! " #$% & ' () *+,! -.-! ++ !/012/+/ 3+./1!/ 33+.+ #%%4 Comprehensive summaries of Uppsala Dissertations from the Faculty of Science and Technology 786 Nuclear pore membrane glycoprotein 210 as a new marker for epithelial cells BY MAGNUS OLSSON Dissertation for the degree of Doctor of Philosophy in Molecular Cell Biology presented at Uppsala University 2003. ABSTRACT Olsson, M. (2003). Nuclear pore membrane glycoprotein 210 as a new marker for epithelial cells. Comprehensive summaries of Uppsala Dissertations from the Faculty of Science and Technology 786. 55 pp. Uppsala. ISBN 91-554-5486-0. Epithelial cell polarisation is a prerequisite for the branching morphogenesis in several organs. Differential screening techniques were used to identify genes, which are upregulated during induction of epithelium in early kidney development. This investigation revealed two separate genes, Nuclear localising protein 1 (Nulp1), a previously undescribed gene with sequence characteristics of the basic helix-loop-helix transcription factor family, and glycoprotein 210 (gp210, POM210), an integral membrane protein constituent of the nuclear pore complex (NPC). Of these, gp210 was found to be upreglated during conversion of mesenchyme to epithelium. The nuclear envelope, which demarcates the nuclear region in the eukaryotic cell, consists of an inner and an outer membrane that are fused at the locations for NPCs. These large macromolecular assemblages are tube like structures connecting the cytoplasmic and nuclear compartments of the cell. NPCs serve as the only conduits for exchange of molecular information between these cellular rooms. Electron microscopy techniques have revealed detailed information about the NPC architecture. A number of its proteins (nucleoporins) have been characterised and embodied as components of the NPC structure. Active, energy dependent nucleocytoplasmic transport of RNAs and proteins is mediated by a group of soluble receptor proteins, collectively termed karyopherins. Gp210 has been suggested to be important for nuclear pore formation. Nevertheless, our analyses showed a limited expression pattern of gp210, with its mRNA and protein largely confined to epithelial cells in the mouse embryo. Furthermore, in several cell lines, gp210 was undetectable. The expression pattern of gp210 was not synchronised with some other nucleoporins, indicating NPC heterogeneity. Characterisation of the structure of the human gp210 gene, including its promoter region, gave insight about possible cell-type specific gene regulatory mechanisms. Regulation of molecular traffic between the nucleus and the cytoplasm leads to transcriptional control. Cell specific configuration of the NPC structure, due to diffential expression of gp210, could be involved in this control. Gp210 could be of importance for the development of epithelial cell polarisation. Key words: Nuclear pore complex, gp210, Nulp1, epithelial cells, kidney development. Magnus Olsson, Department of Cell and Molecular Biology, Biomedical Center, Uppsala University, Box 596, SE-751 24 Uppsala Sweden ” Magnus Olsson 2002 ISSN 1104-232X ISBN 91-554-5486-0 Printed in Sweden by Fyris-Tryck AB, Uppsala 2002 2 3 The thesis is based on the following articles, referred to in the text by their Roman numerals I-IV: I. Olsson, M., Ekblom, M., Fecker, L., Kurkinen, M., and Ekblom, P. (1999). cDNA cloning and embryonic expression of mouse nuclear pore membrane glycoprotein 210 mRNA. Kidney International 3, 827-838. II. Olsson, M., Schéele, S., and Ekblom, P. (2002). Limited expression of the nuclear pore membrane glycoprotein 210 in cell lines and tissues suggests cell-type specific nuclear pores in metazoans. In manuscript. III. Olsson, M., Mason, J. M., English, M. A., Licht, J. D., and Ekblom, P. (2002). Genomic structure and promoter sequence analysis of nucleoporin gp210, a marker for epithelial cells. In manuscript. IV. Olsson, M., Durbeej, M., Ekblom, P., and Hjalt, T. (2002). Nulp1, a novel basic helix-loop-helix protein expressed broadly during early embryonic organogenesis and prominently in developing dorsal root ganglia. Cell & Tissue Research 308, 361-370. Reprints were made with permission from the publishers. 4 TABLE OF CONTENTS ABBREVIATIONS …………………………………………………………6 INTRODUCTION …………………………………………………………8 EPITHELIAL CELLS ………………………………………………....9 KIDNEY ORGANOGENESIS Basic features of kidney organogenesis Regulatory molecules in early kidney development TECHNIQUES Differential hybridization Differential display …………………9 ………………..10 ………………………………………..12 ………………………………………..12 SPECIFIC AIMS OF THIS THESIS ………………………………………..14 THE NUCLEAR PORE COMPLEX The nuclear envelope Lamins and the nuclear lamina Architecture of the nuclear pore complex Nuclear pore complexes in the Annulate lamellae Nuclear pore proteins Function of nucleoporins Integral membrane nuclear pore proteins The nuclear pore complex during mitosis ………………..14 ………………..15 ………………..15 ………………..18 ………………..18 ………………..22 ………………..23 ………………..25 BIDIRECTIONAL TRANSPORT OF MOLECULES BETWEEN THE NUCLEUS AND THE CYTOPLASM Passive diffusion ………………………………………………..26 Active import ………………………………………………..26 Active export ………………………………………………..28 RESULTS AND DISCUSSION ………………………………………………..31 FUTURE PERSPECTIVES ………………………………………………..35 POPULÄRETENSKAPLIG SAMMANFATTNING PÅ SVENSKA ………..35 ACKNOWLEDGMENTS ………………………………………………..37 LITERATURE CITED ………………………………………………..38 5 ABBREVIATIONS bp Base pair cDNA Complementary deoxyribonucleic acid ER Endoplasmic reticulum FG Nup Nucleoporin containing FG, GLFG or FXFG amino acid repeats kb Kilobase kDa Kilodalton mAb Monoclonal antibody mRNA Messenger ribonucleic acid N-linked Asparagine-linked NE Nuclear envelope NES Nuclear export signal NLS Nuclear localizing signal NPC Nucleo pore complex Nup Nucleoporin PCR Polymerase chain reaction 6 7 INTRODUCTION Large-scale DNA sequencing efforts have provided us with the assembled and annotated genome sequences from a diverse set of organisms. Fairly precise estimations of the number of genes contained in several of these genomes have been done. The human genome contains approximately 30.000 different genes (Venter et al., 2001), of which only a fraction are simultaneously expressed in any individual cell (Adams et al., 1995). The major control of gene expression occurs at the transcriptional level, the initial step in gene expression. Transcriptional activation proceeds in coordinated stages, from the packaging of a gene into chromatin and its localization within the nucleus to the recruitment of transcription factors whose inhibitory or activating influences is a result of specific protein-protein or proteinDNA interactions. The ability of each cell to program its genome and determine which genes are to be expressed is central to tissue differentiation, organogenesis, organismal development and disease. Thus, the ability to compare expression of genes in different cell types provides the underlying information we need to analyse these events. A number of techniques have evolved aiming to isolate genes expressed at a given time and under specific stimuli. In this thesis, two of these methods where applied in order to specifically identify genes involved in the induction of epithelial cell polarisation in early kidney development, a generally accepted model system for organogenesis. 8 EPITHELIAL CELLS The ability of cells to generate biochemically and functionally distinct plasma membrane domains is a prerequisite for the development of multicellular organisms. The best-studied examples of polarized cells are columnar epithelial cells, such as the cells lining the lumen of the intestine or of nephrons. The plasma membrane of epithelial cells is typically divided into two domains: an apical cell surface facing the organ lumen and a basolateral cell surface that is in contact with neighboring cells and the underlying extracellular matrix. The two domains are separated by tight junctions, which form a morphological border and act as an intramembrane diffusion barrier. KIDNEY ORGANOGENESIS Basic features of kidney organogenesis Most vertebrate organs, once formed, continue to perform the function for which they were generated until the death of the organism. The kidney is a notable exception to this rule. Vertebrates utilise a progression of more complex kidneys as they grow and develop. In mammals, three successive sets of excretory organs develop during embryogenesis: the pronephros, the mesonephros and the metanephros. The first two are only transiently present during early fetal life in mammals, and the mature kidney develops from the metanephros. The development of the human metanephric kidney starts in the fifth gestational week, when an inductive signal originating from the metanephric mesenchymal blastema induces the uretic bud to grow out from the mesonephric (Wolffian) duct. This stage is reached at embryonic day 10.5 in mouse. The mesenchyme induces the ureteric bud to undergo branching morphogenesis, which eventually will lead to the formation of the ureter and the collecting duct system. In turn, the ureteric bud induces a part of the surrounding mesenchyme cells to condensate and to convert into new epithelium (Gruenwald, 1943; Grobstein, 1953; Erickson, 1968). All different cell types included in the adult kidney apart from the endothelial cells of the vascularization system originate from this initial reciprocal cell-cell interaction. The mesenchymal condensates transform into a defined comma-shaped body and further into the S-shaped tubule (Huber, 1905). This cell configuration is the rudiment for proximal and distal tubules, and for the different epithelial cells of the glomerulus, the basic functional unit of all kidneys. Uninduced mesenchyme cells will become the stromal cell compartments. The first branching of the growing ureteric bud is symmetric and dichotomous. In nephron formation, each tip (ampullae) of the bifurcate merge with the distal tubule and the glomerular podocytes become folded into the glomerulus. The outcome of repeated branching of the ureter with new condensates induced 9 at every branch point is a treelike architecture with a total number of one million nephrons in the mature human kidney. Microsurgical studies on embryonic human material, performed by Osathanondh and Potter gave a detailed description of the branching pattern (Osathanondh and Potter, 1963). According to this work, the branching process is divided into four periods. In the first step two tubules are formed, followed by elongation and formation of new branches. Arcades are formed during the second step. These nephric chain structures results from a repeated joining of new nephrons to the same ampullae, giving four to seven nephric units where only the youngest one are attached to the collecting duct system itself. In the third phase, no branching occurs but the collecting duct system extends themselves past the attachment points of the arcades. At this stage terminal nephrons are induced. These are, unlike the arcadic nephrons, directly connected to the collecting duct. In the fourth step the terminal ureter disappears, and the formation of nephrons stop. Regulatory molecules in early kidney development Functional aspects of regulatory molecules have emerged from gene targeting experiments in mice, and from organ culture techniques. A few examples may be sufficient to illustrate this point. Pax2, a transcriptional regulator of the paired-box family is one of the earliest markers of kidney development and also widely expressed during the development of both ductal and mesenchymal components of the urogenital system. When Pax2 is homozygously mutated, the formation of the Wolffian duct is incomplete, which leads to renal agenesis (Torres et al., 1995). Wilms’ tumour gene 1 (Wt1) is normally expressed in the uninduced mesenchyme, and in the case of null mutants, this cell population undergoes apoptosis at embryonic stage 11, suggesting a role for this gene in the survival of the mesenchyme surrounding the Wolffian duct (Kreidberg et al., 1993). Induction of the outgrowth of the ureteric bud from the mesonephric duct depends on secretion of glial derived neurotrophic factor (GDNF), expressed by the metanephric mesenchyme, and the activation of its receptor c-ret, presented exclusively at the tip of the epithelial bulging (Pachnis et al., 1993). GDNF knockout mice (Pichel et al., 1996), as well as mice lacking c-ret (Schuchardt et al., 1994) have no kidneys due to failure of the ureteric bud to grow out. Eya1, a transcription factor belonging to the homeo box gene family has been shown to be the upstream activator of GDNF (Xu et al., 1999). Conversely, counter induction by the ureteric bud itself induces condensation of the metanephric mesenchyme, which sequentially results in tubulogenesis and formation of nephrons. One or several signalling molecules specifically expressed in a spatially and temporarily correct fashion by the ureteric bud probably initiate this event. A member of the interleukin-6 (IL-6) family, leukaemia-inhibiting factor (LIF) have been shown to act as an inducer of mesenchme to epithelial conversion in organ cultures. Other IL-6 type cytokines 10 shared this ability, and the deletion of their common receptor reduced nephron development (Barasch et al., 1999). Additional factors, such as Bone morphogenetic protein 7 (Bmp7) and members of the Wnt gene family have been suggested to act in a synergy with LIF and to be required for nephrogenic induction (Herzlinger et al., 1994; Vukicevic et al., 1996). Each of these ligands is recognised by a receptor in the blastema and it is believed that transmembrane heparan sulphate proteoglycans serve as presenting agents directing these ligands to their proper location. The activation of specific signal transduction pathways will in turn generate gene products needed for nephrogenesis and the reciprocal induction of branching. Bmp7 and different transcription factors, including Pax2, Hoxa11/Hoxd11, Lim1, Pod1 and Wt1 are likely candidates for co-ordinating this activating cascade (reviewed in Burrow, 2000; Dressler, 1999; Schedl and Hastie, 2000). For example, Hoxa11 and Hoxd11 are members of a paralogous subgroup of genes, which also contains Hoxc11. The kidneys in Hoxa11/Hoxd11 double mutants have been described. In those embryos ureteric bud initiation occur normally but branching is perturbed. The expression of several genes, including Wnt11, Pax2, Wt1 and Bf2 was examined and found to be altered compared to controls, particularly with respect to expression along the dorsoventral axis (Patterson et al., 2001). During kidney development, each formed epithelial cell layer becomes coated with a basement membrane, a thin