeast Mob2p in complex with Cbk1p (PDB code: 4LQP). (C) Structure of human MOB1 in complex with a phosphor-peptide (PDB code: 4JIZ).

binding partners mediates homodimerization in the WW2 domain, which results from the presence of hydrophobic residues, such as Val244, Ala261, and Phe249 and a residue deletion between

Gly250 and Thr251 in the WW2 domain. The function of the SAV1 WW domains, particularly the dimeric WW2 domain, remains largely unknown.

Downloaded from http://abbs.oxfordjournals.org/ at UNIVERSITY OF PITTSBURGH on February 11, 2015

Figure 2. Structures of LATS1 and YAP WW domains (A) Structures of LATS1 WW domains. The LATS1 WW1 domain is a monomer, while its WW2 domain is a dimer. (B) Structures of YAP WW domain in complex with its binding partners. Following structures in PDB were used: 2YSB (SAV1_WW1), 2DWV (SAV1_WW2), 1JMQ (YAP_WW1-WBP1), 2LAX (YAP_WW1-SMAD1), 2LAY (YAP_WW1-SMAD1), 2LTW (YAP_WW1-SMAD7), 2L4J (YAP_WW2), 2LAW (YAP_WW2-SMAD1), and 2LTV (YAP_WW2-SMAD7).

33

Structural dissection of Hippo signaling

LATS1/2–MOB1 complex

pombe [70]. In MEN, Cdc15p kinase first phosphorylates the scaffold protein Nud1p and then the phosphorylated Nud1p binds Mob1p to recruit the Dbf2p–Mob1p complex to spindle pole body [71]. Cdc15p further phosphorylates and activates the Dbf2p–Mob1p complex. The structure of human MOB1 in complex with an optimized Nud1p-like phosphor-peptide TVARIYHpSVVRYAPS (pS, phosphor-serine) was recently solved (PDB code: 4JIZ) (Fig. 3C) [71]. This peptide forms an intermolecular β-sheet with the β2 strand of MOB1, where the strand S-1 exists in the Mob1p structure. The phenolic side chain of residue Tyr-2 from the peptide packs against the side chain of Lys84 in the H3/H4 loop of MOB1. Residues at the +1 to +3 positions, which prefer hydrophobic residues in peptide library screening, form hydrophobic interaction with MOB1. The phosphorylation-dependent MOB1 binding is explained by extensive electrostatic interactions between phosphoserine of the peptide and three basic residues Lys132, Arg133, and Arg136 from MOB1. Although the binding site of phosphor-peptide in MOB1 is conserved in yeast and human, the Nud1p homologue protein has not been identified in fly or mammal. Thus, whether conserved proteins or even similar phosphor-peptides exist in higher organism warrants further investigation.

Linker between Upstream and Downstream of Signaling LATS–YAP/TAZ complex YAP/TAZ have an N-terminal TEAD-binding domain (TBD), a transactivation domain, and a PDZ-binding motif at their C termini. Besides, the longest isoform of YAP contains two WW domains, WW1 and WW2, while TAZ has only one WW domain in human. YAP/TAZ interact with signaling proteins, such as LATS1/2, SMAD1/7, Runt-related transcription factor 2 (RUNX2), DVL, AMOT, and P73, via the WW domains to regulate multiple signaling pathways [4,72–74]. LATS1 and LATS2 have two and one PPxY motif, respectively. The second PPxY motif in LATS1 is conserved in LATS2; and this PPxY motif interacts with YAP [75]. Although the LATS–YAP/TAZ complex has no structural information, several structures of YAP WW1 or WW2 domain alone or in complex with other partners have been solved (Fig. 2B) [76–80]. Like SAV1 WW1 domain, each WW domain of YAP consists of a three-stranded antiparallel β-sheet. The linker between two WW domains of YAP adopts a helix–loop–helix fold [78]. Based on sequence similarity, the structure of TAZ WW domain should resemble that of YAP WW1 domain. The WW1 and WW2 domains of YAP have different preference for binding motifs, which is caused by variant residues in the loop 1 between the first two β strands [78,81]. The WW2 domain binds to PPxY motifs, while the WW1 domain can bind both PPxY motifs and phosphor-serine-proline peptides [76,78]. Residues around PPxY motifs also affect their selective binding with WW domain. In the YAP WW1 and WW domain-binding protein 1 (WBP-1) complex (PDB code: 1JMQ), two prolines and one tyrosine in the WBP-1 PPxY motif contact the hydrophobic cavity in YAP WW1 domain [76,77]. The N- and C-terminal parts of SMAD7 PPxY motif extend its interface with YAP via polar interactions (PDB code: 2LTW) [81]. Compared with the WW1 domain, the YAP WW2 domain has lower affinity for SMAD7 PPxY motif since its loop 1 repels the N- and C-terminal extensions of SMAD7 PPxY motif (PDB code: 2LTV). YAP associates with SMAD1 via its tandem WW domains [78]. The YAP WW1 and WW2 domains recognize the phospho-serine-proline motif and the PPxY motif of SMAD1, respectively (PDB codes: 2LAW, 2LAX, and 2LAY). Besides the hydrophobic

Downloaded from http://abbs.oxfordjournals.org/ at UNIVERSITY OF PITTSBURGH on February 11, 2015

