首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 16 毫秒
1.
2.
A collective secondary structure prediction for the human erythrocyte spectrin 106-residue repeat segment is developed, based on the sequences of nine segments that have been reported in the literature, utilizing a consensus of several secondary structure prediction methods for locating turn regions. The analysis predicts a five-fold structure, with three alpha-helices and two beta-strand regions, and differs from previous models on the lengths of the helices and the existence of beta-strand structure. We also demonstrate that this structural motif can be folded into tertiary structures that satisfy the experimental spectrin data and several general principles of protein organization.  相似文献   

3.
NgBR is a type I receptor with a single transmembrane domain and was identified as a specific receptor for Nogo-B. Our recent findings demonstrated that NgBR binds farnesylated Ras and recruits Ras to the plasma membrane, which is a critical step required for the activation of Ras signaling in human breast cancer cells and tumorigenesis. Here, we first use immunohistochemistry and real-time PCR approaches to examine the expression patterns of Nogo-B and NgBR in both normal and breast tumor tissues. Then, we examine the relationship between NgBR expression and molecular subtypes of breast cancer, and the roles of NgBR in estrogen-dependent survivin signaling pathway. Results showed that NgBR and Nogo-B protein were detected in both normal and breast tumor tissues. However, the expression of Nogo-B and NgBR in breast tumor tissue was much stronger than in normal breast tissue. The statistical analysis demonstrated that NgBR is highly associated with ER-positive/HER2-negative breast cancer. We also found that the expression of NgBR has a strong correlation with the expression of survivin, which is a well-known apoptosis inhibitor. The correlation between NgBR and survivin gene expression was further confirmed by real-time PCR. In vitro results also demonstrated that estradiol induces the expression of survivin in ER-positive T47D breast tumor cells but not in ER-negative MDA-MB-468 breast tumor cells. NgBR knockdown with siRNA abolishes estradiol-induced survivin expression in ER-positive T47D cells but not in ER-negative MDA-MB-468 cells. In addition, estradiol increases the expression of survivin and cell growth in ER-positive MCF-7 and T47D cells whereas knockdown of NgBR with siRNA reduces estradiol-induced survivin expression and cell growth. In summary, these results indicate that NgBR is a new molecular marker for breast cancer. The data suggest that the expression of NgBR may be essential in promoting ER-positive tumor cell proliferation via survivin induction in breast cancer.  相似文献   

4.
NgBR is a transmembrane protein identified as a Nogo‐B‐interacting protein and recently has been shown to be a subunit required for cis‐prenyltransferase (cisPTase) activity. To investigate the integrated role of NgBR in vascular development, we have characterized endothelial‐specific NgBR knockout embryos. Here, we show that endothelial‐specific NgBR knockout results in embryonic lethality due to vascular development defects in yolk sac and embryo proper. Loss of NgBR in endothelial cells reduces proliferation and promotes apoptosis of the cells largely through defects in the glycosylation of key endothelial proteins including VEGFR2, VE‐cadherin, and CD31, and defective glycosylation can be rescued by treatment with the end product of cisPTase activity, dolichol phosphate. Moreover, NgBR functions in endothelial cells during embryogenesis are Nogo‐B independent. These data uniquely show the importance of NgBR and protein glycosylation during vascular development.  相似文献   

5.
UDP-GlcNAc 2-epimerase/ManNAc kinase (GNE) catalyzes the first two committed steps in sialic acid synthesis. In addition to the three previously described human GNE isoforms (hGNE1-hGNE3), our database and polymerase chain reaction analysis yielded five additional human isoforms (hGNE4-hGNE8). hGNE1 is the ubiquitously expressed major isoform, while the hGNE2-hGNE8 isoforms are differentially expressed and may act as tissue-specific regulators of sialylation. hGNE2 and hGNE7 display a 31-residue N-terminal extension compared to hGNE1. On the basis of similarities to kinases and helicases, this extension does not seem to hinder the epimerase enzymatic active site. hGNE3 and hGNE8 contain a 55-residue N-terminal deletion and a 50-residue N-terminal extension compared to hGNE1. The size and secondary structures of these fragments are similar, and modeling predicted that these modifications do not affect the overall fold compared to that of hGNE1. However, the epimerase enzymatic activity of GNE3 and GNE8 is likely absent, because the deleted fragment contains important substrate binding residues in homologous bacterial epimerases. hGNE5-hGNE8 have a 53-residue deletion, which was assigned a role in substrate (UDP-GlcNAc) binding. Deletion of this fragment likely eliminates epimerase enzymatic activity. Our findings imply that GNE is subject to evolutionary mechanisms to improve cellular functions, without increasing the number of genes. Our expression and modeling data contribute to elucidation of the complex functional and regulatory mechanisms of human GNE and may contribute to further elucidating the pathology and treatment strategies of the human GNE-opathies sialuria and hereditary inclusion body myopathy.  相似文献   