sheet of specialised extracellular matrix composed primarily of laminins, collagen IV, nidogens and sulphated proteoglycans. A constant extracellular matrix remodelling occurs through a concerted action of different metalloproteinases and by differential expression of matrix components. Mice mutated in the nidogen-binding domain of the laminin-g1 chain were recently shown to suffer from renal agenesis due to ruptures in the Wolffian duct basement membrane (Willem et al., 2002). These results are in line with previous in vitro investigations (Ekblom et al., 1994), and clearly demonstrate that interactions between basement membrane components are important for early kidney morphogenesis. A unique feature of kidney development is the conversion of mesenchymal cells into epithelium. Changes in gene expression can be precisely timed by taking advantage of the transfilter induction method (Grobstein, 1956; Saxen et al., 1968). This organ culture system mimics the in vivo situation of induction of the undifferentiated metanephric mesenchymes, which is completed in 18-24 h. At this stage, the cells are not yet polarized, but during the next few days polarized epithelial cells attached to a basement membane form. Wellcharacterised cell surface receptors for laminins include different integrin chain heterodimers and dystroglycan. Antibody perturbation of the binding sites for laminin-1 in dystroglycan as well as in integrin a6b1 in organ culture studies blocked kidney tubologenesis (Durbeej et al., 1995; Falk et al., 1996), suggesting that proper interactions between the basal lamina and its receptors are crucial for epithelial morphogenesis and kidney formation. These studies 11 also demonstrate that in vitro organ culture systems can be used to identify components involved in epithelial cell polarisation. Obviously, this process must be controlled at many levels, and it remains an important task to identify components that are specific for the polarised epithelial phenotype. A major aim of the present study was to identify molecules that appear early during the converson of mesenchyme to epithelium. A novel marker for the epithelium was identified using the in vitro organ culture system. TECHNIQUES Differential hybridisation Finding genes that are transcriptionally regulated, in a spatial or temporal fashion is one way to clarify developmental pathways as well as an approach to localise disrupted genetic programmes leading to different malignancies such as cancer. Classical methods for isolating regulated genes include differential screening of cDNA libraries or subtractive enrichment methodologies, and their modifications. They both proceed from two cellular pools of RNA that differ in a biological aspect of interest. Transcripts common to both pools are eliminated and genes exclusively expressed in either of them are eventually cloned and further characterised. In differential hybridisation, duplicate lifts from a plated library are probed with labelled cDNA from the two conditions being compared. A strong signal present in one cDNA pool but not in the other is indicative of a candidate differentially expressed message (Dworkin and Dawid, 1980). Subtractive hybridisation, on the other hand, is carried out with an excess amount of biotinylated cDNA from one sample (driver) that is hybridised with cDNA from another (tracer). Biotinylated driver/tracer hybrids are subsequently removed by streptavidin precipitation of biotin (Sive and St John, 1988). Since both these methods suffer from high complexity of the probe material, usually only high abundance cDNAs are detected and consequently eliminating low abundance genes, which may be of interest. Another major disadvantage is the fact that differences in gene expression only can be studied between two transcriptomes at a time. Differential display Two similar methods for fingerprinting of eukaryotic RNA have been developed; mRNA differential display (Liang and Pardee, 1992) and RNA arbitrarily primed PCR (RAP-PCR) (Welsh et al., 1992). They allow the detection of differentially expressed messages based upon arbitrarily priming and amplification of cDNAs representing two or more physiological states. The two protocols are distinguished in the way cDNA initially is generated. In the differential display method an oligo-dT primer, containing one or two specific additional nucleotides at the 3´end, is used whereas an arbitrarily primer is used in RAP-PCR. Following the reverse transcriptase step, a low stringency PCR amplification will result in 5012 100 gene products with a size of 100-500 bp that are resolved on a polyacrylamide gel and visualised by autoradiography. Analogous cDNA pools obtained from the compared RNA samples are resolved side by side. PCR products with sample-to-sample variations indicate regulation of the corresponding gene. Advantages of RNA fingerprinting methods compared to differential hybridisation and subtractive library constructions are primarily that they require very small amounts of total RNA or polyA+ RNA as starting material. Through visualisation, side-by-side, it is also possible to perform a simultaneous comparison of several RNA pools in which both up regulation and down regulation of gene expression can be detected. However, a large number of primer pair combinations are required to fully cover the wide representation of mRNA species presented in any cell or cell population. In recent years, new high-capacity devices for high-throughput determination of gene regulation have been developed to satisfy the increasing demands for genetic tests. Much attention has focused on the use of the cDNA micro arrays technique (Lockhart et al., 1996; Schena et al., 1995). In this method, oligonucleotides are bound to a microchip, which is then hybridised with labelled mRNA from the sample or samples to be analysed. Technically, there is no limitation in how many genes that can be screened simultaneously and in contrast to other methods described, an expression profile is given for every single gene spotted on the array. One limitation is that only predetermined genes can be processed, although this disadvantage in a near future will be overruled by the completion of the genome projects. 13 SPECIFIC AIMS OF THIS THESIS The molecular map of kidney organogenesis is at present fragmentary and it is not yet possible to identify a complete genetic pathway required for any one of the many steps in nephrogenesis. The specific aims for this thesis have been to identify missing links in this process. In particular, the aim was to identify components upregulated early during conversion of the metanephric mesenchyme to epithelium. Although this work generated publications of two separate genes, my efforts have primarily been centred on one of them, gp210, a nuclear pore membrane protein that unexpectedly was found to be differentially expressed during induction of kidney epithelial cell polarisation. Background and discussion in this thesis will therefore mainly be focused on the nuclear pore complex. THE NUCLEAR PORE COMPLEX The nuclear envelope The eukaryotic cell nucleus is highly organised into discrete functional regions separating different activities such as transcription, RNA processing, DNA synthesis and ribosome assembly. In turn, these events are physically separated, and thereby regulated independently from the translation in the cytosol, by two concentric lipid membranes forming the nuclear envelope (NE). Partitioning of the nucleus in eukaryotic cells give a more sophisticated control over gene expression than is possible in prokaryotes. The nuclear envelope is composed of three biochemically distinct domains. The outer nuclear membrane is continuous with the membrane of the endoplasmic reticulum (ER) and consequently shares a set of proteins and functional properties with this domain. In contrast, the inner nuclear membrane has unique characteristics, mainly because of a close association with the chromatin and the nuclear lamina, a supportive polymeric network located subjacent to the NE. The inner and the outer membranes make contact at several points forming holes occupied by nuclear pore complexes (NPCs), thus defining the third membrane domain. The NPCs provide channels for nucleocytoplasmic transport and serve dual roles as both mediators and regulators of transport. Beside these classical features of the nuclear envelope it also exhibits considerable plasticity. Confocal imaging has demonstrated that the entire envelope is able to form deep invaginations into the nuclear interior (Fricker et al., 1997). These straight or branching membrane tubules bring the envelope to the centre of the nucleus and may facilitate the molecular export to the cytoplasm by reducing the distance to the nearest NPC. 14 Lamins and the nuclear lamina The nuclear lamina is a flattened fibrous matrix assembly of nuclear specific intermediate filaments that provide stability to the nuclear envelope. Associated with the lamina meshwork are several integral membrane proteins of the inner nuclear membrane, including the laminaassociated proteins LAP1 and LAP2, emerin, nurim, otefin, MAN1, nesprin and the lamin B receptor (LBR) (reviewed in Goldman et al., 2002). The targeting of integral proteins to the inner membrane is thought to involve lateral diffusion along the ER and the pore membrane domain, followed by retention at the inner membrane mediated by binding to lamins or chromosomes (Ostlund et al., 1999). In humans, three lamin genes have been identified (LMNA, LMNB1 and LMNB2), encoding in total seven alternatively spliced isoforms. The isoforms are divided into type-A or type-B lamins, depending on their isoelectric point, behaviour during mitosis and expression profile. Nearly all mammalian cells express at least one type-B lamin whereas the major A-type lamin variants, lamin A and C, are developmentally regulated and preferentially expressed in differentiated non-proliferating tissues ( Stewart and Burke, 1987; Rober et al., 1990; Rober et al., 1989). The correlation of the induction of lamins A/C has led to the suggestion that these lamins are involved in the maintenance of the differentiated phenotype. All lamin isoforms possess a tripartite secondary structure with a centrally located alpha-helical rod domain flanked by a non-helical NH2-terminal and globular COOHterminal tail domain. The rod domain mediates the lateral association between lamins required for polymerisation (Stuurman et al., 1996). Attachment of the nuclear lamina to the inner nuclear membrane is mediated by interactions with several integral membrane proteins. These include emerin that primarily interacts with type A lamins (Clements et al., 2000; Fairley et al., 1999), LAP1b that interacts with both lamin A and B (Foisner and Gerace, 1993) and LBR, which bind independently to lamin B (Ye and Worman, 1994). Lamins have been suggested to be involved in various nuclear functions, including the maintenance of chromatin organisation and in DNA replication (reviewed in Goldman et al., 2002). Naturally occurring lamin A/C mutations have been implicated in the pathogenesis of muscular dystrophy, certain forms of lipodystrophy (Shackleton et al., 2000; Bonne et al., 1999) and in nuclear envelope defects (Raharjo et al., 2001; Vigouroux et al., 2001), suggesting that lamin A/C has additional functions that remain to be elucidated morefully. Architecture of the nuclear pore complex The nuclear pore complexes (NPCs) are large, supramolecular assemblages of proteins forming the only conduit for the exchange of information between the nucleus and cytoplasm. NPCs from different organisms share a common fundamental architecture. However, significant alterations in NPC architecture have arisen during evolution that may be 15 correlated with differences in nuclear transport regulation or mitotic behaviour (Yang et al., 1998). The vertebrate NPC (vNPC) has an estimated molecular weight of 125 mDa, which makes it approximately twice as large as the NPCs found in yeast (yNPC) (Reichelt at al., 1990; Rout and Blobel, 1993). Our current knowledge about the vNPC structure is to a large extent based on studies using large amphibian oocytes. Ultrastructural analyses of the NPC using transmission electron, scanning electron and atomic force microscopy have revealed a three-dimensional model visualising the Xenopus NPC basic framework as a cylindrical assembly with octagonal symmetry comprised of several distinct structural elements (Fig. 1). Signal mediated transport is believed to occur through a cylindrical element, referred to as the central transporter, which is located along the central axis of the pore complex (Rutherford et al., 1997). Eight spokes extend from a middle, inner ring that embraces the centrally located transporter. These spokes penetrate the nuclear envelope and project into the membrane lumen where they are bridged by a distal ring formed by eight loops named radial arms (Akey and Radermacher, 1993). The spoke ring complex is sandwiched between the cytoplasmic and the nucleoplasmic coaxial rings (Goldberg and Allen, 1993), where the cytoplasmic coaxial ring is located vertically above the plane of the outer nuclear membrane and consists of a thin smooth ring onto which eight subunits are attached, each supporting one short fibre that extend approximately 35-50 nm into the cytoplasm (Jarnik and Aebi, 1991). A different set of eight fibres extends approximately 40-50 nm from the nucleoplasmic coaxial ring into the nucleus, and is joined at their distal ends to form a basket structure (Ris, 1997; Ris and Malecki, 1993). Tethered to the nuclear basket is Tpr, a coiled coil protein able to form filament bundles, which have been traced into the nuclear interior for up to 350 nm (Cordes et al., 1997b). Tpr has been implicated in nuclear export (Frosst et al., 2002; Bangs et al., 1998) and is also hypothesised to constitute an interchromosomal channel-based nuclear matrix (Zimowska et al., 1997). Additional features of the vertebrate NPC are the “star ring” underlying the cytoplasmic ring and two sets of internal filaments joining the cytoplasmic and nucleoplasmic coaxial rings to the central transporter (Goldberg and Allen, 1996; Goldberg et al., 1997). Studies of the passage of gold particles through the central transporter by transmission electron microscopy indicate a central channel that has the ability to vary its opening diameter, from potentially closed to 26 nm (Feldherr and Akin, 1997). The internal filaments have been proposed to participate in the gating process by open or close the channel at the entrance to the transporter (Kiseleva et al., 1998). However, it should be mentioned that there are controversial opinions regarding the central transporter as a distinct element of the NPC. Alternative views claim that this structure is merely a cargo caught in transit during specimen preparation or collapsed cytoplasmic fibres. 