In the Hippo pathway, MST1/2 kinases phosphorylate MOB1 and LATS1/2 to activate LATS1/2 kinases and enhance their affinities with MOB1 [18,21]. Phosphorylated MOB1 binding further promotes the LATS1/2 autophosphorylation and kinase activity, leading to full activation of LATS1/2. LATS1/2 are members of protein kinase A/G/C-like (AGC) kinases nuclear dbf2-related (NDR) family. LATS1/ 2 contain a UBA domain, an S100B and MOB association (SMA) domain, a kinase domain and an AGC kinase C-terminal domain. Although the structure of LATS1/2 has not been determined, the general mechanism of AGC kinase activation has been studied [64]. AGC kinase activation requires phosphorylation of two conserved motifs, the activation loop in the kinase domain and the hydrophobic motif in the AGC kinase C-terminal domain. The two phosphorylation sites of LATS1 are Ser909 and Thr1079, which are phosphorylated by MST1/2 [19]. The phosphorylated hydrophobic motif binds to the N-lobe of AGC kinases to stabilize the αC helix conformation. Phosphorylation of the activation loop alters conformation of the catalytic center. Phosphorylation of both the activation loop and the hydrophobic motif is required for positioning the αC helix in an active conformation and further activation of AGC kinase [64]. Unlike other AGC kinases, LATS1/2, as well as NDR1/2, have an SMA domain preceding the kinase domain and an autoinhibitory sequence (AIS) in the kinase domain. MOB1 binds to the SMA domain of LATS1/2 and may relieve autoinhibition of the AIS [65,66]. MOB1 is highly conserved in eukaryotes from yeast to mammal, which is composed of a left-handed four-helix bundle and several flexible loops (PDB codes: 1PI1, 1R3B, and 2HJN) (Fig. 3A) [67–69]. Four conserved residues, two cysteines (Cys79 and Cys84 in human MOB1A) and two histidines (His161 and His166 in human MOB1A) coordinate a zinc atom to stabilize the structure of MOB1. Saccharomyces cerevisiae MOB1 (Mob1p) forms a dimer in the crystal (PDB: code 2HJN), which was confirmed by static light-scattering analysis [69]. Three structural elements, short strand S-1, extended strand S0, and helix H0, in the N-terminal region of Mob1p were not observed in the solved structures of human and Xenopus laevis MOB1 due to different fragments used for structure determination. The strand S0 and helix H0, but not strand S-1, are conserved in MOB1. The strand S0 associates with the conserved surface composed of helices H2–H4 and the H4/H5 loop. The helix H0 mediates the homodimerization and is important for Mob1p function in budding yeast. Whether the MOB1 associates as a functional dimer in other species, especially in human, remains to be addressed. Nuclear magnetic resonance (NMR) titration showed that the X. laevis NDR1 peptide (amino acids 68–86, AHARKETEFLRLKRTRLGL), as part of the SMA domain, leads to specific chemical shift perturbations in H3 and H4 helices and H4/H5 loop of MOB1, suggesting this region is involved in the NDR–MOB1 interaction [68]. Several studies imply that the positively charged basic and hydrophobic residues in the SMA domain of NDR kinases contact a negatively charged surface on MOBs [65–68]. Recently, three structures of S. cerevisiae MOB2 (Mob2p) in complex with NDR family kinase CBK1 (Cbk1p) were deposited in PDB (PDB codes: 4LQP, 4LQQ and 4LQS) (Fig. 3B). The SMA domain preceding the Cbk1p kinase domain interacts with helices H1, H2, and H7 of Mob2p through extensive electrostatic and hydrophobic interactions. The Cbk1p–Mob2p interaction might be stabilized by phosphorylation of the C-terminal hydrophobic motif of Cbk1p. LATS1/2 and MOB1 may interact with each other in a similar fashion. The Hippo pathway is known as the mitotic exit network (MEN) in S. cerevisiae and the septation initiation network in Schizosaccharomyces

34 interaction between SMAD1 and YAP WW1 domain, the phosphate group of residue Ser206 phosphorylated by cyclin-dependent kinase 8/ 9 (CDK8/9) in SMAD1 forms a hydrogen bond with the YAP WW1 domain. Further phosphorylation of SMAD1 at Thr202 by glycogen synthase kinase-3 (GSK3) destabilizes the interaction of SMAD1 with YAP. SMAD1 binds to YAP WW2 domain mainly via core residues and C-terminal extension of its PPxY motif. Several YAP-binding proteins have two or more PPxY motifs, which may enhance their affinities with tandem WW domains of YAP [73,79].

YAP/TAZ–14-3-3 complex

TEAD Transcription Complex YAP/TAZ–TEAD complex All TEAD1–4 are composed of an N-terminal TEA/ATTS (AbaA, TEC1p, TEF-1 sequence) domain and a C-terminal YAP-binding domain (YBD) [83,84]. YAP/TAZ form a complex with TEAD YBD via their TBDs to transactivate downstream genes expression, such as CTGF and Cyr61 [4,85]. Structuralwise, TEAD YBD adopts an immunoglobulin-like β-sandwich fold with four short helices (PDB codes: 3JUA, 3KYS, and 3L15) (Fig. 4A) [86–88]. YAP TBD consists of one β strand, one α helix, and one twisted coil [87]. YAP TBD interacts with TEAD YBD via three interfaces [86,87]. The interface 1 constitutes the β strand of YAP TBD and the β7 strand of TEAD YBD, forming an antiparallel β-sheet. Α helix of YAP packs into a hydrophobic groove in TEAD YBD on interface 2. On the interface 3, the twisted coil of YAP TBD embeds into a hydrophobic pocket in TEAD YBD. Mutation studies revealed that the interface 3 is critical for YAP– TEAD complex formation.

