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.
Rhizobitoxine is synthesized by the legume symbiont Bradyrhizobium elkanii and the plant pathogen Burkholderia andropogonis. Rhizobitoxine competitively inhibited 1-aminocyclopropane-1-carboxylate (ACC) synthase bLE-ACS2 from the tomato, a key enzyme in the pathway of ethylene biosynthesis. Based on this inhibition of ACC synthase, we have developed a new assay for rhizobitoxine.  相似文献   

2.
1-Aminocyclopropane-1-carboxylate (ACC) oxidase catalyzes the final step in the biosynthesis of the plant signaling molecule ethylene. It is a member of the ferrous iron dependent family of oxidases and dioxygenases and is unusual in that it displays a very short half-life under catalytic conditions, typically less than 20 min, and a requirement for CO2 as an activator. The rates of inactivation of purified, recombinant ACC oxidase from tomato under various combinations of substrates and cofactors were measured. Inactivation was relatively slow in the presence of buffer alone (t1/2 > 1 h), but fast in the presence of ferrous iron and ascorbate (t1/2 approximately 10 min). The rate of iron/ascorbate-mediated inactivation was increased by the addition of ACC, unaffected by the addition of CO2 at saturation (supplied as bicarbonate) but decreased by the addition of catalase or ACC + CO2 at saturation (supplied as bicarbonate). Iron/ascorbate-mediated inactivation was accompanied by partial proteolysis as observed by SDS-PAGE analysis. The fragmentation pattern was altered when ACC was also included, suggesting that ACC can bind to ACC oxidase in the absence of bicarbonate. N-terminal sequencing of fragments resulted in identification of an internal cleavage site which we propose is proximate to active-site bound iron. Thus, ACC oxidase inactivates via relatively slow partial unfolding of the catalytically active conformation, oxidative damage mediated via hydrogen peroxide which is catalase protectable and oxidative damage to the active site which results in partial proteolysis and is not catalase protectable.  相似文献   

3.
4.
OBJECTIVE: To compare the efficacy of the short-acting insulin analog lispro (LP) with that of regular insulin in IDDM patients treated with an external pump. RESEARCH DESIGN AND METHODS: Thirty-nine IDDM patients (age, 39.4 +/- 1.5 years; sex ratio, 22M/17W; BMI, 24.4 +/- 0.4 kg/m2; diabetes duration, 22.5 +/- 1.6 years) who were treated by external pump for 5.1 +/- 0.5 years were involved in an open-label, randomized, crossover multicenter study comparing two periods of 3 months of continuous subcutaneous insulin infusion with LP or with Actrapid HM, U-100 (ACT). Boluses were given 0-5 min (LP) or 20-30 min (ACT) before meals. Blood glucose (BG) was monitored before and after the three meals every day. RESULTS: The decrease in HbA1c was more pronounced with LP than with ACT (-0.62 +/- 0.13 vs. -0.09 +/- 0.15%, P = 0.01). BG levels were lower with LP (7.93 +/- 0.15 vs. 8.61 +/- 0.18 mmol/l, P < 0.0001), particularly postprandial BG levels (8.26 +/- 0.19 vs. 9.90 +/- 0.20 mmol/l, P < 0.0001). Standard deviations of all the BG values (3.44 +/- 0.10 vs. 3.80 +/- 0.10 mmol/l, P = 0.0001) and of postprandial BG values (3.58 +/- 0.10 vs. 3.84 +/- 0.10 mmol/l. P < 0.02) were lower with LP. The rate of hypoglycemic events defined by BG < 3.0 mmol/l did not significantly differ between LP and ACT (7.03 +/- 0.94 vs. 7.94 +/- 0.88 per month, respectively), but the rate of occurrences of very low BG, defined as BG < 2.0 mmol/l, were significantly reduced with LP (0.05 +/- 0.05 vs. 0.47 +/- 0.19 per month, P < 0.05). At the end of the study, all but two (95%) of the patients chose LP for the extension phase. CONCLUSIONS: When used in external pumps, LP provides better glycemic control and stability than regular insulin and does not increase the frequency of hypoglycemic episodes.  相似文献   