6.
Viral entry by herpes simplex virus (HSV) is executed and tightly regulated by four glycoproteins. While several viral glycoproteins can mediate viral adhesion to host cells, only binding of gD to cellular receptor can activate core fusion proteins gB and gH/gL to execute membrane fusion and viral entry. Atomic structures of gD bound to receptor indicate that the C terminus of the gD ectodomain must be displaced before receptor can bind to gD, but it is unclear which conformational changes in gD activate membrane fusion. We rationally designed mutations in gD to displace the C terminus and observe if fusion could be activated without receptor binding. Using a cell-based fusion assay, we found that gD V231W induced cell-cell fusion in the absence of receptor. Using recombinant gD V231W protein, we observed binding to conformationally sensitive antibodies or HSV receptor and concluded that there were changes proximal to the receptor binding interface, while the tertiary structure of gD V231W was similar to that of wild-type gD. We used a biosensor to analyze the kinetics of receptor binding and the extent to which the C terminus blocks binding to receptor. We found that the C terminus of gD V231W was enriched in the open or displaced conformation, indicating a mechanism for its function. We conclude that gD V231W triggers fusion through displacement of its C terminus and that this motion is indicative of how gD links receptor binding to exposure of interfaces on gD that activate fusion via gH/gL and gB.  相似文献   

7.
The secondary structures of DnaK and the mutant DnaK756 heat-shock proteins from Escherichia coli have been investigated by Fourier transform infrared spectroscopy. The analysis of infrared data showed that DnaK and DnaK756 proteins have different secondary structures that are not affected by the presence of ATP or beta, gamma-methyleneadenosine 5'-triphosphate. The infrared data indicate also that the tertiary structures of DnaK and DnaK756 proteins are different and that DnaK protein undergoes conformational changes in its tertiary structure not only during binding of ATP but also during ATP hydrolysis. Using fluorescence spectroscopy of a single tryptophan located in the N-terminal domain of DnaK protein and fluorescence of 1,1'-bis(4-anilino)naphthalene-5,5'-disulfonic acid, which interacts with hydrophobic domains of DnaK protein, we were able to distinguish between two conformational states of DnaK protein. After binding of triphosphonucleotides, the C-terminal domain of DnaK protein changes in tertiary structure in such a way that fewer hydrophobic segments are exposed on the surface of the protein. After ATP hydrolysis, the number of hydrophobic segments on the surface of the protein is further reduced, and moreover the tertiary structure of the N-terminal domain of the protein changes. These data are discussed in terms of structural and functional relationships of both DnaK and DnaK756 proteins.  相似文献   

8.
Members of the fusion-associated small transmembrane (FAST) protein family are a distinct class of membrane fusion proteins encoded by nonenveloped fusogenic reoviruses. The 125-residue p14 FAST protein of reptilian reovirus has an approximately 38-residue myristoylated N-terminal ectodomain containing a moderately apolar N-proximal region, termed the hydrophobic patch. Mutagenic analysis indicated sequence-specific elements in the N-proximal portion of the p14 hydrophobic patch affected cell-cell fusion activity, independent of overall effects on the relative hydrophobicity of the motif. Circular dichroism (CD) of a myristoylated peptide representing the majority of the p14 ectodomain suggested this region is mostly disordered in solution but assumes increased structure in an apolar environment. From NMR spectroscopic data and simulated annealing, the soluble nonmyristoylated p14 ectodomain peptide consists of an N-proximal extended loop flanked by two proline hinges. The remaining two-thirds of the ectodomain peptide structure is disordered, consistent with predictions based on CD spectra of the myristoylated peptide. The myristoylated p14 ectodomain peptide, but not a nonmyristoylated version of the same peptide nor a myristoylated scrambled peptide, mediated extensive lipid mixing in a liposome fusion assay. Based on the lipid mixing activity, structural plasticity, environmentally induced conformational changes, and kinked structures predicted for the p14 ectodomain and hydrophobic patch (all features associated with fusion peptides), we propose that the majority of the p14 ectodomain is composed of a fusion peptide motif, the first such motif dependent on myristoylation for membrane fusion activity.  相似文献   