VGLLs–TEAD complex Drosophila vestigial (Vg) directly interacts with Sd to activate wingspecific gene expression during Drosophila wing development [89– 92]. Mammalian vestigial-like protein 1 to 4 (VGLL1–4) can act as transcriptional coactivator or corepressor [93–101]. VGLL1 or VGLL2 activates TEAD, myocyte-specific enhancer factor 2 (MEF2) or myoblast determination protein 1 (MYOD1)-dependent transcription, while VGLL4 suppresses TEAD-mediated transcription. VGLL3 is amplified and over-expressed in soft tissue sarcomas, and in contrast, it may act as a tumor suppressor in epithelial ovarian cancer [102,103]. VGLL1–4 proteins bind YBDs of TEADs via their Tondu (TDU) domains [93,94,96,104]. VGLL1–3 proteins have one TDU domain, while VGLL4 has two TDU domains, namely TDU1 and TDU2, respectively. VGLL4 can bind two molecules of TEAD via its tandem TDU domains [96]. The TDUs of VGLL1 and VGLL4 interact with TEAD in a similar manner to YAP (PDB codes: 4EAZ

and 4LN0) (Fig. 4B–D) [96,105]. Unlike YAP, the interface corresponding to interface 3 of the YAP–TEAD complex is lacking in VGLL1–TEAD and VGLL4–TEAD complexes (Fig. 4E). YAP contains a conserved LxxLF (Leu-Xaa-Xaa-Leu-Phe) motif in interface 2, while VGLLs have the VxxHF (Val-Xaa-Xaa-His-Phe) motif. The VGLL1 TDU domain possesses interfaces 1 and 2 [105]. The VGLL4 TDU1 domain has only interface 2, while the TDU2 domain has interfaces 1 and 2 [96]. There is only one helix in the interface 2 of YAP TBD and VGLL1 TDU domain, as well as in the TDU1 domain of VGLL4, while an extra helix is present in interface 2 of VGLL4 TDU2 domain, strengthening its interaction with TEAD. Collectively, interface 2 of VGLL1 or VGLL4 TDU domains appears to play an essential role in their interactions with TEAD.

TEAD–DNA complex The TEA domain of TEAD is a DNA-binding domain that is highly conserved from yeast to mammal [83,84]. The TEA domain binds specific DNA elements such as M-CAT (muscle-CAT, 5′-CATTCCT-3′) in the promoter regions of TEAD target genes and 5′-TGGAATGT-3′ in the simian virus 40 (SV40) enhancer [106–108]. The TEAD1 TEA domain is a helix-turn-helix (HTH) fold and belongs to the homeodomain structural family (PDB code: 2HZD) [109]. The TEA domain is composed of three α helices. NMR titration assay identified the H3 helix and L2 loop immediately preceding H3 in TEAD1 TEA domain as DNA-binding interfaces. The structure of TEAD in complex with DNA has not been experimentally determined. According to DALI search, the structure of TEA domain is similar to those of paired box 6 (PAX6) paired domain (PDB code: 6PAX), telomeric repeatbinding factor 1 (TERF1) HTH myb-type domain (PDB code: 1W0 T) and homeobox-containing protein 1 (HMBOX1) homeobox domain (PDB code: 4J19), all of which bind DNA in each structure [110–113]. Thus, it is very likely that TEAD TEA domain may interact with DNA in a similar manner.

Perspective The Hippo signaling involves multiple protein–protein and protein– DNA interactions, which are mediated by several shared modular protein domains. Examples representing modularity of the Hippo pathway include the WW domain and its partner PPxY motif [72–74]. In the Hippo pathway, many proteins such as SAV1, YAP, TAZ, and KIBRA possess one or more WW domains, while one or more PPxY or PPxF (Pro-Pro-Xaa-Phe) motifs were found in other components, including MST1/2, LATS1/2, CRB1-3, AMOT, angiomotin-like protein 1/2 (AMOTL1/2), fat homolog 1-4 (FAT1-4), and dachsous homolog 1/2 (DCHS1/2). The PDZ domain is another modular domain participating in Hippo signaling [72]. PALS1, Pals1-associated tight junction protein, SCRIB, and DLG have PDZ domain, while YAP, TAZ, KIBRA, CRB1-3, AMOT, and AMOTL1/ 2 contain PDZ-binding motif. These proteins form a complicated network to regulate Hippo signaling. Components of the Hippo core kinase cascade are tumor suppressors, while YAP/TAZ are oncoproteins. Germline or somatic mutations of the Hippo pathway genes are rare in human cancers. However, expression levels and nuclear localization of YAP are elevated in various tumors, including liver, lung, breast, colon, ovarian, and gastric cancers [29,114]. YAP/TAZ also regulate stem and progenitor cell self-renewal and expansion and function in tissue repair and regeneration [4,29]. Therefore, targeting YAP/TAZ may provide novel strategies for anticancer therapy, whereas suppressing the Hippo

Downloaded from http://abbs.oxfordjournals.org/ at UNIVERSITY OF PITTSBURGH on February 11, 2015

LATS1/2 and Wts phosphorylate YAP/TAZ at Ser127/Ser89 and Yki at Ser168, respectively, to promote YAP/TAZ and Yki associations with 14-3-3 proteins, leading to their cytoplasmic retention [17,23–25]. The crystal structure of 14-3-3σ in complex with YAP peptide phosphorylated at Ser127 (124-RAHpSSPASLQ-133) has been reported (PDB code: 3MHR) [82]. The 14-3-3σ exists as a dimer with a typical W-shape. The YAP peptide binds to the amphipathic groove of 14-3-3σ created by helices 3, 5, 7, and 9. The phosphorylated Ser127 of the YAP peptide is coordinated by residues Lys49, Arg56, Arg129, and Tyr130 of 14-3-3σ. The orientation of the YAP peptide in the 14-3-3 cavity indicates a comparable mode II 14-3-3-binding motif, although the N-terminus of the peptide fits more closely to a mode I binding motif from a sequence point of view.

Structural dissection of Hippo signaling

35

Structural dissection of Hippo signaling

core kinases may promote tissue regeneration. Structure determination of Hippo pathway components and related complexes can provide atomic insights into its regulatory mechanism. Based on detailed three-dimensional structures, one can design specific and targetable compounds including small molecules and peptides to inhibit kinases or YAP/TAZ activity for therapeutic purposes like tissue regeneration and cancer treatment. A representative example is targeting the YAP– TEAD complex by a small molecule verteporfin and a peptide SuperTDU to suppress YAP-driven liver overgrowth and gastric cancer in mice, respectively [96,115]. Although a number of new discoveries on the Hippo pathway have been reported, many crucial questions await answers before the regulatory mechanism of Hippo pathway can be fully elucidated. Recently, several groups identified Hippo interactome by affinity purification and mass spectrometry (AP-MS) techniques, which may facilitate further understanding of the Hippo pathway [116–119].