5.
The final step in the biosynthesis of the plant signaling molecule ethylene is catalyzed by 1-aminocyclopropane-1-carboxylate (ACC) oxidase, a member of the non-heme iron(II) dependent family of oxygenases and oxidases, which has a requirement for ascorbate as a co-substrate and carbon dioxide as an activator. ACC oxidase (tomato) has a particularly short half-life under catalytic conditions undergoing metal-catalyzed oxidative (MCO) fragmentation. Sequence comparisons of ACC oxidases with isopenicillin N synthase (IPNS) and members of the 2-oxoglutarate Fe(II) dependent dioxygenases show an aspartate and two of six ACC oxidase conserved histidine residues are completely conserved throughout this subfamily of Fe(II) dependent oxygenases/oxidases. Previous mutagenesis, spectroscopic, and crystallographic studies on IPNS indicate that the two completely conserved histidine and aspartate residues act as Fe(II) ligands. To investigate the role of the conserved aspartate and histidine residues in ACC oxidase (tomato fruit), they were substituted via site-directed mutagenesis. Modified ACC oxidases produced were H39Q, H56Q, H94Q, H177Q, H177D, H177E, D179E, D179N, H177D&D179E, H211Q, H234Q, H234D, and H234E. Among those histidine mutants replaced by glutamine, H39Q, H56Q, H94Q, and H211Q were catalytically active, indicating these histidines are not essential for catalysis. Mutant enzymes H177D, H177Q, D179N, H177D&D179E, H234Q, H234D, and H234E were catalytically inactive consistent with the assignment of H177, D179, and H234 as iron ligands. Replacement of H177 with glutamate or D179 with glutamate resulted in modified ACC oxidases which still effected the conversion of ACC to ethylene, albeit at a very low level of activity, which was stimulated by bicarbonate. The H177D (inactive), H177E (low activity), D179E (low activity), and H234Q (inactive) modified ACC oxidases all underwent MCO fragmentation, indicating that they can bind iron, dioxygen, ACC, and ascorbate. The results suggest that MCO cleavage results from active site-mediated reactions and imply that, while H177, D179, and H234 are all involved in metal ligation during catalysis, ligation to H234 is not required for fragmentation. It is possible that MCO fragmentation results from reaction of incorrectly folded or "primed" ACC oxidase.  相似文献   

