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.
The effect of exposure period and concentration of algae on the frequency of infection of aposymbiotic ciliates by algae obtained from the same clone of Paramecium bursaria syngen 2, was studied. The frequency of infection was roughly proportional to the algal concentration and to the exposure time of ciliates to algae. The relationship of algal concentration to infection frequency closely fitted the Poisson distribution curve for N = 1, suggesting that the minimum number of algae required to infect a single ciliate is 1. However, the data also strongly suggested that the average number of algae required to initiate infection of an average ciliate was ? 1,000. Three possible resolutions of this situation are: (a) the selection by the ciliate of a rare infective variant from a heterogeneous population: (b) the rare escape of an alga from digestion by the ciliate; and (c) the requirement for a large number of algae-ciliate contacts to induce susceptibility in the ciliate. Splitting the exposure of ciliates to algae into 2 periods of 0.5 h, separated by 5 h in the absence of algae, produced a much higher frequency of infection than a single l-h exposure, supporting the suggestion that the large number of algae is required to induce susceptibility in the ciliate which can then be infected by as few as a single algal cell.  相似文献   

2.
Heterogeneity is regarded as the major factor leading to the poor outcomes of glioblastoma (GBM) patients. However, conventional two‐dimensional (2D) analysis methods, such as immunohistochemistry and immunofluorescence, have limited capacity to reveal GBM spatial heterogeneity. Thus, we sought to develop an effective analysis strategy to increase the understanding of GBM spatial heterogeneity. Here, 2D and three‐dimensional (3D) analysis methods were compared for the examination of cell morphology, cell distribution and large intact structures, and both types of methods were employed to dissect GBM spatial heterogeneity. The results showed that 2D assays showed only cross‐sections of specimens but provided a full view. To visualize intact GBM specimens in 3D without sectioning, the optical tissue clearing methods CUBIC and iDISCO+ were used to clear opaque specimens so that they would become more transparent, after which the specimens were imaged with a two‐photon microscope. The 3D analysis methods showed specimens at a large spatial scale at cell‐level resolution and had overwhelming advantages in comparison to the 2D methods. Furthermore, in 3D, heterogeneity in terms of cell stemness, the microvasculature, and immune cell infiltration within GBM was comprehensively observed and analysed. Overall, we propose that 2D and 3D analysis methods should be combined to provide much greater detail to increase the understanding of GBM spatial heterogeneity.  相似文献   

3.
The solution structure of contryphan-Vn, a cyclic peptide with a double cysteine S-S bridge and containing a D-tryptophan extracted from the venom of the cone snail Conus ventricosus, has been determined by NMR spectroscopy using a variety of homonuclear and heteronuclear NMR methods and restrained molecular dynamics simulations. The main conformational features of backbone contryphan-Vn are a type IV beta-turn from Gly 1 to Lys 6 and a type I beta-turn from Lys 6 to Cys 9. As already found in other contryphans, one of the two prolines--the Pro4--is mainly in the cis conformation while Pro7 is trans. A small hydrophobic region probably partly shielded from solvent constituted from the close proximity of side chains of Pro7 and Trp8 was observed together with a persistent salt bridge between Asp2 and Lys6, which has been revealed by the diagnostic observation of specific nuclear Overhauser effects. The salt bridge was used as a restraint in the molecular dynamics in vacuum but without inserting explicit electrostatic contribution in the calculations. The backbone of the unique conformational family found of contryphan-Vn superimposes well with those of contryphan-Sm and contryphan-R. This result indicates that the contryphan structural motif represents a robust and conserved molecular scaffold whose main structural determinants are the size of the intercysteine loop and the presence and location in the sequence of the D-Trp and the two Pro residues.  相似文献   

4.
Entodiniomorph ciliates, belonging to the family Cycloposthiidae, are described from the stomach contents of Hippopotamus amphibius. Monoposthium acanthum gen. n., sp. n., with typical cycloposthiid structure, differs from other genera in having only 1 caudalium. Parentodinium gen. n., represented by P. africanum sp. n. and P. ostrea sp. n., although superficially resembling Entodinium, is considered to be a relatively unspecialized cycloposthiid.  相似文献   