2.

3.

4.

5. 6. 7.

8.

9.

Funding This work was supported by the grants from the 973 program of the Ministry of Science and Technology of China (No. 2012CB910204), the National Natural Science Foundation of China (Nos. 31270808, 31300734, 31470736, and 31470868), the Science and Technology Commission of Shanghai Municipality (Nos. 11JC14140000 and 13ZR1446400), and the ‘Cross and Cooperation in Science and Technology Innovation Team’ Project of the Chinese Academy of Sciences.

References 1.

Pan D. Hippo signaling in organ size control. Genes Dev 2007, 21: 886–897.

10.

11.

12.

13.

Zhao B, Lei QY, Guan KL. The Hippo-YAP pathway: new connections between regulation of organ size and cancer. Curr Opin Cell Biol 2008, 20: 638–646. Zhao B, Li L, Lei Q, Guan KL. The Hippo-YAP pathway in organ size control and tumorigenesis: an updated version. Genes Dev 2010, 24: 862–874. Zhao B, Tumaneng K, Guan KL. The Hippo pathway in organ size control, tissue regeneration and stem cell self-renewal. Nat Cell Biol 2011, 13: 877–883. Saucedo LJ, Edgar BA. Filling out the Hippo pathway. Nat Rev Mol Cell Biol 2007, 8: 613–621. Zhang L, Yue T, Jiang J. Hippo signaling pathway and organ size control. Fly (Austin) 2009, 3: 68–73. Harvey KF, Pfleger CM, Hariharan IK. The Drosophila Mst ortholog, hippo, restricts growth and cell proliferation and promotes apoptosis. Cell 2003, 114: 457–467. Jia J, Zhang W, Wang B, Trinko R, Jiang J. The Drosophila Ste20 family kinase dMST functions as a tumor suppressor by restricting cell proliferation and promoting apoptosis. Genes Dev 2003, 17: 2514–2519. Pantalacci S, Tapon N, Leopold P. The Salvador partner Hippo promotes apoptosis and cell-cycle exit in Drosophila. Nat Cell Biol 2003, 5: 921–927. Udan RS, Kango-Singh M, Nolo R, Tao C, Halder G. Hippo promotes proliferation arrest and apoptosis in the Salvador/Warts pathway. Nat Cell Biol 2003, 5: 914–920. Lai ZC, Wei X, Shimizu T, Ramos E, Rohrbaugh M, Nikolaidis N, Ho LL, et al. Control of cell proliferation and apoptosis by mob as tumor suppressor, mats. Cell 2005, 120: 675–685. Justice RW, Zilian O, Woods DF, Noll M, Bryant PJ. The Drosophila tumor suppressor gene warts encodes a homolog of human myotonic dystrophy kinase and is required for the control of cell shape and proliferation. Genes Dev 1995, 9: 534–546. Xu T, Wang W, Zhang S, Stewart RA, Yu W. Identifying tumor suppressors in genetic mosaics: the Drosophila Lats gene encodes a putative protein kinase. Development 1995, 121: 1053–1063.

Downloaded from http://abbs.oxfordjournals.org/ at UNIVERSITY OF PITTSBURGH on February 11, 2015

Figure 4. Structures of TEAD–YAP/VGLLs complexes (A–D) Structures of TEAD in complex with YAP, VGLL1, and VGLL4. (E) Structural comparison of TEAD–YAP/ VGLLs complexes. Following structures in PDB were used: 3KYS (TEAD1–YAP), 4EAZ (TEAD4–VGLL1), and 4LN0 (TEAD4–VGLL4).