16 Although it is similar to the vNPC, the yNPC is much simpler in structure. It lacks the cytoplasmic and nuclear rings and the lumenal radial arms of the spoke complex. Peripheral filamentous domains are present but in a comparison to vNPC, they are shorter. Overall, the yNPC is smaller (96 nm in diameter by 35 nm high) than the vNPC (145 nm diameter by 80 nm high) and occupies one fifth of the volume (Yang et al., 1998; Fahrenkrog et al., 1998). Figure 1. Schematic model of the vertebrate nuclear pore complex. The figure indicates major NPC constituents. Individual parts are not scaled. Note that only two out of eight central spokes are outlined and that the radial arms connecting the spokes in the perinuclear space are missing in the picture. 17 Nuclear pore complexes in the Annulate lamellae Nuclear pore complexes are not merely present in the nuclear envelope. Annulate lamellaes (ALs) are cytoplasmic stacks of double membrane containing large numbers of tightly packed pore complexes. Each double membrane possesses hexagonally packed pores and is aligned vertically by en face contact between pores. ALs are heavily represented in rapidly growing or differentiating cells, such as tumour cells, male and female gametes and virally infected cells (reviewed in Kessel, 1992). At present, there is no defined function for ALs but it has been suggested that they represent a storage form of pore complexes used in the reassembly of daughter nuclei during cell division. Evaluation of comparison between nuclear and cytoplasmic pores by electron microscopy concludes that they are morphologically indistinguishable and AL pores have been shown to interact with the soluble mediators of nuclear protein import and their karyophilic protein substrates (Cordes et al., 1997a). At the molecular level, however, it has been reported that AL pores in rat cells (line RV) lack POM210 and POM121 (Ewald et al., 1996), two integral membrane components of the NPC, although this finding has been questioned, at least for POM121 (Imreh and Hallberg, 2000). Overexpression of rat POM121 in COS-1 cells (African green monkey kidney cells) induced AL formation, indicating that this protein indeed can be a component of cytoplasmic NPCs, also suggesting that it may have a function in NPC biogenesis (Soderqvist et al., 1996). Nuclear pore proteins Stable protein components of the nuclear pore complex are referred to as nucleoporins or nups. The octagonal symmetry of the NPC results in that each nucleoporin occurs mainly in 8 or multiples of 8 copies per pore. Recently, a comprehensive characterisation of all yNPC constituents, including localisation studies performed by immunoelectron microscopy, was done. This investigation revealed approximately 30 distinct nucleoporins, each with a mainly symmetrical distribution within the vertical plane of the NPC structure (Rout et al., 2000). A similar analysis of the mammalian NPC has been presented, based on mass spectrometry identification of all proteins included in a highly purified NPC fraction from rat liver nuclei (Cronshaw et al., 2002). According to this study, the mammalian NPC seems to be composed of essentially the same number of nucleoporins as its yeast counterpart. This is less than expected considering the striking disparity in NPC size and complexity between the two species. Previously, others have estimated a larger number of nucleoporins (Fontoura and al., 1999; Miller and Forbes, 2000), although the discrepancy may be explained by the presence of pore associated proteins in the preparations, for example transport receptor proteins and molecules involved in RanGTP hydrolysis. To date, more than twenty vertebrate nups have been molecularly described and localised to discrete NPC regions (listed in Table 1). Among 18 these is a group of polypeptides that withhold multiple copies of glycine-leucinephenylalanine-glycine (GLFG), FXFG or FG repeats interrupted by characteristic spacer sequences, preferentially in their COOH-terminal region (Table 1). Many nucleoporins with repeats dispersed through a portion of their sequence have been found to bind directly to nuclear transport receptors (Fornerod et al., 1997; Hu et al., 1996; Radu et al., 1995; Shah and Forbes, 1998; Shah et al., 1998), and appear to represent binding sites for transport complexes during their translocation through the NPC. Interactions between nucleoporins and nucleoporin subcomplexes are thought to be mediated through a-helical coiled coil domains (Fornerod et al., 1996). Predicted coiled coil regions have been found in a number of nucleoporins, including Nup88, CAN/Nup214 and the Nup62 complex (Table 1). Common posttranslational modifications of nucleoporins include different kinds of phosphorylations and glycosylations by attachment of O-linked N-acetylglucosamine to serine and threonine residues (O-GlcNAc). Whereas nups are phosphorylated in a dynamic, cell-cycle dependent manner, levels of O-GlcNAc remains constant (Miller et al., 1999; Macaulay et al., 1995; Favreau et al., 1996). The role for glycosylation of NPC subunits remains elusive. M-phase specific phosphorylation, on the other hand, is shared with lamin components. It has been shown that specific phosphorylation of lamins during cell division destabilise lamin polymers, leading to filament dissolution and that the elimination of these phosphorylation sites prevents polymer disassembly during division (reviewed in (Moir et al., 1995)). Thus, it has been suggested that the function of phosphorylation of nups may parallel that shown in lamins leading to a reversible disassociation of the NPC during mitosis. To elucidate the hierarchy of interactions between metazoan nucleoporins has been a main goal in order to clarify their function. This work has partly been accomplished by a comparison with the yeast NPC and partly by either biochemical methods such as the yeast two-hybrid system or by the separation of sub complexes formed during NPC breakdown during mitosis. During this work, individual nucleoporins as well as interacting nucleporins forming sub complexes have further been localised to the central transporter, the cytoplasmic filaments or to the nucleoplasmic basket. Interestingly, vertebrate nucleoporins seems to have a more asymmetrical distribution within the vertical plane of the NPC, compared to yeast nucleoporins. To date, the nucleoporins that contribute to the nuclear basket composition include Nup153 (Pante et al., 1994; Sukegawa and Blobel, 1993; Walther et al., 2001), Nup98 (Fontoura et al., 1999; Radu et al., 1995), Nup93 (Grandi et al., 1997), Nup50 (Guan et al., 2000) and a new Nup160 sub complex containing Nup160-Nup133-Nup96-Nup107 (Vasu et al., 2001). Nup133 and Nup107 have also been localised to the cytoplasmic face of the NPC. Throughout mitosis, a fraction of Nup133 and Nup107 localizes to the kinetochores, thus revealing an unexpected connection between structural NPCs constituents and kinetochores (Belgareh et al., 2000). Potential members of the central scaffold of the pore are the Nup93-Nup205-Nup188 complex (Miller et al., 2000) and Nup155 (Radu et al., 19 1993), where the latter is found on both the nucleoplasmic and the cytoplasmic side of the pore. Also associated to the central domain of the NPC, at the periphery of the central gated channel is a sub complex consisting of Nup62-Nup58-Nup54-Nup45 (Guan et al., 1995; Hu et al., 1996; Finlay et al., 1991). Three components have been found to be restricted to the cytoplasmic face of the NPC, at or near the cytoplasmic fibrils: Nup88/Nup84 (Fornerod et al., 1997), CAN/Nup214 (Kraemer et al., 1994) and RANBP2/Nup358 (Delphin et al., 1997; Wilken et al., 1995), of which the former two form a sub complex (Bastos et al., 1997). 20 Name Characteristic motifs Localization in NPC Function in Transport References RanBP2/Nup358 Zn fingers Ran binding domain FXFG Cytoplasmic filaments May have an essential role in nuclear import. Wu et al., 1995 Yokoyama et al., 1995 Walther et al., 2002 Nup88 (Nup84) Coiled coil Protein import CAN/Nup214 FG Coiled coil At or near the cytoplasmic filaments. Fornerod et al., 1997 Uv et al., 2000 van Deursen et al., 1996 Nup180 Nup45 (p45) Nup54 (p54) Nup58 (p58) Nup62 (p62 Nup155 Nup96 Nup107 Nup133 Nup160 Nup93 Nup188 Nup205 (p205) Nup50 (Npap60) Nup98 Nup153 NLP-1 Tpr POM121 gp210 (POM210) FG Coiled coil FXFG, Coiled coil Leucine zipper Coiled coil Leucine zipper Ran binding domain FG, FXFG, GLFG RNA binding domain Zn-finger FG-repeats M9-like domain FG, PAFG Coiled coil Leucine zipper Coiled coil NLS Transmembrane FXFG Transmembrane Cytoplasmic coaxial ring Centraly, near the gated channel, at both nuclear and cytoplasmic face. Both cytoplasmic and nuclear face Maps to the Nuclear face of NPC. Nuclear basket and/or close to the nuclear entry of the gated channel. Nuclear basket Protein import and mRNA export Wilken et al., 1993 Role in nuclear Import. Finlay et al., 1991 Hu et al., 1996 Hu and Gerace, 1998 Radu et al., 1993 mRNA export Fontoura et al., 1999 Radu et al., 1994 Vasu et al., 2001 NPC biogenesis Grandi et al., 1997 Miller et al., 2000 Protein export. Stimulates NLS mediated import. Export of multiple types of RNA. Associated with export of most RNAs except tRNA. Importin a/b mediated import. Interacts with CRM-1 Guan et al., 2000 Lindsay et al., 2002 Nuclear basket and nuclear interior. Pore membrane Export of mRNA Cordes et al., 1997 Shibata et al., 2002 Soderqvist et al., 1997 Pore membrane Implicated in NPC biogenesis and pore dilation. Wozniak and Blobel, 1992; Drummond and Wilson, 2002 Nuclear basket Nuclear basket Radu et al., 1995 Powers et al., 1997 Shah and Forbes, 1998 Ullman et al., 1999 Walther et al., 2001 Farjot et al., 1999 Table 1. Description of isolated and characterised mammalian nucleoporins. Nucleoporins are listed in their vertical order of appearance except pore membrane proteins and NLP-1. Major subcomplexes are boxed. Coiled coil, a-helical coiled coil domain; NLS, nuclear localizing sequence. 21 Function of nucleoporins The synthetic lethal screen has been one of the most successful genetic approaches for dissecting function for individual nucleoporin members in yeast. By this method, overlapping functions can be envisioned either as redundant function or as function that depend on each other (reviewed in Fabre and Hurt, 1997). A similar approach is not available for vertebrate cells. Instead, information regarding nucleoporins and their function in nucleocytoplasmic transport, NPC assembly and in aspects of intranuclear organisation are primarily derived from alternative methods, such as antibody perturbations and nuclear assembly reactions in cell free systems. Demembranated sperm chromatin induces nuclear assembly when it is added to Xenopus egg extract. These nuclei are functionally normal with respect to protein import, DNA replication, and disassemble in response to mitotic signals (Lohka and Maller, 1985). Nuclei assembled in vitro in the presence of Xenopus egg cytosol, biochemically depleted from the Nup62 complex, has been shown to be defective in nuclear transport with a linear correlation between the amount of complex present in cytosol and the amount of nuclear transport of which those nuclei are capable (Finlay et al., 1991). The same approach has also successfully been used for Nup153, the Nup93 complex and for Nup358. Consistent with its localisation at the base of the nuclear basket, Nup153 deficient NPCs where found to lack basket components Nup93, Nup98 and Tpr. These nuclei where further shown to have a reduced capacity for basic NLS mediated cargo import and became mobile within the NE (Walther et al., 2001). The latter dysfunction may be explained by the recent finding which indicates a direct interaction between Nup153 and lamin LIII, the major lamin form present in Xenopus oocytes (Smythe et al., 2000). Nuclear import for NLS-containing karyophilic substrate appears fairly normal in Nup93-complex-depleted nuclei but show low NPC frequency and reduced size of the nucleus. Thus, Nup93 or its interacting nucleoporin partners seems to be important for NPC biogenesis (Grandi et al., 1997). Interestingly and somewhat unexpectedly, the cytoplasmic filaments, which have been proposed to act as a docking site for both import and export complexes, seem not to be needed for nuclear import. NPCs devoid of RanBP2/Nup358 or CAN/Nup214 or both show no deficiency in NLS or M9 import signal mediated nuclear accumulation (Walther et al., 2002). Nuclear microinjection of anti-Nup98 specific antiserum revealed this nucleporin as an essential element of the RNA export pathway. Cytoplasmic entry for several classes of RNAs, including snRNAs, 5S RNA, large ribosomal RNAs, and mRNA was prevented. In contrast, the export of tRNA was unaffected (Powers et al., 1997). A limited number of nucleoporins have been analysed by genetic methods in metazoans. CAN/Nup214 -/- mouse embryonic stem (ES) cells are not viable but maternally derived CAN account for normal development of homozygous offspring until between 4.0 and 4.5 days postcoitum when they die due to a blastocoel collapse. In 3.5-day-old mutant 22 embryos, cultured in vitro, progressive depletion of CAN leads to cell cycle arrest in the G2 phase, nuclear accumulation of polyadenylated RNA and in sharp contrast to the results presented by Walther and his colleagues (Walther et al., 2002), impaired NLS-mediated protein uptake (van Deursen et al., 1996). Embryonic lethality is also seen in Nup50 deleted mice, although at a later stage compared to CAN-/- mice. Phenotypically, the homozygous Nup50 deletion is characterised by severe defects in the neural tube development and abnormalities in p27kip1 expression (Smitherman et al., 2000). p27kip1 is a Nup50 interacting cyklin-dependent kinase inhibitor involved in regulation of the cell cycle (reviewed in (Sherr and Roberts, 1999)). On the other hand, microinjection studies implicate Nup50 in the nuclear export of proteins containing leucine-rich export sequences and was, in consistent with this finding shown to directly bind Crm1 (Guan et al., 2000). Apart from being a structural component of the NPC, Nup50 may also serve as a soluble co factor in importin-b mediated nuclear import (Lindsay et al., 2002). A current model for translocation of transport complexes through the NPC involves interaction between the transport receptor and FXFG repeat containing nucleporins (FG Nups). Sequential repeat domains are proposed to act as stepping-stones for