5.
We have analyzed conformational changes that occur at the interface between the light (VL) and heavy (VH) chains in antibody variable fragments upon binding to antigens. We wrote and applied the Tiny Probe program that computes the buried atomic contact surface area of three‐dimensional structures to evaluate changes in compactness of the VL–VH interface between bound and unbound antibodies. We found three categories of these changes, which correlated with the size of the antigen. Upon binding, medium‐sized nonprotein antigens cause an opening of the VL–VH interface (less compact), small antigens or haptens cause a closure of the interface (more compact), whereas large protein antigens have little effect on the compactness of the VL–VH interface. The largest changes in the atomic buried contact surface area at the VL–VH interface occur in residue pairs providing two ‘shock absorbers’ between the edge β‐strands of the VL and VH β‐sheets forming the antibody binding site. Importantly, the correlation between the size of antigens and conformational changes indicates that the VL–VH interface in antibodies plays a significant role in the antigen binding process. Furthermore, as the energy involved in such a motion is significant (up to 3 kcal/mol), these results provide a general mechanism for how residues distant from the combining site can significantly alter the affinity of an antibody for its antigen. Thus, mutations introduced at the VL–VH interface can be used to change antibody binding affinity with antigens. Due to the tightly packed VL–VH interface, the introduction of random mutations is not advisable. Rather our analysis suggests that concerted mutations of residues preceding CDRL2 and following CDRH3 or residues preceding CDRH2 and at the end of CDRL3 are most likely to alter or improve antigen binding affinity. Copyright © 1999 John Wiley & Sons, Ltd.  相似文献   

6.
Molecular models of amylopectin were created and investigated by computer simulation. First, single and double helices of various lengths were constructed. The 1 → 6 branching in double and single helices of amylopectin was studied. Subunits of single helices, double helices, and branch points were used as building blocks of larger systems. The possible makeup of amylopectin unit clusters was investigated via a series of models, including single–single, double–single, and double–double helix systems. The lengths of the single helix section that linked two branch points (internal chains) was systematically varied between values of 0–10 glucose residues. It was found that certain internal chain lengths lead to parallel double helices. Thus, it was postulated that the length of internal chains may determine the degree of local crystallinity. Furthermore, it was noted that some of the low‐energy arrangement of double helices could be superimposed on either the two adjacent and nonadjacent double helices of crystalline A and B starch polymorphs. In other cases, the distance between the double helices is so large that it may in fact be a model for branching between two amylopectin crystals or unit clusters. Results obtained through this work were corroborated, where possible, with information available from crystallographic, branching, and enzymatic studies. © 1999 John Wiley & Sons, Inc. Biopoly 50: 381–390, 1999  相似文献   

7.
Primary (AL) amyloidosis results from the pathologic deposition of monoclonal light chains as amyloid fibrils. Studies of recombinant-derived variable region (VL) fragments of these proteins have shown an inverse relationship between thermodynamic stability and fibrillogenic potential. Further, ionic interactions within the VL domain were predicted to influence the kinetics of light chain fibrillogenicity, as evidenced from our analyses of a relatively stable Vlambda6 protein (Jto) with a long range electrostatic interaction between Asp and Arg side chains at position 29 and 68, respectively, and an unstable, highly fibrillogenic Vlambda6 protein (Wil) that had neutral amino acids at these locations. To test this hypothesis, we have generated two Jto-related mutants designed to disrupt the interaction between Asp 29 and Arg 68 (JtoD29A and JtoR68S). Although the thermodynamic stabilities of unfolding for these two molecules were identical, they exhibited very different kinetics of fibril formation: the rate of JtoD29A fibrillogenesis was slow and comparable to the parent molecule, whereas that of JtoR68S was significantly faster. High-resolution X-ray diffraction analyses of crystals prepared from the two mutants having the same space group and unit cell dimensions revealed no significant main-chain conformational changes. However, several notable side-chain alterations were observed in JtoR68S, as compared with JtoD29A, that resulted in the solvent exposure of a greater hydrophobic surface and modifications in the electrostatic potential surface. We posit that these differences contributed to the enhanced fibrillogenic potential of the Arg 68 mutant, since both Jto mutants lacked the intrachain ionic interaction and were equivalently unstable. The information gleaned from our studies has provided insight into structural parameters that in addition to overall thermodynamic stability, contribute to the fibril forming propensity of immunoglobulin light chains.  相似文献   

