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.
Morphology and crystal structure of a recombinant silk-like molecule, SLP4, were studied. Wide angle x-ray scattering (WAXS) and electron diffraction revealed that SLP4 lyophilized powder and thin films were isomorphic with the silk I crystal structure. Transmission electron microscopy of SLP4 thin films demonstrated a morphology of flat, variable width, crystallites that may aggregate in an epitaxial manner. Theoretical diffraction patterns from silk I crystal structure models were critically compared with SLP4 WAXS data. The analysis concluded that while the crankshaft model is capable of describing details of the SLP4 structural data well, the out-of-register model does not explain the experimental results. In particular, the predicted intensities of the crystallographic reflections for the out-of-register model are inconsistent with the SLP4 WAXS data. © 1998 John Wiley & Sons, Inc. Biopoly 45: 307–321, 1998  相似文献   

2.
Silk cocoons provide protection to silkworm from biotic and abiotic hazards during the immobile pupal phase of the lifecycle of silkworms. Protection is particularly important for the wild silk cocoons reared in an open and harsh environment. To understand whether some of the cocoon components resist growth of microorganisms, in vitro studies were performed using gram negative bacteria Escherichia coli (E. coli) to investigate antibacterial properties of silk fiber, silk gum, and calcium oxalate crystals embedded inside some cocoons. The results show that the previously reported antibacterial properties of silk cocoons are actually due to residues of chemicals used to isolate/purify cocoon elements, and properly isolated silk fiber, gum, and embedded crystals free from such residues do not have inherent resistance to E. coli. This study removes the uncertainty created by previous studies over the presence of antibacterial properties of silk cocoons, particularly the silk gum and sericin. © 2013 Wiley Periodicals, Inc. Biopolymers 101: 237–245, 2014.  相似文献   

3.
It is important to resolve the structure of Bombyx mori silk fibroin before spinning (silk I) and after spinning (silk II), and the mechanism of the structural transition during fiber formation in developing new silk-like fiber. The silk I structure has been recently resolved by (13)C solid-state NMR as a "repeated beta-turn type II structure." Here, we used (13)C solid-state NMR to clarify the heterogeneous structure of the natural fiber from Bombyx mori silk fibroin in the silk II form. Interestingly, the (13)C CP/MAS NMR revealed a broad and asymmetric peak for the Ala Cbeta carbon. The relative proportions of the various heterogeneous components were determined from their relative peak intensities after line shape deconvolution. Namely, for 56% crystalline fraction (mainly repeated Ala-Gly-Ser-Gly-Ala-Gly sequences), 18% distorted beta-turn, 13% beta-sheet (parallel Ala residues), and 25% beta-sheet (alternating Ala residues). The remaining fraction of 44% amorphous Tyr-rich region, 22% in both distorted beta-turn and distorted beta-sheet. Such a heterogeneous structure including distorted beta-turn can be observed for the peptides (AG)(n) (n > 9 ). The structural change from silk I to silk II occurs exclusively for the sequence (Ala-Gly-Ser-Gly-Ala-Gly)(n) in B. mori silk fibroin. The generation of the heterogeneous structure can be studied by change in the Ala Cbeta peak of (13)C CP/MAS NMR spectra of the silk fibroin samples with different stretching ratios.  相似文献   

4.
A new crystalline polymorph of Bombyx mori silk, which forms at the air–water interface, has been characterized. A previous study found this structure to be trigonal, and to be distinctly different than the two previously observed silk crystal structures, silk I and silk II. This new structure was named silk III. Identification of this new silk polymorph was based on evidence from transmission electron microscopy and electron diffraction, coupled with molecular modeling. In the current paper, additional data enables us to refine our model of the silk III structure. Some single crystal electron diffraction patterns indicate a deviation in symmetry away from a perfect trigonal unit cell to monoclinic unit cell. The detailed shape of the powder diffraction peaks also supports a monoclinic cell. The monoclinic crystal structure has an nonprimitive unit cell incorporating a slightly distorted hexagonal packing of silk molecular helices. The chains each assume a threefold helical conformation, resulting in a crystal structure similar to that observed for polyglycine II, but with some additional sheet-like packing features common to the threefold helical crystalline forms of many glycine-rich polypeptides. © 1997 John Wiley & Sons, Inc. Biopoly 42: 705–717, 1997  相似文献   