karyopherinmediated transport reactions (Radu et al., 1995). The yeast NPC contains thirteen FG Nups (Rout et al., 2000). When Allen and co-workers sampled eight of these individually using immobilised FG domains as bait for proteins in whole yeast extract it was revealed that they all bound transport receptors, in a more or less selective manner (Allen et al., 2001). A selective binding efficiency between a FG Nup and a transport complex may indicate the existence of separate translocation routes. The observation that Nup98 disruption impairs M9 signal mediated import but leave the import of ribosomal protein L23a and spliceosome protein U1A intact is in agreement with this theory (Wu et al., 2001). Integral membrane nuclear pore proteins Metazoan NPCs are during interphase tightly connected to the bilayered nuclear envelope and moves sparsely, in an elastic manner, in the plane of the NE. Synchronised movements between NPCs and the lamina meshwork clearly indicate that they belong to a common network (Daigle et al., 2001). In contrast, in vivo data on the dynamics of the NPC from budding yeast suggests a high mobility of NPCs in the NE (Bucci and Wente, 1997), maybe as a consequence of the absence of true lamin genes in the yeast genome. Anchoring of the NPC to the NE has been suggested to be mediated by integral membrane nuclear pore proteins (POMs). To date, only two pore membrane proteins have been described in metazoans, namely POM121 (Hallberg et al., 1993) and gp210 (POM210) (Wozniak et al., 1989; Gerace et al., 1982). Both proteins possess a single transmembrane domain with their C-terminal end protruding into the cytoplasm. However, while POM121 is mainly cytoplamic, most of the mass (95%) of gp210 occurs in the perinuclear lumen 23 (Soderqvist and Hallberg, 1994; Greber et al., 1990). Although the NPC target signal have been determined for gp210 and POM121 (Wozniac and Blobel, 1992; Soderqvist et al., 1997), their respective binding partner(s) within the NPC remains elusive. The early post mitotic appearance of POM121 in the nuclear envelope relative to other nucleporins supports its role as an anchoring molecule (Bodoor et al., 1999). Repeat elements of the XFXFG type is indeed present in the primary sequence of POM121 but whether these represent binding domains for other nucleoporins or for molecules of the soluble transport machinery is unclear. Gp210, on the other hand, lack pentapeptide repeats but contain a hydrophobic domain distinct from its transmembrane region. By referring to segments present in the luminal domain of certain viral fusion proteins, Wozniak and Blobel (1992) suggested that this hydrophobic domain might interact with lipid bilayers and act as a fusogenic peptide during formation of the pore. However, comparison of gp210 primary sequences from several species revealed no significant evolutionary conservation of this putative second hydrophobic domain (Cohen et al., 2001). Evidence for an early role of POM210 in NPC assembly and pore dilation comes from studies where recombinant cytoplasmic tails and tail specific antibodies was used in vitro to inhibit the functionality of this domain. It was shown that inhibited nuclei were growth arrested and that only mini pores was formed, these pores lacked several nucleporins, including Nup214, Nup153 and Nup98, but not Nup62 (Drummond and Wilson, 2002). In vivo, antibodies directed to the lumenal part of POM210 interfered with passive diffusion as well as signal-mediated import, data arguing for a direct or indirect link between POM210 and nucleo pore constituents involved in these events (Greber and Gerace, 1992). Furthermore, POM210 has been shown to form homodimers and even higher orders of oligomerisation, at least in vitro (Favreau et al., 2001). A new family consisting of three putative pore membrane proteins in human was recently isolated and characterised (Coy et al., 2002). The encoded proteins of this family (POMFIL1-3) are distantly related to POM121 and to genes involved in axon guidance. In the comprehensive classification of yeast NPC constituents performed by Rout and his colleagues (Rout et al., 2000), three true pore membrane proteins where found, POM152p (Tcheperegine et al., 1999; Wozniak et al., 1994), Ndc1p (Chial et al., 1998) and a previously uncharacterised protein named POM34p. Neither of them shares any significant sequence homology to any POMs found in vertebrates. Strikingly, examinations of yeast strains mutated in POM152p or Ndc1p show that these genes are not essential for cell survival. Deletion mutants of POM152 are lethal in combination with mutations in Nup170p (Aitchison et al., 1995), which are proposed to help establish the functional diameter of the NPC central channel (Shulga et al., 2000). 24 The nuclear pore complex during mitosis De novo assembly of NPCs takes place partly during interphase, in order to accommodate nuclear growth, and partly postmitoticallay, in order to reassemble nucleoporins that have been dispersed in the cytpolasm during the mitotic nuclear envelope breakdown, a process needed to allow the condensed chromosomes access to the mitotic apparatus. In higher eukaryotes, the nuclear membranes disassemble during prometaphase and reassemble around daughter chromosomes during late anaphase and telophase (reviewed in Gant and Wilson, 1997). Soluble components of the NPC are thought to break down into either single units, or into mitotic sub complexes during this process. Whether nuclear membrane proteins are released into domain specific vesicles during cell division (Chaudhary and Courvalin, 1993; Buendia and Courvalin, 1997), or released throughout all ER membranes is still controversial (Yang et al., 1997). Dispersed nucleporins are in the end of mitosis reutilised to form new pore complexes. This seems to occur in a sequential fashion and high resolution scanning EM has revealed morphology distinct NPC assembly intermediates (Goldberg et al., 1997) (Drummond and Wilson, 2002). In an attempt to resolve the temporal reassociation of different nucleoporins into newly formed NPCs, it was shown that Nup153 and POM121 where reintroduced at late anaphase, thereby preceding that of the Nup62 complex and CAN/Nup214 together with Nup84 in early to middle telophase. Interestingly, POM210, proposed to be important for NPC biogenesis, do not reappear until late telophase/early G1 phase, in parallel with tpr (Chaudhary and Courvalin, 1993; Bodoor et al., 1999). 25 BIDIRECTIONAL TRANSPORT OF MOLECULES BETWEEN THE NUCLEUS AND THE CYTOPLAM Passive diffusion Ions and small proteins lacking a nuclear localising signal (NLS) are able to passively diffuse through the NPC. Macromolecules below the size of 90-100 Å or approximately 60 kDa equilibrates between the nucleus and the cytoplasm with rates inversely proportioned to their size. The actual transport channel for passive diffusion across the pore is not defined but two alternatives have been suggested. First, the central transporter may function as the transit element for both active transport and diffusion (Feldherr and Akin, 1997). Second, Hinshaw and his colleagues reported 8 peripheral channels located within the NPC spoke complex that potentially also could mediate diffusion of small permeants (Hinshaw et al., 1992). Active import Whereas the nuclear pore permit passage of relatively small molecules by diffusion, particles with a diameter as large as 230-250 Å can enter the nucleoplasm or cytoplasm by signal mediated transport (Feldherr and Akin, 1990). This transport requires specific interaction with either the NPC or with members of a conserved super family of soluble guiding receptors referred to as b-karyopherins. b-karyopherins can be classified into importins or exportins depending on the direction in which they transport a cargo. Fourteen transport receptors in the yeast Saccharomy cererevisiae and over 20 in higher vertebrates have been identified (reviewed in Strom and Weis, 2001). Some classes of cargo contain signals that bind directly to a cognate receptor while other types of signals bind indirectly, via adaptor proteins belonging to the karyopherin-a, or importin-a family. Human and mouse posses at least 5 isoforms of importin-a (a 1-5) that recognize different cargo target sequences (reviewed in Jans et al., 2000). They all interact specifically with importin-b1 through a conserved importin-b binding (IBB) domain (Kohler et al., 1999; Cingolani et al., 1999; Gorlich et al., 1996a). The existence of multiple receptors that specifically recognise one or a set of cargoes result in a still unknown number of different transport pathways, each regulating the localisation of one or several macromolecules. The most classical and probably most widely used import pathway is a multi-step process that can be divided into three major steps (Fig. 2A). First, proteins containing a nuclear localising signal (NLS) are recognised in the cytoplasm by a importin-a. Variations in NLS occur but the best defined are those of the SV40 Large T antigen and nucleplasmin. They consist of one (monopartite) or two stretches of basic amino acids (bipartite), respectively (Robbins et al., 1991; Kalderon et al., 1984). Other nuclear import signal-receptor system has been described. For example, tranportin (karyopherin-b2) mediates the nuclear import of a subset of hnRNP proteins bearing an M926 type NLS that is rich in glycine and aromatic residues (Siomi and Dreyfuss, 1995). Transportin binds directly to the M9-containing cargo (Nakielny et al., 1996). Importin-a forms a complex with importin-b1 (karyopherin-b1), which target the heterotrimer to the cytoplasmic face of the NPC where it binds to components of the cytoplasmic fibres that radiate from NPC into the cytoplasm (Pante and Aebi, 1996; Delphin et al., 1997). Interestingly, cytoplasmic fibres seems not to be essential for nuclear import indicating that adjacent nucleo pore domains also is able to mediate the docking phase (Walther et al., 2002). In the second step, proteins are, in a still poorly understood fashion, translocated in an energy dependent manner along the axis of the NPC into the nucleoplasm. Finally, translocation is followed by displacement of the cargo from the carrier. Cargo release is restricted to the nucleus by the interaction of the importin a/b-cargo heterotrimer with Ran (Gorlich et al., 1996b; Moore and Blobel, 1993), a small Ras related GTPase. Ran binds to GTP or GDP, and its nucleotide-bound state is regulated by several cofactors. The conversion of RanGDP into RanGTP requires Ran guanine exchange factor (RanGEF or RCC1), that predominantly is found in the nucleus, bound to chromatin (Bischoff and Ponstingl, 1991). GTP hydrolysis by Ran occurs exclusively in the cytoplasm and is initiated by the RanGTPase activating protein-1 (RanGAP) and co stimulated by the Ran-binding proteins 1 and 2 (RanBP1 and Nup358/RanBP2) (Bischoff and Gorlich, 1997; Bischoff et al., 1994; Yaseen and Blobel, 1999). Nuclear transport factor 2 (NTF2), a RanGDP-binding protein, facilitates nuclear accumulation of Ran and is responsible for the accumulation of Ran in the Nucleus at cellular steady state (Ribbeck et al., 1998). The result of these actions is a sharp gradient where the GTP-bound form of ran is found in the cytoplasm and the GDP-bound form is in the nucleoplasm. This asymmetry is crucial for at least major transport pathways in both directions (Izaurralde et al., 1997). It has been suggested that directionality of the bidirectional transport process is maintained through interactions with RanGTP/GDP and indeed, an artificial reversal of the RanGTP/GDP gradient has been shown to run transport backwards (Nachury and Weis, 1999). After releasing their cargo into the nucleus importin-b is returned to the cytoplasm in a heterodimeric form with RanGTP. Importin-a is translocated back through interaction with the exportin CAS in a RanGTP dependent manner and has thereby completed one import cycle (Kutay et al., 1997). 27 Active export Nuclear export of proteins and RNAs is closely analogous to nuclear import and involves distinct targeting signals, importin homologs, and nucleoporin binding sites, as well as Ran and its modifying factors (Fig. 2B). Nearly all nucleus-encoded RNAs require export to the cytoplasm as part of their maturation process or biosynthetic function. RNAs are transported as ribonucleoprotein (RNP) complexes and receptors have been identified that conduct mRNAs, tRNAs and untranslated small nuclear RNAs (snRNAs) to the cytoplasm, each by largely independent pathways. The giant Balbani ring transcripts in Chironomus tentans offer a unique model system for visualising assembly and transport of RNP particles (Kiesler et al., 2002; reviewed in Daneholt 2001). The best understood export signal in proteins is the leucine-rich nuclear export signal (NES), first identified in the immunodeficiency virus (HIV)-1 Rev protein (Fischer et al., 1995). Crm1 (exportin1), a protein that shares sequence similarities to the karyopherin b family appears to act as a export receptor for proteins containing a NES, including a cisacting nuclear export signal that interacts with a dimeric protein complex binding to the 5' monomethylated cap structure of spliceosomal U snRNAs (Izaurralde et al., 1995). Initially, it was shown in S, pombe, that Crm1 is a target for the antifungal drug leptomycin B (Fukuda et al., 1997). When leptomycin B was found to inhibit nuclear export of the HIV-1 protein Rev, Crm1 was suggested to be the karyopherin that exports Rev from the nucleus (Wolff et al., 1997). The human homolog of Crm1 is a soluble protein that interacts with Nup214/CAN (Fornerod et al., 1997). Crm1 seems to interact directly, without adaptor molecules, to NES elements (Ossareh-Nazari et al., 1997). The role of RanGTP during export is in contrast to the import process not to dissociate the receptor-cargo complex but to stabilise it. This has been confirmed by in vitro studies where the addition of Crm1 and RanGTP but not Crm1 alone to digitonin-permeabilised cells promotes the accumulation of export cargo on the cytoplasmic face of the nuclear pore complex (Black et al., 2001). Thus, Crm1 and RanGTP are necessary and sufficient for assembly of transport complexes in which the cargo contains a NES, as well as their targeting and translocation through the NPC. Following the translocation, the export complex reaches the cytoplasmic side of the NPC where its disassembly releases all units into the cytoplasm. This step includes RanGAP dependent hydrolysis of RanGTP to RanGDP, probably in co-operation with RanBP1 and RanBP2. Since Crm1 has a low affinity for RanGDP this conversion efficiently makes the disassembly process