36

37. Scheel H, Hofmann K. A novel interaction motif, SARAH, connects three classes of tumor suppressor. Curr Biol 2003, 13: R899–R900. 38. Valverde P. Cloning, expression, and mapping of hWW45, a novel human WW domain-containing gene. Biochem Biophys Res Commun 2000, 276: 990–998. 39. Volodko N, Gordon M, Salla M, Ghazaleh HA, Baksh S. RASSF tumor suppressor gene family: biological functions and regulation. FEBS Lett 2014, 588: 2671–2684. 40. Richter AM, Pfeifer GP, Dammann RH. The RASSF proteins in cancer; from epigenetic silencing to functional characterization. Biochim Biophys Acta 2009, 1796: 114–128. 41. Oh HJ, Lee KK, Song SJ, Jin MS, Song MS, Lee JH, Im CR, et al. Role of the tumor suppressor RASSF1A in Mst1-mediated apoptosis. Cancer Res 2006, 66: 2562–2569. 42. Khokhlatchev A, Rabizadeh S, Xavier R, Nedwidek M, Chen T, Zhang XF, Seed B, et al. Identification of a novel Ras-regulated proapoptotic pathway. Curr Biol 2002, 12: 253–265. 43. Katagiri K, Imamura M, Kinashi T. Spatiotemporal regulation of the kinase Mst1 by binding protein RAPL is critical for lymphocyte polarity and adhesion. Nat Immunol 2006, 7: 919–928. 44. Ni L, Li S, Yu J, Min J, Brautigam CA, Tomchick DR, Pan D, et al. Structural basis for autoactivation of human Mst2 kinase and its regulation by RASSF5. Structure 2013, 21: 1757–1768. 45. Record CJ, Chaikuad A, Rellos P, Das S, Pike AC, Fedorov O, Marsden BD, et al. Structural comparison of human mammalian ste20like kinases. PLoS One 2010, 5: e11905. 46. Shi Z, Jiao S, Zhang Z, Ma M, Zhang Z, Chen C, Wang K, et al. Structure of the MST4 in complex with MO25 provides insights into its activation mechanism. Structure 2013, 21: 449–461. 47. Lee SJ, Cobb MH, Goldsmith EJ. Crystal structure of domain-swapped STE20 OSR1 kinase domain. Protein Sci 2009, 18: 304–313. 48. Hwang E, Ryu KS, Paakkonen K, Guntert P, Cheong HK, Lim DS, Lee JO, et al. Structural insight into dimeric interaction of the SARAH domains from Mst1 and RASSF family proteins in the apoptosis pathway. Proc Natl Acad Sci USA 2007, 104: 9236–9241. 49. Constantinescu Aruxandei D, Makbul C, Koturenkiene A, Ludemann MB, Herrmann C. Dimerization-induced folding of MST1 SARAH and the influence of the intrinsically unstructured inhibitory domain: low thermodynamic stability of monomer. Biochemistry 2011, 50: 10990–11000. 50. de Souza PM, Lindsay MA. Mammalian Sterile20-like kinase 1 and the regulation of apoptosis. Biochem Soc Trans 2004, 32: 485–488. 51. Praskova M, Khoklatchev A, Ortiz-Vega S, Avruch J. Regulation of the MST1 kinase by autophosphorylation, by the growth inhibitory proteins, RASSF1 and NORE1, and by Ras. Biochem J 2004, 381: 453–462. 52. Makbul C, Constantinescu Aruxandei D, Hofmann E, Schwarz D, Wolf E, Herrmann C. Structural and thermodynamic characterization of Nore1SARAH: a small, helical module important in signal transduction networks. Biochemistry 2013, 52: 1045–1054. 53. Hwang E, Cheong HK, Mushtaq AU, Kim HY, Yeo KJ, Kim E, Lee WC, et al. Structural basis of the heterodimerization of the MST and RASSF SARAH domains in the Hippo signalling pathway. Acta Crystallogr D Biol Crystallogr 2014, 70: 1944–1953. 54. Liu G, Shi Z, Jiao S, Zhang Z, Wang W, Chen C, Hao Q, et al. Structure of MST2 SARAH domain provides insights into its interaction with RAPL. J Struct Biol 2014, 185: 366–374. 55. Matallanas D, Romano D, Yee K, Meissl K, Kucerova L, Piazzolla D, Baccarini M, et al. RASSF1A elicits apoptosis through an MST2 pathway directing proapoptotic transcription by the p73 tumor suppressor protein. Mol Cell 2007, 27: 962–975. 56. Guo C, Tommasi S, Liu L, Yee JK, Dammann R, Pfeifer GP. RASSF1A is part of a complex similar to the Drosophila Hippo/Salvador/Lats tumorsuppressor network. Curr Biol 2007, 17: 700–705. 57. Callus BA, Verhagen AM, Vaux DL. Association of mammalian sterile twenty kinases, Mst1 and Mst2, with hSalvador via C-terminal coiled-coil domains, leads to its stabilization and phosphorylation. FEBS J 2006, 273: 4264–4276.

Downloaded from http://abbs.oxfordjournals.org/ at UNIVERSITY OF PITTSBURGH on February 11, 2015

14. Kango-Singh M, Nolo R, Tao C, Verstreken P, Hiesinger PR, Bellen HJ, Halder G. Shar-pei mediates cell proliferation arrest during imaginal disc growth in Drosophila. Development 2002, 129: 5719–5730. 15. Tapon N, Harvey KF, Bell DW, Wahrer DC, Schiripo TA, Haber D, Hariharan IK. Salvador promotes both cell cycle exit and apoptosis in Drosophila and is mutated in human cancer cell lines. Cell 2002, 110: 467–478. 16. Wu S, Huang J, Dong J, Pan D. Hippo encodes a Ste-20 family protein kinase that restricts cell proliferation and promotes apoptosis in conjunction with salvador and warts. Cell 2003, 114: 445–456. 17. Huang J, Wu S, Barrera J, Matthews K, Pan D. The Hippo signaling pathway coordinately regulates cell proliferation and apoptosis by inactivating Yorkie, the Drosophila homolog of YAP. Cell 2005, 122: 421–434. 18. Wei X, Shimizu T, Lai ZC. Mob as tumor suppressor is activated by Hippo kinase for growth inhibition in Drosophila. EMBO J 2007, 26: 1772–1781. 19. Chan EH, Nousiainen M, Chalamalasetty RB, Schafer A, Nigg EA, Sillje HH. The Ste20-like kinase Mst2 activates the human large tumor suppressor kinase Lats1. Oncogene 2005, 24: 2076–2086. 20. Hirabayashi S, Nakagawa K, Sumita K, Hidaka S, Kawai T, Ikeda M, Kawata A, et al. Threonine 74 of MOB1 is a putative key phosphorylation site by MST2 to form the scaffold to activate nuclear Dbf2-related kinase 1. Oncogene 2008, 27: 4281–4292. 21. Praskova M, Xia F, Avruch J. MOBKL1A/MOBKL1B phosphorylation by MST1 and MST2 inhibits cell proliferation. Curr Biol 2008, 18: 311–321. 22. Zhang L, Ren F, Zhang Q, Chen Y, Wang B, Jiang J. The TEAD/TEF family of transcription factor Scalloped mediates Hippo signaling in organ size control. Dev Cell 2008, 14: 377–387. 23. Dong J, Feldmann G, Huang J, Wu S, Zhang N, Comerford SA, Gayyed MF, et al. Elucidation of a universal size-control mechanism in Drosophila and mammals. Cell 2007, 130: 1120–1133. 24. Zhao B, Wei X, Li W, Udan RS, Yang Q, Kim J, Xie J, et al. Inactivation of YAP oncoprotein by the Hippo pathway is involved in cell contact inhibition and tissue growth control. Genes Dev 2007, 21: 2747–2761. 25. Lei QY, Zhang H, Zhao B, Zha ZY, Bai F, Pei XH, Zhao S, et al. TAZ promotes cell proliferation and epithelial-mesenchymal transition and is inhibited by the hippo pathway. Mol Cell Biol 2008, 28: 2426–2436. 26. Liu CY, Zha ZY, Zhou X, Zhang H, Huang W, Zhao D, Li T, et al. The hippo tumor pathway promotes TAZ degradation by phosphorylating a phosphodegron and recruiting the SCF{beta}-TrCP E3 ligase. J Biol Chem 2010, 285: 37159–37169. 27. Zhao B, Li L, Tumaneng K, Wang CY, Guan KL. A coordinated phosphorylation by Lats and CK1 regulates YAP stability through SCF (beta-TRCP). Genes Dev 2010, 24: 72–85. 28. Yu FX, Guan KL. The Hippo pathway: regulators and regulations. Genes Dev 2013, 27: 355–371. 29. Johnson R, Halder G. The two faces of Hippo: targeting the Hippo pathway for regenerative medicine and cancer treatment. Nat Rev Drug Discov 2014, 13: 63–79. 30. Lawrence PA, Casal J. The mechanisms of planar cell polarity, growth and the Hippo pathway: some known unknowns. Dev Biol 2013, 377: 1–8. 31. Schroeder MC, Halder G. Regulation of the Hippo pathway by cell architecture and mechanical signals. Semin Cell Dev Biol 2012, 23: 803–811. 32. Halder G, Dupont S, Piccolo S. Transduction of mechanical and cytoskeletal cues by YAP and TAZ. Nat Rev Mol Cell Biol 2012, 13: 591–600. 33. Attisano L, Wrana JL. Signal integration in TGF-beta, WNT, and Hippo pathways. F1000Prime Rep 2013, 5: 17. 34. Bernascone I, Martin-Belmonte F. Crossroads of Wnt and Hippo in epithelial tissues. Trends Cell Biol 2013, 23: 380–389. 35. Varelas X, Sakuma R, Samavarchi-Tehrani P, Peerani R, Rao BM, Dembowy J, Yaffe MB, et al. TAZ controls Smad nucleocytoplasmic shuttling and regulates human embryonic stem-cell self-renewal. Nat Cell Biol 2008, 10: 837–848. 36. Creasy CL, Ambrose DM, Chernoff J. The Ste20-like protein kinase, Mst1, dimerizes and contains an inhibitory domain. J Biol Chem 1996, 271: 21049–21053.