8.
The human insulin receptor (IR) homodimer is heavily glycosylated and contains a total of 19 predicted N-linked glycosylation sites in each monomer. The recent crystal structure of the IR ectodomain shows electron density consistent with N-linked glycosylation at the majority of sites present in the construct. Here, we describe a refined structure of the IR ectodomain that incorporates all of the N-linked glycans and reveals the extent to which the attached glycans mask the surface of the IR dimer from interaction with antibodies or other potential therapeutic binding proteins. The usefulness of Fab complexation in the crystallization of heavily glycosylated proteins is also discussed. The compositions of the glycans on IR expressed in CHO-K1 cells and the glycosylation deficient Lec8 cell line were determined by protease digestion, glycopeptide purification, amino acid sequence analysis, and mass spectrometry. Collectively the data reveal: multiple species of complex glycan at residues 25, 255, 295, 418, 606, 624, 742, 755, and 893 (IR-B numbering); multiple species of high-mannose glycan at residues 111 and 514; a single species of complex glycan at residue 671; and a single species of high-mannose glycan at residue 215. Residue 16 exhibited a mixture of complex, hybrid, and high-mannose glycan species. Of the remaining five predicted N-linked sites, those at residues 397 and 906 were confirmed by amino acid sequencing to be glycosylated, while that at residue 78 and the atypical (NKC) site at residue 282 were not glycosylated. The peptide containing the final site at residue 337 was not recovered but is seen to be glycosylated in the electron density maps of the IR ectodomain. The model of the fully glycosylated IR reveals that the sites carrying high-mannose glycans lie at positions of relatively low steric accessibility.  相似文献   

9.
Aims: Validation of stereology and three‐dimensional reconstruction for monitoring the probiotic effect of Aeromonas hydrophila on the gut development of germ‐free Artemia franciscana nauplii. Methods and Results: Germ‐free Artemia nauplii were cultured using Baker’s yeast and dead Aer. hydrophila. Live Aer. hydrophila were added on the first day to the treatment group. The gut length and volume were monitored on days two and four using stereology and three‐dimensional reconstruction. Both methods showed comparable results. Stereology was least labour intensive to estimate volumes, while three‐dimensional reconstructions rendered architectural and topographical data of the gut. Moreover, a positive effect of probiotic bacterium, Aer. hydrophila is likely. Conclusion: Slight increment in the growth of the digestive tract of A. franciscana nauplii exerted by probiotic bacteria could be detected using stereology and three‐dimensional reconstruction. Significance and Impact of the Study: The gnotobiotic Artemia rearing system is unique to investigate the effects of micro‐organisms on the development of nauplii. However, in the base of this model system, only survival counts and length measurements exist as monitoring tools. Therefore, additional tools such as stereology and three‐dimensional reconstruction are prerequisite to obtain more powerful analysis.  相似文献   

10.
Environmental heterogeneity is regarded as one of the most important factors governing species richness gradients. An increase in available niche space, provision of refuges and opportunities for isolation and divergent adaptation are thought to enhance species coexistence, persistence and diversification. However, the extent and generality of positive heterogeneity–richness relationships are still debated. Apart from widespread evidence supporting positive relationships, negative and hump‐shaped relationships have also been reported. In a meta‐analysis of 1148 data points from 192 studies worldwide, we examine the strength and direction of the relationship between spatial environmental heterogeneity and species richness of terrestrial plants and animals. We find that separate effects of heterogeneity in land cover, vegetation, climate, soil and topography are significantly positive, with vegetation and topographic heterogeneity showing particularly strong associations with species richness. The use of equal‐area study units, spatial grain and spatial extent emerge as key factors influencing the strength of heterogeneity–richness relationships, highlighting the pervasive influence of spatial scale in heterogeneity–richness studies. We provide the first quantitative support for the generality of positive heterogeneity–richness relationships across heterogeneity components, habitat types, taxa and spatial scales from landscape to global extents, and identify specific needs for future comparative heterogeneity–richness research.  相似文献   

