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 15 毫秒
1.
Nucleorhabdoviruses such as Sonchus yellow net virus (SYNV) replicate in the nuclei and undergo morphogenesis at the inner nuclear membrane (IM) in plant cells. Mature particles are presumed to form by budding of the Matrix (M) protein‐nucleocapsid complexes through host IMs to acquire host phospholipids and the surface glycoproteins (G). To address mechanisms underlying nucleorhabdovirus budding, we generated recombinant SYNV G mutants containing a truncated amino‐terminal (NT) or carboxyl‐terminal (CT) domain. Electron microscopy and sucrose gradient centrifugation analyses showed that the CT domain is essential for virion morphogenesis whereas the NT domain is also required for efficient budding. SYNV infection induces IM invaginations that are thought to provide membrane sites for virus budding. We found that in the context of viral infections, interactions of the M protein with the CT domain of the membrane‐anchored G protein mediate M protein translocation and IM invagination. Interestingly, tethering the M protein to endomembranes, either by co‐expression with a transmembrane G protein CT domain or by artificial fusion with the G protein membrane targeting sequence, induces IM invagination in uninfected cells. Further evidence to support functions of G‐M interactions in virus budding came from dominant negative effects on SYNV‐induced IM invagination and viral infections that were elicited by expression of a soluble version of the G protein CT domain. Based on these data, we propose that cooperative G‐M interactions promote efficient SYNV budding.  相似文献   

2.
BST‐2/tetherin is a human extracellular transmembrane protein that serves as a host defense factor against HIV‐1 and other viruses by inhibiting viral spreading. Structurally, BST‐2 is a homo‐dimeric coiled‐coil that is connected to the host cell membrane by N and C terminal transmembrane anchors. The C‐terminal membrane anchor of BST‐2 is inserted into the budding virus while the N‐terminal membrane anchor remains in the host cell membrane creating a viral tether. The structural mechanism of viral budding and tethering as mediated by BST‐2 is not clear. To more fully describe the mechanism of viral tethering, we created a model of BST‐2 embedded in a membrane and used steered molecular dynamics to simulate the transition from the host cell membrane associated form to the cell‐virus membrane bridging form. We observed that BST‐2 did not transition as a rigid structure, but instead bent at positions with a reduced interface between the helices of the coiled‐coil. The simulations for the human BST‐2 were then compared with simulations on the mouse homolog, which has no apparent weak spots. We observed that the mouse homolog spread the bending across the ectodomain, rather than breaking at discrete points as observed with the human homolog. These simulations support previous biochemical and cellular work suggesting some flexibility in the coiled‐coil is necessary for viral tethering, while also highlighting how subtle changes in protein sequence can influence the dynamics and stability of proteins with overall similar structure.  相似文献   

3.
As a target of antiviral drugs, the influenza A M2 protein has been the focus of numerous structural studies and has been extensively explored as a model ion channel. In this study, we capitalize on the expanding body of high‐resolution structural data available for the M2 protein to design and interpret site‐directed spin‐labeling electron paramagnetic resonance spectroscopy experiments on drug‐induced conformational changes of the M2 protein embedded in lipid bilayers. We obtained data in the presence of adamantane drugs for two different M2 constructs (M2TM 22–46 and M2TMC 23–60). M2TM peptides were spin labeled at the N‐terminal end of the transmembrane domain. M2TMC peptides were spin labeled site specifically at cysteine residues substituted for amino acids within the transmembrane domain (L36, I39, I42, and L43) and the C‐terminal amphipathic helix (L46, F47, F48, C50, I51, Y52, R53, F54, F55, and E56). Addition of adamantane drugs brought about significant changes in measured electron paramagnetic resonance spectroscopy environmental parameters consistent with narrowing of the transmembrane channel pore and closer packing of the C‐terminal amphipathic helices.  相似文献   