Structural dissection of Hippo signaling

37

Structural dissection of Hippo signaling

80.

81.

82.

83. 84. 85.

86.

87.

88.

89.

90.

91.

92. 93.

94.

95.

96.

97.

98.

99.

100. 101.

domains from YAP and TAZ, nuclear effectors of the Hippo pathway. Biochemistry 2011, 50: 3300–3309. Sudol M, Shields DC, Farooq A. Structures of YAP protein domains reveal promising targets for development of new cancer drugs. Semin Cell Dev Biol 2012, 23: 827–833. Aragon E, Goerner N, Xi Q, Gomes T, Gao S, Massague J, Macias MJ. Structural basis for the versatile interactions of Smad7 with regulator WW domains in TGF-beta Pathways. Structure 2012, 20: 1726–1736. Schumacher B, Skwarczynska M, Rose R, Ottmann C. Structure of a 14-3-3sigma-YAP phosphopeptide complex at 1.15 A resolution. Acta Crystallogr Sect F Struct Biol Cryst Commun 2010, 66: 978–984. Andrianopoulos A, Timberlake WE. ATTS, a new and conserved DNA binding domain. Plant Cell 1991, 3: 747–748. Burglin TR. The TEA domain: a novel, highly conserved DNA-binding motif. Cell 1991, 66: 11–12. Vassilev A, Kaneko KJ, Shu H, Zhao Y, DePamphilis ML. TEAD/TEF transcription factors utilize the activation domain of YAP65, a Src/Yesassociated protein localized in the cytoplasm. Genes Dev 2001, 15: 1229–1241. Chen L, Chan SW, Zhang X, Walsh M, Lim CJ, Hong W, Song H. Structural basis of YAP recognition by TEAD4 in the hippo pathway. Genes Dev 2010, 24: 290–300. Li Z, Zhao B, Wang P, Chen F, Dong Z, Yang H, Guan KL, et al. Structural insights into the YAP and TEAD complex. Genes Dev 2010, 24: 235–240. Tian W, Yu J, Tomchick DR, Pan D, Luo X. Structural and functional analysis of the YAP-binding domain of human TEAD2. Proc Natl Acad Sci USA 2010, 107: 7293–7298. Halder G, Polaczyk P, Kraus ME, Hudson A, Kim J, Laughon A, Carroll S. The Vestigial and Scalloped proteins act together to directly regulate wingspecific gene expression in Drosophila. Genes Dev 1998, 12: 3900–3909. Paumard-Rigal S, Zider A, Vaudin P, Silber J. Specific interactions between vestigial and scalloped are required to promote wing tissue proliferation in Drosophila melanogaster. Dev Genes Evol 1998, 208: 440–446. Simmonds AJ, Liu X, Soanes KH, Krause HM, Irvine KD, Bell JB. Molecular interactions between Vestigial and Scalloped promote wing formation in Drosophila. Genes Dev 1998, 12: 3815–3820. Guss KA, Nelson CE, Hudson A, Kraus ME, Carroll SB. Control of a genetic regulatory network by a selector gene. Science 2001, 292: 1164–1167. Chen HH, Mullett SJ, Stewart AF. Vgl-4, a novel member of the vestigiallike family of transcription cofactors, regulates alpha1-adrenergic activation of gene expression in cardiac myocytes. J Biol Chem 2004, 279: 30800–30806. Maeda T, Chapman DL, Stewart AF. Mammalian vestigial-like 2, a cofactor of TEF-1 and MEF2 transcription factors that promotes skeletal muscle differentiation. J Biol Chem 2002, 277: 48889–48898. Gunther S, Mielcarek M, Kruger M, Braun T. VITO-1 is an essential cofactor of TEF1-dependent muscle-specific gene regulation. Nucleic Acids Res 2004, 32: 791–802. Jiao S, Wang H, Shi Z, Dong A, Zhang W, Song X, He F, et al. A peptide mimicking VGLL4 function acts as a YAP antagonist therapy against gastric cancer. Cancer Cell 2014, 25: 166–180. Guo T, Lu Y, Li P, Yin MX, Lv D, Zhang W, Wang H, et al. A novel partner of Scalloped regulates Hippo signaling via antagonizing Scalloped-Yorkie activity. Cell Res 2013, 23: 1201–1214. Zhang W, Gao Y, Li P, Shi Z, Guo T, Li F, Han X, et al. VGLL4 functions as a new tumor suppressor in lung cancer by negatively regulating the YAP-TEAD transcriptional complex. Cell Res 2014, 24: 331–343. Koontz LM, Liu-Chittenden Y, Yin F, Zheng Y, Yu J, Huang B, Chen Q, et al. The Hippo effector Yorkie controls normal tissue growth by antagonizing scalloped-mediated default repression. Dev Cell 2013, 25: 388–401. Pobbati AV, Hong W. Emerging roles of TEAD transcription factors and its coactivators in cancers. Cancer Biol Ther 2013, 14: 390–398. Mielcarek M, Piotrowska I, Schneider A, Gunther S, Braun T. VITO-2, a new SID domain protein, is expressed in the myogenic lineage during early mouse embryonic development. Gene Expr Patterns 2009, 9: 129–137.