11.
Reactions of N,N′‐bis (salicylidene)‐1,2‐cyclohexanediamine (H2L) with mixed lanthanide counterions of LnCl3·6H2O and Ln (NO3)3·6H2O afford six H2L lanthanide coordination polymers, e.g. {[Pr(H2L)2(NO3)2Cl]·2CH2Cl2}n ( 1 ); {[Ln(H2L)1.5(NO3)3]2·5CHCl3·mCH3OH}n [Ln = Sm ( 2 ), Eu ( 3 ), Gd ( 4 ), Tb ( 5 ) and Yb ( 6 ); m = 1 ( 2 – 5 ); m = 0 ( 6 )]. X‐ray crystallographic analysis reveals that complex 1 exhibits three‐dimensional diamondoid topologic structure and complexes 2 – 6 are of two‐dimensional structure. Luminescent spectra show that complexes 1 and 6 have characteristic near‐infrared (NIR) emission of praseodymium (III) and ytterbium (III) ions and complexes 2 – 5 emit luminescence in the visible region. Complexes 3 and 6 reveal sensitive luminescence responses to formaldehyde.  相似文献   

12.
Mitochondrial pre‐messenger RNAs in kinetoplastid protozoa are substrates of uridylate‐specific RNA editing. RNA editing converts non‐functional pre‐mRNAs into translatable molecules and can generate protein diversity by alternative editing. Although several editing complexes have been described, their structure and relationship is unknown. Here, we report the isolation of functionally active RNA editing complexes by a multistep purification procedure. We show that the endogenous isolates contain two subpopulations of ~20S and ~35–40S and present the three‐dimensional structures of both complexes by electron microscopy. The ~35–40S complexes consist of a platform density packed against a semispherical element. The ~20S complexes are composed of two subdomains connected by an interface. The two particles are structurally related, and we show that RNA binding is a main determinant for the interconversion of the two complexes. The ~20S editosomes contain an RNA‐binding site, which binds gRNA, pre‐mRNA and gRNA/pre‐mRNA hybrid molecules with nanomolar affinity. Variability analysis indicates that subsets of complexes lack or possess additional domains, suggesting binding sites for components. Together, a picture of the RNA editing machinery is provided.  相似文献   

13.
Analysis of the three-dimensional structures of three closely related mesophilic, thermophilic, and hyperthermophilic alcohol dehydrogenases (ADHs) from the respective microorganisms Clostridium beijerinckii (CbADH), Entamoeba histolytica (EhADH1), and Thermoanaerobacter brockii (TbADH) suggested that a unique, strategically located proline residue (Pro100) might be crucial for maintaining the thermal stability of EhADH1. To determine whether proline substitution at this position in TbADH and CbADH would affect thermal stability, we used site-directed mutagenesis to replace the complementary residues in both enzymes with proline. The results showed that replacing Gln100 with proline significantly enhanced the thermal stability of the mesophilic ADH: DeltaT(1/2) (60 min) = + 8 degrees C (temperature of 50% inactivation after incubation for 60 min), DeltaT(1/2) (CD) = +11.5 degrees C (temperature at which 50% of the original CD signal at 218 nm is lost upon heating between 30 degrees and 98 degrees C). A His100 --> Pro substitution in the thermophilic TbADH had no effect on its thermostability. An analysis of the three-dimensional structure of the crystallized thermostable mutant Q100P-CbADH suggested that the proline residue at position 100 stabilized the enzyme by reinforcing hydrophobic interactions and by reducing the flexibility of a loop at this strategic region.  相似文献   

