首页 | 本学科首页   官方微博 | 高级检索  
文章检索
  按 检索   检索词:      
出版年份:   被引次数:   他引次数: 提示:输入*表示无穷大
  收费全文   111篇
  免费   27篇
  2022年   1篇
  2021年   7篇
  2016年   5篇
  2015年   6篇
  2014年   4篇
  2013年   8篇
  2012年   8篇
  2011年   3篇
  2010年   5篇
  2009年   2篇
  2008年   7篇
  2007年   7篇
  2006年   8篇
  2005年   6篇
  2004年   8篇
  2003年   6篇
  2002年   5篇
  2001年   3篇
  2000年   2篇
  1999年   1篇
  1997年   1篇
  1995年   2篇
  1994年   1篇
  1993年   4篇
  1992年   1篇
  1986年   3篇
  1984年   1篇
  1983年   1篇
  1982年   1篇
  1981年   2篇
  1980年   1篇
  1978年   1篇
  1977年   1篇
  1976年   1篇
  1975年   1篇
  1974年   1篇
  1973年   1篇
  1972年   1篇
  1971年   3篇
  1970年   3篇
  1969年   1篇
  1968年   1篇
  1967年   1篇
  1961年   1篇
  1958年   1篇
排序方式: 共有138条查询结果,搜索用时 31 毫秒
91.
92.
Adeno-associated virus (AAV) has attracted considerable interest as a vector for gene therapy owing its lack of pathogenicity and the wealth of available serotypes with distinct tissue tropisms. One of the most promising isolates for vector development, based on its superior gene transfer efficiency to the liver in small animals compared to AAV type 2 (AAV2), is AAV8. Comparison of the in vivo gene transduction of rAAV2 and rAAV8 in mice showed that single amino acid exchanges in the 3-fold protrusions of AAV8 in the surface loops comprised of residues 581 to 584 and 589 to 592 to the corresponding amino acids of AAV2 and vice versa had a strong influence on transduction efficiency and tissue tropism. Surprisingly, not only did conversion of AAV8 to AAV2 cap sequences increase the transduction efficiency and change tissue tropism but so did the reciprocal conversion of AAV2 to AAV8. Insertion of new peptide motifs at position 590 in AAV8 also enabled retargeting of AAV8 capsids to specific tissues, suggesting that these sequences can interact with receptors on the cell surface. However, a neutralizing monoclonal antibody that binds to amino acids (588)QQNTA(592) of AAV8 does not prevent cell binding and virus uptake, indicating that this region is not necessary for receptor binding but rather that the antibody interferes with an essential step of postattachment processing in which the 3-fold protrusion is also involved. This study supports a multifunctional role of the 3-fold region of AAV capsids in the infection process.  相似文献   
93.
The small single-stranded DNA (ssDNA) bacteriophages of the subfamily Gokushovirinae were traditionally perceived as narrowly targeted, niche-specific viruses infecting obligate parasitic bacteria, such as Chlamydia. The advent of metagenomics revealed gokushoviruses to be widespread in global environmental samples. This study expands knowledge of gokushovirus diversity in the environment by developing a degenerate PCR assay to amplify a portion of the major capsid protein (MCP) gene of gokushoviruses. Over 500 amplicons were sequenced from 10 environmental samples (sediments, sewage, seawater and freshwater), revealing the ubiquity and high diversity of this understudied phage group. Residue-level conservation data generated from multiple alignments was combined with a predicted 3D structure, revealing a tendency for structurally internal residues to be more highly conserved than surface-presenting protein–protein or viral–host interaction domains. Aggregating this data set into a phylogenetic framework, many gokushovirus MCP clades contained samples from multiple environments, although distinct clades dominated the different samples. Antarctic sediment samples contained the most diverse gokushovirus communities, whereas freshwater springs from Florida were the least diverse. Whether the observed diversity is being driven by environmental factors or host-binding interactions remains an open question. The high environmental diversity of this previously overlooked ssDNA viral group necessitates further research elucidating their natural hosts and exploring their ecological roles.  相似文献   
94.