irreversible. An additional molecule that has been shown to be essential for the export cargo release is NXT1 (p15) (Black et al., 2001), a RanGTP binding factor that is able to shuttle between the nucleus and the cytoplasm (Black et al., 1999). As mentioned, several other export receptor molecules have been identified, each with a distinct or in some cases redundant binding and transport capacity. For example, mature tRNA is directly recognized by a karyopherin b-like export 28 receptor, exportin-t, in a RanGTP-dependent manner (Arts et al., 1998; Kutay et al., 1998). Less is known about mechanisms responsible for mRNA export but Ran-dependent pathways involving different soluble proteins likely coexist and converge to the NPC. Heterogeneous nuclear ribonucleoporins (hnRNPs), associating with mRNA during or just after transcription could provide NESs. In particular, hnRNP A1, which contains a NES termed M9 is thought to stimulate mRNA export (Michael et al., 1995). Recent investigations have proposed human TAP (also known as NXF1) as a mRNA export receptor. This protein is distinct from the karyopherin-b family and belongs to a conserved group of putative receptor molecules known as the nuclear export factor (NXF) family (Herold et al., 2000). Tap interacts with components of the NPC, as well as to three proteins implicated in the export of cellular mRNAs: RAE1/hGle2, E1B-AP5 and U2AF (Bachi et al., 2000; Zolotukhin et al., 2002), and has been shown to promote export of viral mRNAs bearing a constitutive transport element (CTE) (Gruter et al., 1998). Furthermore, a TAP homolog in C. Elegans and Mex67p, the yeast homolog of hTAP, are both essential for the sequence non-specific export of bulk polyadenylated RNAs to the cytoplasm in each species, respectively (Tan et al. 2000; SantosRosa et al., 1998). Accumulating data suggests that the small NXT1/p15 protein may function as critical cofactor for TAP mediated mRNA export. It has been shown that NXT1 regulates the ability of TAP to interact with components of the nuclear pore (Wiegand et al., 2002). Nucleocytoplasmic traffic is thus regulated at multiple sites in the cytoplasm, nucleus and the NPC. Variations in the molecular compositon of the NPC and the pore membrane could significantly influence transport, but has not beem much studied. 29 Figure 2. Molecular interactions during nucleocytoplasmic transport. (A) Import receptors bind cargo molecules in the cytoplasm and mediate their transport to the nucleus through the nucleopore complex. Binding of RanGTP to the import complex in the nucleus causes cargo and importin-a release. Recycling and disassembly of the RanGTP-karyopherin b complex upon RanGTP hydrolysis terminates the import cycle. (B) The export cycle is highly similar to the import cycle but uses different transport receptors. Furthermore, the RanGTP promote the receptor-cargo interaction. (Adapted from Kuersten et al., 2001). 30 RESULTS AND DISCUSSION All organ development involves co-ordinated growth and differentiation of epithelial and mesenchymal tissues. A unique feature of the kidney is that the mesenchyme converts into epithelium. The model system can be used to define markers for epithelial polarization. In the first study (I), our aim was to identify such components. By screening of a mouse cDNA l phage library representing the induced state of metanephrogenic mesenchyme and using cDNA probes prepared from either induced or uninduced mesenchymes we were able to identify a gene that was up regulated during this development. Unexpectedly, this gene turned out to be gp210, an integral membrane pore protein that previously had been described in rat (Wozniak et al., 1989), Drosophila melangoaster (Berrios et al., 1995) and Xenopus laevis (Gajewski et al., 1996). mRNA in situ hybridisation revealed that the transcript of gp210 was confined to epithelial tissues and to the condensing mesenchyme in the developing kidney. This epithelial pattern was demonstrated in other embryonic organs, including skin, lung, liver, brain, gut and pancreas, whereas no or little expression could be detected in stromal compartments or in mesenchymal tissues such as cartilage. In vitro, we could demonstrate that regulation of the gp210 gene is not dependent on cell proliferation. Thus, regulatory pathways separated from this cellular process activate the expression of gp210 (I). Nuclear pore complexes are critical for nucleocytoplasmic transport and as such fundamental for cell survival. Present models suggest that a single NPC is able to mediate both nuclear import and export of material with no variations in the NPC configuration exists between different cells types. It is therefore striking that several cell types appear to lack gp210. An extensive examination of the relative mRNA expression of several characterised nucleoporins in different organs and at discrete developmental stages revealed no dramatic pattern variations. By using a non-radioactive northern blot system with deoxyginine labed probes directed to Nup98/96, Nup62, Nup107, Nup54, Nup84, Nup153 and gp210, we where able to observe detectable signals for each individual nucleoporin in brain, liver and kidney. Developmental stages for each organ ranged from embryonic day 12 to adult stages (data not shown). In sharp contrast to all other nups analysed, no gp210 transcripts could be detected in four different cell lines analyzed (II). This finding was confirmed at the protein level using a polyclonal antibody raised to the COOH-terminal end of gp210 (II). Furthermore, a very restricted expression of gp210 was seen in developing mouse embryos, in sharp contrast to the distributon of Nup358, Nup214, Nup153 and Nup62, four nucleoporins collectively targeted by mAb214 (II). In conclusion, it seems that the expression of mouse gp210 is not synchronised with nucleoporins in general. In addition, the apparent absence of the gp210 protein in several specific compartments of the developing embryo and in certain cell lines 31 indicates that gp210 is a major polypeptide of the NPC of only some cell types. Hence, it could be that gp210 is not essential for NPC anchoring and pore formation. The data suggest cell-type specific functions. It is well known that the establishment and maintenance of polarity of epithelia is essential for functions of epithelial sheets. A major issue in cell biology has been to characterize components required for cell polarization (Drubin and Nelson, 1996). Spatially and functionally distinct protein scaffolds, assembled from transmembrane and cytoplasmic proteins, provide clues for polarization. Components known to control polarity include laterally located cell adhesion molecules, basally located receptors for basement membranes (Durbeej et al., 1995), and apically located cytoplasmic proteins (Knust, 2000). The current work on gp210 is descriptive, but introduces the new concept that epithelial cell polarity correlates with expression of a nuclear pore membrane protein. Gp210 could be involved in the development and maintenance of epithelial cell polarization. In order to get insight into mechanisms driving the expression of the gp210 gene, we analysed the promoter region for human gp210. Since neither TATA nor CAAT box elements could be identified in the DNA sequence preceding the ATG start codon, transcription start sites where determined experimentally. Two start sites where found, one major at position –26 and one minor at position –18. We could further demonstrate several putative binding elements for tissue specific transcription factors that where conserved in the corresponding mouse genomic region, including clustered consensus sequences associated with Sp1 and Wt1 (III). Experimentally, we tested whether or not Wt1 was involved in the regulation of gp210. This was done in osteosarcoma (SAOS-2) cells using a model system in which Wt1 expression is controlled by tetracyclin. A modest downregulation of gp210 mRNA expression by Wt1 was noted. This means that the putative regulators of the gp210 gene remain to be identified (III). For a long time, gp210 and POM121 were the only known transmembrane proteins that localised to the pore membrane domain of the nuclear envelope. However, the identification and characterisation of the POMFIL1, POMFIL2 and POMFIL3 enlarged the group of integral membrane proteins associated with the nuclear pore (Coy et al., 2002). These novel proteins are regulated in a tissue and developmental specific manner. It is thus gradually becoming evident that variations of NPC configuration occur in metazoans, depending on cell type. To our knowledge, this possibility was first suggested based on the mRNA data for gp210 (I). Receptor mediated transport of proteins can be targeted for regulation at several steps in their movement across the nuclear envelope. In most known cases, the cargo itself is modified in a way that affects its interaction with import or export receptors. Tethering of the cargo-receptor complex to immobile cellular components or regulation of the soluble 32 transport machinery are other examples of how the cell is able to control a variety of fundamental processes (for a comprehensive review on this subject, see Kaffman and O'Shea, 1999). Modifications of the NPC by cell-type specific nucleoporins could confer an additional level of regulation of protein localization. Antibodies directed to the cytoplasmic tail of gp210 arrests pore formation at an early stage when added to Xenopus nuclear assembly extracts (Drummond and Wilson, 2002). The apparent lack of gp210 protein in many cell types does not fit with the view of gp210 as a crucial mediator of one or several discrete steps in early NPC biogenesis. The abundance of gp210 has been calculated to 16-25 copies per nuclear pore representing 5-6 Mda in some cell types (Gerace et al., 1982). Based on these calculations, it has been suggested that this protein may be a major component of the lumenal spoke domain of the nuclear pore complex (Wozniak et al., 1989; Yang et al., 1998). In vitro, it was shown that gp210 is able to self-associate into dimers and that these dimers have the capability to form higher orders of oligomerisation. Large arrays of gp210 could in vivo provide a basic framework that organises into a circular lumenal skeleton for the pore membrane (Favreau et al., 2001). Since our data (I, II) indicate high amounts of gp210 in epithelial tissues compared to non-detectable levels in surrounding cells, it is tempting to speculate that the extent of the lumenal skeleton vary in a cell type specific manner. This variation could in turn influence both the pore and the NPC itself in multiple ways. Inhibition of signal mediated transport of proteins into the nucleus as well as passive diffusion through the pore complex upon depletion of calcium from the lumen of the endoplasmic reticulum and nuclear envelope has been reported (Greber and Gerace, 1995). Similar effects on passive and active transport have been shown by introducing an antibody, targeting a luminal epitope of gp210, into the perinuclear cisterna (Greber and Gerace, 1992). These results provide direct functional evidence for conformational flexibility of the pore complex. Since the amino-terminal of gp210 contains a putative EF hand Ca2+-binding domain (Greber et al., 1990), it has been speculated that variations in calcium levels may trigger rearrangements within the lumenal domain and result in conformational changes in the inner spoke ring and central transporter (Perez-Terzic et al., 1997; Perez-Terzic et al., 1996). At present, the physiological relevance of conformational changes within the NPC is unknown but they could play an important role in regulation of specific states of the cell cycle or differentiation (Feldherr et al., 1994; Feldherr and Akin, 1993). Our findings revealed that the expression of gp210 is most prominent in epithelial cells of the mouse embryo (I, II). These data do not directly address the function of gp210. However, an intriguing possibility is that an enhanced expression of gp210 induces a reversible or nonreversible conformational change in the NPC with subsequent effects on cell differentiation. The transition of progenitor metanephric mesenchymal cells into cells committed to terminal differentiation is accompanied by a distinct up regulation of gp210. Gene regulation 33 of other nucleoporins is not well documented but has in a few cases revealed interesting information. Nup88 has been shown to be a marker for metastaticy in cutaneous melanoma cell lines (Zhang et al., 2002), and to be selectively over expressed, as compared to nup214 and nup153, in several other malignant tumours (Gould et al., 2000; Gould et al., 2002). Nup88 may, therefore, be considered as a putative marker for tumour growth and is probably related to increased cell cycling. Apart from being structural components of the NPC, nup96 and nup98 are also associated with the nuclear interior. Both proteins are expressed from a single gene, which are activated by INF-g in a STAT1 dependent manner. Nup98 is a target of the vesicular stomatitis virus matrix protein-mediated inhibition of mRNA nuclear export and it has been reported that up regulation of 98, induced by INF-g constitutes a mechanism that reverses M protein-mediated inhibition of gene expression (Enninga et al., 2002). These examples clearly demonstrate that studies of mechanisms that regulate nucleoporin genes could provide functional aspects of this group of proteins. The promoter characterisation of the human and mouse gp210 genes (III) was a first step towards an identification of one or several transcription factor utilised in the activation of this gene. 34 FUTURE PERSPECTIVES Our data presented in paper I-III do not rule out the possibility that all cells in the mouse express some gp210, but it is evident that the expression levels vary drastically depending on the cell type in multicellular organisms. It would therefore be of interest to examine if gp210 is fundamental for cell survival. This could be performed by experimental overexpression, or by gene targeting experiments in mice. Expression of gp210 is drastically increased when mesenchymal cells are converted to epithelial cells. This regulation might reflect a dynamic change of the NPC as the cells acquire more specialised functions. Is gp210 merely a marker for epithelial cells, or does it have functions not needed for other cell types? One approach that eventually could answer this question is to induce high expression of gp210 in a cell line, which shows low levels of this protein. The role of gp210 in nucleocytoplasmic transport, NPC conformation, and the cellular phenotype could be analyzed. The promoter sequence of gp210 