Downloaded from http://abbs.oxfordjournals.org/ at UNIVERSITY OF PITTSBURGH on February 11, 2015

58. van der Weyden L, Adams DJ. The Ras-association domain family (RASSF) members and their role in human tumourigenesis. Biochim Biophys Acta 2007, 1776: 58–85. 59. Vavvas D, Li X, Avruch J, Zhang XF. Identification of Nore1 as a potential Ras effector. J Biol Chem 1998, 273: 5439–5442. 60. Stieglitz B, Bee C, Schwarz D, Yildiz O, Moshnikova A, Khokhlatchev A, Herrmann C. Novel type of Ras effector interaction established between tumour suppressor NORE1A and Ras switch II. EMBO J 2008, 27: 1995–2005. 61. Harjes E, Harjes S, Wohlgemuth S, Muller KH, Krieger E, Herrmann C, Bayer P. GTP-Ras disrupts the intramolecular complex of C1 and RA domains of Nore1. Structure 2006, 14: 881–888. 62. Bork P, Sudol M. The WW domain: a signalling site in dystrophin? Trends Biochem Sci 1994, 19: 531–533. 63. Ohnishi S, Guntert P, Koshiba S, Tomizawa T, Akasaka R, Tochio N, Sato M, et al. Solution structure of an atypical WW domain in a novel beta-clam-like dimeric form. FEBS Lett 2007, 581: 462–468. 64. Pearce LR, Komander D, Alessi DR. The nuts and bolts of AGC protein kinases. Nat Rev Mol Cell Biol 2010, 11: 9–22. 65. Bichsel SJ, Tamaskovic R, Stegert MR, Hemmings BA. Mechanism of activation of NDR (nuclear Dbf2-related) protein kinase by the hMOB1 protein. J Biol Chem 2004, 279: 35228–35235. 66. Hergovich A, Schmitz D, Hemmings BA. The human tumour suppressor LATS1 is activated by human MOB1 at the membrane. Biochem Biophys Res Commun 2006, 345: 50–58. 67. Stavridi ES, Harris KG, Huyen Y, Bothos J, Verwoerd PM, Stayrook SE, Pavletich NP, et al. Crystal structure of a human Mob1 protein: toward understanding Mob-regulated cell cycle pathways. Structure 2003, 11: 1163–1170. 68. Ponchon L, Dumas C, Kajava AV, Fesquet D, Padilla A. NMR solution structure of Mob1, a mitotic exit network protein and its interaction with an NDR kinase peptide. J Mol Biol 2004, 337: 167–182. 69. Mrkobrada S, Boucher L, Ceccarelli DF, Tyers M, Sicheri F. Structural and functional analysis of Saccharomyces cerevisiae Mob1. J Mol Biol 2006, 362: 430–440. 70. Hergovich A, Hemmings BA. Hippo signalling in the G2/M cell cycle phase: lessons learned from the yeast MEN and SIN pathways. Semin Cell Dev Biol 2012, 23: 794–802. 71. Rock JM, Lim D, Stach L, Ogrodowicz RW, Keck JM, Jones MH, Wong CC, et al. Activation of the yeast Hippo pathway by phosphorylation-dependent assembly of signaling complexes. Science 2013, 340: 871–875. 72. Park HW, Guan KL. Regulation of the Hippo pathway and implications for anticancer drug development. Trends Pharmacol Sci 2013, 34: 581–589. 73. Sudol M, Harvey KF. Modularity in the Hippo signaling pathway. Trends Biochem Sci 2010, 35: 627–633. 74. Hergovich A. Mammalian Hippo signalling: a kinase network regulated by protein-protein interactions. Biochem Soc Trans 2012, 40: 124–128. 75. Hao Y, Chun A, Cheung K, Rashidi B, Yang X. Tumor suppressor LATS1 is a negative regulator of oncogene YAP. J Biol Chem 2008, 283: 5496–5509. 76. Macias MJ, Hyvonen M, Baraldi E, Schultz J, Sudol M, Saraste M, Oschkinat H. Structure of the WW domain of a kinase-associated protein complexed with a proline-rich peptide. Nature 1996, 382: 646–649. 77. Pires JR, Taha-Nejad F, Toepert F, Ast T, Hoffmuller U, Schneider-Mergener J, Kuhne R, et al. Solution structures of the YAP65 WW domain and the variant L30 K in complex with the peptides GTPPPPYTVG, N-(n-octyl)-GPPPY and PLPPY and the application of peptide libraries reveal a minimal binding epitope. J Mol Biol 2001, 314: 1147–1156. 78. Aragon E, Goerner N, Zaromytidou AI, Xi Q, Escobedo A, Massague J, Macias MJ. A Smad action turnover switch operated by WW domain readers of a phosphoserine code. Genes Dev 2011, 25: 1275–1288. 79. Webb C, Upadhyay A, Giuntini F, Eggleston I, Furutani-Seiki M, Ishima R, Bagby S. Structural features and ligand binding properties of tandem WW