4.
The influenza A M2 protein forms a proton channel for virus infection and mediates virus assembly and budding. While extensive structural information is known about the transmembrane helix and an adjacent amphipathic helix, the conformation of the N‐terminal ectodomain and the C‐terminal cytoplasmic tail remains largely unknown. Using two‐dimensional (2D) magic‐angle‐spinning solid‐state NMR, we have investigated the secondary structure and dynamics of full‐length M2 (M2FL) and found them to depend on the membrane composition. In 2D 13C DARR correlation spectra, 1,2‐dimyristoyl‐sn‐glycero‐3‐phosphocholine (DMPC)‐bound M2FL exhibits several peaks at β‐sheet chemical shifts, which result from water‐exposed extramembrane residues. In contrast, M2FL bound to cholesterol‐containing membranes gives predominantly α‐helical chemical shifts. Two‐dimensional J‐INADEQUATE spectra and variable‐temperature 13C spectra indicate that DMPC‐bound M2FL is highly dynamic while the cholesterol‐containing membranes significantly immobilize the protein at physiological temperature. Chemical‐shift prediction for various secondary‐structure models suggests that the β‐strand is located at the N‐terminus of the DMPC‐bound protein, while the cytoplasmic domain is unstructured. This prediction is confirmed by the 2D DARR spectrum of the ectodomain‐truncated M2(21–97), which no longer exhibits β‐sheet chemical shifts in the DMPC‐bound state. We propose that the M2 conformational change results from the influence of cholesterol, and the increased helicity of M2FL in cholesterol‐rich membranes may be relevant for M2 interaction with the matrix protein M1 during virus assembly and budding. The successful determination of the β‐strand location suggests that chemical‐shift prediction is a promising approach for obtaining structural information of disordered proteins before resonance assignment.  相似文献   

5.
6.
Cytoplasmic domains of transmembrane bacterial chemoreceptors are largely extended four‐helix coiled coils. Previous observations suggested the domain was structurally dynamic. We probed directly backbone dynamics of this domain of the transmembrane chemoreceptor Tar from Escherichia coli using site‐directed spin labeling and electron paramagnetic resonance (EPR) spectroscopy. Spin labels were positioned on solvent‐exposed helical faces because EPR spectra for such positions reflect primarily polypeptide backbone movements. We acquired spectra for spin‐labeled, intact receptor homodimers solubilized in detergent or inserted into native E. coli lipid bilayers in Nanodiscs, characterizing 16 positions distributed throughout the cytoplasmic domain and on both helices of its helical hairpins, one amino terminal to the membrane‐distal tight turn (N‐helix), and the other carboxyl terminal (C‐helix). Detergent solubilization increased backbone dynamics for much of the domain, suggesting that loss of receptor activities upon solubilization reflects wide‐spread destabilization. For receptors in either condition, we observed an unanticipated difference between the N‐ and C‐helices. For bilayer‐inserted receptors, EPR spectra from sites in the membrane‐distal protein‐interaction region and throughout the C‐helix were typical of well‐structured helices. In contrast, for approximately two‐thirds of the N‐helix, from its origin as the AS‐2 helix of the membrane‐proximal HAMP domain to the beginning of the membrane‐distal protein‐interaction region, spectra had a significantly mobile component, estimated by spectral deconvolution to average approximately 15%. Differential helical dynamics suggests a four‐helix bundle organization with a pair of core scaffold helices and two more dynamic partner helices. This newly observed feature of chemoreceptor structure could be involved in receptor function.  相似文献   

7.
Human MICAL1 is a member of a recently discovered family of multidomain proteins that couple a FAD‐containing monooxygenase‐like domain to typical protein interaction domains. Growing evidence implicates the NADPH oxidase reaction catalyzed by the flavoprotein domain in generation of hydrogen peroxide as a second messenger in an increasing number of cell types and as a specific modulator of actin filaments stability. Several proteins of the Rab families of small GTPases are emerging as regulators of MICAL activity by binding to its C‐terminal helical domain presumably shifting the equilibrium from the free – auto‐inhibited – conformation to the active one. We here extend the characterization of the MICAL1–Rab8 interaction and show that indeed Rab8, in the active GTP‐bound state, stabilizes the active MICAL1 conformation causing a specific four‐fold increase of kcat of the NADPH oxidase reaction. Kinetic data and small‐angle X‐ray scattering (SAXS) measurements support the formation of a 1:1 complex between full‐length MICAL1 and Rab8 with an apparent dissociation constant of approximately 8 μM. This finding supports the hypothesis that Rab8 is a physiological regulator of MICAL1 activity and shows how the protein region preceding the C‐terminal Rab‐binding domain may mask one of the Rab‐binding sites detected with the isolated C‐terminal fragment. SAXS‐based modeling allowed us to propose the first model of the free full‐length MICAL1, which is consistent with an auto‐inhibited conformation in which the C‐terminal region prevents catalysis by interfering with the conformational changes that are predicted to occur during the catalytic cycle.  相似文献   