14.
A set of cranial characters was examined in the fruit bats Rousettus egyptiacus and Eidolon helvum to compare trends and relative importance of major components of bilateral morphometric variation, and their relationship with character size. Using two‐way, sides‐by‐individuals ANOVA , four components of variation were estimated for each bilateral variable: individual variation (I), directional asymmetry (DA), non‐directional asymmetry (NDA) and measurement error (E). Both species exhibit similar major trends of variation in asymmetry across characters, as shown by principal component analysis, using variance components as variables. Degree of interspecific congruence among characters was confirmed by a two‐way ANOVA with species and variance components as fixed factors. Congruence of asymmetry patterns between species suggests that the concept of population asymmetry parameter (PAP) could be extended to higher hierarchies. PAPs above the species level may result from common mechanisms or similar developmental constraints acting on species’ buffering capacities and morphological integration processes.  相似文献   

15.
Bovine seminal ribonuclease, a homodimeric enzyme joined covalently by two interchain disulphide bonds, is an equilibrium mixture of two conformational isomers, MxM and M=M. The major form, MxM, whose crystal structure has been previously determined at 1.9 A resolution, presents the swapping of the N-terminal segments (residues 1-15) and composite active sites formed by residues of different chains. The three-dimensional domain swapping does not occur in the M=M form. The different fold of each N-terminal tail is directed by the hinge loop (residue 16-22) connecting the swapping domain to the body of the protein. Reduction and alkylation of interchain disulphide bridges produce a monomeric derivative and a noncovalent swapped dimer, which are both active. The free and nucleotide-bound forms of the monomer have been crystallized at an alkaline pH and refined at 1.45 and 1.65 A resolution, respectively. In both cases, the N-terminal fragment is folded on the main body of the protein to produce an intact active site and a chain architecture very similar to that of bovine pancreatic ribonuclease. In this new fold of the seminal chain, the hinge loop is disordered. Despite the difference between the tertiary structure of the monomer and that of the chains in the MxM form, the active sites of the two enzymes are virtually indistinguishable. Furthermore, the structure of the liganded enzyme represents the first example of a ribonuclease complex studied at an alkaline pH and provides new information on the binding of a nucleotide when the catalytic histidines are deprotonated.  相似文献   

16.
Interleukin 6 (IL-6) belongs to a large group of secreted proteins called cytokines functioning to mediate and regulate immunity, inflammation, and hematopoiesis with direct effects on cell proliferation, apoptosis, and differentiation. Along with the IL-6 protein, two of its splice variants, IL-6delta2 and IL-6delta4, were reported to be transcribed or expressed in vivo in human, and the mRNAs of IL-6delta3 and IL-6delta5 had been observed in mouse. While the existence of different splice variants of IL-6 has been shown, very little is known on how the structural modifications of IL-6 resulting from the formation of the different splice variants may alter cytokine functions. We have analyzed the potential effects splicing would have on interactions with the cell surface receptor complex. We (1) constructed three-dimensional structures of the IL-6 splice variants, IL-6delta2, IL-6delta3, and IL-6delta4, with the assumption that an interleukin splice variant as a folded protein should retain a functional hydrophobic core; (2) reconstructed the ternary structural complexes consisting of the modeled IL-6 splice variants, the IL-6 receptor molecule (IL-6R) and the dimeric signal-transducing protein, gp130, and (3) analyzed all complexes and made comparisons with the X-ray structure of the wild-type IL-6 complex. We identified three separate sites on IL-6 where interactions are made with IL-6R and with each of the two copies of gp130. The structural consequences of losing an exon lead to a unique pattern of lost interaction with different components of the receptor complex. Thus, in IL-6 and its splice variants, the exons appear to have compartmentalized roles contributing to the combined function of the cytokine. The modeled interactions suggest that splice variants could act as antagonists, and that IL-6delta2, missing the signal peptide, would be a cytoplasmic protein and be released and interact with nearby cell-surface receptors when cells are damaged. We argue that in the case of IL-6, helix E may act as a "silent secondary structure," which only has an active role when it substitutes for a part of the hydrophobic core, for example, replacing helix A in IL-6delta2.  相似文献   