6.
Dihydropyrimidine dehydrogenase catalyzes, in the rate-limiting step of the pyrimidine degradation pathway, the NADPH-dependent reduction of uracil and thymine to dihydrouracil and dihydrothymine, respectively. The porcine enzyme is a homodimeric iron-sulfur flavoprotein (2 x 111 kDa). C671, the residue postulated to be in the uracil binding site and to act as the catalytically essential acidic residue of the enzyme oxidative half-reaction, was replaced by an alanyl residue. The mutant enzyme was overproduced in Escherichia coli DH5alpha cells, purified to homogeneity, and characterized in comparison with the wild-type species. An extinction coefficient of 74 mM-1 cm-1 was determined at 450 nm for the wild-type and mutant enzymes. Chemical analyses of the flavin, iron, and acid-labile sulfur content of the enzyme subunits revealed similar stoichiometries for wild-type and C671A dihydropyrimidine dehydrogenases. One FAD and one FMN per enzyme subunit were found. Approximately 16 iron atoms and 16 acid-labile sulfur atoms were found per wild-type and mutant enzyme subunit. The C671A dihydropyrimidine dehydrogenase mutant exhibited approximately 1% of the activity of the wild-type enzyme, thus preventing its steady-state kinetic analysis. Therefore, the ability of the C671A mutant and, for comparison, of the wild-type enzyme species to interact with reaction substrates, products, or their analogues were studied by absorption spectroscopy. Both enzyme forms did not react with sulfite. The wild-type and mutant enzymes were very similar to each other with respect to the spectral changes induced by binding of the reaction product NADP+ or of its nonreducible analogue 3-aminopyridine dinucleotide phosphate. Uracil also induced qualitatively and quantitatively similar absorbance changes in the visible region of the absorbance spectrum of the two enzyme forms. However, the calculated Kd of the enzyme-uracil complex was significantly higher for the C671A mutant (9.1 +/- 0.7 microM) than for the wild-type dihydropyrimidine dehydrogenase (0.7 +/- 0.09 microM). In line with these observations, the two enzyme forms behaved in a similar way when titrated anaerobically with a NADPH solution. Addition of an up to 10-fold excess of NADPH to both dihydropyrimidine dehydrogenase forms led to absorbance changes consistent with reduction of approximately 0.5 flavin per subunit, with no indication of reduction of the enzyme iron-sulfur clusters. Absorbance changes consistent with reduction of both enzyme flavins were obtained by removing NADP+ with a NADPH-regenerating system. On the contrary, the two enzyme species differed significantly with respect to their reactivity with dihydrouracil. Addition of dihydrouracil to the wild-type enzyme species, under anaerobic conditions, led to absorbance changes that could be interpreted to result from both partial flavin reduction and the formation of a complex between the enzyme and (dihydro)uracil. In contrast, only spectral changes consistent with formation of a complex between the oxidized enzyme and dihydrouracil were observed when a C671A mutant enzyme solution was titrated with this compound. Furthermore, enzyme-monitored turnover experiments were carried out anaerobically in the presence of a limiting amount of NADPH and excess uracil with the two enzyme forms in a stopped-flow apparatus. These experiments directly demonstrated that the substitution of an alanyl residue for C671 in dihydropyrimidine dehydrogenase specifically prevents enzyme-catalyzed reduction of uracil. Finally, sequence analysis of dihydropyrimidine dehydrogenase revealed that it exhibits a modular structure; the N-terminal region, similar to the beta subunit of bacterial glutamate synthases, is proposed to be responsible for NADPH binding and oxidation with reduction of the FAD cofactor of dihydropyrimidine dehydrogenase. The central region, similar to the FMN subunit of dihydroorotate dehydrogenases, is likely to harbor the site o  相似文献   

7.
Analysis of transgenic mice expressing familial amyotrophic lateral sclerosis (ALS)-linked mutations in the enzyme superoxide dismutase (SOD1) have shown that motor neuron death arises from a mutant-mediated toxic property or properties. In testing the disease mechanism, both elimination and elevation of wild-type SOD1 were found to have no effect on mutant-mediated disease, which demonstrates that the use of SOD mimetics is unlikely to be an effective therapy and raises the question of whether toxicity arises from superoxide-mediated oxidative stress. Aggregates containing SOD1 were common to disease caused by different mutants, implying that coaggregation of an unidentified essential component or components or aberrant catalysis by misfolded mutants underlies a portion of mutant-mediated toxicity.  相似文献   

8.
9.
1-Aminocyclopropane-1-carboxylate (ACC) deaminase catalyzes the cyclopropane ring fragmentation and deamination of ACC. Replacement of cysteine with alanine at a reactive thiol site, Cys-162, of ACC deaminase did not affect the enzyme activity, in spite of the previous result that modification of Cys-162 caused complete loss of the enzyme activity. Substitution of glycine or valine for the cysteine residue gave a higher Km for ACC without a significant change of the K0, indicating that changes of the amino acid side chain had structural effects on substrate binding. Replacement of lysine with alanine at the pyridoxal phosphate (PLP) binding site of the ACC deaminase caused a lower content of PLP and loss of detectable activity of ACC deamination. This mutant enzyme, K51A, showed absorption peaks at 330 nm and 405 nm. The peak at 405 nm was shifted to about 425 nm by the addition of ACC, D-, L-alanine, and D-, L-serine. The formation of aldimine complexes indicated by the spectral shift was reversible. It is suggested that lysine 51 affects the formation of holoenzyme and is important in catalysis.  相似文献   