Protoporphyrinogen oxidase (PPOX), the penultimate enzyme in the haem biosynthetic pathway catalysers the six electron oxidation of protoporphyrinogen-IX to protoporphyrin-IX, in the presence of flavin adenine dinucleotide (FAD) and oxygen. In humans, partial defects in PPOX result in variegate porphyria. In this study, the FAD binding region in Myxococcus xanthus PPOX was analysed by engineering and characterising a selection of mutant proteins. Amino acid residues which interact with FAD via their side chains were selected for study. Mutants were characterised and compared with wild type protein. Characterisation included FAD quantitation, analysis of FAD spectra and kinetic assay. Results revealed that Serine 20 mutants could still bind FAD, but polarity in this position is favourable, yet not essential for the integrity of FAD binding. Study of Glutamate 39 mutants suggest that a negative charge at position 39 is clearly favoured for interaction with the ribose ring of FAD, as all non-conservative replacements could not bind sufficient FAD. Asparagine 441 appears not to be directly involved in FAD binding but rather in stabilizing the FAD, and polarity in this position appears important. Tryptophan 408 may play a role in orientating or stabilizing the bound substrate during catalysis, and a non-polar (or slightly polar) residue is favoured at this position; however, aromaticity in this position appears not to be critical. Overall this study sheds further light on how M. xanthus PPOX interacts with FAD.  相似文献   
95.
The amplitude of ultrasonic relaxation in aqueous solutions of disordered polysaccharides shows a marked increase with increasing degree of coil overlap and, at comparatively low concentrations, attains values comparable to those observed in polysaccharide gels. Mechanical spectroscopy studies indicate that, on the ultrasonic timescale, dynamic networks formed by polymer entanglement in solution are indistinguishable from true gels. In both cases the intense relaxations observed are attributed predominantly to motion of solvent within the polymer network. Due to the inherent stiffness of most polysaccharides, formation of a highly entangled network structure (with consequent enhancement of ultrasonic relaxation) occurs at much lower concentrations than for typical synthetic polymers. The onset of coil overlap (c1 transition) is accompanied by an abrupt change in the concentration dependence of ultrasonic velocity. Results for the conformationally rigid polysaccharide xanthan, suggest that velocity measurements may offer a convenient method for determination of c1 in systems where the normal viscometric characterisation is impossible.  相似文献   
96.
Sodium salicylate (1,000 μg/ml) was delivered through a drip irrigation system to agricultural field soils planted to tomato and infested with Pseudomonas putida PpG7, the host of the salicylate catabolic plasmid NAH7. In nonfumigated soils infested with approximately 103 CFU of PpG7 per g in the top 30 cm, population densities were increased up to 112-fold within 14 days of the initial application of salicylate compared with the densities in the respective nonamended soils. Mean season-long population densities of PpG7 in the top 30 cm of soil were significantly increased (P < 0.01) from 216 CFU/g in nonamended soils to 1,370 CFU/g in salicylate-amended soils. In the respective rhizosphere soils, mean population densities of PpG7 were significantly increased (P < 0.01) from 92 to 2,066 CFU/cm of root. Soil fumigation interacted (P < 0.01) with salicylate amendment and further increased the mean population densities of PpG7 in nonrhizosphere soil by an additional 5,689 CFU/g of soil. This fumigation effect was not detected in rhizosphere soils. The effect of salicylate in increasing population densities of PpG7 in soil also was affected by inoculum level, field site, and soil depth. Proportionate differences were greater in soils infested with approximately 103 CFU of PpG7 per g than in comparable soils infested with 105 CFU/g. In low-inoculum soils, increases from salicylate amendments were 26- and 29-fold in rhizosphere and nonrhizosphere soils, respectively, and in high-inoculum soils, the respective increases were 5.6- and 5-fold. No increases of fungi able to utilize salicylate were detected in soils amended with salicylate. However, soil fumigation with metham-sodium significantly reduced (P < 0.01) population densities of fungal salicylate utilizers in rhizosphere and nonrhizosphere soils.  相似文献   
97.