9.
Amyloid fibers, implicated in a wide range of diseases, are formed when proteins misfold and stick together in long rope‐like structures. As a natural mechanism, osmolytes can be used to modulate protein aggregation pathways with no interference with other cellular functions. The osmolyte sucrose delays fibrillation of the ribosomal protein S6 leading to softer and less shaped‐defined fibrils. The molecular mechanism used by sucrose to delay S6 fibrillation was studied based on the two‐state unfolding kinetics of the secondary and tertiary structures. It was concluded that the delay in S6 fibrillation results from stabilization and compaction of the slightly expanded tertiary native structure formed under fibrillation conditions. Interestingly, this compaction extends to almost all S6 tertiary structure but hardly affects its secondary structure. The part of the S6 tertiary structure that suffered more compaction by sucrose is known to be the first part to unfold, indicating that the native S6 has entered the unfolding pathway under fibrillation conditions. Proteins 2015; 83:2039–2051. © 2015 Wiley Periodicals, Inc.  相似文献   

10.
Li M  Song J 《Proteins》2007,68(1):100-108
RTN4 or Nogo proteins are composed of three alternative splice forms, namely 1192-residue Nogo-A, 373-residue Nogo-B, and 199-residue Nogo-C. Nogo proteins have received intense attentions because they have been implicated in a variety of critical cellular processes including CNS neuronal regeneration, vascular remodeling, apoptosis, interaction with beta-amyloid protein converting enzyme, and generation/maintenance of the tubular network of the endoplasmic reticulum (ER). Despite their significantly-different N-terminal lengths, they share a conserved C-terminal reticulon-homology domain consisting of two transmembrane fragments, a 66-residue extracellular loop Nogo-66 and a 38-residue C-tail carrying ER retention motif. Nogo-A owns the largest N-terminus with 1016 residues while the Nogo-B has an N-terminus almost identical to the first 200 residues of Nogo-A. So far, except for our previous determination of the Nogo-66 solution structure, no structural characterization of the other Nogo regions has been reported. In the present study, we initiated a systematically investigation of structural properties of Nogo molecules by a combined use of bioinformatics, CD, and NMR spectroscopy. The results led to two striking findings: (1) in agreement with bioinformatics prediction, the N- and C-termini of Nogo-B were experimentally demonstrated to be intrinsically unstructured by CD, two-dimensional 1H 15N NMR HSQC, hydrogen exchange, and 15N heteronuclear NOE characterization. (2) Further studies showed that the 1016-residue N-terminus of Nogo-A was again highly disordered. Therefore, it appears that being intrinsically-unstructured allows Nogo molecules to serve as double-faceted functional players, with one set of functions involved in cellular signaling processes essential for CNS neuronal regeneration, vascular remodeling, apoptosis and so forth and with another in generating/maintaining membrane-related structures. We propose that this mechanism may represent a general strategy to place the formation/maintenance of membrane-related structures under the direct regulation of the cellular signaling.  相似文献   