10.
We report the development of a high-yield heterologous expression system for the copper-containing nitrite reductase from a denitrifying variant of Rhodobacter sphaeroides. Typical yields of wild-type protein are 20 mg L-1, which can be fully loaded with copper. Nitrite reductase contains an unusual blue-green Type 1 copper center with a redox/electron transfer function and a nearby Type 2 center where nitrite binds and is reduced to nitric oxide. The wild-type enzyme was characterized by: (1) its blue-green Type 1 optical spectrum; (2) its EPR spectrum showing rhombic character to its Type 1 center and nitrite perturbation to its Type 2 center; (3) its 247-mV Type 1 midpoint potential which is low relative to other Type 1 centers; and (4) its kinetics as measured by both steady-state and stopped-flow methods. The Type 2 copper reduction potential as monitored by EPR in the absence of nitrite was below 200 mV so that reduction of the Type 2 center by the Type 1 center in the absence of nitrite is not energetically favored. The mutation M182T in which the methionine ligand of Type 1 copper was changed to a threonine resulted in a blue rather than blue-green Type 1 center, a midpoint potential that increased by more than 100 mV above that of the wild-type Type 1 center, and a somewhat reduced nitrite reductase activity. The blue color and midpoint potential of M182T are reminiscent of plastocyanin, but the Type 1 cupric HOMO ground-state electronic g value and copper hyperfine properties of M182T (as well as cysteine and histidine ENDOR hyperfine properties; see next paper) were unchanged from those of the blue-green native Type 1 center. His287 is a residue in the Type 2 region whose imidazole ring was thought to hydrogen bond to the Type 2 axial ligand but not directly to Type 2 copper. The mutation H287E resulted in a 100-fold loss of enzyme activity and a Type 2 EPR spectrum (as well as ENDOR spectra; see next paper) which were no longer sensitive to the presence of nitrite.  相似文献   

11.
X-ray derived structural data predicted that serine 229 was positioned to act as a proton donor to the developing C2 carbanion during the reduction of enolpyruvyl-UDP-N-acetylglucosamine catalyzed by the bacterial peptidoglycan biosynthetic flavoenzyme MurB. To investigate this effect, a mutant where serine 229 was replaced by alanine was constructed and purified. Kinetic analysis of the two half-reactions for the mutant enzyme revealed a 9-fold decrease in the reduction of EFlox by NADPH and a dramatic 10(7)-fold decrease in the reoxidation of EFlred with the enolpyruvyl substrate. In addition, studies of S229A with the substrate analog, (E)-enolbutyryl-UDP-N-acetylglucosamine, showed a striking bias of the partitioning toward formation of the (Z) geometric isomer as opposed to formation of the reduced product UDP-methylmuramic acid, which was the predominant product in wild-type MurB. These studies provide evidence for the proposed role of this active-site serine as a general acid catalyst.  相似文献   

12.
We review and illustrate the wild-type mitotic cycle of Aspergillus nidulans and report the sequence alterations in six mutant alleles of the A. nidulans benA, beta-tubulin, gene. These alleles confer heat sensitivity and resistance to the antifungal, antimicrotubule compound benomyl, and they have been very important in the study of mitosis and microtubule function in A. nidulans. The mutations are novel and fall at amino acids 50, 134, and 257. We have examined the phenotypes conferred by the mutations at restrictive temperatures. None blocks the assembly of microtubules. One allele, benA33, blocks anaphase A and partially inhibits the disassembly of cytoplasmic microtubules in mitosis. We also often observe abnormal spindle morphologies in strains carrying benA33. Another allele, benA31, causes arrest in mitosis with short mitotic spindles and, thus, appears to inhibit spindle elongation.  相似文献   