5.
6.
The genus Oxytate L. Koch, 1878 comprises a homogeneous group of nocturnal crab spiders that have silk apparatuses even though they do not spin webs to trap prey. We examined the microstructure of the silk spinning apparatus of the green crab spider Oxytate striatipes, using field emission scanning electron microscopy. The silk glands of the spider were classified into three types: ampullate, pyriform and aciniform. The spigots of these three types of silk gland occur in both sexes. Two pairs of major ampullate glands send secretory ductules to the anterior spinnerets, and another two pairs of minor ampullate glands supply the median spinnerets. In addition, the pyriform glands send ductules to the anterior spinnerets (45 pairs in females and 40 pairs in males), and the aciniform glands feed silk into the median (9–12 pairs in females and 7–10 pairs in males) and the posterior (30 pairs in both sexes) spinnerets. The spigot system of O. striatipes is simpler and more primitive than other wandering spiders: even the female spiders possess neither tubuliform glands for cocoon production nor triad spigots for web‐building.  相似文献   

7.
Bombyx mori silk fibroin is a fibrous protein whose fiber is extremely strong and tough, although it is produced by the silkworm at room temperature and from an aqueous solution. The primary structure is mainly Ala-Gly alternative copolypeptide, but Gly-Ala-Ala-Ser units appear frequently and periodically. Thus, this study aims at elucidating the role of such Gly-Ala-Ala-Ser units on the secondary structure. The sequential model peptides containing Gly-Ala-Ala-Ser units selected from the primary structure of B. mori silk fibroin were synthesized, and their secondary structure was studied with (13)C CP/MAS NMR and wide-angle X-ray scattering. The (13)C isotope labeling of the peptides and the (13)C conformation-dependent chemical shifts were used for the purpose. The Ala-Ala units take antiparallel beta-sheet structure locally, and the introduction of one Ala-Ala unit in (Ala-Gly)(15) chain promotes dramatical structural changes from silk I (repeated beta-turn type II structure) to silk II (antiparallel beta-sheet structure). Thus, the presence of Ala-Ala units in B. mori silk fibroin chain will be one of the inducing factors of the structural transition for silk fiber formation. The role of Tyr residue in the peptide chain was also studied and clarified to induce "locally nonordered structure."  相似文献   

8.
Transmission electron microscopy was used to investigate the supramolecular structure of Araneus diadematus (garden spider) cocoon silk. Electron diffraction patterns contain features which are consistent with the presence of non-periodic lattice crystals, i.e. highly frustrated crystalline regions as identified previously in the major ampullate silk (MAS, dragline) of Nephila clavipes spiders. The diffraction patterns further suggest that crystals in A. diadematus cocoon silk may be twisted parallel to the chain direction, offering a potential explanation for the lower tensile stiffness of this fibre relative to MAS.  相似文献   

9.
The metastable state silk I structures of Bombyx mori silk fibroin in the solid state were studied on the basis of 15N- and 13C-nmr chemical shifts of Ala, Ser, and Gly residues. The 15N cross-polarization magic angle spinning (CP/MAS) nmr spectra of the precipitated fraction after chymotrypsin hydrolysis of B. mori silk fibroin with the silk I and silk II forms were measured to determine the 15N chemical shifts of Gly, Ala, and Ser residues. For comparison, 15N CP/MAS nmr chemical shifts of Ala were measured for [15N] Ala Philosamia cynthia ricini silk fibroin with antiparallel β-sheet and α-helix forms. The 13C CP/MAS nmr chemical shifts of Ala, Ser, and Gly residues of B. mori silk fibroin with the silk I and silk II forms, as well as 13C CP/MAS nmr chemical shifts of Ala residue of P. c. ricini silk fibroin with β-sheet and α-helix forms, are used for the examination of the silk I structure. Both silk I and α-helix peaks are shifted to a lower field than silk II (β-sheet) for the Cα carbons of the Ala residues, while both Cβ carbon peaks are shifted to higher field. However, the silk I peak of the 15N nucleus of the Ala residue is shifted to lower field than the silk II peak, but the α-helix peak is shifted to high field. Thus, the difference in the structure between the silk I and α-helix is reflected in a different manner between the 13C and 15N chemical shifts. The Cα and Cβ chemical shift contour plots for Ala and Ser residues, and the Cα plot for the Gly residue, were prepared from the Protein Data Bank data obtained for 12 proteins and used for discussing the silk I structure quantitatively from the conformation-dependent chemical shifts. The plots reported by Le and Oldfield for 15N chemical shifts were also used for the purpose. All these chemical shift data support Fossey's model (Ala: ϕ = −80°, φ = 150°, Gly: ϕ = −150°, φ = 80°) and do not support Lotz and Keith's model (Ala: ϕ = −104.6°, φ = 112.2°, Gly: ϕ = 79.8°, φ = 49.7° or Ala: ϕ = −124.5°, φ = 88.2°, Gly: ϕ = −49.8°, φ = −76.1°) as the silk I structure. © 1997 John Wiley & Sons, Inc.  相似文献   