11.
Unique sequence-binding sites are exposed on the surface of high molecular weight kininogen which complex prekallikrein or factor XI with high affinity and specificity. A sequence comprising 31 residues of the mature kininogen molecule (Asp565-Lys595) retains full binding activity for prekallikrein (K D =20 nM) and assumes a complex folded structure in solution which is stabilized by long-range interactions between N- and C-terminal residues. The sequence Trp569-Lys595 (27 residues) shows only 28% of this binding affinity and lacks the key structural features required for protein recognition (Scarsale, J. N., and Harris, R. B.,J. Prot. Chem. 9, 647–659, 1990). We were thus able to predict that N- or C-terminal truncations of the binding-site sequence would disrupt the conformational integrity required for binding. Two new peptides of 20- and 22- residues have now been synthesized and their solution phase structures examined. These peptides are N- and C-terminal truncations, respectively, of the 27-residue sequence and correspond to the sequences Asp576-Lys595 and Trp569-Asp590 of high molecular weight kininogen. The results of fluorescence emission and circular dichroism (CD) spectroscopies in the range 25–90°C and from differential scanning calorimetry (DSC) all substantiate the idea that the C-terminal truncation peptide binds prekallikrein 35-fold poorer than the 31-residue peptide because it is relatively unoredered and possesses a less stable structure. Surprisingly, the N-terminal truncation peptide (20-mer) shows structural stability even at elevated temperatures and, like the 31-residue peptide, undergoes cold-induced denaturation observable in the DSC. 2D-NMR analysis of the 20-residue peptide revealed two distinct structures; one conformer possesses a more compact, folded structure than the other. However, the predicted structures assumed by either conformer are very different from those of either the 31- or 27-residue peptides. Hence, the binding affinity of the 20-residue peptide is 60-fold poorer than that for the 31-residue peptide because it assumes a nonproductive binding conformation(s).  相似文献   

12.
The structure of human syndecan, an integral membrane proteoglycan, has been determined by cloning its full-length cDNA, which codes for the entire 310-amino acid-long core protein, including the NH2-terminal signal peptide. Similar to mouse syndecan (Saunders, S., Jalkanen, M., O'Farrell, S., and Bernfield, M. (1989) J. Cell Biol. 108, 1547-1556), the core protein of human syndecan can be divided into three domains: a matrix-interacting ectodomain containing putative glycosaminoglycan attachment sites, a 25-residue hydrophobic membrane-spanning domain, and a 34-residue cytoplasmic domain. Several interesting conserved structures were revealed by comparing the human syndecan sequence to the murine one. (i) Although the ectodomains are only 70% identical, all putative glycosaminoglycan attachment sites are identical (two of them belong to the consensus sequence SGXG and three others to (E/D)GSG(E/D), as are also (ii) the single putative N-glycosylation site and (iii) the proteinase-sensitive dibasic RK site adjacent to the extracellular face of the transmembrane domain. Furthermore, (iv) the transmembrane domain is 96% identical, as the only change in human syndecan was an alteration of an alanine residue to glycine; and finally, (v) the cytoplasmic domain is 100% identical, including 3 identically located tyrosine residues. Comparison of transmembrane and cytoplasmic domains to a third cell-surface proteoglycan, 48K5 from human lung fibroblasts (Marynen, P., Zhang, J., Cassiman, J., Vanden Berghe, H., and David, C. (1989) J. Biol. Chem. 264, 7017-7024), indicates that the transmembrane and cytoplasmic domains are similar also in this molecule regardless of the presence of a totally nonhomologous ectodomain. Thus, the transmembrane and cytoplasmic domains are unique for these cell-surface proteoglycans, which we propose to be members of a novel gene family of syndecans.  相似文献   

13.
The Drosophila Toll receptor, which functions in both embryonic patterning and innate immunity to fungi and Gram-positive bacteria, is activated by a dimeric cytokine ligand, Sp?tzle (Spz). Previous studies have suggested that one Spz cross-links two Toll receptor molecules to form an activated complex. Here we report electron microscopy structures of the Toll ectodomain in the absence and presence of Spz. Contrary to expectations, Spz does not directly cross-link two Toll ectodomains. Instead, Spz binding at the N-terminal end of Toll predominantly induces the formation of a 2:2 complex, with two sites of interaction between the ectodomain chains, one located near to the N terminus of the solenoid and the other between the C-terminal juxtamembrane sequences. Moreover, Toll undergoes a ligand-induced conformational change, becoming more tightly curved than in the apo form. The unexpected 2:2 complex was confirmed by mass spectrometry under native conditions. These results suggest that activation of Toll is an allosteric mechanism induced by an end-on binding mode of its ligand Spz.  相似文献   