8.
Retroviral Gag protein plays a critical role during the late stage of virus budding and possesses a so‐called L‐domain containing PT/SAP, PPxY, YxxL or FPIV motifs that are critical for efficient budding. Mason–Pfizer monkey virus (M‐PMV) contains PSAP, PPPY, and YADL sequences in Gag. This study was performed to investigate the roles of these three L‐domain‐like sequences in virus replication in three different cell lines, 293T, COS‐7 and HeLa cells. It was found that the PPxY motif plays an essential role in progeny virus production as a major L‐domain in all three cell lines. The PSAP sequence was shown to function as an additional L‐domain in HeLa cells and to promote efficient release of M‐PMV; however, this sequence was dispensable for M‐PMV production in 293T and COS‐7 cells, suggesting that the role of the PSAP motif as an L‐domain in M‐PMV budding is cell type‐dependent. Viruses possessing multiple L‐domains appear to change the L‐domain usage to replicate in various cells. On the other hand, the YADL motif was required for M‐PMV production as a transport signal of Gag to the plasma membrane, but not as an L‐domain.  相似文献   

9.
The effect molecular crowding, defined as the volume exclusion exerted by one soluble inert molecule upon another soluble molecule, has on the structure and self‐interaction of lipid‐free apoA‐I were explored. The influence of molecular crowding on lipid‐free apoA‐I oligomerization and internal dynamics has been analyzed using electron paramagnetic resonance (EPR) spectroscopy measurements of nitroxide spin label at selected positions throughout the protein sequence and at varying concentrations of the crowding agent Ficoll‐70. The targeted positions include sites previously shown to be sensitive for detecting intermolecular interaction via spin–spin coupling. Circular dichroism was used to study secondary structural changes in lipid‐free apoA‐I imposed by increasing concentrations of the crowding agent. Crosslinking and SDS‐PAGE gel analysis was employed to further characterize the role molecular crowding plays in inducing apoA‐I oligomerization. It was concluded that the dynamic apoA‐I structure and oligomeric state was altered in the presence of the crowding agent. It was also found that the C‐terminal was slightly more sensitive to molecular crowding. Finally, the data described the region around residue 217 in the C‐terminal domain of apoA‐I as the most sensitive reporter of the crowding‐induced self‐association of apoA‐I. The implications of this behavior to in vivo functionality are discussed. © 2016 Wiley Periodicals, Inc. Biopolymers 105: 683–692, 2016.  相似文献   

10.
The protein anosmin‐1, coded by the KAL1 gene responsible for the X‐linked form of Kallmann syndrome (KS), exerts its biological effects mainly through the interaction with and signal modulation of fibroblast growth factor receptor 1 (FGFR1). We have previously shown the interaction of the third fibronectin‐like type 3 (FnIII) domain and the N‐terminal region of anosmin‐1 with FGFR1. Here, we demonstrate that missense mutations reported in patients with KS, C172R and N267K did not alter or substantially reduce, respectively, the binding to FGFR1. These substitutions annulled the chemoattraction of the full‐length protein over subventricular zone (SVZ) neuronal precursors (NPs), but they did not annul it in the N‐terminal‐truncated protein (A1Nt). We also show that although not essential for binding to FGFR1, the cysteine‐rich (CR) region is necessary for anosmin‐1 function and that FnIII.3 cannot substitute for FnIII.1 function. Truncated proteins recapitulating nonsense mutations found in KS patients did not show the chemotropic effect on SVZ NPs, suggesting that the presence behind FnIII.1 of any part of anosmin‐1 produces an unstable protein incapable of action. We also identify the extracellular signal‐regulated kinase 1/2 (ERK1/2) pathway as necessary for the chemotropic effect exerted by FGF2 and anosmin‐1 on rat SVZ NPs.  相似文献   