10.
Condensation of single molecules from solution into crystals represents a transition between distinct energetic states. In solution, the atomic interactions within the molecule dominate. In the crystalline state, however, a set of additional interactions are formed between molecules in close contact in the lattice—these are the packing interactions. The crystal structures of d(CCCT), d(TAACCC), d(CCCAAT), and d(AACCCC) have in common a four-stranded intercalated cytosine segment, built by stacked layers of cytosine · cytosine+ (C · C+) base pairs coming from two parallel duplexes that intercalate into each other with opposite polarity. The intercalated cytosine segments in these structures are similar in their geometry, even though the sequences crystallized in different space groups. In the crystals, adenine and thymine residues of the sequences are used to build the three-dimensional crystal lattice by elaborately interacting with symmetry-related molecules. The packing elements observed provide novel insight about the copious ways in which nucleic acid molecules can interact with each other—for example, when folded in more complicated higher order structures, such as mRNA and chromatin. © 1998 John Wiley & Sons, Inc. Biopoly 44: 257–267, 1997  相似文献   

11.
Spider silk has been studied extensively for its attractive mechanical properties and potential applications in medicine and industry. The production of spider silk, however, has been lagging behind for lack of suitable systems. Our approach focuses on solving the production of spider silk by designing, expressing, purifying and characterizing the silk from cylindriform glands. We show that the cylindriform silk protein, in contrast to the commonly used dragline silk protein, is fully folded and stable in solution. With the help of GFP as a fusion tag we enhanced the expression of the silk protein in Escherichia coli and could optimize the downstream processing. Secondary structures analysis by circular dichroism and FTIR shows that the GFP‐silk fusion protein is predominantly α‐helical, and that pH can trigger a α‐ to β‐transition resulting in aggregation. Structural analysis by small angle X‐ray scattering suggests that the GFP‐Silk exists in the form of a hexamer in solution. © 2013 Wiley Periodicals, Inc. Biopolymers 101: 378–390, 2014.  相似文献   

12.
林森珠  陈格飞  孟清 《生物工程学报》2016,32(12):1704-1714
为建立高效快捷的蛛丝功能化修饰平台,蛋白质内含子的反式剪接技术被首次应用于重组蛛丝的功能化修饰。在体外通过Ssp Dna B的反式剪接作用,在蛋白质水平上将12 k Da泛素相关修饰蛋白(SUMO)与蛛丝蛋白(W2CT)连接形成功能化蛛丝蛋白SUMOW2CT。修饰后SUMOW2CT与W2CT均能形成纳米至微米级的丝纤维,但SUMOW2CT自动成丝速度明显下降且产量约为W2CT的一半。与W2CT丝纤维(W)相似,SUMOW2CT丝纤维(UW)不具有超收缩能力和对2%SDS不耐受,但机械性能低于W2CT丝纤维。功能化蛋白SUMOW2CT形成的丝纤维中SUMO蛋白仍保持着正确三维结构,可被SUMO蛋白酶酶切。外源功能化蛋白质虽在一定程度上降低了丝的形成速度和机械性能,但修饰上的功能化蛋白仍保持着生物活性,表明断裂蛋白质内含子介导的蛛丝修饰平台成功建立,也为蛛丝的功能化修饰和应用奠定了坚实的技术基础。  相似文献   