38

111. Xu HE, Rould MA, Xu W, Epstein JA, Maas RL, Pabo CO. Crystal structure of the human Pax6 paired domain-DNA complex reveals specific roles for the linker region and carboxy-terminal subdomain in DNA binding. Genes Dev 1999, 13: 1263–1275. 112. Kappei D, Butter F, Benda C, Scheibe M, Draskovic I, Stevense M, Novo CL, et al. HOT1 is a mammalian direct telomere repeat-binding protein contributing to telomerase recruitment. EMBO J 2013, 32: 1681–1701. 113. Court R, Chapman L, Fairall L, Rhodes D. How the human telomeric proteins TRF1 and TRF2 recognize telomeric DNA: a view from highresolution crystal structures. EMBO Rep 2005, 6: 39–45. 114. Harvey KF, Zhang X, Thomas DM. The Hippo pathway and human cancer. Nat Rev Cancer 2013, 13: 246–257. 115. Liu-Chittenden Y, Huang B, Shim JS, Chen Q, Lee SJ, Anders RA, Liu JO, et al. Genetic and pharmacological disruption of the TEAD-YAP complex suppresses the oncogenic activity of YAP. Genes Dev 2012, 26: 1300–1305. 116. Couzens AL, Knight JD, Kean MJ, Teo G, Weiss A, Dunham WH, Lin ZY, et al. Protein interaction network of the mammalian Hippo pathway reveals mechanisms of kinase-phosphatase interactions. Sci Signal 2013, 6: rs15. 117. Hauri S, Wepf A, van Drogen A, Varjosalo M, Tapon N, Aebersold R, Gstaiger M. Interaction proteome of human Hippo signaling: modular control of the co-activator YAP1. Mol Syst Biol 2013, 9: 713. 118. Kwon Y, Vinayagam A, Sun X, Dephoure N, Gygi SP, Hong P, Perrimon N. The Hippo signaling pathway interactome. Science 2013, 342: 737–740. 119. Wang W, Li X, Huang J, Feng L, Dolinta KG, Chen J. Defining the proteinprotein interaction network of the human hippo pathway. Mol Cell Proteomics 2014, 13: 119–131.

Downloaded from http://abbs.oxfordjournals.org/ at UNIVERSITY OF PITTSBURGH on February 11, 2015

102. Helias-Rodzewicz Z, Perot G, Chibon F, Ferreira C, Lagarde P, Terrier P, Coindre JM, et al. YAP1 and VGLL3, encoding two cofactors of TEAD transcription factors, are amplified and overexpressed in a subset of soft tissue sarcomas. Genes Chromosomes Cancer 2010, 49: 1161–1171. 103. Gambaro K, Quinn MC, Wojnarowicz PM, Arcand SL, de Ladurantaye M, Barres V, Ripeau JS, et al. VGLL3 expression is associated with a tumor suppressor phenotype in epithelial ovarian cancer. Mol Oncol 2013, 7: 513–530. 104. Vaudin P, Delanoue R, Davidson I, Silber J, Zider A. TONDU (TDU), a novel human protein related to the product of vestigial (vg) gene of Drosophila melanogaster interacts with vertebrate TEF factors and substitutes for Vg function in wing formation. Development 1999, 126: 4807–4816. 105. Pobbati AV, Chan SW, Lee I, Song H, Hong W. Structural and functional similarity between the Vgll1–TEAD and the YAP–TEAD complexes. Structure 2012, 20: 1135–1140. 106. Xiao JH, Davidson I, Ferrandon D, Rosales R, Vigneron M, Macchi M, Ruffenach F, et al. One cell-specific and three ubiquitous nuclear proteins bind in vitro to overlapping motifs in the domain B1 of the SV40 enhancer. EMBO J 1987, 6: 3005–3013. 107. Mar JH, Ordahl CP. M-CAT binding factor, a novel trans-acting factor governing muscle-specific transcription. Mol Cell Biol 1990, 10: 4271–4283. 108. Nikovits W Jr., Kuncio G, Ordahl CP. The chicken fast skeletal troponin I gene: exon organization and sequence. Nucleic Acids Res 1986, 14: 3377–3390. 109. Anbanandam A, Albarado DC, Nguyen CT, Halder G, Gao X, Veeraraghavan S. Insights into transcription enhancer factor 1 (TEF-1) activity from the solution structure of the TEA domain. Proc Natl Acad Sci USA 2006, 103: 17225–17230. 110. Holm L, Rosenstrom P. Dali server: conservation mapping in 3D. Nucl Acids Res 2010, 38: W545–W549.

Structural dissection of Hippo signaling

Structural dissection of Hippo signaling.

The Hippo pathway controls cell number and organ size by restricting cell proliferation and promoting apoptosis, and thus is a key regulator in develo...
814KB Sizes 3 Downloads 6 Views