13.
Crystal structures of the protease of human immunodeficiency virus type 1 (HIV-1) and two mutant proteases, V82D and V82N, have been determined. In all three cases the enzyme forms a complex with the peptidic inhibitor U-89360E. All structures have been determined to 2.3 A resolution and have satisfactory agreement factors: 0.173 for wild type, 0.175 for V82D, and 0.182 for V82N. Comparison of the three crystal structures provides explanations which are consistent with the known kinetic properties of these mutant enzymes with the U-89360E inhibitor [Lin, Y., Lin, X., Hong, L., Foundling, S., Heinrikson, R. L., Thaisrivongs, S., Leelamanit, W., Raterman, D., Shah, M., Dunn, B.M., & Tang, J. (1995) Biochemistry 34, 1143-1152]. Unfavorable van der Waals interactions between the inhibitor and the mutated side chains at position 82 are consistent with diminished affinity for the inhibitor by the mutant enzymes. If a mutation is potentially resistant to an inhibitor, the mutant enzyme should not only have an increased Ki for the inhibitor but should also preserve considerable catalytic capability. The V82D mutant possesses these qualities. In the V82D crystal structure, a water molecule, which connects the protease flap to the inhibitor, is missing or of low occupancy. Absence of this bridge may be important in determining catalytic capability. Moreover, mutation at position 82 induces change in two polypeptide backbone regions, 35-41 and 67-68, which may be related to protease flap mobility.  相似文献   

14.
Structural models of subunit epsilon of the ATP synthase from Escherichia coli have been determined recently by NMR [Wilkens et al. (1995) Nat. Struct. Biol. 2, 961-967] and by X-ray crystallography [Uhlin et al. (1997) Structure 5, 1219-1230], revealing a two-domain protein. In this study, six new epsilon mutants were constructed and analyzed: Y63A, D81A, T82A, and three truncated mutants, tr80(S), tr94(LAS), and tr117(AS). Seven mutants constructed previously were also analyzed: E31A, E59A, S65A, E70A, T77A, R58A, and D81A/R85A. Subunits were purified by isoelectric focusing from extracts of cells that overproduced these 13 mutants. F1 was prepared lacking subunit epsilon by immobilized-Ni affinity chromatography. Three mutants, E70A, S65A, and E31A, showed somewhat higher affinities and extents of inhibition than the wild type. Three mutants, T82A, R85A, and tr94(LAS), showed both lower affinities and extents of inhibition, over the concentration range tested. Two showed no inhibition, D81A and tr80(S). The others, T77A, Y63A, E59A, and tr117(AS), showed lower affinities than wild type, but the extents of inhibition were nearly normal. Results indicate that the C-terminal domain of subunit epsilon contributes to inhibition of ATP hydrolysis, but it is not necessary for ATP-driven proton translocation. Interactions with subunit gamma are likely to involve a surface containing residues S65, E70, T77, D81, and T82, while residues R85 and Y63 are likely to be important in the conformation of subunit epsilon.  相似文献   

15.
Mesoderm formation is critical for the establishment of the animal body plan and in Drosophila requires the snail gene. This report concerns the cloning and expression pattern of the structurally similar gene snail1 from zebrafish. In situ hybridization shows that the quantity of snail1 RNA increases at the margin of the blastoderm in cells that involute during gastrulation. As gastrulation begins, snail1 RNA disappears from the dorsal axial mesoderm and becomes restricted to the paraxial mesoderm and the tail bud. snail1 RNA increases in cells that define the posterior border of each somite and then disappears when somitic cells differentiate. Later in development, expression appears in cephalic neural crest derivatives. Many snail1-expressing cells were missing from mutant spadetail embryos and the quantity of snail1 RNA was greatly reduced in mutant no tail embryos. The work presented here suggests that snail1 is involved in morphogenetic events during gastrulation, somitogenesis and development of the cephalic neural crest, and that no tail may act as a positive regulator of snail1.  相似文献   