13.
Several methods have been proposed for linkage analysis of complex traits with unknown mode of inheritance. These methods include the LOD score maximized over disease models (MMLS) and the "nonparametric" linkage (NPL) statistic. In previous work, we evaluated the increase of type I error when maximizing over two or more genetic models, and we compared the power of MMLS to detect linkage, in a number of complex modes of inheritance, with analysis assuming the true model. In the present study, we compare MMLS and NPL directly. We simulated 100 data sets with 20 families each, using 26 generating models: (1) 4 intermediate models (penetrance of heterozygote between that of the two homozygotes); (2) 6 two-locus additive models; and (3) 16 two-locus heterogeneity models (admixture alpha = 1.0,.7,.5, and.3; alpha = 1.0 replicates simple Mendelian models). For LOD scores, we assumed dominant and recessive inheritance with 50% penetrance. We took the higher of the two maximum LOD scores and subtracted 0.3 to correct for multiple tests (MMLS-C). We compared expected maximum LOD scores and power, using MMLS-C and NPL as well as the true model. Since NPL uses only the affected family members, we also performed an affecteds-only analysis using MMLS-C. The MMLS-C was both uniformly more powerful than NPL for most cases we examined, except when linkage information was low, and close to the results for the true model under locus heterogeneity. We still found better power for the MMLS-C compared with NPL in affecteds-only analysis. The results show that use of two simple modes of inheritance at a fixed penetrance can have more power than NPL when the trait mode of inheritance is complex and when there is heterogeneity in the data set.  相似文献   

14.
Although spider silk has been studied for a number of years the structures of the proteins involved have yet to be definitely determined. X-ray diffraction and solid-state nuclear magnetic resonance (NMR) were used to study major ampullate (dragline) silk from Nephila clavipes. The silk was studied in its natural state, in the supercontacted state and in the restretched state following supercontraction. The natural silk structure is dominated by β-sheets aligned parallel to the fiber axis. Supercontraction is characterized by randomizing of the orientation of the β-sheet. When the fiber is restretched alignment is regained. However, the same reorientation was observed for wetting of minor ampullate silk which does not supercontract. Thus, the reorientation of β-sheets alone cannot explain the supercontraction in dragline silk. Cocoon silk showed very little β-sheet orientation in the natural state and there were no changes upon wetting. NMR and X-ray diffraction data are consistent with the β-sheets arising from the poly-alanine sequences known to be present in the proteins of major ampullate silk as has been proposed previously. © 1997 John Wiley & Sons, Ltd.  相似文献   

15.
Spiders attach silken threads to substrates by means of glue-coated nanofibers (piriform silk), spun into disc-like structures. The organization and ultrastructure of this nano-composite silk are largely unknown, despite their implications for the biomechanical function and material properties of thread anchorages. In this work, the ultrastructure of silken attachment discs was studied in representatives of four spider families with Transmission Electron Microscopy to facilitate a mechanistic understanding of piriform silk function across spiders. Based on previous findings from comparative studies of piriform silk gland morphology, we hypothesized that the fibre-glue proportion of piriform silk differs in different spiders, while the composition of fibre and glue fractions is consistent. Results confirmed large differences in the relative proportion of glue with low amounts in the orb weaver Nephila senegalensis (Araneidae) and the hunting spider Cupiennius salei (Ctenidae), larger amounts in the cobweb spider Parasteatoda tepidariorum (Theridiidae) and a complete reduction of the fibrous component in the haplogyne spider Pholcus phalangioides (Pholcidae). We rejected our hypothesis that glue ultrastructure is consistent. The glue is a colloid with polymeric and fluid fractions that strongly differ in proportions and assembly. We further confirmed that in all species studied both dragline and piriform silk fibres do not make contact with the environmental substrate. Instead, adhesion is established by a thin dense skin layer of the piriform glue. These results advance our understanding of piriform silk function and the interspecific variation of its properties, which is significant for spider biology, web function and the bioengineering of silk.  相似文献   

16.
Taddei P  Asakura T  Yao J  Monti P 《Biopolymers》2004,75(4):314-324
For a deeper insight into the structure of Bombyx mori silk fibroin, some model peptides containing tyrosine (Y), valine (V), and serine (S) in the basic (AG)n sequence were synthesized by the solid-phase method and analyzed by Raman spectroscopy in order to clarify their conformation and to evaluate the formation and/or disruption of the ordered structure typical of B. mori silk fibroin upon incorporation of Y, V, and S residues into the basic (AG)n sequence. The Raman results indicated that the silk I structure remains stable only when the Y residue is positioned near the chain terminus; otherwise, a silk I --> silk II conformational transition occurs. The peptides AGVGAGYGAGVGAGYGAGVGAGYG(AG)3 and (AG)3YG(AG)2VGYG(AG)3YG(AG)3 treated with LiBr revealed a prevalent silk II conformation; moreover, the former contained a higher amount of random coil than the latter. This result was explained in relation to the different degrees of interruption of the (AG)n sequence. The Raman analysis of the AGSGAG-containing samples confirmed that the AGSGAG hexapeptide is a good model for the silk II crystalline domain. As the number of AGSGAG repeating units decreased, the random coil content increased. The study of the Y domain (I850/I830 intensity ratio) allowed us to hypothesize that in the packing characteristic of Silk I and Silk II conformations the Y residues experience different environments and hydrogen-bonding arrangements; the packing typical of silk I structure traps the tyrosyl side chains in environments more unfavorable to phenoxyl hydrogen-bonding interactions.  相似文献   