14.
The effects of chain cleavage and circular permutation on the structure, stability, and activity of dihydrofolate reductase (DHFR) from Escherichia coli were investigated by various spectroscopic and biochemical methods. Cleavage of the backbone after position 86 resulted in two fragments, (1--86) and (87--159) each of which are poorly structured and enzymatically inactive. When combined in a 1 : 1 molar ratio, however, the fragments formed a high-affinity (K(a) = 2.6 x 10(7) M(-1)) complex that displays a weakly cooperative urea-induced unfolding transition at micromolar concentrations. The retention of about 15% of the enzymatic activity of full-length DHFR is surprising, considering that the secondary structure in the complex is substantially reduced from its wild-type counterpart. In contrast, a circularly permuted form with its N-terminus at position 86 has similar overall stability to full-length DHFR, about 50% of its activity, substantial secondary structure, altered side-chain packing in the adenosine binding domain, and unfolds via an equilibrium intermediate not observed in the wild-type protein. After addition of ligand or the tight-binding inhibitor methotrexate, both the fragment complex and the circular permutant adopt more native-like secondary and tertiary structures. These results show that changes in the backbone connectivity can produce alternatively folded forms and highlight the importance of protein-ligand interactions in stabilizing the active site architecture of DHFR.  相似文献   

15.
At fertilization, sea urchin ovoperoxidase (OPO) is incorporated into a nascent fertilization envelope in association with proteoliaisin and plays an essential role in its hardening. By cDNA cloning based on the previously reported partial amino acid sequences of Hemicentrotus pulcherrimus OPO, the cDNA and deduced amino acid sequences of the enzyme precursor were determined. Its 814-residue sequence consists of 125-residue signal- plus pro-peptides and 689-residue mature enzyme and has 92.2 and 81.4% identity with the OPOs of Strongylocentrotus purpuratus and Lytechinus variegatus, respectively. Compared with human myeloperoxidase, it has 30.3% identity and 9.6% similarity and has an additional C-terminal sequence of about 100 residues, suggesting its possible role as the site for interaction with proteoliaisin, if not for the entire sequence. The linker peptide sequence between L- and H-chains (e.g., ASFVTG and FSFFTG) cleaved off upon activation in mammalian promyeloperoxidases is intrinsically lacking in OPO, uniquely rendering the single 70K polypeptide a basic unit. The positions of distal and proximal histidines, distal arginine, and six disulfide bridges are highly conserved among mammalian and sea urchin peroxidases. The secondary structure was predicted by EMBL-PHD on the Internet for the whole sequence of mature OPO, and both secondary and tertiary structures were predicted by Swiss-Model for the partial sequences residues 18-65 and 123-570 with canine myeloperoxidase as a template. The overall architecture of the OPO molecule is close to that of human myeloperoxidase but its secondary structure has some differences based on the sequence variation, as depicted by RasMol. Another software, Swiss-PdbViewer, produced a slightly but significantly different image of the OPO structure for the same predicted atomic coordinates. A discrepancy was also found between the secondary structures of human myeloperoxidase in the PDB file and in its Swiss-PdbViewer presentation.  相似文献   

16.
Sodium dodecylsulfate (SDS) and dodecylphosphocholine (DPC) micelles are often used to mimic the membrane- or receptor-bound states of peptides in NMR studies. From the present examination of a 26-residue analog of exendin-4 (TrEX4) by NMR and CD in water, aqueous 30% trifluoroethanol (TFE), and bound to both SDS and DPC micelles, it is clear that these two lipid micelles can yield very different peptide structures. The Trp-cage fold (also observed in 30% TFE) is present when TrEX4 is bound to SDS micelles; however, tertiary structure is absent in the presence of DPC micelles. The loss of tertiary structure is attributed to an energetically favorable interaction (estimated as 2-3 kcal/mol) of the tryptophan side chain with the phosphocholine head groups. These dramatic structural differences suggest that care must be taken when using either SDS or DPC to mimic the membrane- or receptor-bound states.  相似文献   