Human bocavirus (HBoV) was recently discovered and classified in the Bocavirus genus (family Parvoviridae, subfamily Parvovirinae) on the basis of genomic similarity to bovine parvovirus and canine minute virus. HBoV has been implicated in respiratory tract infections and gastroenteric disease in children worldwide, yet despite numerous epidemiological reports, there has been limited biochemical and molecular characterization of the virus. Reported here is the three-dimensional structure of recombinant HBoV capsids, assembled from viral protein 2 (VP2), at 7.9-Å resolution as determined by cryo-electron microscopy and image reconstruction. A pseudo-atomic model of HBoV VP2 was derived from sequence alignment analysis and knowledge of the crystal structure of human parvovirus B19 (genus Erythrovirus). Comparison of the HBoV capsid structure to that of parvoviruses from five separate genera demonstrates strong conservation of a β-barrel core domain and an α-helix, from which emanate several loops of various lengths and conformations, yielding a unique surface topology that differs from the three already described for this family. The highly conserved core is consistent with observations for other single-stranded DNA viruses, and variable surface loops have been shown to confer the host-specific tropism and the diverse antigenic properties of this family.Human bocavirus (HBoV), a newly discovered member of the family Parvoviridae, was originally isolated in randomly selected nasopharyngeal aspirates (5). Since this initial discovery, HBoV has also been detected worldwide, predominantly in children under the age of 2 years with respiratory infections, in serum, urine, and fecal samples (40). Symptomatic children commonly exhibit acute diseases of the upper and lower respiratory tracts (7, 36, 44, 56) and, possibly, gastroenteritis (31, 56) though the link to gastroenteritis outbreaks has been questioned (12). It is still unclear if HBoV is the sole etiologic agent of respiratory disease as higher rates of coinfections with other respiratory pathogens such as human rhinovirus and Streptococcus spp. are often observed (4). However, Allander et al. recently reported (4) that HBoV was found in 19% of children with acute wheezing, thereby making it the fourth most common virus, after rhinoviruses, enteroviruses, and respiratory syncytial virus, detected in children exhibiting this symptom. These findings suggest that, at high viral load, HBoV could be an etiologic agent of respiratory tract disease (4). HBoV infection is common in the first few years of life, and clinical research suggests it may follow the primary period for acquisition of human parvovirus B19 (B19) though there is no antigenic cross-reactivity between B19 and HBoV (28, 30). By age 5, most people have circulating antibodies against HBoV, as is also true for other respiratory viruses such as respiratory syncytial virus, rhinoviruses, and human metapneumovirus (17). HBoV has also been identified in adults, with ∼63% of samples tested being seropositive, showing a positive correlation with age and a slight positive bias toward women (14).The Parvoviridae is a family of small, nonenveloped viruses that package a single-stranded DNA (ssDNA) genome of ∼5,000 bases. These viruses are subdivided into two subfamilies: Parvovirinae and Densovirinae (Table (Table1).1). The Parvovirinae are further subdivided into five genera, all of whose members infect vertebrates. The Densovirinae (four genera) infect only invertebrates. Phylogenetic analysis places HBoV in the recently classified Bocavirus genus (Table (Table1).1). In addition to HBoV, numerous parvoviruses circulate among the human population. Among these are the following: several dependoviruses; adeno-associated virus (AAV) serotypes AAV1 to AAV3, AAV5, and AAV9; the Erythrovirus B19; and the newly discovered human parvovirus genotypes 4 (Parv4) and 5 (Parv5) (23, 27, 50). Of these, only B19 had been implicated in disease until the discovery of HBoV and Parv4, which has been isolated from patients who present symptoms of acute HIV infection (50).

TABLE 1.