contains several putiative binding sites for cell specific regulator elements. By using reporter gene assays, gel mobility shift assays and antibodies directed to putative regulators, it would be possible to clarify the role of such factors. POPULÄRVETENSKAPLIG SAMMANFATTNING PÅ SVENSKA Det nukleära höljet består av ett yttre och ett inre membran. Tillsammans begränsar de den nukleära regionen i kärnförsedda eukaryota celler och ger upphov till en funktionell miljö för dess kromosomer. Separationen av DNA-replikation och RNA-transkription i kärnan från den cytoplasmatiska protein-translationen ger eukaryota celler en mer sofistikerad kontroll över genernas uttryck än vad som är möjlig hos prokaryoter. Kontrollen förutsätter dock att proteiner och RNA har möjlighet att passera den barriär membranen utgör. Nukleära porkomplex (NPK) är stora proteinkomplex (125 mDa) som i elektronmikroskopi uppvisar en åttafaldig symmetri runt en, i förhållande till det nukleära höljet, vertikal axel. De ligger inbäddade i de nukleära membranen och bildar rörliknande strukturer som förbinder cytoplasma och kärna. Kanalerna är den enda kända förbindelselänken mellan dessa cellulära rum. NPK har även påträffats i cytoplasmatiska membran (s.k. Annulate lamellae), men där är deras funktion ännu inte känd. Man har uppskattat att det i porkomplexen ingår 30 olika protein (nukleoporin, nup), varav ett tjugotal är karakteriserade. Det pågår en intensiv forskning för att utreda den specifika funktionen för vart och ett och för att identifiera nya. 35 Import av proteiner till kärnan styrs av ett antal signalsekvenser varav den klassiska nukleära lokaliseringssignalen (NLS) och importsignalen M9 är de bäst karaktäriserade. På motsvarande sätt styrs RNA ut i cytosolen. Den nukleära export signalen (NES) är en leucinrik aminosyra-sekvens som ofta återfinns på RNA bindande protein. Lösliga receptorer tillhörande genfamiljen b-kayoferiner binder till signalerna och lotsar därefter substratet fram till och igenom porkomplexen. Den här processen är energikrävande. Vissa protein transporteras inte konstitutivt, utan är beroende av posttranslationella modifieringar för sin cellulära distribuering. Selektiv transport av proteiner genom NPK är på så vis ett sätt för cellen att kontrollera genuttryck, och därmed en av flera mekanismer som gör det möjligt för cellen att reglera exempelvis proliferation och differentiering. Den nu rådande modellen beskriver NPK som en homogen, statisk struktur, utan direkt betydelse för selektiviteten under den nukleocytoplasmatiska trafiken. Det finns dock rapporter som tyder på motsatsen och visar att det nukleära porkomplexet är mer dynamiskt än vad som tidigare antagits. Nukleoporin gp210 (glykoprotein 210) är ett integralt membranprotein med en mindre cytoplasmatisk karboxyterminal del, en transmembranregion samt en stor lumenal aminoände. Proteinet är väl karakteriserat och dess funktion har länge associerats med biogenes och stabilitet av NPK. Mina data visar att gp210 uppregleras under en tidig fas av njurutvecklingen. Expressionsstudier på musfoster påvisade även ett begränsat uttrycksmönster, som förutom i njure även i hud, lever och lunga sammanföll med framväxande epitelcellers. Studien genomfördes initialt på RNA-nivå men konfirmerades på proteinnivå med hjälp av en antikropp riktad mot den cytoplasmatiska domänen av gp210. Vidare tyder våra in-vitro data på att ett stort antal celltyper helt saknar gp210. Jämförelser med andra karakteriserade nukleoporiner tyder på att distributionen av gp210 är unik. Den begränsade utbredningen av gp210 under den embryonala utvecklingen och i cellinjer visar att det existerar olika typer av NPK. Gp210 kan alltså ha en annan funktion än den tidigare föreslagna. Länge var gp210 tillsammans med POM121 (POre Membrane protein 121) de enda karakteriserade membranprotein som var lokaliserade till NPK. Så är inte längre fallet. Nyligen isolerades en familj (POMFIL1, POMFIL2 och POMFIL3) av gener med tydliga homologier till POM121. Presenterade data tyder på att åtminstone POMFIL1 och POMFIL2 är membranbundna nukleoporin med ett, liksom gp210, differentiellt uttrycksmönster under embryogenesen. Dessa resultat styrker våra teorier om att det förekommer variationer i sammansättningen av NPK, och att dessa variationer är vävnads och/eller celltyps beroende. Denna variation kan vara av stor betydelse för den molekylära trafiken mellan kärnan och cytoplasman. 36 ACKNOWLEDGMENTS Most of this work was carried out at the Department of Cell and Molecular Biology, Uppsala University and at the Department of Cell and Molecular Biology, Lund University. I would like to express my gratitude to all those that have helped me, in so many different ways, while working on this thesis. Thank you! Especially I would like to like to thank: Professor Peter Ekblom, my supervisor, for your scientific guidance and a constant moral support. Thank you for everything! Tord Hjalt, Donald Gullberg, Erik Forsberg and Ann-Marie Olofsson for taking their time to educate me in different aspects of molecular biology. Marja Ekblom, for letting me inherit ”clone 5” and for being such a helpful and friendly person. My co-authors. Siw Medborg, Agneta Ekman-Lundell and Sigrid Petterson, past and present secretaries. True competence in the interface with the real world. I wish you all luck! Madde, Teet, Calle, Kati, Maria R., Katta, Lotta, Jan, Elke, Ulrika, Hong, Yu-Chen, Hao, Yang, Maria W., Alexander, Fredrik E., Wilhelm, Kinga and Wilma, my colleagues through the years, for all memorable moments. It has been a pleasure to work with you all. Susanne and Maria F., my roommates. The two of you improved Lund! Mats, for years of great friendship and encouragement. Peter and Elik, my good friends, for sharing pain and pleasure during the BMC years, and for being such good losers in badminton. All other old ”biological” friends, especially Henrik Pärn, to whom I also owe an excuse. I couldn’t wait any longer! Erik Lindqvist and his family, for their generous hospitality. My family, always reliable. And Sara, for your love and support. 37 LITERATURE CITED Adams, M. D., Kerlavage, A. R., Fleischmann, R. D., Fuldner, R. A., Bult, C. J., Lee, N. H., Kirkness, E. F., Weinstock, K. G., Gocayne, J. D., White, O., and et al. (1995). Initial assessment of human gene diversity and expression patterns based upon 83 million nucleotides of cDNA sequence. Nature 377, 3-174. Aitchison, J. D., Rout, M.P., Marelli, M., Blobel, G., and Wozniak R. W. (1995) Two novel related yeast nucleoporins Nup170p and Nup157p: complementation with the vertebrate homologue Nup155p and functional interactions with the yeast nuclear pore-membrane protein Pom152p. J. Cell Biol. 131, 1133-1148. Akey, C. W., and Radermacher, M. (1993). Architecture of the Xenopus nuclear pore complex revealed by three- dimensional cryo-electron microscopy. J. Cell Biol. 122, 1-19. Allen, N. P., Huang, L., Burlingame, A., and Rexach, M. (2001). Proteomic analysis of nucleoporin interacting proteins. J. Biol. Chem. 276, 29268-29274. Arts, G. J., Fornerod, M., and Mattaj, I. W. (1998). Identification of a nuclear export receptor for tRNA. Curr. Biol. 8, 305-314. Bachi, A., Braun, I. C., Rodrigues, J. P., Pante, N., Ribbeck, K., von Kobbe, C., Kutay, U., Wilm, M., Gorlich, D., Carmo-Fonseca, M., and Izaurralde, E. (2000). The C-terminal domain of TAP interacts with the nuclear pore complex and promotes export of specific CTE-bearing RNA substrates. Rna 6, 136-158. Bangs, P., Burke, B., Powers, C., Craig, R., Purohit, A., and Doxsey, S. (1998). Functional analysis of Tpr: identification of nuclear pore complex association and nuclear localization domains and a role in mRNA export. J. Cell Biol. 143, 1801-1812. Barasch, J., Yang, J., Ware, C. B., Taga, T., Yoshida, K., Erdjument-Bromage, H., Tempst, P., Parravicini, E., Malach, S., Aranoff, T., and Oliver, J. A. (1999). Mesenchymal to epithelial conversion in rat metanephros is induced by LIF. Cell 99, 377-386. Bastos, R., Ribas de Pouplana, L., Enarson, M., Bodoor, K., and Burke, B. (1997). Nup84, a novel nucleoporin that is associated with CAN/Nup214 on the cytoplasmic face of the nuclear pore complex. J. Cell Biol. 137, 989-1000. Belgareh, N., Rabut, G., Bai, S. W., van Overbeek, M., Beaudouin, J., Daigle, N., Zatsepina, O. V., Pasteau, F., Labas, V., Fromont-Racine, M., Ellenberg, J., and Doye, V. (2001). An evolutionarily conserved NPC subcomplex, which redistributes in part to kinetochores in mammalian cells. J. Cell Biol. 154, 1147-1160. 38 Berrios, M., Meller, V. H., McConnell, M., and Fisher, P. A. (1995). Drosophila gp210, an invertebrate nuclear pore complex glycoprotein. Eur. J. Cell Biol. 76, 1-7. Bischoff, F. R., and Gorlich, D. (1997). RanBP1 is crucial for the release of RanGTP from importin beta-related nuclear transport factors. FEBS Lett. 419, 249-254. Bischoff, F. R., Klebe, C., Kretschmer, J., Wittinghofer, A., and Ponstingl, H. (1994). RanGAP1 induces GTPase activity of nuclear Ras-related Ran. Proc. Natl. Acad. Sci. USA 91, 2587-2591. Bischoff, F. R., and Ponstingl, H. (1991). Catalysis of guanine nucleotide exchange on Ran by the mitotic regulator RCC1. Nature 354, 80-82. Black, B. E., Holaska, J. M., Levesque, L., Ossareh-Nazari, B., Gwizdek, C., Dargemont, C., and Paschal, B. M. (2001). NXT1 is necessary for the terminal step of Crm1-mediated nuclear export. J. Cell Biol. 152, 141-155. Black, B. E., Levesque, L., Holaska, J. M., Wood, T. C., and Paschal, B. M. (1999). Identification of an NTF2-related factor that binds Ran-GTP and regulates nuclear protein export. Mol. Cell Biol. 19, 8616-8624. Bodoor, K., Shaikh, S., Salina, D., Raharjo, W. H., Bastos, R., Lohka, M., and Burke, B. (1999). Sequential recruitment of NPC proteins to the nuclear periphery at the end of mitosis. J. Cell Sci. 112, 2253-2264. Bonne, G., Di Barletta, M. R., Varnous, S., Becane, H. M., Hammouda, E. H., Merlini, L., Muntoni, F., Greenberg, C. R., Gary, F., Urtizberea, J. A., et al. (1999). Mutations in the gene encoding lamin A/C cause autosomal dominant Emery- Dreifuss muscular dystrophy. Nat. Genet. 21, 285-288. Bucci, M., and Wente, S. R. (1997). In vivo dynamics of nuclear pore complexes in yeast. J. Cell Biol. 136, 1185-1199. Buendia, B., and Courvalin, J. C. (1997). Domain-specific disassembly and reassembly of nuclear membranes during mitosis. Exp Cell Res 230, 133-144. Burrow, C. R. (2000). Regulatory molecules in kidney development. Pediatr. Nephrol. 14, 240-253. Chaudhary, N., and Courvalin, J. C. (1993). Stepwise reassembly of the nuclear envelope at the end of mitosis. J. Cell Biol. 122, 295-306. 39 Chial, H. J., Rout, M. P., Giddings, T. H., and Winey, M. (1998). Saccharomyces cerevisiae Ndc1p is a shared component of nuclear pore complexes and spindle pole bodies. J. Cell Biol. 143, 1789-1800. Cingolani, G., Petosa, C., Weis, K., and Muller, C. W. (1999). Structure of importin-beta bound to the IBB domain of importin-alpha. Nature 399, 221-229. Clements, L., Manilal, S., Love, D. R., and Morris, G. E. (2000). Direct interaction between emerin and lamin A. Biochem. Biophys. Res. Commun. 267, 709-714. Cohen, M., Wilson, K. L., Gruenbaum, Y. (2001). Membrane proteins of the nuclear pore complex: Gp210 is conserved in Drosophila, C. elegans and A. thaliana. Gene Ther. Mol. Biol. 6, 47-55. Cordes, V. C., Rackwitz, H. R., and Reidenbach, S. (1997a). Mediators of nuclear protein import target karyophilic proteins to pore complexes of cytoplasmic annulate lamellae. Exp. Cell. Res. 237, 419-433. Cordes, V. C., Reidenbach, S., Rackwitz, H. R., and Franke, W. W. (1997b). Identification of protein p270/Tpr as a constitutive component of the nuclear pore complex-attached intranuclear filaments. J. Cell Biol. 136, 515-529. Coy, J. F., Wiemann, S., Bechmann, I., Bachner, D., Nitsch, R., Kretz, O., Christiansen, H., and Poustka, A. (2002). Pore membrane and/or filament interacting like protein 1 (POMFIL1) is predominantly expressed in the nervous system and encodes different protein isoforms. Gene 290, 73-94. Cronshaw, J. M., Krutchinsky, A. N., , Zhang, W., , Chait, B. T., , and Matunis, M. J. (2002). Proteomic analysis of the mammalian nuclear pore complex. J. Cell Biol. 158, 915-927. Daigle, N., Beaudouin, J., Hartnell, L., Imreh, G., Hallberg, E., Lippincott-Schwartz, J., and Ellenberg, J. (2001). Nuclear pore complexes form immobile networks and have a very low turnover in live mammalian cells. J. Cell Biol. 154, 71-84. Daneholt, B. (2001). Assembly and transport of a premessenger RNP particle. Proc. Natl. Acad. Sci. USA 98, 7012-7017. Delphin, C., Guan, T., Melchior, F., and Gerace, L. (1997). RanGTP targets p97 to RanBP2, a filamentous protein localized at the cytoplasmic periphery of the nuclear pore complex. Mol. Biol. Cell 8, 2379-2390. Dressler, G. R. (1999). Kidney development branches out. Dev. Genet. 24, 189-193. 40 Drubin, D. G., and Nelson, W. J. (1996) Origins of Cell Polarity. Cell 84, 335–344. Drummond, S. P., and Wilson, K. L. (2002). Interference with the cytoplasmic tail of gp210 disrupts "close apposition" of nuclear membranes and blocks nuclear pore dilation. J. Cell Biol. 158, 53-62. Durbeej, M., Larsson, E., Ibraghimov-Beskrovnaya, O., Roberds, S. L., Campbell, K. P., and Ekblom, P. (1995). Non-muscle alpha-dystroglycan is involved in epithelial development. J. Cell Biol. 130, 79-91. Dworkin, M. B., and Dawid, I. B. (1980). Use of a cloned library for the study of abundant poly(A)+RNA during Xenopus laevis development. Dev. Biol. 76, 449-464. Ekblom, P., Ekblom, M., Fecker, L., Klein, G., Zhang, H. Y., Kadoya, Y., Chu, M. L., Mayer, U., and Timpl, R. (1994). Role of mesenchymal nidogen for epithelial morphogenesis in vitro. Development 120, 2003-2014. Enninga, J., Levy, D. E., Blobel, G., and Fontoura, B. M. (2002). Role of nucleoporin induction in releasing an mRNA nuclear export block. Science 295, 1523-1525. Erickson, R. A. (1968). Inductive interactions in the develoment of the mouse metanephros. J. Exp. Zool. 169, 33-42. Ewald, A., Kossner, U., Scheer, U., and Dabauvalle, M. C. (1996). A biochemical and immunological comparison of nuclear and cytoplasmic pore complexes. Journal of Cell Science 109, 1813-1824. Fabre, E., and Hurt, E. (1997). Yeast genetics to dissect the nuclear pore complex and nucleocytoplasmic trafficking. Annu. Rev. Genet. 31, 277-313. Fahrenkrog, B., Hurt, E. C., Aebi, U., and Pante, N. (1998). Molecular architecture of the yeast nuclear pore complex: localization of Nsp1p subcomplexes. J. Cell Biol. 143, 577-588. Fairley, E. A., Kendrick-Jones, J., and Ellis, J. A. (1999). The Emery-Dreifuss muscular dystrophy phenotype arises from aberrant targeting and binding of emerin at the inner nuclear membrane. J. Cell Sci. 112, 2571-2582. Falk, M., Salmivirta, K., Durbeej, M., Larsson, E., Ekblom, M., Vestweber, D., and Ekblom, P. (1996). Integrin alpha 6B beta 1 is involved in kidney tubulogenesis in vitro. J. Cell Sci. 109, 2801-2810. 