17.
Li F 《Journal of virology》2012,86(5):2856-2858
Among different coronavirus genera, the receptor-binding S1 subunits of their spike proteins differ in primary, secondary, and tertiary structures. This study identified shared structural topologies (connectivity of secondary structural elements) in S1 domains of different coronavirus genera. The results suggest that coronavirus S1 subunits share a common evolutionary origin but have attained diverse sequences and structures following extensive divergent evolution. The results also increase understanding of the structures and functions of coronavirus S1 domains whose tertiary structures are currently unknown.  相似文献   

18.
A Cave  M Pages  P Morin  C M Dobson 《Biochimie》1979,61(5-6):607-613
1H NMR and ORD were used to characterize the respective variations of tertiary structure and secondary structure of parvalbumins with calcium content ((Pa(O), without calcium and PaCa2 calcium saturated) and temperature. It has been observed that the tertiary structure can be lost without significant variation of the helical content. Cooperative binding of calcium to Pa(O) has been shown by NMR spectroscopy under low ionic strength conditions and at neutral pH. The present study shows that the calcium binding affinity of parvalbumin is dependent on the tertiary structure. Calcium binding and calcium release functions of parvalbumins in the muscle may be controlled by their tertiary structure.  相似文献   

19.
The amino-terminal ectodomain of the human TSH receptor has been expressed at the surface of CHO cells as a glycosylphosphatidylinositol-anchored molecule containing a 10-residue histidine tag close to its C terminus. The soluble ectodomain could be released from the cells by treatment with a glycosylphosphatidylinositol-phospholipase C and purified to apparent homogeneity by cobalt-Sepharose chromatography. Two nanomoles of material was obtained, which was suitable for analysis by mass spectrometry. This allowed the identification of four out of the six potential N-glycosylation sites as being effectively glycosylated. A proportion of the purified soluble ectodomain displayed specific binding of (125)I-labeled TSH, allowing for the first time performance of classical saturation binding experiments. Two classes of high-affinity binding sites were identified: site A, K(d) 0.014 nM; site B, K(d) 0.83 nM. The significance of site A, whose affinity is much higher than for the holoreceptor at the surface of intact cells, remains to be clarified. The purified ectodomain was capable of inhibiting efficiently the thyroid stimulating activity of immunoglobulins from patients with Graves' disease. It allowed computation of the amounts of these immunoglobulins in patient's serum, giving values up to 10 microg/mL. Contrary to all currently available assays, the soluble ectodomain of the TSH receptor purified in a functionally competent conformation allows direct studies of its interactions with TSH and autoantibodies and opens the way to structural studies.  相似文献   

20.
Chromodomain from heterochromatin protein 1 and polycomb protein is known to be a lysine-methylated histone H3 tail-binding module. Chromo-helicase/ATPase DNA-binding protein 1 (CHD1) is an ATP-dependent chromatin remodeling factor, containing two tandem chromodomains. In human CHD1, both chromodomains are essential for specific binding to a K4 methylated histone H3 (H3 MeK4) peptide and are found to bind cooperatively in the crystal structure. For the budding yeast homologue, Chd1, the second but not the first chromodomain was once reported to bind to an H3 MeK4 peptide. Here, we reveal that neither the second chromodomain nor a region containing tandem chromodomains from yeast Chd1 bind to any lysine-methylated or arginine-methylated histone peptides that we examined. In addition, we examined the structures of the chromodomains from Chd1 by NMR. Although the tertiary structure of the region containing tandem chromodomains could not be obtained, the secondary structure deduced from NMR is well conserved in the tertiary structures of the corresponding first and second chromodomains determined individually by NMR. Both chromodomains of Chd1 demonstrate a structure similar to that of the corresponding part of CHD1, consisting of a three-stranded beta-sheet followed by a C-terminal alpha-helix. However, an additional helix between the first and second beta-strands, which is found in both of the first chromodomains of Chd1 and CHD1, is positioned in an entirely different manner in Chd1 and CHD1. In human CHD1 this helix forms the peptide-binding site. The amino acid sequences of the chromodomains could be well aligned on the basis of these structures. The alignment showed that yeast Chd1 lacks several key functional residues, which are responsible for specific binding to a methylated lysine residue in other chromodomains. Chd1 is likely to have no binding affinity for any H3 MeK peptide, as found in other chromodomain proteins.  相似文献   

设为首页 | 免责声明 | 关于勤云 | 加入收藏

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