Selected properties of representative members of the Parvoviridae
Subfamily (host) and genusMember(s)aNo. of VPsbGroupcMajor VP(s)d
Parvovirinae (vertebrate)
    ParvovirusMVM*, CPV*, FPV*3IVP2
    ErythrovirusB19*, SPV2IIIVP2
    DependovirusAAV2*, AAV4*, GPV3IIIVP3
    AmdovirusAMDV2IIIVP2
    BocavirusHBoV, BPV, CnMV2NAVP2
Densovirinae (invertebrate)
    DensovirusGmDNV*, JcDNV4IIVP4
    IteravirusBmDNV4-6IIVP1-VP4
    BrevidensovirusAaeDNV, AalDNV2-3NAVP1 or VP2/3
    PefudensovirusPfDNV5NAVP1
Open in a separate windowaAalDNV, Aedes albopictus densovirus; AaeDNV, Aedes aegypti densovirus; BmDNV, Bombyx mori densovirus; BPV, bovine parvovirus; CnMV, canine minute virus; CPV, canine parvovirus; FPV, feline panleukopenia virus; GPV, goose parvovirus; JcDNV, Junonia coenia densovirus; PfDNV, Periplaneta fuliginosa densovirus; SPV, simian parvovirus. *, structure determined by X-ray crystallography; †, structure determined by cryo-EM.bThe number of VPs in the virion capsid.cGroup refers to the surface topologies described in Results and Discussion (HBoV is currently the only bocavirus with a known structure; there is no structure available for Periplaneta fuliginosa densovirus or the brevidensoviruses). NA, structural group not assigned.dThe VP(s) that comprises most of the wild-type virion.The HBoV genome, like that of all members of the Bocavirus genus, contains three open reading frames (ORFs). The first ORF, at the 5′ end, encodes NS1, a nonstructural protein. The next ORF, unique to the bocaviruses, encodes NP1, a second nonstructural protein. The third ORF, at the 3′ end, encodes the two structural capsid viral proteins (VPs), VP1 and VP2. The HBoV VPs share 42% and 43% amino acid sequence identity with the corresponding VPs of bovine parvovirus and canine minute virus, respectively (5). More recently, two additional HBoV-like viruses, HBoV-2 and HBoV-3, were identified in stool samples from children (8, 31). The genome organization of these viruses is identical to that of HBoV, with the NS1, NP1, and VP proteins of HBoV-2 and HBoV-3 being, respectively, ∼80 and 90%, ∼70 and 80%, and ∼80 and 80% identical to the respective proteins in HBoV (8, 31).Parvovirus genomes are packaged into a T=1 icosahedral capsid that is assembled from 60 copies of a combination of up to six types of capsid VPs (VP1 to VP6), all of which share a C terminus. VP1 is always a minor component, typically comprising about five copies per capsid, whereas the smallest VP is always the major component. The unique N-terminal region of VP1 (VP1u) contains a conserved phospholipase A2 (PLA2) motif within the first 131 amino acids that is essential for infection (49, 61). Interestingly, Aleutian mink disease virus (AMDV), the only member of the Amdovirus genus, is the only exception in that this motif is absent, which suggests that this virus employs a different mechanism to escape the endosome during infection (54).The X-ray crystal structures of several parvoviruses show that all VPs contain a conserved, eight-stranded β-barrel motif (βB to βI) that forms the core of the capsid (15). There is also a conserved α-helix (αA) observed in all parvovirus structures determined to date. The bulk of the VP consists of elaborate loops between the strands that form the surface of the capsid. For example, the GH loop between the βG and βH strands is ∼230 residues. The composition and topology of these loops encode several important functions, including tissue tropism, pathogenicity, and the antigenic response directed against each parvovirus during infection (2).A number of parvoviruses have been studied by cryo-electron microscopy (cryo-EM) and three-dimensional (3D) image reconstruction in concert with and complementary to X-ray crystallographic studies (reviewed in reference 15). Reported here is the 3D structure of a recombinant HBoV capsid solved to 7.9-Å resolution using cryo-EM. The capsid of HBoV was compared to that of representative members of the Parvoviridae (Table (Table1)1) with known atomic structures (AAV2, minute virus of mice [MVM], B19, and Galleria mellonella densovirus [GmDNV]) or pseudo-atomic models built into cryo-EM reconstructed density (AMDV) to identify similarities and differences. The capsid topology of the newly emerging HBoV incorporates a combination of surface structural features seen in other members of the Parvovirinae and is closest to that of B19, the only other structurally characterized parvovirus that is pathogenic to humans. A pseudo-atomic model of the HBoV VP2, built into the reconstructed density, identified conserved core secondary structure elements, which are known to be important for parvovirus capsid assembly, and variable surface loops, which likely govern host specific interactions.  相似文献   
98.