11.
The retroviral integrase (IN) carries out the integration of a dsDNA copy of the viral genome into the host DNA, an essential step for viral replication. All IN proteins have three general domains, the N‐terminal domain (NTD), the catalytic core domain, and the C‐terminal domain. The NTD includes an HHCC zinc finger‐like motif, which is conserved in all retroviral IN proteins. Two crystal structures of Moloney murine leukemia virus (M‐MuLV) IN N‐terminal region (NTR) constructs that both include an N‐terminal extension domain (NED, residues 1–44) and an HHCC zinc‐finger NTD (residues 45–105), in two crystal forms are reported. The structures of IN NTR constructs encoding residues 1–105 (NTR1–105) and 8–105 (NTR8–105) were determined at 2.7 and 2.15 Å resolution, respectively and belong to different space groups. While both crystal forms have similar protomer structures, NTR1–105 packs as a dimer and NTR8–105 packs as a tetramer in the asymmetric unit. The structure of the NED consists of three anti‐parallel β‐strands and an α‐helix, similar to the NED of prototype foamy virus (PFV) IN. These three β‐strands form an extended β‐sheet with another β‐strand in the HHCC Zn2+ binding domain, which is a unique structural feature for the M‐MuLV IN. The HHCC Zn2+ binding domain structure is similar to that in HIV and PFV INs, with variations within the loop regions. Differences between the PFV and MLV IN NEDs localize at regions identified to interact with the PFV LTR and are compared with established biochemical and virological data for M‐MuLV. Proteins 2017; 85:647–656. © 2016 Wiley Periodicals, Inc.  相似文献   

12.
The mechanisms of interfacial folding and membrane insertion of the Alzheimer's amyloid‐β fragment Aβ(25–35) and its less toxic mutant, N27A‐Aβ(25–35) and more toxic mutant, M35A‐Aβ(25–35), are investigated using replica–exchange molecular dynamics in an implicit water‐membrane environment. This study simulates the processes of interfacial folding and membrane insertion in a spontaneous fashion to identify their general mechanisms. Aβ(25–35) and N27A‐Aβ(25–35) peptides share similar mechanisms: the peptides are first located in the membrane hydrophilic region where their C‐terminal residues form helical structures. The peptides attempt to insert themselves into the membrane hydrophobic region using the C‐terminal or central hydrophobic residues. A small portion of peptides can successfully enter the membrane's hydrophobic core, led by their C‐terminal residues, through the formation of continuous helical structures. No detectable amount of M35A‐Aβ(25–35) peptides appeared to enter the membrane's hydrophobic core. The three studied peptides share a similar helical structure for their C‐terminal five residues, and these residues mainly buried within the membrane's hydrophobic region. In contrast, their N‐terminal properties are markedly different. With respect to the Aβ(25–35), the N27A‐Aβ(25–35) forms a more structured helix and is buried deeper within the membrane, which may result in a lower degree of aggregation and a lower neurotoxicity; in contrast, the less structured and more water‐exposed M35A‐Aβ(25–35) is prone to aggregation and has a higher neurotoxicity. Understanding the mechanisms of Aβ peptide interfacial folding and membrane insertion will provide new insights into the mechanisms of neurodegradation and may give structure‐based clues for rational drug design preventing amyloid associated diseases. Proteins 2010. © 2010 Wiley‐Liss, Inc.  相似文献   

13.
The 173–195 segment corresponding to the helix 2 of the globular PrP domain is a good candidate to be one of the several ‘spots’ of intrinsic structural flexibility, which might induce local destabilization and concur to protein transformation, leading to aggregation‐prone conformations. Here, we report CD and NMR studies on the α2‐helix‐derived peptide of maximal length (hPrP[180–195]) that is able to exhibit a regular structure different from the prevalently random arrangement of other α2‐helix‐derived peptides. This peptide, which has previously been shown to be affected by buffer composition via the ion charge density dependence typical of Hofmeister effects, corresponds to the C‐terminal sequence of the PrPC full‐length α2‐helix and includes the highly conserved threonine‐rich 188–195 segment. At neutral pH, its conformation is dominated by β‐type contributions, which only very strong environmental modifications are able to modify. On TFE addition, an increase of α‐helical content can be observed, but a fully helical conformation is only obtained in neat TFE. However, linking of the 173–179 segment, as occurring in wild‐type and mutant peptides corresponding to the full‐length α2‐helix, perturbs these intrinsic structural propensities in a manner that depends on whether the environment is water or TFE. Overall, these results confirm that the 180–195 parental region in hPrPC makes a strong contribution to the chameleon conformational behavior of the segment corresponding to the full‐length α2‐helix, and could play a role in determining structural rearrangements of the entire globular domain. Copyright © 2008 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