16.
A series of benzodiazepine receptor ligands with different chemical structures were evaluated for their affinities at diazepam-sensitive and diazepam-insensitive binding sites for [3H]Ro 15-4513 (ethyl-8-azido-5,6-dihydro-5-methyl-6-oxo-4H-imidazo-[1,5a][1,4] benzodiazepine-3-carboxylate) in cerebellar GABAA receptors. Rats of Wistar strain and of alcohol-sensitive (ANT) and alcohol-insensitive (AT) lines were used. The ANT rats possess a single point mutation in their GABAA receptor alpha 6 subunit, which makes their diazepam-insensitive sites sensitive to benzodiazepine agonists, unlike those of AT and Wistar rats. All compounds evaluated displayed high-affinity binding to diazepam-sensitive sites (Ki < 50 nM). In contrast, a wider range of affinities were observed at diazepam-insensitive sites which depended upon the basic structure and substitutions. The 7- and 8-halogen substituted imidazobenzodiazepines and 12-halogen substituted diimidazoquinazolines displayed the highest affinities (Ki < 15 nM), while intermediate to low affinities (100 < Ki < 4000 nM) were displayed by imidazoquinazolines, thienopyrimidines, one oxoimidazoquinoxaline, and some cyclopyrrolones. The imidazoquinoxalines evaluated displayed the lowest affinity (Ki > 10000 nM). The oxoimidazoquinoxaline, 6-chloro-3-(5-cyclopropyl-1,2,4-oxadiazol-3-yl)-4,5-dihydro-5-isop ropyl-4-oxo-imidazo[1,5-a]quinoxaline (NNC 14-0578) and suriclone represent the first benzodiazepine receptor full agonists to bind with relatively high affinity (Ki approximately 100 nM) to diazepam-insensitive sites. The 5 position substituted methoxybenzyl, dimethylallyl, and 4-fluorobenzyl oxoimidazoquinoxaline analogs demonstrated a 58-336-fold higher affinity for ANT than AT diazepam-insensitive sites.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

17.
Nerve growth factor (NGF) is a required differentiation and survival factor for sympathetic and a majority of neural crest-derived sensory neurons in the developing vertebrate peripheral nervous system. Although much is known about the function of NGF, the intracellular signaling cascade that it uses continues to be a subject of intense study. p21 ras signaling is considered necessary for sensory neuron survival. How additional intermediates downstream or in parallel may function has not been fully understood yet. Two intracellular signaling cascades, extra cellular regulated kinase (erk) and phosphatidylinositol-3 (PI 3) kinase, transduce NGF signaling in the pheochromocytoma cell line PC12. To elucidate the role these cascades play in survival and differentiation, we used a combination of recombinant adenoviruses and chemical inhibitors to perturb these pathways in sensory neurons from wild-type mice and mice deficient for neurofibromin in which the survival and differentiation pathway is constitutively active. We demonstrate that ras activity is both necessary and sufficient for the survival of embryonic sensory neurons. Downstream of ras, however, the erk cascade is neither required nor sufficient for neuron survival or overall differentiation. Instead, the activity of PI 3 kinase is necessary for the survival of the wild-type and neurofibromin-deficient neurons. Therefore, we conclude that in sensory neurons, NGF acts via a signaling pathway, which includes both ras and PI 3 kinase.  相似文献   