17.
Unlike other known anti‐fluorescein antibodies, the monoclonal antibody 43.1 is directed toward the fluorescein's carboxyl phenyl moiety. It demonstrates a very high affinity (KD ~ 70 pM) and a fast association rate (kon ~ 2 × 107 M?1 s?1). The three‐dimensional structure of the Fab 43.1—fluorescein complex was resolved at 2.4 Å resolution. The antibody binding site is exclusively assembled by the CDR loops. It is comprised of a 14 Å groove‐shaped entrance leading to a 9 Å by 7 Å binding pocket. The highly polar binding pocket complementary encloses the fluorescein's carboxyphenyl moiety and tightly fixes it by multiple hydrogen bonds. The fluorescein's xanthene ring is embedded in the more hydrophobic groove and stacked between the side chains of Tyr37L and of Arg99H providing conditions for an excited state electron transfer process. In comparison to fluorescein, the absorption spectrum of the complex in the visible region is shifted to the “red” by 23 nm. The complex demonstrates a very weak fluorescence (Φc = 0.0018) with two short lifetime components: 0.03 ns (47%) and 0.8 ns (24%), which reflects a 99.8% fluorescein emission quenching effect upon complex formation. The antibody 43.1 binds fluorescein with remarkable affinity, fast association rate, and strongly quenches its emission. Therefore, it may present a practical interest in applications such as molecular sensors and switches. © 2016 Wiley Periodicals, Inc. Biopolymers 105: 234–243, 2016.  相似文献   

18.
19.
Abstract APETALA1 (AP1) and CAULIFLOWER (CAL) are a pair of paralogous genes that were generated through the pre‐Brassicaceae whole‐genome duplication event. AP1 and CAL have both partially redundant and unique functions. Previous studies have shown that the K and C regions of their proteins are essential for the functional divergence. However, which differences in these regions are the major contributors and how the differences were accumulated remain unknown. In the present study, we compared the sequences of the two proteins and identified five gaps and 55 amino acid replacements between them. Investigation of genomic sequences further indicated that the differences in the proteins were caused by non‐synonymous substitutions and changes in exon–intron structures. Reconstruction of three‐dimensional structures revealed that the sequence divergence of AP1 and CAL has resulted in differences between the two in terms of the number, length, position and orientation of α‐helices, especially in the K and C regions. Comparisons of sequences and three‐dimensional structures of ancestral proteins with AP1 and CAL suggest that the ancestral AP1 protein experienced fewer changes, whereas the ancestral CAL protein accumulated more changes shortly after gene duplication, relative to their common ancestor. Thereafter, AP1‐like proteins experienced few mutations, whereas CAL‐like proteins were not conserved until the diversification of the Brassicaceae lineage I. This indicates that AP1‐ and CAL‐like proteins evolved asymmetrically after gene duplication. These findings provide new insights into the functional divergence of AP1 and CAL genes.  相似文献   

20.
Background information. DC (dendritic cells) continuously capture pathogens and process them into small peptides within the endolysosomal compartment, the MIIC (MHC class II‐containing compartment). In MIICs peptides are loaded on to MHC class II and rapidly redistributed to the cell surface. This redistribution is accompanied by profound changes of the MIICs into tubular structures. An emerging concept is that MIIC tubulation provides a means to transport MHC class II—peptide complexes to the cell surface, either directly or through vesicular intermediates. To obtain spatial information on the reorganization of the MIICs during DC maturation, we performed electron tomography on cryo‐immobilized and freeze‐substituted mouse DCs after stimulation with LPS (lipopolysaccharide). Results. In non‐stimulated DCs, MIICs are mostly spherical. After 3 h of LPS stimulation, individual MIICs transform into tubular structures. Three‐dimensional reconstruction showed that the MIICs frequently display fusion profiles and after 6 h of LPS stimulation, MIICs become more interconnected, thereby creating large MIIC reticula. Microtubules and microfilaments align these MIICs and reveal physical connections. In our tomograms we also identified a separate population of MIIC‐like intermediates, particularly at extended ends of MIIC tubules and in close proximity to the trans‐Golgi network. No fusion events were captured between reticular MIICs and the plasma membrane. Conclusions. Our results indicate that MIICs have the capacity to fuse together, whereby the cytoskeleton possibly provides a scaffold for the MIIC shape change and directionality. MIIC‐like intermediates may represent MHC class II carriers.  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