14.
Calsequestrin (CASQ) exists as two distinct isoforms CASQ1 and CASQ2 in all vertebrates. Although the isoforms exhibit unique functional characteristic, the structural basis for the same is yet to be fully defined. Interestingly, the C‐terminal region of the two isoforms exhibit significant differences both in length and amino acid composition; forming Dn‐motif and DEXn‐motif in CASQ1 and CASQ2, respectively. Here, we investigated if the unique C‐terminal motifs possess Ca2+‐sensitivity and affect protein function. Sequence analysis shows that both the Dn‐ and DEXn‐motifs are intrinsically disordered regions (IDRs) of the protein, a feature that is conserved from fish to man. Using purified synthetic peptides, we show that these motifs undergo distinctive Ca2+‐mediated folding suggesting that these disordered motifs are Ca2+‐sensitivity. We generated chimeric proteins by swapping the C‐terminal portions between CASQ1 and CASQ2. Our studies show that the C‐terminal portions do not play significant role in protein folding. An interesting finding of the current study is that the switching of the C‐terminal portion completely reverses the polymerization kinetics. Collectively, these data suggest that these Ca2+‐sensitivity IDRs located at the back‐to‐back dimer interface influence isoform‐specific Ca2+‐dependent polymerization properties of CASQ. © 2014 Wiley Periodicals, Inc. Biopolymers 103: 15–22, 2015.  相似文献   

15.
On consideration that intrinsic structural weakness could affect the segment spanning the α2‐helical residues 173–195 of the PrP, we have investigated the conformational stabilities of some synthetic Ala‐scanned analogs of the peptide derived from the 180–195 C‐terminal sequence, using a novel approach whose theoretical basis originates from protein thermodynamics. Even though a quantitative comparison among peptides could not be assessed to rank them according to the effect caused by single amino acid substitution, as a general trend, all peptides invariably showed an appreciable preference for an α‐type organization, consistently with the fact that the wild‐type sequence is organized as an α‐helix in the native protein. Moreover, the substitution of whatever single amino acid in the wild‐type sequence reduced the gap between the α‐ and the β‐propensity, invariably enhancing the latter, but in any case this gap was larger than that evaluated for the full‐length α2‐helix‐derived peptide. It appears that the low β‐conformation propensity of the 180–195 region depends on the simultaneous presence of all of the Ala‐scanned residues, indirectly confirming that the N‐terminal 173–179 segment could play a major role in determining the chameleon conformational behavior of the entire 173–195 region in the PrP. Copyright © 2008 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

16.
The serine protease inhibitor, plasminogen activator inhibitor Type‐1 (PAI‐1) is a metastable protein that undergoes an unusual transition to an inactive conformation with a short half‐life of only 1–2 hr. Circulating PAI‐1 is bound to a cofactor vitronectin, which stabilizes PAI‐1 by slowing this latency conversion. A well‐characterized PAI‐1‐binding site on vitronectin is located within the somatomedin B (SMB) domain, corresponding to the first 44 residues of the protein. Another PAI‐1 recognition site has been identified with an engineered form of vitronectin lacking the SMB domain, yet retaining PAI‐1 binding capacity (Schar, Blouse, Minor, Peterson. J Biol Chem. 2008;283:28487–28496). This additional binding site is hypothesized to lie within an intrinsically disordered domain (IDD) of vitronectin. To localize the putative binding site, we constructed a truncated form of vitronectin containing 71 amino acids from the N‐terminus, including the SMB domain and an additional 24 amino acids from the IDD region. This portion of the IDD is rich in acidic amino acids, which are hypothesized to be complementary to several basic residues identified within an extensive vitronectin‐binding site mapped on PAI‐1 (Schar, Jensen, Christensen, Blouse, Andreasen, Peterson. J Biol Chem. 2008;283:10297–10309). Steady‐state and stopped‐flow fluorescence measurements demonstrate that the truncated form of vitronectin exhibits the same rapid biphasic association as full‐length vitronectin and that the IDD hosts the elusive second PAI‐1 binding site that lies external to the SMB domain of vitronectin.  相似文献   

17.
For many membrane proteins, the determination of their topology remains a challenge for methods like X‐ray crystallography and nuclear magnetic resonance (NMR) spectroscopy. Electron paramagnetic resonance (EPR) spectroscopy has evolved as an alternative technique to study structure and dynamics of membrane proteins. The present study demonstrates the feasibility of membrane protein topology determination using limited EPR distance and accessibility measurements. The BCL::MP‐Fold (BioChemical Library membrane protein fold) algorithm assembles secondary structure elements (SSEs) in the membrane using a Monte Carlo Metropolis (MCM) approach. Sampled models are evaluated using knowledge‐based potential functions and agreement with the EPR data and a knowledge‐based energy function. Twenty‐nine membrane proteins of up to 696 residues are used to test the algorithm. The RMSD100 value of the most accurate model is better than 8 Å for 27, better than 6 Å for 22, and better than 4 Å for 15 of the 29 proteins, demonstrating the algorithms' ability to sample the native topology. The average enrichment could be improved from 1.3 to 2.5, showing the improved discrimination power by using EPR data. Proteins 2015; 83:1947–1962. © 2015 Wiley Periodicals, Inc  相似文献   