18.
The kinetic mechanism of rabbit muscle glycogen synthase I was investigated by determining isotope-exchange rates at chemical equilibrium between uridine diphosphoglucose (UDPG) and glycogen and between UDPG and uridine 5'-diphosphate (UDP). The rates were followed simultaneously by use of UDPG labeled with 14C in the glucose moiety and with 3H in the uracil group. They were found to be independent of the concentrations of glycogen and the UDPG-UDP pair, averaging 6 X 10(-9) mol min-1 mg-1, with a ratio of UDPG-glycogen exchange to UDPG-UDP exchange of 0.85-0.95. The conclusion is that glycogen synthase has a rapid equilibrium random bi bi mechanism. The previously reported slow activation of glycogen-free synthase in the presence of glycogen was examined kinetically. The activation rate appears to be independent of glycogen concentration over a wide range, while the maximum activation is related to the third or fourth root of the glycogen concentration. This suggest that the slow bimolecular reaction mechanism proposed for human polymorphonuclear leucocyte glycogen synthase I [S?lling, H., & Esmann, V. (1977) Eur. J. Biochem. 81, 129] does not apply to rabbit muscle synthase I. The rate of exchange of glycogen molecules in the complex between glycogen and rabbit muscle synthase I under conditions where the enzyme is catalytically active was estimated by a novel method. The enzyme-glycogen complex was treated with [glucose-14C]UDPG and glycogen of different molecular weight. The distribution of isotope between the two forms of glycogen was determined after their separation by agarose gel chromatography. A rate constant of 0.3 min-1 was estimated for the exchange. It can be calculated, on the basis of the specific activity of the enzyme (20 mumol min-1 mg-1) and its action pattern, that hundreds of individual chains in the glycogen molecule must be available to the enzyme during the average lifetime of the complex. A mechanism is proposed for this process.  相似文献   

19.
Programmed cell death plays an essential role in the normal embryonic development of Drosophila melanogaster. One region of the embryo where cell death occurs, but has not been studied in detail, is the abdominal epidermis. Because cell death is a fleeting process, we have used time-lapse, fluorescence microscopy to map epidermal apoptosis throughout embryonic development. Cell death occurs in a stereotypically striped pattern near both sides of the segment border and to a lesser extent in the middle of the segment. This map of wild-type cell death was used to determine how cell death patterns change in response to genetic perturbations that affect epidermal patterning. Previous studies have suggested that segment polarity mutant phenotypes are partially the result of increased cell death. Mutations in wingless, armadillo, and gooseberry led to dramatic increases in apoptosis in the anterior of the segment while a naked mutation resulted in a dramatic increase in the death of engrailed cells in the posterior of the segment. These results show that segment polarity gene expression is necessary for the survival of specific rows of epidermal cells and may provide insight into the establishment of the wild-type epidermal cell death pattern.  相似文献   

20.
Mechanisms responsible for neuroattenuation of herpes simplex virus (HSV) have been defined previously by studies of mutant viruses in cultured cells. The hypothesis that null mutations in host genes can override the attenuated phenotype of null mutations in certain viral genes was tested. Mutants such as those in infected cell protein (ICP) 0, thymidine kinase, ribonucleotide reductase, virion host shutoff, and ICP34.5 are reduced in their capacity to replicate in nondividing cells in culture and in vivo. The replication of these viruses was examined in eyes and trigeminal ganglia for 1-7 d after corneal inoculation in mice with null mutations (-/-) in interferon receptors (IFNR) for type I IFNs (IFN-alpha/betaR), type II IFN (IFN-gammaR), and both type I and type II IFNs (IFN-alpha/beta/gammaR). Viral titers in eyes and ganglia of IFN-gammaR-/- mice were not significantly different from congenic controls. However, in IFN-alpha/betaR-/- or IFN-alpha/beta/gammaR-/- mice, growth of all mutants, including those with significantly impaired growth in cell culture, was enhanced by up to 1,000-fold in eyes and trigeminal ganglia. Blepharitis and clinical signs of infection were evident in IFN-alpha/betaR-/- and IFN-alpha/beta/gammaR-/- but not control mice for all viruses. Also, IFNs were shown to significantly reduce productive infection of, and spread from intact, but not scarified, corneas. Particularly striking was restoration of near-normal trigeminal ganglion replication and neurovirulence of an ICP34.5 mutant in IFN-alpha/betaR-/- mice. These data show that IFNs play a major role in limiting mutant and wild-type HSV replication in the cornea and in the nervous system. In addition, the in vivo target of ICP34.5 may be host IFN responses. These experiments demonstrate an unsuspected role for host factors in defining the phenotypes of some HSV mutants in vivo. The phenotypes of mutant viruses therefore cannot be interpreted based solely upon studies in cell culture but must be considered carefully in the context of host factors that may define the in vivo phenotype.  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