41 Farjot, G., Sergeant, A., and Mikaelian, I. (1999). A new nucleoporin-like protein interacts with both HIV-1 Rev nuclear export signal and CRM-1. J. Biol. Chem. 274, 17309-17317. Favreau, C., Bastos, R., Cartaud, J., Courvalin, J. C., and Mustonen, P. (2001). Biochemical characterization of nuclear pore complex protein gp210 oligomers. Eur. J. Biochem. 268, 3883-3889. Favreau, C., Worman, H. J., Wozniak, R. W., Frappier, T., and Courvalin, J. C. (1996). Cell cycle-dependent phosphorylation of nucleoporins and nuclear pore membrane protein Gp210. Biochemistry 35, 8035-8044. Feldherr, C., Cole, C., Lanford, R. E., and Akin, D. (1994). The effects of SV40 large T antigen and p53 on nuclear transport capacity in BALB/c 3T3 cells. Exp. Cell Res. 213, 164171. Feldherr, C. M., and Akin, D. (1990). The permeability of the nuclear envelope in dividing and nondividing cell cultures. J. Cell Biol. 111, 1-8. Feldherr, C. M., and Akin, D. (1993). Regulation of nuclear transport in proliferating and quiescent cells. Exp. Cell Res. 205, 179-186. Feldherr, C. M., and Akin, D. (1997). The location of the transport gate in the nuclear pore complex. J. Cell Sci. 110, 3065-3070. Finlay, D. R., Meier, E., Bradley, P., Horecka, J., and Forbes, D. J. (1991). A complex of nuclear pore proteins required for pore function. J. Cell Biol. 114, 169-183. Fischer, U., Huber, J., Boelens, W. C., Mattaj, I. W., and Luhrmann, R. (1995). The HIV-1 Rev activation domain is a nuclear export signal that accesses an export pathway used by specific cellular RNAs. Cell 82, 475-483. Foisner, R., and Gerace, L. (1993). Integral membrane proteins of the nuclear envelope interact with lamins and chromosomes, and binding is modulated by mitotic phosphorylation. Cell 73, 1267-1279. Fontoura, B. M., Blobel, G., and Matunis, M. J. (1999). A conserved biogenesis pathway for nucleoporins: proteolytic processing of a 186-kilodalton precursor generates Nup98 and the novel nucleoporin, Nup96. J. Cell Biol. 144, 1097-1112. Fornerod, M., Boer, J., van Baal, S., Morreau, H., and Grosveld, G. (1996). Interaction of cellular proteins with the leukemia specific fusion proteins DEK-CAN and SET-CAN and their normal counterpart, the nucleoporin CAN. Oncogene 13, 1801-1808. 42 Fornerod, M., van Deursen, J., van Baal, S., Reynolds, A., Davis, D., Murti, K. G., Fransen, J., and Grosveld, G. (1997). The human homologue of yeast CRM1 is in a dynamic subcomplex with CAN/Nup214 and a novel nuclear pore component Nup88. EMBO J. 16, 807-816. Fricker, M., Hollinshead, M., White, N., and Vaux, D. (1997). Interphase nuclei of many mammalian cell types contain deep, dynamic, tubular membrane-bound invaginations of the nuclear envelope. J. Cell Biol. 136, 531-544. Frosst, P., Guan, T., Subauste, C., Hahn, K., and Gerace, L. (2002). Tpr is localized within the nuclear basket of the pore complex and has a role in nuclear protein export. J. Cell Biol. 156, 617-630. Fukuda, M., Asano, S., Nakamura, T., Adachi, M., Yoshida, M., Yanagida, M., and Nishida, E. (1997). CRM1 is responsible for intracellular transport mediated by the nuclear export signal. Nature 390, 308-311. Gajewski, A., Lourim, D., and Krohne, G. (1996). An antibody against a glycosylated integral membrane protein of the Xenopus laevis nuclear pore complex: a tool for the study of pore complex membranes. Eur. J. Cell Biol. 71, 14-21. Gant, T. M., and Wilson, K. L. (1997). Nuclear assembly. Annu. Rev. Cell Dev. Biol. 13, 669-695. Gerace, L., Ottaviano, Y., and Kondor-Koch, C. (1982). Identification of a major polypeptide of the nuclear pore complex. J. Cell Biol. 95, 826-837. Goldberg, M. W., and Allen, T. D. (1993). The nuclear pore complex: three-dimensional surface structure revealed by field emission, in-lens scanning electron microscopy, with underlying structure uncovered by proteolysis. J. Cell Sci. 106, 261-274. Goldberg, M. W., and Allen, T. D. (1996). The nuclear pore complex and lamina: threedimensional structures and interactions determined by field emission in-lens scanning electron microscopy. J. Mol. Biol. 257, 848-865. Goldberg, M. W., Wiese, C., Allen, T. D., and Wilson, K. L. (1997). Dimples, pores, starrings, and thin rings on growing nuclear envelopes: evidence for structural intermediates in nuclear pore complex assembly. J. Cell Sci. 110, 409-420. Goldman, R. D., Gruenbaum, Y., Moir, R. D., Shumaker, D. K., and Spann, T. P. (2002). Nuclear lamins: building blocks of nuclear architecture. Genes Dev. 16, 533-547. 43 Gorlich, D., Henklein, P., Laskey, R. A., and Hartmann, E. (1996a). A 41 amino acid motif in importin-alpha confers binding to importin- beta and hence transit into the nucleus. EMBO J. 15, 1810-1817. Gorlich, D., Pante, N., Kutay, U., Aebi, U., and Bischoff, F. R. (1996b). Identification of different roles for RanGDP and RanGTP in nuclear protein import. EMBO J. 15, 5584-5594. Gould, V. E., Martinez, N., Orucevic, A., Schneider, J., and Alonso, A. (2000). A novel, nuclear pore-associated, widely distributed molecule overexpressed in oncogenesis and development. Am. J. Pathol. 157, 1605-1613. Gould, V. E., Orucevic, A., Zentgraf, H., Gattuso, P., Martinez, N., and Alonso, A. (2002). Nup88 (karyoporin) in human malignant neoplasms and dysplasias: correlations of immunostaining of tissue sections, cytologic smears, and immunoblot analysis. Hum. Pathol. 33, 536-544. Grandi, P., Dang, T., Pane, N., Shevchenko, A., Mann, M., Forbes, D., and Hurt, E. (1997). Nup93, a vertebrate homologue of yeast Nic96p, forms a complex with a novel 205-kDa protein and is required for correct nuclear pore assembly. Mol. Biol. Cell 8, 2017-2038. Greber, U. F., and Gerace, L. (1992). Nuclear protein import is inhibited by an antibody to a lumenal epitope of nuclear pore complex glycoprotein. J. Cell Biol. 116, 15-30. Greber, U. F., and Gerace, L. (1995). Depletion of calcium from the lumen of endoplasmic reticulum reversibly inhibits passive diffusion and signal-mediated transport into the nucleus. J. Cell Biol. 128, 5-14. Greber, U. F., Senior, A., and Gerace, L. (1990). A major glycoprotein of the nuclear pore complex is a membrane-spanning polypeptide with a large lumenal domain and a small cytoplasmic tail. EMBO J. 9, 1495-1502. Grobstein, C. (1953). Inductive epithelio-mesenchymal interactions in cultured organ rudiments of the mouse. Science 118, 52-55. Grobstein, C. (1956). Trans-filter induction of tubules in mouse metanephric mesenchyme. Exp. Cell Res. 10, 424-440. Gruenwald, P. (1943). Stimulation of nephrogenic tissues by normal an abnormal inductors. Anat. rec. 86, 321-335. 44 Gruter, P., Tabernero, C., von Kobbe, C., Schmitt, C., Saavedra, C., Bachi, A., Wilm, M., Felber, B. K., and Izaurralde, E. (1998). TAP, the human homolog of Mex67p, mediates CTE-dependent RNA export from the nucleus. Mol. Cell. 1, 649-659. Guan, T., Kehlenbach, R. H., Schirmer, E. C., Kehlenbach, A., Fan, F., Clurman, B. E., Arnheim, N., and Gerace, L. (2000). Nup50, a nucleoplasmically oriented nucleoporin with a role in nuclear protein export. Mol. Cell Biol. 20, 5619-5630. Guan, T., Muller, S., Klier, G., Pante, N., Blevitt, J. M., Haner, M., Paschal, B., Aebi, U., and Gerace, L. (1995). Structural analysis of the p62 complex, an assembly of O-linked glycoproteins that localizes near the central gated channel of the nuclear pore complex. Mol. Biol. Cell 6, 1591-1603. Hallberg, E., Wozniak, R. W., and Blobel, G. (1993). An integral membrane protein of the pore membrane domain of the nuclear envelope contains a nucleoporin-like region. J. Cell Biol. 122, 513-521. Herold, A., Suyama, M., Rodrigues, J. P., Braun, I. C., Kutay, U., Carmo-Fonseca, M., Bork, P., and Izaurralde, E. (2000). TAP (NXF1) belongs to a multigene family of putative RNA export factors with a conserved modular architecture. Mol. Cell Biol. 20, 8996-9008. Herzlinger, D., Qiao, J., Cohen, D., Ramakrishna, N., and Brown, A. M. (1994). Induction of kidney epithelial morphogenesis by cells expressing Wnt-1. Dev. Biol. 166, 815-818. Hinshaw, J. E., Carragher, B. O., and Milligan, R. A. (1992). Architecture and design of the nuclear pore complex. Cell 69, 1133-1141. Hu, T., and Gerace, L. (1998). cDNA cloning and analysis of the expression of nucleoporin p45. Gene 221, 245-253. Hu, T., Guan, T., and Gerace, L. (1996). Molecular and functional characterization of the p62 complex, an assembly of nuclear pore complex glycoproteins. J Cell Biol 134, 589-601. Huber, C. (1905). On the development and shape of uriniferous tubules in certain higher mammals. Am. J. Anat. 4, 1-98. Imreh, G., and Hallberg, E. (2000). An integral membrane protein from the nuclear pore complex is also present in the annulate lamellae: implications for annulate lamella formation. Exp. Cell Res. 259, 180-190. 45 Izaurralde, E., Kutay, U., von Kobbe, C., Mattaj, I. W., and Gorlich, D. (1997). The asymmetric distribution of the constituents of the Ran system is essential for transport into and out of the nucleus. EMBO J. 16, 6535-6547. Izaurralde, E., Lewis, J., Gamberi, C., Jarmolowski, A., McGuigan, C., and Mattaj, I. W. (1995). A cap-binding protein complex mediating U snRNA export. Nature 376, 709-712. Jans, D. A., Xiao, C. Y., and Lam, M. H. (2000). Nuclear targeting signal recognition: a key control point in nuclear transport? Bioessays 22, 532-544. Jarnik, M., and Aebi, U. (1991). Toward a more complete 3-D structure of the nuclear pore complex. J. Struct. Biol. 107, 291-308. Kaffman, A., and O'Shea, E. K. (1999). Regulation of nuclear localization: a key to a door. Annu. Rev. Cell Dev. Biol. 15, 291-339. Kalderon, D., Roberts, B. L., Richardson, W. D., and Smith, A. E. (1984). A short amino acid sequence able to specify nuclear location. Cell 39, 499-509. Kessel, R. G. (1992). Annulate lamellae: a last frontier in cellular organelles. Int. Rev. Cytol. 133, 43-120. Kiesler, E., Miralles, F., Visa, N. (2002) HEL/UAP56 binds cotranscriptionally to the Balbiani ring pre-mRNA in an intron-independent manner and accompanies the BR mRNP to the nuclear pore. Curr. Biol. 12, 859-862. Kiseleva, E., Goldberg, M. W., Allen, T. D., and Akey, C. W. (1998). Active nuclear pore complexes in Chironomus: visualization of transporter configurations related to mRNP export. J. Cell Sci. 111, 223-236. Knust E. (2000) Control of epithelial cell shape and polarity. Curr. Opin. Genet. Dev. 10, 471-475 Kohler, M., Speck, C., Christiansen, M., Bischoff, F. R., Prehn, S., Haller, H., Gorlich, D., and Hartmann, E. (1999). Evidence for distinct substrate specificities of importin alpha family members in nuclear protein import. Mol. Cell Biol. 19, 7782-7791. Kraemer, D., Wozniak, R. W., Blobel, G., and Radu, A. (1994). The human CAN protein, a putative oncogene product associated with myeloid leukemogenesis, is a nuclear pore complex protein that faces the cytoplasm. Proc. Natl. Acad. Sci. USA 91, 1519-1523. 46 Kreidberg, J. A., Sariola, H., Loring, J. M., Maeda, M., Pelletier, J., Housman, D., and Jaenisch, R. (1993). WT-1 is required for early kidney development. Cell 74, 679-691. Kuersten, S., Ohno, M., and Mattaj, I. W. (2001). Nucleocytoplasmic transport: Ran, beta and beyond. Trends. Cell Biol. 11, 497-503. Kutay, U., Bischoff, F. R., Kostka, S., Kraft, R., and Gorlich, D. (1997). Export of importin alpha from the nucleus is mediated by a specific nuclear transport factor. Cell 90, 1061-1071. Kutay, U., Lipowsky, G., Izaurralde, E., Bischoff, F. R., Schwarzmaier, P., Hartmann, E., and Gorlich, D. (1998). Identification of a tRNA-specific nuclear export receptor. Mol. Cell 1, 359-369. Liang, P., and Pardee, A. B. (1992). Differential display of eukaryotic messenger RNA by means of the polymerase chain reaction. Science 257, 967-971. Lindsay, M., Plafker, K., Smith, A., Clurman, B., and Macara, I. (2002). Npap60/Nup50 Is a Tri-Stable Switch that Stimulates Importin-alpha:beta- Mediated Nuclear Protein Import. Cell 110, 349. Lockhart, D. J., Dong, H., Byrne, M. C., Follettie, M. T., Gallo, M. V., Chee, M. S., Mittmann, M., Wang, C., Kobayashi, M., Horton, H., and Brown, E. L. (1996). Expression monitoring by hybridization to high-density oligonucleotide arrays. Nat. Biotechnol. 14, 1675-1680. Lohka, M. J., and Maller, J. L. (1985). Induction of nuclear envelope breakdown, chromosome condensation, and spindle formation in cell-free extracts. J. Cell Biol. 101, 518523. Macaulay, C., Meier, E., and Forbes, D. J. (1995). Differential mitotic phosphorylation of proteins of the nuclear pore complex. J. Biol. Chem. 270, 254-262. Michael, W. M., Choi, M., and Dreyfuss, G. (1995). A nuclear export signal in hnRNP A1: a signal-mediated, temperature- dependent nuclear protein export pathway. Cell 83, 415-422. Miller, B. R., and Forbes, D. J. (2000). Purification of the vertebrate nuclear pore complex by biochemical criteria. Traffic 1, 941-951. Miller, B. R., Powers, M., Park, M., Fischer, W., and Forbes, D. J. (2000). Identification of a new vertebrate nucleoporin, Nup188, with the use of a novel organelle trap assay. Mol Biol Cell 11, 3381-3396. 47 Miller, M. W., Caracciolo, M. R., Berlin, W. K., and Hanover, J. A. (1999). Phosphorylation and glycosylation of nucleoporins. Arch. Biochem. Biophys. 367, 51-60. Moir, R. D., Spann, T. P., and Goldman, R. D. (1995). The dynamic properties and possible functions of nuclear lamins. Int. Rev. Cytol. 141-182. Moore, M. S., and Blobel, G. (1993). The GTP-binding protein Ran/TC4 is required for protein import into the nucleus. Nature 365, 661-663. Nachury, M. V., and Weis, K. (1999). The direction of transport through the nuclear pore can be inverted. Proc. Natl. Acad. Sci. USA 96, 9622-9627. Nakielny, S. , Siomi, M.C. , Siomi, H. , Michael, W.M. , Pollard, V. and Dreyfuss, G. (1996). Transportin: nuclear transport receptor of a novel nuclear protein import pathway. Exp. Cell Res. 229, 261-266. Osathanondh, V., and Potter, E. (1963). Development of human kidney as shown by microdissection. III. Formation and interrelationships of collecting tubules and nephrons. Arch. Pathol. 76, 290-302. Ossareh-Nazari, B., Bachelerie, F., and Dargemont, C. (1997). Evidence for a role of CRM1 in signal-mediated nuclear protein export. Science 278, 141-144. Ostlund, C., Ellenberg, J., Hallberg, E., Lippincott-Schwartz, J., and Worman, H. J. (1999). Intracellular trafficking of emerin, the Emery-Dreifuss muscular dystrophy protein. J. Cell Sci. 112, 1709-1719. Pachnis, V., Mankoo, B., and Costantini, F. (1993). Expression of the c-ret proto-oncogene during mouse embryogenesis. Development 119, 1005-1017. Pante, N., and Aebi, U. (1996). Sequential binding of import ligands to distinct nucleopore regions during their nuclear import. Science 273, 1729-1732. Pante, N., Bastos, R., McMorrow, I., Burke, B., and Aebi, U. (1994). Interactions and threedimensional localization of a group of nuclear pore complex proteins. J. Cell Biol. 126, 603617. Patterson, L. T., Pembaur, M., and Potter, S. S. (2001). Hoxa11 and Hoxd11 regulate branching morphogenesis of the ureteric bud in the developing kidney. Development 128, 2153-2161. 