18.
Anamorsin is a recently identified molecule that inhibits apoptosis during hematopoiesis. It contains an N‐terminal methyltransferase‐like domain and a C‐terminal Fe‐S cluster motif. Not much is known about the function of the protein. To better understand the function of anamorsin, we have solved the crystal structure of the N‐terminal domain at 1.8 Å resolution. Although the overall structure resembles a typical S‐adenosylmethionine (SAM) dependent methyltransferase fold, it lacks one α‐helix and one β‐strand. As a result, the N‐terminal domain as well as the full‐length anamorsin did not show S‐adenosyl‐l ‐methionine (AdoMet) dependent methyltransferase activity. Structural comparisons with known AdoMet dependent methyltransferases reveals subtle differences in the SAM binding pocket that preclude the N‐terminal domain from binding to AdoMet. The N‐terminal methyltransferase‐like domain of anamorsin probably functions as a structural scaffold to inhibit methyl transfers by out‐competing other AdoMet dependant methyltransferases or acts as bait for protein–protein interactions.Proteins 2014; 82:1066–1071. © 2013 Wiley Periodicals, Inc.  相似文献   

19.
Previous studies reported the presence of choline acetyltransferase (ChAT) mRNA and protein in the mammalian testis. We have now found that none of the ChAT mRNAs produced in the testis is capable of encoding a full‐length ChAT protein. Two ChAT cDNAs were isolated from an adult rat testis cDNA library encoding N‐terminally truncated ChAT proteins of 450 and 414 amino acids (aa), respectively, the former containing a novel N‐terminal extension of 69 residues. Rapid Amplification of cDNA Ends (RACE) analysis revealed a complex pattern of 5′ untranslated mRNA termini generated from the ChAT gene locus in the testis, all representing truncated versions of the ChAT enzyme. Two of these proteins were produced in transfected fibroblasts and found to lack ChAT activity. Neither did they show binding to the ChAT substrates, acetyl CoA and choline, in a competition assay. These results indicate that mammalian testis lacks a bona fide ChAT enzyme but expresses truncated ChAT proteins with a possible unique function to the testis. Mol. Reprod. Dev. 53:274–281, 1999. © 1999 Wiley‐Liss, Inc.  相似文献   

20.
The tryptophan rich basic protein/calcium signal‐modulating cyclophilin ligand (WRB/CAML) and Get1p/Get2p complexes, in vertebrates and yeast, respectively, mediate the final step of tail‐anchored protein insertion into the endoplasmic reticulum membrane via the Get pathway. While WRB appears to exist in all eukaryotes, CAML homologs were previously recognized only among chordates, raising the question as to how CAML's function is performed in other phyla. Furthermore, whereas WRB was recognized as the metazoan homolog of Get1, CAML and Get2, although functionally equivalent, were not considered to be homologous. CAML contains an N‐terminal basic, TRC40/Get3‐interacting, region, three transmembrane segments near the C‐terminus, and a poorly conserved region between these domains. Here, I searched the NCBI protein database for remote CAML homologs in all eukaryotes, using position‐specific iterated‐basic local alignment search tool, with the C‐terminal, the N‐terminal or the full‐length sequence of human CAML as query. The N‐terminal basic region and full‐length CAML retrieved homologs among metazoa, plants and fungi. In the latter group several hits were annotated as GET2. The C‐terminal query did not return entries outside of the animal kingdom, but did retrieve over one hundred invertebrate metazoan CAML‐like proteins, which all conserved the N‐terminal TRC40‐binding domain. The results indicate that CAML homologs exist throughout the eukaryotic domain of life, and suggest that metazoan CAML and yeast GET2 share a common evolutionary origin. They further reveal a tight link between the particular features of the metazoan membrane‐anchoring domain and the TRC40‐interacting region. The list of sequences presented here should provide a useful resource for future studies addressing structure‐function relationships in CAML 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