A series of HIV-1 protease mutants has been designed in an effort to analyze the contribution to drug resistance provided by natural polymorphisms as well as therapy-selective (active and non-active site) mutations in the HIV-1 CRF_01 A/E (AE) protease when compared to that of the subtype B (B) protease. Kinetic analysis of these variants using chromogenic substrates showed differences in substrate specificity between pretherapy B and AE proteases. Inhibition analysis with ritonavir, indinavir, nelfinavir, amprenavir, saquinavir, lopinavir, and atazanavir revealed that the natural polymorphisms found in A/E can influence inhibitor resistance. It was also apparent that a high level of resistance in the A/E protease, as with B protease, is due to it aquiring a combination of active site and non-active site mutations. Structural analysis of atazanavir bound to a pretherapy B protease showed that the ability of atazanavir to maintain its binding affinity for variants containing some resistance mutations is due to its unique interactions with flap residues. This structure also explains why the I50L and I84V mutations are important in decreasing the binding affinity of atazanavir.  相似文献   
99.
Integrins have been implicated as coreceptors in the infectious pathways of several nonenveloped viruses. For example, adenoviruses are known to interact with alphaV integrins by virtue of a high-affinity arginine-glycine-aspartate (RGD) domain present in the penton bases of the capsids. In the case of adeno-associated virus type 2 (AAV2), which lacks this RGD motif, integrin alphaVbeta5 has been identified as a coreceptor for cellular entry. However, the molecular determinants of AAV2 capsid-integrin interactions and the potential exploitation of alternative integrins as coreceptors by AAV2 have not been established thus far. In this report, we demonstrate that integrin alpha5beta1 serves as an alternative coreceptor for AAV2 infection in human embryonic kidney 293 cells. Such interactions appear to be mediated by a highly conserved domain that contains an asparagine-glycine-arginine (NGR) motif known to bind alpha5beta1 integrin with moderate affinity. The mutation of this domain reduces transduction efficiency by an order of magnitude relative to that of wild-type AAV2 vectors in vitro and in vivo. Further characterization of mutant and wild-type AAV2 capsids through transduction assays in cell lines lacking specific integrins, cell adhesion studies, and cell surface/solid-phase binding assays confirmed the role of the NGR domain in promoting AAV2-integrin interactions. Molecular modeling studies suggest that NGR residues form a surface loop close to the threefold axis of symmetry adjacent to residues previously implicated in binding heparan sulfate, the primary receptor for AAV2. The aforementioned results suggest that the internalization of AAV2 in 293 cells might follow a "click-to-fit" mechanism that involves the cooperative binding of heparan sulfate and alpha5beta1 integrin by the AAV2 capsids.  相似文献   
100.
A new adeno-associated virus (AAV), referred to as AAV(VR-942), has been isolated as a contaminant of adenovirus strain simian virus 17. The sequence of the rep gene places it in the AAV serotype 2 (AAV2) complementation group, while the capsid is only 88% identical to that of AAV2. High-level AAV(VR-942) transduction activity requires cell surface heparan sulfate proteoglycans, although AAV(VR-942) lacks residues equivalent to the AAV2 R585 and R588 amino acid residues essential for mediating the interaction of AAV2 with the heparan sulfate proteoglycan receptor. Instead, AAV(VR-942) uses a distinct transduction region. This finding shows that distinct domains on different AAV isolates can be responsible for the same activities.  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