17.
蜘蛛丝的分子结构与力学性能研究   总被引:2,自引:0,他引:2  
蜘蛛丝尤其是蜘蛛大囊状腺产生的拖丝,具有独特的机械性能,是自然界颇具应用潜力的生物材料。现代分子生物学技术使蜘蛛丝蛋白基因得以克隆,通过高分子物理化学手段方法的利用,有利于揭示蜘蛛丝蛋白质序列、分子结构、以及分子结构和力学性能之间的关系。对不同种类蜘蛛丝蛋白的深入研究,将为基因工程方法人工合成并改造蜘蛛丝成为可能。  相似文献   

18.
The mechanism of silk formation inApis mellifera salivary glands, during the 5th instar, was studied. Larval salivary glands were dissected and prepared for light and polarized light microscopy, as well as for scanning and transmission electron microscopy. The results showed that silk formation starts at the middle of the 5th instar and finishes at the end of the same instar. This process begins in the distal secretory portion of the gland, going towards the proximal secretory portion; and from the periphery to the center of the gland lumen. The silk proteins are released from the secretory cells as a homogeneous substance that polymerizes in the lumen to form compact birefringent tactoids. Secondly, the water absorption from the lumen secretion, carried out by secretory and duct cells, promotes aggregation of the tactoids that form a spiral-shape filament with a zigzag pattern. This pattern is also the results of the silk compression in the gland lumen and represents a high concentration of macromolecularly well-oriented silk proteins.  相似文献   

19.
The presence of oil is the major limitation to the regeneration of spun silk from waste silk. A pure culture fermentation process was developed to remove oil from waste silk. Fourteen strains were isolated from natural fermentation liquor of waste silk. The strain D3 showed highest lipase activity and oil‐removing ability. This strain was identified as Rhodococcus sp. on the basis of morphological and biochemical characteristics and 16S rRNA sequence analysis. The strain D3 was used to remove oil from waste silk by pure culture fermentation. The effects of various parameters on oil removal were investigated. A pH of 7.0, a temperature of 35–40°C, an incubation time of three days and an inoculum of 10% were optimum conditions for removing oil from waste silk by stain D3. This study shows that pure culture fermentation is a promising process to improve the oil removal of waste silk.  相似文献   

20.
A banded morphology has been observed for Bombyx mori silk fibroin films obtained from an aqueous hexane interface; the period of the banding is approximately 1 microm. Morphology and diffraction from different regions of the banded structure suggest that it is a free surface formed by a cholesteric liquid crystal. Truncated hexagonal lamellar crystallites of B. mori silk fibroin have been observed in films formed in the surface excess layer of fibroin at the interface between aqueous fibroin and hexane or chloroform. Based on initial crystallographic evidence, a three-fold helical conformation has been ascribed to the fibroin chains within the crystals. The chain conformation and crystalline habit appear to be similar to the silk III structure previously observed at the air-water interface (Valluzzi R, Gido SP. Biopolymers 1997;42:705-717; Valluzzi R, Gido S, Zhang W, Muller W, Kaplan D. Macromolecules 1996;29:8606-8614) but the crystalline packing is different. Diffraction data obtained for the crystallites are similar to diffraction behavior for a collagen-like model peptide. Diffraction patterns obtained from crystallized regions of the banded morphology can be indexed using the same unit cell as the hexagonal lamellar crystallites. Surfactancy of fibroin and subsequent aggregation and mesophase formation may help to explain the liquid crystallinity reported for silk, which is long suspected to play a role in the biological silk spinning process (Valluzzi R, Gido SP. Biopolymers 1997;42:705-717; Willcox, P. J.; Gido, SP, Muller W, Kaplan DL. Macromolecules 1996:29:5106-5110; Magoshi J, Magoshi Y, Nakamura S. In: Kaplan D, Adams W, Farmer B, Viney C, editors, Mechanism of Fiber Formation of Silkworm. Washington, DC: American Chemical Society 1994:292-310; Magoshi J, Magoshi Y, Nakamura S. J Appl Polym Sci Appl Polym Symp 1985;41:187-204; Magoshi J, Magoshi Y, Nakamura S. Polym Commun 1985;26:309.).  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