48 Perez-Terzic, C., Jaconi, M., and Clapham, D. E. (1997). Nuclear calcium and the regulation of the nuclear pore complex. Bioessays 19, 787-792. Perez-Terzic, C., Pyle, J., Jaconi, M., Stehno-Bittel, L., and Clapham, D. E. (1996). Conformational states of the nuclear pore complex induced by depletion of nuclear Ca2+ stores. Science 273, 1875-1877. Pichel, J. G., Shen, L., Sheng, H. Z., Granholm, A. C., Drago, J., Grinberg, A., Lee, E. J., Huang, S. P., Saarma, M., Hoffer, B. J., et al. (1996). Defects in enteric innervation and kidney development in mice lacking GDNF. Nature 382, 73-76. Powers, M. A., Forbes, D. J., Dahlberg, J. E., and Lund, E. (1997). The vertebrate GLFG nucleoporin, Nup98, is an essential component of multiple RNA export pathways. J. Cell Biol. 136, 241-250. Radu, A., Blobel, G., and Wozniak, R. W. (1993). Nup155 is a novel nuclear pore complex protein that contains neither repetitive sequence motifs nor reacts with WGA. J. Cell Biol. 121, 1-9. Radu, A., Blobel, G., and Wozniak, R. W. (1994). Nup107 is a novel nuclear pore complex protein that contains a leucine zipper. J. Biol Chem. 269, 17600-17605. Radu, A., Moore, M. S., and Blobel, G. (1995). The peptide repeat domain of nucleoporin Nup98 functions as a docking site in transport across the nuclear pore complex. Cell 81, 215222. Raharjo, W. H., Enarson, P., Sullivan, T., Stewart, C. L., and Burke, B. (2001). Nuclear envelope defects associated with LMNA mutations cause dilated cardiomyopathy and EmeryDreifuss muscular dystrophy. J. Cell Sci. 114, 4447-4457. Reichelt, R., Holzenburg, A., Buhle, E. L., Jr., Jarnik, M., Engel, A., and Aebi, U. (1990). Correlation between structure and mass distribution of the nuclear pore complex and of distinct pore complex components. J. Cell Biol. 110, 883-894. Ribbeck, K., Lipowsky, G., Kent, H. M., Stewart, M., and Gorlich, D. (1998). NTF2 mediates nuclear import of Ran. EMBO J. 17, 6587-6598. Ris, H. (1997). High-resolution field-emission scanning electron microscopy of nuclear pore complex. Scanning 19, 368-375. 49 Ris, H., and Malecki, M. (1993). High-resolution field emission scanning electron microscope imaging of internal cell structures after Epon extraction from sections: a new approach to correlative ultrastructural and immunocytochemical studies. J. Struct. Biol. 111, 148-157. Robbins, J., Dilworth, S. M., Laskey, R. A., and Dingwall, C. (1991). Two interdependent basic domains in nucleoplasmin nuclear targeting sequence: identification of a class of bipartite nuclear targeting sequence. Cell 64, 615-623. Rober, R. A., Sauter, H., Weber, K., and Osborn, M. (1990). Cells of the cellular immune and hemopoietic system of the mouse lack lamins A/C: distinction versus other somatic cells. J. Cell Sci. 95, 587-598. Rober, R. A., Weber, K., and Osborn, M. (1989). Differential timing of nuclear lamin A/C expression in the various organs of the mouse embryo and the young animal: a developmental study. Development 105, 365-378. Rout, M. P., Aitchison, J. D., Suprapto, A., Hjertaas, K., Zhao, Y., and Chait, B. T. (2000). The yeast nuclear pore complex: composition, architecture, and transport mechanism. J. Cell Biol. 148, 635-651. Rout, M. P., and Blobel, G. (1993). Isolation of the yeast nuclear pore complex. J. Cell Biol. 123, 771-783. Rutherford, S. A., Goldberg, M. W., and Allen, T. D. (1997). Three-dimensional visualization of the route of protein import: the role of nuclear pore complex substructures. Exp. Cell Res. 232, 146-160. Santos-Rosa, H., Moreno, H., Simos, G., Segref, A., Fahrenkrog, B., Pante, N., and Hurt, E. (1998). Nuclear mRNA export requires complex formation between Mex67p and Mtr2p at the nuclear pores. Mol. Cell Biol. 18, 6826-6838. Saxen, L., Koskimies, O., Lahti, A., Miettinen, H., Rapola, J., and Wartiovaara, J. (1968). Differentiation of kidney mesenchyme in an experimental model system. Adv. Morphog. 7, 251-293. Schedl, A., and Hastie, N. D. (2000). Cross-talk in kidney development. Curr. Opin. Genet. Dev. 10, 543-549. Schena, M., Shalon, D., Davis, R. W., and Brown, P. O. (1995). Quantitative monitoring of gene expression patterns with a complementary DNA microarray. Science 270, 467-470. 50 Schuchardt, A., D'Agati, V., Larsson-Blomberg, L., Costantini, F., and Pachnis, V. (1994). Defects in the kidney and enteric nervous system of mice lacking the tyrosine kinase receptor Ret. Nature 367, 380-383. Shackleton, S., Lloyd, D. J., Jackson, S. N., Evans, R., Niermeijer, M. F., Singh, B. M., Schmidt, H., Brabant, G., Kumar, S., Durrington, P. N., et al. (2000). LMNA, encoding lamin A/C, is mutated in partial lipodystrophy. Nat. Genet. 24, 153-156. Shah, S., and Forbes, D. J. (1998). Separate nuclear import pathways converge on the nucleoporin Nup153 and can be dissected with dominant-negative inhibitors. Curr. Biol. 8, 1376-1386. Shah, S., Tugendreich, S., and Forbes, D. (1998). Major binding sites for the nuclear import receptor are the internal nucleoporin Nup153 and the adjacent nuclear filament protein Tpr. J. Cell Biol. 141, 31-49. Shibata, S., Matsuoka, Y., and Yoneda, Y. (2002). Nucleocytoplasmic transport of proteins and poly(A)+ RNA in reconstituted Tpr-less nuclei in living mammalian cells. Genes Cells 7, 421-434. Sherr, C. J., and Roberts, J. M. (1999). CDK inhibitors: positive and negative regulators of G1-phase progression. Genes Dev. 13, 1501-1512. Shulga, N., Mosammaparast, N., Wozniak, R., and Goldfarb, D. S. (2000). Yeast nucleoporins involved in passive nuclear envelope permeability. J. Cell Biol. 149, 1027-1038. Siomi, H., and Dreyfuss, G. (1995). A nuclear localization domain in the hnRNP A1 protein. J. Cell Biol. 129, 551-560. Sive, H. L., and St John, T. (1988). A simple subtractive hybridization technique employing photoactivatable biotin and phenol extraction. Nucleic Acids Res. 16, 10937. Smitherman, M., Lee, K., Swanger, J., Kapur, R., and Clurman, B. E. (2000). Characterization and targeted disruption of murine Nup50, a p27(Kip1)- interacting component of the nuclear pore complex. Mol. Cell Biol. 20, 5631-5642. Smythe, C., Jenkins, H. E., and Hutchison, C. J. (2000). Incorporation of the nuclear pore basket protein nup153 into nuclear pore structures is dependent upon lamina assembly: evidence from cell- free extracts of Xenopus eggs. EMBO J. 19, 3918-3931. 51 Soderqvist, H., and Hallberg, E. (1994). The large C-terminal region of the integral pore membrane protein, POM121, is facing the nuclear pore complex. Eur. J. Cell Biol. 64, 186191. Soderqvist, H., Imreh, G., Kihlmark, M., Linnman, C., Ringertz, N., and Hallberg, E. (1997). Intracellular distribution of an integral nuclear pore membrane protein fused to green fluorescent protein--localization of a targeting domain. Eur. J. Biochem. 250, 808-813. Soderqvist, H., Jiang, W. Q., Ringertz, N., and Hallberg, E. (1996). Formation of nuclear bodies in cells overexpressing the nuclear pore protein POM121. Exp. Cell Res. 225, 75-84. Stewart, C., and Burke, B. (1987). Teratocarcinoma stem cells and early mouse embryos contain only a single major lamin polypeptide closely resembling lamin B. Cell 51, 383-392. Strom, A. C., and Weis, K. (2001). Importin-beta-like nuclear transport receptors. Genome Biol. 2. Stuurman, N., Sasse, B., and Fisher, P. A. (1996). Intermediate filament protein polymerization: molecular analysis of Drosophila nuclear lamin head-to-tail binding. J. Struct. Biol. 117, 1-15. Sukegawa, J., and Blobel, G. (1993). A nuclear pore complex protein that contains zinc finger motifs, binds DNA, and faces the nucleoplasm. Cell 72, 29-38. Tan, W., Zolotukhin, A. S., Bear, J., Patenaude, D. J., and Felber, B. K. (2000). The mRNA export in Caenorhabditis elegans is mediated by Ce-NXF-1, an ortholog of human TAP/NXF and Saccharomyces cerevisiae Mex67p. Rna 6, 1762-1772. Tcheperegine, S. E., Marelli, M., and Wozniak, R. W. (1999). Topology and functional domains of the yeast pore membrane protein Pom152p. J. Biol. Chem. 274, 5252-5258. Torres, M., Gomez-Pardo, E., Dressler, G. R., and Gruss, P. (1995). Pax-2 controls multiple steps of urogenital development. Development 121, 4057-4065. Ullman, K. S., Shah, S., Powers, M. A., and Forbes, D. J. (1999). The nucleoporin nup153 plays a critical role in multiple types of nuclear export. Mol. Biol. Cell 10, 649-664. Uv, A. E., Roth, P., Xylourgidis, N., Wickberg, A., Cantera, R., and Samakovlis, C. (2000). members only encodes a Drosophila nucleoporin required for rel protein import and immune response activation. Genes Dev. 14, 1945-1957. 52 van Deursen, J., Boer, J., Kasper, L., and Grosveld, G. (1996). G2 arrest and impaired nucleocytoplasmic transport in mouse embryos lacking the proto-oncogene CAN/Nup214. EMBO J. 15, 5574-5583. Vasu, S., Shah, S., Orjalo, A., Park, M., Fischer, W. H., and Forbes, D. J. (2001). Novel vertebrate nucleoporins Nup133 and Nup160 play a role in mRNA export. J. Cell Biol. 155, 339-354. Venter, J. C., Adams, M. D., Myers, E. W., Li, P. W., Mural, R. J., Sutton, G. G., Smith, H. O., Yandell, M., Evans, C. A., Holt, R. A., et al. (2001). The sequence of the human genome. Science 291, 1304-1351. Vigouroux, C., Auclair, M., Dubosclard, E., Pouchelet, M., Capeau, J., Courvalin, J. C., and Buendia, B. (2001). Nuclear envelope disorganization in fibroblasts from lipodystrophic patients with heterozygous R482Q/W mutations in the lamin A/C gene. J. Cell Sci. 114, 4459-4468. Vukicevic, S., Kopp, J. B., Luyten, F. P., and Sampath, T. K. (1996). Induction of nephrogenic mesenchyme by osteogenic protein 1 (bone morphogenetic protein 7). Proc. Natl. Acad. Sci. USA 93, 9021-9026. Walther, T. C., Fornerod, M., Pickersgill, H., Goldberg, M., Allen, T. D., and Mattaj, I. W. (2001). The nucleoporin Nup153 is required for nuclear pore basket formation, nuclear pore complex anchoring and import of a subset of nuclear proteins. EMBO J. 20, 5703-5714. Walther, T. C., Pickersgill, H. S., Cordes, V. C., Goldberg, M. W., Allen, T. D., Mattaj, I. W., and Fornerod, M. (2002). The cytoplasmic filaments of the nuclear pore complex are dispensable for selective nuclear protein import. J. Cell Biol. 158, 63-77. Welsh, J., Chada, K., Dalal, S. S., Cheng, R., Ralph, D., and McClelland, M. (1992). Arbitrarily primed PCR fingerprinting of RNA. Nucleic Acids Res. 20, 4965-4970. Wiegand, H. L., , Coburn, G. A., , Zeng, Y., , Kang, Y., , Bogerd, H. P., , and Cullen, B. R. (2002). Formation of Tap/NXT1 heterodimers activates Tap-dependent nuclear mRNA export by enhancing recruitment to nuclear pore complexes. Mol. Cell Biol. 22, 245-256. Wilken, N., Kossner, U., Senecal, J. L., Scheer, U., and Dabauvalle, M. C. (1993). Nup180, a novel nuclear pore complex protein localizing to the cytoplasmic ring and associated fibrils. J. Cell Biol. 123, 1345-1354. 53 Wilken, N., Senecal, J. L., Scheer, U., and Dabauvalle, M. C. (1995). Localization of the Ran-GTP binding protein RanBP2 at the cytoplasmic side of the nuclear pore complex. Eur. J. Cell Biol. 68, 211-219. Willem, M., Miosge, N., Halfter, W., Smyth, N., Jannetti, I., Burghart, E., Timpl, R., and Mayer, U. (2002). Specific ablation of the nidogen-binding site in the laminin gamma1 chain interferes with kidney and lung development. Development 129, 2711-2722. Wolff, B., Sanglier, J. J., and Wang, Y. (1997). Leptomycin B is an inhibitor of nuclear export: inhibition of nucleo-cytoplasmic translocation of the human immunodeficiency virus type 1 (HIV-1) Rev protein and Rev-dependent mRNA. Chem. Biol. 4, 139-147. Wozniac, R. W., and Blobel, G. (1992). The single transmembrane segment of gp210 is sufficient for sorting to the pore membrane domain of the nuclear envelope. J. Cell Biol. 119, 1441-1449. Wozniak, R. W., Bartnik, E., and Blobel, G. (1989). Primary structure analysis of an integral membrane glycoprotein of the nuclear pore. J. Cell Biol. 108, 2083-2092. Wozniak, R. W., Blobel, G., and Rout, M. P. (1994). POM152 is an integral protein of the pore membrane domain of the yeast nuclear envelope. J. Cell Biol. 125, 31-42. Wu, X., Kasper, L. H., Mantcheva, R. T., Mantchev, G. T., Springett, M. J., and van Deursen, J. M. (2001). Disruption of the FG nucleoporin NUP98 causes selective changes in nuclear pore complex stoichiometry and function. Proc. Natl. Acad. Sci. USA 98, 3191-3196. Wu, J., Matunis, M. J., Kraemer, D., Blobel, G., and Coutavas, E. (1995). Nup358, a cytoplasmically exposed nucleoporin with peptide repeats, Ran- GTP binding sites, zinc fingers, a cyclophilin A homologous domain, and a leucine-rich region. J. Biol. Chem. 270, 14209-14213. Xu, P. X., Adams, J., Peters, H., Brown, M. C., Heaney, S., and Maas, R. (1999). Eya1deficient mice lack ears and kidneys and show abnormal apoptosis of organ primordia. Nat. Genet. 23, 113-117. Yang, L., Guan, T., and Gerace, L. (1997). Integral membrane proteins of the nuclear envelope are dispersed throughout the endoplasmic reticulum during mitosis. J. Cell Biol. 137, 1199-1210. Yang, Q., Rout, M. P., and Akey, C. W. (1998). Three-dimensional architecture of the isolated yeast nuclear pore complex: functional and evolutionary implications. Mol. Cell. 1, 223-234. 54 Yaseen, N. R., and Blobel, G. (1999). GTP hydrolysis links initiation and termination of nuclear import on the nucleoporin nup358. J. Biol Chem. 274, 26493-26502. Ye, Q., and Worman, H. J. (1994). Primary structure analysis and lamin B and DNA binding of human LBR, an integral protein of the nuclear envelope inner membrane. J. Biol. Chem. 269, 11306-11311. Yokoyama, N., Hayashi, N., Seki, T., Pante, N., Ohba, T., Nishii, K., Kuma, K., Hayashida, T., Miyata, T., Aebi, U., and et al. (1995). A giant nucleopore protein that binds Ran/TC4. Nature 376, 184-188. Zhang, H., Schneider, J., and Rosdahl, I. (2002). Expression of p16, p27, p53, p73 and Nup88 proteins in matched primary and metastatic melanoma cells. Int. J. Oncol. 21, 43-48. Zimowska, G., Aris, J. P., and Paddy, M. R. (1997). A Drosophila Tpr protein homolog is localized both in the extrachromosomal channel network and to nuclear pore complexes. J. Cell Sci.110, 927-944. Zolotukhin, A. S., Tan, W., Bear, J., Smulevitch, S., and Felber, B. K. (2002). U2AF participates in the binding of TAP (NXF1) to mRNA. J. Biol. Chem. 277, 3935-3942. 55
© Copyright 2025 Paperzz